100% found this document useful (1 vote)
405 views469 pages

2015 - Melese, Richter & Solomon (Military Cost-Benefit Analysis)

Uploaded by

Tato
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
100% found this document useful (1 vote)
405 views469 pages

2015 - Melese, Richter & Solomon (Military Cost-Benefit Analysis)

Uploaded by

Tato
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 469

Military Cost–Benefit Analysis

This is the first comprehensive book dedicated to military cost–benefit analysis.


The aim is to help countries identify affordable defense capabilities that effectively
counter security risks in uncertain and fiscally constrained environments. This vol-
ume offers several new practical tools designed to guide defense investments (and
divestments), combined with a selection of real-world applications.
The widespread employment of cost–benefit analysis offers a unique opportu-
nity to transform legacy defense forces into efficient, effective, and accountable
twenty-first-century organizations. A synthesis of economics, statistics, and deci-
sion theory, cost–benefit analysis is currently used in a wide range of defense
applications in countries around the world: i) to shape national security strategy,
ii) to set acquisition policy, and iii) to inform critical investments in people, equip-
ment, infrastructure, services, and supplies. As sovereign debt challenges squeeze
national budgets, and emerging threats disrupt traditional notions of security, this
volume offers valuable tools to navigate the political landscape, meet calls for
fiscal accountability, and boost the effectiveness of defense investments to help
guarantee future peace and stability.
A valuable resource for scholars, practitioners, students, and experts, this book
offers a comprehensive overview of military cost–benefit analysis that will appeal
to anyone interested or involved in improving national security. It should also be
of general interest to public officials responsible for major programs, projects or
policies.

Francois Melese is Professor of Economics at the Defense Resources Manage-


ment Institute (DRMI) in the Graduate School of Business and Public Policy at
the Naval Postgraduate School, USA.

Anke Richter is Professor of Operations Research at the Defense Resources Man-


agement Institute (DRMI) in the Graduate School of Business and Public Policy
at the Naval Postgraduate School, USA.

Binyam Solomon is a Senior Scientist with Defence Research and Development


Canada.
“At a time when government budgets and public sector procurement are under
immense pressure globally, it is imperative that practitioners and students alike
fully understand the appropriate techniques available to properly evaluate such
procurement in the interests of effective decision-making. This book makes a
unique contribution in this regard providing, for the first time, an accessible hand-
book designed to meet this objective which focuses on the military sector where
some of the most difficult decisions are currently having to be made. Both students
and practitioners in the military sector will find this book essential reading and the
ideas contained here will also resonate with those working in or studying other
parts of the public sector.”
Professor Derek Braddon, Emeritus Professor of Economics
University of the West of England, Bristol, UK

“This is a thorough, timely, and up-to-date treatment of cost-benefit analysis in


the defence environment. Its analysis should be helpful to all those potentially and
actually involved not only in weapons system acquisition but in public procure-
ment more broadly.”
Peter Hall, Emeritus Professor, School of Business,
University of New South Wales, Canberra, Australia
Routledge Studies in Defence and Peace Economics
Edited by Keith Hartley, University of York, UK and
Jurgen Brauer, Augusta State University, USA

Volume 1 Volume 7
European Armaments From Defense to Development?
Collaboration International perspectives on
Policy, problems and prospects realizing the peace dividend
R. Matthews A. Markusen, S. DiGiovanna and
M. Leary
Volume 2
Military Production and Volume 8
Innovation in Spain Arms Trade and Economic
J. Molas-Gallart Development
Theory, policy, and cases in arms
Volume 3
trade offsets
Defence Science & Technology
Edited by Jürgen Brauer and
Adjusting to change
J. Paul Dunne
R. Coopey, M. Uttley and
G. Spiniardi Volume 9
Volume 4 Exploding the Myth?
The Economics of Offsets The peace dividend, regions and
Defence procurement and market adjustment
countertrade Derek Braddon
S. Martin Volume 10
Volume 5 The Economic Analysis of
The Arms Trade, Security and Terrorism
Conflict Tilman Brück
Edited by P. Levine and R. Smith
Volume 11
Volume 6 Defence Procurement and
Economic Theories of Peace Industry Policy
and War Edited by Stefan Markowski and
F. Coulomb Peter Hall
Volume 12 Volume 14
The Economics of Defence Policy Military Cost–benefit
A new perspective Analysis
Keith Hartley Theory and practice
Edited by Francois Melese,
Volume 13
Anke Richter and
The Economics of UN
Binyam Solomon
Peacekeeping
Nadège Sheehan

Other titles in the series include:

The Economics of Regional Security


NATO, the Mediterranean, and Southern Africa
Jürgen Brauer and Keith Hartley
In recognition of the

Defense Resources Management Institute (DRMI)


Graduate School of Business and Public Policy
Naval Postgraduate School
Monterey, California
www.nps.edu/drmi

Celebrating 50 Years
Excellence in Education since 1965
This page intentionally left blank
Military Cost–benefit
Analysis
Theory and practice

Edited by Francois Melese,


Anke Richter, and
Binyam Solomon
First published 2015
by Routledge
2 Park Square, Milton Park, Abingdon, Oxon OX14 4RN
and by Routledge
711 Third Avenue, New York, NY 10017
Routledge is an imprint of the Taylor & Francis Group, an informa business

c 2015
Selection and editorial material produced by US Government employees
are works of the US Government and are not subject to copyright in the
United States. Non-US copyrights may apply. Individual chapters produced
by US Government employees are works of the US Government and are
not subject to copyright protection in the United States. Non-US copyrights
may apply.
Selection and editorial material produced by Binyam Solomon are works
of Her Majesty the Queen in Right of Canada.
The right of Binyam Solomon to be identified as an author of the editorial
material, and of the authors for their individual chapters, has been
asserted in accordance with sections 77 and 78 of the Copyright, Designs
and Patents Act 1988.
All rights reserved. No part of this book may be reprinted or reproduced
or utilised in any form or by any electronic, mechanical, or other means,
now known or hereafter invented, including photocopying and recording,
or in any information storage or retrieval system, without permission in
writing from the publishers.
Trademark notice: Product or corporate names may be trademarks or
registered trademarks, and are used only for identification and explanation
without intent to infringe.
British Library Cataloguing in Publication Data
A catalogue record for this book is available from the British Library
Library of Congress Cataloging-in-Publication Data
Military Cost–benefit Analysis: Theory and Practice /
edited by F. Melese, A. Richter, and B. Solomon.
pages cm
Includes bibliographical references and index.
1. United States–Armed Forces–Appropriations and expenditures–
Evaluation. 2. United States. Department of Defense–Appropriations
and expenditures–Evaluation. 3. Armed Forces–Appropriations and
expenditures–Evaluation–Case studies.
4. Cost effectiveness–Government policy–United States.
5. Operations research–United States.
6. United States–Defenses–Costs.
I. Melese, F., editor. II. Richter, A. (Anke), editor.
III. Solomon, B. (Binyam), editor.
UA25.5.M55 2015
355.6’223–dc23 2014041462

ISBN: 978-1-138-85042-2 (hbk)


ISBN: 978-1-315-72469-0 (ebk)

Typeset in Times
by Sunrise Setting Ltd, Paignton, UK
Contents

List of figures xii


List of tables xvi
List of contributors xviii
Preface xxvii
Abbreviations xxviii
Acknowledgments xxxiv

PART I
Introduction and problem formulation 1

1 Introduction: military cost–benefit analysis 3


FRANCOIS MELESE, ANKE RICHTER, AND BINYAM SOLOMON

2 Allocating national security resources 18


JACQUES S. GANSLER AND WILLIAM LUCYSHYN

3 Measuring defense output: an economics perspective 36


KEITH HARTLEY AND BINYAM SOLOMON

4 The economic evaluation of alternatives 74


FRANCOIS MELESE

PART II
Measuring costs and future funding 111

5 Cost analysis 113


DIANA I. ANGELIS AND DAN NUSSBAUM

6 Advances in cost estimating: a demonstration of


advanced machine learning techniques for cost
estimation 136
BOHDAN L. KALUZNY
x Contents
7 Facing future funding realities: forecasting budgets
beyond the future year defense plan 161
ROSS FETTERLY AND BINYAM SOLOMON

PART III
Measuring effectiveness 195
8 Multiple objective decision-making 197
KENT D. WALL AND CAMERON A. MACKENZIE

9 A new approach to evaluate safety and force


protection investments: the value of a statistical life 237
THOMAS J. KNIESNER, JOHN D. LEETH, AND RYAN S. SULLIVAN

PART IV
New approaches to military cost–benefit analysis 261
10 The role of cost-effectiveness analysis in allocating
defense resources 263
KENT D. WALL, CHARLES J. LACIVITA, AND ANKE RICHTER

11 A risk-based approach to cost–benefit analysis:


strategic real options, Monte Carlo simulation,
knowledge value added, and portfolio
optimization 289
JOHNATHAN C. MUN AND THOMAS HOUSEL

12 Extensions of the Greenfield–Persselin optimal fleet


replacement model: applications to the Canadian
Forces CP-140A Arcturus Fleet 313
DAVID W. MAYBURY

PART V
Selected applications 333
13 Embedding affordability assessments in military
cost–benefit analysis: defense modernization
in Bulgaria 335
VENELIN GEORGIEV

14 Real options in military acquisition: a retrospective


case study of the Javelin anti-tank missile system 348
DIANA I. ANGELIS, DAVID FORD, AND JOHN DILLARD

15 An application of military cost–benefit analysis in a major


defense acquisition: the C-17 transport aircraft 363
WILLIAM L. GREER
Contents xi
16 Cost-effectiveness analysis of autonomous aerial platforms
and communication payloads 401
RANDALL E. EVERLY, DAVID C. LIMMER, AND CAMERON A. MACKENZIE

17 Time discounting in military cost–benefit analysis 424


JASON HANSEN AND JONATHAN LIPOW

Index 432
Figures

2.1 Defense and selected entitlement spending as a percentage of


GDP 23
2.2 DoD total budget authority 24
2.3 Resources for Defense in selected years as a percentage of total
DoD budget 25
4.1 When cost is relatively more important than effectiveness 85
4.2 When effectiveness is relatively more important than cost 85
4.3 Fixed budget approach 88
4.4 Fixed effectiveness approach 89
4.5 Economic (expansion path) optimization approach 91
4.6 Modified budget approach 93
4.7 Modified effectiveness approach 94
4.8 Opportunity cost approach 95
4.9 Decision map to structure an EEoA 96
A4.1 The inappropriate application of cost–benefit ratios 98
5.1 Cost estimating process 115
5.2 Program WBS 116
5.3 Cost element structure 117
5.4 Cost estimating techniques as a function of acquisition phases 118
5.5 Steps used to develop a cost estimating relationship 125
5.6 Distribution of cost growth (difference between estimate and
actual costs) 128
5.7 Cost distribution for two alternatives 129
5.8 Cost risk curves for two alternatives 130
6.1 Defense Acquisition University (DAU) integrated defense
acquisition, technology, and logistics life cycle management
framework chart 137
6.2 Hierarchical clustering with simple distance function 142
6.3 Dendrogram illustrating the arrangement of the clusters
produced by the hierarchical clustering of ships (weighted
distance function) 145
Figures xiii
6.4 Weighted hierarchical clustering: correlation plot of actual
versus predicted ship costs 146
6.5 Example M5 model tree 148
6.6 M5 model tree applied to the SAS-076 dataset 150
6.7 M5 model tree: correlation plot of actual versus predicted
ship costs 153
6.8 Plots of residuals for attributes used in the M5 model tree
linear regression models 155
6.9 Unsmoothed M5 model tree: correlation plot of actual versus
predicted ship costs 156
7.1 Trend in defense and non-defense spending 164
7.2 Projected defense funding, budget 2010 165
7.3 Belgian defense expenditures forecast based on an ARIMA
model 175
7.4 Canadian defense expenditures forecast based on constant
GDP growth 180
7.5 Personnel growth projections with and without inflation
adjustment 184
7.6 Demand based on cost driver 184
A7.1 The autocorrelation function plot—Belgian defense
spending 186
A7.2 The partial autocorrelation function plot—Belgian defense
spending 186
8.1 Generic hierarchy 202
8.2 Sloat Radar objectives hierarchy 204
8.3 Radar range value function 208
8.4 Fitted exponential value function for radar range 210
8.5 Radio weight value function 211
8.6 Fitted exponential value function for radio weight 211
8.7 Components of overall effectiveness for Sloat Radar 219
8.8 Sensitivity to w A and wC with w P = 0.35 220
8.9 Sensitivity to w A and wC with w P = 0.2 220
8.10 Supersonic attack scenario 221
8.11 ECM scenario 222
8.12 Alternatives in cost-effectiveness space 223
8.13 Superior or dominant solution 224
8.14 Efficient solution 224
8.15 Satisficing solution 225
8.16 Marginal reasoning solution 226
8.17 Cost-effectiveness linear preference 228
8.18 WC  W E 228
8.19 W E > WC 229
8.20 W E  WC 229
8.21 Cost-effectiveness: WC /W E = 1.0 230
8.22 Cost-effectiveness: WC /W E = 0.5 230
xiv Figures
8.23 Cost-effectiveness: WC /W E = 0.25 231
8.24 Cost-effectiveness: WC /W E = 0.33 232
8.25 Cost-effectiveness for supersonic attack scenario:
WC /W E = 0.57 233
8.26 Cost-effectiveness for ECM scenario: WC /W E = 0.74 233
A8.1 Linear value function for cost 234
9.1 Labor market equilibrium with discrete risk 241
9.2 Labor market equilibrium with continuous risk 242
9.3 Supply of voluntary enlistments and the Vietnam draft 245
10.1 The two-level hierarchy 265
10.2 Two-dimensional example of primal solution 268
10.3 Two-dimensional example of dual solution 272
10.4 Continuous case with B = 2.5 278
10.5 Continuous case with B = 4.5 279
10.6 Continuous case with B = 6.5 280
10.7 Discrete case alternatives 282
10.8 Discrete case with B = 2.5 283
10.9 Discrete case with B = 4.5 284
10.10 Discrete case with B = 6.5 285
11.1 Why is risk important? 291
11.2 Adding an element of risk 291
11.3 Single-point estimates 292
11.4 Simulation results 294
11.5 Example real options framing 299
11.6 Portfolio optimization and allocation 302
11.7 Efficient frontiers of portfolios 303
11.8 Portfolio optimization (continuous allocation of funds) 304
11.9 Integrated risk analysis process 306
12.1 CP-140 Arcturus data with the critical region indicated 321
12.2 CP-140 Arcturus data with no stochastic discount factor
shown with the critical region indicated 322
12.3 CP-140 Arcturus data hitting time distribution 324
13.1 Variables used in AoA and EEoA 339
13.2 The MoE hierarchy diagram 341
13.3 Example of an MoE hierarchy diagram 341
13.4 Possible approaches to structure an EEoA 345
14.1 The Javelin anti-tank weapon system missile and command
launch unit 349
14.2 Deterministic cost-effectiveness 354
14.3 Probabilistic cost-effectiveness 357
15.1 Aircraft size and capacity comparisons 368
15.2 MRS airlift delivery requirements for MRC-East and
MRC-West combined 374
15.3 Range payload curves for airlifter types 375
Figures xv
15.4 Comparison of cost and effectiveness of alternatives with
current trucks 380
15.5 Comparison of cost and effectiveness of alternatives with new
army FMTV trucks 383
15.6 Comparison of selected alternatives with different
assumptions about MTV loading on militarized 747 383
15.7 Impact of reduced MOG and reduced C-17 use rates 385
15.8 Cost and effectiveness comparisons of new alternatives 387
15.9 Comparison of 59 MTM/D expanded capacity alternatives
with the nominal 52 MTM/D alternatives 388
15.10 Comparisons of alternatives in LRC-Short 389
15.11 Effect of FYDP stretch-out on alternatives in MRC 390
15.12 Effect of discount rates on two alternatives 391
16.1 Objectives hierarchy for aerial platform 407
16.2 Objectives hierarchy for communication payload 408
16.3 Annualized O&S cost as a function of acquisition cost 409
16.4 Cost-effectiveness of aerial platforms for disaster relief
scenario 411
16.5 Cost-effectiveness of communication payloads for disaster
relief scenario 413
16.6 Cost-effectiveness of aerial platforms for long-range scenario 415
16.7 Cost-effectiveness of communication payloads for long-range
scenario 416
16.8 Cost-effectiveness of aerial platforms for tactical user
scenario 418
16.9 Cost-effectiveness of communication payloads for tactical
user scenario 419
Tables

3.1 Defense inputs for a group of nations (2009) 43


3.2 Annual defense expenditures FY 2012–13 46
3.3 MoD resources by budget area FY 2010–11 48
A3.1 Canadian Program Alignment Architecture 2011–12 67
4.1 Six approaches to structuring an EEoA 87
4.2 Three-stage multi-attribute procurement auction—economic
(expansion path) optimization approach (Simon and Melese
2011) 91
5.1 Advantages and disadvantages of cost estimating
methodologies 126
6.1 Categories of ship data 140
6.2 Principal component analysis and optimal macro-attribute
weights 144
6.3 Weighted distance of the Rotterdam LPD to ships in the
SAS-076 dataset 146
6.4 M5 model tree linear regression models 151
6.5 Statistics of attributes used in the M5 model tree linear
regression models 152
6.6 Mean absolute percentage errors of known instances per
individual M5 model tree linear model 154
6.7 Multiple linear regression F-statistics and p-values 157
6.8 Comparison of the M5 model tree and hierarchical clustering
methods 158
7.1 Real GDP per capita growth for selected NATO members 163
7.2 Government annual surplus/deficit per GDP 164
7.3 Long-run coefficients 178
7.4 Error correction representation 179
7.5 Key variables to forecast or simulate 181
8.1 Variable definitions 201
8.2 Evaluation data for Sloat Radar 218
8.3 Values and weights for Sloat Radar 218
Tables xvii
8.4 Discounted life-cycle costs and MoEs for Sloat Radar 223
10.1 Border security alternatives 281
10.2 Coastal security alternatives 281
13.1 MoP specifications for the multi-role fighter 342
13.2 MoE/MoP for multi-role fighter alternatives 343
13.3 Time characteristics assumed for the alternatives 346
14.1 Anti-tank missile guidance system effectiveness 353
14.2 Anti-tank missile cost 354
14.3 Marginal analysis of cost and effectiveness for LBR and FO 355
14.4 Marginal analysis of cost and effectiveness for FO and FLIR 355
14.5 Probability of development success and expected MoE for
Javelin technology options 356
14.6 Expected cost of Javelin guidance technology alternatives
without option to terminate project 357
14.7 Value of Javelin guidance technology options 358
15.1 C-17 decision timelines 365
15.2 Alternatives with 52 MTM/D capacity 372
15.3 Airlifter surge use rates 376
15.4 MOG estimates maximum number of aircraft on ground
simultaneously (by theater and aircraft type) 376
15.5 Summary of O&S costs per PAA 379
15.6 Summary of C-17 acquisition decisions through FY 2010 393
16.1 Global tradeoff weights for aerial platform for each scenario 410
16.2 Global tradeoff weights for communication payload for each
scenario 412
17.1 Future value of $100 425
17.2 Present value of $100 received in the future 425
17.3 Comparative life-cycle costs 426
17.4 Parameter definitions 429
17.5 Expected cost and utility of propulsion systems 429
Contributors

Diana I. Angelis is an associate professor at the Naval Postgraduate School in the


Defense Resource Management Institute with a joint appointment to the Depart-
ment of Systems Engineering. She studied accounting at the University of Florida
and received a BS in Business Administration in 1977 and a BS in Electrical Engi-
neering in 1985. She received her PhD in Industrial and Systems engineering from
the University of Florida in 1996. Her research interests include cost accounting,
activity-based costing, valuation of R&D and acquisition innovation. She was
commissioned an officer in the US Air Force in 1984 and served as a program
engineer until 1989. Dr Angelis is a Certified Public Accountant and retired in
2009 as a Lieutenant Colonel in the US Air Force Reserve. She joined the Defense
Resources Management Institute’s faculty in 1996.
Col. John T. Dillard, USA (Ret.), managed major weapons development efforts
for most of his 26-year career in the US Army. He is now the Acquisition
Academic Area Chair, Graduate School of Business and Public Policy, at the
US Naval Postgraduate School in Monterey, California. Col. Dillard is also a
1988 graduate of the Defense Systems Management College.
Randall E. Everly Commander, US Navy, is a graduate of the US Naval
Academy with a Bachelor of Science in Electrical Engineering. Originally from
Plant City, Florida, Commander Everly is a Naval Aviator. He flew C-2A Grey-
hounds and E-2C Hawkeyes operationally, and instructed in the T-6 Texan
II aircraft. He graduated from the Naval Postgraduate School with Master
of Science in Information Technology Management and a Master of Busi-
ness Administration. Commander Everly currently serves as the Deputy N7
in charge of Training and Education for Commander, Navy Reserve Forces
Command in Norfolk, Virginia.
Col. Ross Fetterly is currently the Canadian Air Force Comptroller. He previ-
ously served as the Military Personnel Command Comptroller. Col. Fetterly has
also been employed as the Section Head in Director Strategic Finance and Cost-
ing (DSFC) within Assistant Deputy Minister (Finance) at National Defence
Headquarters responsible for costing analysis of all capital projects and major
departmental initiatives, as well as the Section Head responsible for Strategic
Contributors xix
Finance. Col. Fetterly completed a tour in February 2009 as the Chief CJ8 at
COMKAF HQ, the NATO Base HQ at Kandahar Airfield, Afghanistan. While
deployed he wrote the paper entitled “Methodology for Estimating the Fiscal
Impact of the Costs Incurred by the Government of Canada in Support of the
Mission in Afghanistan” with staff from the Parliamentary Budget Office. Col.
Fetterly was employed as the Deputy Commanding Officer of the Canadian
Contingent in the United Nations Disengagement Observer Force in the Golan
Heights during the second intifada in 2000–2001. He has served as an Air Force
Squadron Logistics Officer and at military bases across Canada. An Adjunct
Assistant Professor at the Royal Military College of Canada (RMC), he teaches
financial decision-making. Col. Ross Fetterly earned his PhD (War Studies) at
the RMC. His fields of study were Defence Economics, Canadian Defence Pol-
icy and Defence Cost Analysis. His current research focus is defense resource
management.

Dr David N. Ford is an Associate Professor at Texas A&M University and the


US Naval Postgraduate School. In addition to teaching, he researches develop-
ment project strategy, processes, and resources management. Dr Ford earned
his doctorate from the Massachusetts Institute of Technology, and his mas-
ter’s and bachelor’s degrees from Tulane University. He has over 14 years of
engineering and project management experience in industry and government.

The Honorable Jacques S. Gansler, former US Under Secretary of Defense for


Acquisition, Technology, and Logistics, is a Professor and holds the Roger C.
Lipitz Chair in Public Policy and Private Enterprise in the School of Public
Policy at the University of Maryland, and is the Director of the Center for Pub-
lic Policy and Private Enterprise. As the third-ranking civilian at the Pentagon
from 1997 to 2001, Professor Gansler was responsible for all research and
development, acquisition reform, logistics, advance technology, environmen-
tal security, defense industry, and numerous other security programs. Before
joining the Clinton Administration, Dr Gansler held a variety of positions in
government and the private sector, including Deputy Assistant Secretary of
Defense (Material Acquisition), Assistant Director of Defense Research and
Engineering (electronics), executive vice president at TASC, vice president of
ITT, and engineering and management positions with Singer and Raytheon
Corporations. Throughout his career, Dr Gansler has written, published, tes-
tified, and taught on subjects related to his work. He is the author of five books
and over 100 articles. His most recent book is Democracy’s Arsenal: Creating
a 21st Century Defense Industry (MIT Press, 2011). He is a member of the
National Academy of Engineering and a Fellow of the National Academy of
Public Administration. Additionally, he is the Glenn L. Martin Institute Fellow
of Engineering at the A. James Clarke School of Engineering, an Affiliate Fac-
ulty member at the Robert H. Smith School of Business and a Senior Fellow at
the James MacGregor Burns Academy of Leadership (all at the University of
Maryland).
xx Contributors
Dr Venelin Georgiev is a senior researcher in Department of Defense Resource
Management at the Bulgarian Academy of Sciences. He has a distinguished
career in the defense establishment with responsibilities that included design-
ing processes and systems for resource management and managing projects.
Dr Georgiev has held the positions of Chief of Defence Resource Program-
ming Section and State Expert in Defence Acquisition Policy Directorate in
the Bulgarian MoD. His research interests include defense policy, program
and project approaches for economic development, defense resource manage-
ment, risk management, innovation, investment, micro and macro-economics,
administration, and institutional economics.
William L. Greer is Assistant Director of the System Evaluation Division at the
Institute for Defense Analyses (IDA), a non-profit federally funded research
and development center in Alexandria, VA. He specializes in air mobility cost
and effectiveness analyses and has been involved in several recent large studies
that address strategic and tactical airlift demands and the next-generation air-
borne tanker. Other work includes analyses of the Joint Strike Fighter, USAF
long-range bomber force size requirements, ballistic missile defenses and naval
surface forces. Prior to working at IDA, Dr Greer worked at the Office of
the Secretary of Defense in the Pentagon, the Center for Naval Analyses in
Alexandria, VA, and the Chemistry Department of George Mason University in
Fairfax, VA. He has a PhD in chemical physics from the University of Chicago
and a BA in chemistry from Vanderbilt University.
Jason K. Hansen, PhD, is an economist in the modeling and simulation group
at the Idaho National Laboratory (INL). Prior to joining INL, he served on
the faculty at the Naval Postgraduate School in the Defense Resources Man-
agement Institute as an assistant professor. Dr Hansen’s research interests are
in the economics of water resources, energy and public policy. His work is
published in journals such as the Journal of Benefit Cost Analysis, Journal of
American Water Works Association, Hydrogeology Journal, and the Journal of
Environmental Management.
Professor Keith Hartley is Emeritus Professor of Economics at the University of
York where he was previously Director of the Centre for Defence Economics.
He was founding Editor of the journal Defence and Peace Economics (with
Todd Sandler) and remained Editor from 1990 to 2007 and is currently Special
Advisor to the Editor. He has acted as adviser and consultant to the UN, EC,
European Defence Agency, various UK Government Departments and to the
Parliamentary Defence Committee. He was a NATO Fellow and QinetiQ Vis-
iting Fellow. Current research interests are in defense economics and include
procurement, pricing, the costs of conflict, measuring defense output and the
political economy of the aerospace industry.
Thomas Housel specializes in valuing intellectual capital, telecommunications,
information technology, value-based business process reengineering, and
knowledge value measurement in profit and non-profit organizations. He
Contributors xxi
received his PhD from the University of Utah in 1980 and is currently a tenured
Full Professor for the Information Sciences (Systems) Department at the Naval
Postgraduate School in Monterey, California. His current research focuses on
the use of Real Options models in identifying, valuing, maintaining, and exer-
cising options in military decision-making. Prior to joining NPS, he also was
a Research Fellow for the Center for Telecommunications Management and
Associate Professor at the Marshall School of Business at the University of
Southern California. Housel has been the Chief Business Process Engineer for
Pacific Bell. He is Managing Partner for Business Process Auditors, a firm that
specializes in training Big Six consultants, large manufacturing and service
companies in the Knowledge Value Added methodology for objectively mea-
suring the return generated by corporate knowledge assets/intellectual capital.
His latest books include Measuring and Managing Knowledge (McGraw-Hill,
2001) and Global Telecommunications Revolution: The Business Perspective
(McGraw-Hill, 2001).
Bohdan L. Kaluzny is a defense scientist with Defence Research and Devel-
opment Canada, Centre for Operational Research and Analysis. Dr Kaluzny
obtained his doctorate degree in computer science from McGill University in
Montreal and his research interests include polyhedral computation, compu-
tational geometry, combinatorial optimization, multi-criteria decision analysis,
and operational research.
Thomas J. Kniesner was born in Cleveland, Ohio, and received his PhD degree
in Economics from The Ohio State University. He has recently joined the fac-
ulty of Claremont Graduate University as University Professor and continues
to be a Research Fellow at the Institute for the Study of Labor (IZA). Dr Knies-
ner’s specialty is the econometric examination of labor and health economic
issues. His interests are labor supply, workplace safety, and health care costs
and use. He has published articles in over 20 different professional journals
including The American Economic Review, Econometrica, Journal of Political
Economy, Review of Economics and Statistics, Journal of Economic Litera-
ture, Journal of the European Economic Association, Journal of Mental Health
Policy and Economics, Journal of Monetary Economics, Industrial and Labor
Relations Review, Journal of Labor Economics, Journal of Risk and Uncer-
tainty, Labour Economics, International Economic Review, Journal of Policy
Analysis and Management, Journal of Health Policy, Politics and Law, Health
Affairs, The Economics of Neuroscience, and Regulation. He is the co-author
of seven books, including Labor Economics: Theory, Evidence, and Policy,
Simulating Workplace Safety Policy, The Law and Economics of Workers’ Com-
pensation Insurance, and The Effects of Recent Tax Reforms on Labor Supply.
He is has served as co-editor of the Journal of Human Resources, co-editor of
Foundations and Trends in Microeconomics, and associate editor of the Journal
of Risk and Uncertainty.
xxii Contributors
C. J. LaCivita, Professor Emeritus, graduated in l969 from the University of
Detroit with a Bachelor of Electrical Engineering. He received an MBA from
Valdosta State College in 1975, and a PhD in Economics from the University
of California at Santa Barbara in l98l. Professor LaCivita served as a pilot in
the US Air Force for five years. His current research concerns the relation-
ship between accounting costs and economic costs and their use in promoting
more efficient management of defense resources. He is a member of the Ameri-
can Economic Association and the American Society of Military Comptrollers.
He is also a member of the commission that developed and oversees the Cer-
tified Defense Financial Manager Program. He was Assistant Professor of
Economics at the University of North Carolina at Greensboro before joining
the Defense Resources Management faculty in May, l985. He served as Exec-
utive Director from 1993 to 2011 and was the Acting Dean of the School of
International Graduate Studies at the Naval Postgraduate School from 2001 to
2002. He also served as a member of the oversight commission for the Certi-
fied Defense Financial Manager program from its inception through 2008. He
has published widely in a number of areas in economics and defense resources
management.
John D. Leeth is a Professor of Economics at Bentley University. He received
a BA degree in political science from the University of Southern California
and a PhD in economics from the University of North Carolina, Chapel Hill.
Before coming to Bentley University in 1987, he was on the faculty of the
Edwin L. Cox School of Business, Southern Methodist University in Dallas,
Texas. His research areas include workplace safety, executive compensation,
economic inequality, and mergers and acquisitions. His articles have appeared
in various journals including British Journal of Management, Industrial and
Labor Relations Review, International Economic Review, International Jour-
nal of Industrial Organization, Journal of Finance and Quantitative Analysis,
Journal of Risk and Uncertainty, and Regulation. He co-authored the book
Simulating Workplace Safety Policy (Kluwer Academic Publishers, 1995).
David C. Limmer Lieutenant, US Navy, earned his BS in Computer Science
from Gordon College (2000) and his MS in Information Technology Man-
agement and MBA in Information Systems Management from the Naval
Postgraduate School (2014). Lieutenant Limmer was commissioned through
Officer Candidate School in Pensacola, FL and has since served as a Naval
Flight Officer and Information Professional. He has flown the EA-6B and has
been stationed overseas in Japan, Guam, and Italy.
Jonathan Lipow, PhD, is currently an Associate Professor at the Defense
Resources Management Institute (DRMI) located at the Naval Postgraduate
School (NPS). Prior to joining DRMI, he served as a professor at the Hebrew
University of Jerusalem and at Oberlin College, acted as Bank of America’s
country representative in Israel, and worked as an economic advisor to Israel’s
Ministry of Defense. His military service was in the Israel Defense Forces’
Contributors xxiii
Combat Engineering Corps. His research has focused on a wide variety of
topics related to national security and defense economics, and he has published
papers in journals such as Defense and Peace Economics. Southern Economic
Journal, Economics Letters, World Development, and Journal of Economic
Behavior and Organization.
William Lucyshyn is the Director of Research and Senior Research Scholar, at
the Center for Public Policy and Private Enterprise, in the School of Public
Policy, at the University of Maryland. In this position, he directs research on
critical policy issues related to the increasingly complex problems associated
with improving public sector management and operations, and how govern-
ment works with private enterprise. Current projects include: modernizing
government supply chain management, identifying government sourcing and
acquisition best practices, and department of defense business modernization
and transformation. Previously, Mr. Lucyshyn served as a program manager
and the principal technical advisor to the Director of the Defense Advanced
Research Projects Agency (DARPA) on the identification, selection, research,
development, and prototype production of advanced technology projects. Prior
to joining DARPA, Mr. Lucyshyn completed a 25-year career in the US
Air Force. Mr. Lucyshyn received his Bachelor Degree in Engineering Sci-
ence from the City University of New York, and earned his Master’s Degree
in Nuclear Engineering from the Air Force Institute of Technology. He has
authored numerous reports, book chapters, and journal articles.
Cameron A. MacKenzie is an Assistant Professor in the Defense Resources
Management Institute at the Naval Postgraduate School. His research and
teaching focus on decision and risk analysis, with a particular emphasis on
modeling the economic and business impacts caused by disruptions. He has
analyzed the economic impacts caused by the 2011 Japanese earthquake and
tsunami, and he has developed a resource allocation model to help an economic
region recover from a disaster like the Deepwater Horizon oil spill. Previously,
he consulted in the areas of defense and homeland security for former Defense
Secretary William Cohen. He received a BS and BA from Indiana-Purdue Uni-
versity at Fort Wayne (2001), an MA in International Affairs from The George
Washington University (2003), an MS in Management Science and Engineer-
ing from Stanford University (2009), and a PhD in Industrial Engineering from
the University of Oklahoma (2012).
David Maybury received his BSc in Applied Mathematics from the University
of Western Ontario and his PhD in Physics from the University of Alberta.
He was a postdoctoral fellow with the Rudolf Peierls Centre for Theoretical
Physics, University of Oxford, and with the high energy physics group at Car-
leton University. Currently, David is an operational research scientist, focusing
on military applications of financial engineering, with Defence Research and
Development Canada.
xxiv Contributors
Francois Melese, is Professor of Economics and former Executive Director of
the Defense Resources Management Institute at the Naval Postgraduate School
([email protected]). He earned his BA in Economics at the University of Cali-
fornia, Berkeley, his MA at the University of British Columbia, Canada, and
his Doctorate at the Catholic University of Louvain, Belgium. He has over
25 years of experience conducting courses and workshops for military and
civilian officials around the world. In 2005 Dr Melese participated in the US
Defense Department’s Quadrennial Defense Review (QDR), and in 2008 con-
tributed to the Department’s first Strategic Management Plan. He has consulted
extensively, including for the Joint Chiefs of Staff (Comptroller), the Defense
Business Board, and the Office of the Secretary of Defense (Directorate for
Organizational & Management Planning). Dr Melese has earned several awards
for teaching and research, and has published over 50 papers and book chapters
on a variety of topics in economics and management including: public budget-
ing, economic development, energy markets, international trade, applied game
theory, labor markets, public procurement, and defense management. At the
request of NATO Headquarters and the US State Department he has repre-
sented the US as an expert in public budgeting and defense management at
NATO meetings throughout Europe.
Dr Johnathan C. Mun is a research professor at the US Naval Postgraduate
School (Monterey, California) and teaches executive seminars in quantitative
risk analysis, decision sciences, real options, simulation, portfolio optimiza-
tion, and other related concepts. He has also researched and consulted on
many Department of Defense and Department of Navy projects and is con-
sidered a leading world expert on risk analysis and real options analysis. Dr.
Mun received his PhD in Finance and Economics from Lehigh University,
has an MBA in business administration, an MS in management science, and
a BS in Biology and Physics. He is Certified in Financial Risk Management
(FRM), Certified in Financial Consulting (CFC), and Certified in Risk Man-
agement (CRM). He has authored 12 books including Modeling Risk: Applying
Monte Carlo Risk Simulation, Real Options, Optimization, and Forecasting,
First and Second Editions (Wiley 2006, 2010); Real Options Analysis: Tools
and Techniques for Valuing Strategic Investments and Decisions, First and
Second Editions (Wiley 2003, 2006); Real Options Analysis Course: Busi-
ness Cases (Wiley 2003); Applied Risk Analysis: Moving Beyond Uncertainty
(Wiley 2003); Valuing Employee Stock Options (Wiley 2004), and others. He
is the founder and CEO of Real Options Valuation, Inc., a consulting, training,
and software development firm.
Dan Nussbaum, Professor at the Naval Postgraduate School, chairs the NPS
Energy Academic Group, and provides leadership to the US Secretary of
Navy’s Executive Energy Education program. He teaches courses in cost esti-
mating and analysis, and provides cost estimating and business case analyses
for DoD organizations. He designs, develops and delivers distance learning
Contributors xxv
courses in cost estimating and analysis. Prior to this position, Dan was a Prin-
cipal, Booz Allen Hamilton. He has also been the Director, Naval Center for
Cost Analysis, Office of the Assistant Secretary of Navy (Financial Manage-
ment and Comptroller), Washington, DC. He directed all Navy independent
cost estimates as required by Congress and senior Defense leadership on ships,
aircraft, missiles, electronics, and automated information systems. Dr. Nuss-
baum has a BA in Mathematics and Economics, Columbia University; a PhD
in Mathematics, Michigan State University; a Fellowship from National Sci-
ence Foundation in Econometrics and Operations Research at Washington State
University; and a fellowship in National Security from Harvard University’s
Kennedy School of Government.
Anke Richter, is a Professor of Operations Research at the Defense Resources
Management Institute at the Naval Postgraduate School. She received a BA in
Mathematics and French from Dartmouth College (1991) and a PhD in Oper-
ations Research from Stanford University (1996). Her graduate work was sup-
ported by a grant from the Office of Naval Research. Dr. Richter was previously
a Director of Health Outcomes at RTI-Health Solutions, RTI International.
Her research interests include resource allocation for epidemic control, disease
modeling and economic impact assessment, and bio terrorism. Dr. Richter is
a member of the Institute for Operations Research and the Management Sci-
ences (INFORMS) and the International Society for Pharmacoeconomics and
Outcomes Research (ISPOR). She has published in several peer-reviewed jour-
nals, including the Journal of the American Medical Association, Journal of
Clinical Epidemiology, PharmacoEconomics, Medical Decision Making, Clin-
ical Therapeutics and Managed Care Interfaces. While English is Dr. Richter’s
first language, she is also fluent in German and French.
Dr Binyam (Ben) Solomon is a senior scientist with Defence Research and
Development Canada (DRDC), a special operating agency of the Canadian
Department of National Defence (DND). Dr. Solomon is also an Adjunct
Research Professor at Carleton University Canada and Co-Director of the
Institute of Defence Resources Management (IDRM) at the Royal Military
College of Canada. He has previously worked for the Department of National
Defence as Chief Economist. He earned his BA in Mathematics, Statistics and
Economics at the University of Regina, Canada, his Masters in Economics
at the University of Ottawa, Canada, and his Doctorate at the University of
York, United Kingdom. Dr. Solomon has more than 20 years of experience
in economics research consulting and teaching. He has published articles and
reports on various quantitative and defense economics issues ranging from
time series analysis and economic modeling to the defense industrial base,
peacekeeping, and defense resources management. Dr Solomon has repre-
sented Canada at a number of NATO research panels including a NATO System
Analysis and Studies (SAS) on benchmarking (2006–2008), SAS 090 on the
economics of international collaboration (2011) a Russia-NATO workshop on
civil-military institutions (2003) and a NATO-Canada study on allied military
xxvi Contributors
training in Canada (1996). His research interests include economic aspects of
international security, time series econometrics and defense resources manage-
ment. Recently, he was named the Americas Editor of the Defence and Peace
Economics Journal (Routledge).
Ryan Sullivan, Assistant Professor, received a PhD in economics from Syracuse
University in 2010. Dr Sullivan joined the faculty of the Defense Resources
Management Institute (DRMI) in that same year and has taught a variety of top-
ics related to cost–benefit and cost-effectiveness analysis, marginal reasoning,
budgeting, labor economics, and game theory. His research interests include
program cost–benefit analyses with a specialization in value of statistical life
(VSL) studies. He has published in several peer-reviewed journals, including
American Economic Journal: Economic Policy, Defence and Peace Economics,
Economic Inquiry, National Tax Journal, Public Budgeting and Finance, and
Public Finance Review, among others. His work has been discussed in such
prominent outlets as Time Magazine, USA Today, and US News and World
Report. Dr Sullivan is a member of the American Economic Association and
the National Tax Association. He served as a soldier in the US Army National
Guard from 1998 to 2006.
Kent D. Wall, is a Professor at the Defense Resources Management Institute at
the Naval Postgraduate School. He earned a PhD in Control Sciences from
the University of Minnesota. After completing his studies he was awarded two
postdoctoral fellowships in England, the first with the University of Manch-
ester, and the second with the University of London. While in the UK he
lectured at HM Treasury, the Bank of England, Queen Mary College, Impe-
rial College, London School of Economics, and London Business School. He
returned to the US as a Research Associate with the National Bureau of Eco-
nomic Research in Cambridge, Massachusetts. Before coming to the Naval
Postgraduate School he was an Associate Professor of Systems Engineering
at the University of Virginia. His research interests focus on the development
of quantitative aids in decision-making. He has published his work in many
scholarly journals, and has been invited to present special courses at the Uni-
versity of Paris IX (Dauphine). He joined the faculty in August 1985 and served
as Assistant Director for Academic Programs from 1993 to 1998.
Preface

This edited volume should appeal to anyone interested or actively involved in


improving national security. It should also be of general interest to those respon-
sible for major government programs, projects or policies. A valuable resource
for scholars and practitioners, novices and experts alike, this book offers a com-
prehensive overview of military cost–benefit analysis (CBA). The goal is to help
countries identify affordable defense capabilities that effectively counter security
risks, in fiscally constrained environments.
The book consists of five parts: I) Introduction and problem formulation;
II) Measuring costs and future funding; III) Measuring effectiveness; IV) New
approaches to military cost–benefit analysis; and V) Selected applications. The
seventeen chapters that make up these five parts showcase a diversity of interna-
tional experts with both theoretical and hands-on experience. Lifting the veil on
military CBA, this volume offers several new practical tools designed to guide
defense investments (and divestments), combined with a selection of real-world
applications.
Widespread employment of CBA offers a unique opportunity to transform
legacy defense forces into efficient, effective, and accountable twenty-first-century
organizations. A synthesis of economics, management science, statistics and
decision theory, CBA is currently used in a wide range of defense applica-
tions in countries around the world: i) to shape national security strategy, ii) to
set acquisition policy, and iii) to inform critical investments in people, equip-
ment, infrastructure, services and supplies. As sovereign debt challenges squeeze
national budgets, and emerging threats disrupt traditional notions of security, this
volume offers valuable tools to navigate the political landscape, meet calls for
fiscal accountability, and boost the effectiveness of defense investments to help
guarantee future peace and stability.
Abbreviations

ABC activity based costing


ACAS airlift cycle analysis spreadsheet
ACF autocorrelation function
ADF augmented Dickey–Fuller
ADF Australian Defense Force
AFB air force base
AFM airlift flow model
AFSAA air force studies and analysis agency
AIMP aurora incremental modernization program
ALM airlift loading model
AMC Air Mobility Command
AMOS air mobility operations model
AoA analysis of alternatives
APC armored personnel carriers
ARDL autoregressive distributive lag
ARIMA autoregressive integrated moving average
AVF all-volunteer force
BCA business case analysis
BLS Bureau of Labor Statistics
BRAC base realignment and closure
BSC balanced scorecard
CAF Canadian armed forces
CAG Comptroller and Auditor General
CAIG Cost Analysis Improvement Group
CAIV cost as an independent variable
CAPE cost assessment and program evaluation (formerly PA&E)
CARD cost analysis requirements description
CBA cost–benefit analysis
CBO Congressional Budget Office
CBP capability based planning
CBS cost breakdown structure
CC conventional campaign
CCDR contractor cost data reports
Abbreviations xxix
CEA cost-effectiveness analysis
CER cost estimating relationships
CES cost element structure
CFDS Canada first defence strategy
CFOI census of fatal occupational injuries
COCOMO constructive cost model
COEA cost and operational effectiveness analysis
CORA Centre for Operations Research and Analysis
CORE cost-oriented resource estimating
CPS current population survey
CRAF civil reserve air fleet
CSIS Center for Strategic and International Studies
CSS combat service support
C-X cargo-experimental
DAB defense acquisition board
DACIMS defense acquisition automated cost information system
DAES defense acquisition executive summaries
DAMIR defense acquisition management information retrieval
DAS defense acquisition system
DASA Defense Atomic Support Agency
DAU Defense Acquisition University
DCARC Defense Cost and Resource Center
DFARS defense federal acquisition regulation supplement
DMDC Defense Manpower Data Center
DND Department of National Defence
DoD Department of Defense
DOT Department of Transportation
DPS defense policy statement
ECCM electronic counter-countermeasures
ECM electronic counter-measures
EDA European Defence Agency
EEoA economic evaluation of alternatives
EIBA European International Business Academy
EMD engineering and manufacturing development
EPA Environmental Protection Agency
ESBM enhanced scenario-based method
EVMS earned value management system
FAA Federal Aviation Administration
FAR federal acquisition regulation
FDI foreign direct investment
FFRDC federally funded research and development center
FLIR forward looking imaging infra-red
FMTV family of medium tactical vehicles
FO fiber-optic
FPA focal plane array
xxx Abbreviations
FV future value
FY fiscal year
FYDP future year defense plan
GAO Government Accountability Office
GDF guidance for the development of the force
GDP gross domestic product
GMOD German Ministry Of Defence
GSA general services administration
HALE high altitude long endurance
HMMWV high-mobility, multipurpose wheeled vehicle
ICEAA International Cost Estimating and Analysis Association
IDA Institute for Defense Analyses
IEDS improvised explosive devices
IEP Institute for Economics and Peace
IFB invitation for bid
IIE Institute of Industrial Engineers
IOC initial operational capability
IRB industrial regional benefit
IS Islamic state
ISR intelligence, surveillance and reconnaissance
IW irregular warfare
JCIDS joint capabilities integration and development system
JCTD joint capability technology demonstration
JPG joint programming guidance
JROC Joint Requirements Oversight Council
KVA knowledge value added
LBR laser beam-riding
LCAC landing craft
LCC life cycle cost
LCTA lowest cost technically acceptable
LMI logistics management institute
LPD landing platform dock
LRC lesser regional contingency
MADM multi-attribute decision-making
MAIS major acquisition information system
MANET mobile ad hoc network
MASS mobility analysis support system
MAUT multi-attribute utility theory
MCDM multi-criteria decision-making
MCO major combat operation
MDAP major defense acquisition projects
ME military expenditures
MEO most efficient organization
MHE materiel-handling equipment
MM man-months
Abbreviations xxxi
MoD Ministry of Defence
MODM multi-objective decision-making
MoE measure of effectiveness
MOG maximum on ground
MRC major regional contingency
MRS BURU mobility requirements study bottom up review update
MRS mobility requirements study
MRTS marginal rate of technical substitution
MSHA Mine Safety and Health Administration
MTBF mean time between failure
MTM/D million ton-miles per day
MTTR mean time to repair
NAO National Audit Office
NATO North Atlantic Treaty Organization
NDAA non-developmental airlift aircraft
NGO non-government organization
NM nautical mile
NPV net present value
NZDF New Zealand Defense Force
O&M operations and maintenance
O&S operations and support
OECD Organisation for Economic Co-Operation and Development
OLS ordinary least squares
OMB Office of Management and Budget
ONS Office for National Statistics
OSD Office of the Secretary of Defense
OT&E operational test and evaluation
OUSD (A&T) Office of the Under Secretary of Defense (Acquisition and
Technology)
PA&E program analysis and evaluation (now renamed CAPE)
PAA primary authorized aircraft
PAA program alignment architecture
PACF partial autocorrelation function
PALYS protection-adjusted life years
PAX passengers
PBD program budget decision
PCA principal component analysis
PDM programmed depot maintenance
PEM program element monitors
PFC Privates First Class
PGI procedures, guidance, and information
POM program objectives memorandum
PPBE planning, programming, budgeting and execution
PPBES planning, programming, budgeting, and execution system
PPBS planning, programming, and budgeting system
xxxii Abbreviations
PV present value
PVA present value analysis
PWS performance work statement
QALYS quality measures of healthcare based on quality-adjusted
life years
QDR quadrennial defense review
R&D research and development
RAB resource accounting and budgeting
RAID rapid aerostat initial deployment
RDT&E research, development, testing, and evaluation
RFP request for proposal
RFQ request for quotation
RM&A reliability, maintainability, and availability
ROI return on investment
ROK return-on-knowledge
RP risk premium
RTO (NATO) Research Technology Board
SAB Scientific Advisory Board
SAFMA strategic airlift force mix analysis
SAR selected acquisition report
SAS system analysis and studies
SC Schwartz bayesian criterion
SCEA Society of Cost Estimation and Analysis
SDC (US Army Materiel Systems Analysis Activity’s) sample
data collection
SDR strategic defense review
SecDef Secretary of Defense
SIPRI Stockholm International Peace Research Institute
SLEP service life extension program
SMOD Sweden Ministry of Defence
SoS system-of-systems
SOW statement of work
SSC Strategic Systems Committee
SSEB Source Selection Evaluation Board
SSSP steady-state security posture
SUR structural unexpended rate
TAI total aircraft inventory
TCE transaction cost economics
TOW tube-launched, optically tracked, wire-guided
TPFDD time-phased force deployment data
TQM total quality management
TRL technology readiness levels
TWV tactical wheeled vehicles
UAV unmanned aerial vehicle
USAF United States Air Force
Abbreviations xxxiii
USDOD Under Secretary of Defense
USN United States Navy
USTRANSCOM US Transportation Command
VAMOSC visibility and management of operating and
support costs
VSI value of a statistical injury
VSL value of statistical life
WBS work breakdown structure
WEKA the WEKA project (data mining in Java)
WMD weapons of mass destruction
WSARA Weapon Systems Acquisition Reform Act
Acknowledgments

The editors are grateful to the series editors for valuable suggestions, and for
reviewers’ comments provided by Taylor & Francis that offered constructive guid-
ance and support for this project. We would also like to thank the publisher for
excellent editorial direction. Dr Melese is grateful to his wife, Heather, for her
patient role as a sounding board, and especially to RADM James Greene, US Navy
(retired) and his staff at the Acquisition Research Program at the Naval Postgrad-
uate School for their early support of this book. Our recently departed defense
economist colleague, Michael Intriligator, was an inspiration to us all. Finally,
we are indebted to the authors for their outstanding contributions, as well as to
current and past faculty and participants at the Defense Resources Management
Institute (DRMI) and the broader defense economics community from whom we
have learned so much over the years. The views expressed are those of the edi-
tors and authors and do not necessarily represent those of the US Department of
Defense or the Canadian Department of National Defense.
Part I

Introduction and problem


formulation
This page intentionally left blank
1 Introduction
Military cost–benefit
analysis: theory and practice
Francois Melese, Anke Richter, and
Binyam Solomon

1.1 Background
Military cost–benefit analysis (CBA) offers a vital tool to help guide governments
through both stable and turbulent times. As countries struggle with the dual chal-
lenges of an uncertain defense environment and cloudy fiscal prospects, CBA
offers a unique opportunity to transform defense forces into more efficient and
effective twenty-first-century organizations.
Defense reforms typically involve politically charged debates over invest-
ments (in projects, programs, or policies) as well as contentious divestment
decisions—from base realignment and closure (BRAC) to outsourcing and asset
sales. A powerful contribution of CBA is to inform such complex and con-
tentious decisions—carefully structuring the problem and capturing relevant costs
and benefits of alternative courses of action. Lifting the veil on military CBA,
this edited volume reveals several systematic quantitative approaches to assess
defense investments (or divestments), combined with a selection of real-world
applications.
The frameworks and methods discussed in the following chapters should
appeal to anyone interested or actively involved in understanding and applying
CBA to improve national security. These valuable approaches also have broader
government-wide applications, especially in cases where it is difficult to monetize
the benefits of a public project, program, or policy.
Unprecedented government spending to counter the global financial crisis has
placed enormous pressure on public budgets. Combined with alarming demo-
graphics, many countries struggle to fulfill past promises to underwrite health care
expenditures, social security payments, government pensions, and unemployment
programs. As debt burdens grow to finance current operations, the risk of esca-
lating interest payments threatens to crowd out vital future public spending. As
the single largest discretionary item in many national budgets,1 military expen-
ditures make a tempting target. Especially vulnerable are military and civilian
compensation (pay and benefits) and the purchase and operation of equipment,
facilities, services, and supplies.
4 F. Melese, A. Richter, B. Solomon
Anticipating future spending cuts, this book explores both conventional and
unconventional approaches to contemporary defense decisions—from critical
investments in facilities, equipment, and materiel to careful vendor selection to
build, operate, and maintain those investments. Recognizing the value of sys-
tematic quantitative analysis, senior US Army leadership has “directed that any
decisions involving Army resources be supported by a CBA.”2
Faced with severe budget cuts and an uncertain threat environment, defense
officials around the world confront urgent decisions on whether to approve
specific projects (e.g. infrastructure—military housing; training, and mainte-
nance facilities) or programs (e.g. weapon systems—unmanned aerial vehicles
(UAVs), armored personnel carriers (APCs), cyber defense). Military CBA offers
a valuable set of analytical tools to increase the transparency, efficiency, and
effectiveness of critical defense decisions.
A synthesis of economics, management science, statistics, and decision theory,
military CBA is currently used in a wide range of defense applications in coun-
tries around the world: i) to shape national security strategy, ii) to set acquisition
policy, and iii) to inform critical investments in people, equipment, infrastructure,
services, and supplies. This edited volume offers a selection of carefully designed
CBA approaches, and real-world applications, intended to help public officials
identify affordable defense capabilities that effectively counter security risks in
fiscally constrained environments.

1.2 A brief history of cost–benefit analysis


The French engineer Jules Dupuit (Dupuit 1844) is widely credited with an early
concept of CBA called “economic accounting.” The British economist Alfred
Marshall (Marshall 1920) later developed formal concepts that contributed to the
analytical foundations of CBA.3 In a pioneering survey, Prest and Turvey indicate
that as early as 1902 the US River and Harbor Act required the Army Corps of
Engineers to report on the desirability of any project, taking into account both the
cost and the amount of “commerce benefited” (Prest and Turvey 1965). Widespread
application of CBA in the United States is generally attributed to the 1936 Fed-
eral Navigation (Flood Control) Act. This required the Army Corps of Engineers
to carry out projects to improve waterways when “the benefits to whomsoever they
may accrue are in excess of the estimated costs” (Prest and Turvey 1965).
At the heart of CBA is the economists’ concept of “allocative efficiency,” in
which resources are deployed to their highest valued use to maximize social welfare.
A related and intuitively appealing definition called “Pareto efficiency” underpins
CBA. An allocation is Pareto-efficient if no alternative allocation can make at least
one person better off without making someone else worse off (Pareto 1909).
The link between allocations that yield maximum net benefits in CBA and
Pareto efficiency is straightforward: If a public policy, program, or project has
positive net benefits, then it is possible to find a set of transfers (side payments)
that make at least one person better off without making anyone else worse off.
Unfortunately, transfers necessary to achieve Pareto efficiency are difficult to
Introduction: theory and practice 5
implement in practice. Therefore, out of practical necessity, CBA relies on a
related decision rule called the Kaldor–Hicks criterion (Kaldor 1939; Hicks 1940).
This decision rule states that a public policy, program, or project should be
adopted, if and only if gainers could potentially fully compensate losers, and still
be better off.4
Application of this decision rule is relatively straightforward:5 Adopt all
projects that have positive net benefits.6 An important caveat is that the Kaldor–
Hicks criterion only applies when costs and benefits can be monetized. Given
the prevalence of non-monetary benefits in national defense, this poses a serious
challenge for the security sector.
The growing interest in CBA after WWII is often attributed to rapid develop-
ments in operations research and systems analysis—techniques that helped win
the war by combining economics, statistics, and decision theory. Following the
allied victory, Project RAND (launched in 1946 by the Army Air Corps) received
government funding to maintain scientific expertise developed in WWII and to
conduct independent and objective research in national security.
A key contribution of RAND’s research was “systems analysis” pioneered by
Ed Paxson and advanced by Charles Hitch who in 1948 founded RAND’s Eco-
nomics Division. Whereas operations research had a more immediate, wartime
focus (e.g. finding the best short-run solution to a military mission, given a
restricted set of equipment, etc.), systems analysis was more future-oriented,
focused on finding the optimal mix of doctrine, forces and equipment necessary to
accomplish a military goal at the lowest possible cost (or alternatively, for a given
budget, to find the optimal mix that maximizes defense capabilities).
Working at RAND in the immediate post-war era, Hitch teamed up with another
economist, Roland McKean, to publish a pioneering text entitled The Economics
of Defense in the Nuclear Age (Hitch and McKean 1960). The authors emphasized
two main ways in which military CBA can be applied: i) to guide defense policy
(i.e. the allocation of resources between major missions or military goals) and ii) to
guide defense investments (i.e. choices between alternative projects or programs to
achieve a given mission or goal). A significant challenge in applying CBA to defense
decisions is the complex and often controversial task of measuring “benefits.”
At the highest national strategic level, “benefits” of a specific defense policy7
might be measured in terms of its impact on long-term economic growth, peace,
and prosperity—all key contributors to social welfare. For example, suppose
resource costs to achieve specific military goals are viewed as insurance payments
against hazardous states of the world. Suppose further that defense policy deci-
sions that achieve specific military goals reduce risk premiums associated with
domestic and foreign direct investment (FDI). Empirical evidence suggests that
FDI boosts economic growth and in turn contributes to peace and prosperity.8 In
this example, high-level defense decisions could ideally be made with the aim of
increasing social welfare by encouraging investment, boosting GDP, and thereby
generating a virtuous cycle of peace and prosperity.
In reality, this high-level effort to capture monetary benefits of defense policy
as growth in GDP is rarely explored.9 Instead, it typically gives way to a more
6 F. Melese, A. Richter, B. Solomon
familiar perspective that makes up the bulk of chapters in this edited volume—
where non-monetary “measures of effectiveness” (MoEs) of a policy, project or
program substitute for monetary benefits.
Denied the opportunity to conduct controlled experiments or full-scale inde-
pendent field tests to evaluate alternative policies, projects or programs, military
officials and analysts are forced to resort to “proxy” variables. These include cri-
teria and characteristics that reflect multiple objectives and that describe essential
features of the alternatives being analyzed.10 When benefits cannot be monetized,
the terms “systems analysis” or “cost-effectiveness analysis” (CEA) are often used
to describe military CBA.11
A related literature, alternately called multi-criteria decision-making (MCDM)
or multi-objective decision-making (MODM), rapidly evolved after WWII to
address the challenge of measuring non-monetary benefits of defense investments.
The reader is encouraged to explore this literature for details on competing benefit
measurement strategies, some of which are discussed in this volume. These mea-
sures have been in continuous development since the adoption of systems analysis
by the US Department of Defense in the early 1960s.12
Following his election as President in 1960, John F. Kennedy appointed Robert
McNamara Secretary of Defense. McNamara subsequently hired Charles Hitch
as Comptroller to implement the Planning, Programming, and Budgeting System
(PPBS) that Hitch had earlier helped develop at RAND. An output-oriented bud-
geting framework, PPBS relies heavily on systems analysis, or military CBA, to
build a defense budget.
Prior to Hitch’s tenure in the Office of Secretary of Defense, US defense budgets
were largely based on the services’ (Army, Navy, Air Force) proposals for annual
incremental increases in inputs or “appropriation” categories (military personnel,
procurement, operations and maintenance, military construction, etc.), often with
little or no clear connection to defense outputs, joint missions, or national security
goals. Having successfully employed a variant of PPBS called “program budget-
ing” as CEO of Ford Motor Company, McNamara recognized the value of building
a defense budget that focuses on outputs (benefits) as well as costs.
The major innovation of PPBS is “programming,” which bridges the gap
between long-term military planning goals and short-term civilian budget reali-
ties. Designed as a constrained optimization underpinned by systems analysis, the
“programming” phase was intended to produce a cost-effective mix of forces to
maximize national security subject to funding constraints.13
Under certain conditions, the optimal allocation of a budget across various
inputs (e.g. defense resources) that contribute to a common goal (i.e. increasing
national security) requires the marginal contribution of each input towards that
goal, for a given incremental cost, to be the same for any input. Since this deci-
sion rule is independent of the units in which the goal is measured, in principle
it provides a valid test for allocative efficiency, satisfies the condition for Pareto
optimality, and guarantees the most effective use of a defense budget.14
To implement PPBS, Hitch hired a RAND colleague, Dr Alain Enthoven, as
Deputy Comptroller for Systems Analysis. In 1965, the impact of military CBA
Introduction: theory and practice 7
was reinforced when Dr Robert Anthony of the Harvard Business School replaced
Hitch as Comptroller and elevated Enthoven’s position to Assistant Secretary of
Defense for Systems Analysis.15 Throughout his tenure, Secretary McNamara
consistently applied systems analysis to evaluate policy, project and program
proposals from the military services and to build defense budgets submitted to
the Congress.16 Military CBA continues to provide an analytical foundation that
guides PPBS decisions in the United States and in countries around the world.
It is clear that politics influences defense decisions. It is also true that public
officials can manipulate CBA for their own personal strategic interests. Politicians
likely win more votes highlighting a program’s benefits and downplaying its costs,
and public administrators may be similarly rewarded. While it is clear pork-barrel
politics often plays an important role in defense decisions, this book attempts to
take the high road. It encourages the application of military CBA with a strict
focus on national security interests.17
While employment, income distribution, and regional impacts of defense invest-
ment decisions often play a role in political decisions, a clean CBA can inform the
process by revealing the true (opportunity) cost of decisions that drift too far from
the goal of making the best use of scarce resources for the security of the country.
Ideally, a carefully constructed military CBA focused strictly on national secu-
rity concerns could be used to inform voters and counter special interest lobbying
and rent-seeking that often leads defense firms to inefficiently spread production
across key voting districts to promote their programs.18
A risk for any military CBA is that benefit and cost estimates might be strate-
gically manipulated by self-interested agencies or individual decision-makers.19
As Robert Haveman and others have pointed out, politicians facing difficult
re-election tend to prefer projects that concentrate benefits on particular inter-
est groups that offer them support, and to camouflage or defer costs, or to
spread them widely across the population (Haveman 1976). Fortunately, as nations
around the world embrace civilian control of the military, and citizens insist on
greater accountability (including tighter linkages between budgets and security),
an increased premium is placed on transparency in defense decisions.
While politics still dominates major defense decisions, the importance of
military CBA rises alongside growing demands for transparency and account-
ability.20 Costly defense procurement scandals reinforce the need for objective
CBA approaches to improve transparency in vendor selection decisions.21 Mean-
while, painful recovery from the global financial crisis,22 combined with emergent
threats, fuel public demand to carefully apply tools such as military CBA to build
efficient, effective, and accountable security forces.

1.3 Outline
This edited volume reveals how military CBA can reduce budget pressures and
improve defense decisions that contribute to national security. The dual purpose
of CBA is to encourage more efficient and effective allocation of society’s scarce
resources to increase social welfare.23 Governments often employ CBA to rank
8 F. Melese, A. Richter, B. Solomon
(mutually exclusive) portfolios of projects or programs. The typical CBA involves
at least eight steps:

1 Identify key decision-makers (and other stakeholders) to clarify goals, objec-


tives, preferences, and constraints (including realistic funding projections).
2 Carefully structure the problem and identify feasible alternatives that con-
tribute to those goals/objectives and that satisfy the constraints.
3 Determine the relevant time horizon over which the CBA will be conducted
and select an appropriate discount rate.
4 Estimate relevant time-phased costs of each alternative over the relevant
period.
5 Forecast time-phased benefits that will accrue over the relevant period.24 This
edited volume offers alternative approaches to structure a military CBA when
benefits cannot be monetized. If benefits can be monetized, then the project
or program with the highest net present value (NPV) can be recommended.25
6 The sixth step is to recognize uncertainty and conduct sensitivity analyses
to determine whether results change with changes in key parameters (costs,
benefits, budgets, discount rates, etc.).26
7 The seventh step is to report the results of the analysis (rankings of projects,
programs, etc., along with key assumptions).
8 The final step is to make well-informed recommendations.

These eight basic steps of a CBA are explored throughout the chapters of this
edited volume. The book consists of seventeen chapters divided into five parts.

1.3.1 Part I: Introduction and problem formulation


This part includes the first four chapters. Chapter 1 which you are reading offers a
broad overview and outline of the book. Chapter 2 entitled “Allocating national
security resources” sets the strategic tone of the book through the lens of US
global security concerns. The Honorable J. Gansler (former US Under Secre-
tary of Defense for Acquisition, Technology and Logistics) and his co-author W.
Lucyshyn discuss challenges of wide-ranging international threats, domestic bud-
getary restrictions, and ongoing acquisition problems—including questions about
future capacity to support current acquisitions. Revealing a possible mismatch
between the National Security Strategy and the PPBS process, the authors high-
light the need for military CBA at national, departmental, and program levels to
make sound resource allocation decisions.27 They also stress the vital role played
by CBA in the continual process of reassessment and innovation necessary to
maintain critical linkages between resources and requirements and to guarantee
effective forces in a dynamic security environment.
In Chapter 3, a prominent UK pioneer in defense economics, K. Hartley,
and his Canadian senior defense scientist co-author B. Solomon (co-editor of
this volume), confront the challenge of measuring defense outputs. While the
economics approach discussed in “Measuring defense output: an economics
Introduction: theory and practice 9
perspective” is difficult to operationalize into a set of clear and unambiguous
policy precepts, it does provide an important framework to help evaluate the ben-
efits of defense outputs and activities. Combining theory and practice, the chapter
describes attempts to measure defense outputs in the United States, Australia, New
Zealand, and the United Kingdom and other European nations. Later chapters in
this book provide several practical methods to help address challenges posed by
the authors.
While maintaining the strategic themes of Chapter 2 and recognizing mea-
surement challenges discussed in Chapter 3, Chapter 4 by F. Melese offers a
comprehensive set of military CBA approaches to structure public investment
decisions. Entitled “The economic evaluation of alternatives,” six approaches are
introduced that address a significant weakness in many conventional military
“analyses of alternatives” (AoAs).28 Historically, while AoAs correctly focused
on lifecycle costs and operational effectiveness to evaluate alternatives, “afford-
ability” was an afterthought, at best only implicitly addressed in final stages of
the analysis.29 In sharp contrast, the economic evaluation of alternatives (EEoA)
encourages analysts and decision-makers to include affordability explicitly and up
front in structuring a military CBA. EEoA places taxpayers alongside warfighters
in the defense decision-making process. This requires working with vendors to
build proposals based on different funding (budget/affordability) scenarios.30 The
decision map in the concluding section of Chapter 4 offers a comprehensive guide
for practitioners to help structure an EEoA.31

1.3.2 Part II: Measuring costs and future funding


This part consists of three chapters. Chapter 5 entitled “Cost analysis” focuses on
the first of the three main components of an EEoA—costs, budgets, and benefits.
D. Nussbaum, who served as the US Navy’s chief cost analyst, and his co-author,
Professor D. Angelis, discuss approaches to collect, analyze, and estimate costs
of proposed projects, programs, or activities. A unique contribution of this chap-
ter is the explicit recognition of “transaction costs.”32 These include measurement,
monitoring, management, contracting, negotiation, and other costs associated with
government procurement. Depending on the nature of the transaction, it is con-
ceivable that transaction costs could overwhelm the production costs of the desired
product or service. Ignoring transaction costs creates a serious risk of underes-
timating the total costs of a project, program, or activity. In fact, the absence
of transaction cost considerations in military CBAs may help explain the preva-
lence of cost overruns that often negatively impact expected returns on defense
investments. To help address current biases and improve cost estimates in military
CBAs, the authors recommend incorporating transaction cost considerations into
traditional production cost calculations.
Chapter 6 entitled “Advances in cost estimating: a demonstration of advanced
machine leaning techniques for cost estimation” presents recent technical
advancements in cost estimation. The standard methods to estimate costs of
defense systems in early design phases discussed in Chapter 5 include costing
10 F. Melese, A. Richter, B. Solomon
by analogy and parametric approaches. Analogy methods base the costs of new
systems on historical costs of similar or “analogous” systems. The traditional
approach is to ask subject matter experts to make subjective evaluations of dif-
ferences between the new system and the old. This leads to the application of
complexity factors to adjust the analogous (old) system’s cost to produce an esti-
mate for the new system. Rather than apply subjectively obtained complexity
factors, an innovative proposal by Defence Research and Development Canada
scientist B. Kaluzny explores the use of machine learning algorithms to estimate
the costs of systems in early design phases. The author proposes a cost estimation
by analogy approach that involves an agglomerative hierarchical cluster analy-
sis and nonlinear optimization that requires limited subjective input. With limited
information, traditional parametric approaches to cost estimation rely on basic
statistical models to develop cost estimating relationships (CERs) to help identify
major cost drivers. CERs can be as simple as a ratio or involve linear regression
analysis of historical systems or subsystems. The author proposes a new paramet-
ric technique, the M-system of Quinlan (a combination of decision trees and linear
regression models), for learning models that predict numeric values.
Having established the importance of treating affordability (or future funding
constraints) up front in an EEoA, the challenge of forecasting long-term defense
budgets is explored in Chapter 7. Colonel R. Fetterly and B. Solomon begin their
chapter “Facing future funding realities: forecasting budgets beyond the future year
defense plan” by highlighting the importance of strategic management methods,
such as capabilities-based planning, to link existing military capabilities and force
development goals to the future security environment. These strategic management
approaches are coupled with a variety of forecasting models that take into account
a nation’s security threats, income, spillover effects of allies’ defense posture, and
competing demands for a limited public purse. The authors draw on data from a
selection of NATO countries to develop several valuable budget forecasting models.

1.3.3 Part III: Measuring effectiveness


Chapters 8 and 9 offer a standard and novel approach, respectively, to develop mil-
itary MoEs. Chapter 8, entitled “Multiple-objective decision-making,” focuses on
practical, conventional methods used to structure a military CBA when faced with
the challenge of quantifying non-monetary benefits of defense projects, programs,
or policies. Professors K. Wall and C. MacKenzie confront the challenge of
non-monetary benefits leveraging the literature on multiple-objective (and multi-
criteria) decision-making. The authors present a standard approach to help solve
multiple-objective decision problems. Many contemporary decision problems in
defense management and government resource allocation produce multiple, com-
peting benefits. This chapter offers the widely employed analysis of alternatives
(AoA) approach.
Chapter 9 offers a new, cutting-edge approach to conduct a military CBA
focused on force protection investments. If the goal is to evaluate investments
to protect soldiers, then monetizing the benefits of lives saved can help save the
Introduction: theory and practice 11
greatest number of lives. In this chapter, entitled “A new approach to evaluate
safety and force protection investments: the value of a statistical life,” Professors
T. Kniesner, J. Leeth, and R. Sullivan cogently discuss how economists evaluate
the benefits of safety investments by observing tradeoffs people actually make
between safety and other job or product characteristics. The authors present a
widely relevant application of their technique to evaluate the cost-effectiveness of
adding armor protection to tactical wheeled vehicles. The value-of-statistical-life
(VSL) approach presented in this chapter is an innovative military CBA technique
highly recommended for future safety and force protection investments.

1.3.4 Part IV: New approaches to military cost–benefit analysis


In Chapter 10, entitled “The role of cost-effectiveness analysis in allocating
defense resources,” Professor K. Wall joins forces with C.J. LaCivita and Pro-
fessor A. Richter (a co-editor) to present a new CBA approach to solve multi-level
(multi-tiered) resource allocation problems. Their solution method re-interprets
the conventional AoA model with a twist. Applying standard operations research
techniques, they incorporate bounded rationality to realistically portray how
decision-makers can and do cope with the complexities of multi-level con-
strained optimization.33 The bounded rationality formulation employs subjectively
assessed weights derived from the judgment and expertise of a central allocator
(e.g. the Minister of Defense), that offer guidance to lower-level decision-makers
(e.g. the Services: Army, Navy, Air Force) to balance costs and MoEs in building
defense proposals.
Another new, groundbreaking military CBA approach is introduced in
Chapters 11 and 14. Chapter 11 is entitled “A risk-based approach to cost–benefit
analysis: strategic real options, Monte Carlo simulation, knowledge value added,
and portfolio optimization.” In that chapter renowned expert J. Mun and Pro-
fessor T. Housel present their pioneering “real options” approach that estimates
military returns on investment (ROI), combining risk analysis concepts and port-
folio optimization techniques. Two dramatic events unfolded in recent history that
fundamentally transformed the contemporary security landscape—the collapse of
the Soviet Union and the tragedy of 9/11. From a single well-defined “cold war”
nuclear threat, countries now face a wide range of diffuse risks: anything from
failed states, terrorism, and arms proliferation to human trafficking, piracy, and
cyber-attacks. This historic shift in the national security environment prompted
many countries to switch from “threat-based” planning to “capabilities-based”
planning (see Fitzsimmons (2007)). With emerging threats harder to predict,
strategic planners recommend diversification—building broad portfolios of flex-
ible defense capabilities to counter a wide range of possible security concerns.
Chapter 11 offers a new, unconventional approach to military CBA designed to
help build “capability portfolios.” The strategic intent of the United States and
other militaries is to maintain a military edge over rivals. Bureaucratic inertia
and political lobbying by established defense firms, however, often result in too
heavy a focus on prior conflicts. Research and development (R&D) expenditures
12 F. Melese, A. Richter, B. Solomon
represent a real options approach to future contingencies where some, but not
all, research is expected to lead to the development of new systems. R&D pay-
ments are similar to premiums paid for financial options in that they grant the
government the right—but not the obligation—to exploit, defer or abandon R&D
investments. Periodic adjustments are made based on research results, new budget
realities, and the evolving defense environment.34 This innovative chapter intro-
duces hands-on applications of Monte Carlo simulation, real options analysis,
stochastic forecasting, portfolio optimization, and knowledge value added.
The real options approach attempts to make the best possible decisions under
uncertainty and to identify, analyze, quantify, mitigate, and manage risks for
military options. In Chapter 12, entitled “Extensions of the Greenfield-Persselin
optimal fleet replacement model: application to the Canadian Forces CP-140A
Arcturus Fleet,” D. Maybury adapts other recent developments in financial mod-
eling to construct a stochastic fleet replacement/overhaul model to predict the
optimal timing of replacement. The chapter provides an interesting military
application that features a popular maritime surveillance aircraft (the CP-140A
Arcturus, a Canadianized version of the Lockheed P-3 Orion).

1.3.5 Part V: Selected applications


The last five chapters provide a selection of valuable applications and lessons
learned that correspond to the methods and concepts discussed in the preceding
chapters. Chapter 13, entitled “Embedding affordability assessments in military
cost–benefit analysis analysis: defense modernization in Bulgaria,” by V. Georgiev
presents an application of the Economic Evaluation of Alternatives (EEoA) in
Bulgaria’s defense organization. The next two chapters each present real-world
applications of military CBA, and are authored by subject matter experts with
direct experience in high profile defense programs. Former program manager J.
Dillard joins forces with Professors D. Angelis and D. Ford to review develop-
ment of the Javelin anti-tank weapon system in Chapter 14 entitled “Real options
in military acquisition: a retrospective case study of the Javelin anti-tank missile
system.” Study director W. Greer reviews the C-17 strategic airlift program in
Chapter 15 entitled “An application of military cost–benefit analysis in a major
defense acquisition: the C-17 Transport Aircraft.” Whereas the former study pro-
vides a retrospective application of the “real options” approach, the latter offers a
valuable historical perspective of traditional military CBA.
In Chapter 16, entitled “Cost-effectiveness analysis of autonomous aerial plat-
forms and communications payloads,” Commander (USN) R. Everly, Lieutenant
(USN) D. Limmer and Professor C. MacKenzie build a traditional military CBA
to evaluate investments in UAVs. The final chapter, by economists J. Hanson and
J. Lipow, tackles a thorny issue: the so-called “social rate of discount.” The debate
among economists on whether, and by how much, to discount future costs in
public procurement remains unresolved. Chapter 17, entitled “Time-discounting
in military cost–benefit analysis” cogently summarizes the literature and contrasts
it with current guidelines published by the US Office of Management and Budget
Introduction: theory and practice 13
US OMB (1992). This final contribution offers valuable insights and a practical
way forward that could help integrate the existing literature with government
guidelines to improve the quality of military CBAs.

1.4 Conclusion
Tight budgets make for hard choices. The greater the pressure on public budgets
the greater the opportunity to apply military CBA. Today, the impact of govern-
ment deficits and debt on military spending is inescapable.35 As one of the largest
discretionary items in government budgets, military spending is an obvious tar-
get for cuts. While wise use of military power can underpin economic growth,
it is equally clear that economic strength underpins military power. Shrinking
budgets place a renewed premium on affordability. As sovereign debt challenges
squeeze national budgets, and emerging threats challenge existing security forces,
this edited volume offers a valuable set of tools and techniques to help navigate
the political landscape and meet calls to increase the transparency, efficiency, and
effectiveness of the defense sector.

Notes
1 According to the World Bank average, military expenditures globally account for 9.2
percent of government spending. http://data.worldbank.org/indicator/MS.MIL.XPND.
ZS/countries/AU-C5-C7?display=graph (accessed February 7, 2014).
2 Office of the Deputy Assistant Secretary of the Army, U.S. Army Cost–Benefit Anal-
ysis Guide, January 12, 2010, p.6. In general, the US Federal Government’s Office
of Management and Budget (OMB) Circular A-94 provides guidance for the applica-
tion of CBA across the entire Executive Branch. DoD Instruction 7041.3 “Economic
Analysis for Decision Making” provides explicit guidance for the Department of
Defense.
3 Notable among these is the concept of “consumer surplus.”
4 Note that this criterion does not require transfers to actually occur and is occasionally
debated on equity grounds (see Footnote 10).
5 Assuming policies, projects or programs are independent and there are no binding
constraints on inputs.
6 In theory, selecting projects with positive net benefits maximizes aggregate wealth
(e.g. GDP growth) which indirectly helps those that might be made worse off. More-
over, an implicit assumption is that costs imposed on some and benefits accrued to
others will tend to average out across individuals. Where interactions occur among
projects, the general rule is to choose the combination of projects that maximizes net
benefits.
7 For example, benefits of a decision to allocate scarce financial resources among major
military missions.
8 For example, see Brooks (2005) or Gartzke (2007).
9 Chapter 3 in this edited volume offers a notable exception.
10 Examples of proximate criteria or partial measures of effectiveness include speed,
operating range, weapons accuracy, and armor protection.
11 For example, see OMB Circular A-94 Guidelines and Discount Rates for Benefit-Cost
Analysis of Federal Programs published by the US Office of Management and Budget.
Note that this edited volume will continue to use the generic term “military CBA” to
refer to cases where benefits cannot be monetized. Although CEA also produces a rank-
ing, there is no explicit information about whether the highest ranked alternative would
14 F. Melese, A. Richter, B. Solomon
provide positive net social benefits. If all alternatives are mutually exclusive, and the
status quo is among the alternatives, sharing similar scale and phasing of costs and
benefits, then it is possible to apply CEA to select the most efficient policy.
12 Given the vast existing literature on building MOEs, this book instead focuses on the
careful construction of military CBAs. Although we occasionally explore the ques-
tion of developing non-monetary benefit measures, we encourage the reader to review
the extensive literature on multi-criteria decision-making for alternative approaches to
deriving such measures of effectiveness. (For example, see Keeney and Raiffa 1976;
Buede 1986; or Kirkwood 1997.)
13 “The ultimate objective of PPBS shall be to provide operational commanders-in-chief
the best mix of forces, equipment, and support attainable within fiscal constraints”
(DoD Directive 7045, 14 May 22, 1984). The basic questions of systems analysis
are twofold: i) given a fixed budget, which weapon systems are most cost-effective
and, conversely, ii) given a fixed military mission, which system(s) could generate
the desired level of effectiveness at the lowest cost? The basic ideas of PPBS were:
“the attempt to put defense program issues into a broader context and to search for
explicit measures of national need and adequacy;” “consideration of military needs
and costs together;” “explicit consideration of alternatives at the top decision level;”
“the active use of an analytical staff at the top policymaking levels;” “a plan combin-
ing both forces and costs which projected into the future the foreseeable implications
of current decisions;” and “open and explicit analysis . . . made available to all inter-
ested parties, so that they can examine the calculations, data, and assumptions and
retrace the steps leading to the conclusions” (Enthoven and Smith 2005). http://www.
rand.org/content/dam/rand/pubs/commercial_books/2010/RAND_CB403.pdf.
14 In practice, measuring contributions of various inputs towards a defense goal can be
difficult and contentious, and these desirable results only hold under the assumption
(“homotheticity”) that optimal input ratios are independent of the budget and depend
only on relative costs of each input. It is also important to recognize transaction costs
associated with the application of military CBA (or systems analysis). For example, cen-
tralization of decision-making authority under Secretary of Defense Robert McNamara
resulted in a proliferation of management systems to collect data required to evaluate the
costs and benefits of alternative projects and programs (weapon systems). Increasingly
buried in paperwork, the term “paralysis of analysis” was coined by some members of
the defense establishment (personal conversation with A. Enthoven).
15 In 1972, the Systems Analysis division evolved into the Office of Program Analysis and
Evaluation (PA&E) and later, in 2009, into the Office of Cost Assessment and Program
Evaluation (CAPE). Prior to his departure, Dr Hitch launched an OSD-sponsored educa-
tion institution to teach civilian and military managers in DoD (and partners and allies)
basic principles of PPBS and CBA. Today it is known as the Defense Resources Man-
agement Institute (DRMI) located at the Naval Postgraduate School in Monterey, Cali-
fornia. Two co-editors of this volume (Dr Melese and Dr Richter) are faculty members at
this institution which celebrated its 50th anniversary in 2015 (http://www.nps.edu/drmi/;
http://www.dtic.mil/whs/directives/corres/pdf/501035p.pdf).
16 McNamara relied heavily on systems analysis to reach several controversial weapon
decisions. He canceled the B-70 bomber, begun during the Eisenhower years as a
replacement for the B-52, stating that it was neither cost-effective nor needed, and later
he vetoed its proposed successor, the RS-70. McNamara expressed publicly his belief
that the manned bomber as a strategic weapon had no long-run future; the interconti-
nental ballistic missile was faster, less vulnerable, and less costly. Similarly, McNamara
terminated the Skybolt project late in 1962. Begun in 1959, Skybolt was conceived as
a ballistic missile with a 1,000-nautical mile range, designed for launching from B-52
bombers as a defense suppression weapon to clear the way for bombers to penetrate to
targets. McNamara decided that Skybolt was too expensive, was not accurate enough,
Introduction: theory and practice 15
and would exceed its planned development time. He claimed other systems, including
the Hound Dog missile, could do the job at less cost.
17 Our view is that a clean military CBA is a valuable starting point for political debate.
Careful analysis can constrain political attempts to turn defense spending into a jobs
program or an opportunity to redistribute income. Since there exist considerably more
efficient and effective ways to promote job growth and income distribution, if these are
the goals, then they should be stated explicitly and explored using a separate CBA. This
may prove a valuable avenue for future research.
18 An example is the case of the F-35 aircraft program. Lockheed-Martin claimed to
have “created 125,000 US-based direct and indirect jobs in 46 states” http://www.
businessweek.com/news/2014-01-22/lockheed-martin-inflating-f-35-job-growth-claims-
nonprofit-says (accessed March 21, 2014).
19 Since costs (e.g. investment expenditures) tend to occur earlier with benefits appearing
later, discount rates can also be strategically selected to make projects appear more or
less attractive. (See Chapter 17.)
20 For example, see NATO’s Building Integrity initiative at www.nato.int/cps/en/natolive/
topics_68368.htm [last accessed June 24, 2014].
21 For example, see Camm and Greenfield (2005).
22 As public officials face growing resistance to tax increases, pressure increases on
governments to work more efficiently and effectively.
23 National defense satisfies two key characteristics of a “public good.” It is: i) non-rival
and ii) non-excludable. In the former case, unlike private goods, if one person in a
geographic area is defended from foreign attack or invasion, his or her consumption is
non-rival in that others in the area can consume the same level of national security for
little or no additional cost. In the latter case, if one person is defended, others in that
same area cannot be excluded from the security benefits. This leads to a classic free-
rider problem, making it difficult to charge people for national defense. The key here
is that whereas it is generally agreed that the provision of national defense is a public
good that must be funded with taxes, the production of national defense depends on
the relative costs (including transaction costs) and benefits of public and private sector
production, which can be evaluated using military CBA.
24 Since benefits of proposed defense investments are often difficult to monetize, various
approaches have been developed to construct MOEs that capture the value or utility of
alternatives. In theory, the benefits of alternative defense investments could be mone-
tized if their contribution to security and stability encourages foreign direct investment
that contributes to economic growth and social welfare. In practice, precise linkages
between defense investments and economic growth are difficult to establish. As a conse-
quence, the benefits of most military investments are not monetized, and instead various
MOEs are constructed to conduct a CEA that is referred to in this volume as a military
CBA (e.g. see OMB Circular A-94).
25 If benefits can be monetized, then calculate the discounted sum of net benefits (benefits
minus costs) from each alternative over the specified time period, i.e. the discounted
NPV. For example, consider two alternative military projects designed to achieve the
goal of reducing the Navy’s fuel budget. Suppose there is a fixed investment budget
available, and the two mutually exclusive alternatives each require the same identically
phased investment. The first proposal is to invest in a program to convert ship propul-
sion from conventional diesel to a less expensive bio-diesel. The second proposal is to
invest in an energy conservation program at Navy installations. Since the two alterna-
tives require the same identically phased investments, the CBA can simply focus on
the stream of benefits (savings from cheaper fuel in the first case and reduction in
fuel demand in the second) that accrue from each project. Assuming a preference for
present vs. future savings, the discounted present value of each stream of savings can be
calculated to determine the winning project. (See Chapter 17.)
16 F. Melese, A. Richter, B. Solomon
26 Alternatively, a “Real Options” approach could be adopted, conducting Monte Carlo
simulations assigning probability distributions to key parameters. (For example, See
Chapters 11 and 14.)
27 The authors also warn that until new mechanisms such as CBA are adopted to address
the continual failure to fuse requirements with necessary resources, DoD will essentially
continue to create a disjointed, ineffective framework for addressing national security
concerns rather than the vital cohesive plan required to confront the changing dynamics
of modern, global warfare.
28 See Ullman and Ast (2011) and OMB Circular A-11 (2008) for discussions of AoAs.
29 In the US, AoAs are conducted in early phases (milestones) of major defense acquisi-
tions. Since they frequently occur in early development before a project is fully funded,
they rarely incorporate future funding forecasts. Instead, the budget estimate for the
program or project is generated as an output of the AoA. As major budget cuts cre-
ate funding challenges for new defense programs, “affordability” in terms of realistic
budget constraints is gaining increasing importance in AoAs.
30 A key difference between traditional AoAs and EEoA is that instead of modeling
competing vendors as points in cost-effectiveness space, EEoA solicits vendor offers
as functions of optimistic, pessimistic, and most likely funding (budget) scenarios. A
formal mathematical model of EEoA can be found in Simon and Melese (2011).
31 Following the recommended EEoA approach also provides a unique opportunity to
achieve a significant defense reform: to coordinate the Requirements Generation Sys-
tem, Defense Acquisition System, and Planning, Programming, and Budgeting System
(PPBS)—to lower costs, and improve performance and schedules.
32 For example, see Williamson and Masten (1999) or Melese et al. (2007).
33 Finding the optimal mix of forces to accomplish a military goal at the lowest possible
cost, or alternatively, for a given budget, to find the optimal mix that maximizes defense
capabilities.
34 The real options approach builds on what was previously referred to as incremental or
“spiral” development of military programs and projects.
35 The response of many governments to the global recession was to bail out banks and
businesses and to stimulate their economies. Combined with falling tax receipts, this
led to unprecedented increases in government spending. The result in many countries
transformed the financial crisis into a chronic sovereign debt crisis. Annual deficits
soared and cumulative debt loads reached unsustainable levels. Aging demographics in
some countries compounded the problem, placing impossible demands on social welfare
programs and introduced further pressure on government budgets. Combined with an
uncertain threat environment, the fiscal crisis makes a compelling case for widespread
application of military CBAs to ensure future defense decisions to produce efficient,
effective, and accountable security forces.

References
Buede, D. M. 1986. “Structuring Value Attributes.” Interfaces 16(2): 52–62.
Brooks, S. G. 2005. Producing Security: Multinatinal Corporations, Globalization, and the
Changing Calculus of Conflict. Princeton, NJ: Princeton University Press.
Camm, Frank, and Victoria A. Greenfield. 2005. How Should the Army Use Contractors on
the Battlefield: Assessing Comparative Risk in Sourcing Decisions. Santa Monica, CA:
RAND Corporation.
Dupuit, J. 1844. De La mesure de l’utilité des travaux publics. Annales des Ponts et
Chaussées, Memories et Documents 2, 8: 332–375. Translated by R. H. Barback into
English, “On the Measurement of the Utility of Public Works.” 1952. International
Economic Papers 2: 83–110.
Introduction: theory and practice 17
Enthoven, A. C., and K. W. Smith. 2005. How Much is Enough? Shaping the Defense
Program, 1961–1969. Santa Monica, CA: RAND Corporation.
Fitzsimmons, M. 2007. “Whither Capabilities-Based Planning?” Joint Forces Quarterly
44(1): 101–105.
Gartzke, E. 2007. “The Capitalist Peace.” American Journal of Political Science 51(1):
166–191.
Haveman, R. 1976. “Policy Analysis and the Congress: An Economist’s View.” Policy
Analysis 2(2): 235–250.
Hicks, J. R. 1940. “The Valuation of the Social Income.” Economica 7(26): 105–124.
Hitch, C., and R. McKean. 1960. The Economics of Defense in the Nuclear Age. London:
Oxford University Press.
Kaldor, N. 1939. “Welfare Propositions of Economics and Interpersonal Comparisons of
Utility.” Economic Journal 49(195): 549–552.
Keeney, R. L., and H. Raiffa. 1976. Decisions with Multiple Objectives: Preferences and
Value Tradeoffs (Probability and Mathematical Statistics Series). New York: John Wiley
& Sons, Inc.
Kirkwood, C. W. 1997. Strategic Decision Making: Multiobjective Decision Analysis with
Spreadsheets. Belmont, CA: Duxbury Press.
Marshall, A. 1920. Principles of Economics (8th edition). London: MacMillan & Co. Ltd.
Melese, F., R. Franck, D. Angelis, and J. Dillard. 2007. “Applying Insights from Transac-
tion Cost Economics to Improve Cost Estimates for Public Sector Purchases: The Case
of U.S. Military Acquisition.” International Public Management Journal 10(4): 357–85.
Pareto, V. 1909. Manuel d’Economie Politique. Paris: V. Giard et E. Briere. Reprinted,
Paris: M. Girard, 1927; Geneva: Librarie Droz, 1966.
Prest, A. R., and R. Turvey. 1965. “Cost–Benefit Analysis: A Survey.” The Economic
Journal 75(300): 683–735.
Simon, J., and F. Melese. 2011. “A Multiattribute Sealed-Bid Procurement Auction with
Multiple Budgets for Government Vendor Selection.” Decision Analysis 8(3): 170–79.
Ullman, D. G., and R. Ast. September 2011. “Analysis of Alternatives (AoA) Based
Decisions.” Military Operations Research Society, Phalanx 44(3): 24–36.
US Office of Management and Budget (OMB). 1992. Circular No. A-94, Guidelines and
Discount Rates for Benefit-Cost Analysis of Federal Programs. Washington, DC: Execu-
tive Office of the President. Available at www.whitehouse.gov/omb/circulars_a094/ [last
accessed September 18, 2014].
US Office of Management and Budget (OMB). 2008. Circular No. A-11, Preparation,
Submission, and Execution of the Budget. Washington, DC: Executive Office of the Pres-
ident. Available at www.whitehouse.gov/omb/circulars_a11_current_year_a11_toc [last
accessed November 11, 2014].
Williamson, O. E., and S. E. Masten (Eds). 1999. The Economics of Transaction Costs
(Elgar Critical Writings Reader). Northampton, MA: Edward Elgar Publishing.
2 Allocating national security
resources
Jacques S. Gansler and
William Lucyshyn

Most innovations fail. And companies that don’t innovate die . . . In today’s world,
where the only constant is change, the task of managing innovation is vital for
companies of every size in every industry. Innovation is vital to sustain and advance
companies’ current business; it is critical to growing new business. It is also a very
difficult process to manage.
Henry W. Chesbrough 2006, Haas School of Business
Berkeley California

2.1 Introduction
Private sector firms have accepted that they must continuously innovate to stay
competitive in today’s global environment. This innovation is not a one-time
event, but a continuous exercise. Similarly, in a world of emerging and evolv-
ing challenges and threats, ministries of defense must be willing to continuously
innovate. The constancy of change cannot be met with technology, training, or
operational concepts that are outdated and inadequate. Equally important, as pri-
vate sector firms have discovered, are innovations improving business processes.
Within the US Department of Defense (DoD), one business process needs imme-
diate attention—the one used to allocate national security resources. Military
cost–benefit analysis (CBA) has an important role to play. To fully appreciate this
imperative, one must understand and appreciate the environment that DoD finds
itself in today. The willingness of DoD to adopt innovative CBA concepts to over-
haul its resource allocation processes will directly affect the department’s capacity
to meet future threats to US and global security.

2.2 Background
During the cold war, US defense planning was, to a large degree, threat-based.
The monolithic threat posed by the Soviet Union could sensibly, if not always
accurately, be described in terms of troops, fighter aircraft, tanks, and missiles.
The threat was remarkably stable and evolved slowly, simplifying the Pentagon’s
planning challenge—as well as the cost–benefit analysis of investments to cope
with that threat.
Allocating national security resources 19
With the dissolution of the Soviet Union, the relentless influence of the cold
war upon the nation and its national security strategy vanished. Some, believ-
ing that mankind had reached the pinnacle of ideological evolution, wrote of the
end of history with the United States as the world’s lone super-power (Fukuyama
1989).
By the beginning of the 1990s, the United States found itself in the midst of a
very different war—Desert Storm. This war showcased a new manner of warfare,
involving precision strike capabilities used in conjunction with real-time informa-
tion exchange—an early version of net-centric warfare. There was still much to be
learned, however, and many additional advancements to be made. The Gulf War
offered the first glimpse of technological capabilities, and these will continue to
reshape warfare throughout the twenty-first century.
Immediately following Operation Desert Storm, powerful forces began to influ-
ence policy-makers as they anticipated a new era of change. These included the
acceleration of the information technology revolution, globalization, the poten-
tial threat of terrorism, and perhaps most importantly the desire to reap a “peace
dividend.” With no clear rival in sight, the United States began a period of
disinvestment, and it expedited efforts to downsize its military forces (Walker
2014).
Then, on September 11, 2001, the attacks on the World Trade Center and the
Pentagon highlighted a new, global threat from terrorist organizations. The events
of that day created an unprecedented urgency for innovative changes in the US
defense establishment, and they laid the foundation for adjustment to the new
global security environment. As the threats now addressed by the DoD were vastly
different from those it faced in the previous five decades, this adjustment proved
especially vital (Freier 2007).
While numerous earlier reports had highlighted the potential for transformation
of the security environment (Cha 2000), the events of 9/11 provided the catalyst
for reform. The formerly relatively stable Soviet threat was replaced by profoundly
uncertain, fragmented, and complex threats that proved far more difficult to satis-
factorily address. In addition to the threat from global terrorism, there were threats
from failed or failing states that resulted in civil wars, humanitarian catastrophes,
and regional instability. This tremendously unstable international security environ-
ment makes it extremely challenging to predict with any confidence what threats
the United States, its partners, and allies might face in the medium or even the
near-term.

2.3 The new security environment


The new security environment continues to pose unique operational challenges,
making it difficult for a cold war-designed military to respond in a timely and
appropriate manner. Recently, it is hoped that the presence and strength of the
United States’ armed forces can contribute to peace in some of the most important
and dangerous places in the world.1 As the United States’s threats and enemies
continue to adapt and transform, however, so too must its force structure.
20 Hon. J. S. Gansler, W. Lucyshyn
The most recent shifts in the dynamics of warfare and threats to the United
States have not been more evident than in the conflicts in Afghanistan and Iraq.
There, US forces faced a new type of protracted, low-intensity conflict that seems
to have defined a new paradigm for warfare. Counterinsurgency, nation-building,
and protracted conflict—all hallmarks of US efforts in Afghanistan and Iraq—
seem to offer insight into modern conflict and DoD’s future requirements. How,
and how well, DoD will prepare to confront new battlefield trends (including “war
among the people”(Smith 2005)) remains to be seen.
The present security requirements demand that the US military be able to
respond to disruptive, catastrophic, irregular, and traditional threats. These can be
delivered by undetermined weapons and delivery systems to undetermined targets
from often undetermined entities at undetermined times (Perl 2006). The pressure
to counter these potential threats demands the development of a new, holistic view
of security. Consider briefly some of these potential threats:

• Global terrorism. Global terrorism can be initiated at the national level or


come from non-state actors, such as Al Qaeda, Hezbollah, Islamic State (IS),
and other extremists groups. Some of these transnational groups are very
well funded and have access to advanced weapons on the international mar-
ket. Their actions, even on a small scale, can have significant effects. One of
the greatest fears is that they will acquire and use weapons of mass destruc-
tion (biological, chemical, radiological, or nuclear). Their goals can be fairly
narrow, such as the overthrow of legitimate regimes; or very broad, such as
the establishment of a regional “caliphate” (Al-Salhy 2014). Terrorist groups
most often target innocent citizens at home and abroad. While statistics vary
widely in their accounting of terrorist attacks, the general consensus is that
since 9/11 the number of events has rapidly increased each year (The Insti-
tute for Economics and Peace 2012). Most important for this discussion is
to recognize that terrorists cannot be defeated solely by traditional military
means.
• Proliferation of weapons of mass destruction (WMD). The proliferation of
WMD, as well as their delivery systems, increases the possibility of their use,
with the potential for significant casualties. Many countries already possess
WMD, or they have the capacity to produce them. In addition, an increasing
number of countries are in the process of acquiring and developing these capa-
bilities. There is a growing interest among certain terrorist groups in acquiring
such weapons.
• Cyber warfare. Although military systems are increasingly being designed
to be hardened to cyber attacks, most of the nation’s critical infrastructure
is owned and controlled by the private sector, and it continues to be vulnera-
ble. Successful cyber attacks on, for example, central banking systems, power
distribution systems, or hospital computer systems could have devastating
effects. There have already been major cyber attacks against other coun-
tries. Estonia, for example, suffered a major cyber attack in 2007 (believed to
have originated in Russia). Further, prior to Russia’s military incursion into
Allocating national security resources 21
Georgia in 2008, the Georgian government’s communications and informa-
tion systems were disrupted with an extensive cyber attack (Schwirtz 2008).
We can anticipate that future military operations will increasingly utilize
cyber attacks to facilitate and amplify any kinetic conflicts.
• Regional conflicts. There are several recent examples of regional conflicts: the
United States’s “Global War on Terror” that involved large-scale operations
in Iraq and Afghanistan, Israel’s conflicts with Hezbollah in Lebanon in 2006
and with Hamas in 2014, and recent events in Ukraine. These can be fought
against national governments and/or against groups of insurgents. In addition,
there exist several potential regional conflicts, such as between North and
South Korea; between China and Taiwan; between India and Pakistan, some
of which could easily draw in the United States.
• Potential future peer competitors. The most frequently identified potential
future peer competitor is mainland China, but others, such as India or even a
resurgent Russia, cannot be dismissed. It is important to note, however, that
there is no current “peer” to the overwhelming strength of the US military
establishment (Perlo-Freeman 2014). Nonetheless, the United States should
not ignore the potential of a future peer competitor as that would signal weak-
ness in future geopolitical negotiations with any nation with ambitions of
building a strong, regional or global military presence.
• Non-traditional national security challenges. There are several potential non-
traditional security challenges that the United States may face. First, there is
a real possibility of devastating global pandemics (for example, outbreaks of
infectious disease, such as the Avian flu) that can spread quickly and rapidly
overwhelm national healthcare systems. A second non-traditional security
challenge is the increasing global demand for scarce strategic resources such
as lithium, or possibly (with the worst-case global warming scenario), even
drinking water. Third, international criminal activities have taken on a scale
where they generate budgets as large as the GDP of many countries, cor-
rupt the political process, and are often linked with the terrorist community.
Finally, as some nation-states fail to maintain their own internal security,
there is a strong possibility that the United States may be asked to contribute
to stability—for humanitarian reasons—or to assist threatened neighboring
partner or allied countries.

Although military CBA can help DoD as it shifts its portfolio of capabilities to
meet these new challenges, the current force structure is still far from being able
to fully respond to this spectrum of new challenges. As a result, the Department is
faced with difficult investment decisions to bridge this gap, while simultaneously
responding to current operational challenges.

2.3.1 Domestic spending pressures on the defense budget


Similar to many countries, the United States faces several long-term budgetary
challenges to meet the rapidly rising burden of domestic entitlement programs.
22 Hon. J. S. Gansler, W. Lucyshyn
Interest payments on the federal debt, required to meet the nation’s mandatory
spending, will soon directly impact the DoD’s ability to transform its forces.
The United States will suffer an escalating financial burden as baby boomers
age (by 2020, the number of people in US population between the ages of 65
and 84 is expected to rise by nearly 50 percent (US Census Bureau 2004)), and
mandatory spending on programs such as social security and Medicare will nec-
essarily increase.2 Unlike annual appropriations, which specify how much can
be spent on a specific program in that year, the laws that govern these enti-
tlement programs specify formulas to calculate benefits and eligibility criteria
that automatically determine the level of spending, and do not require Congres-
sional action.
Social security remains only the largest of the “must pay” costs expected to sig-
nificantly increase in coming decades. Medicare and Medicaid, as well as interest
on the national debt, will also continue to increase significantly as a percentage
of GDP.3 Based upon US Government Accountability Office (GAO) projections,
93 cents of every federal dollar will be spent in one of these three categories by
2020 (GAO 2010). The Congressional Budget Office (CBO) also projects con-
tinued growth within these three major programs.4 Compounding the problem,
net interest payments on the national debt are projected to grow to 8 percent of
GDP by 2030, placing a substantial additional strain on the federal budget (GAO
2010). While CBO and historical trends suggest further increases in US economic
output and a continually growing GDP, DoD’s budget is projected to continue to
shrink as a percentage of GDP. Since 1971, defense budgets have fluctuated from
7.3 percent of GDP to a historical low of 3.0 percent in 1999, 2000, and 2001.
Although defense spending and budgets have since increased, largely in
response to conflicts in Afghanistan and Iraq, this upward trend is reversing after
the military drawdown. CBO has stated that DoD budgets—given the most likely
drawdown rate—will decrease from its current level of 4.6 percent of GDP to
3.1 percent by 2020 (see Figure 2.1).5
The trend is clear: mandatory domestic spending, specifically social security,
Medicare, and Medicaid, as well as interest on the national debt, will increase
well beyond historical levels, and will inevitably put downward pressure on future
DoD budgets. This will serve to constrain funds available for recapitalization,
modernization, and transformation. Future investments guided by military CBA
will increasingly be constrained by shrinking DoD budgets, requiring hard deci-
sions and a reengineering of current processes to ensure the most efficient use of
scarce resources.6

2.3.2 Internal defense budget challenges


As new military missions and requirements emerge in response to the evolving
security environment, the danger is that funding for equipment, personnel, oper-
ations and maintenance (O&M), and homeland security will increasingly depend
on the crisis of the moment—especially given the new fluid environment which
makes long-term planning difficult. Further, this problem is only exacerbated by
Allocating national security resources 23

14

12

10
Percentage of GDP

0
19

19

19

19

19

19

20

20

20

20

20
70

75

80

85

90

95

00

05

10

15

20
Year

Social Security, Medicare, Medicaid Defense Outlays

Figure 2.1 Defense and selected entitlement spending as a percentage of GDP.


Source: CBO 2010.

the year-by-year Congressional budget process which constrains the ability of


DoD to identify and evaluate long-term funding options.
Further constraining defense budgets is the rising costs of military personnel
compensation, annual healthcare, and facilities programs, which make up a sizable
portion of “defense discretionary” spending already earmarked for future defense
“must pay” requirements. Meanwhile, current operations and support funding
represents nearly two-thirds of the DoD budget. Funds dedicated to future mod-
ernization represent only roughly one-third of the budget. CBO projections reveal
major future increases in spending in areas such as personnel (including health-
care) and O&M, which are projected to rise from US$373 billion in financial year
(FY) 2009 to US$425 billion in FY 2028 (in constant 2010 US dollars).
Even though defense budgets are currently well above the cold war average
(as shown in Figure 2.2), they will likely decline significantly with troop draw-
downs. At the same time, funds invested in research, development, testing, and
evaluation (RDT&E) are projected to decline from US$81 billion in FY 2009 to
US$54 billion in FY 2028 (in constant 2010 US dollars)—a reduction of approxi-
mately 33 percent; which threatens to significantly constrain future investment in
modernization (see Figure 2.3) (CBO 2010).
Although budget projections for defense procurement do not currently indicate
significant cuts, many defense analysts are skeptical. “Generally, the first thing to
go is procurement,” Michael Bayer, Chairman of the Defense Business Board was
24 Hon. J. S. Gansler, W. Lucyshyn

Dramatic Increase
$800,000
with OCO Funding

$700,000
Korean War Vietnam Reagan
War Buildup
$600,000

Pre-9/11
$500,000 Low
CY 2015 $ M

$400,000

$300,000

$200,000

$100,000

$0

20
19
19
19
19
19
20
20
20
20
20
19
19
19
19
19
19
19
19
19
19
19
19
19

17
69
72
75
78
81
84
87
90
93
96
99
02
05
08
11
14
48
51
54
57
60
63
66

Figure 2.2 DoD total budget authority.


Source: National Defense Budget Estimates for FY 2015, Office of the Under Secretary of
Defense (Comptroller), April 2014.
Note: FY 2016–2018 are projected estimates from the President’s FY2015 budget.

quoted as saying. “With the spending declines following the Vietnam War and the
post-Reagan years, the decline in procurement was steeper than the rate of decline
in the overall defense budget” (Boessenkool 2009). In the present environment,
it is likely that increasing O&M costs for legacy systems, along with projected
rising personnel expenses, will be financed at the expense of future R&D and
procurement within DoD.

2.4 Acquisition challenges


In addition to all of these environmental stresses, DoD continues to face several
acquisition challenges. In terms of resource allocation, three of the most critical
are continued program cost and schedule growth, the generation of requirements,
and system-of-systems development.

2.5 Cost and schedule growth


DoD’s major weapon system programs continue to take longer, cost more, and
deliver fewer quantities and capabilities than originally planned.
Michael J. Sullivan, GAO 2008 (Sullivan 2008)
Allocating national security resources 25

80%

70%
70%
64%
66%
Percentage of DoD budget

60%
56%

50%
49%
40%

30%
23% 22%
19%
20% 17% 16%

10% 12% 12% 11% 11% 9%


0%
2009 2010 2013 2020 2028

Procurement R,D,T,&E Operations and Support

Figure 2.3 Resources for Defense in selected years as a percentage of total DoD budget.
Source: CBO 2010.
Note: 2009 and 2010 include supplemental funding of US$74 billion and US$130 billion.

Acquiring major weapon systems efficiently has proved to be difficult. During


2007 for example, programs that comprised DoD’s Major Defense Acquisition
Projects (MDAPs), which make up roughly 80 percent of the DoD’s acquisition
budget in a given year (Younossi et al. 2007), had an average program cost growth
of 26 percent when compared to initial estimates. These increases represented
approximately US$295 billion dollars in additional costs. These programs also
experienced, on average, a 21-month delay in delivering initial capability to the
warfighters (Sullivan 2008).
Unfortunately, these difficulties are not new. DoD has experienced similar
development problems since at least the 1950s (Frank 1997). Despite data limi-
tations, numerous reports issued over the past 50 years have noted high program
cost growth. This growth is defined as the positive difference between actual costs
and initial (budgeted) costs estimated in the original military CBA. Chapter 5 pro-
vides a detailed discussion of cost estimation techniques used in military CBA.
Due to its relative ease of measurement, cost growth provides a simple measure to
help gauge the efficiency of the acquisition process.
DoD has attempted several reforms to control program and unit cost growth, but
most have had limited impact. The V-22 program is a good example. The V-22
Osprey is an innovative, tilt-rotor aircraft developed for Marine Corps, Air Force,
and Navy use. The program’s concept exploration phase began in 1981, and full-
scale development was started in 1986. The V-22 achieved its initial operational
26 Hon. J. S. Gansler, W. Lucyshyn
capability (IOC) in mid-2007. By 2008, total program cost estimates increased from
US$39.1 billion to US$56.1 billion, a 43 percent increase from the 1986 estimate
(with reduced quantities!). This was due in large part to increases in R&D costs,
which grew from an original estimate of US$4.1 billion budgeted to US$12.8 billion
(a 209.6 percent increase). This increase was combined with increases in procure-
ment costs, largely from fixed costs being spread over fewer units as the original
program quantity was cut nearly in half (reduced from 913 to 458). While a 2010
GAO report suggested marked improvement in V-22 performance, pre-operational
performance tests of the V-22 revealed multiple software and practical operational
problems. These problems have continued to drive unit costs up. Unfortunately, this
program is not unique. Other major programs, such as the F-22 and most recently
F-35, have experienced similar problems, suffering from overly optimistic cost and
performance estimates in the original military CBA.
If history is any indication, we can anticipate that many future programs will
result in cost overruns during the R&D phase. This in turn results in increased
total program costs, increased unit costs, stretched-out programs, and ultimately a
reduction in the number of units procured. Unrealistic cost and performance esti-
mates result in flawed military CBAs that require DoD to reassess its programs
and change procurement outcomes. Given the nation’s other pressing financial
obligations, DoD must find ways to improve its military CBA decision-making
(both in the initial selection of programs, as well as in its response to inevitable
change orders) so that required capabilities can be developed and acquired as
economically and efficiently as possible.

2.6 Weapon system requirements


As previously discussed, the current threat environment is pushing the DoD in
several different directions. These include fighting the war on terror, deterring
future conflicts, and investing in force transformation. Consequently, the focus of
twenty-first-century acquisition must lie in several areas, including intelligence,
unmanned systems, professional services, and advanced information systems.
Emphasis on enhancing these “mission-oriented” areas marks a movement away
from “platform-centric” thinking to more “network-centric” thinking in terms
of integrated systems-of-systems (with large numbers of relatively inexpensive,
distributed, sensors and shooters, all interlinked with complex and secure com-
mand, control, and communication systems). This increases the importance of
robust military CBA approaches to help guide the acquisition of highly technical,
game-changing systems. Examples include manned and unmanned (sensor-based)
aircraft systems, net-centric communications systems, and advanced air and
ground weapons systems.
All of these systems rely upon technological superiority to maintain their
dominance, but if they are not constantly updated to remain ahead of potential
threats, they could be rapidly rendered obsolete.
The drive to develop these high-tech systems often requires an extended
period to mature and integrate technology. Yet military operators (users) often
Allocating national security resources 27
believe they must get all of their requirements included up front or risk los-
ing the opportunity to obtain the complete benefits. Chapters 11 and 14 offer
a novel so-called “real options” approach to CBA that attempts to address this
concern.
Insisting on including all requirements up front lends itself to a process of
identifying and developing requirements for these weapon systems that “is overly
cumbersome, but also lacking in the expertise and capacity required to truly
vet joint military requirement” (Reform 2010). Further, once requirements are
approved, there is a great reluctance by the Services to trade them off for cost
reductions or schedule accelerations. This results in programs being stretched out
and running over budget. Compounding the problem is that historically DoD’s
requirements process is not resource-constrained; and, as a result, generates a
demand for more new programs than fiscal resources can support. This concern for
“affordability” is explicitly addressed in the “economic evaluation of alternatives”
presented in Chapter 4.
Major investments in new systems are mostly the responsibility of the military
services who recognize that they compete for funding with the other services. The
result can be unhealthy. To secure funding for their programs, there is an alarming
tendency in military CBAs carried out by the services to keep the program’s cost
estimates aggressively low—often resulting in program cost growth (Arena et al.
2006).7
Improving requirements generation and defense acquisition processes is criti-
cal. An unbiased military CBA framework is needed so that when a requirement
is validated, the expectation should be that it is based on mature technology with
more predictable costs and benefits, and that it is programmatically realistic so
that it fits into forecasted available future funding (see Chapter 4). Today, that is
generally not the case.
Finally, because “traditional” platform-based programs (ships, planes, tanks,
etc.) are strongly supported by the services (and industry and Congress), and many
of these programs are in current production and operational, the DoD budget has
only 10 percent of the total acquisition dollars allocated for equipment optimized
for emerging threats such as counterinsurgency, security assistance, humanitarian
operations, etc. (Gates 2010). Thus, a significant shift in resource allocation is
needed; and this begins with the “requirements process” that establishes security
needs (benefits) for the future.

2.7 System-of-systems
With major advances in information technology, many of DoD’s weapon systems
are now interconnected to form interdependent systems-of-systems. A system-of-
systems (SoS) is a set of individual systems that is integrated to operate optimally
as a single system. SoS development provides the DoD with the unique ability
to integrate and field a full range of assets (new and legacy) to provide required
military capabilities. Although one can conceive of developing an entire SoS from
scratch, most of DoD’s SoS efforts consist of synthesizing existing capabilities
28 Hon. J. S. Gansler, W. Lucyshyn
of legacy systems, combined with selected newly developed systems to address
additional capability needs.
Since it must pull together numerous projects and combine them into a cohe-
sive whole, the size and scale of a SoS complicates any military CBA and the
development of these integrated programs. The challenge for DoD is that, although
the systems will be integrated in this fashion, they are still developed as individual
systems.
In the traditional engineering framework, the goal of a typical development
project is to optimize the performance of a particular system (e.g. platform). Since
there are multiple systems within the overall SoS, each element is individually
optimized (to derive maximum individual performance). In this framework, it is
believed that the whole is equal to the sum of its parts.
A problem with this narrow engineering approach is that it fails to take into
account how changes in one component of the interrelated whole may affect the
performance or requirements (or costs) of other components of the entire system.
If one system relies upon a second system having a minimum threshold capabil-
ity that is not realized, the first may have to be modified to achieve the desired
effectiveness. Conversely, optimizing the performance of an individual platform
may not necessarily increase the performance of the SoS and it may, in some
cases, degrade its overall performance. Due to the size and interconnected nature
of SoS, one problem may have ripple effects throughout. Optimizing each indi-
vidual system can produce a sub-optimal result at the SoS level. The goal must be
to optimize the performance of the SoS as a whole and not individual component
systems.
Currently, a SoS program generally spans multiple organizations. As a result, a
number of authorities may be in a position to influence the decision-making pro-
cess. Managing the component systems should be done using a portfolio approach
with some flexibility to shift resources within the portfolio to meet the SoS
requirement. In this way, there would be a level of flexibility not available with
the “platform-centric” approach. Ideally, these SoSs will be able to leverage the
unique advantages of each component system and result in a capability greater
than the sum of its parts.
In the most likely future of limited resources, however, there is another problem
with optimizing individual systems versus the overall SoS: namely “affordabil-
ity” (see Chapter 4). SoS is a potentially valuable innovation that creates new
challenges for military CBAs. In a resource-constrained environment, the focus
of military CBA should be to optimize overall performance at an affordable price
(often integrating lower-cost individual elements).

2.7.1 Resource allocation within Department of Defense


The biggest issue facing the Defense Department will be the mismatch between
requirements and funds available.
General Charles Krulak, Former Marine Corps Commandant,
Elaine M. Grossman, National Journal, May 30, 2008
Allocating national security resources 29
The nation’s approach to allocating resources within the DoD, imperfect under
the best conditions, will be severely challenged in the face of an increasingly
unpredictable threat environment coupled with expected budgetary constraints.
Yet, even today, a great deal of effort is placed on identifying shortfalls and the
requirements to overcome them. Often this does not include the identification
of the necessary resources or where they will come from. In a most likely
future resource-constrained environment, a premium will be placed on resource-
constrained planning (see Chapters 4 and 7). The relationship between the
congressionally approved DoD budget and yearly outlays places great urgency on
the ability to fuse requirements planning with necessary budget funding. Planning,
that has always been difficult, based on the size and scope of the DoD enterprise,
is growing increasingly more complex and uncertain.
Defense planning takes place through the planning, programming, budgeting,
and execution (PPBE) process. The PPBE is the DoD’s internal management and
information system used to plan the allocation of resources to acquire the capa-
bilities believed necessary to accomplish the DoD’s missions. A key output of the
PPBE process is the defense budget included in the President’s budget submitted
to the Congress. This system has evolved from the planning, programming, and
budgeting system (PPBS), which was first introduced by Secretary of Defense
Robert McNamara in the early 1960s. It was developed to coordinate three major
facets of the DoD: planning, programming, and budgeting. As the name implies,
the system attempted to draw together these three facets of DoD within an internal
budgeting process.
The Congressional Budget and Impoundment Control Act was passed in 1974.
This established the statutory basis for an integrated Congressional budget pro-
cess, and provided for the annual adoption of a concurrent resolution on the budget
as a mechanism to facilitate Congressional budgetary decision-making. It also
established the current US federal fiscal year from October 1st to September 30th.
Changes continued and in 1986 the Department of Defense Authorization Act
was passed, requiring DoD to submit a two-year budget. While this was intended
to help promote more accurate forecasting within the DoD planning process, as
well as to shelter DoD from the unpredictability of annual Congressional review
and approval, the Act placed an even greater emphasis upon DoD’s ability to fore-
cast external global actions, internal US politics, and internal changes in DoD. To
date, the process has largely remained an “annual budget” process (in terms of
commitments).
In 2003, the PPBS was changed to the PPBE process. This change—an
attempt to place more emphasis upon “execution”—did not significantly change
the framework from which DoD plans, programs, budgets, or even executes its
budgets.
The first and arguably the most important step in the PPBE process is plan-
ning. A key element in the planning process is the quadrennial defense review
(QDR). The QDR, conducted every four years, attempts to provide a “compre-
hensive examination” of DoD’s national strategy, internal workings, and budget
plan in an attempt to frame and structure national defense strategy and programs.
30 Hon. J. S. Gansler, W. Lucyshyn
Mandated by Congress in 1997 through the National Defense Authorization Act,
the QDR was intended to establish a cohesive longer-term framework for defense
programming—with direct involvement of the Chairman of the Joint Chiefs of
Staff Popescu (2010).
Critics argue that QDR planners from the various services, in an attempt to pre-
serve their favorite weapon systems or increase their branch’s budget within DoD,
simply compete for prominent positions within this broad strategic framework.
The result is a document where the only non-bureaucratized portion is a set of
strategic, although sometimes unfocused and unconnected, bullet points some-
what detached from the reality of DoD’s year-to-year PPBE process until the next
QDR is launched.
During the planning process, the results of the QDR, along with other plan-
ning guidance (including the National Military Strategy), are reviewed, and the
process ends with the development of the Guidance for the Development of the
Force (GDF) and the Joint Programming Guidance (JPG). The GDF provides
broad guidance in the form of strategic goals and desired capabilities, while
the JPG addresses specific programs and provides dollar thresholds. These plan-
ning documents, along with broad fiscal guidance from the OMB, provide DoD
Components with direction on defense policy, strategy, force and resource plan-
ning, and fiscal matters. The Components use them to develop their Program
Objective Memorandum (POM), which extends six years into the future (Keehan
2008). These Component developed POMs are reviewed to ensure they conform
to guidance provided by the Secretary of Defense in the planning phase, and
changes are made as required. “Although the programming phase is a debate about
means—which program choices best achieve the stated goals—it often reopens
the debate about those goals, revisiting choices made in the planning phase”
(Chu 2003).
While DoD’s PPBE process attempts to address shifting requirements and
defense spending, there remains a mismatch between long-term DoD strategy and
resource allocation mechanisms. For example, the QDR—designed to provide a
long-run framework for US defense planning—inadvertently contributes to this
mismatch when it fails to pair the necessary future funding realities with identi-
fied requirements. Chapter 7 in this volume attempts to address the challenge of
forecasting future budgets.
Direct Congressional action on or intervention in DoD’s PPBE creates another
possible mismatch. Administrations can be unpredictable enough, but Congress
requires another level of forecasting: With 535 members, Congress can radi-
cally change the direction of its funding decisions in off-year elections. With
an exceptional number of opinions and votes to consider, forecasting domestic
actions at the Congressional level could baffle even the most skilled political
analyst.
The high-level significance placed upon requirements generation, in processes
such as the PPBE and the QDR, do not address two overarching problems. First,
DoD finds itself in a world with a large number of complex and uncertain threats
with an unprecedented potential for convergence (e.g. terrorists and WMD), and
Allocating national security resources 31
it has no precise mechanism to address these changes. DoD planning relies, to
a large degree, upon attempts to forecast this unstable threat environment, often
resulting in obsolete projections due to the rapidly changing landscape. This is
complicated and made more difficult within the context of continually changing
Congressional dynamics.
Second, although greater weight has been placed on strategic requirement
planning (QDR) and the execution of those requirements (PPBE), two major facets
of the process—programming and, more importantly, budgeting—are not only
de-emphasized as a result, but remain unconnected to the strategic requirement
or execution phases. Moreover, the POM process has a great deal of inertia and
significantly reduces flexibility. When a new mission need is identified, there is
often a two-year wait until funding for that purpose can be inserted in the next
POM. At that point, funding is often severely restricted over the first few years.
The combination of complex and uncertain threats, downward budgetary
pressure for the foreseeable future, and continuing acquisition challenges creates
a need for a more rigorous, analytic process to compare alternatives at the system
level, SoS level, and the enterprise level. This is one important goal of this book.
Based upon anticipated domestic budgetary challenges, the DoD must seek
ways to acquire and maintain the required force structure at minimum cost. Bas-
ing budget sourcing decisions upon cost minimization alone, however, could result
in just that—cost minimization at the expense of quality or effectiveness of DoD
weapon programs and IOC sustainment needs. Furthermore, the implications of
cost minimization on the execution of DoD operations could prove harmful, result-
ing in a failure to acquire the necessary capability, in both quantity and quality of
advanced weapons systems, essential to address anticipated threats.
DoD must not simply resort to cost-minimizing strategies, but must instead
identify the appropriate cost-performance trade-space and then make the necessary
tradeoffs using military CBA, ensuring the most efficient and effective defense
spending. This CBA approach is needed to ensure that DoD works within bud-
get constraints: it forces a new, long-term planning approach which incorporates
requisite resources into requirements identification and implementation processes.
Military CBA can also help decision-makers to identify and address the level
of uncertainty in their estimates and, then, it can be used to test their sensitiv-
ity to various assumptions. Private sector firms have found that although there
can be inaccuracies associated with some of the estimating involved, these analy-
ses provide a meaningful framework for the selection of capital projects (Gordon
2000).
Some defense organizations already use military CBA, albeit at a much more
tactical level. For example, according to Army Lt. Gen. William G. Webster
Jr., as part of the planning for the drawdown of forces in Iraq, US Central
Command’s (United States CENTCOM) 3rd Army (Logistics) conducted “a
cost–benefit analysis of every piece of equipment in Iraq, including the costs
of transporting it out.” Based on the outcome of the analysis, equipment was
either transported back to the United States, shipped to Afghanistan, or left (or
destroyed) in Iraq Carden (2010).
32 Hon. J. S. Gansler, W. Lucyshyn
Current DoD policy requires Services (or joint sponsors) to conduct an analy-
sis of alternatives (AoA) before they initiate major defense acquisition programs
and begin system development at Milestone B.8 The purpose of the AoA is to
compare the operational effectiveness, costs, and risks of alternative solutions that
address operational needs and requirements. However, a recent GAO report found
that most of programs reviewed “either did not conduct an AoA or conducted an
AoA that focused on a narrow scope of alternatives and did not adequately assess
and compare technical and other risks of each alternative.” GAO found programs
that conducted a limited assessment generally did not do as well as those that
conducted a more robust AoA (Sullivan 2009).
While these limited efforts should be lauded, we believe military CBA can pro-
vide a useful framework for acquisition decisions during all phases, at all levels,
and throughout DoD. Unfortunately, at the enterprise level, the analysis becomes
even more complex as one considers the interdependent nature of many of the sys-
tems. Therefore, new programs need to be considered within a portfolio or part of
a SoS. On occasion, systems may look attractive on an individual basis, but when
considered in the broader perspective, in combination with other new systems or
legacy systems, they may, in fact, not provide the anticipated benefits. A military
CBA conducted at the portfolio level would provide DoD with the ability to make
more effective budgetary tradeoffs and program decisions and to help bridge the
gap between requirements and available and future funding.9

2.8 Conclusion
Challenges facing US National Security begin with an increasingly complex and
uncertain environment of global threats and also include domestic budgetary con-
straints that will significantly limit the resources available to address those threats.
Coupled with top-line budgetary pressures, DoD must also deal with the chal-
lenge of the year-to-year line-item budgeting process and ongoing acquisition
challenges, all of which threaten to severely limit DoD’s capacity to support
current acquisitions. All of this must be done while balancing the need for new sys-
tems to address evolving missions, such as “irregular” operations, missile defense,
and cyber-defense. And, of course, sound acquisition decisions require accurate
forecasts of vendor costs and future defense budgets. This “witches’ brew” of
conditions calls for the wider implementation of military CBA-guided resource
allocation decisions throughout DoD. The goal is for the nation to receive the best
value for every dollar it invests in defense programs.
Until new mechanisms are adopted that explicitly incorporate military CBA to
address failures to fuse requirements with necessary resources, DoD will risk a
disjointed, ineffective approach that may not properly address national security
concerns. A comprehensive military CBA approach such as that proposed in this
volume is vital in confronting the changing dynamics of modern, global warfare.
Even if this approach is adopted, however, DoD must still engage in a contin-
uing process of reassessment and innovation in order to retain the strong bond
between resources and requirements these innovations may bring. This ability
Allocating national security resources 33
to constantly innovate will be central to remaining effective in the face of an
increasingly complex and uncertain national security environment.

Notes
1 For example, in Somalia 1992, Iraq 2003 and Libya 2011.
2 In recent updates of Social Security financial projections, the US Congressional Budget
Office (CBO) notes that the Social Security deficit will grow to be significantly larger by
2020 than previously expected (CBO 2010).
3 Social Security, Medicare, and Medicaid have reached 10.1 percent of GDP (2009) and
are projected to continue increasing as a percentage of GDP to 11.8 percent in 2020,
15.5 percent in 2040, and 21.1 percent in 2075 (CBO 2010). According to GAO, these
programs will near 30 percent of GDP by 2040.
4 Both CBO and GAO have conducted long-term studies surrounding the federal bud-
get. Both have noted the dramatic increases in entitlement spending and interest on
the national debt. GAO projections have been carried out until 2084 based upon CBO
assumptions and projections ending in 2020.
5 This current projection was diverted from the baseline after an assumption that the US
will draw down troops in Iraq, Afghanistan, and elsewhere to 60,000 by 2013. Other
projections based upon different troop levels or different spending freezes, however, place
DoD’s budget even lower as a percentage of GDP.
6 Compounding future funding challenges is the fact DoD has come to rely on “supplemen-
tal” funding which is being significantly reduced. The Administration requested US$33.0
billion in 2010 supplemental funding on top of the US$129.6 billion already provided and
a total of US$159.3 billion for its 2011 overseas contingency operations (OMB 2010).
7 The Cost Analysis Improvement Group (CAIG) launched by what is now known as
CAPE (Cost Assessment and Program Evaluation) was meant to be responsible for pro-
viding independent cost estimates for the US Secretary of Defense: see www.cape.osd.mil
[last accessed November 25, 2014].
8 Milestone B is normally the formal initiation of an acquisition program (see DoD Instruc-
tion 5000.02, November 25, 2013, http://www.acq.osd.mil/osbp/sbir/docs/500002_
interim.pdf [last accessed February 9, 2015]).
9 Chapter 10 offers an interesting theoretical framework that captures aspects of a more
general portfolio approach, while Chapter 4 explicitly addresses the issue of affordability.

References
Al-Salhy, S. and T. Arango. June 12, 2014. “Iraq Militants, Pushing South, Aim at Capital.”
New York Times: A1.
Arena, M. V., R. S. Leonard, S. E. Murray, and O. Younossi. 2006. Historical Cost Growth
of Completed Weapon System Programs (Technical Report Series). Santa Monica, CA:
RAND Corporation.
Boessenkool, A. November 11, 2009. “Analysts: U.S. Procurement Drop Coming as Costs
Mount.” Defense News.
Carden, M. J. April 8, 2010. “Army Command Sustains Logistics for Warfighters.” Fort
Hood Sentinel.
CBO. 2010. The Budget and Economic Outlook: Fiscal Years 2010 to 2020. Congressional
Budget Office.
———. 2010. Long Term Implications of the Fiscal Year 2010 Defense Budget. Congres-
sional Budget Office.
Cha, V. D. 2000. “Globalization and the Study of International Security.” Journal of Peace
Research 37(3): 391–403.
34 Hon. J. S. Gansler, W. Lucyshyn
Chesbrough, H. W. 2006. “Open Innovation: A New Paradigm for Understanding Industrial
Innovation.” In Innovation: Researching a New Paradigm (Chapter 1). Oxford: Oxford
University Press.
Chu, D. S. C., and N. Berstein. 2003. “Decisionmaking for Defense.” In New Challenges
New Tools for Defense Decisionmaking, edited by M. L. Stuart and G. F. Treverton.
Santa Monica, CA: RAND Corporation.
Frank, D. F. Summer 1997. “A Theoretical Consideration of Acquisition Reform.” Acqui-
sition Review Quarterly 4(3): 279.
Freier, N. 2007. Strategic Competition and Resistance in the 21st Century: Irregular,
Catastrophic, Traditional, and Hybrid Challenges in Context. Author via CreateSpace
Independent Publishing Platform (a DBA of On-Demand Publishing LLC, part of the
Amazon group of companies).
Fukuyama, F. Summer 1989. “The End of History.” The National Interest: 3–18.
Gates, R. M. May 3, 2010. Remarks as Delivered by Secretary of Defense Robert M. Gates
(cited May 5, 2010). Gaylord Convention Center, National Harbor, MD. Available at
www.defense.gov/speeches/speech.aspx?speechid=1460 [last accessed November 25,
2014].
Gordon, L. A. 2000. Mangerial Accounting: Concepts and Empirical Evidence (5th
edition). Columbus, OH: McGraw-Hill Higher Education.
Grossman, E. 2008. “Marine General Lays Groundwork for Unprecendented Change.”
Government Executive. May 30,2008. Available at www.govexec.com/defense/2008/
05/marine-general-lays-groundwork-for-unprecedented-change/26953/ [last accessed
November 25, 2014].
The Institute for Economics and Peace (IEP). 2012. Global Terrorism Index 2012: Captur-
ing the Impact of Terrorism for the Last Decade (IEP Report Series). New York: The
Institute for Economics and Peace.
Keehan, M. P. 2008. Planning, Programming, Budgeting, and Execution (PPBE) Proccess
(Teaching Note). Fort Belvoir, VA: Defense Acquisiton University.
Perl, Raphael. 2006. Terrorism Expert Calls for a Redefinition of National Security. US
Department of State, Office of International Information Programs, Washington, DC.
Perlo-Freeman, S. and C. Solmirano. April 2014. “Trends in World Military Expenditure,
2013.” Stockholm International Peace Institute Fact Sheet.
Popescu, I. C. March 1, 2010. “The Last QDR? What the Petagon Should Learn from
Corporations about Strategic Planning.” Armed Forces Journal: 26.
Panel on Defense Acquisition Reform. 2010. Interim Findings and Recommendations.
Washington, DC: House Armed Services Committee.
Schwirtz, M., A. Barnard, and C. J. Chivers. August 9, 2008. “Russia and Georgia Clash
Over Separatist Region.” New York Times.
Smith, R. 2005. The Utility of Force: The Art of War in the Modern World. London: Allen
Lane.
Sullivan, M. J. 2008. Defense Acquisitions: Fundamental Changes are Needed to Improve
Weapon Program Outcomes (GAO-08-1159T). Available at www.gao.gov/products/
GAO-08-1159T [last accessed November 25, 2014].
———. 2009. Many Analyses of Alternatives have not Provided a Robust Assesment of
Weapon System Options (GAO-09-665). Available at www.gao.gov/products/GAO-09-
665 [last accessed November 25, 2014].
United States Census Bureau. 2004. Available at www.census.gov [accessed November 15,
2007].
United States Government Accountability Office. 2010. The Federal Government’s Long-
Term Fiscal Outlook. GAO-10-468SP.
Allocating national security resources 35
US Office of Management and Budget (OMB). 2010. President’s Budget for FY 2011.
Washington, DC: Executive Office of the President.
Walker, D. July 15, 2014. “Trends in U.S. Military Spending.” Council on Foreign
Relations. Available at www.cfr.org/defense-budget/trends-us-military-spending/
p28855 [last accessed November 25, 2014].
Younossi, O., M. V. Arena, R. S. Leonard, C. R. Roll, Jr., A. Jain, and J. M. Sollinger. 2007.
Is Weapon System Cost Growth Increasing? A Quantitative Assessment of Completed
and Ongoing Programs. Santa Monica, CA: RAND Corporation.
3 Measuring defense output
An economics perspective

Keith Hartley1 and Binyam Solomon

3.1 Introduction
Measuring defense output is a necessary step to successfully apply military cost–
benefit analysis (CBA) to evaluate alternative security investments. Other chapters
in this book focus on what economists call “intermediate outputs” (e.g. military
forces). This chapter offers a higher-level “macro” perspective of overall defense
output that encompasses total defense spending.
In most countries, the defense sector absorbs substantial scarce resources that
have many valuable alternative uses (schools, hospitals, etc.). Whereas defense
expenditures are well known within each country, there is no single indicator of
value (or benefit) of overall defense output. This contrasts with the valuation of
private sector outputs in market economies. In defense, the economist’s solution to
measuring output assumes output equals inputs (a convention widely used across
the public sector), or that the value of defense output is roughly equivalent to
expenditures made to produce that output.
In sharp contrast, measuring the value of market outputs is not usually regarded
as a policy problem. Market economies ‘solve’ the problem through market prices
that reflect choices of large numbers of buyers and sellers. Defense, however, dif-
fers in several key ways from private markets, which helps to explain the challenge
in measuring and valuing defense output.
An important step in applying military CBA to evaluate security investments is
to discuss output measures. This chapter examines the measurement of defense
output from an economics perspective. Specifically, the chapter identifies key
questions which need to be addressed in measuring overall defense output. These
include: What is defense output? How can it be valued? Under what conditions is
it a worthwhile investment?
Economic theory offers some policy guidelines for determining the optimal
defense output for any society. As an optimizing problem, the rule is to aim at the
socially desirable or optimal level of defense output. This is achieved by equating
additional or marginal costs of proposed defense expenditures with additional or
marginal benefits. While the economics approach is difficult to ‘operationalize’
into a set of clear, unambiguous policy guidelines, it does provide a framework
for designing valuations for defense outputs and activities.
Measuring defense output 37
This chapter is organized as follows. Section 3.2 reframes the defense output
measurement issue as an economic problem utilizing insights from public goods,
public choice, and principal-agent models. Section 3.3 presents the military pro-
duction function as an early attempt to quantify defense outputs, and reveals
some challenges in operationalizing the model. Section 3.4 discusses more recent
attempts at measuring defense outputs through the transformation of defense bud-
gets reporting. Section 3.5 analyzes both economic and non-economic benefits of
defense, while Section 3.6 surveys international experience in measuring defense
outputs. Nations surveyed include Australia, New Zealand, the United Kingdom,
the United States, and a group of European nations.

3.2 Economic theory as a guide


3.2.1 Public goods
Defense is a classic example of a public good, and its desired outcome in the form
of peace is also a public good. A public good is non-rival and non-excludable.
For example, living as neighbors in the same city, my consumption of air defense
protection does not affect your consumption and, once provided, I cannot exclude
you from its consumption nor can you exclude me. In sharp contrast, private goods
such as motor cars and TV sets are rival and excludable. Your payment and con-
sumption of those items means that I cannot simultaneously use them (unless you
choose to share), and private property rights guarantee your exclusive ownership,
so that you can legally exclude me from using them.
The public goods features of defense provide incentives for free-riding. Since
I cannot exclude you from the benefits and you cannot exclude me, each of us is
inclined to let the other pay for protection. Free-riding is a contentious issue both
within a nation and between nations in a military alliance (e.g. NATO, or US–
Canadian security). This ultimately results in a nation’s citizens failing to reveal
their true preferences for, and valuations of, defense. A challenge for the state
in providing and financing defense is that it does not know the true preferences
of the potential beneficiaries of defense: it cannot easily quantify the volume of
the defense public good demanded by consumers and estimate the true price the
beneficiaries are willing to pay (Engerer 2011).
Theoretical solutions exist to estimate the optimal amount of a public good, but
are difficult to operationalize in practice (Cornes and Sandler 1996). Public opin-
ion polls can be used, but these are a limited mechanism for accurately assessing
society’s opinions on defense spending and defense policy and the willingness of
citizens to pay for defense (Zaller and Feldman 1992).
Alternatively, one can frame the question of “how much defense is enough?,”
presenting it as an optimization where the economic decision rule is to achieve
a socially desirable or “optimal” defense output. In principle, this is found by
equating marginal costs with marginal benefits. This approach is difficult to
‘operationalize’ into a set of clear and unambiguous policy guidelines. As dis-
cussed in the introductory chapter, marginal costs and especially marginal benefits
38 K. Hartley, B. Solomon
of many defense investments are not immediately obvious, and are difficult to
quantify.
The economic model assumes a social welfare function showing society’s pref-
erences between defense (security) and civilian goods: again, this is an attractive
concept but not one which is readily operationalized or easily identifiable for any
society. Moreover, the benefits of defense are complicated by its public good and
free-riding characteristics. Voting systems may also not be reliable and accurate
methods of revealing preferences for specific public goods and services. Typically,
elections involve choices between political parties that offer various tax and spend-
ing policies, where defense budgets and security policies are often buried in a wider
policy platform. Problems can also arise in attempting to aggregate voter preferences
into a ranking for society as a whole (the voting paradox: Tisdell and Hartley 2008).
Further problems arise since the economic model assumes maximizing behaviour
on the part of individuals, when most agents might instead be satisficers, willing to
settle for acceptable solutions short of the optimum (Hartley 2010b).

3.2.2 Markets
There are several major differences between private markets and public (defense)
markets. Private markets involve prices that reveal society’s valuation of outputs,
where these prices reflect market incentive and penalty mechanisms. Goods that
are ‘private’ rather than public are characterized by both excludability and rivalry;
large numbers of private consumers and buyers; rivalry between firms; motivation
and rewards through profits; and a capital market that imposes penalties on poor
economic performance through take-overs and the ultimate sanction, bankruptcy
(with managers often losing their jobs).
Public bureaucracies such as the armed forces lack such incentive and penalty
mechanisms, and they consequently tend to be slow to adjust to change. Often,
change in the armed forces results from budget pressures, new technology, vic-
tories and defeats, and occasionally, views of senior military leaders (Solomon
et al. 2008). In contrast to private markets, there is no market price for publicly
provided defense forces: for example, there are no market prices for submarine or
tank forces.
Although some rivalry exists between suppliers (Navy, Army, Air Force, etc.),
there is no profit motive for public suppliers, nor capital market pressures cor-
responding to takeovers and bankruptcy in private markets. Defense has another
distinctive feature reflected in the state-funding and state provision (ownership) of
its armed forces. Governments are monopsony buyers and monopoly providers of
armed forces.
This contrasts with private markets where there are large numbers of buyers
and rivalry amongst suppliers. State-owned and funded defense markets are less
likely to undertake worthwhile changes (Tisdell and Hartley 2008, Chapter 10).
There is also a unique military employment contract which differs drastically from
private sector employment contracts. The military employment contract requires
military personnel to obey commands which relate to type, duration, location, and
Measuring defense output 39
conditions of work (e.g. world-wide deployments) with significant probability of
injury and even death. Such a contract contains elements resembling indentureship
and command systems.
Each of the armed forces is a monopoly supplier of air, land or sea systems with
monopoly property rights in the air, land or sea domains. There are barriers to new
entry which prevent rival internal armed forces from offering competing products.
For example, armies often operate attack helicopters and unmanned aerial vehi-
cles (UAVs) which are rivals to close air support and surveillance provided by
air forces. Similarly, land-based aircraft operated by air forces are alternatives to
naval carrier-borne aircraft. Efficiency requires a mechanism for promoting such
competition; instead, each service guards its traditional monopoly property rights
in the air, land, or sea domain, thereby creating barriers to new entry.
This has an impact on efficiency. Specifically, is the correct amount of output
being produced? Is the correct mix of inputs being used? As monopolies with
significant barriers to entry, each of the armed forces lacks strongly competing
organizations, and hence has less incentive for efficiency improvements and for
innovation (where efficiency embraces both allocative and technical efficiency).
Allocative efficiency requires the choice of socially desirable output, and tech-
nical efficiency requires the use of least-cost methods to produce that output.
Again, problems arise in determining allocative efficiency (see a discussion below
on principal-agent models). Technical efficiency, however, can be assessed by
allowing activities traditionally undertaken ‘in-house’ by the armed forces to be
‘opened-up’ to competition from private suppliers (market testing leading to mili-
tary outsourcing). Indeed, military cost–benefit formulation of such competitions
can offer improvements in allocative efficiency (e.g. by inviting competition for
different levels of service in order to identify true marginal costs for different
levels of output or service).2
Internal defense markets lack other incentives of private markets. There are no
profit incentives to stimulate and reward military commanders to search for and
introduce productivity improvements or to identify new and profitable opportuni-
ties (for example, the role of entrepreneurs in private markets). The absence of a
capital market also means that military managers are unlikely to lose their jobs
for poor performance and there are no capital market opportunities for promot-
ing and rewarding mergers and take-overs. For example, a military commander
of a regiment cannot merge with another regiment to achieve economies of scale
and scope, nor can an Army regiment acquire Air Force or Navy transport units
where such mergers might offer both cost savings and output improvements (such
as horizontal, vertical, or conglomerate mergers).
Uncertainty dominates defense policy. Defense policy has to respond to a range
of future threats, some of which are unknown and unknowable. Assumptions are
needed about likely future allies and their responses to threats, the location of
threats, new technologies, and the time dimension of threats (e.g. today, in 10–15
years, or 30–50 years ahead where uncertainties are greatest). These uncertain-
ties mean that forces have to be capable of adapting to change, and that today’s
defense investments must be capable of meeting tomorrow’s threats. Admittedly,
40 K. Hartley, B. Solomon
the private sector also faces considerable uncertainty about future markets and new
technologies, and these unknowns extend over lengthy time horizons. Defense is
different, however, in that uncertainties are dependent upon, and determined by,
governments, nation states, and some non-state actors, rather than by the actions
of large numbers of private individuals as consumers, workers, and shareholders.
There is one further key difference between defense and private markets. Defense
aims to avoid conflict, but where conflict arises it often destroys markets and valuable
infrastructure and creates disequilibrium as resources are re-allocated to military
forces to gain strategic advantage, with consequent opportunity costs in civilian
goods and services. War involves the destruction of labor and capital. In contrast,
private markets seek the optimal mix of labor and capital to provide goods and
services through voluntary trading and exchange. Resource allocation is based on
price and profit signals that lead to “creative destruction” reflected in continuous
investment in new innovations, inventions and the output of new goods and services.

3.2.3 Public choice and principal-agent models


Defense decisions are made in political markets, which is another reason for their
departure from the economist’s optimizing solution. Political markets comprise
voters, political parties, bureaucracies, and interest groups—each pursuing their
own self-interest. Voters as taxpayers are “principals.” They want something pro-
vided by government and appoint “agents” to perform the necessary tasks. The
challenge for principals is to design incentive mechanisms to ensure that agents
pursue the aims of the principals rather than their own objectives. For example,
voters as principals seek peace, security, and protection, but the priority of their
agents in defense ministries and the armed forces might be to subsidize costly
local goods and services because doing so offers jobs, new technology, or export
benefits which contribute to re-election of the governing party.
Limitations of the voting system as a means of expressing voter preferences for
defense spending constrain principals (voters) in their attempts to guide their agents
(elected politicians and bureaucrats). Free-riding further affects the willingness of
voters to accurately reveal their preferences for defense. Principals also often lack
the necessary information to make informed and rational defense choices in their
voting decisions. The result is that agents often have a wide scope in shaping national
defense policy and in pursuing their own self-interests (e.g. re-election).3
The principal-agent model has implications for CBA, resource-use, and effi-
ciency in defense markets. It also has implications for measuring the benefits of
defense outputs where these reflect a combination of principal and agent choices.
The model can also be linked to the political market where defense choices are
made. In political markets:

• voters and taxpayers as principals will seek to maximize the benefits


(satisfaction) from their votes;
• political parties are vote-maximizers; governments seek re-election and are
the agents of voters;
Measuring defense output 41
• bureaucracies can be modelled as budget maximizers acting as agents of the
government;
• producer (e.g. defense industry) groups are profit- or rent-seekers in their roles
as agents of the defense procurement agency or military bureaucracy.

The principal-agent and public choice models provide a useful analytical


framework to help understand the military-industrial-political complex and its
influence on defense choices and outputs. As principals, voters are generally
poorly informed about defense policy so they defer to various agents to make
their defense choices—namely, governments, civil servants in defense ministries
and procurement agencies, and the armed forces. In turn, these agents are influ-
enced by powerful producer groups in the form of large defense contractors (e.g.
via lobbying) seeking lucrative defense contracts.
Examples of the influence of the military-industrial-political complex on
defense choices abound. Government ministers will be aware of the vote-
consequences of defense choices (e.g. impacts of base and plant closures and
the benefits of awarding defense contracts to firms in marginal constituencies).
Defense ministries and the armed forces as budget maximizers are tempted to
overestimate the threat and underestimate the costs of their preferred policies and
projects. Exaggerating the threat from terrorism enables the armed forces to obtain
larger defense budgets. Understating costs of a new weapon system in a CBA
allows the project to start, and once started projects build interest groups and
become difficult to stop (a key factor in the ‘optimism bias’ that often leads to
contract cost overruns).
Defense contractors also have an incentive to make optimistic claims to pump
up perceived benefits and increase their chances of being awarded valuable
defense contracts. For example, claiming that the contract will contribute valu-
able jobs, technology, spin-offs, exports, and that it is ‘vital’ to the future of the
national defense industrial base. Rarely is attention given to opportunity costs:
namely, whether the resources used in the defense project might provide even
greater net economic benefits if used in alternative sectors of the economy. Over-
all, public choice and principal-agent models reveal how special interest groups
influence military investment decisions and defense outputs.

3.3 The military production function


Another contribution from economic theory to output measurement comes in the
form of the military production function. This is an input-output relationship that
attempts to relate all defense inputs to a final defense output. Inputs comprise tech-
nology, capital (bases, equipment, spare parts, etc.), and labor (military personnel
in the form of conscripts and/or volunteers, civilians, contractors, etc.). A formal
expression of the function is

Q = f (A, K , L) (3.1)
42 K. Hartley, B. Solomon
where Q is defense output and A, K , and L are inputs of technology (A), capital
(K ), and labor (L).
While the model appears attractive, there are at least four major caveats. First,
a production function assumes factor inputs are combined to minimize costs. This
assumption is unrealistic in view of the lack of efficiency incentives in internal
defense markets: there are few rewards or penalties to achieve least-cost produc-
tion. Second, all defense inputs have to be identified and correctly valued. Third,
defense output is simply asserted without recognizing the problems of identify-
ing and measuring output, including the multi-product nature of overall defense
output. Fourth, the model simply identifies defense outputs resulting from various
inputs: there are no criteria for determining society’s preferred defense output (the
“optimal” defense output).
Two central problems with military production functions arise over inputs and
outputs. Consider the problem of identifying and valuing all relevant inputs. These
comprise technology, capital and, labor, and include the following items:

• technical progress as reflected in inputs embracing new equipment and new


military facilities, including communications;4
• physical capital comprising a mix of equipment, military bases, land, and
logistics (repair and maintenance);
• human capital comprising military personnel reflected both in their numbers
the skills and productivity of the military labor force.5

While measuring inputs is a challenge, identifying, measuring, and valuing


defense output is even more challenging. Economic theory simply asserts the con-
cept of defense output without exploring its definition and multi-product nature.
Few published studies have estimated military production functions. Typically,
such studies have estimated readily identified measures of effectiveness, such as
providing an air defense capability, the numbers of aircraft destroyed, or the num-
ber of aircraft sorties per day. This approach is used in cost-effectiveness studies
that focus on intermediate defense outputs (Hildebrandt 1990, 1999).
For example, a military CBA of air defense would compare the costs and effec-
tiveness of alternatives such as land-based air defense missiles versus manned
fighter aircraft; or anti-submarine capability would compare land-based maritime
patrol aircraft versus naval frigates; or anti-tank capability would compare missiles
and attack helicopters. A different approach used in a more recent study estimated
a military production function where various defense inputs were used to estimate
the probability of winning in various conflict scenarios (Middleton et al. 2011).
A variant of the military production function is a defense R&D production func-
tion. This shows that current defense R&D determines future military equipment
quality, with an impact on future defense output. The relationship between defense
R&D and equipment quality is positive, but subject to diminishing returns and
substantial lags. For example, today’s military equipment quality was determined
by defense R&D spending some 10–15 years ago.
Measuring defense output 43
It can be informative to convert equipment quality into a time advantage. Thus,
over the period 1991–2001, US military equipment was six years ahead of that of
the UK, seven years ahead of France, and twelve years ahead of Sweden (Middleton
et al. 2006). The defense R&D production function can be expressed as

E q = f (R Dd , Z ) (3.2)

where E q is military equipment quality (e.g. British versus US tanks), RDd


represents defense R&D, and Z captures all other factors.
The defense R&D production function needs more theoretical and empirical
work. For example, ‘other factors’ might contribute to equipment quality and
these need to be identified specifically in the model. Furthermore, links between
equipment quality and military capability need identifying, including the role of
variables such as military skills which also contribute to final defense output. The
model is also limited in focusing on aggregate defense R&D spending without
any analysis of the most effective mix of R&D spending. More empirical work is
needed to determine the most cost-effective ratio of research to development work
within total defense R&D budgets, and the impact of that R&D on equipment
quality and future output/capabilities.6
Traditionally, defense outputs were measured on an input basis, where input
costs were assumed to equal the value of outputs. Table 3.1 presents some input
data of the type typically used for measuring some of the inputs in a military
production function.

Table 3.1 Defense inputs for a group of nations (2009)

Country Defense spending Defense Armed Defense R&D


(US$ million, share of Forces (US$ million,
2009 prices) GDP (%) personnel 2000 prices)c
(000s)

Australiaa 20,109 1.8 58 242.7


Canadab 19,869 1.5 67 201.6
Chinaa 98,800 2.0 2,285 Not available
Franceb 54,446 2.1 243 3,643.5
Germanyb 47,466 1.4 254 1,103.2
Indiaa 36,600 2.6 1,325 Not available
Italyb 30,489 1.4 197 64.9
New Zealand 1,358 1.2 9.8 Not available
Spainb 16,944 1.2 134 1,666.5
Swedena 6,135 1.3 13 218.7
UKb 59,131 2.7 197 2,559.9
USAb 574,070 4.0 1,368 65,896.0

Sources: NATO (2010); OECD (2010); SIPRI (2011).


Notes
a Defense spending data in 2008 prices: source SIPRI (2011).
b Data for NATO nations is provided from one source and is on a consistent basis.
c Defense R&D data are in US$ millions 2000 prices and PPP rates.
44 K. Hartley, B. Solomon
3.3.1 Technical spin-offs
Defense R&D can also contribute to wider economic benefits in the form of tech-
nical spin-offs and spill-overs (external benefits: positive externalities or external
economies). Among numerous examples are: jet engines, avionics, radar, compos-
ite materials, the internet, and the application of helicopter rotor blade technology
to wind turbines. These positive externalities might be regarded as part of the ben-
efits of defense investments contributing to defense output that need to be included
in a military CBA. But such views need to be assessed critically. Technology spin-
offs are not the main aim of defense spending, which primarily seeks to provide
peace, protection, and security. Any technical spin-offs can be regarded as a wind-
fall benefit of defense spending. Moreover, a list of spin-off examples fails to
address the central question of the market value of such spin-offs, and whether
there are better alternative uses of defense R&D resources. Consideration also
needs to be given to the wider economic impacts of defense spending.

3.3.2 Defense-growth relationships


A considerable literature has developed on the relationship between defense
spending and the benefits to a nation’s economic growth. There are two alter-
native hypotheses. First is the view that defense spending favourably affects an
economy’s growth rate (a positive impact: Benoit 1973). Second is the contrasting
hypothesis that military expenditure adversely affects a nation’s growth rate (a
negative impact: Deger and Smith 1983). Some of the literature has widened the
possible relationship to include the impact of defense spending on other macroeco-
nomic variables such as employment, unemployment, inflation, exports, and R&D
(Hartley 2010a).
Both hypotheses are dominated by myths, emotion, and special pleading. Plau-
sible explanations can be provided for either a positive or negative impact of
defense spending on growth, and there is evidence supporting both! The divergent
results reflect the need for a properly specified model of economic growth. Typi-
cally, defense spending is simply added to a conventional growth model without
careful consideration of its causal impact on growth.7
A considerable literature has used Granger causality tests to examine the rela-
tionship between military spending and the economy. A critique of this literature
concludes that parameters may not be stable over different time periods, different
countries, and that:

Granger causality test statistics are uninformative about the size and direc-
tion of the predicted effects and Granger causality measures incremental
predictability and not economic causality.
(Dunne and Smith 2010, 440)

This particular critique ends with the need to provide “measures of the political
and strategic determinants of military expenditures, such as threats” (Dunne and
Smith 2010, 440).
Measuring defense output 45
3.4 Assessing defense outputs: problems
and challenges
Defense outputs involve a complex set of variables concerned with security, pro-
tection, and risk management, including risks avoided, safety, peace, and stability.
Private markets routinely provide benefit measures such as sales, labor productiv-
ity, and profitability. Unlike private markets, there are no precise benefit measures
for defense output.
Defense inputs are more easily identified, measured, and valued than outputs, as
reflected in many nations’ annual input-oriented defense budgets. For economists,
questions then arise as to whether annual defense budget information provides suf-
ficient data to assess the efficiency and effectiveness of military expenditure: how
do expenditures on inputs correspond to desired defense outputs? Do defense bud-
gets provide policy-makers and politicians with the sort of data needed to conduct
military CBAs?
Typical questions include assessing the benefits and costs of alternative defense
forces; expanding (or contracting) the Army, Navy, or Air Force; substitut-
ing equipment (capital) for military personnel (labor); or substituting national
guard and reserves for regular (active) forces. Various defense budgets used by
nations include: input budgets, output budgets, management budgets, and resource
accounting budgets.

3.4.1 Input budgets


Input budgets provide some limited information on defense inputs such as the
pay of military and civilian personnel, as well as the cost of land, machinery,
and internal financial transactions, such as write-offs of various types of losses
(see Table 3.2). The information in Table 3.2 and particularly the first ten items
show the inputs used by the Canadian Department of National Defense (DND)
in the production of national security outputs. More than half of the budget is
spent on personnel, but there is no information on the proportion dedicated specif-
ically to civilian, regular, and reserve personnel. The last two items referred to as
“transfers” and “subsidies” detail payments in the form of grants and contributions
to various national and international organizations, capital assistance (subsidies)
to industry, research grants, and other assistance towards research carried on
by non-governmental organizations. These might be considered as intermediate
outputs.
Crucially, however, such budgets have major limitations for assessing efficiency.
First, the budget fails to show any final defense outputs other than under the vague
heading of “defense.” Second, it does not relate inputs to specific intermediate out-
puts (e.g. air defense; anti-submarine defenses). Third, inputs focus on the current
year only and do not reflect life-cycle cost implications of current procurement
decisions. Fourth, inputs are not always valued in terms of market values. For
example, some resources such as military bases and land for training might have
been purchased years ago and are assumed to be free or available at a zero price,
while other resources such as conscripts are not priced at their true labor market
46 K. Hartley, B. Solomon
Table 3.2 Annual defense expenditures FY 2012–13

CDN$000

Personnel 10,438,096
Transportation and communications 768,058
Information 13,666
Professionnal and special services 2,982,038
Rentals 383,972
Purchased repair and maintenance 1,465,091
Utilities, materials and supplies 1,017,831
Construction and/or acquisition of land, buildings and works 500,631
Construction and/or acquisition of machinery and equipment 2,434,609
Transfer payments 181,705
Total subsidies and payments 221,561
Total gross expenditure 20,407,258
Total revenue 429,068
Total net expenditure 19,978,190

Source: Public Accounts Volume II. Receiver General of Canada (RGC 2013).

values. These and other limitations led to the development of output budgeting
(Hartley 2011).

3.4.2 Output budgets


Output budgets, also known as program budgets, are much closer to the
economist’s production function model of defense budgets (Hitch and McKean
1960). Together with their costs, they provide information on some intermediate
outputs of defense such as nuclear strategic forces, air defense, aircraft carriers,
infantry regiments, and reserve forces. Output budgets also provide information
on substitution possibilities (e.g. between nuclear and conventional forces, or
between reserves and regulars).8
There are at least two major limitations with output budgets. First, the expendi-
ture figures used in output budgets are unlikely to be least-cost solutions, because
of the lack of competition and market incentives. Second, whilst they are known
as output budgets, there remains a problem in identifying the overall output of
defense. Often, outputs are defined in terms of the numbers of military person-
nel, aircraft squadrons, warships, and infantry regiments. These published data,
however, are measures of intermediate rather than final outputs such as protection,
security, safety, peace, and stability.
For example, the quantity of military personnel can be a misleading mea-
sure if training, productivity, and readiness for operations is ignored. Similarly,
the numbers of aircraft, tanks, and warships are misleading without data on
their average age and their operational availability. Also relevant are the com-
binations of military personnel and equipment required for effective forces with
an ability to be deployed and sustained in different overseas locations for long
periods.
Measuring defense output 47
3.4.3 Management budgets
Management budgets attempt to focus on efficiency. Top level and lower level
budget holders are identified and awarded cash budgets where delegated financial
powers allow military commanders and managers to combine resources to achieve
agreed objectives. Inevitably, however, there are problems with management bud-
gets. Budget holders (e.g. commanders of bases and units) often face constraints
on their freedom to vary the mix of inputs of capital and labor (equipment and
personnel). It is not unknown for large items of expenditure to be pre-committed,
leaving base and unit commanders with choices about relatively minor items of
expenditure (e.g. window cleaning, catering, and transport).
Efficiency incentives are also reduced if cost savings are automatically trans-
ferred to the defense ministry or the national treasury. Nor can efficiency be
achieved without clearly specified defense output targets. Cost savings can always
be realised by reducing the quantity and/or quality of output, especially if output
targets are not clearly specified!

3.4.4 Resource accounting and budgeting


The UK adopted resource accounting & budgeting (RAB) in 2002 to bring public
sector accounting practices into line with those in the private sector. The key fea-
ture of RAB is that costs are accounted for as they are incurred (as with “accrual
accounting”) rather than when payments are made (Table 3.3).9 There is an annual
balance sheet for the Ministry of Defence (MoD) showing fixed and current assets,
provisions, and liabilities. Data on the value of the MoD’s fixed assets includes val-
uations for fighting equipment and the defense estate (e.g. military bases and land
for training). By revealing the costs of holding assets, RAB provides incentives
for efficiencies such as the disposal of surplus spare parts, land, bases, and even
estates. Canada’s DND example of RAB and output budgets, known as Program
Alignment Architecture (PAA), is provided in Appendix 3.1.10
The adoption of private sector management and accounting practices such as
RAB or PAA by themselves will not lead to overall efficiency in the MoD. The
private sector has a range of mechanisms and incentives for achieving efficiency,
including competition, the profit motive and the capital market with the threat of
take-overs and bankruptcy. Such mechanisms and incentives are absent from the
MoD (and most other public agencies).
All parts of the public and private sectors consist of individuals and groups with
incentives to pursue their own self-interests (recall the principal-agent model).
The task for the MoD is to provide efficiency incentives equivalent to those in the
private sector. Here, the challenge of measuring defense output remains a serious
obstacle to assessing efficiency in the defense sector.
Both the UK RAB and the Canadian PAA are necessary steps towards con-
necting outcomes and resources but they require modifications and refinements.
Both describe processes that tend to equate inputs to outputs. “Administration”
in RAB and “internal services” in PAA are catch-all activities that include
fixed costs and outputs (see Appendix 3.1). For example, government-wide
48 K. Hartley, B. Solomon
Table 3.3 MoD resources by budget area FY 2010–11

Top level budget holders FY 2010–11


current
£ milliona

Departmental expenditure limits (DEL) 48,463


Request for resources 1: Provision of defence capability 44,516
Resource DEL 36,221
Commander-in-Chief Navy Command 2,294
General Officer Commanding (Northern Ireland) *
Commander-in-Chief Land Forces 7,189
Air Officer Commanding-in-Chief RAF Strike Command *
Commander-in-Chief Air Command 2,826
Chief of Joint Operations 480
Defence Equipment & Support 16,869
Central 2,401
Defence Estates 2,136
Administration 2,026
Capital DEL 8,295
Request for resources 2: Conflict prevention 3,946
Resource DEL 2,862
Capital DEL 1,084
Annually managed expenditure (AME) 7,881
Request for resources 1: Provision of defence capability 6,918
Request for resources 2: Conflict prevention 28
Request for resources 3: War pensions & allowances, etc 935

Source: Defence Analytical Services and Advice (DASA) 2011.


Notes
a Current £million, 2011.
* No value or not reported.

initiatives on accountability and environmental stewardship, while important,


do not link directly to security outputs. To make both RAB and PAA more
relevant for the assessment of MoD’s and DND’s efficiency and effective man-
agement of resources, the re-design must be linked to a strategic planning
framework and costing tool that displays the total ownership cost resources
(Solomon et al. 2008).

3.4.5 Challenges
Two growing pressures make it essential to focus on the size of a nation’s defense
budget and the efficiency with which defense resources are used. The first pres-
sure is to reduce defense budgets and reallocate resources to other public spending
programmes, especially education, health, and welfare (including care for the
increasing elderly populations). Second, added pressure on defense budgets comes
from the increasing costs of defense equipment. A simple example shows the
Measuring defense output 49
importance of rising unit equipment costs which affects all nations (all figures
are for unit production costs in 2010 prices):11

• Spitfire unit costs (1940): £154,850


• Typhoon unit costs (2010): £73.2 million
• Typhoon replacement in 2050: £1+ billion

Increasing unit costs and constant or falling defense budgets (in real terms)
suggest difficult defense choices cannot be avoided. The challenge in divestment
discussions as in investment decisions is to establish useful measures of defense
outputs, or in the absence of such measures, to develop proxies to enable CBA.
This chapter, having identified multiple issues and obstacles involved in
developing defense output measures; next reviews the experience of several
nations. Many chapters in this book offer valuable techniques to develop proxy
measures of intermediate outputs at lower levels of the organization in support of
military CBA.

3.5 Defining defense outputs: the benefits of defense


In principle, defense provides an output in the form of goods and services which
offers a stream of current and future benefits to a nation’s citizens and to the citi-
zens of other nations who might also receive such benefits. The benefits are both
economic and non-economic. The economic benefits of defense usually take the
form of services which contribute to national output. The non-economic benefits
of defense include foreign policy benefits, peacekeeping, and its contribution to
a nation’s ‘feel good’ factor. This includes its involvement in being a responsible
international citizen and valued member of the international community.
Economists rarely address the concept of overall defense output apart from
vague references to security. Government statisticians and the national accounts
have traditionally measured defense output on the convention that output equals
input (Office for National Statistics (ONS) 2008). Improving this limited measure
requires that the concept of defense output be further developed and explored.

3.5.1 Security
In principle, defense provides security, which itself is a multi-product output
embracing protection, safety, insurance, peace, economic stability, and risk avoid-
ance or reduction (Solomon et al. 2008). Further dimensions include prosperity,
individual and national freedoms, liberty, and a ‘way of life.’ These are all dif-
ficult to measure and might be influenced by factors other than defense. Also,
these aspects of security are public goods which are not marketed and include
non-marketable services involving no tangible and physical products.
Security is sometimes defined as the absence of threats or risks (Baldwin 1997;
Engerer 2011). A world of no threats or risks, however, does not and cannot
exist: the real world is characterized by continually emerging threats and risks.
50 K. Hartley, B. Solomon
Questions then arise about which threats and risks can be reduced, by whom, and
at what cost.
Recent developments have led to security referring to issues other than military
security (creating fuzzy boundaries). Individuals are faced with threats to their
lives, health, property, other assets, and their prosperity (e.g. from criminals and
terrorists, disease/pandemics and ill health, natural or man-made disasters, and
economic recessions). Threats to individuals augment the threats to nation states
(e.g. military threats from other nations and environmental problems originating
from other nations) which raises questions about which threats should be handled
privately and which publicly.
Where threats are handled publicly, which is the most appropriate and least-
cost solution? For example, military solutions are appropriate for external military
threats whilst internal police forces are more appropriate for internal threats
from criminals (e.g. physical violence to individuals involving injury, death,
and robbery). Threats to an individual’s state of health require dietary, medi-
cal, and care solutions (e.g. from doctors, nurses, and care homes). Threats to
prosperity require government macroeconomic and microeconomic policies to
promote full employment and economic growth (e.g. opportunities for educa-
tion, training, and labor mobility—although some of these activities can be funded
privately).
Technical progress and changing consumer preferences have resulted in shifts
from the public provision of security to private protection measures provided
and financed by individuals (e.g. private security guarding, camera surveillance
of property, and creation of neighborhood watch schemes providing local club
goods). Security also has important geographical dimensions.
For example, defense can be viewed as a means of protecting a nation’s property
rights over its land, sea, and air space. A nation’s defense forces, however, might
also be used to protect other nations’ citizens so that the public good becomes inter-
national which further increases the problem of obtaining and financing the optimal
amount of the international public good (including peace). Overall, security mea-
sures can be analyzed as national or international public goods, club goods, and
private goods—each with different solutions and each embracing different indus-
tries (e.g. security and defense industries: Engerer 2011). These different industries
have different customers, products, and technologies (Sempre 2011).

3.5.2 The economic benefits of defense


Defense contributes to individual and collective security and protection, both of
which are valuable commodities. It protects households and their assets, firms and
their assets, the national infrastructure, national institutions, and national freedoms
(e.g. democracy, freedom of speech and movement, etc). It also protects national
interests, including independence and ‘appropriate sovereignty’ (e.g. protecting
a nation’s interests in a globalized world, including leverage and status in world
politics and diplomacy). How can these commodities be valued? There are at least
three approaches.
Measuring defense output 51
First, estimate a nation’s per capita defense spending and then ask whether
its citizens are willing to pay at least such a sum for the annual protection
offered by its armed forces. Comparisons can be made with other public spending
programmes, such as health and police forces. Second, value-of-life studies can
be used to estimate the value of lives saved and injuries avoided resulting from the
provision of armed forces (see Chapter 9, and Jones-Lee 1990).12
Health economists have developed measures of health output benefit measures
in the form of quality-adjusted life years or QALYS (but these are not valued). The
defense equivalent of QALYS would be protection-adjusted life years (PALYS:
Hartley 2010b). In addition to estimating the benefit of lives saved by defense,
there are further gains from valuing property saved by avoiding damage and
destruction (i.e. estimating both human and physical capital saved).
Third, consider defense as insurance in response to various current and future
known and unknown threats and contingencies. These contingencies can involve
time periods of some 25–50 years into the future: the result lags in the relationship
between inputs and defense outputs, meaning that defense productivity cannot
be based on the standard relationship between inputs and outputs within a cal-
endar year. The insurance approach has private market comparators. Individuals
and firms pay for a variety of insurance policies and other forms of protection.
Examples include households buying insurance for homes, motor cars, driving,
healthcare, international travel, and retirement.
In addition, households buy further protection in the form of household security
(e.g. alarms and guard dogs), purchasing safer motor cars, locating in a safe neigh-
borhood, and joining neighborhood watch schemes (through payments-in-kind).
Similarly, firms make various insurance payments to protect their assets: they
employ security guards and introduce measures to protect their staff and assets
from terrorist attacks. Admittedly, these are private rather than public goods but,
nonetheless, the payments in cash and kind provide some indication of the willing-
ness of households and firms to pay for protection; such willingness to pay might
then be applied to estimating the minimum level of a nation’s defense spending.
Further spending for protection is reflected in expenditure on a nation’s police
forces and internal security. The result is a substantial expenditure for private and
public spending on internal security. Again, such sums provide an estimate of the
lower bound of national defense spending.
By providing security and protection for a nation’s citizens, defense spend-
ing on the armed forces creates the conditions allowing and promoting beneficial
voluntary trade and exchange within and between nations. Protection of national
property rights over land, sea, and air space (and more recently cyber-space) pro-
motes national market exchanges whilst protection of international trade routes
promotes beneficial international trade and exchange. For example, a nation’s
navy protects its international shipping and trade routes, including protection from
piracy.
National and international market exchange contributes to improve society’s
welfare (especially compared to societies that lack well-developed national
markets). In the context of national markets, the armed forces provide a capability
52 K. Hartley, B. Solomon
to respond to national emergencies and provide aid to the civilian community.
Without the armed forces, civil powers would have to provide more resources for
emergencies (a relatively high cost for capabilities only required infrequently), or
ignore such contingencies.
Ideally, defense spending on a nation’s armed forces helps to prevent and avoid
conflict; and where conflict occurs, minimizes its duration and effects on citi-
zens, as well as contributing to more rapid post-conflict recovery. In this context,
defense provides a deterrent aiming to persuade potential adversaries that con-
flict is not worthwhile. Where deterrence fails, defense spending aims to provide
a warfighting capability to achieve a ‘successful’ conclusion by minimizing the
costs of conflict.
These features are economic benefits that can be reflected in cost savings from
avoiding conflict or minimizing its duration, and contributing to post-conflict
recovery and restoration of market activity. Again, it is difficult to measure cost
savings for events which do not occur. Indeed, such problems raise the general
methodological issue of the counter-factual: what would have happened in the
absence of defense spending?
Defense spending as a form of fiscal policy provides some direct national eco-
nomic benefits comprising jobs, new technology, spin-offs, exports, and import-
savings. The armed forces are a source of employment, and their spending in local
areas further increases employment and provides a source of trained and skilled
labor for the rest of the economy. In addition, public spending on the nation’s
defense industrial base contributes to jobs, advancing technology, spin-offs, and
the balance of payments.
As discussed earlier, however, these economic benefits need to be assessed
critically: there are serious doubts about many of these claimed economic ben-
efits. For economists, a major concern arises over the alternative-use value of the
resources employed in the armed forces and national defense industries. It needs
to be asked whether the resources used in the military-industrial complex would
make a greater contribution to jobs, technology, spin-offs, and exports if these
resources were used elsewhere in the economy (Hartley 2010b).
Bilateral military alliances might provide additional economic benefits. For
example, the US–UK special relationship provides the UK with access to US
technology for its nuclear-powered submarines and missiles for its nuclear deter-
rent; with a leading role in the F-35 programme; and enables major UK defense
firms’ (e.g. BAE and Rolls-Royce) access to the US defense market. Also, the
US provides security and protection for the UK which might otherwise require a
larger defense budget. Similarly, Canada benefits from US defense spending and
protection, leading to lower Canadian defense spending.

3.5.3 Non-economic benefits of defense


Defense spending also contributes major non-economic benefits to a nation and
it might be that non-economic benefits are more valuable than many economic
benefits. Non-economic benefits are those which do not explicitly contribute
Measuring defense output 53
to national output. They comprise political, military-strategic, and international
benefits.
These non-economic benefits include the ability to pursue national interests and
foreign policy objectives; to add to a country’s international reputation, standing,
and status in the world (the feel good factor); and to increase its position in
the world power hierarchy. These non-economic benefits might be reflected in a
nation’s position in the United Nations (e.g. membership of the Security Council),
its membership in world economic organizations (e.g. OECD, IMF; G-8, and G-20
groups of nations), its leadership positions in international military alliances (e.g.
NATO), and its ability to influence the behaviour of other nations. For example,
military-strategic benefits can arise from bilateral or multilateral military alliances
(e.g. benefits from standardization of equipment and tactics, some of which are
economic in the form of cost savings).
There are other ways in which a nation can obtain non-economic benefits.

• Prestige and international reputation. A nation may provide military forces


for international peacekeeping and peace enforcement, leading to world
peace. Such peacekeeping contributions, however, are not costless.
• Feel good factor. Further non-economic benefits arise when a nation’s armed
forces contribute to international efforts on humanitarian aid and disaster
relief. These contributions provide a ‘feel good’ factor (national pride) for
the contributing nation’s citizens (e.g. national spending on child protection
and social services: Hartley 2010b).
• Compulsory club membership. Some nations may contribute to peace mis-
sions because they value (or feel compelled by) a relationship with a dominant
ally or neighbor. This type of alliance or relationship can be considered as a
compulsory club membership with clear obligations and benefits (Berkok and
Solomon 2011).

3.6 The evidence: the international experience with measuring


defense output
3.6.1 UK experience
Before 1998, the UK published traditional input and intermediate measures of
defense output. Typically, these comprised numbers of armed forces military per-
sonnel and their formations, including numbers of aircraft squadrons, infantry
regiments, tank units, and warships. The published data on unit numbers were
available in varying degrees of detail (e.g. aircraft squadrons by types of aircraft,
types of warships, etc). The amount of published data and its detail have improved
over time. Data were also published on the numbers of regular and reserve forces
and the numbers of civilian personnel employed by the MoD.
During the cold war, the armed forces focused on preparing for and deterring
a direct military attack on the UK or Western Europe. After the cold war, there
was no longer a direct military threat to the UK. In 1998, the publication of the
54 K. Hartley, B. Solomon
Strategic Defence Review (SDR) marked a significant change in published defense
output measures.
The 1998 SDR represented a pioneering contribution to UK published data on
defense output measures. For the first time, the UK published data on its defense
capabilities, which are a more meaningful indicator of defense output. These
defense capabilities are viewed as planning commitments. The 1998 SDR com-
mitted the UK to be a ‘force for good’ in the world with an associated global
military expeditionary capability. On this basis, the UK armed forces were capa-
ble of supporting continuing commitments (e.g. Northern Ireland at that time)
able to:

• respond to a major international crisis of a similar scale and duration to


the Gulf War (an armoured division, 26 warships, and over 80 combat
aircraft); or,
• undertake a more extended overseas deployment on a lesser scale (e.g.
Bosnia) while retaining the ability to mount a second substantial deploy-
ment if this were made necessary by a second crisis (e.g. combat brigade
and supporting air and naval units).13

These defense capabilities were subject to various constraints of readiness, loca-


tion, duration, and concurrency. Different levels of readiness involve different cost
levels: maintaining continual high readiness is costly. Similarly, location and dura-
tion affect force requirements: regional conflicts outside the NATO area and for an
indefinite duration require different sizes and structures of armed forces compared
with short-term deployments to, say, Bosnia or Kosovo.
Concurrency is a further issue involving the number of operations which can be
conducted at any time including their scale, location, and duration. SDR iden-
tified the core regions of Europe, the Gulf, and the Mediterranean. The UK
was committed to conducting two medium-scale operations concurrently (SDR
1998). Surprisingly, the government’s actual military commitments exceeded
these planning assumptions!
Following the 9/11 terrorist attacks on the US, the 1998 SDR was modified.
A modified policy was announced in 2003/04 (comprising Cmnd. 6041 2003
and Cmnd. 6269 2004). The revised policy adopted new planning assumptions
including:

• the ability to support three simultaneous small to medium-scale operations


where at least one is an enduring peacekeeping mission (e.g. Kosovo),14 or
• the ability at longer notice to deploy forces for large-scale operations while
running a concurrent small scale peace-support operation, or
• the ability to project military force to sub-Saharan Africa and South Asia as
well as a capability to respond to international terrorism; and
• the most demanding operations will be conducted as part of a coalition, usu-
ally involving the United States This requires the UK’s armed forces to be
interoperable with US Forces.15
Measuring defense output 55
Further changes occurred with the Strategic Defence and Security Review
of 2010 (Cmnd. 7498 2010). Following the 2010 planned budget cuts, the
UK’s Defence Planning Assumptions (DPAs) and its defense capabilities were
reduced to:

• an enduring stabilization operation around brigade level (up to 6,500 person-


nel) with air and naval support; and
• one non-enduring complex intervention (up to 2,000 personnel); and
• one non-enduring simple intervention (up to 1,000 personnel); or
• alternatively, three non-enduring operations if not already engaged in an
enduring operation; or
• for a limited time, and with sufficient warning, committing all the UK’s effort
to a one-off intervention with up to three brigades with air and naval support
(about 30,000 personnel);16 together with
• maintaining a ‘residual defense capability’ for unforeseen emergencies or to
reinforce existing operations or to respond to scenarios where the UK acts
alone (House of Commons (HCP) 992 2011, 20).

MoD budgets pay for the UK force elements to be ready for operations as out-
lined in the DPAs. The costs of these missions, however, are funded from the
government’s contingency reserves. Over time, the rising unit costs of defense
equipment and of volunteer military personnel will result in smaller armed forces
and reduced defense capabilities (as defined by the UK MoD).
More important would be an assessment of the costs of achieving these defense
capabilities compared with other nations providing similar capabilities (i.e. is the
UK providing its capabilities at least-cost?). Within MoD, measures of defense
training activities are used to assess performance. These include flying hours, days
spent at sea, and Army personnel data on gains to trained strength and data on
military exercises (ONS 2008).
The MoD publishes an annual performance report which offers some further
insight into its defense capabilities (HCP 992 2011).17 Useful though such infor-
mation might be, it is both qualitative and vague (i.e. “success” in Afghanistan)
and focuses on input costs which are unhelpful data for measuring output by
themselves. On force readiness, the MoD’s Performance Report admits that “Mea-
suring and aggregating readiness is complex, not least because it is based on
judgements of what is required to enable the armed forces to respond to a wide
range of potential challenges” (HCP 992 2011, 21).
The MoD also reports on where there are “critical and serious weaknesses” in
UK Forces.18 The MoD’s Performance Report also included a section on imple-
menting the 2010 Strategic Defence and Security Review which provided further
information on the UK’s defense capabilities.19
The UK’s defense capabilities output measures are an improvement on the tra-
ditional input approach but there are deficiencies at least in terms of publicly
available information. For example, the National Audit Office has reported that the
UK MoD has a good system for defining, measuring, and reporting the readiness
56 K. Hartley, B. Solomon
of its armed forces which compares well with other countries (e.g. Australia,
Denmark, and US: NAO 2005). It is recognized that 100 percent readiness is
too costly. The published data on readiness, however, refers to whether there are
serious or major weaknesses, which is useful but not very illuminating (e.g. with-
out knowing what and where such weaknesses arise and their impact on force
effectiveness). For instance, a statement that 50 percent of UK Forces had no seri-
ous or critical weaknesses suggests that the remaining 50 percent suffered serious
weaknesses, which should be a source of concern!
Moreover, these performance assessments are mostly undertaken by MoD
personnel, which could raise questions of independence and objectivity. An
NAO report on the performance of MoD in 2009–10 presented and reviewed
performance indicators (NAO 2010). The report focused on financial manage-
ment information (e.g. management of stocks and assets) and made no men-
tion of defense output measures. Though there was mention of defense output
indicators, these only included qualitative measures: ‘success on operations;’
the existence of serious and critical weaknesses in readiness; manning levels
in relation to manning balance by Service (with no data); and flying hours
achieved against targets (again, without any data). In relation to the MoD aim
of global and regional reductions in conflict, no output measure was reported
(NAO 2010).
The NAO also publishes value-for-money reports (a type of retrospective
CBA)—for example, on the multi-role tanker aircraft capability—and annual
reports on MoD’s major projects. These project reports assess major defense
projects against their contractual commitments on cost, delivery, and performance,
usually identifying cost overruns, delays, and any failures to meet performance
requirements. Such value-for-money reports are a useful addition to knowl-
edge but do not include the wider industrial and economic benefits of major
projects, nor do they provide any assessment of the “battle-winning” perfor-
mance of defense equipment (e.g. as demonstrated in conflicts such as Afghanistan
or Iraq).
Overall, the UK’s defense capabilities are useful measures of defense output but
deficiencies remain. Some indicators of force readiness are qualitative: readiness
is a variable measure depending on circumstances (readiness for what, when and
where?); capabilities are not all identified; benefit values are not attached to capa-
bilities; and the capabilities cannot be aggregated into a single measure of overall
defense output.
None of the output measures address the contribution of defense to conflict pre-
vention and its contribution to minimizing the costs of conflict, including saving
lives. In fact, MoD economists examined different approaches to capturing output
used in various parts of the MoD. As Davies et al. (2011, 399) put it:

These include a number of partial aggregations and a balanced scorecard


approach covering the three main areas of activity: success in military tasks,
readiness to respond, and preparing for the future. . . and. . . it was confirmed
that no existing technique offered a solution. Although it is hoped that in
Measuring defense output 57
the longer term progress will be made on the direct measurement of defense
outputs and productivity, this remains an elusive goal.

The UK’s experience in measuring outputs in other parts of the public sector
and in the private sector is given in Appendix 3.2, offering selected insights for
defense.

3.6.2 Australian experience


The Defence White Paper of 2009 outlined Australia’s defense policy and force
structure to 2030 (Department of Defence (DoD) 2009). It specified Australia’s
strategic interests including (ranked in order of priority):

• the defense of Australia against armed attack with the capability to act inde-
pendently so as not to be reliant on foreign military forces. This principal
task requires the Australian Defence Force (ADF) to control the air and sea
approaches to Australia;
• the security, stability, and cohesion of Australia’s immediate neighborhood
which is shared with Indonesia, Papua New Guinea, East Timor, New
Zealand, and the South Pacific Island states;
• an enduring strategic interest in the stability of the wider Asia-Pacific region;
and
• a strategic interest in preserving the world international order which restrains
aggression, manages other risks and threats, and addresses the security
impacts of climate change and resource scarcity.

These objectives are to be achieved by Australia acting independently,


by leading military coalitions, and by making tailored contributions to mil-
itary coalitions. As a result of these priorities, the ADF of 2030 will need
to invest especially in its maritime capabilities as well as enhancing its air
capabilities.
Part of the funding for these capability improvements is to be achieved through
efficiencies and savings (of AUD$20 billion) which, it is claimed, will not compro-
mise effectiveness (DoD 2009, 14). Also, to fund ADF of 2030, the government
has committed to real growth in the defense budget of 3 percent to 2017–18 and
then 2.2 percent real growth to 2030 (DoD 2009, 137). The 2009 White Paper rec-
ognizes that defense planning is about managing strategic risks, that uncertainties
remain, and that it is not possible to eliminate all risks (an ideal warning time of
10 years is reported: DoD 2009, Chapter 3, 28).
The 2009 White Paper deals with preparedness embracing readiness, sustain-
ability, and concurrency. It recognizes that preparedness comes at a cost (but
provides no data on the marginal benefits or costs of different levels of prepared-
ness). Sustainability refers to the ability to undertake tasks and operations over
time, whilst concurrency deals with the ability to conduct a number of operations
58 K. Hartley, B. Solomon
in separate locations simultaneously. The White Paper provided an extensive list
of the required capabilities of the ADF including:

• the capabilities needed for sea and air control around Australia;
• the ability to deploy a brigade group for combat operations for a prolonged
period of time in the primary operational environment (for shorter period
beyond that area);
• the ability to deploy a battalion group to a different area of operations in the
primary operational environment;
• the ability to maintain other forces in reserve for short-notice, limited warning
missions;
• the ability to provide tailored contributions to operations in support of
Australia’s wider strategic interests (e.g. special task forces group); and
• the ability to provide assistance to civil authorities (e.g. fisheries protection,
terrorist incidents, support for major events, emergency responses, humani-
tarian and disaster relief in Australia and its neighbors, provision of search
and rescue support, etc).

The list of capabilities is extensive with no ranking and little indication of


the military resources available for each capability. Some of the capabilities are
clearly military; others, including aid to civil authorities, are a general ‘catch-all’
which might be used to justify public support for defense spending. Further data
on capabilities is provided by the annual defense budget.
Australian defense budgets show published data on expenditure on various
overseas operations—the sources of planned cost savings, and capital investment
programmes. There is data on the extra costs of overseas operations and on the
numbers of military personnel by service (permanent, reserves, and numbers of
high readiness reserves). Further budget data are presented on planned perfor-
mance and outcomes for each of three defense outcomes comprising the protection
and advancement of Australia’s national and strategic interests and support for
the Australian community and civil authorities (including expenditure by military
base area). Some limited performance (intermediate output) indicators are pub-
lished, such as the number of unit ready days for the Navy and flying hours for
each of the services (DoD 2011).
A review of defense accountability was published in 2011 with the aim of
improving accountability across defense (Black 2011). The review recommended
the introduction of specific, measurable, and achievable outcomes with individuals
given ownership and made accountable for their outcomes. The review recognized
that there was a lack of specific outcome-based language in defense and an insuf-
ficient use of measurable outcomes. Particular focus was placed on performance
measures for shortfalls in equipment delivery to date, including costs and qual-
ity (e.g. average delays of 28 percent or 2+ years; cost overruns of 52 percent:
Black 2011, 60–61). The review, however, focused on management-organizational
issues (e.g. too many committees) and not on the development of defense output
measures and their consequences.
Measuring defense output 59
In June 2011, the Australian Minister for Defence announced a Defence Force
Posture Review designed to assess whether the ADF is correctly positioned
geographically to meet Australia’s modern and future strategic and security
challenges. These include:

• the rise of the Asia-Pacific and the Indian Ocean rim as regions of global
strategic significance;
• the growth of military power projection capabilities of the Asia-Pacific
countries;
• the growing need for the provision of humanitarian assistance and disaster
relief following extreme events in the Asia-Pacific; and
• energy security and security issues associated with expanding offshore
resource exploitation in Australia’s North West and Northern approaches.

The ADF Force Posture Review considers how the ADF will support Australia’s
ability to respond to a range of activities including deployments on overseas
missions and operations; support of operations in Australia’s wider region; and
engagement with the countries of the Asia-Pacific and Indian Ocean rim in ways
which will help to shape security and strategic circumstances in Australia’s inter-
est. The Force Posture Review also makes recommendations on basing options for
Force 2030. There is also a Submarine Sustainment Review to assess sustainment
of Australia’s Collins Class submarines (ADF 2011).
The Australian Defence White Paper of 2013 identified the capabilities the
ADF will need in the future, reflecting the withdrawal from Afghanistan and the
Solomon Islands. It emphasized air, naval,20 special-forces, intelligence, and cyber
security. The commitment to spending 2 percent of GDP on defense became a
long-run target.
The 2013 White Paper presented a broad description of future forces but pro-
vided little detail on their capabilities. Interestingly, the 2013 White Paper referred
to productivity and the need for “defence to become more efficient and prudent
in its use of resources to remove waste and achieve better economies of scale.”
This is a politically attractive phrase, but does not explicitly address the chal-
lenge of preserving an adequate level of overall defense output and corresponding
measurement challenges (DoD 2013).

3.6.3 New Zealand experience


New Zealand has a considerably smaller defense effort compared with the UK
(see Table 3.1; Hartley 2010b). Nonetheless, it has devoted substantial resources to
measuring its defense output. This section describes and assesses the development
of its output indicators as published in 1991, 1993, 2011, and 2013.
In 1991, the New Zealand Defence Force (NZDF) published defense output mea-
sures in its Annual Plan (NZDF 1991). At this time, the output of the NZDF was
grouped into two main categories: namely, retained outputs and current outputs.
60 K. Hartley, B. Solomon
Retained outputs are military groupings of operational forces which are retained
to provide the government with a basis of military power from which force may
be applied. Current outputs reflect the range of current activities undertaken by
the NZDF which reinforce foreign policy goals and contribute to the well-being of
the nation. Current outputs were further divided into core (military activities which
contribute to military outcomes) and non-core (services provided to the community).
Published intermediate output data were provided for the two final outputs. For
example, retained outputs consisted of eleven intermediate outputs: namely, naval
combat forces; mine countermeasure forces; naval control and protection of ship-
ping organizations; strategic assets (force troops); ready reaction forces; infantry
brigade group and force maintenance; long-range maritime patrol force; offen-
sive air support force; long and medium-range air transport force; medium and
short-range air transport force; and the utility helicopter force.
Each intermediate output included performance targets and performance
achievements. For example, the performance target for the infantry brigade group
required deployment for operations within 90 days, and the performance achieved
was for such a force to be available for sustained low-level operations at 90-days’
notice. Offensive air support required 3,760 flying hours by Skyhawks, but there
was a shortfall of over 400 hours against this target.
Changes were made and announced in 1993 (NZDF 1993). Seven output classes
were identified: protection of New Zealand’s territorial integrity and sovereignty;
provision of military advice; provision of intelligence; provision of ancillary ser-
vices; contribution to regional security; mechanisms for participation in defense
alliances; and contributions to collective security.
Each output class was divided into sub-groups, each with performance targets
and achievements. For example, the sub-group of “countering terrorism” had a per-
formance target of two counter-terrorist exercises but only one such exercise was
conducted. Similarly, for the sub-group called “deterring intrusions,” there was a per-
formance target of sustaining a naval presence for up to 30 days in the New Zealand
area, and it was reported that this capability was demonstrated and achieved.
Whilst an impressive amount of detail was published, there are serious defi-
ciencies with the outputs reported and performance indicators used. First, the
outputs and performance indicators reported are mainly inputs or intermediate
measures of output. Second, several outputs are unusual for defense outputs:
namely, the provision of advice, intelligence, and ancillary services which includes
civil defense assistance, support services to the community, and ceremonial sup-
port for the state. Third, the published data provide no weighting to indicate the
relative importance of the various intermediate defense outputs to overall secu-
rity. Is the provision of advice and intelligence ranked as highly as protection of
New Zealand’s territorial integrity and sovereignty? Fourth, some defense outputs
might more appropriately be the responsibility of other government departments.
The defense outputs were refined and developed over time and with experience.
The New Zealand position in 2011 is reflected in the NZDF Statement of Intent
(NZDF 2011) which outlined the country’s defense policy over the next 25 years.
It specified the primary mission of the NZDF as securing New Zealand against
Measuring defense output 61
external threat, protection of its sovereign interests, and the ability to take action
to meet likely contingencies in the country’s strategic area of interest. This primary
mission recognizes that the country’s national interests affect both the security and
prosperity of the nation.
New Zealand must trade to survive, which requires that New Zealand has unfet-
tered access throughout the Asia-Pacific region to engage in business. “Instability,
conflict and war, even far from New Zealand’s shores, can therefore directly affect
New Zealand’s social and economic well-being” (NZDF 2011, 9). Recognizing
that the primary mission of the NZDF is broad, a number of subsidiary or
intermediate outcomes have been developed.
The NZDF main and intermediate outcomes are currently not linked to a for-
mal set of measures, mainly due to the complexity of measuring outcomes which
deliver security and protection: there is no single measure of success in delivering
protection. “There is no definitive way of knowing what might have happened, but
did not happen, because of the activities of the NZDF” (NZDF 2011, 34).
The NZDF has 37 outputs in 16 output/expenses classes. Its output/expenses
classes include naval combat and support forces, mine countermeasures, land com-
bat and support forces, naval helicopter forces, airborne surveillance, and fixed
wing and rotary transport forces.
Other output categories include unusual components such as military hydrogra-
phy, military advice, and multi-class output appropriations (e.g. support to youth
development and support to military museums: NZDF 2011). The NZDF also
stresses its links with the community, reflected in the provision of skills to society,
promotion of a ‘healthy’ defense industry, and a “buy New Zealand” procurement
policy (NZDF 2011, 11). It is now explained, however, that these links to the
community arise as by-products of the NZDF (NZDF 2011, 11).
The NZDF uses a measure of military capability which shows the combined
effect that inputs have on operational effectiveness. Military capability is assessed
using two elements: namely, preparedness and force components, described by the
acronym PRICIE.21 The elements are comprised of personnel, R&D, infrastruc-
ture, concepts of operations and training, information/technology, and equipment
and logistics (NZDF 2011, 48).
The NZDF recognizes that its output measures often appear as inputs rather than
outputs. Inputs are used as proxies for military capabilities (e.g. 500 flying hours of
a specific type of aircraft will provide a certain military capability), but the actual
measurement systems and capabilities are classified. Following the New Zealand
Defence Review, however, concerns were expressed that the current system is too
input-focused, amid a desire to measure military impacts and outcomes and cross-
sector security outcomes. Where complex relationships are involved, it might not
be possible to easily identify and measure cause and effect (Comptroller and
Auditor General (CAG) 2011). No valuations of output are provided.
These concerns continued in the NZDF Annual Report of 2013 which reported
its aggregate level of preparedness as “substantially prepared” (asserted with-
out supporting data). The 2013 report did publish some output measurements,
including objectives, measures, and outcomes (NZDF 2013). Overall, the NZDF
62 K. Hartley, B. Solomon
has made commendable efforts to recognize and address the challenge of measur-
ing defense output.

3.6.4 European experience


The focus here is on major European defense spending nations that publish their
data in English. These nations are France, Germany, Italy, Spain, and Sweden.

3.6.4.1 France
A new French defense policy was announced in 2008 with the aim of making
the French armed forces more flexible for rapid deployment from the Atlantic to
the Indian Ocean. France aims to provide the necessary resources to ensure the
security of its citizens, to safeguard national independence, and to consolidate the
nation’s military and diplomatic power. Under the new policy, France will be able
to project 30,000 personnel with 70 combat aircraft, one carrier group and two
naval battle groups within a six-month period for up to a year (a force capable
of dealing with one major war or crisis at a time). Nuclear deterrence remains
a key military mission, but terrorism is the most immediate threat, and there are
also public service missions. There will be reductions in the numbers of military
personnel and investment in new equipment. Some equipment is of poor quality:
for example, only 50 percent of Leclerc tanks are mission ready; refuelling aircraft
are 45 years old; and some Puma helicopters are 30 years old (Penketh 2008).
In 2010, the UK and France signed an Anglo-French Defence Treaty with the
potential for greater bilateral co-operation between their armed forces and defense
industries.
By 2014, French defense policy was adjusting to a “difficult financial situa-
tion” leading to further reductions in the numbers of military personnel, ships,
tanks, helicopters, and transport aircraft. A Joint Reaction Force was created, with
2,300 personnel and the ability to deploy up to 15,000 troops in a single over-
seas deployment. Plans were in place to cut the defense budget to 1.5 percent
of GDP.

3.6.4.2 Germany
NATO remains the centrepiece of Germany’s defense policy. The new defense
policy announced in 2011 involved major changes for Germany’s armed forces:
reductions in the defense budget; abolishment of conscription replaced by an
all-volunteer force; improvement of Germany’s expeditionary capabilities; and
closer military co-operation in Europe, especially in procurement and training
(German Ministry of Defense (GMOD) 2011). Under the new policy, Germany
plans to increase the deployment of the Bundeswehr outside Germany from the
current 7,000 to some 10,000 soldiers (but there is no statement of the extent
of geographical coverage of these expeditionary forces). There are also plans
to reduce quantities of equipment (aircraft, helicopters, and ships). In 2014,
however, Germany announced a review of its commitment to overseas military
Measuring defense output 63
deployments, including its military contribution to the international pooling and
sharing of military resources.

3.6.4.3 Italy
Despite cuts in defense spending due to Italy’s austerity programme, Italy retains
an expeditionary capability. Reports suggest the Air Force has been particularly
affected by defense cuts. There were also reports that in 2010, Italy planned to
reduce its involvement in peacekeeping missions in the Balkans and possibly in
Lebanon, concentrating instead on Afghanistan where force levels peaked at 4,000
soldiers (Nativi 2010).
Italy’s Military Policy Review of 2013 aimed to reduce the numbers of military
personnel to 150,000 and civilian personnel to 20,000 by 2024. This represents a
lengthy adjustment period. Italy also identified reductions in the numbers of senior
military personnel within the reduced total numbers (i.e. fewer admirals, generals,
colonels, and naval captains). The plan is for a defense budget of 0.84 percent of GDP.
The 2013 Military Policy Review declared that “available resources will be used
on developing systems that combine high operational efficiency, adequate cost-
effectiveness, and a development/growth margin that allows them to be integrated
into complex and net-centric systems.”

3.6.4.4 Spain
Reductions in defense spending were part of Spain’s austerity programme. The
2011 budget reflected four objectives: the safety of the troops (force protection
via operating and logistics expenditure); operational readiness; the maintenance
of weapons systems; and international operations and fulfilment of Spain’s inter-
national commitments. New tools were announced for improved oversight and
management of defense expenditures.

3.6.4.5 Sweden
A new defense policy was announced in 2009 with an emphasis on mobility and
flexibility of Sweden’s armed forces. The plan is for it to be possible for an entire
operational organization of some 50,000 people to be used within one week after
a decision on heightened alert. In contrast, today only one-third of the national
operational organization is equipped and prepared for an operation within one
year. Some defense capabilities were listed in terms of inputs: numbers of military
personnel (e.g. deployment of 1,700 people for continuous international peace-
support operations) and numbers of Gripen aircraft (100 of the C/D model). An
all-volunteer force will replace compulsory military service and there will be sub-
stantial reserve forces (e.g. four mechanized battalions). For the Army, only a
small proportion of its soldiers will be full-time. Sweden specified its area of
national interest: namely, the Baltic Sea or the northern area (Sweden Ministry
of Defence (SMOD) 2009).
64 K. Hartley, B. Solomon
3.6.5 The European Defence Agency
The European Defence Agency (EDA) publishes defense data for its member
states. These include annual financial data such as levels of defense spending and
shares of GDP for member states, equipment procurement and R&D expenditures,
spending on infrastructure and construction, defense expenditure that is out-
sourced, and expenditures on collaborative equipment programs. Input data is
also published that include numbers of military, civilian and internal security
personnel, personnel expenditures, as well as data on numbers of different types
of equipment (combat aircraft tanks and warships).
Most of the data published is for inputs rather than defense outputs, although
some EDA officials regard indicators such as quantity of military personnel as
intermediate output measures. There is, however, data which are clearly a mea-
sure of intermediate output and proxies for defense output: namely, operation and
maintenance expenditure, operational costs, average numbers of troops deployed,
and the average numbers of sustainable (land) forces (EDA 2011). Comparative
analysis of such data could be useful in revealing variations in efficiency and
effectiveness across member states.

3.6.6 The United States


The United States is unique in its global power commitments and large-scale
defense spending. US national security strategy requires a “comprehensive global
engagement aimed at supporting a just and sustainable international order” (Under
Secretary of Defense (USDOD) 2011, 2–1). The United States remains the only
nation able to project and sustain large-scale military operations over extended
distances. Its main objectives are to prevail in today’s wars, prevent and deter con-
flicts, prepare for a wide range of contingencies, and preserve and enhance the
all-volunteer force (AVF). Three of these objectives refer to actual and potential
threats or outputs/outcomes, but the commitment to the AVF is an input, not an
output! Funding for these objectives is to come partly from efficiency savings,
including cancelation of unwanted and poorly performing equipment programs.
The US DoD publishes a massive amount of data of varying degrees of use-
fulness in terms of measuring defense outputs. For example, its aims include
sustaining military capabilities to fight two wars, confront global terrorism, and
provide humanitarian assistance and disaster relief, but no measures or valuations
are attached to any of these objectives.
It also presents extensive data on performance results relating to its primary
warfighting goals and its supporting goals (e.g. preserving the AVF and imple-
menting the defense agenda). It is claimed that 75 percent of DoD performance
goals were met in 2010 with 25 percent not met—“winning our nation’s wars”
was apparently 100 percent met even though, at the time, the final outcomes in
Afghanistan and Iraq were still unknown! Similarly, for defense of homeland secu-
rity, it was reported that 67 percent of goals were not met, a surprisingly high
failure rate for such a core defense function (i.e. protecting US citizens: USDOD
2011, 7–11).
Measuring defense output 65
US defense budget data only marginally enhances the understanding of defense
outputs in terms of capabilities and valuations. Performance results are recorded
by DoD staff who may have a not entirely unbiased interest in reported outcomes.
Budget data tend to be input-oriented, showing annual expenditure on military
personnel, operations and maintenance, R&D and procurement, and family
housing and military construction—all published by totals and by service.
The 2014 Quadrennial Defense Review (DoD 2014) described “tough choices”
being made in an era of fiscal austerity. These choices were reflected in a reduced
force structure, modernizing the forces, investing in readiness and innovation
to meet future challenges, and protecting the health of the AVF in undertaking
these reforms. Within this framework, the United States retained its capability to
simultaneously engage in two regional conflicts.

3.6.7 Evaluating international experience


Since the great recession, financial austerity played a major role in formulating
defense policy, especially in the US, UK, and the rest of Europe. It is under these
constraints that measuring defense output can be particularly helpful in improving
CBAs to guide difficult defense choices. Unfortunately, the response to austerity
often resulted in indiscriminate across-the-board force reductions to accommodate
politically determined budget choices. While the armed forces cannot ignore bud-
get constraints in policy formulation, sensible policy choices require recognition
of defense output implications of smaller budgets and the need to formulate better
measures of defense output to guide future CBAs.

3.7 Conclusion
This chapter has identified a set of important questions that arise in efforts to
measure defense outputs. Indeed, it has raised more questions than answers, but
this investigation contributes to further the understanding needed to address the
central research questions: What is defense output? How can it be valued? Under
what conditions is it a worthwhile investment?
In its published form, the international experience of measuring defense output
reveals some useful intermediate output measures, usually in the form of spe-
cific defense capabilities.22 These are improvements on the traditional reporting
of inputs that have typically included numbers of military personnel and equip-
ment (e.g. combat aircraft, tanks, and warships). By themselves, input measures
offer little indication of the value of overall defense capabilities such as peace,
protection, deterring conflicts, and insurance against future threats.23
A starting point in answering the central research questions is to apply CBA: to
identify the costs of defense and then ask whether defense provides at least a com-
parable level of benefits in the outputs produced. It is also important to capture
non-economic benefits in addition to measurable economic benefits in measur-
ing the overall benefits of defense spending. For example, if defense spending
66 K. Hartley, B. Solomon
costs $X billion, does it provide overall benefits of a similar value? Similar ques-
tions can be asked about the costs and benefits of conflict and peacekeeping
operations.24
Next, the CBA can focus on incremental (or marginal) changes. If defense
spending is increased or decreased by 10 percent, what are the effects on defense
outputs (benefits)? Such marginal analysis can be assessed as a whole (on overall
defense output), or by each military service (e.g. what would be the impact of a
10 percent increase or decrease in the size of the Army?).
Specifying the important questions is the first stage in any evaluation; but
who raises and answers the questions? In a democracy, elected politicians are
ultimately responsible for determining the size of military expenditures and its
allocation among each of the services. Typically, unelected agents within the
military propose many of these choices.25 This reinforces the importance of devel-
oping meaningful defense output measures to guide future military investment and
divestment decisions.

Appendix 3.1: Program Alignment Architecture (PAA)


The PAA is part of a broader government policy on managing resources
and reporting results that directs all departments to have clearly defined and
measurable strategic outcomes. DND’s articulation is depicted in Table A3.1 and
it intends to show what defense does and the results it is striving to achieve.
As shown in Table A3.1, the third strategic outcome states that “Defense oper-
ations improve peace, stability and security wherever deployed.” This statement
is rather vague on the military capabilities required to achieve the outcome. Is
this a peacekeeping or peace enforcement mission? Is it exclusively the provision
of strategic lift to transport aid and soldiers or air patrol? How quickly could the
Canadian armed forces (CAF) provide an effective fighting force for an operation
in North America? The statement also asserts with certainty an outcome, without
any probabilities attached (e.g. defense operations might improve peace, stability,
and security).
Strategic outcome 4, which states “Care and support to the Canadian armed
forces and contribution to Canadian society,” is equally ambiguous. The care
and support to CAF members can be considered an input (part of recruitment
and retention) while contribution to society could be a grant/subsidy program.
Even if we make the assumption that outcomes and outputs are the same, there
are still challenges in measuring the outcome/output. How do you measure
international peace?

Appendix 3.2: UK experience in other parts of the public


sector and the private sector
Other parts of the UK public sector have addressed the issue of measuring their
outputs. Examples include health, education, public order and safety, transport,
and social protection. The problems of measuring UK public sector outputs were
Measuring defense output 67
Table A3.1 Canadian Program Alignment Architecture 2011–12

Strategic outcome Program Actual Alignment to Government


spending of Canada outcomes
2011–12
CDN$000a

1. Resources are 4,334,325


delivered to meet
Government defence
expectation
2. National Defence is 10,169,909
ready to meet
Government defence
expectations
3. Defence operations 3.1 Situational 599,459 A safe and secure world
improve peace, awareness through international
stability and security engagement
wherever deployed
3.2 Canadian peace, 336,917 A safe and secure
stability and Canada
security
3.3 Continental 202,580 A strong and mutually
peace, stability beneficial North
and security American partnership
3.4 International 1,980,673 A safe and secure world
peace, stability through international
and security engagement
Sub-total 3,119,629
4. Care andsupport to 1,516,338
the Canadian Armed
Forces and
contribution to
Canadian society
5. Internal services 1,078,558
Total 20,218,759

Source: Department of National Defence (DND) 2013.


Note
a Current CDN$000.

reviewed by Atkinson (2005). This review started by recognizing that government


output is generally non-marketed output and it is the absence of market transac-
tions which underlie many of the problems of measuring public sector outputs.
The traditional approach used in national accounts’ statistics is the “output equals
input” convention (Atkinson 2005). The review recognized that in the case of
defense it is hard to identify the exact nature of the output (Atkinson 2005, 12).
Some principles were suggested. Can we borrow from private sector experience
(where the focus is on value-added)? Also, government output should be adjusted
for quality changes (which is a problem for defense).
68 K. Hartley, B. Solomon
The Atkinson Review reported on experience of output measurement in public
sectors such as health, education, public order and safety, and social protection.
In health, it reported on the use of an aggregate output index constructed from
separate series such as total numbers of in-patient and day cases. It recognized
quality issues where healthcare embraces saving lives and extending the lifespan
and preventing illness. It reported on the possibilities of using quality measures of
healthcare based on quality-adjusted life years (QALYS).
Education output was measured by such indicators as examination results and
the numbers of full-time school pupils (but numbers fail to reflect attendance).
Public order and safety embraced police, fire, law courts, and prisons. Outputs
were measured by such indicators as number of nights spent in prison, fires
fought, and the number of crime-related incidents. Social protection includes the
residential care of children and adults and output is measured by the numbers in
residential care (Atkinson 2005). Experiences of measuring outputs in these parts
of the UK public sector provide some guidance for measuring defense outputs.
Measuring health outputs which involve saving lives and preventing illness have
parallels in defense (as mentioned earlier, Chapter 9 of this book is another illus-
tration). The development of QALYS for health might be extended to defense in
the form of protection-adjusted life years (PALYS). Since the Atkinson review, the
ONS has continued to develop and improve output measures for various parts of
the UK public sector. For example, education output measures are now adjusted
for attendance and for quality changes (e.g. annual changes in examination points’
scores: ONS 2010). But often the output measures are aggregate indices with no
valuation of actual outputs.
Experience of measuring output in the UK transport sector has addressed a key
issue raised in defense: namely, the value of life and the value of lives saved by
transport improvements. The value of life is based on a person’s willingness to
pay (e.g. for good healthcare and for road safety improvements). On this basis, the
UK Department of Transport valued a life at £1.57 million and a non-fatal serious
injury at £176,215 per person (2009/10 prices).
Experience in the UK private sector might provide guidance on the possible valua-
tions to be placed on defense output. In the private sector, individuals and households
allocate resources to protection and safety. Examples include insurance policies for
protecting property; household security measures (e.g. cameras, fencing, alarms,
and dogs); car insurance and purchase of safer cars; location of homes in ‘safe’
areas; and the purchase of private medical and life insurance. In addition, there are
public expenditures on protection, including police, fire and rescue services, pris-
ons, as well as healthcare. Expenditures on these ‘comparator sectors’ provide an
indication of society’s willingness to pay for various measures of protection.

Notes
1 This chapter is based on a report prepared by Keith Hartley for the Defence Research
and Development Canada under contract number DND-10/23136. Parts have been
updated to 2014.
2 See also Williamson (2000) for more nuanced discussion.
Measuring defense output 69
3 For example, a nation’s international peacekeeping contributions might provide consid-
erable satisfaction to the country’s prime minister, senior ministers, and civil servants
able to attend international meetings at the UN and to participate in regional meetings.
The principal-agent and public choice analysis raises the general question of who gains
and who pays for these defense policies (e.g. international peacekeeping, national pro-
curement of defense equipment including offsets). Ultimately, taxpayers pay and receive
some defense benefits whilst agents consume some benefits not explicitly supported by
the majority of voters and taxpayers.
4 For example, compare today’s space satellite communications systems with the military
communications facilities in 1914 (e.g. observation balloons).
5 Skills and productivity differ between regular forces, conscript, and reserve forces.
Other complementary or substitutable labor inputs include civilian labour, contractors,
national guard and reserves, and even police forces (e.g. police forces substituted for
British Army troops in policing Northern Ireland: Ridge and Smith 1991).
6 Chapter 14 offers an interesting application of a new “real options” cost–benefit analysis
of R&D investments.
7 The varied results in this field reflect different economic and econometric models,
combinations of variables, time-periods, cross-section and time-series studies; a het-
erogeneous set of countries; and the use of data with varying degrees of reliability and
scope of coverage.
8 With output budgets, a distinction needs to be made between the budget available to
the MoD and the budget released to Parliament and the public. The published version of
the budget does not reveal all the information available to the MoD and the basis for the
choices which are reflected in the publicly released version of the budget (Davies et al.
2011, Chapter 17).
9 As an example, note that in Table 3.3 capital departmental expenditures limit (DEL) is
part of overall resource DEL and reflects investments spending (cash) which appears
in MoD’s balance sheet to be consumed over a number of years. The resource DEL
includes depreciation and cost of capital changes.
10 Formerly known as Program Activity Architecture.
11 Norman Augustine famously forecasted that with continued rising unit costs, by 2054
the entire US defense budget will purchase just one aircraft which would have to be
shared between the Air Force and Navy (the Marines would have it for one day in leap
years). He also forecasted that the UK and France would reach this position two years
earlier (Augustine 1987, 143).
12 See Chapter 9 in this book for a novel application of the statistical value of life.
13 It was not expected that both deployments would involve warfighting or that they would
be maintained for longer than six months. One might be a short warfighting deployment;
the other an enduring non-warfighting operation (SDR 1998, 23).
14 “Small scale” is defined as the UK’s deployment to Macedonia in 2001; “medium-scale”
as Afghanistan (2001); and “large scale” as operation TELIC (Iraq).
15 When part of a coalition, the optimum ratio for prolonged commitments was 3–4 ships
and 5 Army and RAF crews/units for each one deployed (Cmnd. 6269 2004).
16 This is about two-thirds of the force deployed to Iraq in 2003.
17 For example, its 2011 report focused on success in Afghanistan reflected in the costs of
operations; the costs of its force elements (e.g. a ship at an annual cost of £28 million;
a fixed wing combat aircraft at £6.5 million per year); and the direct costs of Service
personnel (£49,000 per Service personnel per year: HCP 992 2011).
18 Interestingly, MoD’s Performance Report included a section on Defence exports where
one aim is to support British industry and jobs (HCP 992 2011). Defence exports are
not an obvious output indicator for the MoD.
19 There is a NATO (NATO 2010) commitment to spend at least 2 percent of GDP on
defense; an aim of achieving savings from contract renegotiations with the defense
70 K. Hartley, B. Solomon
industry; a goal to retain a surface fleet of 19 warships; a commitment to reduce the
force of main battle tanks by 40 percent; and, lastly, a commitment to scrap the Nimrod
MRA4 fleet (at a savings of some £200 million per aircraft: HCP 992 2011).
20 The decision to replace the Collins’ class submarines was confirmed but the life of the
Collins’ class was extended by seven years.
21 Compares to “Lines of Development” in the UK and DOTMILPF (doctrine, orga-
nization, training, material, leadership and education, personnel, and facilities) in
the US.
22 None of the nations reviewed in this chapter addressed the challenges of measuring and
valuing overall defense output. The nearest to an intermediate output measure consisted
of the identification of various defense capabilities, but these were not always com-
prehensive. For example, the UK did not identify all its capabilities, including defense
of the UK homeland and its nuclear deterrent, and no valuations were provided for
the assorted capabilities. Nonetheless, the focus on defense capabilities is an improve-
ment over the traditional focus on input measures such as numbers of personnel and
equipment. The next challenge may be to assign weights to the mix of capabilities and
aggregate them into an index that represents an overall measure of defense output.
23 Nor should it be assumed that there exists a single “best” indicator: performance indi-
cators can often give unexpected and perverse results (e.g. the operation was a success
but the patient died).
24 For example, was the Iraq conflict a worthwhile investment for the US?
25 Governments could use representative samples of voters to form focus groups which
would offer their views on the size of alternative defense budgets and force structures.
These focus groups could receive advice from officials and military personnel. Focus
groups are not an ideal solution (e.g. free-rider problems remain; groups have to be
selected; and they will have their internal momentum and dynamics), but would provide
politicians with an additional mechanism for identifying voter preferences on defense
spending and policy.

References
Atkinson, Sir A. B. 2005. “Measurement of Government Output and Productivity for the
National Accounts.” In Atkinson Review: Final Report. Basingstoke, England: Palgrave
Macmillan.
Augustine, N. R. 1987. Augustine’s Laws. London: Penguin Books.
Australian Defence Force (ADF). August 25, 2011. Australian Defence Force Posture
Review Public Consultation and Submission Process. Canberra, AU: Department of
Defence.
Baldwin, D. A. 1997. “The Concept of Security.” Review of International Studies 23: 5–26.
Benoit, E. 1973. Defense and Economic Growth in Developing Countries. Boston, MA:
Lexington Books.
Berkok, U. G., and B. Solomon. 2011. “Peacekeeping, Private Benefits and Common
Agency.” In Handbook on the Economics of Conflict, edited by D. L. Braddon and K.
Hartley. Northampton, MA: Edward Elgar Publishing.
Black, R. 2011. Review of Defence Accountability Framework. Canberra, AU: Australian
Government Department of Defence.
Comptroller and Auditor General (CAG). June 2011. Central Government: Cost-
effectiveness and Improving Annual Reports. Wellington, NA: Comptroller and Auditor
General.
Cornes, R., and T. Sandler. 1996. The Theory of Externalities, Public Goods and Club
Goods (2nd edition). Cambridge: Cambridge University Press.
Measuring defense output 71
Davies, N, T. Turner, A. Gibbons, D. Jones, S. Davis, and N Bennett. 2011. “Helping
Secure the ‘Biggest Bang for the Taxpayers’ Buck:’ Defence Resource Management
in the United Kingdom.” In Handbook on the Economics of Conflict, edited by D. L.
Braddon and K. Hartley. Northampton, MA: Edward Elgar Publishing.
Defence Analytical Services and Advice (DASA). 2011. U.K. Defence Statistics Com-
pendium 2011. London: The Stationery Office Limited.
Deger, S., and R. P. Smith. 1983. “Military Expenditure and Growth in Less Developed
Countries.” Journal of Conflict Resolution 27(2): 335–353.
Department of Defence (DoD). 2009. Defending Australia in the Asia Pacific Century:
Force 2010. Canberra, AU: Department of Defence.
——. 2011. Defence Portfolio Budget Statements 2011–2012. Canberra, AU: Department
of Defence.
——. 2013. Defence White Paper 2013. Canberra, AU: Department of Defence.
—— 2014. Quadrennial Defense Review. Available at www.defense.gov/pubs/
2014_Quadrennial_Defense_Review.pdf [accessed 10 September, 2014].
Department of National Defence (DND). 2013. Department of National Defence: Report
on Plans and Priorities 2013–14. Ottawa, CA: Department of National Defence.
Dunne, P., and R. P. Smith. 2010. “Military Expenditure and Granger Causality: A Critical
Review.” Defence and Peace Economics 21(5–6): 427–441.
Engerer, H. 2011. “Security as a Public, Private or Club Good: Some Fundamental
Considerations.” Defence and Peace Economics 22(2): 135–145.
European Defence Agency (EDA). March 31, 2011. Defence Data 2009. Brussels, BE:
EDA.
German Ministry of Defense (GMOD). May 2011. Defence Policy Guidelines. Berlin, DE:
GMOD.
Hartley, K. 2010a. “Defense Economics.” In The Economics of Human Systems Integration:
Valuation of Investments in People’s Training and Education, Safety and Health, and
Work Productivity, edited by W. B. Rouse. Hobeken, NJ: Wiley.
——. 2010b. The Case for Defence. Defence and Peace Economics 21(5–6): 409–426.
——. 2011. Defence Output Measures: An Economic Perspective. Defence R&D Canada–
CORA. Contract Report: DRDC CORA CR 2011-178.
Hildebrandt, G. G. 1990. “Services and Wealth Measures of Military Capital.” Defence
Economics 1(2): 159–176.
——. 1999. “The Military Production Function.” Defence and Peace Economics 10(3):
247–272.
Hitch, C. J., and R. N. McKean. 1960. The Economics of Defense in the Nuclear Age.
London: Oxford University Press.
House of Commons (HCP) 992. 2011. Ministry of Defence Annual Report and
Accounts 2010–2011 (including Performance Report). London: The Stationery Office
Limited.
House of Commons Defence Committee. 2003. Delivering Security in a Changing World.
2003 Defence White Paper: Cmnd. 6041, Volumes I–II. London: The Stationery Office
Limited.
House of Commons Ministry of Defence. 2004. Delivering Security in a Changing World:
Future Capabilities. 2004 Defence White Paper: Cmnd. 6269. London: The Stationery
Office Limited.
Jones-Lee, M. E. 1990. “Defence Expenditure and the Economics of Safety.” Defence
Economics 1(1): 13–16.
72 K. Hartley, B. Solomon
Middleton, A., S. Bowns, K. Hartley, and J. Reid. 2006. “The Effect of Defence R&D on
Military Equipment Quality.” Defence and Peace Economics 17(2): 117–139.
Middleton, A., K. Copsey, and K. Hartley. 2011. Estimating a Production Function for
Military Power (Mimeo). Malvern, UK: Economex Limited.
Military Policy Review 2013. Available at www.difesa.it/Primo_Piano/Documents/2013/
gennaio%202013/Direttiva%20Ministeriale_ENG.pdf [accessed 10 September, 2014].
National Audit Office (NAO). 2005. Assessing and Reporting Military Readiness (HCP
72). London: The Stationery Office Limited.
——. 2010. Performance of the MoD 2009–10. Briefing to House of Commons Defence
Committee. London: The Stationery Office Limited.
National Security and Intelligence. 2010. Securing Britain in an Age of Uncertainty: The
Strategic Defence and Security Review. (Cmnd. 7498). London: The Stationery Office
Limited.
Nativi, A. November 19, 2010. “Italy Protects Defense Modernization Spending.” Aviation
Week.
New Zealand Defence Force (NZDF). June 1991. Annual Plan: G55. Wellington, NZ:
NZDF.
——. 1993. New Zealand Defence Force Annual Report 1993. Wellington, NZ: NZDF.
——. 2011. New Zealand Defence Force Statement of Intent 2011–2014. Wellington. NZ:
NZDF.
——. 2013. New Zealand Defence Force Annual Report 1993. Wellington, NZ: NZDF.
North Atlantic Treaty Organization (NATO). 2010. Financial and Economic Data Relating
to NATO Defence. Brussels, BE: NATO, Press and Media.
Office of Under Secretary of Defense (USDOD). 2011. US DoD Fiscal Year 2012 Budget
Request. Washington, DC: Department of Defense.
Office for National Statistics (ONS). 2008. Scoping Paper on the Possible Improvements
to Measurement of Defence in the U.K. National Accounts: U.K. Measurement of
Government Activity. Newport, SW: ONS.
——. 2010. Annual Report 2009–10. Newport, SW: UK Centre for Measuring Government
Activity.
Organisation for Economic Co-operation and Development (OECD). 2010. Main Science
and Technology Indicators. Paris: OECD.
Penketh, A. June 18, 2008. “The Big Question: What is the New French Defence Strategy,
and Should we Follow Suit?” The Independent Newspaper.
Receiver General of Canada (RCG). 2013. The Public Accounts of Canada Volume II.
Ottawa, CA: Government of Canada.
Ridge, M., and R. P. Smith. 1991. “U.K. Military Manpower and Substitutability.” Defence
Economics 2(4): 283–293.
Sempre, C. M. 2011. “The European Security Industry: A Research Agenda.” Defence and
Peace Economics 22(2): 245–264.
Solomon B., P. Chouinard, and L. Kerzner. October 2008. The Department of National
Defense Strategic Cost Model (Volume II: Theory and Empirics, Defence R&D). Ottowa,
CA: Canada-CORA.
Stockholm International Peace Research Institute (SIPRI). 2011. Military Expenditure
Database (Database). Solna, SE: SIPRI.
The Strategic Defence Review (Cmnd. 3999). 1998. (See also Supporting Essays). London:
The Stationery Office Limited.
Sweden Ministry of Defence (SMOD). March 19, 2009. A Functional Defence. (Press
Release by the Sweden Ministry of Defence).
Measuring defense output 73
Tisdell, C., and K. Hartley. 2008. Microeconomic Policy: A New Perspective (2nd edition).
Cheltenham, UK: Edward Elgar Publishing.
Williamson, O. E. 2000. “The New Institutional Economics: Taking Stock, Looking
Ahead.” Journal of Economic Literature 38: 595–613.
Zaller J. R., and S. Feldman. 1992. “A Simple Theory of the Survey Response: Answering
Questions versus Revealing Preferences.” American Journal of Political Science 36(3):
579–616.
4 The economic evaluation
of alternatives
Francois Melese

An AoA [analysis of alternatives] is an analytical comparison of the operational


effectiveness, suitability, and life-cycle cost of alternatives that satisfy established
capability needs.
US Department of Defense (DoD 2006, Section 3.3)

4.1 Introduction: making the case for “affordability”


Military cost–benefit analysis (CBA) is routinely used to shape future forces and
influence military spending. It underpins the “analysis of alternatives” (AoA)1 in
the United States and the concept of “value for money” in the United Kingdom. A
standard definition states: “An AoA is an analytical comparison of the operational
effectiveness, suitability, and life-cycle cost of alternatives that satisfy established
capability needs” (DoD July 7, 2006 Section 3.3). Missing from this definition is
“affordability.”
A weakness in the traditional application of military CBA in the United States is
that while AoAs correctly focus on life-cycle costs and operational effectiveness
of alternatives, forecasts of future funding rarely appear.2 As former chairman
of Lockheed Martin Corporation, Norm Augustine, once noted: “affordability is
rarely considered.”3
As ministries of defense struggle to obtain the best value for every dollar they
invest in defense programs, affordability is a growing concern.4 In an age of fiscal
austerity, where deficits and debt require rethinking defense budgets, affordability
increasingly takes center stage.
The latest US defense guidance emphasizes the importance of affordability, but
offers only nominal suggestions about how to include it in an AoA.5 This chapter
reveals several concrete approaches to structuring public investment decisions that
emphasize affordability. We refer to this set of military CBA approaches as the
“economic evaluation of alternatives” (EEoA). The goal is to encourage analysts
and decision-makers (DMs) to integrate affordability (funding and other resource
constraints) early and explicitly in the acquisition process alongside estimates of
operational benefits and life-cycle costs.
The EEoA approach to military CBA brings taxpayers up front alongside
warfighters in the defense acquisition process. This requires working with vendors
Economic evaluation of alternatives 75
to build proposals based on future funding (budget) scenarios.6 EEoA clearly sep-
arates the life-cycle costs (or “price”) of an alternative, its expected operational
effectiveness or benefit (performance and schedule), and the resources (funding or
budget) likely to be available (i.e. affordability) for the overall program.7

4.2 Integrating requirements generation,


defense acquisition, and PPBS
A focus on affordability offers a unique opportunity to achieve a significant
defense acquisition reform—to closely couple requirements generation systems
and defense acquisition systems with financial management systems such as the
Planning, Programming, and Budgeting System (PPBS). The primary goal of
PPBS is to build an “optimal” mix of forces that maximizes national security
within fiscal (affordability) constraints (DoDD 2013). A brief review of defense
budgeting and acquisition highlights the valuable role EEoA can play in improving
investment outcomes.
The requirements generation process naturally fits into the planning phase of
PPBS. The first step in any investment analysis is to identify the derived demand
for a key capability, program or project.8 The focus of EEoA is on materiel
investments identified to fill critical capability gaps. This is accomplished through
DoD’s Requirements Generation System.
Operator demands (“requirements”) are identified and refined in the planning
phase of the PPBS process. Whenever major “materiel” solutions are recom-
mended, prospective military investments are identified that serve as the basis for
a military cost–benefit AoA. AoAs underpin the development of major acquisition
programs in the Defense Acquisition System (DAS).9
Ideally, the planning phase of PPBS establishes fiscally constrained guidance
and priorities for the military services. This guidance on readiness, sustainability,
modernization, etc., provides direction for DoD Components (military depart-
ments and joint defense agencies) in the programming phase of PPBS. In the US
component programs and new investment proposals are summarized in what the
services (Army, Navy, Air Force, and Marine Corps) call their Program Objectives
Memorandum (POM).
Service POMs provide detailed resource-allocation decisions (funding, per-
sonnel, etc.) for programs projected six years into the future. A comprehensive
summary of service POMs is captured in the Future Year Defense Plan (FYDP)
database. A challenge for EEoA is that evaluations of major investments require
DAS life-cycle cost estimates that project well beyond the six years of the FYDP.
This concern is addressed in Chapter 7.
Each POM is subsequently reviewed by the Joint Staff and senior leadership
in the Office of the Secretary of Defense (e.g. Cost Assessment and Program
Evaluation—CAPE) to ensure that it satisfies strategic planning guidance, and
that the overall FYDP is effective and affordable. Today, in the US, budgeting
in PPBS occurs concurrently with programming. The budgeting phase con-
verts the programming phase’s (output-oriented) benefit or military capability
76 F. Melese
perspective into the (input-oriented) cost format required for Congressional
appropriations.10
The Under Secretary of Defense Comptroller and the Office of Manage-
ment and Budget (OMB) jointly review budget submissions to ensure that
programs are affordable in the budget year—i.e. satisfy current fiscal year con-
straints.11 Once accepted by the Secretary of Defense, this spending plan is
included in the President’s overall federal budget submission to the Congress for
approval.
In principle, AoAs should contribute to requirements generation in the planning
phase of PPBS, helping to identify investment alternatives and the benefits and
costs of those alternatives. Through the DAS’s decision milestones, AoAs should
also influence the programming phase of PPBS.12 According to GAO, however:

[T]he vast majority of capability proposals that enter the [Requirements


Generation] process are . . . approved without accounting for resources [fund-
ing/budgets] that will be needed to acquire the desired capabilities.13
(GAO 2009, 6)

This concern about future financial resource availability (“affordability”) is


the primary focus of EEoA. In considering alternative funding scenarios (that
partly rely on accurate FYDP forecasts), EEoA injects an explicit constrained-
optimization approach into defense investment decisions that parallels one already
embedded in the programming phase of PPBS: choosing an optimal mix of
forces, equipment, and support that maximizes national security subject to fiscal
constraints.

4.3 Integrating affordability assessments into military CBA


In practice, military requirements are often approved without fully accounting for
resources available, including future funding realities. While AoAs attempt to esti-
mate costs and benefits of competing alternatives, affordability is addressed as a
separate exercise, and then only ex-post.14 This results in the following obser-
vation: “at the program level, the key cause of poor outcomes is the approval
of programs with business cases [AoAs] that contain inadequate knowledge
about . . . resources [funding] needed to execute them” (GAO 2009, 7). Ironically,
this directly contradicts the department’s own policy outlined in DoD Directive
5000.01 which explicitly states:

All participants in the acquisition system shall recognize the reality of fiscal
constraints . . . DoD components shall plan . . . based on realistic projections of
the dollars . . . likely to be available [and] the user shall address affordability
in establishing capability needs.
[emphasis added] (DoD 5000.01 2007, Enclosure 1, 5)15
Economic evaluation of alternatives 77
According to the US Defense Acquisition Guidebook, the purpose of an
affordability assessment is to demonstrate that a program’s projected funding
requirements are realistic and achievable:16

In general, the assessment should address program funding over the six-year
programming [FYDP] period, and several years beyond. The assessment
should also show how the projected funding fits within the overall DoD
Component plan.17
[emphasis added] (DoD July 7, 2006, Section 3.2.2)

EEoAs provide a mechanism for analysts and DMs to embed affordability


assessments directly into AoAs. In preparing affordability assessments, the FYDP
offers a valuable source of data.

4.4 An intuitive guide to the economic evaluation


of alternatives
Nesting the requirements generation and DASs within PPBS suggests formulating
the military’s acquisition problem in terms of identifying and funding investments
that maximize benefits (performance or effectiveness) for a given budget. Structur-
ing the military investment problem as a constrained optimization, i.e. maximizing
military effectiveness subject to a budget constraint or, alternatively, minimiz-
ing costs of obtaining a given level of effectiveness, are the first two of the six
approaches recommended in the EEoA.18 Either approach would boost the value
of an AoA, since they both require and promote explicit assessment of higher
level (joint) resource allocation decisions in the planning and programming phases
of PPBS.
Multiple criteria decision-making (MCDM) techniques typically applied to
structure AoAs do not currently lend themselves to this interpretation (see Chap-
ter 8 for example). As a consequence, rather than being constrained by budgets,
a budget for a new program or project is generated as the output of an AoA,
producing so-called “funding requirements.”19 The third of the six possible
approaches proposed to structure an EEoA turns this on its head.20 Instead of gen-
erating a budget through the AoA process, forecasts of optimistic, pessimistic,
and most likely funding scenarios are produced as part of the PPBS process,
and then vendors are asked to build alternatives which fit within that budget
envelope.21
According to the US Defense Acquisition Guidebook, affordability assessments
should also provide details on how excess funding demands will be accommodated
by reductions in other mission areas or in other accounts.22 This marginal analysis
across programs is the last of the six approaches proposed to structure an EEoA,
labeled the “opportunity cost approach.” The fourth and fifth (modified budget and
modified effectiveness) approaches to EEoA involve attempts to level the playing
field across alternative vendors; in effect converting the problem back to the first
two constrained optimization approaches.
78 F. Melese
Whereas funding decisions for major programs take place through the PPBS
process, the GAO finds that:

. . . the process does not produce an accurate picture of the department’s


resource needs [funding/budget requirements] for weapon system pro-
grams . . . Ultimately, the process produces more demand for new weapon
system programs than available resources can support.23
(GAO, March 18, 2009, 6)

EEoA directly responds to these challenges. It also responds to other concerns


highlighted by GAO that continue to confront DoD’s DAS. This includes: “(1)
[to make] better decisions about which programs should be pursued or not pursued
given existing and expected funding; [and] (2) [to develop] an analytical approach
to better prioritize capability needs” (March 18, 2009, Highlights).
In sharp contrast with the MCDM approach used in most contemporary AoAs,
EEoA approaches explicitly identify and emphasize funding (budget or affordabil-
ity) constraints. In generating alternatives under optimistic, pessimistic, and most
likely funding scenarios, EEoA requires explicit input from the PPBS process. As
a consequence, widespread application of EEoA would directly contribute to the
goal of achieving:

. . . greater consultation between requirements, budget, and acquisition pro-


cesses [that] could help improve the department’s . . . portfolio of weapon pro-
grams . . . [and guaranteeing] decision makers responsible for weapon system
requirements, funding, and acquisition execution . . . establish an investment
strategy in concert . . . assuring requirements for specific weapon systems are
clearly defined and achievable given available resources [funding/budgets].
(GAO, July 2, 2008, 10 and 14)

The next section offers a brief description and critical evaluation of the status
quo. This includes two common decision criteria widely used in cost-effectiveness
analyses. The first is the popular “bang-for-the-buck” or benefit/cost ratio.24 The
second criterion is essentially a weighted average of cost and effectiveness, a deci-
sion rule generated by the standard MCDM approach to cost-effectiveness analysis
that underpins most contemporary AoAs (see Chapters 8 and 16, for example).
Section 4.6 illustrates the six approaches to structure an EEoA. The final section
concludes with a decision map to guide analysts and DMs in selecting which of
the six approaches is best suited to their circumstances.

4.5 A critical evaluation of the status quo:


two popular decision criteria
Today, most modern military investment (and divestment) decisions are supported
by some form of CBA. The US DoD applies CBA to anything from milestone
decisions for major defense acquisition programs and major acquisition of infor-
mation systems (MDAPs and MAISs) to outsourcing (OMB Circular A-76; Eger
Economic evaluation of alternatives 79
and Wilsker 2007), public-private partnerships, privatization, asset sales, or base
realignment and closure (BRAC) actions (see OMB Circulars A-94; Federal
Acquisition Regulations (FAR); Defense FARs; DoD 5000 series, etc.).
When benefits can be quantified, but not monetized, analysts rely on so-called
“measures of effectiveness” (MoEs) in which case CBA is generally referred to
as “cost-effectiveness” analysis.25 (US OMB 1992) The most common method-
ology and approach for building MoEs and for structuring cost-effectiveness
analyses is alternately referred to as multiple-criteria decision-making (MCDM),
multi-attribute utility theory (MAUT), or multiple-objective decision-making
(MODM) (see for example French 1986; Keeney and Raiffa 1976; Clemen 1996;
Kirkwood 1997; Parnell 2006; Ramesh and Zionts 1997).
This chapter describes some limitations of the two decision criteria used in
most AoAs and proposes an alternative methodology derived explicitly from
the constrained-optimization approach recommended for an EEoA. The latter
approaches are closer in spirit to the economic origins of cost-effectiveness anal-
ysis found in Gorman (1980); Hitch and McKean (1967); Michael and Becker
(1973); Stigler (1945); Theil (1952); etc.—and most often attributed to Lancaster
(1966a, 1966b, 1971, 1979).
A key difference between the MCDM approach to an AoA and EEoA
approaches is that instead of modeling decision alternatives from competing
vendors as points in cost-effectiveness space, EEoA models alternative vendor
proposals as functions of optimistic, pessimistic, and most likely funding (bud-
get) scenarios. The EEoA approach directly responds to GAO’s observation that
affordability needs to be an integral part of any business case (AoA):

[o]ur work in [uncovering] best practices has found that an executable busi-
ness case [requires] demonstrated evidence that . . . the chosen concept can be
developed and produced within existing resources [funding/budgets].
[emphasis added] (GAO September 25, 2008, 6)

Benchmarking against the private sector, GAO further emphasizes that:

. . .successful commercial enterprises. . .follow a disciplined integrated pro-


cess during which the pros and cons of competing proposals are assessed
based on strategic objectives. . .and available resources [budgets/funding].
[emphasis added] (GAO 2009, 5)

A distinctive feature of defense investment decisions is that multiple criteria


such as cost and effectiveness cannot easily be combined into a single, over-
all objective such as “government profitability.” The challenge of ranking public
investments when benefits cannot be monetized has generated a broad literature
across management science, operations research, and decision sciences. Tech-
niques that mostly fall under the umbrella of MCDM are routinely used by analysts
and DMs (such as AoAs) to guide public investment decisions. (See Chapter 8.)
80 F. Melese
This literature models investment alternatives as bundles of measurable character-
istics (criteria or attributes). The development of MoEs,26 combined with life-cycle
cost estimates (Chapter 5), are used to help rank alternatives. (See Chapters 8
and 16) An ongoing concern is how to integrate costs and effectiveness in the
selection process (see Henry and Hogan 1995; Melese and Bonsper 1996; Melese
et al. 1997; etc.).
In their pioneering work applying economic analysis to defense, Hitch and
McKean (1967, 158) define a “criterion” as the “test by which we choose
one alternative . . . rather than another” (p. 120). They stress that “[t]he choice
of an appropriate economic criterion is . . . the central problem in designing a
[cost-effectiveness] analysis.”
The two most popular decision criteria used to integrate cost and effectiveness
in AoAs are 1) to construct benefit/cost (or MoE/cost) ratios, and 2) to assign
a weight on cost relative to effectiveness and construct a weighted average of
cost and effectiveness (using a linear, separable, additive “value” function) as
demonstrated in Chapter 8. The latter decision criterion is a common prescrip-
tion for AoAs that emerges from MCDM. Both approaches can be problematic.
We first focus on what is arguably the most widely applied criterion—benefit/cost
ratios. Next, we move to the most common MCDM decision criterion—to
assign a relative weight to the cost (price) of alternatives in an overall value
function.
At first glance, the benefit/cost (MoE/cost) ratio or “bang-for-the-buck” cri-
terion is appealing. It turns out, however, to be largely meaningless unless
alternatives are constructed with a specific budget (funding/affordability) scenario
in mind or to achieve a specific level of effectiveness. Meanwhile, the second deci-
sion criterion can also turn out to be misleading in the absence of a specific budget
(funding/affordability) constraint or clear understanding of “opportunity costs.”27

4.5.1 “Bang-for-the-buck” (benefit/cost) ratios


It is well known that the application of a benefit/cost ratio (or “bang-for-the-buck”)
decision criterion to rank alternatives is largely meaningless unless alternatives are
constructed for a specific budget (funding) scenario or to achieve a specific level of
effectiveness. The problem of ranking public investments when benefits cannot be
expressed in dollars has spawned an extensive literature in management science,
operations research, and decision sciences:

1 In a military text entitled Executive Decision Making, the author offers that
“[w]hen we cannot fix cost or effectiveness, we might combine them to help
us choose between alternatives . . . If neither can be fixed . . . we can establish
a cost/effectiveness ratio” (i.e. a cost/benefit ratio) (Murray 2002, 6–3, 6–10).
2 In presenting what they claim is a “novel cost–benefit analysis” for the “com-
prehensive evaluation of competing military systems,” authors in the Acquisi-
tion Review Quarterly define a “merit function” as “a single number . . . [that]
reflects the ratio of benefits derived to dollars spent” (i.e. a benefit/cost ratio).
Economic evaluation of alternatives 81
The article asserts: “with this . . . approach, the cost-effectiveness of compet-
ing systems can be compared and “provides for objective and reliable decision
making,” where “a large system merit [benefit/cost ratio] is preferable to a
small one” (Byrns et al. 1995).
3 Similarly, in a section entitled “Comparing Costs and Benefits,” the US
Department of the Army’s Economic Analysis Manual states: “When the
results yield unequal cost and unequal benefits [. . .] in this situation all alter-
natives [. . .] may be ranked in decreasing order of their benefit/cost ratios”
(DoA February 2001, 32).28

Each of these examples recommends using a benefit/cost ratio as the deci-


sion criterion. Each also neglects to include either an affordability (cost or
budget) constraint or a performance (effectiveness or capability) constraint (see
Appendix 4.1). As legendary RAND analyst Gene Fisher (1971) clearly points
out in his classic text Cost Considerations in Systems Analysis:

The use of [benefit/cost] ratios usually poses no problem as long as the anal-
ysis is conducted in [a] framework . . . with the level of effectiveness or cost
fixed. However, it is common to encounter studies where this has not been
done, with the result that comparisons [are] essentially meaningless.
[emphasis added] (p. 11)

In applying economic analysis to defense, Hitch and McKean (1967) warn:

One common ‘compromise criteria’ is to pick that [alternative] which has the
highest ratio of effectiveness to cost . . . [M]aximizing this ratio is the [deci-
sion] criterion. [While] it may be a plausible criterion at first glance . . . it
allows the absolute magnitude of [effectiveness] or cost to roam at will. In
fact, the only way to know what such a ratio really means is to tighten the
constraint until either a single budget (or particular degree of effectiveness)
is specified. And at that juncture, the ratio reduces itself to the test of maxi-
mum effectiveness for a given budget (or a specified effectiveness at minimum
cost), and might better have been put that way at the outset . . . .29
(pp. 165–7)

The first two ways to structure an EEoA, as well as the fourth and fifth
approaches, follows their advice:

The test of maximum effectiveness for a given budget (or alternatively, mini-
mum cost of achieving a specified level of effectiveness) . . . seems much less
likely to mislead the unwary.30
(p. 167)

The conclusion is that the use of benefit/cost ratios as a decision criterion


in AoAs does not pose a problem as long as the analysis is structured paying
82 F. Melese
close attention to affordability (i.e. a budget/funding constraint) or operational
performance (i.e. a fixed level of effectiveness or MoE).31 Since AoAs typi-
cally evaluate vendor proposals that differ in both costs (price) and benefits
(MoEs), using benefit/cost (or effectiveness/cost) ratios to rank alternatives is
at best “misleading.”32 Partly as a consequence, decision scientists developed
another approach to rank investment options, the MCDM criterion routinely
applied in AoAs. This second widely prescribed decision criterion is examined
below.

4.5.2 Weighted averages of cost and effectiveness: assigning


a weight to cost as a proxy for affordability
MCDM is often used as an umbrella term, and we will do so here. “In the
literature, the terms multi-attribute decision making (MADM), multi-criteria deci-
sion making (MCDM), and multi-objective decision making (MODM) are used
almost interchangeably” (French 1986, 105). In a typical MCDM evaluation, a
DM is asked to identify desired criteria (characteristics or attributes) of a project,
program or system to fill some critical capability gap, given a specific threat sce-
nario. Next, the DM is asked to reveal agreeable tradeoffs among those attributes.
An exercise of this sort helps analysts to uncover the DM’s underlying utility
function—similar to the economist’s indifference curves that reflect a consumer’s
“utility” function—to help generate an MoE for each alternative.33
To uncover a DM’s utility function, beginning with Saaty (1977), deci-
sion scientists bridged an important implementation gap. Multiple-objective (ana-
lytic) hierarchy approaches were developed to help reveal underlying utility
functions.
The objectives hierarchy approach can help a DM to work down from a high-level
objective (provide national security) to a relevant set of sub-objectives (an effective
airlift capability) to specific attributes that underpin those sub-objectives (mobility,
transportability, etc.) and, finally, to measurable characteristics (mobility = speed
(S), range (R); transportability = payload (P), weight (W ), etc.). The outcome in
this example is a utility function for airlift capability: U = U (M(S, R);T (P, W )),
where the desired characteristics (speed, range, payload, weight) might be measured
respectively in mph, miles, cubic feet, and pounds.
The standard procedure in the literature is to define a linear, additive separable
utility function.34 This generates an MoE for each alternative, roughly analo-
gous to a weighted average of its attributes (see Chapter 8 for example). There
exists a vast literature concerned with eliciting preference weights and the nor-
malization of characteristics data that involves several important issues for future
research.35
Momentarily overlooking these issues, it is interesting to note in passing that
maximizing a linear multi-attribute utility function subject to a budget (affordabil-
ity) constraint yields a decision rule analogous to the benefit/cost ratio criterion
discussed above. Under the assumption of a fixed budget and a linear, additive
separable utility function, the benefit/cost decision rule can indeed be used to
Economic evaluation of alternatives 83
rank alternatives. In this case, the winning alternative is the one that generates
the highest MoE per dollar or the biggest “bang-for-the-buck” (or in the dual,
cost minimization, the lowest “cost-per-kill”). With a more general (non-linear)
utility function, the equivalent optimization generates a more complex marginal
benefit/marginal cost decision rule.36
In reality, MCDM techniques that underpin most AoAs do not involve an
explicit discussion of budgets, affordability or funding (resource constraints) to
structure the decision problem. As a consequence, the problem is typically not
structured as described above.
Instead of structuring an AoA as a constrained optimization, the most common
decision-analysis approach is to simply attach a weight to cost and introduce
it directly into the utility function.37 (see Chapter 8 for example). The relative
weight assigned to the cost of alternatives under consideration is the proxy for
affordability. As opposed to a benefit/cost (or effectiveness/cost) ratio, this pop-
ular MCDM approach generates an overall “value” function that is essentially a
weighted average of cost and effectiveness.
Alternatives are ranked and selections made through an unconstrained optimiza-
tion where the best alternative is the one that maximizes the “overall effectiveness”
or “value” function, V = V (MoE, COST).
According to two prominent authors:

Deterministic decision analysis is concerned with finding the most preferred


alternative in decision space by constructing a value function representing a
decision maker’s preference structure, and then using the value function to
identify the most preferred solution.
(Ramesh and Zionts 1997, 421)

The linear, additive separable version of this value function is frequently used to
calculate a positively weighted MoE and negatively weighted cost for each alter-
native (for example, see Beil and Wein 2003; Che 1993; Clemen 1996; French
1986; Hwang and Yoon 1981; Keeney 1994; Keeney and Raiffa 1976; Kirkwood
1997; Liberatore 1987; Pinker et al. 1995; Vazsonyi 1995).
The typical decision sciences’ (MCDM) approach to an AoA38 thus
involves:

Given several alternatives, select the preferred alternative that provides the
best value, i.e. Maximize: V(MoE,COST) = w1*MoE − w2*COST.

This requires two important modeling efforts: 1) Building an effectiveness


(MoE) model (non-cost–benefit factors; MoE = quality, schedule, etc.); and 2)
Building a cost model (costs/prices; estimate total system life-cycle costs, total
ownership costs—see Chapter 5). Once these independent modeling efforts are
completed, the challenge is to integrate the two either using benefit/cost ratios
(discussed earlier) or by assigning a relative weight to cost (a value for w2 in the
example above) as a proxy for affordability.
84 F. Melese
The typical recommendation in the applied literature to integrate cost and effec-
tiveness is to ask the Decision Maker (DM): “How important is cost relative to
effectiveness?” The US FAR and both the General Services Administration (GSA)
and the OMB explicitly promote this approach:39

The solicitation shall state whether all evaluation factors other than cost/price,
when combined [i.e. an MoE], are significantly more important than, approx-
imately equal to, or significantly less important than cost/price.
(GSA, DoD, NASA, March 2005, Section 15.101-1(2))

The specific weight given to cost or price shall be at least equal to all other
evaluation factors combined unless quantifiable performance measures can be
used to assess value and can be independently evaluated.
(OMB Circular A-76, B-8)

An early proponent of this (MCDM) decision methodology offers an example


of administrators evaluating alternative pollution control devices being asked to
answer questions such as: “Which is more important, costs or [reducing] pollutant
concentrations [i.e. effectiveness]?” (Keeney 1994, 797). As the author is quick
to point out, the problem with this approach is that without some estimate of the
total budget available or any knowledge of opportunity costs, one cannot expect
the DM to provide a sensible answer. The author (ironically an early adopter and
prolific advocate of this MCDM approach) warns: “I personally do not want some
administrator to give two minutes of thought to the matter and state that pollutant
concentrations are three times as important as cost”40 (Keeney 1994, 797).
Figures 4.1 and 4.2 offer illustrations of alternative vendor offers. Figure 4.1
reflects a situation where the DM believes costs to be important enough (and thus
assigns a sufficiently large relative weight to cost, reflected in the ratio w2/w1)
that the preferred alternative is A1 (the low-cost option). The opposite case is
illustrated in Figure 4.2.41 Similar graphs are presented in Chapter 8.
How does a DM decide on an appropriate weight to assign to costs? Con-
sider an extreme case. Suppose affordability is not an issue, so funding is not
an issue. In that case the budget is not binding, making costs irrelevant. Clearly
this means a zero weight should be assigned to costs and the alternatives can be
ranked exclusively on the basis of their MoEs (e.g. A2 wins). As a consequence,
any weight applied to the costs of the alternative proposals (w2) must reflect an
implicit concern about affordability (future funding or budget realities).
A key hypothesis in the EEoA is that if a DM pays any attention to costs
(i.e. places a weight on cost), it is because they acknowledge an implicit fund-
ing/budget or affordability constraint, or recognize the opportunity cost of funds
committed to the program. Future funding and opportunity cost concerns are
directly related to the earlier discussion of affordability, and close coordination
of requirements generation, defense acquisition, and PPBS.42
The irony, as Keeney (1994) rightly observed, is that to assign any weight to
costs requires the DM to have some understanding of affordability (funding/budget
Economic evaluation of alternatives 85
realities) and an appreciation of opportunity costs. If this information is known, then
analysts and DMs have no reason to take the MCDM approach and assign a weight to
costs since the more robust, constrained-optimization (mathematical programming)
EEoA approach is available (Simon and Melese 2011).
In fact, it is relatively straightforward to demonstrate that even if the DM had
perfect information about future funding (budget/affordability) and attempted to
interpret that information through a weight assigned to the relative cost (price)
of alternatives (as illustrated in Figures 4.1 and 4.2), vendor rankings that resulted

Which alternative is “best”?


Decision sciences approach
Max V = V(MoE, Cost) = w 1 ×MoE – w 2 × Cost
Ask decision-maker what is more important: MoE or cost?
(d MoE/d Cost = w 2 /w 1)
MoE(Utils)

A2
Effectiveness

If Cost has a sufficiently


A1 greater weight (w2 >>w1),
then low cost, low
effectiveness alternative
A1 wins

Cost

Figure 4.1 When cost is relatively more important than effectiveness.

Which alternative is “best”?


Decision sciences approach
Max V = V(MoE, Cost) = w1 × MoE – w2 × Cost
(d MoE/d Cost = w2 /w1)
MoE(Utils)

A2
Effectiveness

If Performance has a
A1 sufficiently greater weight
(w1 >>w2), then high cost,
high effectiveness
alternative A2 wins

Cost

Figure 4.2 When effectiveness is relatively more important than cost.


86 F. Melese
would only coincidentally correspond to rankings obtained under the full informa-
tion constrained-optimization EEoA approach (where MoE is maximized subject
to the budget constraint).43
This is a damning result that clearly undermines the way MCDM is typically
applied to support AoAs. If there is no guarantee that the MCDM approach will yield
consistent results under full information (including affordability), then using this
criterion with less than perfect information (i.e. in the absence of explicit assump-
tions about future funding/budgets) is clearly problematic. In fact, GAO emphasizes
“[w]ith high levels of uncertainty . . . funding needs are often understated. . .” (GAO
2009, 9). A very real risk in applying the MCDM approach is that if AoAs
“fail to balance needs with resources [funding/budgets] . . . un-executable programs
[are allowed] to move forward, [and] program managers . . . are handed . . . a low
probability of success” (GAO 2009, 10).
In conclusion, the popular MCDM decision sciences approaches that underpin
most AoAs either ignore affordability and apply a benefit/cost (effectiveness/cost)
ratio criterion, or implicitly attempt to capture affordability through a relative weight
assigned to cost in a value function such as: Maximize V = V (MoE, Cost) = w1 ×
MoE − w2 × Cost.44 Again, to quote economists Hitch and McKean (1967):

One ubiquitous source of confusion is the attempt to maximize gain


[w1*MoE] while minimizing cost [w2*Cost] . . . If a person approaches a
problem with the intention of using such a [decision] criterion, he is con-
fused to begin with . . . [A] criterion in which the budget . . . is specified has
the virtue of being aboveboard.
[emphasis added] (pp. 165–7)

Rather than attempt to get a DM to reveal affordability concerns through


a weight assigned to costs (or prices) of alternatives, EEoA recommends a
more transparent approach—to treat “cost as an independent variable” (CAIV).
The relevant CAIV concept follows a definition posted on the Office of the
US Under Secretary of Defense (Acquisition and Technology) website in early
1999. It states that CAIV is the “DoD’s acquisition methodology of mak-
ing . . . performance a function of available budgeted resources” (see Lorell and
Graser 2001, 33).45 According to the Defense Acquisition Guidebook “all par-
ticipants . . . are expected to recognize the reality of fiscal constraints” (July 7,
2006, Section 3.2.4). This reflects a direct concern for affordability. The next
section illustrates how affordability can be explicitly incorporated in military
CBAs through the EEoA.

4.6 Six ways to structure an “economic evaluation of


alternatives”
There are six ways in which analysts and DMs can structure an EEoA. The first,
third, and fourth approaches are in the spirit of CAIV. It is also useful to distinguish
between: i) intra-program analysis approaches (1–5) and the ii) inter-program
Economic evaluation of alternatives 87
analysis approach (6). In the case of intra-program analysis, the DM associated
with the program is assumed to have sufficient information to be able to select an
alternative without reference to competing programs. That is not the case in inter-
program analysis, which requires a higher level “opportunity cost approach.” The
six EEoA approaches are shown in Table 4.1.
Two possibilities are highlighted within the intra-program analysis approach
outlined in Table 4.1. The first possibility is that DMs (analysts) are able to
construct/define/build alternatives (“endogenous alternatives”). The second possi-
bility is that alternatives are already constructed/defined/built (pre-specified) and
must simply be evaluated (“exogenous alternatives”). This section describes each
of the six EEoA approaches in some detail.46
A quote from Hitch and McKean (1967, 167) cited earlier highlights the first
two EEoA approaches: “[A] criterion in which the budget or level of effectiveness
is specified has the virtue of being aboveboard.” Starting with the first of the EEoA
approaches, the fixed budget approach, it is useful to recall another quote from
Hitch and McKean (1967, 167): “The test of maximum effectiveness for a given
budget seems much less likely to mislead the unwary.” [emphasis added]

4.6.1 Fixed budget approach


In his groundbreaking book, Cost Considerations in Systems Analysis, Fisher
(1971) describes the first approach to EEoA:

In the fixed budget case, the alternatives being considered are compared on
the basis of effectiveness likely to be attainable for the specified budget level
(p. 12). The analysis attempts to determine that alternative (or feasible com-
bination . . .) which is likely to produce the highest effectiveness.
(p. 10)47

Table 4.1 Six approaches to structuring an EEoA

I) INTRA-PROGRAM ANALYSIS
A) Build alternatives:
1. Fixed budget approach
2. Fixed effectiveness approach
3. Economic (expansion path) approach: (Construct alternatives as
cost-output/effectiveness relations or “response functions.”)
B) Modify existing alternatives: “Level the playing field”
4. Modified budget approach: GOTO 1.
5. Modified effectiveness approach: GOTO 2.
II) INTER-PROGRAM ANALYSIS
6. Opportunity cost/benefit approach
88 F. Melese

EEoA build alternatives


1. Fixed budget approach
Maximize effectiveness subject to budget constraint
(construct alternatives for given budget)
Benefit/cost
MoE(Utils) = MoE3/B*
A3
A2

Outsourcing opportunity: Effectiveness A1


Can we get more bang for the
same bucks?

B* Cost($)

Figure 4.3 Fixed budget approach.

Drawing on these observations, the fixed budget approach to EEoA leverages


Lancaster’s (1966a, 1966b, 1971, 1979) “characteristics approach to demand the-
ory.” Building on studies by Gorman (1980), Stigler (1945), Theil (1952), and
others who provided early foundations for the MCDM literature, Lancaster offered
economists (and defense analysts) a familiar way to analyze the consumer or
defense DM’s choice problem (to choose among alternative defense investments).
In Lancaster’s model, different vendors generate different bundles of char-
acteristics evaluated by DMs (“consumers”). Applying Lancaster’s model to
choose among alternative bundles of characteristics (say, computers) offered
by different vendors, defense DMs would maximize a utility function defined
over a desired set of multiple attributes or characteristics, subject to a budget
(funding/affordability) constraint.48 In this approach, the cost-effective alternative
is the one that, for a given budget (funding or affordability level), generates the
best mix of characteristics determined by the DM’s utility function.
The fixed budget approach illustrated in Figure 4.3 is the first of six ways pro-
posed to structure an EEoA. In Figure 4.3, the budget (funding or affordability
level) forecasted for the program is set at B*. The three vendor offers received for
this budget are A1, A2, and A3. Given its superior performance (i.e. MoE), vendor
A3 wins the competition which, in this case of a fixed budget, also corresponds to
the vendor with the highest benefit/cost ratio.49

4.6.2 Fixed effectiveness approach


The second way to structure an EEoA is the dual of the first: select the vendor that
minimizes costs of achieving a given MoE. Figure 4.4 offers an illustration. The
least cost vendor for a fixed MoE is A1, which also offers the greatest benefit/cost
ratio.50
A good example of the fixed effectiveness approach for structuring an EEoA
appears in US Federal guidance for public-private (“competitive sourcing” or what
Economic evaluation of alternatives 89

EEoA build alternatives


2. Fixed effectiveness approach
Dual: Minimize costs subject to effectiveness constraint
(construct alternatives for given MoE)

MoE(Utils) Benefit/cost
= MoE*/$B1

MoE* A1 A2 A3 Outsourcing opportunity:


Can we spend less bucks for
Effectiveness

the same bang?


OMB Circular A-76: MoE is defined
by the Statement of Work(SOW);
Invitation for Bid(IFB)

Cost($)

Figure 4.4 Fixed effectiveness approach.

the British call “market testing”) competitions conducted under OMB Circular
A-76, which “requires . . . a structured process for [evaluating] the most efficient
and cost-effective method of performance for commercial activities” (OMB May
29, 2003). This process, sometimes referred to as Lowest Cost Technically Accept-
able (LCTA), involves four steps: 1) develop a Statement of Work (SOW) or
Performance Work Statement (PWS) to define desired performance/effectiveness;
2) construct the Most Efficient Organization for the in-house competitor; 3) issue
an Invitation for Bid (IFB) for well-defined, routine commercial activities; and
4) compare bids or proposals, and select the “least cost.” This offers an illustration
of the fixed effectiveness approach: minimizing costs of achieving a given MoE
(e.g. defined by the SOW or PWS).51

4.6.3 Economic (expansion path) approach


In an earlier quote, Hitch and McKean (1967) strongly hint at the third, and most
general, approach to structure an EEoA:

The test of maximum effectiveness for a given budget seems much less likely
to mislead the unwary.
(p. 167)

As a starter, . . . several budget sizes can be assumed. If the same [alternative]


is preferred for all,. . . budgets, that system is dominant. If the same [alterna-
tive] is not dominant, the use of several . . . budgets is nevertheless an essential
step, because it provides vital information to the decision maker.
(p. 176) [emphasis added]
90 F. Melese
This third way to structure an EEoA provides the underlying foundation for the
first five EEoA approaches. It is described intuitively here and formally modeled
in Simon and Melese (2011).
The model involves a three-step process that includes multiple players. For ease
of exposition, assume three players: the military buyer and two competing private
vendors.
The first step is for the military buyer to publish a solicitation for bids (or
“request for proposals”). This solicitation states all significant non-price factors
(e.g. criteria, attributes, characteristics) that the agency expects to consider in
evaluating bid proposals, and is combined with an affordability assessment—
e.g. optimistic (B1), pessimistic (B2), and most likely estimates of the budget
(funding) for the program.
Once the solicitation is issued in the form of a Request for Proposal (RFP) or
IFB, interested vendors submit their offers and the selection process begins.52 Each
vendor is assumed to have different production and cost functions which they use to
build proposals with the desired attributes.53 Assuming the award is made without
discussions (pursuant to FAR 52.212-1 and 52.215-1), the military buyer employs a
(secret) scoring rule to rank vendors that is only revealed after award of the contract.54
The model is illustrated in Figure 4.5 and briefly described here. This third
EEoA approach describes the military buyer and competing vendors as follows:

Military buyer goal


Select an alternative that:
Maximizes MoE = Utility function = U (non-cost factors/attributes), subject to
BUDGET constraint.
Vendor goal
Select a mix of non-cost factors that:
Maximizes Q = Production function = Q (non-cost factors/attributes), subject to
sum of costs of attributes = c1 × a1 + c2 × a2 + · · · ≤ budget.
Military buyer requirements
MoE: Build-effectiveness model (non-cost factors: performance = quality, sched-
ule, etc.)
COST: Build-cost model (costs/prices: estimate total system life-cycle costs, total
ownership costs—see Chapter 5);
AFFORDABILITY: Estimate budget (forecast future funding available for the
program—see Chapter 7).
Private vendor requirements
Production function: possible attribute mixes given vendor-specific technology;
Total costs: vendor-specific costs of producing each attribute;
Vendor proposal constructed as a function of buyer’s revealed funding/budget
constraint.

This three-stage procurement auction briefly summarized in Table 4.2 is formally


modeled in Simon and Melese (2011).
EEoA build alternatives
3. Expansion path approach (Engel/response curves)
(Alternatives are cost-effectiveness relations, not points)
Explore impact of budget cuts/increases (Identify vendor responses)

Alternative 1
A1
MoE(Utils)
Alternative 2
A2
Effectiveness

“Knee of the Curve”

B2 B1 Budget ($)
Source Selection Decision: A2 for pessimistic budget; A1 for optimistic budget

Figure 4.5 Economic (expansion path) optimization approach.

Table 4.2 Three-stage multi-attribute procurement auction—economic (expansion path)


optimization approach (Simon and Melese 2011)

1) First stage: (cost as an independent variable – CAIV)


– The DoD provides notional budget guidance, (B), to possible vendors for the
program. DoD searches for the optimum product (procurement) and/or service
(R&D; O&M) package that it can obtain at that price, B. DoD also reveals
optimistic and pessimistic budget guidance.
– The DoD defines the set of characteristics/attributes that it values: This is known
to vendors. DoD’s precise utility function over those characteristics, however, is
unknown to vendors (secret scoring rule).
2) Second stage: (target costing)
– Vendors have different costs and production functions for generating the products
or services defined as bundles of characteristics.
– Each vendor maximizes its output offer (an optimal mix of the desired
characteristics) subject to their particular budget constraint (which includes
DoD’s budget guidance and the vendor’s individual costs to produce a unit of
each characteristic).
– This is the product and/or service package (output) a particular vendor is able to
propose for each possible budget, (B), given their production function (technical
production possibilities) and their vendor-specific costs of generating those
characteristics.
3) Third stage: (selection)
– With the latest budget forecast, DoD selects from the optimized characteristic
bundles proposed by each vendor the bundle/alternative (total product/service
package) that maximizes DoD’s utility function (or MoE).
92 F. Melese
This third and most general EEoA approach follows Hitch and McKean’s
(1967) earlier recommendation:

As a starter . . . several budget sizes can be assumed. If the same [alternative]


is preferred for all . . . budgets, that system is dominant . . . If the same [alterna-
tive] is not dominant the use of several . . . budgets is nevertheless an essential
step, because it provides vital information to the decision maker.
(p. 176)

This approach is illustrated in Figure 4.5 with two notional forecasted future
funding/budget levels: B1 (pessimistic) and B2 (optimistic).
The expansion paths for each vendor reveal combinations of attributes each ven-
dor can offer at different budget levels (e.g. pessimistic, optimistic, and most likely).
In Figure 4.5, each vendor’s expansion paths are transformed (through the govern-
ment’s utility function) into cost-utility or cost-effectiveness response functions (A1
and A2). These response functions reveal each vendor’s proposal under different
budget scenarios and represent the most general description of “alternative vendor
proposals” in the EEoA. Given a range of likely budgets for the program, the most
effective vendor over that range of budgets can be selected by the buyer.55
A key difference between traditional AoAs and this economic (expansion
path) approach is that instead of modeling competing vendors as points in
cost-effectiveness space, EEoA solicits vendor offers as functions of optimistic,
pessimistic, and most likely funding (budget) scenarios. This approach explicitly
addresses a key concern voiced by GAO:

A cost estimate is . . . usually presented to decision-makers as a . . . point esti-


mate that is expected to represent the most likely cost of the program but
provides no information about the range of risk and uncertainty or level of
confidence associated with the estimate.
(GAO 2009, 9)

Whereas the first three ways to structure an EEoA assume alternatives can be
constructed or built based on budget or effectiveness considerations (endogenous
alternatives), the last three approaches assume alternatives are exogenously deter-
mined, so that DMs must evaluate pre-specified alternatives. The most interesting
cases are where an alternative (vendor proposal) costs more but offers greater
utility (MoE), while others cost less and offer less utility (MoE).

4.6.4 Modified budget approach


Suppose alternatives are provided that have been developed exogenously—for
example, on the basis of a manpower or squadron constraint (one computer per
person, or a certain number of aircraft per squadron, etc.). If the overall budget or
desired level of effectiveness (MoE) for a program is not available, and analysts
and DMs have not structured the problem explicitly recognizing affordability, then
Economic evaluation of alternatives 93

EEoA: Modify existing alternatives


(“Level the playing field”)
4. Modified budget approach (GOTO 1 & 3)
Modify alternatives to equalize budget
(Identify vendor MoE responses to budget increase)
Revealed budget

MoE(Utils) A1*Cost-effectiveness relation


A1*

A2
Effectiveness

A1

Cost Budget ($)

Figure 4.6 Modified budget approach.

it is likely these pre-specified alternatives solicited from different vendors have


different costs and yield different MoEs.
The first step commonly prescribed in evaluating these alternatives is to create a
scatter plot of effectiveness versus cost (see Figure 4.6 and illustrations in chapters
throughout this book). In the absence of any other information, given two alter-
natives A1 and A2 that differ in both costs and effectiveness, leveling the playing
field requires the DM to determine the highest cost alternative he/she is willing to
consider (say the price of A2). This can then be used as an implicit budget for the
program. The fourth way to structure an EEoA recognizes that the highest-cost
(highest-utility) alternative DMs are willing to consider (A2 in Figure 4.6) reveals
their (implicit) budget constraint.
Then, in order to “level the playing field,” the DM asks the lower cost, lower
utility vendor (A1) how it might use the difference in funds (between their costs
and the high cost proposal) to increase the utility/MoE of their proposal (say from
A1 to A1*).56 Note that this effectively returns the problem to the first (and third)
way of structuring an EEoA—“maximize effectiveness (MoE) for a given budget.”

4.6.5 Modified effectiveness approach


Similarly, the fifth way to structure an EEoA levels the playing field for a spec-
ified choice of utility or effectiveness (MoE) revealed by one of the vendors.
This returns the problem to the second (and third) way of structuring an EEoA—
“minimize the cost of achieving a given level of MoE.”
For example, in Figure 4.7, anchoring the desired MoE at a target level such as
that revealed (offered) by vendor 2 (A2), the government asks vendor 1 how much
94 F. Melese

EEoA: Modify existing alternatives


(“Level the playing field”)
5. Modified effectiveness approach (GOTO 2 & 3)
Modify alternatives to equalize MoE
(Identify vendor COST responses to higher MoE requirement)

MoE(Utils)

A1* A2
Equal MoE
Effectiveness

A1

Cost
Budget ($)

Figure 4.7 Modified effectiveness approach.

it would cost to achieve the same target level of MoE.57 In Figure 4.7, vendor 1 is
preferred since the response (A1*) minimizes the budget required for the program.

4.6.6 Opportunity cost (or effectiveness) approach


Finally, suppose DMs find themselves in a situation where (due to legal, infor-
mation or other constraints): i) alternatives cannot be modified to obtain response
functions, and ii) future funding is unknown and a specific desired target level
MoE cannot be determined. In this case, it is likely some alternatives (bundles)
offered by vendors will cost more but offer more effectiveness, while others cost
less and offer less effectiveness. For example, consider Program A in Figure 4.8.
The sixth and final approach to structure an EEoA involves marginal analysis
and an inter-program comparison called the “opportunity cost approach.” Rather
than modify the alternatives to level the playing field as in the fourth and fifth
EEoA approaches, the opportunity cost approach requires a more challenging
inter-program analysis to choose between lower-cost, lower-effectiveness alterna-
tives (say A1 in Figure 4.8) and higher-cost, higher-effectiveness alternatives (A2).
The main challenge in selecting an alternative in this context is that the DM
must reach beyond the immediate program (A) into higher-level inter-program
considerations. If alternatives are exogenously determined and it is not possible
to level the playing field, then to find the most cost-effective solution requires
information about other competing programs (Program B in Figure 4.8).
This involves a higher-level inter-program analysis illustrated in Figure 4.8.
The DM must consider the incremental (marginal) loss in utility (MoE) in other
Economic evaluation of alternatives 95

EEoA: Inter-program analysis


6. Opportunity cost approach
A) Question: Where is the extra money coming from if I
buy the high cost alternative?
B) Question: Where is the extra money going if I buy the
low cost alternative?

MoE(Utils) Program A MoE(Utils) Program B

A2 B2

A1 B1

Cost($) Cost($)

Figure 4.8 Opportunity cost approach.

programs that must be sacrificed (e.g. a budget cut in Program B shifting the deci-
sion from B2 =>B1) for funds to be released to purchase greater marginal utility
(MoE) in the program under review (e.g. boosting the budget of program A to shift
the decision from A1 =>A2). “[T]he assessment should provide details as to how
excess funding . . . demands will be accommodated by reductions in other mission
areas, or in other . . . accounts” (DoD July 7, 2006, Section 3.2.2).
Alternatively, the DM can explore how much additional utility the extra money
might generate somewhere else if the low-cost alternative (A1 in program A)
was chosen. These are tough but useful questions that break through the sub-
optimization of most traditional AoAs. As a consequence, EEoA encourages
critical communication between different layers of the organization and a seam-
less interface between the requirements generation system, the acquisition system,
and PPBS.58
The bottom line is that it is often more transparent, efficient, and effective to
develop MoEs that are independent of costs. Equally important are theroles of budget
(funding) forecasts and opportunity costs in helping structure defense investment
decisions. Structuring an Economic Evaluation of Alternatives (EEoA) using one
of the six approaches summarized in Figure 4.9 can help achieve the primary goals
of defense acquisition, to lower costs and improve performance and schedules.

4.7 Conclusion: a decision map for decision-makers


This chapter has identified several major challenges that face a conventional military
cost–benefit AoA. It also critically examined key assumptions and decision crite-
ria typically used by the military to structure investment decisions. An alternative
set of military cost–benefit approaches is proposed. Based on microeconomic
Decision Map to Structure an Economic Evaluation of Alternatives (EEoA)

Dr. F. Melese
CAN YOU BUILD
Naval Postgraduate School
ALTERNATIVES?
[email protected]

NO:
Identify/Plot MOE &
YES
Cost/Budget of each
Alternative

IS THE DESIRED
MOE Can you Modify
SIMPLE TO Alternatives?
DEFINE/MEASURE?

NO: YES:
YES Do you have Level the NO
a BUDGET? Playing Field

Build Alternatives that Modify Alternatives Modify Alternatives Opportunity Cost


Yield Equal NO: Equalize Budget Equalize MoE Approach Build Inter-
YES:
Effectiveness i) PWS & RFP Choose desired Budget Choose desired MoE Program Evaluation
SOW & IFB (solicit Build Alternatives from list of vendors and fromlist of vendors and
ii) Build Vendors’ of Alternatives
prices from vendors Equal Budget let vendors compete let vendors compete Marginal Benefit,
that offerequal MoE) Response Functions
on MoE. on Cost. Marginal Cost

(3) Select (6) Select


(1) Select lowest (2) Select biggest Bang for the Buck (4) Select biggest (5) Select lowest Marginal Bang for
Buck bid Bang bid based on chosen Bang bid Buck bid the Buck relative
Budget or MoE to other programs

Figure 4.9 Decision map to structure an EEoA.


Economic evaluation of alternatives 97
foundations, the six alternative approaches to structure a cost–benefit analysis are
collectively referred to as the economic evaluation of alternatives (EEoA).
A significant weakness in the multi-criteria decision-making (MCDM) approach
that underpins many contemporary AoAs is that while they correctly focus on life-
cycle costs and the operational effectiveness of alternatives, “affordability” is an
afterthought, only implicitly addressed through a weight assigned to costs. In con-
trast, EEoA encourages analysts and DMs to embed affordability directly into AoAs.
This requires working with vendors to build investment alternatives based on differ-
ent funding (budget/affordability) scenarios. Supported by a static, deterministic,
multi-stage, constrained-optimization, microeconomic production (procurement
auction) model, EEoA explicitly addresses affordability (see Simon and Melese
2011).
The key difference between the traditional MCDM approach to AoAs and
the EEoA approach is that instead of modeling decision alternatives from com-
peting vendors as points in cost-effectiveness space, EEoA models alternatives
as functions of optimistic, pessimistic, and most likely funding/budget scenar-
ios. In demonstrating how to embed affordability directly into an AoA, EEoA
represents an important step in the long-running effort to achieve a significant
defense reform: to integrate the requirements generation system and the defense
acquisition system with PPBS to lower costs and improve performance.
The primary goal of this chapter was to demonstrate ways to improve defense
investments. The decision map illustrated in Figure 4.9 offers a summary guide
for analysts and DMs to select which of the six EEoA approaches is best suited
to their circumstances. Carefully structuring military cost–benefit analyses using
EEoA should improve defense decisions, leading to better use of scarce resources
and greater national, regional, and international security.

Appendix 4.1
A simple example helps to illustrate the danger in using benefit/cost ratios with-
out anchoring either the budget (here measured in dollars) or a specified measure
of effectiveness (MoE) (here measured in “utils”). Suppose Alternative A1 in
Figure A4.1 costs $10 million and yields an MoE of 10 utils, while Alternative A2
costs $1 billion and yields an MoE of 900 utils. Applying a benefit/cost ratio cri-
terion indicates that A1 has a bigger “bang-for-the-buck” since it returns 1 util per
million dollars, while A2 only offers 0.9 utils per million dollars. (In Figure A4.1,
the slope of any ray from the origin represents the (constant) benefit/cost ratio
anywhere along that ray: the steeper the slope, the greater the benefit/cost
ratio.)
Using benefit/cost ratios to rank alternatives, however, is dangerous in this
case since it ignores the absolute magnitude of the costs involved. Now suppose
the situation was reversed and that A2 offered a higher benefit/cost ratio than
A1. Anyone who chooses A2 in a simple ranking of alternatives strictly on the
basis of “bang-for-the-buck,” ignoring affordability, would be in for an unpleasant
surprise: a $1 billion vs. $10 million commitment!
98 F. Melese

Is A1 really superior to A2 ?
Lesson: Danger in using Benefit/cost (Bang/Buck) or
Cost/Benefit (Buck/Bang) ratios without anchoring Budget or MoE

RANKING:
MoE(Utils)
Benefit/Cost (A1) > Benefit/Cost (A2)

A2
Effectiveness

Efficient Set
A1
Marginal Benefit/Marginal Cost
“The perceived benefits of the
higher priced proposal shall
merit the additional cost…”
www.arnet.gov FAR 15.101-1(2)c

Cost

Figure A4.1 The inappropriate application of cost–benefit ratios.

Since affordability and opportunity costs are always a concern in public


investment decisions (and especially so through the interdependence of the
Requirements Generation System, DAS, and PPBS), it is imperative that ana-
lysts and DMs explore budget and opportunity cost implications of going with
the high-cost alternative (for example, the extra $990 million to obtain an addi-
tional 890 utils of MoE) or, equivalently, of the savings in going with the low-cost
alternative.

Notes
1 This study often uses the term “analysis of alternatives” (AoA) in its broad, generic sense.
Although focused on defense acquisition, the results of the study apply to any public-
sector procurement where benefits can be quantified but not monetized. It should be clear in
context whenever the term AoA references major defense acquisition programs (MDAPs)
or the acquisition of major automated information systems (MAISs).
2 For examples of the theoretical foundations of AoAs, see Clemen (1996), French
(1986), Keeney (1982, 1992, 1994), Keeney and Raiffa (1976), or Kirkwood (1997).
3 As quoted in E. Newell “Business group urges reform of Pentagon contract requirements
process,” Government Executive, www.govexec.com [last accessed July 27, 2009].
4 The US Government Accountability Office (GAO) emphasizes that a major challenge
facing the Department of Defense (DoD) is to “achieve a balanced mix of weapon
systems that are affordable” (GAO, 2009, 5). Section E1.1.4, “Cost and Affordabil-
ity” of US DoD Directive 5000.01, states: “All participants in the acquisition system
shall recognize the reality of fiscal constraints. They shall view cost as an independent
variable, and the DoD Components shall plan programs based on realistic projections of
the dollars and manpower likely to be available in future years . . . The user shall address
affordability in establishing capability needs” [The Defense Acquisition System, DoDD
5000.01 (certified current as of November 20, 2007)].
5 “Affordability means conducting a program at a cost constrained by the maximum
resources the Department can allocate for that capability” [quote from “Better Buying
Economic evaluation of alternatives 99
Power” Memorandum for Acquisition Professionals from Ashton Carter, Undersecre-
tary of Defense for Acquisition, Technology and Logistics (dated September 14, 2010)].
Available at www.acq.osd.mil/docs/USD_ATL_Guidance_Memo_September_14_2010
_FINAL.PDF [last accessed November 11, 2014].
6 Instead of modeling decision alternatives from competing vendors as points in cost-
effectiveness space, EEoA generates vendor proposals as functions of optimistic,
pessimistic, and most likely funding (resource/budget) scenarios.
7 The DoD uses the Planning, Programming, Budgeting And Execution (PPBE) process
as its principal decision support system to provide the best possible mix of forces, equip-
ment, and support within fiscal constraints. Two other major decision support systems
complement the PPBE process: a Requirements Generation System to identify military
investment opportunities and the Defense Acquisition System (DAS) to develop and
procure new weapon systems. In explicitly addressing affordability up front, EEoA pro-
vides a unique opportunity to achieve a significant acquisition reform to lower costs and
to improve performance and schedules by tightly integrating requirements generation
and defense acquisition within the PPBE.
8 Based on US strategic guidance (the National Security Strategy, National Military Strat-
egy, Quadrennial Defense Review, Strategic Planning Guidance, etc.), the Requirements
Generation System reviews existing and proposed capabilities and identifies critical capa-
bility gaps. To fill those capability gaps, senior leadership examines the full range of “doc-
trine, organization, training, materiel, leadership and education, personnel and facilities”
(DOTMLPF). (JCIDS 2007, A-1; DoD 5000.2 2008, 14) The DAS provides principles
and policies that govern major defense acquisition decisions and milestones. To ensure
transparency and accountability and to promote efficiency and effectiveness, various
instructions (e.g. FAR; DFARS; DoD Directive 5000.01; DoD Instruction 5000.02; etc.)
specify statutory and regulatory reports (e.g. AoAs) and other information requirements
for each milestone and decision point of major defense investments.
9 Major Defense Acquisition Program (MDAP) and Major Acquisition Information Sys-
tems (MAIS) proposals that emerge from the planning process enter the DAS and are
incorporated into the programming phase of PPBE.
10 Translating the budget implications of these decisions into the usual Congressional
appropriation categories [military personnel, procurement, operations and maintenance
(O&M), military construction, etc.] generates the nation’s defense budget and FYDP.
While DoD’s biennial defense budget projects funding only two years into the future, it
includes more financial detail than the six-year POMs.
11 Office of Management and Budget (OMB) Circular A-11 titled Preparation and Sub-
mission of Budget Estimates is the official guidance on the preparation and submission
of budget estimates to Congress. The Army’s acquisition guidance emphasizes “the
requirement for presenting the full funding for an acquisition program—that is the
total cost [for] a given system as reflected in the most recent FYDP [. . .] pertains to
all acquisition programs” (DoA July 15, 1999, 41).
12 Current DoD directives require that an AoA be performed at key milestone decision
points (i.e. A, B, C) for all MDAPs and MAISs. “Affordability Analysis should be con-
ducted as early as possible in a system’s life cycle so that it can inform early capability
requirements trades and the selection of alternatives to be considered during the Analy-
sis of Alternatives (AoA)” (USD (AT&L), November 25, 2013, 120). The EEoA offers
a mechanism to embed affordability assessments into AoAs.
13 “A 2008 DoD directive established nine joint capability-area portfolios, each managed
by civilian and military co-leads [. . .]. However, without [. . .] control over resources
[funding/budgets], the department is at risk [. . .] of not knowing if its systems are being
developed within available resources [funding/budgets]” (GAO 2009, 11).
14 “Typically, the last analytical section of the AoA plan deals with the planned approach
for the cost-effectiveness comparisons of the study alternatives” (DoD July 7, 2006,
100 F. Melese
Section 3.3). Note that there is no mention of “affordability,” but instead only an
ex-post cost-effectiveness tradeoff that implies a concern for affordability. Moreover,
this tradeoff occurs at the end of a process in which alternatives under consid-
eration have been developed independently of any cost/budget/funding/affordability
constraint.
15 One of the multiple cost analysis tasks prescribed in the US Air Force AoA Handbook
is to: “provide funding and affordability constraints. . .” and “describe the appropriate
CAIV [cost as an independent variable] methodology for the AoA” (US Air Force 2008,
32 and 49). Yet in its explanation of the final vendor selection process, the Hand-
book states “There is no formula for doing this; it is an art whose practice benefits
from experience” (p. 41). In sharp contrast, Figure 4.9 at the end of this chapter offers
explicit guidance to help analysts and decision-makers structure an EEoA with a focus
on affordability.
16 “Since this assessment requires a DoD Component corporate perspective, the
affordability assessment should not be prepared by the program manager nor should
it rely too heavily on the user. It requires a higher-level perspective capable of balancing
budget tradeoffs (affordability) across a set of users” (DAU, 2006, Section 3.2.2).
17 A first step in the program’s affordability assessment is to portray the projected annual
modernization funding (Research, Development, Test and Evaluation (RDT&E) plus
procurement, measured as Total Obligation Authority (TOA)) in constant dollars for
the six-year programming period and for twelve years beyond. Similar funding streams
for other acquisition programs in the same mission area also would be included. What
remains to be determined is whether this projected funding growth is realistically afford-
able relative to the DoD component’s most likely overall funding. This chapter proposes
structuring the EEoAs not only for a “most likely” budget, but also for an “optimistic”
(higher) and “pessimistic” (lower) budget.
18 These dual constrained optimization approaches represent the first two of six ways
proposed in this study to structure an EEoA. See Figure 4.9: A decision map.
19 A Senate report (Acquisition Advisory Panel 2007) states that “Awards are made on the
basis of the solicitation of factors and sub-factors by a Source Selection Official who,
using his or her discretion and independent judgment [e.g. guided by an AoA], makes
a comparative assessment of [. . .] competing proposals, trading off relative benefits and
costs” (Chapter 1, Commercial Practices, 65). The Senate Committee’s recommenda-
tion is that “Regulatory guidance [. . .] be provided in the FAR [Federal Acquisition
Regulations] to [include] a minimum weight to be given to cost/price” (p. 102). Missing
in this discussion is an explicit and realistic acknowledgement of “affordability”—the
resources, funding, or budgets available for the procurement—something only indirectly
and implicitly addressed in assigning a “weight” to cost.
20 The first and second EEoA approaches are the standard fixed budget or fixed effec-
tiveness techniques. In the former case, the recommended alternative offers the greatest
effectiveness for a fixed budget and in the latter case the lowest cost to achieve the
desired level of military effectiveness. More interesting and realistic cases involve deci-
sions among alternatives that offer somewhat greater (less) effectiveness for somewhat
higher (lower) costs.
21 This is in the spirit of the Army’s Acquisition Procedures, which explicitly states that
“Cost as an Independent Variable (CAIV) applies to all defense acquisition programs
[. . .and] treats cost as an input to, rather than an output of, the materiel requirements
and acquisition processes.” The Army guidance emphasizes “CAIV is focused on [. . .]
meeting operational requirements with a solution that is affordable [. . . and that does]
not exceed cost constraints [and to] establish CAIV-based cost objectives (development,
procurement, and sustainment costs) early in the acquisition process. Moreover, the
“RFP must [. . .] solicit from potential suppliers an approach [. . .] for meeting CAIV
objectives” (DoA July 15, 1999, 63).
Economic evaluation of alternatives 101
22 This offers an alternate approach to defense investment decisions based on explicit
funding (resource/budget/affordability) scenarios that supports the “. . . long-standing
DoD policy to seek full funding of acquisition programs . . .” [Defense Acquisi-
tion Guidebook, Chapter 3.23: “Full Funding, Defense Acquisition” University,
https://akss.dau.mil/dag/DoD5000.asp?view=functional (accessed April 18, 2009)].
23 The cost of many programs reviewed by the GAO exceeded planned funding/budget
levels (GAO July 2, 2008).
24 The inverse cost/benefit ratio is also used and sometimes calculated as a “cost-per-kill.”
For an example, see the case of the Javelin anti-tank missile in Chapter 14.
25 Fisher (1965) argues that “numerous terms [. . .] convey the same general meaning
[. . .] ‘cost–benefit analysis,’ ‘cost-effectiveness analysis,’ ‘systems analysis,’ ‘oper-
ations analysis,’ etc. Because of such terminological confusion, [. . .] all of these
terms are rejected and ‘cost-utility analysis’ is employed instead” (p. 185). Although
the terms “cost–benefit” and “cost-effectiveness” are used interchangeably here, the
assumption throughout is that neither “benefits” nor “effectiveness” can be measured
in monetary terms.
26 “[E]ffectiveness analysis . . . is built upon the hierarchy of military worth, the assumed
scenarios and threats, and the nature of the selected alternatives . . . .In many AoAs
involving combat operations [a] typical classification would consist of four levels: (1)
system performance, based on analyses of individual components of each alternative or
threat system, (2) engagement, based on analyses of the interaction of a single alterna-
tive and a single threat system, and possibly the interactions of a few alternative systems
with a few threat systems, (3) mission, based on assessments of how well alternative
systems perform military missions in the context of many-on-many engagements, and
(4) campaign, based on how well alternative systems contribute to the overall military
campaign.” (DAU, Section 3.3.3.5).
27 Ironically, if a budget scenario is specified, there is no need to take the MCDM approach
that underpins most AoAs since it is possible to adopt the EEoA approach. The EEoA
approach constructs alternatives to fit within a budget envelope, converting the problem
into a straightforward MOE maximization (see next section).
28 A fourth example involves a recent landmark RAND study on capabilities-based plan-
ning. The author falls into the same trap. In a section entitled “Choosing Among
Options in a Portfolio,” Paul Davis (2002) develops “A Notional Scorecard for Assess-
ing Alternatives in a Portfolio Framework,” where alternatives differ in both their costs
and effectiveness. Nevertheless, the decision criterion recommended by the author
to select an alternative in “[t]he last column is the ratio of effectiveness over cost”
(pp. 45–6).
29 The authors continue: “Of course, if the ratios did not alter with changes in the scale
of achievement (or cost, the higher ratio would indicate the preferred system, no matter
what the scale [. . .]. But to assume that such ratios are constant is inadmissible some of
the time and hazardous the rest” (Hitch and McKean 1967, 167).
30 The third (and most general) approach to EEoA follows another of Hitch and McKean’s
(1967) recommendations: “As a starter [. . .] several budget sizes can be assumed. If
the same [alternative] is preferred for all [. . .] budgets, that system is dominant [. . .].
If the same [alternative] is not dominant the use of several [. . .] budgets is never-
theless an essential step, because it provides vital information to the decision maker”
(p. 176).
31 An additional (necessary and sufficient) condition is a linear, separable, additive
objective function.
32 “Usually, ratios are regarded as potentially misleading because they mask important
information” (DoD July 7, 2006, Section 3.3.1).
33 “Measures of Effectiveness [. . .] provide the details that allow the proficiency of
each alternative in performing the mission tasks to be quantified [. . .]. A measure of
102 F. Melese
performance typically is a quantitative measure of a system characteristic (e.g. range,
etc.) chosen to enable calculation of one or more measures of effectiveness . . . The cost
analysis normally is performed in parallel with the operational effectiveness analysis. It
is equal in importance in the overall AoA process [. . .]. [I]ts results are later combined
with the operational effectiveness analysis to portray cost-effectiveness comparisons”
(DoD July 7, 2006, Section 3.3.1).
34 It is often implicitly asserted certain key assumptions are satisfied for this utility func-
tion to be valid, such as “additive independence” (see French 1986; Keeney and Raiffa
1976; and Keeney 1994).
35 For example, one issue is that normalization is not necessary, and worse, can be mislead-
ing. The author is aware of several applications where relative weights were assigned
to different attributes based on soliciting acceptable tradeoffs among measurable char-
acteristics from decision-makers, but then later those same weights were applied to
normalized values of the characteristics to obtain MOEs (personal correspondence with
DoD officials).
36 This corresponds to the first of the six EEoA approaches presented in this chapter—the
fixed budget approach.
37 “In the European Union, a legislative package intended to simplify and modernize exist-
ing public procurement laws was recently adopted. As before, the new law allows for
two different award criteria: lowest cost and best economic value. The new provisions
require that the procurement authority publishes ex-ante the relative weighting of each
criteria used when best economic value is the basis for the award” (see EC 2004a and
EC 2004b).
38 “An AoA is an analytical Comparison of the operational Effectiveness, suitability,
and Life-cycle Cost of Alternatives that satisfy established Capability needs” [Defense
Acquisition Guidebook, Chapter 3.3: “Analysis of Alternatives,” Defense Acquisition
University. https://akss.dau.mil/dag/DoD5000.asp?view=functional (accessed April 18,
2009)].
39 According to US Federal Acquisition Regulations (FAR), “source selection” is the deci-
sion process used in competitive, negotiated contracting to select the proposal that
offers the “best value” to the government: “In different types of acquisition, the rel-
ative importance of cost or price may vary” (General Services Administration 2005,
Section 15.101).
40 Surprisingly, the author has continued to write prolifically in this field and continued to
promote this decision criterion, apparently never taking the time to reflect back on these
key observations.
41 Note that the slope of the straight-line indifference curves that reflect the DM’s relative
preference (or tradeoffs) between effectiveness (MOE) and cost are given by the relative
weights assigned, the ratio –w2/w1.
42 The Army’s Economic Analysis (EA) Manual states that “good EA should go beyond
the decision-making process and become an integral part of developing requirements in
the PPBE process” (DoA February, 2001, 12).
43 The weight on cost in the unconstrained optimization (MCDM approach) roughly
corresponds to the Lagrangian multiplier (shadow price) of the budget constraint in
the constrained optimization (the EEoA approach). (See Chapter 10 for an alternative
interpretation.)
44 In a section describing Building a Model, Fisher (1965) comments: “Since by definition
a model is an abstraction from reality, the model must be built on a set of assump-
tions. These assumptions must be made explicit. If they are not, this is to be regarded
as a defect of the model design” (p. 190). It is easy to inadvertently conceal the impor-
tance of affordability (budget/funding) issues in the MCDM decision sciences approach
that underpins many AoAs. In sharp contrast, the EEoA approach encourages explicit
affordability (budget/funding) assumptions.
Economic evaluation of alternatives 103
45 The OMB Circular A-109 for Major Systems Acquisition mentions the goal of
“design-to-cost.” “Under the CAIV philosophy, performance and schedule are consid-
ered dependent on the funds available for a specific program.”
46 The technical model can be found in Simon and Melese (2011). The static, determin-
istic, multi-stage, constrained-optimization, microeconomic production (procurement
auction) model developed in that study underpins the third, and most general, approach
to EEoA, called the “expansion path approach.”
47 In a footnote, Fisher (1971, 10) adds: “[T]he fixed budget situation is somewhat analo-
gous to the economic theory of consumer [optimization] . . . For a given level of income
[budget/funding] the consumer is assumed to behave in such a way that he maximizes
his utility.”
48 Note that we refer to the usual deterministic “utility function” that is a conventional
term used in the economics literature. This is in contrast to the way a utility function
is typically defined in the decision sciences and operations management literature as a
stochastic function. The “value function” described in the latter literature is similar to
our “utility function,” except that costs can enter into a value function and are excluded
from our utility function since they appear as part of the budget constraint.
49 Note that in the first and second EEoA approaches, since either the budget (funding
level) or MOE (level of effectiveness) is anchored in the constrained optimization, the
benefit/cost ratio decision criterion can be used as a decision rule in the selection pro-
cess. The steeper the slope from the origin through an alternative (A1, A2, A3), the
bigger the “bang-for-the-buck.”
50 An example of this second approach to EEoA is RAND Corporation’s AoA study for
the KC-135 recapitalization program: “in this AoA, the most ‘cost-effective’ alternative
[fleet] means precisely the alternative whose effectiveness meets the aerial refueling
requirement at the lowest cost” (Kennedy et al. 2006, 7).
51 Title 10, Subtitle A, Part IV, Chapter 146, Sec. 2462 of the US Code reads: “A func-
tion of the Department of Defense . . . may not be converted . . . to performance by a
contractor unless the conversion is based on the results of a public-private competition
that . . . examines the cost of performance of the function by Department of Defense
civilian employees and the cost of performance of the function by one or more con-
tractors to demonstrate whether converting to performance by a contractor will result in
savings to the Government over the life of the contract” (January 3, 2007). This law con-
cerning public-private competitions offers another illustration of the fixed effectiveness
approach, minimizing the cost of achieving a given MOE.
52 Budget announcements are analogous to an agency exploring in order to uncover its
true reservation price for the product (given competing demands for scarce budgets).
Adoption of this approach of evaluating vendor proposals under different “reservation
prices” could eventually lead to greater use of fixed-price contracts. Note also that
although requiring vendors to offer multiple bids based on different budget scenarios
increases vendors’ bidding costs, the government’s ability to make a superior vendor
selection with the extra information should more than compensate for any incremen-
tal price increase charged by vendors or for extra costs the government might incur to
compensate bidders for their higher bidding costs.
53 The vendors’ constrained optimizations define distinct expansion paths, one for each
vendor (for example, two vendor expansion paths are illustrated in Figure 4.5). From
the Envelope Theorem, the Lagrangian multiplier in each vendor’s constrained opti-
mization problem reveals the marginal product [the extra output or attribute mix they
are capable of producing) if the military buyer (DoD) relaxes its funding constraint (i.e.
corresponding to a more optimistic budget)].
54 The buyer reveals desired attributes/characteristics of the investment to the sellers, but
not the relative importance weights, and requests a single offer from each seller based on
a pre-specified budget (affordability) constraint, and then he selects the one he prefers
104 F. Melese
among the submitted offers. “We call this procedure a ‘single-bid auction with secret
scoring rule”’ (Asker and Cantillon 2004, 1). A more general construct is presented in
Simon and Melese (2011).
55 For example, if the PPBS process ultimately results in narrowing the budget/expenditure
constraint for the program around B2, then alternative 2 (A2) is selected. Note that if the
planning process allows the optimistic budget assumption, B1, to persist, then A1 would
have been selected, but when reality struck and budget B2 was all that was available,
choosing A1 would have turned out to be highly inferior decision (see Figure 4.5). This
EEoA approach relies upon and reinforces the importance of closely coupled iterative
interactions between the requirements generation system, defense acquisition system,
and PPBS.
56 Alternatively, different valuable uses for the money saved by choosing the lower-cost
alternative could be brought into the effectiveness calculation. Some will recognize this
search for the “next best alternative use of funds” as the standard economic definition
of opportunity costs. This sets the stage for the sixth way to structure an EEoA.
57 Note that the response function (expansion path) for vendor 1 includes points A1 and A1*.
58 Fisher (1965, 182) quotes Secretary of Defense Robert McNamara: “Suppose we have
two tactical aircraft which are identical in every important measure of performance
[MOE] except one—aircraft A can fly ten miles per hour faster than Aircraft B. Thus,
if we need about 1,000 aircraft, the total additional cost would be $10 million. If
we approach this problem from the viewpoint of a given amount of resources, the
additional combat effectiveness . . . of Aircraft A would have to be weighed against
the additional combat effectiveness which the same $10 million could produce if
applied to other defense purposes—more Aircraft B, more or better aircraft munitions,
or more ships, or even more military family housing . . . This kind of determina-
tion is the heart of the planning-programming-budgeting . . . problem with the Defense
Department.”

Bibliography
Acquisition Advisory Panel. Report to the US Congress, Commercial Practices (Chap-
ter 1), Jan. 2007, pp. 31–157. Available at www.acquisition.gov/comp/aap/24102_
GSA.pdf [last accessed November 22, 2014].
Asker, J., and E. Cantillon. December 2004. Equilibrium in Scoring Auctions. FEEM
(Fondazione Eni Enrico Mattei) Working Paper No. 148.04.
Beil, D. R. and L. M. Wein. 2003. “An Inverse-Optimization-Based Auction Mechanism to
Support a Multiattribute RFQ Process.” Management Science 49(11): 1529–1545.
Byrns, Jr., E. V., E. Corban, and S. A. Ingalls. 1995. “A Novel Cost–Benefit Analysis
for Evaluation of Complex Military Systems.” Acquisition Review Quarterly, Winter:
1–20.
Carter, A. Undersecretary of Defense for Acquisition, Technology and Logistics (AT&L).
2010. Memorandum for Acquisition Professionals. Office of Secretary of Defense
(OSD), Washington DC.
Che, Y.-K. 1993. “Design Competition through Multidimensional Auctions.” RAND Jour-
nal of Economics 24(4): 668–80.
CJCS. May 2007. Joint Capabilities Integration and Development System (JCIDS) (CJCSI
3170.01F). Washington, DC: Author.
Clemen, R. T. 1996. Making Hard Decisions: An Introduction to Decision Analysis (2nd
edition). Belmont, CA: Duxbury Press.
Corner, J. L. and C. W. Kirkwood. 1991. “Decision Analysis Applications in the Operations
Research Literature, 1970–1989.” Operations Research 39(2): 206–219.
Economic evaluation of alternatives 105
Davis, P. K. 2002. Analytic Architecture for Capabilities-Based Planning, Mission-System
Analysis, and Transformation. Santa Monica, CA: National Defense Research Institute,
RAND.
Defense Acquisition University. 2006. Defense Acquisition Guidebook. Section 3.2.2:
AoA Study Plan—Affordability. Available at https://dag.dau.mil/Pages/Default.aspx
[last accessed November 18, 2014].
Defense Federal Acquisition Regulation Supplement (DFARS) and Procedures, Guidance,
and Information (PGI). Available at www.acq.osd.mil/dpap/dars/dfarspgi/current/ [last
accessed September 12, 2014].
DoA. July 15, 1999. Pamphlet 70–3. Washington, DC: US Army Headquarters.
——. February 2001. Economic Analysis Manual. Arlington, VA: US Army Cost and
Economic Analysis Center.
DoD. May 22, 1984. The Planning, Programming, and Budgeting System (PPBS) (DoD
Directive 7045.14). Washington, DC: Author.
——. April 9, 1987. Implementation of the Planning, Programming, and Budgeting System
(PPBS) (DoD Directive 7045.07). Washington, DC: Author.
——. November 7, 1995. Economic Analysis for Decision making (DoD Directive 7041.3).
Washington, DC: Author.
——. 1998. DoD Guide to Cost as an Independent Variable (DA PAM 70–3). Washington,
DC: Author.
——. July 7, 2006. Department of Defense (DoD) Defense Acquisition Guidebook. Chapter
1.1 (Integration of the DoD Decision Support System); Chapter 1.2 (Planning, Program-
ming, Budgeting and Execution Process; Office of Management and Budget (OMB)
and the PPBE Process). Available at https://acc.dau.mil/CommunityBrowser.aspx [last
accessed March 16, 2009].
DoDD. November 20, 2007. Department of Defense Directive Number 5000.01, USD
(AT&L) (SUBJECT: The Defense Acquisition System).
——. January 25, 2013. Department of Defense Directive Number 7045.14, USD (C)
(SUBJECT: The Planning, Programming, Budgeting and Execution Process).
DoDI. December 8, 2008. Department of Defense Instruction Number 5000.02, USD
(AT&L) (SUBJECT: Operation of the Defense Acquisition System).
Dyer, J. S., and R. K. Sarin. 1979. “Measurable Multiattribute Value Functions.” Operations
Research 27: 810–822.
Dyer, J. S., P. C. Fishburn, R. E. Steuer, J. Wallenius, and S. Zionts. 1992. “Multiple Crite-
ria Decision Making, Multiattribute Utility Theory: The Next Ten Years.” Management
Science 38(5): 645–654.
Edwards, W., D. von Winterfeldt, and D. Moody. 1988. “Simplicity in Decision Analy-
sis.” In Decision Making: Descriptive, Normative, and Prescriptive Interactions, edited
by D. Bell, H. Raiffa, and A. Tversky, pp. 43l–464. Cambridge, England: Cambridge
University Press.
Eger, R. J., and A. L. Wilsker. 2007. “Cost Effectiveness Analysis and Transporta-
tion: Practices, Problems and Proposals.” Public Budgeting and Finance 27(1):
104–116.
European Commission. March 2004a. Coordinating the Procurement Procedures of Enti-
ties Operating in the Water, Energy, Transport and Postal Services Sectors (Article 55 of
the European Commission Directive 2004/17/EC of the European Parliament and of the
Council of March 31, 2004).
——. March 2004b. On the Coordination of Procedures for the Award of Public Works
Contracts, Public Supply Contracts and Public Service Contracts (Article 53 and 54 of
106 F. Melese
the European Commission Directive 2004/18/EC of the European Parliament and of the
Council of March 31, 2004).
Ewing, P. L., W. Tarantino, and G. S. Parnell. 2006. “Use of Decision Analysis in the
Army Base Realignment and Closure (BRAC) 2005 Military Value Analysis.” Decision
Analysis 3(1): 33–49.
Federal Acquisition Streamlining Act of 1994 (Public Law 103–355).
Federal Acquisition Reform Act of 1995 (Public Law 104–106).
Federal Activities Inventory Reform Act of 1998 (Public Law 105–270).
Fisher, G. H. 1965. “The Role of Cost-Utility Analysis in Program Budgeting.” In Pro-
gram Budgeting, edited by D. Novick, pp. 61–78. Cambridge, MA: Harvard University
Press.
——. 1971. Cost Considerations in Systems Analysis. New York: American Elsevier
Publishing.
Fowler, B. W. and P. P. Cason. 1998. “The Cost Exchange Ratio: A New Aggregate
Measure of Cost and Operational Effectiveness.” Military Operations Research 3(4):
57–64.
French, S. 1986. Decision Theory: An Introduction to the Mathematics of Rationality (Ellis
Horwood Series in Mathematics and Its Applications). Cambridge: Ellis Horwood.
Gansler, J. S. June 2003. Moving Toward Market-Based Government: The Changing Role
of Government as the Provider (New Ways to Manage Series). Arlington, VA: IBM
Endowment for the Business of Government.
GAO. July 2, 2008. Defense Acquisitions: A Knowledge-Based Funding Approach Could
Improve Major Weapon System Program Outcomes (GAO-08-619). Washington, DC:
Author.
——. September 25, 2008. Defense Acquisitions: Fundamental Changes are Needed
to Improve Weapon Program Outcomes (GAO-08-1159T). Testimony of M. Sullivan,
Director of Acquisition and Sourcing Management before the Subcommittee on Federal
Financial Management, US Senate. Washington, DC: Author.
——. December 2008. Defense Logistics: Improved Analysis and Cost Data Needed to
Evaluate the Cost-Effectiveness of Performance-Based Logistics (GAO-09-41). Wash-
ington, DC: Author.
——. March 18, 2009. Defense Acquisitions: DoD must Prioritize its Weapon System
Acquisition and Balance them with Available Resources (GAO-09-501T). Testimony of
M. Sullivan, Director of Acquisition and Sourcing Management before the Committee
on the Budget, House of Representatives. Washington, DC: Author.
General Services Administration, Department of Defense, National Aeronautics and Space
Administration. March 2005. Federal Acquisition Regulation (FAR). Washington, DC:
Author.
Golabi, K., C. W. Kirkwood, and A. Sicherman. 1981. “Selecting a Portfolio of Solar
Energy Projects Using Multiattribute Preference Theory.” Management Science 27:
174–189.
Gorman, W. 1980. “The Demand for Related Goods: A Possible Procedure for Analysing
Quality Differentials in the Egg Market.” Review of Economic Studies 47: 843–856.
Henry, M. H. and W. C. Hogan. 1995. “Cost and Effectiveness Integration.” Military
Operations Research Society’s PHALANX, 28/1.
Hitch, C. J. and R. N. McKean. 1967. The Economics of Defense in the Nuclear Age.
Cambridge, MA: Harvard University Press.
Howard, R. 1988. “Decision Analysis: Practice and Promise.” Management Science 34(6):
679–695.
Economic evaluation of alternatives 107
Hwang, C.-L, and K. Yoon. 1981. Multiple-Attribute Decision Making: Methods and
Applications. New York: Springer-Verlag.
Keeney, R. L. 1982. “Decision Analysis: An Overview.” Operations Research 30(5):
803–38.
——. 1992. Value-Focused Thinking: A Path to Creative Decision Making. Cambridge,
MA: Harvard University Press.
——. 1994. “Using Values in Operations Research.” Operations Research 42(5): 793–813.
Keeney, R. L. and H. Raiffa. 1976. Decisions with Multiple Objectives: Preferences and
Value Tradeoffs. New York: John Wiley & Sons.
Kennedy, M., L. H. Baldwin, M. Boito, K. M. Calef, J. S. Chow, J. Cornuet, M. Eisman, C.
Fitzmartin, J. R. Gebman, E. Ghashghai, J. Hagen, T. Hamilton, G. G. Hildebrandt, Y.
Kim, R. S. Leonard, R. Lewis, E. N. Loredo, D. M. Norton, D. T. Orletsky, H. S. Perdue,
R. A. Pyles, C. R. Roll, Jr., W. Stanley, J. Stillion, F. Timson and J. Tonkinson. 2006.
Analysis of Alternatives (AoA) for KC-135 Recapitalization. Santa Monica, CA: Project
Air Force, RAND.
Kirkwood, C. W. 1997. Strategic Decision Making: Multiobjective Decision Analysis with
Spreadsheets. Belmont, CA: Duxbury Press.
Lancaster, K. J. 1966a. “A New Approach to Consumer Theory.” Journal of Political
Economy 74(2): 132–157.
——. 1966b. “Change and Innovation in the Technology of Consumption.” American
Economic Review 56(2): 14–23.
——. 1971. Consumer Demand: A New Approach. New York: Columbia University Press.
——. 1979. Variety, Equity, and Efficiency. New York: Columbia University Press.
Larsen, R. and D. Buede. 2002. “Theoretical Framework for the Continuous Early
Validation Method.” Systems Engineering 5(3): 223–241.
Liberatore, M. J. 1987. “An Extension of the Analytic Hierarchy Process for Industrial
R&D Project Selection and Resource Allocation.” IEEE Transaction on Engineering
Management 34(1): 12–18.
Lorell, M. and J. Graser. 2001. An Overview of Acquisition Reform Cost Saving Estimates.
Santa Monica, CA: RAND Project Ari Force.
Melese, F. May 17, 2007. “Outsourcing for Optimal Results: Six Ways to Structure an
Analysis of Alternatives.” In Proceedings of the 4th Annual Acquisition Research Sym-
posium (Creating Synergy for Informed Change). Monterey, CA: Naval Postgraduate
School.
Melese, F. and D. Bonsper. December 1996. “Cost Integration and Normalization Issues.”
Military Operations Research Society’s PHALANX 29(4).
Melese, F., R. Franck, D. Angelis, and J. Dillard. January 2007. “Applying Insights from
Transaction Cost Economics to Improve Cost Estimates for Public Sector Purchases:
The Case of US Military Acquisition.” International Public Management Journal 10(4):
357–385.
Melese, F., J. Lowe, and M. Stroup. 1997. “Integrating Cost and Effectiveness: An Eco-
nomic Perspective.” Military Operations Research Society’s PHALANX 30(3): 13–17.
Michael, R. T. and G. S. Becker. 1973. “On the New Theory of Consumer Behavior.”
Swedish Journal of Economics 75(4): 378–396.
Murray, C. H. 2002. Executive Decision Making (6th edition). Newport, RI: National
Security Decision Making Department, US Naval War College.
Office of Aerospace Studies. 2008. Analysis of Alternatives (AoA) Handbook. Air Force
Materiel Command, Kirtland AFB, New Mexico. Available at http://www.ndia.org/
Divisions/Divisions/SystemsEngineering/Documents/Forms/AllItems.aspx?RootFolder
108 F. Melese
=https%3a%2f%2f2.zoppoz.workers.dev%3a443%2fhttp%2fwww%2endia%2eorg%2fDivisions%2fDivisions%2fSystemsEngine-
ering%2fDocuments%2fCommittees%2fMission%20Analysis%20Committee%2fSupp-
ort%20Documentation&FolderCTID=0x0120003F6EF07406A5714EB4933D54BF88
1105 [last accessed November 19, 2014].
OMB. April 5, 1976. Major Systems Acquisitions (OMB Circular A-109). Washington, DC:
Author.
——. October 29, 1992. Guidelines and Discount Rates for Benefit-Cost Analysis of
Federal Programs (OMB Circular A-94). Washington, DC: Author.
——. May 29, 2003. Performance of Commercial Activities (Public-Private Competitions)
(OMB Circular A-76, Revised). Washington, DC: Author.
Parkes, D. C. and J. Kalagnanam. 2005. “Models for Iterative Multiattribute Procurement
Auctions.” Management Science 51(3): 435–451.
Parnell, G. S. 2006. “Value-Focused Thinking Using Multiple Objective Decision Analy-
sis.” In Methods for Conducting Military Operational Analysis: Best Practices in Use
Throughout the Department of Defense, edited by A Loerch and L. Rainey, Chapter 19.
Alexandria, VA: Military Operations Research Society.
Pinker, A., A. H. Samuel, and R. Batcher. December 1995. “On Measures of Effectiveness.”
Military Operations Research Society’s PHALANX, 28/4.
Quade, E. S. 1989. Analysis for Public Decisions (3rd edition). Englewood Cliffs, NJ:
Prentice Hall.
Ramesh, R. and S. Zionts. 1997. “Multiple Criteria Decision Making.” In Encyclopedia
of Operations Research and Management Science, edited by S. Gass, and C. Harris,
pp. 538–543. Boston, MA: Kluwer.
Retchless, T., B. Golden, and E. Wasil. 2007. “Ranking US Army Generals of the 20th
Century: A Group Decision-Making Application of the Analytic Hierarchy Process.”
Interfaces 37(2): 163–175.
Ruefli, T. 1971. “PPBS—An Analytical Approach.” In Studies in Budgeting, edited by R.
Byrne. Amsterdam: North Holland.
——. 1974. “Analytic Models of Resource Allocation in Hierachical Multi-Level Systems.”
Socio-Economics Planning Science 8(6): 353–363.
Saaty. T. L. 1977. “A Scaling Method for Priorities in Hierarchical Structures.” Journal of
Mathematical Psychology 15(3): 234–281.
——. 1980. The Analytic Hierarchy Process. New York: McGraw-Hill.
Simon, J., and F. Melese. 2011. “A Multi-Attribute Sealed-Bid Procurement Auction
with Multiple Budgets for Government Vendor Selection.” Decision Analysis 8(3):
170–179.
Smith, J. E. and D. von Winterfeldt. 2004. “Decision Analysis in Management Science.”
Management Science 50(5): 561–574.
Stigler, G. J. 1945. “The Cost of Subsistence.” Journal of Farm Economics 27: 303–314.
Theil, H. 1952. “Qualities, Prices and Budget Enquiries.” Review of Economic Studies 19:
129–147.
Thomas, C. and R. Sheldon. 1999. “The Knee of the Curve.” Military Operations Research
4(2): 17–24.
US Code. January 2007. Title 10—Armed Forces, Subtitle A, Part IV, Chapter
146, Section 2462. US Government Printing Office. Available at www.gpo.gov/
fdsys/granule/USCODE-2006-title10/USCODE-2006-title10-subtitleA-partIV-chap146
[last accessed November 19, 2014].
US Congress. 1997. Clinger-Cohen Act—Information Technology Management Reform Act
of 1996 (Public Law 104–106). Washington, DC: Chief Information Officers Council.
Economic evaluation of alternatives 109
USD (AT&L). November 20, 2007. The Defense Acquisition System (DoD Directive
5000.01). Washington, DC: Author.
——. November 25, 2013. Operation of the Defense Acquisition System (DoD Instruc-
tion 5000.02). Washington, DC: Author. Available at: http://www.acq.osd.mil/osbp/
sbir/docs/500002_interim.pdf [last accessed February 9, 2015].
Vazsonyi, A. December/January 1995. “Multicriteria Decision Making.” Decision Line
26(1): 13–14.
Willard, D. 1998. “Cost-Effectiveness, CAIV, and the Knee of the Curve.” Military
Operations Research Society’s PHALANX 31(3).
Winterfeldt, D. and W. Edwards. 1986. Decision Analysis and Behavioral Research.
Cambridge, England: Cambridge University Press.
Zahedi, F. 1986. “The Analytic Hierarchy Process—A Survey of the Method and its
Applications.” Interfaces 16: 96–108.
Zionts, S. 1980. “Methods for Solving Management Problems Involving Multiple Objec-
tives.” In MCDM: Theory and Applications, edited by G. Fandal, and T. Gaul,
pp. 540–558. New York: Springer-Verlag.
This page intentionally left blank
Part II

Measuring costs and


future funding
This page intentionally left blank
5 Cost analysis
Diana I. Angelis and Dan Nussbaum

5.1 Introduction
Military cost–benefit analysis (CBA) requires careful consideration of future
costs. Cost estimating is the process of collecting and analyzing historical data and
applying quantitative models, techniques, tools, and databases to predict the future
cost of an item, product, program or task. The term “cost analysis” is broadly used
to include not only the process of estimating (measuring) the cost of a project but
also the process of discovering, understanding, modeling and evaluating the rel-
evant information necessary to estimate the cost as well as the cost uncertainty
and risk.
Some might make a distinction between the cost estimating techniques and
analysis used for commercial off-the-shelf items (where a competitive market
determines the price), and products developed specifically for defense (where
there is no obvious market price). When comparing total life-cycle costs of alter-
natives, the techniques and practices presented in this chapter apply equally to
both cases. The only distinction is whether the acquisition price is determined by
the market or through contract negotiations.
It is important to note that cost analysis and estimating, as applied to finan-
cial management throughout government and industry, provides for the structured
collection, analysis, and presentation of life-cycle cost (LCC) data to assist in
decision-making for weapon systems capabilities throughout their useful life.
Examples of the components of LCC analysis conducted during each phase are:

• research and development (R&D)—complexity and innovation studies;


• production—performance, scale, and process studies;
• operations and support (O&S)—crew levels, training, fuel consumption
support, reliability and maintainability, and logistics studies;
• disposal—feasibility and tradeoff studies.

There are three main applications of cost estimating in the Planning, Program-
ming, Budgeting, and Execution System (PPBES) process: (1) preparation and
justification of budgets, (2) making choices among alternatives, and (3) source
selection (contracting). Descriptions of these applications are given below.
114 D. I. Angelis, D. Nussbaum
• Preparation and justification of budgets. Cost estimates are performed within
all phases of the acquisition process to determine the budget amounts required
to fund the program (or “funding requirements”) throughout its life-cycle.
• Making choices among alternatives. This is further subdivided into the
following areas:
◦ acquisition cost forecasting;
◦ analysis of alternatives (AoA).

• Source selection (contracting), through its phases of:


◦ acquisition planning;
◦ contract formulation;
◦ contract management and administration;
◦ contract close-out.

Cost estimates for the AoA requires forecasts of time-phased cash flows that
occur over the life of a system. This forms the basis for estimates of the total LCC
of a system. The LCC estimates are usually developed using constant dollars,1
and cost comparisons of alternatives are based on the present value2 of the total
LCC.
Cost estimating for budgeting assumes that the choice has been made and we are
interested in the near-term implementation cost of the chosen alternative. Budget
estimates must include the effects of expected inflation over the budget years and
are therefore stated in current (or nominal) dollars.
While the planning phase of PPBES is conducted with some resource con-
straints, it is not an intensive consumer of cost estimating support. The execution
phase is primarily focused on actual costs rather than estimated costs and the
difference between the two can be a source of much consternation, and is
one of the main reasons credible and accurate cost estimates are so important.
Additionally, the execution phase often uses the earned value management sys-
tem (EVMS), which has its roots in cost estimates, both to identify planned
vs actual cost variances, retrospectively, as well as to forecast cost variances,
prospectively.3

5.2 Cost estimating process


According to the International Cost Estimating and Analysis Association
(ICEAA), formerly called the Society of Cost Estimating and Analysis (SCEA),4
estimating is “the art of approximating the probable worth or cost of an activity
based on information available at the time.” An estimate is a judgment, opinion,
forecast or prediction. A cost estimate therefore is a prediction of the likely future
cost of a process, product, project, service, program or system. The Cost Estimat-
ing and Assessment Guide published by the Government Accountability Office
(GAO) provides a framework for understanding the cost estimating process as
displayed in Figure 5.1.
Initiation and research Assessment Analysis Presentation
Your audience, what you Cost assessment steps are The confidence in the point or range Documentation and
are estimating , and why iterative and can be of the estimate is crucial to the presentation make or
you are estimating it are accomplished in varying order decision maker break a cost estimating
of the utmost importance or concurrently decision outcome
Analysis, presentation, and updating the estimate steps
can lead to repeating previous assessment steps

Develop
Develop Develop
Develop
Develop Develop
Develop
Develop
Develop Develop
Develop
Develop Conduct a Present Update the
Define the Develop the Develop Develop
Develop Document
'-------"' Conduct risk and estimate to estimate to
estimate's estimating the
sensitivity uncertainty management reflect actual
purpose plan Develop the point estimate
Obtain analysis for approval costs/changes
estimate and compare
the
data it to an independent

cost estimate

Figure 5.1 Cost estimating process.


Source: GAO 2009.
116 D. I. Angelis, D. Nussbaum
In the defense acquisition process, a cost estimate is a prediction or forecast
of the complete costs of a complex program or weapon system, and it is often a
time-phased estimate. The first step in any cost estimate of complex systems is
to understand the attributes of the program or weapon system whose cost is to be
estimated.
Traditionally, this requires understanding and describing the weapon system
in terms of physical and technical parameters, operational and support concepts,
mission requirements, and interfaces with other systems. Understanding the pro-
gram’s schedule and acquisition profile is also important in developing a cost
estimate. The goal is to understand the relationship between key weapon systems’
attributes and cost. The next step is to develop an explicit framework for the cost
estimate.

5.3 Cost breakdown structure


A cost estimate is organized using a cost breakdown structure (CBS). The CBS
is a hierarchy that captures the totality of the inputs and activities required to
develop, produce, operate, and retire a system. A CBS is developed by subdividing
a product, process or service into its major cost elements and sub-elements. At the
highest level, the system CBS can be organized according to the life-cycle phases
of a system: R&D, investment, operating and support, and disposal. Under each
phase, a hierarchical structure documents the activities and resources required to
complete the work associated with that phase.
For example, the hierarchy for the investment phase (commonly referred to as
the work breakdown structure or WBS) defines at an aggregated level what is to be
procured and consists of at least three program levels with associated descriptions.
The WBS contains uniform terminology, definitions, and placement in the input-
oriented family tree structure. Guidance on work breakdown structures for US
Department of Defense (DoD) acquisitions is provided in MIL-HDBK 881.5 An
example of a program WBS is shown in Figure 5.2.
The level of detail in the WBS evolves as the system is defined and developed
throughout its life-cycle. After a program WBS has been approved, the contractor
will then extend a contract WBS to the appropriate lower levels to provide better

Aircraft
Level 1
System

Air Peculiar
Training Level 2
Vehicle Support
Equipment

Fire Com m u­ Equip­ Level 3


Receiver Services Depot
Control nication ment

Figure 5.2 Program WBS.


Source: US DoD 2005.
Cost analysis 117
definition of the complete scope of the contract. When integrated with the pro-
gram WBS, the extended contract provides a complete WBS for the acquisition
program. Similar to the economists’ production function, WBS reveals all the key
inputs required to generate the ultimate output/product. Further details on WBS,
including examples for various weapons classes, can be found in MIL-HDBK 881.
Operations and support (O&S) costs are also organized in a hierarchy (com-
monly known in the US DoD as a cost element structure or CES). The Cost
Analysis and Program Evaluation (CAPE) team in the Office of the Secretary of
Defense (OSD) defines standard O&S cost structures and cost elements that cover
the full range of O&S costs that could occur in any defense system in their Operat-
ing and Support Cost-Estimating Guide.6 An example of a cost element structure
is shown in Figure 5.3.
The final step is to develop a cost estimate. From the CBS perspective, the total
cost estimate of the system is obtained by adding up the cost estimates of the
individual elements (inputs) in the hierarchy across the levels of the CBS. To do
this, a cost estimate must be generated for each element of the CBS. The level of
detail of the cost estimate depends on the amount of information available about
the system. In early stages of development, only rough estimates are available of
the system parameters and attributes; therefore, only the highest levels of the CBS
cost can be estimated using “rough order of magnitude” or “top-down” techniques
such as analogy and parametric approaches.
It should be noted that parametric estimates are usually underpinned by the sta-
tistical technique of regression analysis, and they result in equations called cost
estimating relationships (CERs) that relate cost to underlying technical and per-
formance parameters.7 For example, the estimated cost of a new aircraft may be
modeled as a function of the weapon system’s key inputs and attributes, such as

Aircraft

Mission Unit Level


Maintenance Support
Personnel Consumption

Intermediate
Consumable Mat/ Sustaining
POL/Energy Other Maintenance
Repair Parts Support

Depot Indirect
Maintenance
Maintenance Support

Contractor
Operational
Support

Support

Figure 5.3 Cost element structure.


Source: OSD 1992.
118 D. I. Angelis, D. Nussbaum
empty weight, speed, useful load, wing area, power, and landing speed. This model
estimates the total cost of the aircraft based on a set of parameters independent of
the WBS structure.
As more information about the system becomes available, more detailed esti-
mates can be developed for lower levels of the CBS using a variety of techniques.
A cost estimate is now developed for each element at a given level of the CBS.
For example, the cost of the aircraft wing may be estimated using historical data
for weight, area, and materials. While this estimate is still based on the parametric
technique, the level of detail (a wing) corresponds to a CBS element. As the system
matures, the work and resource requirements are sufficiently well defined to use
“bottom-up” techniques such as engineering estimates and actual costs obtained
from prototypes and early production models. An engineering (production func-
tion) estimate would be based on the labor, material, and overhead required to
complete a particular element of the CBS. Obviously, this sort of estimation
requires a great deal of information about how the system is built—data that
is usually not available until early phases of production. Figure 5.4 shows the
relationship between estimating techniques and the evolution of a system. We will
have more to say about cost estimating methodologies later in this chapter.
While the CBS framework provides an excellent accounting system for cost
estimates, it does have a few drawbacks. First, although it captures the func-
tional relationship between inputs and cost elements, it does not explicitly show
potential correlations among those cost elements. If there is a correlation between
two cost elements that are assumed to be independent, the variability of the
cost estimate can be significantly increased. Second, the program CBS is input-
oriented, not relationship-oriented; it therefore largely overlooks transaction costs
such as search and information costs, decision and contracting costs, and monitor-
ing and enforcement costs (for example, see Melese et al. 2007). While it might be

GROSS ESTIMATES

DETAILED ESTIMATES

PARAMETrC System l
EXTRAPOLATION
FROM ACTUALS
Development
Demonstration I
&l

ANALOGY ENGINEERING

Concept Technology Production, Deployment,


Refinement Demonstration~ Operations & Support
:

Figure 5.4 Cost estimating techniques as a function of acquisition phases.


Source: Angelis et al. (2007).
Cost analysis 119
argued that these costs are implicitly considered in the cost estimates of the vari-
ous CBS components, it is likely that most are underestimated or ignored, thereby
possibly leading to overly optimistic cost estimates, that is to say, underestimating
future costs.

5.4 Functions of cost estimating


Cost estimates are designed to inform users about the cost realism, the cost valid-
ity, and the cost reasonableness of proposed alternatives, contracts, and budgets.
Initial descriptions of these functions are given next.
Cost realism helps to determine and communicate a realistic view of the likely
cost outcome, and it can form the basis of the plan for executing the work. Cost
realism also clarifies questions about technical competencies, technology matu-
rity, and budgetary defensibility. More specifically, cost realism addresses the
following questions.

• Is it clear from the cost estimate or proposal that the estimator/vendor


understands the technical requirements of the project?
• Is the cost proposal consistent with the technologies and technology readi-
ness levels (TRLs) described in the Cost Analysis Requirements Description
(CARD) and the technical proposal?
• Are the budgets that support the program defensible?

In the acquisition arena, cost validity supports decisions on program viabil-


ity, program structure, and resource requirements, while in the arena of “making
choices among alternatives,” cost validity measures the degree of confidence that
the government has in its cost estimates. Often, the metric is a level of uncertainty
such as standard error or one of the statistically grounded goodness of fit measures,
such as a correlation coefficient (R 2 ).
In the “source selection” arena, cost validity measures the degree of government
confidence in the vendor’s ability to perform within the proposed cost. This assists
in both the source selection process and in the engineering processes of design
and analyzing tradeoffs in the design phase. In both of these arenas, i.e. in the
acquisition arena, as well as in the “making choices among alternatives” arena,
cost reasonableness addresses two questions, namely:

• Is there a complete identification of cost drivers? The underlying thought is


that no cost estimate can be credible if it does not address all the elements of
cost that the project will incur.
• Is the proposed profit margin commensurate with the understanding of the
risk in the program? While the usual mechanism for adjudicating and bal-
ancing the inherent risk of the project between buyer (government) and seller
(vendor) is the contract type (e.g. cost plus fixed fee, firm fixed price, etc.),
an analysis of the vendor’s proposed profit margin can also provide insights
120 D. I. Angelis, D. Nussbaum
into whether the buyer and seller have a shared appreciation of the intrinsic
program risk.

5.5 Environment of cost estimating


In pursuing its objectives, the US DoD leverages large administrative and decision
systems: the requirements generation system, known as the Joint Capabilities Inte-
gration & Development System (JCIDS); the financial and resources management
system, known as PPBES; and the Defense Acquisition System (DAS). Cost esti-
mating, as described next, plays an important role in each of these three systems,
and, therefore, cost estimating plays a critical role in supporting administrative
and decision processes within DoD.
Cost estimating in the DoD is organized within either the financial manage-
ment/Comptroller or the engineering communities. The dual structure represents
historic legacies and philosophies. At the OSD level, the CAPE office was estab-
lished in 2009 as directed by the Weapons Systems Acquisition Reform Act of
2009.8 It is useful to think of the CAPE as the functional manager for cost esti-
mating processes within DoD. According to the Defense Acquisition University
website:9

The Director, CAPE is the principal staff assistant to the Secretary of Defense
for Cost Assessment and Program Evaluation. The Director’s principal
responsibilities include:
• analyzing and evaluating plans, programs, and budgets in relation to
US defense objectives, projected threats, allied contributions, estimated
costs, and resource constraints;
• reviewing, analyzing, and evaluating programs, including classified pro-
grams, for executing approved policies;
• providing leadership in developing and promoting improved analyti-
cal tools and methods for analyzing national security planning and the
allocation of resources;
• ensuring that the costs of DoD programs, including classified programs,
are presented accurately and completely;
• assessing effects of DoD spending on the US economy, and evaluating
alternative policies to ensure that DoD programs can be implemented
efficiently.

Each service has, as part of its Assistant Secretary for Financial Management
and Comptrollership, its own service cost center, as follows:

• Department of Army—Office of the Deputy Assistant Secretary of Army for


Cost and Economics (DASA-CE);10
• Department of Navy—Naval Center for Cost Analysis (NCCA);11
• Department of Air Force—Deputy Assistant Secretary for Cost and Eco-
nomics).12
Cost analysis 121
Within each service, there are cost estimating organizations located in the
acquisition communities. To a large extent, these organizations support program
managers and program executive officers in the PPBES and acquisition processes.
For example:

• In the Army Material Command there are cost estimating organizations within:
◦ Communications and Electronics Command (CECOM);
◦ Aviation and Missile Command (AMCOM);
◦ Tank and Automotive Command (TACOM).

• In the Air Force Systems Command, there are cost estimating organizations
within:
◦ Space and Missile Command (SMC);
◦ Aeronautical Systems Center (ASC);
◦ Electronics Systems Center (ESC).

• Within the Navy and Marine Corps, there are cost estimating organizations
within:
◦ Naval Air Systems Command (NAVAIR);
◦ Naval Sea Systems Command (NAVSEA);
◦ Space and Naval Warfare Systems Command (SPAWAR);
◦ Marine Corps Systems Command (MCSC).

5.6 Data collection and sources


Perhaps the most difficult part of developing a credible cost estimate is find-
ing reliable data to inform and support the estimate. Although the cost estimate
is a prediction about the future, that prediction relies on historical information,
including data about the physical, technical, and operational parameters of similar
systems and the cost of similar systems, efforts or components. In addition, other
information including prior experience with schedule, production, and operations
is used to support cost estimates.
The first problem is finding relevant historical data (if it exists). Once data has
been collected, the cost analyst must verify the validity of the information and
assess how it relates to the current system being estimated. To do this, the analyst
needs to have an understanding of what was included in the historical records.
The historical data needed for a cost estimate will likely come from a
number of different sources. Seldom is all the technical and cost data avail-
able in one location. Typical sources of cost data include accounting records,
cost reports (often required for earned value management), cost propos-
als, documentation associated with prior cost estimates, contracts and budget
documents.
Sources for technical data vary by the type of information required, but
often include acquisition documents, such as the CARD,13 system specifications,
122 D. I. Angelis, D. Nussbaum
contract documents, operational and maintenance records, vendor descriptions and
numerous other records and documents. In addition to system documents, cost and
technical data may be provided by subject matter experts, developed using spe-
cial studies or market analysis or even collected from different websites. Finding
data sources is a challenging process and can be disappointing when cost analysts
discover that desired data was never collected or recorded.
Once potential data sources are identified, the data must be collected. Extract-
ing and organizing data from different sources can prove to be a time-consuming
effort. Often historical records are not automated and the data must be manually
transferred to a computer for further analysis. Even when the records are digitized,
they may only be copies of original documents that must be reviewed individu-
ally to find and extract the relevant data. Often the data is contained in different
databases that must be combined to provide useful information. Unfortunately,
much of the historical cost and technical data collected by contractors to prepare
proposals for new weapon systems is considered proprietary and may be difficult,
if not impossible, to obtain.
In general, the acquisition cost collected in DoD databases is not standardized.
For a given category, the data collected may be inconsistent over time and across
programs. If there are changes in data categories, there is no mapping from the
old categories to the new categories. Differences between databases need to be
reconciled prior to using the data for cost estimates.
The US Defense Acquisition Management Information Retrieval (DAMIR) sys-
tem contains various data sources that the acquisition community uses to manage
acquisition programs and thus can be a useful source of technical, cost, and sched-
ule data. This database includes information on contract performance and program
cost from a variety of reports, such as Selected Acquisition Reports (SAR) and
Defense Acquisition Executive Summaries (DAES), as well as other reports.
The SAR is prepared by the program manager and documents estimates of the
program cost, schedule, and performance for major acquisition programs. The
DAES reports program assessments, unit cost, current estimates of the program
baseline parameters, exit criteria, and vulnerability assessments. Historically, pro-
gram managers only report on the six largest, currently active contracts (excluding
subcontracts) that exceed US$40 million in then-year dollars for a given pro-
gram; thus, the totals are moving targets. In addition, information reported in
this database does not necessarily track to the information reported for the same
program in the OSD budget.
Contractor cost data in the US is documented in Contractor Cost Data Reports
(CCDR) and includes the Cost Data Summary Report (DD form 1921, CDSR)
and the Functional Cost Hour Report (DD form 1921-1, FCHR). While there
are inconsistencies in reporting program costs from contract to contract and from
company to company, the reports are based on the program WBS, so the report-
ing is usually consistent over time for a given contract. The Defense Acquisition
Automated Cost Information System (DACIMS) houses all CCDR data collected
by the Defense Cost and Resource Center (DCARC).
Cost analysis 123
Information on operations and maintenance for major weapon systems is
maintained in historical databases managed by each service:

• Navy—Visibility and Management of Operating and Support Costs


(VAMOSC);
• Air Force—Air Force Total Ownership Cost (AFTOC);
• Army—Operating and Support Management Information System (OSMIS).

What are known as Budget Item Justification sheets in the OSD budget can also
provide useful information on cost and schedule for major weapon systems, albeit
at an aggregated level.
A critical step when using historical data is to first “normalize” the data before
using it to estimate future costs. Appendix 5.1 offers a comprehensive overview
of data challenges and the importance of data normalization in cost estimates.

5.7 Cost estimating methodologies


Estimating future costs is always challenging, and is one of the most difficult
tasks facing analysts. The three basic methods recognized for use within the
professional cost estimating community are:

• analogy—using costs of similar systems to determine the cost of the subject


system;
• parametric—generalized relationships between system characteristics and
costs;
• engineering (also called bottom-up or industrial engineering analysis).

These approaches rely on statistical properties and logical relationships, and are
largely based on historical data. One view of these three approaches is often heard
at cost estimating conferences, such as the one put on annually by the International
Society of Cost estimating and Analysis (ICEAA, formerly called the Society of
Cost Estimating and Analysis (SCEA)):

• “It’s like one of these” (Analogy)—subjectively compares the new system


with one or more existing similar systems for which there is accurate cost and
technical data.
• “This pattern holds” (Parametric)—sometimes known as the statistical
method, this technique generates an estimate based on system performance
or design characteristics. It uses a database of elements from similar systems.
It differs from analogy in that it uses multiple systems and makes statistical
inferences about the CERs.
• “It’s made up of these” (engineering)—a “bottom-up” method of cost analysis
that is the most detailed of all the techniques and the most costly to implement.
Each WBS element must be costed to build an estimate for the entire program.
124 D. I. Angelis, D. Nussbaum
Analogy-based cost estimating, which can also be called “It’s like one of these,”
subjectively compares the new system with one or more existing similar systems
for which there is accurate cost and technical data. With this approach, an ana-
lyst selects a system that is similar to or related to the system for which costs
are being estimated, and adjusts for differences between the two systems. This
approach works well for derivative or evolutionary improvements. Therefore, a
relevant starting baseline must exist to apply the method successfully. The anal-
ogy estimating approach is faster than the other two approaches, yielding more
immediate cost estimates.
The parametric relationships used for estimating costs are called cost estimating
relationships (CERs). According to the Parametric Cost Estimating Handbook,14
“a parametric cost estimate is one that uses CERs and associated mathematical
algorithms (or logic) to establish cost estimates.” Parametric methods are usually
based on a statistical technique that attempts to explain the correlation between the
dependent variable (typically cost) as a function of several explanatory variables
such as technical and performance characteristics, independent variables derived
from project parameters, such as:

• intended performance and in-service date (hedonic);


• overall system design characteristics (parametric);
• sub-system design characteristics (synthetic);
• inputs to program work packages (resource-based).

The relationships between these variables and cost are frequently determined
using a statistical technique, such as ordinary least squares (OLS) regression.
The flowchart in Figure 5.5 shows the steps taken in developing a CER.
An example, taken from the Constructive Cost Model (COCOMO)15 used to
estimate the cost of developing software, is:

MM = 2.4(KDSI)∗∗ 1.05

where MM = man-months of development effort and KDSI = thousands of


delivered source instruction.
This estimating equation indicates the expected number of man-months (MM)
of development effort for a programming project (the dependent variable),
depends exclusively on delivered source instructions in thousands (the indepen-
dent or explanatory variable). More specifically MM is equal to 2.4 times the
delivered source instructions (in thousands) to the power 1.05.
Parametric analysis has some advantages for estimating costs. These advantages
include:

• after the basic parametric estimating relationship has been defined, it is


straightforward to apply;
Cost analysis 125

Define Estimating
Collect
“Hypothesis”
“Relationship”
Data Evaluate and
Analyze Data
Normalize Data
for Candidate
Relationships

Perform Statistical
(Regression)
Analysis
Test
Relationships
Select Cost
Estimating
Relationship

Figure 5.5 Steps used to develop a cost estimating relationship.


Source: Nussbaum class notes, OA 4702, NPS 2013.

• the analyst is not required to be a technical expert; however, to apply the


method, the analyst must first obtain values for all the explanatory variables
(e.g. cost drivers);
• parametric relationships generated using OLS also provide information on the
uncertainty underlying the forecasted value. That is, the power of the underly-
ing statistical methodology permits a result of y ± e where e is related to the
error terms of the regression. This uncertainty value can be just as informative
as the predicted value.

The bottom-up approach relies on detailed engineering analysis to determine


an estimate. To apply this approach to estimate future aircraft engine production
costs, an analyst would need the detailed design and configuration information for
various engine components and accounting information for all material, equip-
ment, and labor. A conceptual engine design is built from scratch (hence the name
“bottom-up”). This approach generates a fairly detailed forecast.
One of the advantages of this approach is that many issues can be addressed, and
the effect of each issue can be well understood. For example, we could isolate the
effect of choosing a new material in construction or a new manufacturing method.16
The bottom-up method, however, has some drawbacks. First, the analysis is time
consuming. Often, a great deal of time must be spent generating the conceptual
design and corresponding cost estimate. Some companies have automated the pro-
cess by creating sophisticated database tools, but these systems can be expensive
to develop.
A second drawback of the bottom-up approach is that the analyst must be an
expert in the design of the technology being employed. Specific design details
126 D. I. Angelis, D. Nussbaum
Table 5.1 Advantages and disadvantages of cost estimating methodologies

Model Description Advantages Limitations


category

Analogy Compare project with Estimates are Truly similar projects


past similar projects based on actual must exist
experience
Expert Consult with one or Little or no Experts tend to be
judgment more experts historical data is biased; knowledge
needed; good for level is sometimes
new or unique questionable
projects
Bottom-up Individuals assess each Accurate estimates Methods are
component and then are possible time-consuming;
component estimates because of detailed data may
are summed to detailed basis of not be available,
calculate the total estimate (BOE); especially early in a
estimate promotes program; integration
individual costs are sometimes
responsibility disregarded
Parametric Perform overall Models are usually Models can be
models estimate using fast and easy to inaccurate if not
design parameters use, and useful properly calibrated
and mathematical early in a and validated; it is
algorithms program; they are possible that
also objective historical data used
and repeatable for calibration may
not be relevant to
new programs

Source: Nussbaum class notes, OA 4702, NPS 2013, [email protected].

must be considered to apply the method correctly. The user (i.e. cost estimator)
must also understand design tradeoffs and the current state of technology.
A third disadvantage is that the system must be well defined, i.e. there is little
allowance for unknown factors. For example, a component’s cost must be esti-
mated even though that component might represent a first-of-a-kind technology.
Finally, the user of the bottom-up approach must have access to, or maintain
an extensive and detailed database of, development, production, operating, and
support costs for the particular technology.
Table 5.1 summarizes the advantages and disadvantages of each approach. A
fourth approach, “expert judgment,” is included, which can be used when histori-
cal data is not available. Note, however, that there are no statistics associated with
an estimate based upon this method.

5.8 Cost uncertainty and risk


A cost estimate, by definition, is uncertain. That is, since we do not know the
future, we will have to make assumptions in order to estimate a future cost. A good
Cost analysis 127
cost estimate must therefore include information about the uncertainty surround-
ing the estimate. This requires the statistical analysis (using either actual data or
simulations) to estimate a distribution for the cost estimate, as well as a confidence
interval around the point estimate.
We arrive at a point estimate by evaluating the available data and calculating a
mean. This represents the most likely cost for a given project or program. The vari-
ability of the cost data provides the information necessary to build a confidence
interval which consists of the estimated mean cost plus or minus some number
of standard deviations. The number of standard deviations chosen depends on the
desired level of confidence. For example, we can state that we are 95 percent confi-
dent that a construction project will cost between US$4 million and US$6 million
dollars with the most likely cost being US$5 million dollars. In this case, our esti-
mate of the mean is US$5 million and the standard deviation (or standard error) is
US$0.5 million dollars. To build a 95 percent confidence interval, we use plus or
minus two standard deviations to arrive at our statement of cost uncertainty.
Statisticalanalysiscan also beused to assessthecostrisk.Firstwemustdefinewhat
we mean by “cost risk.” The standard quantitative risk assessment paradigm focuses
on thefollowing tripletof questionsarticulated byKaplanand Garrick(1981):

1 What can go wrong?


2 What are the associated likelihoods?
3 What are the consequences?

Once these critical questions have been answered, the greater challenge of risk
management is to address and control the following three issues articulated by
Haimes (1991):

1 What can be done and what options are available?


2 What are the trade-offs in terms of costs, benefits, and risks?
3 What are the impacts of current decisions on future options?

We will focus on the first set of questions when assessing cost risk. For cost
estimates, we pose the following three questions:

1 What can go wrong with the cost estimate?


2 What is the probability of a cost overrun?
3 What are the consequences of a cost overrun?

For a cost estimate, the answer to the first question is: the actual cost of the pro-
gram exceeds the program budget (a cost overrun). Note that while it is certainly
possible that the actual cost of the program is less than the budget, this is not a
“risk” since it is a desirable outcome. In addition, there are a number of reasons
why cost overruns are much more likely than cost underruns, as data and analysis
from RAND has shown (see Figure 5.6). The next section reviews these reasons
in examining biases in cost estimating.
128 D. I. Angelis, D. Nussbaum
16

14

12

10
Frequency

0
0.75–1.00 1.00–1.25 1.25–1.50 1.50–1.75 1.75–2.00 2.00–2.25 2.25–2.50
CGF range

Figure 5.6 Distribution of cost growth (difference between estimate and actual costs).
Source: Arena et al. (2006).

The answer to the second question (what is the probability of a cost overrun?)
depends on how much money is programmed in the budget. The real question is:
how much money should we put in the budget? We could start with the mean esti-
mate for the cost of the project. If we budget the mean and the cost distribution
is approximately normal, the risk of exceeding the budget (for any particular pro-
gram) is about 50 percent. In other words, the risk of a cost overrun is 50 percent,
not a very attractive number if you are the program manager.
If we budget less than the mean for each program in a portfolio of programs,
there is a higher risk of individual program cost overruns (not necessarily for the
total portfolio), but it permits funding more programs in a portfolio. Of course,
funding a program at such a high level of risk may imply to program managers
that a cost overrun is expected. It also reduces the amount of funds available for
contract negotiations, management reserves, and performance incentives.
We can reduce the risk of cost overruns if we budget an amount greater than the
mean. This will lower the risk of an individual program exceeding its budget, but
it may leave dollars on the table if actual costs are less than the amount budgeted.
Of course, this less risky approach means fewer programs can be funded in the
portfolio. It also provides less incentive to program managers to control costs, but
it increases the funds available for management reserves, contract negotiations,
and performance incentives.
The third question (what are the consequences of a cost overrun?) may provide
some insight into why defense programs are more often than not underfunded
(resulting in cost overruns). What are the consequences of a cost overrun in a
defense program? Cost overruns tend to result in either a reduction in the number
of units purchased (for example, cost overruns in the B-2 program forced the US
Cost analysis 129
Air Force to cut the number of aircraft procured from 132 to 21) or an extension
of the schedule, so that less units are procured each year. Understanding produc-
tion costs is essential to evaluating the impact of such actions. Often, cutting the
quantity of units produced increases the unit cost, resulting in less capability, as
fewer units can be purchased for a given budget.
Funding shortfalls may also lead to a reduction in scope or capabilities and
sometimes even program cancellations, but such outcomes are less frequent. Much
more likely is that one program will be cut to fund another (higher priority)
program to maintain a robust portfolio of capabilities. While some of these conse-
quences, such as cancellation of the program, are highly undesirable, historically
they have been rare and as such present less risk. As a result, the US DoD has been
willing to accept the risk of cost overruns, perhaps judging that the consequences
are manageable and thus not very risky.
It is common for CBA to focus on a comparison of the capabilities (effective-
ness) and total cost of each alternative. It is also useful to compare the cost risk of
different alternatives. To do this, we need the cost distribution for each alternative.
In some cases, one alternative dominates the other, meaning that regardless of the
funding (budget) level, the risk of a cost overrun is always lower for the dominant
alternative. In other cases, the comparison is not as clear. Consider the alternatives
A and B shown in Figure 5.7.
Alternative A has a lower cost risk up to a certain funding level (US$1,770),
and alternative B has a lower risk beyond that level. Figure 5.8 illustrates how risk
curves can be used to highlight this information and help decision-makers make
better choices.

4.5

4.0

3.5
Alternative A
3.0
Alternative B
Probability

2.5

2.0

1.5

1.0

0.5

0.0
0

2000

4000

6000

8000

10000

12000

14000

dollars

Figure 5.7 Cost distribution for two alternatives.


130 D. I. Angelis, D. Nussbaum

1.0

0.8
Alternative A
Alternative B
0.6
Probability

0.4

0.2

0.0
0

2000

4000

6000

8000

10000

12000

14000
dollars

Figure 5.8 Cost risk curves for two alternatives.

5.9 Cost estimating biases


The two factors commonly used to explain “unrealistic cost estimates” are bad
incentives (due to optimism bias (psychological explanation) or perverse bureau-
cratic incentives and strategic misrepresentation of budgets (political-economic
explanation)) and bad estimates (imperfect forecasting techniques such as an omit-
ted variable bias (methodological explanation)). It is helpful to examine these
factors in the context of the PPBE process used by the DoD to build the nation’s
defense budget.
In theory, “programming” involves a constrained optimization in which invest-
ment and operating decisions are made to maximize national security subject to
fiscal constraints. In the US, the Secretary of Defense (including CAPE and the
DoD Comptroller) and the Chairman of the Joint Chiefs of Staff have good rea-
sons to prefer accurate cost estimates given their global optimization perspective.
In contrast, given their sub-optimizing perspectives, Combatant Commanders and
the military services (for instance) may be more subject to a bias toward optimism
and to perverse bureaucratic incentives. Since their primary focus is on perfor-
mance and military capability, they have reason to be less critical (and realistic) in
their cost estimates.
According to McNicol (2005, S-4), “statistical analysis is consistent with
the well-established presumption that the military services tend to prefer [. . .]
optimistic procurement cost estimates.” Program Element Monitors (PEMs) in
the US, for example, act as advocates for the funding of their programs and
are responsible for defending those programs throughout the PPBE cycle. Since
all programs compete for the DoD’s limited resources in particular funding
Cost analysis 131
categories, PEMs are well aware that resource allocation is a competitive, zero-
sum game. Given their job descriptions, PEMs are drawn to lower cost estimates
that fit their advocacy agendas, rather than to higher cost estimates which make
defending their programs more difficult. Moreover, since they do not always pos-
sess a higher-level perspective, they may not fully appreciate how underestimating
costs can impact other programs.
Since this makes their programs more attractive to the DoD and the Congress,
defense companies also have good, strategic reasons at the beginning of a program
to underestimate costs. Once a program is accepted as a “program of record,” and
the DoD has accepted that the future capabilities associated with the program will
become available, the incentive to underestimate costs declines.
Given the complexity and uncertainty of many major, technologically advanced
weapon systems, contracts are necessarily incomplete; i.e. they cannot cover
all possible contingencies. As a consequence, the incumbent company can be
confident it will recoup overruns from multiple change-orders (generally for added
scope) anticipated over the life of the contract. In fact, firms have an incen-
tive to anticipate, but strategically omit, some of these elements in the original
contract negotiation. Moreover, by strategically hiring workers in key Congres-
sional districts, the company can pressure Congress to retain the contract and
approve compensation for cost overruns to preserve jobs. This combination of
strategic behaviors could explain some of the systematic bias in initial cost
estimates.
In an early attempt to address the factors that lead to biased cost estimates,
US Secretary of Defense Melvin Laird directed that independent parametric cost
estimating be made a part of the DoD acquisition process. Legislation to this
effect followed. The OSD launched an independent Cost Analysis Improvement
Group (CAIG) that began operations in the spring of 1972. Eventually, the CAIG
was assigned the statutory responsibility of providing independent life-cycle cost
estimates for all major weapon acquisition programs.
Whereas independent cost estimation is an important step that clearly attenuates
the impact of psychological and political-economic factors, the challenge of devel-
oping a more comprehensive cost estimating/forecasting methodology remains.
Cost estimating techniques must properly anticipate transaction costs such as mea-
surement, monitoring, management, and re-negotiation costs that may overwhelm
initial production cost estimates and which vary according to the nature of the
transaction. If transaction cost considerations are not properly anticipated, the
result will be that cost estimates will continue their downward bias, and programs
will continue to suffer cost overruns. One possible means to improve cost estima-
tion is to add transaction cost considerations to the current production cost focus
in cost estimating methods.

5.10 Summary
In this summary section, we provide recommendations on future research direc-
tions. In many disciplines, research is driven by both demand-side and supply-side
132 D. I. Angelis, D. Nussbaum
considerations. The demand-side represents the research which is done in response
to what users of the discipline are demanding, while the supply-side represents
those new techniques driven by researchers internal to the profession.
In cost estimating, the impetus for research is driven almost exclusively from
the demand-side, which encompasses changes in various ministries of defense
and military acquisition strategies. From this perspective, the most important
developments in the US are:

• the passage of the Weapon Systems Acquisition Reform Act (WSARA) in


2009;
• continued problems in estimating the costs of science and technology and
software;
• renewed emphases on affordability; and
• the ability to provide quick turnarounds of cost estimates in the early stages
of a program.

There are also some internal challenges within the professional community to
make sure that the next generation of cost estimators, with the right academic
background, are properly identified and developed in order to provide a better
level of service to the next generation of defense leadership.
Finally, there is the ever-present challenge of data availability. In this regard,
there have been some major improvements in data collection in the US over
the past 15 years, such as the development and maintenance of extensive data
collections like VAMOSC17 and DCARC18 (see Appendix 5.1). More recently,
OSD/CAPE has been working to consolidate all DoD cost data sources into the
Cost Assessment Data Enterprise, a unified initiative to collect, organize and use
cost data across the cost community. These and similar databases must be contin-
uously updated and maintained to support future cost estimates that are a critical
input in military CBA.

Appendix 5.1: Data normalization


Using historical data to predict future costs requires that the data not only be rel-
evant to the system under consideration, but that it match the assumptions of the
cost estimating model. For example, cost data is recorded in accounting records
based on transactions that occur over time. The cost of each transaction reflects the
buying power of the currency at the time the transaction was completed. Over time,
if a country is experiencing inflation or deflation, the buying power of the same
amount of currency changes. As a result, cost information contained in accounting
records incorporates inflation or deflation. These effects must be removed before
they are used in a cost estimating model; otherwise the cost estimate will be biased.
There are many other factors that can corrupt historical data and lead to
unreliable cost estimates. Data errors can arise from entry mistakes or software
problems. Numbers are transposed, incompletely recorded, entirely missing or
duplicated. In some instances, data is missing due to lack of documentation (so
Cost analysis 133
that it never enters the accounting system); in other cases, human or software
mistakes lead to incorrect data. Data that undergoes any kind of aggregation or
transformation is particularly subject to data errors. A simple check for possible
outliers in a data set is a good first step in cleaning up the data. This process may
reveal anomalies that require further investigation.
Anomalies in cost data can result from valid but unusual circumstances. For
example, the production process may be interrupted due to a shortage of materials
or a strike, leading to lower production rates and costs during a given period.
An accelerated delivery schedule may lead to higher prices for materials (due to
rush orders) and increased overtime expenses. A system modification may lead
to significantly higher maintenance costs in one period. Understanding how such
anomalies affect cost is critical to proper data normalization.
Data normalization is the process of making data recorded under different
circumstances comparable. The circumstances may include time, location, quan-
tity, technology, and operating tempo as well as many other variables that differ
between data points. As mentioned earlier, cost data collected over time must be
“normalized” to constant dollars to remove the effects of inflation or deflation.19
In addition, cost data collected in different time periods may differ due to changes
in accounting rules (related to what amount should be recorded or included in
the cost of an item) or changes in the accounting system or software (which, for
example, may lead to the same cost being recorded in a different category from
one period to the next).
Over time we can expect that production or operating processes will change,
perhaps becoming more efficient, and that materials and supplies used may change
and either increase or decrease the cost of production and operation. In addi-
tion, quantities produced or operated will likely change and the cost of inputs
such as labor, raw materials, and equipment will differ from period to period.
These differences must be accounted for prior to using the data for a cost
estimate.
One of the most important differences in costs over time in DoD arises from the
difference between peacetime and wartime operations. The higher wartime opera-
tional tempo consumes significantly more resources, which, in turn, leads to higher
costs. In addition, the attrition level is higher during wartime, and this may impact
both acquisition costs and operating and sustainment costs. These differences
should be recognized when estimating system life-cycle costs. Of course, pre-
dicting the timing of conflicts is an almost impossible task. For this reason, DoD
uses peacetime assumptions when preparing cost estimates for weapon systems
and budgets.
Often the cost of a new system is estimated using historical data from similar or
related systems. The systems will differ to some degree in terms of configuration,
materials, technology, missions, manning, and numerous other characteristics that
lead to differences in development, acquisition, and operational cost. While such
differences are unavoidable, it is the job of the cost analyst to adjust (or “nor-
malize”) data to account for the most significant differences to ensure costs are
comparable.
134 D. I. Angelis, D. Nussbaum
Notes
1 Constant (or real) dollars are amounts that reflect the buying power in a base year and
do not include inflation or deflation over time.
2 Present value analysis converts amounts paid or received in future years to an equivalent
amount paid or received today. See Appendix 5.1 for a more detailed discussion.
3 Earned value management (EVM) is a project management technique for measuring
project performance and progress by combining into a single system measures of scope,
schedule, and cost (adapted from Wikipedia).
4 See www.iceaaonline.com/
5 Department Of Defense Work Breakdown Structures For Defense Materiel Items. www.
acq.osd.mil/pm/currentpolicy/wbs/MIL_HDBK-881A/MILHDBK881A/WebHelp3/
MILHDBK881A.htm
6 Operating and Support Cost-Estimating Guide (OSD CAIG), Oct 2007. http://dcarc.pae.
osd.mil/reference/osd_ces/O_S_Cost_Estimating_Guide_Oct_2007.pdf
7 An application of ordinary least squares regression appears later in the chapter.
8 Text available at http://www.govtrack.us/congress/billtext.xpd?bill=s111-454, or
description available at
http://en.wikisource.org/wiki/Weapon_Systems_Acquisition_Reform_Act_of_2009
http://www.govtrack.us/congress/billtext.xpd?bill=s111-454)
9 See https://acc.dau.mil/CommunityBrowser.aspx?id=24463
10 See http://asafm.army.mil/offices/office.aspx?officecode=1400
11 See www.ncca.navy.mil/
12 See www.saffm.hq.af.mil/organizations/index.asp
13 See https://acc.dau.mil/CommunityBrowser.aspx?id=28858&view=w&lang=en-US
14 Parametric Cost Estimating Handbook. http://cost.jsc.nasa.gov/PCEHHTML/pceh.htm)
15 See http://sunset.usc.edu/csse/research/COCOMOII/cocomo_main.html
16 All representatives from engine producing companies interviewed during the course of
writing this chapter indicated that they employ some form of bottom-up approach to
estimating aircraft engine costs.
17 Visibility and Management of Operating Support Costs (VAMOSC). https://www.vam-
osc.navy.mil/
18 Defense Cost and Resource Center (DCARC). http://dcarc.cape.osd.mil/csdr/default.
aspx
19 For example, we can’t directly compare the nominal cost of fuel purchased by the US
Navy in 1990 (US$1.35 billion) and 1997 (US$1.52 billion). We must first delete the
effect of inflation since a barrel of fuel in 1990 cost US$30 while the same barrel of
fuel in 1997 cost US$38. We can adjust the 1997 cost of US$1.52 billion to reflect
1990 prices using a price index (38/30 = 1.27); thus, the cost in real dollars is US$1.20
billion. Now we can compare US$1.35 billion in 1990 to US$1.20 billion in 1997.

Bibliography
Angelis, Diana I., J. Dillard, R. Franck, and F. Melese. 2007. Applying Insights from Trans-
action Cost Economics (TCE) to Improve DoD Cost Estimation. Acquisition Research
Sponsored Report Series, NPS-GSBPP-07-008, Naval Postgraduate School.
Arena, Mark V., R. S. Leonard, S. E. Murray, and O. Younossi. 2006. Historical Cost
Growth of Completed Weapon System Programs. Santa Monica, CA: Rand Corporation.
Defense Acquisition University (DAU). Weapon Systems Acquisition Reform Act of 2009
(WSARA 09), DTM 09-027.
Fisher, Gene H. 2007. Cost Considerations in Systems Analysis. Santa Monica, CA: Rand
Corporation.
Cost analysis 135
Garvey, Paul R. 2000. Probability Methods for Cost Uncertainty Analysis. New York:
Marcel Dekker.
Government Accountability Office. 2009. Cost Estimating and Assessment Guide. Avail-
able at www.gao.gov/new.items/d071134sp.pdf
Haimes, Y. Y. 1991. Total Risk Management. Risk Analysis 11(2): 169–171.
Kaplan, S. and B.J. Garrick. 1981. On The Quantitative Definition of Risk. Risk Analysis
1: 11–27.
McNicol, D. L. 2005. Cost Growth in Major Weapon Procurement Programs (2nd edition,
IDA Paper No. P-3832). Alexandria, VA: Institute for Defense Analyses.
Melese, F., R. Franck, D. Angelis, and J. Dillard. 2007. Applying Transaction Cost
to Improve Cost Estimates for Public Sector Purchases: The Case of US Military
Acquisition. International Public Management Journal 10(4): 357–385.
Office of the Secretary of Defense. 1992. Cost Analysis Improvement Group Operating and
Support Cost-Estimating Guide. Available from www.dtic.mil/pae.
Shrull, D., ed. 1998. The Cost Analysis Improvement Group: A History. McLean, VA:
Logistics Management Institute (LMI).
Stewart, R. D. 1991. Cost Estimating (2nd edition). New York: John Wiley and Sons, Inc.,
Chapter 4.
US Air Force. 2007. Cost Risk and Uncertainty Analysis Handbook. Available at https://acc.
dau.mil/CommunityBrowser.aspx?id=316093
US Department of Defense. 2005. Work Breakdown Structures for Defense Materiel Items
(MIL-HDBK-881A). Available at www.everyspec.com.
6 Advances in cost estimating
A demonstration of advanced machine
learning techniques for cost estimation
Bohdan L. Kaluzny

6.1 Introduction

Defense cost analysis is an amalgam of scientific rigor and sound judgment.


It requires knowledge, insight, and the application of statistical principles, as
well as the critical interpretation of a wide variety of information known with
imprecision.
(Flynn and Garvey 2011)

6.1.1 Background
The International Cost Estimating and Analysis Association (ICEAA) Cost Esti-
mating Handbook of Knowledge (CEBoK)1 categorizes different costing tech-
niques commonly applied for defense cost estimation. These include costing by
analogy, costing by parametric approaches, engineering build-up approaches, and
extrapolation from actuals. Figure 6.1 shows when each of these costing tech-
niques are commonly applied relative to a program’s life cycle. The figure shows
the approximate proportional usage of cost techniques by phase. At the beginning
of a program there is more emphasis on using analogies and parametrics. As the
program matures, it becomes more defined, additional data is collected, and the
estimates get more detailed. A challenge to cost analysts is developing good cost
estimates during early design phases using costing by analogy and parametrics.
Cost estimation by analogy is typically accomplished by forecasting the cost of
a new system based on the historical cost of similar or analogous systems. This
requires a reasonable correlation between the new and historical system. The cost
of the historical system is adjusted by undertaking a technical evaluation of the
differences between the systems, deducting the cost of components that are not
comparable to the new design and adding estimated costs of the new components.
Usually subject matter experts are required to make a subjective evaluation of
the differences between the new system of interest and the historical system, and
subjectively chosen complexity factors are often used to adjust the analogous sys-
tem’s cost to produce an estimate. This subjectivity is a disadvantage of traditional
analogy methods.
Advances in cost estimating 137

Program Life Cycle


Phase A
Phases / C PhaseD Operations &
Concept Support
Design Build
Development

Program Life
Program Life
Program Life Program Life
Program Life

Gross Estimates Program Life

Figure 6.1 Defense Acquisition University (DAU) integrated defense acquisition,


technology, and logistics life cycle management framework chart.

Parametric approaches to cost estimation use regression or other statistical


methods to develop Cost Estimating Relationships (CERs). Strengths in paramet-
ric approaches are their potential to capture major portions of an estimate quickly
and with limited information, their basis on consistent and quantitative inputs,
and standard tests of validity, including a coefficient of correlation indicating
the strength of association between the independent variables and the dependent
variables in the CER. A major disadvantage of typical parametric approaches is
that they may not provide low-level visibility (cost breakdown) and changes in
sub-systems are not reflected in the estimate if they are not quantified via an inde-
pendent variable. CERs may be mathematically simple (e.g. a simple ratio) or
involve linear regression analysis (of historical systems or subsystems) on a sin-
gle or multiple parameters (weight, length, density, etc.). Miroyannis (2006) noted
that CERs are often insufficient and that other cost-driving factors must be incor-
porated to develop estimates of sufficient quality at the preliminary design phase.
Furthermore, the relationship between the parameter(s) and cost may not be best
expressed in linear form. While the field of regression analysis offers a multitude
of alternative approaches (see Ryan 1997), linear regression is the most popular
and easiest to understand.
This chapter explores recent advancements in costing by analogy and
parametrics that were presented in the context of estimating the development
and production costs of naval ships by the North Atlantic Treaty Organization
(NATO) Research Technology Board (RTO) System Analysis and Studies (SAS)
task group 076 on “NATO Independent Cost Estimating and the Role of Life Cycle
Cost Analysis in Managing the Defence Enterprise.” The NATO RTO SAS-076
Task Group (2012) report and associated publications (Kaluzny et al. 2011;
Kaluzny and Shaw 2011) detail the data collection and normalization efforts and
provide cost estimation results. The primary focus of this chapter is to highlight
and further explore the methods published in the latter references. For this reason
138 B. L. Kaluzny
actual costs (data, normalization, results, etc.) are not discussed and only appear on
occasion in figures and tables to facilitate understanding of the methods.

6.1.2 Leveraging machine learning


Witten and Frank (2005) define data mining as the process of discovering pat-
terns, automatically or semi-automatically, in large quantities of data. Data mining
often employs machine learning methods whose focus is on prediction, based on
known properties learned from data. In this chapter we define machine learning
techniques as mathematical algorithms for “best-fit” prediction. The application
of data mining and machine learning techniques to cost estimation, or cost pre-
diction, is natural and not new—regression analysis is a predominant machine
learning technique utilized by cost analysts. However, advancements in the fields
of data mining and machine learning have brought forward additional tools that
can be leveraged to improve cost estimation. In this chapter two machine learning
algorithms are discussed in detail and exemplified:

• A combination of hierarchical cluster analysis, principal component analy-


sis, and nonlinear optimization can be used for a cost estimation by analogy
approach that is void of subjective adjustment factors. The approach also con-
siders multiple analogous systems rather than just one. Hierarchical cluster
analysis identifies the historical systems that are the “nearest neighbors” to the
new system. A hierarchy of clusters grouping similar items together is pro-
duced: from small clusters of very similar items to large clusters that include
more dissimilar items. A matrix of distances (measure quantifying the sim-
ilarity of two systems) among systems is calculated expressing all possible
pairwise distances among them. These distances are then used to predict the
cost of a new instance.
• Using an algorithm combining decision trees and regression analysis, a para-
metric approach for cost estimation incorporating a multitude of cost-driving
factors, while remaining a “top down” approach applicable in early design
phases of the procurement cycle, is described. The Quinlan (1992) M5 model
tree algorithm combines features of decision trees with linear regression mod-
els to both classify similar systems (based on attributes) and build piecewise
multivariate linear regression models.

6.1.3 Outline
The remainder of this chapter is structured as follows. Section 6.2 provides a very
brief description of the dataset used to exemplify the machine learning algorithms
discussed herein. Section 6.3 presents a machine learning algorithm for costing by
analogy and Section 6.4 exhibits the application of M5 model trees to parametric
cost estimation. Section 6.5 provides a combined discussion. Finally, the chapter
concludes in Section 6.6.
Advances in cost estimating 139
6.2 Data
Data mining approaches are data intensive. Both the quality and quantity of the
data drive the quality of the outputs of machine learning algorithms. While the
appropriate level of data quality and quantity will be case-specific, to illustrate
the application of machine learning algorithms a database of 59 naval ships com-
piled by the NATO RTO SAS-076 working group is considered. This dataset
is subsequently referred to as the SAS-076 ship dataset. The database captures
over 100 descriptive, technical, and cost attributes for 59 naval ships from 18
classes (of ships), from seven nations, commissioned between 1950 and 2010.
The normalized costs also span multiple orders of magnitude.
The SAS-076 ship dataset contains military and civilian auxiliary (coast guard
or similar) vessels that were judged to be more or less similar to Landing Platform
Dock (LPD) ships. A benefit of employing machine learning algorithms is that
they allow for, or can excel on, a greater variability in the input dataset—variability
that could be questioned when using traditional costing by analogy or parametric
approaches. The fidelity of costing by analogy and parametric approaches also
depends on the size of the input dataset. In this case, the availability of ship cost
data limited the size of the SAS-076 ship dataset. While the decision as to which
ships were or were not included in the dataset was subjective, it was primarily
driven by the availability of cost data.
Using machine learning terminology, each ship is an instance and together the
ships listed in the SAS-076 dataset form a set of instances for input to a machine
learning scheme. Instances are things that are to be classified, associated or clus-
tered. Each instance is characterized by the values of a set of predetermined
features or attributes. The matrix of instances versus attributes forms the input
dataset (e.g. the SAS-076 ship dataset). The value of an attribute can be numeric,
nominal, ordinal, interval or ratio. Numeric attributes measure numbers (real or
integers) and are not necessarily continuous. Nominal, or categorical, attributes
take on values from a prespecified, finite set of possibilities. These can take the
form of distinct symbols or even words like “yes” or “red.” Ordinal quantities are
nominal quantities with a rank order (e.g. fast > slow > very slow). Interval quan-
tities are ordinal but also measured in fixed and equal widths (e.g. temperature in
degrees Celsius). Ratio quantities are relative; specified by an inherent zero-point
(e.g. the distance between two objects).
Descriptive, technical, and cost data were gathered for each of the ships and
broken down into categories, as shown in Table 6.1. Attribute units are expressed
by either nominal values (e.g. “fixed pitch,” “controlled pitch,” “yes,” “no”), or
by numerical units such as meters, megawatts, knots, hours, nautical miles, etc.
Unknown or missing data is permissible and is denoted by empty or “?” entries.

6.3 Data mining for cost estimation by analogy


This section describes how hierarchical cluster analysis is used for a cost estima-
tion by analogy approach that is void of the subjectivityinherent (of traditional
140 B. L. Kaluzny
Table 6.1 Categories of ship data

Category Number of attributes

I Description 6
II Construction 8
III Dimensions 5
IV Performance 8
V Propulsion 9
VI Electrical power generation 3
VII Lift capacity 35
VIII Flight deck 19
IX Armament 13
X Countermeasures 5
XI Radars/sonars/sensors 13
XII Combat data systems 1
XIII Weapons control systems 1
XIV Other capabilities 7
XV Cost data 3
Total 136

analogy approaches) in quantifying the cost of the technical and other differences
between the historical system and the new system. The approach also considers
multiple analogous systems rather than just one.

6.3.1 Hierarchical cluster analysis


Hierarchical cluster analysis is a data mining approach that facilitates cost esti-
mation by analogy by identifying the systems that are the “nearest neighbors”
to the new system. Hierarchical cluster methods produce a hierarchy of clusters
grouping similar items together: from small clusters of very similar items to large
clusters that include more dissimilar items. In particular, agglomerative hierar-
chical methods work by first finding the clusters of the most similar items and
progressively adding less similar items until all items have been included into a
single large cluster. Hierarchical agglomerative cluster analysis begins by calculat-
ing a matrix of distances among systems expressing all possible pairwise distances
among them. Initially each system is considered a group, albeit of a single item.
Clustering begins by finding the two systems that are most similar, based on the
distance matrix, and merging them into a single group. The characteristics of this
new group are based on a combination of the systems in that group. This proce-
dure of combining two groups and merging their characteristics is repeated until
all the systems have been joined into a single large cluster.
Hierarchical cluster analysis is a useful means of observing the structure of the
dataset. The results of the cluster analysis are shown by a dendrogram (tree), which
lists all of the samples and indicates at what level of similarity any two clusters
were joined. The x-axis is a measure of the similarity or distance at which clusters
join. The resulting clustering can be used to estimate the cost of a new system by
Advances in cost estimating 141
taking a weighted average of the cost of historical systems based on the relative
distances between the new system and the historical systems.

6.3.2 Application to SAS-076 ship dataset


Using the SAS-076 ship dataset, hierarchical clustering is used to define a ship
distance function which takes as input ship attributes for a pair of ships and out-
puts a single value indicating the distance, or similarity, between the two ships.
Formally, define

dijk = distance between ship i and j with respect to attribute k. (6.1)

For numeric attributes, dijk is normalized to lie in the [0, 1] range with dijk = 1 indi-
cating that ships i and j lie at opposite ends of the observed spectrum for attribute
k (e.g. shortest and longest length ships), and dijk = 0 indicating that the ships are
the same with respect to attribute k. For nominal attributes, dijk is binary—set to 0
if ship i and j are the same with respect to attribute k, and 1 if they are not.
A variety of distance metrics can be used to calculate similarity of two ships
based on the attribute distances dijk . Using a simple Euclidean distance metric, the
aggregate distance between two ships i and j , is expressed as

di j = di2j k , (6.2)
k∈A

where A is the subset of attributes considered.


Computing the distances between all pairs of ships using Equation (6.2), the
cost of a ship i can be estimated by computing the weighted-average cost of the
other ships. Let C j be the known cost of ship j , then

 Cj 1
C̃i = · 1 (6.3)
d2
j  =i i j di2j
j  =i

is the estimated cost of ship i . Figure 6.2 plots the actual ship costs, Ci , versus the
costs predicted, C̃i , by the analogy method using hierarchical clustering based on
a simple distance metric. The solid line running at 45 degrees represents perfect
prediction. The measures of worth of the analogy method using hierarchical clus-
tering analysis (simple distance metric) for predicting ship costs are R = 0.48 and
R 2 = 0.23. The mean absolute percentage error is 49 percent indicating that this
raw approach does poorly in learning the known ship costs and would likely yield
poor cost predictions.
An assumption in the above approach is that all attributes are of equal impor-
tance; the (normalized) differences or similarities for each attribute contribute
equally to the measure of similarity between ships. To potentially improve the
142 B. L. Kaluzny

700

600
Predicted Cost

500

400

300

200

100

0
0 100 200 300 400 500 600 700
Actual Cost

Figure 6.2 Hierarchical clustering with simple distance function: correlation plot of
actual versus predicted ship costs.

predictive capability of the method, attribute weights can be defined. Let

wk = weight of attribute k. (6.4)

Using a weighted Euclidean distance metric, the aggregate distance between two
ships i and j , is expressed as

d̂i j = (wk · di j k )2 , (6.5)
k∈A


where k∈A wk = 1 and wk ≥ 0 for all k. As before, let C j be the known cost of
ship j , then

 Cj 1
Ĉi = · 1 (6.6)
2
j  =i d̂i j d̂i2j
j  =i

is the estimated cost of ship i using weighted attributes. The optimal allocation of
weights is determined by minimizing the prediction error for the known ships,
 2
minimize Ci − Ĉi . (6.7)
i

The resulting mathematical optimization is a nonlinear convex program. With


a large set of attributes the mathematical program may be too computationally
Advances in cost estimating 143
intensive to solve in a reasonable amount of time. To reduce the dimensional-
ity of the problem, a smaller subset of attributes can be selected. While there
are numerous attribute selection algorithms (see Witten and Frank 2005), prin-
cipal component analysis (PCA), a tool in exploratory data analysis and for
making predictive models, is recommended. PCA involves a mathematical pro-
cedure that transforms a number of possibly correlated variables into a smaller
number of uncorrelated variables called principal components. The first princi-
pal component accounts for as much of the variability in the data as possible,
and each succeeding component accounts for as much of the remaining variabil-
ity as possible. The reader is referred to Jolliffe (2002) for mathematical details
of PCA.
Using PCA, the ship dataset was reduced to a set of 16 macro-attributes account-
ing for 95 percent of the original set’s variability. Each macro-attribute is a linear
combination of the original attributes. For example, macro-attribute A2 is

A2 = 0.204 × length
+ 0.196 × beam width
+ 0.183 × vehicle space
+ 0.18 × no. of expeditionary fighting vehicles
+ 0.165 × 1 if has a well deck, otherwise 0
+ 0.165 × width of the well deck
+ 0.164 × length of the well deck
+ 0.159 × no. of large personnel landing craft
(6.8)
+ 0.156 × no. of Chinook helicopters supported
+ 0.155 × full load displacement
+ 0.153 × no. of combat data systems
+ 0.150 × light load displacement
+ 0.144 × well deck capacity
+ 0.143 × no. of elevators
+ 0.142 × vehicle fuel capacity
etc.

(Only the top 15 attributes—in terms of PCA coefficient size—are enumerated in


Equation (6.8).)
Table 6.2 lists the macro-attributes and the respective percentage of data
variability (cumulative) each accounts for.
Using the top ten macro-attributes, accounting for over 80 percent of the origi-
nal dataset’s variance, an optimal solution to Equation (6.7) for the macro-attribute
weights was determined. The weights are listed in the last column of Table 6.2.
144 B. L. Kaluzny
Table 6.2 Principal component analysis and optimal macro-attribute weights

Percentage of data variability accounted

Macro-attribute Proportion Cumulative Optimal weight

A1 17 17 0
A2 12 29 0.452
A3 11 41 0
A4 9 49 0
A5 8 57 0.334
A6 7 64 0
A7 6 69 0
A8 5 74 0
A9 4 78 0
A10 3 81 0.214
A11 3 85
A12 3 88
A13 3 90
A14 2 92
A15 2 94
A16 1 95

Only macro-attributes A2, A5, and A10 have non-zero weights. This is a typi-
cal extreme output of mathematical optimization software. There may exist other
optimal solutions with other non-zero macro-attributes weights.
The hierarchical cluster analysis using the weighted distance function on the top
ten macro-attributes determined by PCA (effectively the three macro-attributes
assigned non-zero weight) is visualized in Figure 6.3. The figure illustrates the
resulting dendrogram indicating the relative grouping of ships. As expected, ships
within the same class (but different rank) are closely grouped together. The
weighted distance function is used to determine the distances of each ship to a
particular ship. For example, plugging in the macro-attributes weights of Table 6.2
into Equation (6.5) results in the list of computed distances to the Rotterdam LPD
presented in Table 6.3.
Figure 6.4 plots the actual ship costs, Ci , versus the costs predicted, Ĉi , by the
analogy method using hierarchical clustering based on a weighted distance met-
ric (the figure is based on Table 6.3 and Equation (6.6)). The solid line running
at 45 degrees represents perfect prediction. The measures of worth of the analogy
method via hierarchical clustering analysis (weighted distance matrix) for predict-
ing ship costs are R = 0.93 and R 2 = 0.86. The latter coefficient of determination
indicates that 86 percent of the total variation in the ship costs can be explained
by an average cost of the known ships weighted by an optimized distance met-
ric. The mean absolute percent error is 16 percent, a significant improvement
over the hierarchical clustering based on the simple distance metric that yielded a
49 percent error.
Hermitage
Alamo
Monticello
‘Spiegel Grove’
‘Point Defiance’
‘Fort Snelling’
‘Plymouth Rock’
Thomaston
‘Gunston Hall’
Comstock
‘Fort McHenry’
Germantown
‘Whidbey Island’
Rushmore
Ashland
Tortuga
‘Fort Fisher’
‘Mount Vernon’
Pensacola
Portland
Anchorage
Denver
Juneau
Duluth
Cleveland
Dubuque
Ogden
Austin
Ponce
Trenton
Nashville
Coronado
Shreveport
‘La Salle’
Vancouver
Raleigh
‘Pearl Harbour’
‘Oak Hill’
‘Carter Hall’
‘Harpers Ferry’
Ocean
BPC3
Mistral
Tonnerre
Svalbard
‘Johan de Witt’
Preserver
Protecteur
Oden
Atle
Carlskrona
Siroco
‘Cardigan Bay’
‘Mounts Bay’
‘Lyme Bay’
‘Largs Bay’
Bulwark
Albion
Rotterdam

Figure 6.3 Dendrogram illustrating the arrangement of the clusters produced by the hierarchical clustering of ships (weighted distance function).
146 B. L. Kaluzny
Table 6.3 Weighted distance of the Rotterdam LPD to ships in the SAS-076 dataset

Name Distance Name Distance Name Distance

Rotterdam 0.000 Vancouver 0.312 Nashville 0.639


Largs Bay 0.034 La Salle 0.316 Trenton 0.646
Lyme Bay 0.034 Harpers Ferry 0.405 Ponce 0.650
Mounts Bay 0.035 Carter Hall 0.431 Whidbey Island 0.655
Cardigan Bay 0.035 Oak Hill 0.435 Germantown 0.659
Oden 0.039 Pearl Harbour 0.439 Fort McHenry 0.664
Carlskrona 0.044 Anchorage 0.546 Gunston Hall 0.668
Johan de Witt 0.046 Portland 0.550 Comstock 0.673
Atle 0.052 Pensacola 0.553 Tortuga 0.678
Albion 0.057 Mount Vernon 0.557 Rushmore 0.682
Bulwark 0.058 Fort Fisher 0.561 Ashland 0.687
Siroco 0.067 Austin 0.601 Thomaston 0.971
Svalbard 0.068 Ogden 0.606 Plymouth Rock 0.975
Protecteur 0.128 Duluth 0.610 Fort Snelling 0.979
Preserver 0.129 Cleveland 0.612 Point Defiance 0.983
Ocean 0.227 Dubuque 0.617 Spiegel Grove 0.987
Tonnerre 0.244 Denver 0.621 Alamo 0.992
Mistral 0.246 Juneau 0.626 Hermitage 0.996
BPC3 0.266 Coronado 0.630 Monticello 1.000
Raleigh 0.309 Shreveport 0.634

700

600
Predicted Cost

500

400

300

200

100

0
0 100 200 300 400 500 600 700
Actual Cost

Figure 6.4 Weighted hierarchical clustering: correlation plot of actual versus predicted
ship costs.

The standard deviation, indicating the variation that the predictions have from
the actual costs, is computed as


 n
 (yi − ŷi )2
i=1
. (6.9)
n
Advances in cost estimating 147
To summarize, as an analogy costing approach, hierarchical agglomerative clus-
ter analysis, principal component analysis, and nonlinear optimization were used
to calculate a matrix of distances among the systems of a dataset. These distances
can then be used to predict the cost of new systems. The fidelity of the estimation
model is very dependent the dataset. It is a “top down” approach applicable in
early design phases of the procurement cycle.

6.4 Data mining for parametric cost estimation


Parametric approaches to cost estimation use regression or other statistical meth-
ods to develop CERs, which are equations used to estimate a given cost element
using an established relationship with one or more independent variables. The typ-
ical ship costing CER expresses the cost (dependent variable) as a function of the
ships length (independent variable).
There are several advantages of a parametric approach to cost estimation,
including:

• it can capture major portions of an estimate quickly and with limited infor-
mation (which is often all that is available in early phases of the procurement
cycle);
• the CER is objective—it is based on consistent and quantitative inputs;
• once the CER is established it is easy to perform sensitivity analyses (deter-
mine the change in cost subject to changes in the independent input variables);
and
• CERs established using regression analyses include standard tests of validity,
including a coefficient of correlation indicating the strength of association
between the independent variables and the dependent variable in the CER.

A critical consideration in parametric cost estimating is the similarity of the sys-


tems in the underlying database, both to each other and to the system which is
being estimated. A major disadvantage of the approach is that it may not provide
low-level visibility (cost breakdown) and changes in sub-systems are not reflected
in the estimate if they are not quantified via an independent variable.
This section describes a parametric approach for ship cost estimation that
incorporates a multitude of cost-driving factors, while remaining a “top down”
approach applicable in the early design phases of the procurement cycle. It
combines features of decision trees with linear regression models to both clas-
sify similar ships (based on attributes) and build piecewise multivariate linear
regression models.

6.4.1 The M5 model tree system


In 1992, Quinlan (1992) pioneered the M5 system for learning models that predict
numeric values. The M5 system combines features of decision trees with linear
regression models. M5 model trees are similar to regression trees; a decision
tree induction algorithm is used to build an initial tree, recursively splitting the
148 B. L. Kaluzny
dataset based on the value of a chosen splitting attribute. The splitting attribute is
selected to minimize the prediction error down each branch. Whereas the nodes of
a regression tree each contain a constant value (prediction), each model tree node
is a multivariate linear regression model. The attributes defining these regression
models are the attributes that are involved in the tree’s branching decisions.
Figure 6.5 depicts a simple example of a M5 model tree. The subfigure on the
left shows a two-dimensional space of independent variables x 1 and x 2 . This is a
graphical representation of a dataset with two attributes (e.g. ship length and crew
size), where each two-dimensional point in the x 1 –x 2 plane represents an instance
(a ship). The dependent variable, Y , represents the numerical value associated with
each instance (i.e. ship cost)—the values for the known instances are not shown
in the subfigure. To build a model to predict the Y value of a new instance, the
algorithm proceeds in four parts. First, a decision tree is constructed using the
same procedure as for regression trees: the tree is constructed recursively by split-
ting the training set (known ships) by selecting the attribute value that minimizes
the error in the estimation of the known instances. For example, in Figure 6.5 the
algorithm determines that splitting the dataset on attribute value x 2 > 2 yields a
lower estimation error (predicted versus known cost of the ships) in comparison
to a model without the split. The process is repeated on the half-spaces generated.
The recursion stops when the estimation error is small (e.g. less than 5 percent) or
when the sub-space contains only a small number of instances. The decisions to
divide the space are represented by a model tree as shown in the right subfigure.
The second part of the algorithm constructs multivariate linear regression mod-
els for each region, in other words at each node of the model tree, using only the
attributes that are referenced by tests somewhere in the subtree of this node.2 In
the example illustrated, there would be eleven linear regression models, one for
each of the six leaves and five internal nodes. Part three of the algorithm applies a
tree pruning routine which eliminates subtrees of a node if the estimation error is
higher in the lower branches than the estimation error when using the node’s inter-
nal regression model. For example, if the combined estimation error of the linear
regression models of leaves 5 and 6 was worse than the estimation error of the
decision node labelled “x 2 < 1,” then the algorithm would eliminate Models 5 and
6 and the subtree rooted at node “x 2 < 1” would collapse to a leaf. The last step

x2
LM3 LM2 x2 > 2
4
Yes No
3
x1 > 2.5 x1 < 4
LM1 Yes No Yes No
2
LM6 LM3 LM4
x2 < 3.5 x2 < 1
1 LM4 Yes No Yes No
LM5 LM1 LM2 LM5 LM6
1 2 3 4 5 6 x1
Y (output)

Figure 6.5 Example M5 model tree.


Advances in cost estimating 149
of the algorithm is a smoothing process. With the goal of improving prediction
accuracy, the linear regression models of adjacent leaves are adjusted so that they
are continuous and smooth (piecewise connected). In the example, the algorithm
would adjust Model 1 to be piecewise connected to Model 2, Model 2 to be piece-
wise connected to Model 3, etc. This process is particularly effective when some
of the linear regression models are constructed from few training cases.
To predict a value for a new instance (e.g. a new ship design), the M5 model
tree is followed from the root down to a leaf using the instance’s attribute values to
make routing decisions at each node. The leaf contains a linear regression model
based on a subset of the attributes, and this is evaluated for the new instance to
output a predicted value.

6.4.1.1 Discussion
Regression trees are more accurate than straightforward linear regression, but the
trees are often cumbersome and difficult to interpret (Witten and Frank 2005).
Model trees are more sophisticated than regression trees as they approximate con-
tinuous functions by linear “patches”—M5 model trees are analogous to piecewise
linear functions. M5 model trees have an advantage over regression trees with
respect to compactness and prediction accuracy due to the ability of model trees
to exploit local linearity in the data. Regression trees will not predict values lying
outside the range of learned training cases, while M5 model trees can extrapolate.
The M5 model tree algorithm is optimized both to learn known cases and predict
unknown cases. The trees are also smaller, easier to understand, and their average
error values on the training data are lower.
M5 model trees tackle the difficulties in considering a multitude of cost-driving
factors; they determine the right set of independent variables (ship attributes) by
construction. The effect of the M5 system is somewhat similar to the use of indica-
tor variables in standard linear regression analysis, with the key difference being
that appropriate indicators are identified and included based on an optimization
algorithm.
M5 model trees have been shown to excel even when limited data is available
(Wang and Witten 1997) or learn efficiently from large datasets. They can handle
datasets which include systems with notable differences, missing data, and noise
(as is the case for the SAS-076 ship dataset). The decision tree can branch on any
variable type: nominal (e.g. military versus non-military) or numeric (e.g. tonnage
less than 15,000 or greater than 15,000).
The use of M5 model trees for numeric prediction has increased since com-
prehensive descriptions, implementations, and refinements of Quinlan’s method
became available (Wang and Witten 1997; Malerba et al. 2004; Torgo 2000, 2002;
Dobra 2002). Recently Chen (2006) and Sharma and Litoriya (2012) discussed
the benefits of the system for estimating the cost of software development.
There are many other types of prediction methods that employ data mining tech-
niques. For example, neural networks are commonly used for predicting numeric
quantities, but suffer from the disadvantage of producing opaque structures that
150 B. L. Kaluzny
mask the nature of the solution. Their arbitrary internal representation means that
there can be variations between networks trained on the same data. By comparison,
the M5 system is transparent and the model tree construction is repeatable. The
final tree and linear regression models are straightforward and clear.

6.4.2 Application to SAS-076 ship dataset


An easy-to-use implementation of Quinlan’s M5 model tree is available as part
of the WEKA project.3 Witten and Frank’s (2005) textbook provides documenta-
tion on both the theory and implementation (data formatting, execution, etc.) of
the algorithm. The descriptive attributes, the complete set of technical attributes,
and a normalized cost attribute for each of the ships in the SAS-076 dataset were
used as input to the M5 model tree algorithm. Using WEKA,4 the construction of
the M5 model tree for the SAS-076 ship database takes on order of seconds of
computation.
Figure 6.6 shows the resulting M5 model tree. The root of the decision tree
divides the ships into two categories based on the number of air-cushioned landing
craft (LCAC) that the ship is designed to carry. Internal nodes of the tree further
split the dataset on attributes such as the number (#) of torpedo decoy systems
on board, the ship’s rank in class, the number of propeller shafts, the maximum
number of helicopters supported, and the ship’s range in terms of total distance
in nautical miles. The tree branches out to seven leaves, where seven correspond-
ing linear regression models are fitted. The regression models are presented in
Table 6.4. In addition to the ship attributes used in the decision tree for branch-
ing, the linear regression models use the ship attributes of length (in meters), crew
size, and number of elevators on board. Table 6.5 provides the minimum, median,
mean, and maximum values found in the SAS-076 ship dataset for the attributes
used by the M5 model.
The linear regression models output the log-transformed (decadic) cost of a
ship. The individual linear regression models are mostly intuitive: the cost of a

# of LCAC
< 2 or ? ≥2

# of Torpedo Decoys Rank in Class

0 ≥ 1 or ? ≥6 ≤5

# of Propeller Shafts LM3 LM7 Maximum # of Helicopters Supported

≤1 ≥2 ≤1 ≥2

LM1 LM2 LM4 Range (total distance in nmi)


≤ 8800 > 8800 or ?

LM5 LM6

Figure 6.6 M5 model tree applied to the SAS-076 dataset.


Advances in cost estimating 151
Table 6.4 M5 model tree linear regression models

LM1 LM2

Log(Cost) = 7.8009 Log(Cost) = 7.7916


− 0.0083 × rank in class − 0.0083 × rank in class
+ 0.0017 × length (m) + 0.0017 × length (m)
+ 0.0006 × crew size + 0.0006 × crew size
− 0.0450 × no. of propeller − 0.0450 × no. of propeller
shafts shafts
+ 0.0304 × no. of LCAC + 0.0304 × no. of LCAC
+ 0.0218 × no. of elevators + 0.0218 × no. of elevators
+ 0.1897 × no. of torpedo + 0.1897 × no. of torpedo
decoys decoys

LM3 LM4

Log(Cost) = 7.8869 Log(Cost) = 8.1555


− 0.0189 × rank in class − 0.0226 × rank in class
+ 0.0017 × length (m) + 0.0031 × length (m)
+ 0.0006 × crew size + 0.0001 × crew size
− 0.0203 × no. of propeller + 0.0216 × no. of LCAC
shafts + 0.0155 × no. of elevators
+ 0.0304 × no. of LCAC
+ 0.0218 × no. of elevators
+ 0.1285 × no. of torpedo
decoys

LM5 LM6

Log(Cost) = 8.2675 Log(Cost) = 8.4549


− 0.0349 × rank in class − 0.0303 × rank in class
+ 0.0029 × length (m) + 0.0018 × length (m)
+ 0.0001 × crew size + 0.0001 × crew size
+ 0.0216 × no. of LCAC + 0.0216 × no. of LCAC
+ 0.0064 × no. of elevators + 0.0038 × no. of elevators

LM7

Log(Cost) = 8.4544
− 0.0202 × rank in class
+ 0.0012 × length (m)
+ 0.0001 × crew size
+ 0.0216 × no. of LCAC
+ 0.0155 × no. of elevators

ship increases as the length, number of elevators, crew size, number of LCAC sup-
ported, or number of torpedo decoy systems increase(s). The regression models
also predict a shipbuilding learning curve as the cost of constructing a ship
decreases as a function of the ship’s rank in class. However, the negative coeffi-
cient preceding the number of ship propeller shafts in the regression models LM1,
LM2, and LM3 is counter-intuitive. It seems unlikely that a ship will cost less if
152 B. L. Kaluzny
Table 6.5 Statistics of attributes used in the M5 model tree linear regression models

Attribute Minimum Median Mean Maximum

Rank 1 3 3.6 12
Length 103.7 173.8 170.2 203.4
Range (total distance) 6,000 8,000 9,933 30,000
Crew size 15 352 321 491
No. of propeller shafts 0 2 1.7 4
No. of LCAC 0 2 1.9 4
No. of elevators 0 0 0.6 3
No. of helicopters supported 0 6 5 18
No. of torpedo decoy systems 0 0 0.5 1

the number of propeller shafts is increased. The SAS-076 ship data explains this
anomaly: all but one of the ships captured in the SAS-076 ship dataset have at
most two propeller shafts; only Sweden’s Atle icebreaker has four. The Atle is
also the lowest costing ship. Since there is no constraint on keeping the coeffi-
cients positive, the linear regression model LM2 uses the propeller shaft attribute
to model Atle’s low cost. The combination of Atle’s low cost and unique propeller
shaft attribute drives the negative coefficient sign. The smoothing process of the
M5 model tree construction spreads this feature to the adjacent linear regression
models LM1 and LM3. The M5 model can be potentially adjusted in an attempt to
remove such anomalies by disabling the particular attribute (number of propeller
shafts); however, there is no guarantee that the regenerated model will not sub-
stitute this attribute with another, also allocated a negative coefficient. Similarly,
removing the instance (e.g. Atle) from the dataset provides no guarantees. Rather
than subjectively diminishing the dataset, anomalies are noted and discussed as
part of the results.
Figure 6.7 plots the actual ship costs versus the costs predicted by the M5 model
tree. The worth of a regression-based model is most commonly measured by the
coefficient of correlation, the quantity that gives the quality of a least squares
fitting to the original data. Let yi be the actual cost of ship i and ŷi the predicted
cost (using the M5 model tree), then the coefficient of correlation, R, of the M5
system is calculated as
  
n yi ŷi − yi ŷi
R=
  2

 2  2 , (6.10)
n yi −
2
yi n ŷi − ŷi

where n = 59, the number of ships in the dataset, and each of the sums are over
these 59 ships. By design R can range from −1 to +1 with R = 1 indicating
a perfect positive linear correlation. A correlation greater than 0.8 is generally
described as strong (Ryan 1997). The value R 2 , known as the coefficient of deter-
mination, is a measure of how well the regression line represents the data—it
is a measure determining how certain one can be in making predictions from
Advances in cost estimating 153

700

600
Predicted Cost

500

400

300

200

100

0
0 100 200 300 400 500 600 700
Actual Cost

Figure 6.7 M5 model tree: correlation plot of actual versus predicted ship costs.

the model. R 2 is also the ratio of the explained variation (by the model) to the
total variation of the dataset. The measures of worth of the M5 model tree for
predicting ship costs are a strong R = 0.9646 and R 2 = 0.9304. The coefficient
of determination indicates that 93 percent of the total variation in the ship costs
can be explained by the linear relationships described by the M5 model tree lin-
ear regression equations. The remaining 7 percent of the total variation remains
unexplained.
The mean absolute percentage error quantifies the amount by which the esti-
mated cost differs from the actual cost. The mean absolute percentage error is
computed as follows:

1  |yi − ŷi |
i=59
. (6.11)
59 i=1 yi

The mean absolute percentage error for the M5 model tree applied to the SAS-076
ship dataset is 11 percent. Mean absolute percentage errors specific to the
individual M5 model tree linear regression models are shown in Table 6.6. In
addition to absolute percentage errors, the standard deviation is also indicative of
a model’s accuracy. For the purpose of determining the standard deviation of a
M5 model tree output prediction, the standard deviation over all the training data
is more reflective of the M5 system than just the standard deviation of the train-
ing cases reaching the particular leaf node used for the prediction. By M5 model
tree construction, each of the training cases influence the structure of the final
model tree.
Figure 6.8 plots the residuals for attributes used in the M5 model tree linear
regression models, providing a visual confirmation that the residuals are randomly
154 B. L. Kaluzny
Table 6.6 Mean absolute percentage errors of known instances per individual M5 model
tree linear model

LM1 LM2 LM3 LM4 LM5 LM6 LM7

Mean percentage error 24 22 10 13 12 7 6


No. of instances 3 3 16 9 12 6 10

scattered (indicating that a linear fit is suitable) for all attributes except for the
rank in class. The residuals for the rank in class attribute indicate more accurate
predictions the higher the ship’s rank in class. This indicates that a linear fit may
not be the most suitable. Typically the construction of lower-ranked ships of the
same class benefit from labor learning curves and economies of scale. Learning
curves are nonlinear, taking the general form of Cr = a · r b , where Cr is the cost of
unit (of rank) r , and a and b are fitted parameters (see Goldberg and Touw (2003)
for more information). Nonlinear regression may be more suitable to model the
relationship of ship cost to the rank in class attribute, however this is outside the
capabilities of the M5 model tree algorithm.
The results presented in Table 6.4 show that linear regression model LM5 con-
tributes the greatest to the standard deviation. LM5 models the nine highest costing
ships. Figure 6.7 reveals that in eight of these cases, LM5 underestimates the actual
cost. By the piecewise linear M5 model tree construction, LM5 is influenced by
the adjacent linear regression models LM4 and LM6. To evaluate the degree of
this influence, a separate multiple linear regression model was fitted to the ships
reaching the LM5 leaf using the same five ship attributes as LM5. The resulting
CER is as follows:

Log10 (Cost) = 8.6376 − 0.0651 × rank in class


(6.12)
+0.0637 × number of LCAC,

with an R value of 0.95 (R 2 = 0.90), and mean absolute percentage error of 7


percent. This result indicates that it is indeed possible to derive better models
for subsets of the data. This does not depreciate the M5 model tree algorithm,
whose strength is its optimization in predicting unknown cases (rather than simply
memorizing the known). Using the linear regression model of Equation (6.12) in
place of LM5 would leave adjacent linear models LM4, LM5, and LM6 sharply
discontinuous. The M5 system applies a smoothing process to construct piecewise
linear regression models. Experiments by Wang and Witten (1997) show that this
smoothing substantially increases the accuracy of predictions of unseen cases. The
M5 system smoothing process may be a reason for the underestimation of the most
expensive ships. When the smoothing procedure is disabled, the same model
tree depicted in Figure 6.6 is generated, but the linear regression models
differ.
0.15 0.15 0.15
0.10 0.10 0.10
0.05 0.05 0.05
0.00 0.00 0.00
–0.05 –0.05 –0.05

(Predicted)
(Predicted)
(Predicted)
–0.10 –0.10 –0.10

Log(Actual)–Log
Log(Actual)–Log
Log(Actual)–Log
–0.15 –0.15 –0.15
2 4 6 8 10 12 100 120 140 160 180 200 5000 10000 15000 20000 25000 30000
Rank Length Range (total distance)

0.15 0.15 0.15


0.10 0.10 0.10
0.05 0.05 0.05
0.00 0.00 0.00
–0.05 –0.05 –0.05

(Predicted)
(Predicted)

(Predicted)
–0.10 –0.10 –0.10

Log(Actual)–Log
Log(Actual)–Log

Log(Actual)–Log
–0.15 –0.15 –0.15
100 200 300 400 500 0 1 2 3 4 0 1 2 3 4
Crew size Number of propeller shafts Number of LCAC

0.10 0.10 0.10


0.05 0.05 0.05
0.00 0.00 0.00
–0.05 –0.05 –0.05

(Predicted)
(Predicted)

(Predicted)
–0.10 –0.10 –0.10

Log(Actual)–Log
Log(Actual)–Log

Log(Actual)–Log
–0.15 –0.15 –0.15

0.0 0.5 1.0 1.5 2.0 2.5 3.0 0 5 10 15 0.0 0.2 0.4 0.6 0.8 1.0
Number of elevators Number of helicopters supported Number of torpedo decoy systems

Figure 6.8 Plots of residuals for attributes used in the M5 model tree linear regression models.
156 B. L. Kaluzny
Figure 6.9 plots the actual ship costs versus the costs predicted by the M5 model
tree when the smoothing procedure is disabled. The R value is 0.985 (R 2 = 0.97).
The unsmoothed M5 model tree does better in learning the costs of the most
expensive known ships, but likely at the expense of prediction accuracy of unseen
cases.

6.4.3 Comparison to linear regression models


The best CER returned by applying simple linear regression on the SAS-076
dataset is

Log10 (Cost) = 6.94 + 0.01 × length of ship (in meters), (6.13)

with an R value of 0.75 (R 2 = 0.56). One typically tests the hypothesis that the
linear regression fit is a better fit than fitting to just the mean length of the ships.
Define the total variation to be the variance when a model is fit to just the mean
length of the ships. Residual variation is defined as the variance when the lin-
ear model is fit. The F-statistic is the ratio of the model and residual variance
and represents the proportional increase in error of fitting a mean model versus
the linear model. The p-value is the probability of rejecting the null hypothesis
that the linear fit is equal to the mean fit alone, when it is in fact true. A sig-
nificant p-value (e.g. less than 5 percent) implies that the linear fit is a better
fit than the mean alone. In this case the F-statistic is 74.2 and the p-value is
6.7 × 10−12.
Applying multiple linear regressions with a greedy attribute selection method
(step through the attributes removing the one with the smallest standardized

700

600
Predicted Cost

500

400

300

200

100

0
0 100 200 300 400 500 600 700
Actual Cost

Figure 6.9 Unsmoothed M5 model tree: correlation plot of actual versus predicted
ship costs.
Advances in cost estimating 157
coefficient until no improvement is observed in the estimate of the error given
by the Akaike (1980) information criterion), yields the CER

Log10 (cost) = 5.7184 − 0.0229 × rank in class


+ 0.0125 × length (in meters)
+ 0.0329 × beam (in meters)
+ 0.1073 × draught (in meters)
(6.14)
− 0.0001 × full load displacement (in tonnes)
+ 0.0011 × crew size
− 0.0830 × number of propeller shafts
− 0.0223 × number of guns of calibre ≥ 75,

with an R value of 0.91 (R 2 = 0.83). Table 6.7 lists the F-statistics and p-values
associated with each of the model attributes. The rank in class and number of
propeller shaft attributes have a p-value near 10 percent, indicating that the fitted
linear model should not be claimed to be a better fit than just using the mean values
for these two attributes.
While the CER produced by applying simple linear regression is straightfor-
ward to understand, the negative coefficient signs of the multiple linear regression
CER makes its interpretation non-trivial.

6.5 Discussion
Figure 6.4 shows that the analogy method using hierarchical clustering based
on a weighted distance metric underestimates the cost of seven of the eight
most expensive ships. This was also a characteristic of the parametric estimation
presented in the preceding section. In the latter it was conjectured that the under-
estimation was likely a result of the smoothing and pruning functions of the M5
model tree algorithm, designed to optimize the predictive capability of the method.
However, the underestimation of expensive ships in both methods is potentially an
indication that the attributes (and their values) of the SAS-076 ship dataset do

Table 6.7 Multiple linear regression F-statistics and p-values

Attribute F-statistic p-value

Rank in class 2.7 1.0 × 10−01


Length 169.6 1.1 × 10−17
Beam 9.9 2.8 × 10−03
Draught 6.3 1.5 × 10−02
Full load displacement 43.7 2.5 × 10−08
Crew size 15.0 3.1 × 10−04
Number of propeller shafts 2.6 1.1 × 10−01
Number of guns of calibre ≥ 75 4.5 4.0 × 10−02
158 B. L. Kaluzny
Table 6.8 Comparison of the M5 model tree and hierarchical clustering methods

M5 model tree Hierarchical clustering

Coefficient of correlation 0.96 0.93


Coefficient of determination 0.93 0.86
Mean absolute percentage error 11 16
Ability to learn known cases  
Optimized to predict unknown cases  ×
Uses entire dataset  ×

not provide enough information to help distinguish the highest costing ships from
their peers.
Multiple cost estimations are a good thing, especially during early design and
developmental phases of a system. Multiple independent cost estimates are ideal,
and strongly endorsed as best practice (see NATO RTO SAS-054 Task Group
(2007) report). The two machine learning approaches described in this chapter
are independent of each other; however, given their reliance on a common source
of data (in this instance) the independence of the cost estimates they produce can
be argued. Given multiple estimates, decision-makers must exercise sound judg-
ment to assess and select a final figure. Properties of the machine learning methods
discussed can provide guidance:

• The M5 model tree algorithms are optimized both to learn known cases
and predict unknown cases. The attribute weights used in the hierarchical
clustering method are optimized to learn the known cases.
• The hierarchical clustering approach uses PCA to reduce the dimensionality
of the attribute space. Due to computational limitations, the weight optimiza-
tion method could only be applied on the top ten macro-attributes, accounting
for 80 percent of the original dataset’s variability. In comparison, the M5
model tree algorithm is computationally superior as it efficiently learns from
large datasets.
• The M5 model tree results in a better correlation measure, lower mean abso-
lute percentage error, and smaller standard deviation in estimating the known
cases.

Table 6.8 synthesizes the predictions and compares properties of the M5 model
tree and hierarchical clustering methods.
Given this information, decision-makers could be advised that estimates gener-
ated by the M5 model tree should be considered to be the primary estimates and
the hierarchical clustering estimates as secondary estimates.

6.6 Conclusion
Two novel approaches to cost estimation using machine learning algorithms are
demonstrated. Both approaches incorporate a multitude of cost-driving factors and
Advances in cost estimating 159
allow for a greater variability in the input dataset—variability that could be cur-
tailed when using traditional approaches. As with other parametric and analogy
approaches, the fidelity of the estimation models are very dependent the dataset,
especially if the size of the dataset is small. Both are “top down” approaches
applicable in early design phases of the procurement cycle.
As an analogy costing approach, hierarchical agglomerative cluster analysis,
principal component analysis, and nonlinear optimization were used to calculate
a matrix of distances among the systems of a dataset. These distances can then be
used to predict the cost of new systems.
The parametric approach combined features of decision trees with linear regres-
sion models to both classify similar systems (based on attributes) and build
piecewise multivariate linear regression models. The attributes of a new system
can then be used to trace down the tree and as input to the resulting regression
models which output a cost prediction.

Acknowledgements
Portions of this chapter reproduce work originally presented in the NATO RTO
SAS-076 Task Group (2012) report, Journal of Cost Analysis and Parametrics
article by Kaluzny et al. (2011), and Defence Research and Development Canada
report by Kaluzny and Shaw (2011). The author is indebted to Sorin Barbici,
Göran Berg, Renzo Chiomento, Dimitrios Derpanis, Ulf Jonsson, R.H.A. David
Shaw, Marcel Smit, and Franck Ramaroson for their contributions.

Notes
1 See www.iceaaaonline.org [last accessed November 20, 2014].
2 The linear models are also tested to see if eliminating a set of attributes reduces the
estimated error (a measure of accuracy of the model on unseen cases).
3 Data mining software in Java, available at www.cs.waikato.ac.nz/ml/weka/ [last accessed
November 20, 2014].
4 The program’s default parameters for the M5 model tree algorithm were used.

References
Akaike, H. 1980. “Likelihood and the Bayes Procedure.” Trabajos de Estadística y de
Investigaciín Operativa 31: 143–166.
Chen, Z. 2006. Reduced-Parameter Modeling for Cost Estimation Models. PhD thesis,
Faculty of the Graduate School, University of Southern California.
Dobra, A. 2002. “Secret: A Scalable Linear Regression Tree Algorithm.” In Proceedings of
the Eighth ACM SIGKDD International Conference on Knowledge Discovery and Data
Mining. New York: ACM Press, pp. 481–487.
Flynn, B. and Garvey, P. 2011. “Weapon Systems Acquisition Reform Act (WSARA) and
the Enhanced Scenario-Based Method (ESBM) for Cost Risk Analysis.” In Proceedings
of 44th Annual Department of Defense Cost Analysis Symposium Williamsburg, Virginia.
Goldberg, M. and Touw, A. 2003. Statistical Methods for Learning Curves and Cost Analy-
sis. Topics in Operations Research Series. Linthicum MD, USA: Institute for Operational
Research and Management Sciences.
160 B. L. Kaluzny
Jolliffe, I. 2002. Principal Component Analysis. New York: Springer.
Kaluzny, B., Barbici, S., Berg, G., Chiomento, R., Derpanis, D., Jonsson, U., Shaw, R.,
Smit, M., and Ramaroson, F. 2011. “An Application of Data Mining Algorithms for
Shipbuilding Cost Estimation.” Journal of Cost Analysis and Parametrics 4: 2–30.
Kaluzny, B. and Shaw, R. 2011. A Parametric Approach to Cost Estimation During Options
Analysis. Technical Memorandum DRDC CORA TM 2011-069, Defence Research and
Development Canada.
Malerba, D., Esposito, F., Ceci, M., and Appice, A. 2004. “Top-Down Induction of Model
Trees with Regression and Splitting Nodes.” IEEE Transactions on Pattern Analysis and
Machine Intelligence 26(5): 612–625.
Miroyannis, A. 2006. Estimation of Ship Construction Costs. PhD thesis, Department of
Mechanical Engineering, Massachusetts Institute of Technology.
NATO RTO SAS-054 Task Group 2007. Methods and Models for Life Cycle Costing.
Technical Report RTO-TR-SAS-054, NATO Research & Technology Organization.
NATO RTO SAS-076 Task Group 2012. NATO Independent Cost Estimating and the Role
of Life Cycle Cost Analysis in Managing the Defence Enterprise. Technical Report
NATO RTO-TR-SAS-076 AC/323)SAS-076)TP/420, NATO Research & Technology
Organization.
Quinlan, J. 1992. “Learning with Continuous Classes.” In Proceedings AI’92, edited by A.
Adams and L. Sterling. Singapore: World Scientific, pp. 343–348.
Ryan, T. 1997. Modern Regression Methods (2nd edition). New York: John Wiley & Sons,
Inc.
Sharma, N. and Litoriya, R. 2012. “Incorporating Data Mining Techniques on Software
Cost Estimation: Validation and Improvement.” International Journal of Emerging
Technology and Advanced Engineering 2(3): 301–309.
Torgo, L. 2000. Inductive Learning of Tree-based Regression Models. PhD thesis, Univer-
sidade do Porto.
Torgo, L. 2002. “Computationally Efficient Linear Regression Trees.” In Proceedings
of International Federation of Classification Societies (IFCS) 2002: Classification,
Clustering and Data Analysis: Recent Advances and Applications.
Wang, Y. and Witten, I. 1997. “Inducing Model Trees for Continuous Classes.” In
Proceeding of the 9th European Conference on Machine Learning Poster Papers,
pp. 128–137.
Witten, I. and Frank, E. 2005. Data Mining: Practical Machine Learning Tools and
Techniques (2nd edition). Massachusetts: Morgan Kaufmann.
7 Facing future funding realities
Forecasting budgets beyond the
future year defense plan

Ross Fetterly and Binyam Solomon1

7.1 Introduction
Introducing future funding realities is a critical component of any military cost–
benefit analysis (CBA). Building a national defense force for future generations,
planning missions that will be assigned to military forces in the coming decades,
developing a capital equipment program to meet anticipated future operational
requirements, and building the infrastructure and logistics system to support
them are by their very nature long-term tasks. It is significant, therefore, that in
defense departments around the world one of the biggest challenges is to plan
defense budgets over the medium to long-term.2 Indeed, “despite ferociously
complex planning processes, coherence between strategic guidance, capability
development and the budget remains elusive” (Thompson 2007, 3).
Forecasting future funding envelopes for defense programs is a daunting task.
The challenge of allocating defense budgets among labor (military and civilian
personnel) and capital (weapon systems, platforms and infrastructure) inputs is
further complicated by three additional factors. First, the long-term planning hori-
zon necessary in defense is in stark contrast to the relatively short-term planning
outlook of most national governments, a consequence of the four to five year
election cycles that exist in most modern democracies. Second, defense directly
competes for discretionary funding with publicly popular programs such as health-
care and infrastructure spending. Third, defense faces external shocks in the form
of dramatic changes in the international strategic environment, such as the end of
the cold war in 1989, and the rise of multiple inter-state and intra-state conflicts
throughout the world after September 11, 2001. Although some of these prob-
lems were anticipated by analysts, it is still difficult to predict the scale, scope and
impact of those changes (Dorff 2005).
Given these challenges, no one strategic management approach or forecasting
model can provide a completely robust framework for all nations that face differ-
ent external threats, economic growth trajectories, and mutual defense alliances.
One key element required to enhance defense resources management is the devel-
opment of a more comprehensive approach to forecast and manage budgets within
defense departments.
162 R. Fetterly, B. Solomon
In this chapter three forecasting approaches and two management strategies are
presented to help develop future funding lines (i.e. budget envelopes) for defense
investments. The development of funding lines would allow defense planners to
conduct more rigorous cost–benefit and tradeoff analyses using reasonable proxies
of future budget constraints. While most multi-year planning exercises tend to be
fiscally informed, future cost constraints are often not anchored by a reasonable
approximation of a funding line.
The risk is that a capability acquired in the absence of careful budget esti-
mates may impact other programs if funds are “borrowed” from those programs to
fund the new program that exceeded the actual funding constraint. Alternatively,
unanticipated budget cuts can seriously impact program performance sacrificed to
accommodate the smaller budget.
Recognizing these challenges, a United States Government Accountability
Office (GAO) report on cost growth in US defense acquisition stated, “the goal
is to get a better match between budgeted funds and costs in order that the true
impact of investment decisions is known” (GAO 2005). The major finding of a
Center for Strategic and International Studies (CSIS) study adds that “resource
constraints require a more integrated approach in defense” (CSIS 2005).
The rest of the chapter is structured as follows. The first section presents stylized
facts about defense and fiscal frameworks. The second section examines manage-
ment of the defense budget while the third section discusses defense cost drivers.
The fourth section introduces various forecasting models and contrasts their uses
to the challenges of optimal defense resources management. The last section con-
cludes the chapter and highlights some future directions for research in forecasting
defense budgets beyond the future year defense plan (FYDP).3

7.2 Defense and the economy: stylized facts


The development of a defense budget is a complex and time-consuming process.
The results of this process are influenced by the dynamics of the political pro-
cess, various organizations within the defense establishment, government officials,
and the external environment. Each of these diverse groups has differing objec-
tives (incentives) and responsibilities (see Chapter 3 for example). For the central
government the challenge is to allocate tax revenues among competing demands.
For most NATO member nations, the standard of living of their citizens has
grown significantly in the last three decades. As shown in Table 7.1, most of these
nations displayed real (inflation-adjusted) GDP per capita growth of about 23 per-
cent during the 1980s and 1990s and a more moderate rate of growth of about
13 percent during the 2000s. The next half decade to 2015 is forecasted to mod-
erate even further to 9 percent. This moderation in the standard of living further
complicates the central government’s budget allocation process.
A nation’s standard of living can be improved by increasing the labor participation
rate of its citizens (demographics), the level of effort (hours worked), and/or produc-
tivity. The former two have natural limits in the form of demographics, immigration
absorption and available hours of work.Productivity can be improved by increasing
Facing future funding realities 163
Table 7.1 Real GDP per capita growth for selected NATO members

Country 1980s % 1990 % 2000s % 2010–15 %

Belgium 19.3 21.4 7.8 7.0


Canada 16.0 18.9 8.2 6.0
Czech Republic n.a. n.a. 30.0 16.0
Denmark 20.7 22.2 4.1 9.0
France 18.8 15.7 5.5 7.3
Germany 20.3 17.0 9.2 9.8
Greece 1.9 16.0 23.3 4.9
Italy 23.5 15.3 −2.8 3.2
Luxembourg 46.5 36.3 15.2 6.5
Netherlands 16.9 24.2 9.6 8.6
Norway 21.5 31.2 7.6 6.3
Portugal 35.8 28.6 2.1 3.3
Slovak Republic n.a. n.a. 48.8 20.9
Slovenia n.a. n.a. 26.1 13.5
Spain 25.9 25.4 7.1 7.6
United Kingdom 26.0 22.6 9.1 8.4
United States 23.0 21.7 7.1 8.7

Source: IMF (various years) World Economic Outlook.


n.a. = not applicable.

capital intensity and through innovation. These require investment in infrastructure,


research and education. Meanwhile, increased demand for social programs by an
aging population constrains budget resources available for defense.
Table 7.2 presents annual government fiscal positions as a proportion of income
(GDP). The 2008 financial crisis has significantly eroded government finances.
The selected NATO member nations with the exception of Norway are expected to
face budgetary shortfalls at least until 2015. While stimulus spending and bailout
packages to combat the financial crisis contributed to the worsening fiscal bal-
ance, automatic stabilizers such as transfer payments that automatically kick in
during economic downturns (e.g. unemployment insurance, food stamps, income
supplements and other services for the poor) further deplete tax revenues.
Competition for shrinking government funding can have large impacts on dis-
cretionary items such as national defense budgets. A powerful example of this
comes from the Canadian experience of the 1990s. A newly elected Liberal gov-
ernment began its mandate in 1993 saddled with an inherited CAN$40 billion (and
rising) public debt, which accounted for 70 percent of GDP by 1995 (combined
public sector debt was close to 100 percent). In addition to the usual strategies of
raising taxes and cutting and re-engineering transfer payments, the new govern-
ment also formulated a new defense policy to reflect post-cold war realities (Stone
and Solomon 2005).
As shown in Figure 7.1, both defense and non-defense spending were cut, albeit,
at a different rate. Non-defense was cut by about 9 percent while defense absorbed
a reduction between 24 percent (economy-wide inflation-adjusted) and 28 percent
164 R. Fetterly, B. Solomon
Table 7.2 Government annual surplus/deficit per GDP

Country 2000 2010 2015

Belgium −0.037 −4.784 −5.181


Canada 2.945 −4.933 −0.249
Czech Republic −3.722 −5.411 −5.207
Denmark 1.804 −4.556 −1.09
France −1.474 −8.002 −2.199
Germany 1.314 −4.496 −1.414
Greece −3.692 −7.883 −1.956
Italy −0.864 −5.111 −3.003
Luxembourg 5.968 −3.793 −0.711
Netherlands 1.972 −6.002 −4.055
Norway 15.374 11.14 11.756
Portugal −1.093 −7.253 −5.818
Slovak Republic −12.3 −8.018 −1.763
Slovenia −1.235 −5.685 −0.829
Spain −0.997 −9.272 −4.414
United Kingdom 1.345 −10.171 −2.417
United States 1.1 −11.085 −6.499

Source: IMF (various years) World Economic Outlook.

150
Other Depts DND (GDP Defl) 93–94 = 100 DND (DSI Defl)
140

130

120

110

100

90

80

70
93–94 95–96 97–98 99–00 01–02 03–04 05–06 07–08 09–10

Figure 7.1 Trend in defense and non-defense spending: 93–94 = 100.


Source: Authors’ calculations.

(defense-specific inflation). There is evidence that inflation inflicts considerable


damage to the purchasing power of defense dollars (Solomon 2003; Kirkpatrick
2004). Defense inflation is discussed in some detail in subsequent sections.
By the end of 1997–98, the Government of Canada managed to balance the
books and non-defense spending resumed its original growth path beginning in
1999. Defense spending, on the other hand, only reached its pre-reduction levels
in 2007, about eight years later. The main point to take away from this Canadian
Facing future funding realities 165
example is that a country’s budgetary constraints affects the level of funding avail-
able for defense, and defense tends to absorb a disproportionately large share of
the cuts. Since defense often accounts for a large share of the discretionary budget,
this is to be expected. Figure 7.2 shows how the recent financial crisis has forced
the Government of Canada to revise its 20-year defense projections by CAN$1
billion beginning in 2013. Similarly in the United States, the US GAO long-term
simulations show that “absent policy changes the federal government faces an
unsustainable growth in debt” (GAO 2010).
In the US and elsewhere, the growing cost in non-discretionary (mandatory)
budget items as a percentage of total government spending (medical care, social
security, and other payments in economies with aging populations) is irrevoca-
bly changing the dynamics of federal budgets. Certainly, federal programs based
on entitlements are relatively inflexible in the short-term and difficult to change
politically. The implication for defense is that governments will be increasingly
constrained in the way they fund defense budgets in response to conflicts or shifts
in the international strategic environment, largely due to the increasing proportion
of social welfare programs in the budget. Consequently, restoring a more equitable
overall balance between revenues and expenses, as well as controlling cost growth
in social welfare and defense programs, will be critical to a long-term fiscal policy
capable of funding essential national security programs. The challenge for most
Western governments is to achieve long-term sustainable defense funding in an
environment where fiscal policy is likely to be “tightened sharply over the next
several years” (World Bank 2010).

National Defence Budgeta


billions of dollars
24

22
Pre-Budget 2010
Post-Budget 2010
20

18

16

14

12

10
2005–2006 2008–2009 2011–2012 2014–2015 2017–2018

Figure 7.2 Projected defense funding, budget 2010.


Source: Department of Finance-DoF (2010).
Note
a Excludes incremental funding for deployed operations in Afghanistan and in support of
the 2010 Olympics.
166 R. Fetterly, B. Solomon
Senior finance department officials over the coming decade will find that
national budgetary deficits have become to some degree structural in nature, and as
a result will have a substantial continuing negative impact on the fiscal framework.
The challenge, therefore, is to adjust and shape spending practices in defense to
accommodate a medium-term environment of anticipated budgetary constraints.
As a consequence, affordability needs to be brought up alongside operational
performance as the principal criteria under which all defense programs will be
assessed. In response to this multi-faceted and complex challenge, a broad-based
and integrated response by defense departments is required.

7.3 Defense resources management


The defense sector has a number of very distinctive characteristics. Indeed, the
relationship between factor inputs and outputs in the military production function
is very specific. The inputs into the production function that produces “defense
capability” include capital, labor and technology—which provide the personnel,
infrastructure and equipment for national defense forces. From an economics per-
spective, the goal is to find the optimal mix of these inputs that maximizes national
security subject to future funding constraints (see Chapter 3).
A distinguishing characteristic of defense budgets is that although flexibil-
ity in the current year may be minimal, current decisions on specific programs
can have a significant impact on future year budgets. Therefore, today’s budget
decisions have a crucial long-term impact on the production of defense output.
The long-term challenge is to manage over-programming (GAO 1994) in-year,
while simultaneously dealing with systematic fiscal optimism in defense plan-
ning (Jordan 2000), and addressing a multitude of unanticipated in-year events:
all while managing a capital program that extends several decades into the
future.
Various NATO governments have introduced management strategies to tackle
the defense resources management challenge described above. Some have intro-
duced strategic-level and comprehensive approach such as the Planning, Program-
ming, Budgeting and Execution (PPBE) process adopted by the US Department of
Defense (DoD) and capability based planning (CBP) in Australia, Canada and the
United Kingdom. These management strategies are discussed extensively in the
literature and also in other chapters of this book. In this chapter, we concentrate
on CBP and other short-term management strategies as they are being applied in
Canada, with the goal of extracting more general lessons for other countries faced
with similar challenges of forecasting future defense budget constraints.

7.3.1 Business planning


The need for a comprehensive long-term resources management strategy stems
from the fact that governments do not explicitly incorporate the temporal asym-
metry between defense, the holder of the largest discretionary budget and capital
outlays, and other departments that deal with transfer payments and recurring
operations. Thus in times of fiscal constraints, the temptation is to delay capital
Facing future funding realities 167
acquisition, and cut equipment maintenance funding and life-cycle maintenance
of infrastructure to provide short-term financial relief to national governments.
However, in the long-term, dramatic swings in defense funding can increase the
cost of defense investments and result in sub-optimal allocation of scarce defense
resources. Moreover, long-term budgetary planning is even more essential in
defense “because of the lengthy lead times required to field new weaponry and
force structure” (Jones and McCaffery 2005). In addition, effective management
of defense budgets is complicated by the fast paced and changing international
security environment, dominated by asymmetric threats.
The reality of simultaneously maintaining aging equipment, supporting new
equipment, adapting to a transforming military force, and coping with shifting
activities and activity levels compounds budgetary forecasting difficulties. The
most effective approach to mitigate these issues is to develop an institutional level
and in-depth knowledge and understanding of defense costs, as well as the estab-
lishment of a prioritized capability plan. Armed with this knowledge, politicians
and defense decision-makers can develop a long-term affordable plan to manage
and efficiently allocate resources to maximize national security.

7.3.2 Capability based planning


Using Canada as a case study, the central government decision-making process
can be illustrated as follows. The federal government’s policies and priorities are
often communicated, in broad terms, through the Speech from the Throne. Allo-
cating specific funds through the Federal Budget (often announced a few months
after the Speech) operationalizes specific priorities outlined in the Speech. While
the Speech gives the general short-term direction of the federal government, policy
statements and White Papers provide more precise expectations and directions of
the central government to specific departments or policy areas. The 2005 Defence
Policy Statement (DPS) and 2008 Canada First Defence Strategy (CFDS) are
examples of such direction, and a specific guide for defense spending.
CBP greatly enhances the information available to decision-makers and was
developed with minimal changes in the financial reporting mechanism. CBP pro-
vides the tools through which the Defence Services Program can be linked to
government policy and expectations (Chouinard and Wood 2007). CBP is employed
through a process that creates scenario specific force-wide capability goals.
From an economics perspective, this process is a high-level constrained opti-
mization whose objective is to select a mix of capabilities that achieves the
optimal allocation of resources (maximize national security subject to fiscal
constraints). CBP also explicitly encourages output-based planning by linking
capability decisions to high-level strategic goals. By focusing on forecasted
requirements for future capabilities to achieve the nation’s strategic goals, deci-
sions to upgrade existing systems or invest in new systems critically depend on
how these investments are likely to impact future capabilities and strategic goals.
CBP consists of three interrelated steps. The first step assesses the future
security environment. Step 2 deals with the linkage of the “concept of operations”
168 R. Fetterly, B. Solomon
to the capability framework and to validate the planning process. In Step 3,
an operations research technique is applied to create a priority list of relevant
activities for the given force development scenario (Department of National
Defence (DND) 2008). In general, CBP assists senior managers to establish pan-
Canadian Forces priorities within the budget, and to shift resources among defense
programs—and across the military services—from less to more productive uses
in response to changes in the future security environment. Another economic
interpretation of the CBP is linking the future security environment and the pol-
icy intentions of political leaders to decisions on force structure, readiness, and
modernization. This problem is equivalent to choosing a point on a “production
possibility frontier” that represents intertemporal tradeoffs between “consump-
tion” (current readiness) and “investment” (modernization or future readiness). In
fact, the resources allocation challenge of the DND is a subset of the Government
of Canada’s own challenge of funding its policies and programs within available
fiscal resources (Solomon et al. 2008).
CBP provides a clear linkage between military outputs and defense policy out-
comes that policy and decision-makers can use. However, CBP does not answer
the question of cost efficiency since there is no mechanism for ensuring gen-
uine competition among the various service providers or appropriate incentives
for minimizing costs. Certainly the right incentives in the form of gain-sharing
(a portion of savings given as a reward for minimizing costs) may induce culture
change in a public sector faster than resource budgeting (Melese 1997).

7.3.3 Lessons from capability based planning


CBP is now gradually starting to become a significant focus for planning staff at
national defense headquarters in Western nations. This also includes increasing
the effectiveness in the management of capital equipment life-cycle operations
and support costs (GAO 2010). This shift in emphasis towards a more integrated
approach to defense management requires increased involvement by leaders in
defense headquarters in the financial management of defense resources. Fur-
thermore, senior management need more comprehensive cost information when
making resource allocation decisions. This, in turn, alters and challenges the
manner in which the various staffs prepare senior defense officials for internal
decision-making forums.
This shift in emphasis can be categorized as conceptual, structural and financial
in nature. In essence, these developments are related to changes in how defense
departments decide to purchase new capital equipment, what they purchase, and
how departmental procurement organizations manage the acquisition process. In
order to support the development of a management framework that itself supports
a departmental investment plan that clearly responds to the direction provided in
government policy, the following mechanisms are suggested:

• A long-term cost model that integrates procurement, as well as operating and


personnel costs. Such a model can support the estimation and analysis of
Facing future funding realities 169
the affordability of future plans. In addition, structural changes are needed
within defense procurement organizations to consolidate the management
and oversight of many of the large complex capital projects under one
organization.
• Enhanced Comptrollership and financial management within defense depart-
ments. The unique role and activities of defense departments within national
government financial structures can be a challenge if defense is restricted to
only utilize the government’s standard financial reporting and management
processes.
• A greater degree of integration of financial information within strategic
levels of defense decision-making. This allows common linkages for con-
ceptual, structural and financial changes needed within defense headquarters.
This strengthening of defense administration processes also needs to extend
beyond defense departments.
• A foundation that incorporates departmental priorities with fiscal realities
through the alignment of financial resources with anticipated future require-
ments (DND 2007). This can be achieved through a range of assumptions
in which funding gaps are identified. This planning process articulates pre-
vious decisions, as well as the funding available in the fiscal framework to
support future requirements. The foundation is elaborated by a framework
that provides a series of financial envelopes for the primary defense cost
drivers. This includes personnel costs for military and departmental civil-
ians, capital equipment acquisition, as well as operations and maintenance
(O&M) costs. Estimate resource requirements with the explicit understand-
ing that conventional analysis would likely understate Defense Department
projections of future funding needs. The approach by the Congressional Bud-
get Office of including unbudgeted costs (CBO 2010) in projecting defense
resource requirements is an effective planning benchmark.
• Inclusion of medium-term transformation and modernization costs that incor-
porates the many new capabilities that will be introduced over the medium
term (say next ten years), and retirements of obsolete older systems. As a
planning document, it can be expected that funding is likely to be reallocated
between financial envelopes as plans are further defined.

7.3.4 Absorption rate


Forecasting a likely funding line into the future, and building a capability portfo-
lio constrained by this budget envelope, is only one element of a comprehensive
strategy of managing cost growth. The other key element is absorption rate, or the
demand for defense services.
An era of high operational tempo and demand for Western military forces began
in the early 1990s and has continued through the present. The paradox is that
the initial period in the post-cold war era brought a protracted series of substan-
tial budget cuts to defense at the same time that there was increasing demand
for peacekeeping and peace making operations. Despite a subsequent decade of
170 R. Fetterly, B. Solomon
budget increases to reflect this increased demand for defense services, the result is
continuing pressures on resources and capabilities.
To manage the budget reductions during the 1990s, while maintaining the
maximum level possible of military capability, the capital equipment budget and
personnel (military and departmental civilians) were reduced dramatically, numer-
ous bases were closed, and other major expenditures such as infrastructure repairs
and maintenance were deferred. Substantial re-investment in defense by Western
nations since September 11, 2001 strained the capacity of defense departments
to absorb and spend resources effectively. Whereas cuts to defense capability can
be made rapidly, rebuilding capability can take a number of years to regain the
experience and expertise needed to operate effectively.
The nature of defense activities is such that defense departments are prime can-
didates to employ multi-year strategies that over the long-term could be a more
optimal use of government funding than is currently the case. The unprecedented
deployment rate of military personnel in coalition operations overseas will likely
continue in response to the current international strategic environment, although
perhaps at a more sustainable rate in the future.
Any supplemental funding made available to defense at the end of a fiscal
year must go to capital, the operating budget or personnel. The ability to absorb
incremental funding on personnel is small due to lengthy hiring and recruiting
processes. For capital investments, unless the projects are well advanced and
ready for contract, the absorption rate is low to medium. The critical mecha-
nism for absorbing any supplemental funding is in maintenance O&M activities.
The “absorption rate”, or ability to spend any uncommitted or new incremental
funding, can be increased to the extent that a strategic planning process is in
place to identify and evaluate valuable capabilities and projects and to have them
ready for contracts as soon as funding comes available. The ability of defense
systems to reveal funding pressures in maintenance O&M activities, as well as
to ensure capital projects are in a position to be funded, are essential strate-
gies to anticipate opportunities for supplemental funding and to ensure value for
money.
The wide range of cost drivers in defense ranging from personnel costs, activity
rates and major equipment acquisition programs all contribute to demands on the
defense budget. Managing within a fixed funding allocation from government, a
significant possibility exists that considerable funds could lapse unexpended at the
end of the fiscal year. Uncontrollable events, slippage and delays in approvals lead
to unexpended funds that lapse at year-end.
Applying a Structural Unexpended Rate (SUR) as an over-programming mech-
anism to manage expected in-year under-expenditure is a strategy defense plan-
ners use through a balanced blend of risk management, spending controls and
resources management. The objective of applying over-programming at the SUR
is to plan early in the year to spend the entire defense budget incorporating
over-programming (called the planning level), rather than strictly managing to
the spending authority level. Implementing this over-programming mechanism
reduces the potential for lapsed funds at year-end and more importantly avoids
Facing future funding realities 171
a year-end spending rush while expending the funds wisely with due regard for
economy and efficiency.
In effect, enabling defense departments to “plan to spend” upcoming in-year
surpluses at the beginning of the fiscal year and paying down the SUR
over-programming throughout the year increases the effectiveness of resource
allocations. This is particularly true in cases where these funds are targeted to areas
that can generate future savings—such as increasing preventative maintenance of
infrastructure on defense installations, or increasing funding to the capital program
to replace equipment fleets before maintenance costs becomes excessive.
Legislative requirements to fully spend funding allocations prior to the end of
the fiscal year can be a powerful disincentive to effective resources management.
Government authority to carry-forward a certain level of the defense budget to the
next year could support more efficient and effective use of resources. In the current
defense resource environment, even a modest carry-forward authority for senior
resource managers will not provide the necessary flexibility to manage multi-year
projects or realign multi-year spending.

7.4 Cost drivers


In the event that the above-mentioned strategic management approaches, par-
ticularly CBP, are not truly institutionalized, the benefits of optimal resource
allocation will not be realized and the effectiveness of the armed forces risks being
eroded along with its purchasing power. One fundamental resources manage-
ment problem has come from high-level commitments by Western governments
to overseas operations, with a concurrent reticence to make critical disinvestment
decisions to relinquish capabilities and realign force structure based on these new
commitments within fiscal constraints.
Indeed, with the economic downturn in late 2008, changed fiscal realities have
moved the necessity to address a resources/capability gap to the near-term. Specif-
ically, the baseline funding in many Western countries to achieve a stable capital
equipment program and to sustain existing capabilities is likely insufficient at cur-
rent funding levels. To be sure, the ability of national security establishments
to “craft, implement, and adapt effective long-term strategies against intelligent
adversaries at acceptable costs has been declining for some decades” (Krepinevich
and Watts 2009). The present constraining factors on the defense budget are
defense-specific inflation, increasing personnel numbers and associated costs, as
well as aging equipment fleets and infrastructure. These factors increase pressure
on the defense budget and have the potential to draw limited defense resources
away from new defense programs. This challenge is amplified by the steady
growth in the cost of maintaining advanced military forces.

7.4.1 Personnel
Defense establishments are first and foremost people-driven organizations. As
a consequence, personnel costs in a defense budget are often both the most
significant and the least flexible expenditure category. The cost of military
172 R. Fetterly, B. Solomon
personnel and departmental civilians has an overwhelming impact on other
expenditure categories and drives much of the remaining defense spending.
Consequently, the preliminary and possibly most significant step in defense
decision-making is the determination of the required number of military
personnel. Forecasted Army, Navy, Air Force, and Marine Corps personnel, to a
large extent, guide the force planning process. The requirement for military bases,
quantities of military equipment, training facilities, as well as O&M expenses all
flow from decisions regarding personnel strengths. Government costs for military
personnel includes direct pay, environmental and operational allowances, and the
employer share of programs such as pensions and unemployment insurance.

7.4.2 Capital
The acquisition of capital equipment in defense is generally the most visible aspect
of defense spending, because of the substantial annual allotment of taxpayer dol-
lars and the expectations of industry to obtain contracts related to those planned
expenditures. To maintain a viable and effective military capability, it is essen-
tial to have a well-conceived capital equipment replacement plan that is supported
by a stable long-term financial commitment. And in order to develop a coherent
and effective plan for the acquisition of new and modernized equipment there are
always tradeoffs that must be made. Future tradeoffs include:

• number of personnel versus investment in high-technology equipment (capi-


tal/labor substitution);
• multi-purpose versus single purpose forces (i.e. specialized capabilities);
• regular force versus reserve force mix;
• the optimum balance between force elements (air, land, maritime and special
forces).

In terms of goods and services procured and personnel employed, defense depart-
ments spend a large percentage of the overall federal operating budget and capital
funds. This is significant because these categories of expenditure are highly visible,
generate employment, and to a certain extent are discretionary; accordingly these
expenditures are particularly vulnerable to spending cuts during budget crises.

7.4.3 Defense-specific inflation


Protection from the loss of purchasing power resulting from inflation in defense
goods and services, and from required compensation increases for military and
civilian personnel, are essential elements in forecasting future defense funding.
However, given the specific nature of defense, the GDP deflator does not cap-
ture the anomalies of the “defense market” and is thus an inadequate measure of
the inflation adjustment required to maintain stable defense purchasing power.
The traditional defense perspective is that defense-specific goods and services
experience greater price fluctuations during inflationary periods, which should
Facing future funding realities 173
be accommodated in both internal and external fiscal planning and budgetary
considerations.
Defense-specific inflation is different for three reasons. First, the defense market
is different from markets that produce civilian goods and services included in bas-
kets underlying most aggregate measures of inflationary price changes. Second,
imperfections in some segments of the defense market provide the potential for
greater price fluctuations than prevail elsewhere in the economy. Finally, exchange
rate fluctuations will have greater effect on defense, as defense is the direct final
consumer, and as such is also susceptible to importing inflation, whereas in the
economy as a whole, other factors tend to dampen the price level effects of
exchange rate fluctuations4 (Solomon 2003).

7.4.4 International deployments


Another potential cost driver is activities related to international deployments.
Defense departments in NATO alliance nations all face a common set of finan-
cial pressures. The cost of ongoing overseas operations, investing in readiness,
and intense investment in preparation for the future environment are creating sig-
nificant internal budget pressures. To “succeed in this era of fiscal constraint, new
ways of thinking, constructive change, and basic reforms are essential. Everything
must be on the table and subject to scrutiny” (Dodaro 2010).
Increasing interoperability and specialization among NATO allies could assist
in increasing alliance effectiveness and efficiency while lowering costs (Berkok
2005). Also, the growing involvement of non-governmental organizations (NGOs)
and international organizations, as well as regional organizations in assisting failed
and failing states since the end of the cold war has been significant and not always
fully recognized. This has resulted in defense planning “becoming denationalized
and far more plural” (Nelson 2001).

7.5 Modelling future defense budgets


Proper application of military CBA and CBP and other management strategies
require an information management system that can attribute direct and indirect
costs of military capabilities. Activity-based costing is one such approach. In the
absence of such a system, hybrid systems can be used.
For example, Solomon et al. (2008) develop a strategic cost model by attributing
expenditures in a given fiscal year to capabilities based on basic attribution rules.
Thus, a typical intermediate output such as land training consists of assets such as a
Land (Army) Warfare Centre, direct fire equipment such as tanks, artillery schools
and common training. The civilian and military personnel that work in each of
these organizations as well as their O&M costs make up the total cost of army
training. These are linked to a number of final army outputs (capabilities) such
as light infantry direct fire and brigade headquarters that use such training. The
attribution of this training activity is based on the proportion of personnel each of
the army capabilities will potentially use for the training. This proportion is based
174 R. Fetterly, B. Solomon
on the total cost of personnel in each of these final capabilities in the reference year
(Solomon et al. 2008). Unfortunately, such information management systems and
databases are not available in most nations and often decisions must be made in a
data- and time-constrained environment. The modelling section presented below
acknowledges these constraints.
Future defense budgets can be modelled using a variety of econometric tech-
niques. The relevance and appropriateness of each model depends on the research
question, data constraints and forecast horizon. The three econometric models
used in this chapter are selected with the above mentioned goals and constraints
in mind. We begin with the simplest model when data and governing theoretical
frameworks are severely lacking.

7.5.1 Defense budget forecast model #1: an application


of a univariate autoregressive integrated moving
average (ARIMA) model
Univariate (single variable) ARIMA models which are based on the Box–Jenkins
time series modelling process discussed in Appendix 7.1, describe a single series,
such as defense expenditures over time, as a function of its own past values (Box
and Jenkins 1970). As an illustration of the univariate ARIMA forecasting model
we use time series data of Belgian military expenditures (ME) in real dollars.
The data is extracted from the Stockholm International Peace Research Institute
(SIPRI) and NATO databases.
The ARIMA modelling process starts with the identification of the time series
and a differencing order of 1 was used to make the series mean stationary. The
autocorrelation and partial auto-correlation (ACF and PACF) functions of the
Belgian time series are calculated and shown in Appendix 7.1 (Figures A7.1
and A7.2). The initial identification seems to point to an autoregressive model
of order 1. Currently, there are a number of statistical software packages that can
automatically select models based on a series of diagnostic tests. One such soft-
ware is Autobox (AFS 2000). As indicated in Appendix 7.1, Autobox is also one
of the time series forecasting software packages that include automatic outlier
detections. In fact the software package identified 1988 as a level shift and 1991
as transitory (pulse) shocks each of which is explicitly incorporated in the final
forecasting model.
It is preferable if the modeller has a priori information on particular events that
may affect the time series in question. For example, if there were any events or
policy announcements in Belgium that might have affected defense expenditures,
then these should be explicitly incorporated in the model. Of the two dates selected
by the software, 1991 is consistent with Belgium’s decision to reduce defense
spending after the fall of the Soviet Union (Amara 2008).
The forecast horizons based on the univariate ARIMA models are presented
in Figure 7.3. The first model uses the real Belgian defense expenditures without
prior adjustments and the forecasts indicate almost flat growth (a small increase
in the range of 0.7 percent to 0.5 percent) in real defense expenditures. The
Facing future funding realities 175

60
Forercast Lower80% Upper80%
55

50
€000

45

40

35

30
2008 2009 2010 2011 2012 2013 2014 2015 2016 2017 2018 2019

Figure 7.3 Belgian defense expenditures forecast based on an ARIMA model (000
euros).

model based on defense expenditures in logarithmic scale forecasts a more modest


increase of about 1 percent in real terms for the next decade.
Stationarity (stationary time series with or without differencing) is an important
aspect of the ARIMA class of models. This characteristic constrains the predic-
tion of future values to be linear functions of the observations. Such a restriction,
often a sufficient approximation of reality, does not cover every aspect of real life
situations and thus a broader class of models should be entertained. The linearity
assumption is somewhat relaxed when using a transformation of the original series
as was done here for Belgian defense expenditures.
Unfortunately, forecasts based on univariate models cannot tell us whether
growth is due to increases in the armed forces (people), or major weapon systems
(capital). In addition, univariate models are not suitable for conducting simulation
exercises on what effects changes in threat, income or government policies might
have on overall defense budgets. For such broader simulation models with pol-
icy prescriptions we turn to a second set of models that rely on strong theoretical
frameworks based on the neoclassical social welfare maximization models.

7.5.2 Defense budget forecast model #2: military demand models


Military demand models assume defense budgets are a result of the derived
demand for national security, S. The most common theoretical specification for
the demand for ME is based on the neoclassical model of welfare maximization
(Smith 1989, 1995):

W = f (C, S, Z n ). (7.1)

This standard welfare function includes economic variables such as consumption


C and the population of a country,5 as well as security S and other variables Z n ,
176 R. Fetterly, B. Solomon
which parameterize shifts in the welfare function, such as doctrinal changes or the
politics of the party in power.
The limits on welfare maximization include the standard income constraint:

I = pc C + pm M (7.2)

where I is aggregate income often proxied by GDP; M and C are real military
expenditures and consumption, and the price for each are denoted by pm and
pc respectively. Note that alternate welfare function specifications will use mil-
itary and non-military expenditures and prices in the budget constraint model. For
detailed discussions on the empirical estimation of demand for military spending
models see Hartley and Sandler (1990) and Smith (1995).6
The other key constraint in the welfare maximization problem is the security
function. Security can be represented as

S = f (M1 , . . . , Mn , Z s ). (7.3)

In a Nash–Cournot specification the i th country will take the military expendi-


tures, M, of its (n − 1) allies as given when establishing its own security needs.
Note, however, that if other countries are hostile and pose a threat, this affects
country i ’s national security posture.
In a Nash equilibrium, the first-order condition from maximizing Equation (7.1)
subject to Equation (7.2) and (7.3) implies

S = f (I, p, Q, Z n , Z s ) (7.4)

Thus the security of a nation depends on its military forces and other countries’
forces and other strategic variables that may affect the security environment.
Looking at the ME of other countries, allies’ military expenditure may be seen
as “spillin”, thus


n
Q = Mit + M j t − Mit
j =1

where i = j . The resulting partial equilibrium determination of one country’s


forces, given those of the other nations, leads to the derived demand for ME, which
defines military budget forecast model #2:

MEt = β0 + β1 ( pm / pc ) + β2 I + β3 T h + β4 Q + β5 Z + εt . (7.5)

Recall that the ME of other nations elicit varied security postures depending
on whether one is an ally or a threat. The reaction is captured as either Th
(threat) or Q (“spillin”). Aspects that are unique to a nation, such as external con-
flict, new security policy or other exogenous effects are summarized as strategic
variables (Z ).
Facing future funding realities 177
The empirical work on demand for military spending falls largely into two
groups. The first group, popularized by the important work of Murdoch and San-
dler (1982), supports a well-specified theoretical model that generates the military
budget forecasting equation. The second group proposes an approach where the
estimating equation is determined by the best fit to the data, based on rigorous
statistical testing. Chowdhury (1991) is one example of the latter.
Recent studies have departed from these two groups in favor of a compromise
strategy. This strategy, adopted in this chapter, starts with a model based on a
general theory of the demand for military spending, followed by rigorous statis-
tical testing to investigate the relative importance of strategic and other economic
factors for the country(ies) under investigation. Smith (1989) and most recently
Dunne et al. (2003) and Solomon (2005) employ such a strategy.7
For this application of model #2, we use Belgian and Canadian ME for the
period 1955 to 2007 from SIPRI annual publications. To convert US and other
NATO members’ military spending data into constant dollars the IMF database
was consulted. Note that the GDP deflator of each country was used. All figures are
in US dollars. Belgium, Denmark, France, Germany, Italy, Luxembourg, Nether-
lands, Norway, and the United Kingdom were included in the NATO Europe
sample. This variable (NATO_Europe) proxies the potential spill-ins from allied
contribution. If the sign on this allied spill-in (Q) variable is negative, this sug-
gests that the nation (Belgium or Canada) is a free rider. However, Canada and
Belgium articulate their defense posture within the context of a multilateral secu-
rity arrangement and thus a positive response to allies’ ME is expected. To reflect
the choices faced by decision-makers between socio-economic and security needs,
government expenditures were obtained from the World Bank (removing ME from
the aggregate). A negative association is theoretically predicted.
Various databases were consulted to identify potential enemies, militarized
inter-state disputes and peacekeeping related expenditures and missions. Threat
variables include USSR (sample period 1952–1990 and Russia for the period
1991–2006) on which data were obtained from the SIPRI (various years) year
book, and Soviet/Russia Nuclear Equivalent Megatons and Intercontinental Ballis-
tic Missiles (data from Jim Finan, Royal Military College, Canada). The latter was
found to be a better proxy for the Soviet threat since the smaller NATO nations’
primary concern was strategic capability and the potential nuclear fallout.
Data on foreign aid were obtained from the Organization for Economic Coop-
eration and Development (OECD). In addition various dummy variables were
used to proxy various country-specific security policies. These were proxied
according to the dates of the introduction or suspension of government docu-
ments that articulated the nations’ security strategy. NATO doctrine changed from
Mutual Assured Destruction (1952–1969 = 1, 0 elsewhere) to Flexible Response
(1970–2001 = 1, 0 elsewhere; 1970 was used as the period of shift based on
statistical properties) and was also used as a possible determinant of military
spending.
Other standard variables from social welfare maximization, such as income
(GDP) and the relative price between military and civilian good are anticipated
178 R. Fetterly, B. Solomon
to be significant and show positive and negative effects respectively. This is
because defense is considered a normal good, and also reflects the tendency
for governments to economize on military spending when it becomes relatively
expensive. Note, however, that this variable is only estimated for Canada, for
which military price data is readily available. In addition, public finance the-
ory suggests that military spending may crowd out other government programs;
as such, a significant and negative coefficient for non-military expenditures is
anticipated.
Preliminary tests were conducted on the specification of the variables. Levels as
opposed to shares of the variables were used, as these better explain the nature of
the demand for ME (see also Hartley and Sandler 1995).8 The estimated long-run
coefficients for Belgium and Canada are presented in Table 7.3.
Examining results for Canada, country income, and US and European allies,
both contribute positively to military spending. Surprisingly the threat from inter-
continental nuclear attack is negative; this is discussed later in this section and
turns out to be somewhat consistent with the theoretical prediction of the model.
The price variable is not significant for Canada, indicating that rising military
prices have little or no impact on government ME (resource allocation) decisions.
On the other hand, the significance of the income variable for Canada suggests
that income constraints have important long-run impacts.
The long-run coefficients for Belgium are mostly not significant, and only the
threat variable proxied by nuclear arsenals was significant. This suggests that for
Belgium, short-run effects tend to govern defense spending.
Table 7.4 presents the short-run effects through the error correction representa-
tion of the demand model;9 see also Appendix 7.2.
It is important to point out that the short-term dynamics of the demand for
military spending models in Canada are governed mostly by a complementary
reaction to allied military spending, and budgetary inertia. The latter is evident for
all the nations considered, and especially when observed from the perspective of
peace support operations and defense procurement.
Peacekeeping and other military expeditionary operations, to be effective as a
military response to the threat of international instability, are by necessity medium

Table 7.3 Long-run coefficients

Constant Income Price Spill-in Policy Threat

Belgium 8.6 −1.6 n.a. −0.9 0.3∗∗ 0.11∗∗


NT
Canada −5.7∗∗∗ 0.22∗∗∗ 0.49 1.23∗∗∗ n.a. −0.13∗∗∗
US 0.27∗∗ NICM

Notes
∗∗∗
1% significance level ∗∗ 5% and ∗ 10%; n.a. = not applicable; n.s. = not significant.
Sample period 1955–2007.
Policy = Defense policy changes (dummy variables); NICM = Nuclear Intercontinental Missile
(USSR/Russia); NT = NICM plus surface; US = US Military Expenditures.
Facing future funding realities 179
Table 7.4 Error correction representation

Variables Belgium Canada

CONSTANT 1.28∗∗ −2.14∗∗∗


ME 0.85∗∗ 0.32∗∗∗
ME−1 n.a. −0.30∗∗∗
Q −0.14 0.45∗∗∗
Q(US) n.a. 0.10∗∗
Th(NTOT) 0.02∗∗ n.a.
Th(NICM)) n.a. 0.02
Th(NICM))−1 n.a. −0.071∗∗∗
Th(NEX) 0.02∗∗ n.a.
Z (policy) 0.02 n.a.
PRICE n.a. 0.18
GDP 0.24 0.08∗∗∗
Z (NONDEFENCE) −0.21 n.a.
ECM−1 −0.14∗∗ −0.37∗∗∗

Notes
Key: See Equation (7.5) for variable descriptions, and also Table 7.3.

10%, ∗∗ 5%, ∗∗∗ 1% significance.
 denotes the first difference of the variable; X −1 implies X −1 − X −2 .

to long-term propositions. Once commitments have been made by a nation to a


coalition expeditionary mission, it is difficult to disengage. Similarly defense pro-
curement is a lengthy process requiring several years to administer and deliver.
Another unique aspect of this dynamic is Canada’s short-term reaction to the
USSR/Russia’s strategic threat. While the change in the strategic threat is not
significant for Canada, it is the lagged change effect that is negative and signif-
icant. This suggests that Canada reacted to the nuclear threat after engaging in
other security changes not directly related to military spending (hardening targets,
infrastructure spending, etc.).
Belgian short-run dynamics are governed by inertia and the proxies for
threat (total nuclear arsenal and nuclear explosions). For the smaller allies, the
USSR/Russia strategic assets (nuclear power) were the main concern. This is con-
sistent with Hilton and Vu’s (1991) assertion that non-nuclear allies in NATO may
have perceived the Soviet threat beyond their capacity and thus tended to ignore it
in their decision-making. The error correction factor of 0.14 indicates a moderate
to long period adjustment to a shock. The general indication for Belgium is that
other country-specific factors govern the demand for ME. (This also suggests that
bureaucratic and political interpretation may provide important clues).
In general, models such as these based on the demand for ME can be used to
simulate future budget scenarios by individually calibrating GDP (income) growth
or changes in alliance posture or threat. As an example, Figure 7.4 shows a possi-
ble Canadian defense expenditures path given real GDP growth rate of 2 percent
for the next 10 years, holding other variables at constant growth. This implies that
defense will grow by about 0.44 percent in real terms given the elasticity estimate
180 R. Fetterly, B. Solomon

13
CANME2 Forecast
12

11
$B (nominal Canadian $)

10

4
19 te
19 6
19 8
19 0
19 2
19 4
19 6
19 8
19 0
19 2
19 4
19 6
19 8
19 0
19 2
19 4
19 6
19 8
19 0
19 2
19 4
19 6
20 8
20 0
20 2
20 4
06
20 8
20 0
20 2
20 4
16
5
5
6
6
6
6
6
7
7
7
7
7
8
8
8
8
8
9
9
9
9
9
0
0
0

0
1
1
1
a

20
D

Figure 7.4 Canadian defense expenditures forecast based on constant GDP growth
(2 percent).

of 0.22. Given the long-run estimate of GDP inflation at 1.5 percent, the forecast
for nominal defense growth is around 1.9 percent.

7.5.3 Defense budget forecast model #3: Budget components


based models
The preceding two models are based on strategic or macro-level assessments of
ME. In this section, we offer a micro-level perspective that decomposes the ME
data into the subcomponents that make up the defense budget. Broadly defined,
these budget components (or defense appropriations categories) include people,
capital and O&M costs. As shown in Table 7.5 each of these budgetary compo-
nents are further divided to include type of personnel (military officers, conscripts,
civilian and part-time soldiers), capital (equipment and infrastructure) and O&M
attributed to people, equipment and infrastructure. Note that the variations in the
primary inputs of labor and capital are due to technology and other cost drivers
discussed earlier.
This is also the model we recommend when a strategic management system
(such as CBP) is fully institutionalized, with an information management database
linked to activity- and capability based costing as discussed in earlier sections. The
model application again uses Canadian data. Following the budget identity equa-
tion in Treddenick (2000) for Canada, the defense budget can be disaggregated
into several primary input variables including personnel (P), weapons (capital or
equipment (E)), infrastructure (I ), and O&M.
Facing future funding realities 181
Table 7.5 Key variables to forecast or simulate

Personnel
Military (officers and enlisted civilian)
Reserve or other non-regular
Inflation_Pay
Operations & maintenance
People related
Equipment
Infrastructure
Inflation_O&M
Capital
Equipment
Infrastructure
Technology
Procurement policy
Defense inflation
Statutory and transfers
Personnel related
Other transfers
Exogenous
Federal budgets
Defense White Papers
Other government initiatives
Grand total

Personnel costs can be further disaggregated into military and civilian personnel
costs,

Pt = pm Pmil + pci Pciv (7.6)

where ( pm Pmil ) is number of military personnel times the wage rate, and similarly
civilian personnel costs are ( pci Pciv ). Equipment costs are represented as

E t = (1 − αe)E t−1 + E Ct (7.7)

and infrastructure costs as

I f t = (1 − βw)I f t−1 + I Ft , (7.8)

where α and β are parameters.


While the above depicts possible weapons accumulation, with E Ct depicting
weapons expenditures, and I Ft infrastructure expenditures, both at time t, it is
possible to extend the mathematical formulation to include lagged adjustments to
a desired/planned future stock.
In some nations, there are legal limits on the armed forces, in others where
conscription has been replaced by voluntary forces, recruitment and retention
182 R. Fetterly, B. Solomon
requires careful calibration. In the Canadian case, the 2008 CFDS explicitly con-
strains the size of the personnel component to about 68,000 to 70,000 regular
forces (enlisted and officers), 23,000 civilians, and about 30,000 reserves. Given
an activity-based costing or an information management system that can be manip-
ulated to calculate the cost of a military or civilian personnel, one can forecast
future personnel-related costs.
The key cost drivers for personnel are inflation (both economy-wide and
defense-specific), demography, and economic opportunities. The last two affect
the available pool of recruits and/or retention of the most desired cohort. From
the perspective of forecasting future defense budgets, the relevant question is the
current strength of the armed forces and if this number is:

1 below the required goal, how long will it take to achieve the desired level given
the attrition rates in basic training and regular attrition in the existing force?
2 above the required level, how would the reduction be achieved (e.g. to reduce
recruitment until regular attrition reduces the number, or design an accelerated
reduction program)?

In essence, the challenge is to estimate the growth/reduction in personnel


until a steady state is achieved, and then use inflation rates to estimate future
budgetary requirements for that steady state force. As mentioned in earlier
sections, if one assumes defense-specific inflation, personnel component costs
must be increasing faster than the civilian sector even after adjusting for produc-
tivity (quality) gains. Since measuring productivity is difficult in the government
sector, the best inflation rate to use is the economy-wide inflation rate, and infor-
mation from negotiated wage increases between public sector unions and the
government.
For capital projects the assumption is that the defense department uses a strate-
gic management technique that involves military CBA discussed elsewhere in this
book, CBP or other management techniques. The key message from these man-
agement techniques is that the rationale for acquiring a platform or a weapon
system is based on a clear need linked to defense policy or other relevant directive
and the legal control of expenditures (Melese et al. 2004). Again, for forecasting
purposes we assume that the list of capital programs for the forecast period are
selected using such management techniques and the relevant forecast drivers are
as follows.

• National policies on procurement. As discussed earlier, nations tend to lever-


age defense acquisitions to satisfy multiple objectives that are often mutually
exclusive and unrelated to defense objectives. As indicated in Solomon
et al. (2008), industries likely impose premiums to address government man-
dates like the Canadian Industrial Regional Benefit Program (IRB) that are
estimated to be in the range of 13–22 percent.
Facing future funding realities 183
• Technology. Kirkpatrick (1995) and Pugh (2007) offer excellent sources for
the impact of technology on the cost of weapon systems. Specifically, Kirk-
patrick (1995) has found that equipment costs have been rising by about 10
percent per year in real terms. Nations may want to customize the impact of
technology by examining their own nation’s trends in costs. Solomon (2009,
Table 6.8) has found a rate of increase of about 6–7 percent for selected air
force platforms in Canada.
• Defense inflation. For those nations that have estimates of defense inflation
the forecasted increase in this inflation rate can be used, and for those nations
that do not explicitly measure defense inflation, the GDP deflator is a good
proxy. Note also that for forecasting purposes one can also use the mid-point
of the inflation target used by the Central Bank. For example, if the Bank of
Canada inflation target is currently in the range 1–3 percent, 1.5 percent can
be used to project future inflation rates.

Finally, O&M costs can be forecast using projected inflation rates. Note that
O&M expenditures depend on the number of personnel (P), infrastructure (I ),
and weapon stocks (E):10

O Mt = δ0 + pOM I f + pOM E + pOM (Pmil + Pciv ). (7.9)

The model is applied to Canada. Baseline personnel numbers and expected cap-
ital acquisitions are taken from Canada’s departmental plans and priorities (DND
2009). As shown earlier in Figure 7.2, the Federal Budget has reduced its own
20-year projection for defense, and this information is coupled with the fact that
personnel numbers will hold steady.
Personnel-related costs are projected 20 years out (Figure 7.5). Projections for
capital will depend on the ability of DND to re-scope and re-phase projects given
the absorption challenges discussed previously. Specifically a report to Parliament
indicated DND did not have the capacity to absorb the announced increases to
its budget, and the result was a funding lapse of approximately $300 million
(1.6 percent of the total budget; DND 2009). Projections on personnel, readi-
ness, infrastructure and equipment are constrained by the CFDS 20-year funding
ceiling but year to year growth is based on the 1.5 percent and then 2 percent
increase expected by DND (DoF 2010). Figure 7.6 shows overall future funding
projections for the defense budget by calibrating policy, technology and inflation
projections.11
It should be reiterated that the funding line (budget envelope) projected here
may or may not equal derived demand since the funding line will always be con-
strained by legislated personnel numbers, approved inflation funding, and force
acquisition and elimination decisions based on CBP management decisions.
On the demand side the attribution rules generate, in input-output model par-
lance, technology coefficients that are updated annually. Long-term projections
remain fixed on these updated technology coefficients. To the extent the mod-
eller receives a priori information about a new in-service support arrangement
184 R. Fetterly, B. Solomon

14.00
Pers Cost Pers Cost +Inflation
12.00

10.00
$ Billions

8.00

6.00

4.00

2.00


2008–09
2009–10
2010–11
2011–12
2012–13
2013–14
2014–15
2015–16
2016–17
2017–18
2018–19
2019–20
2020–21
2021–22
2022–23
2023–24
2024–25
2025–26
2026–27
2027–28
Figure 7.5 Personnel growth projections with and without inflation adjustment.
Source: Initial Personnel Numbers from DND (2009). Inflation projection based on DND (2008).

35
Demand
30

25
$ Billions

20

15

10

0
2008–09
2009–10
2010–11
2011–12
2012–13
2013–14
2014–15
2015–16
2016–17
2017–18
2018–19
2019–20
2020–21
2021–22
2022–23
2023–24
2024–25
2025–26
2026–27
2027–28

Figure 7.6 Demand based on cost driver.

or training that may affect factors such as the attribution of training, material
and infrastructure support, these can be incorporated in long-term projections.
For example, if the future security environment and capability planning favors
an emphasis on the Arctic, then one example of “early warning” data is in-service
support arrangements for Arctic patrol capability.
Facing future funding realities 185
While this final model is closely linked to the strategic management strategy
we outlined in preceding sections, each of the forecasting models discussed here
have their merits depending on the maturity and availability of strategic and
management information systems.

7.6 Conclusions and future directions


Long-term defense budget planning—under any circumstances—is challenging.
The unique requirement to deal simultaneously with a continually shifting inter-
national security environment and the peaks and valleys of global business cycles
adds to the complexity of trying to manage a multi-billion dollar defense portfo-
lio. Under these conditions, the optimal strategy is to focus staff energies on the
development and execution of a flexible and balanced approach that is aligned
with government policy, has the necessary funding, and is supported by exercising
careful and proficient management of the established plan.
CBP and other strategic management strategies discussed elsewhere in this
book are important, but institutionalizing such systems has proven problematic.
The literature on transaction cost economics and public choice may offer some
important clues. Of the three defense budget forecasting models introduced, mod-
elling the derived demand for military expenditures to forecast future defense
budgets may prove the most useful. When future threats are not easily defined,
a nation’s defense posture is often governed by its bilateral and multilateral secu-
rity arrangements, income constraints, and policy directives. Introducing future
funding realities is a critical component of any military CBA.

Appendix 7.1: Box–Jenkins method


The process (Box–Jenkins) is a three-stage iterative process that reduces a given
time series with underlying structure to pure randomness (white noise). The three
stages are identification, estimation, and forecasting.
The identification stage entails choosing a preliminary model based on sim-
ple visual aids such as plots of the time series and selected key statistics such
as the autocorrelation and partial autocorrelation of the time series. These key
statistics are used to make the series stationary, where the mean and variance
are constant over time. Mean stationarity is often achieved by differencing, and
variance stationarity by applying the correct power transformation such as the log-
arithmic or square root of the original time series (Kennedy 1998). To determine
the order of the differencing, the number of time periods between the relatively
high autocorrelation is usually a good indicator (see Figures A7.1 and A7.2 for
example). For variance stationarity the power transformation can be obtained
from a flexible family of transformations introduced by Box and Cox (1964;
Kennedy 1998).
Once a tentative model is identified, a nonlinear least squares or a maximum
likelihood estimation based on the Marquardt Algorithm, is employed (Box–
Jenkins 1970, 504–505). A pure auto regressive (AR 1) process sometimes known
186 R. Fetterly, B. Solomon

1.2
ACF
1

0.8
Autocorrelation

0.6

0.4

0.2

0
1 2 3 4 5 6 7 8 9 10 11 12
–0.2
Lags

Figure A7.1 The autocorrelation function plot—Belgian defense spending.


Source: Box and Jenkins (1970).

-
1.5

1
0.5 I
I I •I -I •
-
0
I I -
Autocorrelation

,
-o.s 1
" n 1 '>

-1
-1.5
-2

-2.5
-3
• PACF
-3.5
Lags

Figure A7.2 The partial autocorrelation function plot—Belgian defense spending.


Source: Box and Jenkins (1970).

as a “random walk” model can be estimated by linear methods. Whether the tenta-
tive model is appropriate for forecasting is assessed through a series of diagnostic
checks or tests known in Box–Jenkins parlance as necessity, invertibility, and
sufficiency tests. Each parameter included in the model should be statistically sig-
nificant (necessary) and each factor must be invertible. In addition, the residuals
from the estimated models should be random (model sufficiency).
The test for necessity is performed by examining the T-ratios for the indi-
vidual parameter estimates. Parameters with non-significant coefficients may be
deleted from the model in order to have a parsimonious model. Invertibility is
Facing future funding realities 187
determined by extracting the roots from each factor in the model. All the roots
must lie outside of the unit circle. If one of the factors is non-invertible, then the
model must be adjusted. The appropriate adjustment is dictated by the type of the
factor that is non-invertible. For example, a non-invertible autoregressive factor
usually indicates under-differencing, while a non-invertible moving average fac-
tor may indicate over-differencing. A non-invertible moving average factor could
also represent the presence of a deterministic factor. Since the model structure is
not really clearcut, the overall model must be considered when adjusting for non-
invertibility (Box and Jenkins 1970). The residuals are then tested for white noise
by studying the autocorrelation and partial autocorrelation of the residuals. Fur-
thermore, a test statistics Q or “portmanteau test” is performed on the residuals
autocorrelations of all lags. If the model is mis-specified or inappropriately fitted,
the Q test tends to be inflated (Box and Jenkins 1970; Kennedy 1998).
The general multiplicative Box–Jenkins model as articulated in its standard
form is (see Box and Jenkins 1970)

ϕ p (B) P (B s )∇sD Z t = θq (B)


Q (B s )at (A7.1.1)

where s denotes periodicity of the seasonal component (4 for quarterly and in our
case annual which is 1); B denotes the backward operator, i.e.

B Z t = Z t−1
(B )Z t = (B s )Z t−s ;
s

∇ d = (1 − B)d is the ordinary difference operator of order d; D ∇s = (1 − B s ) D


is the seasonal difference operator of order D; ϕ p (B) and  p (B s ) are station-
ary autoregressive operators (they are polynomials in B of degree p and in B s of
degree P, respectively); θq (B) and
Q (B s ) are invertible moving average opera-
tors (they are polynomials in B of degree q and in B s of degree Q, respectively);
at is a purely random process (Box and Jenkins 1970). The general multiplicative
model (1.0) is said to be of order ( p, d, q) (P, D, Q)s .
After the selection of the appropriate model that passes the relevant diagnostic
checks, the time series model is forecasted. Specifically, the forecasting function
of a typical random walk multiplicative model (1,0) can be expressed as

Z t+l = Y1 Z t+l−1 + · · · + Ym Z t+l−m − at+l − P1 at+l−1 . . . Pn at+l−n (A7.1.2)

where m = ( p + s.P + d + s.D) and n = (q + s.Q); ψ(B) = ϕ p (B s )∇ d ∇sD is


the general autoregressive operator; π(B) = θq (B)
Q (B) is the general moving
average operator.
Looking at the explicit form of the forecasting function, for l > n = (q + s.Q),
the conditional expectation of (2) at time t is

Z t (l) − 1 Z t (l − 1) − m Z t (l − m) = 0.
188 R. Fetterly, B. Solomon
For l > m and the solution of this difference equation is

(t) (t)
Z t (l) = b0 (t) f0 (l) + b1 f 1 (l) + · · · + bm−1 fm−1

for l > n − m. Note also that f represents functions of the lead time l and includes
polynomials, exponentials, etc. The general autoregressive operator ψ(B) defined
above, which determines the mathematical form of the forecasts function, i.e. the
nature of f . In other words, it determines whether the forecasting function is to
be a polynomial, a mixture of sines and cosines, a mixture of exponentials or
some combinations of these functions. As indicated in Box and Jenkins (1970),
the above is called the “eventual forecast function”, and since n > m it provides
the forecasts only for lead times l > n − m.

Outlier detection
Most time series based forecasting software packages include a routine that
automatically detects the presence of outliers (isolated events such as strikes,
earthquakes, etc.) or “shocks,” and whether these shocks are transitory or induce a
level shift. Often such events are modelled as “dummy” variables with a series of
zeroes and one(s) for the time period(s) of the isolated event. The automatic out-
lier detection algorithm is obviously better than the theory-based method for cases
when the modeller does not have a priori knowledge of when the event may have
occurred. In addition, if the outlier causes a structural shift (such as new institu-
tional legislation, defense White Papers, etc.) and if the impact of the event takes
a longer time lag to affect behaviour, then the modeller risks biasing the effect of
the outlier by choosing the day the event occurred as the break in the series.
On the other hand, theory-free methods have the “potential for finding spurious
significance” during the testing of the hypothesis (AFS 2000). However, since many
of these events may or may not be identifiable to the modeller, a theory-free detection
algorithm is often desirable. The algorithm begins by first fitting a univariate ARIMA
model to the series divided by a series of regressions at each time period to test the
hypothesis that there is an intervention. After identifying an outlier the residuals are
modified and the test will resume until all outliers are uncovered.
Some forecasters (see the 1990 AFS manual) argue that fitting a univari-
ate ARIMA model does not necessarily imply that the series is homogeneous
(the error term of the series is random about a constant mean). In such situations,
the identification of the outliers become dubious due to the recursive process. The
suggested solution is the testing of a more rigorous specification, i.e., the mean of
the errors must be near zero for all time sections (AFS 2000). As an alternative,
an endogenous determination of the timing of structural breaks can be more infor-
mative. Lanne et al. (2002) developed an Augmented Dickey–Fuller (ADF) type
test on such a method by considering shift functions of a general nonlinear form
that is used to add to the deterministic term. Lütkepohl and Saikkonen (2002) also
argue that a shift may occur in various forms ranging from a shift that is spread
Facing future funding realities 189
out over a number of periods to a smooth transition to a new level. This method is
employed in the next section to supplement dummy variable selection.

Appendix 7.2: The autoregressive distributed lag (ARDL) approach


to cointegration
As discussed in Appendix 7.1, socio-economic time series are not trend stationary.
In econometric parlance, such variables are said to be integrated of order 1 (I (1)
for short) and regression analysis based on these variables may lead to spurious
results with high R 2 .
One typical cointegration technique for estimating time series based models is
the autoregressive distributed lag (ARDL) approach to cointegration introduced
by Pesaran et al. (1996) and Pesaran and Shin (1999). Estimating a model using
the ARDL approach to cointegration means estimating a model of the form


k
θ (L, p)yt = α1 + α2 Tt + βi (L, qi )x it + εt (A7.2.1)
i=1

where: x it are exogenous variables; Tt is a deterministic time trend; and


(L, p)
and βi (L, qi ) are polynomial lag operators, with maximum lag of p and qi
respectively (Pesaran and Pesaran 1997). The order of the distributed lag func-
tion on yt and the forcing variable x t are selected using the Schwartz Bayesian
Criterion (SC).
The Pesaran and Shin (1999) test consists of adding, in the first differenced
version of Equation (7.5), lags of first differences of the variables so as to
orthogonalize the relationship between the explanatory variables and the residual
term g. Testing for cointegration then amounts to a F-test on the joint statistical
significance of adding level regressors of the variables suspected to be cointe-
grated. Under the null hypothesis of no cointegration, the distribution of such an
F-statistic is non-standard; so the usual critical values do not apply. Pesaran et al.
(2001), tabulate the relevant critical bounds for I (0) and I (1). Instead of the con-
ventional critical values, this test involves two asymptotic critical value bounds,
depending on whether the variables are I (0) or I (1) or a mixture of both. If the test
statistic exceeds their respective upper critical values, then we can reject the null of
no cointegration regardless of the order of integration of the variables and there is
evidence of a long-run relationship; if below we cannot reject the null hypothesis
of no cointegration; and if it lies between the bounds, inference is inconclusive.
The error-correction form of (7.9) can be represented as


n 
n 
n
M E t = β0 + αM E t−i + χ( pm / pc )t−i + φQ t−i
i=1 i=1 i=1


n 
n 
n
+ γ T h t−i + ϕIt−i + ηM E t−i + εt .
i=1 i=1 i=1
190 R. Fetterly, B. Solomon
A variable addition test where the lagged values of the level variables repre-
sented as

θ1 M E t−1 + θ2 ( pm / pc )t−1 + θ3 Q t−1 + θ4 Z t−1 + θ4 It−1 .

One can use the variable addition test to fit a more parsimonious model of demand
for military expenditures. For example, if by dropping a variable x from the model
it ceases to be cointegrated, then we can infer that x has a significant effect on
the demand for military spending in the long run. However, if the reverse holds,
then one can remove the variable. The use of the F-test to eliminate variables is
similar to one employed by Beenstock (1998) to test significance of variables for
cointegration in Israel’s demand model. Specifically, if the null hypothesis testing
for long-run relations is sensitive to the removal of the variable in question, then
one can safely assume that the variable in question is likely a long-run forcing
variable explaining military expenditure.

Notes
1 The views expressed in this chapter are the views of the authors and do not necessar-
ily reflect the views or policies of the Canadian Armed Forces or the Department of
National Defence.
2 National defense is a complex, high consequence of error, capital-intensive, knowledge-
dependent, national security instrument. Categorization of defense in this manner results
from the diversity of tasks assigned to the defense department and the military forces, the
elevated level of technology employed within these organizations, the aggregate size of
the budget, and the considerable annual capital investment in weapon systems. Indeed,
management of the defense budget “is a highly constrained exercise in pricing the exe-
cutability of programs within the parameters of affordability and political feasibility”
(Jones 1991, 20).
3 While most of the examples used in this chapter have a Canadian flavor, selected NATO
member nations are also discussed where data and information are available.
4 There is considerable debate on this issue. While Solomon (2003) makes the case for
defense inflation for Canada as does Kirkpatrick (2008) for the United Kingdom, a
special issue of the Royal United Services Institute (RUSI) dedicated a special issue on
the subject (RUSI 2009) that needs to be considered to assess whether a separate index
of inflation is required in defense costing.
5 The addition of population may seem peculiar given the fact that defense is a pub-
lic good; however, consumption probably is not and as such, per capita consumption
is assumed to matter for welfare. In other words, demand for defense spending may
increase if it has a high-income elasticity and increase in population reduces the tax cost
faced by the median voter (Dunne et al. 2003).
6 Most empirical studies in the literature, citing practical and conceptual difficulties asso-
ciated with military price indices, make the assumption that the relative price differential
between the military and civilian price is constant. Solomon (2003, 2005) discusses the
arguments for and against this view as well as the implications for empirical work.
7 The conventional econometric method for estimating time series models is based on
cointegration. (See Appendix 7.2.)
8 The log-transformation of the variables was preferred, based on non-nested tests. As
such, these results are presented.
9 The ARDL (autoregressive distributed lag) estimates and the lag order were selected
using the Schwarz Bayesian Criterion. The error correction term enters the equation with
Facing future funding realities 191
the expected negative sign and is significant at the 1 percent and 5 percent significance
for both countries considered. The results from the model in general do not point to any
econometric problems, as the model successfully passes all the diagnostic tests.
10 If the information system or costing model allows further disaggregation then personnel
can be limited only to operational and support personnel.
11 Initial calibrations are based on the official data from DND Report on Plans and Pri-
orities (FY 2008–09) Section 3 Supplementary Tables A&B as well as Table 13 on all
approved major capital projects.

References
Amara, J. 2008. “NATO Defense Expenditures: Common Goals Or Diverging Interests? A
Structural Analysis.” Defence and Peace Economics 19(6): 449–469.
Automatic Forecasting. Systems-AFS. 2000, 1990 and 1986. User’s Manual for Autobox
Software. Hatboro: Pennsylvania.
Beenstock, M. 1998. “Country Survey XI: Defence and the Economy in Israel.” Defence
and Peace Economics 9(3): 171–222.
Berkok, U. 2005. “Specialization in Defence Forces.” Defence and Peace Economics 16(3):
191–204.
Box, G. E. P. and D. R. Cox. 1964. “An Analysis of Transformations.” Journal of the Royal
Statistical Society Series B 26: 211–243.
Box, G. E. P. and G. Jenkins. 1970. Time Series Analysis: Forecasting and Control (rev. 1st
edn). San Francisco: Holdenday.
Center for Strategic and International Studies. 2005. European Defense Integra-
tion: Bridging the Gap between Strategy and Capabilities (Washington: Cen-
ter for Strategic and International Studies). Available at www.csis.org/media/
csis/pubs/0510_eurodefensereport.pdf [accessed April 12, 2007].
Chouinard, P. and I. Wood. 2007. The Department of National Defence Strategic Cost
Model: Development. DND Centre for Operational Research & Analysis, Ottawa, p. 2.
Chowdhury, A. R. 1991. “A Causal Analysis of Defence Spending and Economic Growth.”
Journal of Conflict Resolution 35: 80–97.
Congressional Budget Office. 2010. Long-Term Implications of the Fiscal Year
2010 Defense Budget (CBO, Washington). Available at www.cbo.gov/ftpdocs/
108xx/doc10852/01-25-FYDP.pdf [accessed July 23, 2010].
Department of Finance. 2010. The Federal Budget 2010. Ottawa: Public Works and
Government Services Canada.
Department of National Defence. 2007. Investing in the Future: A Ten Year Strategic Invest-
ment Plan Framework for the Department of National Defence. Ottawa: Department of
National Defence.
———. 2008. Canada First Defence Strategy. Ottawa: Department of National Defence.
———. 2009. Report on Plans and Priorities. Ottawa: Department of National Defence.
Dodaro, Gene L. 2010. Maximizing DOD’s Potential to Face New Fiscal Challenges and
Strengthen Interagency Partnerships. Government Accountability Office, Washington,
p. 2. Available at www.gao.gov/cghome/d10359cg.pdf [accessed July 5, 2010].
Dorff, Robert H. 2005. “Failed States After 9/11: What Did We Know and What Have We
Learned?” International Studies Perspectives 6(1): 20–34.
Dunne J. P., E. Nikolaidou and N. Mylonidis. 2003. “The Demand for Military Spending in
the Peripheral Economies of Europe.” Defence and Peace Economics 14(6): 447–460.
GAO. 1994. Future Years Defense Program: Optimistic Estimates Lead to Billions
in Over-programming. GAO, Washington. Available at www.legistorm.com/showFile/
192 R. Fetterly, B. Solomon
L2xzX3Njb3JlL2dhby9wZGYvMTk5NC83/ful24551.pdf [accessed July 28, 2010].
———. 2005. Defense Acquisitions: Improved Management Practices Could Help Min-
imize Cost Growth in Navy Shipbuilding Programs GAO, Washington). Available at
www.gao.gov/new.items/d05183.pdf [accessed July 4, 2010].
———. 2010. The Federal Government’s Long-Term Fiscal Outlook: January 2010.
GAO, Washington. Available at www.gao.gov/new.items/d10468sp.pdf [accessed July
19, 2010].
Hartley, K. and Sandler, T. (Eds). 1990. The Economics of Defence Spending: An
International Survey. London: Routledge.
———. 1995. Handbook of Defence Economics, Volume 1 Amsterdam: North Holland.
Hilton, B. and A. Vu. 1991. “The McGuire Model and the Economics of the NATO
Alliance.” Defence Economics 2(1): 105–121.
International Monetary Fund (IMF). Various years. World Economic Outlook Database.
Jones, L. R. 1991. “Policy Development, Planning, and Resource Allocation in the
Department of Defense.” Public Budgeting and Finance 11(3): 15–27.
Jones, L. R. and Jerry L. McCaffery. 2005. “Reform of the Planning, Programming, Bud-
geting System, and Management Control in the US Department of Defense: Insights
from Budget Theory.” Public Budgeting & Finance 25(3): 1–19.
Jordan, Leland G. 2000. “Systematic Fiscal Optimism in Defense Planning.” Acquisition
Review Quarterly (Winter): 47–62.
Kennedy, P 1998. A Guide to Econometrics (4th edition). Cambridge: MIT Press.
Kirkpatrick, D. 1995. “The Rising Unit Costs of Defence Equipment – the Reasons and the
Results.” Defence and Peace Economics 6(4): 263–288.
———. 2004. “Trends in the Costs of Weapon Systems and the Consequences.” Defence
and Peace Economics 15(3): 259–273.
———. 2008. “Is Defence Inflation Really as High as Claimed?” RUSI Defence Systems
11(2): 66–71.
Krepinevich, Andrew F., and Barry D. Watts. 2009. Regaining Strategic Competence.
Center for Strategic and Budgetary Assessments, Washington, p. vii. Available at www.
csbaonline.org/4Publications/PubLibrary/R.20090901.Regaining_Strategi/
R.20090901.Regaining_Strategi.pdf [accessed August 21, 2010].
Lanne, M., H. Lütkepohl, and P. Saikkonen. 2002. “Comparison of Unit Root Tests for
Time Series with Level Shifts.” Journal of Time Series Analysis 23: 667–685.
Lütkepohl, H. and P. Saikkonen. 2002. “Testing for a Unit Root in a Time Series with a
Level Shift at Unknown Time.” Econometric Theory 18: 313–348.
Melese, F. 1997. “Gain-Sharing, Success-Sharing and Cost-Based Transfer Pricing: A New
Budgeting Approach for the Department of Defense.” Military Operations Research
3(1): 23–50.
Melese, F., J. Blandin, and S. O’Keefe. 2004. “A New Management Model for Government:
Integrating Activity Based Costing (ABC), the Balanced Scorecard (BSC), and Total
Quality Management (TQM) with the Planning, Programming and Budgeting System
(PPBS).” International Public Management Review 5(2): 103–131.
Murdoch, James C. and Todd Sandler. 1982. “A Theoretical and Empirical Analysis of
NATO.” Journal of Conflict Resolution 26(2): 237–263.
Nelson, Daniel S. 2001. “Beyond Defence Planning.” Defence Studies 1(3): 25–36.
Pesaran, M. H. and B. Pesaran. 1997. Working with Microfit 4.0: Interactive Econometric
Analysis. Oxford: Oxford University Press.
Pesaran, M. H. and Yongcheol Shin. 1999. “An Autoregressive Distributed Lag Modelling
Approach to Cointegration Analysis.” In Econometrics and Economic Theory in the 20th
Facing future funding realities 193
Century, edited by S Strom. The Ragnar Frisch Centennial Symposium, Chapter 11.
Cambridge: Cambridge University Press.
Pesaran, M. H., Y. Shin, and R. J. Smith. 1996. “Testing for the Existence of a Long-
Run Relationship.” DAE Working Papers Amalgamated Series, No.9622, University of
Cambridge.
———. 2001. “Bounds Testing Approaches to the Analysis of Level Relationships.”
Journal of Applied Econometrics 16: 289–326.
Pugh P. G. 2007. Source Book of Defence Equipment Costs. London: Pugh.
Royal United Services Institute (RUSI). 2009. “Defence Inflation: Myth or Reality.”
Defence Systems 12(1): 12–21.
SIPRI Yearbooks. Various years. Armaments, Disarmament and International Security.
Stockholm International Peace Research Institute. Available from http://www.sipri.
org/yearbook [last accessed February 9, 2015].
Smith, R. 1989. “Models of Military Expenditure.” Journal of Applied Econometrics 4(4):
345–359.
———. 1995. “The Demand For Military Expenditure.” In Handbook of Defence Eco-
nomics, Volume 1, edited by K. Hartley and T. Sandler. Amsterdam: North Holland.
Solomon, B. 2003. “Defence Specific Inflation: A Canadian Perspective.” Defence and
Peace Economics 14(1): 19–36.
———. 2005. “The Demand For Military Expenditures in Canada.” Defence and Peace
Economics 16(3): 171–189.
———. 2009. “The Defence Industrial Base in Canada”, In The Public Management of
Defence in Canada, edited by Craig Stone. Toronto: Breakout Educational Network.
Solomon, B., P. Chouinard, and L. Kerzner. 2008. The Department of National Defence
Strategic Cost Model: Volume 2 – Theory and Empirics. DRDC-CORA TR2008-03,
Ottawa.
Stone, J. C. and B. Solomon. 2005. “Canadian Defence Policy and Spending.” Defence and
Peace Economics 16(3): 145–169.
Thompson, M. 2007. Improving Defence Management. Canberra: Australian Strategic
Policy Institute.
Treddenick, J. M. 2000. “Modelling Defense Budget Allocations: An Application to
Canada.” In The Economics of Regional Security: NATO, The Mediterranean, and South-
ern Africa, edited by Jurgen Brauer and Keith Hartley, pp. 43–70. Amsterdam: Harwood
Academic Publishers.
World Bank 2010. Global Economic Prospects Summer 2010: Fiscal Headwinds and
Recovery. World Bank, Washington, p. 2. Available at http://siteresources.worldbank.
org/INTGEP2010/Resources/FullReport-GEPSummer2010.pdf [accessed July 5, 2010].
This page intentionally left blank
Part III

Measuring effectiveness
This page intentionally left blank
8 Multiple objective decision-making
Kent D. Wall and Cameron A. MacKenzie

8.1 Introduction
Imagine the the following decision problem for the fictitious country of Drmecia,
a small nation with many security problems and limited funds for addressing these
problems:

The Army of Drmecia operates a ground based air-search early warning radar
base located near the capital city, Sloat. The radar exhibits low availability
because of reliability and maintainability problems. This exposes the country
to a surprise air attack from its enemy, Madland. The Army obtained pro-
curement funds two years ago and installed another radar of the same type
as a back-up, or standby, radar. Unfortunately, the second radar unit also has
exhibited low availability and too often both radars are off-line. The com-
mander of Sloat Radar Base is now requesting funds for a third radar, but the
Army Chief of Staff is very concerned. He knows that the defense budget is
under increasing pressure and funds spent on another radar may have to come
from some other Army project. In addition, two new radar models are now
available for procurement at higher cost than the existing radars. It may be
better to purchase one of these than another radar of the existing model.
The Chief of Staff knows he must have a strong case if he is to go to the
Minister of Defense and ask for additional funds. He wants to know which
radar is best. He wants to know what he is getting for the extra money spent.
He wants to know what is the most cost-effective course of action.
The director of analysis for the Army of Drmecia knows that making a
cost-effective decision requires two things: (1) a cost analysis and (2) an
effectiveness analysis. The analysis of cost is a familiar problem and he has a
staff working on it. The effectiveness part of the problem, however, is a con-
cern. What is effectiveness in this situation? How can effectiveness be defined
and quantified? How can the cost analysis be integrated with effectiveness
analysis so that a cost-effective solution can be found?

The goal of this chapter is to develop a method to think about and quantify
effectiveness in the public sector, specifically defense. Effectiveness can best be
198 K. D. Wall, C. A. MacKenzie
measured in the public sector by developing a framework for solving decision
problems with multiple objectives. The framework will provide you with a practi-
cal tool for quantitative investigation of all factors that may influence a decision,
and you will be able to determine why one alternative is more effective than others.
This analytical ability is very important because many real-life decision problems
involve more than a single issue of concern. This holds true for personal-life deci-
sions, private sector business decisions, and public sector government resource
allocation decisions.
Examples of personal-life decision problems with multiple objectives are plenti-
ful: selecting a new automobile; choosing from among several employment offers;
or deciding between surgery or medication to correct a serious medical problem. In
the private sector, maximizing profit is often the sole objective for a business, but
other objectives may be considered, such as maximizing market share, maximizing
share price performance, and minimizing environmental damage.
Public sector decision-making almost always involves multiple objectives. State
and local government budget decisions are evaluated, at least, in terms of their
impacts on education programs, transportation infrastructure, public safety, and
social welfare. The situation is more complex at the Federal level because national
defense issues enter the picture. For example, consider the ways in which the
US Department of Defense (DoD) evaluates budget proposals. Top-level decision-
makers consider the effects of a budget proposal on (1) the existing force structure,
(2) the speed of modernization, (3) the state of readiness, and (4) the level of risk,
among other factors. Other national defense objectives include the ability to deter
aggression, project force around the world, and defeat an enemy in combat regardless
of its location or military capability. This multidimensional character permeates all
levels of decision-making within a planning, programming, budgeting and execution
system (PPBES). Acquisition decisions, training and doctrine policy changes, and
base reorganization and consolidation choices all have multiple objectives (Buede
and Bresnick 1992; Parnell et al. 1998; Ewing et al. 2006).
Government decisions in general, and defense resource allocation decisions
in particular, have an added evaluation challenge. Outcomes are difficult, if not
impossible, to represent in monetary terms. First, benefit cannot be expressed
in terms of profit. Unlike the private sector, the public sector is not profit moti-
vated and this single monetary measure of benefit is not relevant. Second, market
mechanisms often do not exist for “pricing out” the many benefits derived from
public sector decisions. Thus, it is not possible to convert all the benefits into
monetary terms and conduct a cost–benefit analysis (CBA). In national defense,
benefits are often characterized in terms such as deterrence, enhanced security,
and increased combat capability (Dyer et al. 1998; Parnell et al. 2001). No mar-
kets exist that generate a price per unit of deterrence or a unit increase in national
military security.
Solving these decision problems requires a structured systematic approach that
aids the discovery of all the relevant objectives and makes it easy to work with
objectives expressed in many different units of measure. It also must allow the
decision-maker to account for the relative importance of the objectives and the
Multiple objective decision-making 199
importance of marginal changes within individual objective functions. Following
such an approach can allow decision-makers to perform cost-effectiveness analy-
sis (CEA) to compare programs and alternatives.

8.2 Problem formulation


A systematic approach to decision-making requires developing a model that
reflects a decision-maker’s preferences and objectives. We assume the decision-
maker is rational and seeks the most attractive or desirable alternative. Goals
describe what the decision-maker is trying to achieve, and objectives determine
what should be done in order to achieve the decision-maker’s goal(s). A model
based on the decision-maker’s objectives will allow the decision-maker to evaluate
and compare alternatives.
For a decision problem with a single objective, the model for decision-making is
straightforward. For example, consider the simple problem of choosing the “best”
alternative where “best” is interpreted as the greatest range. Not surprisingly, the
decision-maker should choose the alternative with the greatest range.
Things are never this straightforward when there are multiple objectives
because the most desirable or most effective alternative is not obvious. For exam-
ple, consider the selection of the Sloat Radar as discussed in the introduction. The
decision-maker knows that range and interoperability are important. Maintenance
and reliability are also concerns. How do we define effectiveness in this situation?
Do we focus only on range? Do we focus only on interoperability? How can we
consider all four objectives at the same time?

8.2.1 Issues of concern and objectives


The issues of concern to the decision-maker can always be expressed as objec-
tives. For example, in selecting a radar system suppose two alternatives are equal
in all respects except for range. In this case, the decision-maker chooses the
radar with more range. This concern with range translates to an objective: the
decision-maker wants to maximize range. Likewise, suppose that two alternatives
are equal in all respects except for required maintenance, and the decision-maker
selects the alternative with less required maintenance. The concern with required
maintenance translates to an objective: the decision-maker will want to minimize
required maintenance.
The rational decision-maker can always be shown to act in a way consistent
with a suitably defined set of objectives. Hence, decision problems characterized
by many issues of concern are decision problems in which the decision-maker
attempts to pursue many objectives. In other words, the decision-maker confronts
a Multiple Objective Decision Problem (Keeney and Raiffa 1976; Keeney 1996).
The set of objectives is fundamental to the formulation of this type of decision
problem. If we have a complete set of objectives that derive from all the issues of
concern to the decision-maker, we have a well-formulated problem. Furthermore,
if these objectives are defined in sufficient detail, they tell us how to evaluate
the alternatives in terms that have meaning to the decision-maker. In the radar
200 K. D. Wall, C. A. MacKenzie
example, each alternative can be evaluated in terms of its range, its interoperability,
its required maintenance, etc. This knowledge is fundamental to the solution of
this type of decision problem. It tells us how to quantify things or how to measure
the attainment of the objectives. We seek to represent the alternatives by a set of
numbers that reveal to the decision-maker how well each alternative “measures
up” in terms of the objectives.
Let us introduce some mathematical notation to help with the formulation, and
Table 8.1 defines all the variables used in this chapter. Let there be M objectives
for the decision-maker. Let these be defined in enough detail so that we know
how to measure the attainment of each objective. Let these measures be denoted
x i where 1 ≤ i ≤ M. For example, x 1 = range, x 2 = interoperability, x 3 = required
maintenance, etc. Let there be N alternatives indexed by the letter j where 1 ≤
j ≤ N. Thus, each alternative can be represented by a set of M numbers:

{x 1 ( j ), x 2( j ), x 3 ( j ), . . . , x M ( j )}.

These numbers become the attributes for the j th alternative, and each of the M
numbers represents how well the j th alternative meets each one of the decision-
maker’s M objectives. Evaluating an alternative means determining the numerical
value of each of the measures. These serve as the raw data upon which the decision
is made.
The key to solving the decision problem is to explain how this set of M num-
bers is viewed in the mind of the decision-maker. This is done by constructing
a collective measure of value that reflects how much the decision-maker prefers
one collection of attributes vis-à-vis another collection of attributes. For example,
how does the decision-maker value the set of attributes for the first alternative,
{x i (1)}, compared to the set of attributes for the second alternative, {x i (2)}, where
1 ≤ i ≤ M? Is the the set {x i (1)} more “attractive” than the set {x i (2)}?
It is challenging for a decision-maker to compare alternatives if each alternative
has several attributes (i.e., if M is greater than 3 or 4). Developing a function
that combines the M attributes into a single number provides a method for the
decision-maker to compare alternatives based on the attributes that describe each
alternative. This single number is called the Measure of Effectiveness (MoE).

8.2.2 Effectiveness
The function that measures the effectiveness of the j th alternative is called a
value function, v( j ). This function incorporates the preferences of the decision-
maker and converts each collection of attributes as represented by the set {x i ( j )}
into a number that represents the attractiveness or desirability of the collection in
the mind of the decision-maker. It measures the extent to which the j th alterna-
tive helps the decision-maker pursue all the objectives while taking into account
the relative importance of each. Effectiveness is a number. It quantifies “how far
we go” towards achieving our goals as measured by the value of the objectives.
The objectives are not all of equal importance, however, and v( j ) also takes into
account the relative importance of each objective.
Multiple objective decision-making 201
Table 8.1 Variable definitions

Variable Definition

a Assessed parameter in the exponential value function


b1 Assessed parameter for squared term in the exponential
value function
b2 Assessed parameter for cubic term in the exponential value function
c( j ) Life-cycle cost of the j th alternative
cideal Dollar amount at which the value of cost equals 1
ctoo much Dollar amount at which cost is too high or value of cost equals 0
i A single objective or attribute
j A single alternative
k= c 1 Reciprocal of too high of cost
too much
wi Importance weight for the ith attribute
wA Importance weight for availability
wC Importance weight for complexity
wE Importance weight for electronic counter-countermeasures (ECCM)
wI Importance weight for interoperability
wL Importance weight for cognitive load
wP Importance weight for performance
wR Importance weight for range
wU Importance weight for ease-of-use
v( j ) Measure of effectiveness for the j th alternative
vc (c( j )) Value function for cost of the j th alternative
vi (xi ) Value function for the ith attribute
xi Measurement for the ith attribute
xi ( j ) Measurement for the ith attribute for the j th alternative
xideal Ideal measurement for an attribute
xmax Maximum measurement for an attribute
xmin Minimum measurement for an attribute
xtoo little Too little for an attribute for more-is-better case
xtoo much Too much for an attribute for less-is-better case
z Swing in attribute that contributes least to an objective
zi Difference between xi and either xmax or xmin
K Normalization constant for exponential value function
K1 Normalization constant for quadratic exponential value function
K2 Normalization constant for cubic exponential value function
M Number of objectives or attributes
N Number of alternatives
V ( j) Payoff function for cost-effectiveness of the j th alternative
V ∗ and V ∗∗ Payoff values of cost-effectiveness
WC Importance weight for cost
WE Importance weight for effectiveness

The decision-maker desires to maximize effectiveness by choosing the alter-


native with the highest v( j ) subject of course to cost considerations. The funda-
mental problem in formulating the decision problem is defining v( j ) based on
the set of attributes {x 1 ( j ), x 2 ( j ), x 3( j ), . . . , x M ( j )} and then integrating cost with
the MoE.
202 K. D. Wall, C. A. MacKenzie
Formulating a decision problem with multiple objectives requires four pieces of
information (Kirkwood 1997):

1 we need a list of the relevant objectives;


2 we need to know how to value the measures associated with each individual
objective;
3 we need to know the relative importance of these objectives;
4 we need to know the relative importance between the cost and effectiveness of
each alternative.

We proceed to address each of these needs in the rest of the chapter.

8.3 Discovering the relevant objectives


We must know what matters in a decision problem or the consequences that a
decision-maker considers when thinking about each solution alternative. These
are the issues of concern and are represented by objectives. We cannot judge
alternatives without knowing the objectives of the decision-maker.
Discovering all the relevant objectives is the first step in solving the decision
problem (Keeney 1996; Hammond et al. 1999; Clemen and Reilly 2001). It is
helpful to employ a graphical construction called an hierarchy or “tree structure.”
(Think of a tree: a trunk with a few main branches that have many more smaller
branches that have even more smaller branches—now turn that picture upside
down and you have a picture of an hierarchy.) Figure 8.1 depicts a hierarchy with
several levels. The hierarchy begins at the top most level with a single over-arching
objective that captures in its definition all that the decision-maker is trying to do.
For multiple objective problems in public policy, the overall objective is to maxi-
mize effectiveness. The objectives in the first level below the overall objective tell
us how we define the top level hierarchy. We say the overall objective is refined or
defined in more detail by the objectives listed on the next level down. The lower
level provides more detail as to what is meant by the objective at the next higher
level.

0 vera II 0 bjective

Ob jec ti e 1 .2.m Ob jec ti e 1 .2.m Ob jec ti e 1 .2.m

Ob jec ti e 1 .2.m Ob jec ti e 1 .2.m I Ob jec ti e 1 .2.m I

Ob"ectiveOb"ective Ob"ective Ob"ective

Ob"ective 1.2 .l.k.l Ob"ective I .2 . l. k.r

Figure 8.1 Generic hierarchy.


Multiple objective decision-making 203
The definition is not operational unless it is useful in measuring things. We must
have a way to develop the hierarchy in enough detail so that we get specific enough
in our definition of the objectives that it becomes obvious to us (and everyone else)
how “to measure things.” What we need is a method of construction that takes us
from the top level down to the lowest level where measurement is obvious. There
are several ways of doing this but we will only discuss the “top-down” method
and the “bottom-up” method.

8.3.1 The top-down approach


In the top-down approach you start with the obvious, to maximize effective-
ness. This becomes the top-level objective. Because this objective can mean many
things to many people and is too vague to be operational, the next step is to seek
more detail. You proceed by asking the question: “What do you mean by that?”
The answer to this question will allow you to write down a set of sub-objectives,
each of which derives from a more detailed interpretation by the decision-maker
of just what overall effectiveness means to him or her. For example, in the case
of Sloat Radar the decision-maker may say: “Maximizing availability is part of
maximizing overall effectiveness, I’ve got to have a radar that works almost all the
time.” He may also say that “I need a high performance radar that works almost
all the time, so maximizing radar performance is also part of maximizing overall
effectiveness.” Finally, he may say that he needs a radar that is easy to use. Thus
minimizing radar complexity for the user is also important.
The top-down approach continues in this way until there is no doubt what the
objectives mean because we will be able to measure their value for each alterna-
tive. For example, in the first level down from the top we know exactly what we
mean by acting so as “to maximize availability.” Availability has a well-known
precise definition: availability is the probability that a system will work at any
given time. It can be measured either directly or by computation using the system’s
Mean Time Between Failure (MTBF) and its Mean Time To Repair (MTTR). Each
alternative can be evaluated in terms of its availability, and maximizing availability
is pursued by seeking the system that exhibits the highest availability.
Performance, however, does not have a precise agreed-upon definition that
tells us how we can measure it. Here we must ask once again: “What do you
mean by that?” The answer to this question will allow us to understand what is
meant by maximizing performance. Suppose, for example, that the response to
this question is: “High performance means great effective radar range, resistance
to electronic countermeasures, and high interoperability.” We do not need to ask
any other questions here because we know how to measure (evaluate) range for
each alternative. We know how to evaluate resistance to electronic countermea-
sures: a simple “yes” or “no” answer will do. Either an alternative has electronic
counter-countermeasures (ECCM) capability or it does not. Finally, we know how
to measure interoperability: we can count the number of communication links that
can be operated by each alternative (so it can feed target information to the various
anti-air forces).
204 K. D. Wall, C. A. MacKenzie
Minimizing user complexity also requires refinement so we need to ask: “What
do you mean by that?” The answer in our example may be that to minimize user
complexity we need to minimize the cognitive load placed on the radar operator
(measured by how many things the operator has to watch, sense, react to, adjust,
and refine) and maximize the ease-of-use of the radar by the operator. Each alter-
native could be evaluated by cognitive psychologists and human-machine interface
industrial engineering specialists. In this case each evaluation could result in a sim-
ple rating scheme that uses “high,” “medium,” or “low” coding. Therefore both
objectives are specific enough to be measurable.
The end result of this top-down approach for the Sloat radar example is the
objectives hierarchy depicted in Figure 8.2. The key to the top-down approach is
repeated application of the question “What do you mean by that?” to provide more
and more detail. You stop refining the structure when the answer to the question
defines a quantity that can be measured, quantified or evaluated. When you have
reached this point for each part of the hierarchy then you have completed the
process and obtained what you desire—a complete description of what you mean
by effectiveness and a way to measure it.

8.3.2 The bottom-up approach


The bottom-up approach starts where the top-down approach ends: with a collec-
tion of very detailed and specific measures. These are structured, or grouped, into
a hierarchy in a way that assures we are not forgetting anything important and
we are not double-counting objectives. The key to this approach is the repeated
application of the question: “Why is that there?”
The construction of the list is accomplished in many ways. First, it can be con-
sidered a Christmas “wish list.” The decision-maker can be asked to list everything
he or she would like to have in an alternative. For example, with Sloat Radar
the decision-maker may respond with the following. “I’d like to have maximum
effective range and complete immunity to electronic ‘jamming.’ I’d also like to
have it very easy to use so my least technically adept soldier could operate it.”
Second, specific measures can be found in the “symptoms” listed in the original
problem statement. If the Army is upset with the existing radar availability then

To Maximize Effectiveness

To Max. Availability
I
To Max. P e rformanc e ToM in . Complexity

To Max.
Inter-
I
To Max To Max.
Effective
~
To Min To Max.
ECCM Cognitive Ease-of-Use
Operability Range Load

Figure 8.2 Sloat Radar objectives hierarchy.


Multiple objective decision-making 205
it is obvious that the decision-maker would also be interested in improving or
maximizing availability.
For each item listed the decision-maker is asked: “Why is that there?” The
response to this question will provide information that aids in grouping the mea-
sures. This makes it easy to define higher level objectives for each group. For
example, electronic countermeasure resistance and range are included so that
“when the radar works, it has enough capability to do its job better than any other
alternative.” This response may bring to mind performance issues, and we group
these two measures under an objective that seeks to maximize performance. Once
performance is included as a higher objective it may provoke consideration of
other ways one interprets performance, and this may make the decision-maker
think of high interoperability. The same process applied to the ease-of-use mea-
sure would lead to considering cognitive loading on the operator and a general
concern with user complexity.
Finally, this process aids the decision-maker in clarifying the higher level objec-
tives. A structure emerges that helps insure nothing is forgotten and nothing is
double-counted. For example, once the higher objectives of maximizing perfor-
mance and minimizing complexity are evoked, the decision-maker will be able
to see what is important: (1) the radar has got to work almost all the time, i.e.,
maximize availability; (2) when it works it must be the best, i.e., maximize perfor-
mance; and (3) it should be easy to operate or else all the other stuff is not worth
anything, i.e., minimize complexity. The result is a hierarchy of objectives as in
Figure 8.2.

8.3.3 Which approach to use


Each approach has its advantages and disadvantages. The top-down approach
enforces a structure from the very beginning of the exercise. It is, however, often
difficult to think in general terms initially. Humans find it easier to focus on
specifics, like range and availability. The bottom-up approach is more attractive in
this respect, but this will not produce a logical structure without additional work.
Producing an extensive wish list will, most likely, produce redundant measures.
Oddly enough, the longer the list, the more likely something important will be
forgotten. Long lists are harder to critically examine and locate omissions. This is
where structure helps.
The best approach is perhaps a combination of the two. First, construct a hier-
archy with the top-down approach, and then “reverse” direction—construct a
hierarchy using the bottom-up approach. Critical examination of the results pro-
vides information for a more complete final hierarchy. The result should be a
hierarchy that is (Keeney 1996):

1 mutually exclusive (where each objective appears once);


2 collectively exhaustive (where all important objectives are included);
3 able to lead to measures that are operational (can actually be used).
206 K. D. Wall, C. A. MacKenzie
The bottom level of the hierarchy composes the M objectives or attributes
necessary to build the function for effectiveness. The Sloat Radar hierarchy has
M = 6 attributes (availability, interoperability, ECCM, range, cognitive load, and
ease-of-use), which will be used to compare alternatives.

8.3.4 The types of effectiveness measures


The effectiveness measures that result from development of the objectives hierar-
chy can be of three general types: (1) natural measures; (2) constructed measures;
and (3) proxy measures (Kirkwood 1997).

8.3.4.1 Natural measures


Natural measures are those that can be easily counted or physically measured.
They use scales that are most often in common use. Radar range, interoperability
(as measured by the number of communication links provided), and availability (as
measured by the probability that the system will be functioning at any given time)
are examples. Weight, payload, maximum speed, size, volume, and area are other
examples. Whenever possible, we should try to refine our objectives definitions in
the hierarchy to obtain this type of measure.

8.3.4.2 Constructed measures


This type of measure attempts to measure the degree to which an objective is
attained. It often results when no commonly used physical or natural measure
exists. For example, cognitive load is a constructed measure. It is assessed as
high, medium, or low depending on the number of visual and audible signals to
which the operator must attend. Suppose indicators of operator cognitive abilities
as measured by aptitude scores, training rigor, and education level indicate sig-
nificant error in operator function if the combination of audio and visual cues
is greater than 5. A radar with 5 or more cues is then assessed as having a
high cognitive load. Similarly, a combination of cues between 3 and 4 consti-
tutes a medium cognitive load, and a combination of cues less than 3 constitutes
a low cognitive load. Constructed measures are the next best thing to natural
measures.

8.3.4.3 Proxy measures


These measures are like natural measures in that they usually are countable or physi-
cally measurable. The difference is that they do not directly measure the objective of
concern. In one sense we could consider radar range to be a proxy for measuring how
much an alternative helps to maximize warning time to react better to an airborne
threat. The actual reaction time cannot be directly measured because we would need
to know the exact attack speed of the threatening bombers. We do know, however,
that the farther out we can detect an attack, the more reaction time we would have.
So maximizing range is a proxy measure for maximizing time to react.
Multiple objective decision-making 207
8.4 Decision-maker preferences
After building the objectives hierarchy, we know what is important and have
created a list of the relevant objectives. Each alternative can be evaluated using
the list of individual effectiveness measures for the different attributes, but these
measures involve a variety of incommensurable units. We must develop a way to
convert all these disparate measures to a common unit of measure. This common
unit of measure represents value in the mind of the decision-maker.
Calculating the MoE for the j th alternative as represented by the function v( j )
requires two types of preference information: (1) information that expresses pref-
erence for more or less of a single attribute and (2) information that expresses
relative importance between attributes. The first represents preference informa-
tion within a single attribute, while the second represents preference information
across different attributes (Kirkwood 1997).
The first is important because of the way a decision-maker values marginal
changes in a single effectiveness measure. For example, in the Sloat Radar case, a
decision-maker may value the increment in range from 200 km to 400 km more than
twice as much as the same increment from 400 km to 600 km. The same increment in
range (of 200 km) from 600 km to 800 km may be valued very little. This valuation
is quite reasonable. More range is preferred to less range, but there is a point beyond
which additional range has little additional value to the decision-maker. We already
will have “enough” range for the purposes of early warning against air attack. Such
information is very important for another reason. It allows us to construct an indi-
vidual attribute value function that provides a scaling function to convert the natural
units of measurement into units of value on a scale of 0–1. This has the advantage
of removing any problem caused by the units of measurement for the individual
attribute. For example, measuring radar range in terms of meters, kilometers, or
thousands of kilometers can influence the numerical analysis and ultimately the
answer. We need a process that is independent of the units of measurement. This
is one of the important things we achieve with an individual attribute value func-
tion. Such a function also makes the process independent of the range of values of
the individual effectiveness measures. This preference information will provide the
answer to the question: “How much is enough?”
The second type of information is important because a decision-maker val-
ues some individual effectiveness measures of attributes more than others. For
example, in the Sloat Radar problem, the decision-maker may feel it much more
important to increase availability than to increase range. Such information implies
that a decision-maker would be willing to obtain more availability at the expense
of less range. This preference information will provide the answer to the question:
“How important is it?”

8.4.1 How much is enough (of an individual attribute)?


Once we know all the relevant effectiveness measures, we need to describe how
a decision-maker values marginal changes in each measure. This is done by
constructing a function that converts the nominal measurement scale into a 0–1
208 K. D. Wall, C. A. MacKenzie
value scale where 0 is least preferred and 1 is most preferred. There are several ways
of doing this but only one method is presented here. It is a three-step procedure.
First, divide the nominal scale into “chunks,” or intervals, over which the
decision-maker can express preference information. For example, consider radar
range. The decision-maker may find it difficult to express preferences for each
additional kilometer of range but may find it easier to do so if we consider range
in 100 km chunks.
Second, ask the decision-maker to tell you, using a scale of 0-10, how valuable
is the first 100 km of radar range. Then ask how valuable is the next 100 km of
range (the second increment from 100 km to 200 km). Repeat this process for each
additional 100 km of range until you reach the upper limit of interest. Suppose
this exercise gives the following sequence when applied to range measurements
between 0 km and 1000 km: 10, 10, 8, 7, 5, 3, 1, 0.5, 0.1, 0.05. These are the
increases in marginal effectiveness to the decision-maker.
Third, use this marginal information to create a cumulative value function
scaled to give value scores between 0 and 1. The radar range example gives a total
cumulative value (the sum of the marginal increments) equal to 10 + 10 + 8 +
7 + 5 + 3 + 1 + 0.5 + 0.1 + 0.05 = 44.65. Divide all marginal increments by this
sum: 10/44.65, 10/44.65, 8/44.65, 7/44.65, etc. We obtain a sequence of smaller
numbers that add to one. The cumulative value, calculated by adding the marginal
values, is what we use to determine the value of a measure. For example, the value
of 300 km range is equal to the value derived from increasing the range from 0 to
100 km plus the additional value derived from increasing the range from 100 km
to 200 km plus the additional value derived from increasing the range from 200
km to 300 km. The cumulative value of these normalized marginal increments is
presented graphically in Figure 8.3.

0.9

0.8

., 0.7
::I

~., 0.6

> 0.5
~
:i
E 0.4
::I
(J
0.3

0.2

0.1

0
0 100 200 300 400 500 600 700 800 900 1000
Range(km)

Figure 8.3 Radar range value function.


Multiple objective decision-making 209
This process achieves two important things. First, if we repeat it for each
attribute, we have a common unit of measure standardized to a 0–1 interval. Sec-
ond, and most importantly, we now have a way of valuing radar range. A range of
700 km has a value of 0.985 and a range of 800 km has a value of approximately
0.997, which informs us that the decision-maker prefers 800 km to 700 km but that
this increase is not highly valued (only an increase in value of 0.012). On the other
hand, increasing the range from 200 km (with a value of 0.448) to 300 km (with a
value of 0.628) is much more valuable to the decision maker. The increase in value
here is 0.18. Thus we know the decision-maker prefers the 100 km increment from
200 to 300 km more than the going from 700 to 800 km.
The analysis of effectiveness is facilitated by converting this graphical informa-
tion to algebraic form so that we can obtain a model of preferences that allows us
to quantify these preferences for every possible numerical measure for an attribute.
We construct a value function for an individual attribute where x i represents the
numerical value for the i th attribute and vi (x i ) is the value function for the i th
attribute (Kirkwood 1997). A useful algebraic form for capturing the important
characteristics of Figure 8.3 is the exponential function:

1 − e−a[xi −xmin ]
vi (x i ) = (8.1)
K

where K = 1 − e−a[xmax −xmin ] is a normalization constant. This function requires


three parameters: x min , x max , and a. The first two are known already because the
decision-maker has specified the range of interest over which x i will vary: [x min ≤
x i ≤ x max ]. The parameter a can be obtained using least-squares regression using
the Solver “add-in” tool in Excel. The parameter can also be estimated graphically
by adjusting a until the curve generated by the function vi (x i ) corresponds to the
points representing the decision-maker’s preferences in the cumulative value func-
tion as ascertained by the three-step method. More flexible but more complicated
forms of this exponential function can be used (see Appendix 8.1). The purpose
of these algebraic forms is to match the original cumulative value function given
to us by the decision-maker as closely as possible.
Calculating a numerical value for a in vi (x i ) that corresponds to the infor-
mation for the range of the radar, as depicted in Figure 8.3, is done in Excel
using the Solver add-in tool. Figure 8.4 presents vi (x i ) corresponding to the data
in Figure 8.3 when we use an exponential form with a = 0.0033, x min = 0, and
x max = 1,000.
The preceding example illustrates a case where “more is better.” Exactly the
same approach is used when “less is better.” For example, consider a man-portable
field radio. Here less weight is preferred to more weight. Now the decision-maker
is asked a series of questions relating to marginal value but we “start from the
other end” of the measurement scale—we start with the heaviest weight and work
backwards to the lightest weight. For example, suppose it is decided that field
radios weighing more than 20 kg are useless because they weigh too much for
a typical soldier to carry. Imagine we use marginal changes in weight of 2 kg.
210 K. D. Wall, C. A. MacKenzie

,............, ..----
~
!-A"""
v
0.9

.,
0.8

0.7
v-
::I
0.6 ~
~.,
> 0.5 /
~
:;
E 0.4
r/-
::I
(.)
0.3 /
0.2 ~~-
0.1 I
0 J
0 100 200 300 400 500 600 700 800 900 1000
Range(km)

Figure 8.4 Fitted exponential value function for radar range.

We begin by asking the decision-maker to give us a number (on a scale of 0–10)


expressing the value of a 2 kg reduction (from a weight of 20 kg to a weight of 18
kg). Then we ask what value is associated with a further 2 kg reduction in weight
(to 16 kg). We repeat the process until we get an answer for the value attached to
the last 2 kg of weight. Suppose we obtain the sequence: 10, 9, 8, 5, 3, 1, 0.4, 0.1,
0.0. These sum to 36.5 so we normalize the decrements by this total. The result is
presented in Figure 8.5.
As in the case where more is better, the graph of cumulative value can be used
to fit an exponential function for conducting analysis of effectiveness. Once again,
exponential functions are very useful and can be used to represent this information:

1 − e−a[xmax −xi ]
vi (x i ) = (8.2)
K

where K is the same as before. Figure 8.6 depicts the results of fitting the cumu-
lative value function for radio weight in Excel using the Solver add-in tool where
a = 0.21, x min = 0, and x max = 20. Once again, closer fit can be obtained by using
more terms in the exponent as described in Appendix 8.1.
A linear function can also be used to approximate the exponential function for
either the more-is-better case or the less-is-better case. The linear function can use
the same cumulative value function as described previously, or just two values can
be assessed from the decision-maker. The two values necessary are a number cor-
responding to “not enough performance” and a number corresponding to “good
enough performance.” If more is better, the number corresponding to not enough
performance is “too little” or x too little . If less is better, the number corresponding
Multiple objective decision-making 211

I I
I I I I I
0.9

I I I I I I
0.8

.,
I I I I I I I
0.7
::I

~.,
I I I I I I I
0.6

I I I I I I I I
> 0.5
i
'S

I I I I I I I I
E 0.4
::I
(J

I I I I I I I I
0.3

I I I I I I I I I
0.2

I I I I I I I I I
0.1

0
0 2 6 8 10 12 14 16 18 20
Weight(kg)

Figure 8.5 Radio weight value function.

...........,
0.9 N KJ
0.8 r-.....
., 0.7 ~
"
::I

~., 0.6

> ~

"n'
0.5
i
'S
E 0.4
::I
(J
0.3

0.2

0.1
\
0 \
0 2 6 8 10 12 14 16 18 20
Weight(kg)

Figure 8.6 Fitted exponential value function for radio weight.

to not enough performance is “too much” or x too much . In both cases, the number
corresponding to good enough performance is “ideal” or x ideal . For example, in
the more-is-better example of radar range, 0 km can represent not enough per-
formance or too little and 600 km can represent good enough performance or the
ideal. We choose 600 km rather than 1,000 km as the ideal because the assessed
value at 600 km is 0.963 and the increase in value from 600 to 1,000 km is only
0.037. In the less-is-better example of the radio’s weight, we choose 20 kg as too
212 K. D. Wall, C. A. MacKenzie
much and 8 kg as ideal. Eight kilograms is chosen as ideal because the value at
8 kg is 0.986 and the increase in value from 8 to 0 kg is only 0.014. The linear
value function is defined for more is better in which [x too little ≤ x i ≤ x ideal ]:

x i − x too little
vi (x i ) = , (8.3a)
x ideal − x too little

and for less is better in which [x ideal ≤ x i ≤ x too much ]:

x i − x too much
vi (x i ) = . (8.3b)
x ideal − x too much

If x i ≥ x ideal in the more-is-better case or if x i ≤ x ideal in the less-is-better case,


vi (x i ) = 1. Similarly, vi (x i ) = 0 if x i ≤ x too little in the more-is-better case or x i ≥
x too much in the less-is-better case.

8.4.2 How important is it (relative to the other attributes)?


After assessing the preferences of the decision-maker for changes in each attribute,
we need to know the preferences of the decision-maker among different attributes.
The model for the MoE uses importance weights or tradeoff weights denoted by
wi , the importance weight for the i th attribute (Kirkwood 1997).
 M Each importance
weight must satisfy 0 ≤ wi ≤ 1 and together must satisfy i=1 wi = 1. A larger
value of wi signifies that the effectiveness measure for the i th attribute is more
important.
The conditions imposed on the weights will automatically be satisfied if we
apply the conditions to each of the partial hierarchies that appear within the
overall objectives hierarchy. For example, the overall hierarchy in Figure 8.2
is actually composed of three hierarchies. In the first hierarchy, defined by the
top level and the first level underneath it, maximizing overall effectiveness is
supported by maximizing availability, maximizing performance, and minimiz-
ing user complexity. The relative importance of each of these three objectives
to maximizing overall effectiveness can be expressed with three weights such that
0 ≤ w A , w P , wC ≤ 1 and w A + w P + wC = 1 ( A = availability, P = performance and
C = complexity). In the second hierarchy, maximizing performance is supported
by maximizing interoperability, maximizing ECCM capability, and maximizing
range. The relative importance of each of these can be expressed with three more
weights such that 0 ≤ w I , w E , w R ≤ 1 and w I + w E + w R = 1 (I = interoperability,
E = ECCM, and R = range). Finally, minimizing complexity is supported by min-
imizing cognitive load and maximizing ease-of-use. The relative importance of
these two objectives can be expressed with two weights such that 0 ≤ w L , wU ≤ 1
and w L + wU = 1 (L = cognitive load and U = ease-of-use).
Assigning weights in this fashion guarantees that the weights for all M objec-
tives or attributes sum to one and are all within the range of 0–1. The Sloat
Radar example has M = 6 attributes, and we need six importance weights,
w1 , w2 , w3 , w4 , w5 , and w6 to correspond to each one of the attributes (availability,
Multiple objective decision-making 213
interoperability, ECCM, range, cognitive load, and ease-of-use). These six “global”
weights can be derived from the “local” weights described in the previous paragraph.
Availability stands alone in the hierarchy, and w A = w1 where w1 is the global weight
for availability. Performance is composed of threeattributes, andw P = w2 + w3 + w4
where w2 is the global weight for interoperability, w3 is the global weight for ECCM,
and w4 is the global weight for range. Complexity is composed of two attributes,
and wC = w5 + w6 where w5 and w6 are the global weights for cognitive load and
ease-of-use, respectively. Using the fact that w I + w E + w R = 1 for the performance
sub-hierarchy and w L + wU = 1 for the complexity sub-hierarchy, we can express
the global weights for the six attributes:

w1 = w A (8.4a)
w2 = w P · w I (8.4b)
w3 = w P · w E (8.4c)
w4 = w P · w R (8.4d)
w5 = wC · w L (8.4e)
w6 = wC · wU . (8.4f)

Since each group of local weights sum to one, we know the global weights as
calculated above will also sum to one:

w1 + w2 + w3 + w4 + w5 + w6 = w A + w P + wC = 1.

Obtaining the weights requires solicitation of decision-maker preferences using


a different set of questions from that used above. There are many ways to do this,
and we consider five methods.

8.4.2.1 Direct assessment


The most obvious way to gain the information required is to directly ask the
decision-maker
M for M numbers, wi , such that 0 ≤ wi ≤ 1 for {1 ≤ i ≤ M} and
i=1 w i = 1. If there are no more than three or four effectiveness measures this is a
feasible approach. Unfortunately many real-life problems have many more effec-
tiveness measures and direct assessment becomes too difficult. The Sloat Radar
example has six effectiveness measures, and this presents a formidable problem to
the decision-maker. All is not lost, however, if we utilize the structure given us in
the objectives hierarchy.
Consider the Sloat Radar hierarchy. We can ask the decision-maker to consider
first just the relative importance of the three components that make up overall
effectiveness: availability, performance, and user complexity. Directly assessing
three importance weights at this level presents no insurmountable problem for the
decision-maker. Next we can ask the decision-maker to consider the components
of performance and tell us the relative importance of interoperability, ECCM capa-
bility, and range. Finally, we can ask the same question for the components of user
214 K. D. Wall, C. A. MacKenzie
complexity that requires only two importance weights be assigned. The results of
this series of questions allows us to compute the complete set of six weights using
Equations (8.4a)–(8.4f).
Using the hierarchy and this “divide and conquer” approach makes direct
assessment feasible in many situations where otherwise it would appear over-
whelming.

8.4.2.2 Equal importance


Direct assessment may result in the decision-maker stating that “all individual
measures are equally important.” In this case we must ask the decision-maker if
this statement applies to all M attributes or to each part of the hierarchy. Different
importance weights result depending on the interpretation of the statement.
If the decision-maker is referring to all M attributes, then clearly wi = 1/M. In
the Sloat Radar example this means that each measure would be given a weight
of 1/6. It is important to realize that this weighting has implications for the cor-
responding weights in the hierarchy. As discussed earlier, the importance weight
for performance equals the sum of global weights for interoperability, ECCM, and
range, and w P = 1/6 + 1/6 + 1/6 = 1/2. The weight for complexity equals the
sum of the weights for cognitive load and ease-of-use, and wC = 1/6 + 1/6 = 1/3.
Clearly, equal importance across the bottom level measures does NOT imply
equal weights throughout the hierarchy. Thus, it is very important to go back to the
decision-maker and ask: “Do you really believe performance is three times more
important than availability?” “Do you really believe that minimizing complexity
is twice as important as maximizing availability?” When phrased this way the
decision-maker may be led to reassess the importance weights.
If the decision-maker is referring to the weights at each level in the objectives
hierarchy, then we get a different set of weights for wi . Figure 8.2 tells us that on the
first level down availability, performance and complexity are all equally important:

w A = w P = wC = 1/3.

This also means that all the component objectives for performance are equally
important:

w I = w E = w R = 1/3.

Finally, suppose the decision-maker says that maximizing ease-of-use and mini-
mizing cognitive load are equally important, so then we know

wU = w L = 1/2.

These local weights translate into the following global weights:

w1 = 1/3, w2 = 1/3 · 1/3 = 1/9, w3 = 1/3 · 1/3 = 1/9,


w4 = 1/3 · 1/3 = 1/9, w5 = 1/3 · 1/2 = 1/6, w6 = 1/3 · 1/2 = 1/6.
Multiple objective decision-making 215
A very different picture emerges. Availability is three times more important than
interoperability, ECCM capability, and range. Availability is twice as important as
cognitive load and ease-of-use.
The structure of the hierarchy is a powerful source of information that we will
continually find helpful and informative.

8.4.2.3 Rank sum and rank reciprocal


Sometimes the decision-maker will provide only rank order information. For
example, in assessing the weights making up the performance measure, the
decision-maker may tell us: “Range is most important, interoperability is second
most important, and ECCM is least important.” In such cases we can use either
the rank sum or rank reciprocal method. Range has rank one, interoperability has
rank two, and ECCM has rank three (Stillwell et al. 1981).
In the rank sum method, range has an un-normalized weight of 3, interoper-
ability 2 and ECCM 1, based on the decision-maker’s ranking. Dividing each
of these weights by the sum 3 + 2 + 1 = 6 returns the normalized weights:
w R = 3/6, w I = 2/6, and w E = 1/6.
In the rank reciprocal method, the reciprocals of the original ranks give

Range = 1/1,
Interoperability = 1/2,
ECCM = 1/3,

which sum to 11/6. Dividing the reciprocal ranks by their sum gives three numbers
that satisfy the summation condition and represent a valid set of weights: w R =
6/11, w I = 3/11, and w E = 2/11.
Both the rank sum and rank reciprocal methods only require rank order infor-
mation from the decision-maker. The rank sum method returns weights that are
less dispersed than the rank reciprocal method, and less importance is placed on
the first objective in the rank sum method.

8.4.2.4 Pairwise comparison


Decision-makers find it easier to express preferences between objectives where
there are only two objectives. This has led to a procedure for soliciting pref-
erence information based on pairwise comparisons. This allows the importance
weights over M measures to be established by comparing measures only two at
a time. This is a special case of the swing weighting method, which is described
below.

8.4.2.5 Swing weighting


One method helpful with direct assessment that incorporates more than rank order
information is swing weighting (von Winterfeldt and Edwards 1986; Clemen
and Reilly 2001). Although more complex than the other methods, the swing
216 K. D. Wall, C. A. MacKenzie
weighting method can provide a more accurate depiction of the true importance
the decision-maker places on each objective. Swing weighting is also sensitive to
the range of values that an attribute takes, and different values for attributes can
result in different weights. The method has four steps. First, the decision-maker
is asked to consider the increments in overall effectiveness that would be repre-
sented by shifting, or “swinging,” each individual effectiveness measure from its
least preferred value to its most preferred value. Second, the decision-maker is
asked to order these overall increments from least important to most important.
Third, the decision-maker quantitatively scales these increments on a scale from
0 to 100. Finally the sum-to-one condition is used to find the value of the least
preferred increment.
Let us illustrate this process using the performance part of the Sloat Radar hier-
archy. We need to find three importance weights: w I , w E , and w R . First, suppose
that the least preferred under performance are 0 km for range, 0 communica-
tion links for interoperability, and no ECCM. This “least-preferred alternative”
with 0 km in range, no communication links, and no ECCM has a value of 0
for performance. The most preferred under performance are 1,000 km for range,
5 communication links, and with ECCM. The decision-maker believes that a
swing in range from 0 to 1,000 km will contribute most to performance, a swing
from no communication links to five links will contribute second most to per-
formance, and a swing from no ECCM capability to having ECCM capability
will contribute the least to performance. Because range is the most preferred
swing, we arbitrarily assign a value of 100 to a hypothetical alternative with
1,000 km range, 0 communication links, and no ECCM. We ask the decision-
maker for the value of swinging from 0 to 5 communication links relative to
the alternative with a value of 100 and the least-preferred alternative with a
value of 0. The decision-maker may say that a hypothetical alternative with 0
km range, 5 communication links, and no ECCM has a value of 67. Finally, the
decision-maker may conclude that a hypothetical alternative with 0 km range,
0 communication links, and having ECCM capability has a value of 33 relative
to the other alternatives. Thus, we know the un-normalized weights are 100 for
range, 67 for interoperability, and 33 for ECCM. Dividing each weight by the
sum of the un-normalized weights gives the importance weight for each attribute:
w R = 100/(100 + 67 + 33) = 1/2, w I = 67/(100 + 67 + 33) = 1/3, and w E = 33/
(100 + 67 + 33) = 1/6.
These weights express something very important in effectiveness analysis.
When x i is at its least preferred value, vi (x i ) = 0. When it is at its most preferred
value vi (x i ) = 1. The weights, {wi , 1 ≤ i ≤ M}, capture the relative importance
of these changes. The ratio wi /wi expresses the relative importance between
the changes from worst to best in the measures x i and x i for the i th and i th
attributes.
Swing weighting incorporates three types of preference information: ordinal
(rank), relative importance, and range of variation of the individual effectiveness
measures. Pairwise comparison uses rank and relative importance. The rank sum
and rank reciprocal methods only use rank information.
Multiple objective decision-making 217
8.5 Effectiveness analysis
The quantification of decision-maker preferences completes our model of effec-
tiveness. We can begin assessing or analyzing the overall effectiveness of each
alternative. All the ingredients are present: The objectives hierarchy tells us what
is important and defines the individual measures of effectiveness. The individual
value functions tell us how the decision-maker values marginal increments in these
measures and scales them to a 0–1 interval. Finally, the importance weights tell
us the relative importance of the individual attributes. The weights are combined
with the values for a single alternative to calculate the overall effectiveness of the
j th alternative:

v( j ) = w1 · v1 (x 1 ( j )) + w2 · v2 (x 2 ( j ))

M
+ w3 · v3 (x 3 ( j )) + · · · = wi · vi (x i ( j )). (8.5)
i=1

The MoE is calculated as the weighted sum of all the individual value functions.
Nothing of relevance is left out, and everything that enters the computation does
so according to the preferences of the decision-maker (Keeney and Raiffa 1976;
Kirkwood 1997).
Evaluating the overall effectiveness of an alternative requires computing its
v( j ). The result allows us to order all the alternatives from best to worst according
to their MoE. The “best” alternative is now the alternative with the largest MoE.
While this provides us a way to find “the” answer, we have a far more powerful
tool at hand.
We can use Equation (8.5) to investigate why we get the answers we do. For
example, what makes the best alternative so desirable? What individual effective-
ness measures contribute most to its desirability? An alternative may be the best
because of very high effectiveness in only one single measure. This may tell us
that we could be “putting all our eggs in one basket.” If this alternative is still
under development, uncertainties about its development may make this alternative
risky, and identifying the second best alternative may be important.
We can also assess the sensitivity of the answer to the importance weights. We
can find out how much the weights must change to give us a different answer.
For example, would a change of only 1 percent in one of the higher level weights
change the ordering of the alternatives? How about a change of 10 percent? This is
of practical significance because the decision-maker assigns weights subjectively,
and this always involves a lack of precision.
Most important of all is the ability to assess the effect of uncertainty in the
future condition. The decision-maker’s preferences are a function of the future
condition. If the decision-maker believes the future condition will change, then the
preferences, value function and weights may change. The effects of uncertainties
in the problem formulation can be readily evaluated once we have our model.
218 K. D. Wall, C. A. MacKenzie
We illustrate each of these situations using the Sloat Radar example. Suppose
there are four alternatives: (1) “do nothing” (keep the two existing radars at Sloat
which is labeled as Sloat 2); (2) purchase a third radar for Sloat of the same type
(which is labeled as Sloat 3); (3) purchase the new SkyRay radar; and (4) pur-
chase the new Sweeper radar. The latter two alternatives are new, with better range,
availability, interoperability and ECCM. These come at the cost of higher cogni-
tive load and less ease-of-use. The costs of procurement for the two new radars
are also higher than purchasing an existing radar. The data on which each of these
four alternatives can be assessed are depicted in Table 8.2.

8.5.1 The components of success


The decision-maker preferences for the importance weights are w A = 0.60, w P =
0.35, wC = 0.05, w I = 0.50, w E = 0.0, w R = 0.50, w L = 0.50, and wU = 0.50.
Individual value functions for availability and range are specified by value func-
tions similar in shape and form to that portrayed in Figure 8.4. The individual
value function for interoperability is similar but specified for integer values,
0–5. The individual value function for ECCM capability is binary, taking the
value 0 for no ECCM capability and the value 1 for having ECCM capability.
Cognitive load and ease-of-use are evaluated using a constructed scale of three cat-
egories: low, medium, and high. For cognitive load, v L (low) = 1, v L (medium) =
0.5, and v L (high) = 0. For ease-of-use vU (low) = 0, vU (medium) = 0.5, and
v H (high) = 1. Combining these evaluations and preferences gives the values
depicted in Table 8.3.

Table 8.2 Evaluation data for Sloat Radar

Alternative Availability Interoperability ECCM Range Cognitive Ease-of-use


(km) load

Sloat 2 0.79 2 data links No 250 Low Medium


Sloat 3 0.87 2 data links No 250 Low Medium
SkyRay 0.98 4 data links No 700 Medium Low
Sweeper 0.97 2 data links Yes 500 High Low

Table 8.3 Values and weights for Sloat Radar

Alternative Availability Interoperability ECCM Range Cognitive Ease-of-use


(km) load

Sloat 2 0.962 0.375 0 0.413 1.0 0.5


Sloat 3 0.982 0.375 0 0.413 1.0 0.5
SkyRay 0.998 0.938 0 0.909 0.5 0
Sweeper 0.997 0.375 1 0.774 0 0
Global 0.60 0.175 0 0.175 0.025 0.025
weights
Multiple objective decision-making 219

1.000

0.900

0.800

0.700

0.600

=> 0.500

0.400

0.300

0.200

0.100

0.000
Sloat 2 Sloat 3 SkyRay Sweeper

l2l Range a lnteroperability o ECCM o Availability • Complexity

Figure 8.7 Components of overall effectiveness for Sloat Radar.

Multiplying each of the global weights by the each of the values and adding
them together calculates an MoE for each radar, as pictured in Figure 8.7. SkyRay
is the most effective alternative and we know why. It has the highest value for
interoperability and range. Sweeper is second best because it has the second high-
est value for range. All alternatives possess approximately the same availability
rating. Interoperability and range are the attributes that are most important in
discriminating among these alternatives.

8.5.2 Sensitivity to preferences


The original importance weights produce an ordering of alternatives in which
SkyRay is most effective, Sweeper next most effective, followed by Sloat 3 and
then Sloat 2. Is this ordering robust to reasonable changes in the weights? If
not, which weights are most influential? Questions like these are answered using
the model to recompute the MoE of each alternative with different importance
weights.
First let us fix w P = 0.35 and vary both w A and wC over a reasonable range
of values subject to the condition that w A + w P + wC = 1. The result is shown
in Figure 8.8. SkyRay remains the most effective alternative for the range of
weights examined, which means that the most effective alternative is fairly robust
to changes in the importance weights for availability relative to complexity.
The ordering of the other alternatives changes if w A ≤ 0.55 or, correspondingly,
when wC ≥ 0.10. As availability becomes less important relative to complexity,
Sweeper becomes the least effective radar, and Sloat 3 becomes the second most
effective alternative. When determining the second most effective alternative, the
220 K. D. Wall, C. A. MacKenzie
decision-maker needs to ask himself: “How confident am I that w A > 0.55 when
w P = 0.35?”
Attention now focuses on the influence of w P . We repeat the above analysis for
w A and wC for different values of w P . If w P = 0.25, SkyRay remains the most
effective alternative as long as w A ≥ 0.5 and wC ≤ 0.25. As depicted in Figure 8.9,
if w P = 0.2, Sloat 3 becomes the most effective if w A ≤ 0.55 or wC ≥ 0.25. The
question to ask is: “How likely is it that w A ≤ 0.55, w P ≤ 0.2, and wC ≥ 0.25?” If
the decision-maker responds that he will not put that much importance on com-
plexity relative to availability and performance, then SkyRay remains the most
effective alternative.

1.0
b.
b.
-~ ~
0.9 ·u

b.

...---:::::::: -~~-
b.
0.8
-~ ~

e>
0.7
... :::----
~
0.6

I -+-Sloat 2 -+- Sioat3 t:::. SkyRay -o-Sweeper I


0.5
0.45 0.50 0.55 0.60 0.65

Figure 8.8 Sensitivity to w A and wC with w P = 0.35.

1.0

I -+-Sloat 2 -+- Sioat3 t:::. SkyRay -o- Sweeper I


0.9
b.

~ :t
0.8
~
-v·· ~

~
b.
e>
b.
0.7

~
0.6
o-----
0.5
0.45 0.50 0.55 0.60 0.65

Figure 8.9 Sensitivity to w A and wC with w P = 0.2.


Multiple objective decision-making 221
8.5.3 Uncertainty in the future conditions
A new future condition or planning scenario may change the decision-maker’s
preferences and, consequently, the ordering of the alternatives. Investigating this
type of sensitivity is very important when there is uncertainty in the future con-
dition. Let us assume that the situation described in Section 8.5.1 and depicted in
Figure 8.7 corresponds to a particular planning scenario we will call the baseline.
Now suppose that new intelligence estimates suggest the potential enemy
will re-equip its bomber fleet with supersonic attack aircraft. This new scenario
may change decision-maker preferences to place more importance on perfor-
mance because of the enemy’s increase in capability. For example, the importance
weights defining the measure of overall effectiveness may change to w A = 0.30,
w P = 0.65, and wC = 0.05. Furthermore, the importance weights defining the com-
ponents of performance could change to w I = 0.25, w E = 0.0, and w R = 0.75. This
preference structure gives the result depicted in Figure 8.10. The ordering of the
alternatives is unchanged and the picture resembles that in Figure 8.7. There are
differences, however, in the composition of the MoE for each alternative. Range
plays a more important role, and availability plays a less important role. Never-
theless, the original ordering of alternatives is robust to this change in the future
condition.
Now assume that intelligence reports indicate that the potential enemy cannot
afford to reequip its bomber fleet with new aircraft. Instead, they have decided
on an avionics upgrade to give the existing bomber fleet radar jamming capability
or electronic countermeasures (ECM). Under this future condition the decision-
maker may still favor performance over availability, and w A = 0.30, w P = 0.65,
and wC = 0.05. ECCM becomes important, and the weights within performance
may change to w I = 0.10, w E = 0.75, and w R = 0.15. The resulting MoE are

1.000

0.900

0.800

0.700

0.600
e> 0.500

0.400

0.300

0.200

0.100

0000
Sloat 2 Sloat 3 SkyRay Sw eeper

Figure 8.10 Supersonic attack scenario.


222 K. D. Wall, C. A. MacKenzie

1.000

0.900

0.800

0.700

0.600

G 0.500
>
0400

0.300

0.200

0.100

0.000
Sloat 2 Sloat 3 Sky Ray Sweeper

Figure 8.11 ECM scenario.

depicted in Figure 8.11. Sweeper is now the most effective alternative for this
future condition because it is the only radar with ECCM capability.

8.6 Cost-effectiveness
Computing the MoE for each alternative allows the decision-maker to order the
alternatives from most effective to least effective. This is only half the story,
though, because cost also matters. Ultimately the decision-maker will have to inte-
grate cost information with effectiveness and engage in cost-effectiveness analysis
(CEA). This section presents the conceptual elements used in the analysis and
outlines the process followed in the analysis.

8.6.1 Conceptual framework


The decision-maker ultimately pursues two overall objectives when searching for
a solution: (1) maximize effectiveness and (2) minimize cost. An alternative j
will be evaluated in terms of its MoE, v( j ), and its discounted life-cycle cost,
c( j ). Drawing a picture of effectiveness and cost on a scatter plot can help us
think in these two dimensions at the same time. Cost is plotted on the horizontal
axis, and effectiveness is plotted on the vertical axis. If the four alternatives in
the Sloat Radar example have the discounted life-cycle costs and MoEs (with the
importance weights from the baseline) as in Table 8.4, Figure 8.12 depicts this
information graphically.
This framework makes it easy for the decision-maker to see which alternative
is cheapest or most expensive and which alternative is most effective or least
effective. It shows the distance between alternatives—how much more effective
and/or costly is one alternative compared with another. Such a picture provides
all the information the decision-maker needs to choose the most cost-effective
Multiple objective decision-making 223
Table 8.4 Discounted life-cycle costs and MoEs for Sloat Radar

Alternative Cost (millions of dollars) Effectiveness (MoE)

Sloat 2 23.9 0.752


Sloat 3 41.6 0.765
SkyRay 63.5 0.934
Sweeper 70.4 0.799

1.00

0
0.90 SkyRay

0.80 0
0 0 Sweeper
Sloat2 Sloat3
0.70
e>
0.60

0.50

0.40

0.30
$0 $10 $20 $30 $40 $50 $60 $70 $80 $90

System Discounted Life-Cycle Cost ($M)

Figure 8.12 Alternatives in cost-effectiveness space.

alternative, the exact meaning of which requires understanding the many ways the
solution can be interpreted.

8.6.2 Solution concepts


Multiple ways exist to define a solution that minimizes cost and maximizes effec-
tiveness. The solutions are distinguished from each other based on the amount of
additional information required of the decision-maker. First, we present two con-
cepts that do not require any more information from the decision-maker. Second,
we develop solution concepts that require a little additional information from the
decision-maker. Finally, we conclude with a definition for the most cost-effective
solution. This last concept requires eliciting additional decision-maker preference
information and is the most demanding.

8.6.2.1 Superior solution


A superior solution is a feasible alternative that has the lowest cost and the great-
est effectiveness. It does not matter if you are a decision maker who places all
emphasis on minimizing cost or a decision-maker who places all emphasis on
224 K. D. Wall, C. A. MacKenzie
maximizing effectiveness. Both types of decision makers would select the superior
solution if one exists. Figure 8.13 illustrates this concept. Alternative 1 is superior
to all the others, and it is called the dominant solution because this alternative is
the cheapest and the most effective. We do not need any preference information
beyond that which we have already obtained in order to define effectiveness. No
alternatives exist in the “northwest” quadrant relative to alternative 1 because no
alternative is either cheaper or more effective than alternative 1.

8.6.2.2 Efficient solution


The efficient solution concept builds upon the superior solution concept. An effi-
cient solution is one that is not dominated by, or inferior to, any other feasible
alternative, and an alternative is not efficient because there exists another alter-
native that is superior to it. An example of this type of solution is found in
Figure 8.14. Alternatives 4 and 6 are dominated by, or are inferior to, alternatives
3 and 5, respectively. Consequently, alternatives 4 and 6 should not be selected.
Alternatives 2, 3, 5, and 7 are all efficient.

<fJ 4
•7
r----------... 6
c
3
u
4
~

"'
>- 4

System Cost

Figure 8.13 Superior or dominant solution.

<fJ
i;'l r •7
. •s
r----------...i 6
c

·"u
Q)

• 3
~ ~---------- . . 4
E
Q)

"'
>-
(f)
4

System Cost

Figure 8.14 Efficient solution.


Multiple objective decision-making 225
This solution concept does not necessarily yield unique answers. If only one
efficient solution exists, it is the superior solution. If more than one efficient solu-
tion exist, a set of non-unique solutions or alternatives exists. This may not be a
bad circumstance, however, because the decision-maker has flexibility, and several
alternatives can be defended as efficient.

8.6.2.3 Satisficing solution


A satisficing solution is a feasible solution that is “good enough” in the sense that
it exceeds a minimum level of effectiveness and does not exceed a maximum cost.
This solution concept requires the decision-maker to state a desired MoE and a
maximum life-cycle cost. An alternative that satisfies both of these requirements
simultaneously is satisfactory. Like the efficient solution concept, the satisficing
solution concept often yields non-unique answers. Figure 8.15 shows a situation
where both alternatives 3 and 5 are satisfactory solutions because both alternatives
are cheaper than the maximum cost (costmax ) and more effective than the minimum
effectiveness level (v( j )min ). Because both alternatives are satisfactory, a decision-
maker will need another solution concept to decide between the two alternatives.

8.6.2.4 Marginal reasoning solution


If a superior solution does not exist, marginal reasoning may be used to select
an alternative. This solution concept begins with an efficient set of alternatives,
such as those from Figure 8.14, which are reproduced in Figure 8.16. Suppose
the decision-maker is considering alternative 3. Alternatives 2 and 5 are its closest
neighbors: alternative 2 is similar in cost and alternative 5 is similar in effec-
tiveness. Two questions can be asked: (1) Is the marginal increase in cost worth
the marginal gain in effectiveness when moving to alternative 5? and (2) Is
the marginal savings in cost worth the marginal decrease in effectiveness when
moving to alternative 2? If the decision-maker prefers alternative 5 over alter-
native 3, the additional cost is worth paying to obtain the additional increase
in effectiveness. If the decision-maker prefers alternative 3 to alternative 5, the
marginal increase in effectiveness is not worth the marginal increase in cost. If the

I
I
I
max System Cost

I
·~
·~I e 6
- - - - - - ~;L - . - - ~------------ v(i) m;o
I
I
I
·~ I
I
I
I

co st max System Cost

Figure 8.15 Satisficing solution.


226 K. D. Wall, C. A. MacKenzie

"'"'
Ql
c: -------• 5 •7
--------, 3
Ql
.2:
t3
~ I
w I
E I
Ql I
Ui
>-
2• 1
(f) I
I
I
I
I

S y stem Cost

Figure 8.16 Marginal reasoning solution.

decision-maker cannot decide between alternative 3 and alternative 5, the marginal


increase in cost is balanced by the marginal increase in effectiveness, or equiva-
lently, the cost savings just compensate for the decrease in effectiveness. It is a
“toss up” between the two.
This reasoning can be applied to each alternative in the efficient set. After all
the alternatives have been considered, the decision-maker will arrive at one of two
situations: (1) one alternative is identified as the most preferred alternative; or (2)
two or more alternatives are identified as equivalent and, as a group, are more
preferred than than all the rest.

8.6.2.5 Importance weights for effectiveness and cost


The most formal solution concept asks the decision-maker to determine the rel-
ative importance of cost to effectiveness. This means there exists in the mind of
the decision-maker a payoff function that combines the two issues of concern:
(1) maximizing effectiveness and (2) minimizing cost. The first objective is rep-
resented by the MoE v( j ). The second objective is a function of the cost measure
c( j ) and represents a less-is-better preference relation. The decision-maker needs
to determine a value function for cost similar to the example given in Section 8.4:
“How much is enough?” where a value function for the weight of a radio is con-
structed. The value function for cost vc (c( j )) can be an exponential function as
in Equation (8.2) or a linear function as in Equation (8.3b) in which less is bet-
ter. Figure A8.1 in Appendix 8.1 depicts a linear value function for cost in which
cideal = $0 and ctoo much = $90 million.
The payoff function for the cost-effectiveness of the j th alternative is denoted
by V ( j ) and follows a simple additive form:

V ( j ) = W E · v( j ) + WC · vc (c( j )), (8.6)

where W E and WC are the importance weights for effectiveness and cost,
respectively. Upper case letters distinguish these weights from those used in
Multiple objective decision-making 227
defining v( j ). W E represents the importance to the decision-maker of maximizing
effectiveness and WC represents importance to the decision-maker of minimiz-
ing cost. The weights must satisfy 0 ≤ W E ≤ 1, 0 ≤ WC ≤ 1, and W E + WC = 1.
The relative importance of these two conflicting objectives is captured by the
ratio WC /W E .
Equation (8.6) represents the preferences of a rational decision-maker. When
confronted by a choice between two alternatives equal in effectiveness, the
decision-maker will choose the less expensive alternative because vc (c( j ))
decreases as the cost c( j ) increases. Similarly, when confronted by a choice
between two alternatives equal in cost, the decision-maker will choose the alter-
native with the greater MoE because v( j ) is greater for the alternative that is more
effective.
The overall cost-effectiveness function V ( j ) defines a preference structure that
allows us to develop valuable insights. We can see this graphically by rearranging
the cost-effectiveness function and drawing a straight line to represent V ( j ) on
the cost-effectiveness graph. Express v( j ) as a function of V ( j ) and c( j ):

1 WC
v( j ) = V ( j) − · vc (c( j )).
WE WE

For simplicity, we assume a linear individual value function—see Equa-


tion (8.3b)—for vc (c( j )) in which less is better. If we assume cideal = 0, then
vc (c( j )) = [c( j ) − ctoo much ]/[−ctoo much ] = 1 − k · c( j ) where k = 1/ctoo much . For
a fixed value of V ( j ), we see that v( j ) is a linear function of c( j ):

1 WC
v( j ) = V ( j) − · [1 − k · c( j )]
WE WE
1 WC WC
= V ( j) − + · k · c( j ).
WE WE WE

This is the equation for a straight line when we interpret c( j ) as the independent
variable and v( j ) as the dependent variable. The intercept for the line is given
by V ( j )/W E − WC /W E and the slope is given by k · WC /W E . This line repre-
sents all combinations of cost and effectiveness corresponding to a given level
of overall cost-effectiveness, V ( j ). As V ( j ) increases (for example, V ∗∗ > V ∗ in
Figure 8.17), this “isoquant” line shifts up and to the left. As we move closer to
the northwest corner of the cost-effectiveness plot, we move to greater levels of
overall cost-effectiveness.
Suppose the decision-maker is much more interested in minimizing cost
than maximizing effectiveness. This decision-maker would select weights where
WC  W E . The slope of the lines representing constant overall cost-effectiveness
would be very steep. This situation is depicted in Figure 8.18. Alternative 3 is
the best in this situation because it lies on the highest achievable isoquant of
cost-effectiveness. A decision-maker who places more emphasis on maximiz-
ing effectiveness would choose weights that result in less steep lines of constant
228 K. D. Wall, C. A. MacKenzie
Patient

!113 !113
!113 Patient
Patient
Patient

!113

System Cost

Figure 8.17 Cost-effectiveness linear preference.

/ / / / /
/ / / / /
/ / / / /
/ / / / /
/ / / / /
/ / / / /

"'"'c
/ / / / /
/ / / // • 7 /
Q)

Q) /
/
/
/ • s// / /
/
>
t5 /
/
/
/
!113 /
/
/
/
/
/

/
/
/

:Jg / / / / /
w / / / / /
/ / / / /
E / / / / /
Q)
/ / / /
1i5
>- / • 2 / / /
(/) / / / /
/ / / /
/ / / /
/ / / /
/ / / /

System Cost

Figure 8.18 WC  W E .

overall cost effectiveness, which could yield a picture like Figure 8.19. Now alter-
native 5 is the most cost-effective. Finally, consider a decision-maker who places
considerable importance on the maximization of effectiveness, which implies that
W E  WC . The slope of the lines is very small resulting in nearly flat isoquants of
cost-effectiveness as in Figure 8.20. Alternative 7 is now the most cost-effective
alternative.
The above three cases illustrate the importance of the efficient set, and the
most cost-effective alternative is selected from the set of efficient alternatives. The
most cost-effective depends on the decision-maker’s preferences for cost reduction
versus effectiveness maximization. The answer to the decision problem requires
elicitation of decision-maker preferences—we do not know what we mean by
“cost-effectiveness” until we incorporate decision-maker preferences over cost
and effectiveness.
Multiple objective decision-making 229

------ --- --- --- ---


--- --- --- --·-s-- -- ---- --- ---
----.. -3-- --- ---
-- --
--- --- --- -- -- -- ---
--- ---
--- -----.2 --- -- -- -- ---- -- --
--- -- --- --- -- --- -- -- S y stem Cost

Figure 8.19 W E > WC .

------------------------------
(/)

ill~ -------------- · ~------ -*"9------


"
~
~
E ------------------------------
w
0)
1il
:>.
(fJ ------------------------------
• 2

S y stem Cost

Figure 8.20 W E  WC .

8.6.3 Cost-effectiveness for Sloat Radar


Figure 8.12 shows that no superior solution exists among the four alternatives of
the Sloat Radar problem and Sweeper is dominated by SkyRay. The efficient set is
composed of Sloat 2 (the “do nothing” alternative), Sloat 3 (purchase a third radar
of the type already installed), and SkyRay. The question is which of the three
non-dominated alternatives in the efficient set is most cost-effective? The answer
cannot be given until we solicit additional preferences from the decision-maker.
A marginal reasoning solution may begin with Sloat 2 by asking the decision-
maker whether he is willing to spend an additional $17.7 million to achieve an
increase in 0.013 in effectiveness (the difference in MoEs between Sloat 3 and
Sloat 2). The increase in effectiveness occurs because Sloat 3 is available more
frequently than Sloat 2. If the decision-maker responds that he is unwilling to
spend the extra money for a small increase in effectiveness, we can ask him if
he is willing to spend an additional $39.6 million to achieve an increase in 0.182
in effectiveness (the difference in MoEs between SkyRay and Sloat 2). SkyRay
is more effective than Sloat 2 because it has more range, is available more often,
230 K. D. Wall, C. A. MacKenzie

1.0 ,---------------------
//-/y----/-
//~_ _/_/_/~--/-/~/c----
//~/,---/-/~
/ ----/-/'/~---
//_.---/-/c
/ r--/-/'
/

0.9 S,weepef /

0.8
// o/ //"" / ,"b /" / " ,., / "
S,weepef
/ / Sl9af"2 "/ "stoat 3/" /" "/ "/
0.7
/// ,"" ,./' //" ,.,., /// ,-/' //"
G
> ,-"" ,/'/ ,/'/ ,"" ,-"" ,/'/ ,"" ,-/'
0.6 // // ," // ," /// // ,"
/ / / // ," / / /
/// ,"" /// // /" ," //" //"
/ / / ," ," / / /

0.5

0.4 :///::::>/::::::-:::::>,/ /<::::>::::>-/ ,"


0. 3
/
/
/
/

~--~--~----~--~--~~-.~---.------.------.------,-----~
/ /
/
," /

$0 $10 $20 $30 $40 $50 $60 $70 $80 $90


System Discounted Life-Cycle Cost ($M)

Figure 8.21 Cost-effectiveness: WC /W E = 1.0.

1.0

0.9 S,weepef /

0.8
S,weepef /
0.7 "stoat
Sl9af"2 "stoat ,.,.,
G
>
0.6

0.5

0.4

0.3
$0 $10 $20 $30 $40 $50 $60 $70 $80 $90
System Discounted Life-Cycle Cost ($M)

Figure 8.22 Cost-effectiveness: WC /W E = 0.5.

and has greater interoperability. Framing the questions in this manner can help the
decision-maker understand precisely what additional capability he is getting for
an increased cost.
The more formal manner of answering which alternative is the most cost-
effective requires elicitation of the final set of importance weights, WC and W E .
Multiple objective decision-making 231
Suppose the decision-maker believes cost and effectiveness are equally important.
These preferences translates to WC /W E = 1.0 resulting in the picture given in
Figure 8.21. The decision-maker would select the “do nothing” alternative because
at least one isoquant line lies to the right of Sloat 2 and to the left of Sloat 3 and
SkyRay.
If the decision-maker believes maximizing effectiveness is twice as important as
minimizing cost, WC /W E = 0.5, which results in the picture given in Figure 8.22.
Sloat 2 is still most cost-effective because at least one isoquant line separates
Sloat 2 from Sloat 3 and SkyRay. If the decision-maker believes maximizing
effectiveness is four times more important than cost reduction, WC /W E = 0.25.
These preferences give the picture shown in Figure 8.23. SkyRay is the most cost-
effective because at least one isoquant line lies below SkyRay and above Sloat 2
and Sloat 3.
This method also can serve to find the ratio between WC and W E which would
make the decision-maker indifferent between two alternatives. For example, when
WC /W E = 1/3 we obtain the situation in Figure 8.24. Here Sloat 2 and SkyRay
are almost of equal cost-effectiveness. The analysis of cost-effectiveness afforded
by the model reduces the management question to one of asking the decision-
maker: “Do you feel that effectiveness is more than three times as important as
cost?” If the answer is yes, SkyRay should be selected. If the answer is no then
do nothing (Sloat 2) is the best alternative. The model helps to focus attention on
critical information.
Now that we have found the critical value of WC /W E 1/3, it is of interest
to consider the effects of uncertainty in the future condition. Two other plan-
ning scenarios besides the baseline were considered during the MoE discussion.

1 . 0.---------------
~-~-=---------------
----
~-----------------
----
~-~--.
--- D ---- ---
--- -----
-
------
0.9
------ --- --- --- ----- -------sk;Ray --- --- ---
----- --- ------- --- --------D ----
0.8
-------D
----- 0 ------
--- ---

0.7 ---- Sloat 2 -------sfoat-3 -----


--------- --------- --- --- ---
---- --- --- --- --- ... ----
--- --- ---
G
>
0.6
......... -- ... -- --- --- --- --- ---
--- __
--- ---
............
--- --- --- --- --- ---
--- ---- ---
0.5
--- --- ---
... -------
--- --- _... ---- --- ...... -- ...... ---
---
--- -----
0.4
------ --- --- --- ---
--- --- -------
0 . 3~~--~----~------~-----r------r---~,_----~------~--~~
__ ... -------
$0 $10 $20 $30 $40 $50 $60 $70 $80 $90
System Discounted Life-Cycle Cost ($M)

Figure 8.23 Cost-effectiveness: WC /W E = 0.25.


232 K. D. Wall, C. A. MacKenzie

$90

0.9 Kind
0.8

Kind
0.7 Kind Kind
G
> ~~~
0.6

0.5

0.4

$90
$90 $10 $20 $30 $40 $50 $60 $70 $80 $90
System Discounted Life-Cycle Cost ($M)

Figure 8.24 Cost-effectiveness: WC /W E = 0.33.

If the supersonic attack scenario is highly likely, then we use different weights
in the definition of effectiveness. The v( j ) values change for all alternatives and
Figure 8.25 shows that SkyRay is the most cost-effective if WC /W E ≤ 0.57. The
decision-maker should be asked: “Is effectiveness more than 1.75 times as impor-
tant than cost?” If he answers yes, SkyRay is the best alternative. The ECM
planning scenario changes the importance weights in effectiveness, and the result-
ing cost-effectiveness is shown in Figure 8.26. If WC /W E = 1/3, Sweeper is the
best alternative. Under this scenario, Sweeper remains the best alternative as long
as the decision-maker values effectiveness at least 1.4 times as much as cost, or
WC /W E ≤ 0.74.

8.7 Summary
Multiple objective decision problems arise very frequently. Their solution requires
the decision-maker to first determine what really matters and list the issues or
consequences of concern. This process of discovery should be pursued through
the construction of an objectives hierarchy. It is the single most important step
towards a solution. Without doing this we run the risk of not knowing “what is the
real problem” and of not asking “the right question.”
The lowest levels of the hierarchy define the individual measures of effective-
ness. These are the natural, constructed or proxy measurement scales by which
we begin to quantify the effectiveness of an alternative. The proper integration of
these measures into a single MoE requires quantification of decision-maker pref-
erences. First we need to know decision-maker preferences over marginal changes
in the individual effectiveness measures—the answer to “How much is enough?”
Multiple objective decision-making 233

1.0

0.9
,....skyRa~
0.8

0.7
,....skyRa~
e>
0.6

,....skyRa~ ,....skyRa~
0.5

0.4

0.3
$0 $10 $20 $30 $40 $50 $60 $70 $80 $90
System Discounted Life-Cycle Cost ($M)

Figure 8.25 Cost-effectiveness for supersonic attack scenario: WC /W E = 0.57.

1.0 /
/
/
/
/
/ /
/
/
/
/ / / / /
/ / / / /
/ / / / /
/ / / / / /
/ / / / / /
/ / / / / /
0.9 /
/
/

/
/
/
,."""' p "' ,."'
/
/
/
/

/ / / / / /
/ / /

/
/
/
/

/
/
/
/

/
/'
/
;>w'e..ep~r / /
/
/
/

/
/
/

0.8 /
/
/

/
/
/
/
/
/
/
/
/
/
/
/
/

/ / / / /
/ / / / / / /
/ / / / / / /
/ / / / / / /
/ / / / / / / /
/ / / / / / / /
/ / / / / / / /
0.7 /
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/

/ / / / / / / /
/ / / / / / /

e>
/

,."'
/ / / / / / / /

/
/
/

/
/
/
,." ,.,. /
/
/
/
/
/
/
/ / / / / / /
/ / / / / / / / /
0.6 /
/
/

/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/ / / /
/ / / / / / / / /
/ / / / / / / / /
/ / / / / / / / /
/ / / / / / / / / /
/ / / / / / / / / /
/ / / / / / / / / /
0.5 /
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/

,-0
/ / / / / / / / /
/ / / / / / / / /
/ / / / / / / / /
/ / / / / / / / /

,....skyRa~.........
/ / / / / / / / /
/ / / /

,.",.,. ,.o ",.,. . . "


/ / / /
/ / / / /
0.4 /
/
/
/
/
/
0 /
/
/

/
/
/ /
/
/

/ / / / / /

/
/
/
/

/
, / ~16at-2 ' / / /
/
/
/
Sloa~ )- '
/
/
/
/

/
/
/
/

/
/
/
/

/ / / / / / / / /
0.3
$0 $10 $20 $30 $40 $50 $60 $70 $80 $90
System Discounted Life-Cycle Cost($M)

Figure 8.26 Cost-effectiveness for ECM scenario: WC /W E = 0.74.

These changes can represent increases (when more is better) or decreases (when
less is better). This information allows us to convert the individual effectiveness
measures into individual value measures on a 0–1 scale. Second we need to know
decision-maker preferences across the individual measures—the answer to “How
234 K. D. Wall, C. A. MacKenzie
important is it?” This information allows us to combine the individual scaling
functions using a weighted sum defined by the importance weights.
Next, we combine the resulting MoE with the discounted life-cycle cost. This
provides all the information needed to conduct the analysis of cost-effectiveness.
Viewing things in a two-dimensional framework expedites thinking about the
solution by allowing visual inspection and making use of humans’ ability at
pattern recognition. In this framework, five solution concepts apply: (1) the supe-
rior solution, (2) the efficient solution, (3) the satisficing solution, (4) marginal
reasoning, and (5) weighting cost versus effectiveness. The first concept should
always be sought, and a superior solution (if one exists) is the best alternative
regardless of the preferences for cost reduction versus effectiveness maximiza-
tion. The efficient solution and the satisficing solution often yield non-unique
answers and gives the decision-maker flexibility over which alternative to select.
The set of efficient solutions is intrinsically important because because the most
cost-effective alternative will come from this set. Selecting the most cost-effective
alternative requires eliciting additional decision-maker preferences. We must
know the decision-maker’s preferences over cost minimization and effectiveness
maximization before determining the most cost-effective alternative.

Appendix 8.1: Individual value functions


There are many functions for describing decision-maker preferences over marginal
changes in a single effectiveness measure. The exponential function provides a
good approximation to many preferences:

1 − e−azi
vi (z i ) =
K

0.9

"'"'
""
0.8

""""
0.7

0.6
.,

""""
:::l 0.5
~
0.4

""""
0.3

0.2

""
0.1

0
$0 $20 $40 $60 $80 $100 $120

Discounted System Life Cycle Cost{$M}

Figure A8.1 Linear value function for cost.


Multiple objective decision-making 235
where either

z i = x i − x min

(the more-is-better case) or

z i = x max − x i

(the less-is-better case) and K = 1 − e−a[xmax −xmin ] is a constant that standardizes


the range of variation so that 0 ≤ vi (z i ) ≤ 1.
This function is flexible enough to include the linear value function. It can be
shown using calculus that, in the limit as a → 0, this function becomes

zi
vi (z i ) = .
x max − x min

This describes a decision-maker who places equal value on marginal changes of


equal amounts. For example, let a = 0.001, x min = 0 and x max = $90 million. We
obtain the function pictured in Figure A8.1. This represents a decision-maker who
values a $10 million cost savings the same, no matter if it is a reduction in cost
from $90 to $80 million or from $70 to $60 million.
This exponential form can be expanded to include a quadratic term,

2
1 − e−azi −b1 zi
vi (z i ) = ,
K1

or a cubic term,
2 3
1 − e−azi −b1 zi −b2 zi
vi (z i ) = ,
K2
2
where K 1 = 1 − e−a[xmax −xmin ]−b1 [xmax −xmin ] and K2 = 1 −
2 3
e−a[xmax −xmin ]−b1 [xmax −xmin ] −b2 [xmax −xmin ] . The quadratic form with a = 0 allows
value functions with an “S-shape” to be represented. The cubic form with a = b = 0
permits us to incorporate value functions that appear almost like “step” functions.
Higher-order terms can be specified but in practice are almost never needed.
The additional terms permit a wider range of preference behavior to be modeled
but are hardly ever worth the increased complexity.

References
Buede, Dennis M., and Terry A. Bresnick. 1992. “Applications of Decision Analysis to the
Military Systems Acquisition Process.” Interfaces 22:110–25.
Clemen, Robert T., and Terence Reilly. 2001. Making Hard Decisions with DecisionTools.
Mason, OH: South-Western.
236 K. D. Wall, C. A. MacKenzie
Dyer, James S., Thomas Edmunds, John C. Butler, and Jianmin Jia. 1998. “A Multiat-
tribute Utility Analysis of Alternatives for the Disposition of Surplus Weapons-Grade
Plutonium.” Operations Research 46:749–62.
Ewing Jr., Paul L., William Tarantino, and Gregory S. Parnell. 2006. “Use of Decision Anal-
ysis in the Army Base Realignment and Closure (BRAC) 2005 Military Value Analysis.”
Decision Analysis 3:33–49.
Hammond, John S., Ralph L. Keeney, and Howard Raiffa. 1999. Smart Choices: A Practical
Guide to Making Better Life Decisions. New York: Broadway Books.
Keeney, Ralph L. 1996. Value-Focused Thinking: A Path to Creative Decisionmaking.
Cambridge, MA: Harvard University Press.
Keeney, Ralph L., and Howard Raiffa. 1976. Decisions with Multiple Objectives: Prefer-
ences and Value Tradeoffs. New York: John Wiley & Sons.
Kirkwood, Craig W. 1997. Strategic Decision Making: Multiobjective Decision Analysis
with Spreadsheets. Belmont, CA: Brooks/Cole.
Parnell, Gregory S., Harry W. Conley, Jack A. Jackson, Lee J. Lehmkuhl, and John M.
Andrew. 1998. “Foundations 2025: A Value Model for Evaluating Future Air and Space
Forces. Management Science 44:1336–50.
Parnell, Gregory S., Ronald E. Metzger, Jason Merrick, and Richard Eilers. 2001. “Mul-
tiobjective Decision Analysis of Theater Missile Defense Architectures.” Systems
Engineering 4:24–34.
Stillwell, William G., David A. Seaver, and Ward Edwards. 1981. “A Comparison of Weight
Approximation Techniques in Multiattribute Utility Decision Making.” Organizational
Behavior and Human Performance 28:62–77.
von Winterfeldt, Detlof, and Ward Edwards. 1986. Decision Analysis and Behavioral
Research. Cambridge, UK: Cambridge University Press.
9 A new approach to evaluate
safety and force protection
investments
The value of a statistical life
Thomas J. Kniesner, John D. Leeth,
and Ryan S. Sullivan
9.1 Introduction
A fundamental tenet of economics is that public decisions should be evaluated in
terms of benefits and costs. Such decisions include laws, rules, regulations, policies,
and investments, and all are aimed at improving civilian safety and reducing military
casualties. Economic welfare is only increased if the benefits of safety improve-
ments exceed their costs. Whereas calculating monetary costs of safety investments
is relatively well established—determined through engineering or accounting stud-
ies (see Chapter 5, for example)—monetizing benefits can be highly controversial.
As we explain in this chapter, assessing the benefit of fewer fatalities and injuries
requires both an accurate forecast of fatalities (and injuries) prevented or eliminated
and an estimate of the monetary value of lives saved or injuries avoided.
Some contend that no monetary value can be placed on human life. The associ-
ated implication is that life is infinitely precious, so any action to improve safety
regardless of the cost is worthwhile. The dilemma facing government officials can
be stated as follows: although everyone believes saving lives is moral, societal
resources are limited, so actions to save lives in one area come at the expense of
other worthwhile goals, including the possibility of saving lives elsewhere. This
chapter reveals how pricing lives can allow government officials to determine if
the resources devoted to improving safety in one area might be better allocated to
other areas or other programs. We urge governments with limited budgets to con-
sider using the value of a statistical life (VSL) approach described in this chapter in
any cost–benefit analysis (CBA) of alternative force protection investments. This
will not only ensure resources are allocated to their highest valued use, but more
importantly, may result in more lives saved.
Any military that operates as if human life had infinite value would soon be
out of business. The focal message of our chapter is that hard choices must be
made because complete safety is impossible. Approving every force protection
investment (in armaments, technology or training) to reduce casualties or injuries
would rapidly exhaust military budgets.
In reality, most of us do not behave as if our life has infinite value. Many make
conscious decisions to drive small cars, knowing large cars are significantly safer,
because small cars are cheaper to own and operate. We routinely engage in risky
238 T. J. Kniesner, J. D. Leeth, R. S. Sullivan
activities (skiing, biking, roller-blading, etc.), because the perceived benefits com-
pensate for the risks. People smoke, drink alcoholic beverages, and eat unhealthy
food, even though health hazards are well known, presumably because the imme-
diate pleasure outweighs any possible long-run consequences. In short, people
routinely make cost–benefit decisions that influence their own safety, and those
decisions do not reflect an infinite value of life.
This chapter examines how economists evaluate the costs and benefits of safety
improvements, and then provides a contemporary case study of a major recent
force protection investment. The study examines the cost-effectiveness of adding
armor protection to tactical wheeled vehicles (TWVs) to protect soldiers. We
demonstrate how the conclusions depend on the measure called the “value of a
statistical life,” or VSL.
A common approach to value the benefits of greater safety is to observe actual
tradeoffs people make between safety and other valuable job or product charac-
teristics (Viscusi and Aldy 2003). Consider two identical jobs that only differ in
terms of safety: one job is completely safe, but in the other job there is a 1 in
100,000 (or 0.00001) chance of a workplace injury resulting in death. The com-
pletely safe job pays US$80 less per year than the job with the slight chance of a
workplace fatality.
This scenario offers two complementary perspectives. Workers in the safe job
essentially sacrifice US$80 annually in the form of a lower wage to eliminate
a 0.00001 chance of injury on the job that could result in death over the next
year. Alternately, workers in the dangerous job demand a wage premium of
US$80 extra annually to accept a 0.00001 increase in the chance of a deadly
injury.
One possible interpretation is that if 100,000 workers were each willing to pay
US$80 per year (a total of US$8 million) to avoid a 0.00001 probability of injury,
then one life per year would be saved on average. Collectively, the US$8 million
is the value workers implicitly place on that one life and implicitly on their own
individual lives. Because the actual life that is saved is unknown, but in a sense
is drawn randomly from the 100,000 workers, economists refer to the resulting
figure (US$8 million) as the VSL.
The VSL in this case can be obtained by dividing the wage gain in the riskier
occupation (US$80) by the higher probability of death (0.00001), resulting in
a dollar figure (US$80/0.00001 = US$8 million) standardized to one life saved.
Although the monetary value is expressed as the amount per life, it is derived by
observing the actual amount of money workers sacrifice to avoid a slightly higher
chance of death, and so might be better thought of as the value of mortality risk
reduction (US Environmental Protection Agency (EPA) 2010b).1
After describing the theory behind VSLs, this chapter briefly summarizes vari-
ous approaches used to generate VSL estimates and provides a broad overview of
empirical results from the literature. VSLs can vary by income and demographic
characteristics (age, gender, race), so that no single estimate is appropriate in all
situations. In fact, the range of VSL estimates reported in the literature turns
out to be sizeable, anywhere from US$0 to US$40 million. However, careful
Value of a statistical life 239
empirical modeling and improved risk information reduce the likely relevant range
of VSLs in the United States to between US$4 million and US$10 million. For
example, US Government agencies such as the EPA and US Department of Trans-
portation (DOT) generally evaluate safety improvements using a VSL of around
US$8 million.2
The chapter concludes with a force protection case study that examines the cost-
effectiveness of two large-scale vehicle replacement programs recently imple-
mented by the US military to increase vehicle armor protection for troops in
day-to-day operations. The principal aim of the replacement programs was to
reduce fatality and injury risks faced by US military personnel.
This case study illustrates a common problem when evaluating safety or force
protection investments. Estimates of the benefits of risk reduction are based on
engineering studies that do not consider changes in behavior that might mitigate
or completely counteract the safety investment. For example, requiring drivers
to wear seatbelts might reduce deaths or injuries in an accident, but could also
encourage those drivers to drive faster or more recklessly, thus increasing the num-
ber of accidents. The behavioral response places drivers at greater risk, as well
as increasing risks to other drivers, passengers, and pedestrians (Peltzman 1975).
Simply estimating the lower probability of death or injury from using a seatbelt
in a specific accident scenario will tend to overestimate the benefits of requiring
seatbelt usage if behavioral responses that increase the total number of accidents
are not included.
The case study presented in this chapter shows that the move from a lightly
armored vehicle to a moderately armored vehicle reduced infantry fatalities at
a relatively low cost of US$1 million to US$2 million per life saved for some
military units. Given an estimated VSL in the range of US$4 million to US$10 mil-
lion, such a force protection investment easily passes the cost-effectiveness test.
Interestingly, switching to a more costly, highly armored third alternative did
not appreciably reduce fatalities in any of the different military units examined,
despite engineering studies that demonstrated the new vehicle was clearly safer
against specific risks such as improvised explosive devices (IEDS).
An explanation for the surprising lack of life-saving benefits observed from
the added armor protection in the most costly vehicle may come from offset-
ting behavior of our own troops or from actions by the enemy in response to
the change. Both may have contributed to counteract the added force protec-
tion investment of additional armor, ultimately reflecting diminishing returns that
essentially neutralized the value of the most costly safety investment. This result
points to the importance of properly framing a cost–benefit study, especially
when lives are at stake, and the value of VSL to help guide force protection
investments.

9.2 The economic approach to valuing safety investments


The economic theory behind valuing safety is built on the assumption that people
prefer not to gamble excessively with their lives or their health. Thrill seekers may
240 T. J. Kniesner, J. D. Leeth, R. S. Sullivan
seek danger for its own sake, but are in the minority. The vast majority dislike
risk and are only willing to accept it if they are compensated in some way. All
else being equal, a worker might choose a more hazardous workplace if he or she
receives a higher wage, or a consumer might be willing to purchase a less safe
product if its price is sufficiently low.3
On the other side of the market, it is costly for firms to improve workplace and
product safety. To avoid the added costs, firms typically assess the possibility of
paying higher wages (i.e. compensating wage differentials) to workers to accept
slightly greater risks, or offering lower prices to customers to accept slightly more
hazardous products.
Consider a labor market where there are only two types of jobs: a job with zero
risk of workplace injury (π = 0) and a job with a high risk of workplace injury
(π = πhigh ). The jobs are identical in every other respect. With the exception of
thrill seekers, no worker would willingly choose the dangerous job over a safe job
for the same wage. At best, the worker in the dangerous job is uninjured and earns
the same as in the perfectly safe job; at worst, the worker is injured and bears the
pain and cost of the injury and corresponding loss of income.4
For workers to accept the gamble of the dangerous job, the expected payoff
if uninjured must be high enough to outweigh the expected loss if injured. This
payoff might include a higher wage, better fringe benefits, or a more pleasant work
environment, but for simplicity suppose the only compensation for the riskier job
is a higher wage. Differences among workers in their economic circumstances,
family situations, and general tastes and preferences will cause some workers to
demand a very large compensating wage differential to accept extra risk, while
others might be willing to accept a relatively small wage premium.
Programs that improve workplace safety have costs. Firms may need to pur-
chase additional equipment or protective devices, install machine guards, slow
down the pace of production or stop production entirely to service equipment,
hire consultants to advise management or train workers in safe procedures. Those
firms may also have to devote valuable management time to monitor safety.
For some employers safety-enhancing efforts may be quite expensive, but for
other employers the costs may be slight. Because of inherent dangers in produc-
tion, similar to the military, firms in mining, logging, fishing, and construction will
need to spend relatively more than firms in manufacturing, retail trade, or finan-
cial services to achieve the same level of safety for their workers. To be willing to
bear the costs of expanding safety programs firms must anticipate corresponding
economic benefits such as greater worker job satisfaction (leading to lower com-
pensation payments) or the ability to produce greater output, charge higher prices,
or pay smaller insurance premiums.
If wages were the same in both safe and risky occupations, all workers (with the
exception of risk seekers) would choose to work in establishments with complete
safety. Workers would gain nothing from accepting any job risk, and firms would
earn no economic returns from investing in safety. However, the excess supply of
workers choosing safe jobs will cause their wages to fall, while excess demand by
firms for workers required for risky jobs will cause wages in risky jobs to rise.
Value of a statistical life 241
The resulting wage gap between the two types of jobs will cause some workers
to improve their welfare by accepting a risky job and some firms to realize they can
expand profits by investing to reduce job hazards. The decisions of workers and firms
between the two types of jobs will continue until no firm can expand its profits by
adjusting its expenditures on safety, and no worker can expand his or her expected
utility by changing jobs.5
An example is provided by all-volunteer military forces. Assuming the military is
the riskier occupation, all else being equal, compensating wage differentials (which
might include extra job satisfaction, pride in serving the country, education and train-
ing opportunities) are required to recruit and retain individuals that have a choice of
safer occupations. In turn, the military as an employer can reduce the premium (com-
pensating wage differential) it pays soldiers through its force protection investments
that reduce risk of death or injury.
Because the wage difference (whigh − w0 ) seen in Figure 9.1 reflects both sides
of the market, it measures both the marginal cost (the added cost of safety pro-
grams) and the marginal benefit (as measured by worker preferences) of improving
safety. Notice that if job risk represents the chance of a fatal workplace injury,
then the implied VSL in this market is equal to (whigh − w0 )/(πhigh − 0), which is
equivalent to the earlier VSL calculation (US$80/0.00001 = US$8 million).
The process that sorts workers and firms among the various levels of risk is
the same when there are three or more distinct levels of risk. Wages rise as
risk expands, and the higher wage encourages less risk-averse workers to accept
higher-risk jobs, and firms with lower costs of providing safety to eliminate
hazards. In the limit, risk becomes continuous and the resulting equilibrium rela-
tionship between wages and risk is captured by a function described by a curve
such as the one in Figure 9.2.6

π=0 π = πhigh

Wage Wage Sh

whigh
S0

w0 Dh

D0

Labor Labor

Figure 9.1 Labor market equilibrium with discrete risk.


242 T. J. Kniesner, J. D. Leeth, R. S. Sullivan

Wage
W (π)

Probability of injury (π)

Figure 9.2 Labor market equilibrium with continuous risk.

As is well known in economics, firms maximize profits by using inputs to the


point where their marginal benefits equal their marginal costs. In the case of safety
inputs, marginal costs might include the expense of buying additional equipment,
slowing production, or paying for safety training, while marginal benefits include
the lower wages firms must pay to attract workers as risks fall.
Because the costs of providing a given level of safety vary among workplaces, the
optimum level of safety in different industries (occupations) will also vary. In equi-
librium, all firms tend to provide their workers a level of safety where the marginal
benefits of additional safety investments match the marginal costs. Firms with lower
inherent costs of reducing hazards provide a safer workplace and locate to the left
along the wage relationship shown in Figure 9.2, and firms with higher costs provide
less safety and locate to the right.
Similar to firms, people maximize their welfare by pursuing actions to the point
where the marginal benefit of an action equals its marginal cost. In terms of safety
the marginal benefit of accepting a lower-risk job is the lower chance of injury,
and the marginal cost is the corresponding drop in the wage. All workers max-
imize their welfare by comparing the marginal benefit of added safety against
the additional cost. More risk-averse workers will value the potential benefits of
safety investments relatively more (or are willing to accept relatively lower wages
to avoid safety risks), while the opposite is true of less risk-averse workers.
Firms tend to supply more or less workplace safety based on compensating wage
differentials they must offer to recruit and retain workers, and their costs to produce
a safe work environment. In turn, workers tend to sort themselves into job risk cat-
egories based on market wage differentials and their personal preferences (relative
risk aversion) regarding safety.
In the end, wages must adjust along the entire risk spectrum to equilibrate sup-
ply and demand for labor. If there is a shortage of workers at any risk level, wages
Value of a statistical life 243
will rise at that risk level (in that industry), causing some firms to alter their expen-
ditures on safety and some workers to relocate to the now better-paying jobs. The
movement of firms and workers will create labor shortages elsewhere and cause
wages to rise at the other levels of risk. A surplus of labor at any level of risk
will cause wages to fall at that risk level, which will cascade throughout the entire
risk spectrum as firms and workers adjust to the new lower wage. Equilibrium is
established in the market when there is no excess demand or supply of labor at
any risk level.
The slope of the wage function shown in Figure 9.2 measures the income
workers must sacrifice to lower their chance of injury by a small amount (or alter-
natively, the amount they must be compensated to accept greater risk). Because
workers maximize their well-being by equating marginal cost with marginal ben-
efit, the slope of the wage function also equals the value workers place on small
improvements in safety. Similar to the earlier case with only two levels of risk,
a VSL can be calculated for a representative worker by dividing the wage pre-
mium necessary for a worker to accept a small increase in the probability of death
(measured by the slope of the wage function evaluated at the mean level of risk),
by the associated increase in the probability of death. An implicit value of injury
can be found in a similar manner by dividing an estimated market wage increase
required for a representative worker to accept a small increase in the probability
of a work-related non-fatal injury or illness, by the change in that probability of
injury or illness. The resulting calculation is known as the value of a statistical
injury (VSI).
Market equilibrium in product markets is similar to that in labor markets, except
the end result is an inverse (negative) relationship between price and risk. The
added cost of producing safer products implies that the price of the product must
rise as the product’s risk (probability of injury from using the product) falls,
compensating manufacturers for the extra costs.
Consumers value both additional safety and more goods and services, mean-
ing that, all else being equal, they are only willing to purchase riskier products if
they are sufficiently cheaper than safer products. The interaction of firms and con-
sumers in the marketplace generates a price function where firms sort themselves
based on their abilities to produce safe products, while consumers sort themselves
based on their relative valuations of increased safety. In equilibrium the slope of
the market price function provides an estimate of consumers’ willingness to pay
for added safety, which can also be used to generate a VSL or VSI (Viscusi and
Aldy 2003).

9.3 Value of statistical life estimates


The majority of VSL estimates are derived from estimates of a market wage
equation such as that presented in Appendix 9.1. Variation in wages is typically
explained by workplace risk, controlling for other job characteristics and various
demographic variables. The relationship obtained between wages and job risk,
controlling for other factors, generates the data needed to estimate a VSL.
244 T. J. Kniesner, J. D. Leeth, R. S. Sullivan
Viscusi (2004) provides a good example. For his measure of risk Viscusi relies
on data drawn from the Census of Fatal Occupational Injuries (CFOI), which are
discussed in more detail in Viscusi (2013). Since 1992, the Bureau of Labor Statis-
tics (BLS) has compiled yearly data on all workplace fatalities in the United States
by examining death certificates, medical examiner reports, Occupational Safety
and Health Administration (OSHA) reports, and workers’ compensation records.
In his sample data the average fatality risk is 4/100,000, with the lowest risk level
0.6/100,000, and the highest nearly 25/100,000.7
Viscusi then combines the fatality risk data with individual worker demographic
data from the 1997 merged outgoing rotation groups of the Current Population
Survey (CPS).8 His estimating equation assumes a linear relationship between
log hourly wages and fatal injury risk (see Equation (9A.1.1) in Appendix 9.1).
The VSL calculated for all workers from his baseline regression analysis is
US$12.1 million.9
House prices provide another avenue for determining implied VSLs. All else
being equal, houses located in less desirable (or riskier) areas should sell for less
than houses located in more desirable areas. If desirability reflects the absence
of pollution or other health hazards, then the smaller chance of premature death,
injury or illness and the resulting higher price can be used to estimate a VSL.
Gayer et al. (2000) estimate a house market price equation to determine the
implicit value of avoiding cancer. See Appendix 9.2 for the estimating equation
of the relationship between house prices and environmental risks. The authors use
risk information and individual housing data drawn from an area surrounding a
Superfund Hazard site in the greater Grand Rapids region of Michigan to estimate
the relationship between housing prices and environmental (cancer) risk.10 They
find that in the period following the EPA’s release of its report, a 1.81 per 1,000,000
reduction in cancer risk (the mean level in the sample) raises the price of the
average house by about US$25. The estimates imply that the household valuation
of reducing one cancer is about US$13.9 million. Given an average number of
household members of about 2.6, the individual value of reducing one cancer (or
implicit VSL) is estimated to be about US$5.4 million.
Economists have examined a variety of products and actions to determine
reasonable values for VSLs. When consumers purchase products that directly
decrease risk such as bike helmets or smoke detectors, the approach is quite
straightforward: simply divide the cost of the product by the lower likelihood of
death. In some cases actions to expand safety may not require a monetary outlay,
such as using a seat belt or driving more slowly, but even here it is often possible to
generate a VSL by combining the safety effects of the action with dollar estimates
of the opportunity cost of time (Viscusi and Aldy 2003).
In a particularly interesting and relevant example of the indirect approach,
Rohlfs (2012) develops VSL estimates by examining two tactics used by young
men during the Vietnam War to avoid the draft: enrolling in college or volun-
tarily enlisting. Although one might not necessarily view voluntarily enlisting in
the military as a sensible strategy to avoid the draft, by volunteering men could
avoid the possibility of having their careers interrupted through required military
Value of a statistical life 245
service and they could choose their branch of service, obtain specialty training,
and in some cases enter as officers. Enlisting also improved one’s chance of sur-
vival. From 1966 to the end of US involvement in the Vietnam War the death
rates of volunteers and officers were significantly lower than the death rates for
draftees.
To see how Rohlfs determines the value of avoiding the draft by voluntarily
enlisting in the military consider Figure 9.3, which shows two supply curves. The
first is the supply of men to the military if there were no draft. The second is the
supply of men volunteering to serve in the military given a draft. At every wage,
more men are willing to enlist in the military with a draft than without a draft
to avoid the costs of having their careers interrupted, to be able to choose their
branch of service and type of training, and to expand their odds of survival. The
supply of men with the draft is everywhere to the right of the supply curve without
a draft.
Given the level of pay in the military, the number of men who would enlist
without a draft is E nd and the number of men who would enlist with a draft is
E d . Rohlfs estimates the increase in the number of enlistees during the Vietnam
War era, all else being equal, and then uses those estimates with outside estimates
of the supply elasticity of men willing to enlist in the military to determine the
vertical distance between the two supply curves in Figure 9.3. The vertical distance
represents the increase in military pay necessary to generate the same increase in
enlistments as generated from the draft. It represents the dollar value of avoiding
the draft for the marginal enlistee during that period (an individual just indifferent
between enlisting and taking a chance with the draft).

Military
pay

Supply without Value of draft


a draft avoidance for the
marginal volunteer

P*
Supply with
a draft

Observed difference
in quantity supplied

End Ed Voluntary enlistment

Figure 9.3 Supply of voluntary enlistments and the Vietnam draft.


246 T. J. Kniesner, J. D. Leeth, R. S. Sullivan
Rohlfs uses a similar procedure to determine the dollar value of avoiding the
draft by enrolling in college for the marginal student (an individual just indiffer-
ent between attending college and taking his chance with the draft), except that in
the case of college he examines two demand curves instead of two supply curves.
Based on college enrollments the cost of avoiding the draft for the marginal stu-
dent was about US$30,000, and based on enlistments the cost of avoiding the draft
for the marginal volunteer was about US$117,000. Rohlfs generates the implied
VSLs by assuming the only cost of being drafted is the higher likelihood of death.
He refers to the resulting values as upper bounds on the true VSLs. Specifically, he
divides his estimates of the value of avoiding the draft by the associated differences
in mortality risk and finds VSLs for young men that are in the US$1.6 million to
US$5.2 million range using information from college enrollments, and VSLs that
are in the US$7.4 million to US$12.1 million range using information from mil-
itary enlistments. Rohlfs suggests that the higher range derived from information
on voluntary enlistments may more accurately reflect the true VSL of young men
at the time.11

9.4 Contingent valuation estimates


Besides using market data, VSLs can also be estimated using data from surveys
that question individuals about their willingness to pay for small reductions in
risk. Generally, the surveys used in contingent valuation (or so-called “stated pref-
erence”) studies describe a base risk level that can be reduced by engaging in
some activity, purchasing some product or service, or purchasing a different, safer
product than the product originally described.12
In its simplest form, survey data provide direct values of people’s willingness
to pay for improvements in safety. Unlike studies using market wages or housing
prices, contingent valuation studies attempt to uncover preferences directly.
Following the law of diminishing returns the marginal benefit of further gains
in safety declines as safety improves. Attempting to estimate the value functions
of workers or consumers that underlie market wage or housing price functions is
quite difficult.13 Because the price of safety is implicit, identifying the parameters
of the structural equations underling equilibrium is more difficult than estimating
a standard simultaneous equation system such as the supply and demand for a
homogenous product such as milk (Brown and Rosen 1982; Epple 1987; Kahn
and Lang 1988).
As with all surveys, one must ensure that respondents carefully consider the
alternatives presented. Specifically, researchers should attempt to create unbi-
ased surveys. For example, respondents tend to place a higher value on a safety
device or policy when it is the only option available on a survey. Also, one must
consider the order of the survey questions. Respondents often place higher val-
ues on safety devices or policies when they appear first in a set of questions.
Finally, “willingness to say” does not always correspond to “willingness to pay.”
Such complications can undermine survey results and bias any subsequent VSL
estimates.14
Value of a statistical life 247
Due to inherent problems with contingent valuation surveys, many economists
prefer to rely on VSLs generated from market data that reflect consumers’
(or workers’) actual willingness to pay (or to be compensated) for a marginal
improvement (deterioration) in safety.15

9.5 Generating VSL estimates


The range of VSL estimates derived from the empirical literature is quite large,
on the order of US$0–US$40 million (Mrozek and Taylor 2002; Viscusi and Aldy
2003; Bellavance et al. 2009).16 The variation can create a dilemma for policy-
makers. A proper evaluation of any program designed to improve health and safety
requires accurate estimates of benefits from the program. Employing too low
a VSL in workforce protection and other safety investments will underestimate
the benefits of mortality risk reduction, resulting in the rejection of worthwhile
programs. Setting too high a VSL will tend to overestimate the benefits of spe-
cific mortality risk reduction investments, resulting in accepting extremely costly
programs and thereby diverting resources away from other more valuable uses.
One approach to solving the dilemma of the wide range of VSL estimates is
to use meta-analysis to combine the various estimates into a single value that
corresponds roughly to the mean of the estimates. For example, the US EPA
recommends using US$8.0 million to value the lives saved from eliminating envi-
ronmental hazards, based on the mean of a Weibull statistical distribution fitted to
26 separate VSL estimates (US EPA 2010a). The US DOT uses an US$8.6 million
VSL to evaluate the benefits of improving transportation safety, based on the mean
value of nine fairly recent labor market VSL studies (US DOT 2008).17
Kniesner et al. (2012) offer another approach and demonstrate how using the
best available data and econometric tools can significantly narrow the range of
VSL estimates.18 They demonstrate how systematic econometric modeling nar-
rows the estimated VSL from a range of US$0–US$40 million to US$4 million–
US$10 million, which the authors demonstrate facilitates the choice of the proper
labor market-based VSL to use for public policy (Kniesner et al. 2012).
Government agencies constantly make choices on how to spend scarce resources
that impact public safety. Achieving complete safety, even if technologically fea-
sible, is never possible due to resource scarcity. Economists regularly employ the
monetary benefit metric of a VSL to help evaluate safety and force protection invest-
ments. As discussed earlier, the critical value of a VSL is revealed by individuals’
implicit willingness to pay for reduced harm via compensating wage differentials,
or lower prices paid for products that involve somewhat greater hazards.
The manner in which a death might occur and personal characteristics affect-
ing the willingness to accept danger (for example age or education) are important
in military applications, and both can play important roles in evaluating the cost-
effectiveness of increased personal safety or military force protection investments.
The next section offers an application of VSL in a cost–benefit study of a mil-
itary force protection investment to improve the armor of TWVs. The same
principles that find applications in evaluating government investments in safety
248 T. J. Kniesner, J. D. Leeth, R. S. Sullivan
improvements (EPA, DOT, DOL), and in private labor and product markets, are
equally relevant and applicable to military force protection investments.

9.6 Case study: Force protection investment—armoring


tactical wheeled vehicles
The following case study summarizes the methods and results presented in a
recently published article by Rohlfs and Sullivan (2013a) entitled “The Cost-
Effectiveness of Armored Tactical Wheeled Vehicles for Overseas US Army
Operations.” In their study, Rohlfs and Sullivan use a VSL cutoff of US$7.5 mil-
lion to determine the cost-effectiveness of three separate armored tactical wheeled
vehicles. In addition to summarizing the methods and findings, we offer a dis-
cussion of policy implications and limitations of using VSL cutoffs in evaluating
tradeoffs between money spent (cost) and lives saved (benefits) for this and other
military programs.
Due to the sensitive nature of the data used in Rohlfs and Sullivan (2013a), the
alternative vehicles in the case study are labeled with the generic titles of Type 1,
2, or 3, the months of data are denoted 1 through 71, and the theater of operations
is labeled Theater A. The US Army maintains each of the three vehicle types
primarily for troop transport, and the vehicles vary in terms of maneuverability,
weight, and speed. However, the primary difference among the vehicles is their
armor protection.19

9.6.1 Background and data


As discussed in Rohlfs and Sullivan (2013a), over the course of the Global War on
Terror the US Army engaged in two large-scale force protection investments that
involved vehicle replacement programs to reduce fatality and injury risks for US
troops. Both programs affected wheeled ground vehicles that US forces used for
day-to-day operations. The first replacement program involved replacing a large
number of lightly armored Type 1 TWVs, which cost roughly US$50,000 apiece,
with medium armored Type 2 TWVs that cost more than three times as much, or
US$170,000.20
The second replacement program began at the time the first program was end-
ing, and it involved replacing about 9,000 Type 2 TWVs with heavily armored
Type 3 TWVs that cost more than three times as much as the Type 2s, around
US$600,000 apiece. The total investment in the Type 3 TWV force protection pro-
gram was over US$50 billion, almost 10 percent of the US base defense budget at
the time.21
To conduct an (ex-post) military CBA of alternative force protection invest-
ments requires estimates of costs (procurement, maintenance, fuel) and benefits
(lives saved). The maintenance and fuel costs of Type 2 and Type 3 TWVs are
typically higher than for Type 1 TWVs. Rohlfs and Sullivan estimated in-theater
fuel and maintenance costs of US$5.50 per mile for Type 1 TWVs, and US$10.50
per mile for Type 2 and Type 3 TWVs.
Value of a statistical life 249
Type 3 TWVs are also much heavier and therefore more expensive to transport
overseas in comparison to the other types. Moreover, the physical depreciation (dete-
rioration) of the vehicle types is higher in conflict conditions than in peacetime,
and thus impacts expected lifetime costs, which requires further adjustments. For
purposes of their study, Rohlfs and Sullivan (2013a) assume a one-way trip and
three-year lifespan for all vehicle types, and that the US Army uses each vehicle
in combat for all three years. The estimated costs of an average three-year vehi-
cle deployment (including procurement and transportation) were US$143,000 for a
Type 1 TWV, US$345,000 for a Type 2 TWV, and US$780,000 for a Type 3 TWV.22
To evaluate the force protection benefits of the three alternatives, Rohlfs and
Sullivan (2013a) use casualty and unit-level characteristics in Theater A pro-
vided by the Defense Manpower Data Center (DMDC) for all Army units over
a 71-month time period. They group the data at the battalion level to match
up closely with how the Army assigned vehicles and tasks. They also group
the data into four distinct unit types: (i) infantry, (ii) armored and cavalry,
(iii) administrative and support, and (iv) other.
The Army incurred 34 deaths and 266 combat-related injuries in Theater A on
average each month. The standard deviation in deaths across months is 23, while
the standard deviation in combat-related injuries is 166. Total Army casualties in
Theater A often changed by more than 100 from one month to the next.
Unit-level controls include detailed information on the number of troops; the
fractions that are officers, Privates or Privates First Class (PFC), high school grad-
uates, male, black, or Hispanic; average days of deployment experience; age by
unit and month; and the unit’s name and home state within the United States.

9.6.2 Armored vehicle case study: econometric methods


and discussion of results
To evaluate the cost-effectiveness of vehicle Types 1–3, Rohlfs and Sullivan
(2013a) first examine the relationship between the three vehicle types and over-
all unit fatalities. Specifically, they use standard econometric methods including
regression analysis to estimate changes in unit-level casualty rates as a function
of each unit’s complement of vehicle types while controlling for other factors.
Details of the modeling approach can be found in Appendix 9.3. The results are
summarized below.23

• Preferred estimates indicate that the medium armor provided by Type 2


TWVs over the lightly protected Type 1 TWVs reduced fatalities at a cost
of US$1 million to US$2 million per life saved for infantry units, well below
the US$7.5 million VSL threshold set out in the study.
• For armored and cavalry units, and administrative and support units, the esti-
mates indicate that the switch from Type 1 to Type 2 TWVs did not reduce
fatalities.
• The switch from Type 2 to Type 3 TWVs did not appreciably reduce fatalities
for any of the four unit types, despite the higher cost and heavier armor of
Type 3 TWVs.
250 T. J. Kniesner, J. D. Leeth, R. S. Sullivan
In terms of evaluating tradeoffs between money spent (costs) and lives saved
(benefits), the policy implications are relatively straightforward. It does not appear
that the extra armor protection provided by Types 2 or 3 TWVs over Type 1 TWVs
provided enough benefit (in terms of lives saved) to warrant the additional costs for
armored and cavalry or administration and support units. Thus, results from Rohlfs
and Sullivan (2013a) suggest it would have been more cost-effective to use Type 1
TWVs in day-to-day operations for armored and cavalry, and administration and
support units, and to deploy Type 2s only to infantry units. Money saved by follow-
ing recommendations from the cost–benefit calculations could presumably have
been better used in alternative force protection investments to save more lives, or
to increase social welfare through higher compensating wage differentials.
Whereas for infantry units the results suggested Type 2 TWVs were signifi-
cantly more cost-effective in comparison to Type 1 TWVs (fatalities were reduced
at a relatively low cost of US$1 million to US$2 million per life saved), the addi-
tional armor protection provided by Type 3 over Type 2 TWVs did not appreciably
reduce casualties. The findings suggest the preferred policy would have favored
infantry units using Type 2 TWVs, rather than the less expensive Types 1s or the
more expensive Type 3s.
Surprisingly, Type 3 vehicles did not appear to reduce fatalities or injuries
appreciably for any of the four unit types, despite Type 3’s substantially greater
armor protection. The results in Rohlfs and Sullivan (2013a) therefore suggest
the significantly higher investment required for Type 3s was not cost-effective in
operations for any of the unit types. An important caveat is that political, psy-
chological, operational or other benefits may exist that broaden the scope of the
calculations beyond the simple evaluation of vehicle-provided force protection.
Such additional considerations naturally complicate any policy recommendations
that might come from a basic force protection CBA.
One key distinction between the study by Rohlfs and Sullivan (2013a) and other
authors who have studied the topic (citations are withheld due to security con-
cerns), is that the other studies tend to focus exclusively on differences in casualty
rates from IEDs among the various vehicle types (with a specific focus on Type 3
TWVs). In contrast, Rohlfs and Sullivan measure the full effects of changing
vehicle type, taking into account the intrinsic properties of the vehicles and any
behavioral responses by the units.
For instance, other studies highlight findings that show the fraction of IED
attacks resulting in death is higher for Type 1 and 2 vehicles than for Type 3
vehicles. Some authors used the IED results to suggest Type 3 TWVs saved thou-
sands of lives in comparison to Type 1 or 2 TWVs. It should be noted, however,
that several studies incorrectly assume soldiers in Type 3 TWVs are attacked by
IEDs at the same rate as soldiers in other vehicle types. The other researchers also
implicitly assume that the design of the extra protection provided by Type 3 TWVs
does not impact the tendency for soldiers to expose themselves to attacks that do
not involve IEDs or that occur outside those vehicles.
As discussed in Rohlfs and Sullivan (2013b), it appears that several implicit
assumptions used in previous studies may not be appropriate. When soldiers
Value of a statistical life 251
adopted the heavily armored Type 3 vehicles, insurgents may have directed their
efforts away from direct IED attacks on vehicles, and more toward indirect attacks
on soldiers once they were outside their vehicles. Moreover, providing troops with
safer vehicles may have increased their willingness to take certain risks, much as
wearing a motorcycle helmet makes motorcycle riders somewhat more likely to
take risks on the road.
In the (SDC) data, there was strong evidence of a substitution effect among
armored and cavalry units. When the Army replaced their Type 1 and 2 vehi-
cles with the heavier Type 3 vehicles, soldiers made greater use of their TWVs
and less use of vehicles with more firepower, such as tanks. Anecdotal evidence
also suggests that soldiers avoided certain dirt roads and bridges when driving
Type 3 TWVs due to their weight. If the behavioral responses just described
occurred more generally—changing the rate at which units were exposed to dif-
ferent types of attacks—then a more appropriate way to measure the vehicles’
full benefits is to use the approach in Rohlfs and Sullivan (2013a), who exam-
ine the impact of introducing more heavily armored vehicles on a unit’s overall
fatalities.
A limitation of the Rohlfs and Sullivan (2013a) study is that it does not monetize
all the capabilities provided by each of the three vehicles. The only benefits from
the vehicles that their study monetizes are the vehicles’ impacts on fatalities and
injuries.
Other benefits or costs not accounted for in the study could potentially alter
their estimates. For example, some vehicles may have greater maneuverability or
less weight in comparison to others, which may or may not enhance operational
combat capabilities. Increasing the scope of the cost–benefit calculations to cap-
ture additional capability attributes could cause the vehicles to be more or less
cost-effective than indicated in Rohlfs and Sullivan.
They also acknowledge their estimates “should be interpreted with the caveat
that they do not capture other potential benefits, such as increasing mission suc-
cess.” A more complete CBA would require monetizing any “additional benefits
and costs” as outlined in US OMB (1992). For the reasons listed above, policy
recommendations generated from this CBA might need to be adjusted if the scope
of the research is broadened to include additional benefits and costs in evaluating
the three alternative vehicles.
Capturing secondary and tertiary costs and benefits of force protection invest-
ments makes it more difficult to provide precise recommendations on optimal
vehicle investments to policy-makers. A valuable extension of the original Rohlfs
and Sullivan (2013a) study would be to examine other vehicle attributes that may
or may not impact mission success. However, given that force protection (saving
lives) was the Army’s stated goal, and the principal stated aim of the vehicle pro-
curement programs, advocating the provision of light armored Type 1 TWVs to
administrative and support units, and medium armored Type 2 TWVs to infantry
units, would appear to be a reasonable recommendation. The increase in armor
protection provided by investments in Type 3 TWVs over Type 2 TWVs does not
appear to justify their cost strictly in terms of life-saving capabilities.24
252 T. J. Kniesner, J. D. Leeth, R. S. Sullivan
9.7 Concluding observations and other military applications
The VSL offers a new approach to evaluate safety and force protection invest-
ments. Applying VSL to improve force protection investments, such as the
evaluation of alternative TWVs, raises an important issue. If the cost of an invest-
ment to save a life is substantially higher than the VSL threshold (US$7.5 million
in the study), could the money be better used (e.g. save more lives) in alternative
ways? A number of other possible military force protection investments exist and
constitute important topics for future research.
Worthy topics include investments in defensive weapons systems to protect
civilian populations or state-of-the-art body armor to protect soldiers. Another
possible application of VSL could be more contentious. It concerns one of the
costliest areas in the military—the substantial resources spent to care for wounded
warriors.
Creating a list of VSL values to be applied in medical interventions based
on patient characteristics (injury types, severity, ease of repair) might serve as
a type of triage tool. This could potentially help evaluate tradeoffs between money
spent and lives saved across numerous medical procedures to achieve the greatest
possible total health benefits for the military from its limited resources.
A related opportunity to apply this cost–benefit technique is to evaluate the
optimal number of medics to assign to different military units. For example, it is
typical for infantry platoons (of 40 or so persons) to have only one medic with
them in the field. Unfortunately, this means that some squads (usually around nine
people) frequently go on patrols without medics. Would it be cost-effective to
hire additional medics to help save more lives, rather than spend resources on
alternative life-saving investments? This is currently an open question that could
be informed using VSL.
In the current, fiscally constrained environment, the military must make difficult
choices about which life-saving programs to fund. Budget constraints limit the
ability to fund every safety and force protection proposal. This chapter offers a
variety of techniques and examples on how VSL can be used to conduct military
CBA on safety and force protection investments. Application of this powerful tool
will improve the management of scarce defense resources and help determine if
life-saving investments in one area might be better spent on other methods or
programs to reduce overall military casualties.

Appendix 9.1: Calculating VSL through wage differentials


Estimating equation of the relationship between
wages and workplace risk
The standard estimating equation underlying most labor market VSL
calculations is


m 
n
ln(w) = c + αi πi + β j X j + ε, (9A.1.1)
i=1 j =1
Value of a statistical life 253
where ln(w) is the natural logarithm of wage; πi are measures of workplace risk;
X i are demographic variables (such as education, race, marital status, and union
membership) and job characteristics (such as wage replacement under workers’
compensation insurance, and industry, occupation, or geographic location indi-
cators); ε is an error term; and c, αi , and β j are parameters to be estimated.
Many studies also include interaction terms between workplace risk and various
X i to determine variation in risk compensation by type of worker (union/nonunion,
white/black, native/immigrant, young/old) or by level of income replacement from
workers’ compensation insurance (Viscusi and Aldy 2003).

Appendix 9.2: Calculating VSL through house price


differentials
Estimating equation of the relationship between
house prices and environmental risks
As with labor market studies, the underlying framework for generating the VSL
values is a market price equation that relates house price to environmental hazards,
holding other factors constant. Specifically,


m 
n
ln( p) = c + αi E i + β j C j + ε, (9A.2.1)
i=1 j =1

where ln( p) is the natural logarithm of price; the E i are environmental hazards;
and C j are structure characteristics, neighborhood, and other control variables
that affect price (for instance, number of rooms, number of bathrooms, lot size,
neighborhood median income, and distance to central city) (Gayer et al. 2000).

Appendix 9.3: Calculating fatality and cost effects of


armoring tactical wheeled vehicles
Estimating equations of the relationships among fatalities,
expenditures, and vehicle types
For a given unit i in month t, suppose that fatalities are determined according to
the following linear equation:


3
fatalitiesit = αcjf × q j it + βcf xit + εitf , (9A.3.1)
j =1

where q1it , q2it , and q3it represent the quantities of each of three types of vehicles
possessed by unit i in month t; xit is a vector of control variables that might include
other vehicle quantities, troop characteristics, or fixed effects for month, province,
month by province, or unit; εitf is random error; and the coefficients are allowed
to vary by the unit’s classification c as infantry, armored or cavalry, administrative
254 T. J. Kniesner, J. D. Leeth, R. S. Sullivan
and support, or other. Let vehicle types one, two, and three be defined as Type 1,
Type 2, and Type 3 TWVs.
Because the focus of the Rohlfs and Sullivan (2013a) study is the effects of
replacing one vehicle type with another, it is convenient to rearrange the terms in
Equation (9A.3.1) to obtain the following specification:

fatalitiesit = (αc2f − αc1f ) × q2it + (αc3f − αc1f ) × q3it + αc1f



3
q j it + βcf xit + εit .
f
× (9A.3.2)
j =1

Equation
3
(9A.3.2) serves as their main regression specification, with
j =1 q j it , q2it , q3it , and varying formulations of xit as the regressors. The differ-
ences (αc2f − αc1f ) and (αc3f − αc1f ) measure the effects of replacing Type 1 TWV
with a Type 2 or Type 3 TWV.
Rohlfs and Sullivan (2013a) next focus on the costs related to the vehicles by
estimating changes in expenditures by unit as a function of the different vehicle
types and their usage rates. Their expenditure modeling technique follows.
To relate effects on fatalities to dollar costs, let p0 j denote the purchase price
(including shipping cost), and let p1 j denote the additional cost per mile driven
for a type j vehicle. Let mileage j it denote the miles driven of type j vehicles by
unit i in month t, and define expenditureit as unit i ’s month t expense for all three
vehicle types

 3
1
expenditureit = p0 j × q j it + p1 j × mileage j it (9A.3.3)
j =1
36

Expenditures can be written as a linear function symmetrically to the fatalities


equations:

expenditureit = (αc2
e
− αc1
e
) ∗ q2it + (αc3
e
− αc1
e
) ∗ q3it

3
+ αc1
e
∗ q j it + βce xit + εite (9A.3.4)
j =1

where (αc2
e
− αc1e
) and (αc3
e
− αc1
e
) represent the monthly costs of replacing a Type 1
TWV with a Type 2 or Type 3 TWV. Expressing the costs of changing vehicle type
through Equation (9A.3.4) helps to ensure that the same factors are held constant
and the same sets of observations are compared to one another for the cost as
for the fatalities estimation. The mileage and expenditure data are constructed
from the SDC sample and use the same sampling weights as the vehicle counts;
consequently, any measurement error in vehicle quantities that affects the fatalities
regressions does not affect the coefficients in Equation (9A.3.4).
Value of a statistical life 255
The cost per life saved from replacing a Type 1 with a Type 2 TWV is
f f
−(αc2e
− αc1
e
)/(αc2 − αc1 ), which expressed verbally is negative one times the ratio
of the difference between the coefficient for Type 2 TWVs and the coefficient for
Type 1 TWVs from Equation (9A.3.4), the expenditures equation, divided by the
corresponding difference in coefficients from Equation (9A.3.3), the fatalities equa-
tion. The cost per life saved from replacing a Type 2 TWV with a Type 3 TWV
is −(αc3e
− αc2
e
)/(αc3f − αc2f ), which is computed as negative one times the differ-
ence between the coefficient for Type 3 TWVs and the coefficient for Type 2 TWVs
from Equation (9A.3.4)divided by the corresponding difference in coefficients from
Equation (9A.3.3). Other dependent variables considered in the study are combat-
related injuries and miles driven by different vehicles. Rohlfs and Sullivan calculate
standard errors for the ratios by estimating Equations (9A.3.3)and (9A.3.4) together
using seemingly unrelated regression and applying the delta method.
Measuring effects on injuries helps to identify a benefit of changing vehicle
type not included in the cost per life saved calculations. To interpret the effects on
injuries better, Rohlfs and Sullivan (2013a) compute the ratios of injuries reduced
per life saved as (αc2i
− αc1
i
)/(αc2f − αc1f ) and (αc3
i
− αc2
i
)/(αc3f − αc2f ), where the i
superscript denotes coefficients from the injury equation. Measuring effects on the
mileage variables helps to determine the extent to which the replacement policies
led units to alter their behavior.

Notes
1 Rarely do you see two jobs that are otherwise identical except for the chance of a work-
place fatality. Economists derive VSL estimates econometrically using large datasets
that allow them to control for other factors affecting wages or product prices to generate
the ceteris paribus impact of risk. It is worth noting, though, that sometimes risks are
explicitly priced for workers by employers—for example, with so-called hazardous duty
pay in the military.
2 Implementing a safety improvement that costs more than the benefits generated can
result in more lives being lost because the funding could be better spent on alter-
native safety improvements. Kniesner and Leeth (2004), for instance, examine safety
inspections of underground coal mines by the Mine Safety and Health Administration
(MSHA) and find that inspections fail the cost-benefit test. They show that eliminat-
ing one safety inspection per mine per year (MSHA inspects underground coal mines
quarterly) and moving the funding to other activities such as heart disease screening
or purchasing defibrillators, would save 39,700 more miners’ lives. Formally, the cost-
benefit test shows whether resources are allocated efficiently among competing uses,
but many times a competing use can be an alternative means of improving safety and
saving lives.
3 The “all else equal” requirement means that one might not always see a simple positive
relationship between an occupation’s pay and its level of risk. Loggers face greater risk
on the job than financial analysts, but financial analysts earn more than loggers. The
additional education required to become a financial analyst outweighs the added risk
of logging. Economists attempt to control for other factors affecting wage rates using
large datasets and multiple regression techniques. Two popular reality TV shows, The
History Channel’s Ice Road Truckers and the Discovery Channel’s Deadliest Catch,
provide nice anecdotal examples of the relationship between pay and risk holding all
else equal. The two shows follow workers who face the most extreme hazards in their
256 T. J. Kniesner, J. D. Leeth, R. S. Sullivan
professions (driving across frozen lakes and rivers in the Arctic and fishing for Alaskan
king crabs in the Bering Sea) and document both the dangers they face and the very high
monetary rewards they receive for facing the dangers.
4 To use an analogy, it is as if the worker is being given the choice between a dollar with
certainty or a gamble where there is a 1 − π chance of a dollar payoff and a π chance
of a 0 payoff. Regardless of risk preferences, rational workers would always choose the
dollar payoff with certainty.
5 The wage gap sorts firms and workers into the two submarkets. The firm on the margin is
indifferent between offering safe or hazardous employment. For the so-called marginal
firm, the costs of the safety programs necessary to eliminate job hazards just equal
the benefits of greater safety, the drop in wage. Likewise, the worker on the margin
is indifferent between the two types of jobs: the utility gain from the higher wage just
offsets the utility loss from the greater likelihood of a drop in income, medical expenses,
and pain and suffering from an injury or illness.
6 See Kniesner and Leeth (1995, Chapter 3) for a formal derivation.
7 Viscusi uses the CFOI data to determine mortality risk by first grouping fatalities by
2-digit industry (72 industries) and then by 1-digit occupation (10 occupations) to gen-
erate a total of 720 industry-occupation cells. He then calculates the frequency of a
workplace death by dividing the average number of fatalities by the average number
of employees within each industry-occupation cell, using data from the period 1992 to
1997. By averaging over six years, Viscusi minimizes possible errors in characterizing
risk resulting from the random fluctuation of workplace fatalities over time.
8 As is fairly typical, he excludes from his sample agricultural workers, part-time work-
ers, workers younger than 18, and workers older than 65. Viscusi also does not include
interactions between fatal injury risk and other control variables (see discussion of
Equation (9A.1.1) in Appendix 9.1).
9 Although the VSL function depends on the values of the right-hand side in Equation
(9A.1.1) in Appendix 9.1, most commonly considered is the mean VSL. With a fatality
risk measure of deaths per 100,000 workers and a work year of 2,000 hours, the value of
a statistical life is VSL = α × exp(ln(w)) × 100,000 × 2,000. If α is small then it is the
case that α × exp(ln(w)) ≈ αw. (See Appendix 9.1.) All dollar figures presented have
been adjusted for inflation to 2010 using the CPI-U.
10 As part of a Remedial Investigation of a Superfund site, the EPA releases an assess-
ment of the site’s risks across the various affected areas, including the elevated risk of
contracting cancer (Gayer et al. 2000).
11 The range of values arises from differences in empirical specifications underlying the
estimates of costs and risk. Rohlfs argues credit constraints prevented many young men
from attending college during the period, causing the enrollment data to understate the
cost of the draft and the VSL of draft-age men.
12 One example of such an approach is Vassanadumrongdee and Matsuoka (2005). They
determine the VSL in Bangkok, Thailand using a survey instrument that describes the
current level of air pollution in the city as resulting in annual deaths of 430 in 1,000,000.
Individuals are then told that they can reduce their chance of dying from air pollution
by 30 in 1,000,000 (or 60 in 1,000,000) if they take an annual health exam that detects
respiratory system impairments. Each individual is randomly assigned one of four prices
and asked if they would be willing to pay the amount for the annual checkup. Based on
the hypothetical price that would need to be paid for the checkup and the reduction
in the chance of premature death from air pollution from taking the health exam, they
determine a VSL in Bangkok of about US$1 million.
13 In terms of Figure 9.2, market studies estimate the market wage equation W (π),
whereas contingent valuation studies estimate the preferences underlying the wage
equation. In equilibrium, the slope of the market wage function equals the marginal
benefit to the worker from a slightly safer job. The equality of the two values means it
Value of a statistical life 257
is unnecessary to estimate underlying preferences to determine appropriate VSLs, but
for more substantial improvements in safety market-determined values will overstate
people’s willingness to pay for greater safety.
14 Contingent valuation surveys are also more challenging to construct than other surveys
because many people have difficulty properly understanding complex probability con-
cepts and are unaccustomed to the notion of trading income for expanded safety. For
example, people generally understand the simple difference in payoffs between betting
on a race horse with 4:1 winning odds versus betting on a horse with 2:1 winning odds.
They might have a harder time, however, understanding whether or not they would sup-
port a government policy that reduces mercury levels in the local water supply and thus
decreases the risk of death for each constituent by 0.00001 at a cost of US$50 per tax-
payer. If respondents do not fully understand the probability concepts in surveys then
the VSL estimates from the survey may be inaccurate.
15 For further discussion see Carson et al. (2001) and Hausman (2012).
16 VSL estimates vary depending on both the set of control variables included in the esti-
mating wage or price equation and the underlying population examined (Viscusi 2013).
In the United States, VSL estimates are higher for union workers than for non-union
workers, higher for whites than blacks, and higher for women than men (Viscusi and
Aldy 2003; Viscusi 2003; Leeth and Ruser 2003). VSLs for native workers are roughly
the same size as for immigrant workers, except for non-English speaking immigrants
from Mexico who appear to earn little compensation for bearing very high levels of
workplace risk (Hersch and Viscusi 2010). VSLs rise as people age, at least until the
mid-40s, and then gradually decline (Kniesner et al. 2006; Aldy and Viscusi 2008). Not
surprisingly given that safety is generally considered to be a normal good, VSLs are
larger for higher income groups within a country at a point in time, within a country
over time as the country’s income increases, and within developed countries than within
less developed countries (Mrozek and Taylor 2002; Viscusi and Aldy 2003; Costa and
Kahn 2004; Bellavance et al. 2009; Kniesner et al. 2010). VSLs can vary across pop-
ulations because of differences in risk preferences. Groups more willing to take risks
will locate further to the right along the market wage function, and if the wage function
is concave from below as shown in Figure 9.2, more risk-tolerant groups will have a
smaller reduction in wage for a given increase in safety and a lower VSL. Alternatively,
some populations may have a lower VSL because they are less careful than others and
their lower safety productivity causes the wage function they face to lie below and to
be flatter (resulting in a lower VSL) than the wage function faced by others. The lower
ability of some workers to produce safety makes it desirable for employers to offer them
smaller wage premiums for accepting risk. Discrimination may also result in two sepa-
rate market wage functions, with the disadvantaged group facing not only lower wages
at every level of risk but also smaller increases in wages for given changes in risk than
for the advantaged group.
17 A difficulty with meta-analyses is that they include results from studies with known
problems, which “imparts biases of unknown magnitude and direction” (Viscusi 2009,
p. 118).
18 The authors devote particular attention to measurement errors, which have been noted in
Black and Kniesner (2003), Ashenfelter and Greenstone (2004), and Ashenfelter (2006).
Although they do not have information on subjective risk beliefs, they use very detailed
data on objective risk measures and consider the possibility that workers are driven by
risk expectations. Published industry risk beliefs are strongly correlated with subjective
risk values, and they follow the standard practice of matching to workers in the sample
an objective risk measure. Where Kniesner et al. (2012) differ from most previous stud-
ies is the pertinence of the risk data to the worker’s particular job, and theirs is the first
study to account for the variation of the more pertinent risk level within the context of a
panel data study. Their work also distinguishes job movers from job stayers. They find
258 T. J. Kniesner, J. D. Leeth, R. S. Sullivan
that most of the variation in risk and most of the evidence of positive VSLs stems from
people changing jobs across occupations or industries rather than from variation in risk
levels over time in a given job setting.
19 Rohlfs and Sullivan (2013a) used a variety of publicly available datasets from the
Department of Defense (DoD) in their empirical research. They obtained vehicle data
from the Theater A portion of US Army Materiel Systems Analysis Activity’s Sam-
ple Data Collection (US AMSAA, SDC, 2010) and cost information from cost experts
within the Army Material Command (AMC) and Cost Analysis and Program Evalua-
tion (OSD-CAPE) departments in DoD. Some of the data used in Rohlfs and Sullivan
(2013a) are considered For Official Use Only (FOUO). For another author to use the
data requires that person to have an official defense-related government purpose. Rohlfs
and Sullivan have made arrangements with the US Naval Postgraduate School (NPS) to
sponsor potential replicators of their project, so that replicators could obtain access to
the data.
20 All prices presented in the replacement vehicle case study are in 2010 dollars.
21 This is most likely an underestimate of the actual cost of the program because it does
not include many costs such as maintenance and fuel. It should also be noted that the
US$50 billion figure is for all Services, not just the Army, which was the focus of the
study by Rohlfs and Sullivan (2013a).
22 The three cost estimates use 484 miles per month as a constant rate of usage, which is
the rate observed for the average vehicle in the SDC data.
23 See Tables I–IV in Rohlfs and Sullivan (2013a) for details on their findings.
24 Exceptions, however, could be made for particularly combat-intensive units to receive
heavily armored Type 3 TWVs depending on the combat environment and their specific
capability requirements.

References
Aldy, J. E., and W. K. Viscusi. 2008. “Adjusting the Value of a Statistical Life for Age and
Cohort Effects.” Review of Economics and Statistics 90: 573–581.
Ashenfelter, O. 2006. “Measuring the Value of a Statistical Life: Problems and Prospects.”
Economic Journal 116: C10–C23.
Ashenfelter, O., and M. Greenstone. 2004. “Estimating the Value of a Statistical Life: The
Importance of Omitted Variables and Publication Bias.” American Economic Association
Papers and Proceedings 94: 454–460.
Bellavance, Francois, Georges Dionne, and Martin Lebeau. 2009. “The Value of a Statis-
tical Life: a Meta-Analysis with a Mixed Effects Regression Model.” Journal of Health
Economics 28.2: 444–464.
Black, D. A., and T. J. Kniesner. 2003. “On the Measurement of Job Risk in Hedonic Wage
Models.” Journal of Risk and Uncertainty 27: 205–220.
Brown, J. N., and H. S. Rosen. 1982. “On the Estimation of Structural Hedonic Price
Models.” Econometrica 50: 765–768.
Carson, R. T., N. Flores, and N. F. Meade. 2001. “Contingent Valuation: Controversies and
Evidence.” Environmental and Resource Economics 19: 173–210.
Costa, D. L., and M. E. Kahn. 2004. “Changes in the Value of Life, 1940–1980.” Journal
of Risk and Uncertainty 29: 159–180.
Epple, D. 1987. “Hedonic Prices and Implicit Markets: Estimating Supply and Demand
Functions for Differentiated Products.” Journal of Political Economy 95: 59–80.
Gayer, T., J. T. Hamilton, and W. K. Viscusi. 2000. “Private Values of Risk Tradeoffs
at Superfund Sites: Housing Market Evidence on Learning about Risk.” Review of
Economics and Statistics 82: 439–451.
Value of a statistical life 259
Hausman, J. 2012. “Contingent Valuation: From Dubious to Hopeless.” Journal of Eco-
nomic Perspectives 26: 43–56.
Hersch, J., and W. K. Viscusi. 2010. “Immigrant Status and the Value of Statistical Life.”
Journal of Human Resources 45: 749–771.
Kahn, S., and K. Lang. 1988. “Efficient Estimation of Structural Hedonic Systems.”
International Economic Review 29: 157–166.
Kniesner, T. J., and J. D. Leeth. 1995. Simulating Workplace Safety Policy. Boston, MA:
Kluwer Academic Publishers.
——. 2004. “Data Mining Mining Data: MSHA Enforcement Efforts, Underground Coal
Mine Safety, and New Health Policy Implications.” Journal of Risk and Uncertainty 29:
83–111.
Kniesner, Thomas J., W. Kip Viscusi, and James P. Ziliak. 2006. “Life-Cycle Consumption
and the Age-Adjusted Value of Life.” Contributions in Economic Analysis & Policy 5.1.
——. 2010. “Policy Relevant Heterogeneity in the Value of Statistical Life: New Evidence
from Panel Data Quantile Regressions.” Journal of Risk and Uncertainty 40: 15–31.
Kniesner, T. J., W. K. Viscusi, C. Woock, and J. P. Ziliak. 2012. “The Value of a Statistical
Life: Evidence from Panel Data.” Review of Economics and Statistics 94: 74–87.
Leeth, J. D., and J. Ruser. 2003. “Compensating Wage Differentials for Fatal and Nonfatal
Injury Risk by Gender and Race.” Journal of Risk and Uncertainty 27: 257–277.
Mrozek, J. R., and L. O. Taylor. 2002. “What Determines the Value of Life? A Meta-
Analysis.” Journal of Policy Analysis and Management 21: 253–270.
Peltzman, S. 1975. “The Effects of Automobile Safety Regulation.” Journal of Political
Economy 83: 677–725.
Office of Management and Budget. 1992. Circular No. A-94 Revised. The White House.
October 29. Available at www.whitehouse.gov/omb/circulars_a094 [last accessed
November 18, 2013].
Rohlfs, C. 2012. “The Economic Cost of Conscription and an Upper Bound on the Value
of a Statistical Life: Hedonic Estimates from Two Margins of Response to the Vietnam
Draft.” Journal of Benefit-Cost Analysis 3: 1–37.
Rohlfs, C., and R. Sullivan. 2013a. “The Cost-Effectiveness of Armored Tactical Wheeled
Vehicles for Overseas US Army Operations.” Defence and Peace Economics 24: 293–316.
Rohlfs, C., and R. Sullivan. 2013b. “A Comment on Evaluating the Cost-Effectiveness of
Armored Tactical Wheeled Vehicles.” Defence and Peace Economics 24: 485–494.
US Army Material Systems Analysis Activity. 2010. Theater A Sample Data Collection.
US Defense Manpower Data Center. 2008. Unit Master File. Data Request System number
23242.
US Environmental Protection Agency. 2010a. Guidelines for Preparing Economic
Analyses. Available at http://yosemite.epa.gov/ee/epa/eerm.nsf/vwAN/EE-0568-50.pdf/
US$file/EE-0568-50.pdf [last accessed November 18, 2013].
——. 2010b. Valuing Mortality Risk Reductions for Environmental Policy: A
White Paper. Available at http://yosemite.epa.gov/ee/epa/eerm.nsf/vwAN/EE-0563-
1.pdf/US$file/EE-0563-1.pdf [last accessed November 18, 2013].
US Department of Transportation. 2008. Revised Departmental Guidance 2013: Treat-
ment of the Value of Preventing Fatalities and Injuries in Preparing Economic Analyses.
Available at www.dot.gov/sites/dot.dev/files/docs/VSL%20Guidance%202013.pdf [last
accessed November 18, 2013].
Vassanadumrongdee, S., and S. Matsuoka. 2005. “Risk Perceptions and Value of a Statis-
tical Life for Air Pollution and Traffic Accidents: Evidence from Bangkok, Thailand.”
Journal of Risk and Uncertainty 30: 261–287.
260 T. J. Kniesner, J. D. Leeth, R. S. Sullivan
Viscusi, W. K. 2003. ‘Racial Differences in Labor Market Values of a Statistical Life’.
Journal of Risk and Uncertainty 27: 239–256.
——. 2004. “The Value of Life: Estimates with Risks by Occupation and Industry.”
Economic Inquiry 42: 29–48.
——. 2009. “The Devaluation of Life.” Regulation and Governance 3: 103–127.
——. 2013. “Estimating the Value of a Statistical Life Using Census of Fatal
Occupational Injuries (CFOI) Data.” Monthly Labor Review, October. Avail-
able at http://www.bls.gov/opub/mlr/2013/article/using-data-from-the-census-of-fatal-
occupational-injuries-to-estimate-the-1.htm [last accessed November 18, 2013].
Viscusi, W. K., and J. Aldy. 2003. “The Value of a Statistical Life: A Critical Review of
Market Estimates throughout the World.” Journal of Risk and Uncertainty 27: 5–76.
Part IV

New approaches to
military cost–benefit
analysis
This page intentionally left blank
10 The role of cost-effectiveness
analysis in allocating
defense resources
Kent D. Wall, Charles J. LaCivita, and
Anke Richter

10.1 Introduction
Defense resource allocation is a multi-level resource allocation process with multi-
ple levels within the organization making budget requests and granting approvals.
Not only are there multiple levels within a defense organization, but the resource
allocation process may iterate between the levels. For example, in the US PPBES
system, each of the services presents its desired set of capabilities (weapons
systems, personnel, new acquisitions, maintenance and readiness requirements,
training, research and development, etc.) requiring a certain portion of the total
Department of Defense (DoD) budget to the Office of the Secretary of Defense
(OSD). OSD is then responsible for reviewing these submissions, combining the
requested needs and desires into an overall proposal that will become the total
requested defense budget and which will be included in the President’s budget
request to Congress. Given system incentives, the requests always exceed the
expected available budget and OSD must decide which mix of capabilities will
provide the strongest defense force for the nation within the budgetary limitations.
As they decide to allocate different levels of funds to the military services, the
services have the possibility to adjust their submission packages of capabilities.
Each service strives for an optimal allocation among capabilities within its organi-
zation given a specific budget allocation. OSD strives for an optimal allocation of
budget among the services to attain what it views as the best mix of capabilities.
The process iterates among these two decision-makers (and two decision-making
levels) until the services and OSD are in agreement on the capabilities that will be
provided at a given allocation of the budget. (For an explanation and documenta-
tion of the PPBS process, please refer to the websites and documents provided at
the end of the references.) Making the process more complex, different types and
amounts of information are available to these two levels of decision-makers. Oper-
ations research, especially cost-effectiveness analysis (CEA), can provide valuable
insight to this iterative programming process.
While the preceding paragraph emphasizes the interactions seen at the top level
of the PPBES system, similar situations arise throughout the defense organiza-
tion. At lower levels, there is competition among capability providers in a service
to be included in the service’s budget proposal to OSD; for example, the armored
264 K. D. Wall, C. J. LaCivita, A. Richter
division and the infantry division in the Army; the infantry division and the field
artillery in the Army; the surface navy and aircraft carriers in the Navy; or infor-
mation operations and the surface navy in the Navy. At the higher level, this occurs
when the DoD budget is presented to Congress as part of the overall budget for
the country.
The situation is also seen outside of defense. Excellent examples are found
in multi-national aid programs. For example, the majority of HIV/AIDS inter-
vention programs implemented in Africa are funded by external sources (e.g.
donor governments; UN agencies; the Global Fund to Fight AIDS, Tubercu-
losis, and Malaria; and the World Bank). Each of these sources chooses to
allocate funding to individual countries based on their proposed interventions.
The individual countries then allocate these funds to local organizations and pro-
grams to conduct these interventions. Finally, the individual organization tasked
with implementing the interventions may make further allocation decisions to
spend monies on prevention and treatment efforts with individual populations
or geographical regions. Lasry et al. (2007) provides more information for these
applications.
Normative or prescriptive theory tells us that decision-makers ultimately must
choose among alternatives based on their benefits and their costs. For decision-
makers in the public sector benefits are most often expressed in terms of effective-
ness because of the difficulties in “pricing out” the “products” associated with
each alternative. National defense and homeland security are prime examples.
There are no markets for placing monetary value on the “products” associated
with an aircraft carrier battle group, national missile defense system or counter
terrorism policy. The decision-maker somehow must integrate information on
cost and effectiveness and decision problems are focused on striking a balance
between these two attributes. Cost-benefit analysis (CBA) is replaced by CEA
and examination of tradeoffs between effectiveness and cost becomes a crucial
task for the decision-maker. Unfortunately the integration of cost and effective-
ness is not straightforward and there are many views on how it should be done
(see Willard 1998; Fowler and Cason 1998; Thomas and Sheldon 1999; and the
excellent literature review by Ruefli 1974). This issue is even further compounded
and complicated in a multi-level decision process.
Mathematical programming techniques can provide vital insight into these
types of optimization problems. For any optimization problem (maximize effec-
tiveness subject to budget constraints) it is possible to develop the dual problem
formulation (minimize the shadow prices—the value of each constraint—subject
to desired levels of effectiveness). With basic convexity assumptions, the opti-
mal solution of the primal problem is an optimal solution of the dual problem as
well. This relationship has been extensively explored in the operations research
literature (see for example Bazaraa and Shetty 1979; Hillier and Lieberman 1990;
Luenberger 1973; Solow 1984) and it has been shown how the problem can be
solved either by working with the primal formulation or the dual formulation.
Frequently, one formulation is much easier to solve than the other. This is indeed
the case when considering the multi-level resource allocation process as it can
Defense resources management 265
take advantage of (rather than be hindered by) the different information available
at each decision-making level.
This chapter explores these two potential solution techniques in a multi-level
resource allocation problem. The goal of the resource allocation is to maximize
the overall effectiveness in a public sector organization through the allocation of a
single limited resource (the budget). We begin with a formulation of the resource
allocation decision problem in its primal form and explore practical aspects of its
solution, paying special attention to information requirements. The next section
presents the dual formulation of the problem and its solution. The practical imple-
mentation of the dual solution and its information requirements are compared with
those of the primal solution. The third section considers the resource allocation
problem from the perspective of CEA. Here we show that a decision-maker who
claims to be motivated by cost-effectiveness concerns is actually implementing
the dual solution. The desire to find “the best tradeoff between cost and effec-
tiveness” is, in reality, a desire to solve the resource allocation problem via the
dual formulation. Two illustrative examples are provided, one involving contin-
uous alternatives and one involving discrete alternatives, before we end with a
summary of our main points.

10.2 The primal decision problem


The public sector organization is viewed as a maximizer of overall effectiveness
subject to a resource constraint. This objective is pursued by allocation of resources
“downward” through its hierarchical structure. To simplify the exposition and fix
ideas, we consider a single resource (money) and the simple two-level hierarchy
depicted in Figure 10.1. The first level is occupied by one unit, the top-level decision-
maker or manager, who decides how to allocate the resource to the units on the lower
level. The second-level units are themselves decision-makers—they decide how best
to use the allocated resources given to them by the upper level.
For example, in the DoD, the Secretary of Defense would be the first-level
decision-maker who is responsible for allocating the defense budget to the second-
level decision-makers, the military services. In this manner, in Figure 10.1 DM0
is the Office of the Secretary of Defense, DM1.1 is the Army, DM1.2 is the Navy,
DM1.3 is the Air Force, etc. The military services are themselves decision-makers
as they choose how best to allocate the funds within their organizations to provide
the greatest amount of military capability.

DM0

DM1.1 DM1.2 ... DM1.N

Figure 10.1 The two-level hierarchy.


266 K. D. Wall, C. J. LaCivita, A. Richter
A mathematical formulation of the first-level decision-maker’s problem is easy.
Let V denote the overall effectiveness and let x i (i = 1, 2, . . . , N) denote the funds
allocated to the i th second-level unit. The decision-maker then desires to:

max V (x) (10.1)


x

subject to

x i = B, (10.2)

where x = [x 1 , x 2 , . . . x n ] is the vector of budget of allocations and B is the avail-


able budget. The decision-maker’s decision behavior is modeled as the solution
to a mathematical programming problem in primal form. To solve this system
optimally, the first-level decision-maker must be all-knowing, all-seeing, and in
possession of prodigious computation capability. His optimal choice is determined
by the first-order necessary conditions for the solution

∂ V (x∗ )
− λ = 0; i = 1, 2, . . ., N, (10.3)
∂ xi

where x∗ is the vector of optimal budget allocations and λ is the Lagrange mul-
tiplier associated with the budget, i.e. the marginal effectiveness of another unit
of budget. The standard prescriptive result is clear. The first-level decision-maker
should select an allocation such that marginal effectiveness across all second-level
units is equal.
The main issue with this formulation is the amount and type of information that
is available at the different levels of an organization. The “all-knowing” and “all-
seeing” decision-maker necessary to formulate and solve Equation (10.3) does
not exist. At the first level, the decision-maker has a strategic view (a joint view)
over all of the lower levels, the capabilities they produce, and how the entire
organization and portfolio of capabilities interact and contribute to the overall
effectiveness of the defense force. This level of decision-maker is less knowl-
edgeable of the operational details; exactly how this capability is best produced.
That information resides with the second-level decision-makers (in our example,
the military services). They have all of the operational knowledge, which is why,
when given a budget allocation, they will make a further allocation that enables
them to best produce the needed capabilities. The second-level decision-makers
will typically have less of a strategic/joint view. These different levels of informa-
tion can be captured in the mathematical formulation of the problem presented in
Equations (10.1)–(10.3).
Let yi represent the units of output produced by each second-level decision-
maker if it is given an allocation x i . This function, yi (x i ) is known only to the
second-level decision-makers. The first-level decision-maker knows the x i and can
observe the yi but does not know the details of the production function, i.e. the
first-level decision-maker does not know how yi would change with a slightly
Defense resources management 267
larger or slightly smaller allocation, x i . The first-level decision-maker assesses and
values the yi and their relative strategic contribution to the overall effectiveness of
the organization. Using the concepts of multi-criteria decision-making (see Belton
and Stewart 2003; Kirkwood 1997), we can create an approximation for V (x)
representing the first-level decision-maker’s decision process.
Let vi express the decision-maker’s valuation of the contribution yi makes to
overall effectiveness. For example, if x i dollars are allocated to the Air Force for
anti-aircraft defense then yi interceptor squadrons can be purchased and operated
over the time corresponding to the planning period. The first-level decision-maker
values this contribution towards anti-aircraft defense by assigning to the yi a num-
ber, vi (yi ). Let wi represent the relative importance of each measure to the overall
value function, V (x), as seen through the eyes of the first-level decision-maker:
in this case, the relative importance of anti-aircraft defense versus other military
capabilities to overall national defense. Thus we assume the decision-maker bases
allocation decisions on the maximization of V (x) expressed as

V (x) w1 v1 (y1 ) + w2 v2 (y2 ) + · · · + w N v N (y N ).

Given this approximation, the first-level decision-maker is confronted with the


problem of selecting x so as to

max{w1 v1 (y1 ) + w2 v2 (y2 ) + · · · + w N v N (y N )}, (10.4)


x

subject to Equation (10.2) and yi = yi (x i ). The theory of constrained optimization


again gives the answer. The manager behaves so as to produce the maximiz-
ing budget allocation, x i∗ (i = 1, 2, . . . , N), that obeys the first-order necessary
conditions:
dvi d yi ∗
wi (x ) − λ = 0; i = 1, 2, . . . , N. (10.5)
d yi d x i i

These equations, together with the budget constraint, provide enough information
to determine the best allocations and the value for the Lagrange multiplier.
While this formulation captures the information differential between the lev-
els of decision-makers, it is not an easier problem to solve than what was
presented originally. To find the optimal solution requires the calculation of
the N-dimensional system defined by Equation (10.5). Second, the first-level
decision-maker requires complete knowledge of the yi (x i ) and the marginal prod-
ucts, d yi (x ik )/d x i . The decision-maker must be both knowledgeable and under-
standing of this detailed information and be able to use it to determine individual
second-level individual marginal effectiveness. This is counter to our assump-
tion that the first-level decision-maker does not know the operational details of
the second-level decision-maker’s organization. Finally, if Equation (10.5) is not
satisfied, the first-level decision-maker is given little direction on how to re-
allocate funds to satisfy the first-order necessary conditions. A sensible guide
268 K. D. Wall, C. J. LaCivita, A. Richter
might be to allocate more funds to the sub-units that display higher marginal
effectiveness and to allocate less funds to those sub-units with low marginal
effectiveness. This, however, still requires the first-level decision-maker to be in
possession of the detailed knowledge assumed to reside only with the second-level
decision-makers.
The only way to approach this problem is through trial and error in an iter-
ative approach. Given an initial allocation, x0 , satisfying Equation (10.2), the
top-level decision-maker then observes the resulting {yi ; 1 ≤ i ≤ N} and evaluates
Equation (10.4). Using the portion of Equation (10.5) which the top-level decision-
dv
maker knows, wi dyi , he guesses what an improved allocation of x might be. The
i
decision-maker does not know for a fact that his choice will indeed lead to a bet-
ter allocation because he cannot estimate the second portion of Equation (10.5),
(x i∗ ), since only the second-level decision-maker has this information. The new
dyi
dx i
allocation is implemented and the first-level decision-maker observes the new
{yi ; 1 ≤ i ≤ N} and evaluates Equation (10.4). This continues until an allocation is
found for which Equation (10.5) is satisfied or when changes to Equation (10.4)
become sufficiently small.
This entire concept can be explained graphically using a 2-dimensional exam-
ple as in Figure 10.2. In the case of perfect, complete information with a single
decision-maker, the iterative solution process to find the optimal allocation can be
detailed on this graphic. The budget constraint, x 1 + x 2 = B, appears in the fourth
quadrant. The first and third quadrants depict the marginal effectiveness of each x i
assuming an additive form; i.e.,

∂V dvi d yi
Di V = = (x i ). (10.6)
∂ x i d yi d x i

D1V
slope =wfw 1

Quadrant II Quadrant I

vv'0 ·- ------

vl-------~-:r-: __·-+:- ::::::j::::::-;----


D1V D1V
D1V
D1V
D1V
D1V
D1V
D1V
=wfw
slope =wfw
=wfw
1 1 1 D1V
D1V
D1V
Quadrant Ill Quadrant IV
D1V

Figure 10.2 Two-dimensional example of primal solution.


Defense resources management 269
The second quadrant depicts the relationship between the allocation chosen by
the decision-maker and the Lagrange multiplier. The decision-maker begins by
specifying x 10 + x 20 = B. This allocation generates yi0 = yi (x i0 ) and the decision-
maker uses this and Equation (10.6) to compute the marginal effectiveness of each
sub-unit. Since D1 V > D2 V the first-level decision-maker decides on x 11 > x 10 and
x 21 < x 20 such that x 11 + x 21 = B. The cycle repeats, and eventually an allocation is
found for which 0 ≤ V − V k ≤ . Now in the case of a multi-level resource allo-
cation with differential information, the functions for the marginal effectiveness
of each x i , Equation (10.6) are not know to the first-level decision-maker. Hence
quadrants I and III are blank and there is no guidance for the decision-maker’s
search along the budget line.

10.3 The dual decision problem


The primal formulation of the multi-level resource allocation problem is very
difficult to solve directly. Another approach is to see if it is easier to solve the
dual formulation of the decision problem. The dual formulation of the first-level
decision-maker’s problem posed by Equations (10.2) and (10.4) is (see Luenberger
1973, 318–320; or Everett 1963)
 
min ϕ(λ) = min max [wi vi (yi (x i )) − λx i ] + λB , (10.7)
λ λ x

where

ϕ(λ) = max {wi vi (yi (x i )) − λx i } + λB (10.8)
x

is called the dual function. The dual formulation represents the decision problem
confronting the first-level decision-maker as a minimization. The decision-maker
seeks that value of λ for which the dual function attains its minimum.
The dual function is composed of N separate unconstrained, one-dimensional,
maximization problems that represent the decision problems faced by each of the
second-level decision-makers. We find it convenient to exchange the summation
and maximization so we may define this function in terms of a weighted sum of
individual dual functions, ϕi (λ),
 
ϕ(λ) = max{wi vi (yi (x i )) − λx i } + λB = ϕi (λ) + λB. (10.9)
xi

While the second-level decision-makers’ problem is

max{wi vi (yi (x i )) − λx i }, (10.10)


xi

this equation is not solved in this form because its solution, x i∗ , satisfying the
first-order condition
dvi d yi ∗
(x ) = λ/wi ; i = 1, 2, . . . , N
d yi d x i i
270 K. D. Wall, C. J. LaCivita, A. Richter
requires knowledge of dvi /d yi , wi and λ. These are known only to the first-level
decision-maker.
The second-level decision-maker knows only yi (x i ) and d yi /d x i . However, we
can define the decision problem of the second-level decision-maker as

max{yi (x i ) − λi x i } (10.11)
xi

given λi where
 
dvi −1
λi = λ wi = λm −1
i . (10.12)
d yi

This λi is provided to the second-level decision-maker by the first-level decision-


maker. In essence, λi is the marginal price that must be paid by the second-level
decision-maker to produce more effectiveness. It is inversely proportional to the
preferences of the first-level decision-maker as expressed by: (1) the marginal
productivity, dvi /d yi ; and (2) the tradeoff weight, wi . If m i is larger (smaller)
than m j then λi will be smaller (larger) than λ j . This marginal price captures the
logic that in order for one second-level decision-maker to produce more capabil-
ity, he is taking budget away from something else the first-level decision-maker
wants to obtain from another second-level unit. Hence λi is different across i , the
set of second-level decision-makers. Also note that this problem is completely
equivalent to that of Equation (10.10) because it has exactly the same first-order
condition and, hence, solution x i∗ .
The significant difference is the information set in possession of the second-
level decision-maker. This decision-maker knows only yi (x i ), d yi /d x i . The
Lagrange multiplier, λi , the price to be paid for additional effectiveness of the
i th unit, is given to the second-level decision-makers by the first-level decision-
maker. The second-level decision-maker does NOT have explicit knowledge of λ,
wi or dvi /d yi .
The resulting set of decisions {x i∗ (λi ); 1 = 1, 2, . . . , N} produce the set
{ϕi (λi ); 1 = 1, 2, . . . , N} that are aggregated by the first-level decision-maker
according to Equation (10.9) to define the dual function ϕ(λ) = ϕ(x∗ (λ)). Note:
at the top-level {wi ; dvi /d yi ; 1 = 1, 2, . . . , N} are considered fixed and are left
implicit for simplicity of notation.
Given the dual function, the global solution to Equation (10.7) is achieved by
finding that value of λ satisfying

∇ϕ(λ∗ ) = 0, (10.13)

where ϕ(λ∗ ) = ϕ(x∗ (λ∗ )). The most striking aspect of the dual formulation is the
fact that this gradient is the budget constraint itself (see Luenberger 1973, 313):

∇ϕ(λ) = x i∗ (λ) − B. (10.14)
Defense resources management 271
In other words, when λ = λ the resulting {x i∗ (λ∗ /wi ); 1 = 1, 2, . . ., N} satisfy the

budget constraint. The solution to a 1-dimensional problem provides the answer


to an N-dimensional allocation problem.
An iterative process involving information exchange between the first and
second-level decision-makers provides a practical solution. This exchange is a
price directive mechanism in which the top-level transmits downward to each sub-
unit its implicit marginal cost of effectiveness, λi . The sub-units respond by using
this parameter to solve their individual maximization problems and then transmit
upward their decisions, x i∗ (). The top-level decision-maker then checks to see if
Equation (10.13) is satisfied. This check is, however, nothing more than a check
on the budget constraint, Equation (10.14). If the aggregate decision is “over”
budget (“under”) then the top-level decision-maker increases (decreases) λ and
transmits back down “new” λi to each second-level decision-maker. This inter-
level exchange continues until no further change in the λi is required; i.e., the
budget constraint is met within the required precision: |∇ϕ(λ)| ≤ δ for some small
number δ.
The key parameter is, of course, the Lagrange multiplier, λ. It is the shadow
price of the allocated asset; i.e. the value of one unit of V in terms of one unit
of B. Equation (10.12) shows that the individual λi are a function of λ and the
marginal effectiveness of the i th sub-unit in the mind of the first-level decision-
maker. This means that the λi transmitted to the lower-level decision-makers
will vary according to the preferences of the first-level decision-maker. Different
Lagrange multipliers are transmitted to each second-level unit. Thus, if m i > m j
then λi < λ j and the i th second-level decision-maker will receive a lower “price”
for effectiveness than the j th second-level decision-maker.
The search for the optimal value of λ takes the form of an iterative scheme using
a simple adjustment mechanism:

λk+1 = α k λk .


For each iteration we now replace x i (λ) with x ik (λk ) and λi with λki . Given
 k x i k(λ )
k k

we check the budget constraint. If x i (λ ) − B < 0 set α < 1. If


k k k
x i (λ ) −
B > 0 set α k > 1 (hence λk+1 > λk when the budget is in deficit and λk+1 < λk
when the budget is in surplus). Recompute the different shadow prices for each
sub-unit, λk+1
i = m −1
i λ
k+1
. Set k = k + 1 and transmit the new “prices” downward
to the second level and receive the “new” x ik (λk ). Go to the budget check and
repeat until convergence to a satisficing solution is obtained; i.e., |∇ϕ(λk )| ≤ δ.
Figure 10.3 provides a two-dimensional illustration using the same set-up
employed in illustrating the primal problem solution. The first, third and fourth
quadrants of the graph are exact replicas of Figure 10.2. The second quadrant still
depicts the relationship between the values of λ and the marginal effectiveness
of each sub-unit. Now, however, the values of λ are those chosen by the first-
level decision-maker. These, by definition, all lie on a line with slope equal to
w2 /w1 . After specifying an initial value, λ0 , the first-level decision-maker gives
each second-level decision-maker his Lagrange multiplier, λ0i . The second-level
272 K. D. Wall, C. J. LaCivita, A. Richter

slope =wfw 1

Quadrant II Quadrant I

Quadrant Ill
Quadrant IV

Figure 10.3 Two-dimensional example of dual solution.

decision-makers solve their maximization problems and transmit to the first-level


decision-maker their proposal, yi0 , and its cost, (x i∗ )0 . The corresponding budget,
B 0 , is observed to be far below the allowed amount so λ is reduced to λ1 . This
produces new Lagrange multipliers, λ1i , that are used by each subunit in another
round of proposals and costs. There is still a surplus so the Lagrange multiplier is
lowered further to λ2 . At this value a deficit is observed so the first-level decision-
maker raises the Lagrange multiplier to λ3 . Now the budget is almost completely
consumed. Another adjustment of λ could be made but the first-level decision-
maker may deem the current situation “good enough” and terminate adjustment of
λk when 0 ≤ |B − B k | ≤ δ.
A graphical comparison of Figures 10.2 and 10.3 shows that the primal
approach always satisfies the budget constraint but requires iteration to seek sat-
isfaction of Equation (10.5). The dual solution always satisfies Equation (10.5)
but requires iteration to satisfy the budget constraint. Secondly, the dual problem
requires a search over only one-dimension whereas the primal problem requires
an N-dimensional search. Finally, the gradient of the dual solution is known
and is easy to use to guide the search for λ∗ . We argue that it is far easier
to check for satisfaction of Equation (10.2) and iterate with the aid of Equa-
tion (10.14) than it is to attempt to satisfy the traditional first-order necessary
condition.

10.4 Maximizing overall cost-effectiveness


Let us examine the problem formulation when individuals claim that decisions are
made on the basis of overall cost-effectiveness. These decisions express a set of
preferences that mimic those in standard microeconomic
 A theory: allocation x A is
preferred to x if either V (x ) = V (x ) while
B A B
x i < x i or V (x A ) > V (x B )
B
Defense resources management 273
 
while =
x iA x iB .
If two alternatives have the same overall effectiveness then
the less costly alternative is said to be more cost-effective. Likewise, if two alter-
natives incur the same cost then the alternative with more overall effectiveness is
said to be more cost-effective. If both costs and effectiveness vary, the decision-
maker searches for that value of x that maximizes the overall cost-effectiveness
measure, J :

J (x) = ω E V (x) − ωC xi . (10.15)

Overall “cost-effectiveness,” J , is a weighted difference between the effectiveness,


V , and the cost. Here ω E and ωC are non-negative weights, summing to one, that
express the decision-maker’s preference for maximizing effectiveness relative to
minimizing cost. A decision-maker who bases choice on maximizing overall cost-
effectiveness will select that alternative representing the “best” solution to a simple
two-attribute decision problem Equation (10.15) derived from the pursuit of two
overriding, generally conflicting objectives: (1) the maximization of effectiveness;
and (2) the minimization of cost. The solution is, of course, a function of the two
key parameters, ω E and ωC , representing the decision-maker’s preferences. The
values set for these will depend on the budget constraint.
The resource allocation problem is now viewed as an unconstrained maximiza-
tion problem:

max ω E [w1 v1 (y1 (x 1 )) + w2 v2 (y2 (x 2 )) + · · · + w N v N (y N (x N ))]
x
  
− ωC xi

or
 
max [ω E wi vi (yi (x i )) − ωC x i ] .
x

The solution, as before, is determined by the first-order necessary conditions:

dvi d yi (x i∗ ) 1 ωC
− = 0; 1≤i ≤ N
d yi d x i wi ω E

or

d yi (x i∗ ) ωC
− m −1 = 0; 1≤i ≤ N (10.16)
d xi i
ωE

where

dvi
m i = wi .
d yi
274 K. D. Wall, C. J. LaCivita, A. Richter
This solution is identical to that obtained in the dual formulation if we define
ωC /ω E = λ. Thus, the maximization of overall cost-effectiveness is identi-
cal to the dual function maximization problem given by Equation (10.8):
x i∗ (m −1 ∗ ∗ ∗
i λ ) = x i (λi ). The identical solution method as presented in Section 10.2
can be used.
We find it both reasonable and natural for decision-makers to rationalize deci-
sions by claiming to have found the “correct,” “best,” or “most appropriate”
tradeoff between cost and effectiveness. They are stating that they have found the
most effective solution given the opportunity cost of the resource (as captured by
the Lagrange multipliers). The dual problem iterative solution we have presented
is the basis for the communication between levels in a hierarchical organization.
The first-level decision-maker transmits downward the current opportunity cost
of effectiveness. In return the second-level decision-makers send the first level
their solutions: yi (x ik ) and x ik . The first-level decision-maker checks to
see if the
resource constraint is satisfied. If the constraint is satisfied (i.e., |B − x i | ≤ δ)
then a solution has been found. If not then the first-level decision-maker sends
revised opportunity cost information to the second level and the process repeats.
No marginal information is exchanged. The first-level decision-maker merely
checks the budget requests against the constraint and returns to the second-level
decision-makers new information on the tradeoff between effectiveness and cost.
All evaluations of the second-level decisions and how they integrate with the first-
level decision-maker’s overall tradeoff between effectiveness and cost are done
implicitly.

10.5 Illustrative examples


We present two examples to demonstrate the ability of a coordinated maximization
of overall cost-effectiveness in solving the multi-level resource allocation prob-
lem. The first example is the continuous case where there are an infinite number
of possible alternatives for each of the second-level decision-makers. This is the
situation illustrated in Figures 10.1 and 10.2. This is the world of continuous dif-
ferentiability, smooth curves, very “tidy” mathematics and a closed-form solution
of the second-level maximization of the cost-effectiveness function. The second
example assumes each second-level decision-maker only has a finite number of
possible alternatives to offer. This is the discrete case and is more representative
of many real-life situations. Here we no longer have a differentiable representation
for yi (x i ) and hence no closed-form solution. A solution must proceed by direct
search using function values; i.e., the numerical values of yi (x i ) proposed by the
second-level decision-makers. Our illustrations use a very simple adjustment pro-
cedure for the revision of the λ, which ensures that λ
first-level decision-maker’s
increases if x i > B and decreases if x i < B. In real-life this revision may
be more involved. However, even with our very basic implementation of the solu-
tion procedure, the process finds the correct answer in very few iterations. This
implies that the adjustment of λ need not be a complex procedure to be effective
and makes the entire process more usable.
Defense resources management 275
Both cases relate to the following scenario. Imagine the Ministry of Defense of
a small country using a planning system in which its force elements, both exist-
ing and prospective, are organized into programs. The country has a coastline
and land borders to it north, east and south. There are two major threat environ-
ments: (1) one originating with an aggressive regime in the country on its eastern
border; and (2) one to the west originating from across the sea. Suppose there
are two main programs, Program 1 for border security and Program 2 for coastal
security. The director of the joint planning staff is the first-level decision-maker
and the program managers of the two programs are the second-level decision-
makers. The director of the joint planning staff is given a budget, B, by the
Minister of Defense and must decide how to allocate these funds between the
two programs. The instructions from the Minister of Defense convey to the direc-
tor of joint planning a preference for addressing border security more than coastal
security problems. In addition, the director of joint planning knows that both pro-
grams will provide a decreasing marginal contribution to overall national military
effectiveness. The director believes, however, that Program 1 will have a higher
initial marginal rate of contribution to overall national military effectiveness than
does Program 2. We will assume these preferences are represented by the value
function:

V (x 1 , x 2 ) = 0.6[1 − e−2.95y1(x1 ) ] + 0.4[1 − e−2.10y2(x2 ) ]. (10.17)

Thus, the director of joint planning prefers maximization of border security


effectiveness to maximization of coastal security effectiveness by a ratio of 3:2.
Furthermore, the initial marginal rate of contribution to overall national military
effectiveness from y1 is 40 percent more than that from y2 .
The director of the joint planning staff requests proposals from the program
managers along with instructions indicating the cost (price) of effectiveness. Each
program staff then executes analyses and makes a recommendation. The recom-
mendations become requests for funds by each program office. The information
contained in V (x 1 , x 2 ) in Equation (10.17) is understood to reside completely
in the mind of the first-level decision-maker and is, therefore unknown by the
second-level decision-makers. The first-level decision-maker, however, provides
instructions to the second-level decision-makers, using the Lagrange multiplier, as
detailed in the examples that follow, that allow them to respond in a way consistent
with the preferences captured in V (x 1 , x 2 ).

10.5.1 The continuous case


Let Program 1 (Border Security) be described by the effectiveness production
function

y1 (x 1 ) = 1 − e−1.9x1
276 K. D. Wall, C. J. LaCivita, A. Richter
and let Program 2 (Coastal Security) be described by the effectiveness production
function

y2 (x 2 ) = 1 − e−0.9x2 .

This information resides internal to each program office and its staff. The director
of the joint planning staff does not know, nor chooses to know, these details. The
director possesses only the “strategic” information—that contained in V (x). The
planning process begins with the director setting a value for λ0 based on his/her
assessment of the given budget, B. If B represents a significant growth over pre-
vious budgets then λ0 will be smaller. This will allow the program managers to
propose more capabilities as the marginal cost of providing effectiveness is low.
Likewise, if B represents a significant reduction from past experience then λ0 will
be larger. This will force the program managers to reduce the amount of capabil-
ities they propose as the marginal cost of effectiveness is high. The director now
uses Equation (10.16) to construct the λi resulting in

λ1 = [0.6 · 2.95 · e−2.95y1 ]−1 λ0


λ2 = [0.4 · 2.10 · e−2.10y2 ]−1 λ0 ,

where we set yi equal to the previous proposed value given to the director by the
program managers. Initially, in the beginning of the current planning period, these
would be the result of the previous planning period solution.
We now illustrate the planning process derived from the overall cost-
effectiveness paradigm (based on the dual formulation) by simulating the process
using the following steps. Let a1 = 2.95, a2 = 2.1, c1 = 1.9, c2 = 0.9 and λ0 = 1/B.

STEP 1: (Initialization) Set k = 0, λk = λ0 , y1 = y2 = y.


e−ai
STEP 2: λki = [wi · a1 · ] λ = m −1
yi −1 k
i λ .
k

λki
STEP 3: x ik = −ci−1 n ci
.
STEP 4: B k = + x 1k x 2k .
 |B k−B| ≤ δ then STOP. Otherwise
k
STEP 5: Check for budget satisfaction.
 If
If B > B + δ then set λ = 1 + β ·
k k+1 B−B
B
λk and k = k + 1 and go to
STEP 2.   k 
If B k < B − δ then set λk+1 = 1 − β · B B−B λk and k = k + 1 and go to
STEP 2.

Step 2 represents the director’s determination of each program managers’s


opportunity cost of effectiveness using the director’s tradeoff weights for over-
all national military effectiveness and an estimate of the marginal contribution
to V (x) of each yi . We assume this estimate is the result of a simple mental
calculation: the evaluation of vi /yi = ai e−ai yi where yi = y = 0.001.
Step 3 represents the individual program managers and the processes by which
Defense resources management 277
they determine the most cost-effective proposal given their opportunity cost of
effectiveness defined by the director. In the case where yi (x i ) ∈ C 1 we always
have a closed form solution to the program manager’s maximization of the cost-
effectiveness function. This produces the x ik . Step 4 and 5 represent the director’s
budget check in terms of a satisficing decision rule (where η is the director’s per-
missible percentage budget error). In the following simulation experiments we use
δ = η · B where η = 0.02 while β = 1.0.
Figures 10.4, 10.5 and 10.6 depict simulations for three budgets: B = 2.5, 4.5,
and 6.5. Each simulation is presented in four panels: the top left shows λk , the top
right gives the decisions x 1k (solid line) and x 2k (dashed line), the lower left traces
V (xk ) and the lower right presents B k = x 1k + x 2k . Convergence to the solution {λ∗ ,
x∗1 , x∗2 } is asymptotic whenever V (xk ) is continuous in xk . Convergence is always
finite, however, when the decision maker employs a satisficing rule for the budget
like |B K − B| ≤ δ. In each case the larger is B; the smaller is λ∗ . When B = 2.5
we obtain convergence in two iterations and when B = 4.5 we obtain convergence
in two iterations. When B = 6.5 we obtain convergence in eight iterations. Rapid
convergence can always be achieved by increasing β. When β = 2.0, we get con-
vergence in three iterations, as in Figure 10.6. The convergence of λ is monotonic,
either from above or below, depending upon whether λ0 = 1/B results in λ0 > λ∗ or
λ0 < λ∗ .

10.5.2 The discrete case


The border security program manager and staff develop an efficient set of six
alternatives described in Table 10.1.
The coastal security manager and staff develop an efficient set of six alternatives
given in Table 10.2.
As in the previous example, all the data in the tables are internal to each program
office. The director of the joint planning staff is not in possession of these details
and delegates to the program managers all responsibility for developing the data,
executing the analyses and making recommendations.
The discrete case is both more representative of real-life resource allocation
problems and more difficult. First, there may be no allocation (x 1k , x 2k ) such that
B = B k = x 1k + x 2k , while V ((x 1k , x 2k ) ≥ V (x 1 , x 2 ) for all other budget feasible
allocations. There may be, however, several satisficing solutions; i.e., allocations
for which |B − B k | ≤ δ while V ((x 1k , x 2k ) ≥ V (x 1 , x 2 ). Any such alternatives are
satisfactory solutions. Second, the discrete environment does not admit a closed-
form solution for x i (λi ). Each program manager must perform a direct search
over his/her alternatives {yi j , x i j } to identify that x i∗j that maximizes yi j − λi x i j for
given λi . Here the search is over j ∈ Ni , the index set for the i th program manager.
Figure 10.7 presents the complete set of alternatives in this resource allocation
problem. There are 36 combinations of program proposals in cost-effectiveness
space representing the set of possibilities open to choice by the director. It must be
understood that this information is NOT known a priori by the director. The director
is only presented with proposals—one from each program office so is only aware
Effectiveness Lambda
0.0 0.2 0.4 0.6 0.8 1.0 0.00 0.05 0.10 0.15 0.20 0.25 0.30
0
0

2
2

4
4

Iteration
Iteration

Figure 10.4 Continuous case with B = 2.5.


6
6

8
8

Cost Budget Allocations X1 and X2


0 1 2 3 4 5 6 7 8 0 1 2 3 4 5

0
0

2
2

4
4

Iteration
Iteration
6

6
Budget = 2.50
8

8
Effectiveness Lambda
0.0 0.2 0.4 0.6 0.8 1.0 0.00 0.05 0.10 0.15 0.20 0.25 0.30
0
0

2
2

4
4

Iteration
Iteration

Figure 10.5 Continuous case with B = 4.5.


6
6

8
8

Cost Budget Allocations X1 and X2


0 1 2 3 4 5 6 7 8 0 1 2 3 4 5

0
0

2
2

4
4

Iteration
Iteration
6

6
Budget = 4.50
8

8
Effectiveness Lambda
0.0 0.2 0.4 0.6 0.8 1.0 0.00 0.05 0.10 0.15 0.20 0.25 0.30
0
0

2
2

4
4

Iteration
Iteration

Figure 10.6 Continuous case with B = 6.5.


6
6

8
8

Cost Budget Allocations X1 and X2


0 1 2 3 4 5 6 7 8 0 1 2 3 4 5

0
0

2
2

4
4

Iteration
Iteration
6

6
Budget = 6.50
8

8
Defense resources management 281
Table 10.1 Border security alternatives

Alternative x1 y1 (x1 )

1 0.50 0.15
2 0.75 0.40
3 0.81 0.49
4 1.50 0.68
5 1.79 0.75
6 3.50 0.85

Table 10.2 Coastal security alternatives

Alternative x2 y2 (x2 )

7 0.40 0.11
8 1.00 0.55
9 1.55 0.76
10 2.00 0.83
11 2.80 0.91
12 3.95 0.98

of the combinations as they are presented sequentially (by the program managers).
There are 20 clearly dominated alternatives—alternatives where there exist at least
one other alternative that is either less costly yet equally (or better) effective or
more effective yet equally (or less) costly. We see that for small B there are clearly
defined solutions but the envelop of the set of alternatives, the efficient set, becomes
relatively “flat” for B > 4.0 making the identification of efficient solutions difficult.
We present Figure 10.7 solely for the benefit of the reader to understand the challenge
of each example allocation problem presented below.
As in the previous case we have w1 = 0.6, w2 = 0.4, a1 = 2.95, and a2 = 2.1.
The simulation is composed of the following steps.

STEP 1: (Initialization) Set k = 0, λk = λ0 , y10 = y20 = δ.


0
STEP 2: λki = [wi ∗ a1 ∗ e−ai yi ]−1 λk = m −1
i λ .
k

STEP 3: x i = x i j = the solution of maxxi j {yi j − λki x i j }
k

STEP 4: B k = x 1k + x 2k .  
STEP 5: If |B k − B| ≤ δ then STOP, y1 (x 1k ), y2 (x 2k ) is a satisfactory solution.
Otherwise   k

If B k > B + δ then set λk+1 = 1 + β · B−B B
λk and k = k + 1 and go to
STEP 2.   k 
If B k < B − δ then set λk+1 = 1 − β · B B−B λk and k = k + 1 and go to
STEP 2.
282 K. D. Wall, C. J. LaCivita, A. Richter

1.0
0.9
Possible Alternatives in C–E Space
0.8
0.7
0.6
0.5
0.4
0.3

Budget = 4.00
0.2
0.1
0.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0 7.5 8.0
Iteration

Figure 10.7 Discrete case alternatives.

Steps 1, 2, 4, and 5 again represent the director and Step 3 the program man-
agers’ maximization of cost-effectiveness, which is now a direct search over their
respective finite set of solutions. Figures 10.8, 10.9, and 10.10 show the simulation
results for three budgets: B = 2.5, 4.5, and 6.5, respectively. As in the continuous
case, each simulation is presented in four panels: (1) the top left presents λk ; (2)
the top right presents the decisions x 1k (solid line) and x 2k (dashed line); (3) the
lower left presents V (x); and (4) B k = x 1k + x 2k occupies the lower right panel.
All simulations find a satisficing solution. Convergence depends on δ = ηB and
β. We use η = 0.01 and β = 0.45 for B = 2.5 and obtain convergence in three
iterations. We use η = 0.01 and β = 1.10 for B = 4.5 and obtain convergence
in three iterations. When B = 6.5 convergence is obtained in three iterations if
η ≥ 0.02 with β = 1.0 or η = 0.01 with β = 2.0.
The convergence indicated here compares favorably with the existing PPBES
process at the DoD. Undoubtedly, convergence to a satisficing solution within a
three or four level hierarchy would take longer to achieve.

10.6 Summary
The dual formulation of a multi-level public sector resource allocation problem
provides a sound theoretical basis for decision-makers who rationalize their deci-
sions with statements like “. . .the chosen alternative possesses the best balance
between cost and effectiveness.” They are expressing nothing less than the first-
order necessary conditions for a solution of the dual formulation of the problem.
The pivotal variable in such problems is the Lagrange multiplier, or “shadow
price”. This represents the marginal cost of effectiveness.
4

1.0
X2

0.8
3
and
X1

0.6
2

Lambda
0.4
1

0.2
Budget Allocations
0

0.0
0 2 4 6 8 1 3 5 7 9 11
Iteration Iteration

1.0
Budget = 2.50

0.8
6
5

0.6
4
Cost

0.4
3

Effectiveness
2

0.2
1
0

0.0
1 3 5 7 9 11 1 3 5 7 9 11
Iteration Iteration

Figure 10.8 Discrete case with B = 2.5.


4

1.0
X2

0.8
3
and
X1

0.6
2

Lambda
0.4
1

0.2
Budget Allocations
0

0.0
0 2 4 6 8 10 1 3 5 7 9 11
Iteration Iteration

1.0
Budget = 4.50

0.8
6
5

0.6
4
Cost

0.4
3

Effectiveness
2

0.2
1
0

0.0
1 3 5 7 9 11 1 3 5 7 9 11
Iteration Iteration

Figure 10.9 Discrete case with B = 4.5.


4

1.0
X2

0.8
3
and
X1

0.6
2

Lambda
0.4
1

0.2
Budget Allocations
0

0.0
0 2 4 6 8 10 1 3 5 7 9 11
Iteration Iteration

1.0
Budget = 6.50

0.8
6
5

0.6
4
Cost

0.4
3

Effectiveness
2

0.2
1
0

0.0
1 3 5 7 9 11 1 3 5 7 9 11
Iteration Iteration

Figure 10.10 Discrete case with B = 6.5.


286 K. D. Wall, C. J. LaCivita, A. Richter
Both the traditional primal formulation and our dual formulation require an iter-
ative procedure to obtain a solution. In the primal problem the budget constraint is
always satisfied and one iterates on the budget allocation until all marginal effec-
tiveness values are equal (to the optimum value of the Lagrange multiplier). In
the dual problem the marginal effectiveness measures are always equal and one
iterates on the Lagrange multiplier (the magnitude of the marginal effectiveness)
until the budget constraint is met.
Mathematically the approaches are equivalent, but behaviorally there is a sig-
nificant distinction. The primal solution requires the first-level decision-maker be
in possession of all marginal information and be able to solve an N-dimensional
search using this information. This conflicts with our assumptions on the dif-
ferential information available at each level of the organization. The first-level
decision-maker has neither the expertise nor the specialized and detailed oper-
ational information possessed by the second-level decision-makers. The dual
approach only requires the first-level decision-maker to know the values, yi and
x i . No marginal information is required, i.e. the first-level decision-maker does not
need to know how the capabilities proposed by the second-level decision-makers
will change based on a small change in the budget allocation. The first-level
decision-maker solves only a one-dimensional problem, searching for the value of
λ that produces satisfaction of the budget constraint, which is a much more man-
ageable task. The second-level decision-makers are delegated complete authority
to decide how to “spend” the resource, consistent with the opportunity cost trans-
mitted to them by the first-level decision-maker. We believe this is consistent with
the decentralization found in many organizations.
Of particular importance is the relative speed with which an optimal solution to
the multi-level resource allocation problem can be found implementing the dual
formulation. Very few iterations are needed to find acceptable results. The conver-
gence was rapid using an extremely basic adjustment mechanism for λ. In real life,
these adjustments can be made more adroitly using the first-level decision-maker’s
expertise and historical knowledge. This is important because in real applications,
with the usual budget and time constraints, only a limited number of iterations will
be possible.
We believe the dual formulation is the best way to solve a multi-level resource
allocation problem and is, in fact, the way individuals naturally approach these
types of problems when they wish to make cost-effective decisions.

Acknowledgement
This work benefits from the insights and interpretations of Assistant Professor David
L. Rose who has since passed away. He was an inspiring teacher and brilliant mind
that is fondly remembered, and sorely missed, by all who knew him.

Bibliography
Bazaraa, Mokhtar S. and C.M. Shetty. 1979. Nonlinear Programming: Theory and Algo-
rithms. Toronto, Canada: John Wiley and Sons.
Defense resources management 287
Belton, Valerie and T.J. Stewart. 2003. Multiple Criteria Decision Analysis: An Integrated
Approach. Norwell, MA: Duxbury Press.
Everett III, Hugh. 1963. “Generalized Lagrange Multiplier Method for Solving Problems
of Optimum Allocation of Resources.” Operations Research 11: 399–417.
Fowler, B.W. and P.P. Cason. 1998. “The Cost Exchange Ratio: A New Aggregate Measure
of Cost and Operational Effectiveness.” Journal of the Military Operations Research
Society 3: 57–64.
Hillier, Frederick S. and Gerald J. Lieberman. 1990. Introduction to Operations Research
(5th edition). New York: McGraw-Hill Publishing Company.
Kaplan, Stanely and B. John Garrick. 1981. “On the Quantitative Definition of Risk.” Risk
Analysis 1: 11–27.
Kirkwood, Craig. W. 1997. Strategic Decision Making: Multiobjective Decision Analysis
with Spreadsheets. Belmont, CA: Duxbury Press.
Lasry, A., G.S. Zaric, and M.W. Carter. 2007. “Multi-level Resource Allocation for HIV
Prevention: A Model for Developing Countries.” European Journal of Operational
Research 180: 786–799.
Libby, R. and P.C. Fishburn. 1977. “Behavioral Models of Risk Taking in Business
Decisions: A Survey and Evaluation.” Journal of Accounting Research 15: 272–292.
Luenberger, David G. 1973. Introduction to Linear and Nonlinear Programming. Reading,
MA: Addison-Wesley.
Payne, J.W., D.J. Laughhunn, and R. Crum. 1980. “Translation of Gambles and Aspiration
Level Effects in Risky Choice Behavior.” Management Science 10: 1039–1060.
Petty II, J.W. and D.F. Scott. 1981. “Capital Budgeting Practices in Large American
Firms: A Retrospective Analysis and Update.” In Financing Issues in Corporate Project
Selection, edited by F.G.J. Derkinderen and R.L. Crum. Boston, MA: Martinus Nyhoff.
Quade, E.S. 1989. Analysis for Public Decisions. Englewood Cliffs, NJ: Prentice Hall.
Ruefli, T. 1971. “PPBS – An Analytical Approach.” In Studies in Budgeting, edited by
R.F. Byrne, A. Charnes, W.W. Cooper, O.A. Davis, and D. Gilford. Amsterdam: North-
Holland.
Ruefli, T. 1974. “Analytic Models of Resource Allocation in Hierarchical Multi-Level
Systems.” Socio-Economic Planning Sciences 8: 353–363.
Solow, Daniel 1984. Linear Programming: An Introduction to Finite Improvement Algo-
rithms. New York: Elsevier Science Publishing Company.
Thomas, C. and R.S. Sheldon. 1999. “The ‘Knee of the Curve’ – Useful Clue but
Incomplete Support.” Journal of the Military Operations Research Society 4: 17–29.
Willard, D. 1998. “Cost-Effectiveness, CAIV, and the Knee of the Curve.” Phalanx 31:
24–27.

Documents and websites relating to the PPBS process


Bright, T. 2010. PPBE at the Pentagon: Changes Coming? Available at http://view.dau.mil/
mp_pbl/dauvideo/Active_Directory-CN_Admin___MP_OU_Service_Accounts_and_
Groups_OU_Ft_Belvoir_DC_DAU-ADS_DC_dau_DC_mil/MediaBin/Download/3614/
PPBE_at_the_Pentagon_FEB_2010_DAU_2-16-10.pdf [last accessed November 25,
2014].
Defense Acquisition University. 2012. Program Budget Decision (PBD) Cycle Oct.–Nov.
Available at https://dap.dau.mil/acquipedia/Pages/ArticleDetails.aspx?aid=d97f7732-
4fab-414b-b49b-75caa037e304 [last accessed November 25, 2014].
288 K. D. Wall, C. J. LaCivita, A. Richter
Defense Acquisition University. Defense Acquisition Portal. The Budgeting PPBE Phase.
Available at https://dap.dau.mil/aphome/ppbe/Pages/Budgeting.aspx [last accessed
November 25, 2014].
DoD. Directive Number 7045.14.2013. Available at www.dtic.mil/whs/directives/corres/
pdf/704514p.pdf [last accessed November 25, 2014].
Tomasini, R. Planning, Programming, Budgeting and Execution (PPBE) Process. Defense
Acquisition University. Available at http://www.google.com/url?sa=t&rct=j&q=&esrc=
s&source=web&cd=8&sqi=2&ved=0CE8QFjAH&url=https%3A%2F%2F2.zoppoz.workers.dev%3A443%2Fhttp%2Fwww.dau.mil
%2Fhomepage%2520documents%2FPPBE%2520Process%2520Brief%2C%2520with
%2520Carter%2520Efficiency%2520Initiatives %2C%2520Tomasini%2C%2520Dec%
252010.pptx&ei=MPI6U7bhIIKS2QWK2oHIBw&usg =AFQjCNEYOx6DNS_h9Ig1
AXVoyPJcvEwf_w&sig2=v195LBreynL0rB-O8b8gtA [last accessed November 25,
2014].
Travers, O. 2011. Department of Defense Planning, Programming, Budgeting and
Execution (PPBE) Process/Army Planning, Programming, Budgeting and Execution
(PPBE) Process: An Executive Primer. Available at https://www.documentcloud.org/
documents/293931-dodarmyppbeprimernov-2011.html [last accessed November 25,
2014].
11 A risk-based approach to
cost–benefit analysis
Strategic real options, Monte Carlo
simulation, knowledge value added,
and portfolio optimization
Johnathan C. Mun and Thomas Housel

11.1 Introduction
Games of chance have been popular throughout history. Biblical accounts even
record Roman soldiers gambling for Christ’s robes. In earlier times, chance was
something that occurred in nature, and humans were simply subjected to it as a
ship is to random waves. Until the Renaissance, the future was thought to be a
chance occurrence of completely random events, beyond the control of humans.
With the advent of games of chance the study of risk and uncertainty increas-
ingly mirrored real-life events. Although games of chance were initially played
with great enthusiasm, rarely did anyone consciously determine the odds. It was
not until the mid-1600s that the concept of chance was seriously studied: The
first such endeavor can be credited to Blaise Pascal, a father of the study of
choice, chance, and probability. Fortunately, after many centuries of mathematical
and statistical innovations by visionaries such as Pascal, Bernoulli, Bayes, Gauss,
Laplace, and Fermat, and with the explosion of computing technology, uncertainty
can increasingly be explained with much more elegance through rigorous, method-
ological applications. Today, with the assistance of powerful software packages,
we have the ability to apply these techniques to increase the value of cost–benefit
analysis (CBA), or what might be called “risk–benefit analysis.”
Humans have struggled with risk for centuries, but through trial and error and
through the evolution of human knowledge, have devised ways to describe, quan-
tify, hedge, and even take advantage of risk.1 In the US military, risk analysis, real
options analysis, and portfolio optimization techniques are enablers of a new way
of approaching problems of estimating return on investment (ROI) and estimating
risk-value tradeoffs of strategic real options.2
Several advanced quantitative risk-based concepts are introduced in this chap-
ter, including the application of strategic real options analysis, Monte Carlo risk
simulation, stochastic forecasting, portfolio optimization, and knowledge value
added (KVA). These methodologies rely on standard metrics and existing tech-
niques, such as ROI and discounted cash flow discussed in Chapter 17, and
complement these techniques by pushing the envelope of analytics while not
290 J. C. Mun, T. Housel
replacing them outright. These advanced techniques are used to help make the best
possible decisions, to allocate budgets, forecast outcomes, and create portfolios
with the highest strategic value or ROI (even when conditions surrounding these
decisions are risky or uncertain). These new techniques can be used to identify,
analyze, quantify, value, forecast, and manage (hedge, mitigate, optimize, allo-
cate, and diversify) risks associated with strategically important military options
(see Chapter 14 for example).

11.2 Why is risk important in making decisions?


Before we embark on a journey to review these advanced techniques, let us first
consider why risk is critical when making decisions. Risk is an important part
of any decision-making process. For instance, suppose projects are chosen based
simply on an evaluation of returns (i.e., monetary benefits from the project’s
output) alone, or costs alone; clearly, the higher-return and lower-cost project will
be chosen first over lower-return and higher-cost projects. Projects that produce
valuable outputs that offer higher returns also often bear higher risks. Projects with
significantly lower returns and higher risks will tend to be abandoned.
An exclusive focus on cost reduction can result in stifled innovation and creativ-
ity. The goal is not simply cost reduction to reduce risks. In this case, the simplest
approach is to fire everyone and sell off all the assets. The real question is: how do
costs compare to desired outputs/benefits—that is, “costs compared to what?”
To encourage a focus on improving processes and innovative technologies, a
new way of calculating ROI that includes a unique numerator is required. ROI
is a simple ratio that requires unique estimates of the numerator (i.e., monetary
value such as revenues) and the denominator (i.e., monetized costs and invest-
ments). ROI estimates, however, must be placed within a larger context and a
longer-term view that includes estimates of risk and the ability of management to
adapt as they observe the performance of their investments over time. Therefore,
instead of relying purely on immediate ROIs or costs, a project, strategy, new pro-
cess innovation, or new technology should be evaluated based on its total strategic
value—including updated returns, costs, and strategic options as well as its risks.
Figures 11.1 and 11.2 illustrate possible errors in judgment that could occur when
risks are ignored.
Figure 11.1 lists three mutually exclusive projects with their respective implemen-
tation costs, expected net returns (net of the costs to implement), and risk levels (all
in present values).3 Clearly, for the severely budget-constrained decision-maker, the
cheaper the project the better, resulting in the selection of Project X. Assuming the
budget is not a binding constraint, the returns-driven decision-maker will choose
Project Y with the highest benefits. Project Z will be chosen by the risk-averse
decision-maker as it provides the least amount of risk, while still providing a pos-
itive net return. The upshot is that, with three different projects and three different
decision-makers, three different decisions will be made. Who is correct and why?
Figure 11.2 demonstrates when Project Z should be chosen. For purposes of
illustration, suppose all three projects are independent and mutually exclusive,
Real options: a risk-based approach 291

Name of Project Cost Returns Risk


Project X $50 $50 $25
Project Y $250 $200 $200
Project Z $100 $100 $10

Project X for the cost and budget-constrained manager


Project Y for the returns driven and nonresource-constrained manager
Project Z for the risk-averse manager
Project Z for the smart manager

Figure 11.1 Why is risk important?

Looking at bang for the buck, X (2), Y (1), Z (10), Project Z should
be chosen–with a $1,000 budget, the following can be obtained:
Project X: 20 Project Xs returning $1,000, with $500 risk
Project Y: 4 Project Xs returning $800, with $800 risk
Project Z: 10 Project Xs returning $1,000, with $100 risk
Project X: For each $1 return, $0.5 risk is taken
Project Y: For each $1 return, $1.0 risk is taken
Project Z: For each $1 return, $0.1 risk is taken
Project X: For each $1 of risk taken, $2 return is obtained
Project Y: For each $1 of risk taken, $1 return is obtained
Project Z: For each $1 of risk taken, $10 return is obtained
Conclusion:
Risk is important. Forgoing risks results in making the wrong decision.

Figure 11.2 Adding an element of risk.

and that an unlimited number of projects from each category can be chosen (only
constrained by a budget of US$1,000). With this US$1,000 budget, 20 project Xs
can be chosen, yielding US$1,000 in net returns and US$500 risks, etc.
It is clear from Figure 11.2 that project Z is the best project: That is, for the same
net return (US$1,000), the least amount of risk is undertaken (US$100). Another
way of viewing this selection is that for each US$1 of risk, US$10 in returns are
obtained on average. This example illustrates the concept of bang for the buck or
getting the best value (net monetary benefits) with the least amount of risk.
An even more striking example is provided when there are several different
projects with identical single-point average net benefit or costs of US$10 million
each. Without risk analysis, a decision-maker should, in theory, be indifferent to
choosing any of the projects. With risk analysis, however, a better decision can be
made. For instance, suppose the first project has a 10 percent chance of exceeding
US$10 million; the second, a 15 percent chance; and the third, a 55 percent chance.
Then additional critical information is obtained on the riskiness and uncertainty
of the project or strategy, and a better decision can be made.
292 J. C. Mun, T. Housel
11.3 From traditional risk assessment to
Monte Carlo risk simulation
Military and business leaders have dealt with risk throughout the history of war
and commerce. In most cases, decision-makers viewed risks of a particular project,
acknowledged its existence, and moved on. Little quantification was performed. In
fact, most decision-makers only focus on single-point estimates of a project’s net
benefit or profitability. Figure 11.3 shows an example of a single-point estimate.4
The estimated net revenue of US$30 is simply that: a single-point estimate whose
probability of occurrence is close to zero.5
In the simplest model shown in Figure 11.3, the effects of interdependencies are
ignored, and, in modeling jargon, we have the problem of garbage-in, garbage-out
(GIGO). As an example of interdependencies, the units sold are probably negatively
correlated to the price of the product (since demand curves slope downwards), while
average variable costs may be positively or negatively correlated to units sold (i.e.,
depending on whether there are decreasing or increasing returns to scale in produc-
tion). Ignoring these effects in a single-point net revenue estimate with uncertain
sales will yield grossly misleading results. There are numerous interdependencies
in military options as well—for example, complex issues in logistics, beginning
with the supplier all the way through to the warrior in the field.
In the commercial example illustrated in Figure 11.3, if the unit sales variable
becomes 11 instead of 10, the resulting revenue may not simply be US$35. The
net revenue may actually decrease if there is an increase in variable cost-per-unit,
while the sale price may actually be slightly lower driving the increase in unit
sales. Ignoring these interdependencies will reduce the accuracy of the model.

Single Point Estimate




Unit Sales
Sales Price
Total Reve nue
;ot)
---
~
$100
lnterdependendes
rne::m GIGO

• Variable Cost/Unit _$5_


• Total Fixed Cost $20
• Total Cost $20
• Net Revenue $20

Estimate
Point Estimate

How confident are you of the analysis outcome?


This may be dead wrong!

Figure 11.3 Single-point estimates.


Source: Mun (2010).
Real options: a risk-based approach 293
One traditional approach used to deal with risk and uncertainty is the appli-
cation of scenario analysis. For example, scenario analysis is a central part of
the capabilities-based planning approach in widespread use for developing DoD
strategies (Mun 2010). In the single-point estimate example described above, sup-
pose three scenarios were generated: the worst-case, nominal-case, and best-case
scenarios. When different values are applied to the unit sales, the resulting three
scenarios’ net revenues are obtained. As discussed earlier, if problems of inter-
dependencies are not addressed, then the net revenue obtained in each case is
unreliable and not much confidence can be placed in the analysis.
In the military planning case, problems are exacerbated by the lack of objective
ways to estimate benefits in monetary units. Without a monetary benefits analy-
sis such as that illustrated in Chapter 8, it becomes difficult (if not impossible)
to compare the net benefits of various scenarios. In addition, interdependencies
must be interpreted in a largely subjective manner. This makes it impossible to
apply powerful mathematical statistical tools that enable more objective portfo-
lio analysis.6 This is a particularly challenging problem for top leaders in the
DoD required to make judgments for selection among alternatives (often referred
to as “trades”) about the potential benefits and risks of numerous projects and
technology investments.
A related approach is to perform sensitivity analysis. Each variable is perturbed
a pre-specified amount (e.g., unit sales is changed ±10 percent; sales price is
changed ±5 percent; and so forth), and the resulting change in net benefits is cap-
tured. This approach is useful for understanding which variables drive or impact
the result the most. Performing such analyses by hand or with simple Excel spread-
sheets is tedious and provides incremental benefits at best. A related approach that
has the same goals but uses a more powerful analytical framework is the use of
computer-modeled Monte Carlo risk simulations and tornado sensitivity analy-
sis, where all perturbations, scenarios, and sensitivities can be run hundreds of
thousands of times automatically (Mun 2010).
Computer-based Monte Carlo risk simulations can be viewed as an extension of
the traditional approaches of sensitivity and scenario testing. Simulations focus on
critical success drivers (variables that affect bottom-line variables the most, which
themselves are uncertain).
Applying simulation, interdependencies are captured using estimates of mul-
tiple regression analysis and other analytical tools. The uncertain variables are
then simulated tens of thousands of times automatically to simulate all poten-
tial permutations and combinations, generating a spectrum of possible outcomes.
The resulting net revenues–monetary benefits from these simulated potential out-
comes are tabulated and analyzed. In its most basic form, simulation is simply
an enhanced version of traditional approaches—such as sensitivity and scenario
analysis—but automatically performed thousands of times while accounting for
all the dynamic interactions between the simulated variables.
Based on specific assumptions, the resulting net revenue outcomes from a sim-
ulation can be displayed as in Figure 11.4. The interpretation is that there is a 90
percent probability net revenues will fall between US$19.44 and US$41.25, with
294 J. C. Mun, T. Housel

Net Revenue - Risk Simulator Forecast Distribution ~

Histogram Preferences II Statistics II Options I

Net Revenue·· Histogram/Cumulative Probability (5000 Trials)


600 1.00
()
500 0.80 §
c
[) 400
c 0.60 ~
~ 300
a
(l)

Q)
o.4o
tt 200 0"
m
100 0.20 ~
'<

9.92 15.92 25 .92 35 .92 40 .92 55 .£00 J


Type l rwo·Tail vJI. 19.4416 1 l 41 .2516 1 Certainty(%) ! 90 1

Figure 11.4 Simulation results.


Source: Mun (2010).

a 5 percent worst-case scenario of net revenues falling below US$19.44. Rather


than having only three scenarios, the simulation created 5,000 scenarios, or trials,
where multiple variables were changed simultaneously according to user-specified
underlying functional relationships (relating unit sales, sale price, and variable
cost-per-unit).
Monte Carlo risk simulation, named for the famous gambling capital of Monaco,
is a very potent methodology. For the practitioner, simulation opens the door to
solve difficult and complex but practical problems with relative ease.7 In simple
terms, Monte Carlo risk simulation creates virtual futures by generating thousands
and even hundreds of thousands of sample paths of outcomes and analyzes their
prevalent characteristics. In practice, Monte Carlo risk simulation methods are used
for risk analysis, risk quantification, sensitivity analysis, and forecasting.
An alternative to simulation is the use of highly complex stochastic closed-form
mathematical models (Mun 2010). A well-informed analyst would use all avail-
able statistical tools to obtain the same answer in the easiest and most practical
way possible. In all cases, when modeled correctly, Monte Carlo risk simulation
provides similar answers to the more mathematically elegant methods. In addition,
there are many real-life applications where closed-form models do not exist and
the only recourse is to apply simulation methods. So, what exactly is Monte Carlo
risk simulation and how does it work?
Monte Carlo risk simulation, in its simplest form, is a random-number gener-
ator that is useful for forecasting, estimating, and analyzing risk. A simulation
calculates numerous scenarios of a model by repeatedly picking values from a
user-predefined probability distribution for the uncertain variables and it uses
Real options: a risk-based approach 295
those values for the model. As all those scenarios produce results, the collec-
tion of results can be used to generate a forecast (based on underlying formulas,
functions, and user-specified probability distributions of key variables built into
the simulation).
Think of the Monte Carlo risk simulation approach as picking golf balls out of a
large basket repeatedly with replacement. The size and shape of the basket depend
on the distributional input assumption (e.g., a normal distribution with a mean of
100 and a standard deviation of 10, versus a uniform distribution or a triangu-
lar distribution) where some baskets are deeper or more symmetrical than others,
allowing certain balls to be pulled out more frequently than others. The number
of balls pulled repeatedly depends on the number of trials simulated. Each ball
is indicative of an event, scenario, or condition that can occur. For a large model
with multiple related assumptions, imagine the large model as a very large basket
wherein many baby baskets reside. Each baby basket has its own set of colored
golf balls bouncing around. Sometimes these baby baskets are linked with each
other (if there is a correlation between the variables), forcing the golf balls to
bounce in tandem—whereas in other uncorrelated cases, the balls are bouncing
independently of one another. The balls that are picked each time from these inter-
actions within the model (the large basket) are tabulated and recorded, providing
a forecast output result of the simulation.

11.4 Knowledge value added analysis


As the US military is not in the business of making money, referring to revenues
throughout this chapter may appear somewhat misleading. For non-profit orga-
nizations, especially the military, an alternative approach to those illustrated in
Chapters 8 or 9 to capture benefits is to apply KVA as a proxy estimate to gener-
ate an equivalent monetary “benefit” (or “revenue”) to permit ROI analysis (Cook
et al. 2007; Housel et al. 2007). ROI is a basic CBA ratio with the exception
that the estimated profit (revenue minus cost) is in the numerator and cost in the
denominator. KVA generates ROI estimates by developing a market comparable
price-per-common-unit of output, multiplied by the quantity of output, to achieve
a total revenue estimate.
The primary purpose of KVA methodology is to describe all organizational out-
puts in common units. The use of common units provides a means to compare the
outputs of all assets (human, machine, and IT). For example, the purpose of a mili-
tary process may be to collect signal intelligence or to plan for a ship modification
and upgrade. KVA would describe the outputs of both processes in common units;
thus making their performances comparable.
KVA measures the value provided by human capital assets and IT assets by ana-
lyzing an organization, process or function at the process level. It provides insights
into each dollar of IT investment by monetizing the outputs of all assets, including
intangible assets (e.g., such as that produced by IT and humans). By capturing the
value of knowledge embedded in an organization’s core processes (i.e., employ-
ees and IT), KVA identifies the actual cost and revenue of a process, product,
296 J. C. Mun, T. Housel
or service. Because it identifies every process required to produce an aggregated
output in terms of historical prices and costs-per-common-unit of output of those
processes, KVA allows unit costs and unit prices to be calculated. The methodol-
ogy has been applied in 45 areas within the DoD: from flight scheduling, to ship
maintenance and modernization (Rios et al. 2006; Housel et al. 2008; Rabelo and
Housel 2008; Kendall and Housel 2004).
As a performance tool, the KVA methodology:

• compares all processes in terms of relative productivity;


• allocates revenues and costs to common units of output;
• measures value added by IT by the outputs it produces; and
• relates outputs to the cost of producing those outputs in common units.

Based on the tenets of complexity theory, KVA assumes that humans and tech-
nology in organizations add value by taking inputs and changing them (measured
in units of complexity) into outputs through core processes. The amount of change
an asset within a process produces can be used as a measure of its marginal
value-added or benefit. The additional assumptions in KVA include the following:

• Describing all process outputs in common units (e.g., using a knowledge


metaphor for descriptive language in terms of time it takes an average
employee to learn how to produce the outputs) allows historical revenue and
cost data to be assigned to those processes.
• All outputs can be described in terms of the time required to learn how to
produce them.
• Learning time, a proxy for procedural knowledge required to produce process
outputs, is measured in common units of time. Consequently, units of learning
time = common units of output or “Knowledge” (K ).
• A common unit of output makes it possible to compare all outputs in terms
of cost-per-unit as well as price-per-unit; revenue can now be assigned at the
sub-organizational level.
• Once cost and revenue streams have been assigned to sub-organizational out-
puts, normal accounting and financial performance and profitability metrics
can be applied (Rodgers and Housel 2006; Pavlou et al. 2005; Housel and
Kanevsky 1995).

Describing processes in common units also permits market-comparable data


to be generated, a feature particularly important for non-profit organizations like
the US military. Using a market-comparables approach, data from the commer-
cial sector can be used to estimate price-per-common-unit, allowing a monetary
measure of benefits or revenue estimates of process outputs for non-profit orga-
nizations. This approach also provides a common-units basis to define benefit
streams regardless of the process analyzed.
KVA differs from other non-profit ROI models because it allows for revenue
estimates, enabling the use of traditional accounting, financial performance, and
Real options: a risk-based approach 297
profitability measures at the sub-organizational level. KVA can rank processes by
the degree to which they add value to the organization or its outputs. This ranking
assists decision-makers in identifying the value-added of various processes.
Value is quantified in two key metrics: Return-on-knowledge (ROK: rev-
enue/cost) and ROI ((revenue-investment cost)/investment cost). The outputs from
a KVA analysis become the input into the ROI models and real options analy-
sis discussed next. By tracking the historical volatility of price- and cost-per-unit
as well as ROI, it is possible to establish risk (as compared to uncertainty)
distributions, which is important to accurately estimate the value of real options.
The KVA method has been applied to numerous core military processes across
the services (Rios et al. 2006; Housel et al. 2008; Housel and Nelson 2005). This
KVA research has provided a means for simplifying real options analysis for DoD
processes discussed below. Current KVA research will provide a library of market-
comparable price- and cost-per-unit of output estimates. This research will enable
a more stable basis for comparisons of performance across core processes. The
data also provide a means to establish risk-distribution profiles for integrated risk
analysis approaches such as real options and KVA, which have been linked directly
to the Real Options Super Lattice Solver and Risk Simulator software for rapid
adjustments to real options valuation projections (Housel et al. 2010).

11.5 Strategic real options analysis


An important step in performing strategic real options analysis is the application of
Monte Carlo risk simulation. By applying Monte Carlo risk simulation to simul-
taneously change all critical inputs in a correlated manner within a model, one
can identify, quantify, and analyze risk.8 The question then is: what next? Simply
quantifying risk is useless unless you can manage it, reduce it, control it, hedge it,
or mitigate it. This is where strategic real options analysis can be used. Think of
real options as a strategic road map for making decisions.
Suppose you are driving from point A to point B, and you only have or know
one way to get there: a straight route. Further suppose there is a lot of uncertainty
as to what traffic conditions are like further down the road, and you risk being
stuck in traffic (there is a 50 percent chance this will occur).
Simulation will provide you the 50 percent figure. Knowing that half the time
you will get stuck in traffic is valuable information, but the question now is: so
what? (especially if you must ultimately arrive at point B). Suppose you discover
several alternate routes to get to point B. You could still drive the straight route;
but, if you hit traffic, make a left, right, or U-turn to avoid congestion—mitigating
the risk, and getting you to point B faster and more safely. There is potential value
in these newly discovered real options.
The real options approach can help guide how much you would be willing to
pay for a strategic road map or global positioning satellite investment to reveal
and possibly exploit alternative routes. In military situations with high risk, real
options can help identify strategies to mitigate risks. In fact, both businesses and
the military have applied real option approaches for hundreds of years without
298 J. C. Mun, T. Housel
fully realizing it—the military calls this approach courses of action. For example,
do we take Hill A so that it provides us the option and ability to take Hill B
and Valley C, or how should we take Valley C? Or do we avoid taking Valley C
altogether? And so forth.
It is useful to discuss the formal structure (and subsequent analytics) that real
options analysis provides. Using real options analysis, we can quantify and value
each strategic pathway and frame strategies to hedge, mitigate, or sometimes take
advantage of risk.
In the past, corporate investment decisions were relatively straightforward,
applying traditional cost–benefit (NPV) analysis as in Chapter 17. Buy a new
machine that is more efficient (where the discounted stream of future mone-
tary benefits exceeds the investment cost); make more products costing a certain
amount (if benefits outweigh the costs); or hire a larger pool of sales associates (if
the marginal increase in forecast sales revenues exceeds the additional salary and
implementation costs). Need a new manufacturing plant? Show that the construc-
tion costs can be recouped quickly and easily by the increase in revenues it will
generate through new and more improved products, and the initiative is approved.
Many real-life conditions, however, are more complex.
Suppose a company decides to go with an e-commerce strategy, but multiple
strategic paths exist. Which path does it choose? What are the options? If the
wrong path is chosen, how does it get back on the right track? How does the
company value and prioritize the paths that exist? If you are a venture capitalist
firm with multiple business plans to consider for possible investment, how do you
value a start-up firm with no proven track record? How do you structure a mutually
beneficial investment deal? What is the optimal timing to a second or third round
of financing?
Real options analysis is useful not only in valuing a firm through its strategic
business options: it can also be used as a strategic business tool in capital invest-
ment and acquisition decisions. For instance, should the military invest millions
in a new open architecture initiative and, if so, what are the values of the various
strategies for such an investment, and how do we proceed? How should the mili-
tary choose among several alternative IT infrastructure projects? Should it invest
billions in a risky R&D initiative? The consequences of a wrong decision can be
disastrous, and lives could be at stake.
In a traditional analysis, these questions cannot be answered with any certainty.
In fact, some of the answers generated through the use of traditional analysis
can be misleading because most models assume a static, one-time decision-
making process. The real options approach is sequential in nature, taking into
consideration the strategic options certain projects create under uncertainty and a
decision-maker’s flexibility in exercising or abandoning these options at different
points in time, as the level of uncertainty decreases or is revealed over time.
Traditional analysis assumes a static investment decision and assumes that
strategic decisions are made initially with no recourse to choose other pathways
or options in the future. In contrast, the real options approach incorporates a learn-
ing model such that the decision-maker makes better and more-informed strategic
Real options: a risk-based approach 299
decisions when some levels of uncertainty are resolved through the passage of
time, actions, and events (see Chapter 14 for an interesting application).
The use of the KVA methodology to monitor the performance of given options,
and the adjustments to real options as leaders learn more from the execution of
given options, provides an integrated methodology to help military leaders hedge
their bets while taking advantage of new opportunities over time. Real options
analysis can be used to frame strategies to mitigate risk, increase value, and find
optimal strategic pathways to pursue. It can likewise generate options to enhance
the value of a project while managing risks. Imagine real options as your guide
when navigating through unfamiliar territory, providing road signs at every turn
to help make the best and most informed driving decisions. This guidance is the
essence of real options.
Figure 11.5 illustrates a very basic real options framing exercise—clearly more
complex situations can be developed. From the options that are framed, Monte

Phase
Phase Milestone 4
Phase II Milestone 3
Exit
Milestone 2 Exit Stop after
Phase I
Phase III
Stop after
Milestone 1
Phase II
Strategy A Exit
Divide the development Stop after
into multiple milestones, Phase I
Exit
stage-gating the
development and Do nothing Contract
mitigating the risk. Outsourcing
Phase II
Larger Deployment
Exit
POC
Small-Scale Internal development rather
Proof of than off-the-shelf applications
Start Strategy B
Exit
Concept
Apply a quick proof-of-
concept (POC) of the Abandon
technology, replicate and Exit
scale to larger force-wide Do nothing
deployment. If POC fails, Expan
abandon project and
reduce future risks. R&D new technology,
Outsource or obtain off- Buy expand into other
the-shelf platforms if application areas
successful. Purchase
technology Exit
Strategy C
Purchase technology Sell IP and technology,
rather than developing it Exit abandon project
in-house. If application fails,
Do nothing
abandon and sell the
company’s intellectual
property and technology. If
successful, find other
applications for technology.

Figure 11.5 Example real options framing.


Source: Mun (2010).
300 J. C. Mun, T. Housel
Carlo risk simulation and stochastic forecasting, coupled with traditional tech-
niques, are applied. Then, real options analytics are used to solve and value each
strategic pathway allowing better-informed decisions.9
As mentioned previously, real options analysis can generate options to enhance
the value of a project while managing risks. Sample options include the option
to expand, contract, abandon, or utilize sequential compound options (phased
stage-gate options, options to wait and defer investments, proof-of-concept stages,
milestone development, and R&D initiatives).
Sample applications in the military include using real options for acquisitions
(see Chapter 14), spiral development, and various organizational configurations,
as well as the strategic importance of integrated and open architectures as real
options multipliers. Under guidance of the US Office of Management and Budget
(OMB) Circular A-76, comparisons using real options analysis could be applied
to enhance outsourcing competitions between the Government’s Most Efficient
Organization (MEO) and private-sector alternatives. Real options can be used
throughout strategic planning (the Joint Capabilities Integration Development Sys-
tem (JCIDS) requirements generation process in the United States) and in the
Defense Acquisition System.10 Many other possible applications exist in military
decision-making and portfolio analysis.

11.6 Portfolio optimization


In most decisions, there are variables over which leadership has control (control
variables), such as how to establish supply lines, modernize a ship, use network
centricity to gather intelligence, and so on. Similarly, business leaders have options
in what they charge for a product, how much to invest in a project, or which
projects they should choose in a portfolio when they are constrained by bud-
gets or other resources. These decisions could also include allocating financial
resources, building or expanding facilities, managing inventories, and determining
product-mix strategies.
Such decisions might involve thousands or millions of possible alternatives.
Considering and evaluating each of them would be impractical or even impossible.
These control variables are also called decision variables. Finding the optimal val-
ues for decision variables can make the difference between reaching an important
goal and missing that goal. An optimization model can provide valuable assistance
in incorporating relevant variables when analyzing decisions and finding the best
solutions for making decisions. Optimization models often provide insights that
intuition alone cannot.
An optimization model has three major elements: an objective, decision vari-
ables, and constraints (see Mun 2010, Chapter 4). In short, the optimization
methodology finds the best combination or permutation of decision variables (e.g.,
best way to deploy troops, build ships, and choose which projects to execute) in
every conceivable way such that the objective is maximized (e.g., strategic value,
enemy assets destroyed, or ROI), or minimized (e.g., risk and costs), while still
satisfying the constraints (e.g., time, budget, and other scarce resources).
Real options: a risk-based approach 301
Obtaining optimal values from complex models generally requires iterative or
ad hoc search algorithms. This search often involves running one iteration for an
initial set of values, analyzing the results, changing one or more values, rerunning
the model, and repeating the process until a satisfactory solution is obtained. This
process can be very tedious and time consuming even for small models and it is
often not clear how to adjust values from one iteration to the next.
A more rigorous method systematically enumerates all possible alternatives
(Mun 2010). This approach guarantees optimal solutions if the model is correctly
specified.11 Figures 11.6 to 11.8 illustrate a sample portfolio analysis where, in the
first case, there are 20 total projects to choose from (if all projects were executed,
it would cost US$10.2 billion), and where each project has its own ROI or mon-
etary benefit measure, cost, strategic ranking, and comprehensive, tactical, and
total military scores (obtained from field commanders through the Delphi Method
to elicit their thoughts about the strategic value of a particular project or initiative;
see Mun 2010).
The constraints are full-time equivalence resources, budget, and strategic score.
In other words, to build the most strategic portfolio possible there are 20 projects
or initiatives to choose from, where we want to select the top 10—subject to
having enough money to pay for them, the people to do the work, etc.12 All the
while, Monte Carlo risk simulation, real options, and forecasting methodologies
are applied in the optimization model (e.g., the values for each project are shown
in Figure 11.6, linked from their own models with simulation and forecasting
methodologies applied, and the best strategy for each project is chosen using real
options analysis) (Mun 2006, 2010). Or perhaps the projects shown are nested
within one another. For instance, you cannot exercise Project 2 unless you execute
Project 1, but you can only exercise Project 1 without having to do Project 2, and
so forth. The results are shown in Figure 11.6.
Figure 11.7 shows the optimization process done in series with some of the con-
straints relaxed. For instance, what would be the best portfolio and the strategic
outcome if a budget of US$3.8 billion were imposed? What if it were increased
to US$4.8 billion, US$5.8 billion, and so forth? The efficient frontiers depicted
in Figure 11.7 illustrate the best combination and permutation of projects in the
optimal portfolio. Each point on the frontier is a portfolio of various combinations
of projects that provides the best allocation possible given the different budgets
(requirements and constraints) (Note: This can be thought of as a stochastic ana-
logue to the deterministic economic evaluation of alternatives (EEoA) approach
to CBA introduced in Chapter 4).
Finally, Figure 11.8, another example, shows the top ten projects (based on a
budget constraint) that were chosen and how the total budget is optimally allocated
to provide the best and most well-balanced portfolio.

11.7 Integrated risk management


We are now able to put all the pieces together into an Integrated Risk Management
framework to see how these different techniques are related in a risk analysis and
Strategy Return to Profitability Military Tactica FTE Comprehensive
Project Name ENPV NPV Cost Selection
Ranking Rank Ratio Index Score l Score Resources Score

Project 1 $458.00 $150.76 $1,732.44 1.20 381.67 1.09 0 8.10 2.31 1.20 1.98
Project 2 $1,954.00 $245.00 $859.00 9.80 199.39 1.29 1 1.27 4.83 2.50 1.76
Project 3 $1,599.00 $458.00 $1,845.00 9.70 164.85 1.25 0 9.88 4.75 3.60 2.77
Project 4 $2,251.00 $529.00 $1,645.00 4.50 500.22 1.32 0 8.83 1.61 4.50 2.07
Project 5 $849.00 $564.00 $458.00 10.90 77.89 2.23 0 5.02 6.25 5.50 2.94
Project 6 $758.00 $135.00 $52.00 7.40 102.43 3.60 1 3.64 5.79 9.20 3.26
Project 7 $2,845.00 $311.00 $758.00 19.80 143.69 1.41 1 5.27 6.47 12.50 4.04
Project 8 $1,235.00 $754.00 $115.00 7.50 164.67 7.56 1 9.80 7.16 5.30 3.63
Project 9 $1,945.00 $198.00 $125.00 10.80 180.09 2.58 1 5.68 2.39 6.30 2.16
Project 10 $2,250.00 $785.00 $458.00 8.50 264.71 2.71 1 8.29 4.41 4.50 2.67
Project 11 $549.00 $35.00 $45.00 4.80 114.38 1.78 0 7.52 4.65 4.90 2.75
Project 12 $525.00 $75.00 $105.00 5.90 88.98 1.71 0 5.54 5.09 5.20 2.69
Project 13 $516.00 $451.00 $48.00 2.80 184.29 10.40 0 2.51 2.17 4.60 1.66
Project 14 $499.00 $458.00 $351.00 9.40 53.09 2.30 1 9.41 9.49 9.90 4.85
Project 15 $859.00 $125.00 $421.00 6.50 132.15 1.30 1 6.91 9.62 7.20 4.25
Project 16 $884.00 $458.00 $124.00 3.90 226.67 4.69 1 7.06 9.98 7.50 4.46
Project 17 $956.00 $124.00 $521.00 15.40 62.08 1.24 1 1.25 2.50 8.60 2.07
Project 18 $854.00 $164.00 $512.00 21.00 40.67 1.32 0 3.09 2.90 4.30 1.70
Project 19 $195.00 $45.00 $5.00 1.20 162.50 10.00 0 5.25 1.22 4.10 1.86
Project 20 $210.00 $85.00 $21.00 1.00 210.00 5.05 0 2.01 4.06 5.20 2.50

Total $14,185.00 $3,784.00 99.00 10 58.58 62.64 73.50 33.15


Profit/Rank $143.28
Profit*Score $470,235.60 Maximize < =$3800 < =100 x <=10 <=80

Figure 11.6 Portfolio optimization and allocation.


Source: Mun (2010).
Comprehensive Tactical Allowed ROI-RANK
Budget Score Score Military Score Projects Objective

$3,800.00 33.15 62.64 58.58 10 $470,235.60


$4,800.00 36.33 68.85 66.86 11 $521,645.92
$5,800.00 38.40 70.46 75.69 12 $623,557.79
$6,800.00 39.94 72.14 82.31 13 $659,947.99
$7,800.00 39.76 70.05 86.54 14 $676,279.81

Comprehensive Score Military Score


90.00 86.54
42.00
85.00 82.31
39.94 39.76
40.00 80.00
38.40 75.69
38.00 75.00
36.33
36.00 70.00 66.86

34.00 33.15 65.00


58.58
32.00 60.00

30.00 55.00
$3,500 $4,000 $4,500 $5,000 $5,500 $6,000 $6,500 $7,000 $7,500 $8,000 $8,500 $3,500 $4,500 $5,500 $6,500 $7,500 $8,500

Tactical Score ROI-RANK


74.00
72.14 $700,000.00 $676,279.81
$659,947.99
72.00
70.46 70.05 $650,000.00
70.00 $623,557.79
68.85

68.00 $600,000.00

66.00 $550,000.00

64.00 62.64 $521,645.92


$500,000.00 $470,235.60
62.00
$450,000.00
60.00 $3,500 $4,000 $4,500 $5,000 $5,500 $6,000 $6,500 $7,000 $7,500 $8,000 $8,500
$3,500 $4,000 $4,500 $5,000 $5,500 $6,000 $6,500 $7,000 $7,500 $8,000 $8,500

Figure 11.7 Efficient frontiers of portfolios.


Source: Mun (2010).
Required Required Returns Risk Return to Allocation
Asset Class Annualized Volatility Allocation Return to
Minimum Maximum Ranking Ranking Risk Ranking Ranking
Description Returns Risk Weights Risk Ratio
Allocation Allocation (Hi-Lo) (Lo-Hi) (Hi-Lo) (Hi-Lo)
Selected Project 1 10.54% 12.36% 11.09% 5.00% 35.00% 0.8524 9 2 7 4
Selected Project 2 11.25% 16.23% 6.86% 5.00% 35.00% 0.6929 7 8 10 10
Selected Project 3 11.84% 15.64% 7.78% 5.00% 35.00% 0.7570 6 7 9 9
Selected Project 4 10.64% 12.35% 11.23% 5.00% 35.00% 0.8615 8 1 5 3
Selected Project 5 13.25% 13.28% 12.09% 5.00% 35.00% 0.9977 5 4 2 2
Selected Project 6 14.21% 14.39% 11.04% 5.00% 35.00% 0.9875 3 6 3 5
Selected Project 7 15.53% 14.25% 12.30% 5.00% 35.00% 1.0898 1 5 1 1
Selected Project 8 14.95% 16.44% 8.90% 5.00% 35.00% 0.9094 2 9 4 7
Selected Project 9 14.16% 16.50% 8.37% 5.00% 35.00% 0.8584 4 10 6 8
Selected Project 10 10.06% 12.50% 10.35% 5.00% 35.00% 0.8045 10 3 8 6

Portfolio Total 12.6919% 4.52% 100.00%


Return to Risk Ratio 2.8091

Figure 11.8 Portfolio optimization (continuous allocation of funds).


Source: Mun (2010).
Real options: a risk-based approach 305
risk management context. This framework is comprised of eight distinct phases
of a successful and comprehensive risk analysis implementation, going from a
qualitative management-screening process, to creating clear and concise reports
for management. The process was developed by the authors based on previ-
ous successful implementations of risk analysis, forecasting, real options, KVA
cash-flow estimates, valuation, and optimization projects both in the consulting
arena and in industry-specific problems (Mun 2010). These phases can be per-
formed either in isolation or combined in sequence for a more robust integrated
analysis.
Figure 11.9 shows the integrated risk analysis process. We can segregate the
process into the following eight simple steps:

1 qualitative management screening;


2 time-series and regression forecasting;
3 base case KVA and NPV analysis;
4 Monte Carlo risk simulation;
5 real options problem framing;
6 real options modeling and analysis;
7 portfolio and resource optimization;
8 reporting and update analysis.

11.7.1 Qualitative management screening


Qualitative management screening is the first step in any integrated risk analy-
sis process. Decision-makers have to decide which projects, assets, initiatives, or
strategies are viable for further analysis in accordance with the organization’s mis-
sion, vision, goal, or overall strategy. The organization’s mission, vision, goal, or
overall strategy may include strategies and tactics, competitive advantage, techni-
cal, acquisition, growth, or synergistic or global threat issues. That is, the initial
list of projects should be qualified in terms of meeting the leadership’s agenda.
Often the most valuable insight is created as leaders frame the complete problem
to be resolved. This is the step where risks to the organization are identified and
carefully described.

11.7.2 Time-series and regression forecasting


The future is then forecasted using time-series analysis, stochastic forecasting, or
multivariate regression analysis (if historical or comparable data exist). Otherwise,
other qualitative forecasting methods may be used (subjective guesses, growth-rate
assumptions, expert opinions, Delphi Method, and so forth).13

11.7.3 Base case KVA and net present value analysis


For each project that passes the initial qualitative screens, a KVA-based discounted
cash-flow model is created (Rios et al. 2006). This model serves as the base case
analysis where a NPV and ROI are calculated for each project using the forecasted
List of projects Base case projections Develop static Dynamic Monte
1 and strategies to 2 for each project
3 financial models
4 Carlo simulation
evaluate
Risk
Risk Simulator
Simulator
A
B Simulation
C Time Series Forecasting
D
E

RISK ANALYSIS

RISK MODELING

RISK PREDICTION

RISK IDENTIFICATION
Start with a list of projects

Traditional analysis stops here!


or strategies to be …with the assistance of …the user generates a …Monte Carlo simulation is added
evaluated… these projects time-series forecasting, traditional series of static base to the analysis and the financial
have already been through future outcomes can be case financial (discounted cash model outputs become inputs into
qualitative screening predicted... flow) models for each project… the real options analysis…

Framing Options analytics, Portfolio optimization Reports presentation


5 Real Options 6 simulation, optimization 7 and asset allocation 8 and update analysis
Simulation Lattice Period First Cash Flow Discounted Value of Discounted Value of the Interest Rate

DCF Value Opportunity Cost


Risk Starting (t) (t + 3) Future Cash Flows Costs to Invest (monthly basis)

Simulator Phase II Options

Optimization Retirement
13 296,916 9,851,788 6,086,684 3,765,104 0.949% 0.87%

Effects of Waiting Personal Financials


13 158,350 4,741,612 4,869,348 -127,735 0.949% 0.87%

Defray cost Effects of Going Private Loans


19 132,757 3,246,855 5,921,771 -2,674,916 0.949% 0.87%

Other opportunities Revenue enhancement Academic Loans


19 146,850 3,715,300 4,288,179 -572,878 0.949% 0.87%

Loss revenues Cost reduction


- Optimal Exercise
share
+ Decision Standard

realized label
Loss cost reduction Strategic options value
Value of the

+ Deviation of Actualized Flexibility


+ competition
Loss of market Strategic
"Discounted Value Option Value at t Option Value at t = 0 Decision To Invest
+ Volatility- +
leadership competitiveness
Actualized CF Parameter
Market Value + technical risk
of the Costs to
High cost outlay
+ + + Cash Flows
+ +
prevalence Project Value developmental testing Invest"
-
- -
+
market size + Real Options Super
- - R&D Cost
4,130,101 Execute Investment
elasticity price Post-P3 Cost 9,851,788 1.263
+
+ Lattice Solver 2,324,992 Wait to Invest
4,741,612 1.263
Time to P3

23,699 Wait to Invest


3,246,855 1.263
…stochastic optimization is the

RISK HEDGING
1,154,349 Wait to Invest
3,715,300 1.263

RISK MITIGATION
…the relevant projects next optional step if multiple
are chosen for real …real options analytics are projects exist that require efficient
RISK MANAGEMENT

RISK DIVERSIFICATION
options analysis and the calculated through binomial lattices asset allocation given some …create reports, make
project or portfolio real and closed-form partial-differential budgetary constraints… useful for decisions, and do it all
options are framed… models with simulation… strategic portfolio management… again iteratively over time…

Figure 11.9 Integrated risk analysis process.


Real options: a risk-based approach 307
values in the previous step. This step also applies if only a single project is under
consideration. This NPV is calculated using the traditional approach utilizing the
forecast revenues and costs and discounting the net of these revenues and costs
at an appropriate risk-adjusted rate (see Chapter 17). The ROI and other financial
metrics are generated here.

11.7.4 Monte Carlo risk simulation14


Because the static discounted cash flow produces only a single-point estimate,
there is often little confidence in its accuracy given that future events that affect
forecasted cash flows are highly uncertain. To better estimate the actual value
of a particular project, Monte Carlo risk simulation can be employed. Usually,
a sensitivity analysis is first performed on the discounted cash-flow model; that is,
setting the NPV or ROI as the dependent variable, we can change each significant
explanatory variable and note the change in the dependent variable of interest.
Explanatory variables can include revenues, costs, tax rates, discount rates, capital
expenditures, depreciation, and so forth, which ultimately flow through the model
to affect the NPV or ROI figure of interest. Each significant explanatory variable
can be changed by a set amount to determine the effect on the resulting NPV.
A graphical representation can then be created in Risk Simulator (which is often
called a tornado chart because of its shape), where the most sensitive precedent
(explanatory) variables are listed first in descending order of magnitude. Armed
with this information, the analyst can then decide which key variables are highly
uncertain in the future and which can be assumed to be deterministic. The set of
key uncertain variables that drive the NPV (and, hence, the decision) are called
critical success drivers. These critical success drivers are prime candidates for
Monte Carlo risk simulation. Because some of these critical success drivers may
be correlated, a correlated and multidimensional Monte Carlo risk simulation
may be required. Typically, these correlations can be obtained through histori-
cal data. Running correlated simulations provides a much closer approximation to
the variables’ real-life behaviors.

11.7.5 Real options problem framing15


The risk information obtained somehow needs to be converted into actionable
intelligence. Once risk has been quantified using Monte Carlo risk simulation,
what do we do about it? The answer is to use real options analysis to hedge these
risks, to value these hedges, and to position the organization to take advantage of
the risks.
The first step in real options analysis is to generate a strategic map through the
process of framing the problem. Based on the overall problem identification occur-
ring during the initial qualitative management-screening process, certain strategic
option alternatives would have become apparent for each particular project. These
so-called strategic optionalities may include (among other things) the option to
expand, contract, abandon, switch, choose, and so forth. Based on the identifi-
cation of strategic optionalities that exist for each project or at each stage of the
308 J. C. Mun, T. Housel
project, the analyst can then choose from a list of options to analyze each in more
detail. Real options can be added to the projects to hedge downside risks and to
take advantage of upside swings.

11.7.6 Real options modeling and analysis


Through the use of Monte Carlo risk simulation, the resulting stochastic dis-
counted cash-flow model will have a distribution of values. Thus, simulation
models analyze and quantify the various risks and uncertainties of each project.
The result is a distribution of NPVs and the project’s volatility. In the real options
approach, we assume that the underlying variable is the future profitability of the
project, which is the future cash-flow series.
An implied volatility of the future free cash flow or underlying variable can
be calculated through the results of a Monte Carlo risk simulation previously
performed. Usually, the volatility is measured as the standard deviation of the
logarithmic returns on the free cash-flow stream. In addition, the present value
of future cash flows for the base case discounted cash-flow model is used as the
initial underlying asset value in real options modeling (Mun 2006). Using these
inputs, real options analysis is performed to obtain the project’s strategic option
values. (Note: an alternative approach is introduced in Chapter 14.)

11.7.7 Portfolio and resource optimization16


Portfolio optimization is an optional step in the analysis that generalizes and
extends the “opportunity cost” approach in the EEoA discussed in Chapter 4. If the
analysis is done on multiple projects, decision-makers should view the results as
a portfolio of rolled-up projects because the projects are, in most cases, correlated
with one another, and viewing them individually will not present the true picture.
As most organizations do not only have single projects, portfolio optimization is
crucial. Given that certain projects are related to others, there are opportunities for
hedging and diversifying risks through a portfolio approach. Because firms have
limited budgets and have time and other resource constraints, while at the same
time they have requirements for certain overall levels of returns, risk tolerances,
and so forth, portfolio optimization takes into account all these to create an opti-
mal portfolio mix. The analysis will provide the optimal allocation of investments
across multiple projects.

11.7.8 Reporting and update analysis


The analysis is not complete until reports are generated. Not only are the results
presented, but the process used to obtain the results should also be described.
Clear, concise, and precise explanations transform a difficult black-box set of ana-
lytics into transparent steps. Decision-makers will be reluctant to accept results
coming from black boxes if they do not understand where the assumptions or
data originate and what types of mathematical or analytical models have been
applied. Risk analysis assumes that the future is uncertain and that decision-
makers have the ability to make midcourse corrections when these uncertainties
Real options: a risk-based approach 309
become resolved or risks become known; the analysis is usually done ahead of
time and, thus, ahead of such uncertainty and risks.
Therefore, when these risks become known over the passage of time, actions,
or events, the analysis should be revisited to incorporate the decisions made at
each step and/or to revise any input assumptions. Sometimes, for long-horizon
projects, several iterations of real options analysis should be performed, where
future iterations are updated with the latest data and assumptions. Understanding
the steps required to undertake an integrated risk analysis is important because it
provides insight not only into the methodology itself, but also into how it evolves
from traditional analyses, showing where the traditional approach ends and where
the new analytics begin.

11.8 Conclusion
The leadership of ministries of defense can take advantage of more advanced ana-
lytical procedures to make strategic investment decisions and manage portfolios
of projects using KVA and the real options approach. In the past, due to the lack
of technological maturity, this would have been extremely difficult; hence, busi-
nesses and the government had to resort to intuition, past experience, and relatively
unsophisticated models. Now, with the assistance of new technology and more
mature methodologies, financial leaders have every reason to take the analysis a
step further.
Corporations such as 3M, Airbus, AT&T, Boeing, BP, Chevron, Johnson &
Johnson, Motorola, and many others have already successfully applied these tech-
niques. As emphasized in this chapter (and in the real options application in
Chapter 14) the military has the same opportunity. The relevant software appli-
cations, books, case studies, and public seminars have been created, including
specific case studies developed exclusively for the US Navy.
The only real barrier to implementation is a lack of exposure to the potential
benefits of these new methods. Hopefully, this chapter has revealed the poten-
tial benefits of these analytical techniques and tools that can complement and
enhance CBA techniques currently used by leadership. In order to fully prepare
for twenty-first-century challenges and create highly effective and flexible mil-
itary forces, strategic real options, KVA, and risk analysis can be leveraged by
leadership to improve decision making. Combining real options with KVA can
help ensure maximum strategic flexibility and improve risk-based CBA.

Notes
1 To the people who lived centuries ago, risk was simply the inevitability of chance occur-
rence beyond the realm of human control, albeit many phony soothsayers profited from
their ability to convincingly profess their clairvoyance by simply stating the obvious
or reading the victims’ body language and telling them what they wanted to hear. We
modern-day humans—ignoring for the moment the occasional seers among us and even
with our fancy technological achievements—are still susceptible to risk and uncertainty.
We may be able to predict the orbital paths of planets in our solar system with astound-
ing accuracy or the escape velocity required to shoot a man from the Earth to the Moon
310 J. C. Mun, T. Housel
or drop a smart bomb within a few feet of its target thousands of miles away, but when
it comes to, say, predicting a firm’s revenues the following year, we are at a loss.
2 There are many new US Department of Defense (DoD) requirements for using more
advanced analytical techniques. For instance, the Clinger-Cohen Act of 1996 mandates
the use of portfolio management for all federal agencies. The Government Accountabil-
ity Office’s Assessing Risks and Returns: A Guide for Evaluating Federal Agencies’ IT
Investment Decision-Making (Version 1) (February 1997) requires that IT investments
apply ROI measures. DoD Directive 8115.01 issued October 2005 mandates the use of
performance metrics based on outputs, with ROI analysis required for all current and
planned IT investments. DoD Directive 8115.bb (2006) implements policy and assigns
responsibilities for the management of DoD IT investments as portfolios within the DoD
enterprise—where a portfolio is defined to include outcome performance measures and
an expected ROI. The DoD Risk Management Guidance Defense Acquisition guidebook
requires that alternatives to the traditional cost estimation need to be considered because
legacy cost models tend not to adequately address costs associated with information
systems or the risks associated with them.
3 Risks can be computed many ways, including volatility, standard deviation of lognormal
returns, value at risk, and so forth. For more technical details, see Mun’s Modeling Risk
(2nd edition, 2010).
4 We will demonstrate how KVA combined with the traditional market comparables
valuation method allows for the monetization of benefits (i.e., revenue).
5 On a continuous basis, the probability of occurrence is the area under a curve, e.g., there
is a 90 percent probability revenues will be between US$10 and US$11. The area under
a straight line, however, approaches zero. Therefore, the probability of hitting exactly
US$10.00 is close to 0.00000001 percent.
6 See Chapter 14 for an interesting application of real options analysis in the case where
benefits are not measured in monetary units.
7 Perhaps the most famous early use of Monte Carlo risk simulation was by the Nobel
physicist Enrico Fermi (sometimes referred to as the father of the atomic bomb) in 1930,
when he used a random method to calculate the properties of the newly discovered neu-
tron. Monte Carlo methods were central to the simulations required for the Manhattan
Project: in the 1950s Monte Carlo risk simulation was used at Los Alamos for early
work relating to the development of the hydrogen bomb and became popularized in the
fields of physics and operations research. The Rand Corporation and the US Air Force
were two of the major organizations responsible for funding and disseminating informa-
tion on Monte Carlo methods during this time, and today there is a wide application of
Monte Carlo risk simulation in many different fields—including engineering, physics,
research and development, business, and finance.
8 The outcomes from a Monte Carlo risk simulation include probabilities and various risk
statistics that can be used to make better decisions.
9 The pathways can be valued using partial differential closed-form equations, lattices,
and simulation. Various software packages are available for this task, such as the
author’s The Real Options SLS software by Real Options Valuation, Inc., available at
www.realoptionsvaluation.com [last accessed November 21, 2014].
10 For example, in the US DOTMLPF vs New Program/Service solutions, joint integra-
tion, analysis of material alternatives (AMA), analysis of alternatives (AoA), and spiral
development.
11 Suppose that an optimization model depends on only two decision variables. If each
variable has 10 possible values, trying each combination requires 100 iterations (102
alternatives). If each iteration takes a very short amount of time (e.g., two seconds), then
the entire process could take approximately three minutes of computer time. Instead of
two decision variables, however, consider six; then consider that trying all combina-
tions requires 1,000,000 iterations (106 alternatives). It is easily possible for complete
Real options: a risk-based approach 311
enumeration to take many years to carry out. Therefore, solving complex optimizations
has always been a fantasy until now: With the advent of sophisticated software and
computing power, coupled with smart heuristics and algorithms, such analyses can now
be completed within minutes.
12 There are 2 × 1018 possible permutations for this problem and, if tested by hand, it
would take years to complete. Using Risk Simulator, the problem is solved in about
five seconds, or several minutes if Monte Carlo risk simulation and real options are
incorporated in the analysis.
13 See Mun (2010, Chapters 8 and 9) for details on forecasting and using software to run
time-series, extrapolation, stochastic process, ARIMA, and regression forecasts.
14 See Mun (2010, Chapters 4 and 5) for details on using software to run Monte Carlo
simulations.
15 See Mun (2006) for more technical details on framing and solving real options
problems.
16 See Mun (2010, Chapters 10 and 11) for details on using software to perform portfolio
optimization.

References
Cook, G. R., T. J. Housel, J. Mun, and P. Pavlou. January 2007. “Return on Investment
in Non-revenue Generating Activities: Applying KVA and Real Options to Government
Operations.” Review of Business Research 7(1): 16–19.
DoD. US Department of Defense Directive 8115. October 10, 2005. Information Tech-
nology Portfolio Management. NUMBER 8115.01, ASD(NII)/DoD CIO. Available
at http://dtic.mil/whs/directives/corres/pdf/811501p.pdf [last accessed November 27,
2014].
DoD. US Department of Defense Directive 8115. October 30, 2006. Information Technol-
ogy Portfolio Management Implementation. NUMBER 8115.02, ASD(NII)/DoD CIO.
Available at http://www.dtic.mil/whs/directives/corres/pdf/811502p.pdf [last accessed
November 27, 2014].
GAO. February 1997. Assessing Risks and Returns: A Guide for Evaluating Federal
Agencies’ IT Investment Decision-Making. Version 1. GAO/AIMD-10.1.13. Available
at www.gao.gov/assets/80/76295.pdf [last accessed November 27, 2014].
Housel, T. J., and V. Kanevsky 1995. “Reengineering Business Processes: A Complexity
Theory Approach to Value Added.” INFOR 33: 248–262.
Housel, T. J., and S. K. Nelson. 2005. “Knowledge Valuation Analysis: Applica-
tions for Organizational Intellectual Capital.” Journal of Intellectual Capital 6(4):
544–557.
Housel, T. J., W. Little, and W. Rodgers. October 29–31, 2007. “Estimating the Value
of Non-Profit Organizations Using the Market Comparables Approach.” In Refereed
Proceedings of the Collected Papers of the European Institute for Advanced Studies
in Management (EIASM): 3RD Workshop on Visualising, Measuring and Managing
Intangibles & Intellectual Capital. Ferrara, Italy, pp. 156–162.
Housel, T. J., E. Tarantino, S. Hom, and J. Mun. 2008. “Shipyard Planning Case.” In Making
Cents Out of Knowledge Management, edited by J. Liebowitz, pp 68–90. Lanham, MD:
Scarecrow Press.
Housel, T. J., V. A. Kanevsky, W. Rodgers, and W. Little. May 28, 2010. “The Use of Mod-
ern Portfolio Theory in Non-Profits and Their IT Decisions.” In Refereed Proceedings
of the IC6: World Conference on Intellectual Capital for Communities in the Knowledge
Economy. Paris, France.
312 J. C. Mun, T. Housel
Kendall, T., and T. J. Housel. July 18–21, 2004. “Resolving the ROI on IT Issue (Best Ses-
sion Paper).” In Refereed Proceedings of the 8th World Multi-conference on Systemics,
Cybernetics, and Informatics. Orlando, FL.
Mun, J. 2006. Real Options Analysis: Tools and Techniques for Valuing Strategic Invest-
ments and Decisions (2nd edition). Hoboken, NJ: Wiley Finance.
Mun, J. 2010. Modeling Risk: Applying Monte Carlo Risk Simulation, Strategic Real
Options, Stochastic Forecasting, and Portfolio Optimization (2nd edition). Hoboken, NJ:
Wiley Finance.
OMB (Executive Office of the President; Office of Management and Budget). August
4, 1983, revised 1999. Circular N0. A-76. Washington DC 20503. Available at
http://www.whitehouse.gov/sites/default/files/omb/assets/omb/circulars/a076.pdf [last
accessed September 30, 2014].
Pavlou, P. A., T. J. Housel, W. Rodgers, and E. Jansen. 2005. “Measuring the Return on
Information Technology: A Knowledge-based Approach for Revenue Allocation at the
Process and Firm Level.” Journal of the Association of Information Systems 6: 199–226.
Rabelo, L., and T. J. Housel. May 20, 2008. “Estimating the Knowledge Value-Added
of Information Technology Investments.” In Refereed Proceedings of the Institute of
Industrial Engineers (IIE) Annual Conference. Vancouver, British Columbia, Canada,
pp. 1849–1854.
Rios, C., T. J. Housel, and J. Mun. 2006. “Real Options and KVA in Military Strategy
at the United States Navy.” In Modeling Risk: Applying Monte Carlo Simulation, Real
Options Analysis, Forecasting, and Optimization Techniques, Case Study, pp. 441–452.
Hoboken, NJ: Wiley Finance.
Rodgers, W., and T. J. Housel. 2006. “Improvement of Global Performance Measures
Related to Intangible Assets.” Collected Papers of the 32nd European International
Business Academy (EIBA) Conference (Refereed Proceedings). December 7–9, Fribourg
Switzerland.
12 Extensions of the Greenfield–Persselin
optimal fleet replacement model
Application to the Canadian
Forces CP-140A Arcturus Fleet
David W. Maybury

12.1 Introduction
Today’s Canadian Forces operate increasingly aging military platforms, often
retaining fleets for unprecedented long service lives (DND 2014). The unfamiliar
territory of operating platforms well beyond expected service lifetimes presents
the Canadian Department of National Defence (DND) with a central cost–benefit
analysis (CBA) problem: to replace a fleet of aging vehicles or to continue an
ongoing maintenance regime. As vehicles age, we expect that the overall oper-
ating and maintenance costs (O&M) will eventually reach a level prohibitive to
continued operation, thereby suggesting vehicle replacement as the best alterna-
tive. Aircraft platform replacement decisions arising primarily from aging effects
present a relatively new problem for global militaries. Most replacement decisions
in the past have occurred largely through new capability requirements, thus limit-
ing aging effect data. Only recently have selected platforms yielded sufficient data
to allow for comprehensive studies.
The United States Air Force (USAF) faces substantial challenges in operat-
ing airframes with exceptionally long service lives and, having recognized age
as a factor in O&M costs as early as the 1960s (Johnson 1962), the USAF
started comprehensive fleet lifetime O&M studies in the 1990s. While limited
data in early studies contributed to a confused literature in the 60s and 70s,
from Johnson (1962) to Marks and Hess (1981), large datasets and the applica-
tion of more sophisticated analytical methods available to the USAF in recent
years has enabled investigators to draw the cautious conclusion that age effects
impact O&M costs. In particular, it was demonstrated in Ramsey et al. (1998)
that heavy-maintenance workloads have increased with the chronological age of
the KC-135 tanker aircraft, and an earlier RAND study (Hildebrandt and Sze
1990) estimated that, for every year increment in the age of a USAF mission
design series, O&M costs increase on average by 1.7 perent. Further studies on
age effects using commercial airline data (Ramsey et al. 1998; Dixon 2005), data
from US Navy aircraft (Johnson 1993; Jondrow et al. 2002; Stoll and Davis 1993;
Francies and Shaw 2000) and those of the USAF (Hildebrandt and Sze 1990;
Kiley 2001; Pyles 2003; Keating and Dixon 2003) in the areas of workloads,
material consumption, repairs per flight hour, mean time between failures, and
314 D. W. Maybury
program depot maintenance all show positive growth with age. Investigators warn
(Pyles 2003) that changing accounting practices, budget sluggishness, and rela-
tively fixed maintenance-personnel requirements plague USAF and USN studies,
creating a difficult environment in which to extract age effects from other pres-
sures. Raymond Pyles’s 2003 RAND corporation investigation for the USAF on
the impact of aging airframes on maintenance represents the most comprehen-
sive study to date (Pyles 2003). Using regression methods that address several
issues simultaneously, Pyles discovered a positive relationship between mainte-
nance requirements—in nearly all activities—with airframe age. Not only did
Pyles find that different maintenance tasks realize different age effects for the same
airframe, but he also found that age effects correlate with aircraft complexity and
fly-a-way costs.
In light of growing evidence that suggests age impacts maintenance and hence
O&M costs, DND requires a strategy to determine the optimal replacement time
for a fleet of vehicles after which it becomes increasingly disadvantageous both
economically and operationally to maintain the fleet. As a case study for construct-
ing and implementing a strategy for fleet replacement, we consider the CP-140A
Arcturus maritime surveillance aircraft. Acquired from the Lockheed Aircraft
Corporation by DND in 1993, the CP-140A is the sister aircraft of the CP-140
Aurora, with both airframes based on the P-3 Orion. The Arcturus fleet con-
sists of three vehicles based with 14 Wing at CFB Greenwood and, unlike the
Aurora, the Arcturus does not possess anti-submarine warfare capability. Today,
the Arcturus finds a primary role in pilot training, drug trafficking interdiction,
search-and-rescue, and sovereignty patrols.
As part of the Canada First Defence Strategy announced in 2008, the Gov-
ernment of Canada has undertaken a significant renewal process for Canada’s
military. In late 2008, the Assistant Deputy Minister (Materiel) and the Director
General Aerospace Equipment Program Management sought operations research
input from the Centre for Operations Research and Analysis (CORA) on deci-
sions surrounding the CP-140A Arcturus and the CP-140 Aurora fleets. The
Government of Canada had recently altered the decade-long Aurora Incremental
Modernization Program (AIMP) for CP-140 life extension to help position DND
for the eventual acquisition of a new long-range maritime patrol surveillance air-
craft. The Assistant Deputy Minister (Materiel) wished to better understand the
expected service life of this platform.
Since acquisition, the CP-140A Arcturus fleet has seen annual O&M costs per
vehicle increase dramatically since procurement. Discerning pure age effects from
the upward spiralling annual O&M costs presents a highly non-trivial problem.
Instead of attempting to isolate each age effect from usual inflationary pres-
sures, operational tempo, budget processes, training programs, and random events
(such as the discovery of unanticipated maintenance problems), we apply the
Greenfield–Persselin model (Greenfield and Persselin 2002, 2003) by assuming
that random events occur independently of airframe age, but that the aircraft’s
resiliency diminishes with time. Our position maintains that random events trigger
O&M costs that evolve proportionally—the aircraft becomes more susceptible to
Optimal fleet replacement 315
damage as it ages. By focusing on economic considerations that include tradeoffs
among costs, operational availability, and the effects of uncertainty, we find an
optimal replacement time in a stochastic setting.
In addition to addressing fleet replacement issues, the model we apply also con-
tains forecasting capabilities, yielding insight into future replacement decisions.
Most importantly—and we cannot overstress this point—the model implicitly
attaches an option value to the capacity to delay a decision. Models based solely on
deterministic net present value (NPV) arguments implicitly assume that a military
investment choice is either reversible, in that the military can cancel the invest-
ment if circumstances become unfavourable, or if irreversible, that the military
investment choice occurs as a now or never proposition. Deterministic approaches
oversimplify military fleet replacement decisions by not placing an option value
on the capacity to delay.
In this paper, we model the CP-140A Arcturus O&M costs as a geometric ran-
dom walk with a stochastic discount factor represented by operational availability
(Ao ). We apply the techniques of dynamic programming to calculate an optimal
replacement time by translating the Greenfield–Persselin model into an explicit
first passage time problem. The model we construct represents a prototype for a
fleet replacement strategy using data from the CP-140A Arcturus as a case study.
The methods we use build on the Greenfield–Persselin model and we apply the
model directly to help inform Canadian military leadership on fleet-wide decisions
regarding Canada’s maritime surveillance capabilities.
We organize the paper in four parts. Following the introduction, we present
the Greenfield–Persselin model in Section 12.2. Section 12.3 applies the model to
the CP-140A Arcturus data and provides a mean first passage time approach to the
replacement problem. Finally, in Section 12.4, we present a discussion concerning
military impact and future research.

12.2 The Greenfield–Persselin model


Empirical evidence suggests a positive relationship between the age of military
platforms and their O&M costs (Johnson 1962; Nelson 1977; Marks and Hess
1981; Ramsey et al. 1998; Hildebrandt and Sze 1990; Dixon 2005; Johnson 1993;
Stoll and Davis 1993; Francies and Shaw 2000; Jondrow et al. 2002; Kiley 2001;
Pyles 2003; Keating and Dixon 2003). From a modeling point of view, the most
important feature emerging from RAND studies suggests that “many of the prob-
lems associated with aging material have emerged with little or no warning”
(Pyles 2003). Thus, the RAND studies lead us to consider random shocks in addi-
tion to constant growth in O&M costs. Based on the ideas from the literature
on forest harvesting and renewable resource management, Greenfield and Pers-
selin suggest an infinite horizon generational replacement model in the presence
of stochastic noise.
Greenfield and Persselin view the problem of fleet replacement similarly to the
problem of deciding to harvest a forest for commercial purposes. As each timber
stand grows, it gains in value and, at some point in the future, the forester will
316 D. W. Maybury
make the decision to harvest the stand followed by replanting. Thus, over an infi-
nite time horizon the forester faces an infinite succession of harvest-replanting
cycles. The forester must determine the holding period before harvesting. If all
else remains equal over each interval, the first optimal holding period will be iden-
tical to all other periods. The military faces a similar problem in that, as a military
platform ages, its O&M costs increase, eventually leading to a necessary replace-
ment. We can imagine that the military enters a series of repair-replace cycles and,
like the forester, the military must determine the optimal replacement time.
In a deterministic universe in which we exactly know the fleet O&M costs evo-
lution function, Greenfield and Persselin construct a simple fleet replacement time
model. Let us assume that a military platform has an acquisition cost, p, in an
environment of a continuously compounding interest rate discount, r , with an
instantaneous (known) O&M cost rate, m(t). The variable t denotes the chrono-
logical age of the platform. We assume that the military looks to minimize total
ownership costs, thereby determining an appropriate replacement time. The NPV
life-cycle cost of a single fleet generation as a function of time reads
 s
NPV(s) = p + dt m(t) exp(−r t), (12.1)
0

where s denotes the replacement time. We now imagine that the replacement pro-
cess repeats ad infinitum. Like the simplified forester’s problem, if all else remains
equal, each generation faces the same O&M cost structure. Thus, we can write the
total ownership cost of successively owning a series of fleets as


  s
c(s) = p+ dt m(t) exp(−r t) exp(−r si ), (12.2)
i=0 0

where s denotes the replacement time of each generation. The total ownership
cost, Equation (12.2), represents a converging geometric series yielding
s
p+ dt m(t) exp(−r t)
c(s) = 0
. (12.3)
1 − exp(−r s)

Differentiating this expression with respect to the replacement time, s, we find that
the military minimizes total ownership cost by applying the first order condition
which yields
 ∗ 
s
0
dt m(t) exp(−r t) + p exp (−r s ∗ )
m(s ∗ ) = r p + r . (12.4)
1 − exp (−r s ∗ )

As an example, if we know that the instantaneous O&M cost rate has the form
m(t) = b exp(αt), then we find that
Optimal fleet replacement 317

b(1 − exp(−(r − α)s)


p+
c(s) = r −α , (12.5)
1 − exp(−r s)

and

b(1 − exp (−(r − α)s ∗ ))


+ p exp (−r s ∗ )
∗ ∗
m(s ) = b exp(αs ) = r p + r r − α .
1 − exp(−r s ∗ )
(12.6)

The replacement time, s ∗ , can be found numerically.


While the deterministic model version of the Greenfield and Persselin model
assumes that the military replaces the fleet with exactly the same vehicle at exactly
the same price in constant dollars, we caution that the model does not assume this
situation reflects reality. The model simply determines the total ownership cost
associated with the specific vehicle under consideration given an infinite number
of generations. The Greenfield–Persselin infinite horizon generational replace-
ment model determines the total ownership cost curve associated with the fleet
under consideration. In this regard, the model determines the length of the holding
period for each generation in the infinite series.
As Greenfield and Persselin elucidate, the problem with the model outlined
thus far stems not from its lack of a replacement option comparison, but from the
assumption of an exactly known O&M cost evolution. While the model recog-
nizes the irreversibility of the replacement decision, it implicitly assumes that the
decision to replace occurs as a now or never proposition. According to the solu-
tion given in Equation (12.6), for every year the fleet is retained past t = s ∗ , the
military incurs a known marginal cost. Thus, the model demands replacement at
exactly time t = s ∗ , the now or never situation. In reality, we do not know the actual
O&M cost evolution function which is further complicated by random shocks to
the fleet. Since the costs evolve stochastically, there is a chance that the follow-
ing year will see a decrease in O&M costs from random occurrences. The chance
that the O&M costs can actually decrease in the following year—no matter how
small that decrease might be—must factor into our decision to replace the fleet.
We require a model that places an option value on the capacity to wait.
Greenfield and Persselin made three key insights into the deterministic replace-
ment model. They noticed that Equation (12.5) contains three important factors:

• the constant p in the numerator represents the part of the total ownership cost
associated with the initial procurement;
• the second term in the numerator represents a NPV contribution to the total
ownership cost arising from O&M activity; and
• the denominator represents a discount factor.
318 D. W. Maybury
Greenfield and Persselin promote the O&M cost function, the NPV, and the
discount factor of Equation (12.5) to a stochastic variable and use dynamic pro-
gramming to calculate the expected replacement time. By calculating the expected
discount factor and the expected NPV at the expected replacement time, Green-
field and Persselin produce a boundary in the O&M costs which signals the
optimal replacement point (Greenfield and Persselin 2002, 2003).
While the Greenfield–Persselin model strictly uses O&M costs as the random
process that triggers replacement, we feel that the strict reliance on monetary
considerations masks the military’s true objective: maximize a military advan-
tage at the best possible cost. By relying only on pure monetary issues, the
Greenfield–Persselin model can call for long holding periods even in cases where
the fleet fails to provide the military with operational availability. A similar prob-
lem occurs in economic modeling. Few people value money for its own sake, but
rather value the consumption possibilities that wealth represents. For this reason,
economic modeling usually resorts to utility functions based on consumption as
opposed to wealth itself. We extend the Greenfield–Persselin model by including
a stochastic discount factor with the O&M cost process. Determining an objective
value function that attempts to measure degrading capability will not only prove
exceedingly difficult to construct, but will also destroy the attractive feature of the
Greenfield–Persselin model in that the fleet is always compared to itself. We pro-
pose using operational availability (Ao ) as the stochastic discount factor. By using
Ao , we penalize the fleet for failing to provide the military with a vehicle deemed
necessary even if the lack of availability reduces costs.
We model the O&M costs per Ao (m u ) as a geometric random walk

dm u = αm u dt + σ m u d W (t), (12.7)

where d W (t) is the Weiner process, α gives the geometric growth rate, and σ sets
the strength of the driving uncertainty. Following the Greenfield–Persselin model,
we can calculate the expected replacement time, the expected discount factor at
replacement, and the expected NPV at replacement using dynamic programming.
Each part can be used to assemble an expected total ownership cost function. The
expected replacement time, a(m ∗u , m u ), satisfies the Bellman relation
 
da(m ∗u , m u ) = dt + a(m ∗u , m u ) + E da(m ∗u , m u ) , (12.8)

where m ∗u is the O&M cost per Ao at replacement. (Mathematical details appear in


Appendix 12.1.) Notice that the replacement time is no longer a function of time,
but rather a function of O&M costs per Ao . The solution to Equation (12.8), given
the boundary condition a(m ∗u , m ∗u ) = 0, reads
 

ln m ∗u /b
a(m u , b) = , (12.9)
α − (1/2)σ 2
where m u (t = 0) = b, the initial O&M costs per Ao . We see that volatility in the
O&M costs per Ao process, represented by the parameter σ ≥ 0, increases the
Optimal fleet replacement 319
lifetime of the fleet relative to a deterministic approach. The model values the
capacity to wait in an uncertain environment. We will return to this idea in the
following section. Applying the Greenfield–Persselin model, the expected total
cost per Ao function, c E (m u ), becomes
 β1 −1
p+ b
r−α
1 − mbu
c E (m u ) =  β1 , (12.10)
1 − mbu

where

2
α 1 α 1 2r
β1 = − − + − + , (12.11)
σ2 2 σ 2 2 σ2

and using the first order condition we find that

 β1 −1  β1
b
1 − mb∗ + p mb∗
β1 β1 r−α u u
m ∗u = (r − α) p + (r − α)  β1 ,
β1 − 1 β1 − 1
1 − mb∗
u

(12.12)

which determines the O&M per Ao level at replacement.


While Greenfield–Persselin focus on the expected fleet replacement time, the
model can be reinterpreted entirely as a first passage time problem to the barrier
m ∗u , which is set by Equation (A12.1.22). By focusing on a first passage time inter-
pretation of the problem, along with the use of a stochastic discount factor, we find
that the Greenfield–Persselin model class represents an excellent tool in assessing
fleet lifetimes and in providing military leadership with a decision window for
an aging fleet. We demonstrate the power of this model class with the CP-140A
Arcturus fleet.

12.3 An application: the CP-140 Arcturus


Using the data obtained through the DND Cost Factors Manual and the Canadian
Forces fleet data base PERFORMA, we analyze the CP-140A Arcturus for opti-
mal replacement in the presence of stochastic noise. The dataset contains 13 years
of data (1995–2007) on each of the three aircraft that comprise the fleet, including
acquisition costs. The O&M costs include petroleum, oil, lubricants, engineer-
ing services, repairs, overhauls, and sparing costs. Since any amortization costs
connected to the fleet do not involve the disbursement of funds, and since usual
military activity in the form of salaries or unit support will require funding inde-
pendent of vehicle age, we do not include costs associated with these effects
in O&M.
320 D. W. Maybury
Based on RAND studies, we take the position that fleet resiliency diminishes
with time (e.g., metal fatigue, increased chance of corrosion discoveries), and
hence problems triggered by random events—which occur independently of the
CP-140A’s age—compound. We apply the Greenfield–Persselin model to cap-
ture these effects with a stochastic discount factor represented by Ao . Finally,
we assume a constant interest rate discount of 4.40 percent, which provides a
good approximation to a risk-free interest rate in Canada over the period in
consideration (see Chapter 17). We keep the discount rate fixed throughout the
study.
To summarize our model input assumptions, we have:

• effects of new technology, modernization and replacement choice are not


considered;
• salaries supporting DND civilian or military personnel are not included in
O&M;
• direct squadron support costs are not included in O&M;
• amortization costs are not included;
• a fixed discount rate of 4.40 percent;
• the fleet becomes less resilient to random events with time;
• random events occur independently of vehicle life.

We do not include the actual dataset for the CP-140A Arcturus in this report.1
Instead we provide a normalized dataset that sets the initial O&M per Ao at 4.4 in
dimensionless units. We apply a standard likelihood analysis technique to estimate
the model parameters of the geometric random walk in Equation (12.7), yielding
(at two standard deviations)

α̂ = 0.2 ± 0.1 σ̂ = 0.4 ± 0.1, (12.13)

which generates an estimate for the O&M per Ao barrier of

16 < m ∗u < 39. (12.14)

Notice that we have large uncertainty in the estimated parameters. This reflects
the limited costing data we have CP-140A fleet—we have a small sample size.2 In
Figure 12.1, we see the O&M per Ao sample path (in normalized units) with the
calculated barrier and its associated uncertainty represented by a hatched bar. The
first passage interpretation becomes immediately clear in the presence of real data
(normalized in this case). The uncertainty in the estimated barrier, m ∗u , implies
uncertainty in the expected replacement time. Since the expectation is calculated
conditionally on parameter estimates, the expectation itself becomes uncertain.
To capture the subtleties associated with uncertainty in the parameter estimates,
we treat the problem as a first passage of time to the region delineated by the
upper and lower estimate bounds of m ∗u . We interpret the region between the two
boundaries as a critical region for the decision-maker. Once the process enters the
Optimal fleet replacement 321

95% confidence contour

critical region
100
O&M costs/Ao

10

1
5 10 15 20 25
Years after procurement

Figure 12.1 CP-140 Arcturus data with the critical region indicated.

critical region, the model calls for the decision-maker to consider a replacement
or an overhaul initiative.
By interpreting entry of the critical region as a flag for initiating a decision
process, we see in Figure 12.1 that the model calls for replacement consideration
of the CP-140A Arcturus at year ten. Using Equation (12.8), the expected time to
enter the critical region from the first year of data is at year 11. The uncertainty
envelope gives the 95 percent confidence band for the future sample path.
The application of the original Greenfield–Persselin model, which does not use
a utility discount, yields Figure 12.2. Without the utility discount, and again using
standard likelihood techiques, we find the parameter estimates

α̂ = 0.1 ± 0.1, σ̂ = 0.2 ± 0.1, (12.15)

with barrier

9 < m ∗u < 20. (12.16)

The expected time to the critical region jumps to 23 years, representing more than
a 100 percent increase over the model without the stochastic discount factor. We
see that the effect of a stochastic discount factor dramatically changes the predic-
tions of the model. Again, we display the uncertainly envelope at the 95 percent
confidence band.
322 D. W. Maybury

100

95% confidence contour


critical region
O&M costs

10

1
5 10 15 20 25 30
Years after procurement

Figure 12.2 CP-140 Arcturus data with no stochastic discount factor shown with the
critical region indicated.

Given the current level of the O&M per Ao process, the first passage of time
interpretation allows us to calculate the expected sojourn time within the critical
region, and the probability that the process will exit through the upper boundary
before the lower boundary. Both pieces of information provide a clearer picture of
the decision problem to military leadership. The sojourn time calculation can give
decision-makers an estimate of the amount of time remaining before the model
definitively calls for fleet replacement or a major overhaul.
To calculate the expected time of a drift diffusion process within a region, we
can use the backward Fokker-Planck equation (see Gardiner 2003 for details),
which for our model yields the differential equation for the expected time as

d T (m̃ u ) 1 d 2 T (m̃ u )
k1 + k2 = −1, (12.17)
d m̃ u 2 d m̃ 2u

where m̃ u represents the log-transformed process, a < m̃ u < b, T (a) = T (b) = 0,


and where in our model k1 = (1/2)σ 2 and k2 = α − (1/2)σ 2 . In fact, we can use
Equation (12.17) to derive the Greenfield–Persselin expected replacement time
result, Equation (12.8), by considering m̃ u ∈ (−∞, m̃ ∗u ], where m̃ ∗u represents the
log-transformed boundary. Thus, the Greenfield–Persselin model is in fact a first
passage time problem in the log-transformed space and it is this characterization
that bears fruit in the presence of real data. Once the fleet enters the critical region,
Optimal fleet replacement 323
using Equation (12.17), the expected sojourn time reads
⎡  −k2 Ũ −k2  ⎤
m̃ u
1 ⎢ (Ũ − L̃) e −e
k k
1 1

T (m̃ u ) = ⎢ ⎣ −k −k
− m̃ u + Ũ ⎥
⎦, (12.18)
k2 2 L̃ 2 Ũ
e 1 −e 1
k k

where L̃, and Ũ denote the log-transformed upper and lower boundaries of the crit-
ical region respectively. We find that the expected sojourn time within the critical
region at year 11 for the CP-140A Arcturus in Figure 12.1 is 0.8 years.
As an aside, we can also calculate the probability that the O&M per Ao process
will exit through the upper barrier of the critical region before the lower barrier by
once again invoking the backward Fokker-Planck equation. In this case, we have

d 2 πU (m̃ u ) dπU (m̃ u )


k1 2
+ k2 = 0, (12.19)
d m̃ u d m̃ u

where πU (m̃ u ) represents the probability of exit through Ũ before L̃ with the
conditon that

πŨ (m̃ u ) + π L̃ (m̃ u ) = 1, (12.20)

and the condition that πŨ (Ũ ) = 1 and πŨ ( L̃) = 0. The probability that the process
will exit through the upper boundary Ũ before the lower boundary is given by the
solution to Equation (12.19), namely,
−k2 −k2
k1 m̃ u k1 L̃
e −e
πŨ (m̃ u ) = −k2 −k2
, (12.21)
Ũ L̃
e k1 −e k1

−k2 −k2
m̃ L̃
e k1 u −e k1
π L̃ (m̃ u ) = 1 − −k2 −k2
, (12.22)
Ũ L̃
e k1 − e k1

where m̃ u gives the current log-transformed value of the O&M per Ao process. The
probability of exiting through the top of the critical region before the lower bound-
ary can give the decision-maker insight into delaying the procurement initiative
once the process has entered the critical region.
In addition to calculating the sojourn time and probabilities of exit, we can also
calculate the distribution of the hitting time to any level. In particular, since we are
working in the log-transformed space, the probability density function for the first
passage time to level L̃ is given by
# $
( L̃ − m u ) ( L̃ − m u − (α − 1/2σ 2 )t)2
p( L̃, m u ; t) = √ exp − , (12.23)
σ 2πt 3 2σ 2 t
324 D. W. Maybury

0.08

0.07

0.06
Probability density

0.05

0.04

0.03

0.02

0.01

0
0 5 10 15 20 25 30 35
Years after procurement

Figure 12.3 CP-140 Arcturus data hitting time distribution.

where α and σ are the growth and volatility parameters of the model. In
Figure 12.3 we show the distribution of the hitting time to the lower barrier in
Figure 12.1 from the first year of data. The hatched region in Figure 12.3 denotes
the 60 percent confidence region around the expectation of 11 years. The hitting
time calculation allows decision-makers to estimate the probability of reaching
a given level which can shed light on the risks of holding the fleet longer than
anticipated.

12.4 Discussion and conclusion


The stochastic discount extended Greenfield–Persselin model applied to the CP-
140A Arcturus fleet replacement problem captures the random features in costs
and performance that systems exhibit over long periods of time. Through dynamic
programming the model extracts a critical region which delineates the opti-
mal replacement interval. We note that replacement under uncertainty contains
non-intuitive features, such as the lengthening of the expected time of replace-
ment relative to the deterministic approach. Uncertainty plays a large role in
determining the optimal replacement time.
In 2009, DND used this model as a decision input tool for restructuring the CP-
140A Arcturus and CP-140 Aurora fleets until their expected replacement in the
early 2020s. Our principle findings on the service life of CP-140A Arcturus fleet
included the following points.
Optimal fleet replacement 325
• Including the stochastic discount factor, the CP-140A entered the critical
region for decision at year 11 (calendar year 2006).
• If we ignore the stochastic discount factor, the expectation time for replace-
ment/overhaul jumps by more than 100 percent to 23 years.
• Under the assumption that the Ao represents an appropriate discount factor,
the model calls for the initiation of a replacement/overhaul decision process
for the CP-140A Arcturus fleet.

We should not view the critical region solely in terms of fleet replacement. The
critical region simply denotes the place in O&M per Ao space where the expected
total ownership costs begin to realize an expected marginal cost. We imagine that
several external variables, such as queuing problems or sparing issues, can cre-
ate an adverse growth environment and corrections in these areas might radically
drive the sample path away from the critical region in a non-stochastic manner.
Furthermore, we might find it advantageous to undertake a reset-the-clock over-
haul rather than acquiring a new fleet once the sample path has drifted into the
critical region. Ideally, we see the stochastic model presented in this paper used in
conjunction with other decision tools.
We have treated the O&M cost per Ao as a single stochastic variable in a
drift-diffusion model. The statistical tests we performed on the data gave us no
compelling reason to reject this approach. In reality, the O&M costs and Ao both
evolve stochastically and potentially feed back on one another. Thus, we can
imagine the fleet replacement/overhaul problem as a coupled system of stochastic
differential equations with O&M cost following a geometric random walk, and the
Ao tracking a different process, such as a mean reverting process (e.g. an Ornstein–
Uhlenbeck process with boundaries). We might also advance the model by using a
stochastic interest rate with an appropriate industry specific military deflator (see
Chapter 17). Finally, if military consensus is available, the stochastic discount
factor could be promoted to a nonlinear function of Ao with the inclusion of poten-
tially other relevant performance and economic measures. In particular, since the
fleet will require routine maintenance work, we might imagine a stochastic dis-
count function of Ao that turns on slowly as Ao diminishes. We view this area of
research as an exciting possibility.
The application of the Greenfield–Persselin model and the model extension pre-
sented above rest on the ability to capture the statistical properties of the sample
path in the data. Describing the sample path as a geometric random walk rep-
resents an approximation. If the data supports it, we can use more sophisticated
models, such as jump diffusion (or more general Lévy processes). These modifica-
tions would require a new first passage time calculation along with a re-calculation
of the critical region. Again, this represents exciting opportunities for further work.
However, as the O&M cost per Ao is a price process that evolves over time, a geo-
metric random walk is a good starting place. If on examination of the data, the
sample path cannot be approximated by a geometric random walk, new meth-
ods will be required. Finally, as in all operational research studies, this model is
an approximation of reality that endeavors to provide decision support—in this
326 D. W. Maybury
case a military cost–benefit analysis. The overall replacement/overhaul decision
will require additional input, including examination of replacement alternatives,
budget feasibility, and military obsolescence.
The CP-140A Arcturus fleet represents a special case in that the homogenous
fleet consists of only three aircraft all delivered to the military at the same time.
The application of the model to the CP-140A did not include complications from
multiple aircraft variants, aircraft with different service lives, or aircraft that have
seen life extension modifications. More complicated fleets will require special
treatment of these circumstances to apply the Greenfield–Persselin model more
generally.
Operational tempo, unforeseen engineering and maintenance problems, and
budgeting processes all factor into the evolution of O&M costs of vehicle fleets.
Separating specific variables that directly relate to age from other effects presents
a daunting task. By treating the problem stochastically under the assumption
that fleet resiliency diminishes with time, we can determine the critical time to
consider fleet replacement. Most importantly, the stochastic approach implicitly
attaches an option value to the capacity to delay in making the replacement
decision—as more information becomes available to the decision-makers, bet-
ter choices can be made. Deterministic methods that focus only on marginal
costs or NPV inputs ignore this important feature. The Greenfield–Persselin
model, with a stochastic discount factor, balances economic considerations, uncer-
tainty, and fleet performance in establishing a replacement time. We feel that
the Greenfield–Persselin model class represents an important step in under-
standing optimal fleet replacement times in military life cycle cost–benefit
problems.

Appendix 12.1: The mathematical details of the


Greenfield–Persselin model
The detail presented here follows Greenfield and Persselin (2002) closely. In Equa-
tion (12.5), which gives the relationship between total ownership costs and the
replacement interval, we see that the numerator contains an initial procurement
cost and an NPV part, and the denominator represents a discount factor. To pro-
mote the relationship to its stochastic counterpart, we need the expected age at
replacement, the expected NPV at replacement, and the expected discount factor at
replacement. Once we have these three expressions we can rewrite Equation (12.5)
in its stochastic setting.

Expected age
First, we imagine a pre-determined cut-off cost per utility (perhaps set indepen-
dently by policy) m ∗u , above which DND is not willing to pay. Once we find
m u > m ∗u , we consider fleet replacement or a reset-the-clock overhaul. We imag-
ine that the process starts at initial condition b below m ∗u and, using Bellman’s
recursive equation from the last section, we see that the expected age, a(m ∗u , m u )
Optimal fleet replacement 327
evolves as

a(m ∗u , m u ) = dt + (a(m ∗u , m u ) + E(da(m ∗u , m u ))). (A12.1.1)

Note that Equation (A12.1.1) contains two parts: an additive time increment in
the expected age from the current time increment, and a term that represents
the expected age at replacement resulting from the time increment. After some
manipulation, we can re-express Equation (A12.1.1) as

E(da(m ∗u , m u ))
+ 1 = 0; (A12.1.2)
dt
Expanding the differential using Ito’s Lemma we find

1 2 2 d 2 a(m ∗u , m u ) da(m ∗u , m u )
E(da(m ∗u , m u )) = σ mu + αm u , (A12.1.3)
2 dm 2u dm u

where we have used E(d W ) = 0. Thus, we have the differential equation for the
expected age of replacement as

1 2 2 d 2 a(m ∗u , m u ) da(m ∗u , m u )
σ mu + αm + 1 = 0. (A12.1.4)
2 dm 2u dm u

We should note that, while we obtained Equation (A12.1.4) through dynamic pro-
gramming, we could have obtained the same result from a mean first passage time
approach. If we regard the cutoff, m ∗u , as a barrier, the problem amounts to find-
ing the expected time for m u to diffuse from some initial starting point to the
barrier. The stochastic formalism requires the backward Fokker-Planck equation
(Gardiner 2003) and, in this case, results in Equation (A12.1.4). Thus, we can
approach the replacement problem from either perspective, and we will use either
formalism as appropriate. Since m u = 0 is an absorbing point in the geometric
random walk, the solution to Equation (A12.1.4) reads

a(m ∗u , m u ) = A + B ln(m u ). (A12.1.5)

If we assume the initial condition m u (t = 0) = b along with the boundary condition


that the expected age at m ∗u vanishes, a(m ∗u , m ∗u ) = 0, we find the expected time for
platform replacement is
 

ln m ∗u /b
a(m u , b) = . (A12.1.6)
α − (1/2)σ 2

Again, we can obtain the same result from calculating the mean first passage time
of a(m ∗u , m) from b through the barrier m ∗u .
Equation (A12.1.6) connects with the deterministic approach since setting σ = 0
recovers the deterministic result. In the deterministic case, we find the exact age
328 D. W. Maybury
at which to replace the vehicle as a function of the pre-determined boundary, m ∗u ,
namely

1  ∗ 
a(m ∗u , b) = ln m u /b . (A12.1.7)
α
The deterministic case determines m ∗u through the minimization of the total own-
ership cost function. Thus, we see that uncertainty in the system, characterized by
σ , demands that we hold the platform longer than the deterministic result. We see
that the stochastic approach implicitly values the capacity to wait by recognizing
that there is a chance that next year’s O&M cost per utility will decrease. The
term, σ , reflects the size of this chance and thereby increases the waiting time. We
also notice that in order for the Equation (A12.1.6) to make sense, we must have
α − (1/2)σ 2 > 0, otherwise the expectation time to cross the barrier m ∗u becomes
infinite (implying that it will never be advantageous to replace the vehicle).

Expected discount factor


The expected discount factor at replacement calculation proceeds in a similar fash-
ion to the expected age calculation except that we no longer have an additive
increment since the discount factor appears multiplicative in the total ownership
relation. Thus, we have the following Bellman equation for the expected discount
factor:
 
ρ(m ∗u , m u ) = exp(−r dt) ρ(m ∗u , m u ) + E(dρ(m ∗u , m u )) . (A12.1.8)

Rearranging, we find

E(dρ(m ∗u , m u )) + (1 − exp(r dt))ρ(m ∗u , m u ) = 0 (A12.1.9)

and, with the use of Ito’s Lemma, we have


1 2 2 d 2 ρ(m ∗u , m u ) dρ(m ∗u , m u )
E(dρ(m ∗u , m u )) = σ mu + αm u dt = 0.
2 dm 2u dm u
(A12.1.10)

Altogether, we find that the expected discount factor obeys

1 2 2 d 2 ρ(m ∗u , m u ) dρ(m ∗u , m u )
σ mu + αm u − rρ(m ∗u , m u ) = 0. (A12.1.11)
2 dm 2u dm u

Equation (A12.1.11) is a Euler-Cauchy differential equation and therefore has the


solution

ρ(m ∗u , m u ) = Am βu 1 + Bm βu 2 . (A12.1.12)
Optimal fleet replacement 329
Substituting our solution into Equation (A12.1.11) we obtain β1 , β2 as the roots of
the quadratic equation

1 2
σ (β − 1)β + αβ − r = 0, (A12.1.13)
2

namely,


α 1 α 1 2 2r
β1,2 = − − ± − + 2. (A12.1.14)
σ2 2 σ2 2 σ

The boundary condition

ρ(m ∗u , 0) = 0, (A12.1.15)

demands that the coefficient associated with negative root vanishes, and the
boundary condition

ρ(m ∗u , m ∗u ) = 1 (A12.1.16)

implies that A = (m ∗u )−β1 . Thus, the full solution for the expected discount factor
at replacement, in light of the boundary conditions, reads
β1
mu
ρ(m u ) = . (A12.1.17)
m ∗u

Expected net present value


We can approach the expected NPV of the life-cycle cost similarly. In this case,
the recursive Bellman equation contains an additive term as the O&M cost per
utility accrues over time:

 
M(m ∗u , m u ) = m u dt + exp(−r dt) M(m ∗u , m u ) + E(d M(m ∗u , m u )) .
(A12.1.18)

Applying Ito’s Lemma once more, we find the differential equation

1 2 2 d 2 M(m ∗u , m u ) d M(m ∗u , m u )
σ mu + αm u − r M(m ∗u , m u ) + m u = 0.
2 dm 2u dm u
(A12.1.19)

We have an inhomogeneous Euler-Cauchy differential equation, and applying the


boundary conditions M(m ∗u , m ∗u ) = 0 (i.e. the life-cycle cost goes to zero as the
330 D. W. Maybury
cycle has ended) and M(m ∗u , 0) = 0, and including the starting value, b, at the
beginning of the cycle, we find
# β1 −1 $
b b
M(m ∗u , b) = 1− . (A12.1.20)
r −α m ∗u

If we now put the expected net present life-cycle value together with the
expected discount factor, we find the expected ownership cost curve as a function
of the cost:
 β1 −1
p + r−α
b
1 − mbu
c(m u ) =  β1 . (A12.1.21)
1 − mbu

Differentiating with respect to m u we find the transcendental equation for m ∗u , i.e.


the value of m u that minimizes the expected total ownership cost in the presence
of stochastic noise, as

 β1 −1  β1
b
1 − mb∗ + p mb∗
β1 β1 r−α u u
m ∗u = (r − α) p + (r − α)  β1 .
β1 − 1 β1 − 1
1− b
m∗u

(A12.1.22)

In the deterministic model, in which σ = 0, we find the analogous relation


 r/α−1  r/α
b
r−α
1 − mb∗ + p mb∗
u u
m ∗u = r p + r  r/α . (A12.1.23)
1 − m∗
b
u

Note that Equation (A12.1.23) conveys the same information as Equation (12.6)
since the deterministic case links the replacement age and the O&M cost per utility
in a one-to-one manner.
Comparison of Equation (A12.1.22) and Equation (A12.1.23) reveal the essen-
tial difference concerning the value of waiting. The deterministic case balances
the marginal costs and the known savings from delay through a small increment in
time. On the other hand, the stochastic case balances the marginal costs and sav-
ings from delaying until the O&M costs per utility increase by a small increment.
In the stochastic case, the time interval required to increment the O&M costs per
utility by a given amount is random.
Optimal fleet replacement 331
Notes
1 DND did not want the actual cost data included with this paper.
2 This paper elucidates an application of the Greenfield–Persselin model with extension to
a Canadian Forces fleet that was already undergoing a replacement decision process by
DND, and it was for this reason that we chose the CP-140A fleet.

References
Dixon, M. 2005. The Costs of Aging Aircraft: Insights from Commercial Aviation. RAND
Dissertation, Pardee RAND Graduate School.
DND (last access July 2014). Ch-124 Sea King: Intelligence, Surveillance and Recon-
naissance Aircraft. http://www.rcaf-arc.forces.gc.ca/en/aircraft-current/ch-124.page.
Department of National Defence 2, 1.
Francies, P. and G. Shaw 2000. Effect of Aircraft Age on Maintenance Costs. Briefing,
Center For Naval Analyses, Alexandria VA.
Gardiner, C. 2003. Handbook Of Stochastic Methods (3rd edition). Heidelberg:
Springer.
Greenfield, V. and D.M. Persselin. 2002. An Economic Framework For Evaluating Military
Aircraft Replacement. Santa Monica, CA: RAND MR-1489-AF.
——. 2003. “How Old Is Too Old? An Economic Approach To Replacing Military
Aircraft.” Defence and Peace Economics 14(5): 357–368.
Hildebrandt, G. and M. Sze. May 1990. An Estimation Of USAF Aircraft Operating And
Support Cost Relations. Santa Monica, CA: RAND, N-3062-ACQ.
Johnson, J. August 1993. Age Impacts On Operating And Support Costs: Navy Air-
craft Age Analysis Methodology. Patuxent River, MD: Naval Aviation Maintenance
Office.
——. June 1962. Forecasting The Maintenance Manpower Needs Of The Depot System.
Santa Monica, CA: RAND, RM-3163-PR.
Jondrow, J., R. Toast, M. Ye, J. Hall, P. Francis, and B. Measell. 2002. Support
Costs And Aging Aircraft: Implications For Budgeting And Procurement. CNA
Corporation.
Keating, E. and M. Dixon. 2003. Investigating Optimal Replacement Of Aging Air Force
Systems. Santa Monica, CA: RAND MR-1763-AF.
Kiley, G. 2001. The Effects Of Aging On The Costs Of Maintaining Military Equipment.
Washington, DC: Congressional Budget Office.
Marks, K. and R. Hess. 1981. Estimating Aircraft Depot Maintenance Costs. Santa Monica,
CA: RAND R-2731-PA&E.
Nelson, J. 1977. Life Cycle Analysis Of Aircraft Turbine Engines. Santa Monica, CA:
RAND, R-2103-AF.
Pyles, R. 2003. Aging Aircraft: USAF Workload And Material Consumption Life Cycle
Patterns. Santa Monica, CA: RAND MR-1641.
Ramsey, T., C. French, and K. Sperry. 1998. Aircraft Maintenance Trend Analysis. Briefing,
Oklahoma City Air Logistics Center, and Boeing.
Stoll, L. and S. Davis. July 1993. Aircraft Age Impact On Individual Operating And Support
Cost Elements. Patuxent River, MD: Naval Aviation Maintenance Office.
This page intentionally left blank
Part V

Selected applications
This page intentionally left blank
13 Embedding affordability
assessments in military
cost–benefit analysis
Defense modernization in
Bulgaria
Venelin Georgiev

13.1 Introduction
The economic reality of scarce resources combined with the human need for
security requires hard choices. The challenge is to make rational, informed and
effective decisions to meet existing needs with available resources1 in a way that
maximizes security. Defining, motivating, and supporting such decisions with mil-
itary cost–benefit analysis (CBA) is a complex process, the implementation of
which would be impossible without appropriate tools—quantitative and qualita-
tive analytical methods and models for evaluating the effectiveness of the use
of resources to meet the needs of people and society. The limited availability
and competing demands for financial resources on the one hand, and the exis-
tential human need for security on the other, requires careful resource allocation
decisions, and useful tools to justify and support the implementation of those
decisions.
The tools to manage defense resources include various management systems
(requirements generation, financial management, defense acquisition, etc.), and
involves making routine and reasonable decisions. The overall effectiveness of
defense spending depends on the resulting output, namely the security of society,
guaranteed by the defensive capabilities obtained as a product of careful manage-
ment of defense resources and innovative and technologically appropriate defense
investments (see Chapter 3). The efficiency and effectiveness of management deci-
sions depends on objective, scientifically sound and generally accepted military
CBA methods and models such as those described in previous chapters, applied in
practice by competent experts, managers and politicians.
Resource decisions are made within a process that in itself needs to be trans-
parent to decision-makers, e.g., to allow the preservation of a clear audit trail
from national security objectives, through defense objectives, to taxpayers’ money
(Tagarev 2009). Among the various questions related to resources management,
the most important are:

• how to create affordable, i.e. resource constrained plans;


• how to deal with uncertainty; and
336 V. Georgiev
• how to support the senior civilian leadership of a defense ministry in the exer-
cise of its authority and obligations as agents of democratic control of the
armed forces.

Decisions concerning required capabilities (“defense requirements”) are often


resource informed (i.e. assessed as realistic), but not necessarily resource con-
strained (i.e. fitting within defense budget forecast—see Chapter 7). But once
program decisions are made, the cost of the defense program for each future year
cannot exceed the defense budget forecast for the respective year (Tagarev 2006,
2009). This chapter discusses the case of a recently inducted member of NATO,
Bulgaria. An example of a military CBA is presented that involves the evaluation
of a new multi-role fighter for the Bulgarian Air Force.

13.2 Description of the current situation


The coherence and integration between resources and financial management sys-
tems does not always provide for effective management. For example, the US
Government Accountability Office (GAO) concludes that defense processes for
determining required capabilities, in the planning, programming and budget-
ing system (PPBS), and for developing, manufacturing and the acquisition of
defense products, are not sufficiently integrated. This leads to more defense invest-
ment programs and projects going forward than can be funded from the defense
budget (GAO 2009). As a result, there are significant risks and problems in try-
ing to achieve maximum military effectiveness with existing defense budgets.2
Additional analytical instruments—methods, models and other decision support
tools—are needed in order to overcome these deficiencies.
What are the empirical results from the implementation of expensive projects
to modernize the armed forces? In many cases investment projects do not yield
the intended effects in terms of defense capabilities, making it necessary to spend
additional resources to extend existing or start new acquisition projects. How-
ever, unforeseen cost increases in ongoing investment projects often reduce the
possibilities of funding new ones.
Another related shortcoming is the fact that delayed implementation of invest-
ment projects leads to delays in expected delivery of defense capabilities needed
by the armed forces to fulfill security missions and operations. The consequences
in terms of quality and effectiveness of project management outcomes, partly
derived from weaknesses in the traditional application of military CBA, are illus-
trated in a review of the investment portfolio of the US Department of Defense
(DoD) in 2007, according to which (GAO 2009):

• the total number of implemented investment programs is 95;


• the cost of research and development programs have increased unexpectedly
by 40 percent;
• the total costs of investment programs has increased by 26 percent;
• the absolute value of the total contingency programs is 295 billion dollars;
Affordability assessments 337
• the number of programs with more than 25 percent unexpected increase in
the total cost represents 44 percent of all investment programs; and
• the average delay in implementation of investment programs is 21 months.

The risk is that there exists an imbalance between the massive investment costs
required to modernize armed forces, and the quality and effectiveness of cur-
rent military CBA methods and models applied to the management of acquisition
projects, especially relating to the evaluation and selection of projects for future
funding.
Another example is provided by the recent experience of a new NATO member,
Bulgaria. The country suffered a significant cut in its 2010 defense budget both in
absolute terms and as a percentage of its gross domestic product (GDP). The new
defense capital budget amounted to only 0.8 percent of the total defense budget
in 2010, compared to an average of 20–22 percent in previous years. This sharp
reduction in the amount of investment funding, among other things, increased
demands on the decision-making process. It created a strong impetus to improve
financial and project management, and in particular to increase the application of
military CBA methods that incorporate careful economic analysis in the selection
and evaluation of force modernization projects.
Bulgaria (BG) faced many of the same problems experienced by the US
described above. For example, within a single planning cycle, cost estimates of
the country’s commitments to NATO defense and security in the form of force
goals increased several times, from the time commitments were originally made to
the actual assessment of costs in the programming and budgeting cycle. Adverse
effects from implementing such a mechanism to plan, evaluate and implement
defense programs and projects can be classified in several ways.
First, attempts to reduce investment costs can lead to the development of incom-
plete and therefore less effective investment projects.3 Here it is sufficient to
mention two examples of false savings. The acquisition of ships for the BG
Navy was made without appropriate energy units to provide power while docked.
Another example involves the purchase of training aircraft for the BG Air Force
that did not include the necessary computer simulator courses required to support
the initial training of pilots. The results of such oversights is that life-cycle costs
of defense investments increase unexpectedly, and/or that it becomes necessary to
launch new investment projects for the acquisition of additional equipment, train-
ing, services or supplies. Ultimately, this leads to lower quality investments that
reduce the desired capability (effectiveness) of the program.
Second, incentives matter. There is strong motivation for program managers to
complete investment projects, even if they prove costly, since cancelling them can
be a career killer. This behavior can bias cost estimates. If optimistic (lower) cost
estimates are systematically used in military CBAs, and future defense funding
constraints are not explicitly incorporated into the analysis, then more investment
projects are started than can realistically be financed by the defense budget. As
a consequence, in the implementation phase financial resources have to be trans-
ferred from one investment project to another and from one major program to
338 V. Georgiev
another, negatively impacting overall capabilities and the performance of defense
forces. Ideally, such inter-program decisions should be conducted ex-ante through
a properly structured CBA and not ex-post under duress.
It is often necessary to reduce the scope, or to suspend or eliminate invest-
ment projects to accommodate budget shortfalls. This was the experience of recent
projects for the acquisition of transport aircraft and helicopters by the BG Air
Force (BGAF), and for automotive equipment by the BG Land Forces. The ulti-
mate result of management decisions based on biased or incomplete analysis is a
delay in providing required defense capabilities, often accompanied by a reduction
in the level and quality of those capabilities.
The issue of reporting the feasibility of investment projects in terms of future fund-
ing realities is becoming more urgent. The 2010 White Paper on Defense and the
Armed Forces of Bulgaria declares the intention of Ministry of Defense (MoD) lead-
ership to develop a long-term investment plan incorporating all force modernization
projects aimed at building and maintaining required defense capabilities. Acknowl-
edging future funding constraints, the plan calls for the MoD to include investment
projects that are “financially feasible in the period until 2020” (MoD 2010).
Ideally, military CBAs would normally require investment plans that ensure
financial feasibility of modernization projects throughout their life cycles, and
thus a longer horizon of investment planning may be necessary. The designers of
these plans should take account of multiple risks, including those related to future
financing opportunities, the depreciation of the defense capital stock over time,
economic viability, the dynamics of innovation and technology, possible changes
in the security environment, and the commitments of the country to international
security (see Chapter 7).
Accepting that innovation is a “new idea that works” (Tzvetkov 2004), analysts
and practitioners need to look for opportunities to utilize military CBA to resolve
analytically challenging tasks in defense resources management, or to approach
existing analytical tasks in more effective ways. The new “real options” approach
introduced in Chapters 11 and 14 in this volume offer an example.4 No matter
whether innovation is of the “push” type, implementing ideas coming from science
and research, or “pulled” from needs identified in practice, the objective is to pro-
vide defense forces that maximize national (and collective) security of individuals
and society.
Implementation of innovative ideas can increase the competitiveness of the
defense sector in the contest for the allocation of public funds among areas such
as health, education, internal order, etc. The economic evaluation of alternatives
(EEoA) presented in Chapter 4 is an example of one such innovative method for
performing CBA in evaluating and selecting investment projects to modernize the
armed forces, under realistic budget scenarios.
One of the major differences between the traditional analysis of alternatives
(AoA) and EEoA is the content of the reported input variables in the analy-
sis (Figure 13.1). While examining operational effectiveness and the life-cycle
costs of defense investments (similar to a conventional AoA), EEoA adds a new
variable—the uncertainty of existing and projected future funding constraints.
Affordability assessments 339

METHOD
(tool)
INPUT OUTPUT
(variables) (results)

AoA – comparing cost and


Life Cycle Cost
effects of the examined
Operational alternatives and choosing
effectiveness Proposal for decision
the preferred one

Life Cycle Cost EEoA – generating possible


scenarios for funding and,
Operational based on that, comparing the
effectiveness examined alternatives and Proposal for decision
choosing the preferred one
Budget
limitations

Figure 13.1 Variables used in AoA and EEoA.

Uncertainty is represented by possible budget/funding scenarios, thus explicitly


incorporating the affordability of alternatives under examination. One possible
strategy is to incorporate budget scenarios for ‘optimistic,’ ‘pessimistic’ and ‘most
likely’ funding options.

13.3 Evaluation of a multi-role fighter for Bulgaria


Today, military CBAs offer a practical tool for decision support in capabilities-
oriented force planning processes to determine the optimal structure of the armed
forces, and for analyzing and selecting options based on defense budget realities,
risk management, etc. A valuable contribution is the development of various mea-
sures of effectiveness (MoE or “effects”) such as those discussed in Chapter 8 of
this volume.
In response to observed capability gaps defined on the basis of scenario
planning against current and future threats, planners identify weapons system
requirements, (as well as doctrine, training, readiness levels, etc.). From these
requirements, selected criteria are identified to help develop MoEs.
Consider the following example. The goal is to find the most suitable multi-role
fighter for the BGAF. The procurement of a new multi-role fighter for the BGAF
requires an evaluation of several possible alternatives using military CBA. The
process consists of several phases.
The first phase of any military CBA includes the definition of goals and objec-
tives and proper scoping of the analysis. Precise, clear and, if possible, quantitative
340 V. Georgiev
definitions of military goals and objectives can help to define program require-
ments and narrow the scope of the analysis. In this case the goal is to acquire
required operational capabilities (ROC) to perform BGAF missions and tasks
through the procurement of a new multi-role aircraft, or modernization of the cur-
rent fleet. The objective is to obtain necessary capabilities for the implementation
of new missions and tasks such as air superiority, air interdiction, close air support,
kill boxing, reconnaissance, etc.
The main factors that influence the scope of the analysis are the defined needs
and requirements that have to be met, the relevant time period for the evaluation,
and the efforts and expenditures necessary to complete the military CBA. A “pri-
mary” analysis is performed when an alternative (proposal) is evaluated with the
objective of minimizing costs of satisfying existing requirements. A “secondary”
analysis involves new needs or requirements that must be met or when existing
systems can no longer satisfy current requirements. For the purpose of this mili-
tary CBA, the project can be classified as a secondary analysis—required to satisfy
new needs of the BGAF.

13.3.1 Comparing costs of competing alternatives


Several methods exist to compare the cost of competing alternatives. Among
the more frequently applied are net present value (NPV; see Chapter 17), sav-
ings/investment ratios (SIR) and rate of return (ROR) etc. (see DOA 2001). The
main factors that determine the method to be used for the particular project and its
analysis are the amount of financial resources available, the benefit/effectiveness
measures, and the duration of the period of the analysis.
Due to the fact that the cost and the effectiveness are different for the differ-
ent alternatives considered in this example, but the duration of the period for
analysis is the same for each of the alternatives, the methods recommended by
DOA (2001) for comparison of the alternatives are NPV, or annual benefit/cost
ratios (ABCR). Given the severe warnings in Chapter 4 about using benefit/cost
ratios when neither effectiveness nor costs are fixed, the EEoA offers other
approaches that will be reviewed after the discussion of MoEs and measures of
performance.

13.3.2 Measures of effectiveness and measures of performance


Higher level MoEs represent the customers’ (or operators’) perspective, usually
of a qualitative nature. They describe the customers’ expectations of a product,
project or system. The MoE can be represented in the form of a hierarchical
diagram (see Figure 13.2). Weights can be attached to each component at each hor-
izontal level in the hierarchy. Evaluation of alternate designs can be made through
the use of a method such as the weighted objective decision matrix or similar
methods (see Chapter 8). The use of MoEs enables analysts to propose the best
alternative to solve the problem under consideration.
For purposes of the military CBA for a multi-role fighter, possible measures of
effectiveness (MoEs) include survivability, vulnerability, weapon system features,
Affordability assessments 341

Overall Measure

Components of
Level 1 Level 1 Level 1 measures

More specific
Level 2 Level 2 Level 2 Level 2 Level 2 Level 2 Level 2
measures

Figure 13.2 The MoE hierarchy diagram.

Overall
Level 1 Level 2
measure

Survivability

Vulnerability Development and construction


(effects) of the
new multi-role
Total combat
capabilities

fighter

Life-cycle cost Investment and procurement

Weapon system features Operation and maintenance

Armaments flexibility Modernization and disposal

Exploitation and upgrade

Figure 13.3 Example of an MoE hierarchy diagram.

armament flexibility, exploitation and upgrade options, shown as Level 1 in


Figure 13.3. Life-cycle costs involves all the Level 2 factors.
Measures of performance (MoP) express system design specifications derived
from higher level MoEs, as described in Table 13.1. MoPs are quantitative char-
acteristics of more interest to operators and their purpose is to demonstrate what
companies have done to achieve the MoE requirements (see Table 13.1).
The next phase of a military CBA includes identifying alternatives. This list
of alternatives should be as comprehensive as possible, including alternatives that
may not yet be available. An incomplete list of alternatives could be reason to
question the validity of a military CBA. Alternatives that are not yet available
should be discussed and documented, but it is not necessary to include them in
subsequent work on cost and effectiveness estimation. As a general rule, alterna-
tives can be included when they satisfy stated requirements and their realization is
feasible given the period of analysis, and defense budget limitations.
342 V. Georgiev
Table 13.1 MoP specifications for the multi-role fighter

Specifications (MoP)

1 Maximum speed 2,000–2,300 km/h


2 Maximum altitude 15,000–18,000 m.
3 Range 2,000–2,500 km.
4 G-max 9
5 Combat radius 500 km.
6 Combat endurance >2 hrs.
7 Weight – empty 15–20 tons
8 Payload 3–5 tons
9 Maximum wing loading up to 4 tons
10 Thrust–weight ratio 0.95–1.2
11 Reverse thrust of power plant yes
12 Active power thrust vector control yes
13 Max vertical velocity 300–350 m/s
14 Supersonic cruising speed yes
15 Takeoff and landing distance 750–1,000 m.
16 Navigation/weapon delivery system yes
17 Weapons loads up to 8 pylons
18 Systems for self-defense yes
19 Interoperability with land and air based systems yes
20 Ability to transfer data in real time at a distance more yes
than 150 km
21 Container for air reconnaissance yes
22 Container for electronic warfare and reconnaissance yes
23 Aerial refueling yes
24 Catapult system at V0 and H0 yes
25 Exploitation system yes

Source: Guide for Aircrafts’ Tactical and Technical Characteristics, 2000, pp. 66–82.

A set of possible alternatives to consider for military modernization could


include producing systems within the MoD, purchasing or leasing new systems,
purchasing or leasing used systems, or modifying/upgrading existing systems.
New alternatives might also be generated from various combinations of these
alternatives.
For the multi-role fighter, the list of possible alternatives includes moderniza-
tion of existing fighters (MiG-21 or MiG-29), as well as purchase of new or used
fighters (F-16, F-18, Gripen, etc.). The first alternatives from the list (modern-
ization of existing fighters) are documented as part of the analysis, but were not
pursued.
The next phase of military CBA requires gathering data for the cost and effec-
tiveness of each alternative, storing them in a database and executing the necessary
computations. The sources used to gather data and the results of the computations
should be carefully documented as they can greatly influence the results. This
phase of military CBA is one of the most critical parts of the analysis. It requires
Affordability assessments 343
Table 13.2 MoE/MoP for multi-role fighter alternatives

Specification Requirements Alt1 Alt2 Alt3


F-18 E/F F-16 C/D GRIPEN

1 Navigation/weapon delivery yes yes yes yes


system
2 Maximum speed 2,000–2,300 km/h 1,900 2,142 2,126
3 Combat radius 500 km 722 1,252 800
4 G-max 9 9 9 9
5 Maximum wing loading Min. 4 tons 7.7 7.2 6.5
6 Thrust–weight ratio 0.95–1.2 – 1.1 –
7 Combat endurance >2 hrs 2.5 – –
8 Interoperability with land and yes – – –
air based systems
9 Systems for self-defense yes yes yes yes
10 Max vertical velocity 300–350 m/s – 294 80
11 Payload 3–5 tons – – –
12 Maximum altitude 15,000–18,000 m 15,200 15,300 19,000
13 Range 2,000–2,500 km 3,700 4,215 3,300
14 Ability to transfer data in real yes – – –
time at a distance more than
150 km
15 Container for air yes yes – –
reconnaissance
16 Weapons loads Min. 8 pylons 9 9 8
17 Exploitation system yes yes yes yes
18 Container for electronic yes yes – –
warfare and reconnaissance
19 Catapult system at V0 and H0 yes yes yes yes
20 Reverse thrust of power plant yes yes – yes
21 Aerial refueling yes yes yes yes
22 Supersonic cruising speed yes yes – yes
23 Active power thrust vector yes yes – yes
control
24 Takeoff and landing distance 750–1000 m 430–620 450–650 400–600
25 Weight – empty 15–20 tons 16.65 11.84 9.7

Source: Guide for Aircrafts’ Tactical and Technical Characteristics, 2000, pp. 66–82.

sufficient time and resources since it can have a major impact on the final results
(see Chapter 8). Analysts should take into account all relevant cost factors and the
total expected effectiveness of each alternative. MoE/MoP results can be reported
as shown in Table 13.2.
As discussed earlier in Chapter 5, analysts use a variety of methods to calcu-
late life-cycle costs. Here, the Government Asset Management Committee (2001)
recommends engineering methods when there is detailed and accurate capital and
operational cost data available; direct cost estimation of particular cost elements;
analogous cost methods when there exists data from similar systems of analogous
size, technology, and operational characteristics; and parametric cost estimation
when historical cost data is available (see Chapter 5).
344 V. Georgiev
13.3.3 Approaches to compare and rank alternatives
The next phase of the military CBA involves comparing and ranking alternatives.
The literature refers to this as “the heart of the analysis.” In the vocabulary of the
EEoA, the multi-role fighter example is an “intra-program” analysis, and the alter-
natives under consideration could be constructed as competing fleets of aircraft
(see Chapter 4).
One of the main challenges in defense management is to achieve a balanced
portfolio of acquisition projects that, among other requirements, are feasible in
terms of budgetary constraints. It is possible to address this challenge through
EEoA (see Chapter 4). This method clearly distinguishes among three key vari-
ables: life cycle costs, operational effectiveness, and the budgets expected to be
available to finance the program/project (funding scenarios). One of the ideas
behind EEoA is to improve coordination between capabilities-based planning,
programmatic defense resources management, and defense acquisition.
An innovative idea of EEoA is to use different budget scenarios to reflect fore-
casts for the future programming period (also see Chapter 7). EEoA prescribes
an explicit optimization approach to make military investment decisions within
realistic funding constraints. This parallels the optimization approach implicit in
PPBS, designed to build an optimal mix of forces, equipment and supplies which
maximizes national security within fiscal constraints.
In general there are six ways to structure an EEoA, as illustrated in
Figure 13.4.5

• In the first five approaches the program or project can be analyzed inde-
pendently of other programs/projects in the defense portfolio. In the sixth
approach, the analysis of costs and effectiveness takes account of competing
(or complementary) programs/projects included in the portfolio.
• The first three approaches assume that alternatives can be constructed by
decision-makers in cooperation with analysts (endogenous alternatives); an
example would be comparing alternatives fleets of multi-role fighters gener-
ated by experts and the programming team (see Chapter 15 for example).
• The last three approaches focus on evaluating previously defined alternatives
(exogenous alternatives) e.g. different companies producing multi-role air-
craft participating in the contest for selection of a supplier of such aircraft for
the needs of the BGAF.

The next phase of a military CBA includes carefully recording assumptions


and constraints. In general, a military CBA is performed in conditions of incom-
plete information, which requires numerous assumptions. The goal is to reduce
the extent of uncertainty in the analysis. Assumptions and constraints are often
defined before choosing the alternatives. Assumptions and constraints define the
environment for the analysis and influence the final results of any military CBA.
This is extremely important for decision-makers to recognize and for analysts to
communicate.
Affordability assessments 345

Type of
EEoA

Intra-program
analysis

Build New
Alternatives

Fixed budget approach

Fixed effectiveness approach

Expansion path approach

Modify Existing
Alternatives

Modified budget approach

Inter-program Modified effectiveness approach


analysis

Modify Existing
Alternatives

Opportunity cost/benefit approach

Figure 13.4 Possible approaches to structure an EEoA (see Chapter 4).

For example assumptions and constraints for the multi-role fighter could involve
future budgets, the time-span of life-cycle costs and effectiveness, technical and
tactical characteristics, operations and maintenance costs, etc. In discounting
future costs (see Chapter 17) there are several key milestones such as a mis-
sion’s life, base year, lead-time, etc. The mission’s life is the period over which
the system is required. The base year is the year to which costs and benefits for the
alternatives will be discounted. Start year is the year in which initial investments
for the considered alternatives are made. Lead-time is the time from the beginning
of the start year to the beginning of the economic life of an asset for each of the
alternatives. Possible values of these time characteristics for the multi-role fighter
example are shown in Table 13.3.
The next to last phase of a military CBA involves exploring assumptions and
constraints through sensitivity analysis. Sensitivity analysis answers “what-if”
346 V. Georgiev
Table 13.3 Time characteristics assumed for the alternatives

Time characteristics Possible values

Mission’s life period 15–20 years


Base year 2010
Start year 2012
Lead-time 2–3 years

Source: Guide for Aircrafts’ Tactical and Technical Characteristics, 2000, pp. 66–82.

questions and helps to generate confidence in the results. Variations in key param-
eters such as costs, future budgets, or MoE/MoPs can be explored. If a change in
the value of a parameter impacts the ranking of the alternatives, this needs to be
carefully recorded.6
The final phase of any military CBA is to report the results and make recom-
mendations. It is an essential phase—the best alternative needs to be identified and
recommendations regarding its implementation formulated. Recommendations to
decision-makers are extremely important because sometimes the rankings do not
clearly demonstrate which alternatives are truly superior options.
The final results from the application of the AoA approach should be pre-
sented to decision-makers in the most attractive and intuitive manner possi-
ble. The report should include the goals, assumptions and constraints, list of
alternatives, the database for life-cycle cost and effectiveness, appropriate cost-
effectiveness charts, and results of the sensitivity analysis. The report should end
with conclusions and recommendations.

13.4 Conclusion
In conclusion we can summarize by observing that the application of the EEoA
framework offers several significant advantages in military CBA. The multi-
criteria approach to decision-making, which is a basis for the traditional AoA,
focuses on life-cycle costs and operational effectiveness while leaving affordabil-
ity as a secondary consideration. In contrast, EEoA expands the range of variables
in the analysis by explicitly including future funding constraints. In practice,
this can be accomplished through the development and implementation of finan-
cial scenarios for the acquisition project. Given future funding challenges, it is
imperative that all defense acquisition decisions explicitly include affordability.

Notes
1 The amount of financial resources is limited by various factors, such as the size of gross
domestic product, the state of the economy as a whole, the amount of funds allocated
from the state budget in the interest of building and maintaining defense capabilities, the
allocation of earmarked funds for security in areas which include personnel, maintenance,
training, participation in operations, investments, etc.
2 Another consequence is the adoption of investment programs and projects in defense
without having sufficiently comprehensive, detailed and reliable estimates of costs and
Affordability assessments 347
associated resources. In a study conducted in 2009, the Business Executives for National
Security (BENS) concluded that most of the requirements for defense capabilities are
determined in the absence of adequate input as to the suitability of the investment
programs and projects in terms of budgetary constraints.
3 In the language of project management, too often the scope of the investment project
does not match the scope of the defense products, which is the subject of investment.
4 The development and implementation of improved or new methods of military CBA
approaches in the evaluation and selection of investment projects to modernize the armed
forces can be considered a critical part of the innovation process in defense.
5 A detailed discussion of the six approaches to structure an EEoA appears in Chapter 4.
This chapter offers a decision map which significantly eases the practical work of ana-
lysts in choosing the preferred approach to structure an EEoA. Depending on specific
characteristics of the particular acquisition, this decision map facilitates the analytical
process and can serve as a valuable guide for analysts and decision-makers.
6 An appropriate way to find parameters that impact the results is to evaluate their influ-
ence, for example, in terms of percentages of the costs or effectiveness of the alternatives.
From a practical point of view, all factors that are related to spending greater than 20 per-
cent from the total spending for the alternative need to be considered in the sensitivity
analysis.

References
DOA. 2001. Economic Analysis Manual. US Army Cost & Economic Analysis Center.
GAO. March 2009. Michael J. Sullivan, Defense Acquisitions: DoD Must Prioritize
its Weapon System Acquisitions and Balance Them with Available Resources, Testi-
mony, before the Committee on the Budget, House of Representatives, GAO-09-501T,
Washington, DC.
Government Asset Management Committee. 2001. Life Cycle Costing. Sofia: Military
Publishing House.
Guide for Aircrafts’ Tactical and Technical Characteristics. 2000. Sofia: Military Publish-
ing House.
Ministry of Defense. 2010. White Paper on Defence and Bulgarian Armed Forces. Sofia:
Ministry of Defense.
Tagarev, Todor. Spring–Summer 2006. “Introduction to Program-based Defense
Resource Management.” Connections: The Quarterly Journal 5:1 55–69. Avail-
able at www.pfpconsortium.org/file/introduction-to-program-based-defense-resource-
management-by-todor-tagarev [last accessed February 8, 2012].
Tagarev, Todor. 2009. “Introduction to Program-based Force Development.” In Defence
Management: An Introduction, edited by Hari Bucur-Marcu, Philipp Fluri, and Todor
Tagarev. Geneva: DCAF, pp. 75–92.
Tzvetkov, Tzvetan. 2004. Innovations and Investments in Defence. Sofia: Stopanstvo.
14 Real options in military
acquisition
A retrospective case study of the
Javelin anti-tank missile system
Diana I. Angelis, David Ford, and
John Dillard

14.1 Introduction
This chapter offers an application of the real options approach to cost–benefit
analysis (CBA). The Javelin “fire-and-forget” anti-tank missile was originally
designed to replace the US Army’s DRAGON missile system. Introduced in the
late 1970s, the DRAGON had a wired command link to guide its missile to the
target that needed to be optically tracked by a gunner. Besides placing the gunner
at risk, the overall effectiveness of the DRAGON was somewhat compromised
due to limited range, lethality, and reliability problems. It also proved challeng-
ing for gunners to aim the missile and track the target. The goal was to replace
the DRAGON with a lighter weapon with increased range, lethality, and ideally a
“fire-and-forget” tracking and guidance system.
The development of the Javelin, shown in Figure 14.1, required advances in
several immature technologies to achieve the desired benefits. Target locating and
missile guidance subsystems represented particularly troublesome issues. Three
alternative technologies were initially considered. Each of the three technologies
represented realistic options that promised the desired benefits/effectiveness.
Instead of investing in a single technology, the Army decided to award three
“Proof of Principle” (technology demonstration phase) contracts of US$30 million
to each of three competing contractor teams to develop their technologies, fol-
lowed by a “fly-off” missile competition. The Army paid US$90 million in August
1986 for these three options, each with the potential, but not a guarantee, of deliv-
ering the desired benefits. Spreading the US$90 million over three competitors,
the Army in effect acquired the right, but not the obligation, to purchase the most
successful technology at a later date—a textbook example of a “real option.”
Section 14.2 offers a brief review of real options theory introduced in
Chapter 11 of this volume. Section 14.3 follows with a more detailed description
of the three Javelin guidance technology alternatives. A multi-criteria model is
developed in Section 14.4 to measure the benefits or “effectiveness” of the three
Real options: the Javelin anti-tank missile 349

Figure 14.1 The Javelin anti-tank weapon system missile and command launch unit.
Source: US Marine Corps photo 2/10/2010 by Corporal Andres Escatel.

guidance technologies. Combining cost estimates with overall measures of effec-


tiveness (MoE) offers the ability to graph the cost-effectiveness of the three
alternatives. Cost-effectiveness graphs are generated in Section 14.5 under both
deterministic and probabilistic assumptions. A decision tree is used in Section 14.6
to model the value of each option, given probabilities of success and esti-
mated costs to recover from failure. The Army’s actual decision is revealed in
Section 14.7, and the final section, Section 14.8, summarizes the case study.

14.2 Real options


An option offers the right, but not the obligation, to take specific future actions
depending on how uncertain future conditions evolve (Brealey and Meyers 2000).
The “real options” approach to CBA applies options theory to tangible assets. Real
options theory offers a powerful tool in structuring and valuing flexible strate-
gies to address uncertainty (Courtney et al. 1997). The central premise of real
options theory is this: if future conditions are uncertain, and switching strategies
at a later date would incur substantial costs, then adopting flexible strategies and
delaying decisions can have significant value when compared to committing to
a particular strategy early in pre-project planning (Amram and Kulatilaka 1999).
Real options theory helps answer questions such as: “What are possible future
actions?”; “When should we choose between these actions to maximize value
based on the evolution of uncertain conditions?”; and “How much is the right
to delay choosing an alternative worth at any given time?”
A real option compares one or more strategies that might be used in the future to
a reference strategy that could be committed to today. Uncertain future conditions
350 D. I. Angelis, D. Ford, J. Dillard
are monitored and converted into signals that can be compared to trigger condi-
tions that exercise a decision rule to determine if the reference strategy should be
abandoned and an alternative strategy adopted (i.e. to “exercise” the option). Wait-
ing to see how uncertainty evolves (i.e. learning) in order to make better strategy
choices is a key feature of the real options approach. The real options approach
requires defining both the exercise signal and the exercise decision in the context
of a set of observable variables.1
Real options can be described along several dimensions (ownership, value,
complexity, availability, etc.). A common typology separates real options accord-
ing to the managerial action that can be applied. This includes options to postpone
a decision (hold and phasing options), change the amount of investment (growth,
scaling or abandonment options), or alter the form of involvement (switching
options).
The use of real options can focus on the monetary valuation of acquiring
flexibility, or on the design and impacts of real options on decision-making (man-
agerial real options). Both of these aspects of real options can improve project
management and add value.
A wide variety of mathematical models have been used to estimate the monetary
value of options to help select strategies (e.g. see Chapter 11, and Garvin and
Cheah 2004). These models use the various benefits and costs of an option to
estimate its value. Although some real options can be purchased and exercised at
no cost (e.g. the option to have salaried employees work overtime), others involve
significant costs to obtain, maintain or exercise flexibility that might add value.
The cost of an option is a payment for access to flexibility to change the strategy.
The total costs of exercising an option must also include costs of changing the
strategy if the option is exercised. Option maintenance costs include any benefits
lost by delaying the decision. One simple and intuitive approach is to estimate the
value of an option as the difference between the values of the project with and
without the option, e.g. recognizing that uncertainty impacts future performance
versus assuming a single specific and known (deterministic) future.
Managerial real options can improve decision-making. This facilitates option
design and implementation by structuring risky circumstances faced by practi-
tioners into real options. For example, Ceylan and Ford (2002) describe the use
of options to manage technology development risk in the development of the
National Ignition Facility by the Department of Energy. Employing managerial
real options addresses many of the challenges of using real options valuation
models to improve risk management (Triantis 2005; Garvin and Ford 2012).
Structuring development program risk management challenges as real options
requires describing challenges with standard real options parameters and struc-
tures (Miller and Lessard 2000). This can improve management’s understanding
of the risks of alternative decisions, and it sets the stage for the design of risk
mitigation strategies. The design of options can be improved by assessing and
selecting individual and interacting sets of option parameters and their values,
such as evaluating the exercise cost that would make an option very attractive or
never beneficial.
Real options: the Javelin anti-tank missile 351
Operationalizing real option components, such as the operational changes
required to change strategeis and implement project monitoring signals, can enhance
the implementation of real options. Real options can improve program planning and
management by helping managers to recognize, design, and use flexible alternatives
to manage dynamic uncertainty. A retrospective opportunity to apply the real options
approach is provided by evaluating the technology development that was required
for the success of the Javelin anti-tank missile system.

14.3 Javelin technology options


The Javelin is a “fire-and-forget” anti-tank missile system. This “fire-and-forget”
capability came from the Joint Army/Marine Corps Source Selection Board’s deci-
sion to select the development team’s AAWS–M design. This design coupled a
forward looking imaging infra-red (FLIR) system with a beyond state-of-the-art
onboard software tracking system.2 As mentioned earlier, the CBA involved three
teams (alternatives) that were formed to develop competing guidance technolo-
gies for the Javelin. Only one team would then be chosen for follow-on advanced
development and then production.
Ford Aerospace, teamed with its partner Loral Systems, offered the laser beam-
riding (LBR) missile. Hughes Aircraft, teamed with Boeing, offered a fiber-optic
(FO) guided missile. Texas Instruments, teamed with Martin-Marietta, offered a
forward-looking infra-red (FLIR) missile system.
With the Ford/Loral LBR system, the gunner would identify the target visually
and point a laser beam at the target throughout the flight. After launch, the missile
continuously corrected its flight to match the line of the laser (to “ride” the laser
beam) to the target.
The Hughes/Boeing FO system included a coil of very long and fine optical
fiber that connected the launch unit, operated by the gunner, to a camera in the
nose of the missile. The gunner would fly the missile to the target using a joystick
controller device.
The TI/Martin-Marietta FLIR scanned the view in front of the gunner and
generated a thermal-based image of the target area. Once observed through the
Command Launch Unit, or thermal sight, the gunner switched to a starting array
in the missile to acquire the target by narrowing brackets in the viewfinder around
the target with a simple thumb switch. After launch, the missile would continu-
ously correct its flight path using a tracking algorithm which employed optical
correlaters oriented upon visible and distinct target features.
Once the 27-month “Proof of Principle” phase ended, each team enjoyed
generally successful missile flight tests. Each flew over a dozen missiles and
achieved target hit rates over 60 percent. Each candidate’s system, however,
offered significant advantages and disadvantages.
The Ford/Loral LBR required an exposed gunner and man-in-loop throughout
its rapid flight. It was cheapest at an estimated US$90,000 “cost-per-kill,” a figure
comprised not only of average unit production cost estimates but also reliability
and accuracy estimates. It was fairly effective in terms of potential combat
352 D. I. Angelis, D. Ford, J. Dillard
utility with diminishing probability-of-hit at increasing range. “Top-attack” on
armor would be dependent upon precision fusing, detonation, and accuracy of
downward-firing explosively formed projectiles from shaped charges.
The Hughes/Boeing FO guide prototype enabled an unexposed gunner (once
launched) and also required man-in-loop throughout its slower flight. It was
judged as likely costlier, but less affected by accuracy throughout its range with an
automatic lock and guidance in its terminal stage of flight—and it even offered tar-
get switching. Although requiring more intensive gunner training, it could attack
targets from above at the most vulnerable part of a tank’s armor.
The TI/Martin-Marietta FLIR prototype offered completely autonomous “fire-
and-forget” flight to target after launch, but was perceived as both the costliest and
technologically riskiest alternative. It would be easiest, however, to train and be
effective at maximum ranges by means of its target acquisition sensor and guid-
ance packages. It used “top-attack” as a more effective means of armored target
defeat, but would also have a flat trajectory capability for “direct fire” against
targets under cover of bridges, trees, etc.
The challenge was to evaluate the operational effectiveness and cost of each
alternative, and this is explored in Section 14.4.

14.4 Effectiveness of technology options


The benefits of each alternative anti-tank missile guidance system can be measured
using a simple multi-criteria effectiveness model based on acquisition objectives
originally specified for the anti-tank missile: lethality, tactical advantage, gunner
safety, and procurement.3 This multi-criteria effectiveness model is based on con-
cepts discussed earlier in the book originally developed by decision scientists (for
example, see Chapter 8; Keeney 1982, 1988; Buede 1986).
The first three objectives clearly relate to the operational effectiveness of the
missile. Meanwhile, the procurement objective attempts to recognize the burden
of procurement and technology issues that make some alternatives more difficult
to procure than others. This concern is captured as “ease of procurement.”4
Under each objective, specific metrics are identified to measure how well the
objective is achieved by a particular alternative. The objectives and their corre-
sponding metrics (measures) appear in Table 14.1. The weights assigned to each
objective, and to measures that define those objectives, reveal the relative impor-
tance of the objectives, and the relative importance of each measure with respect
to an objective. In this example, it is assumed that lethality, tactical advantage, and
gunner safety are equally important. Since they each receive a weight of 0.3, they
are three times as important as procurement, which receives an importance weight
of 0.1. All weights at each level of the hierarchy sum to one.
For lethality, the relative weight placed on the probability of a kill (i.e. the prob-
ability of a hit times the probability of a kill given a hit or P(H) × P(K/H)) is 0.7.
The relative weight placed on “top-attack” capability is 0.3. For tactical advantage,
the most important attribute is the heaviness of the system, which receives a rela-
tive weight of 0.4. This is followed by time to engage (0.3), then time of flight
Real options: the Javelin anti-tank missile 353
Table 14.1 Anti-tank missile guidance system effectiveness

Weight Objective Weight Measure LBR FO FLIR

Value Score Value Score Value Score

0.3 Lethality
0.7 P(H) × P(K) 5 1.05 4 0.84 7 1.47
0.3 Top attack 6 0.54 7 0.63 9 0.81
0.3 Tactical
advantage
0.4 Weight 9 1.08 5 0.6 3 0.36
0.3 Time to engage 8 0.72 7 0.63 5 0.45
0.2 Time to flight 7 0.42 5 0.3 5 0.3
0.1 Redirect 10 0.3 10 0.3 0 0
capability
0.3 Gunner
safety
0.2 Required training 5 0.3 1 0.06 10 0.6
0.8 Exposure after 2 0.48 8 1.92 10 2.4
launch
0.1 Procurement 1 Ease of 8 0.8 6 0.6 4 0.4
procurement
MoE 5.69 5.88 6.79

(0.2), and then redirect capability (0.1). Gunner safety is measured mostly by
the gunner’s exposure to enemy fire after launch (0.8) and also by the amount
of training required (0.2).
In this example, actual values for each measure (computed or subjectively
assigned) for each of the three competing technology alternatives (LBR; FO;
FLIR) are converted (normalized) on a scale of 0 to 10. For instance, in the case
of gunner safety, the objective is to minimize the amount of time the gunner is
exposed to enemy fire (measured in minutes). With LBR the gunner must stay in
place until the target is hit, leading to a longer exposure time, so LBR receives a
low value of 2. The FO system allows the gunner to hide while guiding the mis-
sile, so he is exposed for a shorter time and, thus, FO receives a better value of 8.
The FLIR system allows the gunner to conceal himself immediately after launch
(“fire-and-forget”) and, thus, is given the maximum value of 10. The other values
shown in Table 14.1 are derived in similar fashion.
The overall MoE achieved by each of the three alternatives under the previous
assumptions is shown at the bottom of Table 14.1. Individual scores shown in
Table 14.1 for each measure are calculated by taking the normalized value for that
measure, multiplying it by the relative weight for that measure, and multiplying
again by the relative weight placed on the higher level objective.5 The overall MoE
for any given alternative is the sum of the individual metric scores.6
The overall MoEs reported in Table 14.1 suggest that FLIR offers the greatest
benefit/effectiveness (6.79), with LBR and FO closely tied for second (offering
354 D. I. Angelis, D. Ford, J. Dillard
overall MoEs of 5.69 and 5.88, respectively). This ranking turns out to be con-
sistent with the Army’s revealed preference for the three guidance technologies:
They preferred FLIR over the other two guidance systems, and they perceived
the FO system as being slightly better than the LBR system (J. Dillard; personal
communication, November 12, 2012).

14.5 Cost-effectiveness analysis


The previous section focused on the overall operational effectiveness of the three
guidance alternatives. Assuming budgets for the program are constrained, then
selection of the best alternative also requires cost estimates, including technology
development and manufacturing (production) costs.
The Army’s estimates of the cost-per-kill for each alternative are reported in
Table 14.2. This table also reveals estimates of the total program cost for each
alternative assuming an Army requirement of 2,000 missiles.
A cost vs effectiveness graph, in which the MoE is plotted against total program
cost for each alternative, is shown in Figure 14.2.
From Figure 14.2 we can see that no single alternative dominates another, mean-
ing there is no alternative that is both cheaper and more effective than another. In

Table 14.2 Anti-tank missile cost

LBR (US$m) FO (US$m) FLIR (US$m)

Cost/kill 0.09 0.11 0.15


Program cost 180 220 300

Program Cost
10

-
7
.&

• + LBR
4 • Fo
3 A FUR

0
$- $50 $100 $150 $200 $250 $300 $350
$Millions

Figure 14.2 Deterministic cost-effectiveness.


Real options: the Javelin anti-tank missile 355
this case, as discussed in Chapters 4 or 10, it is useful to examine marginal benefits
and marginal costs to help evaluate alternatives.
The LBR alternative is the least costly and least effective. We compare it to the
FO alternative in Table 14.3 and note that the marginal cost of choosing FO over
LBR is US$40 million. Table 14.3 also shows differences in values for each of the
effectiveness measures used to calculate the MoE between the two alternatives. A
positive change represents an increase in effectiveness while a negative difference
indicates a decrease in effectiveness. A similar analysis for FO vs FLIR is shown
in Table 14.4.
While Figure 14.2 offers an overall picture of differences in cost vs effective-
ness of the three alternatives, Tables 14.3 and 14.4 allow decision-makers to see
specifically what they gain and lose at the margin as they go from one technol-
ogy to the next. Arguably this is captured in an aggregated way in the overall
MoE, but it is also important for decision-makers to understand specific benefits
they receive for the money (value-for-money).7 This marginal analysis defines the
tradeoff space for decision-makers, but it is not the solution.8

Table 14.3 Marginal analysis of cost and effectiveness for LBR and FO

Marginal analysis LBR FO Difference

Program cost (US$ million) 180 220 40


P(H) × P(K) 5 4 −1
Top-attack 6 7 +1
Weight 9 5 −4
Time to engage 8 7 −1
Time of flight 7 5 −2
Redirect capability 10 10 0
Required training 5 1 −4
Exposure after launch 2 8 +6
Ease of procurement 8 6 −2

Table 14.4 Marginal analysis of cost and effectiveness for FO and FLIR

Marginal analysis FO FLIR Difference

Program cost (US$ million) 220 300 80


P(H) × P(K) 4 7 +3
Top-attack 7 9 +2
Weight 5 3 −2
Time to engage 7 5 −2
Time of flight 5 5 0
Recall capability 10 0 −10
Required training 1 10 +9
Exposure after launch 8 10 +2
Ease of procurement 6 4 −2
356 D. I. Angelis, D. Ford, J. Dillard
The previous cost vs effectiveness analysis assumed all guidance technology
development efforts would be equally successful and achieve the estimated MoEs.
At the start of the “Proof of Principle” effort for the project, however, there was
significant uncertainty and no assurance that any of the technologies would be
successfully developed.
In fact, early assessments of the probability of success differed among the three
alternatives. A notional assessment of the probability of success for each option
is given in Table 14.5. Table 14.5 also calculates the “expected MoE” based on
the given probability of success for each option. The expected MoE is simply the
overall MoE shown in Table 14.1 multiplied by the probability of success.
Given this new calculation of expected MoEs, the previous cost vs effectiveness
graph can be revised as shown in Figure 14.3. Here we conclude the LBR system
is superior since it is both cheaper and has a higher expected MoE than the other
two (FO and FLIR), and therefore it dominates those guidance systems. Interest-
ingly, the Army’s actual cost and operational effectiveness analysis (COEA) in
fact found LBR was the preferred alternative (J. Dillard; personal communication,
November 12, 2012).

14.6 Value of technology real options: a retrospective assessment


The value of a real option is derived from the difference between the expected
net value (i.e. expected benefits minus expected costs) of an investment and the
net value of that investment given that it succeeds. The option value lies in the
flexibility to terminate the project if it is not successful.
To develop a simple option valuation model for the Javelin guidance technolo-
gies, we note benefits appear as the MoEs in Table 14.1. In the absence of a real
options approach, each of the alternatives has an expected cost that is simply based
on the uncertainty associated with its technological development.
If the development effort fails, we assume the Army will have to pay some
additional cost to complete the technology and to achieve the anticipated level
of effectiveness (MoE). We refer to this additional cost as the “cost to fix” the
technology.
If technology development is successful, then the cost to exercise the option will
be the total program cost from Table 14.2 and, in Table 14.6, is shown as “cost to
implement.” If the technology development fails, the cost of the alternative will be
the “cost to implement” plus the “cost to fix” the technology. The total expected
cost of each alternative is given in Table 14.6.

Table 14.5 Probability of development success and expected MoE for Javelin technology
options

LBR FO FLIR

P (success) 0.6 0.5 0.4


Expected MoE 3.414 2.94 2.716
Real options: the Javelin anti-tank missile 357

Program Cost
10

6
~ 5 + LBR
::E
• Fa
• --
4


3 .A FUR

0
$- $50 $100 $150 $200 $250 $300 $350
$Millions

Figure 14.3 Probabilistic cost-effectiveness.

Table 14.6 Expected cost of Javelin guidance technology alternatives without option to
terminate project

LBR FO FLIR

Probability of success 0.6 0.5 0.4


Cost to implement (US$ million) 180 220 300
Probability of failure 0.4 0.5 0.6
Cost to fix (US$ million) 50 70 90
Total cost to implement 230 290 390
given failure (US$ million)
Expected cost (US$ million) 200 255 354

The values shown in Table 14.6 assume that we do not use a real options
approach. Instead, we will pick one of the technologies based on the cost vs ben-
efit analysis presented in Section 14.5. Whichever technology we choose, we will
have to pay an additional cost to achieve the anticipated effectiveness (MoE) if the
development phase fails.
The real options approach allows the Army to pay for the option to find out if
technology development succeeds before making its final choice. If development
succeeds, the MoE shown in Table 14.1 is achieved and the Army can proceed with
the project if they prefer that option based on the cost vs effectiveness analysis
presented in the previous section. If development fails, the particular development
project can be terminated and there is no further cost.
358 D. I. Angelis, D. Ford, J. Dillard
The value of the option is given by the difference between the expected cost
of the technology development project with no option (from Table 14.6) and the
expected cost of the project with the option to terminate. Values of the options for
the three alternatives appear in Table 14.7.
Differences in values of the options reflect the different levels of uncertainty
associated with each technology. The values shown in Table 14.7 are maximums
in the sense that if we pay any more than the option value we would have been
better off not using an option. If we pay less than the option value, we experience
real cost savings by not expending funds on an unsuccessful technology. Note the
more uncertain the technology (i.e. probability of failure), the greater the value of
an option to terminate the project if technology development fails.
Let us suppose that the Army preferred the LBR technology (based on the cost
vs effectiveness analysis presented in Section 14.5). Then, according to Table 14.7,
they should pay no more than US$92 million for the option to terminate the
project. Since the Army was in effect buying options for all three technologies,
this indicates that the total amount spent on options should not exceed the value
of the option for the preferred technology. Since the maximum value of the LBR
option is US$92 million, if the Army allocated that option value equally across all
the alternatives, they should spend no more than roughly US$30 million for each
option, which is precisely what they did.
Given some technologies are more uncertain than others, a better approach is to
allocate the US$92 million based on the level of uncertainty the Army is trying to
resolve. Using the probability of failure data in Table 14.7 as a notional measure
of risk, 27 percent of the total option value would be allocated to LBR, 33 percent
to FO, and 40 percent to FLIR.
Returning to the previous example, if ex-ante the Army prefers the LBR
technology, then the total cost of the option should not exceed US$92 million. Rec-
ognizing uncertainty (i.e. risk of failure), this means the Army should be willing to
pay roughly US$25 million for the LBR option, US$30 million for the FO option,
and US$37 million for the FLIR option. Doing so allocates the dollars based on
risk while keeping the total cost no more than the maximum option value of the
preferred alternative. Again, since US$92 million is the maximum value of the

Table 14.7 Value of Javelin guidance technology options

LBR FO FLIR

Probability of success 0.6 0.5 0.4


Cost to implement (US$ million) 180 220 300
Probability of failure 0.4 0.5 0.6
Cost if project is terminated (US$ million) 0 0 0
Expected cost with option (US$ million) 108 110 120
Expected cost w/o option (US$ million) 200 255 354
Value of option (US$ million) 92 145 234
Real options: the Javelin anti-tank missile 359
real option, to realize cost savings from this strategy, the Army would need to pay
less than US$92 million for all three options.

14.7 The Army’s choice revealed


The LBR candidate emerged as the winner of the original CBA based on initial
assessments of relative weights on the multiple criteria. In a strange twist, the
Source Selection Evaluation Board (SSEB) instead chose the FLIR candidate,
partly as a result of a significant shift in the importance (relative weight) attached
to gunner safety that drove a strong preference toward “fire-and-forget” in the
overall MoE.9 While time of flight and gunner survivability were not explicitly
stated requirements in the original (AAWS-M) Joint Required Operational Capa-
bility document, “fire-and-forget” nevertheless translated into greatly enhanced
gunner survivability, and it ultimately appealed to decision-makers (particularly
to operator representatives). In June 1989, a full-scale development contract was
awarded for the AAWS-M project to the FLIR team of Texas Instruments and
Martin-Marietta.
The Office of the Secretary of Defense viewed the program as having an accept-
able risk because of its 27-month technology development phase (which included a
real options approach to develop a technical solution) and the subsequent 36-month
plan for full-scale development. At the program office level, however, FLIR was
known to be a high risk program in several technical areas. Focal plane array (FPA)
technology was still immature, despite successful technology development phase
results.10 Since it was recognized as technologically risky, the government funded
its own night-vision laboratory. In effect, it purchased an option by partially fund-
ing several companies to produce these devices. In 1991, the only five known FPA
makers in the world were Rockwell International, Loral, Santa Barbara Research
Corporation, Sofradir (a French firm), and Texas Instruments.11
As an additional gauge of technological maturity, a comparative baseline test
was mandated at the second milestone upon the decision to launch the Javelin
program into full-scale development. That test would pit the immature FPA tech-
nology against existing tube-launched, optically tracked, wire-guided (TOW), and
DRAGON (legacy systems) night-viewing optics. Results of this test showed the
Javelin’s immature FPAs still outperformed the DRAGON and were almost as
effective as the much larger TOW anti-tank missile system.12
About 18 months into the Engineering and Manufacturing Development (EMD)
phase, serious technical problems doubled the expected cost of development and
added about 18 more months to the originally planned 36 months to complete. This
required formal re-baselining that took the better part of the next year to accom-
plish.13 The program was restructured, given a new baseline, and was subsequently
completed largely within the new parameters. The additional 18 months added to
the original 36-month phase helped to resolve the uncertainties and complexities
of system development.
Today, Javelin is viewed as a successful weapon system that delivers valu-
able operational effectiveness, despite its earlier developmental challenges. It is
360 D. I. Angelis, D. Ford, J. Dillard
routinely used in combat operations and has enjoyed repeated full-rate production
contracts.14 In the decade from 2003 to 2013, more than 1,000 Javelin missiles
were fired in Iraq and Afghanistan with close to 98 percent reliability.

14.8 Conclusions
Several observations can be drawn from the analysis of the Javelin guidance tech-
nology acquisition process. The first is that the benefit of weapon systems or, in
this case, missile guidance systems, is not measured in dollars. This makes using
a traditional option valuation model based on monetary benefits minus costs diffi-
cult (see Chapter 11 for example), if not impossible. Instead, a useful approach is
to turn to principles of multi-criteria decision-making to develop MoEs for each
alternative. The overall MoE can then be joined with cost estimates to illustrate a
trade space of alternatives available to decision-makers.
Secondly, we note that the three proposed guidance systems involved dif-
ferent levels of risk. This information can be used to calculate an “expected
MoE” for each alternative, incorporating uncertainty into the analysis. This prob-
abilistic MoE can be combined with the alternative’s expected cost to present a
risk-adjusted trade space for decision-makers.
Thirdly, we show that a real options approach not only allows uncertainty to be
incorporated explicitly in the analysis, but it also enables calculation of the value
of real options based upon various risks. This leads to different option values for
different alternatives based on technological maturity.
Using this real options approach, we conclude that the Army might have mit-
igated its risks by offering each development team a different amount of money
to develop their proposed technology, based on probabilities of success derived
from technological risk assessments. We note that the final “cost to fix” the FLIR
guidance technology ultimately selected by the Army turned out to be signifi-
cantly higher than the US$30 million originally paid to develop the technology.
This is in line with predictions of the real options model that recognized the FLIR
technology as the riskiest investment.
In conclusion, the real options approach to CBA captures the value of flexibility
in acquisition decisions. Faced with significant uncertainty, implementing a real
options approach can result in more efficient use of scarce resources.

Notes
1 In the classic example of stock purchase options, the exercise decision rule is to sell a
stock if the price rises above a certain price, and the exercise signal is the stock price.
The decision delay is incurred while the option holder waits to see if the stock price
rises above the exercise price.
2 The Advanced Anti-Armor Weapon System – Medium (AAWS-M) project inspired the
Javelin program. The joint Army and Marine Corps operational requirements document
for the Javelin was formally approved-amended in 1986–88. The FLIR approach con-
sisted of a laser beam-riding system, a fiber-optic guided system, and a forward-looking
infra-red system.
Real options: the Javelin anti-tank missile 361
3 The services procuring the Javelin system did not actually use this exact methodology
for the selection of the Javelin guidance technology, but used something similar for a
weighted decision analysis of the three alternatives.
4 This burden is sometimes termed “transaction costs” as opposed to “production costs.”
Note prior discussions in this volume about the interpretation of placing a weight on pro-
duction costs. Ideally, the operational performance of investments should be evaluated
independently of production costs in the spirit of “cost as an independent variable”
(CAIV). The next section discusses a way to combine cost and effectiveness measures.
5 For example, required training has a weight of 0.2 and is a metric that supports gunner
safety, which has a weight of 0.3. LBR received a value of 5 for this metric, so the score
for LBR is 0.3 = (5) × (0.2) × (0.3). The score for FO is 0.060 = (1) × (0.2) × (0.3)
and the score for FLIR is 0.6 = (10) × (0.2) × (0.3). All other scores in Table 14.1 are
calculated in a similar manner.
6 For example, the MoE for LBR is 5.69 = 1.05 + 0.54 + 1.08 + 0.72 + 0.42 + 0.30 +
0.30 + 0.48 + 0.80. Note that the MoE is calculated on a scale of 0 to 10 where an
“ideal” alternative would receive an MoE of 10.
7 As discussed in Chapter 4, and because the MoE combines different measures that are
not necessarily substitutes, it is not possible to use benefit/cost ratios (i.e. the ratio of
MoE to cost) to rank alternatives.
8 Solving the cost-effectiveness problem as defined in this case could be handled by “lev-
eling the playing field” as discussed in Chapter 4, using the fourth or fifth approaches
to structuring an economic evaluation of alternatives (EEoA)—the modified budget or
modified effectiveness approaches—or the sixth (“opportunity cost”) approach.
9 As part of the capability formulation process, technical constraints are deliberately
avoided in requirements documents to allow and encourage a maximum range of
alternative solutions to a particular requirement or capability deficiency.
10 It would be gauged today at approximately Technology Readiness Level 5. DoD
assesses the maturity of critical technologies on a scale from 1 (lowest) to 9 (highest)
as described in the Technology Readiness Assessment Guidance (2011) http://www.acq.
osd.mil/ddre/publications/docs/TRA2011.pdf [last accessed December 4, 2014].
11 The two-partner TI/Martin-Marietta Joint Venture in the full-scale development phase
was also free to maximize competition at the subcontractor level. In their make-or-buy
decision, Texas Instruments elected to make the focal plane array for both of its uses
in the command launch unit and in the missile. The company had made these devices
for other programs, but not in these two distinct configurations (scanning and starting
arrays).
12 Around focal plane array attainment of specified sensitivity and production yield,
system weight, tracker algorithm, and other areas.
13 This constituted a “Nunn-McCurdy breach” of cost and schedule thresholds, which
required Congressional notifications and formal re-baselining taking the better part of
the next year to accomplish. Over that next year, the program sought a new baseline
with many different revised program estimates—climbing from 36 months duration and
US$298 million in cost, to 48 months duration and US$372 million in cost, and finally
to 54 months and US$443 million for the total cost and duration of this phase.
14 The system design has continued to be upgraded, not as blocks of capability, but with
software, warhead, and producibility enhancements.

Bibliography
Amram, M. and N. Kulatilaka. 1999. Real Options: Managing Strategic Investment in an
Uncertain World. Cambridge, MA: Harvard Business School Press.
Brealey, R. and S. Myers. 2000. Principles of Corporate Finance (6th edition). New York:
McGraw Hill.
362 D. I. Angelis, D. Ford, J. Dillard
Buede, D. M. 1986. “Structuring Value Attributes.” Interfaces 16(2): 52–62.
Ceylan, B. K. and D. N. Ford. 2002. “Using Options to Manage Dynamic Uncertainty in
Acquisition Projects.” Acquisition Review Quarterly 9(4): 243–58.
Courtney, H., J. Kirkland, and P. Viguerie. 1997. “Strategy under Uncertainty.” Harvard
Business Review 75(6): 67–79.
Dillard, J. and D. Ford. 2009. “From Amorphous to Defined: Balancing the Risks of Spiral
Development.” Accounting Review (52).
Ford, D. and J. Dillard. 2009. “Modeling the Integration of Open Systems and Evolutionary
Acquisition in DoD Programs.” Acquisition Review Quarterly (51).
Garvin, M. and C. Cheah. 2004. “Valuation Techniques for Infrastructure Decisions.”
Construction Management & Economics 22(5): 373–383.
Garvin, M. and D. Ford. 2012. “Real Options in Infrastructure Projects: Theory, Practice
and Prospects.” Engineering Project Organization Journal 2(1–2): 97–108.
Keeney, R. 1982. “Decision Analysis: An Overview.” Operations Research 30(5): 803–838.
——. 1988. “Structuring Objectives for Problems of Public Interest.” Operations Research
36(3): 396–405.
Lyons, J., D. Long, and R. Chait. 2006. Critical Technology Events in the Development of
the Stinger and Javelin Missile Systems: Project Hindsight Revisited. Washington, DC:
National Defense University Center for Technology and National Security Policy.
Miller, R. and D. Lessard. 2000. The Strategic Management of Large Engineering Projects.
Cambridge, MA: MIT Press.
Triantis, A. 2005. “Realizing the Potential of Real Options: Does Theory Meet Practice?”
Journal of Applied Corporate Finance 17(2): 8–16.
15 An application of military
cost–benefit analysis in a
major defense acquisition
The C-17 transport aircraft
William L. Greer

15.1 Introduction
This case study illustrates the traditional approach to military cost–benefit analysis
(CBA) and acquisition decisions within the Department of Defense (DoD) described
in Chapter 8. It specifically discusses decisions made in 1993–95 to replace C-141
transport aircraft with the C-17 strategic airlifter. It also includes a discussion of
events leading up to the study, the analytical approach taken, the study participants,
critical points in the analyses, decisions reached, and lessons learned.
The C-17 Globemaster II aircraft has been one of the more successful military
transportation assets in history. It provides rapid delivery of cargo or passengers
almost anywhere in the world. Currently it is flown by the US Air Force (USAF)’s
Air Mobility Command (AMC) and by the armed forces of several other coun-
tries, including the United Kingdom’s Royal Air Force, the Royal Australian Air
Force, the Canadian Force’s Air Command, NATO, and the Qatar Emiri Air Force.
The United Arab Emirates recently ordered C-17s, and many other countries are
exploring the possibility of doing so. In addition to its transoceanic strategic-
lift capability, the C-17 can also perform tactical airlift, medical evacuation, and
airdrop missions. It played a significant role in moving goods, vehicles, and per-
sonnel during Operation Enduring Freedom in Afghanistan and Operation Iraqi
Freedom in Iraq, and it has contributed to emergency relief operations around the
world. It is widely considered a success story.
That was not always the case. The program came very close to being cancelled
in the early 1990s. Key decisions made around that time serve as a valuable exam-
ple of the useful role of CBA in major defense investment decisions. This chapter
describes cost-effectiveness analyses (CEAs) that informed the decision to retain
the C-17 program, and serves as a case study in how such assessments can assist
decision-makers. The focus is on the analyses but, just as important, it provides
insights into how these analyses were received and used to make the decision to
invest in the C-17.
The analyses depicted here are drawn from documents (Greer et al. 1993;
Bexfield et al. 2001) prepared by the Institute for Defense Analyses (IDA) for the
DoD. Studies such as these were referred to at that time as cost and operational
effectiveness analyses (COEAs), currently called analyses of alternatives (AoAs).
364 W. L. Greer
These studies include many of the basic considerations of cost and effectiveness
analysis discussed in this book.

15.1.1 Background
In the 1970s, the USAF launched a plan for a new cargo aircraft to carry most
of the largest equipment the Army needed (e.g. tanks, large-tracked and wheeled
vehicles, and Patriot batteries) into small remote airfields—many with runways
shorter than 3,000 feet. Although at the time airlifters existed that could carry large
equipment (the C-5) and land in small airfields (the C-130), no existing airlifters
could perform both these tasks.
In December 1979, the DoD initiated the Cargo-Experimental (C-X) competi-
tion for a new strategic airlift aircraft. The C-X was to be an aircraft that could
deliver a full range of combat equipment over intercontinental distances, operate
from a 3,000-foot runway, possess survivability features, have excellent reliabil-
ity, maintainability, and availability, and have a low life-cycle cost. Early in 1980,
the DoD issued the request for proposals (RFPs) for the new C-X program. Boe-
ing, Lockheed Martin, and McDonnell Douglas responded. After reviewing the
various designs, in August 1981 the DoD selected the aircraft design proposed
by McDonnell Douglas.1 The winning design incorporated many features already
demonstrated on the YC-15, a McDonnell Douglas aircraft that had been devel-
oped and flight tested in the 1970s. This chosen design was later designated the
C-17. Because of the earlier testing of the YC-15, it was viewed as a program with
low development and cost risk.
In addition to carrying large cargo and landing in small airfields, the C-17 was
also designed to back up and turn in a small radius, thereby optimizing the use of
limited ground space. It was also designed to conduct troop and cargo airdrops,
replacing the aging C-141s that previously conducted those missions. It was an
aircraft for all missions.
In 1982, the DoD conducted a large transportation assessment, the Mobility
Study, which established the air mobility requirements at that time, based on per-
ceived wartime needs. To meet these requirements, the USAF determined that 210
C-17s were needed. These requirements were codified in the 1983 Air Mobility
Master Plan. Largely because of budget rather than effectiveness considerations,
the number 210 was subsequently reduced by the Office of the Secretary of
Defense (OSD) to 120 during the 1990 Major Airlift Review. Thus the number
of C-17s under consideration in the early 1990s was 120.
Several complicating issues arose at that time. The earlier assessments of low
development and cost risk proved wrong. Costs for the C-17 continued to rise,
and the several aircraft delivered for testing did not demonstrate the reliability and
performance expected. Static wing loading tests revealed that the wings failed at
lower loads than predicted, forcing even higher costs to strengthen them.
Lockheed Martin, manufacturer of the C-5 and C-141, made an unsolicited pro-
posal in 1991 to upgrade and extend the life of the C-141s at a cost that appeared
lower than that of the troubled C-17 program. An added complication and a renewed
Acquisition of the C-17 aircraft 365
sense of urgency was presented when it was found that C-141s had developed a large
number of microscopic cracks in their wing structures—a condition that temporarily
grounded the entire C-141 fleet until repairs could be made.
This raised a number of questions for DoD: Should it continue with the C-17
program despite escalating costs and lower performance than desired? Should it
cancel the C-17 program and extend the life of C-141s, continuing with wing
repairs until a new C-X design was chosen? Or should it investigate other alterna-
tives? These were the issues confronting Congress and the DoD in 1993 and are
the topics of this case study.
Congress mandated that the DoD conduct an independent assessment of the
C-17 program and report its findings to the US Congress (Senate and House
Armed Services Committees). The legislation further restricted fiscal year (FY)
1994 spending on the C-17 until such a study was completed. After reviewing the
expertise and capabilities in several federally funded research and development
centers (FFRDCs), the OSD selected the IDA to conduct the study. IDA com-
pleted its study in 1994. The findings were used by the DoD and duly reported to
the congressional defense committees.
Table 15.1 summarizes the history of the C-17 program up to the point at which
analyses were conducted to determine whether to go ahead with the program.
With the CEAs in hand, the OSD finally approved the 120-aircraft C-17 pro-
gram in 1995. The road to this decision point was not an easy one. It required
numerous analyses, the main one of which is the cost-effectiveness case study

Table 15.1 C-17 decision timelines

Year Events

1970s • USAF decided new cargo aircraft, C-X, was needed to carry
heavy and large army equipment into austere airfields
1981 • McDonnell Douglas selected to build C-X, called C-17
1982 • Mobility Study published, calling for more airlift capability
1983 • USAF Airlift Master Plan estimated that 210 C-17s needed to
meet needs identified in mobility study
1990 • Major Airlift Review reduced C-17 requirement from 210 to
120
1991 • Lockheed Martin proposed C-141 service life extension
program (SLEP)
1992 • Joint Staff issued report, concluding that C-17 is more
cost-effective than either new C-5 or C-141 SLEP
• Joint Staff issued Mobility Requirements Study
1993 • Problems with aging C-141s continued
• C-17 exhibited engineering problems and cost increases
• Congress restricted FY94 C-17 spending pending further study
• DoD asked IDA to conduct COEA on C-17 program
1994 • IDA completed C-17 COEA
• COEA sent to Congress after OSD review
• Additional DoD studies were authorized
366 W. L. Greer
summarized in this section. At the end of this section, we also summarize more
recent decisions affecting the C-17 program.

15.1.2 Organizations involved


In 1993, strong views existed on the C-17 program. Proponents for the C-17
resided, among other places, within the Army Material Command and at McDon-
nell Douglas. Detractors included Lockheed Martin, manufacturer of competing
aircraft like the C-5 and C-141, the Boeing Corporation, manufacturer of the
commercial 747 and 767 aircraft, and other potential competitors to the C-17.
Numerous subcontractors were also interested in the outcome. Senior OSD offi-
cials were responsible for generating objective, unbiased analysis to support the
Secretary of Defense on whether to continue the C-17 program. The Secretary was
interested in a study that sorted through the claims and evaluated the alternatives.
The report would be delivered to the Defense Acquisition Board (DAB), the OSD
board that recommends acquisition actions to the Secretary of Defense, who, after
review, would propose it to the House of Representatives and Senate defense com-
mittees. The CBA needed to be independent, clear, and able to withstand intense
scrutiny.
The IDA study team consisted of experts in airlifter operations, computer mod-
eling and simulation, aircraft and engine design, and cost estimating. The study
team engaged in frequent interactions with aircraft and engine manufacturers,
USAF operational commands, the AMC, the USAF Air Staff, the Joint Staff,
and OSD.

15.1.3 Methodology
The approach taken in this study consisted of the following steps:

• identify the alternatives (fleets of aircraft to replace C-141s and upgrade


strategic-lift capability);
• establish airlift requirements (criteria and attributes);
• estimate effectiveness of each alternative in the airlift mission (measures of
effectiveness or MoEs);
• estimate the total ownership cost, sometimes referred to as the total system
life-cycle cost, of each alternative;
• arrange the cost and effectiveness information to facilitate decision-making.
Some steps (costing and determining effectiveness) were performed concur-
rently; and
• perform sensitivity analyses as needed.

15.1.3.1 Identify fleet alternatives


The alternatives were initially selected to have equal lift capability (equal
effectiveness—see Chapter 4 in this book) as measured by one of the com-
mon capacity measures used in the airlift community: million ton-miles per day
Acquisition of the C-17 aircraft 367
(MTM/D) (AFPAM 10-1403 2003). When divided by the distance an airlifter must
fly from the United States to the theater of interest, the MTM/D is a rough mea-
sure of the steady-state rate of delivery of cargo to the theater, provided sufficient
infrastructure and parking spaces are available, including en-route and in-theater.
Since the infrastructure and parking assumptions are rarely satisfied, fleet-wide
MTM/D offers only a partial approach to identify comparable alternative fleets of
aircraft that offer similar airlift capability. More detailed analyses further differen-
tiate among the alternatives, identifying both higher-cost-higher-effectiveness and
lower-cost-lower-effectiveness alternatives.

15.1.3.2 Establish airlift requirements


The requirements for airlift were taken from the Mobility Requirements Study
(MRS) (Joint Staff J4 Logistics 1992). This study, completed by the Joint Staff
and OSD, identified the type and tonnage of cargo and troops that were to be
moved by air (as well as by land and sea) per day. Cargo was identified by a com-
modity code and by dimensions that allowed it to be sorted into three conventional
cargo categories called outsize, oversize, and bulk.2 The MRS established lower
limits for acceptable cargo delivery rates that would meet medium-risk criteria
in warfighting simulations. The final cargo delivery schedule is referred to as the
Time-Phased Force Deployment Data (TPFDD).

15.1.3.3 Estimate effectiveness of each alternative


The effectiveness analysis involved estimating for each alternative fleet of airlift
aircraft how much cargo and how many troops from the TPFDD were delivered to
the theater within a fixed period of time. To make these estimates, the study used
a detailed simulation of the aircraft loading, aircraft movement, and cargo flow.
The specific models used were the Airlift Loading Model (ALM) for loading, the
detailed simulation model Mobility Analysis Support System (MASS) for airlift
movement and cargo flows, and the Airlift Cycle Analysis Spreadsheet (ACAS)
for more aggregate but faster assessments of cargo and passenger movement.3 The
primary output of MASS and ACAS used in the report was the total cargo tonnage,
separated into the three categories of bulk, oversize, and outsize cargo, delivered
in 30 days by each alternative fleet.

15.1.3.4 Estimate costs


The study team also provided cost estimates including the development, procure-
ment, and the annual operations and support (O&S) costs for each airlifter in
each alternative fleet. Total life-cycle costs over a 25-year period were then esti-
mated for each alternative. Information from the C-17 Program Office, contractors,
and OSD was used to arrive at independent cost assessments, all conducted in
accordance with standard costing methods.4
An attempt was made to make the alternatives as comparable as possible in
all other measures. Specifically, all alternatives were sized to be capable of a
368 W. L. Greer
brigade-size airdrop. All alternatives possessed comparable intra-theater tactical
airlift capabilities. Since the C-17s could also substitute for C-130s in tactical
airlift roles when in-theater, the study argued that mixed fleets with C-17s need
not have as many C-130s for comparable overall strategic and tactical capability
as those without them. Adjustments were made to the C-130 fleet size needed for
each alternative with corresponding adjustments in the associated O&S costs.

15.1.3.5 Arrange cost and effectiveness information to facilitate


decision-making
Finally, once cost and effectiveness measures were estimated for all alternatives,
a way of displaying them simultaneously to help decision-makers quickly sort out
favorable ones from less favorable ones was developed. As discussed earlier, this
step is crucial in any study of this type, since it establishes the framework for the
decisions that must ensue. Such a framework must be free from biases (or must
clearly identify biases that would otherwise be hidden) but must emphasize the
critical issues. The organization of the cost and effectiveness results is vital to the
utility of the analyses. Examples will be provided later.

15.2 Alternatives
Alternatives refer to the alternative fleets, each made up of a number of different
types of aircraft. This section first discusses the different types of aircraft in the
alternative fleets and then describes the composition of the alternative fleets, i.e.
how many of which types of aircraft are in each. Figure 15.1 provides a schematic
of the aircraft under consideration to indicate the relative sizes and capacities of
these airlifters. Capacity is significant for determining how much each airlifter can
carry, and size proves important for airbases with restricted space or facilities.

c-130 c-1418 Kc-10 c-11 747 c-s


Bulk Bulk Bulk Bulk Bulk Bulk
Oversize Some Oversize Some Oversize
Oversize Outsize Oversize Outsize
MAX
PAYLOAD 41,220 89,000 166,000 169,000 243,000 239,000
(Lb)

Figure 15.1 Aircraft size and capacity comparisons.


Acquisition of the C-17 aircraft 369
The military aircraft types considered in the study included: C-5A, C-5B, KC-
10A, C-17, C-5B+, C-141, C-Y, militarized 747-400F, and militarized 767-300F.
In addition, aircraft from the Civil Reserve Air Fleet (CRAF) that can be called
up for duty to transport troops and bulk cargo during war conditions were also
incorporated into fleet mixes. Each is discussed briefly in turn.

15.2.1 Core aircraft types


The 1993 core airlift fleet is summarized below. These aircraft would remain in
the fleet no matter which new airlifter program was selected and are referred to as
members of the core fleet, common to any alternatives.

15.2.1.1 C-5A/B
At the time of the study in 1993, 76 C-5As and 50 C-5Bs made up the total aircraft
inventories of the active, guard, and reserve fleets.5 Operating costs and utilization
rates are slightly different for the two types, but they carry the same payloads the
same distances. These are the largest airlifters of the USAF. Each C-5 can take off
or land with a maximum payload of nearly 120 tons. At the time the study was
conducted, the average cargo weight carried by the C-5s was 68.9 tons (these esti-
mates are updated frequently by AMC, so more current numbers should be used in
future studies), with a block speed6 of 423 knots. C-5s can carry very heavy or very
large outsize cargo, such as M-1 Abrams tanks, Patriot battery radars, or CH-47
helicopters. They can also carry smaller oversize cargo, such as wheeled vehi-
cles (5-ton and smaller trucks, M-2 Bradley or Stryker vehicles, or high-mobility
multi-purpose wheeled vehicles (HMMWVs)). Of course, the C-5 can also carry
palletized bulk cargo such as ammunition, supplies and food, as well as troops.

15.2.1.2 KC-10A
The KC-10 is a dual-purpose aircraft. It serves as the largest airborne tanker in the
fleet and as a cargo carrier. For purposes of this study, AMC estimated that 23 of
the KC-10s would be dedicated to cargo missions. A single KC-10 can lift 83 tons
of cargo, although it can carry only a fraction of oversize cargo and cannot carry
any outsize cargo. The average payload assumed was 41.7 tons, and the aircraft
moves with a 445-knot block speed.

15.2.1.3 Civil Reserve Air Fleet


The CRAF consists of a number of different aircraft owned and operated by
various airfreight companies such as Federal Express (FedEx) and commercial
passenger airlines such as United Airlines and American Airlines. The CRAF pro-
gram was initiated in 1952 to provide on-call transportation assets in emergencies.
These companies contract to provide aircraft and crew during specified wartime
conditions in return for a guaranteed market share of DoD’s peacetime airlift (US
Congressional Budget Office CBO). There are three stages of wartime CRAF
conditions. The Commander US Transportation Command (USTRANSCOM),
370 W. L. Greer
with the concurrence of the Secretary of Defense, has the authority to activate
each stage.

• Stage I: minor operations—a pre-set number of commercial aircraft on 24-


hour notice. Stage I has only been used a few times, such as in 1990 to
support Operation Desert Shield deployments to Southwest Asia and in 2003
to support Operation Iraqi Freedom;
• Stage II: theater war—additional aircraft are released for more serious
emergencies. Stage II has also only been used once, for Operation Desert
Storm;
• Stage III: national emergency or two-theater war—even more aircraft made
available. This stage has never been used. Nonetheless, this is the stage
assumed for this study, i.e. supporting two nearly simultaneous major wars
abroad.

15.2.2 Proposed aircraft types


Proposed aircraft types represent various alternatives that compete to augment the
core fleet. These alternatives include proposals to extend the life of and/or replace
the C-141. This section first reviews a program to extend the life of the C-141s
and then introduces a variety of new aircraft alternatives.

15.2.2.1 C-141 service life extension program


The C-141, which can carry oversize and bulk cargo, was scheduled for retirement
because of its age and because outsize cargo carriers were thought to be more
useful to the airlift fleet of the future. To maintain the C-141 in the fleet would
have required a Service Life Extension Program (SLEP). The biggest cost items
in the C-141 SLEP were to re-wing the aircraft and add a new cab top to replace
serious corrosion damage. The maximum payload for the C-141 is nearly 45 tons,
although the average cargo carried is 27.5 tons at a 410-knot block speed.

15.2.2.2 C-Y
The C-Y is a hypothetical new replacement airlifter for the C-141. The C-Y would
have modern off-the-shelf engines and avionics, but would carry the same load as
the C-141. It was used in the analyses as a potential replacement for the C-141 as
each C-141 reached its terminal fuselage life span at 45,000 flying hours, a point
reached before the end of the 25-year period of the analysis. Thus, two acquisition
and operating costs are associated with the C-141 SLEP: the cost of SLEP itself
plus the cost of the replacement airlifter when the C-141 reaches the end of its
operational life.

15.2.2.3 C-17
The C-17 is smaller than the C-5, but larger than the C-141. The C-17 can carry
outsize cargo, but not as much as a single C-5. The C-17 requires no more room at
Acquisition of the C-17 aircraft 371
an airbase than the C-141 does, which is important when airbase space is limited.
It can also operate from shorter runways than can the C-5, affording an opportunity
to use more airfields in-theater and en-route. The C-17 can take off and land with a
maximum payload of 85 tons, although the average payload assumed in the study
was 48.3 tons (again, AMC sources have updated data). The C-17 block speed was
estimated at 423 knots.

15.2.2.4 C-5B+
This is a new aircraft proposed as an alternative to the C-17. It is identical to the
C-5B except that new GE CF6 engines would be used to reduce the high noise
and pollution levels associated with the C-5 TF39 engines. The study assumed the
same performance as the C-5B. Since the current C-5 line has been shut down,
part of its cost involves a restart of the C-5 manufacturing line at Lockheed.

15.2.2.5 Militarized 747 and 767


A different approach to replacing C-141s would be to purchase commercial air-
craft, such as Boeing 747-400s or 767-300s, modified (militarized) to handle some
oversize as well as the traditional bulk cargo.7 Unlike the CRAF, these aircraft
would be owned and operated by the USAF. The 747s were estimated to be able
to carry 73.1 tons each on average, while each 767 could carry 44.2 tons. Both air-
craft types can fly at 445-knot block speeds. Two different commercial derivatives
were used to explore whether large or small options might be best. As will be seen
in the results, the larger 747s were significantly more cost-effective than the 767s.
When the study began, militarized commercial aircraft were not included as
alternatives to the C-17. Some within the Air Force objected that these commer-
cial (militarized) aircraft could not carry outsize cargo and only limited amounts
of oversize cargo and therefore should not be considered candidates to replace
the C-141. Since the C-141 SLEP was an option as a replacement for the C-17,
and C-141s cannot carry outsize cargo, logic and fairness dictated that these mil-
itarized commercial derivatives be included. This inclusion turned out to play a
crucial role in the ultimate decisions made, as will be demonstrated. It also under-
scores the importance of not prematurely eliminating alternatives, and including
as broad a range as possible, since a wide set of alternatives can help highlight
important variables to include in the CBA.

15.2.3 Alternative fleets


A total of 26 alternative fleets were considered8 in the preliminary analysis illus-
trated in Table 15.2. The numbers in the table indicate how many aircraft of each
type were included in each alternative. Each entry consists of two numbers, sep-
arated by a diagonal line. The larger number represents total aircraft inventory
(TAI); the smaller number represents the primary aircraft authorized (PAA).9 Each
of the alternatives, identified with a numerical label ranging from 1 to 26, was
designed to have the capacity to deliver 52 MTM/D.
372 W. L. Greer
Table 15.2 Alternatives with 52 MTM/D capacity

Categories Alternative Numbers of aircraft (TAI/PAA)a


number
C-17 C-141 SLEP C-5B+ 747 767

One-aircraft 1 120/102 – – – –
alternatives 2 – 263/225 – – –
3 – – 102/87 – –
Two-aircraft 8 94/80 58/49 – – –
alternatives 9 94/80 – 21/18 – –
with reduced 10 94/80 – – 18/17 –
numbers of 11 94/80 – – – 30/28
C-17s
25 70/60 108/92 – – –
26 70/60 – 42/36 – –
18 70/60 – – 34/32 –
22 70/60 – – – 56/53

23 47/40 160/137 – – –
24 47/40 – 62/53 – –
21 47/40 – – 49/47 –
17 47/40 – – – 83/78

4 26/20 212/181 – – –
5 26/20 – 82/70 – –
6 26/20 – – 66/63 –
7 26/20 – – – 108/103
Two-aircraft 12 – 136/116 49/42 – –
alternatives 13 – 136/116 – 40/38 –
with no C-17s 14 – 136/116 – – 65/62
15 – – 49/42 42/40 –
16 – – 49/42 – 69/66
19 – – – 32/30 83/79
Illustrative 20 47/40 93/79 – 21/20 –
three-aircraft
alternative

Note
a TAI = total aircraft inventory; PAA = primary aircraft authorized.

The entries in Table 15.2 represent alternative additions to the core fleet. Each
alternative also includes the then-current core fleet of C-5s, KC-10s, and Stage III
CRAF forces. The alternatives are sorted into sets of categories: one-aircraft alter-
natives; two-aircraft alternatives with reduced numbers of C-17s; two-aircraft
alternatives with no C-17s; and an illustrative three-aircraft alternative. Alternative 1
has 120 C-17s; all other alternatives have fewer, or no, C-17s.
Note that the one-aircraft alternatives involve only one new type of aircraft
added to the core fleet. As already noted, Alternative 1 contains the full 120 C-17
fleet as programmed at that time by the USAF. Alternatives 2 and 3 proposed
Acquisition of the C-17 aircraft 373
substitution of the C-17s with a SLEPed C-141 fleet or a new-start C-5 fleet,
respectively.
The alternatives that included a reduced number of C-17s proposed a number
of different options including some C-17s in addition to one other aircraft. For
example, Alternative 8 has 94 C-17s and 58 C-141 SLEPs. The smallest number
of C-17s considered in this category was 26 C-17s, the minimum number already
obligated to DoD under contract without incurring termination penalties.
The two-aircraft alternatives with no C-17s include mixes of the C-141 SLEP,
C-5B+, 747, and 767 aircraft, taken two at a time, that provide the same MTM/D
as 120 C-17s.
The study also briefly examined a few three-aircraft alternatives, but it was felt
that DoD would probably not embrace a program that called for three new airlifter
types unless the benefits turned out to be dramatic for a reasonable cost. Although
the results for these alternatives did not turn out to offer dramatic benefits, they
served as a useful example for purposes of comparison.

15.3 Key inputs and models


This section summarizes the important data and considerations needed for the
analyses. It also points to the central role played by the MASS and ACAS models
to develop measures of effectiveness and the cost-oriented resource estimating
(CORE) model for cost measures.

15.3.1 Scenarios and delivery requirements


The careful choice of scenarios is essential to any analysis, since this provides the
context within which the system will be employed and a context for the evaluation
of each alternative’s effectiveness. The airlift analyses were conducted for two dif-
ferent types of scenarios: one massive and geographically distant, the other small,
close, but intense.
The selection of scenarios offers one of many opportunities for participants to
influence a military CBA. Proponents and opponents of various alternatives can
attempt to manipulate the scenarios in order to focus on features that show their
solution to its best advantage. It is critical to guard against this kind of bias. Scenar-
ios must be selected to explicitly address all important features, since the purpose
of the analysis is to provide objective and unbiased insights to the decision-maker
on how the system is likely to perform in critical scenarios relevant to national
defense.
In this case, the first scenario considered involved a concurrent, sequential
deployment of forces from the United States to meet the needs first of a Major
Regional Contingency (MRC)-East and subsequently of a MRC-West.10 The des-
ignations East and West refer to the US coast from which most of the airlifters
would stage for the respective MRC operations.11 Over 550,000 tons of cargo was
required within a 90-day period for both MRC scenarios.
The second scenario involved a Lesser Regional Contingency (LRC)12 that
began with a brigade-size airdrop followed by air delivery of cargo and additional
374 W. L. Greer

250
Cargo Required (ktons)

30 days
200 90 days

150

100

50

0
Outsize Oversize Bulk
Cargo Type

Figure 15.2 MRS airlift delivery requirements for MRC-East and MRC-West
combined.

troops to the target country. In this second scenario, CRAF was not activated and
the reserves were not mobilized. A total of 50,000 tons of cargo was required in
four days, in addition to that delivered in the initial brigade airdrop. This case
allowed the analytic team to test the robustness of the alternatives to very different
requirements from those addressed by the MRCs.
The MRC scenarios dominated the analyses conducted. These scenarios
required the full support of the airlifter fleet, including full augmentation by
CRAF. Most of the analyses discussed in this chapter refer to the first set of sce-
narios. It turns out that the LRC scenario added little to the insights obtained from
the two MRC scenarios.
The airlift requirements from the mobility requirements study (MRS) are sum-
marized graphically in Figure 15.2 for two different time periods: 30 days and
90 days. Virtually all outsize and oversize cargo was required to be in-theater
within the first 30 days. There are only minor differences between requirements
for 30-day delivery and for 90-day delivery. To meet continuing resupply require-
ments, the remaining bulk cargo in each case continues to build over time. This
emphasis on rapid delivery of outsize and oversize cargo will become the basis of
measures of merit (or effectiveness) in the study.

15.3.2 Aircraft characteristics


Several key aircraft characteristics are essential in determining load and delivery
of required cargo. These are discussed in this section.

15.3.2.1 Range-payload relationships


Figure 15.3 summarizes the range-payload curves for each of the strategic airlifter
types. The payload is limited at short flying distances (ranges) by the maximum
cargo weight the aircraft can lift. The maximum range is defined where an aircraft
carries its maximum fuel load at takeoff and no cargo. At intermediate distances,
Acquisition of the C-17 aircraft 375

140
747
120
Maximum Payload (tons)

C-5
100
C-17 KC-10
80
767
60
C-141
40

20

0
0 500 1000 1500 2000 2500 3000 3500 4000 4500 5000 5500 6000
Range (nmi)

Figure 15.3 Range payload curves for airlifter types.

the range-payload relationship illustrated in Figure 15.3 depends on tradeoffs


between cargo and fuel loads.
Once these relationships were established, they were used to determine the
best refueling locations en route to allow an airlifter to carry as much cargo as
possible. While actual operations would undoubtedly provide opportunities for
in-flight refueling, payloads were assigned with the conservative assumption that
the airlifter would have to use en-route bases for refueling.13 As Figure 15.3
reveals, the 747 and KC-10 airlifters are the least constrained by the range-payload
relationships.

15.3.2.2 Use rate


Surge use rates14 are the upper limit in terms of flying hours that a fleet of aircraft
can fly on any given day, applied to the first 45 days of operations. Table 15.3
reports surge use rates for each airlifter type analyzed. These numbers became
major sources of disagreement as the study proceeded, resulting in sensitivity anal-
yses using lower C-17 use rates and higher C-5B+ use rates. While each excursion
requires time and resources, the use of sensitivity analysis to investigate key rela-
tionships and variables is one of the advantages of analysis. Sensitivity studies
are usually conducted on a range of issues with priorities established by the key
decision-makers. The need for sensitivity analyses should be anticipated and built
into study timelines.

15.3.2.3 Maximum on ground and ground times


The space available at airbases is referred to as “maximum on ground” (MOG). It
measures the maximum number of each specific type of airlifter that can be located
at a given base to receive service at the same time.15 The values for each air-
lifter type were taken from a list of worldwide MOG allotments for each airbase.
376 W. L. Greer
Table 15.3 Airlifter surge use rates

Airlifter category Airlifter type Surge use rate


(hrs/day/PAA)

Military C-5A/B 11.0 (average)


C-141 SLEP 12.5
C-17 15.2
C-5B+ 12.5
KC-10 12.5
Commercial derivative Militarized 747 or 767 12.5
CRAF All 10.0

Table 15.4 MOG estimates maximum number of aircraft on ground simultaneously (by
theater and aircraft type)

MRC theater C-17 C-141 SLEP, 767 C-5, KC-10, 747

East 26 26 15
West 24 20 11

Table 15.4 summarizes MOG estimates for the airlifters analyzed at an aggregate
theater-wide level. Note that the largest airlifters consume nearly twice the MOG
space of the smaller ones. For example, a maximum of 26 C-17 aircraft can be on
the ground simultaneously in the East-MRC theater of operation.16
Another important input is the ground time required to load and unload cargo
and refuel aircraft. This reflects the amount of time each aircraft uses its allocation
of MOG. Ground times vary by mission (en-route refueling, on-load, and off-load)
and aircraft type. The C-5s had the longest ground times (3.75 hours on-load;
3.25 hours en route and off-load), especially when compared to 2.25 hours for en
route, on-load, and off-load for the C-17 and C-141. Boeing’s commercial deriva-
tives had the longest on-load times (4–5 hours) but the shortest en route times
(1.5 hours).17

15.3.3 Loading and transportation effectiveness analysis


The main models used for the estimates of how much cargo and how many pas-
sengers were delivered were ALM, MASS/AFM, and ACAS. These models are
summarized briefly to give a sense of the assumptions used in the analysis.18
Further detail on each of the models can be found in Appendix 15.1.

15.3.3.1 Airlift loading model


The ALM was used to determine the average cargo payload carried by each air-
craft in the fleet. The MASS/Airlift Flow Model (AFM) then used this result to
simulate the transport. ALM used cargo from the entire TPFDD for these esti-
mates. The cargo was differentiated according to 26 different categories called
“commodities.” Each commodity could contain vehicles, pallets, and passengers
Acquisition of the C-17 aircraft 377
(PAX). The sequence in which the aircraft were to be loaded had to be specified
as well as aircraft cargo storage dimensions, cargo door opening sizes, number
of passenger seats, and maximum payload allowable by weight. Loading stopped
when the weight limit was reached, or, more commonly, when cargo fully filled
the cargo space in the aircraft.

15.3.3.2 MASS/AFM
The MASS/AFM was a large Monte Carlo simulation19 that scheduled and exe-
cuted each airlift mission by tail number. It incorporated a great deal of realism,
such as specific routes used, airbases with MOG limitations, times for loading
and off-loading, and crew duty days. It used the ALM loading results along with
the nature of the TPFDD cargo and PAX at individual airbases to load the cargo.
Only that part of the TPFDD available at a specific base on a specific day (plus
any undelivered leftover cargo from previous days) was available for loading. The
output was tons delivered per day.

15.3.3.3 Airlift cycle analysis spreadsheet


The ACAS model was a simplified approximation to the analyses produced by
MASS/AFM and was used to conduct the large number of excursions needed in
the study. It used aggregate data and aggregate MOG values, calibrated against
MASS/AFM to provide confidence for these numerous excursions.

15.3.4 Cost analysis


The total costs for the various airlifters were estimated from numerous sources
and models. The study included both the acquisition costs for new aircraft (C-17,
C-5B+, C-Y, 747, or 767) and for the C-141 SLEP, as well as 25-year operating
costs. All costs were expressed in constant FY 1993 dollars and were discounted
to net present value (NPV) using the then-current standard discount factor of
4.5 percent per year.20
The study team used detailed cost information from numerous sources—the
C-17 Program Office, airframe and engine contractors, and the OSD Cost Analy-
sis Improvement Group (CAIG)—to arrive at its own independent cost estimates.
When no data were available, historical data trends were used to supplement these
data sources. In general, IDA cost estimates for the C-17 were somewhat higher
than those of the Program Office, an issue that slowed progress during parts of
the study while these differences were debated. The CAIG agreed with the IDA
estimates, and these were the ones retained in the report.

15.3.4.1 General
The basic approach was first to establish acquisition schedules and C-141 retire-
ment schedules for all alternatives and to make assumptions regarding the
active/reserve aircraft mix for airlift. The study team used the C-17 schedule
programmed at that time by the USAF, with C-141s retired at such a rate as
378 W. L. Greer
to maintain a constant MTM/D capacity. Other airlift alternatives (C-5B+ and
commercial derivatives) were acquired at rates consistent with their manufactures’
capabilities, with the C-141s also retired at a rate to keep MTM/D constant.21

15.3.4.2 Acquisition
The study used the USAF C-17 acquisition schedule, with a maximum acquisition
of 16 per year. Historical cost data for the C-17 were used to generate a cost-learning
curve to predict future costs for C-17s. Appropriate adjustments to historical trends
were added as needed, such as the cost for extra weight to strengthen the wings
and higher costs associated with more realistic engine estimates for all lots after the
first ones. IDA estimated the full C-17 program acquisition cost at approximately
US$23 billion, about US$3 billion over the Program Office estimate. In the end, the
higher-cost IDA results were used in the report, with a comparison to the Program
Office estimates included as an excursion.
At the time the study was conducted, 26 C-17s were under contract with long-
lead funding. The US government would have incurred a cost penalty for breach of
contract if it failed to buy 26 production C-17s. Thus alternatives that included less
than 26 C-17s were levied an additional termination fee associated with breaking
the contract (US$1.5 billion).
Lockheed Martin Aeronautical Systems provided cost estimates for the alter-
native of restarting the C-5 line as well as for refurbishing the C-141s in a C-141
SLEP. IDA used these and other historical Lockheed data and made adjustments as
required. Examples of adjustments include adding hush kits to dampen the sound
of the engines and adding extra material to the fuselage. The cost to restart the C-5
line was estimated at US$750 million. New C-5 engines were used to reduce noise
levels below those mandated by the Federal Aviation Administration (FAA).
A recently completed Scientific Advisory Board (SAB) report had recom-
mended that the weep hole cracks in then-current C-141s could be repaired and
their life extended to 45,000 flying hours. When the C-141 reached its 45,000-
flying-hour life limit, the SAB determined it had to be replaced, and the cost of a
new C-Y replacement was imposed to include a US$3.5 billion development cost.
Boeing supplied basic cost data for the 747 and 767 commercial derivative air-
craft alternatives. Production rates from 6 to 12 per year for military acquisition
were used, based on Boeing production capabilities and competing markets for
new 747 or 767 aircraft. An additional 20 percent of the acquisition price was
added to account for reinforced floors and rollers and widened side doors, for
an estimate of US$155 million for the militarized 747 and US$88 million for the
militarized 767. Since the relevant production lines were already open, no develop-
mental costs were imposed. Over the course of the study, these same commercial
derivative aircraft also became known as “non-developmental airlift aircraft.”

15.3.4.3 Operating and support


The primary tool used to estimate O&S costs was the USAF CORE model. This
is the standard model used for O&S cost estimates in the cost community (AFI
Acquisition of the C-17 aircraft 379
Table 15.5 Summary of O&S costs per PAA

Aircraft Number of flying CORE cost estimates


hours per year ($million/year/PAA)

C-17 1,427 10.8


C-5B or B+ 660 9.49
C-5A 325 6.75
C-141 SLEP 1,178 7.72
747 900 8.08
767 900 4.98
KC-10A 550 4.62
C-130E/H 637 4.01

65-503 1994). Use of the model requires aircraft to be sorted according to whether
they are in the active, associate or reserve fleets. It takes into account the number of
flying hours per year, the cost of the fuel burned, the cost associated with personnel
needed (crew, maintenance), software costs, training costs, the cost of spares, and
the cost of contractor logistics support.
O&S cost estimates from the USAF CORE model for all the aircraft types
are summarized in Table 15.5. The commercial derivative aircraft were assumed
to have contractor logistics support, the cost of which was included in the cost
estimates for alternatives containing commercial aircraft derivatives.

15.4 Initial cost and operational effectiveness results


This section summarizes the initial results presented to the Strategic Systems
Committee (SSC) and to the DAB in August 1993, along with their reaction.22
Throughout the early stages of the analyses, key decision-makers were periodi-
cally apprised of the methodology, inputs, and results. Frequent meetings were
held between the study team and working levels representing the interested DoD
organizations. AMC was in continuous contact, providing all the MASS/AFM
computer runs with inputs provided from IDA. This is an essential part of any
analysis, which is both a process and a product. Close communication allows
the analyst to understand the critical issues as viewed by the various players and
decision-makers, as well as to provide the decision-makers with an opportunity to
help guide the analysis.
The initial cost and operational effectiveness results are summarized in
Figure 15.4. These results were developed using cargo and equipment data speci-
fied in the airlift portion of the TPFDD files from the mobility requirements study
that had been completed the year before.

15.4.1 Initial results


All 26 alternative fleets of aircraft under consideration are displayed in Figure 15.4
according to their effectiveness (tons of outsize cargo23 delivered in a 30-day
period) and cost (FY 1993 25-year discounted life-cycle cost). The numbers beside
380 W. L. Greer

Current trucks
70~c=========~==========~~--------~~-----------,--------~
C-17 :
0 COMMERCIAL DERIVATIVES -----------L-------------------------
65 6, C-141 SLEP i TPFDD
D C-5B+
Outsize cargo delivered in 30 days (kilotons)

60
Equal cost
effectiveness

55 Alternative 1
cost

55

1 26
24
55 18 8
10
5 9
11 25
21 22
55 MRS 3
6 20
17 15
16 4 23
13 12
55 7
2
19
14
55
55 55 55 55 55 55

25-year cumulative cost ($B)

Figure 15.4 Comparison of cost and effectiveness of alternatives with current trucks.

each symbol identify the alternatives. To assist with visual interpretation, alterna-
tives that contain any C-17s contain a small black dot; those with any C-5B+ have
a square; those with any C-141 SLEP have a triangle; and those with commercial
derivatives such as 747s or 767s have an open circle. For example, Alternative 1
with only a small black dot includes only C-17s (120 according to Table 15.2). In
contrast, alternative 18 contains both C-17s and commercial derivatives (70 C-17s
and 34 747s according to Table 15.2) as indicated by the dot inside the larger open
circle.
For reference, the outsize cargo requirement of the 30-day scenario set by the
MRS is shown as a horizontal dashed line. Any fleet alternatives below the line
would require additional aircraft to satisfy this mission requirement. The vertical
dashed line shows the cost of the 120-C-17 Alternative 1. Any fleet alternatives
to the right have a higher cumulative cost than this fleet made up exclusively
of C-17s. The sloping dashed line indicates a locus of alternatives that have the
same cost-effectiveness as Alternative 1 (i.e. same “bang for the buck” or ratio of
effectiveness to cost).24
This chart clarifies several things. First, the C-141 SLEP alternatives (i.e. the
triangles) appear to be among the worst candidates relative to the C-17 fleet
(Alternative 1), both from a cost and from an effectiveness perspective. To make
them more effective by adding either C-141 aircraft or another aircraft that is
Acquisition of the C-17 aircraft 381
part of the mix, additional costs would be incurred.25 This seems to eliminate a
whole category of alternatives immediately. Second, the fleet alternatives of C-17s
combined with commercial derivatives (open circles with concentric black dots)
appear to be close in cost-effectiveness. This catches our attention and elevates
our consideration of militarized 747s and 767s. In particular, the 747 appears the
better of the two commercial derivatives. Finally, the C-5B+ (squares) alternatives
appear slightly less cost-effective in all their mixes than the C-17 mixes, but run a
close second. They cannot easily be dismissed from this chart.
Based on these initial results, the DAB was impressed with the potential for
using commercial derivatives such as the 747 at lower cost than the C-17 program.
The study team was tasked with examining several excursions and reporting back
as soon as possible. At the same time, C-17 proponents perceived the results as a
threat to their program and reviewed the assumptions of the analyses in detail.

15.4.2 Changing assumptions: transportation of new army trucks


As often happens when unexpected results are presented, those most affected by
the conclusion that commercial aircraft could be the most attractive alternatives
began to search for flaws in the methodology and errors in the inputs. Hence they
made a major discovery that seriously challenged the analysis to date. This dis-
covery concerned the ability of commercial aircraft derivatives to transport new
army trucks.
The preliminary analyses had assumed current (meaning circa 1993) army
trucks (hence the subtitle in Figure 15.4 of “Current Trucks”) were to be airlifted.
Actually, the Army planned to replace their 2.5- and five-ton trucks with 10,843
new ones, members of the family of medium tactical vehicles (FMTV). Since
trucks constituted a major component of the TPFDD, with thousands being air-
lifted to theater, the effect on the analyses could potentially be significant. These
new trucks differed in two significant ways as far as air transportability is con-
cerned: axel loads and truck height. Each is discussed next with its effect on
transportability in the different alternatives.
The axel loads on some of the new trucks are much higher than those of the
corresponding current trucks. These loads are so heavy, in fact, that a stronger
reinforced floor would be needed on the militarized 747s to carry them. Thus
additional costs for reinforcing commercial derivatives would be necessary if these
trucks are to be carried. Since the CRAF 747s are not reinforced, they cannot carry
any new trucks at all, thus reducing their utility in moving some of the oversize
cargo.
Even more significant is truck height. There are two versions of the new trucks:
one designed for airdrop or low-altitude extraction by parachute and the other for
normal loading and unloading from a stationary airlifter. The difference is that the
airdrop-capable versions have collapsible cab tops that permit them to be easily
loaded through the nose of a 747; the others (the vast majority) have fixed non-
removable cab tops whose heights prevent the trucks from entering through the
nose door. The nose door height cannot be changed because the 747s nose space
382 W. L. Greer
is needed for cockpit and crew and therefore limits door height. The side doors
would be tall enough, but not wide enough (without alteration). The new FMTVs
would, however, fit through the doors of C-17s, C-5s, C-141s, and KC-10s.
The study team investigated possible ways to address this impasse. First, the
Army could spend an additional US$700 per truck to ensure that all trucks have
collapsible cab tops. This would ensure that they fit on 747s. The Army objected,
not just on the basis of cost, but also on operational utility. The Army deliberately
chose non-collapsible cab tops for the next-generation trucks to afford greater pro-
tection in chemical environments. Thus this solution did not seem reasonable.
Second, if the cab tops could not be changed, perhaps the 747 side doors could
be widened. Boeing engineers estimated that the cost to widen the doors and pro-
vide additional roller reinforcements would be about US$10 million per aircraft.
The materiel-handling equipment could raise the trucks to the appropriate height
from the ground, and the trucks would have to maneuver with a sharp turn as
they entered the fuselage. Thus, greater care in moving the trucks through the side
doors would be needed, with some risk and longer loading times. This seemed
technically possible, so it, along with the extra costs entailed, became the new
commercial derivative design: wider side doors with additional roller and floor
reinforcements.
The new trucks also affected CRAF. Civilian aircraft forced into service could
not be expected to carry the new trucks because of their extra weight.
The results under the new assumptions are shown in Figure 15.5. The results
assume either collapsible cab tops for the new trucks (FMTVs) or an extended
side door for the Boeing 747.
Figure 15.5 is similar to Figure 15.4, except that it captures the fact that all
alternatives will have lower delivery capabilities, since CRAF cannot carry the
new army trucks.26
Figure 15.5 also contains a new alternative, Alternative 21a. This alternative
was added after a second briefing to the DAB. At the DAB it was noted that,
with only three additional 747s, Alternative 21 (with 47 C-17s and 49 747s) could
be raised up to the MRS line and therefore meet the requirement and level the
playing field.27 With this, Alternatives 10 (with 94 C-17s), 18 (with 70 C-17s),
and 21a (with 47 C-17s) could be seen as providing at least as much capability as
demanded by the MRS, but at lower cost than Alternative 1 (the programmed fleet
of 120 C-17s). This chart reinforces the notion that there are less costly alternatives
to the C-17 program, most involving militarized commercial derivatives.
Some decision-makers wondered what the results would be if the new trucks
could not be accommodated aboard militarized 747s. That was a fair question
deserving an answer. Figure 15.6 summarizes results for that situation for a
selected set of alternatives.
As Figure 15.6 clearly illustrates, the viability of the commercial derivatives
is critically associated with the ability to load the Army’s FMTV trucks onto the
modified aircraft. Under these assumptions, the combined fleets of 747 and C-17
(Alternatives 10, 18, and 21) that earlier had cost-effectiveness ratios similar to
Alternative 1 (made up exclusively of C-17s), no longer do so and even fall below
New, collapsible cab top trucks or extended 747 side door
70~r=========~===========r~--------~~ ~ ----------,---------~
C-17 :
0 COMMERCIAL DERIVATIVES -----------~-------------------------
TPFDD
65 ~ C-141 SLEP
0 C-5B+
Outsize cargo delivered in 30 days (kilotons)

* 21a adds 3 new 747s to Alternative 21


Alternative 1
60 cost

Equal cost
55 effectiveness

50 ~-­
-~.,...,-',.,.

45 1
t -------[il-----
26
18 24

-----------------------e·~~::::~t~-~--~--~----~-A------
10 8
21a* 11 5
MRS 21 9 25
40
c6 :
~--- -o 22 3 Q

35 ~~-~--
-----· e - 17
®C
16
15
i
:
4
20
23~
/ (i) 13 I 12
0 7
19
o : 2
30~----------~----------~------~(\
14
~~:--~~~------~--------~
45 50 55 60 65 70
25-year cumulative cost ($B)

Figure 15.5 Comparison of cost and effectiveness of alternatives with new army
FMTV trucks.

60~r=======~=========c~------~-------,--------,

~~~MERCIAL DERIVATIVES;~
j
Q Equal cost
55 O
~ C-141 SLEP effectiveness
C-58+
Outsize cargo delivered in 30 days (kilotons)

50
Alternative 1
cost _,,.; .... "' --
_......... -......... -
.... -,.."""

45
............
..
_1 -------[:] 26
I

E1 E1 9
18 10
@- ,''- ® : 24
MRS
40 ------------------- =: -:(!f ~ ~:::: =------~ - - lG -----------------------
21
15
3
5

----- c®18 10
&
35 ---------- • 23

c
13
---- ®
21 15
.0 2
30 Fleet Composition (TAI)
Legend:
.0
Alternatives
FMTV can load on modified 747
13 A/C 1 2 3 5 9 10 13 15 18 21 23 24 26
C-17 120 – – 26 94 94 – – 70 47 47 47 70
25 C-5B+ – – 94 82 21 – – 49 – – – 62 42
FMTV do not load on modified 747 747 – – – – – 18 40 42 34 49 – – –
C-141 1
20L_________ ________ _ L_ _ _ _ _ _ _ _ __ SLEP
_ _ _–_ _263
_L _ –_ _–_ _– _ _– 136 – – – 160 – –

60
~ ~L_ ~

45 50 55 65 70

25-year cumulative cost ($B)

Figure 15.6 Comparison of selected alternatives with different assumptions about


MTV loading on militarized 747.
384 W. L. Greer
the minimum effectiveness MRS scenario line. If commercial 747 derivatives
cannot be loaded with new army trucks, then their effectiveness plummets.

15.5 Sensitivity analyses of other underlying assumptions


Additional analyses were conducted to gauge the sensitivity of the results to other
assumptions. These are discussed here.

15.5.1 MOG and use


The analyses presented so far include specific assumptions about the space avail-
able at “en-route” and in-theater airbases and about utilization rates of aircraft.
The space available for aircraft at airfields en route to a final destination is a func-
tion of the assumptions concerning other activities at those bases and in the theater
and thus is subject to considerable uncertainty. The assumption regarding the uti-
lization rates of aircraft depends on funding decisions and maintenance policies
to provide specified levels of operational performance and could, in principle, be
strategically altered. To gauge the sensitivity of the results to specific assumptions
and data inputs, the study subjected both of these key assumptions to sensitivity
analysis.
Although the values used in the study allowed for reasonably robust operating
conditions, C-17 proponents cited less-than-ideal operating conditions as a key
advantage of their aircraft. The C-17, with a MOG value close to that of the C-141,
requires less space than the C-5 or 747. Since considerable subjectivity is inherent
in the values used for any study, it seemed reasonable to explore how sensitive the
results were to changes in this input. The study team thus considered the following
three MOG conditions:

• robust, corresponding to the AMC values used in previous charts;


• moderate, with conditions in MRC-East that approximated the 1990 Desert
Shield MOG values;
• constrained, which included the moderate MOG for MRC-East just noted and
reduced the MOG in MRC-West by removing airfields near the conflict border
and reducing MOG elsewhere in the theater by 50 percent.

In all cases, as the MOG was reduced, the ability of the C-17 to land and take
off on short airfields, to back up, and to take up less space than C-5s and the com-
mercial derivatives made it more favorable. This is illustrated in Figure 15.7 for
five of the alternatives. For clarity, only the most robust and constrained extremes
are shown while the moderate MOG case lies in between.
Changes in the MOG have a powerful influence on the results. In the constrained
case, none of the alternatives meet the MRS requirements, although Alternative 1
with the most C-17s suffers less than the other fleet alternatives with C-17s and
other, more space-intensive airlifters. In terms of systems thinking, this clearly
points to the possibility of investing in adequate infrastructure to improve MOG
as a new alternative that can be integrated with various combinations of aircraft.
Acquisition of the C-17 aircraft 385

50~------~------~------~------~------~------~------~----~

----~------------~---------------
Robust MOG
C-17 UTE equal cost
rate effectiveness
45 reduction 1 26
Outsize cargo delivered in 30 days (kilotons)

Robust MOG
MRS based on
- robust MOG
®-----------
- 15.2 C–17 UTE rate
16 [!] 26
40 -~:::~~~-~~~:~::: __________ c:J 3___________________________________ _
® 16 1

23
35
___ _ --
_,..._.,.,.
_.,.,._ ....

.
...
_,.._ ...
....

Constrained MOG
equal cost _____1 -----
30 C-17 UTE
effectiveness
------ rate
reduction

Constrained MOG
16 26
25
Fleet Composition (TAI)
Alternatives

A/C 1 3 18 23 26

20
C-17
C-5B+
120


94
70

47

70
42 D3 Note: C-17 surge UTE rates
23
747 – – 34 – –
C-141
- baseline 15.2 hrs/day
SLEP – – – 160 – - excursion 12.5 hrs/day
15L-----~------~------_L ______ _ L_ _ _ _ _ _~------~------L-----~

52 54 56 58 60 62 64 66 68
25-year cumulative cost ($B)

Figure 15.7 Impact of reduced MOG and reduced C-17 use rates.

It also cautions against assuming that lower cost alternatives always provide the
best payoffs.
Another key assumption in the analysis was questioned by 747 proponents who
objected to the assumption of higher C-17 use rates. The basic analyses assumed
a 12.5-hour-per-day use rate for the 747s and a 15.2-hour-per-day rate for the
C-17s under wartime surge conditions. The use rate attainable by the C-17 was
still unproven when the study was conducted, so some doubted it would exceed
that achievable by commercial aircraft.
Figure 15.7 also illustrates the effect of a smaller use rate for the C-17. The
reduced use rate was assumed to be 12.5 hours/day—the same as the assumed use
rates for the C-5B+ and the militarized commercial derivatives. The smaller use
rate reduced effectiveness and also cost (because of fewer flying hours per year
associated with such a lower use rate). The overall effect was noticeable but was
not nearly as dramatic as that seen in the MOG assumption excursions. While it
was reasonable for the study team to be skeptical about such a large projected use
rate for an unfinished aircraft, deviations from that projection did not dramatically
affect the cost-effectiveness measures for the C-17. Interestingly, subsequent tests
two years after this original study validated the 15.2 hours/day value, rendering the
arguments moot. Study teams, however, can never know these things in advance
and are well served to look carefully at all key assumptions that could influence
the results.
386 W. L. Greer
15.5.2 Assuming airlift capability other than 52 MTM/D
After the first DAB meeting in August 1993, there was a growing sense on the part
of some decision-makers that commercial derivatives in conjunction with some
number of C-17s might be the best solution. When the initial results were briefed
to the DAB, additional airlift capability assumptions were requested. These cases
were developed to determine how few C-17s might be acceptable if that program
was reduced from the 120 aircraft originally assumed in Alternative 1.
IDA was asked to assess how well a number of new alternatives would fare.
The result of one set of analyses is summarized in Figure 15.8. In this case, 10
new alternative fleets are introduced, with MTM/D values ranging from 46 up to
55. An attention to cost was also used to make virtually all the new alternatives
no more costly than Alternative 1. Figure 15.8 illustrates cost and effectiveness
measures for Alternative 1, various mixes of C-17s and commercial derivatives,
and mixes of C-17s and C-5B+s. Since by the time the DAB was expected to
reach its decision it was expected that 40 C-17s would already be delivered or
under construction, many of the new alternatives included 40 C-17s, and none
involves fewer than that number. As Figure 15.8 shows, the set of mixes of C-17s
with commercial derivatives is roughly equal in cost-effectiveness to Alternative
1 with exclusively C-17s. Moreover, for roughly the same budget, the commercial
and C-17 fleet mixes are well above the effectiveness of the C-17 and C-5B+
mixes.28 This chart confirmed the sense derived from earlier charts that a mixed
fleet of C-17s and commercial derivatives might have merit—provided they could
carry the new army trucks.
Some decision-makers felt that a fleet size capable of generating a transportation
capability of 52 MTM/D was too small. After all, the C-17 program had been born
when the requirement was 66 MTM/D. As previous charts have shown, all alterna-
tives fell significantly below the effectiveness demanded by the TPFDD line. Thus
IDA was subsequently asked to examine the possibility of alternatives that generated
higher MTM/D. Figure 15.9 summarizes the results of two different sets of alterna-
tives: those with 52 MTM/D and those with 59 MTM/D, a value with historical roots
when new airlift programs were initiated several years prior to this study period. The
59 MTM/D alternatives are all identified with the prefix “E” on the chart. It is inter-
esting that the same frontier line observed earlier still carries over to these higher
MTM/D cases. Namely, all alternatives fall on or below the equal cost-effectiveness
ratio line that runs through Alternative 1, within statistical uncertainty.29

15.5.3 Another scenario: LRC-short


The foregoing analyses assumed two different MRCs. There may be a need, how-
ever, for forces to meet LRCs. The one selected for analysis was LRC-Short,
involving sustained airlift to a relatively close theater (short range, hence the name
Short), following an airdrop on Day 1. The baseline assumption is that all major
airbases are seized by the airdropped troops and are available for operations. Since
the MRC cases showed MOG to be a critical input, the study also examined an
excursion in which 50 percent of these same airbases are unavailable.
Acquisition of the C-17 aircraft 387

New, collapsible cab top trucks or extended 747 side door


50,-------,-------,--------,-------,-------,r-------,-------,-------~

Fleet Composition (TAI)


C-17/Commercial Derivative Mixes
Alternatives Alternative 1
48 cost
A/C 1 110 111 112 115 116
C-17 120 40 40 40 80 100 Equal cost
Outsize cargo delivered in 30 days (kilotons)

46 747 – 55 64 74 27 14 effectiveness
MTM/D 52 52 54 55 52 52

1
44

112
,/®_/-; 116
42
.A!> 115
125
--------
40 MRS __...- ---- II
-----------:=~~~~~ ~--------------------~-----------------------
111
124
------ 110
®
Fleet Composition (TAI)
38 C-17/C-5B+ Mixes
Alternatives
---------------- A/C 1 121 122 123 124 125

36 123 C-17 120 40 40 40 80 100

El C-5B+
MTM/D

52
30
46
35
47
42
48
18
50
16
52
122
121
34

32 C-17
COMMERCIAL DERIVATIVES 0
C-58+ D
30L_----~----~------~----_L--~==~====~====~====~
50 52 54 56 58 60 62 64 66
25-year cumulative cost ($B)

Figure 15.8 Cost and effectiveness comparisons of new alternatives.

Figure 15.10 summarizes the results for LRC-Short. In this case, the effective-
ness measure is the total cargo delivered (not just outsize) in eight days after the
initial first-day airdrop. The 25-year total cost is the same as before. The four-day
TPFDD is used for purposes of comparison, although it has not been achieved
even in eight days by any alternative.
A comparison of the amount of cargo delivered for robust (baseline) MOG con-
ditions and for a 50 percent loss in airfields shows the same general results that
appeared before in the case of the MRC scenarios: alternatives with fleets with C-5s
and 747s suffering a greater reduction in capability than those with C-17s when
MOG is constrained. Those with the C-5B+ seem to suffer the greatest reductions.

15.5.4 Cost excursions


All the sensitivity analysis excursions discussed to this point principally focused
on different effectiveness assumptions including scenarios, use rates, MOG, and
the composition of alternative fleets of aircraft. The cost estimates included in the
study represented the best forecasts generated by the cost team. While there are
many uncertainties in the cost estimates, one that became obvious to the DAB was
the yearly acquisition cost assumptions for each of the alternatives.
388 W. L. Greer

New, collapsible cab top trucks or extended 747 side door


70~------~--------~~--------~--------~------~--------~------~
I
Fleet CompositionI (TAI)
-------------------r----------------------------------------------
I (52 MTM/D)
Basic Alternatives TPFDD

65 A/C 1 3 5 9 10 15 18 21 24 26
C-17 120 – 26 94 94 – 70 47 47 70 _..
C-5B+ – 94 82 21 – 49 – – 62 42 Equal cost – _..
Outsize cargo delivered in 30 days (kilotons)

effectiveness
60
747 – – – – 18 42 34 49 – –
... --
----
.... --
Fleet Composition (TAI)
______ ....... ----
Expanded Alternatives (59 MTM/D)
.......
55 A/C E-1 E-2 E-3 E-4 E-7 E-8 E-9 E-10 E-3
.,.•,,.,. .... --
E-1 ... --
C-17 176 120 120 120 94 94 70 47
C-141 SLEP – 124 – – – – – – __ . . . . -- ®E-4 E-7
E-2
50 C-5B+ – – 48 39 69 – –
.... --

747 – – – – 57 73 89
E-8
___ .,. .......... "' .... "' @
E-10 / (!) E-9
45 ~[!] 25
18 10
_.. -f
1
24
-"'"' I r:1
Sl ..--"' @ :5 L!J[!] 9
MRS
40 --------~~~---- ~ - ~ --------------------------------------------
,_.--(!}
21 3 :
. . ----"' [!]
15
:I
35 I
I Alternative 1
I
I cost
I
I
1
30L---------~--------~ ---------L--------~----------L---------~--------~
50 55 60 65 70 75 80 85
25-year cumulative cost ($B)

Figure 15.9 Comparison of 59 MTM/D expanded capacity alternatives with the nominal
52 MTM/D alternatives.

In the cost estimates shown earlier, each alternative was developed estimating
the costs of efficient production-lines, independent of annual budgets. A more real-
istic approach would be to constrain all annual expenses over the six-year period
of the Future Years Defense Program (FYDP) to be no greater than those associ-
ated with the C-17 program itself, as expressed in the then-current C-17 Selected
Acquisition Report (SAR). This would force many of the alternatives analyzed to
stretch their acquisition profiles to remain under the spending cap.30 To test how
stretching out programs and the concomitant increases in acquisition costs would
influence the outcome of the analyses, a sensitivity analysis was conducted. The
results for selected fleet alternatives appear in Figure 15.11.
As Figure 15.11 shows, fleet alternatives with mixes of C-17 and commer-
cial derivatives all show increases in total cost of US$1–2 billion. The cost of
Alternative 1 also increases slightly, since the cost team did not feel that the full
120 C-17 program could stay within the cost caps of the C-17 acquisition bud-
get. The C-5B+ fleet alternatives do not increase in cost in this excursion, since
the start-up time for the new C-5 line takes several years and removes it from
competition with other aircraft types in those alternatives. As the figure shows, even
Acquisition of the C-17 aircraft 389

New collapsible cab top trucks or extended 747 side door


55~--------~--------~--------~----------~------~

4-Day (and Beyond) TPFDD


50 ------------------------------------------------------------------

45
Total cargo delivered in 8 days (kilotons)

21
® ®18 1
40 10 ® 5
® [!] Baseline MOG

c
35 15 1
3 ®
30
0
® ®
25 18
® 10
21 One-half MOG
20 excursion

[!] 5
15

c 0
Fleet Composition (TAI)
15 3 Alternatives
10 A/C 1 3 5 10 15 18 21
C-17 C-17 120 – 26 94 – 70 47

5
0 COMMERCIAL DERIVATIVES C-5B+ – 94 82 – 49 – –
fj. C-141 SLEP 747 – – – 18 42 34 49
0 C-58+
0
45 50 55 60 65 70
25-year cumulative cost ($B)

Figure 15.10 Comparisons of alternatives in LRC-Short.

though total lifetime costs do increase if the SAR serves to limit near-term fund-
ing, the C-17/commercial-derivatives fleets still dominate the cost-effectiveness
comparisons.
The relative contribution from the different cost elements can be seen in
Figure 15.12. Figure 15.12 also illustrates the effect of different discounting
assumptions on the cost estimates. Two of the alternatives are compared in
Figure 15.12: Alternative 1 with 120 C-17s, and Alternative 2 with no C-17s but
with 263 C-141 SLEP/C-Ys instead. Figure 15.12 also illustrates the effect of dis-
counting through two examples: no discounting and the baseline OMB-directed
4.5 percent discounting. Note that the large effect of O&S costs that accrue over
the 25-year life-cycle cost period for all alternatives. Discounting emphasizes
near-term costs, which has a bigger effect (reduction) on long-term O&S cost
estimates than on near-term acquisition cost estimates.

15.6 Reported findings


The C-17 Cost–Benefit Analysis (i.e. COEA) reported the following findings,
briefed to the DAB in November 1993. The findings incorporated several items
reported in earlier briefings and added the sensitivity analysis excursions discussed
in the previous section.
390 W. L. Greer

New, collapsible cab top trucks or extended 747 side door


70~r=========~==========~~--------~~----------,---------~
C-17 :
Q COMMERCIAL DERIVATIVES ___________
TPFDD L ________________________ _

65 ~ C-141 SLEP
0 C-58+
Outsize cargo delivered in 30 days (kilotons)

Baseline Stretched profile


60

Alternative 1
cost
55

__ ....
_.,.,. ..... ""-
50
- -
................ ........
Equal cost

.... -
........

- -
effectiveness
45 __1 ........ ........

~ . . @-® ® :5
18 10, - " ' I

40 ---------------------~;
MRS 21 ~~::_=---- ~ --~ ---------------------- -
3
_........ _........ [] : Fleet Composition (TAI)
Alternatives
~~- 15 I
..-- I A/C 1 3 5 10 15 18 21
35 --~
:
I
~ ..... .,.,.,...... C-17 120 – 26 94 – 70 47

-- II C-5B+ – 94 82 – 49 – –
I 747 – – – 18 42 34 49
30L-------------~------------~------------~~------------~------------~
70
1

45 50 55 60 65
25-year cumulative cost ($B)

Figure 15.11 Effect of FYDP stretch-out on alternatives in MRC.

• The cost and performance of the planned C-17 fleet alternative (#1) make it
the preferred military airlifter. It is more resistant to airfield constraints (i.e.
MOG) than the new C-5 and possesses a higher use rate. It is far superior in
both cost and effectiveness to the C-141 SLEP.
• The next most attractive alternatives after the 120 C-17s would be mixed fleets
of C-17s and modified commercial aircraft with specially reinforced floors
and some concession for the height of new FMTV army trucks, such as wider
side doors.
• If new army trucks cannot be loaded on commercial derivatives, then the next
most attractive alternatives to the 120 C-17 program would be mixes of C-17s
and new C-5s.

A mixed fleet of commercial derivatives and some number of C-17s (less than
120) became a serious contender to Alternative 1, the original/baseline 120 C-17
fleet, reducing the total cost of the program. The study did caution that overall
effectiveness could be compromised. The introduction of 747s would provide
fewer aircraft for certain unique military operations such as airdrops, deliveries
to remote and inadequate airfields, low-altitude parachute extractions, and rapid
off-loading of cargo while the airlifter is still moving down the runway (combat
Acquisition of the C-17 aircraft 391

C-130 O&S
C-Y O&S
100 C-141 O&S
C-17 O&S
25-Year Cumulative Cost (FY 93 $B)

C-Y Acquisition
80
C-141 SLEP Acquisition
C-17 Acquisition

60

40

20

0
Alt 1 Alt 2 Alt 1 Alt 2
No Discounting Baseline 4.5% Discounting

Figure 15.12 Effect of discount rates on two alternatives.

off-loading) with engines running. It also provided a fleet less well able to adapt
to limited MOG conditions. Nonetheless, its lower cost recommended it for more
detailed consideration.

15.7 The rest of the story: ex-post decisions subsequent to


the analysis

15.7.1 Decisions 1994–1995


The decision ultimately reached by the DAB in January 1994 was to place a cap
on C-17 production at the 40 already under development and to establish a non-
developmental airlift aircraft (NDAA) program to investigate the feasibility of
acquiring and operating militarized commercial derivatives in lieu of C-17s. The
analyses presented in the IDA study had a major impact on this decision. After
the results of the NDAA study became available, a new decision would be made.
The DoD also reached a financial settlement in 1994 with the C-17 manufacturer
McDonnell Douglas over past claims, setting the stage for a clean slate and new
decisions.
From February through May of 1994, the IDA analysts responsible for the study
briefed congressional staffers and the Government Accountability Office (GAO).
392 W. L. Greer
In its report that same year (US GAO 1994), the GAO favored the acquisition
of 40 C-17s and 64 747s over the 120 C-17 case. In May 1994, the DoD formally
released the IDA report to Congress in response to the initial congressional request
for such a study.
In 1995, many events converged that influenced the ultimate decision. In
January of that year, the C-17 reached its initial operational capability (IOC).
It subsequently underwent formal reliability, maintainability, and availabil-
ity (RM&A) testing in July, achieving the performance needed to support a
15.2 hours/year use rate. During the same year, the AMC initiated, per OSD guid-
ance, a study of NDAA options called the Strategic Airlift Force Mix Analysis
(SAFMA) USAF Air Mobility Command (1995). The aim of that study was to
determine the minimum number of C-17s acceptable and to provide industry with
an opportunity to make proposals for NDAA competitors to the C-17. The RFPs
was released in March. Meantime, also in March, the Joint Staff and OSD released
a new mobility requirements study to replace the older MRS used in the IDA
C-17 Cost-Benefit Analysis (COEA) report. The new study was titled the Mobil-
ity Requirements Study Bottom Up Review Update (MRS BURU) OSD (1995)
and established new cargo and troop lift requirements for the MRCs. This study
provided the new delivery requirements for the SAFMA. SAFMA used essentially
the same MASS model (updated) that IDA used two years earlier.
AMC conducted the SAFMA using fleet mixes of C-17s and Boeing 747s, the
only contractor-proposed alternative. Data from the C-17 RM&A as well as new
lower price estimates from McDonnell Douglas for the C-17 were available. IDA
played an auxiliary role in this process, providing an independent validation of
the models and approaches taken, although the actual analyses were conducted by
AMC as part of a proprietary source selection process.
In November 1995, the results of the RM&A and the SAFMA were presented
to the DAB. At this meeting, it was decided to proceed with the 120 C-17 program
and not acquire any militarized 747s.
Thus, in a lengthy set of decisions, OSD arrived back at the position it started
from two years before: buy 120 C-17s. Something very significant, however, had
happened in the interim: the introduction of competition had compelled the C-17
manufacturer to reduce costs and improve performance in order to prevail. By
revealing feasible alternatives, the IDA cost-effectiveness study contributed to the
creation of what economists call a “contestable market” (Baumol et al. 1983,
492–496). The threat of the entry of new competitors was sufficient to reveal
efficiencies that reduced aircraft costs. The new restructured C-17 program was
several billion dollars lower in total cost than the older one, a significant savings
for the government and its taxpayers.

15.7.2 Decisions post-1995


The C-17 program has prospered since the days of these early analyses. It has
grown to nearly twice the size of the program discussed here and is now larger
than the size of the original fleet envisioned in 1983 of 210 aircraft. Table 15.6
Acquisition of the C-17 aircraft 393
Table 15.6 Summary of C-17 acquisition decisions through FY 2010

Year C-17 acquisition decisions Results

2002 60 additional C-17s bought Total 180


2005 OSD Mobility Capabilities Study (OSD 2005) shows no –
additional C-17s beyond 180 are needed.
2007 10 additional C-17s bought Total 190
2008 15 additional C-17s bought Total 205
2009 IDA study on size and mix of airlift force (Greer et al. –
2009) shows no additional C-17s beyond 205 are needed.
2009 SecDef declares that no more C-17s are needed. –
2009 8 additional C-17s bought Total 213
2010 OSD/USTRANSCOM Mobility Capabilities & –
Requirements Study (OSD 2010) shows no additional
C-17s beyond 213 are needed
2010 10 additional C-17s to be bought (production through 2013) Total 223
2010 Quadrennial Defense Review (OSD QDR 2010) indicates –
that no more C-17s are needed
2010 Subcommittee of Senate Homeland Security & –
Governmental Affairs hearings (US Senate Hearing
2010) on C-17 requirements, also affirmed that no
additional C-17s are needed
2011–14 No additional C-17s in budget –

reports subsequent decisions made to augment the C-17 fleet from the original
120 aircraft to 223.
The impetus for adding C-17s beyond the 120 initially approved came after
the attacks on September 11, 2001, and the decision to strike back at al-Qaeda
and the Taliban in Afghanistan. Airfield limitations in Afghanistan and Iraq sup-
ported the case for additional C-17s that could more easily operate in austere
environments.
Additional studies have continued to the present time, all addressing the issue
of how many C-17s (and other airlift aircraft) are needed. Table 15.6 summarizes
the conclusions reached in those studies. The programmed number in 2011 is 223
C-17s.

15.8 Lessons learned


Many lessons can be learned from this study. A selection is presented below as the
conclusion of this chapter.

15.8.1 The importance of an open process to combat special interests


The results were certain to be controversial to some interested parties, so it was
necessary to maintain credibility and impartiality. The C-17 COEA benefited from
a continuously open process, with frequent feedback and critical commentary
as the study developed. The analysts were able to maintain their independence
and objectivity throughout but made their results available to others as the study
394 W. L. Greer
evolved. In this way, all parties had an opportunity to see how the results were
evolving and to recommend alternative approaches, inputs, and/or sensitivities.

15.8.2 The value of analysis in promoting competition


The emergence of viable competitors to the C-17 and the willingness of decision-
makers to honestly consider these alternatives created positive incentives for the
C-17 manufacturers to meet and exceed their cost and performance objectives.
Even though in the end DoD did not select the 747, its presence, looming as
an alternative on the horizon, appears to have had a salutary effect on the C-17
program and the airlift capabilities of DoD. While this cannot be proven unequiv-
ocally, most observers of the decision process feel that the analysis encouraged
implicit competition that may have saved the government billions of dollars.

15.8.3 The power of sensitivity analysis and other excursions


Every piece of data and every assumption in the analysis needs to be scrutinized to
understand its effect on the conclusions. The problem must be clearly structured
up front and reviewed as the analysis proceeds from an independent and broader
systems perspective to avoid overlooking important factors. Future army trucks
having trouble fitting in 747s is a good example. Total systems’ thinking is key. It
is easy for analysts to become focused on the particular program—buy C-17s—
while the effect of other programs that could have a profound effect—new rigid
top trucks—are overlooked.
Another example is MOG. The infrastructure investments needed to improve
airfield airlift capability could overwhelm the cost of the aircraft themselves. In
fact an important tradeoff exists between improving airfields and lower aircraft
costs or improving aircraft so they can operate in more constrained airfields. These
insights and the importance of changes in assumptions and input values can only
be determined by a robust set of alternatives and sensitivity analysis excursions.
Not all of these can be foreseen at the start of a study, but ensuring time and
resources are included in the overall study plan to conduct sensitivity analyses can
have powerful payoffs.

Appendix 15.1: Additional details on loading and transportation


models and acquisition costs
The main models used for estimating how much cargo and how many passengers
were delivered were ALM, MASS/AFM, and ACAS. These models are summa-
rized briefly to give a sense of the assumptions used in the analysis. As noted
earlier, these models have subsequently evolved to more advanced versions, but it
is important to know how the transportation problem was approached in 1993.

Airlift loading model


The ALM was used to determine the average cargo payload carried by each aircraft
in the fleet. The MASS/AFM then used this result to simulate the transport. ALM
Acquisition of the C-17 aircraft 395
used cargo from the entire TPFDD for these estimates. The cargo was differenti-
ated according to 26 different categories called “commodities.” Each commodity
could contain vehicles, pallets, and passengers. Examples include Air Force air-
craft, Air Force support, Marine prepositioning, armored infantry, mechanized
infantry, Combat Service Support (CSS), engineer, CSS medical, and ammunition.
Pallets were of a common size and weight, but weight and dimensions differen-
tiated the vehicles from one another. The sequence in which the aircraft were to
be loaded had to be specified, as well as aircraft cargo storage dimensions, cargo
door opening sizes, number of passenger seats, and maximum payload allowable
by weight. The ALM then used user-selected rules for loading the aircraft (e.g.
vehicles were sorted according to width, with the widest loaded first and pallets
last for a given commodity code). Loading stopped when the weight limit was
reached or, more commonly, when cargo fully filled the cargo space in the aircraft.
Most airlifters reached volume constraints before they reached weight constraints,
as they do in real life. From these simulations, the analysts calculated the aver-
age payload carried by each aircraft type, detailed by commodity type and by the
cargo classes of out/over/bulk/PAX.
While ALM does an excellent job of estimating optimum average cargo loads,
these loads are rarely achieved in actual operations. Differences in actual units—in
the way cargo arrives at a port of debarkation and in the proficiency of the cargo
loaders—often results in actual loads falling slightly below those predicted by the
model. These inefficiencies can sometimes be noted during real exercises and oper-
ations. By establishing community agreement at the outset on the use of the same
methodology for all aircraft, the study treated each alternative in a consistent manner.

MASS/AFM
The MASS/AFM was a large Monte Carlo simulation31 that scheduled and exe-
cuted each airlift mission by tail number. It incorporated a great deal of realism,
such as specific routes used, airbases with MOG limitations, times for loading and
off-loading, and crew duty days. It used the ALM loading results along with the
nature of the TPFDD cargo and PAX at individual airbases to load the cargo. Only
that part of the TPFDD available at a specific base on a specific day (plus any unde-
livered leftover cargo from previous days) was available for loading. User-selected
priorities for loading out/over/bulk/PAX classes were considered for each airlifter.
The AMC analysts ran the simulations at Scott AFB with inputs from IDA. The
primary outputs for the study were tons delivered per day by commodity type and
class (out/over/bulk/PAX).
As with ALM, MASS/AFM may produce results that would be viewed as opti-
mistic in real operations. For example, the inefficiencies that often exist with high
levels of congestion at en-route bases are not included (although MOG does limit
the number of aircraft being serviced at the base). Again, the decision-makers
and study team need to make reasonable assumptions and focus on the higher
objectives of the study. If an assumption appears to have the potential for making
a large impact on the results, sensitivity analysis may be required.
396 W. L. Greer
Airlift cycle analysis spreadsheet
The ACAS model was a simplified approximation to the analyses produced by
MASS/AFM and was used to conduct the large number of excursions needed in
the study. It used aggregate data and aggregate MOG values, calibrated against
MASS/AFM to give confidence, for these numerous excursions.

Acquisition costs
The study used the Air Force C-17 procurement schedule, with a maximum acqui-
sition of 16 per year. It also used multi-year procurement arrangements to reduce
uncertainty and cost. From the historical cost data for the C-17, IDA derived a cost-
learning curve to predict future costs for C-17s as more manufacturing experience
is attained. This stage was a delicate one, requiring the cooperation of McDonnell
Douglas to supply competition-sensitive information with the understanding that
only aggregate levels of detail would appear in the final report. Publishing a sec-
ond proprietary document for use by the government, while using the aggregate
results in the COEA, attained the desired balance between a need for discussion
of the methodology and a display of the aggregate results and the need to pro-
tect the legitimate interests of the source. Appropriate adjustments to historical
trends were added as needed, such as the cost for extra weight to strengthen the
wings and higher costs associated with more realistic engine estimates for all lots
after the first ones. During the study, the C-17 Program Office disagreed strongly
with the IDA estimates. The basic disagreement arose from different assumptions
about the learning curves. IDA used data for the entire C-17 program, adjusting
the first few low-cost lots for their later prices while the Program Office used
later prices only. IDA estimated the full C-17 program acquisition cost at approx-
imately US$23 billion, about US$3 billion over the Program Office estimate. In
the end, the higher-cost IDA results were used in the report, with a comparison to
the Program Office estimates included as an excursion.
At the time the study was conducted, three test C-17s (not counted as part of
the 120) had been produced, an additional six production aircraft were undergo-
ing operational test and evaluation (OT&E), and a total of 17 production aircraft
had either been delivered or were in various stages of assembly at the Long Beach
McDonnell Douglas facility. Moreover, aircraft through production number 26
were under contract with long-lead funding. Failing to buy 26 production C-17s
would have incurred a cost penalty by the US government for a breach of con-
tract. Thus alternatives with fewer than 26 C-17s were levied an additional cost
associated with breaking the contract (US$1.5 billion).
Lockheed Martin Aeronautical Systems provided cost estimates for restarting
the C-5 line as well as for refurbishing the C-141s in a C-141 SLEP. Again, only
aggregate data appeared in the final report, with company-sensitive data relegated
to the proprietary government-use document. IDA used these data and other his-
torical Lockheed data and made adjustments as required. Examples of adjustments
include adding hush kits to dampen the sound of the engines and extra material
to the fuselage. A recently completed SAB had recommended that the weep hole
Acquisition of the C-17 aircraft 397
cracks in then-current C-141s could be repaired and their life extended to 45,000
flying hours. When the C-141 reached its 45,000-flying-hour life limit, the SAB
determined it had to be replaced, and the cost of a new C-Y replacement was
imposed to include a US$3.5 billion development cost.
For the C-5 restart, a cost to restart the line was imposed on any alternatives
using C-5B+ aircraft. IDA estimated that US$750 million would be needed to
restart the line in Marietta, Georgia. New C-5 engines were used to reduce noise
levels below those mandated by the FAA.
Boeing supplied basic cost data for the 747 and 767 commercial deriva-
tive aircraft. Production rates from six to 12 per year for military acquisition
were used, based on Boeing production capabilities and competing markets for
new 747 or 767 aircraft. An additional 20 percent of the acquisition price was
added to account for reinforced floors and rollers and widened side doors, for
an estimate of US$155 million for the militarized 747 and US$88 million for
the militarized 767. Since these lines are open, no developmental costs were
imposed.

Notes
1 McDonnell Douglas was subsequently merged with the Boeing Company in 1997.
2 Bulk cargo is carried on standard USAF 463L pallets and is transportable by all air-
lifters. A pallet holds bulk cargo measuring no more than 104 × 84 × 96 . Oversize
cargo is larger than the pallet dimensions and consists typically of wheeled vehicles.
Oversize cargo is larger than a single pallet but less than 1,090 × 117 × 105 . All
oversize cargo can fit on a C-141 (or larger aircraft). Outsize cargo is the largest of
these three, fitting only on C-5s and C-17s.
3 The use of a detailed model (i.e. MASS) along with a more aggregate one (i.e. ACAS)
is a common approach that balances study resources and time with precision in results.
Calibration of ACAS to specific MASS results allowed ACAS to be used in rapid sen-
sitivity excursion analyses. Comparisons of ACAS runs with MASS runs for these
same sensitivity excursions showed no loss in precision for using the significantly faster
ACAS.
4 Costs also addressed contract claims that McDonnell Douglas had made as well as
penalty charges associated with terminating the C-17 line short of then-contracted
production levels.
5 In 2010, a total of 111 C-5s were in the USAF inventory. Of these, 49 were C-5Bs.
6 Block speed is the distance flown by an aircraft divided by the time spent once blocks
are removed from the aircraft wheels and it taxies down the runway until it lands and
comes to a complete stop with blocks again in place.
7 Bulk cargo would be no problem in the commercial non-militarized version, since that is
what CRAF carries. The nature of the militarization would be to strengthen the floors to
handle higher pressure points (such as under the wheels of fully loaded five-ton trucks)
and possibly to modify the doors to permit wider cargo or equipment to be loaded from
the side.
8 A considerable number of additional alternatives were added later in the study, not all of
which had a capacity of 52 MTM/D, but these were introduced after the more powerful
comparison displays (to be shown later) were developed. At that point, equal MTM/D
became an interesting but marginally useful criterion. Actual performance in realistic
scenarios became the more important gauge. To start the analyses, IDA first focused on
these 26 equal-MTM/D alternatives.
398 W. L. Greer
9 The TAI entries indicate how many new aircraft are to be procured: The PAA values
are the numbers available to deliver cargo and troops in the scenarios studied and the
numbers used in O&S costing estimates.
10 Today the acronym MRC has been replaced by MCO (Major Combat Operation) and
CC (Conventional Campaign), but the meaning is equivalent. Here, we will use the
MRC notation used in the original 1993 study.
11 MRC-East refers to a theater in Southwest Asia, supported mainly by flights out of
East Coast US bases. Similarly, MRC-West refers to a theater in the Western Pacific
with deployments from the western US airbases. It was assumed that CRAF Stage II
would be authorized at the onset of MRC-East, with Stage III activated when MRC-West
deployments began.
12 Today, LRC has been replaced by IW (Irregular Warfare) or a vignette from the SSSP,
the Steady State Security Posture collection of lower intensity conflicts.
13 This assumption is further supported by the heavy demand for tankers to support fighter
deployments to the MRC theaters as well as the potential for supporting combat aircraft
employment during and following deployment.
14 The use rate is the average number of hours per day that an aircraft flies. Its estimate
depends on the amount of time it spends on the ground which, in turn, involves the
mission capable rate, which is a measure of reliability. It is used in the models to place
an upper limit on the amount of time an aircraft flies.
15 The MOG-value of each airbase is specified by aircraft type and is determined by the
material handling equipment (MHE) available, fuel availability, and the number of ser-
vice positions allotted to airlift (in-theater bases are often mostly occupied by fighter
aircraft). Large aircraft may require more than one service position.
16 It suffices to show the difference in the number of spaces available for small airlifters
as contrasted with larger ones. Detailed databases at the appropriate classification are
maintained by AMC.
17 MOG was usually the most constraining at en-route bases. En-route MOG constraints
may cause longer, less efficient routes to be used.
18 These models have subsequently evolved to more advanced versions, but it is important
to know how the transportation problem was approached in 1993.
19 MASS/AFM has now been replaced at AMC by the Air Mobility Operations Model
(AMOS).
20 The Office of Management and Budget (OMB) changes the discount factor every
year (US OMB 1992). For cost-effectiveness analyses of Federal programs, OMB
mandates the use of discounted dollars.
21 All costs prior to FY 1994 were considered “sunk” and ignored. Only future expendi-
tures were considered.
22 The SSC provided a review of the study prior to presentation to the higher-level DAB.
All offices represented in the DAB had representatives in the SSC.
23 As noted earlier in Figure 15.2, virtually all the required outsize cargo must be delivered
within a 30-day time period. This category of cargo differentiates best among the alterna-
tives because of its criticality to the early stages of battle. Excursions (not shown here) at
20 days showed no new insights, so 30-day deliveries were displayed in the study.
24 As discussed elsewhere in this book, using cost-effectiveness ratios to evaluate alterna-
tives is generally not appropriate and can in fact be very misleading without a budget
constraint or fixed level of effectiveness.
25 These are examples of the modified budget approach and the modified effectiveness
approach for cost-effectiveness comparisons in which neither cost nor effectiveness are
initially constrained to be equal for all alternatives. These are described by Francois
Melese (Melese 2010, 40–45).
26 The reader may wonder how oversize cargo such as trucks can influence the rate of
delivery of outsize cargo. The interaction is complex, involving a competition for space
Acquisition of the C-17 aircraft 399
aboard aircraft that does carry outsize (i.e. C-17s and C-5s) and between outsize cargo
and oversize cargo. If no oversize trucks can be carried by CRAF, those same trucks
begin to crowd out any outsize cargo that C-17s and C-5s would have otherwise carried,
reducing their outsize delivery capability.
27 This is an example of the modified effectiveness (“level the playing field”) approach
described by F. Melese in Chapter 4 of this book.
28 This is an example of the fixed budget approach described by Francois Melese (2010,
33–35).
29 The use of cost-effectiveness ratios can be misleading unless the analyses are con-
structed under the same budget or same effectiveness frameworks (Melese 2010,
17–20). In this case, the use of a common effectiveness framework, i.e. same MTM/D
confers meaning.
30 Note that the economic evaluation of alternatives (EEoA) approaches introduced in
Chapter 4 of this book make a clear distinction between the “life-cycle costs” or “price”
of an alternative, its operational effectiveness (schedule and performance), and the
resources (funding or budget) likely to be available for the overall program.
31 MASS/AFM has now been replaced at AMC by the Air Mobility Operations Model
(AMOS).

References
Baumol, W. J., J. C. Panzar, and R. D. Willig. 1983. “Contestable Markets: An Uprising in
the Theory of Industry Structure: Reply.” American Economic Review 73(3): 492–496.
Bexfield, J. M., T. A. Allen, and W. L. Greer. September 2001. C-17 COEA Case Study
(IDA Document D-2688). Alexandria, VA: Institute for Defense Analyses.
Greer, W. L., J. N. Bexfield, J. R. Nelson, D. A. Arthur, P. B. Buck, W. C. Devers, J. L.
Freeh, B. R. Harmon, D. Y. Lo, H. J. Manetti, A. Salemo, J. A. Schwartz, J. W. Stahl,
V. Suchorebrow, and D. M. Utech. December 1993. Cost and Operational Effectiveness
Analysis of the C-17 Program (IDA Report R-390). Alexandria, VA: Institute for Defense
Analyses.
Greer, W. L. et al. February 2009. (U) Study on Size and Mix of Airlift Force (IDA Paper
P-4425) (Classified study). Alexandria, VA: Institute for Defense Analyses. (Subject is
SECET//NOFORN.)
Joint Staff J4 Logistics. 1992. (U) Mobility Requirements Study (Classified study). Wash-
ington, DC: Joint Staff J4.
Melese, F. January 4, 2010. The Economic Evaluation of Alternatives (EEoA) (NPS-
GSBPP-10-002). Monterey, CA: Naval Postgraduate School.
Office of the Secretary of Defense (OSD). 1995. (U) Mobility Requirements Study Bottom
Up Review Update (Classified study). Washington, DC: OSD.
Office of the Secretary of Defense (OSD). 2010. Quadrennial Defense Review
Report. Washington, DC: OSD. Available at www.defense.gov/qdr/qdr%20as%
20of%2029jan10%201600.pdf [last accessed July 23, 2014].
——. 2005. (U) Mobility Capabilities Study (Classified study). Washington, DC: OSD.
——. 2010. (U) Mobility Capabilities and Requirements Study – 2016 (Classified study).
Washington, DC: OSD.
Secretary of the Air Force. February 4, 1994. US Air Force Cost and Planning Factors (AF
Instruction 65-503). Washington, DC: Secretary of the Air Force.
——. December 2003. Air Mobility Planning Factors (Air Force Pamphlet 10-1403). Scott
Air Force Base, IL: USAF Air Mobility Command. (December 2003 is the most recent
version of AFPAM 10-1403).
400 W. L. Greer
US Congressional Budget Office (CBO). October 2007. Issues Regarding the Current and
Future Use of the Civil Reserve Air Fleet. Washington, DC: US CBO.
US Government Accountability Office (GAO). July 11, 1994. Airlift Requirements:
Commercial Freighters Can Help Meet Requirements at Greatly Reduced Costs
(GAO/NSIAD 94-209). Washington, DC: US GAO. Available at www.gpo.gov/fdsys/
pkg/GAOREPORTS-NSIAD-94-209/content-detail.html [last accessed July 22, 2014].
US Office of Management and Budget (OMB). October 29, 1992. Guidelines and
Discount Rates for Benefit-Cost Analysis of Federal Programs (OMB Circular No.
A-94) (Revised annually). Washington, DC: OMB. Available at www.whitehouse.gov/
omb/circulars_default [last accessed July 22, 2014].
USAF Air Mobility Command (USAF). 1995. (U) Strategic Airlift Force Mix Analysis
(Classified study). Scott Air Force Base, IL: USAF Air Mobility Command.
US Senate Homeland Security and Governmental Affairs; Subcommittee on Federal
Financial Management, Government Information, Federal Services, and International
Security. July 13, 2010. Hearing on the Cost-Effectiveness of Procuring Weapon Sys-
tems in Excess of Requirements. Statement by Dr. William L. Greer, Assistant Director,
System Evaluation Division, Institute for Defense Analyses, Alexandria, VA.
16 Cost-effectiveness analysis of
autonomous aerial platforms
and communication payloads
Randall E. Everly, David C. Limmer,
and Cameron A. MacKenzie

16.1 Introduction
This chapter demonstrates the use of multi-criteria decision-making techniques
(discussed in Chapter 8) to analyze the cost-effectiveness of autonomous aerial
platforms and communication payloads for communication missions in the mili-
tary. We compare the cost-effectiveness of 17 aerial platforms and 9 communica-
tion payloads across three mission scenarios.
Unmanned aerial vehicles (UAVs) can help to supply the information
technology connectivity that is increasingly being demanded by technology
advancements on the battlefield. Autonomous vehicles are well suited for the
communication mission, which is often mundane and tedious. The advent of
lightweight construction materials, high energy-density lithium battery technol-
ogy, and more efficient microprocessors increase UAV capability (Department of
Defense 2013). The larger payload capacities and longer endurance of modern
UAVs combined with communication payloads that are more capable and effi-
cient and weigh less make them more suitable to the communications relay role
than their predecessors. UAVs are also more flexible than more permanent infras-
tructure such as relay and cellular towers, and they can be quickly repositioned to
support warfighters on the move.
Scrutiny on discretionary spending in public budgets, including defense
budgets, will continue to increase in many countries. Future acquisitions will
need to be executed thoughtfully with a clear consideration of the value and cost.
Although other studies (see Ferguson and Harbold 2001; Collier and Kacala 2008)
have analyzed the costs and benefits of UAVs for different missions, no study
has undertaken an extensive cost-effectiveness of the most modern UAVs and
other aerial platforms that can support communication requirements for military
operations in austere environments. This chapter seeks to remedy that deficiency
by performing a cost-effectiveness analysis of aerial platforms and communica-
tion payloads for use as a communication relay in support of distributed military
operations.
We conduct a multi-objective analysis of alternatives to compare the cost-
effectiveness of selected aerial platforms and communication payloads across
three scenarios. This chapter considers 13 different UAVs, 4 alternative aerial
platforms, and 9 communication payloads suitable for the communication relay
402 R. E. Everly, D. C. Limmer, C. A. MacKenzie
mission. UAVs range in size from the hand-launched Raven to the Triton with
its 130-foot wingspan, and communication payloads vary in weight from less
than a pound to over 250 pounds. We follow the approach detailed in Chapter 8
to analyze the cost-effectiveness of these alternatives. An objectives hierarchy
lists the desirable attributes for aerial platforms and communication systems, and
value measures and tradeoff weights lead to a numerical measure of effective-
ness (MoE). The annualized life-cycle costs (LCC) are estimated for the aerial
platforms, and acquisition costs are calculated for the communication payloads.
Selecting the most cost-effective alternative involves consideration of the costs
and MoEs. After selecting the most cost-effective aerial platforms and commu-
nication payloads for a specified scenario, we discuss whether the selected aerial
platforms are compatible with the communication payloads.
Section 16.2 reviews a few applications of cost-effectiveness analysis in defense
and discusses previous studies of UAV effectiveness. We briefly outline in
Section 16.3 the different UAVs, alternative aerial platforms, and communication
payloads. Section 16.4 discusses the cost-effectiveness model. An objectives hier-
archy is presented for the aerial platform and for the communication payload in
Section 16.5. After detailing the methodology for estimating costs for the aerial
platforms in Section 16.6, we analyze the cost-effectiveness of all the alterna-
tives for three different scenarios in Section 16.7, and offer concluding thoughts
in Section 16.8. We rely on subject matter experts to collect and analyze data for
several alternatives, and develop value measures and tradeoff weights based on
our own research, expertise, and experiences. Consequently, the results should
not be viewed as definitive, but the analysis does provide insights into select-
ing the best aerial platform for different communication relay missions. Defense
decision-makers could incorporate their own preferences within this framework to
determine the most cost-effective aerial platform and communication payload for
a given scenario.

16.2 Literature review


Cost-effectiveness analysis seeks to evaluate costs versus the benefits of an alter-
native when the benefits cannot easily be measured in monetary units. The MoE
can be a single numerical measure of performance if only one objective deter-
mines the effectiveness for a decision problem. Frequently, decision-makers have
multiple objectives, and combining several objectives into a single MoE requires a
systematic process or mathematical equation. Value-focused thinking encourages
decision-makers to articulate their values and identify and structure their objec-
tives (Keeney 1996). Multi-objective or multi-criteria decision-making provides
a framework for developing value measures over those objectives and tradeoff
weights among objectives in order to combine several objectives into a single MoE
for a given alternative (Keeney and Raiffa 1976; Kirkwood 1997).
Some multiple objective analyses include minimizing cost as an objective
and developing a multi-attribute value function over all the objectives, includ-
ing cost. For example, a study to help the US Army determine which bases to
Cost-effectiveness of platforms and payloads 403
close develops a value function for 40 attributes, including a few cost-related
attributes (Ewing et al. 2006). The decision-maker selects the set of bases to
remain open that maximizes the total value subject to Army capability require-
ments. A cost-effectiveness analysis for the Army (Buede and Bresnick 1992)
constructs MoEs for several air defense alternatives and depicts the MoE and the
cost of each alternative on a two-dimensional chart. Similarly, our analysis sepa-
rates the cost of each alternative from its MoE and compares the cost-effectiveness
of the alternatives on a two-dimensional chart.
Several studies have analyzed the use of UAVs for different military missions,
but many are financial in nature. The LCC of the the Triton UAV determines the
financial implications of the US Navy purchasing this system for its Broad Area
Maritime Surveillance (Lawler 2010). Yilmaz (2013) compares the Reaper and the
Guardian UAVs for US border security and the Israeli-manufactured Heron UAV
for patrolling Turkey’s border but focuses primarily on the costs of the systems as
opposed to their effectiveness. Some business case analyses (Thiow Yong Dennis
2007; Fry and Tutaj 2010) determine the number of UAVs of a specified type
that is necessary to achieve a mission, and the decision-maker should choose the
UAV type that can accomplish the mission at the lowest total system cost. These
approaches are similar to the second way to structure an economic evaluation of
alternatives as described in Chapter 4.
A multiple objective analysis (Ferguson and Harbold 2001) compares the
Global Hawk UAV, the manned research vehicle Proteus, and three solar powered
aircraft that could be used for communications relay, intelligence, surveillance,
and reconnaissance. A score is assessed for each platform based on the following
objectives: instantaneous access area, endurance, survivability, feasibility, flexi-
bility, responsiveness, and acquisition cost. A more recent analysis (Collier and
Kacala 2008) measures the effectiveness of airships, UAVs, and tactical satellites
based on a multi-objective framework across a range of operating environments
and mission sets. The decision-maker should select a mix of different aircraft to
maximize the total fleet effectiveness subject to a budget constraint (see Chapters 4
and 10).
In this chapter, we apply the principles of multi-criteria decision-making to
evaluate the effectiveness of several aerial platforms, including UAVs and air-
ships. Like Ferguson and Harbold (2001), we focus on a communication relay
mission, but we use LCC for the aerial platforms as opposed to acquisition costs
because LCC more accurately represents total costs (see Chapter 5). We exam-
ine more aerial platforms, which have multiplied since the Ferguson and Harbold
(2001) study. A unique contribution of this chapter is the examination of the cost-
effectiveness of communication payloads that can be integrated with the aerial
platform. Collier and Kacala (2008) study multiple scenarios—a technique that
we adopt in this chapter. Unlike Lawler (2010), Thiow Yong Dennis (2007),
and Fry and Tutaj (2010), who develop models so that the benefits or effective-
ness is equal across the UAV alternatives, our analysis determines unique MoEs
for different aerial platforms and communication payloads and different mission
scenarios.
404 R. E. Everly, D. C. Limmer, C. A. MacKenzie
16.3 Alternatives
Over 50 countries’ militaries are developing or operating about a thousand differ-
ent UAV platforms (Parsons 2013). The US Department of Defense currently has
approximately 11,000 UAVs in its inventory, consisting of almost 150 different
platforms (American Institute of Aeronautics and Astronautics 2013). Israel, who
first deployed UAVs for military operations, manufactures approximately 45 dif-
ferent UAV platforms (Dobbing and Cole 2014). We focus on 13 UAVs and 4 other
aerial platforms that can be used for communication missions. We separately con-
sider several communication payloads that can be fitted to these aerial platforms.
Much of the information and many of the specifications for the aerial platforms
come from Nicholas and Rossi (2011).
UAVs can be divided into five categories: small, medium, large, high alti-
tude long endurance (HALE), and vertical takeoff and landing. Small UAVs
are hand-launched and include the Wasp III and RQ-11 Raven. Both UAVs are
hand-launched with ranges of 3–5 miles.
We examine two medium-size UAVs, the RQ-Shadow 2000 and the RQ-21
Blackjack. The Shadow, with a wingspan of 20 ft, is launched by catapult and
usually lands using a runway with a hook at the bottom of the aircraft to snag a
wire across the runway. The Shadow is primarily used for reconnaissance, surveil-
lance, target acquisition, and battle damage assessment. The RQ-21 Blackjack,
with a 16 ft wingspan, is a catapult-launched, vertically arrested UAV with multi-
mission capability and six payload bays with power and Ethernet that can be fitted
with cameras, communication capabilities, or other custom payloads.
Large UAVs have wingspans exceeding 25 ft and include the MQ-1 Predator,
the MQ-1C Gray Eagle, the MQ-9 Reaper, and the X-47B Unmanned Combat
Air System Air Carrier Demonstration (UCAS-D). The Predator was originally
designed for intelligence, surveillance, and reconnaissance (ISR) but has been
enhanced to be capable of taking on many roles including targeting, forward air
control, laser designation, weapons delivery, and bomb damage assessment (Gen-
eral Atomics Aeronautical 2013). The Gray Eagle is a larger version of the
Predator used by the US Army (General Atomics Aeronautical 2012a). The
Reaper, also known as Predator B, is a hunter-killer UAV designed to eliminate
time-sensitive targets via onboard 500-pound bombs and Hellfire missiles (Gen-
eral Atomics Aeronautical 2012b). Still in the demonstration and testing phase,
the UCAS-D is designed to take off and land on aircraft carriers and to perform
persistent surveillance with strike capability (Naval Air Systems Command 2014).
HALE UAVs are typically large UAVs with a maximum ceiling of approxi-
mately 60,000 feet and an endurance of 24 hours or more. The RQ-4 Global Hawk
is operated by the US Air Force (USAF) for long-range ISR. The MQ-4C Triton
is the U.S Navy’s version of the Global Hawk and has deicing capability.
Vertical takeoff and landing UAVs (which can be small, medium or large) have
rotary blades. These UAVs include the MQ-8 Fire Scout, which is used on ships
to provide situation awareness, targeting support, and communications relay, and
the YMQ-18A Hummingbird used for ISR and carrying cargo.
Cost-effectiveness of platforms and payloads 405
Non-UAV aerial platforms considered as part of this study include rapidly
erected towers and tethered balloons. Rapidly erected towers present an interest-
ing alternative to UAVs because data link and power can be supplied via cables
from the ground, which can offer near limitless endurance. Two possible towers
are the Rapid Aerostat Initial Deployment (RAID) Tower System and Cerberus
Tower (Army Deploys 2008). Tethered balloons, such as the TIF-25K and PTDS-
74K aerostats, are capable of carrying power and data through the tether and can
go significantly higher than towers (US Army Aerostat-Based PTDS 2010; Raven
Aerostar 2014).
In addition to identifying the most cost-effective aerial platform, we also evalu-
ate the cost and effectiveness of 9 communication payloads that can be used with
the aerial platforms. Wave Relay is a mobile ad hoc network (MANET) solution
that continuously adapts to fluctuations in terrain and the environment to maximize
connectivity and communication performance (Persistent Systems 2012, 2014).
We consider two Wave Relay configurations: the Wave Relay Gen4 Board and the
Wave Relay Quad Radio Router. WildCat II is a tactical MANET product and can
interact with two separate ground networks, bridge ground and airborne networks,
or act as backhaul (TrellisWare Technologies 2014a). Ocelot is a small form mod-
ule manufactured by the same company that manufactures WildCat II (TrellisWare
Technologies 2014b). The Xiphos 1RU and the Xiphos 6RU radios are 4G tactical
broadband solutions for mobile communication users. The radios claim a range
of 5 to 7 miles on the ground and up to 62 miles via airborne deployment with
clear line of sight (Oceus Networks 2014). The Falcon III RF-7800W OU440 is
rated for operational uses up to 15,000 ft and includes a high level of encryption
capability (Harris 2012b). The Falcon III AN/PRC-1176 is a manpack capable
of UHF/VHF analog voice and digital data (Harris 2012a). Finally, the Direct
Data Link is designed specifically for small UAVs and provides Internet Protocol
(IP)-based communication between a ground station and the aircraft (AeroVi-
ronment 2013). The Raven UAV can come equipped with the Direct Data
Link.

16.4 Cost-effectiveness model


The model for evaluating the effectiveness of the aerial platforms and commu-
nication payloads follows the approach described in Chapter 8, and interested
readers are encouraged to read that chapter for more extensive details. Measur-
ing effectiveness begins with the development of an objectives hierarchy, and the
bottom level of the hierarchy consists of measurable attributes. If the attributes
are mutually preferentially independent, an additive value function can be used to
measure the effectiveness of an alternative (Kirkwood 1997). The MoE of the j th
alternative v( j ) is calculated via the following equation:


M
v( j ) = wi vi (x i ( j )), (16.1)
i=1
406 R. E. Everly, D. C. Limmer, C. A. MacKenzie
where wi is the global tradeoff weight for attribute i , vi (·) is a value function for
attribute i , x i ( j ) is the level of attribute i for alternative j , and M is the total
number of attributes. The range for vi (·) is between 0 and 1, inclusive.
If the attribute is a numerical measure, we assess the value function using
marginal analysis by asking what the incremental or marginal change in value is
when the attribute is increased or decreased. We fit an exponential function to the
assessed values for attribute i using a least-squares approach (see Appendix 16.1
for more details). If the attribute is composed of categorical ratings, such as
“low,” “medium” and “high,” we directly assign values from 0 and 1 for each
category.
Local tradeoff weights are weights assessed for all attributes within a single
objective. We frequently use a swing weight procedure to assess local tradeoff
weights among attributes within an objective (von Winterfeldt and Edwards 1986).
The swing weight procedure assumes that all attributes are at the worst level, and
the decision-maker is asked which attribute he or she desires to move to the best
level. The selected attribute receives a score of 100 and the desirability of “swing-
ing” the other attributes from worst to best are assessed relative to the score of 100
for the most preferred attribute and a score of 0 when all the attributes are at their
worst level. Local tradeoff weights for each attribute are calculated by normalizing
the scores so that these weights sum to one. Multiplying the local weights by the
tradeoff weights for each objective higher up in the objectives  Mhierarchy returns
global weights wi for i = 1, . . . , M for each attribute so that i=1 wi = 1.
After the MoE for each alternative is calculated via Equation (16.1) and Equa-
tion (A16.1.1) in the Appendix, we depict the MoE and the cost of each alternative
as a point on a two-dimensional chart. We analyze the cost-effectiveness of each
alternative that belongs to the efficient solution set, where the efficient set is
composed of alternatives that are not dominated by another alternative. Tradeoffs
between the different alternatives are discussed.1

16.5 Objectives hierarchy


The first step to measure the effectiveness of an alternative is to create an objec-
tives hierarchy describing what should be done to achieve an effective solution
for the aerial platform and communication payload. The aerial platform and com-
munication payload have separate objectives hierarchies, and the bottom level for
each hierarchy consists of attributes that can be measured or observed for each
alternative. To keep the vocabulary simple, we call the first and second levels of the
hierarchy “objectives” and the bottom or third level of the hierarchy “attributes.”
The hierarchies are based on our opinion after extensive research and discussions
with subject matter experts.

16.5.1 Objectives hierarchy for aerial platform


Maximizing the mission effectiveness of the aerial platform is divided into
four objectives: maximizing performance, flexibility, readiness, and survivability
Cost-effectiveness of platforms and payloads 407

Maximize effectiveness

Maximize performance Maximize flexibility Maximize readiness Maximize survivability

Maximize Maximize Minimize Maximize Maximize Maximize Minimize Maximize


range useful launch man tech- all- observ- stealth
load require- porta- nology weather ability
ments bility maturity capability
Maximize Maximize level
en- ceiling Minimize
durance recovery
requirement
Maximize
cruise speed

Figure 16.1 Objectives hierarchy for aerial platform.

(Figure 16.1). Performance describes the capability of an aerial platform to per-


form the mission and is composed of five attributes. We seek to maximize
range (the distance an aerial platform can travel as stated by the manufacturer),
endurance (the maximum number of hours a platform can remain aloft), ceiling
(the platform’s mean sea level as stated by the manufacturer), cruise speed (the
highest sustained operational speed a platform can achieve), and useful load (the
amount of weight a platform can carry).
Flexibility describes the ability to employ the aerial platform in different con-
ditions. Flexibility is further divided into three attributes: launch requirement,
recovery requirement, and man portability. Launch requirement describes the type
of launch required by the aerial platform, and aerial platforms that need less space
launching are favored over assets that require more space for launch. Recovery
requirement describes the type of method utilized for recovering the aerial plat-
form, and platforms that need less space to be recovered are favored over assets
that require more space for landing. Man portability is a binary attribute, and a
platform that is man portable can be transported and launched by a single oper-
ator. We seek to minimize the launch and recovery requirements and maximize
man portability in order to maximize flexibility.
Readiness determines how prepared an aerial platform is to support the mission.
Technology maturity level and all-weather capability are the only two attributes
that define readiness. Other attributes that could also be used to measure readiness
include the mishap rate and availability, but data for these attributes are generally
not available for the alternatives. Technology maturity level is measured by the
number of years in service for the aerial platform based on initial operating capac-
ity, and it serves as proxy measure for the reliability of a platform. Platforms with
less technology maturity lack the field testing and refinement of older and more
tested designs. All-weather capability is a binary attribute, and the platform must
have deicing capability to be all-weather capable.
Survivability describes the ability of an aerial platform to accomplish its mis-
sion without being harmed by the enemy. We want to minimize observability and
408 R. E. Everly, D. C. Limmer, C. A. MacKenzie
maximize stealth in order to maximize survivability. We define observability as
the ability of enemy combatants to detect the aerial platform with the naked eye.
Stealth is a binary attribute. If a platform uses radar absorbent materials similar to
modern stealth aircraft, we consider that the platform has stealth capability.

16.5.2 Objectives hierarchy for communication payload


As depicted in Figure 16.2, the objectives of maximizing performance, flexibil-
ity, and readiness will maximize the effectiveness of a communication payload.
Performance describes the ability of the communication payload to provide a com-
munications link during optimum conditions. We seek to maximize power output
and receiver sensitivity (which measure the payload’s range) and to maximize
throughput. Throughput measures the connection speed between the communica-
tion payload and a single user, as reported by the vendor in raw number of bits per
second. Power output is measured in watts, and receiver sensitivity is measured in
decibel-milliwatts.
Flexibility describes the ability of the communication payload to operate in
varying conditions and fulfill different mission requirements. Flexibility is com-
posed of mesh capability, power consumption, weight, and traffic type. Mesh
capability is a binary attribute because the radio can either execute a mesh
topology—where nodes in the communication network can send and receive data
and serve as a relay for other nodes—or not. We seek to minimize power con-
sumption by the communication payload, which is measured in watts. Weight,
measured in pounds, is based on the radio manufacturer’s specifications. Traffic
type is voice (VHF and UHF transmission), data (e.g., video, imagery), or both.
Readiness determines how prepared a communications payload is to per-
form the mission. As with the aerial platforms, we use technology maturity
level as a proxy measure for readiness because data on the availability of the
communications payloads are not available.

Maximize effectiveness

Maximize performance Maximize flexibility Maximize readiness

Maximize Maximize Maximize Maximize Maximize


through- receiver mesh traffic technology
put sensi- capability type maturity level
tivity
Minimize Minimize
Maximize power weight
power output con-
sumption

Figure 16.2 Objectives hierarchy for communication payload.


Cost-effectiveness of platforms and payloads 409

30
Data points
Best−fit line
Annual O&S cost ($ millions)

25 Global Hawk

20
Triton
15

10
Reaper
Fire
5 Scout
Gray Eagle
Predator
0
0 50 100 150 200 250
Acquisition cost ($ millions)

Figure 16.3 Annualized O&S cost as a function of acquisition cost.

16.6 Cost estimation


Cost-effectiveness analysis should use the LCC for each alternative, which tradition-
ally includes research and development, acquisition, operations and sustainment
(O&S), and disposal costs (see Chapter 5). The program acquisition unit cost
includes research and development, procurement, and military construction costs.
Disposal costs are not included in this analysis either because the aerial platforms
will be used for more than 15 years or because they are assumed to be minimal.
We use the annualized LCC for each aerial platform where the LCC equals the
program acquisition unit cost plus the annual O&S costs multiplied by the planned
service life of each platform. The acquisition cost is based on a single UAV, tower,
or aerostat. We are not able to differentiate the cost of the individual aerial platform
from the launch and recovery equipment, ground control stations and other support
equipment. The annual O&S cost for a platform is estimated as a linear function
of the acquisition cost based on known O&S costs for six UAVs, as depicted in
Figure 16.3. The annual O&S cost is 10.1 percent of the acquisition cost.
Since data on O&S costs for the communication payload alternatives are not
available, we compare the costs for communication payloads only using the acqui-
sition costs. Since O&S costs are not included for communication payloads, the
costs used in this analysis underestimate the true LCCs. Even if more accurate
LCCs for communication payloads were used, their costs would still be minimal
compared to the costs of the aerial platforms.

16.7 Cost-effectiveness analysis


The aerial platforms and their communication payloads are evaluated under three
different scenarios. The three scenarios are disaster relief, long range, and tactical
410 R. E. Everly, D. C. Limmer, C. A. MacKenzie
user. Each scenario poses unique communication challenges, and a decision-
maker’s preferences differ among the scenarios. Consequently, value functions
for each attribute and the tradeoff weights change under each scenario. We
describe each scenario and analyze the cost-effectiveness of aerial platforms and
communication payloads for each scenario.

16.7.1 Disaster relief scenario


Under the disaster relief scenario, we envision a situation similar to Hurricane
Katrina or Typhoon Haiyan in which communication and cellular infrastructure
has been rendered useless, and a nation’s military is ordered to provide aid and
coordinate relief efforts in the area. Communications links for rescue teams are
needed for a fast and successful response. The disaster area is a circular area with a
100 nautical mile (NM) diameter. Task force headquarters is stationary and located
at an airfield on the immediate perimeter of the disaster area. Throughput demand
is expected to be more than 200 Mbps at various times and the number of con-
current users could be between 50 and 100. UHF, VHF, and data relay support is
needed.

16.7.1.1 Cost-effectiveness of aerial platform


The global weights for the disaster relief scenario are presented in Table 16.1.
We use the swing weight procedure to determine the tradeoff weights among
the attributes that define performance. The local weights for the performance
attributes are 0.28 for ceiling, 0.22 each for useful load and endurance, 0.17 for
range, and 0.11 for cruise speed. Having a sufficient ceiling is necessary to ensure
coverage of the entire disaster area. Endurance is important because of the long
communication requirement, and useful load is important because of the large

Table 16.1 Global tradeoff weights for aerial platform for each scenario

Objective Attribute Tradeoff weights for each scenario

Disaster relief Long range Tactical user

Performance Range 0.150 0.090 0.000


Useful load 0.200 0.090 0.155
Endurance 0.200 0.090 0.155
Ceiling 0.250 0.113 0.000
Cruise speed 0.100 0.113 0.000
Flexibility Launch requirement 0.000 0.000 0.115
Recovery requirement 0.000 0.000 0.058
Man portability 0.000 0.000 0.173
Readiness Technology maturity level 0.100 0.200 0.138
All-weather capability 0.000 0.300 0.000
Survivability Observability 0.000 0.000 0.166
Stealth 0.000 0.000 0.041
Cost-effectiveness of platforms and payloads 411
number of first responders who need connectivity. Comparing among the objec-
tives of performance, flexibility, readiness, and survivability, we do not assign
any weight to either flexibility or survivability because the scenario assumes that
headquarters is at an operational airfield and no enemy are present. We consider
performance nine times more important than readiness for this scenario. Because
we assume that icing is not a problem for this scenario, technology maturity level
is the only attribute within flexibility to receive a non-zero weight.
Figure 16.4 depicts the cost-effectiveness of each aerial platform. Ceiling and
useful load account for almost half of the MoE of the most effective solutions
(the Predator, Reaper, Hummingbird, and Global Hawk). The Triton, the Navy’s
variant of the Global Hawk, scores equally as well as those four UAVs for ceil-
ing and useful load but is not scheduled for its first flight until 2015. Thus, the
Triton performs badly on the readiness objective. If the Triton’s technology matu-
rity level equals that of the Global Hawk, its MoE would be similar to that of
Global Hawk.
As can be seen from Figure 16.4 which depicts the annualized LCC on a loga-
rithmic scale, the efficient solution—which eliminates any alternative that is both
less effective and more expensive than another alternative—includes the Raven,
Cerberus Tower, TIF-25K, Predator, Hummingbird, Reaper, and Global Hawk.
The Raven’s endurance of only 1.5 hours and payload of 6.5 ounces make it
an ineffective platform for this scenario. The Cerberus Tower’s height of 30 feet
allows for approximately a 7 NM radio horizon, which is inadequate coverage for
this scenario.
Because the Global Hawk’s effectiveness is only slightly greater than the
Reaper’s MoE and it costs US$30 million more per year, we prefer the Reaper
to the Global Hawk. We also prefer the Predator to the Hummingbird because
the Predator is only slightly less effective and costs US$64,000 less per year. The
Hummingbird has a faster cruising speed but less endurance than the Predator.

1
Hummingbird Global Hawk
Reaper
Predator
0.8 Gray
Fire Scout Triton
Eagle
PTDS 74K UCAS−D
TIF−25K
0.6 Shadow
Cerberus
MoE

Tower Blackjack
RAID Tower
0.4
Raven

0.2 Wasp
T−Hawk

0
10−2 10−1 100 101 102
Annualized LCC ($ millions)

Figure 16.4 Cost-effectiveness of aerial platforms for disaster relief scenario.


412 R. E. Everly, D. C. Limmer, C. A. MacKenzie
Comparing the Predator and Reaper, we ask if annually spending US$1.62 million
is worth the increase in effectiveness from 0.85 (Predator) to 0.86 (Reaper). The
Reaper is judged more effective than Predator because the former has a higher
cruising speed, but in our opinion, the higher speed is not worth the additional
cost for this scenario. We examine the tradeoffs between the TIF-25K aerostat
and the Predator. Although the Predator’s MoE is 0.19 more than the MoE for the
TIF-25K, it also costs US$1.9 million more per year. The TIF-25K’s endurance of
14 days surpasses that the Predator’s endurance of 40 hours. Selecting either alter-
native can be justified. The TIF-25K offers a low cost aerial platform with long
endurance and adequate payload and ceiling, and the Predator offers more maneu-
verability and a larger coverage area, albeit at the expense of less endurance and a
higher cost.

16.7.1.2 Cost-effectiveness of communication payload


The global tradeoff weights for the communication payload attributes for disas-
ter relief are depicted in Table 16.2. We use a swing weight method to determine
the local tradeoff weights for the three attributes (throughput, power output, and
receiver sensitivity) within the performance objective. Improving power output
from the worst (0 watts) to best (500 watts) level is most important and has a local
tradeoff weight of 0.4. Improving throughput is almost as important, and we assign
a local weight of 0.36. Finally, improving receiver sensitivity is least important,
and its local weight equals 0.24. Within the flexibility objective, we assess that
having mesh capable technology is most important, and we judge having mesh
capable technology is twice as important as having a radio that can carry both
voice and data (traffic type). The payload’s weight and power consumption are
not important for this scenario. Comparing the objectives of performance, flexi-
bility, and readiness, we rank performance most important, flexibility second most
important, and readiness least important. We use the rank-sum method to calculate
the local tradeoff weights for these three objectives.2

Table 16.2 Global tradeoff weights for communication payload for each scenario

Objective Attribute Tradeoff weights for each scenario

Disaster relief Long range Tactical user

Performance Throughput 0.180 0.000 0.150


Receiver sensitivity 0.120 0.240 0.090
Power output 0.200 0.240 0.090
Flexibility Mesh capability 0.221 0.000 0.102
Traffic type 0.109 0.000 0.102
Power consumption 0.000 0.165 0.144
Weight 0.000 0.165 0.212
Readiness Technology 0.170 0.190 0.110
maturity level
Cost-effectiveness of platforms and payloads 413

1
Xiphos 6RU

0.8 Xiphos 1RU


Wave Relay
Falcon III RF−7800W
Quad
0.6 WildCat II
MoE

Wave Falcon III AN/PRC


Relay Ocelot
0.4

Digital Data Link


0.2

0
100 101 102 103 104
Acquisition cost ($ thousands)

Figure 16.5 Cost-effectiveness of communication payloads for disaster relief


scenario.

As can be seen from Figure 16.5, the efficient solution set comprises the Wave
Relay, Wave Relay Quad, Falcon III RF-7800W, Xiphos 1RU, and Xiphos 6RU.
(Figure 16.5 shows the acquisition cost for each radio on a logarithmic scale
to more clearly see the differences in costs.) If the decision-maker places more
importance on receiver sensitivity and traffic type relative to throughput, the Fal-
con III AN/PRC will replace the Falcon III RF-7800W as a member of the efficient
solution set. Regardless, either of the Falcons is only slightly more effective than
the Wave Relay Quad, and the Falcons cost US$18,000 more, and we prefer the
Wave Relay Quad to either of the Falcons. The Xiphos 1RU and 6RU are most
effective with MoEs of 0.79 and 0.93 respectively. Both radios offer high power
output, high throughput, and excellent scalability for multiple users. Their superior
capabilities are worth the additional cost.

16.7.1.3 Solution compatibility


Both the TIF-25K and the Predator have enough useful load to carry either of
the Xiphos radio units. However, available power for the radio units is a differ-
entiator. The Predator has 1,800 watts available for payloads. The Xiphos 6RU’s
power consumption is over 3,000 watts. The Predator cannot normally support the
Xiphos 6RU. The TIF-25K aerostat is capable of sending power to its payload
from a ground based source up the tether. The TIF-25K should be able to power
the Xiphos 6RU radio unit. This may require high voltage power sent from the
ground up to the aerostat to avoid high power losses over the long transmission
line of the tether, but it is feasible. If a decision-maker prefers the more effective
Xiphos 6RU because of its better throughput and power output, we believe the
TIF-25K is the most cost-effective aerial platform. If the decision-maker prefers
414 R. E. Everly, D. C. Limmer, C. A. MacKenzie
the less expensive Xiphos 1RU, either the Predator with its greater coverage or the
TIF-25K with its long endurance can serve as an effective aerial platform.

16.7.2 Long-range scenario


In the long-range scenario, the military leadership at headquarters needs to establish
UHF (voice) communications with a ship 340 NM away. The ship has lost its satel-
lite communication capability, and we assume that headquarters is based at a large
airfield near the coast. No enemy resistance is expected. The weather is cloudy with
the possibility of icing conditions at higher altitudes. We assume that headquarters
will need to communicate with the ship for approximately four hours.

16.7.2.1 Cost-effectiveness of aerial platform


Table 16.1 displays the global tradeoff weights for the long-range scenario. As
with the disaster relief scenario, the objectives for flexibility and survivability
receive zero weight because the military is operating from a large airfield and
enemies are not present. Within the performance objective, we believe it is most
important to improve cruising speed, and cruising speed receives a local weight
equal to 0.24. Improving the ceiling is only slightly less important, and the local
tradeoff weight for ceiling equals 0.23. Useful load, endurance, and range each
receive local weights equal to 0.18. Within the readiness objective, the deicing
capability may be crucial, and we assess local weights of 0.6 for all-weather
capability and 0.4 for technology maturity level. The attributes that determine per-
formance (useful load, ceiling, endurance, range, and cruising speed) have value
functions in which the value of 0 corresponds to satisfactory requirements to com-
plete the communication mission. Consequently, we are willing to trade off more
performance in favor of readiness than we were in the disaster relief scenario. Both
the performance and readiness objectives receive a tradeoff weight equal to 0.5.
The long-range scenario requires that aerial platforms satisfy a minimum ceil-
ing of 19,102 ft. This number is calculated as a function of the 340 NM distance
between the ship and military headquarters. The Shadow, Raven, Wasp, and
T-Hawk UAVs, the TIF-25K aerostat, and the RAID and Cerberus Towers are
excluded from this analysis because their ceilings do not meet this requirement.
The Blackjack’s ceiling is 19,500 ft, but its range of 55 NM is so small that it
would need a much higher ceiling to be effective for this scenario. Consequently,
the Blackjack is not feasible for this scenario.
The efficient solution set for the long-range scenario consists of the Predator,
Reaper, and Triton (Figure 16.6). The Triton and UCAS-D are the most effec-
tive because of their deicing capability. Since the first flight for the UCAS-D is
scheduled for 2019, its technology maturity level performs poorly, but even if it
had the same technology maturity level as the Triton, the latter would still be
more effective because the Triton has a higher ceiling than the UCAS-D. We esti-
mate the annualized LCC of the UCAS-D at US$95 million, but its price point
may come down to that of the Triton as the technology becomes more mature.
Cost-effectiveness of platforms and payloads 415

Triton
0.8
UCAS−D
Reaper
Global Hawk
0.6
Predator Hummingbird
MoE

Gray Eagle
0.4 Fire Scout
PTDS 74K

0.2

0
10−2 10−1 100 101 102
Annualized LCC ($ millions)

Figure 16.6 Cost-effectiveness of aerial platforms for long-range scenario.

If the UCAS-D’s LCC is closer to that of the Triton, the UCAS-D could be a
cost-effective alternative for the long-range scenario.
The Triton is annually US$23–25 million more expensive than the Reaper and
Predator, but the Triton’s MoE is 0.21 greater than that of the Reaper and 0.30
greater than that of the Predator. This gap in effectiveness is primarily due to
the Triton’s deicing capability. If icing conditions are not a factor, we prefer the
Reaper to the Triton. The Reaper costs US$1.6 million more than the Predator, but
its MoE is 0.09 greater than that of the Predator. All three UAVs will be evaluated
for compatibility with the most cost-effective communication payloads.

16.7.2.2 Cost-effectiveness for communication payload


Since the headquarters needs to be able to talk with the ship, the communication
payload must support voice transmission. Given this requirement, the Falcon III
AN/PRC and the Wildcat II are the only feasible alternatives. Within the flexibil-
ity objective, we are indifferent between trading off between power consumption
and weight, and the traffic type receives no importance because the two feasible
alternatives can support both voice and data transmission (Table 16.2). We are also
indifferent between trading off between power output and receiver sensitivity, and
throughput is not important for this scenario. When comparing among the higher-
level objectives, we prefer to increase performance, then to increase flexibility,
and finally to improve readiness. We assign local weights of 0.48, 0.33, and 0.19
to those three objectives, respectively.
Figure 16.7 depicts the cost-effectiveness of the two feasible radio units. The
Falcon III AN/PRC and the Wildcat II are about equally effective according
to our preferences, and we prefer the less expensive Falcon III AN/PRC. If a
decision-maker is unwilling to trade off as much flexibility for performance as
416 R. E. Everly, D. C. Limmer, C. A. MacKenzie

0.8
Falcon III AN/PRC WildCat II

0.6
MoE

0.4

0.2

0
100 101 102 103
Acquisition cost ($ thousands)

Figure 16.7 Cost-effectiveness of communication payloads for long-range scenario.

we are, the Wildcat II will be rated more effective than the Falcon III AN/PRC.
A decision-maker may prefer the lighter Wildcat II that consumes less power than
the Falcon III AN/PRC. The cost difference between the two communication pay-
loads is about US$25,000, which may not be an important discriminator when the
annual cost of the UAV is US$2 million or more.

16.7.2.3 Solution compatibility


The three identified cost-effective aerial platforms for the long-range scenario are
the Predator, Reaper, and Triton UAVs. The Predator offers an effective low cost
option. The Reaper offers twice the ceiling of the Predator and a turbine engine
for a reasonable increase in cost over the Predator. The Triton offers unparalleled
mission effectiveness but is also the most expensive of the three options by several
orders of magnitude. Budget considerations as discussed in Chapters 4 and 7 may
influence which UAV is selected. All three UAVs can easily support the Falcon
III AN/PRC, which is a 12-pound, 55-watt 117G radio. Because these UAVs can
easily support the Falcon III AN/PRC, a decision-maker may not want to purchase
the more expensive Wildcat II for this scenario.

16.7.3 Tactical user scenario


This tactical user scenario involves a small special operations team or a single
soldier trying to establish communications with higher command to coordinate
extract. The tactical user, who is close to completing his mission and needs to radio
back to base to arrange a helicopter extract, is about 10 NM away from the base,
and a 500 ft mountain is blocking radio signals back to base. The tactical user lacks
satellite communication capability. No extraction plans or time of next contact
were established during the last contact between the base and the tactical user
Cost-effectiveness of platforms and payloads 417
more than 24 hours ago. Only a small number of end users need to be supported.
Although the primary method of communications is UHF/VHF, video may also
be necessary. Additionally, the enemy is believed to be in the vicinity. The tactical
user prefers to have possession and control of the aerial platform.

16.7.3.1 Cost-effectiveness of aerial platform


If the 500 ft mountain is located midway between the user and headquarters, the
tactical user scenario requires a minimum ceiling of 1,000 ft for the aerial plat-
form. This requirement eliminates the T-Hawk, RAID Tower, and Cerberus Tower
as feasible alternatives. After meeting this requirement, an aerial platform with a
higher ceiling will not necessarily perform better, and we assign a tradeoff weight
equal to 0 to the ceiling attribute. The attributes range and cruising speed are not
important for this scenario and also receive a weight of 0. We determine an equal
preference for trading off between useful load and endurance, and each attribute
receives a local weight of 0.5 within performance. Within the objective of flexibil-
ity, we use the rank-sum method to assign local weights after determining that man
portability is the most important, launch method is the second most important, and
recovery method is the least important. We determine that observability is much
more important than stealth because the enemy is relatively low tech and decreas-
ing observability is four times as important as moving from no stealth to stealth
capability. This scenario has good weather, and deicing capability is unimportant,
and technology maturity level has a local weight of 1.0.
The global weights for the attributes for the tactical user scenario are depicted in
Table 16.1. We use a swing weight procedure to determine the relative importance
of performance, flexibility, readiness, and survivability. Improving the flexibility
of the aerial platform so that the platform can be carried and launched by a single
individual is slightly more important than improving the performance (increas-
ing the useful load and ceiling). Less important is improving the survivability
(decreasing the observability), and least important is increasing the readiness
(improving the technology maturity level).
The Raven is the superior solution for the tactical user scenario because it is the
most effective and least expensive alternative (Figure 16.8). The Raven’s small
size allows it to be carried by a single individual. It can only remain in the air for
1.5 hours, but that should be enough to allow the user to communicate with higher
command. If man portability is a requirement, the only other feasible platform is
the Wasp. The Wasp’s smaller size and lower weight may make it more portable
than the Raven, but the Raven has more endurance and range.In our opinion, hav-
ing more endurance is more important than the smaller size, and we select the
Raven as the most cost-effective platform for this scenario.

16.7.3.2 Cost-effectiveness of communication payload


Because we are selecting a UAV that can be carried by a single user, we place
a lot of importance on maximizing flexibility, as depicted in Table 16.2. Using a
swing weight procedure, we assign tradeoff weights of 0.56 for flexibility, 0.33 for
418 R. E. Everly, D. C. Limmer, C. A. MacKenzie

0.8
Raven Hummingbird
Cerberus Fire Scout
Tower RAID Tower
0.6 PTDS 74K
TIF−25K Shadow
Reaper Global Hawk
MoE

Wasp
Blackjack
Predator UCAS−D
0.4 T−Hawk Gray Eagle
Triton

0.2

0
10−2 10−1 100 101 102
Annualized LCC ($ millions)

Figure 16.8 Cost-effectiveness of aerial platforms for tactical user scenario.

performance, and 0.11 for readiness. Within the flexibility objective, we believe
it is important to minimize the weight of the radio, followed by minimizing
power consumption. The attribute for weight has a local tradeoff weight of 0.38,
and power consumption has a local weight of 0.26. Mesh capability and traffic
type each have local weights of 0.18. For the performance objective, increasing
throughput is the most important attribute, to allow the tactical user to transmit
some video. The local weight for throughput is 0.45, and the both of the local
weights for power output and receiver sensitivity are 0.27.
Figure 16.9 displays the cost-effectiveness for the communication payloads.
The Wave Relay radio is the most effective and the least expensive alternative. The
Wave Relay’s MoE is the largest because of the radio’s small weight and relatively
high throughput. If a decision-maker places more importance on power consump-
tion and receiver sensitivity relative to throughput, the Ocelot will become more
effective than the Wave Relay although the former costs almost three times as
much as the Wave Relay. Both communication payloads weigh less than 0.2 lbs.
If a decision-maker is unwilling to trade off as much performance in favor of
flexibility than in our assessment, the WildCat II may be the most effective. The
WildCat II weighs 3.4 lbs, which likely makes it too heavy for the small UAVs.

16.7.3.3 Solution compatibility


The Raven and Wasp III UAVs have useful loads of 0.41 and 0.2 lbs, respectively,
which excludes use of the WildCat II radio. The Wasp III would also have trouble
carrying either radio. The Raven can easily accept the weight of the Ocelot and
the Wave Relay card at 96 grams. However, with additional wiring, dedicated
battery and antenna, either the Ocelot or Wave Relay could exceed the Raven’s
useful load, which could result in degraded performance of the airframe. Even
Cost-effectiveness of platforms and payloads 419

0.8
Wave Relay
Ocelot WildCat II
0.6 Digital Xiphos 6RU
MoE

Data Falcon III RF−7800W Xiphos 1RU


Link
0.4 Falcon III AN/PRC
Wave Relay Quad

0.2

0
100 101 102 103 104
Acquisition cost ($ thousands)

Figure 16.9 Cost-effectiveness of communication payloads for tactical user scenario.

though both communication payloads possess a small form factor, the payload
compartment of the Raven may not be able to support either radio. A previous
study (Menjivar 2012) tested a Raven with a Wave Relay payload. The Wave Relay
radio had to be taped onto the outside of the UAV instead of being secured inside
the payload compartment, which is not ideal for performance of the UAV or the
communication link.
One communication payload that would work with the Raven is AeroViron-
ment’s Digital Data Link. AeroVironment offers the Digital Data Link with the
Raven UAV from the factory. Digital Data Link’s price is positioned between the
Wave Relay and the Ocelot at US$5,000, but it has a slightly lower MoE than either
one at 0.55. However, compatibility between the Digital Data Link and the Raven
is assured, whereas the Ocelot and the Wave Relay units needs further testing to
ensure compatibility.

16.8 Conclusions
This chapter demonstrates the use of traditional multi-criteria decision-making
techniques to analyze the cost-effectiveness of aerial platforms and communi-
cation payloads for communication missions in the military. We compared the
cost-effectiveness of 17 aerial platforms and 9 communication payloads across
three mission scenarios.
The first scenario requires long endurance and high bandwidth capability to
complete a disaster relief mission. The most cost-effective aerial platforms are the
TIF-25K aerostat and the Predator UAV. The TIF-25K combines endurance mea-
sured in weeks and a high useful load with a moderate ceiling. Additionally, the
TIF-25K can utilize a tether to power its payloads from the ground. The Predator
is a very capable UAV platform with a relatively long endurance, long range, and
420 R. E. Everly, D. C. Limmer, C. A. MacKenzie
reasonable cost of ownership. The most cost-effective communication payloads
are the two configurations of the Oceus Networks Xiphos (1RU and 6RU) due to
their high throughput, power output, and excellent scalability. The Xiphos radios
can be relatively heavy (78 lbs to 276 lbs depending on configuration) and power
hungry (855 to 3,275 watts depending on configuration). We select either a TIF-
25K aerostat with a Xiphos 6RU connected to ground power or a Predator with a
Xiphos 1RU configuration.
In the second scenario, a long-range relay is needed to connect users across a
340 NM range, which requires a minimum altitude of 19,102 ft due to radio hori-
zon and UHF capability. The most cost-effective aerial platforms are the Predator,
Reaper, and Triton. Due to its UHF capability, sensitivity, and power output, the
Falcon III AN/PRC is the most cost-effective communication payload. Each of the
three aerial platforms could be the best choice, depending on the decision-maker’s
preference for price (Predator), better altitude and performance (Reaper), or out-
standing performance and deicing capability (Triton). The extra performance and
capability of the Triton comes with a much higher ownership cost than the Predator
or Reaper.
A covert, tactical situation where portability is preferred is the final scenario.
The most cost-effective aerial platform is the Raven, which combines man porta-
bility and adequate range and endurance with the lowest cost of any aerial platform
in this study. The Wave Relay is the most cost-effective communication pay-
load for this scenario because it weighs very little and possesses a relatively
high throughput for its small form factor. However, compatibility between the
Raven and Wave Relay cannot be confirmed at this time. The slightly less effec-
tive and more expensive Digital Data Link is compatible with the Raven and can
be an acceptable solution in the interim until more compatibility testing can be
completed between the Raven and the Wave Relay communication payload.
As with any analysis, our cost-effectiveness analysis relies on a number of
assumptions, which can motivate future research. We assess compatibility between
the most cost-effective aerial platforms and communication payloads within each
scenario based primarily on the manufacturer’s provided specifications. Actual
field testing would provide a proof of concept and help validate the findings of
this chapter. Since complete, transparent, and reliable cost data was not always
available, further research of the fully captured LCCs for UAVs, military towers,
and aerostats will enable a more accurate and detailed analysis. Other missions,
such as ISR, may present an interesting and useful subject for future research,
and other aerial platforms currently under development could be considered.
Finally, other decision-makers may have different objectives constraints, value
functions, and tradeoff weights, which would change the alternatives’ MoEs and
the cost-effectiveness analysis.
This chapter offers a framework for comparing the cost-effectiveness of dissim-
ilar aerial platforms and communication payloads across different mission sets.
Several militaries around the world currently operate UAVs, and their use for com-
munication will likely grow in the future. This research develops value functions
and weighting parameters that change based upon mission requirements and are
Cost-effectiveness of platforms and payloads 421
easily adapted to other mission sets including the traditional ISR mission. Our
analysis demonstrates that the cost-effectiveness of an alternative depends on the
mission requirements, and suggests the military should continue to purchase a
wide array of aerial platforms and communication payloads that can be used for
different missions.

Acknowledgments
We would like to thank Glen Cook and John Gibson of the Naval Postgraduate
School for their advice and feedback during this research and for sharing their
expertise in communication systems. Copies of the spreadsheets used to ana-
lyze the cost-effectiveness of the aerial platforms and communication payloads
can be downloaded and altered to meet a decision-maker’s own preferences from
Cameron MacKenzie’s web page at https://faculty.nps.edu/camacken/ under the
tab “Thesis Supervision.”

Appendix 16.1
Exponential value functions are used for the numerical attributes. The value of
attribute i at a given level x i can be calculated based on whether the decision-
maker prefers “more” or “less” of an attribute:

⎪ 1 − exp(a[x i − x min ]b )

⎨ if more is preferred
K
vi (x i ) = (A16.1.1)

⎩ 1 − exp(a[x max − x i ] )
⎪ b
if less is preferred,
K

where K = 1 − exp(a[x max − x min ]b ) is a normalizing constant, x min is the attribute


level for which v(x min ) = 0 if more is preferred and v(x min ) = 1 if less is preferred,
x max is the attribute level for which v(x max ) = 1 if more is preferred and v(x max ) = 1
if less is preferred, and a ∈ R and b > 0 are parameters selected to minimize the
sum of the squared distances between vi (x i ) and the assessed values.
In order for the value function to be defined over the domain of the attribute and
so that 0 ≤ vi (x i ) ≤ 1, we require that vi (x i ) = 0 for all x i ≤ x min and vi (x i ) = 1 for
all x i ≥ x max if more is preferred and require that vi (x i ) = 0 for all x i ≥ x max and
vi (x i ) = 1 for all x i ≤ x min if less is preferred.

Notes
1 See Chapter 4 for an alternative perspective.
2 The rank-sum only requires the decision-maker to rank the objectives rather than deter-
mining the exact tradeoff between multiple objectives. For three objectives, the most
important objective receives an unnormalized weight of 3, the second most important
receives a 2, and the least important receives a 1. Dividing each weight by the sum of
these weights returns normalized weights of 1/2, 1/3, and 1/6 for the first, second, and
third objectives, respectively.
422 R. E. Everly, D. C. Limmer, C. A. MacKenzie
References
AeroVironment. 2013. “Digital Data Link.” Available at https://www.avinc.com/uas/
small_uas/ddl/ [last accessed August 29, 2014].
American Institute of Aeronautics and Astronautics. 2013. “UAV Roundup 2013.”
Aerospace America, July/August. Available at www.aerospaceamerica.org/Documents/
AerospaceAmerica-PDFs-2013/July-August-2013/UAVRoundup2013t-AA-Jul-Aug
2013.pdf [last accessed September 15, 2014].
“Army Deploys 300th RAID Tower, Supporting Forward Base Protection by Per-
sistent Surveillance and Dissemination System PSDS2.” November 23, 2008.
Defense Update, November 23. Available at http://defense-update.com/features/2008/
november/231108_psds2_raid_sensors.html [last accessed August 29, 2014].
Buede, Dennis M., and Terry A. Bresnick. 1992. “Applications of Decision Analysis to the
Military Systems Acquisition Process.” Interfaces 22: 110–125.
Collier, Corey M., and Jeffrey C. Kacala. 2008. “A Cost-Effectiveness Analysis of Tacti-
cal Satellites, High-Altitude Long-Endurance Airships, and High and Medium Altitude
Unmanned Aerial Systems for ISR and Communication Missions.” Master’s thesis,
Naval Postgraduate School.
Department of Defense. 2013. “Unmanned Systems Integrated Roadmap
FY2013-2038.” Reference number 14-S-0553. Available at www.defense.gov/
pubs/DOD-USRM-2013.pdf [last accessed August 29, 2014].
Dobbing, Mary, and Chris Cole. 2014. Israel and the Drone Wars: Examining Israel’s
Production, Use and Proliferation of UAVs. Oxford, UK: Drone Wars UK. Available
at http://dronewarsuk.files.wordpress.com/2014/01/israel-and-the-drone-wars.pdf [last
accessed November 27, 2014].
Ewing Jr., Paul L., William Tarantino, and Gregory S. Parnell. 2006. “Use of Decision Anal-
ysis in the Army Base Realignment and Closure (BRAC) 2005 Military Value Analysis.”
Decision Analysis 3: 33–49.
Ferguson, Charles R., and Douglas A. Harbold. 2001. “High Altitude Long Endurance
(HALE) Platforms for Tactical Wireless Communications and Sensor Use in Military
Operations.” Master’s thesis, Naval Postgraduate School.
Fry, Jereed N., and Steven E. Tutaj. 2010. “A Business Case Analysis for the Vulture
Program.” MBA Professional Report, Naval Postgraduate School, Monterey, CA.
General Atomics Aeronautical. 2012a. “MQ-1C Gray Eagle.” Available at www.
ga-asi.com/products/aircraft/gray_eagle.php [last accessed August 29, 2014].
——. 2012b. “MQ-9 Reaper/Predator B.” Available at www.ga-asi.com/products/
aircraft/predator_b.php [last accessed August 29, 2014].
——. 2013. “MQ-1 Predator.” Available at www.ga-asi.com/products/aircraft/
predator.php [last accessed August 29, 2014].
Harris. 2012a. “AN/PRC-177G(V)1(C).” Available at http://rf.harris.com/capabilities/
tactical-radios-networking/an-prc-117g/default.asp [last accessed August 29, 2014].
——. 2012b. “RF-7800W-OU440.” Available at http://rf.harris.com/capabilities/
tactical-radios-networking/rf-7800w.asp [last accessed August 29, 2014].
Keeney, Ralph L. 1996. Value-Focused Thinking: A Path to Creative Decisionmaking.
Cambridge, MA: Harvard University Press.
Keeney, Ralph L., and Howard Raiffa. 1976. Decisions with Multiple Objectives: Prefer-
ences and Value Tradeoffs. New York: John Wiley & Sons.
Kirkwood, Craig W. 1997. Strategic Decision Making: Multiobjective Decision Analysis
with Spreadsheets. Belmont, CA: Brooks/Cole.
Cost-effectiveness of platforms and payloads 423
Lawler, Paul P. 2010. “Cost Implications of the Broad Area Maritime Surveillance
Unmanned Aircraft System for the Navy Flying Hour Program and Operation and
Maintenance Budget.” Master’s thesis, Naval Postgraduate School.
Menjivar, Jose D. 2012. “Bridging Operational and Strategic Communication Architectures
Integrating Small Unmanned Aircraft Systems as Airborne Tactical Communication
Vertical Nodes.” Master’s thesis, Naval Postgraduate School, Monterey, CA.
Naval Air Systems Command. 2014. “Aircraft and Weapons: Unmanned Combat Air Sys-
tem Demonstration.” Available at www.navair.navy.mil/index.cfm [last accessed May
24, 2014].
Nicholas, Ted G., and Rita Rossi. 2011. U.S. Unmanned Aircraft Systems, 2011. Fountain
Valley, CA: Data Search Associates.
Oceus Networks. 2014. “Oceus Networks Xiphos.” Available at www.oceusnetworks.
com/products/xiphos [last accessed August 29, 2014].
Parsons, Dan. 2013. “Worldwide, Drones Are in High Demand.” National
Defense Magazine, May. Available at www.nationaldefensemagazine.org/archive/
2013/May/Pages/Worldwide,DronesAreinHighDemand.aspx [last accessed September
15, 2014].
Persistent Systems. 2012. “Technology.” Available at www.persistentsystems.com/
pdf/WaveRelay_ WhitePaper_Technology_01.pdf [last accessed August 29, 2014].
——. 2014. “Quad Radio Router.” Available at www.persistentsystems.com/quad/ [last
accessed August 29, 2014].
Raven Aerostar. 2014. “TIF-25K teathered aerostat.” Available at http://raven
aerostar.com/solutions/aerostats/tif-25k [last accessed August 29, 2014].
Thiow Yong Dennis, Lim. 2007. “A Methodological Approach for Conducting a Business
Case Analysis (BCA) of the Global Observer Joint Capability Technology Demonstra-
tion (JCTD).” Master’s thesis, Naval Postgraduate School.
TrellisWare Technologies. 2104a. “TW-130 WildCat II.” Available at https://www.
trellisware.com/wp-content/uploads/TW-130-WildCat-II-Product-Bulletin.pdf [last
accessed March 5, 2014].
——. 2014b. “TW-600 Ocelot.” Available at https://www.trellisware.com/wp-content/
uploads/TW-600_Ocelot_Product_bulletin.pdf[last accessed March 9, 2014].
“US Army Aerostat-Based PTDS Provide IED Warning.” June 8, 2010. Defense Industry
Daily. Available at www.defenseindustrydaily.com/133M-to-Lockheed-Martin-for-US-
Army-Aerostat-based-Warning-System-05835/ [last accessed August 29, 2014].
von Winterfeldt, Detlof, and Ward Edwards. 1986. Decision Analysis and Behavioral
Research. Cambridge, UK: Cambridge University Press.
Yilmaz, Bahadir. 2013. “Cost and Effectiveness Analysis on Unmanned Aerial Vehicle
(UAV) Use at Border Security.” In SPIE Proceedings, edited by Edward M. Cara-
pezza. Paper presented at the Sensors, and Command, Control, Communications, and
Intelligence (C3I) Technologies for Homeland Security and Homeland Defense XII,
Baltimore, Maryland, April 29. doi:10.1117/12.2015914.
17 Time discounting in military
cost–benefit analysis
Jason Hansen and Jonathan Lipow

17.1 Introduction
The proverb “a bird in the hand is worth two in the bush” can be found in John
Ray’s 1670 A Handbook of Proverbs, and remains just as valid today. People place
higher value on what is certain—a bird in hand ready to eat—than they place on
the uncertain—two birds in the bush that may be ready to escape. Since the past
and present are by definition certain, uncertainty is really a characteristic of the
future. Hence, since people value certainty over uncertainty, they also value the
present over the future.
Careful accounting for the difference in value that people place on the present
as opposed to the future is an important requirement of any properly conducted
cost–benefit analysis (CBA). Costs incurred in the present tend to be weighted
more heavily than those that can be deferred to the future. Similarly, monetary
and other benefits received in the present tend to be weighted more heavily that
those that will only be enjoyed at some future date.
The standard method used in military CBA to adjust costs and benefits to reflect
their timing is known as “present value analysis” (PVA) or “time discounting.”
This chapter briefly reviews how to apply PVA to CBA, paying extra attention to
the most problematic aspect—the choice of an appropriate discount rate. The next
section offers a simple example of PVA applied to CBA. Section 17.3 discusses
the challenge of selecting an appropriate discount rate. Risk is explicitly incorpo-
rated into the discussion in Section 17.4. Concluding comments in Section 17.5
reinforce the importance of conducting sensitivity analysis on discount rates and
other key parameters in a military CBA.

17.2 Present value analysis


Suppose you have $100 in cash. You can spend that money today, or you can
deposit it in a risk-free savings account offered by a local bank that offers cus-
tomers 10 percent annual interest. This means you have a choice between spending
$100 today, or having $110 to spend in a year’s time. Of course, if you left your
$110 in the account for another year, you would have even more money to spend
the following year. Your $100 grows over time as described in Table 17.1.
Time discounting in war and peace 425
Table 17.1 Future value of $100

Year Growth $ Future value $

1 100 × (1 + 10%) = 110.00


2 100 × (1 + 10%) (1 + 10%) = 121.00
3 100 × (1 + 10%) (1 + 10%) (1 + 10%) = 133.10

The equation that defines this growth is

FV = PV(1 + r )n , (17.1)

where the future value (FV) is the value n years in the future of a sum of money
whose present value (PV) is deposited in an account today earning interest at an
annual rate of r . Solving for PV, we find that

FV
PV = . (17.2)
(1 + r )n

Equation (17.2) is the foundation of PVA.1 It determines the PV of a sum (FV)


that we expect to receive n years in the future. Table 17.2 calculates the PVs for
$100 that we expect to receive one, two or three years in the future.
These PVA relationships can be applied to a simple CBA problem. The CBA
problem here is structured as a “fixed effectiveness” approach, the second of
the six economic evaluation of alternatives (EEoA) approaches described in
Chapter 4. Where competing alternatives are designed to offer the same util-
ity/benefit/effectiveness, the analysis focuses on minimizing costs. In the US
Department of Defense this approach is also known as “lowest price technically
acceptable” (LPTA).
Consider a CBA that involves the comparison of program costs of two equally
capable helicopters—Alpha and Beta—that have a useful life of six years.2 The
procurement cost is $425 million for Alpha and only $400 million for Beta. As
discussed in Chapter 5, total life-cycle costs should also include future operations
and maintenance (O&M) costs.
Once acquired, operating and maintaining the Alpha will cost $50 million per
year for three years, but then will increase to $100 million per year. For Beta, these

Table 17.2 Present value of $100 received in the future

Year Conversion $ Present value $

1 100/ (1 + 10%) =1
90.90
2 100/ (1 + 10%)2 = 82.64
3 100/ (1 + 10%)3 = 75.19
426 J. Hansen, J. Lipow
Table 17.3 Comparative life-cycle costs

Alternatives

Alpha Beta

Year Cash ($m) PV ($m) Cash ($m) PV ($m)


(1) (2) (3) (4) (5)

Procurement 0 425.00 425.00 400.00 400.00


Operations and 1 50.00 45.45 100.00 90.91
maintenance 2 50.00 41.32 100.00 82.64
3 50.00 37.57 100.00 75.13
4 100.00 68.30 50.00 34.15
5 100.00 62.09 50.00 31.05
6 100.00 56.45 50.00 28.22
Total 875.00 736.18 850.00 742.11
Interest rate = 10%

operation and maintenance costs are reversed: $100 million per year for the first
three years but only $50 million per year for the following three years.
Calculating the six year life-cycle costs (LCC) for each alternative, columns (2)
and (4) in Table 17.3 reveals that Alpha’s LCC price is $875 million, while Beta’s
is only $850 million. But is Beta really the “lowest price” alternative?
Comparing columns (2) and (4) in Table 17.3, we see that Beta costs less to
procure but more to operate initially. Overall, however, helicopter Beta requires a
lot more cash than Alpha over the first three years, and a lot less in the final three
years. One possible interpretation is that if Alpha was selected, there would be
excess funds freed up during those early years that could be deposited in a bank
earning interest. Funds earned in that account could then be used to defray the
higher costs that Alpha would incur in later years. The discounted present value
calculations displayed in columns (3) and (5) of Table 17.3 take these additional
funds into account. This suggests that (assuming a 10 percent interest rate) Alpha,
not Beta, may be the lowest price option.

17.3 Choosing an appropriate discount rate


As we have seen, applying PVA to CBA is relatively straightforward. What is dif-
ficult is to choose an appropriate discount rate, or the value of r . For convenience,
r was previously referred to as an interest rate. However, for public investment
decisions, it is more accurate to characterize r as the discount rate, or even the
“social rate of discount,” since its value is unlikely to equal any particular interest
rate observed in the capital markets or offered by banks.
What value should be used for r and how should that value be identified? A
survey of over 2,000 economists did not come close to answering those ques-
tions (Weitzman 2001). While economists clearly cannot identify the correct value
for r , there is a surprisingly wide consensus about what, at least in theory, the value
Time discounting in war and peace 427
of r ought to reflect. The foundation of this consensus is Ramsey’s seminal 1928
analysis of the optimal savings rate (Ramsey 1928).
Ramsey reasoned that the optimal savings rate is the one that maximizes aggre-
gate utility (benefits) for all present and future citizens. This can only be achieved
if the marginal rate of substitution (MRS) of any two consumption goods equals
the marginal rate of technical substitution (MRTS) in the production of the two
goods. By choosing current aggregate consumption and future aggregate con-
sumption as the goods in question, Ramsey then solves for the value of r that
leads MRS to equal MRTS:

r = α + ηλ, (17.3)

where α is the pure rate of time discount, λ is the rate at which per capita real con-
sumption is expected to grow over the long term, and η is the marginal elasticity
of utility with respect to real consumption. Each variable is interpreted carefully
below.
We can think of α as the portion of r that discounts future utility or well-
being (Heal 2007). It captures the reality that people are impatient, and “like
small children. . . hate waiting for anything,” as well as capturing the degree to
which people fear unexpected death or similar, equally unforeseen as undesirable,
events (Hansen and Lipow 2013). The larger α, the greater the desire to enjoy
benefits earlier rather than later.
We can think of ηλ as the portion of r that discounts future consumption bene-
fits (Heal 2007). The value of η captures the marginal utility of consumption, and
is also a measure of risk aversion. A value of η = 0 implies that a person is risk
neutral so that their marginal utility of consumption is constant over time. Hansen
and Lipow (2013) point out that:

this is the equivalent of assuming that if the prosperous Fitzwilliam Darcy


(from Jane Austen’s Pride and Prejudice) found $100 lying on the ground, it
would offer him just as much utility as it would offer the impoverished Oliver
Twist (from Charles Dickens’ Oliver Twist) should he find the same sum.
Clearly, this is wrong – people are risk averse and η does not equal zero.3

The value of λ captures the degree to which we expect consumption opportuni-


ties to be available in the future. If the value of λ is high, then future consumption
is expected to be much higher than it is today. As a result, the incremental utility
that this future consumption will provide will tend to be lower, and the correspond-
ing discount rate higher (i.e. one tends to discount future consumption benefits
relatively more, if it is expected there will be more available). If the value of λ
is low, then future consumption is expected to be much lower. As a result, future
incremental utility will be higher, and the discount rate correspondingly lower (i.e.
one tends to discount future consumption benefits relatively less, if it is expected
there will be less available).
428 J. Hansen, J. Lipow
If there is a widespread agreement amongst economists on the appropriate for-
mula for the discount rate, r = α + ηλ, why can’t a consensus value for the discount
rate be determined? The answer is that there are bitter disagreements over the
correct values for the individual variables: α, η and λ. Different assumptions and
perspectives lead to parameter estimates that result in real (inflation adjusted) dis-
count rates that range between 1 percent and 7 percent (Weitzman 2001). The
contentious challenge of converging to a consensus value has led many economists
to simply recommend using the government’s cost of borrowing as an imperfect
proxy for the discount rate (US OMB 1992).

17.4 Risk adjusted discount rates


Ramsey’s analysis assumes a world of certainty, and many CBA practitioners con-
duct PVAs using various estimates for the value of r identified in equation (17.3).
This may offer a false sense of security. To illustrate why, consider the following
example described in Hansen and Lipow (2013).
Suppose a country of very like-minded people wish to maximize their collec-
tive well-being over two periods, subject to the constraint of being adequately
defended by their nation’s navy. To obtain an adequate defense capability, the
country must add an additional submarine to the fleet in period two. In order to do
so, the navy must build it during period one.
The navy can propel the submarine using one of two alternatives. One possi-
bility is to equip the submarine with a nuclear reactor. The second is to equip it
with diesel-electric propulsion. For simplicity, this CBA assumes the two alter-
native propulsion systems—nuclear and diesel-electric—are equally effective (i.e.
the fixed effectiveness approach).
Let nuclear and diesel each have known and identical costs in the first period,
where a representative agent’s share (per capita) of the initial cost is g1 . Define the
uncertainty in period two with two possible states of nature; war with probability
π, and peace with probability (1 − π).
During war submarines steam for many hours at high speeds, while peacetime
requires little steaming.4 Operating cost does not vary between the two states of
nature for nuclear propulsion, but for diesel-electric technology more steaming
means more cost. Let g̃2 be the random variable describing per capita operating
costs in the second period.
Consider the expected utility, or well-being, of the representative citizen. In
each period the agent receives income, c, that can be consumed but not stored.
Assume that the agent’s benefit or “well-being” can be represented by a log-utility
function. Then the expected utility of a representative agent is

EU = ln(c − g1 ) + γ E[ln(c − g2 )], (17.4)

where E is the expectations operator and γ = 1/(α + 1). In (17.4) the natural log
operator represents a common method in economics to represent an individual’s
utility, or well-being. The first operator on the right hand side of the equation
Time discounting in war and peace 429
Table 17.4 Parameter definitions

Symbol Definition Value

π Probability of war 0.30


α Pure rate of time discount 0.00
g1 Agent share of initial cost 1.20
c Citizen income 2.00

computes utility in period one based on the certain sum of money the citizen
retains. The second piece of the equation does the same thing for period two,
except that now the amount the citizen retains is uncertain so the expectations
operator must be applied. The sum across both periods estimates the citizen’s
utility.
To operationalize this simple model, consider the parameters in Table 17.4.
In Table 17.5, we assume values for operating costs of both alternative propul-
sion systems, g2 , in both states of nature that can occur in period two (war and
peace). We then plug values from Tables 17.4 and 17.5 into Equation (17.4) to
obtain the expected utility of the two alternatives reported in Table 17.5.
From Table 17.5, diesel has the lowest expected cost in period two. Since
cost and utility in the first period are identical for nuclear and diesel, a conven-
tional CBA focused exclusively on minimizing expected costs of propulsion would
recommend diesel-electric propulsion, regardless of the discount rate applied.
However, nuclear propulsion is the choice that maximizes the expected utility,
or well-being, of the representative citizen.
As the example above demonstrates, failure to take risk into account can eas-
ily lead to situations where CBA rejects the superior project or program while
endorsing the inferior. The most popular methodology for correcting this problem
is to modify the discount rate, r , by adding an additional “risk premium.” Breeden
(1979) shows the value of the risk premium (RP) that maximizes expected utility
is given by
 
1
RPx = − × cov(u c , X), (17.5)
E(u c )

where X is the uncertain benefit or cost, and u c is the marginal utility of


consumption (Varian 1992, p. 379).

Table 17.5 Expected cost and utility of propulsion systems

Propulsion type g2 in peace g2 in war Expected cost Expected utility

Nuclear 0.3000 0.3000 0.3000 0.3075


Diesel 0.1500 0.6450 0.2985 0.2986
430 J. Hansen, J. Lipow
This approach to adjusting the discount rate is widely practiced in the pri-
vate sector, and Bazelon and Smetters (1999) strongly endorse the use of similar
techniques for public sector CBAs, concluding that “there is little rationale for
the government to discount future costs and benefits of any particular project or
program differently than the private market.”
So how does the private sector estimate the RP? The standard way to estimate
the RP is given by adjusting Equation (17.5) as follows:

RPx = βx [E(R) − r ], (17.6)

where R is a proxy for wealth or consumption (often, simply the return on a stock
index such as the S&P 500), r is Ramsay’s discount rate (often proxied by rates
on government bonds), and βx is the covariance of X and R.
Consistent with this approach, the US government’s guidelines for CBA (US
OMB 1992, 12) acknowledges that “the absolute variability of a risky outcome can
be much less significant than its correlation with other significant determinants of
social welfare, such as real national income.” Surprisingly, however, Circular A-94
does not then go on to mandate use of the formula given in (17.5).
A possible explanation is that the formula works fairly well when the covariance
of X and R can be estimated using historical data, but is much more difficult to
apply when trying to evaluate new investments with little or no historical data. In
such circumstances, a popular graduate finance textbook explains that since “you
cannot hope to estimate the relative risk of assets with any precision. . . examine
the project from a variety of angles and look for clues to its riskiness” (Brealey
and Myers 1991, p. 200).

17.5 Conclusion
For the government analyst, “looking for clues” can be quite a challenge. For
most military CBAs, “clues” may not even exist.5 So what can be done to account
for risk in a military CBA? While far from perfect—indeed far from adequate—
Hansen and Lipow’s (2013) proposed approach has the virtue of being relatively
easy to implement in the United States. The analyst first identifies the discount
rate mandated by OMB Circular A-94 (US OMB 1992). This value for r , which
is based on government bond prices, is given by Appendix C of A-94 and is
updated every year. The analyst then makes a qualitative judgment as to whether
each flow of cost or benefit associated with an investment project is positively or
negatively correlated with per capita consumption, and documents this judgment
by writing a paragraph that explains his or her reasoning. The analyst then adds
1 percentage point to the mandated value of r used to discount benefits and costs
that are positively correlated with consumption, while subtracting 1 percentage
point from the value of r used to discount flows that are negatively correlated with
consumption.
While such an approach is likely superior to current practice, the resultant esti-
mates of the appropriate discount rates will still not precisely reflect the true social
Time discounting in war and peace 431
rate of discount. Barring a methodological breakthrough, the accounting for time
and risk in discounting are likely to remain an Achilles’ heel of military CBA.
This reinforces the value of sensitivity analysis, including Monte Carlo simulation
and other modeling efforts discussed in earlier chapters, to generate a thorough
understanding of the sensitivity of military CBA to key parameters.

Notes
1 The formula to use in PVA is

FV
PV = .
(1 + r )n

Here PV is the present value of a future value FV in n years, discounted into present
terms at the rate of discount r .
2 For simplicity of exposition, assume the helicopters are fully depreciated at the end of
this six year life-cycle and have no scrap value.
3 Evans (2005), for example, estimates η for 20 OECD countries and derives estimates that
range from 1.08 to 1.82, with values ranging from 1.15 to 1.45 for the United States.
4 Naval pedants would be correct to point out that, formally, a diesel-electric submarine
does not use steam propulsion, and hence can’t “steam.”
5 What exactly is the covariance of the S&P and the costs associated with building and
operating an aircraft carrier? That is not an easy question. What is the covariance of
the S&P and the benefits (or flows of effectiveness) that we expect to receive from that
aircraft carrier? That isn’t a question at all—it is a riddle worthy of the Sphinx.

References
Bazelon, C., and Smetters, K. 1999. “Discounting Inside the Washington DC Beltway.” The
Journal of Economic Perspectives, 13(4): 213–228.
Breeden, D. T. 1979. “An Intertemporal Asset Pricing Model with Stochastic Consumption
and Investment Opportunities.” Journal of Financial Economics, 7(3): 265–296.
Brealey, R. A., and Myers, S. C. 1991. Principles of Corporate Finance (4th edition).
New York: McGraw-Hill.
Evans, D. J. 2005. “The Elasticity of Marginal Utility of Consumption: Estimates for 20
OECD Countries.” Fiscal Studies, 26(2): 197–224.
Hansen, J., and Lipow, J. 2013. “Accounting for Systematic Risk in Benefit–Cost Analysis:
A Practical Approach.” Journal of Benefit-Cost Analysis, 4(3): 361–373.
Heal, G. 2007. “Discounting: A Review of the Basic Economics.” The University of
Chicago Law Review, 74(1): 59–77.
Ramsey, F. P. 1928. “A Mathematical Theory of Saving.” The Economic Journal, 38(152):
543–559.
Varian, H. R. 1992. Microeconomic Analysis (3rd edition). New York: Norton.
US Office of Management and Budget. 1992. Guidelines and Discount Rates for Benefit–
Cost Analysis of Federal Programs, OMB Circular No. A-94 Revised.
Weitzman, M. L. 2001. “Gamma Discounting.” American Economic Review, 91(1):
260–271.
Index

absorption rate 169–70 analysis (COEA) 356, 365, 389, 392–3,


accountability 7, 48, 58, 99 396; cost and schedule growth 24; cost
acquisition 24–7, 46, 75–6, 113–14, 352, as an independent variable (CAIV) 86,
360 91, 100, 361; cost effectiveness 42,
activity-based costing 173, 182 78–80, 222–3, 229–34, 272–4, 405–6,
affordability 12–13, 27–8, 74–7, 82–6, 90, 409; cost estimating relationships
132, 346 10, 117, 124, 137; cost estimation 131,
air mobility 363–4, 392 136–40, 147, 158, 343, 409;
air transport 60, 381 cost reduction 228, 231, 234, 290, 306
allocative efficiency 4, 6, 39 cost–benefit analysis 4, 11–12, 97, 113,
alternatives 77–81, 87–95, 340–6 161, 264, 335, 348, 363, 424
analysis of alternatives (AoA) 32, 74, 98, cost-effectiveness analysis 6, 11–12,
114, 298, 338 78–80, 222, 263, 354, 401–3
attribute 90–2, 139–44, 152–9, 402 CP-140A Arcturus 12, 314–15, 319–21,
autoregressive distributed lag (ARDL) 323–6
approach to cointegration 189–90 critical region 320–5
cyber warfare 20
Bellman relation 318, 328, 329
benefits 49–50, 52–3, 56–7, 80–2, 98, data mining 138–40, 149
240–2, 264 decision tree 10, 138, 147–50, 159, 349
Box–Jenkins method 175, 185–7 defense budget 21–4, 48–9, 161–3, 166,
budgets 3, 6–10, 13, 45–9, 173–5, 237 174–80
Bulgaria 12, 335–9 defense outputs 43, 45, 49, 51, 60, 64–5
defense specific inflation 171–3, 182
C-17 363–71, 375–81, 391–3 development program 336, 350
C-5 364–6, 369–73, 375–81, 396–7, 399 disaster relief 53, 58, 59, 64, 409–14
Canada 43, 67–8, 163–8, 183, 314–15, 320 discount rate 8, 15, 307, 426–30
capability based planning 166–8 DRAGON 348, 359
Civil Reserve Air Fleet (CRAF) 369, 372, drift diffusion process 322
374, 376, 381–2, 397–9 dynamic programming 315, 318, 324,
cointegration 189–90 327
communication payload 402, 406, 408–9,
412, 415, 419–20 economic evaluation of alternatives 9, 74,
constrained optimization 77, 83, 85–6, 97, 77, 86, 338, 361, 399, 425
100, 130 effectiveness 10–14, 42, 80–94, 199–204,
constructed measure 206 206–10, 340–3, 352–3
cost 24–6, 113–15, 171, 173, 350; cost endogenous alternatives 87, 92, 344
analysis 113, 119–20, 123, 136, 197, equipment 5, 41, 43, 161, 180–1, 364–5
377; cost and operational effectiveness exogenous alternatives 87, 344
Index 433
financial crisis 3, 7, 16, 163, 165 Nash–Cournot specification 176
first passage time 315, 319–20, 322–3, 325, national security 21, 32, 64, 166, 171, 175
327 natural measure 206
fixed budget approach 87–8, 102, 345 nonlinear optimization 10, 138, 147, 159
fixed effectiveness approach 87–9, 103,
345, 425, 428 objectives 199–202, 215, 222, 402
FLIR 351–3, 355–60 objectives hierarchy 204, 206–7, 212, 217,
forecasting 130, 136, 161–2, 167, 172, 185, 232, 405–8
187–8, 289, 305–6 operational effectiveness 32, 61, 74–5,
free-riding 37–8, 40 338–9, 344
operations 6, 21–3, 166, 336, 345
GDP deflator 172, 177, 183 opportunity cost 41, 84, 87, 94–6, 308, 345
geometric random walk 315, 318, 320, 325, optimal replacement 314–16, 324
327
global terrorism 19–20, 64 parametrics 136–7
Greenfield–Persselin model 315–22, Pareto efficiency 4
324–6 peer competitors 21
planning programming budgeting
hierarchical cluster analysis 138–40, 144 execution (PPBE) 29–31, 99, 130, 166
Planning, Programming and Budgeting
immature technology 348, 359 System (PPBS) 6–7, 14, 29, 75–8, 104
infrastructure 4, 20, 40, 61, 181, portfolio optimization 300, 302, 304,
183, 401 306, 308
integrated risk management 301 principal-agent models 40–1
inter-program analysis 87, 94–5 procurement 25, 91, 181–2, 317, 321–4,
intra-program analysis 86–7, 344–5 341, 352–3
investment 3–5, 75, 78–80, 170, 298, production possibility frontier 168
336–8, 341, 361 productivity 45, 162, 182, 296
program alignment architecture (PAA)
Javelin 348–9, 351, 356–60 47, 66
property rights 39, 50–1
knowledge value added 289, 295 protection-adjusted life years (PALYS)
51, 68
life cycle cost 75, 222–3, 316, 329, 337, prototype 118, 315, 352
339, 341, 344, 425–6 proxy measure 206, 232, 407, 408
public choice 40–1, 69, 185
machine learning 138–9, 158 public goods 37–8, 49–51
maintenance 133, 171, 181, 199–200, 314,
326, 341, 350 Quadrennial Defense Review 29, 65, 393
marginal analysis 66, 77, 94, 355, 406 quality-adjusted life years (QALYS) 51, 68
military expenditures 13, 66, 176, 178, quantitative analysis 4
185, 190
military production function 41–3, 166 rank sum method 215, 412
modified budget approach 87, 92–3, 345 real options 11–12, 297–300, 305–9, 338,
modified effectiveness approach 77, 87, 349–51, 357
93–4, 345, 361 regional conflicts 21, 54, 65
monetary valuation 350 regression analysis 117, 125, 137–8, 149,
Monte Carlo simulation 293–5, 297, 248
305–8, 310, 377, 395 rent-seeking 7
multi-criteria decision-making (MCDM) 6, requirements generation 75–7, 97, 120,
82, 97, 360 300, 335
multi-objective decision-making (MODM) resource accounting and budgeting (RAB)
6, 79, 82 47–8
434 Index
return on investment (ROI) 11, 289–90, uncertainty 8, 119, 221, 289, 293, 298, 424
295–7, 300–3 Univariate Autoregressive Integrated
risk management 45, 127, 170, 301, 305–6 Moving Average (ARIMA) Model
174–5, 188
sensitivity 31, 217, 221, 408; sensitivity unmanned aerial vehicle (UAV) 39, 401–5,
analysis 293, 307, 345, 347, 375, 384, 416
394, 424 utility function 82–3, 88, 90–2, 318, 428
social welfare 5, 177, 198, 249, 430
stationarity 175, 185 value for money 56, 74, 170, 355
strategies 132, 163, 166, 170, 185, 349 value function 80, 83, 86, 103, 200–1, 207,
structural unexpended rate 170 234–5
value of statistical life 11, 237–9, 241,
supplies 46, 133, 337
243–4, 246–9
system-of-systems 24, 27
systems analysis 5–7, 14, 81, 87
weapon system requirements 26, 78
weapons of mass destruction (WMD)
technical efficiency 39 20, 30
technological maturity 309, 359–60 weights 70, 102, 142–4, 158, 212–19,
technology development 350–1, 354, 414–15, 417; importance weight 201,
356–9 212–19, 221–2, 226, 230, 232, 234;
tradeoffs 11, 31–2, 119, 168, 250, 264 swing weighting 215–16; tradeoff
training 47, 50, 61, 113, 173, 240 weight 212, 270, 276, 402, 406, 410,
transparency 4, 7, 13, 99 412, 417–18

You might also like