0% found this document useful (0 votes)
34 views393 pages

Tapan K. Sengupta, Swagata Bhaumik - DNS of Wall-Bounded Turbulent Flows-Springer Singapore (2019)

The document discusses Direct Numerical Simulation (DNS) of wall-bounded turbulent flows, focusing on the transition from receptivity to fully developed turbulence over flat plates. It emphasizes the importance of spatio-temporal analysis in understanding flow instability and turbulence, while also addressing historical and theoretical aspects of fluid dynamics. The authors aim to provide a comprehensive account of their research, making it accessible for both undergraduate students and researchers in hydrodynamics.

Uploaded by

Pung Kang Qin
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
0% found this document useful (0 votes)
34 views393 pages

Tapan K. Sengupta, Swagata Bhaumik - DNS of Wall-Bounded Turbulent Flows-Springer Singapore (2019)

The document discusses Direct Numerical Simulation (DNS) of wall-bounded turbulent flows, focusing on the transition from receptivity to fully developed turbulence over flat plates. It emphasizes the importance of spatio-temporal analysis in understanding flow instability and turbulence, while also addressing historical and theoretical aspects of fluid dynamics. The authors aim to provide a comprehensive account of their research, making it accessible for both undergraduate students and researchers in hydrodynamics.

Uploaded by

Pung Kang Qin
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/ 393

Tapan K.

Sengupta · Swagata Bhaumik

DNS of Wall-
Bounded
Turbulent
Flows
A First Principle Approach
DNS of Wall-Bounded Turbulent Flows
Tapan K. Sengupta Swagata Bhaumik

DNS of Wall-Bounded
Turbulent Flows
A First Principle Approach

123
Tapan K. Sengupta Swagata Bhaumik
Department of Aerospace Engineering Department of Mechanical Engineering
Indian Institute of Technology Kanpur Indian Institute of Technology Jammu
Kanpur, Uttar Pradesh Jammu, Jammu & Kashmir
India India

ISBN 978-981-13-0037-0 ISBN 978-981-13-0038-7 (eBook)


https://doi.org/10.1007/978-981-13-0038-7
Library of Congress Control Number: 2018940403

© Springer Nature Singapore Pte Ltd. 2019


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part
of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations,
recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission
or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar
methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specific statement, that such names are exempt from
the relevant protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publisher nor the
authors or the editors give a warranty, express or implied, with respect to the material contained herein or
for any errors or omissions that may have been made. The publisher remains neutral with regard to
jurisdictional claims in published maps and institutional affiliations.

Printed on acid-free paper

This Springer imprint is published by the registered company Springer Nature Singapore Pte Ltd.
part of Springer Nature
The registered company address is: 152 Beach Road, #21-01/04 Gateway East, Singapore 189721,
Singapore
Preface

A long habit of not thinking a thing is wrong gives it a


superficial appearance of being right. Reason obeys itself,
and ignorance submits to whatever is dictated to it.
Tom Paine

There are many books and monographs on instability and transition, including
one by one of the authors of this monograph. Despite these books, this monograph
is written with a specific purpose of presenting a detailed account of what consti-
tutes direct numerical simulation (DNS) of flow from receptivity stage to fully
developed turbulent flow over flat plate, while simulating the classical transition
experiments for zero pressure gradient boundary layer. This has been made possible
with the work of the authors’ colleagues and students, proposing new theoretical
and computational results, which establish a deterministic route to turbulence, as in
the classical experimental efforts [3, 5, 7].
The Navier–Stokes equation governs fluid flows and Reynolds explained that the
exact solution of the Navier–Stokes equation, even when it exists, is unable to
maintain its stability with respect to omnipresent small disturbances in the flow. An
equilibrium solution ensures satisfaction of conservation principles of mass,
momentum, and energy balance, yet the flow in the famous pipe flow experiment of
Reynolds disintegrated into sinuous motion, eventually leading to turbulent flow.
This onset has been attributed to the instability of flows ever since. However, even
today the transition to turbulence in pipe flow is not completely understood.
Another canonical problem which attracted the attention of researchers in the
beginning is the flow over a flat plate. In studying instability of this flow field,
Rayleigh developed his stability equation, but it was not amenable to study con-
vective disturbances, which is the signature of transition for zero pressure gradient
boundary layer flow. By the beginning of twentieth century, it became evident that
the instability problem is intractable without the inclusion of the effects of second
derivative of velocity profile and viscous actions. These latter effects have been
termed as resistive instability, which is in contradiction to the assumption made by

v
vi Preface

early pioneers, who mistakenly considered the action of viscous forces to attenuate
disturbances and was considered not central to the study of instability.
This was the reason for Orr and Sommerfeld to independently come up with the
famous Orr–Sommerfeld equation, which was investigated by Tietjens [14] and
Heisenberg [4]. However, it was the definitive attempt by Tollmien that paved the
understanding the concept of critical Reynolds number [15]. Following this lead,
Schlichting also studied the instability of zero pressure gradient boundary layer,
while making some implicit assumption which connected the temporal and spatial
growth of disturbances, which has since been addressed by invoking the concept of
group velocity [8]. One of the central themes of Tollmien and Schlichting’s work
was to show the presence of growing waves in limited part of the flow, and this is
now called the Tollmien–Schlichting (TS) waves. This eigenvalue analysis is
performed by neglecting the growth of boundary layer. We would like to mention
that studying flow instability by temporal and spatial theory and artificially patching
the two together is fraught with danger. This has been demonstrated by the authors
and colleagues that there is really no need to make this artificial demarcation into
spatial and temporal routes. The best course of action is to adopt the spatio-temporal
route to avoid this ambiguity. While studying instability of mixed convection
boundary layer, authors came across a perplexing situation, when a singularly
cooled plate demonstrated simultaneous presence of temporal and spatial instability
[10]. The problem was resolved because of the availability of DNS results obtained
by high-accuracy method, which showed the flow to follow primarily temporal
growth, and not the spatial route (the growth should be strictly stated to be via
spatio-temporal route). This emphasizes the need to perform spatio-temporal
analysis, which has been practiced and yielded new insights, which form the
contents of this treatise.
However, despite Taylor demonstrating the strength of linear stability of flow
inside concentric cylinder [12], TS wave could not be detected experimentally. This
created doubts about the existence of TS wave and, in turn, on the relevance of
viscous linear theory. This problem was resolved by Dryden’s group, by demon-
strating for the first time the existence of TS wave by vibrating a ribbon inside the
boundary layer, with the results announced after the Second World War [3, 5]. This
experimental approach may be termed as frequency response of the fluid dynamical
system. It would be appropriate to note that exciting the flow at any frequency will
not lead to flow instability, as the vibrating diaphragm experiment of Taylor [13] at
2 Hz demonstrated. The frequencies to be excited were present in the works of
Tollmien and Schlichting [8, 15], but Taylor’s experimental vibration was at too
low a frequency [13]. This aspect of how to experimentally study eigenvalue
problem is not clearly understood. In a recent work, researchers have explained the
relationship between frequency response and instability for an associated problem
of bifurcation [6].
Viewing fluid flows as dynamical systems, it is thus unwise to ignore the
receptivity aspect. Receptivity was emphasized by Schubauer and Skramstad in
reporting their classic experimental results [7]. Thereafter, reported TS waves as
Preface vii

early marker of flow instability was readily embraced by the research community to
conclude prematurely that TS waves are the cause for flow transition. It is worth-
while remembering that vortical excitation validated instability theory, while the
acoustic excitation did not! Also, instability theory does not require any specific
excitation, except in prescribing the qualitative nature of boundary conditions. The
main feature of this monograph is to relate receptivity with flow instability and
show how different routes of excitation lead to different types of disturbance
evolution. This has been achieved here primarily by DNS to explain theoretical
aspects of the flow.
Viscous instability results also suffered credibility due to the use of a parallel
flow assumption for the equilibrium flow, along with linearization. This criticism
was partly silenced by experimental verification provided [7], but whether TS waves
can cause transition or not was not known for a long time, till the definitive routes of
transition were identified for 2D zero pressure gradient flow [9]. It has been shown
by the present authors that when the Navier–Stokes equation is computed in 2D
framework to mimic the experiment [7] for moderate- to high-frequency wall
excitation, created TS wave packet remains passive, while the spatio-temporal
wave-front (STWF) causes transition. The STWF is the first wave front that is
created upon starting off the excitation, as in an experiment, which has the property
to regenerate other STWF in its wake [9]. It is interesting to note that STWF was
originally proposed in research on electromagnetic wave propagation [2]. In fluid
dynamics, the major success in tracking STWF from the spatio-temporal dynamics
of Orr–Sommerfeld equation was reported [11], following the developed technique
of Bromwich contour integral method. While these have been obtained for the
frequency response of disturbance evolution, corresponding role of STWF during
impulse response has been reported only recently [1], which also shows that the
same physical mechanism of STWF explains the features of geophysical phenom-
ena like tsunami and rogue waves.
This book has evolved into an account of the research interests of the authors
over the years. Efforts are made to keep the treatment at an elementary level
requiring rudimentary knowledge of calculus, Laplace–Fourier transform, and
complex analysis, which should be equally amenable to undergraduate students, as
well as serious researchers in the field of hydrodynamics and mixed convection.
This monograph shows to readers that without good computing, this subject will be
poorer in linking spatio-temporal growth, instability at low frequencies, and the
actual physical mechanisms of transition. In providing computational results from
receptivity to a fully developed turbulent stage of 2D and 3D disturbance flows, we
also definitely provide the basis of experimental approach for transition to turbu-
lence. In doing so, we state that turbulence is deterministic in its origin, as it is
implicitly assumed in transition experiments [3, 5, 7]. Trained with high-accuracy
computing methods, users of this monograph will be able to further contribute in
this rich field of nonlinear dynamics.
Thus, the emphasis of the monograph is on DNS of transitional and turbulent
flows, basis of which are laid out in Chap. 2. The applications of DNS technique are
used to explain receptivity and instability in hydrodynamics and mixed convection
viii Preface

flows in Chap. 3. The power of DNS is demonstrated by using results to explain


vortex-induced instability in a nonlinear framework to explain receptivity and
instability in Chap. 4. This is done by developed disturbance enstrophy transport
equation [16, 17]. This provides a nonlinear framework based on Navier–Stokes
equation without making any assumption on the equilibrium flow. Another non-
linear framework to study instability in this chapter is by using enstrophy-based
proper orthogonal decomposition. The last two chapters are to demonstrate the
power and accuracy presented by the DNS techniques described in earlier chapters
to show that STWF is the precursors of transition to turbulence, for both 2D and 3D
disturbance flow simulations. The resulting turbulent flows are as have been created
experimentally [3, 5, 7]. The transition to turbulence is shown to be experimentally
obtained for the integral properties and presented in textbooks. In closing this
discussion, we note that the subject has matured very rapidly in recent times with
the advent of very high-accuracy methods of computing, which will lead to an
explosive growth of activities in the subject field. The contents of the monograph
can be adopted as text for a high-level course on DNS and transition to turbulence.
The contents of this monograph are based on the doctoral thesis work of one
of the authors and the other graduate students who have been associated with the
first author over the last three decades. The contents have been specifically enriched
by faculty colleagues in NUS, Singapore, specifically by Profs. Y. T. Chew,
T. T. Lim, B. C. Khoo, K. S. Yeo, and S. Chang and Prof. Pierre Sagaut (UPMC),
Prof. Julio Soria (Monash University), Prof. W. Schneider (TUW, Vienna), Prof.
K. R. Sreenivasan (ICTP and NYU), Prof. M. Klocker (University of Stuttgart),
Dr. J. M. Kendall (Jet Propulsion Lab, USA), Dr. B. R. Noack (Pprime, University
of Poitier, France), Prof. M. Deville and Prof. F. Gallaire (EPFL, Switzerland),
Prof. S. Girimaji (TAMU, USA), Prof. P. J. Strykowski (University of Minnesota),
and Prof. A. Tumin (University of Arizona). Much of our work on receptivity has
been influenced by the works of Prof. H. Fasel (University of Arizona). The first
author acknowledges the influence of Prof. M. Gaster and Prof. D. G. Crighton
(University of Cambridge) for encouraging him to develop a theory of receptivity.
Many of the contributors of this book are now faculty colleagues in different parts.
It is our pleasant duty to acknowledge the contributions by Sandeep Nijhawan,
Manish Ballav, A. P. Sinha, Vivek Rana, Manoj T. Nair, K. Venkatasubbaiah,
A. K. Rao, Manojit Chattopadhyay, Z. Y. Wang, A. Dipankar, S. Sarkar, S. De,
Yogesh Bhumkar, Vijay Vedula, Manoj Rajpoot, S. Unnikrishnan, R. Bose,
N. A. Sreejith, Ashish Bhole, and Soumyo Sengupta. In a very recent development,
the authors have been looking at “exact” nonlinear theories of instability for
incompressible flows, based on disturbance mechanical energy and disturbance
enstrophy, with the former covered in the text here. However, we refrain from
providing the details on the theory based on disturbance enstrophy and is currently
being probed for different types of flow fields and would be described elsewhere.
The authors acknowledge Aditi Sengupta and V. K. Suman for helping with
the development of this instability theory based on disturbance enstrophy [17].
Preface ix

The authors would like to acknowledge the competent help provided in typing this
material by Mrs. Baby Gaur and Mrs. Shashi Shukla. Specifically, Mrs. Gaur takes
exceptional care in typing the text and preparing figures at all times. We also
acknowledge various helps provided to us by Mr. Mukesh Kumar.

Kanpur, India Tapan K. Sengupta


Jammu & Kashmir, India Swagata Bhaumik

References

1. Bhaumik, S., & Sengupta, T. K. (2017). Impulse response and spatio-temporal wave-packets:
The common feature of rogue waves, tsunami and transition to turbulence. Physics of Fluids,
29, 124103.
2. Brillouin, L. (1960). Wave propagation and group velocity. New York: Academic Press.
3. Dryden, H. L. (1955). Fifty years of boundary-layer theory and experiment. Science, 121
(3142), 375–380.
4. Heisenberg, W. (1924). Üeber stabilität und turbulenz von flüssigkeitsströmen. Annalen der
Physik, 74, 577–627.
5. Klebanoff, P. S., Tidstrom, K. D., & Sargent, L. M. (1962). The three-dimensional nature of
boundary-layer instability. Journal of Fluid Mechanics, 12, 1–34.
6. Lestandi, L., Bhaumik, S., Avatar, G. R. K. C., Mejdi, A., & Sengupta, T. K. (2018). Multiple
Hopf bifurcations and flow dynamics inside a 2D singular lid driven cavity. Computers &
Fluids.
7. Schubauer, G. B., & Skramstad, H. K. (1947). Laminar boundary layer oscillations and the
stability of laminar flow. Journal of the Aeronautical Sciences, 14(2), 69–78.
8. Schlichting, H. (1933). Zur entstehung der turbulenz bei der plattenströmung. Nachrichten
von der Gesellschaft der Wissenschaften zu Göttingen, Mathematisch-Physikalische Klasse,
181–208.
9. Sengupta, T. K., & Bhaumik, S. (2011). Onset of turbulence from the receptivity stage of fluid
flows. Physical Review Letters, 107(15), 154501.
10. Sengupta, T. K., Bhaumik, S., & Bose, R. (2013). Direct numerical simulation of transitional
mixed convection flows: Viscous and inviscid instability mechanisms. Physics of Fluids, 25,
094102.
11. Sengupta, T. K., Rao, A. K., & Venkatasubbaiah, K. (2006). Spatio–temporal growing wave
fronts in spatially stable boundary layers. Physical Review Letters, 96(22), 224504.
12. Taylor, G. I. (1923). Stability of a viscous liquid contained between two rotating cylinders.
Philosophical Transactions of the Royal Society (London), A223, 289–343.
13. Taylor, G. I. (1939). Some recent developments in the study of turbulence. In J. P. Den Hartog
& H. Peters (Eds.), Proceedings of the Vth International Conference on Applied Mechanics.
14. Tietjens, O. (1925). Beiträge zur enstehung der turbulenz. ZAMM, 5, 200–217.
15. Tollmien, W. (1931). Über die enstehung der turbulenz. I, English translation. NACA TM 609.
16. Sengupta, T. K., Sharma, N. & Sengupta, A. (2018). Non-linear instability analysis of the
two-dimensional Navier-Stokes equation: The Taylor-Green vortex problem. Physics of
Fluids (Accepted).
17. Sengupta, A., Suman, V. K., Sengupta, T. K. & Bhaumik, S. (2018). An enstrophy-based
linear and non-linear receptivity theory. Physics of Fluids (Accepted).
Contents

1 DNS of Wall-Bounded Turbulent Flow: An Introduction . ........ 1


1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ........ 1
1.2 Why Deterministic Study Is More Relevant than
Stochastic Approaches? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2.1 Historic Developments . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Present State of Art in the Field . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4 Different Transition Routes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.5 Role of Equilibrium Flow in DNS . . . . . . . . . . . . . . . . . . . . . . . . 6
1.6 What Is Instability? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.7 Temporal and Spatial Instability . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.8 Some Instability Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.8.1 Kelvin–Helmholtz Instability . . . . . . . . . . . . . . . . . . . . . 10
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2 DNS of Navier–Stokes Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.1 Fluid Dynamical Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.1.1 Equation of Continuity . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.1.2 Momentum Conservation Equation . . . . . . . . . . . . . . . . . 19
2.1.3 Energy Conservation Equation . . . . . . . . . . . . . . . . . . . . 22
2.1.4 Alternate Forms of the Energy Equation . . . . . . . . . . . . . 23
2.1.5 Vorticity Transport Equation for Incompressible
Flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ... 23
2.1.6 Derived Variable Formulation for 2D Incompressible
Navier–Stokes Equation . . . . . . . . . . . . . . . . . . . . . . ... 27
2.2 Spatial and Temporal Scales for Transitional
and Turbulent Flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.3 Numerical Methods for Developing DNS/LES . . . . . . . . . . . . . . . 31
2.3.1 Waves – Building Blocks of a Disturbance Field . . . . . . . 31
2.3.2 Resolution of Spatial Discretization . . . . . . . . . . . . . . . . . 34
2.3.3 High-Accuracy Compact Schemes for Evaluation of
First Derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . ... 39

xi
xii Contents

2.3.4 Boundary Closure Schemes for Compact Schemes to


Evaluate First Derivative . . . . . . . . . . . . . . . . . . . . . . .. 42
2.3.5 Compact Schemes for Second Derivative Evaluation . . .. 44
2.3.6 Compact Schemes for Interpolation and First Derivative
Evaluation in Staggered Grids . . . . . . . . . . . . . . . . . . . .. 51
2.3.7 Discretization of Self-Adjoint Terms . . . . . . . . . . . . . . .. 56
2.3.8 Computing Methods for Unsteady Flows: Dispersion
Relation Preserving (DRP) Methods . . . . . . . . . . . . . . .. 61
2.3.9 Numerical Amplification Factor for 1D Convection
Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 61
2.3.10 Quantification of Dispersion Error and Error
Propagation Equation . . . . . . . . . . . . . . . . . . . . . . . . . .. 64
2.3.11 Dispersion Relation Preserving Schemes . . . . . . . . . . . .. 67
2.4 Design of DRP Schemes for DNS/LES . . . . . . . . . . . . . . . . . . .. 75
2.5 Numerical Filtering: Error Control via Stabilization
and Dealiasing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
2.5.1 Use of Filters for DNS and LES . . . . . . . . . . . . . . . . . . . 76
2.5.2 Use of 1D Filters at the Outflow . . . . . . . . . . . . . . . . . . . 83
2.5.3 Two-Dimensional Higher Order Filters . . . . . . . . . . . . . . 86
2.5.4 Adaptive 2D Filter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
2.6 Role of Solenoidality Errors in Velocity-Vorticity
Formulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 95
2.6.1 Computation of Physically Unstable Flows in Square
Lid Driven Cavity . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 97
2.6.2 Solution of 3D Cubic Lid Driven Cavity Using
Velocity-Vorticity Formulation and Role of
Solenoidality Error for Vorticity . . . . . . . . . . . . . . . . . . . 102
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
3 Receptivity and Instability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
3.1 Linear Stability/Receptivity Theories: Classical Approaches
and Signal Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
3.1.1 The Equilibrium Flow Equation . . . . . . . . . . . . . . . . . . . 122
3.2 Linear Stability Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
3.3 Linear Receptivity Analysis of Parallel Boundary Layer . . . . . . . 128
3.3.1 Receptivity of Blasius Boundary Layer . . . . . . . . . . . . . 131
3.3.2 Near-Field of the Localized Excitation:
The Local Solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
3.3.3 The Spatio-Temporal Wave-Front (STWF) . . . . . . . . . . . 136
3.4 Stability and Transition of Mixed Convection Flows . . . . . . . . . . 137
3.5 Governing Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
Contents xiii

3.6 Equilibrium Boundary Layer Flows . . . . . . . . . . . . . . . . . . . . . . . 140


3.6.1 Schneider’s Similarity Solution . . . . . . . . . . . . . . . . . . . . 141
3.6.2 Ambiguities of Spatial and Temporal Linear Theories:
Example of Mixed Convection Problem . . . . . . . . . . . . . 148
3.7 Equilibrium Solution for Mixed Convection Flows:
Isothermal Wall Case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
3.7.1 Governing Equation for Flow over Isothermal Wall . . . . . 151
3.7.2 Mixed Convection Governing Equations and Boundary
Conditions for DNS . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
3.7.3 Linear Viscous Instability: Spatial and Temporal
Routes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
3.7.4 Linear Spatial Viscous Theory for Mixed Convection
Flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
3.7.5 Linear Temporal Viscous Theory for Mixed Convection
Flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
3.8 Receptivity of Mixed Convection Flows . . . . . . . . . . . . . . . . . . . 168
3.8.1 Receptivity of Cold Adiabatic Flat Plate Cases . . . . . . . . 169
3.8.2 Receptivity of Hot Isothermal Wedge Case . . . . . . . . . . . 172
3.8.3 DNS of Instability of Mixed Convection Flows:
New Theorems of Instability . . . . . . . . . . . . . . . . . . . . . . 172
3.9 Nonlinear and Nonparallel Effects: Receptivity by DNS . . . . . . . . 185
3.9.1 Computational Domain and Grid Resolution . . . . . . . . . . 187
3.9.2 The Equilibrium Flow . . . . . . . . . . . . . . . . . . . . . . . . . . 187
3.9.3 Typical Morphology of the Disturbance Field . . . . . . . . . 189
3.9.4 Receptivity to SBS Excitation: Nonlinear and
Nonparallel Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
3.9.5 Effects of Nonlinearity: Role of Amplitude
of Excitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
3.9.6 Comparison of Results for a1 ¼ 0.002, 0.005
and 0.01 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
3.9.7 Results for a1 ¼ 0:1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199
3.9.8 Wall-Normal Variation of ud and vd . . . . . . . . . . . . . . . . 204
3.9.9 Further Evidences of Nonparallel, Nonlinear Effects
and Bypass Transition . . . . . . . . . . . . . . . . . . . . . . . . . . 210
3.9.10 Effects of the Location of Exciter . . . . . . . . . . . . . . . . . . 213
3.9.11 Effects of Frequency of Excitation . . . . . . . . . . . . . . . . . 216
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
xiv Contents

4 Nonlinear Theoretical and Computational Analysis


of Fluid Flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
4.1 Nonlinear Instability Theory Based on Mechanical Energy . . . . . . 225
4.2 Vortex-Induced Instability: Application of a Nonlinear
Theory for Total Mechanical Energy . . . . . . . . . . . . . . . . . . . . . . 227
4.2.1 An Experimental Observation of Vortex-Induced
Instability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
4.3 Numerical Simulation of Vortex-Induced Instability . . . . . . . . . . . 229
4.3.1 Velocity-Vorticity Formulations for 2D Flows . . . . . . . . . 230
4.3.2 Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
4.3.3 Numerical Methods and Grids . . . . . . . . . . . . . . . . . . . . 232
4.3.4 Grid Generation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234
4.3.5 Numerical Results and Discussion . . . . . . . . . . . . . . . . . . 235
4.3.6 The Instability Mechanism . . . . . . . . . . . . . . . . . . . . . . . 238
4.4 Enstrophy Transport Equation: A New Approach to Nonlinear
Receptivity Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240
4.4.1 Enstrophy Transport Equation . . . . . . . . . . . . . . . . . . . . . 243
4.4.2 Enstrophy Cascade for General Inhomogeneous
Flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
4.5 Theory of Instability for Enstrophy: Creation of Rotationality . . . . 248
4.6 Proper Orthogonal Decomposition . . . . . . . . . . . . . . . . . . . . . . . . 252
4.6.1 Some Useful Mathematical Relations . . . . . . . . . . . . . . . 253
4.6.2 Method of Snapshots . . . . . . . . . . . . . . . . . . . . . . . . . . . 255
4.6.3 TS Wave Instability over Zero-Pressure Gradient
Boundary Layer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271
5 Dynamics of the Spatio-Temporal Wave-Front
in 2D Framework . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275
5.2 From Linear Theory to Turbulence via Deterministic Routes . . . . . 276
5.2.1 Governing Equations, Numerical Schemes
and Simulation Parameters . . . . . . . . . . . . . . . . . . . . . . . 277
5.3 Small Amplitude Disturbance at Moderate Frequency . . . . . . . . . . 280
5.3.1 Growth and Speed of STWF . . . . . . . . . . . . . . . . . . . . . 282
5.3.2 Spatial Spectrum and Scale Selection of STWF . . . . . . . . 285
5.3.3 Wall-Normal Variation of the Disturbance Velocity . . . . . 286
5.4 Dynamics of STWF for High Amplitude Wall-Excitation and
Turbulent Spot Regeneration Mechanism . . . . . . . . . . . . . . . . . . . 288
Contents xv

5.5 Low Frequency Excitation: Interaction of Near-Field Solution


and Primary STWF . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 292
5.5.1 Low Frequency Excitation: Dominant Role
of the Near-Field Solution . . . . . . . . . . . . . . . . . . . . . . . 296
5.6 Dynamics of the STWF for Excited Flow Over an Airfoil . . . . . . 300
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 304
6 3D Routes of Transition to Turbulence by STWF . . . . . . . . . . . . . . 307
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 307
6.1.1 Governing Equations and Numerical Methods . . . . . . . . . 307
6.1.2 Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . 310
6.1.3 Initial Condition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 311
6.1.4 Grid Generation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 312
6.1.5 Numerical Method and Solution Technique . . . . . . . . . . . 313
6.2 Gaussian Circular Patch (GCP) Excitation . . . . . . . . . . . . . . . . . . 314
6.3 Spanwise Modulated (SM) Excitation . . . . . . . . . . . . . . . . . . . . . 327
6.4 Routes of Flow Transition: K- and H-Type Routes . . . . . . . . . . . . 334
6.5 Formation of Turbulent Spots and Fully Developed
Turbulent Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 337
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 343
Appendix A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 347
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 353
Symbols Description

Chapter 1
f Function in physical plane
F Bilateral Laplace transform of f
k Wavenumber
p Pressure
Re Reynolds number
Recr Critical Reynolds number
Retr Transitional Reynolds number
T Temperature
U1 ; U2 Velocity in Kelvin-Helmholtz instability problem
Uj Streamwise velocity
v Specific volume
x; y; t External streamline fixed coordinates
zs Interface displacement
a Wavenumber in x-direction
b Wavenumber in y-direction
d Displacement thickness
e Small parameter
g; ^
g Interface displacements
c Wavenumber vector inclination with x-axis
m Kinematic viscosity

x Circular frequency
/j Velocity potentials
q1 ; q2 Density in Kelvin-Helmholtz instability problem

xvii
xviii Symbols Description

Chapter 2
a1 ; a2 Wave amplitude
Aj Quotient for RK method
c physical phase speed
cN Numerical phase speed
Dx Divergence of vorticity
Dv Divergence of velocity field
e Internal energy per unit mass
Eðk; eÞ Energy spectrum
Fb Turbulence burst frequency
G Spectral amplification factor
H Shape factor
h Grid-spacing
h1 ; h2 Scale factors of transformation
k Complex wavenumber
kc Cutoff wavenumber
l Landau coefficient
Nc CFL number
pm Mechanical pressure
Re Reynolds number
TF Transfer function
uj Velocity components
u; v; w Cartesian velocity components
Uc Convection velocity
Vg Group velocity
VgN Numerical group velocity
xj Coordinates
b0 Fixed frequency
b1 ; b2 Frequency bandwidth of excitation
d Displacement thickness
dij Kronecker delta function
gK Kolmogorov length scale
C-form Error analysis by space-time discretization
j Thermal conductivity
k Bulk viscosity
k Wavelength
l Molecular viscosity
r Gradient operator
x Circular frequency
Xf Frequency ranges in temporal scale
U0 Dissipation function
P-form Error analysis by spatial discretization
W Stream function
Symbols Description xix

q Density
R Mean strain
r Stress
sij Stress tensor
! Velocity vector
V
n; g Transformed coordinates

Chapter 3
A; Am Amplitude of wall excitation
Br Bromwich contour
c Complex conjugate of phase speed
ci Temporal growth rates
cr Phase speed, real part
Gr Grashof number
F Excitation frequency (viscous scale)
HðtÞ Heaviside function
K Buoyancy parameter
Ki ; Ka Isothermal and adiabatic, buoyancy parameter
L Length
Pr Prandtl number
q Instantaneous quantity
Ri Richardson number
vd Wall-normal excitation velocity
Uc Convective velocity
ud Streamwise disturbance velocity
Ue Boundary layer edge velocity
Ue ðXÞ Non-uniform edge velocity
xd Disturbance vorticity
wex Exciter width
X Streamwise distance
a1 Amplitude control parameter of wall exciter
x0j Circular frequency
ar ; ai Wavenumber, real and imaginary part
b Complex frequency
b0 Fixed frequency
ðbÞcr Critical circular frequency
Red Reynolds number based on displacement thickness
ðRed Þcr Critical Reynolds number
dð~xÞ Dirac-delta function
d Displacement thickness
dTL Temperature scale
DT Temperature scale
C Contour
xx Symbols Description

CF Fjortoft integral, Theorem II


r2 Laplacian operator
0
x0 ; x Circular frequency
/ Amplitude of v-velocity
Uv Viscous dissipation
w; x Stream function, vorticity
2 Pressure scale
qU1
h Momentum thickness
H Temperature
H0 Wall-normal derivative of temperature
Hw Wall temperature
U1 ; T1 Free stream temperature, velocity denoted
n; f Transformed orthogonal coordinate

Chapter 4
am ; aj Amplitude functions
d Core diameter of convecting vortex
E Total mechanical energy
Ed Disturbance mechanical energy
H Constant height of convecting vortex
R Regular POD modes
Red Reynolds number
Rij Cross-correlation
u; v; w Velocity components
u0 ; v 0 Disturbance velocity components
Uc Convective speed
ud Disturbance velocity components
! Velocity–vorticity vectors
V ;x
ðbÞL Circular frequency
bx ; by Grid stretching parameter
DT Non-dimensional frequency
C Convecting vortex strength
ej Normalization factor in POD
e Dissipation function
m Kinematic viscosity
^
X Average enstrophy
X1 ; X2 ; X3 Higher moments of enstrophy
Xd Disturbance enstrophy
/i Deterministic vectors for POD
/j POD eigenfunctions
w; x Stream function, vorticity
n; f Transformed plane coordinates
Symbols Description xxi

Chapter 5
F Non-dimensional excitation frequency
k Wavenumber
km Dominant wavenumber
m Falkner–Skan pressure gradient parameter
u Streamwise component of velocity
Uc Convective speed of disturbance at outflow
ud Streamwise disturbance velocity amplitude
udm Maximum amplitude of ud
a1 Amplitude control function of excitation
f0 Similarity functions
w; x Stream function and vorticity
wwp Perturbation stream function
U1 Free stream speed

Chapter 6
Cf Skin friction coefficient
Dv Divergence of velocity
Ed Disturbance mechanical energy
Ff Non-dimensional frequency
h1 ; h2 ; h3 Scale factors of transformation
kx ; k1 Streamwise wavenumber
x1 ; x2 Wall-excitation stretch
ud , vd , wd Disturbance velocity components
bx ; by Grid stretching parameters
d and h Displacement and momentum thickness
k2 Second eigenvalue of the velocity strain matrix
! Velocity vector
V
! Vorticity vector
X
Xn ; Xg and Xf Vorticity components
ðn; g; fÞ Transformed plane coordinates
List of Figures

Fig. 1.1 Kelvin–Helmholtz instability at the interface of two flowing


fluids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 11
Fig. 2.1 Ranges of temporal scales excited in flows of engineering
interest for different speeds and the corresponding frequency
ranges (Xf ). Also, indicated is the band for mean frequencies
of large eddies in high Reynolds number flows . . . . . . . . . . . .. 32
Fig. 2.2 Schematic of grid-spacing. All the nodes are equispaced
and the distance between adjacent grid-points is h . . . . . . . . . .. 35
Fig. 2.3 a ðkeq =kÞreal and b ðkeq =kÞimg plotted for CD2 , CD4 , CD6 , UD1
and UD2 spatial discretization schemes as a function of kh . . .. 37
Fig. 2.4 ðkeq =kÞreal plotted for the CD22 - and CD22a -schemes as a
function of kh. These schemes are given by Eqs. (2.70) and
(2.74), respectively . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 39
Fig. 2.5 ðkeq =kÞ plotted as a function of kh for the interior stencil of
OUCS3 (Eq. (2.83)) and the eighth order compact schemes
proposed by Lele (Eq. (2.81)). Here, we consider only the
central stencil for periodic problems. For OUCS3, we take the
central stencil, i.e., g ¼ 0 in Eq. (2.83) . . . . . . . . . . . . . . . . . . .. 42
Fig. 2.6 a Real and b imaginary part of ðkeq =kÞ plotted for OUCS3
scheme with explicit boundary closure schemes given
by Eqs. (2.92) and (2.93). For the interior OUCS3 stencil here
we consider g ¼ 0 in Eq. (2.83) . . . . . . . . . . . . . . . . . . . . . . . .. 44
Fig. 2.7 a ðkeq =kÞ for the periodic tridiagonal compact stencil,
Eq. (2.97) plotted for different values of a2 . a2 ¼ 2=11 yields a
sixth order accurate scheme. b Integral objective function I
given by Eq. (2.102) plotted as a function of a2 for c ¼ p . . . .. 46
Fig. 2.8 a ðk1eq =kÞ and b ðk2eq =kÞ2 plotted for the first and second
derivatives corresponding to CCD scheme (Eqs. (2.113) and
(2.114)) for periodic problems. Stencils of OUCS3-scheme

xxiii
xxiv List of Figures

Eq. (2.83) (first derivative) and the scheme due to Lele


Eq. (2.97) (for second derivative) are also included in frames
a and b, respectively. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 49
Fig. 2.9 Real and imaginary parts of ðk1eq =kÞj and ðk2eq =kÞ2j are plotted
at indicated grid points for NCCD scheme [98] with explicit
boundary closure schemes given by Eqs. (2.129)–(2.130) . . . .. 51
!
Fig. 2.10 a Staggered representation of variables in 2D ð V ; xÞ-
formulation. Here, x refers to the component of vorticity
perpendicular to the 2D plane. b Staggered arrangement
in 3D of u-, v- and w-velocity components and xx -, xy -
and xz -vorticity components shown on an elementary cell . . . .. 52
Fig. 2.11 a Transfer function of the staggered interpolation scheme
given by Eq. (2.131) plotted as a function of kh for indicated
values of aI . b Integrated phase error plotted as a function
of aI for the stencil for indicated values of c . . . . . . . . . . . . . .. 54
Fig. 2.12 a Spectral resolution of the scheme given by Eq. (2.135),
shown for indicated values of aII . b Integrated phase error
plotted as a function of aII for the stencil for the indicated
values of c . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 55
Fig. 2.13 Contour plots for the real part of keq 2
=k2 for different values of
aI in (kh; k0 h)-plane for 50th node, when total number points is
N ¼ 100 for the SACD scheme given by Eq. (2.140). The
maximum and the minimum value for each cases have been
indicated . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 58
Fig. 2.14 Contour plots for the real part of keq 2
=k2 for different values
0
of aII in (kh; k h)-plane for 50th node, when total number
points is N ¼ 100, using OSCS scheme twice to evaluate the
second derivative, as given by Eqs. (2.147)–(2.150).
The maximum and the minimum values for each cases
have been indicated . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 60
Fig. 2.15 Comparison of numerical amplification factor (jGj),
normalized numerical group velocity ðVgN =cÞ and phase error
ð1  cN =cÞ contours for the near-boundary node j ¼ 2 (left
column) and the central node j ¼ 51 (right column) for the
1D convection equation solved by the OUCS3-RK4 method . .. 66
Fig. 2.16 Comparison of the numerical amplification factor (jGj)
for interior nodes by indicated space-time discretization
schemes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 72
Fig. 2.17 Comparison of normalized numerical group velocity ðVgN =cÞ
for interior nodes for the indicated combinations of space-time
discretizations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 73
List of Figures xxv

Fig. 2.18 Transfer function (TF) of the periodic 1D filtering stencil


plotted as a function of kh for 2nd-, 4 and 6th-order filters
for indicated values of af . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 79
Fig. 2.19 a Real and b imaginary parts of transfer function TFj
for non-periodic problems with indicated filter coefficient a
for the indicated nodes for the 6th-order interior filter with
near-boundary filters of [122] . . . . . . . . . . . . . . . . . . . . . . . . . .. 80
Fig. 2.20 a Real and b imaginary part of TFj plotted for a 5th-order
upwinded filter with indicated af and g. The figure is taken
from [90] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 82
Fig. 2.21 Stream-function w- (frames a–e) and vorticity x-contours
(frames f–j) shown at indicated times for 95  x  120 before
applying 1D 2nd-order filter at the outflow. . . . . . . . . . . . . . . .. 84
Fig. 2.22 Stream-function w- (frames a–e) and vorticity x-contours
(frames f–j) shown at indicated times for 95  x  120 after
applying 1D 2nd-order filter at the outflow from t ¼ 15 . . . . . .. 85
Fig. 2.23 Spectral domain of interest for finite grid computations.
The shaded areas (EBF and HDG) contribute to aliasing
in evaluating product terms to zones HOG and EOF . . . . . . . .. 87
Fig. 2.24 The transfer functions of the 2nd-order 2D filter are shown
in frames a and b for indicated a2f . . . . . . . . . . . . . . . . . . . . . .. 91
Fig. 2.25 Schematic of the problem solved for free stream excitation
by an infinite array of convecting vortices over a semi-infinite
flat plate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 92
Fig. 2.26 Vorticity contours of the unfiltered case plotted at indicated
times for the free stream convecting vortex problem with
c ¼ 0:2, a ¼ 10, H ¼ 2:0 at indicated times in frames a, b and
c. d @@xx2w plotted of the filtered case at t ¼ 113 . . . . . . . . . . . .
2
.. 93
Fig. 2.27 a A typical variation of a2f is shown over a 9  9 filter tent in
the transformed plane. The corresponding 3D perspective plot
of the variation of a2f in the filter tent is shown in frame b . . .. 94
Fig. 2.28 Vorticity contours plotted at indicated times for the free stream
convecting vortex problem with c ¼ 0:2, a ¼ 10, H ¼ 2:0
times in frames a, b and c after applying adaptive 2D filter
from t ¼ 100. @@xx2w plotted of the filtered case at t ¼ 113
2

in frame d . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 95
Fig. 2.29 a Wall vorticity at t ¼ 113 of the unfiltered free stream
convecting vortex problem shown in Fig. 2.27 and b its
Fourier transform plotted as a function of streamwise distance
x and wavenumber k, respectively. c Wall vorticity at t ¼ 13
of the case where 2D adaptive filtered is used for the free
stream convecting vortex problem shown in Fig. 2.28 and d its
Fourier transform plotted as a function of streamwise distance
x and wavenumber k, respectively . . . . . . . . . . . . . . . . . . . . . .. 96
xxvi List of Figures

Fig. 2.30 Flow in a lid-driven cavity for Re ¼ 9000. a Maximum


absolute divergence error plotted as a function of time for
schemes A and B. b, c Vorticity time-history at x ¼ 0:95,
y ¼ 0:95 plotted for schemes A and B, respectively.
d, e Vorticity contours at indicated times shown
for schemes A and B, respectively . . . . . . . . . . . . . . . . . . . . . .. 99
Fig. 2.31 Flow in a lid-driven cavity for Re ¼ 10000. a Maximum
absolute divergence error time-history plotted for the indicated
schemes with different e. b Vorticity time-history at x ¼ 0:95
and y ¼ 0:95 shown plotted for (i) scheme A with e ¼ 104 ,
(ii) scheme A with e ¼ 106 , (iii) scheme A with e ¼ 109 and
(iv) scheme B with e ¼ 106 . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
Fig. 2.32 Flow inside a cubic lid-driven cavity for Re ¼ 1000. (i, ii)
Time-history of vorticity divergence error plotted for
computations using Laplacian and conservative rotational
forms of VTE, respectively. (iii, iv) Time-history of velocity
divergence error plotted for Laplacian and conservative
rotational forms of VTE, respectively . . . . . . . . . . . . . . . . . . . . . 104
Fig. 2.33 (i, ii) u- and v-velocity components at the mid-plane plotted for
Laplacian and rotational forms of VTE and (iii, iv) compared
with the results of Albensoeder and Kuhlmann [2] for flow
inside a cubic lid-driven cavity at Re ¼ 1000 . . . . . . . . . . . . . . . 105
Fig. 2.34 Stream-traces and contours of normal component of vorticity
are shown plotted at mid-ðx; yÞ plane (frames (i) and (ii)),
mid-ðy; zÞ plane (frames (iii) and (iv)) and mid-ðz; xÞ plane
(frames (v) and (vi)) for Re ¼ 1000 after steady state is
reached. All the plots correspond to the simulations done
using Laplacian form of VTE . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
Fig. 2.35 Stream-traces and contours of normal component of vorticity
are plotted at mid-ðx; yÞ plane (frames (i) and (ii)), mid-ðz; yÞ
plane (frames (iii) and (iv)) and mid-ðz; xÞ plane (frames
(v) and (vi)) for Re ¼ 1000 after steady state is reached. All the
plots correspond to the simulations done using conservative
rotational form of VTE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
Fig. 2.36 Flow inside a cubic lid-driven cavity for Re ¼ 3200. (i, ii)
Time-history of vorticity divergence error plotted for
computations using Laplacian and conservative rotational
forms of VTE, respectively. (iii, iv) Time-history of velocity
divergence error plotted for Laplacian and conservative
rotational forms of VTE, respectively . . . . . . . . . . . . . . . . . . . . . 108
Fig. 2.37 Flow inside a cubic lid-driven cavity for Re ¼ 3200. (i, ii)
u- and v-velocity components at the mid-plane and time
averaged between t ¼ 1900 and 2000 by using Laplacian
form are compared with the experimental time averaged results
List of Figures xxvii

of [51]. (iii, v) u- and (vi, vi) v-velocity components at


mid-plane obtained by using rotational form and Laplacian
form are compared at t ¼ 250 and t ¼ 2000 as indicated
in the frames . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
Fig. 2.38 Stream-traces shown plotted for a Laplacian form of VTE
and b rotational form of VTE used, at t ¼ 250 and 350
on mid-planes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
Fig. 2.39 Contours of vorticity component normal to the indicated
mid-planes shown plotted a Laplacian form of VTE and
b rotational form of VTE are used at t ¼ 250 and t ¼ 350 . . . . 111
Fig. 2.40 Stream-traces shown plotted for a Laplacian form of VTE
and b rotational form of VTE used at t ¼ 1800 and 2000
on mid-planes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
Fig. 2.41 Contours of vorticity component normal to the indicated
mid-planes shown plotted for a Laplacian form of VTE and
b rotational form of VTE used at t ¼ 1800 and 2000 . . . . . . . . . 113
Fig. 3.1 The solution of the Blasius equation plotted as a function
0 00 000
of the similarity variable g for a f , b f and c f . . . . . . . . . . . 123
Fig. 3.2 Harmonic excitation of a parallel boundary layer
corresponding to the location of the exciter . . . . . . . . . . . . . . . . 129
Fig. 3.3 Bromwich contour along with the region of convergence
shown in a a- and b x0 -plane . . . . . . . . . . . . . . . . . . . . . . . . . . 130
Fig. 3.4 Streamwise disturbance velocity plotted as function of ~x at
~y ¼ 0:278, for the case of harmonic excitation of a parallel
boundary layer. The location of the exciter corresponds to
~ ¼ 1500 and non-dimensional frequency of excitation is
Re
 0 ¼ 0:025, with the result shown for the viscous time scale
x
of t ¼ 801:11 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
Fig. 3.5 Streamwise disturbance velocity plotted as function of ~x at
~y ¼ 0:278, for the case of harmonic excitation of a parallel
boundary layer. The location of the exciter corresponds to
~ ¼ 750 and non-dimensional frequencies of excitation are
Re
ax  0 ¼ 0:10 and b x  0 ¼ 0:15, with the result shown for the
viscous time scale of t ¼ 801:11. . . . . . . . . . . . . . . . . . . . . . . . . 133
Fig. 3.6 Bromwich contour in a-plane . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
Fig. 3.7 Variation of mean flow quantities obtained
for K ¼ 1  106 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
Fig. 3.8 Variation of mean flow quantities obtained
for K ¼ 3  103 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
Fig. 3.9 Variation of mean flow quantities obtained
for K ¼ 9  102 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
Fig. 3.10 Buoyancy-induced pressure gradient term KgH plotted
as a function of g . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
xxviii List of Figures

Fig. 3.11 ðKgH=2XÞ contours plotted in the (x, y)-plane for the
indicated values of K. gfpg is marked by a dotted line . . . . . . . . 148
Fig. 3.12 Location of inflection point (gip ) plotted as a function
of K . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
Fig. 3.13 Schematic diagram for mixed convection flow over a wedge.
Identical computational domains are used for computing
equilibrium flow and its receptivity to wall–excitation.
[Reproduced from Direct numerical simulation of transitional
mixed convection flows: Viscous and inviscid instability
mechanism. Sengupta et al., Physics of Fluids, 25, 094102
(2013), with the permission of AIP Publishing.] . . . . . . . . . . . . . 153
Fig. 3.14 a Variation of mean flow quantities obtained from boundary
layer solution for flow past an adiabatically cooled horizontal
flat plate buoyancy parameter Ka ¼ 0:001 and 0:01.
b Streamwise velocity component and temperature obtained
from DNS for flow past an adiabatically cooled plate
Ka ¼ 0:001 and 0:01 plotted as a function of the similarity
variable g at indicated x-locations along with the
corresponding similarity solution. [Reproduced from Direct
numerical simulation of transitional mixed convection flows:
Viscous and inviscid instability mechanism. Sengupta et al.,
Physics of Fluids, 25, 094102 (2013), with the permission
of AIP Publishing.] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
Fig. 3.15 a Variation of mean flow quantities obtained from similarity
solution for the case of flow over adiabatic flat plate with
buoyancy parameter Ka ¼ 0:05. b Variation of mean flow
quantities obtained from similarity solution for the case of flow
over isothermal wedge with buoyancy parameter Ki ¼ 0:04.
[Reproduced from Direct numerical simulation of transitional
mixed convection flows: Viscous and inviscid instability
mechanism. Sengupta et al., Physics of Fluids, 25, 094102
(2013), with the permission of AIP Publishing.] . . . . . . . . . . . . . 158
Fig. 3.16 Vorticity (x) contours plotted in (x, y)-plane at indicated times
while computing equilibrium flow cold adiabatic plate with
buoyancy parameter Ka ¼ 0:05. [Reproduced from Direct
numerical simulation of transitional mixed convection flows:
Viscous and inviscid instability mechanism. Sengupta et al.,
Physics of Fluids, 25, 094102 (2013), with the permission
of AIP Publishing.] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160
Fig. 3.17 Vorticity (x) contours plotted in (x, y)-plane at indicated times
while computing equilibrium flow of heated isothermal wedge
with buoyancy parameter Ki ¼ 0:04. The dotted line in each
frame indicate the location in y direction where Fj/rtoft
integrand C changes sign from positive to negative. . . . . . . . . . . 161
List of Figures xxix

Fig. 3.18 a, b Neutral curve obtained from linear spatial analysis shown
in ðRed ; b0 Þ-plane buoyancy parameter for: a flow over cold
adiabatic plate and b flow over heated isothermal wedge.
Frames c, d show neutral curve obtained from linear temporal
instability analysis shown in the ðRed ; kreal Þ-plane for
indicated values of buoyancy parameter for the cases shown in
frames (a) and (b). [Reproduced from Direct numerical
simulation of transitional mixed convection flows: Viscous and
inviscid instability mechanism. Sengupta et al., Physics of
Fluids, 25, 094102 (2013), with the permission of AIP
Publishing.] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
Fig. 3.19 Temporal growth rate contours ci ¼ bi =kreal plotted in
ðRed ; kreal Þ-plane for indicated mixed convection cases.
[Reproduced from Direct numerical simulation of transitional
mixed convection flows: Viscous and inviscid instability
mechanism. Sengupta et al., Physics of Fluids, 25, 094102
(2013), with the permission of AIP Publishing.] . . . . . . . . . . . . . 167
Fig. 3.20 a Disturbance streamwise velocity ud at a height of y ¼ 0:0061
plotted at indicated times for flow past a horizontal adiabatic
cold plate with Ka ¼ 0:001 (frames (i  v)) and 0:01
(frames (vi  x)) is excited by SBS strip placed at x ¼ 1:5.
Here, non-dimensional frequency and amplitude of excitation
are F ¼ 9:76  105 and a1 ¼ 0:001, respectively. The
vertical dashed lines in the frames indicate onset and decay of
disturbance locations (as P1 and P2 in Fig. 3.18a) as per local
linear spatial instability theories. b ud is plotted at a height of
y ¼ 0:0061 for the indicated times for flow past an isothermal
heated wedge with Ki ¼ 0:025 excited by SBS strip placed at
x ¼ 21:0. Here, F ¼ 1:65  105 and a1 ¼ 0:001, are used
and the vertical dashed lines in the frames indicate onset and
decay of disturbance locations (as Q1 and Q2 in Fig. 3.18b)
obtained by linear spatial instability theory. Reproduced from
[Direct numerical simulation of transitional mixed convection
flows: Viscous and inviscid instability mechanism. Sengupta
et al., Physics of Fluids, 25, 094102 (2013)], with the
permission of AIP Publishing . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
Fig. 3.21 a Rayleigh integrand U for flow over adiabatic plate parameter
Ka ¼ 0:05 for i ci ¼ 0:01 and ii ci ¼ 0:1. b Fj/rtoft
integrand CF for flow over adiabatic plate with buoyancy
parameter Ka ¼ 0:05 for i ci ¼ 0:01 and ii ci ¼ 0:1.
Reproduced from [Direct numerical simulation of transitional
mixed convection flows: Viscous and inviscid instability
mechanism. Sengupta et al., Physics of Fluids, 25, 094102
(2013)], with the permission of AIP Publishing . . . . . . . . . . . . . 176
xxx List of Figures

Fig. 3.22 a Rayleigh integrand U for flow over isothermal wedge with
buoyancy parameter Ki ¼ 0:04 for i ci ¼ 0:01 and
ii ci ¼ 0:1. b Fj/rtoft integrand CF for flow over isothermal
wedge with buoyancy parameter Ki ¼ 0:04 for i ci ¼ 0:01
and ii ci ¼ 0:1. [Reproduced from Direct numerical
simulation of transitional mixed convection flows: Viscous
and inviscid instability mechanism. Sengupta et al., Physics
of Fluids, 25, 094102 (2013), with the permission
of AIP Publishing.] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
Fig. 3.23 Critical value ðcr Þcrit plotted for cases of a flow over
isothermal wedge in (Ki ; ci )-plane and b for flow over
adiabatic cold plate in ðKa ; ci Þ-plane. Here, ðcr Þcrit represents
the critical phase speed such that for all cr [ ðcr Þcrit , no
zero-crossing point of U exists in the flow. [Reproduced from
Direct numerical simulation of transitional mixed convection
flows: Viscous and inviscid instability mechanism. Sengupta
et al., Physics of Fluids, 25, 094102 (2013), with the
permission of AIPR Publishing.] . . . . . . . . . . . . . . . . . . . . . . . . . . 181
Fig. 3.24 Contours of IC ¼ CF d~y plotted in ðcr ; ci Þ-plane for flow past
a adiabatic cold plate with indicated values of buoyancy
parameter Ka . [Reproduced from Direct numerical simulation
of transitional mixed convection flows: Viscous and inviscid
instability mechanism. Sengupta et al., Physics of Fluids, 25,
094102 (2013), with the permission of AIP Publishing.] . . . . . . 182
Fig. 3.25 Contours of /real ¼ 0 (solid lines) and /img ¼ 0 (broken lines)
plotted at the wall in ðcr ; ci Þ-plane for flows past adiabatic cold
plate with indicated buoyancy parameter Ka and disturbance
wavenumber kreal . [Reproduced from Direct numerical
simulation of transitional mixed convection flows: Viscous and
inviscid instability mechanism. Sengupta et al., Physics of
Fluids, 25, 094102 (2013), with the permission of AIP
Publishing.] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
Fig. 3.26 Contours of /real ¼ 0 (solid lines) and /img ¼ 0 (broken lines)
plotted at the wall in ðcr ; ci Þ-plane for flows past isothermal hot
plate with indicated buoyancy parameter Ki . We have chosen
disturbance wavenumber kreal ¼ 0:2. [Reproduced from Direct
numerical simulation of transitional mixed convection flows:
Viscous and inviscid instability mechanism. Sengupta et al.,
Physics of Fluids, 25, 094102 (2013), with the permission of
AIP Publishing.] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
Fig. 3.27 Theoretical neutral curve in ðRed ; x0 Þ-plane. According to
linear theory with parallel flow assumption, three frequencies
shown corresponding to OA, OB and OC are all stable. The
List of Figures xxxi

discrete symbols are the experimental data from Schubauer and


Skramstad [50] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
Fig. 3.28 Fourier transform of the input disturbance by the SBS strip
given by Eqs. (3.100) and (3.101) for the amplitudes of
excitation a1 ¼ 0:05 and 0.1. The exciter is located from
x1 ¼ 0:22 to x2 ¼ 0:264 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
Fig. 3.29 The similarity functions a f 0 ðgÞ, b f 00 ð gÞ and c f 000 ð gÞ to the
equilibrium flow obtained by DNS, plotted as a function of the
similarity variable g, at indicated streamwise locations. The
corresponding solution of the Blasius equation is also shown
for comparison . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188
Fig. 3.30 Streamwise disturbance velocity ud obtained as a receptivity
solution at t ¼ 8, for y = 0.008, identifying the local,
asymptotic and wave-front components due to a strip
excitation near the leading edge. The frequency of excitation
of the SBS strip on the wall is F ¼ 3:0  104 . . . . . . . . . . . . . 189
Fig. 3.31 Fourier transform of the receptivity solution shown in
Fig. 3.30, identifying the wavenumbers of the local,
asymptotic and wave front components in the legend . . . . . . . . . 190
Fig. 3.32 ud plotted as a function of x for t ¼ 20 at y ¼ a 0:00521,
c 0.00662, e 0.00808, g 0.00958 and i 0.01137. The signals are
for the SBS strip excitation amplitude of a1 ¼ 0:05, and
non-dimensional frequency, F ¼ 3:0  104 with exciter
location between x ¼ 0:22 and x ¼ 0:264. Fourier transform
of the signals shown in right frames . . . . . . . . . . . . . . . . . . . . . . 191
Fig. 3.33 ud =jud jmax plotted as a function of time for x ¼ 0:3 at
a y ¼ 0:00385, c 0.00662, e 0.00808, g 0.00958 and i 0.0144.
The figure is shown for the SBS strip excitation amplitude of
a1 ¼ 0:05 and non-dimensional frequency F ¼ 3:0  104
with exciter location between x1 ¼ 0:22 and x2 ¼ 0:264.
Fourier transforms plotted for the signals in right frames.
Here, jud jmax is the maximum absolute value of ud for
20  t  21 at the corresponding (x, y)-location . . . . . . . . . . . . . 193
Fig. 3.34 ud plotted as a function of x for t ¼ 20 at a y ¼ 0:00124,
c 0.00252, e 0.00385, g 0.03 and i 0.032. The ZPG flow past a
flat-plate is excited by a SBS strip located between x1 ¼ 0:22
and x2 ¼ 0:264 with a1 ¼ 0:05 and the frequency
F ¼ 3:0  104 . Fourier transform of the signals are shown
in right frames . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194
Fig. 3.35 a Mean and disturbance components of streamwise velocity
shown along with estimates of critical layer for the second
spatial mode. b ud and c vd plotted as function of y for
indicated streamwise locations at time-instants when
corresponding ud at inner maxima attains its maximum value.
xxxii List of Figures

In all the three frames, the lines 1a and 1b indicates the range
of heights where maximum local spatial growth is noted.
Presented data are for the SBS excitation with a1 ¼ 0:05
and F ¼ 3:0  104 with exciter location between
0.22 and 0.264 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196
Fig. 3.36 Normalized disturbance streamwise velocity ud at t ¼ 20 and
indicated heights plotted of x for a a1 ¼ 0:002, b a1 ¼ 0:005
and c a1 ¼ 0:01. The normalization is done with respect to
a1 ¼ 0:01 case. The exciter is located between 0.22 and 0.264
which excites the flow with F ¼ 3:0  104 . . . . . . . . . . . . . . . 197
Fig. 3.37 ud =jud jmax plotted as a function of time for 20  t  21 and
x ¼ 0:35 at indicated heights for a a1 ¼ 0:002, b a1 ¼ 0:005,
c a1 ¼ 0:01. The exciter is located between 0.22 and 0.264
and F ¼ 3:0  104 . jud jmax is the maximum absolute value of
ud for 20  t  21 at corresponding (x, y)-location . . . . . . . . . . . 199
Fig. 3.38 w-contours shown in the physical plane for t ¼ 20 for the case
of a1 ¼ 0:10; F ¼ 3:0  104 located between x1 ¼ 0:22 and
x2 ¼ 0:264. Note the micro-bubbles near the exciter
on the plate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
Fig. 3.39 ud plotted for y ¼ 0:00245; 0.00521 and 0.035, respectively,
from top to bottom at t ¼ 20. Results presented are for the SBS
strip excitation with a1 ¼ 0:10; Ff ¼ 3:0  104 and the
exciter located between x1 ¼ 0:22 and x2 ¼ 0:264. b Fourier
transform of the signals shown in a . . . . . . . . . . . . . . . . . . . . . . 201
Fig. 3.40 Normalized disturbance velocity u^d plotted against x at t ¼ 15
for amplitudes of excitation at y ¼ 0:00252, which is the first
grid line above the plate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
Fig. 3.41 Normalized disturbance velocity u^d (normalized with respect
to a1 ¼ 0:01 case) plotted against x at t ¼ 15 for amplitudes of
excitation at y ¼ 0:0081 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202
Fig. 3.42 a Time variation of ud =jud jmax plotted at x ¼ 0:3 and
y ¼ 0:00124; 0.00662 and 0.034 for a1 ¼ 0:1,
Ff ¼ 3:0  104 . The exciter located between 0.22 and 0.264.
In frame b, Fourier transform of the signal shown in a. Here,
jud jmax is the maximum absolute value of ud for 13  t  14 at
the corresponding (x, y)-location . . . . . . . . . . . . . . . . . . . . . . . . . 203
Fig. 3.43 Time variation of xd plotted (left) for a1 ¼ 0:1 and
Ff ¼ 3:0  104 at x ¼ 1:34 for the three indicated heights
close to the plate. Corresponding Fourier transforms shown
on the right . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204
Fig. 3.44 Time variation of xd plotted at x ¼ 0:27 for a1 ¼ 0:1 and
Ff ¼ 3:0  104 (left) for the heights shown in Fig. 3.43.
Corresponding Fourier transforms are shown on the right . . . . . 205
List of Figures xxxiii

Fig. 3.45 Normalized streamwise disturbance velocity ud plotted as a


function of y=d at x ¼ 0:3 and indicated time instants. Here,
C1, C2, C3, C4 and C5 represent the cases with a1 ¼ 0:002,
0.005, 0.01, 0.05 and 0.1, respectively. The exciter is put
between 0.22 and 0.264 with F ¼ 3:0  104 . . . . . . . . . . . . . . 206
Fig. 3.46 Normalized streamwise disturbance velocity ud plotted as a
function of y=d at x ¼ 0:5 and indicated time instants. Here,
C1, C2, C3, C4 and C5 represent the cases with a1 ¼ 0:002,
0.005, 0.01, 0.05 and 0.1, respectively. The exciter is between
0.22 and 0.264 with F ¼ 3:0  104 . . . . . . . . . . . . . . . . . . . . . 207
Fig. 3.47 Normalized wall-normal disturbance velocity vd plotted as a
function of y=d at x ¼ 0:3 and indicated time instants. Here,
C1, C2, C3, C4 and C5 represent the cases with a1 ¼ 0:002,
0.005, 0.01, 0.05 and 0.1, respectively. The exciter is between
0.22 and 0.264 with F ¼ 3:0  104 . . . . . . . . . . . . . . . . . . . . . 208
Fig. 3.48 Normalized wall-normal disturbance velocity vd plotted as a
function of y=d at x ¼ 0:5 and indicated time instants. Here,
C1, C2, C3, C4 and C5 represent the cases with a1 ¼ 0:002,
0.005, 0.01, 0.05 and 0.1, respectively. The exciter is between
0.22 and 0.264 with F ¼ 3:0  104 . . . . . . . . . . . . . . . . . . . . . 209
Fig. 3.49 ud plotted against x at t ¼ 15 for y ¼ 0:00662, a1 ¼ 0:01
and the indicated frequencies of excitation . . . . . . . . . . . . . . . . . 210
Fig. 3.50 xd plotted as a function of x at t ¼ 15 for y ¼ 0:0066
with a1 ¼ 0:01 for the displayed frequencies . . . . . . . . . . . . . . . 211
Fig. 3.51 xd plotted against x at t ¼ 15 for the indicated frequencies
of excitation at y ¼ 0:0112 with a1 ¼ 0:01 . . . . . . . . . . . . . . . . . 212
Fig. 3.52 ud plotted against x at t ¼ 15 for the indicated frequencies
of excitation at y ¼ 0:0112 with a1 ¼ 0:01 . . . . . . . . . . . . . . . . . 213
Fig. 3.53 Stream function and vorticity contours for the two different
strip exciter cases having different width. a and b correspond
to exciter located between 0.22 and 0.264; c and d correspond
to exciter located between 0.2 and 0.29 . . . . . . . . . . . . . . . . . . . 214
Fig. 3.54 a jud j=jud jmax and b jvd j=jvd jmax plotted as a function of y
for the indicated exciter locations times after the complete
establishment of the TS wave-packet. The profiles are plotted
at a location of x ¼ 0:075 downstream of the center of the
exciter. Here, F ¼ 3:0  104 and a1 ¼ 0:01 . . . . . . . . . . . . . . . 215
Fig. 3.55 ud plotted as a function of x for the indicated values of F
at y ¼ 0:02549 and t ¼ 20. Here, the exciter of width
Dxex ¼ 0:05 is located at xex ¼ 0:25 with a1 ¼ 0:01.
The vertical straight lines in a and b indicate the entry
and exit point of the corresponding Ff ¼ Const: line into
the neutral curve . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218
xxxiv List of Figures

Fig. 4.1 Schematic of the experimental set-up. a Side view and b top
view as seen in the tunnel. Broken line boundary in b indicates
the computational domain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228
Fig. 4.2 Bypass transition created by a counter-rotating vortex for
ðU1 ¼ 16:26 cm s1 , c ¼ 0:154; H ¼ 9 cm and X ¼ þ 5 r.p.
s.) for the experimental arrangement shown in Fig. 4.1 . . . . . . . 229
Fig. 4.3 Computational domain used in studying vortex-induced
instability by an isolated convecting (at a speed, c) vortex
(of circulation, C) in the free stream over a flat plate supporting
a zero pressure gradient boundary layer . . . . . . . . . . . . . . . . . . . 230
!
Fig. 4.4 Staggered grid system used in ð V ; xÞ-formulation for 2D
problems. The velocity components are at the mid-point of the
elemental surface over which it is normal . . . . . . . . . . . . . . . . . . 233
Fig. 4.5 Stream function contours plotted in the computational domain
at the indicated times. Arrowheads at the top show the
streamwise location of the convecting vortex . . . . . . . . . . . . . . . 236
Fig. 4.6 Vorticity contours plotted in the computational domain at the
indicated times, as in Fig. 4.5. Same contour values are plotted
in all the frames. Arrowheads at the top show the streamwise
location of the convecting vortex . . . . . . . . . . . . . . . . . . . . . . . . 238
Fig. 4.7 Contours of the right-hand side of the disturbance energy
Eq. (4.6). The negative values indicating disturbance energy
sources are plotted as dark contours . . . . . . . . . . . . . . . . . . . . . . 240
Fig. 4.8 (i) Disturbance stream function versus x caused by circulatory
and displacement effects, by a translating and rotating circular
cylinder, while the disturbance source is at xc ¼ 5 and H1 ¼ 6
moving at the indicated propagation speed. (ii) Induced
pressure gradient for the case of frame (i) for the indicated
propagation speed. (iii) The same pressure gradient shown as a
function of time for various convecting and non-convecting
free stream vortex cases. (iv) Variation of Falkner–Skan
parameter m with x shown at two representative times for the
case of c ¼ 0:2. The time Tsep indicates a large time for
plotting m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241
Fig. 4.9 The growth rate of total enstrophy Eq. (4.19) for the
vortex-induced problem defined in Fig. 4.3 for which the
physical variables are shown in Figs. 4.5 and 4.6 for the
identical time instants. The marked regions are for the
instability of enstrophy, i.e., where rotationality increases with
time at the indicated time instants . . . . . . . . . . . . . . . . . . . . . . . . 247
Fig. 4.10 The growth rate of Xd Eq. (4.19) for the vortex-induced
instability problem defined in Fig. 4.3, for which the physical
variables are shown in Figs. 4.5 and 4.6 for the identical time
List of Figures xxxv

instants. On the left frames shown is the case with positive Xd


and on the right is the case of negative Xd shown at the
indicated time instants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 250
Fig. 4.11 The nonlinear growth rate of Xd for the vortex-induced
instability defined in Fig. 4.3, for which the physical variables
are shown in Figs. 4.5 and 4.6 for the identical time instants.
On the left frames shown are the regions where Xd is positive
and on the right where Xd is negative, with overlap regions
where rotationality increases with time at the indicated time
instants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
Fig. 4.12 Vorticity contours at indicated times are shown plotted in
frames (a) to (b) for ZPG flow past a flat plate excited by SBS
strip-exciter. The location of the midpoint of the exciter strip is
pointed by an arrow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256
Fig. 4.13 Cumulative enstrophy content are shown plotted against
corresponding mode number for zero-pressure gradient flow
over semi-infinite flat-plate with wall-excitation . . . . . . . . . . . . . 257
Fig. 4.14 Time dependent amplitude function aj ðtÞ of first eighteen POD
modes for zero-pressure gradient flow over semi-infinite
flat-plate plotted as a function of time. The pair-forming modes
are shown together in a frame as indicated. The numbering
of the modes follow the conventions used in [62, 64] . . . . . . . . 260
Fig. 4.15 Fast Fourier transform of amplitude functions aj ðtÞ of first
eighteen POD modes are shown plotted as a function of
non-dimensional frequency bDt . . . . . . . . . . . . . . . . . . . . . . . . . 261
Fig. 4.16 Contours of eigenfunctions /j ðx; yÞ corresponding to the first
ten POD modes for zero-pressure gradient flow semi-infinite
flat-plate with wall-excitation are shown plotted . . . . . . . . . . . . . 262
Fig. 4.17 Contours of eigenfunctions /j ðx; yÞ corresponding to the first
eighteen POD modes for zero-pressure gradient flow
semi-infinite flat-plate with wall-excitation are
shown plotted . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263
Fig. 4.18 Modulus of amplitude function jAj ðtÞj of the instability modes
shown plotted as a function of time for zero-pressure gradient
flow over semi-infinite flat-plate . . . . . . . . . . . . . . . . . . . . . . . . . 267
!
Fig. 4.19 Modulus of space dependent function jfj ð X Þj of the indicated
instability modes are shown plotted as a function of time for
zero-pressure gradient flow over semi-infinite flat-plate . . . . . . . 268
Fig. 4.20 Reconstructed signals of a5 ðtÞ with indicated frequency band
about the peak at bDt ¼ 7:234  102 shown in frames (ii) to
(vii). The original signal a5 ðtÞ is shown in the top frame . . . . . . 269
Fig. 4.21 Reconstructed signals of a11 ðtÞ with indicated frequency band
about the peak at bDt ¼ 7:234  102 are shown plotted in
xxxvi List of Figures

frames (ii) to (vii). The original signal a11 ðtÞ is shown in the
top frame . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270
Fig. 5.1 Schematic of 2D receptivity to SBS wall excitation . . . . . . . . . . 277
Fig. 5.2 ud plotted as a function of x at y ¼ 0:0057 and indicated times
for a1 ¼ 0:002, F ¼ 1:0  104 and xex ¼ 1:5. The
spatio-temporal wave front in the top frame and the TS
wave-packet in the last frame is indicated by # and asterisk
symbol, respectively . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
Fig. 5.3 a Variation and b location of maximum peak-to-peak
amplitude of the STWF for ud at y ¼ 0:0057 with time
a1 ¼ 0:002, F ¼ 1:0  104 and xex ¼ 1:5. Six stages
of the growth of the wave front are identified and marked
in frame (a) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283
Fig. 5.4 The stream-function w- (left column) and vorticity x-contours
(right column) are plotted at indicated times for a1 ¼ 0:002,
F ¼ 1:0  104 and xex ¼ 1:5 . . . . . . . . . . . . . . . . . . . . . . . . . . 284
Fig. 5.5 The spatial Laplace transform of the STWF at y ¼ 0:0057
shown for indicated times. The excitation parameters are
Ff ¼ 104 , xex ¼ 1:5 and a1 ¼ 0:002 . . . . . . . . . . . . . . . . . . . . . 285
Fig. 5.6 E11 corresponding to the STWF plotted as a function of k;
for SBS excitation case with a1 ¼ 0:002; Ff ¼ 104
and xex ¼ 1:5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 286
Fig. 5.7 ud corresponding to the STWF plotted as a function of y=d
at indicated time and location. Here, d is the local
displacement thickness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287
Fig. 5.8 a Variation of the maximum peak-to-peak amplitude udm
corresponding to the STWF at y ¼ 0:0057 with time shown
for indicated values of a1 . In the inset shown in frame a,
the variation of udm for a1 ¼ 0:01 is shown for initial times . . . . 288
Fig. 5.9 w- and x-contours plotted at indicated times for a1 ¼ 0:05
with F ¼ 1:0  104 and xex ¼ 1:5 . . . . . . . . . . . . . . . . . . . . . . 289
Fig. 5.10 ud corresponding to the STWF plotted as a function of x at
y ¼ 0:0057 and indicated times for a1 ¼ 0:01, F ¼ 1:0  104
and xex ¼ 0:5 in frames (a) to (g) . . . . . . . . . . . . . . . . . . . . . . . . 290
Fig. 5.11 ud corresponding to the STWF plotted as a function of x at
y ¼ 0:0057 and indicated times for a1 ¼ 0:05, F ¼ 1:0  104
and xex ¼ 0:5 in frames (a) to (g) . . . . . . . . . . . . . . . . . . . . . . . . 291
Fig. 5.12 Maximum amplitude of ud at y ¼ 0:0057 corresponding to the
TS wave-packet plotted as a function of time for indicated
values of a1 with F ¼ 1:0  104 and xex ¼ 1:5 . . . . . . . . . . . . 291
Fig. 5.13 ud plotted as a function of x at y ¼ 0:0057 and indicated times
for a1 ¼ 0:002, F ¼ 0:5  104 and xex ¼ 1:5 . . . . . . . . . . . . . . 293
List of Figures xxxvii

Fig. 5.14 Variation of the maximum peak-to-peak amplitude of ud


corresponding to the TS wave-packet at y ¼ 0:0057 plotted
as a function of time for Ff ¼ 0:5  104 , a1 ¼ 0:002
and xex ¼ 1:5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295
Fig. 5.15 Spatial Laplace transform of ud within the range 2:5  x  50
at y ¼ 0:0057 plotted at indicated times with Ff ¼ 0:5  104
and a1 ¼ 0:002. The exciter is located at xex ¼ 1:5 . . . . . . . . . . 296
Fig. 5.16 Time variation of the amplitude of the main peak k10 and
sub-peaks k11 to k17 as marked in Fig. 5.14 are plotted . . . . . . . 297
Fig. 5.17 ud at y ¼ 0:0057 plotted as a function of x at indicated times
for a1 ¼ 0:003, xex ¼ 1:5 and Ff ¼ 0:5  104 . . . . . . . . . . . . . 298
Fig. 5.18 ud at y ¼ 0:0057 plotted as a function of x at indicated times
for a1 ¼ 0:003, xex ¼ 1:75 and Ff ¼ 0:5  104 . . . . . . . . . . . . 299
Fig. 5.19 Stream function contours for flow past SHM1 airfoil shown at
the indicated time instants for Re ¼ 10:3  106 , when the
airfoil is kept at zero angle of attack (AOA). The top frame
shows the flow field around the full airfoil, while the bottom
frames show the zoomed view of the flow field near the trailing
edge, showing the presence of small bubbles indicative of
bypass transition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301
Fig. 5.20 Variation of azimuthal component of velocity (u) at indicated
height of 1:242  106 from the top surface of the SHM1
airfoil at the indicated time instants . . . . . . . . . . . . . . . . . . . . . . 302
Fig. 5.21 Variation of azimuthal component of velocity (u) at indicated
height from the top surface SHM1 airfoil at the indicated time
instants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
Fig. 5.22 Variation of azimuthal component of velocity (u) at a height
of 1:242  106 from the top surface of the SHM1 airfoil
at the indicated time instants. Wall excitation corresponds
to an SBS frequency of F ¼ 1:11441  105 , with an
amplitude of 0.001 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 304
Fig. 6.1 Schematic diagram of the 3D receptivity problem for a 2D
ZPG boundary layer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 308
Fig. 6.2 a The amplitude function Am ðrÞ plotted as a function of r=rmax .
b ud plotted in ðx; zÞ-plane at y ¼ 0:00189 and t ¼ 15 for 2D
ZPG boundary layer excited by a circular wall bump exciter
with non-dimensional frequency Ff ¼ 5  105 . . . . . . . . . . . . . 315
Fig. 6.3 ud plotted as a function of x at z ¼ 0, y ¼ 0:00189 and
indicated time-instants in the right column for 2D ZPG
boundary layer circular wall bump exciter with
non-dimensional frequency Ff ¼ 5  105 . The spectrum for
the plotted ud in the left column is shown in the right . . . . . . . . 316
xxxviii List of Figures

Fig. 6.4 ud -contours at indicated time instant and y ¼ 0:00189 shown


in ðx; zÞ-plane for a circular wall bump exciter for indicated
frequencies of excitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 317
Fig. 6.5 See caption of Fig. 6.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 317
Fig. 6.6 See caption of Fig. 6.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 318
Fig. 6.7 See caption of Fig. 6.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 318
Fig. 6.8 See caption of Fig. 6.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319
Fig. 6.9 See caption of Fig. 6.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319
Fig. 6.10 Xf -contours plotted in ðx; yÞ-plane for z ¼ 0 at indicated times
for a circular wall bump exciter with Ff ¼ 0:5  104
(left frames) and Ff ¼ 1:0  104 (right frames) . . . . . . . . . . . . 320
Fig. 6.11 ud plotted as a function of y=d at a z ¼ 0 and b z ¼ 0:6625.
The time instants and the streamwise locations correspond to
the maximum amplitude of the wave-front at z ¼ 0 and
z ¼ 0:6625, respectively for y ¼ 0:00189 . . . . . . . . . . . . . . . . . . 321
Fig. 6.12 Iso-surface of Q shown plotted at a t ¼ 24:96 and b t ¼ 27:28
for a circular wall bump exciter for Ff ¼ 0:5  104 . . . . . . . . . 322
Fig. 6.13 ud -contours plotted in the ðx; zÞ-plane at indicated time
for a zmax ¼ 2 and b zmax ¼ 4 when excited by a circular wall
bump exciter for Ff ¼ 1:0  104 . . . . . . . . . . . . . . . . . . . . . . . 323
Fig. 6.14 See the caption for Fig. 6.13 . . . . . . . . . . . . . . . . . . . . . . . . . . . 324
Fig. 6.15 See the caption for Fig. 6.13 . . . . . . . . . . . . . . . . . . . . . . . . . . . 324
Fig. 6.16 See the caption for Fig. 6.13 . . . . . . . . . . . . . . . . . . . . . . . . . . . 325
Fig. 6.17 See the caption for Fig. 6.13 . . . . . . . . . . . . . . . . . . . . . . . . . . . 325
Fig. 6.18 See the caption for Fig. 6.13 . . . . . . . . . . . . . . . . . . . . . . . . . . . 326
Fig. 6.19 Amplitude of ud corresponding to the wave-front plotted as a
function of time at indicated spanwise locations of z ¼ 0, 0.65
and 0.9 for y ¼ 0:00189. The frequency of excitation is
Ff ¼ 0:5  104 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 327
Fig. 6.20 Perspective plot of ud at a fixed height shown for SM exciter.
The frequency of excitation is F ¼ 0:5  104 . . . . . . . . . . . . . . 328
Fig. 6.21 ud -contours plotted in the ðx; zÞ-plane at y ¼ 0:00189 and
indicated time, when 2D ZPG boundary layer excited by
spanwise modulated exciter strip for Ff ¼ 1:0  104 . . . . . . . . 329
Fig. 6.22 Xn -contours at indicated streamwise locations and t ¼ 27:24
plotted in the ðy; zÞ-plane for spanwise modulated excitation
for Ff ¼ 1:0  104 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 330
Fig. 6.23 ud plotted at y ¼ 0:00189, indicated times and spanwise
locations for spanwise modulated excitation for
Ff ¼ 1:0  104 . All the indicated z-locations correspond to
the peak positions of the excitation . . . . . . . . . . . . . . . . . . . . . . . 332
Fig. 6.24 ud at y ¼ 0:00189 and spanwise locations plotted for a t ¼ 10
and b t ¼ 15. Maximum amplitude of the STWF udm at
y ¼ 0:00189 plotted as a function of time for z ¼ 0:25, 0.125
List of Figures xxxix

and 0.375. Here, spanwise modulated excitation is provided


for Ff ¼ 1:0  104 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 333
Fig. 6.25 Perspective view of k2 ¼ 0:0025 iso-surface is shown
in (x; z)-plane at t ¼ 25 after the onset of excitation
corresponding to GCP exciter case for a Ff ¼ 1:0  104
and b Ff ¼ 0:5  104 . Flow is from left to right . . . . . . . . . . . 335
Fig. 6.26 Perspective and top view (in (x; z)-plane) of k2 ¼ 0:015
iso-surface are shown for the SM exciter case for (a, b)
Ff ¼ 104 and (c, d, e) Ff ¼ 0:5  104 . Flow is from left
to right . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 336
Fig. 6.27 xz -contours at z ¼ 0 station plotted in (x; y)-plane at a t ¼ 30,
b t ¼ 35, c t ¼ 40 and d t ¼ 50 for GCP exciter case for
F ¼ 0:5  104 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 338
Fig. 6.28 a u þ plotted as a function of y þ at indicated x-stations for
GCP exciter case for F ¼ 0:5  104 . b Non-dimensional
Reynolds stress plotted as a function of y þ at indicated
x-stations and c nondimensional streamwise component rms
velocity shown at the indicated streamwise locations . . . . . . . . . 339
Fig. 6.29 a–c Instantaneous and d time-averaged skin-friction coefficient
Cf along z ¼ 0 spanwise station plotted as a function of x.
In d Time-averaged Cf during the interval 40  t  50 are
plotted as function of x. For comparison, corresponding
correlations for laminar and turbulent flows are shown in all
these frames. e Stream-trace is shown at t ¼ 40 to indicate the
recirculating region at the onset of transition . . . . . . . . . . . . . . . 340
Fig. 6.30 a–e Displacement thickness d , and momentum thickness h,
at z ¼ 0 section plotted at indicated time instants. f–j Shape
factor H at z ¼ 0 section plotted at indicated time instants . . . . . 341
Fig. 6.31 a–c Time averaged compensated streamwise spectral density
for streamwise, wall-normal and spanwise velocity
components plotted as a function of streamwise wavenumber
kx for z ¼ 0, 0.38 and 0.48 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342
List of Tables

Table 3.1 Streamwise pressure gradient induced by buoyancy effects for


a heated flat plate for different buoyancy parameters . . . . . . . . . 147
Table 3.2 First ten most dominant unstable modes from inviscid
instability analysis tabulated for flows past adiabatic cold plate
with Ka ¼ 0:05 and Ka ¼ 0:001. We have tabulated only
for disturbance wavenumber of kreal ¼ 0:2 and 0.05.
[Reproduced from Direct numerical simulation of transitional
mixed convection flows: Viscous and inviscid instability
mechanism. Sengupta et al., Physics of Fluids, 25, 094102
(2013), with the permission of AIP Publishing.] . . . . . . . . . . . . 184
Table 3.3 Onset and end x-locations for the spatial growth of ud at y ¼
0:00662 and y ¼ 0:00808 for different values of excitation
amplitude a1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202
Table 3.4 Location of the exciter xex , corresponding Red ,
non-dimensional wavenumbers ar ; kr and spatial growth rates
ai ; ki at the exciter location as obtained from the linear stability
theory are tabulated. Here, ar and ai are defined with respect to
local d , whereas kr and ki is defined with respect to L. For all
the cases listed here, F ¼ 3:0  104 and a1 ¼ 0:01 . . . . . . . . . 214
Table 3.5 The onset and end of the spatial growth or the near-neutral
behaviour of the TS wave-packet listed for all heights, for the
indicated locations of exciter. Here, F ¼ 3:0  104 and
a1 ¼ 0:01. The blank cells indicate those heights, where no
spatial growth is noticed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
Table 3.6 Non-dimensional frequency of excitation F, corresponding
non-dimensional wavenumbers ar ; kr and spatial growth rates
ai ; ki at the exciter location as obtained from the linear stability
theory along with the entry (xN1 ) and exit (xN2 ) point of the
F ¼ Const: line into the neutral curve are tabulated. Here, ar
and ai is defined with respect to local d , whereas kr and ki is

xli
xlii List of Tables

defined with respect to reference length scale L. For all the


cases listed here, the exciter is located at xex ¼ 0:25 with
a1 ¼ 0:01 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
Table 3.7 The onset and end of the spatial growth or the near-neutral
behaviour of the TS wave-packet for all the heights shown in
Fig. 3.55 for indicated F  104 cases. The blank cells indicate
heights, where no spatial growth is observed . . . . . . . . . . . . . . . 219
Table 4.1 Eigenvalues and cumulative enstrophy of the POD modes . . . . . 258
pffiffiffi
Table 4.2 Normalization factor ej and its square root ej for instability
modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 266
Chapter 1
DNS of Wall-Bounded Turbulent Flow:
An Introduction

1.1 Introduction

This book covers the topic of direct numerical simulation (DNS) of wall-bounded
turbulent flow by first principle approach. It is mandatory to track the flow from
laminar state to fully developed turbulent state, and which has been solved very
recently for specific cases. While one can attempt DNS of Navier-Stokes equation
(NSE) for fully developed turbulent flow, there would be ambiguity about specifying
initial and boundary conditions. Moreover, such an exercise even if it is successful,
then also it may not answer most of the fundamental questions related to the evolution
of the flow field. First, by erroneous numerical methods and artificial excitations,
one may achieve fully developed turbulence, but that approach should be avoided.
Such simulations would not enlighten one in following the physical processes during
transition. A detailed review and critique of some of the methods used for DNS and
large eddy simulation (LES) of transitional and turbulent flows is given in [24].
This is of fundamental relevance for the contents of this book. Here, we present a
more fundamental approach that studies the topic by considering the flow evolution
from laminar to fully developed state by a thoroughly analyzed numerical method.
The aim is to computationally reproduce the physical phenomena as recorded in
classical transition experiments by Schubauer and Skramstad [19] for 2D transition
and Klebanoff et al. [10] for the 3D routes. These experiments were designed after lots
of soul searching in the fluid dynamic community, when experimentalists failed to
detect some wave-solutions predicted for instability studies by Heisenberg, Tollmien
and Schlichting. The reader is directed to the introductory discussion in [23] about
the historic development of the subject of instability studies and a brief account will
be provided later.
The qualitative features of stability of zero pressure gradient flow over a flat
plate has been studied in many complimentary ways, after the governing viscous
fluid flow equations were written down in the form of NSE in the first half of the
nineteenth century. The correctness of this equation is now well accepted in the
continuum regime of fluid flow. Apart from the fact that correct governing equations

© Springer Nature Singapore Pte Ltd. 2019 1


T. K. Sengupta and S. Bhaumik, DNS of Wall-Bounded Turbulent Flows,
https://doi.org/10.1007/978-981-13-0038-7_1
2 1 DNS of Wall-Bounded Turbulent Flow: An Introduction

are needed, one has also to ensure that correct boundary and initial conditions are used
in the solution process. One notes that the no-slip boundary condition is a modeling
approximation and has never been proven rigorously. Batchelor [2] has noted for
Newtonian fluid flow that the absence of slip at a rigid wall is now amply confirmed
by direct observation and by the correctness of its many consequences under normal
conditions. Despite this, we do not question the correctness of the no-slip condition
for continuum flows.
However, it became immediately apparent to Stokes that the closed-form solution
of NSE for circular pipe flow did not match with the experimental observation. This
led all the leading fluid dynamicists of the day to understand that mere existence
and uniqueness of the governing equation is not adequate. It is equally important
that such solutions should be stable, if the solution is to represent a physical flow
condition. This is due to the fact that such physical fluid dynamical systems are
always visited upon by omnipresent background disturbances. It is often likely that
such disturbances will destabilize the nominal solution, as obtained by solving the
governing equation. This nominal solution is called the equilibrium solution, as it
is obtained from the solution of the governing equation describing the dynamic
equilibrium of participating forces and moments. Obtaining the equilibrium solution
and testing it for its instability is central to most fluid dynamical studies for practical
systems [23]. The instability is related to growth of background disturbances, which
eventually is responsible for creation of turbulence.
To present the state of art on the subject topic, we focus upon the most fundamental
canonical problem, namely transition of fluid flow over a flat plate subjected to zero
pressure gradient. This zero pressure gradient flow can be simplified by boundary
layer assumption into an ordinary differential equation by similarity transformation,
as was performed by Blasius and the resultant velocity is given by the Blasius profile
[22]. In the present book, equilibrium flow will be obtained from the solution of
NSE, and not from the approximate boundary layer solution.

1.2 Why Deterministic Study Is More Relevant


than Stochastic Approaches?

The reason for numerically simulating deterministic experiments here, is on purpose.


In real life situation, the flow becomes turbulent due to stochastic inputs, which may
be difficult to quantify. Although some investigators have attempted using random
noise or trip to trigger transition, results of such simulations do not unambiguously
show route(s) to turbulence [18, 20, 33]. It is worthwhile noting that no proof exists
that turbulence is unique and hence making a flow turbulent by artificial numerical
means may simulate a broadband spectrum flow field, but whether it is the right
one is not certain. Also in many practical applications, the nature and extent of
transitional flow matters and knowledge about such flows are essential. For example,
in turbomachineries one notices a very large streamwise extent of transitional flows
1.2 Why Deterministic Study Is More Relevant than Stochastic Approaches? 3

in the turbine stages. To calculate fluid dynamic and thermal loads for such devices,
one needs exact knowledge of the flow field. This point of view is espoused in the
presentation here.
To understand transitional and turbulent flows, it is therefore paramount to under-
stand the flow physics accurately, by the applications of computational tools. Also
viewing the flow to be a dynamical system, where one can correlate causes with
effects; this helps us develop understanding of transitional and turbulent flows. In
this context, one appreciates the role of kernel experiments in exploratory studies
of fluid flow [31]. It is in this spirit, experiments at NBS helped the fluid dynamic
community understand the role of instability theory via the experiments in [10, 19]
for the canonical zero pressure gradient boundary layer.
For example, all earlier studies viewed it as a problem of linear hydrodynamic
instability. As the background disturbances are not always distinguishable due to
the infinitesimally small amplitudes, it is quite natural to study the instabilities by
linearizing the disturbance equations, and solving the resultant eigenvalue problem.
The disturbance equations are often obtained from the governing equation, by split-
ting the flow variable into equilibrium quantity and a disturbance component. Formal
application of perturbation theory to derive this equation is well described in [6, 23].
Despite this, in the presented approach, we will avoid linearization of the governing
equation and study fully the nonlinear instability and receptivity of the flow.

1.2.1 Historic Developments

At the beginning, leading scientists of the day in fluid dynamics, like Rayleigh,
Kelvin and Helmholtz proposed that one can study the onset of turbulence problem by
studying disturbance growth by considering the disturbance field to be inviscid, even
if the equilibrium state whose instability has to be studied, is obtained by including
viscous term(s) in governing equations. This was justified from the point of view that
the action of viscosity is essentially dissipative and will attenuate the disturbance. In
contrast, the inviscid equation will provide more critical estimate of flow instability
via disturbance growth. Instead of providing a quantitative measure of instability of
flows (whose growth in the streamwise direction was neglected), Rayleigh enunciated
a theorem which provided a necessary condition (and not sufficient!) for instability
to be that the velocity profile must have an inflection point inside the viscous region.
This is known as the Rayleigh’s inflection point theorem for the growth of disturbance
in time or temporal instability.
However at the onset of such an attempt for the zero pressure gradient flow over
flat plate, it was realized that this flow does not suffer temporal instability, as the cor-
responding velocity profile does not have an inflection point. This failure of temporal
inviscid instability led researchers to look for growing disturbance in space, a natural
extension of the temporal theory. Also for this canonical flow, it was noted that the
growing disturbances convect in space. However, it is easy to show that Rayleigh’s
stability equation does not account for spatial inviscid growth of disturbances.
4 1 DNS of Wall-Bounded Turbulent Flow: An Introduction

This necessitated inclusion of viscous terms in disturbance equation, which was pro-
posed by Orr and Sommerfeld [16, 32] independently, and is now known as
the Orr-Sommerfeld equation. It is now becoming more and more apparent, that the
action of viscous diffusion is not merely attenuating the solution, it can by itself con-
tribute to instability, as noted in combustion. Moreover, viscous actions can also add
to phase shift of components and under suitable conditions, such phase shift can make
the disturbance grow. Recently a theory of instability is being proposed which is based
on enstrophy. In many work on turbulence, researchers have identified enstrophy with
dissipation, based on work for homogeneous periodic flow.
It was Heisenberg [7] under the guidance of Sommerfeld, attempted to solve the
Orr-Sommerfeld equation for complex wavenumber (for variation in the streamwise
direction) as the eigenvalue, for prescribed Reynolds number (based on displace-
ment thickness of the equilibrium boundary layer) and real frequency of excitation.
Subsequently, Tollmien [34] and Schlichting [21], under the guidance of Prandtl,
also solved this eigenvalue problem analytically. All of these researchers predicted
the instability via growing wave-like solutions in the streamwise direction. This is
now known as Tollmien-Schlichting (TS) wave and has been the pacemaker of insta-
bility research for decades. We emphasize that the TS wave has been predicted for
an equilibrium flow, which does not grow in streamwise direction. This may hold
true for fully developed channel flow, but is strictly not correct for boundary layers.
Right from the beginning, it was assumed that the analysis can be performed locally
for boundary layers, with the current boundary layer thickness considered to be held
constant in the streamwise direction, for the purpose of solving the Orr-Sommerfeld
equation. In an actual growing boundary layer, obtained eigenvalue will continually
change for the complex wavenumber. The real part of this complex wavenumber
indicates the wavelength of the TS wave, while the imaginary part indicates growth
or attenuation rate of the TS wave. Thus, in conjunction with the local parallel flow
assumption, the TS wave solution would indicate a wave-packet, instead of a wave,
whose streamwise extent is determined by the imaginary part of the wavenumber.
This approach did not find immediate acceptance, as it was difficult to detect TS
wave experimentally. This situation prevailed till the experimental reporting of TS
wave by Schubauer and Skramstad [19] for 2D zero pressure gradient boundary layer
representing the equilibrium flow. The success of this experiment, in the light of con-
tinuing failure by others, was related to shifting of attention from natural transition
to forced excitation of the equilibrium flow by vibrating a ribbon inside the bound-
ary layer. To accentuate the relationship of cause and effect, a tunnel was designed
for extremely negligible background disturbances and then providing deterministic
excitation at a fixed frequency, as was the case of instability studies. Whereas the
natural disturbances are not monochromatic disturbances, these are not expected to
create TS wave of eigenvalue analysis.
Along with the successful reporting of 2D TS wave for 2D equilibrium flow,
natural extension of deterministic route of 2D transition was investigated, where
the imposed deterministic excitation was made three-dimensional (3D) and resultant
transition was investigated by Klebanoff et al. [10]. This was considered important,
as in most turbulent flows (excepting some flows, including atmospheric dynamics),
1.2 Why Deterministic Study Is More Relevant than Stochastic Approaches? 5

three-dimensionality is important due to vortex stretching. This work was followed by


many other forms of 3D instability mechanisms, which were also considered impor-
tant for transition to turbulence for the canonical zero pressure gradient boundary
layer [8].

1.3 Present State of Art in the Field

The reader is reminded that the experiment in [19] only demonstrated the exis-
tence of TS wave and did not show that TS wave eventually causes transition. This
was an implicit assumption right from the beginning of instability studies, which
presupposes that any instability waves eventually lead to causing turbulence. Over
the last two decades, it has been shown that apart from TS wave, there exists the
spatio-temporal wave-front (STWF). This has been made possible first by solving
the Orr-Sommerfeld equation in spatio-temporal framework [26, 28, 29]. This was
performed to show how a so-called linearly stable boundary layer supports mas-
sive wave-packets with sharp wave-fronts, something like tsunami or rogue waves.
The existence of STWF has also been established from the solution of NSE. Fur-
thermore, it has been shown subsequently that such a STWF via non-linear growth
causes 2D turbulence [25, 30]. The STWF results from finite start-up time of the
exciter in experiments/computations. However, once the STWF is set-up, it has been
shown that this structure has self regeneration mechanism [25]. One of the assuring
features of this finding is that the energy spectrum displays k −3 variation, as has been
predicted by Kraichnan [11] and Batchelor [1] for 2D turbulence.
After the experimental verification of TS wave [19], and reporting of experimental
results for 3D transition routes (see [8] for a review), computing these scenario
was reported by only a few groups of researchers. We will discuss the merits and
drawbacks of these efforts in greater details. In some of these, governing equation
itself has been changed; in some the numerical methods are flawed; while in some
of these the initial and boundary conditions were not correct. Problem arises when
generic codes have been adopted to simplify the simulation efforts and in the process
unphysical routes of transition to turbulence have been documented for very small
computational domains. Instead of noticing 3D TS wave and subsequent secondary
or nonlinear stages of transition in the experiments of [10], the readers have been
told that the simulations represent bypass transition (see for example, [18, 36–38] for
various reviews). While in most of the cited works, numerical artifacts created flow
fields, which look like turbulent flows. Authors in [3, 4] have solved the problem
for two primary routes of 3D transition described in the literature for wall excitation
in [10] without any numerical artifacts. Of specific interest will be a comparison
between the work in [18] with [4]. In the present book, all the essential details of of
the work in [3, 4] are provided with explanation.
6 1 DNS of Wall-Bounded Turbulent Flow: An Introduction

1.4 Different Transition Routes

So far we have discussed transition and turbulence, which is caused by wall exci-
tation. This type of transition is stated to be caused by wall modes [23] in the lin-
earized analysis, and in the same reference, transition caused by convecting vortices
in the free stream is said to be due to free stream modes. Experimentally, Lieb et
al. [13] have shown how these two types of modes are coupled and the theoretical
aspect of this coupling is given in [23] for the Orr-Sommerfeld equation. In the context
of nonlinear instability, solution of NSE displays a k −3 spectrum, when the Fourier
transform of streamwise disturbance velocity is plotted as a function of streamwise
wavenumber, for the inhomogeneous 2D flow past a flat plate subjected to periodi-
cally passing train of vortices at a fixed height. Thus one cannot distinguish between
wall and free stream excitation for 2D turbulence. The problem of vortex-induced
instability caused by a single convecting 2D vortex at a fixed speed and height over
the plate can cause instability at a streamwise location, where growing TS wave is
not seen by wall excitation [14, 27]. Thus, the vortex-induced instability is a sub-
critical phenomenon (discussed in Chap. 4). It was Morkovin [15] who coined the
term bypass transition, for all those cases where transition is not associated with the
creation of TS wave. Originally, it was noted that Couette flow, Poiseuille flow in
a pipe were noted to be stable by linear spatial theory and it was thought that such
flows, along with bluff body flows, suffer transition by bypass route. Even the crit-
ical Reynolds number obtained by linear theory for channel flow (Recr = 5772) is
far above the value for which the flow is noted to become turbulent (at Retr ≈ 1000).
Thus, it becomes apparent that apart from the zero pressure gradient boundary layer,
most flows do not even show the existence of TS wave. With varied pressure gradient,
as in flow over aerofoil, one does not notice monochromatic TS wave also. Finally,
even for zero pressure gradient boundary layer, one does not notice TS wave, when
the amplitude of wall excitation is significantly higher. However, the original con-
notation of flow instability is with respect to vanishingly small excitation. Thus, we
should explain bypass transition, when this can be attributed to physical processes and
not due to numerical means by which stable laminar flow is tripped to turbulent state
[18, 37]. The TS wave is thus noted only with extremely carefully designed exper-
iment and the resultant turbulence provides the true test of computational ability of
capturing turbulence by first being able to reproduce these deterministic disturbances
[10, 19]. In this respect the work reported in [3, 4] has shown the correct methodology
for DNS of turbulent flow. This approach is specifically described in this book.

1.5 Role of Equilibrium Flow in DNS

In discussing about the role of analytical solution, Landau and Lifshitz [12] noted that
the flow that occurs in nature must not only follow the equations of fluid dynamics, but
also be stable. If solutions are not observable, then the corresponding equilibrium
1.5 Role of Equilibrium Flow in DNS 7

flows are not stable. This also tells us about the implication of flow instability in
the context of observation of Stokes with respect to pipe flow experimental results,
following a series of simplifying assumptions. Moreover, the flow was considered
to be laminar, while the experimental results corresponded to fully developed tur-
bulent pipe flow. The mismatch between the two results were attributed to flow
instability in [12]. The laminar flow steady solution can be viewed as the equilibrium
solution, and the following instability caused the growth of infinitesimally small per-
turbations present in the surroundings. This sensitive dependence on the disturbance
environment makes the subject of instability very interesting, as well as challeng-
ing in understanding the state of the flow. The smallness of background disturbances
allows one to study the problem of growth of these by small perturbation theory. This
greatly helps, if the governing nonlinear equations can be solved for the equilibrium
solution with ease and then its stability can be studied by linearizing the governing
equation for the perturbation field. In theoretical instability approach, this leads to
an eigenvalue problem, as we have already discussed above for flow past a plate
experiencing zero pressure gradient. One of the drawback of eigenvalue analysis is
that this does not require any information on input, that triggers instability. At the
same time, Schubauer and Skramstad [19] have clearly noted in their experiment,
that the vibrating ribbon placed inside the boundary layer can create TS wave, while
an acoustic exciter placed outside the boundary layer did not cause any instability.
This brings in the concept of receptivity, by which one can say that wall-bounded
shear layer are receptive to vortical excitation (more strongly when it is placed inside
it) and not so receptive to acoustic excitation, when it is applied from outside. Nat-
urally, the amplitude of excitation should be a factor in determining the causation
of instability. Interestingly, this was noted by Osborne Reynolds and he recorded it
unambiguously [17].
The dye experiment performed [17] is perhaps the first recorded experimental
observations of flow instability. Reynolds took pipes of different diameters fitted
with a trumpet shaped mouth-piece or bell-mouth, which accelerated the flow on
entry to the pipe. Such acceleration creates a favorable pressure gradient, atten-
uating background disturbances accompanying the oncoming water in the pipe.
He performed experiments in night, to avoid noise from daytime vehicular traffic.
Reynolds observed that the rapid diffusion of dye with surrounding fluid depends on
the non-dimensional parameter V a/ν, with V as the center-line velocity in the pipe
of diameter a, and ν is the kinematic viscosity. This non-dimensional parameter is the
Reynolds number (Re). Reynolds found that the flow can be kept orderly or laminar
up to Re = 12, 830 in his set up. He noted this value to be extremely sensitive to the
disturbances in the flow before it enters the tube and he noted prophetically that this
at once suggested the idea that the condition might be one of instability for distur-
bance of a certain magnitude and stable for smaller disturbances. The relationship
of instability with disturbance amplitude is often attributed to non-linear instability.
The implicit assumption in this point of view is that linear instability does not require
any knowledge for the amplitude and spectrum of the input disturbances. This must
be viewed as a serious shortcomings of linear theory, which will be used only for
qualitative understanding of flow instability.
8 1 DNS of Wall-Bounded Turbulent Flow: An Introduction

When it comes to computational efforts in understanding instability and transition


to turbulence, one of the strengths lies in being able to obtain equilibrium flow with-
out making too many simplifying assumptions. On the other hand, no computations
can be made without the attendant numerical error. Thus, knowledge and ability to
reduce errors must be one of the central themes of DNS and LES. Since a steady
equilibrium flow becomes unsteady, before becoming fully turbulent, obtaining a
steady flow accurately is equivalent to obtaining the equilibrium flow correctly. At
the same time, we all understand that there are many flows, for range of parameter
values, which cannot be computed without the flow displaying inherent unsteadi-
ness. A simple example is a flow past a circular cylinder for Reynolds number in
excess of 65 or so, for which the computed flow display alternately shed vortices.
Vortex shedding itself is a manifestation of instability of steady laminar flow for
a critical Reynolds number. It is well to remember that neither computationally or
experimentally, one obtains a single fixed critical Reynolds number. In computations
and experiments, background disturbances in experiment and numerical errors act
as seed of instabilities for any flow, provided sufficient numerical accuracy is main-
tained. Numerical errors are due to discretization of continuum equations and for
computing space-time dependent problems, additional error arises due to coupling
of space and time discretizations considered together. Such a relationship for space
and time scales appear as dispersion relation for linear constant coefficient partial
differential equations. The coupled space-time discretization error is thus, termed as
dispersion error. It has been variously investigated for different model equation, for
the purpose of calibrating space-time discretization considered together. For other
flows, there may be other sources of errors, like aliasing error, Gibbs’ phenomenon
etc. which can also provide the seed to destabilize a given flow. Thus, following the
approach used in experiments on transition to turbulence, computationally also, one
must design numerical methods which systematically removes error sources, and
thereby we improve our ability to obtain equilibrium flows. Such equilibrium flows
should be studied for their receptivity to different generic types of errors, which
mimic acoustic, vortical and entropic disturbances.

1.6 What Is Instability?

To study a physical problem analytically, one first obtains governing equations which
model the phenomenon adequately. If the auxiliary equations pertaining to initial and
boundary conditions are well posed, then obtaining the solution is straightforward.
Mathematically, one is concerned with the existence and uniqueness of the solution.
Yet not every solution of equations of motion, even if it is exact, is observable
in nature. This is at the core of many physical phenomena where observability of
solution is of fundamental importance. If the solution is not observable, then the
corresponding basic flow is not stable. Here, the implication of stability is in the
context of the solution with respect to infinitesimally small perturbations.
1.6 What Is Instability? 9

In studying the stability of a problem with respect to ambient disturbances, it


is hardly ever possible that one can incorporate all the contributing factors in a
given physical scenario for posing a physical problem. Occasionally these neglected
causes can be incorporated by process noise and results are made to correlate with
the physical situation. This is possible when the causes are statistically independent
and then it follows upon using the Central Limit Theorem.

1.7 Temporal and Spatial Instability

Instability of an autonomous system is strictly for time-dependent systems that


display growth of disturbances with time. This may also mean that either we are
studying the stability of a flow at a fixed spatial location or the full system displays
identical variation in time for each and every spatial location. Fluid flow instabilities
are treated as if the disturbance growth is either in space or in time. This approach is
merely for expediency and not based on any assessment of reality. It is ideally suited
to treat the growth in a spatio-temporal framework, but no theoretical framework
exists, except by the use of the Bromwich contour integral method [23]. This method
is used in a linear framework and is capable of tracking growth in space and time
simultaneously. This method will be described and results shown in Chaps. 3 and
5. A disturbance originating from a fixed location in space can grow as it convects
downstream. Such a disturbance is termed unstable if it grows as it moves down-
stream. This is the convective instability. This type of instability is considered for
wall-bounded shear layers, exemplified by the classic vibrating ribbon experiment
of [19]. This experiment was performed in a very quiet facility to create TS waves
by vibrating a ribbon inside a flat plate boundary layer, a detailed description of
which is provided in Chap. 3. This experiment was the first to show the existence
of a viscous unstable wave which was predicted earlier theoretically in [7, 21, 34].
Hence, the existence of TS waves was doubted before the experimental results in [19]
were published. Additionally, this experiment was also the first to display receptivity
of wall-bounded shear layers to vortical disturbances created inside a shear layer,
while showing the inadequacy of acoustic excitation in creating TS waves. One of the
major aims of the book is to show that the classical picture of convective instability
is not correct, and the transition is shown to be created by STWF.
In many flows, it can happen that the disturbance can grow first in time at a fixed
location, before it is convected downstream. Such growth of disturbances both in
space and time are seen in many free shear layers and bluff-body flows. If we subject
the equilibrium solution of such an unstable fluid dynamical system to a localized
impulse, then the response field spreads both upstream and downstream of the loca-
tion with respect to the local flow where it originated, while growing in amplitude.
Such instabilities have been termed absolute instabilities. Here, an additional distinc-
tion needs to be made between convective and absolute instabilities.On application of
10 1 DNS of Wall-Bounded Turbulent Flow: An Introduction

an impulse, both situations display disturbances in upstream and downstream direc-


tions. However, in a convectively unstable system, the growth of the disturbance is
predominantly in one direction, while for an absolutely unstable system the growth
will be omni-directional.

1.8 Some Instability Mechanisms

Here two simple cases of instabilities are considered to emphasize the concepts
described above. We begin by distinguishing the difference between static and
dynamic instability by considering the stability of the atmosphere as an example.
When a parcel of air in the atmosphere is moved rapidly from an equilibrium
condition and its tendency to come back to its undisturbed position is noted, then
we term the atmosphere statically stable. The movement of the packet is considered
as impulsive, to preclude any heat transfer from the parcel to the ambience. This
tendency of static stability, when if exists, is due to the buoyancy force caused by
the density differential due to temperature variation with height and such a body
force acts upon the displaced air parcel. In static stability studies, we do not look
for detailed time-dependent motion of the parcel following the displacement (as the
associated accelerations are considered negligible). We refer the reader to detailed
dynamical instability of atmosphere studied in [23].

1.8.1 Kelvin–Helmholtz Instability

This arises when two layers of fluids (may not be of the same species or density)
are in relative motion. Thus, this is an interfacial instability and the resultant flow
features due to imposed disturbance will be much more complicated due to relative
motion. The physical relevance of this problem was seized upon by Helmholtz [35],
who observed that the interface as a surface of separation tears the flow asunder.
Sometime later Kelvin [9] posed this problem as one of instability and solved it. We
follow this latter approach here. The basic equilibrium flow is assumed to be inviscid
and incompressible - as two parallel streams having distinct density and velocity -
flowing one over the another, is depicted in Fig. 1.1.
Before any perturbation is applied, the interface is located at z = 0 and subsequent
displacement of this interface is expressed parametrically as

z s = η̂(x, y, t) = εη(x, y, t) (1.7)

where ε is a small parameter, defined to perform a linearized perturbation analy-


sis. One can view the interface itself as a shear layer of vanishing thickness. For
the considered inviscid irrotational flow, velocity potentials in the two domains are
given by
1.8 Some Instability Mechanisms 11

Interface

Fluid 2

x
Fluid 1

Fig. 1.1 Kelvin–Helmholtz instability at the interface of two flowing fluids

φ̃ j (x, y, z, t) = U j x + εφ j (x, y, z, t) (1.8)

The governing equations in either of the flow domains are given by

∇ 2 φ̃ j = 0 (1.9)

The potentials must satisfy the following far-stream boundary conditions,

φ j s are bounded as z → ±∞ (1.10)

Another set of boundary conditions is applied at the interface, which is the no-fluid
through the interface condition, i.e.,

∂ η̂ ∂ φ̃ j ∂ η̂ ∂ φ̃ j ∂ η̂ ∂ φ̃ j
− =− − (1.11)
∂t ∂z ∂x ∂x ∂y ∂y

In addition, in the absence of surface tension, pressure must be continuous across


the interface. Upon linearization, the interface boundary condition of Eq. (1.11)
simplifies to
∂η ∂η ∂φ j
+ Uj − = 0 for j = 1, 2 (1.12)
∂t ∂x ∂z

where φ̃ j and φ j are as related in Eq. (1.8). Defining the pressure on either flow
domain by the unsteady Bernoulli’s equation, one can write
 
∂ φ̃ j 1
pj = Cj − ρj + (∇ φ̃ j )2 + g η̂ (1.13)
∂t 2

Simplifying and retaining up to 0(ε) terms, we get the following conditions


12 1 DNS of Wall-Bounded Turbulent Flow: An Introduction

1 1
0(1) condition : C1 − ρ1 U12 = C2 − ρ2 U22 (1.14a)
2 2
   
∂φ1 ∂φ1 ∂φ2 ∂φ2
0(ε) condition : ρ1 + U1 + gη = ρ2 + U2 + gη (1.14b)
∂t ∂x ∂t ∂x

One can consider a very general interface displacement given in terms of a bilateral
Laplace transform as
 
η(x, y, t) = F(α, β, t) ei(αx+βy) dα dβ (1.15)

Correspondingly, the perturbation velocity potential is expressed as


 
φ j (x, y, z, t) = Z j (α, β, z, t) ei(αx+βy) dα dβ (1.16)

Writing k 2 = α 2 + β 2 and using Eq. (1.16) in Eq. (1.9), one gets the solution that
satisfies the far-stream boundary conditions of Eq. (1.10) as

Z j = f j (α, β, t) e±kz for j = 1 and 2 (1.17)

Using Eq. (1.15) in the interface boundary condition of Eq. (1.12) one gets

Ḟ + iαU1 F − k f 1 = Ḟ + iαU2 F + k f 2 = 0 (1.18)

where the dots denote differentiation with respect to time. If we denote the density
ratio ρ = ρ2 /ρ1 , then the linearized pressure continuity condition of Eq. (1.14b)
gives
∂φ1 ∂φ2 ∂φ1 ∂φ2
−ρ + U1 − ρU2 + (1 − ρ)gη = 0 (1.19)
∂t ∂t ∂x ∂x
Using Eqs. (1.15) and (1.16) in the above one gets

f˙1 − ρ f˙2 + iαU1 f 1 − iαρU2 f 2 + (1 − ρ)g F = 0 (1.20)

Eliminating f 1 and f 2 from Eq. (1.20) using Eq. (1.18), one gets, after simplifi-
cation,

(1 + ρ) F̈ + 2iα(U1 + ρU2 ) Ḟ − {α 2 (U12 + ρU22 ) − (1 − ρ)gk}F = 0 (1.21)

This ordinary differential equation for the time variation of the interface displace-
ment F can be understood better in terms of its Fourier transform defined by

F(., t) = F̂(., ω̄) ei ω̄t d ω̄ (1.22)
1.8 Some Instability Mechanisms 13

One obtains the following dispersion relation by substitution of Eq. (1.22) in


(1.21) as

− ω̄2 (1 + ρ) − 2α ω̄(U1 + ρU2 ) + (1 − ρ)gk − α 2 (U12 + ρU22 ) = 0 (1.23)

This provides the characteristic exponents in Eq. (1.22) as



α(U1 + ρU2 ) gk(1 − ρ 2 ) − α 2 ρ(U1 − U2 )2
ω̄1,2 =− ∓ (1.24)
(1 + ρ) (1 + ρ)

Based on this dispersion relation the following sub-cases are considered:


CASE 1: If the interface is disturbed in the spanwise direction only, i.e., α = 0,
then 
(1 − ρ)
ω̄1,2 = ∓ gβ (1.25)
(1 + ρ)

Thus, the streaming velocities U1 and U2 do not affect the response of the system.
If in addition, ρ > 1, i.e., a heavier liquid is over a lighter liquid, then the buoyancy
force causes temporal instability (if β is considered real) as is the case for Rayleigh-
Taylor instability (see Chandrasekhar [5]).
CASE 2: For a general interface perturbation if gk(1 − ρ 2 ) − α 2 ρ(U1 − U2 )2 <
0, then the interface displacement will grow in time. Thiscondition can be alternately
1−ρ 2
stated as a condition for instability as (U1 − U2 )2 > gk
α2 ρ
.
Thus, for a given shear at the interface given by (U1 − U2 ) and for a given oblique
disturbance propagation direction at the interface indicated by the wavenumber vector
k, instability would occur for all wavenumbers k ∗ , given by
 2  
∗ k∗ g ρ1 ρ2
k > −
α (U1 − U2 )2 ρ2 ρ1

Note that the wavenumber vector makes an angle γ with the x-axis, such that
cos γ = kα∗ and the above condition can be conveniently written as
 
∗ g ρ1 ρ2
k > − (1.26)
(U1 − U2 )2 cos2 γ ρ2 ρ1

The lowest value of wavenumber (k ∗ = kmin ) would occur for two-dimensional


disturbances, i.e., when cos γ = 1 and this minimum is given by
 
∗ g ρ1 ρ2
kmin = − (1.27)
(U1 − U2 )2 ρ2 ρ1
14 1 DNS of Wall-Bounded Turbulent Flow: An Introduction

CASE 3: Consider the case of shear only of the same fluid in both domains, i.e.,
ρ = 1. The characteristic exponents then simplify to

U1 + U2 iα
ω̄1,2 = −α ∓ (U1 − U2 ) (1.28)
2 2
The presence of an imaginary part with a negative sign implies temporal instability
for all wave lengths. Also, note that since the group velocity and phase speed in the
y direction are identically zero, therefore the Kelvin–Helmholtz instability for pure
shear always will lead to two-dimensional instability.

References

1. Batchelor, G. K. (1969). Computation of the energy spectrum in homogeneous two-dimensional


decaying turbulence. Physics of Fluids, 12(suppl. II), 233–239.
2. Batchelor, G. K. (1988). An introduction to fluid dynamics. UK: Cambridge University Press.
3. Bhaumik, S. (2013). Direct numerical simulation of inhomogeneous transitional and turbulent
flows. Ph. D. Thesis, I. I. T. Kanpur.
4. Bhaumik, S., & Sengupta, T. K. (2014). Precursor of transition to turbulence: Spatiotemporal
wave front. Physical Review E, 89(4), 043018.
5. Chandrasekhar, S. (1960). Radiative transfer (p. 393). New York: Dover Publications Inc.
6. Drazin, P. G., & Reid, W. H. (1981). Hydrodynamic stability. UK: Cambridge University Press.
7. Heisenberg, W. (1924). Über stabilität und turbulenz von flüssigkeitsströmen. Annalen der
Physik (Leipzig), 379, 577–627 (Translated as ‘On stability and turbulence of fluid flows’.
NACA Tech. Memo. Wash. No 1291 1951)
8. Kachanov, Y. S. (1994). Physical mechanisms of laminar-boundary-layer transition. Annual
Review of Fluid Mechanics, 26, 411–482.
9. Kelvin, L. (1871). Hydrokinetic solutions and observations. Philosophical Magazine, 4(42),
362–377.
10. Klebanoff, P. S., Tidstrom, K. D., & Sargent, L. M. (1962). The three-dimensional nature of
boundary-layer instability. Journal of Fluid Mechanics, 12, 1–34.
11. Kraichnan, R. H. (1967). Inertial ranges in two-dimensional turbulence. Physics of Fluid, 10(7),
1417–1423
12. Landau, L. D., & Lifshitz, E. M. (1959). Fluid mechanics (Vol. 6). London: Addison - Wesley.
Pergamon Press.
13. Leib, S. J., Wundrow, D. W., & Goldstein, M. E. (1999). Effect of free stream turbulence and
other vortical disturbances on a laminar boundary layer. Journal of Fluid Mechanics, 380,
169–203.
14. Lim, T. T., Sengupta, T. K., & Chattopadhyay, M. (2004). A visual study of vortex-induced
subcritical instability on a flat plate laminar boundary layer. Experiments in Fluids, 37, 47–55.
15. Morkovin, M. V. (1991). Panoramic view of changes in vorticity distribution in transition insta-
bilities and turbulence. D.C. Reda, H.L. Reed & R. Kobayashi (Eds.), Transition to turbulence
(Vol. 114, pp. 1–12) ASME FED Publication.
16. Orr, W. M. F. (1907). The stability or instability of the steady motions of a perfect liquid and
of a viscous liquid. Part I: A perfect liquid. Part II: A viscous liquid. Proceedings of the Royal
Irish Academy, A27, 9–138.
17. Reynolds, O. (1883). An experimental investigation of the circumstances which determine
whether the motion of water shall be direct or sinuous and of the law of resistance in parallel
channels. Philosophical Transactions of the Royal Society, 174, 935–982.
References 15

18. Sayadi, T., Hamman, C. W., & Moin, P. (2013). Direct numerical simulation of complete H-
type and K-type transitions with implications for the dynamics of turbulent boundary layers.
Journal of Fluid Mechanics, 724, 480–509.
19. Schubauer, G. B., & Skramstad, H. K. (1947). Laminar boundary layer oscillations and the
stability of laminar flow. Journal of Aerospace Sciences, 14(2), 69–78.
20. Schlatter, P., & Örlü, R. (2012). Turbulent boundary layers at moderate Reynolds numbers:
Inflow length and tripping effects. Journal of Fluid Mechanics, 710, 5–34.
21. Schlichting, H. (1933). Zur entstehung der turbulenz bei der plattenströmung. Nach. Gesell. d.
Wiss. z. Gött., MPK, 42, 181–208.
22. Schlichting, H. (1979). Boundary layer theory (7th ed.). New York: McGraw Hill.
23. Sengupta, T. K. (2012). Instabilities of flows and transition to turbulence. Florida, USA: CRC
Press, Taylor & Francis Group.
24. Sengupta, T. K. (2015). A critical assessment of simulations for transitional and turbulent
flows. In T. K. Sengupta, S. K. Lele, K. R. Sreenivasan & P. A. Davidson (Eds.), In the IUTAM
Symposium Proceedings on Advances in Computation, Modeling and Control of Transitional
and Turbulent Flows (pp. 491–532). World Scientific Publishing Company.
25. Sengupta, T. K., & Bhaumik, S. (2011). Onset of turbulence from the receptivity stage of fluid
flows. Physical Review Letters, 154501, 1–5.
26. Sengupta, T. K., Ballav, M., & Nijhawan, S. (1994). Generation of Tollmien-Schlichting waves
by harmonic excitation. Physics of Fluids, 6(3), 1213–1222.
27. Sengupta, T. K., De, S., & Sarkar, S. (2003). Vortex-induced instability of an incompressible
wall-bounded shear layer. Journal of Fluid Mechanics, 493, 277–286.
28. Sengupta, T. K., Rao, A. K., & Venkatasubbaiah, K. (2006). Spatio-temporal growing wave
fronts in spatially stable boundary layers. Physical Review Letters, 96(22), 224504.
29. Sengupta, T. K., Rao, A. K., & Venkatasubbaiah, K. (2006). Spatio-temporal growth of dis-
turbances in a boundary layer and energy based receptivity analysis. Physics of Fluids, 18,
094101.
30. Sengupta, T. K., Bhaumik, S., & Bhumkar, Y. (2012). Direct numerical simulation of two-
dimensional wall-bounded turbulent flows from receptivity stage. Physical Review E, 85(2),
026308.
31. Smith, C. R. (1993). Use of ‘Kernel’ experiments for modeling of near-wall turbulence. In
R.M.C. So, C.G. Speziale & B.E. Launders (Eds.), Near wall turbulent flows (pp. 33–42).
Amsterdam, Holland: Elsevier.
32. Sommerfeld, A. (1908). Ein Beitrag zur hydrodynamiscen Erklarung der turbulenten Flus-
sigkeitsbewegung. In Proceedings of the 4th International Congress of Mathematicians, Rome
(pp. 116–124).
33. Spalart, P. (1988). Direct numerical study of leading edge contamination. AGARD CP, 438,
5.1–5.13.
34. Tollmien, W. (1931). Über die enstehung der turbulenz. I, English translation. NACA TM 609.
35. von Helmholtz, H. (1868). On discontinuous movements of fluids. Philosophical Magazine,
36(4), 337–346.
36. Wu, X. (2017). Inflow turbulence generation methods. Annual Review of Fluid Mechanics, 49,
23–49.
37. Wu, X., & Moin, P. (2009). Direct numerical simulation of turbulence in a nominally-zero-
pressure-gradient flat-plate boundary layer. Journal of Fluid Mechanics, 630, 5–41.
38. Wu, X., Moin, P., & Hickey, J. P. (2014). Boundary layer bypass transition. Physics of Fluids,
26, 091104.
Chapter 2
DNS of Navier–Stokes Equation

2.1 Fluid Dynamical Equations

The governing equations of fluid mechanics are all based on identical fundamen-
tal classical principles of dynamics, namely conservation of mass, momentum and
energy. Based on these principles, Claude-Louis Navier and George Gabriel Stokes,
derived the governing equation for viscous fluids by applying Newton’s second law
to fluid motion, along with the assumption that stresses arising in the fluids are due
to diffusing viscous effects and the pressure gradient. These governing equations
are known as Navier–Stokes Equation (NSE). It is to be noted that there is no fixed
version of NSE, an appropriate version and the formulation has to be chosen based
on the need and the characteristics of the fluid-dynamical problem to be considered.
For example, for low-speed applications without heat-transfer, the appropriate ver-
sion is the incompressible NSE, without considering the energy equation. For such
equations, it can be shown that the kinetic energy of the fluid in a control volume is
automatically conserved from the momentum equation. For flows dominated by heat-
transfer, an additional energy equation is to considered which essentially determines
the temperature at each point. Temperature gradient induces gradient in fluid-density,
which in turn affects the flow if gravitational effects are also present. In incompress-
ible NSE, such buoyancy effects due to temperature induced density-gradient are
generally modeled via Boussinesq approximation [87]. In contrast, when the flow
speed is on the higher side and the compressibility effects are quite dominant, the
appropriate governing equations are the compressible verisons of NSE, where all
the conservation constraints (mass, momentum and energy) are to be considered.
Additional equations are also required for compressible NSE to close the system
of equations which is given by the thermodynamic equation of state relating fluid
temperature, density and pressure. To derive the appropriate governing equations,
following general steps are to be followed:

© Springer Nature Singapore Pte Ltd. 2019 17


T. K. Sengupta and S. Bhaumik, DNS of Wall-Bounded Turbulent Flows,
https://doi.org/10.1007/978-981-13-0038-7_2
18 2 DNS of Navier–Stokes Equation

(i) First, appropriate physical principles have to be decided, which are required to
be satisfied, e.g., if one focuses upon the incompressible flow without heat transfer,
then only conservation principles of interest are mass and momentum.
(ii) Based on (i), an appropriate model for the flow problem has to be chosen.
(iii) Finally, based on (i) and (ii), appropriate governing mathematical equations
should be derived. Once, appropriate governing equations are derived, suitable com-
putational tools are to be used to numerically solve these equations, as generally these
equations do not admit any closed–form analytical solution (save some very few sim-
plified cases). Development and use of highly accurate numerical methodologies are
very important, particularly for transitional and turbulent flows. These computations
require special approaches, which are distinctly different from the conventional CFD
methodologies one usually comes across for engineering applications.
For proper characterization of transitional flows, one has to accurately track dis-
turbance evolution in a base or equilibrium flow, whose properties would determine
the characteristics of amplification and propagation of perturbation. The equilibrium
flow can be obtained by considering the full nonlinear NSE or some simplified form
of it. Historically, even inviscid irrotational assumptions are also made to obtain an
equilibrium flow. However, recent efforts have shown the importance of using full
nonlinear NSE to obtain the equilibrium flows, especially in the context of analyz-
ing its receptivity to imposed excitations. Typically, there are two ways by which
traditionally transitional flows are analyzed. First, through eigenvalue approach by
considering corresponding linearized NSE. In this approach, which have been the
focus of most of the earlier studies in the field, one is not required to characterize
the disturbance environment. In contrast, to study various aspects of the response
of a dynamical system to a specific class of disturbances, one adopts what has been
identified as receptivity analysis. Receptivity analysis can also be performed in the
linearized framework and has been advanced in [26, 30, 31] via the Bromwich con-
tour integral method. Irrespective of the amplitude of perturbations, one can also use
the full NSE for receptivity study. This is undertaken here in Chaps. 3–6. Thus, it
is imperative that readers appreciate the nuances of solution techniques of various
space-time dependent equations.
Unlike in solid mechanics problems, fluid dynamical systems are characterized
by a relatively larger number of degrees of freedom, as the later is characterized
by very low physical dissipation. However, it cannot be completely ignored from
the whole flow domain, as the delicate balance of dissipation and energy transfer to
other spectral components determines the characteristics of transitional and turbulent
flows. In the spectral plane the maximum dissipation occurs at higher wavenumbers
corresponding to smallest eddies. Thus, for simulating transitional flows one would
have to be very careful in resolving very high wavenumbers. The importance of this
is often lost on CFD practitioners, who insist on performing large eddy simulations
(LES) of transitional flows, by considering only the energy spectrum.
The conservation principles for mass, momentum and energy for fluid flows are
illustrated next.
2.1 Fluid Dynamical Equations 19

2.1.1 Equation of Continuity

This follows from the conservation of mass, which for a control volume essentially
states that net rate of creation/destruction of mass inside it is balanced by the amount
of mass flow rate through the control surfaces. For unsteady compressible flows, this
leads to
Dρ → ∂ρ
− −

+ρ ∇ · V = + ∇ · (ρ V ) = 0 (2.1)
Dt ∂t
where D/Dt stands for the material derivative, i.e.,
 
D ∂ −

= + V ·∇
Dt ∂t

For incompressible flows, density is treated as constant i.e., ρ = constant and sub-
sequently Eq. (2.1) simplifies to


∇ ·(V ) = 0 (2.2)

Any vector field that satisfies the equivalent condition given by Eq. (2.2) is called
the divergence-free or solenoidal field.

2.1.2 Momentum Conservation Equation

Similarly, this stems from the principle of conservation of translational momentum


following Newton’s second law of motion applied to a control volume. Following a
control volume in motion is essentially like considering a control-mass system for
which the mass does not change with time. Hence, the rate of change of momentum
inside such a control-mass system is basically the mass multiplied by the acceleration
experienced by the fluid. Therefore, net rate of change of momentum inside the
control volume is balanced by summation of net force exerted on the control volume
and the total rate of of momentum crossing (entering or leaving) the corresponding
control surface. The constituent forces acting on the elementary control mass are
body forces acting on the control volume (which do not depend on the geometry of
the body, e.g., gravitational/buoyancy/electromagnetic forces) and the surface forces
acting directly on the control surfaces caused by normal and shear stresses. For an
elementary volume element of size (d x d y dz), if the local density is ρ, then the
body force acting in the x-direction can be expressed as

Fx = ρ f x (d x d y dz) (2.3)

where f x is the associated acceleration in the x-direction. If τi j denotes the stress


tensor acting in the j-direction on a plane whose normal is in the i-direction, then
20 2 DNS of Navier–Stokes Equation

one can account for all the contributory stresses giving rise to surface force in the
x-direction as  
∂ p ∂τx x ∂τ yx ∂τzx
− + + + d x d y dz (2.4)
∂x ∂x ∂y ∂z

Net force term acting on the fluid element along x-direction is, therefore, the vector
summation of Eqs. (2.3) and (2.4). As the mass of the fluid element is (ρ d x d y dz)
and the acceleration of the moving element is given by its substantive or the material
derivative of the velocity vector, the momentum conservation equation for the fluid
element along x-direction can, therefore, be expressed as
 
Du ∂ p ∂τx x ∂τ yx ∂τzx
ρ = − + + + + ρ fx (2.5)
Dt ∂x ∂x ∂y ∂z

Similarly y- and z-components of the momentum equation can also be obtained


as  
Dv ∂ p ∂τx y ∂τ yy ∂τzy
ρ = − + + + + ρ fy (2.6)
Dt ∂y ∂x ∂y ∂z
 
Dw ∂ p ∂τx z ∂τ yz ∂τzz
ρ = − + + + + ρ fz (2.7)
Dt ∂z ∂x ∂y ∂z

Equations (2.5)–(2.7) are known as the Cauchy equations, which in vector form
can be written as


DV −

ρ = −∇ p + ∇ · τi j δi j + ρ F (2.8)
Dt
where δi j is the Kronecker delta function. By using the vector form of the continuity
equation, the x-component of the Cauchy equations in the conservation form, can
be expressed as
 
∂(ρu) −
→ ∂ p ∂τx x ∂τ yx ∂τzx
+ ∇ · (ρ u V ) = − + + + + ρ fx (2.9)
∂t ∂x ∂x ∂y ∂z

Similarly, one can also express the other two components of the momentum equa-
tion in conservative form. For these system of equations to be useful, one needs
information about the general stress system in the flow field. For Newtonian fluids,
Stokes [109] has shown that stress is related to the strain rates as
 
∂vk ∂vi ∂v j
τi j = λδi j +μ + − pδi j (2.10)
∂ xk ∂x j ∂ xi

where μ is the molecular viscosity, λ is the second coefficient of viscosity and p is


the thermodynamic pressure or hydrostatic stress. If pm is the mechanical pressure,
2.1 Fluid Dynamical Equations 21

then by definition pm = τ3ii . Stokes [109] again hypothesized a relationship between


μ and λ by relating thermodynamic and mechanical pressure from

3λ + 2μ = 0 (2.11)

However, in recent times, various researchers have shown that this hypothesis have
serious shortcomings [104]. Stokes’ hypothesis states that bulk viscosity κ̄ = λ + 23 μ
is set to zero for any flow following the assumption that for compressible flows,
thermodynamic pressure and mechanical pressures are identical [15, 26, 27, 35].
However, several researchers have criticized this assumption [15, 26, 27, 35, 69,
103]. Emanuel [26] and Buresti [15] have weakly justified Stokes’ hypothesis for
mono-atomic rigid molecule gases. However, Liebermann [57], Karim and Rosen-
head [47] and Rosenhead [79] have opined that λ is independent of μ and also, it can
be orders of magnitude higher in amplitude with a reverse sign. Emanuel [26] and
Gad-el-Hak [27] have also questioned the Stokes’ hypothesis and its applicability for
flows such as re-entry into planetary atmosphere. Cramer [19] have shown that many
common fluids including diatomic gases can have bulk viscosities (κ̄) thousand times
larger than λ, prompting Rajagopal [69] to suggest that new approach in obtaining
NSE for all fluids should be adopted without using the simple relationship given by
Eq. (2.11), between λ and μ.
Using the constitutive relation given in Eq. (2.10), one can write the final form of
the NSE for compressible flow as

∂(ρu) ∂(ρu 2 ) ∂(ρuv) ∂(ρuw) ∂p


+ + + =− +
∂t ∂x ∂y ∂z ∂x
       
∂ −
→ ∂u ∂ ∂v ∂u ∂ ∂u ∂w
λ∇ · V + 2μ + μ + + μ + + ρ fx
∂x ∂x ∂y ∂x ∂y ∂z ∂z ∂x
(2.12)

∂(ρv) ∂(ρuv) ∂(ρv2 ) ∂(ρvw) ∂p


+ + + =− +
∂t ∂x ∂y ∂z ∂y
       
∂ −
→ ∂v ∂ ∂v ∂u ∂ ∂v ∂w
λ∇ · V + 2μ + μ + + μ + + ρ fy
∂y ∂y ∂x ∂x ∂y ∂z ∂z ∂y
(2.13)

∂(ρw) ∂(ρuw) ∂(ρwv) ∂(ρw2 ) ∂p


+ + + =− +
∂t ∂x ∂y ∂z ∂z
       
∂ −
→ ∂w ∂ ∂w ∂u ∂ ∂v ∂w
λ∇ · V + 2μ + μ + + μ + + ρ fz
∂z ∂z ∂x ∂x ∂u ∂y ∂z ∂y
(2.14)
22 2 DNS of Navier–Stokes Equation

For incompressible flows, these equations can be further simplified. As for incom-


pressible flows ∇ · V = 0, the terms associated with the bulk viscosity coefficient
κ̄ drop out in Eqs. (2.12)–(2.14). Further, for incompressible flows with very mild
or almost absent temperature variation, μ can be treated as a constant. Using these
assumptions, the vector form of the NSE for incompressible flow is obtained as


∂V −
→ −
→ ∇p −
→ − →
+ ( V · ∇) V = − + ν ∇2 V + F (2.15)
∂t ρ

This is also called the equations in primitive variables or the primitive variable
formulation of incompressible NSE.

2.1.3 Energy Conservation Equation

Here, this is the first law of thermodynamics stated for a control volume system which
is: The rate of change of energy inside the control volume must be the summation
of heat transfer across the control surface and the work done by body and surface
forces. Detailed derivations can be noted in many text books on fluid mechanics and
CFD; see, e.g., [81]. The constituent terms of the energy equation are illustrated next.
Terms Due to Work Done by Body and Surface Forces:
The rate of work done by body forces for the control volume of mass ρ (dx dy dz) is
given in a mixed form as
 

→ ∂ −
→ − →
−∇ · ( p V ) + (u i τ ji ) d x d y dz + ρ F · V d x dy dz (2.16)
∂x j

Terms Due to Heat Transfer:


The net flux of heat is due to volumetric heating such as absorption or emission of
radiation and heat transfer across the control surface due to thermal conduction. If
one defines the rate of volumetric heat addition per unit mass as q̇, then the volumetric
heating of the element is
= ρ q̇ d x dy dz (2.17)

The directional conductive heat transfer through the control surfaces is related to
corresponding temperature gradient, using Fourier’s law by

∂T
q˙j = −κ
∂x j

where κ is the thermal conductivity. Hence the total heat interaction term is obtained
by using Newton’s law to provide
2.1 Fluid Dynamical Equations 23
  
∂ ∂T
ρ q̇ + κ d x d y dz (2.18)
∂x j ∂x j

Terms Due to the Rate of Change of Energy:


2 2
For a moving fluid element, E = e + V2 , where E, e and V2 represent total, internal
and kinetic energy per unit mass. The time rate of change of E is given by the
substantive derivative as  
D V2
ρ e+ d x d y dz (2.19)
Dt 2

The Final Form:


The final form of the energy equation is obtained by collating all the terms illustrated
above to obtain the non-conservation or convective form of the energy equation as
    
D V2 ∂ ∂T −
→ ∂ −
→ − →
ρ e+ = ρ q̇ + κ − ∇ · (p V ) + (u i τ ji ) + ρ F · V
Dt 2 ∂x j ∂x j ∂x j
(2.20)

2.1.4 Alternate Forms of the Energy Equation

Occasionally, the energy equation is written in terms of internal energy only. This is
obtained as
   2  2   
De ∂ ∂T −
→ −
→ ∂u ∂v ∂w 2
ρ = ρ q̇ + κ − p∇ · V + λ(∇ · V )2 + 2μ + +
Dt ∂x j ∂x j ∂x ∂y ∂z

      
∂u ∂v 2 ∂u ∂w 2 ∂v ∂w 2
+μ + + + + + (2.21)
∂y ∂x ∂z ∂x ∂z ∂y

The energy equation in conservation form is obtained from the above by noting

De ∂ −

ρ = (ρe) + ∇ · (ρe V )
Dt ∂t
Terms involving λ and μ constitute the dissipation term Φ0 in Eq. (2.21) .

2.1.5 Vorticity Transport Equation for Incompressible Flows

In fluid flows, vorticity, −



ω is defined as the curl of the velocity field, ı.e., −

ω =
−→
∇ × V . One of the problems of primitive variable formulation is the absence of
24 2 DNS of Navier–Stokes Equation

unique and definitive boundary conditions for pressure and this is often avoided by
using an alternative formulation that does not have explicit pressure dependent terms.
For NSE corresponding to incompressible flows, one can eliminate it by taking a curl
of Eq. (2.15) and using the kinematic relation between vorticity and velocity. This
gives the non-conservative form of vorticity transport equation (VTE) as

∂−
→ω −
→ −

+ ( V · ∇)−
→ ω · ∇) V + ν∇ 2 −
ω = (−
→ →
ω (2.22)
∂t
Note that the first term on the right hand side of Eq. (2.22) is the vortex stretch-


ing term, which is identically zero for 2D flows. Using the vector identity ∇ × ( A ×

→ −
→ −
→ −
→ −
→ −
→ −
→ −
→ −

B ) = A (∇ · B ) − B (∇ · A ) + ( B · ∇) A − ( A · ∇) B and divergence-free


condition on velocity and vorticity (Dv = ∇ · V = 0 and Dω = ∇ · − →
ω = 0), one
gets the Laplacian form of VTE as,

∂−
→ω −

ω × V ) = ν∇ 2 −
+ ∇ × (−
→ →
ω (2.23)
∂t
The viscous term in the Laplacian form of VTE can be further modified by noting
that
∇2−
→ω = ∇(∇.− →
ω ) − ∇ × (∇ × −
→ω)

Thus, ∇ 2 −→ω = −∇ × (∇ × − →ω ), since ∇ · −→


ω = 0.
Modifying the right hand side of Eq. (2.23) using the above relation, one gets the
rotational form of VTE as,

∂−
→ω −

+ ∇ × (−

ω × V ) + ν∇ × (∇ × −

ω)=0 (2.24a)
∂t

→ −

which can also be written in a concise form, by denoting H = (−

ω × V + ν∇ × −→
ω)
and thus,
∂−
→ω −

+∇ × H =0 (2.24b)
∂t
One notes that ideally the vorticity field should always be divergence free, i.e.,


Dω = ∇ · Ω = 0, which comes from the kinematic relation between velocity and
vorticity. For 2D flows, this condition is always satisfied (Dω = ∇ · − →ω = 0). How-
ever, for numerical computation of 3D flows in derived variable formulation, this is
not automatically guaranteed. For wall-bounded flows, vorticity is continually gen-
erated at the no-slip wall and if the generated wall vorticity is not made divergence
free, this will be a continual source of error, if left unchecked. Therefore, one needs
to chose judiciously the form of VTE for 3D simulations. This can be performed
by deriving the governing equation for Dω for the three forms of VTE as given in
Eqs. (2.22)–(2.24a).
2.1 Fluid Dynamical Equations 25

For the non-conservative form of the VTE (Eq. (2.22)), the evolution equation of
Dω is given as

∂ Dω −→ 1 2
+ V · (∇ · Dω ) − −

ω · (∇ · Dv ) = ∇ Dω (2.25a)
∂t Re
One readily identifies that this evolution equation of Dω represents a unsteady
convection-diffusion equation with a source term given as Sω = − →
ω · (∇ · Dv ). In
order to have the source term Sω = 0, one must have Dv = 0 at all the points inside
the computational domain. The necessary and essential conditions to ensure Dω = 0
in the computational domain for all times are, (i) Dω = 0 at t = 0 in the complete
domain, (ii) Dω = 0 on the boundary of the domain for all t > 0 and (iii) the diver-
gence of velocity Dv is zero for all t > 0 in the full domain. The evolution equation
of Dω for the Laplacian form of VTE (Eq. (2.23)), represents a unsteady diffusion
equation given as
∂ Dω 1 2
= ∇ Dω (2.25b)
∂t Re
Hence, the satisfaction of only first two conditions are necessary for Dω to be zero
for all times inside the computational domain. For the rotational form of VTE given
by Eq. (2.24a), the evolution equation of Dω is given simply by,

∂ Dω
=0 (2.25c)
∂t
which requires only the satisfaction of the first condition to ensure that Dω = 0 for
t ≥ 0, inside the computational domain. Hence, one notes that the rotational form of
VTE is much superior to the other two forms and requires very much less stringent
condition to preserve the solenoidality condition on vorticity vector for 3D flows.
In velocity-vorticity formulation of the incompressible NSE, one augments the
transport equations by deriving an auxiliary relation from the kinematic definition of
vorticity. For example, taking a curl of vorticity expressed in the inertial frame, one
obtains


∇ 2 V = −∇ × − →ω (2.26)

Any form of VTE given in Eqs. (2.22)–(2.24a) along with the velocity Pois-
son equation, Eq. (2.26), constitute the governing velocity-vorticity equations in the
derived variable formulation. In velocity–vorticity formulation of incompressible
NSE, the satisfaction of the divergence free condition on velocity (i.e., Dv = 0)
is subject to the accuracy up to which Eq. (2.26) is solved [8]. Therefore, some
researchers have not used Eq. (2.26) for computing the velocity field. For example,
authors in [31–33, 65] have directly used the continuity equation and the kinematic
definition of vorticity, along with the VTE. However, such approaches lead to an
over–determined system of linear algebraic set of equations and consequently, spe-
cial treatment is required to solve these numerically.
26 2 DNS of Navier–Stokes Equation



Another approach is the use of vector potential Ψ , instead of the velocity vector.

→ −
→ −

Vector potential Ψ is defined such that the velocity field V is the curl of Ψ i.e.,

→ −

V =∇×Ψ (2.27)

→ −

One notes that for any given velocity field V , if an uniquely defined Ψ is obtained


then Dv = 0. Taking curl of Eq. (2.27) and also imposing the restriction that Ψ is

→ −

divergence free, i.e., ∇ · Ψ = 0, one gets the equation for Ψ as


∇ 2 Ψ = −−

ω (2.28)

In 2D, only relevant component of both vector potential and vorticity is the one which
is normal to the plane of the flow and consequently, these equations are simplified
to yield stream function-vorticity formulation ((ψ, ω)-formulation) for 2D flows.
For 2D flows, the (ψ, ω)-formulation possess certain significant advantages over the
primitive variable formulation. These are due to (i) lesser number of unknowns as


compared to three variables in ( p, V )-formulation; (ii) exact satisfaction of mass
conservation everywhere in the flow field; (iii) obtaining the primary quantity, namely
the vorticity, directly from the governing equation and not by numerical differentia-
tion of computed quantities and (iv) removing the ambiguity regarding the prescrip-
tion of a suitable boundary condition for pressure. However, it is not straight-forward


to specify boundary conditions in terms of ψ for 3D problems, in general. A gen-


eral mathematical formulation of the boundary conditions on ψ for 3D flows are
given in Hirasaki and Hellums [40, 41]. Solution of 3D NSE using vector potential


( Ψ ) - vorticity (−

ω ) formulation for the flows are solved in Wong and Reizes [129] for
3D duct flows for Re = 10 and 100; for cubic lid-driven cavity (LDC) at Re = 500
and 3200 in Weinan and Liu [126]; for computing nonlinear stability of rotating
Hagen–Poiseuille flow (RHPF) in Ortega-Casanova and Fernandez-Feria [64]; in
Holdeman [43] for cubic LDC problem at Re = 400 and fully developed flow inside
3D open duct problems using finite element method. The difficulty in prescribing the


boundary condition in Ψ for multiply-connected domains are addressed in Wong and
Reizes [130], where an alternative is proposed to determine the boundary condition,
by introducing an over-relaxation factor. This approach was shown to accelerate
the convergence of several stable free and forced convection cases inside annular
cavities.


Equation (2.28) is derived based on the assumption that ∇ · Ψ = 0. However, it

→ −

is not guaranteed for the solution of Ψ obtained from Eq. (2.28). For Ψ obtained
from Eq. (2.28) to be divergence-free, the associated vorticity field also needs to be


solenoidal. If the computed Ψ is not solenoidal, then following Helmholtz decom-


position, one can express Ψ as


→ −

Ψ = ∇Φ + Ψ̃
2.1 Fluid Dynamical Equations 27



where Ψ̃ is a solenoidal vector field such that at the boundary of the domain the

→ −

normal components of Ψ and ψ̃ are identical. Using vector identity, ∇ × ∇Φ = 0,

→ −
→ −

one readily notes from this relation that V = ∇ × Ψ = ∇ × Ψ̃ . Therefore, even
if the computed vector potential field is not divergence-free, it does not affect the
computation of the associated velocity field.

2.1.6 Derived Variable Formulation for 2D Incompressible


Navier–Stokes Equation

For 2D incompressible flows, stream-function and vorticity has only one component,
which is perpendicular to the plane of the flow. The non-dimensional governing 2D
Navier–Stokes equation in (ψ, ω)-formulation is given as

∂ω ∂ψ ∂ω ∂ψ ∂ω 1 2
+ − = ∇ ω (2.29)
∂t ∂y ∂x ∂x ∂y Re

∇ 2 ψ = −ω (2.30)

Equations (2.29) and (2.30) are the VTE and stream-function equation (SFE), respec-
tively. Often one need to solve these equations in transformed computational domain,
so that effects of grid-point clustering at suitable locations can be taken care of. In
the transformed computational (ξ, η)-plane, Eqs. (2.29) and (2.30) convert to the
following form [81, 83]
    
∂ω ∂ψ ∂ω ∂ψ ∂ω 1 ∂ h 2 ∂ω ∂ h 1 ∂ω
h1h2 + − = + (2.31)
∂t ∂η ∂ξ ∂ξ ∂η Re ∂ξ h 1 ∂ξ ∂η h 2 ∂η

   
∂ h 2 ∂ψ ∂ h 1 ∂ψ
+ = −h 1 h 2 ω (2.32)
∂ξ h 1 ∂ξ ∂η h 2 ∂η

In Eqs. (2.31) and (2.32), h 1 and h 2 refer to the scale-factors along transformed ξ -
and η-directions for an orthogonal grid. These are given as,
 2  2
∂x∂y
h1 = +
∂ξ∂ξ
   
∂x 2 ∂y 2
h2 = + (2.33)
∂η ∂η

The contra-variant components of the velocity vector are given by


28 2 DNS of Navier–Stokes Equation

1 ∂ψ
u=
h 2 ∂η
1 ∂ψ
v=− (2.34)
h 1 ∂ξ

In Eqs. (2.29) and (2.31), we used the nondimensional Reynolds number which is
defined as
Ur e f L r e f
Re =
ν
where, Ur e f , L r e f and ν are reference integral velocity and length scales and kinematic


viscosity, respectively. For the velocity-vorticity formulation (( V , ω)-formulation),
VTE can be written in either conservative or nonconservative formulation. The non-
conservative formulation of VTE in transformed computational (ξ, η)-plane can be
obtained from Eqs. (2.31) as
    
∂ω ∂ω ∂ω 1 ∂ h 2 ∂ω ∂ h 1 ∂ω
h1h2 + h2u + h1v = + (2.35)
∂t ∂ξ ∂η Re ∂ξ h 1 ∂ξ ∂η h 2 ∂η

while the corresponding conservative form of VTE is given as


    
∂ω ∂(h 2 uω) ∂(h 1 vω) 1 ∂ h 2 ∂ω ∂ h 1 ∂ω
h1h2 + + = + (2.36)
∂t ∂ξ ∂η Re ∂ξ h 1 ∂ξ ∂η h 2 ∂η


The continuity equation (i.e., ∇ · V = 0) in the (ξ, η)-plane is given as
 
1 ∂(h 2 u) ∂(h 1 v)
+ =0 (2.37)
h1h2 ∂ξ ∂η


In the ( V , ω)-formulation, attendant velocity components are obtained from the
velocity Poisson equation (Eq. (2.26)) and for u- and v-components, these equations
in the transformed plane are given as
    
1 ∂ 1 ∂(h 2 u) 1 ∂ 1 ∂(h 1 u)
+ +
h 1 ∂ξ h 1 h 2 ∂ξ h 2 ∂η h 1 h 2 ∂η
    
1 ∂ 1 ∂(h 1 v) 1 ∂ 1 ∂(h 2 v) 1 ∂ω
− =−
h 1 ∂ξ h 1 h 2 ∂η h 2 ∂η h 1 h 2 ∂ξ h 2 ∂η
    
1 ∂ 1 ∂(h 2 v) 1 ∂ 1 ∂(h 1 v)
+ + (2.38)
h 1 ∂ξ h 1 h 2 ∂ξ h 2 ∂η h 1 h 2 ∂η
    
1 ∂ 1 ∂(h 2 u) 1 ∂ 1 ∂(h 1 u) 1 ∂ω
− =
h 2 ∂η h 1 h 2 ∂ξ h 1 ∂ξ h 1 h 2 ∂η h 1 ∂ξ

One notes from Eq. (2.38) that unlike SFE in Eq. (2.32), the Poisson Equations
for velocity components (Eq. (2.38)) contain mixed-derivatives. Consequently the
2.1 Fluid Dynamical Equations 29

discretized matrix for Eq. (2.38) would contain more terms along any row which
generally slows down the convergence of these equations, when an iterative solver
is used to obtain the solution of the resultant linear algebraic equations. However,
when ξ - and η-directions are along physical x- and y-directions, respectively, so that
h 1 = h 1 (ξ ) and h 2 = h 2 (η), Eq. (2.38) can be simplified to
   
∂ h 2 ∂u ∂ h 1 ∂u ∂ω
+ = −h 1
∂ξ h 1 ∂ξ ∂η h 2 ∂η ∂η
   
∂ h 1 ∂v ∂ h 2 ∂v ∂ω
+ = h2
∂ξ h 2 ∂ξ ∂η h 1 ∂η ∂ξ

The above equation is in self-adjoint form, similar to Eq. (2.32) for stream-function
ψ.

2.2 Spatial and Temporal Scales for Transitional


and Turbulent Flows

Transitional and turbulent flows are inherently unsteady, exhibiting wide-range of


space and time scales. For turbulent flows, the existence of broadband spectra is a
consequence of nonlinearity in the governing equations where transfer of energy takes
place from one wavenumber (or frequency) component to neighboring wavenumber
(or frequency) components. For 3D flows, the vortex-stretching terms (the first term
on the right-hand side of Eq. (2.22)) are responsible for transferring energy from
large to small scales. As these features are intrinsic to NSE, therefore these are also
noted for later stages of transitional flows. For 2D flows also, there can exist an
enstrophy cascade (see Davidson [23] for details), which causes a reverse migration
of energy from small to large scale, a process known as the inverse cascade process.
Scales of turbulent flows are related to eddy sizes in the evolving flow field. The
largest scale is associated with the integral dimension of the fluid dynamical system,
denoted by l, at which the flow is fed with energy. For homogeneous and isotropic
turbulent flows, Kolmogorov has shown that (see Tennekes and Lumley [114]) the
smallest excited length scale (known as the Kolmogorov scale) is given by
1
η K = (ν 3 /ε) 4 (2.39)

If we define u as the representative velocity-scale in the large scale (associated


with kinetic energy per unit mass), then we can define a corresponding Reynolds
number given by
ul
Re =
ν
Thus the largest and the smallest length scales of turbulent flows are related by
30 2 DNS of Navier–Stokes Equation

l 3
= (Re) 4 (2.40)
ηK

For turbulent flows, most energetic structures occur at the lower wavenumbers
scales. Consequently, the peak in the energy spectrum is obtained at lower wavenum-
bers. However, the dissipation peak is located at a higher wavenumber as compared
to the peak of energy spectrum. This is due to the fact that the dissipation is given
by ν||∇u||22 . This can be shown from the energy budget of the disturbance field. In
general, the energy spectrum depends on the wavenumber (k), dissipation (ε) and
kinematic viscosity (ν).
The above description elaborates how wide range of scales are excited particularly
in turbulent flows, at high Reynolds numbers. For computations at high Reynolds
numbers via DNS, one needs to resolve these wide-range of scales. If the cut-off
wavenumber is represented by kc (related to η K ), then Eq. (2.40) can also be written
as,
3
kc l ≈ Re 4 (2.41)

This equation is often used to state grid requirements for DNS. For 3D flows, this
shows that the resolution requirement scales as (Re3/4 )3 or roughly about Re2 . In
deriving Kolmogorov’s scaling theory, it is said that there exist length scales shorter
than those are directly excited (l), but larger than the Kolmogorov scale (η K ), for
which the energy spectrum is independent of the viscous dissipation mechanism.
At these intermediate scales − the inertial subrange − the structure of the energy
spectrum (E(k)) is determined solely by nonlinear energy transfer (via the stretching
term given by the first term on the right-hand side of Eq. (2.22)) by a cascade process
and the overall energy flux is shown to depend as

E(k, ε) = Ck k − 3 ε 3
5 2
(2.42)

The existence of an inertial subrange suggests some form of universality of flow


structure, at this length scale. This is exploited in LES where the flow is computed
by resolving all the way up to the inertial subrange and anything smaller than this is
modeled via sub-grid scale (SGS) stress models.
An inhomogeneous turbulent flow-field is generally identified as having large-
scale low-frequency coherent structures and seemingly random high-frequency tur-
bulent fluctuations. These organized coherent structures in turbulent flows carry
about 20% of total turbulent kinetic energy (TKE) and hence their role in determin-
ing turbulent flow dynamics cannot be underestimated. In fact, several researchers
have identified the significant role these coherent structures play in determining var-
ious physical phenomena like noise radiation [113], low frequency oscillation of
the separation zone in turbulent shock-boundary layer interactions [115], mixing
and entrainment in turbulent free-shear layers and wakes [76]; swirling jets [66],
low-frequency buffeting for transonic flow past an airfoil [21] and many other
similar applications. Therefore, accurate prediction of the various dynamical fea-
tures of these large-scale coherent structures are essential to obtain vital information
2.2 Spatial and Temporal Scales for Transitional and Turbulent Flows 31

of the flow-physics. However, the dynamics of these structures are also strongly
dependent on the random incoherent fluctuations. Reynolds and Hussain [76] have
showed that one has to fairly accurately model the drain of energy from the coherent
structures to incoherent fluctuation to predict the dynamics of the former.
For wall-bounded flows, these organized structures show up as peaks and valleys
in the near-wall region and in terms of wall units, they have lengths between 100
and 2000 units in the streamwise direction and have a spacing of about 50 units in
the spanwise direction. These high energetic events occur at a height of about 20
to 50 units (i.e., in and around the buffer layer) − all these are statistical estimates.
Additionally, these near-wall events are interspersed by bursting of these structures.
After bursts, new intermediate scale motions ensue in the buffer layer; those are also
streamwise and/or hair-pin vortices. Thus, the unsteadiness of the turbulent boundary
layer is characterized by bursting frequency, even for an attached shear layer. For
example, in the zero pressure gradient boundary layer, this critical frequency is
roughly between 20 and 100% of the turbulence burst frequency (Fb ), where Fb =
U∞ /5δ, with δ as the shear layer thickness. For adverse pressure gradient flows, this
critical frequency is between 6 and 28% of Fb .
In Fig. 2.1, various temporal scales excited in typical engineering flows are dis-
played for different speed regimes. For high Reynolds number flows, only the mean
frequencies of large eddies are shown. It is quite clear that ranges of non-dimensional
frequencies span over three orders of magnitude and numerical methods must resolve
these scales for high Reynolds number flows.

2.3 Numerical Methods for Developing DNS/LES

Previous discussion highlights the importance of resolving broadband spatial and


temporal scales for accurate computations of transitional and turbulent flows. There-
fore, for any numerical computation of transitional and turbulent flows, the comput-
ing methodologies should be spectrally accurate. Moreover, the numerical methods
should also preserve the spectral relationship between spatial and temporal scales i.e,
the numerical methods should preserve the physical dispersion relationship. In com-
puting the fluid dynamics governing equations accurately, therefore, both space-time
dependence of the problem has to be considered together. This aspect of computing is
very often overlooked, where spatial and temporal discretization is often decoupled,
which can lead to serious dispersion errors [95, 101].

2.3.1 Waves – Building Blocks of a Disturbance Field

We begin by first highlighting the aspects of disturbance evolution which takes a


flow from the laminar to the turbulent state. Irrespective of the mechanism being
linear or non-linear, it is always possible to explain disturbance as being composed
of Fourier–Laplace transforms. For space-time dependent systems, such disturbances
32 2 DNS of Navier–Stokes Equation

Fig. 2.1 Ranges of temporal scales excited in flows of engineering interest for different speeds and
the corresponding frequency ranges (Ω f ). Also, indicated is the band for mean frequencies of large
eddies in high Reynolds number flows

can often be viewed as plane waves. Waves can be dispersive or non-dispersive, gov-
erned by hyperbolic or non-hyperbolic partial differential equations. The prototype
of hyperbolic waves is often taken as the one dimensional convection equation

∂u ∂u
+c =0 (2.43)
∂t ∂x
where u can represent any disturbance quantity. Consider the propagation of distur-
bances subject to initial conditions

u(x, 0) = f (x) for −∞< x <∞ (2.44)

The unbounded spatial domain makes this a Cauchy problem. In Eq. (2.44), f (x)
is considered continuous functions. The exact solution of Eq. (2.43) is given as
2.3 Numerical Methods for Developing DNS/LES 33

u(x, t) = f (x − ct)

The nature of this solution indicates that the initial disturbance propagates to the
right at a convection speed of c, while maintaining identical amplitude and shape at
all time instants. The building blocks of any arbitrary aggregation of plane waves can
be understood by defining certain wave parameters for a single-periodic function
 

u(x, t) = a sin (x − ct) (2.45)
λ̄

One can identify this as a specific solution for Eq. (2.43). In the above, the quantity
in the square brackets represents the phase of the wave and a represents the amplitude
of the wave. The quantity λ̄ is the wavelength, since u does not change when x is
changed by λ̄, with t held fixed. One defines wavenumber k (= 2π λ̄
), which provides
the number of full waves in a length 2π . Thus, the representation of Eq. (2.45) can
be alternately written as
u(x, t) = a sin [k(x − ct)] (2.46)

Keeping one’s gaze fixed at a single point, the least time after which u(x, t) retains
the same value determines the time period T , and this is also the time required for
the wave to travel one wavelength: T = λ̄c . The number of oscillations at a point per
unit time is the frequency given by f 0 = T1 . One can define a circular frequency ω̄
by noting
ω̄ = kc (2.47)

Thus, c (= ω̄/k) has a dimension of speed and is appropriately called the phase
speed, the rate at which the phase of the wave propagates. Such movement is not
always physical and most often illusory. Equation (2.47) is known as the physical
dispersion relation for obvious reasons.
One consequence of the dispersion property is the group velocity, Vg . The physical
implications of the group velocity Vg is that it is the speed with which energy travels in
a system displaying a wide-band spectrum. This has been recognized by researchers
across many disciplines of science and engineering and studied in [3, 13, 46, 58,
127]. Rayleigh [73] laid the foundation for group velocity as opposed to phase speed
− although it was discussed earlier in [38]. The carrier waves use the phase speed
for phase variation, while the group velocity is associated with the propagation of
amplitude. According to Brillouin [13], group velocity is the velocity of energy
propagation and this was identified as signal speed by Rayleigh [74]. For linear
dispersive systems following some conservation law, the group velocity is defined
as
d ω̄
Vg = (2.48)
dk
Therefore, for 1D convection equation, the dispersion relation shows
34 2 DNS of Navier–Stokes Equation

d ω̄
= Vg = c (2.49)
dk
This signifies that the phase speed and group velocity are indistinguishable for
non-dispersive systems like the one given in Eq. (2.43). However, this is not true
for any general dispersive system. For such a system, the phase speed and the group
velocity are not identical. For example, for deep water waves, the dispersion relation
is given as [127]
ω̄ = gk (2.50)

where g is the acceleration due to gravity. For such waves, the group velocity and
the phase speed are given as

d ω̄ 1 g
Vg = =
dk 2 k
ω̄ g
c= = (2.51)
k k

Therefore, the group velocity is half the phase speed for deep water waves. Similarly,
consider propagation of transverse elastic waves in a non-ideal string, for which the
dispersion relation is given as

T 2
ω̄2 = k + αk 4 (2.52)
μ

where T is the tension in the string, μ is the string’s mass per unit length and the
parameter α depends on the stiffness of the string. One notes that for this system
also the group velocity and the phase speed are not identical. A more detailed dis-
cussion from physical and mathematical stand-point on group velocity can be found
in [82, 83, 127].

2.3.2 Resolution of Spatial Discretization

NSE basically governs space-time evolution of the primary variables. As noted in


all of its forms described till now, it contains various forms of spatial derivatives.
Therefore, the resolution of the discretization scheme to estimate these derivatives
would invariably affect the accuracy of the numerical solution of NSE. Here, we talk
about the resolution of the spatial discretization schemes for the estimation of first
and second derivatives. We first begin with the numerical discretization of the first
derivative.  
Let, u j = ∂∂ux represents the first derivative of the variable u = u(x) at the
j
jth-node. Let, all the nodes be equispaced with h representing the distance between
2.3 Numerical Methods for Developing DNS/LES 35

ui 3 ui 2
ui 1 ui ui+1 ui+2 ui+3
(i 3) (i 2) (i 1) i (i +1) (i + 2) (i + 3)

Fig. 2.2 Schematic of grid-spacing. All the nodes are equispaced and the distance between adjacent
grid-points is h

adjacent nodes, as shown in Fig. 2.2. The 2nd-order central difference (C D2 ) scheme
to evaluate u j is given as

u j+1 − u j−1
u j = (2.53)
2h
Similarly, the 4th-order accurate C D4 -scheme for spatial discretization of the first
derivative is given as

−u j+2 + 8u j+1 − 8u j−1 + u j−2


u j = (2.54)
12h
To evaluate the resolution of a particular scheme, we express the function in terms
of its Fourier–Laplace transform by

uj = U (k)eikx j dk (2.55)

From Eq. (2.55), one notes that ideally

(u j )exact = ikU (k)eikx j dk (2.56)

For C D2 -scheme in Eq. (2.53), u j can be written in spectral form as

eikh − e−ikh
(u j )C D2 = U (k)eikx j dk (2.57)
2h

which gives

sin(kh)
(u j )C D2 = i U (k)eikx j dk = ikeq U (k)eikx j dk (2.58)
h

where keq is the equivalent wavenumber. The term keq /k basically defines the reso-
lution of any spatial discretization scheme. For C D2 and C D4 schemes, this term is
given from Eqs. (2.53) and (2.54) as
36 2 DNS of Navier–Stokes Equation
 
keq sin(kh)
= (2.59)
k C D2 kh
   
keq (4 − cos(kh) sin(kh)
= (2.60)
k C D4 3 kh

Note that for both Eqs. (2.59) and (2.60), (keq /k) −→ 1 as kh −→ 0. This is the
Consistency condition. This must be true for any spatial discretization scheme for
as kh −→ 0, one approaches the continuum from the discrete limit, and this implies
(keq /k) −→ 1. Similar expression of resolution can also be derived for other higher-
order discretization schemes, like C D6 and C D8 schemes. Note from Eqs. (2.59)
and (2.60) for C D2 and C D4 schemes, (keq /k) is a real quantity. This is true for any
central discretization scheme in uniform grids.
For upwinded spatial discretization schemes, (keq /k) is a complex quantity. For
example, for first-order upwind scheme U D1 ,

u j+1 − u j
u j = (2.61)
h

and (keq /k) for U D1 scheme is given as


     
keq sin(kh) (cos(kh) − 1) keq keq
= +i = +i (2.62)
k U D1 kh kh k r eal k img

where, (keq /k)r eal and (keq /k)img are the real and imaginary part of (keq /k). Simi-
larly for second order upwind U D2 scheme,

3u j − 4u j−1 + u j−2
u j = (2.63)
2h
the complex resolution can be obtained as
 
keq sin(kh)(2 − cos(kh)) (cos(kh) − 1)2
= −i (2.64)
k U D2 kh kh

In Fig. 2.3, real and imaginary parts of (keq /k) are plotted as a function of modi-
fied wavenumber, kh, for indicated spatial discretization schemes. Here, (keq /k)r eal
denotes the resolution of the particular spatial discretization scheme. On the other
hand, value of (keq /k)img signifies the added numerical diffusion of the scheme.
This can be understood considering the semi-discrete form (where time-integration
is considered as exact and spatial discretization is as given for the chosen numerical
schemes) of the 1D convection equation (Eq. (2.43)). As for any numerical scheme
 
∂u
= ikeq U (k, t)eikx j dk (2.65)
∂x j
2.3 Numerical Methods for Developing DNS/LES 37

(a) (b)

Fig. 2.3 a (keq /k)r eal and b (keq /k)img plotted for CD2 , CD4 , CD6 , UD1 and UD2 spatial
discretization schemes as a function of kh

Expressing the variable u(x, t) in terms of its Fourier–Laplace amplitude U (k, t),
Eq. (2.43) can be written as
 
∂U
+ ikeq U (k, t) eikx j dk = 0 (2.66)
∂t

Therefore, for any wavenumber component k, the equation for U (k, t) is given as
 
∂U
+ ikeq U (k, t) = 0 (2.67)
∂t

The exact solution of Eq. (2.67) can be obtained as

U (k, t) = U (k, 0) e−i(keq /k)r eal kt e(keq /k)img kt (2.68)

Equation (2.68) clearly shows that U (k, t) grows or decays exponentially with time,
when (keq /k)img is positive or negative, respectively. It can be further shown that (by
considering the spatial discretization schemes for second derivatives, which is illus-
trated next), effect of upwinding converts the 1D-convection equation (Eq. (2.43))
to

∂u ∂u ∂ 2u
+c =β 2 (2.69)
∂t ∂x ∂x

where β > 0 if (keq /k)img < 0. Thus, the role of upwinding is to induce a diffusive
term when (keq /k)img is negative. A positive value of (keq /k)img is referred to as anti-
diffusion,which is potentially destabilizing for the overall scheme via such spatial
discretization.
Next, the resolution of the spatial discretization schemes for the evaluation of
second derivatives are provided. The second order central difference scheme for the
38 2 DNS of Navier–Stokes Equation

evaluation of second derivatives is given as

u j+1 − 2u j + u j−1
u j = (2.70)
h2
where double-prime denotes the spatial second derivative. We refer this scheme as
C D22 to distinguish it from the second order central difference scheme (Eq. (2.53))
for the first derivative. Following Eq. (2.65), one can express the exact and the
numerically obtained second derivative as
 
∂ 2u j
= −k 2 U (k)eikx j dk (2.71)
∂x2 exact
 
∂ 2u j
u j = = −keq
2
U (k)eikx j dk (2.72)
∂x2 num

where keq is the modified wavenumber for the evaluation of the second deriva-
tive. Therefore, similar to the cases for the numerical evaluation of first derivatives,
here also (keq /k) would denote the resolution of the scheme to approximate second
derivatives. The spectral resolution of the scheme of (2.70), in terms of the modified
wavenumber keq , is given as
 2
keq sin2 (kh/2)
= (2.73)
k (kh/2)2

It is noted that the scheme given by Eq. (2.70) is formally second order accurate
in terms of the truncation error, as obtained from the Taylor series expansion. One
can also approximate the second derivative by the following approximation while
maintaining second order accuracy in terms of the truncation error as

u j+2 − 2u j + u j−2
u j = (2.74)
4h 2
2
We refer this scheme as C D2a to distinguish it from the C D22 scheme given in
Eq. (2.70). Though both the schemes are formally second order accurate, the distinc-
tion between these become clear when one plots the corresponding spectral resolution
as a function of wavenumber. For scheme (2.74), the spectral resolution in terms of
(keq /k)2 is given as
 2
keq sin2 (kh)
= (2.75)
k (kh)2

In Fig. 2.4, (keq /k)r eal is plotted for both the schemes as a function of kh. It becomes
readily apparent from this figure that C D22 scheme (Eq. (2.70)) has higher resolution
2
as compared to the C D2a scheme (Eq. (2.74)), specially at higher wavenumbers.
2.3 Numerical Methods for Developing DNS/LES 39

Fig. 2.4 (keq /k)r eal plotted


for the C D22 - and 1
C D2a2 -schemes as a function

of kh. These schemes are 0.8


given by Eqs. (2.70) and

(keq/k)real
0.6
(2.74), respectively
0.4 2
CD2
CD22a
0.2

0
0.5 1 1.5 2 2.5 3
kh

The resolution (keq /k) for C D2a 2


scheme becomes zero at the Nyquist limit i.e.,
at kh = π . In contrast, (keq /k) = (2/π ) at kh = π for the C D22 -scheme given by
Eq. (2.70).

2.3.3 High-Accuracy Compact Schemes for Evaluation


of First Derivatives

In the previous subsection, we have described various explicit schemes to discretize


first and second derivatives. Here, we introduce compact schemes, in which the
derivatives are treated as unknowns as function of the variables. The derivatives at
the grid points are expressed as an auxiliary implicit equation involving the unknown.
There are two essential features: high spectral accuracy and relatively compact sten-
cil, to relate derivatives with the unknown. These methods provide near-spectral
accuracy and robustness, at the same time being less costly than the spectral scheme.
Historically, these methods owe their origin to the entire class of centered explicit
spatial discretization using Padé schemes for ODE, described in [50]. Application
of such schemes to PDEs can be found in many places and for CFD, these can be
found in [53, 81].
The derivatives following the compact schemes are obtained by solving a system
of linear algebraic equations. The compact representation of these schemes indicates
that the resultant matrix is strictly tri-diagonal or penta-diagonal, so that the linear
algebraic equations can be easily solved. The general stencil following compact
schemes to evaluate the nth derivative of the variable u j at the jth-node for a uniform
grid of spacing h can be given as

N2 M2
1
αk u (n)
j+k = βl u j+l (2.76)
k=−N1
hn l=−M1
40 2 DNS of Navier–Stokes Equation

where u (n)
j represents the nth derivative of u j . It is apparent that the resultant matrix is
band-limited with bandwidth defined by N1 and N2 points towards left and right of the
jth-node. Let us consider the general stencil of a compact scheme for evaluation of
first derivative (denoted by the superscript  ) with N1 = N2 = 2 and M1 = M2 = 3.
Such a scheme can be expressed as

1
α+2 u j+2+ α+1 u j+1
+ α0 u j
+ α−1 u j−1
+ α−2 u j−2
= β+3 u j+3 + β+2 u j+2 +
h

β+1 u j+1 + β0 u j + β−1 u j−1 + β−2 u j−2 + β−3 u j−3 (2.77)

One notes that if α+2 = α−2 = 0 then the resultant matrix for Eq. (2.77) would
be tridiagonal and one can easily employ Thomas’ tridiagonal matrix algorithm
(TDMA) to solve the above system of equations to obtain u j . Using Taylor series
expansion of both right- and left-hand side expressions, one can derive the consistency
and other conditions. By matching the coefficient of u j , the first consistency condition
(corresponding to first order accuracy) can be obtained as

(β+3 + β+2 + β+1 + β0 + β−1 + β−2 + β−3 ) = 0 (2.78)

Similarly, by matching the coefficient of u j from left- and right-hand sides, the second
condition (corresponding to second order accuracy) can be obtained as

(α+2 + α+1 + α0 + α−1 + α−2 )


= 3(β+3 − β−3 ) + 2(β+2 − β−2 ) + (β+1 − β−1 ) (2.79)

The general condition for ( p + 1)th-order of accuracy (where p ≥ 2), in terms of


truncation error, can be given as

3p
2 p−1 (α+2 + (−1) p−1 α−2 ) + (α+1 + (−1) p−1 α−1 ) = (β+3 + (−1) p β−3 )
p
2p 1
+ (β+2 + (−1) p β−2 ) + (β+1 + (−1) p β−1 ) (2.80)
p p

Equation (2.77) contains 12 unknown coefficients (α+2 , ....., α−2 , β+3 , ......, β−3 ) to
be determined and therefore, 11th-order of accuracy can be obtained from the com-
pact scheme for spatial discretization, described in Eq. (2.77). However, in practice
hardly such higher order accurate compact schemes are used. Instead, to obtain a
better accuracy scheme, some of the higher order conditions are sacrificed and addi-
tional conditions are imposed based on the resolution (keq /k) in the spectral plane, as
followed to develop high-accuracy compact schemes like OU C S3 in [92, 96]. One
also notes from Eq. (2.77) that if α+k = α−k , (β+l + β−l ) = 0 and β0 = 0, where
k = 1, 2 and l = 1, 2, 3, then one essentially obtains a central spatial discretization
scheme. Following tridiagonal compact scheme is proposed for the evaluation of first
derivative:
2.3 Numerical Methods for Developing DNS/LES 41

u j+3 − u j−3
αu j+1 + u j + αu j−1 = c +
6h
u j+2 − u j−2 u j+1 − u j−1
b +a (2.81)
4h 2h
If c = 0, then for 4th-order accuracy one must have

2 1
a= (α + 2), b = (4α − 1)
3 3
while α = 1/3 makes the scheme 6th-order accurate. If c = 0 then sixth order accu-
racy can be obtained if

1 1 1
a= (α + 9), b= (32α − 9), c= (−3α + 1)
6 15 10
If α = 3/8, then a formally eighth-order accurate scheme can be obtained. If the
scheme given by Eq. (2.81) is applied to a periodic problems, its resolution in the
spectral plane is given as
 
keq 1 a sin(kh) + (b/2) sin(2kh) + (c/3) sin(3kh)
= (2.82)
k kh 1 + 2α cos(kh)

The advantage gained in terms of accuracy, while using pentadiagonal compact


schemes, can be effectively leveraged if spectrally optimized tridiagonal compact
schemes are developed.
In [92, 96], high-accuracy optimized upwind compact scheme OU C S3 is
described, which has the stencil at interior nodes given as

2
1
p j−1 u j−1 + u j + p j+1 u j+1 = qn u j+n (2.83)
h n=−2

η∗ η∗ η∗
where, p j±1 = D ± 60 ; q±2 = ± F4 + 300 ; q±1 = ± E2 + 30 ; q0 = − 11η
150

; D=
0.3793894912; E = 1.57557379 and F = 0.183205192. Here, η is the upwind
parameter and a value of zero implies the corresponding central scheme. This scheme
is second order accurate. Additional coefficients of the scheme are found by max-
imizing the resolution in the spectral plane. The elaborate methodology of spectral
optimization by which these coefficients are obtained for non-periodic stencils are
provided in [92, 96]. For periodic problems, this essentially implies that one maxi-
mizes the following objective function
γ  
keq
Iγ = 1− d(kh) (2.84)
0 k

where 0 ≤ γ ≤ π . In [96], γ = π was chosen. Despite only second order accuracy


of the scheme (Eq. (2.83)), its resolution in the spectral plane is significantly higher
42 2 DNS of Navier–Stokes Equation

(a) (b)
1
1.04
0.8 Lele’s Scheme (8th order)
nd
OUCS3 scheme (2 order)
1.02
(keq/k)real

(keq/k)real
0.6
1
0.4 th
Lele’s Scheme (8 order)
nd
OUCS3 scheme (2 order) 0.98
0.2
0.96
0
1 2 3 0.5 1 1.5 2
kh kh

Fig. 2.5 (keq /k) plotted as a function of kh for the interior stencil of OUCS3 (Eq. (2.83)) and the
eighth order compact schemes proposed by Lele (Eq. (2.81)). Here, we consider only the central
stencil for periodic problems. For OUCS3, we take the central stencil, i.e., η = 0 in Eq. (2.83)

than explicit schemes like C D2 , C D4 and C D6 . In Fig. 2.5, (keq /k) for compact
schemes proposed by Lele [53] with eighth order accuracy and OUCS3 scheme
(Eq. (2.83)) with second order accuracy are compared for interior stencil applied to
periodic problems. For OUCS3 scheme, we have used the central stencil, i.e., η = 0
in Eq. (2.83). Figure 2.5 shows that all these schemes maintain (keq /k) 1 up to
significantly larger range of kh. For OUCS3 scheme, (keq /k) 1, up to kh ≈ 2.3,
after which it starts to fall off and becomes zero at the Nyquist limit. Even though
the scheme proposed in Lele [53] is eighth order accurate, its spectral resolution is
slightly inferior than the central OUCS3 stencil, which is only second order accurate
in terms of truncation error.

2.3.4 Boundary Closure Schemes for Compact Schemes


to Evaluate First Derivative

So far we have only discussed about the compact schemes applied to periodic
problems. For non-periodic problems, one needs stencils at near boundary points
( j = 1, 2, (N − 1) and N ) to close the implicit system of equations, Eq. (2.81)
or (2.83).
Adams [1] used the following stencils for non-periodic problem

1
2u 1 + 4u 2 = (−5u 1 + 4u 2 + u 3 ) at j = 1 (2.85)
h
3
u 1 + 4u 2 + u 3 = (u 3 − u 1 ) at j = 2 (2.86)
h
1
u j−1 + 3u j + u j+1 = (−u j−2 − 28u j−1
12h
+ 28u j+1 + u j+2 ) for 3 ≤ j ≤ (N − 2) (2.87)
2.3 Numerical Methods for Developing DNS/LES 43

Similar stencils can be obtained at j = N − 1 and j = N , following the stencils


given by Eqs. (2.85) and (2.86), respectively. In the above, near-boundary point
stencil at j = 2 has fourth order accuracy, while the interior points have sixth order
formal accuracy.
From, Eqs. (2.85)–(2.87), it is clear that for non-periodic problems, use of compact
stencil to evaluate first derivative result in solving the following linear algebraic
equation [81, 83, 96]

1
[A]{u  } = [B]{u} (2.88)
h

where both [A] and [B] are (N × N ) square matrices. Denoting [C] = [A]−1 [B],
one obtains
1
u = [C]u (2.89)
h
It is pertinent to note that though, [A] and [B] are sparse matrices (here, [A] is tridi-
agonal and [B] contains non-zero elements only at three diagonal columns around
the main diagonal entry), the matrix [C] is in general a non-sparse matrix. Equa-
tion (2.89) implies that the derivative at the jth node is evaluated as

N
1
u j = C jl u l
h l=1


where u l = u(xl ) = U (k) eikxl dk is the function value at the lth node. Using
spectral representation, one can alternately write the numerical derivative as

1
u j = C jl U (k) eik(xl −x j ) eikx j dk (2.90)
h

From Eq. (2.90), one obtains


  N
[keq ] j i
=− C jl eik(xl −x j ) (2.91)
k kh l=1

Although in physical plane computations C jl ’s are real, ([keq ] j /k) is in gen-


eral complex, with real and imaginary parts representing numerical phase and added
numerical diffusion, respectively, as illustrated before. It is determined by the numer-
ical method fixing the entries of [C]. Following Eq. (2.91), one can obtain the real
and imaginary parts of the complex resolution, (keq /k), at every grid point ranging
from j = 1 to N . This is the global spectral analysis (GSA) of implicit compact
schemes for non-periodic problems introduced in [81, 83, 96].
It can be shown by plotting (keq /k) that implicit boundary closure schemes cause
numerical anti-diffusion at near-boundary points.
44 2 DNS of Navier–Stokes Equation

j=1
(a) Central OUCS3 stencil (b) j=2
3 j=3
1.2 j=50
2 j=99
1 j=100
(keq/k)real

(keq/k)img
1 j=101
0.8 j=1
j=2 0
0.6 j=3
j=50
0.4 -1
j=99
j=100
0.2 -2
j=101
-3
0 1 2 3 1 2 3
kh kh

Fig. 2.6 a Real and b imaginary part of (keq /k) plotted for OUCS3 scheme with explicit boundary
closure schemes given by Eqs. (2.92) and (2.93). For the interior OUCS3 stencil here we consider
η = 0 in Eq. (2.83)

The problem of anti-diffusion at near-boundary points for compact schemes with


implicit boundary closure can be addressed by prescribing explicit boundary closure
schemes at those points, i.e., at j = 1, 2, (N − 1) and N . In [96], special explicit
boundary closure schemes have been derived while deriving OU C S1, OU C S2,
OU C S3 schemes, whose interior stencil is given by Eq. (2.83). Following boundary
closure schemes have been proposed in [96],
 
1
u 1 = −3u 1 + 4u 2 − u 3 at j = 1 (2.92)
2h
     
 1 2β∗ 1 8β∗ 1
u2 = − u1 − + u 2 + 4β∗ + 1 u 3
h 3 3 3 2
  
8β∗ 1 2β∗
− + u4 + u 5 at j = 2 (2.93)
3 6 3

with β∗ as a floating parameter. Similar closure schemes for j = (N − 1) and N can


also be written. The non-zero parameter β∗ represents added fourth order diffusion,
which reduces anti-diffusion at near-boundary points. In [96], it is suggested that
for better global numerical properties, one requires β∗ = −0.025 for j = 2 and
β∗ = 0.09 for j = N − 1. Figure 2.6 shows real and imaginary parts of (keq /k) for
OUCS3 scheme with explicit boundary closure schemes given by Eqs. (2.92) and
(2.93). One readily notes that use of explicit boundary closure scheme have reduced
the problem of anti-diffusion at near-boundary points significantly.

2.3.5 Compact Schemes for Second Derivative Evaluation

Similar to the evaluation of first derivatives by compact schemes, one can also
use such schemes to evaluate second derivatives. Lele [53] proposed the following
2.3 Numerical Methods for Developing DNS/LES 45

compact stencils for the evaluation of second derivative in an uniform grid

β2 u j+2 + α2 u j+1 + u j + α2 u j−1 + β2 u j−2 =


u j+3 − 2u j + u j−3 u j+2 − 2u j + u j−2 u j+1 − 2u j + u j−1
c +b +a (2.94)
9h 2 4h 2 h2

Comparing the coefficient of u j from right- and left-hand side of Eq. (2.94), one gets
the consistency condition corresponding to second order accuracy as

(a + b + c) = (1 + 2α2 + β2 ) (2.95)

From the Taylor series expansion of Eq. (2.94), one gets a generalized equation for
the (2 p + 2)th-order term (with p = 1, 2, 3, 4) as

(2 p + 2)! 2 p
(32 p c + 22 p b + a) = (2 β2 + α2 ) (2.96)
(2 p)!

If β2 = c = 0, one obtains a tridiagonal compact scheme for the evaluation of


second derivative as
u j+2 − 2u j + u j−2
α2 u j+1 + u j + α2 u j−1 = b +
4h 2
u j+1 − 2u j + u j−1
a (2.97)
h2
For sixth order accuracy of the stencil, Eq. (2.97), one requires

(a + b) = (1 + 2α2 ) for second order accuracy


(a + 4b) = 12α2 for fourth order accuracy
(a + 16b) = 30α2 for sixth order accuracy (2.98)

Solving Eqs. (2.98), one gets α2 = 2/11, a = 12/11 and b = 3/11 for sixth order
accuracy. Therefore, stencil Eq. (2.97) and above values of a, b and α2 , defines a
sixth order scheme for periodic problems. Similarly a single parameter, fourth order
accurate, tridiagonal compact scheme can be derived for

4
a= (1 − α2 ) (2.99)
3
1
b= (10α2 − 1) (2.100)
3
Following Eq. (2.72), the resolution of the scheme Eq. (2.97) for periodic problems
can be expressed in terms of the modified wavenumber keq as
46 2 DNS of Navier–Stokes Equation

(a) (b)
1.2
1.2
1
(keq/k)real

1 0.8
α=0.4

I(α)
0.6
α=0.3
0.8
α=0.2 0.4
α=0.1
0.6 0.2
α=2/11
0
0.5 1 1.5 2 2.5 3 0.1 0.2 0.3 0.4
kh α2

Fig. 2.7 a (keq /k) for the periodic tridiagonal compact stencil, Eq. (2.97) plotted for different
values of α2 . α2 = 2/11 yields a sixth order accurate scheme. b Integral objective function I given
by Eq. (2.102) plotted as a function of α2 for γ = π

 2
keq 1 (b sin2 (kh) + 4a sin2 (kh/2))
= (2.101)
k (1 + 2α2 cos(kh)) (kh)2

Therefore, an optimized tridiagonal compact scheme, for second derivative evalu-


ation and applicable to periodic problems, in uniform grid can be derived, if one
obtains α2 by minimizing the following objective function

γ  2
keq
I = 1− d(kh) (2.102)
0 k

where γ ≤ π . In Fig. 2.7a, (keq /k)2 for the periodic tridiagonal compact stencil
Eq. (2.97) is plotted for different values of α2 . As noted earlier, α2 = 2/11 yields
a sixth order accurate scheme. Figure 2.7b shows the integral objective function I
given by Eq. (2.102), as a function of α2 for γ = π . Figure 2.7b shows that I is
minimum when α2 = 0.26. Therefore, the fourth order accurate stencil Eq. (2.97)
with α2 = 0.26 provides better spectral accuracy, than corresponding sixth order
accurate stencil with α2 = 2/11 for periodic problems.
For non-periodic problems, one requires boundary closure stencils at near-
boundary points. Lele [53] prescribed following boundary closure stencils for the
compact scheme

1
u 1 + 11u 2 = (13u 1 − 27u 2 + 15u 3 − u 4 ) at j = 1 (2.103)
h2
1  1 12
u 1 + u 2 + u 3 = (u 3 − 2u 2 + u 1 ) at j = 2 (2.104)
10 10 10h 2

Similar stencils can be derived at j = (N − 1) and j = N . One notes that Lele [53]
proposed implicit boundary closure. As noted in Sect. 2.3.4 for the first derivative,
implicit boundary closure schemes lead to anti-diffusion at near boundary points
2.3 Numerical Methods for Developing DNS/LES 47

and spreads over a wider range of near-boundary points. For the scheme given in
Eq. (2.94) with β2 = c = 0, one can significantly remove numerical problems, if
explicit second-order accurate schemes are used at the boundary nodes. The com-
posite compact stencil can then be expressed as

b
α2 u j+1 + u j + α2 u j−1 = (u j+2 − 2u j + u j−2 )
4h 2
a
+ (u j+1 − 2u j + u j−1 ) for 3 ≤ j ≤ (N − 3) (2.105)
h2
1
u j = (u j+1 − 2u j + u j−1 ) for j = 2 and j = (N − 1) (2.106)
h2
1
u 1 = 2 (2u 1 − 5u 2 + 4u 3 − u 4 ) for j = 1 (2.107)
h
1
u N = 2 (2u N − 5u N −1 + 4u N −2 − u N −3 ) for j = N (2.108)
h
Equation (2.88) expresses the linear algebraic equation corresponding to compact
stencils for first derivative evaluation applied to non-periodic problems in equivalent
matrix-vector form. Similar matrix-vector form for non-periodic compact stencils
for second derivatives can be written as
1
[A]{u  } = [B]{u} (2.109)
h2

where both [A] and [B] are (N × N ) square matrices. Denoting [D] = [A]−1 [B],
one obtains
1
{u  } = [D]{u} (2.110)
h2
Though, [A] and [B] are sparse matrices ([A] is tridiagonal and [B] contains non-
zero elements only along three diagonals including the main diagonal entry), the
matrix [D] is not a sparse-matrix. From Eq. (2.110) the derivative at the jth node is
evaluated as
N
1
u j = 2 D jl u l
h l=1

Using spectral representation, one can alternately write the second derivative as [81,
83],
1
u j = D jl U (k) eik(xl −x j ) eikx j dk (2.111)
h2

Following Eq. (2.111), one can perform a full domain spectral analysis for non-
periodic compact stencils and express (keq /k)2 at jth-node as
48 2 DNS of Navier–Stokes Equation

 2 N
[keq ] j 1
=− D jl eik(xl −x j ) (2.112)
k (kh)2 l=1

Therefore, ([keq ] j /k)2 is in general complex, with real and imaginary parts repre-
senting numerical resolution error and artificial numerical dispersion, respectively.
Next, the combined compact differencing (CCD) scheme is described [18]. This
scheme has implicit sixth order stencils, where both first and second derivatives are
evaluated simultaneously. The CCD scheme [18] for uniform grid-points are given
as
a1∗
u j + α1∗ (u j+1 + u j−1 ) + β1∗ h(u j+1 − u j−1 ) = (u j+1 − u j−1 ) (2.113)
2h
β2∗  a2∗
u j + α2∗ (u j+1 + u j−1 ) + (u j+1 − u j−1 ) = 2 (u j+1 − 2u j + u j−1 ) (2.114)
2h h
for sixth order accuracy α1∗ = 7/16, β1∗ = −1/16, a1∗ = 15/8, α2∗ = −1/8, β2∗ =
9/4 and a2∗ = 3. To obtain spectral resolution of the CCD scheme (both first and sec-
ond derivatives), one has to express u j by its Fourier–Laplace transform, as explained
before. Let, U (k), U1 (k) and U2 (k) be the Fourier–Laplace amplitude of u j , u j and
u j , respectively. Then considering a periodic problem, the equations for U1 (k) and
U2 (k) in terms of U (k) can be obtained from Eqs. (2.113) and (2.114) as
a1∗
(1 + 2α1∗ cos(kh))U1 (k) + 2iβ1∗ h sin(kh)U2 (k) = 2i sin(kh)U (k)
2h
(2.115)
β2∗ a2∗
2i sin(kh)U1 (k) + (1 + 2α2∗ cos(kh))U2 (k) = 2 (1 + 2 cos(kh))U (k)
2h h
(2.116)

solving Eqs. (2.115) and (2.116) for U1 (k) and U2 (k), one obtains

U1 (k) 2i sin(kh) (a1∗ − 2a2∗ β1∗ ) + 2(a1∗ α2∗ − 2a2∗ β1∗ ) cos(kh)
=
U (k) 2h (1 + 2α1∗ cos(kh))(1 + 2α2∗ cos(kh)) + 2β1∗ β2∗ sin2 (kh)
(2.117)

U2 (k) 1 a1∗ β2∗ sin2 (kh) + a2∗ (1 + 2 cos(kh))(1 + 2α1∗ cos(kh))


= 2 (2.118)
U (k) h 2β1∗ β2∗ sin2 (kh) + (1 + 2α1∗ cos(kh))(1 + 2α2∗ cos(kh))

One can define k1eq and k2eq , the modified wavenumbers for the evaluation of first
and second derivatives, respectively, following

(U1 (k)/U (k)) = ik1eq and (U2 (k)/U (k)) = −k2 2eq
2.3 Numerical Methods for Developing DNS/LES 49

(a) First Derivative (b) Second Derivative


1
1
0.8
(keq/k)real 0.9

(keq/k)real
0.6
0.8
CCD
0.4 OUCS3
0.7 CCD
Lele’s Scheme
0.2 0.6

0 0.5
1 2 3 1 2 3
kh kh

Fig. 2.8 a (k1eq /k) and b (k2eq /k)2 plotted for the first and second derivatives corresponding
to CCD scheme (Eqs. (2.113) and (2.114)) for periodic problems. Stencils of OUCS3-scheme
Eq. (2.83) (first derivative) and the scheme due to Lele Eq. (2.97) (for second derivative) are also
included in frames a and b, respectively

Therefore, the resolution of the periodic CCD scheme for the evaluation of first and
second derivatives are given as
 
k1eq sin(kh) (a1∗ − 2a2∗ β1∗ ) + 2(a1∗ α2∗ − 2a2∗ β1∗ ) cos(kh)
=
k kh (1 + 2α1∗ cos(kh))(1 + 2α2∗ cos(kh)) + 2β1∗ β2∗ sin2 (kh)
(2.119)

 2
k2 2eq 1 a1∗ β2∗ sin2 (kh) + a2∗ (1 + 2 cos(kh))(1 + 2α1∗ cos(kh))
=
k (kh)2 2β1∗ β2∗ sin2 (kh) + (1 + 2α1∗ cos(kh))(1 + 2α2∗ cos(kh))
(2.120)

In Fig. 2.8, (k1eq /k) (frame (a)) and (k2eq /k)2 (frame (b)) are plotted for the first
and second derivatives, as obtained for CCD scheme given by Eqs. (2.113) and
(2.114) for periodic problems. In the figures, corresponding results for OUCS3-
scheme (Eq. (2.83) for first derivative) and for the scheme due to Lele (Eq. (2.97) for
second derivative) are also included. The superior resolution characteristics of CCD-
scheme is evident from this figure. The (keq /k)2 for the second derivative shows
slight overshoot close to the Nyquist limit. For NSE simulations with uniform grid-
points, this is very effective because (a) it results in very high accuracy in terms of
discretization of diffusion terms and (b) high value of (keq /k)2 at high wavenumber
range signifies adding numerical diffusion selectively at these high values of kh.
The last aspect is computationally very helpful, as it leads to the damping of the
amplitudes of high wavenumber components, thereby controlling several numerical
sources of error, like spurious numerical errors due to aliasing [107].
For non-periodic problems, the CCD-scheme also requires boundary closure
schemes. Chu and Fan [18] proposed the following stencils at j = 1 and j = N ,
while using CCD stencil (Eqs. (2.113) and (2.114)) from j = 2 to (N − 1).
50 2 DNS of Navier–Stokes Equation
 
1 7 1
u 1 + 2u 2 − hu 2 = − u 1 + 4u 2 − u 3 (2.121)
h 2 2
6 1
u 1 + 5u 2 − u 2 = 2 (9u 1 − 12u 2 + 3u 3 ) (2.122)
h  h 
   1 7 1
u N + 2u N −1 − hu N −1 = u N − 4u N −1 + u N −2 (2.123)
h 2 2
6 1
u N + 5u N −1 − u N −1 = 2 (9u N − 12u N −1 + 3u N −2 ) (2.124)
h h
In terms of the matrix-vector form, the linear algebraic equations, Eqs. (2.113)–
(2.114) and (2.121)–(2.124), are obtained as

1
[A]{u  } + h[B]{u  } = [C]{u} (2.125)
h
1 1
[D]{u  } + [E]{u  } = 2 [F]{u} (2.126)
h h
where [A] and [E] are symmetric tridiagonal matrices, whose diagonal elements are
unity, whereas [B] and [D] are skew-symmetric tridiagonal matrices, whose diagonal
elements are all zeros. Solving Eqs. (2.125) and (2.126), one obtains

1
{u  } = [A − B E −1 D]−1 [C − B E −1 F]{u} (2.127)
h
1
{u  } = 2 [E − D A−1 B]−1 [F − D A−1 C]{u} (2.128)
h
Following Eqs. (2.127) and (2.128) a full domain spectral analysis can be carried out,
as shown previously through Eqs. (2.91) and (2.112), respectively, for first and second
derivatives. As noted earlier, for CCD also, implicit boundary closure schemes lead
to numerical problems at near-boundary points. To overcome this difficulty explicit
boundary closure schemes are proposed in [98] at j = 2 and (N − 1), while the
CCD-stencil is applied for 3 ≤ j ≤ (N − 2). This scheme with explicit boundary
closure is termed as NCCD (acronym for New CCD) scheme. The explicit boundary
stencils proposed in [98] at j = 2 are given as
     
1 2β2 1 8β2 1
u 2 = − u1 − + u 2 + 4β2 + 1 u 3
h 3 3 3 2
  
8β2 1 2β2
− + u4 + u5 (2.129)
3 6 3
1
u 2 = 2 (u 1 − 2u 2 + u 3 ) (2.130)
h
where β2 = −0.025. Note that the closure scheme, Eq. (2.129), corresponds to the
closure scheme of OUCS3 (see Eq. (2.93)). Similar stencils can be obtained at
2.3 Numerical Methods for Developing DNS/LES 51

(a) NCCD: First Derivative (b) NCCD: First Derivative


0.5
1

0.8
(keq/k)real

(keq/k)img
0
0.6 j=2
j=3
j=50 j=2
0.4 j=99 -0.5 j=3
j=100 j=50
0.2 j=99
j=100
0 -1
1 2 3 1 2 3
kh kh
(c) NCCD: Second Derivative (d) NCCD: Second Derivative
0.2

1
0.1
(keq/k)real

j=2 (keq/k)img
0.8 j=3 0
j=50 j=2
j=99 j=3
j=100 -0.1 j=50
0.6 j=99
j=100
-0.2
1 2 3 1 2 3
kh kh

Fig. 2.9 Real and imaginary parts of (k1eq /k) j and (k2eq /k)2j are plotted at indicated grid points
for NCCD scheme [98] with explicit boundary closure schemes given by Eqs. (2.129)–(2.130)

j = (N − 1), for which β N −1 = 0.09 is prescribed. One notes that because of the
use of such explicit stencils at j = 2 and (N − 1), no stencil is required at j = 1
or N . In Fig. 2.9, (k1eq /k) j and (k2eq /k) j are plotted at indicated grid points for
NCCD scheme [98] (with explicit boundary closure schemes given by Eqs. (2.129)
and (2.130)), respectively. One clearly notes from this figure that the numerical prob-
lems (in terms of added numerical anti-diffusion and dispersive effects) are mitigated
to a considerable degree in the NCCD scheme. Moreover, the NCCD scheme also
substantially improves the resolution of the second derivative at j = 2 and (N − 1),
compared to the original CCD scheme.

2.3.6 Compact Schemes for Interpolation and First


Derivative Evaluation in Staggered Grids

Sometimes while solving NSE numerically, one needs staggered arrangement of


variables, to achieve higher accuracy of the solution. For the velocity-vorticity for-
mulation of NSE on orthogonal curvilinear grids, staggered grid arrangement of
variables results in less error, as compared to solutions using non-staggered or collo-
cated arrangement of variables [44]. In [70], it was shown while analyzing linearized
52 2 DNS of Navier–Stokes Equation



Fig. 2.10 a Staggered representation of variables in 2D ( V , ω)-formulation. Here, ω refers to the
component of vorticity perpendicular to the 2D plane. b Staggered arrangement in 3D of u-, v- and
w-velocity components and ωx -, ω y - and ωz -vorticity components shown on an elementary cell

shallow water equations, that staggered arrangement of relevant variables provides


best numerical accuracy.
If staggered grid is used for the solution of NSE, as in [8, 9], then one needs
to interpolate and evaluate derivatives at the mid-point locations. For example, if
the variables are defined at integral points, i.e., at jth nodes, then using staggered
arrangement of variables, interpolation or evaluation of derivatives are required
at ( j + 1/2)th nodes. As an example of staggered arrangement of variables in
Fig. 2.10b, staggered arrangement of velocity and vorticity components are shown for
3D problems, while using velocity-vorticity formulation of NSE [7–9]. The velocity
components are defined at the center of the sides (for 2D, Fig. 2.10a) or at the center
of the faces of the elementary cube in the transformed computational plane (for 3D,
Fig. 2.10b). For 2D problems, the only relevant component of vorticity is ωz , i.e., the
vorticity component perpendicular to the 2D plane and in Fig. 2.10a, it is defined at
the corner points of each elementary square cell. For 3D problems, however, all the
three components of vorticity are to be considered and these components are defined
at the center of each sides of the cell as shown in Fig. 2.10b.
The interpolation in staggered grids are carried out by the optimized compact
mid-point interpolation scheme [62]. This interpolation scheme is given as

aI bI
α I û j−1 + û j + α I û j+1 = (u j− 21 + u j+ 21 ) + (u j− 23 + u j+ 23 ) (2.131)
2 2

where uˆj ’s are the interpolated values at the jth-location obtained from the known
u j±n/2 values at the ( j ± n/2)th-locations. For 4th order accuracy, one should have
a I = 18 (9 + 10α I ) and b I = 18 (6α I − 1), with α I as a free parameter. Furthermore, a
2.3 Numerical Methods for Developing DNS/LES 53

choice of α I = 103
yields 6th order accuracy for the interpolation scheme. However,
here α I = 0.42 is obtained by optimizing the integrated phase error of the scheme
in the spectral plane to achieve better DRP properties as given in [7]. Using Fourier–
Laplace transform one can express, û j = Û (k) eikx j dk and u j = U (k) eikx j dk.
Then resolution properties and the performance of an interpolation scheme can be
characterized by a transfer function given by T F = Û (k)/U (k), which is the ratio
of the Fourier amplitude of the interpolated function to that of the original function.
From Eq. (2.131), this transfer function is obtained as,
 
a I cos(kh/2) + b I cos(3kh/2)
T F(kh) = (2.132)
1 + 2α1 cos(kh)

with uniform grid spacing h. Then to minimize the error in the spectral plane, one
can define an objective function given as,
γ
Iinter p = 1 − T F(kh) d(kh) (2.133)
0

which is nothing but the error integrated over the range of kh up to γ , which is
less than or equal to the Nyquist limit. The T F given above is shown plotted as a
function of kh in Fig. 2.11a, for the indicated values of α I . The objective function
Iinter p is shown plotted in Fig. 2.11b, as a function of α I for indicated values of γ . It
is noted clearly that this objective function is minimum at α I = 0.42 for γ = π . This
optimized fourth order scheme has been used here, for all reported computations. For
non-periodic problems, as the flow inside a square lid-driven cavity, the interpolation
scheme given by Eq. (2.131) has to be supplemented by the following boundary
stencil given as
1
û 1 = (u 1 + u 23 )
2 2
1
û N = (u 1 + u N + 21 ) (2.134)
2 N− 2
Such explicit boundary closure does not induce any numerical instability at near-
boundary points.
The evaluation of first derivative in staggered grids can be performed by a scheme,
which is a variation of the scheme given in [62] for derivative evaluation in staggered
grid. The general interior stencil for first derivative (indicated by primed quantity) is
given by

bI I aI I
α I I u j−1 + u j + α I I u j+1 =(u j+3/2 − u j−3/2 ) + (u j+1/2 − u j−1/2 )
3h h
(2.135)

where u j s defined at the jth-location are the first derivatives of the known function
u j±n/2 , defined at the ( j ± n/2)th-locations. Here h is the uniform grid-spacing.
In [62], a fourth-order, single parameter (α I I ) version was considered with b I I =
54 2 DNS of Navier–Stokes Equation

Fig. 2.11 a Transfer


function of the staggered
(a) 1.2
interpolation scheme given
1
by Eq. (2.131) plotted as a
function of kh for indicated 0.8
values of α I . b Integrated
phase error plotted as a 0.6
function of α I for the stencil αI = 0.31

TF
αI = 0.32
for indicated values of γ 0.4
αI = 0.35
αI = 0.38
0.2 αI = 0.30 (6th order)

0
0 0.5 1 1.5 2 2.5 3
kh

γ=1
γ = π/2
(b) γ = 3π/4
100 γ=π
Iinterp (Log scale)

-1
10

-2
10

10-3
6th order
10-4

0 0.1 0.2 0.3 0.4 0.5


αΙ

(22 α I I − 1)/8 and a I I = (9 − 6 α I I )/8 and a sixth order version with α I I = 9/62
was used. However to obtain better spectral properties, an optimized version of the
integrated phase error in the spectral plane for Eq. (2.135) has been developed in
[7]. Using Fourier–Laplace transform, one can define an equivalent wavenumber keq
[98, 107] for the compact scheme in Eq. (2.135) as
 
2 a I I sin(kh/2) + (b I I /3) sin(3kh/2)
keq = (2.136)
h 1 + 2α I I cos(kh)

Therefore, one can define an objective function to be optimized as


γ
keq
Icomp = 1− d(kh) (2.137)
0 k

This objective function Icomp is plotted as a function of α I I in Fig. 2.12b for the
indicated values of γ , while in Fig. 2.12a, keq /k as a function of kh is plotted for
different indicated values of α I I . It is noted from Fig. 2.12 that the optimum value
of α I I = 0.216 is obtained for γ = π . For non-periodic problems, a second order
2.3 Numerical Methods for Developing DNS/LES 55

(a)

(b)

Fig. 2.12 a Spectral resolution of the scheme given by Eq. (2.135), shown for indicated values of
α I I . b Integrated phase error plotted as a function of α I I for the stencil for the indicated values of γ

stencil can be used at j = 1 and N to avoid numerical problems. Such explicit closure
schemes can be given as

1
u1 = (u 3 − u 21 )
h 2
1
u N = (u 1 − u N − 21 ) (2.138)
h N+ 2
56 2 DNS of Navier–Stokes Equation

2.3.7 Discretization of Self-Adjoint Terms

When NSE is solved in the transformed plane (as discussed in Sect. 2.1.6), one notes
the appearance of self-adjoint terms corresponding to viscous diffusion. For example,
consider the following two terms that appear in Eq. (2.31),
   
∂ h 2 ∂ω ∂ h 1 ∂ω
and
∂ξ h 1 ∂ξ ∂η h 2 ∂η

These self-adjoint terms can be generally represented as (g f  )j at the jth node in
terms of the functions g j and f j . Here, the function g j is one of the grid transformation
terms as (h 1 / h 2 ) j and (h 1 / h 2 ) j . For most of the 2D receptivity cases reported here,
these terms are discretized by second order central difference scheme C D2 as,
     
1 g j+1 + g j f j+1 − f j g j + g j−1 f j − f j−1
(g f  )j = −
h 2 h 2 h
(2.139)
A new compact scheme for the discretization of these self-adjoint terms has been
developed in [88], which improves the effectiveness of diffusion discretization over
conventional C D2 -scheme given in Eq. (2.139). This scheme is named in [88] as the
self-adjoint compact difference scheme (SACD). Denoting D j = (g f  )j , the scheme
proposed in [88] is given as

βI
α I D j−1 + D j + α I D j+1 = 2 g j−1/2 f j−1
h
  
+ g j−1/2 + g j+1/2 f j + g j+1/2 f j+1 (2.140)

where β I = 1 + 2α I , g j−1/2 = (g j−1 + g j )/2 and g j+1/2 = (g j + g j+1 )/2. The


scheme given in Eq. (2.140) requires boundary closure scheme. For the variable
f j defined in the nodes ranging from j = 1 to N , one only requires to evaluate the
diffusion terms in nodes ranging from j = 2 to N − 1. The boundary closure of the
SACD scheme is given by the conventional C D2 scheme given by Eq. (2.139).
The spectral analysis of the SACD scheme is done by taking the Fourier–Laplace
transform of the functions, f and g. Here, the function f for a given node j is
transformed in the wavenumber
 plane as: f j = F(k)eikx j dk. In a similar fashion,
 ik  x j
we represent, g j = G(k )e dk  . Hence, for the jth node, the exact derivative
can be written as,
1 
Dj = − k◦2 F(k)G(k  )ei(k+k )x j dk dk  (2.141)
h2

where k◦2 = (kk  + k 2 )h 2 . From Eq. (2.140) for the interior nodes subject to the
boundary closure schemes given by Eq. (2.139), D j can be written in the matrix
form as,
2.3 Numerical Methods for Developing DNS/LES 57

1
[A]{D} = ([B1 ]{ f j g j+1/2 } + [B2 ]{ f j g j−1/2 }) (2.142)
h2
It has been shown in [88], that the expression for equivalent wavenumber at jth
node, keq | j obtained for the discretized form can be expressed as,

 N 
2
keq |j = − (C1 jl P jl + C2 jl Q jl ) (2.143)
l=1

so that,

1 
Dj = − 2
keq F(k)G(k  )ei(k+k )x j dk dk  (2.144)
h2
 
where P jl = eik(x j −xl ) eik (x j −xl + 2 ) , Q jl = eik(x j −xl ) eik (x j −xl − 2 ) , [C1 ] = [A−1 B1 ] and
h h

[C2 ] = [A−1 B2 ]. The quantity (keq 2


/k◦2 ) is in general a complex quantity with the
real part indicating resolution of the scheme and the imaginary part represent-
ing the dispersion error. In Fig. 2.13, the real part of (keq 2
/k◦2 ) is plotted in the

(k h, kh)-plane for different values of α I for j = 50, with N = 100. The coefficient
α I = 0, represents the explicit C D2 scheme given by Eq. (2.139). The imaginary
part of (keq2
/k◦2 ) is found to be very negligible for this case, even at near-boundary
points [88].  2 2
Ideally it is desirable to have the real part of keq /k◦ equal to one for all values of
k and k  in the (kh, k  h)-plane. From Fig. 2.13, it is observed that at the Nyquist limit
(kh = π ), the value of of (keq 2
/k◦2 ) is 4/π 2 , for the explicit scheme (α I = 0). Clearly,
there is an improvement in the resolution properties as α I is increased. This is noted
from the value of (keq 2
/k◦2 ) at the Nyquist limit that increases to 0.49 for α I = 0.05;
0.6086 for α I = 0.1; 0.68 for α I = 0.125; 0.74 for α I = 0.15 and 0.95 for α I = 0.2.
It is noted that the compact scheme over-estimates the second derivative for some
combinations of kh and k  h, in all the cases shown in Fig. 2.13. This is a desirable
property for effective control of aliasing errors for DNS/ LES. For α I = 0.1, the
resolution at Nyquist limit is about 50% higher than that is achieved by the explicit
scheme. For the same α I , the compact scheme over-estimates the derivative by only
10.0%- that too only for a very small range of values in (kh, k  h)-plane. However,
too much overestimation of (keq 2
/k◦2 ) can also alter the physical dispersion property
of the problem. In [88], the bi-directional wave equation given by

∂2 f 2∂ f
2
= c (2.145)
∂t 2 ∂x2
has been solved
 in the transformed
 ξ -coordinate. The transformation function is given
tanh{β̄(1−2ξ )}
by x(ξ ) = 1 − tanh(β̄)
. In the transformed plane, Eq. (2.145) is transformed
to
58 2 DNS of Navier–Stokes Equation

2 /k 2 for different values of α in (kh, k  h)-plane for


Fig. 2.13 Contour plots for the real part of keq ◦ I
50th node, when total number points is N = 100 for the SACD scheme given by Eq. (2.140). The
maximum and the minimum value for each cases have been indicated
2.3 Numerical Methods for Developing DNS/LES 59
 
∂2 f c2 ∂ 1 ∂f
= (2.146)
∂t 2 xξ ∂ξ xξ ∂ξ

where 1/xξ represents the metric of the transformation. In [88], the squared integral

error S E = Nj=1 ( f exact − f num )2 is plotted in (α I , β̄)-plane and it is found that
least error is committed for α I = 0.1. This is proposed in [88] as the optimum value
of α I for the SACD scheme given by Eq. (2.140).
We have discussed optimized staggered compact scheme (OSCS) in Sect. 2.3.6.
This is the optimized version of the staggered compact schemes proposed in [62].
One can also use these schemes twice to evaluate the self-adjoint diffusive terms.
For the boundary closure, C D2 scheme can be used. According to the proposed

methodology, one has to first evaluate f j+1/2 (the first derivative of the function f at
the ( j + 1/2)th point) using the following stencil given as
   
   f j+2 − f j−1 f j+1 − f j
α I I f j−1/2 + f j+1/2 + α I I f j+3/2 = b I I + aI I
3h h
(2.147)

subject to the following boundary stencils at j = 3/2 and j = N − 1/2 given as

 f2 − f1
f 3/2 =
h
f N − f N −1
f N −1/2 = (2.148)
h
As mentioned in Sect. 2.3.6, Eq. (2.147) is a single parameter (α I I ) family of fourth
order schemes for a I I = (9 − 6α I I )/8 and b I I = (−1 + 22α I I )/8. After evaluating

f j+1/2 from Eqs. (2.147) and (2.148) one proceeds to evaluate D j using the stencil
given as
 
σ j+3/2 − σ j−3/2
α I I D j−1 + D j + α I I D j+1 = bI I
3h
 
σ j+1/2 − σ j−1/2
+a I I (2.149)
h

using the following boundary stencils at j = 2 and j = N − 1 given as

σ5/2 − σ3/2
D2 =
h
σ N −1/2 − σ N −3/2
D N −1 = (2.150)
h

where σ j+1/2 = g j+1/2 f j+1/2 and g j+1/2 = (g j + g j+1 )/2. The optimized value of
α I I = 0.22 is mentioned in Sect. 2.3.6, whereas α I I = 9/62 makes Eqs. (2.147) and
60 2 DNS of Navier–Stokes Equation

2 /k 2 for different values of α in (kh, k  h)-plane for


Fig. 2.14 Contour plots for the real part of keq ◦ II
50th node, when total number points is N = 100, using OSCS scheme twice to evaluate the second
derivative, as given by Eqs. (2.147)–(2.150). The maximum and the minimum values for each cases
have been indicated

(2.150) as 6th-order accurate [62]. A similar full-domain analysis can be performed


for twice repeated OSCS scheme for evaluating self-adjoint terms. In Fig. 2.14, the
real part of (keq
2
/k◦2 ) is plotted for j = 50, with N = 100. It has been found that,
α I I = 0.22 provides the least error for the OSCS scheme. Although the proposed
methodology requires the solution of tridiagonal matrix twice, as compared to only
once for SACD scheme, it produces lesser error compared to the latter. This method-
ology is used for some 2D receptivity calculations and all 3D receptivity calculations,
reported in later chapters.
2.3 Numerical Methods for Developing DNS/LES 61

2.3.8 Computing Methods for Unsteady Flows: Dispersion


Relation Preserving (DRP) Methods

We have discussed the dispersion relation applicable to space-time dependent prob-


lems. In the context of transitional and turbulent flows, we have also noted the range
of length and time scales in terms of spectra of such flows. It is noted that DNS would
require preservation of the physical dispersion relation in the numerical sense. We
have noted the role of group velocity in transporting energy of a dynamical system.
One of the main aims of the present section is to relate group velocity with the disper-
sion property of the system numerically. However, apart from matching physical and
numerical group velocity, one would also like to ensure the overall accuracy of the
numerical solution and relate the accuracy with the properties of adopted numerical
methods.
In this context, it has been noted that Eq. (2.43) serves well the purpose of a
model equation to check the accuracy of any numerical method involving space and
time discretization. We have already established that this is a non-dissipative and
non-dispersive equation, which convects the initial solution unattenuated to the right
at the speed c, representing both phase speed and group velocity.
This model equation is non-dispersive and provides a simple, yet a tough test for
any combined space-time discretization method’s dispersion property. We emphasize
that the study of spatial and temporal discretizations separately is improper. Such
studies can be found variously in [52, 111, 116]. Vichnevetsky and Bowles [120]
have reported the dispersion property of a single finite difference and a finite element
method by heuristic approaches. A proper estimate of the numerical dispersion prop-
erty has been advanced in [92, 95, 101]. The explanation here is based on analysis
of numerical schemes developed over the full computational domain in [96].

2.3.9 Numerical Amplification Factor for 1D Convection


Equation

Analyses of numerical methods and associated error propagation study have been
performed following different routes by many researchers. From the discretized gov-
erning equation, one can work backwards to obtain an equivalent differential equa-
tion, that has been effectively solved by the discrete equation, an approach advocated
in [108]. According to this, if the truncation error terms are retained for space-time
dependent problems, then the corresponding differential form is called the Γ -form.
If the retained truncation error terms are for space derivatives alone, then the corre-
sponding differential form is called the Π -form. While most analyses in the literature
have used the Π -form, only in a few, the Γ -form has been utilized [93, 95]. The
present approach follows the same route, where both space and time discretizations
are considered simultaneously. The irrelevance of Π -form has been clearty explained
in [95].
62 2 DNS of Navier–Stokes Equation

In the classical approach attributed to von Neumann, the evolving error of the
discretized differential equation with linear constant coefficients is assumed to fol-
low an identical discrete equation. In this approach, the difference between an exact
and a computed solution arises due to round-off error and error in the initial data.
For periodic problems, error is furthermore decomposed into Fourier series and the
individual normal modes are investigated. This approach is also extended for nonlin-
ear systems and linear systems with variable coefficients. For nonlinear systems, the
principle of superposition does not hold and additionally problem of aliasing arises
due to nonphysical transfer of energy to wrong wavenumbers and frequencies in the
computational plane, whenever products are numerically evaluated with finite grid
resolution.
There have been many efforts in analyzing error dynamics, using the method
attributed to von Neumann, as in [16, 20]. The main assumption for linear problems
that the error and the signal follow the same dynamics appears intuitively correct. It
has been unambiguously shown in [95], that this assumption is flawed for any discrete
computing and the difference is due to dispersion and phase errors, and when the
numerical method causes physical amplification and/or attenuation.
We demonstrate the inadequacy of von Neumann analysis with the help of
Eq. (2.43) in analyzing space-time discretization schemes. This equation helps in
testing numerical methods for solution accuracy, error propagation and most impor-
tantly, the dispersion error − as has been variously attempted in [92, 95, 101, 120,
132].
We represent the unknown, by its Laplace transform  at the jth node of a uni-
formly spaced discrete grid of spacing h as u(x j , t) = U (kh, t) eikx j dk. For the
1D convection equation, one can define the spectral amplification factor G as

U (kh, t n + Δt)
G(Δt, kh) = (2.151)
U (kh, t n )

Note that G is a complex quantity, as the Fourier–Laplace amplitude U (kh, t) is


complex. In the continuum limit of h, Δt → 0, one must have

|G| ≡ 1 (2.152)

However, real and imaginary parts of G depend on the differential equation


solved and the discretization methods adopted for spatial and temporal derivatives.
Other important numerical properties are obtained via the spectral representation of
Eq. (2.43) and using Eq. (2.90) to represent the first derivative. This gives
 
dU c ik(xl −x j )
+ U C jl e eikx j dk = 0 (2.153)
dt h

Since the above equation is true for all wavenumbers, the integrand must be zero
for any k. The implicit condition of Eq. (2.153) can be reinterpreted as
2.3 Numerical Methods for Developing DNS/LES 63

  N
dU cdt
=− C jl eik(xl −x j ) (2.154)
U h l=1

The first factor on the right-hand side of Eq. (2.154) is nothing but the CFL
number (Nc ). As the right-hand side of Eq. (2.154) is node dependent, we express
the left-hand side by the nodal numerical amplification factor (G j ) given by

N
G j = G|(x=x j ) = 1 − Nc C jl eik(xl −x j ) (2.155)
l=1

for the Euler time discretization scheme. From Eq. (2.91), one can replace the
summed up quantity on the right-hand side as

G j = G|(x=x j ) = 1 − i Nc [keq h] j (2.156)

This numerical property can not be expressed in terms of spatial discretization


alone. Euler time integration is thus, not numerically stable, as |G j | > 1. Two time
level, multi-stage higher order methods are needed ideally for accuracy and stability -
as explained with respect to the (R K 4 ) method [83, 95]. If one denotes the right-hand
side of Eq. (2.43), by L(u) = −c ∂∂ux , then the steps used in (R K 4 ) are given by

Δt
Step 1 : u (1) = u (n) + L[u (n) ]
2
Δt
Step 2 : u (2) = u (n) + L[u (1) ]
2

Step 3 : u (3) = u (n) + Δt L[u (2) ]

Δt
Step 4 : u (n+1) = u (n) + {L[u (n) ] + 2L[u (1) ] + 2L[u (2) ] + L[u (3) ]}
6
For the R K 4 time integration scheme, G j is obtained as [95]

A2j A3j A4j


G j = 1 − Aj + − + (2.157)
2 6 24
where
N
A j = Nc C jl eik(xl −x j )
l=1

While G j is a source of error, additional error arises due to dispersion, whose effects
are subtle and often misunderstood.
64 2 DNS of Navier–Stokes Equation

2.3.10 Quantification of Dispersion Error and Error


Propagation Equation

Using Fourier transform for the analysis of a numerical method, one obtains a
numerical dispersion relation by expressing the equivalent differential equation in
the wavenumber-circular frequency plane. This dispersion relation is different from
the physical dispersion relation and the difference between the two gives rise to
dispersion error, as explained with the 1D model convection equation for a typical
space-time dependent system.
If we represent the initial condition for Eq. (2.43) as

u(x j , t = 0) = u 0j = A0 (k) eikx j dk (2.158)

then the general solution at any arbitrary time can be obtained as

u nj = A0 (k) [|G j |]n ei(kx j −nβ j ) dk (2.159)

G
where |G j | = (G r2j + G i2j )1/2 and tan(β j ) = − G ri jj , with G r j and G i j as the real and
imaginary parts of G j , respectively. Thus, the phase of the solution is determined by
nβ j = kc N t, where c N is the numerical phase speed. Although the physical phase
speed is a constant for all k for the non-dispersive system, this analysis shows that
c N is, in general, k-dependent, i.e., the numerical solution is dispersive, in contrast
to the non-dispersive nature of Eq. (2.43). The implications of this simple difference
are profound, as demonstrated below.
The general numerical solution of Eq. (2.43) is denoted as

ū N = A0 [|G|]t/Δt eik(x−c N t) dk (2.160)

The numerical dispersion relation is now given as ω N = c N k, instead of the phys-


ical dispersion relation, ω̄ = ck. Non-dimensional phase speed and group velocity
(from the general definition given in Eqs. (2.48) and (2.49)) at the jth node are
expressed as  
cN βj
= (2.161)
c j ω̄Δt
 
Vg N 1 dβ j
= (2.162)
c j h Nc dk

If the computational error is defined as e(x, t) = u(x, t) − ū N , then one obtains


the governing equation for its dynamics in the following manner. Using Eq. (2.160)
one obtains
2.3 Numerical Methods for Developing DNS/LES 65

∂ ū N
= ik A0 [|G|]t/Δt eik(x−c N t) dk (2.163)
∂x

∂ ū N
=− ik c N A0 [|G|]t/Δt eik(x−c N t) dk
∂t

Ln |G|
+ A0 [|G|]t/Δt eik(x−c N t) dk (2.164)
Δt

Thus, the error propagation equation for Eq. (2.43) is given by [95],
 
∂e ∂e c N ∂ ū N dc N 
+c = −c[1 − ] − ik  A0 [|G|]t/Δt eik (x−c N t) dk  dk
∂t ∂x c ∂x dk

Ln |G|
− A0 [|G|]t/Δt eik(x−c N t) dk (2.165)
Δt

This is the correct error propagation equation, as opposed to that is obtained using
the assumption made in von Neumann analysis, where the right-hand side is assumed
to be identically equal to zero. This is on the premise that c N ∼ = c, i.e., there is no
dispersion error and the numerical method is perfectly neutral, so that the last term
on the right-hand side of Eq. (2.165) is identically zero. Error can grow even faster
when the numerical solution displays sharp spatial variation, due to the first term on
the right-hand side of Eq. (2.165).
To understand the ramifications of Eq. (2.165) for the model equation, one needs
to look at the numerical properties of a specific combination of spatial and temporal
discretization methods. For this purpose, we show the properties of OUCS3 scheme
[96]. The stencil of OUCS3 scheme is given in Eq. (2.83). This scheme is for a non-
periodic problem and here we show some typical results for a central and another
node close to the inflow boundary, when R K 4 time integration strategy is used to
solve Eq. (2.43). In Fig. 2.15, |G|, Vg N /c and (1 − c N /c) are plotted as contours in
the indicated ranges of kh and Nc for a node adjacent to the boundary ( j = 2) and
one in the center of the domain ( j = 51).
Results show that the numerical properties of j = 2 are significantly different
from those for j = 51. It is noted from the top frame for numerical amplification
contours that the scheme is stable for Nc ≤ 1.301, i.e., |G| ≤ 1 for interior nodes. The
scheme is neutrally stable for very small values of Nc and a limited range of kh − a
property absolutely essential for DNS, as one notes that the last term on the right-hand
side of Eq. (2.165) vanishes for the neutrally stable case. In the middle frame of Fig.
2.15, Vg N /c contours display significant dispersion effects for high ks, which would
invalidate long time integration results, even when |G j | = 1 is ensured by computing
with vanishingly small Nc . In fact, above kh ≥ 2.4 the numerical solution will travel
in the wrong direction, as Vg N ≤ 0 for Nc ∼ = 0. Such spurious numerical waves are
termed q-waves, which propagate upstream, in contrast to the physical or p-waves
66 2 DNS of Navier–Stokes Equation

Fig. 2.15 Comparison of numerical amplification factor (|G|), normalized numerical group velocity
(Vg N /c) and phase error (1 − c N /c) contours for the near-boundary node j = 2 (left column) and
the central node j = 51 (right column) for the 1D convection equation solved by the OUCS3-R K 4
method

traveling downstream for Eq. (2.43) [91]. In Eq. (2.165), we note that the first term
on the right hand side affects error evolution via the numerical property (1 − ccN ).
In the bottom frame of Fig. 2.15, contours of this quantity are shown. The bottom
two frames of Fig. 2.15 indicate effects of dispersion error, which cannot be simply
eliminated or reduced by grid refinement alone.
2.3 Numerical Methods for Developing DNS/LES 67

One can show the connection between q-waves as the extreme form of dispersion
error for different numerical methods used in CFD. Even for a large range of k,
corresponding p-waves travel at incorrect group velocity, constituting dispersion
error. The q-wave is entirely due to dispersion in many discrete computation methods
[91].
Flow transition was thought often to occur for external flows by TS wave packets
created as physical instability to small disturbances inside a shear layer. However,
many flows do not show TS wave and are collectively stated to have suffered bypass
transition. In [82], one particular form of bypass transition is explained theoretically
and experimentally, when the flow is excited by a disturbance source remaining
always outside the shear layer. In this scenario, the resultant disturbance occurring
inside the shear layer is seen to propagate upstream, with respect to the source
of disturbance convecting outside the shear layer. Thus, there are flow transition
scenarios where physical disturbances travel upstream and it is necessary to compute
these, without the interference of q-waves created numerically.
For attached flows, q-waves have been reported in [68] as very high k events. Thus,
for DNS of bypass transitional flows, one would like to capture p-waves associated
with bypass transition, while avoiding q-waves which arise solely due to dispersion
error at high wavenumbers. For the zero pressure gradient boundary layer, upstream
propagating damped waves were detected theoretically in [99]. We emphasize that
in many flows, disturbances appear to move downstream with respect to an inertial
frame − but an observer moving with the source at the convection speed will see the
response inside the shear layer move upstream.

2.3.11 Dispersion Relation Preserving Schemes

Significant progress has been made in developing numerical methods for solving
space-time dependent problems with schemes, which preserve physical dispersion
relation over a wide range of parameters. For discrete methods, closeness between
the physical and numerical dispersion relation can be realized only for limited ranges
of space and time steps. In [96], a spectral analysis was developed to characterize
numerical schemes in the full domain, with numerical group velocity used as a mea-
sure to quantify dispersion error. Space-time discretization was treated independently
in many earlier works. This has been corrected in [95, 100] and with the derivation
of Eq. (2.165), a DRP scheme was identified as essential for DNS and acoustics
problems. These results have been used in [71, 100] to optimize time discretization
schemes for high accuracy spatial discretization schemes, by minimizing numerical
error in the spectral plane.
Here, we have chosen representative methods from finite difference, finite volume
and finite element approaches to explain q-waves and quantifying DRP properties.
For the finite difference method, OUCS3 scheme [96] is chosen. In solving Eq. (2.43),
the first spatial derivative indicated by a prime is obtained from the general repre-
sentation
68 2 DNS of Navier–Stokes Equation

1
[A] {u  } = [B] {u} (2.166)
h
For the explicit discretization method, [A] is the identity matrix and the [C] matrix
in Eqs. (2.90) and (2.91) is identical to the [B] matrix. For implicit schemes, [C]
is equal to [A]−1 [B] matrix. Having obtained [C] for spatial discretization, one
evaluates the complex amplification factor G j for any time integration schemes. In
particular, for the R K 4 method one can obtain G j from Eq. (2.157) with

N
A j = Nc C jl eik(xl −x j )
l=1

This approach is used to obtain G j for explicit C D2 and the OUCS3 schemes for
spatial discretization (Eq. (2.83)) with R K 4 time integration scheme.
For a representative finite volume scheme, the QUICK scheme [54] is considered,
which uses flux vector splitting of the convection term as [83],
 
∂u
d x + c (u +j+1/2 − u +j−1/2 ) = 0 (2.167)
∂t

In the above, R K 4 time integration is used for the jth cell and the superscript sign
for the second set of terms indicates right moving quantities showing the balance of
incoming and outgoing fluxes through the cell interfaces. One of the representative
flux quantities is given by (see [42, 83] for details)

1
u +j−1/2 = u j−1 + [(1 − κ)(u j−1 − u j−2 ) + (1 + κ)(u j − u j−1 )] (2.168)
4

where κ = 1/2 for the QUICK scheme. Other flux terms u +j+1/2 can be similarly
obtained. These expressions help one obtain the [C] matrix in Eq. (2.90), which
in turn yields G j . Readers are referred to [97] for a similar analysis for a compact
scheme based flux vector splitting finite volume method.
For the finite element method, the streamwise upwind Petrov–Galerkin (SUPG)
method described in [36] is analyzed. Among finite element formulations, Galerkin
methods belong to the class of solutions for PDEs in which solution residue is mini-
mized, giving rise to the well-known weak formulation of problems. In this approach,
dependent variable u(x, t) for a 1D space-time dependent problem is expressed in
the form
N
u(x, t) = φ j (x) u j (t) (2.169)
j=1

where φ1 , . . . , φ N are the chosen low order polynomials as the basis functions, local-
ized about elements. Note that the weak formulation considers space-time depen-
dent terms together, giving rise to better DRP properties for Galerkin formulations.
2.3 Numerical Methods for Developing DNS/LES 69

However, it has been identified in [83, 91] that the Galerkin methods display insta-
bility near the inflow when linear basis functions are used (G1FEM) for Eq. (2.43).
Similarly, observations on the amplification factor holds for solving the 1D convec-
tion equation by quadratic basis functions by G2FEM.
One method of removing the problems of Galerkin FEM is to make the discrete
equation dissipative. Authors in [24, 123, 124] have proposed dissipative Galerkin
procedures. This procedure was furthermore adopted in [75], who advocated an opti-
mal procedure following an approximate form of phase error proposed in [111] for 1D
convection equation. Brooks and Hughes [14] have adopted the same methodology
in the Petrov–Galerkin formulation and called it the “Streamline Upwind/Petrov–
Galerkin (SUPG)” formulation. Raymond and Garder [75] have also referred to
“ghost waves” following [22]. A complete analysis of the “ghost” or q-waves is pre-
sented here for the SUPG formulation, following the work on q-wave [83, 91] for
different discrete computing methods.
The discretized form of Eq. (2.43) for G1FEM is obtained using linear basis
functions on a uniform grid as
 
h du j+1 du j du j−1 c
+4 + + (u j+1 − u j−1 ) = 0 (2.170)
6 dt dt dt 2

for any interior jth-node. For boundary nodes ( j = 1, N ) one obtains,


 
h du 1 du 2
j =1: 2 + + c(u 2 − u 1 ) = 0 (2.171)
3 dt dt
 
h du N du N −1
j=N: 2 + + c(u N − u N −1 ) = 0 (2.172)
3 dt dt

The derivations are as given in [83, 91].


Three aspects are evident from the above discrete equations for G1FEM: (i) the
non-dissipative nature of the discrete equation for the interior nodes; (ii) instability at
j = 1 due to the one-sided nature of the stencil, with the information propagating to
the boundaries from the interior, which is contrary to the physical description given
by Eq. (2.43) and (iii) the overly dissipative nature of the discrete equation at j = N .
Using the hybrid Fourier–Laplace representation of u in Eqs. (2.170)–(2.172),
(1)
one obtains the effectiveness of the derivative discretization keq for the G1FEM as
[81, 83]  
(1) 3 eikh − 1
j =1: keq = (2.173)
i h eikh + 2

(1) 3 sin kh
2≤ j ≤ N −1: keq = (2.174)
[h(2 + cos kh)]
 
(1) 3 1 − e−ikh
j=N: keq = (2.175)
i h 2 + e−ikh
70 2 DNS of Navier–Stokes Equation

If one adopts Euler time stepping for discretized Eqs. (2.170)–(2.172), the numer-
ical amplification factor for G1FEM is given, as in [81, 83]
 
(1) eikh − 1
j =1: G = 1 − 3Nc (2.176)
eikh + 2
 
(1) sin kh
2≤ j ≤ N −1: G = 1 − 3i Nc (2.177)
2 + cos kh
 
1 − e−ikh
j=N: G (1) = 1 − 3Nc (2.178)
2 + e−ikh

If instead one uses quadratic basis functions (G2FEM) for a uniformly spaced
grid, then the discrete equations for the Galerkin approximation of Eq. (2.43) are as
given in [81, 83]. Here, we report only the equation for the interior nodes as
 
h du l+2 du l+1 du l du l−1 du l−2
3≤ j ≤ N −2: − +4 + 24 +4 −
15 dt dt dt dt dt
 
c
+ u l−2 − 8u l−1 + 8u l+1 − u l+2 = 0 (2.179)
6

Using hybrid Fourier–Laplace transform of the unknown u in the above, one


(2)
obtains the effectiveness of the derivative discretization keq for G2FEM for an interior
node as

(2) 5 sin kh(4 − cos kh)


3≤ j ≤ N −2: keq = (2.180)
h(12 + 4 cos kh − cos 2kh)

If one adopts Euler time stepping for the discretized equations for G2FEM, one
obtains the numerical amplification factor for G2FEM as

sin kh(4 − cos kh)


3≤ j ≤ N −2: G (2) = 1 − 5i Nc (2.181)
(12 + 4 cos kh − cos 2kh)

It is seen that G1FEM suffers from numerical instability at j = 1 and G2FEM


suffers the same for j = 1 and 2. Also, it is shown in [81, 83] that there is overstability
at j = N for G1FEM and at j = (N − 1) and N for G2FEM.
To apply Galerkin methods for wave propagation problems, numerical instabilities
occurring at and near boundaries must be cured − as mentioned in [24, 75, 123, 124].
Brooks and Hughes [14] adopted the prescription in [75] and renamed it as SUPG
method. In Grescho and Sani [36], the discrete equation obtained by the SUPG
method at an interior node (2 ≤ j ≤ (N − 1)) for Eq. (2.43) is given as
2.3 Numerical Methods for Developing DNS/LES 71
      
h β du j−1 du j β du j+1 c
1+ +4 + 1− + u j+1 − u j−1
6 2 dt dt 2 dt 2
 
= βc u j+1 − 2u j + u j−1 (2.182)

where√β is the stream-wise diffusion parameter. An optimal value of β is given


as 1/ 15 in [14] following the work in [75]. In [75], the analytical solution of
Eq. (2.182) was taken from [52, 111] in arriving at the optimal value. However, the
analytical solution was derived on the basis of a semi-discrete analysis, considering
no error being committed in time discretization. Hence, this value is not universal
and it varies from one time discretization method to another, a fact often ignored
among practitioners. The right-hand side of Eq. (2.182) represents the lowest order
SU P G
dissipation term. The keq of the SUPG method for discrete Eq. (2.182) can be
obtained as   
6 sin kh − 2iβ(1 − cos kh)
SU P G
keq = (2.183)
h 4 + 2 cos kh − iβ sin kh

Adopting Euler time stepping for Eq. (2.182), the numerical amplification factor
for the SUPG method can be obtained as
 
4(1 − cos kh)(2 + cos kh) − sin2 kh
G SU P G
= 1 − 6β Nc
4(2 + cos kh)2 + β 2 sin2 kh
 
(4 + 2 cos kh) sin kh + 2β 2 (1 − cos kh) sin kh
− 6i Nc (2.184)
4(2 + cos kh)2 + β 2 sin2 kh

In Figs. 2.16 and 2.17, we specifically show G j (Nc , kh) and Vg N /c, respectively,
for the chosen four methods: (a) R K 4 -OUCS3, (b) R K 4 -C D2 , (c) R K 4 -QUICK and
(d) Euler-SUPG obtained using Eqs. (2.155), (2.157) and (2.162) [83, 91]. In these
figures, properties of these methods for only the interior nodes are shown. From the
contour plots in Fig. 2.16, one notes a narrow range of Nc available for which the
OUCS3 and C D2 schemes have neutral stability for a full range of available kh.
The region in the (Nc , kh)-plane where this is feasible is marked by hatched lines in
Fig. 2.16. In comparison, neither the QUICK nor the SUPG scheme has a neutrally
stable region and would not qualify for DNS.
The main interest here is to compare the dispersion property of these methods in
terms of numerical group velocity and also find the reason for which q-waves are
created numerically in solving Eq. (2.43). The Vg N /c contours shown in Fig. 2.17,
indicate that all these methods produce dispersion error, for the computational param-
eters, when the kh and Nc combination takes values away from the origin. The Vg N /c
contours show for these four methods, presence of line(s) along which Vg N is zero.
For the C D2 and OUCS3 schemes, this is a kh = constant line, parallel to the Nc -
axis, while for the other two methods, these lines are curved. From Eq. (2.162), the
condition of zero group velocity corresponds to
72 2 DNS of Navier–Stokes Equation

Fig. 2.16 Comparison of the numerical amplification factor (|G|) for interior nodes by indicated
space-time discretization schemes
2.3 Numerical Methods for Developing DNS/LES 73

Fig. 2.17 Comparison of normalized numerical group velocity (Vg N /c) for interior nodes for the
indicated combinations of space-time discretizations
74 2 DNS of Navier–Stokes Equation
  
d Gi j
tan−1 − =0
dk Gr j

This can be further simplified to

∂G i j ∂G r j
Gr j = Gi j
∂k ∂k
For the R K 4 time integration and central spatial schemes, this condition further
simplifies to
dkeq
=0
dk
as A j = −ikeq Nc . Above this line for C D2 − R K 4 and OU C S3 − R K 4 methods,
the numerical group velocity of the methods for the solution of Eq. (2.43) is negative,
i.e., the numerical waves would propagate upstream, despite the physical require-
ment of downstream movement. The region in the (Nc , kh)-plane where q-waves
are created is marked by hatched lines in Fig. 2.17. Thus, in solving Eq. (2.43) by
the R K 4 -C D2 method, q-waves are created for kh > π/2, and for the R K 4 -OUCS3
method, q-waves are created for kh > 2.391, for any choice of time steps. This criti-
cal value, (kh)cr , for the OUCS3 method is significantly higher in comparison to the
other three methods under consideration. In Fig. 2.17 for the C D2 method, Vg N /c
contours are symmetric about (kh)cr and also due to the symmetry of the stencil,
minimum and maximum values of Vg N /c are same. Maximum and minimum val-
ues of numerical group velocities are lowest for the C D2 method, followed by the
OUCS3 method. In comparison, the QUICK and SUPG methods have a significantly
higher magnitude of maximum and minimum group velocity − essentially due to
the addition of excessive numerical diffusion. It is often wrongly presumed that the
addition of diffusion does not alter the dispersion property of numerical methods.
The presence of q-waves indicates effects similar to what is noted as unsteady flow
separation, which is often the harbinger of bypass transition, as shown in [82].
The presence or absence of q-waves depends not only upon the existence of a
region with Vg N < 0, but also on the bandwidth of the implicit filter (as given by the
real part of keq /k) and added diffusion (as given by the imaginary part of keq /k) of
the basic numerical method. It is easy to reason that excessive filtering and damping
removes q-waves. The appearance of q-waves is more likely for the OUCS3 method
as compared to other methods, due to its lower filtering and less added diffusion. For
the C D2 method, q-waves will appear for kh > π/2, where the low-pass property of
the method will filter the signal that is 38% at each time step, and in contrast, for the
OUCS3 method at (kh)cr = 2.391, corresponding filtering is equal to only 7%, and
there is an additional small attenuation due to low numerical diffusion. It is seen that
at kh = 2.4, the C D2 method filters the signal (while spatially discretizing) by more
than 70%. In comparison to these two methods, the QUICK and SUPG methods add
excessive numerical diffusion, which prevents the appearance of q-waves at the cost
of numerical accuracy. For the SUPG method, q-waves are created for very small
2.3 Numerical Methods for Developing DNS/LES 75

values of Nc and very high values of kh. However, at these high values of kh, the
SUPG method introduces a very large amount of numerical diffusion, due to which
q-waves are not seen, once again.
The above discussion shows the relative merits of different methods with respect
to dispersion properties, including creation of spurious upstream propagating waves,
as example of an extreme form of dispersion error. It is shown that high accuracy
compact schemes used with the R K 4 scheme provide the best option in terms of the
DRP property. There are many other subtle issues of DRP schemes.

2.4 Design of DRP Schemes for DNS/LES

There has been considerable effort to improve the accuracy of numerical meth-
ods in solving space-time-dependent problems by developing DRP schemes. Earlier
attempts are reported in [39, 112]. In many earlier attempts, DRP methods were
attempted to be obtained by considering the spatial discretization alone, mainly by
minimizing the truncation error [12, 112]. In [5, 45, 67, 72], space-time discretiza-
tions have been considered together to minimize the error between the numerical
amplification factor and the true amplification factor. Based on this approach, Hu
et al. [45], proposed a class of low-dissipation and low-dispersion Runge–Kutta
(LDDRK) schemes. Correct numerical dispersion relation for combined space-time
discretization schemes was identified for convection equation in [25, 83, 93], where
the use of DRP schemes for DNS and acoustics problems (see also [106]) was empha-
sized. To obtain a scheme which would provide ideal DRP property in [71, 100],
space-time discretization has been considered together to minimize dispersion and
phase error, while keeping the scheme neutrally stable for 1D convection equation.
Such a scheme should be used for DNS and acoustics.
Following Eq. (2.157), numerical amplification factor G num for p-stage explicit
Runge–Kutta scheme can be expressed as
p
gnum = 1 + (−1) j a j A j (2.185)
1

One notes that for 1D convection equation, exact amplification factor is given as
G exact = ei Nc kh , Eq. (2.185) is a polynomial approximation of G exact and a j = 1/j!
minimizes the corresponding truncation error in temporal space. Hu et al. [45]
considered the 1D convection equation to minimize the difference between G num
and G exact . Similarly in [5, 67], an optimization was performed by minimizing
|G num − ei Nc kh | ||u 0 ||2 , where n = t/Δt and ||u 0 ||2 defines the L 2 -norm of the ini-
tial condition u 0 (x). In [71, 100] following objective functions were chosen to obtain
the DRP scheme:
76 2 DNS of Navier–Stokes Equation

γ
F(a j , Nc ) = |G num − G exact |2 d(kh) (2.186)
0

where γ = απ with α = [0, 1]. This condition is identical to what was used in [45].
However, in addition to minimizing F, additional explicit constraints were also
imposed in [71, 100], to ensure neutral stability, minimum phase and dispersion
errors. Following the source terms in the correct error propagation equation in Eq.
(2.165). These constraints are given as [71, 100],
γ1
F1 (a j , Nc ) = |1 − |G num ||d(kh) ≤ ε1 (2.187)
0
γ2
cN
F2 (a j , Nc ) = |d(kh) ≤ ε2
|1 − (2.188)
0 c
γ3
Vg N
F3 (a j , Nc ) = |1 − |d(kh) ≤ ε3 (2.189)
0 c

where εi and γi are constants chosen to satisfy the numerical properties of the basic
method. Although, a small tolerance in εi was prescribed, main focus was to look for
numerical parameters, which ensure neutral stability for large range of kh and Nc .
Grid-search technique was employed to solve the constrained optimization problem
and to locate the feasible values of parameters a j for a fixed value of Nc . While
optimized a j ’s are functions of Nc , a near optimal values of these parameters were
searched that could be used over a longer range of Nc .

2.5 Numerical Filtering: Error Control via Stabilization


and Dealiasing

Computing of transitional or turbulent flows is very challenging. If not treated care-


fully, the high wavenumber components (components whose wavenumbers are near
to kh = π ) causes numerical instability due to aliasing error. Hence, the use of
filters are quite essential to periodically remove these critical high wavenumber fluc-
tuations, without affecting the actual dispersion property of the solution. Controlled
amount of explicit fourth-order diffusion is already added to the first derivative, but
their excessive use is not recommended, as this might alter the numerical dispersion
relation. Instead the use of filters, that attenuates the high wavenumber components
are prescribed in Bhumkar [10], without altering the physical dispersion relation,
and is recommended.

2.5.1 Use of Filters for DNS and LES

Filtering is always implicitly present in all discrete computations [53, 81, 83]. In
numerical computations, the nature of filter used has to be low-pass, which should
2.5 Numerical Filtering: Error Control via Stabilization and Dealiasing 77

leave smaller wavenumber components of the variable unaffected, while filtering the
higher wavenumber components, which can create numerical problems as explained
above. In traditional LES, the governing equation is analytically filtered before dis-
cretization. In [34, 48, 59, 61, 80, 110, 119] the effects of different types of filters
used in traditional LES are illustrated. The filtering operation in traditional LES
requires adding additional empirical stresses (termed as SGS or Leonard stresses) to
be incorporated into the governing equation at the formulation stage itself. Such an
operation can lead to numerical instabilities due to aliasing operation as traditional
LES requires the variables to be multiplied with a spatial kernel as


→ → −
− → −

f˜( X , t) = G( X − X  ) f ( X  , t)d 3 X  (2.190)

In contrast, the spatial filters [10, 28–30, 77, 78, 89, 90, 122] are applied at
the end of time advancement of the governing variables, without the need for any
artificial modeling of SGS stresses. Though, no strict rules exist on the frequency
and order of filtering, significant advantages can be achieved, if LES is performed
using spatial filters, where one can completely dispense with empirical SGS model.
Moreover, it can also help avoiding numerical instabilities due to aliasing operation
involved in traditional LES, as explained above.
Explicit Padè type filters have been proposed in [10, 28–30, 83, 89, 90, 122]
to control instabilities arising from mesh non-uniformities and the application of
numerical boundary conditions and advanced the use of spatial filters as a tool for
LES. These spatial filters are central in nature for the interior nodes, while one-sided
stencil have been proposed for non-periodic problems [29, 121] near the boundary.
However, it has been shown in [90] by analysis, using matrix-spectral theory that one-
sided stencils at the boundary can be destabilizing near the inflow of a computational
domain. Such numerical instability can affect also interior points.
One dimensional filters have been proposed and developed in [122]. The general
interior stencil of these 1D filters are given as,

N
an
α f û j−1 + û j + α f û j−1 = (u j+n + u j−n ) (2.191)
n=0
2

which is defined as a (2N )th-order filter. Here, û j is the filtered value of the variable
u j at jth grid-point It is to be emphasized that the definition of the order of filter is
meaningless activity, as in Eq. (2.191). one does not even match the filtered and unfil-
tered variable itself. Original definition in [29, 30, 122] have been inspired by explicit
derivative evaluation, even though one is performing an implicit filtering  process.
Writing, u j and û j in terms of Fourier–Laplace transforms as u j = U (k)eikx j dk

and û j = Û (k)eikx j dk, one can define the corresponding transfer function for the
Û (k)
periodic filtering stencil as T F(k) = U (k)
. For the filter stencil given in Eq. (2.191),
the transfer function for the corresponding periodic problem is obtained as
78 2 DNS of Navier–Stokes Equation
N
a0 + n=1 an cos(nkh)
T F(k) = (2.192)
1 + 2α f cos(kh)

−2 −2
For a (2N )th-order 1D filter, one matches up to the coefficients of u 2N j and û 2N
j
from the Taylor series expansion of Eq. (2.191). This gives N equations for the
(N + 1) coefficients a0 to a N if α f is treated as a parameter. The additional equation
required to determine the values of a0 to a N in terms of the filtering coefficients α f
is found by imposing the condition that T F(k) = 0 at kh = π . The values of the
coefficients a0 to a N in terms of the filter coefficient α f for 2nd- to 10th-order filters
are tabulated in Table IV of [122]. It should be noted that, the value of α f should
range from +0.5 to −0.5, with α f = 0.5 indicates no filtering [83]. Generally positive
values of α f is considered for computation purpose. In Fig. 2.18, T F(k) is shown
plotted as a function of kh for 2nd-, 4th- and 6th-order filter at indicated values of
α f . One finds from Fig. 2.18, that the 2nd-order filter not only affects the high-kh
components, but also attenuates the mid wavenumber components for α f below 0.47.
To implement 4th and higher order filters for non-periodic problems, authors in
[122] proposed the use of one-sided boundary filters, along with the central interior
filters. For example, at a near boundary point j, they have proposed filter stencils of
the following type to retain the tridiagonal form the resultant matrix

11
α f 1 û j+1 + û j + α f 1 û j−1 = an,i u n , iε2, . . . , 5
n=1
10
α f 1 û j+1 + û j + α f 1 û j−1 = a N −n,i u N −n , iε(N − 4), . . . , (N − 1) (2.193)
n=0

For a 6th-order boundary stencil at j = 3, these coefficients are given as [122]: a1,3 =
−1/64 + α f /32, a2,3 = 3/32 + 13α f /16, a3,3 = 49/64 + 15α f /32, a4,3 = 5/16 +
3α f /8, a5,3 = −15/64 + 15α f /32, a6,3 = 3/32 − 3α f /16 and a7,3 = −1/64 + α f /
32. For such non-periodic filtering stencils, one can express the filtering operation as

[A]{û j } = [B]{u j } (2.194)

where the special boundary closure filters given in Eq. (2.193) define the first and last
few rows of [A] and [B] matrices. Expressing, û j and u j by corresponding Fourier–
Laplace transforms, and equating the contributions at the jth-node, one obtains

N N
a jl eik(xl −x j ) Û (k) = b jl eik(xl −x j ) U (k) (2.195)
l=1 l=1

where Û and U represent the Fourier–Laplace transform of û j and u j , respectively.


Thus, a transfer function for the jth-node is defined in the spectral plane as,
2.5 Numerical Filtering: Error Control via Stabilization and Dealiasing 79

(a) αf = 0.450 (b) αf = 0.460


1 1

0.8 0.8
nd
0.6 2 -order 0.6 2nd-order
TF

TF
4th-order 4th-order
0.4 6th-order 0.4 6th-order

0.2 0.2

0 0
0.5 1 1.5 2 2.5 3 0.5 1 1.5 2 2.5 3
kh kh
(c) αf = 0.470 (d) αf = 0.480
1 1

0.8 0.8

0.6 2nd-order 0.6 2nd-order


TF

4th-order TF 4th-order
0.4 6th-order 0.4 6th-order

0.2 0.2

0 0
0.5 1 1.5 2 2.5 3 0.5 1 1.5 2 2.5 3
kh kh
(e) αf = 0.490 (f) αf = 0.495
1 1

0.8 0.8

0.6 2nd-order 0.6 2nd-order


TF

TF

4th-order 4th-order
0.4 6th-order 0.4 6th-order

0.2 0.2

0 0
0.5 1 1.5 2 2.5 3 0.5 1 1.5 2 2.5 3
kh kh

Fig. 2.18 Transfer function (TF) of the periodic 1D filtering stencil plotted as a function of kh for
2nd-, 4 and 6th-order filters for indicated values of α f

N
a jl eik(xl −x j )
T F j (k) = l=1 (2.196)
N
l=1 b jl eik(xl −x j )

Let, u nj = u(x j , t n ) be the solution of 1D convection equation at the jth-node


and nth time step. When a filter is applied after obtaining u n+1 j to obtain the filtered
variable û n+1
j , then the full-step would have an equivalent amplification factor given
as
Ĝ j (kh, Nc ) = T F j (kh)G j (kh, Nc ) (2.197)

In Fig. 2.19a, b, real and imaginary parts of T F j are plotted, respectively for the sixth
order filter with α f = 0.45. The boundary closure stencils are given by Eq. (2.193).
80 2 DNS of Navier–Stokes Equation

Fig. 2.19 a Real and b imaginary parts of transfer function T F j for non-periodic problems with
indicated filter coefficient α for the indicated nodes for the 6th-order interior filter with near-
boundary filters of [122]

In Fig. 2.19a, one notices higher bandwidth of filters for the near-boundary points
and overshoot of T F j for j = 2 and N − 1. The imaginary part of T F j shown in
Fig. 2.19b, indicates numerical instability for j = 2 and 3 and excessive attenuation at
j = N − 1 and N − 2. One also notes that the stability property near the boundaries
is accentuated for higher order filters, with more number of points affected near
the boundaries. At interior points, imaginary part of T F j is identically zero. The
imaginary part of T F j is associated with dispersive effect. To study dispersive effects
of a filter, one needs to look at specific space-time dependent differential equations
like 1D convection equation. Similar studies have been carried out in [83, 90].
2.5 Numerical Filtering: Error Control via Stabilization and Dealiasing 81

Because of the displayed property of filter transfer function in Fig. 2.19, G j


(kh, Nc ) will be attenuated strongly near the Nyquist limit (kh = π ), where by
design, T F j is forced to be equal to zero. For the near-boundary points of a
non-periodic problem, T F j (kh) is a complex number and therefore, it would
alter G j (kh, Nc ) leading to attenuation/amplification and dispersion. Following
Eq. (2.163), the filtered solution of 1D convection equation can be written as

û N = A0 [|Ĝ|]t/Δt eik(x−c N t) dk (2.198)

and corresponding numerical and physical group velocities can be obtained as


 
cN β̂ j
= (2.199)
c j Nc kh
 
Vg N 1 ∂ β̂ j
= (2.200)
c j Nc ∂(kh)
 
−1
where β j = −tan [Ĝ j ]imag /[Ĝ j ]r eal . For periodic problems dispersion proper-
ties are not affected, as T F j is purely a real quantity (See Eq. (2.191)) and therefore,
[Ĝ j ]imag /[Ĝ j ]r eal = [G j ]imag /[G j ]r eal . This is also true for interior points of a non-
periodic filter, irrespective of the choice of α f , the filter coefficient. For these cases,
central filtering does not alter the numerical dispersion properties of the computed
solution and this can be considered a favorable feature.
For non-periodic stencils, as the order of the filter is increased, additional numer-
ical problems are noted at the inflow boundary points, while points near the outflow
boundary display massive dissipation. To circumvent these issues, application of 1D
upwinded filter was proposed in [90]. The general stencil at the interior points for
the proposed upwinded filter is given as [90]

α f û j−1 + û j + α f û j−1 = η f j+4 + (a3 /2 − 5η) f j+3 + (a2 /2 + 10η) f j+2


+(a1 /2 − 10η) f j+1 + (a0 + 5η) f j + (a1 /2 − η) f j−1 + (a2 /2) f j−2
+(a3 /2) f j−3 (2.201)

This is formally a 5th-order upwinded filter whose coefficients are obtained in terms
of filter coefficient α f and the upwinding constant η as: a0 = (22 + 20α f )/32 − 10η,
a1 = (15 + 34α f )/32 + 15η, a2 = (−6 + 12α f )/32 − 6η and a3 = (1 − 2α f )/32
+ η. For η = 0, one recovers the 6th-order central filter, while for η = 0, a complex
transfer function is obtained which would alter the dispersion relation.
In Fig. 2.20, real and imaginary parts of the transfer function are plotted for a 5th-
order upwinded filter [90]. In applying this upwinded filter, central filters have been
used at near-boundary nodes, where Eq. (2.201) is not applicable. The above filter
were used from j = 5 to j = (N − 4); a second order central filter was employed at
82 2 DNS of Navier–Stokes Equation

(a)
1

0.8

Tjr 0.6 α f = 0.45, η = 0.001

J = 1, N
0.4 J = 2, N-1
J = 3, N-2
0.2 J = 4, N-3
J = 5, N-4, N/2

0
0 0.5 1 1.5 2 2.5 3
kh

(b)
0

-0.02

α f = 0.45, η = 0.001
-0.04
Tji

J = 1, N
-0.06 J= 2, N-1
J = 3, N-2
J = 4, N-3
-0.08
J = 5, N-4, N/2

-0.1
0 0.5 1 1.5 2 2.5 3
kh

Fig. 2.20 a Real and b imaginary part of T F j plotted for a 5th-order upwinded filter with indicated
α f and η. The figure is taken from [90]

j = 2 and (N − 1); a fourth order central filter at j = 3 and (N − 2) and sixth order
central filter at j = 4 and (N − 3) were used. Such an upwinded filter completely
eliminates the problem of instability near the inflow and excessive damping near
the outflow of the type shown in Fig. 2.19 for 6th-order central filter. It is readily
seen that there is hardly any difference among T F jr values for different nodes in
Fig. 2.20a, while for the near-boundary points, T F ji is equal to zero in Fig. 2.20b -
due to the use of central filters at these nodes. In Fig. 2.20b for all the other nodes
from j = 5 to (N − 4), the transfer function has the same imaginary value shown
by the solid line. Further details on the use of upwinded filter is provided in [90].
Specifically, the authors have shown that:
1. The upwind filter allows one to add controlled amount of dissipation in the interior
of the domain to remove instability and excessive damping.
2. Absolute control over the imaginary part of the transfer function allows one to
mimic the hyper-viscosity that the SGS models often need for traditional LES.
3. The upwind filters through the presence of the imaginary part of the transfer
function, alters the amplification property differentially for the real and imaginary
part, which in turn changes the numerical group velocity for the upwind filters.
2.5 Numerical Filtering: Error Control via Stabilization and Dealiasing 83

4. Use of upwinded filter either stabilize the computation or allows one to adopt
larger time-step without sacrificing the accuracy of computation for flow prob-
lems, where bypass transition takes place. These types of computations are very
difficult to handle in traditional computations without the use of filters, due to the
sensitive dependence on the numerical dispersion properties at high wavenum-
bers. In this sense, such an approach bears resemblance with process adopted in
DES [4, 117] without requiring the use of different equations in different part of
the domain.

2.5.2 Use of 1D Filters at the Outflow

Here, we show another application of 1D filter to aid in specifying boundary con-


ditions at the outflow, particularly for turbulent flow-field. Consider a situations,
where for wall-bounded turbulent flow-field, massive unsteady separation bubbles
have formed on the wall. To aid these separation bubbles to smoothly pass through
the outflow boundary, one has to resort to special treatment. For incompressible tur-
bulent flows, the outflow boundary conditions are generally specified in terms of
Neumann type boundary conditions or Sommerfeld type convective boundary con-
ditions at the outflow. The Sommerfeld type convective boundary condition on the
primary variable vorticity in derived variable formulation is given as

∂ω ∂ω
+ Uc =0 (2.202)
∂t ∂x
where Uc is the artificial convective velocity. When the outflow boundary is located in
a turbulent zone where massive vortices pass through this boundary, such boundary
conditions are not adequate for these. For such scenarios, one can use 1D filter at the
outflow to allow the vortices to smoothly pass out of the domain, by damping their
amplitude, without affecting the solution much, in the interior of the computational
zone. This sort of application of the 1D filter is equivalent to the buffer-domain
technique used in the literature [49, 60], where the flow Reynolds number is varied
at very thin layer close to the outflow. A specific example is presented here, where
1D filter was used to smoothly convect the massive vortices out of the computational
domain, without affecting the solution at the interior.
In one of the computations of transition of zero pressure gradient boundary layer
(ZPGBL), wall excitation is provided by SBS strip, exciting the flow harmonically
with a non-dimensional frequency of F f = 1.0 × 10−4 and amplitude of excitation
is 1% of the free-stream speed U∞ . The details of the flow transition for such situa-
tions are provided in details in later chapters. Such an excitation generates STWF at
later stages, which induces massive separation bubbles on the wall. The domain of
computation is taken to be from −0.05 ≤ x ≤ 120 and 0 ≤ y ≤ 1.5 (in nondimen-
sional units) in the streamwise and wall-normal directions, respectively. First, the
computations are performed without applying the 1D filter at the outflow boundary.
84 2 DNS of Navier–Stokes Equation

(a) -contour, t = 615 (f) -contour, t = 615


0.3 0.3

0.2 0.2
y

y
0.1 0.1

095 095
100 105 x 110 115 120 100 105 x 110 115 120

(b) -contour, t = 618 (g) -contour, t = 618


0.3 0.3

0.2 0.2
y

y
0.1 0.1

095 095
100 105 x 110 115 120 100 105 x 110 115 120

(c) -contour, t = 620 (h) -contour, t = 620


0.3 0.3

0.2 0.2
y

y
0.1 0.1

095 095
100 105 x 110 115 120 100 105 x 110 115 120

(d) -contour, t = 622 (i) -contour, t = 622


0.3 0.3

0.2 0.2
y

y
0.1 0.1

095 095
100 105 x 110 115 120 100 105 x 110 115 120

(e) -contour, t = 623 Breakdown takes place (j) -contour, t = 623 Breakdown takes place
0.3 0.3

0.2 0.2
y

0.1 0.1

095 095
100 105 x 110 115 120 100 105 x 110 115 120

Fig. 2.21 Stream-function ψ- (frames a–e) and vorticity ω-contours (frames f–j) shown at indicated
times for 95 ≤ x ≤ 120 before applying 1D 2nd-order filter at the outflow

In Fig. 2.21, the ψ- (frames (a)–(e)) and ω-contours (frames (f)–(j)) are plotted for
this case at t = 615, 618, 620, 622 and 623 before the application of the 1D, 2nd-
order, filter at the outflow, as described next. One-dimensional 2nd-order filter at the
outflow is used only for the last 10 streamwise grid-points, whenever its application
becomes necessary. The filter coefficient in this zone is smoothly varied from 0.5 at
the onset of the filtered zone to 0.495 at the outflow boundary as
 
xi − xst 2
α f (i) = 0.5 − 0.05 (2.203)
xout − xst

where xst is the x-location of the beginning of the filtered zone and i is the streamwise
index.
Only a part of the computational domain (95 ≤ x ≤ 120) near the outflow bound-
ary at x = 120 is shown in Fig. 2.21. One notices several separation bubbles on the
wall in Fig. 2.21. The tallest of these separation bubbles have a height of y = 0.2
which translates into roughly 3.5 times the local displacement thickness δ ∗ at its loca-
tion. Hence, these bubbles pierce through the boundary layer and is responsible for
the generation of turbulence. The leading separation bubble, though is quite smaller
than the tallest one, convects downstream and hits the outflow boundary at t = 620 as
shown in Fig. 2.21d and (i). A separation indicate a massive concentration of vortical
2.5 Numerical Filtering: Error Control via Stabilization and Dealiasing 85

(a) -contour, t = 615 (f) -contour, t = 615


0.3 0.3

0.2 0.2
y

y
0.1 0.1

095 095
100 105 x 110 115 120 100 105 x 110 115 120

(b) -contour, t = 618 (g) -contour, t = 618


0.3 0.3

0.2 0.2
y

y
0.1 0.1

095 095
100 105 x 110 115 120 100 105 x 110 115 120

(c) -contour, t = 620 (h) -contour, t = 620


0.3 0.3

0.2 0.2
y

y
0.1 0.1

095 095
100 105 x 110 115 120 100 105 x 110 115 120

(d) -contour, t = 622 (i) -contour, t = 622


0.3 0.3

0.2 0.2
y

y
0.1 0.1

095 095
100 105 x 110 115 120 100 105 x 110 115 120

(e) -contour, t = 623 Breakdown does not take place (j) -contour, t = 623 Breakdown does not take place
0.3 0.3

0.2 0.2
y

0.1 0.1

095 095
100 105 x 110 115 120 100 105 x 110 115 120

Fig. 2.22 Stream-function ψ- (frames a–e) and vorticity ω-contours (frames f–j) shown at indicated
times for 95 ≤ x ≤ 120 after applying 1D 2nd-order filter at the outflow from t = 15

structures and the radiative outflow boundary condition prescribed in Eq. (2.202) is
simply not sufficient enough to handle such large vortices and convect these out of
the outflow boundary, even if the artificial convective velocity Uc is equal to the free-
stream speed, U∞ . The inability of the radiative boundary condition (Eq. (2.202)) to
convect the vortices smoothly out of the domain causes accumulation of vorticity in a
narrow strip near the outflow boundary, and this causes the simulation to breakdown
just after t = 623. The ω-contours shown in Fig. 2.21j at t = 623 shows substantial
accumulation of vorticity at the outflow boundary, which manifests itself as a mas-
sive separation bubble in the respective ψ-contour shown in Fig. 2.21e at the outflow.
Now the 1D, 2nd-order filter is applied at the outflow from t = 615. The unfiltered
solution at t = 615 is prescribed as the initial solution for this modified approach.
The results are shown in Fig. 2.22. The corresponding ψ- (frames (a) to (e)) and ω-
contours (frames (f) to (j)) obtained after applying the 1D, 2nd-order filter at the last
ten streamwise grid-points near outflow are shown in Fig. 2.22 at t = 615, 618, 620,
622 and 623. One notes that by doing so, the previous problem of accumulation of
vorticity at the outflow is solved. Now, the radiative boundary condition (Eq. (2.202))
handles this attenuated level of vorticity at the outflow properly and the simulation
does not breakdown, as it happened in the previous case shown in Fig. 2.21. One
also observes that, the application of the 1D filter does not alter the flow in any way
for x ≤ 115, as all the minute details shown in Figs. 2.21 and 2.22 are identical for
86 2 DNS of Navier–Stokes Equation

both ψ and ω for x ≤ 115. After this application of the filter at the outflow, it was
possible to continue this simulation even for much longer time up to t = 655 till the
time the solution breaks down at an interior point due to aliasing. This problem of
aliasing is cured by the application of 2D adaptive filter first proposed in Bhumkar
and Sengupta [11], which is described next.

2.5.3 Two-Dimensional Higher Order Filters

Despite displaying several advantages, 1D filters also show strong directionality in


the computed solution, for 2D fluid flow problems. In [90], it was shown that filter-
ing in azimuthal direction for flows past an airfoil or a circular cylinder performing
rotational oscillations, causes unphysical smearing of vortical structures in that direc-
tion. Researchers have identified aliasing error as a potent source of error occurring
during discrete computations. The origin of the aliasing problem, while solving the
NSE, occurs whenever one encounters a term which is a product of more than one
quantities. Two such terms in NSE are due to nonlinear convection terms and the
linear self-adjoint diffusion terms, that appear while solving the NSE in transformed
plane. In [89], a new 2D filter is proposed, which is capable of removing the aliasing
problem more effectively than 1D filters, without introducing any directionality to
the solution. These 2D filters are significantly different from the conventional 1D
Padè type filters described in Sect. 2.5.1.
First, we describe the origin of the aliasing error and why it occurs in any dis-
crete computations. As stated, aliasing errors occur whenever the product terms are
evaluated numerically. Consider the product of two terms in the physical space as

w(x j ) = u(x j ) v(x j ) (2.204)

Now, each of these individual terms can be represented by respective bilateral


Fourier–Laplace transform as

w(x j ) = W (k)eikx j dk

u(x j ) = U (k1 )eik1 x j dk1

v(x j ) = V (k2 )eik2 x j dk2 (2.205)

From Eqs. (2.204) and (2.205), one notes that


km km
w(x j ) = U (k1 )V (k2 )ei(k1 +k2 )x j dk1 dk2 (2.206)
−km −km
2.5 Numerical Filtering: Error Control via Stabilization and Dealiasing 87

Fig. 2.23 Spectral domain


of interest for finite grid
computations. The shaded
areas (E B F and H DG)
contribute to aliasing in
evaluating product terms to
zones H OG and E O F

For any discrete computation, the maximum resolvable wavenumber km is given by


the Nyquist limit, where km = π/ h, with h representing the uniform grid-spacing.
Therefore, the limit of integration for k1 and k2 in Eq. (2.206) varies from −km to
km for both k1 and k2 . This zone is shown in Fig. 2.23 by the solid box (ABC D).
One notes from Eq. (2.206) that wavenumber component of w(x j ) is represented
by k = k1 + k2 and as both k1 and k2 varies from −km to km , k varies from 2km to
−2km . For example, in the triangular zone E B F in the top right corner of Fig. 2.23,
km ≤ k ≤ 2km . Similarly, inside the triangular zone H DG (in the bottom left corner
of Fig. 2.23), −2km ≤ k ≤ −km . However, for any discrete computation, one can
resolve scales only up to the Nyquist limit, i.e., −km ≤ k ≤ km . Therefore, the region
where (k1 + k2 ) ≥ |km |, should fold back inside the Nyquist limit, while maintain
the phase relation in either of this zones. Defining a modified wavenumber k̄ =
(k1 + k2 ) − 2km = k − 2km , one notes that
eikx j = ei k̄x j
Moreover, when km ≤ k ≤ 2km , −km ≤ k̄ ≤ 0. This implies that such a transfor-
mation in wavenumber retains the phase relationship, while maintaining the mod-
ified wavenumber k̄ to lie within the Nyquist limit. In essence, the region E B F
would be aliased to the region H OG. Similar arguments can also be provides for
the zone H DG, which would be aliased to the region E O F. Mapping the integral
in Eq. (2.206) from (k1 , k2 )-plane to (k, k2 )-plane, one obtains
0 k+km
w(x j ) = U (k − k2 )V (k2 )eikx j dk2 dk
−km −km
km km
+ U (k − k2 )V (k2 )eikx j dk2 dk
0 k−km
88 2 DNS of Navier–Stokes Equation

−km k+km
+ U (k − k2 )V (k2 )eikx j dk2 dk
−2km −km
2km km
+ U (k − k2 )V (k2 )eikx j dk2 dk (2.207)
km k−km

Derivation of Eq. (2.207) is left as an exercise to the reader. One notes from Eq.
(2.207) that the first and the last term falls within the Nyquist limit for wavenumber
component k, whereas the last two integrals lie outside of it. Following the transfor-
mations k̄ = k + 2km and k̄ = k − 2km in the third and fourth integrals, respectively,
one rewrites Eq. (2.207) as
0 k+km
w(x j ) = U (k − k2 )V (k2 )eikx j dk2 dk
−km −km
km km
+ U (k − k2 )V (k2 )eikx j dk2 dk
0 k−km

km k−km
+ U (k − k2 − 2km )V (k2 )eikx j dk2 dk
0 −km
0 km
+ U (k − k2 + 2km )V (k2 )eikx j dk2 dk (2.208)
−km k+km

In Eq. (2.208), k̄ has been replaced by k for uniformity. One clearly notes the effect
of aliasing in the last two integrals of Eq. (2.208). Expressing w(x j ) by its Fourier–
Laplace transform within the Nyquist limit and equating it with the right hand side
of Eq. (2.208), the spectral amplitude of it is given as
k+km km
W (k) = U (k − k2 )V (k2 )dk2 +
−km k−km
U (k − k2 )V (k2 )dk2 + Walias (k) (2.209)

where the aliased component Walias (k) is given as


k−km
Walias (k) = U (k − k2 − 2km )V (k2 )dk2
−km
km
+ U (k − k2 + 2km )V (k2 )dk2 (2.210)
k+km

Equation (2.210) shows that aliasing causes the shifting of the resultant spectral
amplitude to a different wavenumber than the combined wavenumber of constituent
terms. This problem also arises while solving NSE in uniform grid because of the
presence of the nonlinear convection term. A more comprehensive discussion on the
aliasing error and its effects are given in [83] and are not repeated here for brevity.
2.5 Numerical Filtering: Error Control via Stabilization and Dealiasing 89

Two-dimensional filter, only removes the high wavenumber components that are
responsible for aliasing. General form of the proposed 2D filters in [89] is given by,

û i, j + α2 f (û i−1, j + û i+1, j + û i, j−1 + û i, j+1 )



an
= (u i±n, j + u i, j±n ) (2.211)
n=0
2

In Eq. (2.211), û i, j is the filtered variable, while u i j is the unfiltered counterpart,


which is being filtered here. The variable M̂ in Eq. (2.211) represents the order of
the 2D-filter, which is equal to one for a 2D, 2nd-order filter and every increase
of M̂ by one increases the order of the filter by two. Let, Ni and N j be the total
number of grid-points along i- and j-directions, respectively. Then the indices i
and j varies from i = 2 to (Ni − 1) and j = 2 to (N j − 1), for the 2nd-order, 2D-
filter, respectively. Similarly, for 4th-order, 2D-filter, Eq. (2.211) is applicable from
i = 3 to (Ni − 2) and j = 3 to (N j − 2). Similarly, higher order filters can also be
obtained. Equation (2.211) implies that application of 2D-filter amounts to solving a
linear algebraic equation with the left-hand side giving rise to a pentadiagonal matrix.
Any iterative method employed to solve this linear algebraic equation, requires the
diagonal dominance of the resultant matrix. This can be ensured by prescribing
|α2 f | ≤ 1/4 [89].
Expressing the filtered (û) and unfiltered (u) variables by respective bi-directional
Fourier–Laplace transform as

u j,l = U (k x , k y )ei(kx x j +k y yl ) dk x dk y

û j,l = Û (k x , k y )ei(kx x j +k y yl ) dk x dk y

one obtains the transfer function of the 2D-filter given by Eq. (2.211) as
 M̂
a0 + n=1 an [cos(nk x h x ) + cos(nk y h y )]
T F(k x h x , k y h y ) = (2.212)
1 + 2α2 f [cos(k x h x ) + cos(k y h y )]

where k x and k y are the wavenumbers and h x and h y are grid-spacing along i and
j-directions, respectively. The choice of α2 f = 1/4 indicates no filtering of the vari-
able. The coefficients ak of the 2D-filter can be obtained as a function of the filter
coefficient α2 f by expressing the right-hand and left-hand sides of Eq. (2.211) by
Taylor series expansion and equating coefficients of appropriate terms on either sides
of Eq. (2.211). For example, for 2nd-order filter, one has to match the coefficients of
unfiltered and filtered quantities at the (i, j)th node, for consistency of the variable
(consistency condition), which gives the following condition
a0 + 2a1 = 1 + 4α2 f (2.213)
90 2 DNS of Navier–Stokes Equation

for obtaining a0 and a1 for 2nd-order, 2D-filter, the above equation is supplemented
by an additional condition which requires that the transfer function should be zero
when k x h x = k y h y = π , i.e, at point B in Fig. 2.23. This condition is imposed,
so that the 2D-filter can effectively control the aliasing error without distorting or
unphysically attenuating the physically relevant components. This condition imposes
the relation
a0 = 2a1 (2.214)

Following Eqs. (2.213) and (2.214), one obtains

1
a0 = + 2α2 f
2
1
a1 = + α2 f (2.215)
4
For 4th-order, 2D-filter, the consistency condition is given as

a0 + 2a1 + 2a2 = 1 + 4α2 f (2.216)

For this filter, one also has to equate the coefficients of 2nd-order partial derivatives
(including the mixed derivatives) at (i, j)th-node, which gives the following relation

a1 + 4a2 = 2α2 f (2.217)

Equations (2.216) and (2.217) are to be augmented by vanishing conditions of the


transfer function at point B (see Fig. 2.23), which gives

a0 − 2a1 + 2a2 = 0 (2.218)

Equations (2.216)–(2.218) gives the coefficients (a0 , a1 , a2 ) for the 4th-order 2D-
filter as
(5 + 12α2 f ) (1 + 4α2 f ) (−1 + 4α2 f )
a0 = , a1 = , a2 = (2.219)
8 4 16
Similar conditions can also be derived for the 6th-order 2D-filter and the correspond-
ing coefficients are obtained as [89],

(11 + 20α2 f ) (15 + 68α2 f )


a0 = , a1 =
16 64
(−3 + 12α2 f ) (1 − 4α2 f )
a2 = , a3 = (2.220)
32 64
In Fig. 2.24, the transfer function of a 2nd-order, 2D filter is shown, for indicated
values of α2 f as a function of nondimensional wavenumbers k x h x and k y h y . As
noted earlier, the region corresponding to (k x h x + k y h y ) > π contributes to aliasing
2.5 Numerical Filtering: Error Control via Stabilization and Dealiasing 91

(a) α2f = 0.20 (b) α2f = 0.15


3 3 0.05
0.2
0.4 0.2
0.6 0.4
2 2 0.6
0.8
kyhy

kyhy
0.9 0.8
0.95 0.9
1 1
0.95
0.99 0.99
0.999 0.999
0 0
0 1 2 3 0 1 2 3
kxhx kxhx
(c) α2f = 0.10 (d) α2f = 0.05
3 3
0.05 0.05
0.2 0.2

2
0.4 2 0.4
kyhy

kyhy

0.6 0.6
0.8
1 1 0.8
0.9 0.9
0.95
0.95
0.99
0.99
0 0.999 0
0 1 2 3 0 1 2 3
kxhx kxhx

Fig. 2.24 The transfer functions of the 2nd-order 2D filter are shown in frames a and b for indicated
α2 f

error. It is clear from Fig. 2.24, that the transfer function of the 2D-filter effectively
removes the aliasing error without unduly attenuating the mid-wavenumber region
in each direction [89]. It has also been shown in Bhumkar and Sengupta [11] that
such 2D filters can be used in an adaptive manner on a small patch of the grid to
only affect the zone, where some high wavenumber fluctuations of the solution have
been developed. Such adaptive filters have been used while performing receptivity
calculations in [6, 85, 86] for effective de-aliasing. It is impossible to simulate
transitional or turbulent flows correctly without effective de-aliasing of the solution.
In the next subsection, a brief description of the application procedure of such filters
are provided.

2.5.4 Adaptive 2D Filter

Adaptive 2D filters have been first proposed in [11], which can be effectively used
over small localized patches of computational domain to avoid local numerical insta-
92 2 DNS of Navier–Stokes Equation

Fig. 2.25 Schematic of the problem solved for free stream excitation by an infinite array of con-
vecting vortices over a semi-infinite flat plate

bilities occurring due to aliasing error during physical secondary and higher order
instabilities. For this purpose, the filter operation given by Eq. (2.212) is used in adap-
tive manner at some selective points, based on a predetermined criterion. The resultant
linear algebraic equations can be solved iteratively using Bi-CGSTAB method [118].
Here, one particular transitional and turbulent flow problem is described, where
application of adaptive 2D-filter is essential to carry out the numerical calculations
accurately. The problem is regarding the receptivity of the boundary layer to the
free-stream convecting vortices. The schematic diagram of the problem is shown in
Fig. 2.25. Here, an infinite row of convecting vortices migrate at a fixed height, with a
fixed speed. The imposition of infinite overhead vortex system requires an associated
image vortex system of opposite sign (see Fig. 2.25), so that the wall-normal velocity
component at the wall is zero. The separation between the successive vortices (a),
the strength of each vortex Γ , convection speed c, and its height over the flat plate H ,
are the parameters of the problem. Here a case is described, where c = 0.2, a = 10,
H = 2.0 and Γ = 0.25. To compute the receptivity of the shear-layer to the free-
stream vortices, 2D incompressible NSE given by Eqs. (2.31) and (2.32) are solved
using OUCS3 scheme for spatial discretization scheme for convective terms and
optimized 3-stage Runge–Kutta scheme ORK3 [100] is used for time-integration.
The overhead convecting vortices causes vortical eruptions inside the boundary
layer, which are directly responsible for transition to turbulence. While calculating
this case, in a computational domain with 4500 × 400 points spanning from −0.05
to 90 in the streamwise direction and 1.5 in the wall-normal direction, the problem
of aliasing was faced. This resulted in the complete breakdown of the simulation
at t = 113. In Fig. 2.26, the vorticity contours are plotted at t = 111, 112 and 113
in frames (a), (b) and (c), respectively. One notices the unsteady vortical eruptions
2.5 Numerical Filtering: Error Control via Stabilization and Dealiasing 93

Fig. 2.26 Vorticity contours of the unfiltered case plotted at indicated times for the free stream
convecting vortex problem with c = 0.2, a = 10, H = 2.0 at indicated times in frames a, b and c.
d ∂∂ xω2w plotted of the filtered case at t = 113
2

and associated induction of high wavenumber components nearby, due to aliasing.


In Fig. 2.26d, the quantity ∂ 2 ωw /∂ x 2 is plotted as a function of x. One observes high
wavenumber oscillations at some selective locations underneath the overhead vortex.
The point where the actual breakdown of the solution has taken place at t = 113 is
also marked in Fig. 2.26c, d.
To rectify these issues of aliasing error causing breakdown of the numerical solu-
tion, 2D adaptive filter is used very close to the wall. One identifies a rectangular/
square sub-domain in the transformed (ξ, η)-plane where filtering is performed, with
variation of α2 f with respect to the indices i and j is given as
  
(l1 − l0 ) i max − 2i + i min
α2 f = l 1 − 1 + cos π
2 i max − i min
 
π j − jmin
cos (2.221)
2 jmax − jmin

where, l1 = 0.25, l2 = 0.05 and i max ; i min ; jmax ; jmin are the upper and lower limits of
i and j indices, respectively i.e., i min ≤ i ≤ i max and jmin ≤ j ≤ jmax . In Fig. 2.27a
a typical variation of α2 f is plotted over a 9 × 9 filter-tent, where i max = 9, i min = 1,
jmax = 9 and jmin = 1. The corresponding 3D perspective plot of the variation of
α2 f is shown in Fig. 2.27b. Maximum filtering is performed at (i = (i max + i min )/2,
94 2 DNS of Navier–Stokes Equation

Fig. 2.27 a A typical


(a)
variation of α2 f is shown
j+4
over a 9 × 9 filter tent in the 0.250 0.225 0.250

transformed plane. The j+3 0.200


corresponding 3D j+2 0.175

perspective plot of the 0.150


variation of α2 f in the filter j+1
0.125
tent is shown in frame b j
0.100

j-1
0.075
j-2

j-3

j-4
i-4 i-3 i-2 i-1 i i+1 i+2 i+3 i+4

(b)
i j

0.2

0.1
α2f

2
2
4 4
6 6
8 8

j = jmin )th point of the tent, which progressively reduces to no filtering at the edges
of it, where α2 f = 0.25.
In applying this filter, the streamwise second derivative of the wall-vorticity
(∂ 2 ωw /∂ x 2 , where ωw is the vorticity at the wall) is calculated and if this value
exceed a certain predefined value, then the 2D adaptive filter is activated at that par-
ticular x-location. There are several reasons to chose such a criterion. First, it was
noted that breakdown of simulations happened due to the occurrence of the high-
wavenumber fluctuations in the vorticity and for wall-bounded boundary layers,
maximum vorticity generally occurs on the wall itself. Secondly, as the differen-
tiation operation magnifies the high wavenumber amplitudes, the chosen criterion
would exactly highlight the points on the wall, where high wavenumber oscillations
appear in the vorticity. These are the reasons behind choosing the quantity ∂ 2 ωw /∂ x 2
as a test-parameter. The specification of the cutoff value for the test-parameter was
performed, based on experience and there is no definitive rule to prescribe it.
The same case was run with applying 2D adaptive filter at the selected locations as
described in the last paragraph, whenever ∂ 2 ωw /∂ x 2 exceeded 100. This case was run
from results at t = 100, of earlier unfiltered simulation. The results of the filtered case
is shown in Fig. 2.28. One notices that the application of adaptive filter has been able to
2.5 Numerical Filtering: Error Control via Stabilization and Dealiasing 95

Fig. 2.28 Vorticity contours plotted at indicated times for the free stream convecting vortex problem
with c = 0.2, a = 10, H = 2.0 times in frames a, b and c after applying adaptive 2D filter from
t = 100. ∂∂ xω2w plotted of the filtered case at t = 113 in frame d
2

significantly control the high wavenumber oscillations, while completely preserving


flow features at lower wavenumber. Fig. 2.28d reveals that the amplitude of the high
frequency oscillations are also substantially reduced due to the application of this
filter. To further reveal the effectiveness of the 2D adaptive filter, the wall vorticity and
its FFT at t = 113 is plotted in Fig. 2.29 for both unfiltered (frames (a) and (b)) and
filtered case (frames (c) and (d)). One finds in Fig. 2.29b, that the aliasing effectively
increasing the high wavenumber components as indicated. Fig. 2.29d also testifies
to the fact that 2D adaptive filtering not only suppresses the amplitude of the high
wavenumber components by an order of magnitude, but also keep the low wavenum-
ber perturbation amplitudes intact, which only determines the nature of transition.
Similar application of 2D adaptive filter to control aliasing errors during the compu-
tations of flows past circular cylinder performing rotary oscillation and flows past a
natural laminar aerofoil at a very high Reynolds number have been reported in [11].

2.6 Role of Solenoidality Errors in Velocity-Vorticity


Formulations

One of the major issues of high accuracy finite difference computing of NSE
using different formulations is the identification of an error metric and different
investigators have tried to identify different derived quantities as a marker for the
96 2 DNS of Navier–Stokes Equation

(a) Unfiltered case, t = 113

ωw 200

-200

-400

20 40 60 80
x
(b) Unfiltered case, t = 113
Due to aliasing error
104
Ωw (k)

3
10
2
10
1
10
100 k 101 102

(c) Filtered case, t = 113


200
ωw

-200

-400

20 40 60 80
x
(d) Filtered case, t = 113
4
10 No aliasing
Ωw (k)

3
10
2
10
1
10
100 k 101 102

Fig. 2.29 a Wall vorticity at t = 113 of the unfiltered free stream convecting vortex problem
shown in Fig. 2.27 and b its Fourier transform plotted as a function of streamwise distance x and
wavenumber k, respectively. c Wall vorticity at t = 13 of the case where 2D adaptive filtered is
used for the free stream convecting vortex problem shown in Fig. 2.28 and d its Fourier transform
plotted as a function of streamwise distance x and wavenumber k, respectively
2.6 Role of Solenoidality Errors in Velocity-Vorticity Formulations 97

accuracy of the numerical solution. Here, the solenoidality of velocity and vorticity
as the error metrics are focused upon, and the relevance of these for some benchmark
2D and 3D flow problems are presented. In order to develop an accurate method using

→ →
(V ,− ω )-formulations for simulating transitional flows, the effects of solenoidality
of velocity on flow evolution has to be assessed and for this purpose, one must focus
on the flow regimes, when instabilities play a dominant role for both internal and
external flows. For 3D flows, solenoidality of vorticity vector also has to be consid-
ered. This has prompted the development of the conservative rotational formulation
of VTE, which has been given previously in Eq. (2.24a). Here, application of this
formulation is described along with the development of new optimum versions of
derivative evaluation and interpolation of unknowns.

2.6.1 Computation of Physically Unstable Flows in Square


Lid Driven Cavity

Vorticity is generated at no-slip walls or due to flow instabilities in free shear layers,
for incompressible flows. For turbulent flows, vorticity and other derived variables
play a central role in understanding some of the physical mechanisms of creating
turbulence. Hence, VTE is found very relevant for the analysis and solution of viscous
incompressible flows. Attendant velocity field can be obtained from the solution of the
Poisson equation for stream-function (for 2D case) or by solving Poisson equations
for the velocity components, as described previously in Sects. 2.1.5 and 2.1.6.
As an illustrative example, first the flow inside a 2D square lid-driven cavity
(LDC) is considered. For this problem, each side of the square and the lid-velocity is
used as the reference length and velocity scale to nondimensionalize the governing
equations. This determines the Reynolds number for the flow. A staggered vari-
able arrangement, as used in Guj and Stella [37] and Napolitano and Pascazio [63]
has been adopted. The staggered arrangement of variables are shown in Fig. 2.10a,
where the velocity components are defined at the center of each sides and the vor-
ticity components are defined at each integral grid-nodes. It has been shown that
numerical error is smaller in staggered grid than non-staggered (or collocated) grid


using ( V , ω)-formulation [44]. It is known that unlike (ψ, ω)-formulation, velocity-
vorticity formulation does not identically ensure satisfaction of solenoidality of the
velocity field everywhere inside the flow domain. The governing equations are 2D
incompressible NSE in physical x- and y-domain with uniform distribution of grid-
points. Therefore, h 1 = h 2 = 1 in Eqs. (2.36)–(2.39) and these are given as
 
∂ω ∂ ∂ 1 ∂ 2ω ∂ 2ω
+ (uω) + (vω) = + 2 (2.222)
∂t ∂x ∂y Re ∂ x 2 ∂y
98 2 DNS of Navier–Stokes Equation

∂ 2u ∂ 2u ∂ω
+ =− (2.223)
∂x2 ∂ y2 ∂y

∂ 2v ∂ 2v ∂ω
+ = (2.224)
∂x2 ∂ y2 ∂x


where the velocity V [= (u, v)] is expected to satisfy the divergence-free condition
given as
∂u ∂v
Dv = + =0 (2.225)
∂x ∂y

As the purpose is to ascertain the role of non-solenoidality of velocity field in velocity-


vorticity formulation, an alternate equation derived by differentiating Eq. (2.225)
with respect to y is used, which is used to obtain one of velocity components. This
ensures solenoidality of velocity identically everywhere inside the flow domain. In
the present case, Eq. (2.224) is replaced by the equation given by

∂ 2v ∂ 2u
= − (2.226)
∂ y2 ∂ x∂ y

In evaluating the role of solenoidality of velocity field, the LDC problem is solved
employing two schemes as described next.
Scheme-A: In this approach, Eq. (2.222)–(2.224) are solved. Here first Eq. (2.222)
is solved for time advancing the vorticity field, followed by Eqs. (2.223) and (2.224)
for the velocity components. Thus, the solution obtained by this scheme will not
identically satisfy solenoidality of the velocity field.
Scheme-B: Here, the vorticity field is time advanced by solving Eq. (2.222) first.
The velocity components are obtained by solving Eqs. (2.223) and (2.226). Solving
Eq. (2.226) instead of Eq. (2.224), ensures solenoidality of the velocity field by this
scheme.
In both the schemes, time advancement of vorticity is carried out using 4th-order,
four-stage Runge–Kutta (R K 4 ) scheme, while convective derivatives, i. e., ∂∂x (uω)
and ∂∂y (vω) terms are discretized using OUCS3 scheme, and second derivative terms
are discretized using second-order central difference schemes. To evaluate convective
derivatives in the staggered grid, it is required to interpolate u- and v-components at
the location of the vorticities; which is carried out by the optimized compact mid-
point interpolation scheme of [8]. This is described in Sect. 2.3.2. To obtain u- and
v-velocity components, either scheme A or scheme B is employed by using central
difference explicit schemes for all pure and mixed second derivatives.
Results are presented here for the 2D LDC problem, for Re = 9000 and Re =
10, 000. Both these Reynolds numbers belong to post-critical range for the 2D
flow [105]. Results obtained in [105] using (ψ, ω)-formulation are accurate for this
post-critical Reynolds number and is used here for validation and comparison pur-
poses. The Poisson equations are solved using Bi-CGSTAB [118] iterative scheme.
2.6 Role of Solenoidality Errors in Velocity-Vorticity Formulations 99

Fig. 2.30 Flow in a lid-driven cavity for Re = 9000. a Maximum absolute divergence error plotted
as a function of time for schemes A and B. b, c Vorticity time-history at x = 0.95, y = 0.95 plotted
for schemes A and B, respectively. d, e Vorticity contours at indicated times shown for schemes A
and B, respectively
100 2 DNS of Navier–Stokes Equation

A pre-determined tolerance value for the solution error (ε) is used in the Bi-
CGSTAB solver. Maximum solenoidality error occurring inside the domain for the
scheme A, is directly dependent on the level of ε chosen. In computing both cases, a
grid with (257 × 257) uniformly distributed points have been used. After every 100
time-steps, Dv is computed using 2nd-order central difference scheme and maximum
absolute value of Dv is stored. In Fig. 2.30a, time history of Dv are compared for
Re = 9000 for schemes- A and -B, for ε = 10−6 used with the Bi-CGSTAB solver.
In Fig. 2.30b, c, vorticity value at a point with co-ordinates x = 0.95 and y = 0.95
(with the origin at the lower left corner) is plotted as a function of time. Even though
the levels of Dv shown in Fig. 2.30a, are orders of magnitude different, this does not
affect either the onset time of instability or the amplitude of vorticity fluctuations
in Fig. 2.30b, c. As noted in [105], this flow for Re = 9000 also shows nonlinear
saturation of disturbance amplitude. Despite the difference in Dv , these two schemes
show transient variation that also looks almost similar in Fig. 2.30b, c. This seems to


counter the observation of [17, 128] for ( p, V )-formulation; who have talked about
numerical instabilities due to non-satisfaction of solenoidality for velocity vector
during transient and later stages of flow evolution computed by NSE. Recently it has


also been pointed out in [125] that using ( p, V )-formulation with pseudo-spectral
DNS slowly evolving spurious numerical turbulence is induced inside the flow-field,
when the effect of machine round-off on the divergence-free condition is not carefully


controlled. In contrast, for the ( V , ω)-formulation, this has not seen to be the case,
as shown in Fig. 2.30.
In Fig. 2.30d, e, vorticity contours at indicated times plotted for these two schemes
are also indistinguishable. Both schemes show the presence of a triangular vortex
at the center of the cavity, at t = 700, as reported earlier for 2D supercritical LDC
flow in [98, 105, 107]. But one notices a small time lag among the satellite vortices,
as shown in the third frames of Fig. 2.30d, e. This is due to different phase and
dispersion errors for the two schemes.
Similar results are shown in Fig. 2.31a, b for the LDC flow for Re = 10, 000. In
Fig. 2.31a, time history of Dv has been plotted for schemes A and B. For scheme-A,
different tolerance limit given by ε for the solution of Poisson equation are used. It
is noted that tightening the tolerance limit by orders of magnitude, Dv decreases by
similar orders, yet scheme-B shows negligibly small levels of Dv in Fig. 2.31a(ii).
In Fig. 2.31b, time histories of disturbance vorticity are shown for different val-
ues of ε. For scheme-A three different levels of tolerance, ε = 10−4 , ε = 10−6 and
ε = 10−9 are considered with Bi-CGSTAB solver. In using scheme-B, a tolerance
value of ε = 10−6 has been used in solving the Poisson equation for u-component
of velocity. Note that the solution of Eq. (2.226) is obtained directly via exact solu-
tion of tridiagonal matrix system. From the displayed figures, one again notices that
maximum solenoidality error is markedly different for all the four cases, even though
the vorticity time history at x = 0.95 and y = 0.95 are almost similar.
2.6 Role of Solenoidality Errors in Velocity-Vorticity Formulations 101

Fig. 2.31 Flow in a lid-driven cavity for Re = 10000. a Maximum absolute divergence error time-
history plotted for the indicated schemes with different ε. b Vorticity time-history at x = 0.95 and
y = 0.95 shown plotted for (i) scheme A with ε = 10−4 , (ii) scheme A with ε = 10−6 , (iii) scheme
A with ε = 10−9 and (iv) scheme B with ε = 10−6

Presented results in Figs. 2.30 and 2.31, using velocity-vorticity formulations


(which does not satisfy solenoidality of velocity vector to machine zero) show iden-
tical results for a long time without showing any tendency for numerical instability
for a flow, which demonstrates physical instability and its nonlinear saturation. This
shows that the velocity-vorticity formulation is insensitive to divergence error beyond
a certain level. It is because of the following reasons.
102 2 DNS of Navier–Stokes Equation

In velocity-vorticity formulation of scheme- A, these two quantities are related by


the vector Poisson equation as,


∇ 2 V = −∇ × −

ω (2.227)

Irrespective of the fact whether Dv is zero or not, this equation uses solenoidality
in its derivation. As a consequence, taking divergence of Eq. (2.227) and using the

→ −

vector identity ∇ · (∇ × G ) = 0 for any arbitrary vector G one obtains

∇ 2 Dv = 0 (2.228)

Thus, if one chooses Dv as the error metric in velocity-vorticity formulation, it


becomes time-independent and its property is given by the minmax theorem of the
governing Laplace equation. If any time dependence or instability has to occur in
this formulation, then it must come from the kinematic boundary condition. In the
absence of any such time-dependent boundary forcing, the governing equation for
this metric is perfectly stable, as noted in Figs. 2.30 and 2.31.

2.6.2 Solution of 3D Cubic Lid Driven Cavity Using


Velocity-Vorticity Formulation and Role
of Solenoidality Error for Vorticity


In solving 2D flows using (ψ, ω)- or ( V , ω)-formulations, one notes that the non-
zero component of the vorticity lies in a plane perpendicular to the flow. Thus, the
vorticity in 2D flows always satisfies solenoidality condition (Dω = ∇ · − →
ω = 0).
However for 3D flows this is not automatic, i.e., Dω = 0. Thus, one would like
to investigate the role of Dω for 3D flows. This is described in Sect. 2.1.5. Here,

→ →
the role of Dω in ( V , − ω )-formulations for 3D flows are estimated, by showing
results for flows inside 3D cubic lid driven cavity. In Sect. 2.1.5, three different
formulations of VTE is provided, namely, (a) the non-conservative form, Eq. (2.22),
(b) the Laplacian form, Eq. (2.23), and (c) the rotational form, Eq. (2.24b), and also
the corresponding evolution equation for divergence of vorticity (Dω ) is illustrated
through Eqs. (2.25a)–(2.25c). It was noted that for the rotational form of VTE, one
only requires Dω to be zero, while prescribing the initial conditions. In contrast, for
the Laplacian form of VTE, the evolution of Dω follows a 3D-unsteady diffusion
equation, Eq. (2.25b), and therefore, to maintain the solenoidality of the vorticity
field, not only it has to be zero inside the computational domain, but one must also
has to ensure that no spurious generation of it takes place from the boundary. The
non-conservative formulation of VTE is most erroneous in terms of preserving the
divergence of vorticity Dω , as explained with the help of Eq. (2.25c). To maintain
Dω = 0 identically at all time-instants, it not only requires all the constraints on Dω ,
2.6 Role of Solenoidality Errors in Velocity-Vorticity Formulations 103

but also requires that Dv = 0, which acts as a source term for the evolution of Dω
(see Eq. (2.25c)).

→ →
The Laplacian and rotational variants of ( V , − ω )-formulation of NSE is solved
for flow inside the cubic LDC and results compared for sub-critical and post-critical
Reynolds numbers. For this purpose, the staggered grid used in [131] is employed,
as shown here in Fig. 2.10b on an elementary cell. The velocity components are
defined at the center of the plane, on which the component is perpendicular; while
the vorticity components are placed at mid-location on the edges of the cube which are
parallel to the direction of the vorticity components. Such grid staggering for 3D flows
require additional evaluation of derivatives of vorticity components (see Eqs. (2.23)
and (2.24a)). For example, the x-component of VTE would require evaluation of
derivatives of ω y and ωz at the location of ωx . This derivative evaluation has been
performed using optimized staggered compact scheme, Eq. (2.135), as described in
Sect. 2.3.2.

→ →
For the Laplacian variant of ( V , − ω )-formulation, the diffusion terms, i.e., terms


appearing in ∇ 2 ω are numerically evaluated using second order central difference

→ →
scheme. In contrast, in the rotational variant of ( V , −ω )-formulation, one requires to
evaluate the terms appearing in ∇ × (∇ × − →
ω ), which is performed by applying Eq.
(2.135) twice in each of the x-, y- and z-directions. Boundary conditions used for
this cubic cavity case appear directly from the zero-normal and no-slip conditions
at the wall, and is discussed elaborately in the literature [31, 63, 131]. A total of
81 × 81 × 81 uniformly distributed grid-points have been chosen with a time step
of Δt = 10−3 . As was reported for the previous 2D-simulations, here also, R K 4
scheme is used for time integration.
First, a set of results for the sub-critical Reynolds number of Re = 1000 (based
on the side of the cube and lid velocity) is computed using Laplacian and rotational

→ →
forms of ( V , − ω )-formulation. Time histories of maximum value of Dω and Dv in
the full domain, are shown in Fig. 2.32. It is noted that the values do not change
beyond t = 70 and hence the displayed histories are for the transient state only.
Displayed values of Dω are obtained using the compact scheme and it is clearly
noted that the rotational form is superior over the Laplacian form of the NSE in

→ →
(V ,− ω )-formulation. Although, one notes almost similar time variation of Dv for
both the forms (evaluated using simple C D2 scheme). It is to be noted that in literature
[37, 63, 131] results have been reported for the cubic LDC using Laplacian form. In
[2, 55, 56], this flow is computed using primitive variable formulation by spectral
element method with Gauss–Lobatto–Chebyshev collocation points.
In Fig. 2.33, u- and v-velocity components obtained at the mid-plane are shown

→ →
using these two variants of ( V , − ω )-formulation. The rotational form results match
identically with the Laplacian form for this sub-critical Reynolds number, despite the
large difference in Dω noted in Fig. 2.32. Also, the results obtained using Laplacian
form are seen to match with the results of [2], obtained using primitive variable
by spectral element method. The stream-traces and vorticity contours are shown in
Figs. 2.34 and 2.35 for different components, obtained using Laplacian and rotational
forms, and one cannot see visually any differences. Whatever differences are noted
104 2 DNS of Navier–Stokes Equation

Fig. 2.32 Flow inside a cubic lid-driven cavity for Re = 1000. (i, ii) Time-history of vorticity
divergence error plotted for computations using Laplacian and conservative rotational forms of
VTE, respectively. (iii, iv) Time-history of velocity divergence error plotted for Laplacian and
conservative rotational forms of VTE, respectively

for Dω , these occur at the top left corner only of the displayed (x, y)-plane. This
will become more evident when post-critical Reynolds number flows are studied.
Figs. 2.34 and 2.35 also show that the flow is symmetric about the mid (x, y)-plane.
Symmetry of the flow about the mid (x, y)-plane is retained, even during the transient
stage of flow development to steady state for this Re. At this plane of symmetry, the
cross-flow velocity w is found to be exactly zero. The strong similarity for the results
of the simulations using Laplacian and rotational forms of VTE are even noted for
all the intermediate times despite large difference in the vorticity solenoidality error
levels in Fig. 2.32. In Fig. 2.36, maximum values of Dω and Dv in the computational

→ →
domain, are compared between the rotational and the Laplacian forms of ( V , − ω )-
formulation for Re = 3200, which is a post-critical Reynolds number flow, as it
does not reach a steady state even after t = 2000. Also, the experimental results of
2.6 Role of Solenoidality Errors in Velocity-Vorticity Formulations 105

1 (i) 0.3 (ii)


0.2
0.8
0.1

0.6 0
y

v
Rotational Form -0.1
0.4 Laplacian Form
-0.2 Rotational Form
Laplacian Form
0.2 -0.3

-0.4
0
-0.2 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
u x

(iii) (iv)
1 0.3 Laplacian Form
Albensoeder & Kuhlmann
0.2
0.8
0.1

0.6 0
y

-0.1
0.4
-0.2

0.2 Laplacian Form -0.3


Albensoeder & Kuhlmann
-0.4
0
-0.2 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
u x

Fig. 2.33 (i, ii) u- and v-velocity components at the mid-plane plotted for Laplacian and rotational
forms of VTE and (iii, iv) compared with the results of Albensoeder and Kuhlmann [2] for flow
inside a cubic lid-driven cavity at Re = 1000

[51] for a rectangular parallelepiped (with square cross section) of aspect ratio of
L/D = 3 : 1, has shown this flow for Re = 3200 to be time-dependent. The present
computational results with the experimental results of [51] at the mid-plane are

→ →
compared. For this Re, both forms of ( V , − ω )-formulation show identical level of
Dv , although this does not decrease monotonically as was noted for Re = 1000 in
Fig. 2.32. For Dω , levels of error remain the same for these two forms, with rotational
form showing billion times smaller value of Dω .
In Fig. 2.37, computed velocity components in the mid (x, y)-plane are compared
with the results reported in [51] for the same (x, y)-plane. For comparison purpose,
computed results are time averaged between t = 1900 and t = 2000. The match
appears quite well, considering the fact that the presented computed results are for a
cubic cavity, while the experimental results are for a rectangular cavity with square
cross-section whose length is three-times longer than its width. Results are also
compared between rotational and Laplacian forms at t = 250 and t = 2000. Though
the u-velocity profiles do not show any distinguishable differences at t = 250 and t =
106 2 DNS of Navier–Stokes Equation

(i) Streamtrace at midsection (x,y)-plane (ii) ωz contour at midsection (x,y)-plane


Min = -131.06, Max = 84.59
1 1

0.8 0.8

0.6 0.6
y

y
0.4 0.4

0.2 0.2

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x x
(iii) Streamtrace at midsection (z,y)-plane (iv) ωx contour at midsection (z,y)-plane
Min = -4.65, Max = 4.65
1 1

0.8 0.8

0.6 0.6
y

0.4 0.4

0.2 0.2

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
z z
(v) Streamtrace at midsection (x,z)-plane (vi) ωy contour at midsection (x,z)-plane
1 Min = -4.41, Max = 4.41
1

0.8 0.8

0.6 0.6
z

0.4 0.4

0.2 0.2

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x x

Fig. 2.34 Stream-traces and contours of normal component of vorticity are shown plotted at mid-
(x, y) plane (frames (i) and (ii)), mid-(y, z) plane (frames (iii) and (iv)) and mid-(z, x) plane (frames
(v) and (vi)) for Re = 1000 after steady state is reached. All the plots correspond to the simulations
done using Laplacian form of VTE
2.6 Role of Solenoidality Errors in Velocity-Vorticity Formulations 107

(i) Streamtrace at midsection (x,y)-plane (ii) ωz contour at midsection (x,y)-plane


Min = -131.00, Max = 84.60
1 1

0.8 0.8

0.6 0.6
y

y
0.4 0.4

0.2 0.2

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x x
(iii) Streamtrace at midsection (z,y)-plane (iv) ωx contour at midsection (z,y)-plane
Min = -4.62, Max = 4.62
1 1

0.8 0.8

0.6 0.6
y

0.4 0.4

0.2 0.2

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
z z
(v) Streamtrace at midsection (x,z)-plane (vi) ωy contour at midsection (x,z)-plane
1 Min = -4.50, Max = 4.50
1

0.8 0.8

0.6 0.6
z

0.4 0.4

0.2 0.2

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x x

Fig. 2.35 Stream-traces and contours of normal component of vorticity are plotted at mid-(x, y)
plane (frames (i) and (ii)), mid-(z, y) plane (frames (iii) and (iv)) and mid-(z, x) plane (frames (v)
and (vi)) for Re = 1000 after steady state is reached. All the plots correspond to the simulations
done using conservative rotational form of VTE
108 2 DNS of Navier–Stokes Equation

Fig. 2.36 Flow inside a cubic lid-driven cavity for Re = 3200. (i, ii) Time-history of vorticity
divergence error plotted for computations using Laplacian and conservative rotational forms of
VTE, respectively. (iii, iv) Time-history of velocity divergence error plotted for Laplacian and
conservative rotational forms of VTE, respectively

2000, but there are differences for the v-velocity profiles at t = 250, specifically near
the rear wall. Also, one notices that in frame (vi), v-velocity profile corresponding
to the rotational form has slightly higher maximum and slightly lower minimum.
To further investigate the role played by the solenoidality error in vorticity, the
stream-traces and contours of normal components of vorticity at three different mid-
sections are plotted, as has been done in Figs. 2.34 and 2.35, at different times, in
Figs. 2.38, 2.39, 2.40 and 2.41. In Fig. 2.38 stream-traces are shown plotted for
(a) Laplacian and (b) rotational forms at t = 250 and t = 350, while in Fig. 2.40,
these correspond to t = 1800 and t = 2000. In Figs. 2.39 and 2.41, contours of
normal components of vorticity are shown plotted at three different mid-sections at
t = 250, 350 and at t = 1800, 2000, respectively, for these forms. It can be noticed
from Figs. 2.38, 2.39, 2.40 and 2.41 for this post-critical Reynolds number, that the
plotted contours and stream-traces show significant differences. These differences
are more visibly apparent in the mid (y, z)- or (z, x)-planes. The main differences in
2.6 Role of Solenoidality Errors in Velocity-Vorticity Formulations 109

1 (i) 0.5 (ii)


0.4 Simulation with L/D=1.0
0.8 0.3 Experiment with L/D=3.0

Simulation results are time 0.2


averaged between t = 1900
0.6 0.1
and t=2000
0

v
y

0.4 -0.1
Simulation results are time
-0.2 averaged between t = 1900
-0.3 and t=2000
0.2 Simulations with L/D=1.0
Experimental results of -0.4
Koseff & Street with
L/D=3.0 -0.5
0
0 0.2 0.4 x 0.6 0.8 1
-0.5 0 u 0.5 1

1 (iii) 0.5 (iv)


0.4
0.8 0.3
Laplacian Form
0.2 Rotational Form

0.6
t = 250 0.1
0
v
y

0.4 -0.1
Laplacian Form -0.2 t = 250
Rotational Form
0.2 -0.3
-0.4
-0.5
0
0 0.2 0.4 x 0.6 0.8 1
-0.4 -0.2 0 0.2 0.4 0.6 0.8 1
u

1 (v) 0.5 (vi)


0.4
Laplacian Form
0.8 0.3 Rotational Form
t = 2000 0.2

0.6 0.1
0
v
y

Laplacian Form -0.1


0.4
Rotational Form
-0.2 t = 2000
0.2 -0.3
-0.4
-0.5
0
-0.5 0 0.5 1 0 0.2 0.4 x 0.6 0.8 1
u

Fig. 2.37 Flow inside a cubic lid-driven cavity for Re = 3200. (i, ii) u- and v-velocity components
at the mid-plane and time averaged between t = 1900 and 2000 by using Laplacian form are
compared with the experimental time averaged results of [51]. (iii, v) u- and (vi, vi) v-velocity
components at mid-plane obtained by using rotational form and Laplacian form are compared at
t = 250 and t = 2000 as indicated in the frames
110 2 DNS of Navier–Stokes Equation

Fig. 2.38 Stream-traces shown plotted for a Laplacian form of VTE and b rotational form of VTE
used, at t = 250 and 350 on mid-planes
2.6 Role of Solenoidality Errors in Velocity-Vorticity Formulations 111

(a) Laplacian Form


(i) t = 250, ωz contours in (x,y)-plane (ii) t = 250, ωx contours in (y,z)-plane (iii) t = 250, ωy contours in (z,x)-plane
Min = -139.037, Max = 105.089 Min = -9.5896, Max = 9.5896 Min = -4.026, Max = 4.026
1 1 1

0.8 0.8 0.8

0.6 0.6 0.6


y

z
0.4 0.4 0.4

0.2 0.2 0.2

0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x z x
(iv) t = 350, ωz contours in (x,y)-plane (v) t = 350, ωx contours in (y,z)-plane (vi) t = 350, ωy contours in (z,x)-plane
Min = -139.037, Max = 105.089 Min = -10.134, Max = 10.134 Min = -4.092, Max = 4.093
1 1 1

0.8 0.8 0.8

0.6 0.6 0.6


y

z
0.4 0.4 0.4

0.2 0.2 0.2

0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x z x

(b) Rotational Form


(i) t = 250, ωz contours in (x,y)-plane (ii) t = 250, ωx contours in (y,z)-plane (iii) t = 250, ωy contours in (z,x)-plane
Min = -138.969, Max = 106.35 Min = -10.192, Max = 10.192 Min = -4.489, Max = 4.489
1 1 1

0.8 0.8 0.8

0.6 0.6 0.6


y

0.4 0.4 0.4

0.2 0.2 0.2

0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x z x
(iv) t = 350, ωz contours in (x,y)-plane (v) t = 350, ωx contours in (y,z)-plane (vi) t = 350, ωy contours in (z,x)-plane
Min = -138.831, Max = 106.746 Min = -10.596, Max = 10.596 Min = -5.361, Max = 5.361
1 1 1

0.8 0.8 0.8

0.6 0.6 0.6


y

0.4 0.4 0.4

0.2 0.2 0.2

0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x z x

Fig. 2.39 Contours of vorticity component normal to the indicated mid-planes shown plotted a
Laplacian form of VTE and b rotational form of VTE are used at t = 250 and t = 350
112 2 DNS of Navier–Stokes Equation

Fig. 2.40 Stream-traces shown plotted for a Laplacian form of VTE and b rotational form of VTE
used at t = 1800 and 2000 on mid-planes
2.6 Role of Solenoidality Errors in Velocity-Vorticity Formulations 113

Fig. 2.41 Contours of vorticity component normal to the indicated mid-planes shown plotted for
a Laplacian form of VTE and b rotational form of VTE used at t = 1800 and 2000
114 2 DNS of Navier–Stokes Equation

the flow topology is the location and the nature of the wall stagnation points, as seen
in Fig. 2.38. As noted in Fig. 2.38a, right and left wall stagnation points are of saddle
type, for both frames (i) and (iv). The nature of singularity on the right wall is same
for both the forms in Fig. 2.38a, b. However, there is qualitative difference in the
nature of singularity noted on the left wall between these two forms. For the rotational
form, it is node at t = 250, that changes into a saddle at t = 350. For the Laplacian
form, no such time variation is noted during this time interval. The singular points
on the left wall changes more rapidly for the rotational form, from being a node to a
saddle point. For example, it changes again to a node at t = 450. This phenomenon
happens alternately without a single time period. Similar alteration of the singular
point on the left wall takes place with the Laplacian form more infrequently. During
the initial stages, the rotational form also captures a coherent vortical structure at the
top left corner of the cavity (as can be seen in the frame (iv) of Fig. 2.38b), which
is absent in the Laplacian form. Hence, it can be concluded that the rotational form
is more receptive in capturing flow instabilities, as compared to the Laplacian form.
Undoubtedly, this points to the higher accuracy of the rotational form over the other
forms, using velocity and vorticity. This is also noted by the fact that the rotational
form looses symmetry earlier (at around t = 550), as compared to the Laplacian
form, which looses symmetry at around t = 650. A possible explanation for this
is that the rotational form is more receptive to the numerical errors and hence it
looses symmetry at an earlier time instant. The maximum and the minimum values
of w-velocity at t = 250, in the mid (x, y)-plane for the Laplacian form are noted
as, 3.5 × 10−7 and −1.25 × 10−7 , respectively. In contrast, for the rotational form,
corresponding values are 1.23 × 10−6 and −2.45 × 10−6 , respectively. Thus, apart
from the earlier onset time of instability, the rotational form also displays higher
growth rate of disturbances by an order of magnitude. As noted earlier for the flow
at Re = 1000, w-velocity is exactly zero at the mid (x, y)-plane of symmetry, for
both the formulations.
In Fig. 2.39, contours of normal components of vorticity are shown plotted in
three mid-planes. Frames (i) and (iv) in Fig. 2.39a, show higher amount of numerical
fluctuations at top left corners for the Laplacian form. Such fluctuations at grid
scale level are due to aliasing error and in the conservative rotational form, the
quantum of it is significantly lower in the frames (i) and (iv) of Fig. 2.39b. Similar
observations are noted in Fig. 2.41 for Re = 3200 case, where normal component of
vorticity contours are plotted at the later time, t = 1800 and t = 2000. In Fig. 2.40,
corresponding stream-traces at three different indicated planes are shown. In this
figure, one notices significant lack of symmetry in the mid (y, z)- and (z, x)-planes.
But, this asymmetry is more pronounced for the rotational form. As a result of such a
high level of asymmetry, one can clearly observe the presence of multiple limit cycles
in the stream-trace plots in the mid (x, y)-plane of Fig. 2.40b. Comparing frames
(iii) and (vi) of Fig. 2.40a, with corresponding frames in Fig. 2.40b, higher number
of vortices along y-axis can be observed near the rear wall for the rotational form.
From these observations of the simulated flow inside a cubic cavity for Re = 3200,
one notes the rotational form to capture details of unsteady flows relatively more
accurately, than the Laplacian form.
References 115

References

1. Adams, Y. (1977). Highly accurate compact implicit method and boundary conditions. Com-
puter Physics, 24, 10–22.
2. Albensoeder, S., & Kuhlmann, H. C. (2006). Nonlinear three-dimensional flow in the lid-
driven square cavity. Journal of Fluid Mechanics, 569, 465–480.
3. Auld, B. (1973). Acoustic fields and waves in solids. New York: Wiley-Interscience.
4. Barone, M. F., & Roy, C. J. (2006). Evaluation of detached eddy simulation for turbulent wake
applications. AIAA Journal, 44(12), 3062–3071.
5. Bernardini, M., & Pirozzoli, S. (2009). A general strategy for the optimization of Runge-
Kutta schemes for wave propagation phenomena. Journal of Computational Physics, 228,
4182–4199.
6. Bhaumik, S. (2013). Direct numerical simulation of inhomogeneous transitional and turbulent
flows. Ph.D. Thesis, I. I. T. Kanpur.
7. Bhaumik, S., Sengupta, T. K. (2011). On the divergence-free condition of velocity in two-
dimensional velocity-vorticity formulation of incompressible Navier–Stokes equation. In
AIAA-2011-3238, AIAA CFD Conference, Honolulu, Hawaii, USA.
8. Bhaumik, S., & Sengupta, T. K. (2015). A new velocity-vorticity formulation for direct numer-
ical simulation of 3D transitional and turbulent flows. Journal of Computational Physics, 284,
230–260.
9. Bhaumik, S., & Sengupta, T. K. (2014). Precursor of transition to turbulence: Spatiotemporal
wave front. Physical Review E, 89(4), 043018.
10. Bhumkar, Y. G. (2011). High performance computing of bypass transition. Ph.D. Thesis, I. I.
T. Kanpur.
11. Bhumkar, Y. G., & Sengupta, T. K. (2011). Adaptive multi-dimensional filters. Computers
Fluids, 49(1), 128–140.
12. Bogey, C., & Bailly, C. (2004). A family of low dispersive and low dissipative explicit schemes
for flow and noise computations. Journal of Computational Physics, 194, 194–214.
13. Brillouin, L. (1960). Wave propagation and group velocity. New York: Academic Press.
14. Brooks, A. N., & Hughes, T. J. R. (1982). Streamline upwind/Petrov-Galerkin formulation for
convection dominated flows with particular emphasis on the incompressible Navier–Stokes
equations. Computer Methods in Applied Mechanics and Engineering, 32, 199–259.
15. Buresti, G. (2015). A note on Stokes hypothesis. Acta Mechanica, 226, 3555–9.
16. Charney, J. G., Fj∅rtoft, R., & Von Neumann, J. (1950). Numerical integration of the
barotropic vorticity equation. Tellus, 2(4), 237–254.
17. Chorin, A. J. (1968). Numerical solution of the Navier–Stokes equation. Mathematics of
Computation, 22, 745–762.
18. Chu, Peter C., & Fan, Chenwu. (1998). A three-point combined compact difference scheme.
Journal of Computational Physics, 140(2), 370–399.
19. Cramer, M. S. (2012). Numerical estimates for the bulk viscosity of ideal gases. Physics of
Fluids, 24(066102), 1–23.
20. Crank, J., & Nicolson, P. (1947). A practical method for numerical evaluation of solutions of
partial differential equations of the heat conduction type. Proceedings of Cambridge Philo-
sophical Society, 43(50), 50–67.
21. Crouch, J. D., Garbaruk, A., Magidov, D., & Travin, A. (2009). Origin of transonic buffet on
aerofoils. Journal of Fluid Mechanics, 628, 357–369.
22. Cullen, M. J. P. (1974). A finite-element method for a non-linear initial value problem. J. Int.
Math. Appl., 31, 233–247.
23. Davidson, P. A. (2004). Turbulence: An introduction for scientists and engineers. Oxford,
UK: Oxford University Press.
24. Dendy, F. E. (1974). Sediment trap efficiency of small reservoirs. Transaction of the American
Society of Agricultural Engineers, 17(5), 898–908.
25. Dipankar, A., & Sengupta, T. K. (2006). Symmetrized compact scheme for receptivity study
of 2D transitional channel flow. Journal of Computational Physics, 215(1), 245–253.
116 2 DNS of Navier–Stokes Equation

26. Emanuel, G. (1990). Bulk viscosity of a dilute polyatomic gas. Physics of Fluids A, 2, 2252–
2254.
27. Gad-el-Hak, M. (1995). Questions in fluid mechanics. Stokes’ hypothesis for a Newtonian,
isotrspic fluid. Journal of Fluids Engineering, 117, 3–5. (Technical Forum).
28. Gaitonde, D. V., Shang, J. S., & Young, J. L. (1999). Practical aspects of higher-order numerical
schemes for wave propagation phenomena. International Journal for Numerical Methods in
Engineering, 45, 1849.
29. Gaitonde, D. V. & Visbal, M. R. (1999). Further development of a Navier-Stokes solution pro-
cedure based on higher order formulas. In 37th Aerospace Sciences Meeting and Exhibition,
AIAA 99–0557, Reno, NV.
30. Gaitonde, D. V., & Visbal, M. R. (2000). Padé− type higher-order boundary filters for the
Navier–Stokes equations. AIAA Journal, 38(11), 2103.
31. Gatski, T. B. (1991). Review of incompressible fluid flow computations using the vorticity-
velocity formulation. Applied Numerical Mathematics, 7, 227–239.
32. Gatski, T. B., Grosch, C. E., & Rose, M. E. (1982). A numerical study of the two-dimensional
Navier-Stokes equations in vorticity-velocity variables. Journal of Computational Physics,
48, 1–22.
33. Gatski, T. B., Grosch, C. E., & Rose, M. E. (1989). A numerical solution of the Navier–Stokes
equations for three-dimensional, unsteady, incompressible flows by compact schemes. Journal
of Computational Physics, 82, 298–329.
34. Ghosal, S., & Moin, P. (1995). The basic equations for the large eddy simulation of turbulent
flows in complex geometry. Journal of Computational Physics, 118, 24.
35. Graves, R. E., & Argrow, B. M. (1999). Bulk viscosity: Past to present. Journal of Thermo-
physics and Heat Transfer, 13(3), 337–342.
36. Grescho, P. M., & Sani, R. L. (1998). Incompressible flow and the finite element method.
Chichester, UK: Wiley.
37. Guj, G., & Stella, F. (1993). A vorticity-velocity method for the numerical solution of 3D
incompressible flows. Journal of Computational Physics, 106, 286–298.
38. Hamilton, W. R. (1839). The collected mathematical papers (Vol. 4). Cambridge: Cambridge
University Press.
39. Haras, Z., & Ta’asan, S. (1994). Finite difference scheme for long time integration. Journal
of Computational Physics, 14, 265–279.
40. Hirasaki, G. J., & Hellums, J. D. (1968). A general formulation of the boundary conditions on
the vector potential in three-dimensional hydrodynamics. Quarterly of Applied Mathematics,
26(3), 331–342.
41. Hirasaki, G. J., & Hellums, J. D. (1970). Boundary conditions on the vector and scalar poten-
tials in viscous three-dimensional hydrodynamics. Quarterly of Applied Mathematics, 28(2),
293–296.
42. Hirsch, C. (1990). Numerical computation of internal and external flows (Vol. I and II).,
Computational methods for inviscid and viscous flows. Chichester, UK: Wiley.
43. Holdeman, J. T. (2012). A velocity-stream function method for three-dimensional incom-
pressible fluid flow. Computer Methods in Applied Mechanics and Engineering, 209, 66–73.
44. Huang, H., & Li, M. (1997). Finite-difference approximation for the velocity-vorticity for-
mulation on staggered and non-staggered grids. Computer & Fluids, 26(1), 59–82.
45. Hu, F. Q., Hussani, M. Y., & Manthey, J. L. (1996). Low-dissipation and low-dispersion
Runge-Kutta schemes for computational acoustics. Journal of Computational Physics, 124,
177–191.
46. Jenkins, F., & White, H. (1973). Fundamentals of physical optics. New York: McGraw-Hill.
47. Karim, S. M., & Rosenhead, L. (1952). Review of Modern Phys., 24, 108–16.
48. Kennedy, C. A., & Carpenter, M. H. (1994). Several new numerical methods for compressible
shear-layer simulations. Applied Numerical Mathematics, 14, 397.
49. Kloker, M., Konzelmann, U., & Fasel, H. (1993). Outflow boundary conditions for spatial
Navier–Stokes simulations of transitional boundary layers. AIAA Journal, 31, 620.
50. Kopal, Z. (1966). Numerical analysis. New York, USA: Springer.
References 117

51. Koseff, J. R., & Street, R. L. (1984). On end wall effects in a lid-driven cavity flow. Journal
of Fluids Engineering, 106, 385–398.
52. Kreiss, H., & Oliger, J. (1972). Comparison of accurate methods for the integration of hyper-
bolic equations. Tellus, 24, 199–215.
53. Lele, S. K. (1992). Compact finite difference schemes with spectral-like resolution. Journal
of Computational Physics, 103(1), 16–42.
54. Leonard, B. P., Leschziner, M. A. & McGuirk, J. (1978). Third order finite-difference method
for steady two-dimensional convection. In C. Taylor, K. Morgan, & C. A. Brebbia (Eds.)
Numerical Methods in Laminar and Turbulent Flows (pp. 807–819). London: Pentech Press.
55. Leriche, E. (2006). Direct numerical simulation in a lid-driven cubical cavity at high Reynolds
number by a Chebyshev spectral method. Journal of Scientific Computing, 27, 335–345.
56. Leriche, E., & Gavrilakis, S. (2000). Direct numerical simulation of the flow in a lid-driven
cubical cavity. Physics of Fluids, 12(6), 1363–1376.
57. Liebermann, L. N. (1949). The second viscosity of liquids. Physical Review, 75(9), 1415.
58. Lighthill, M. J. (1978). Fourier analysis and generalized functions. Cambridge, UK: Cam-
bridge University Press.
59. Mathew, J., Lechner, R., Foysi, H., Sesterhenn, J., & Friedrich, R. (2003). An explicit filtering
method for large eddy simulation of compressible flows. Journal of Computational Physics,
15(8), 2279.
60. Meitz, H. L., & Fasel, H. F. (2000). A compact-difference scheme for the Navier–Stokes
equations in vorticity-velocity formulation. Journal of Computational Physics, 157, 371–403.
61. Najjar, F. M., & Tafti, D. K. (1996). Study of discrete test filters and finite difference approx-
imations for the dynamic subgrid-scale stress model. Journal Computational Physics, 8(4),
1076.
62. Nagarajan, S., Lele, S. K., & Ferziger, J. H. (2003). A robust high-order compact method for
large eddy simulation. Journal of Computational Physics, 19, 392–419.
63. Napolitano, M., & Pascazio, G. (1991). A numerical method for the vorticity-velocity Navier-
Stokes equations in two and three dimensions. Computer & Fluids, 19, 489–495.
64. Ortega-Casanova, Joaqun, & Fernandez-Feria, Ramn. (2008). A numerical method for the
study of nonlinear stability of axisymmetric flows based on the vector potential. Journal of
Computational Physics, 227(6), 3307–3321.
65. Oswald, G. A., Ghia, K. N. & Ghia, U. (1988). Direct solution methodologies for the unsteady
dynamics of an incompressible fluid. In S. N. Atluri, G. Yagawa (Eds.) International Confer-
ence on Computer Engineering Science, vol. 2 Atlanta. Berlin: Springer.
66. Oberleithner, K., Sieber, M., Nayeri, C. N., Paschereit, C. O., Petz, C., Hege, H.-C., et al.
(2011). Three-dimensional coherent structures in a swirling jet undergoing vortex breakdown:
stability analysis and empirical mode construction. Journal of Fluid Mechanics, 679, 383–414.
67. Pirozzoli, S. (2007). Performance analysis and optimization of finite-difference schemes for
wave propagation problems. Journal of Computational Physics, 222, 809–831.
68. Poinsot, T., & Veynante, D. (2005). Theoretical and numerical combustion (2nd ed.). PA:
Edwards.
69. Rajagopal, K. R. (2013). A new development and interpretation of the Navier-Stokes fluid
which reveals why the “Stokes assumption” is inapt. International Journal of Non-Linear
Mechanics, 50, 141–151.
70. Rajpoot, M. K., Bhaumik, S., & Sengupta, T. K. (2012). Solution of linearized rotating shallow
water equations by compact schemes with different grid-staggering strategies. Journal of
Computational Physics, 231, 2300–2327.
71. Rajpoot, M. K., Sengupta, T. K., & Dutt, P. K. (2010). Optimal time advancing dispersion
relation preserving schemes. Journal of Computational Physics, 229(10), 3623–3651.
72. Ramboer, J., Broeckhoven, T., Smirnov, S., & Lacor, C. (2006). Optimization of time inte-
gration schemes coupled with spatial discretization for use in CAA applications. Journal of
Computational Physics, 213, 777–802.
73. Rayleigh, L. (1889). Scientific papers (Vol. 1). Cambridge: Cambridge University Press.
74. Rayleigh, L. (1890). Scientific papers (Vol. 2). Cambridge: Cambridge University Press.
118 2 DNS of Navier–Stokes Equation

75. Raymond, W. H., & Garder, A. (1976). Selective damping in a Galerkin method for solving
wave problems with variable grids. Monthly Weather Review, 104, 1583–1590.
76. Reynolds, W. C., & Hussain, A. K. M. F. (1972). The mechanics of an organized wave in
turbulent shear flow. Part 3. Theoretical models and comparisons with experiments. Journal
of Fluid Mechanics, 54(2), 263–288.
77. Rizzetta, D. P., Visbal, M. R., & Blaisddell, G. A. (2003). A time-implicit high-order com-
pact differencing and filtering scheme for large-eddy simulation. International Journal for
Numerical Methods in Fluids, 42, 655.
78. Rizzetta, D. P., Visbal, M. R. & Morgan, P. E. (2008). A high-order compact finite-difference
scheme for large-eddy simulation of active flow control. In 46th aerospace sciences meeting
and exhibition, AIAA 2008-526, Reno, NV.
79. Rosenhead, L. (1954). Proceedings of Royal Society London A, 226, 1–6.
80. Sagaut, P. (2002). Large eddy simulation for incompressible flows. Berlin: Springer.
81. Sengupta, T. K. (2004). Fundamentals of computational fluid dynamics. Hyderabad (India):
Universities Press.
82. Sengupta, T. K. (2012). Instabilities of flows and transition to turbulence. Florida, USA: CRC
Press, Taylor & Francis Group.
83. Sengupta, T. K. (2013). High accuracy computing methods: fluid flows and wave phenomenon.
New York, USA: Cambridge University Press.
84. Sengupta, T. K., Ballav, M., & Nijhawan, S. (1994). Generation of Tollmien-Schlichting waves
by harmonic excitation. Physics of Fluids, 6(3), 1213–1222.
85. Sengupta, T. K., & Bhaumik, S. (2011). Onset of turbulence from the receptivity stage of fluid
flows. Physics Review Letter, 154501, 1–5.
86. Sengupta, T. K., Bhaumik, S., & Bhumkar, Y. G. (2012). Direct numerical simulation of two-
dimensional wall-bounded turbulent flows from receptivity stage. Physical Review E, 85(2),
026308.
87. Sengupta, T. K., Bhaumik, S., & Bose, R. (2013). Direct numerical simulation of transitional
mixed convection flows: Viscous and inviscid instability mechanisms. Physics of Fluids, 25,
094102.
88. Sengupta, T. K., Bhaumik, S., & Usman, S. (2011). A new compact difference scheme for sec-
ond derivative in non-uniform grid expressed in self-adjoint form. Journal of Computational
Physics, 230(5), 1822–1848.
89. Sengupta, T. K., & Bhumkar, Y. G. (2010). New explicit two-dimensional higher order filters.
Computers & Fluids, 39, 1848–1863.
90. Sengupta, T. K., Bhumkar, Y., & Lakshmanan, V. (2009). Design and analysis of a new filter
for LES and DES. Computers & Structures, 87, 735–750.
91. Sengupta, T. K., Bhumkar, Y., Rajpoot, M. K., Suman, V. K., & Saurabh, S. (2012). Spurious
waves in discrete computation of wave phenomena and flow problems. Applied Mathematics
and Computation, 218, 9035–9065.
92. Sengupta, T. K., & Dey, S. (2004). Proper orthogonal decomposition of direct numerical
simulation data of by-pass transition. Computers & Structures, 82, 2693–2703.
93. Sengupta, T. K., & Dipankar, A. (2004). A comparative study of time advancement methods
for solving Navier–Stokes equations. Journal of Scientific Computing, 21(2), 225–250.
94. Sengupta, T. K., & Dipankar, A. (2005). Subcritical instability on the attachment-line of an
infinite swept wing. Journal of Fluid Mechanics, 529, 147–171.
95. Sengupta, T. K., Dipankar, A., & Sagaut, P. (2007). Error dynamics: Beyond von Neumann
analysis. Journal of Computational Physics, 226, 1211–1218.
96. Sengupta, T. K., Ganeriwal, G., & De, S. (2003). Analysis of central and upwind compact
schemes. Journal of Computational Physics, 192, 677–694.
97. Sengupta, T. K., Jain, R., & Dipankar, A. (2005). A new flux-vector splitting compact finite
volume scheme. Journal of Computational Physics, 207, 261–281.
98. Sengupta, T. K., Lakshmanan, V., & Vijay, V. V. S. N. (2009). A new combined stable and
dispersion relation preserving compact scheme for non-periodic problems. Journal of Com-
putational Physics, 228(8), 3048–3071.
References 119

99. Sengupta, T. K. & Nair, M. T. (1997). A new class of wave Blasius boundary layer. In
Proceedings 7th Asian Congress of Fluid Mechanics
100. Sengupta, T. K., Rajpoot, M. K., & Bhumkar, Y. G. (2011). Space-time discretizing optimal
DRP schemes for flow and wave propagation problems. Computers & Fluids, 47(1), 144–154.
101. Sengupta, T. K., Rao, A. K., & Venkatasubbaiah, K. (2006). Spatio-temporal growing wave
fronts in spatially stable boundary layers. Physical Review Letters, 96(22), 224504.
102. Sengupta, T. K., Rao, A. K., & Venkatasubbaiah, K. (2006). Spatio-temporal growth of dis-
turbances in a boundary layer and energy based receptivity analysis. Physics of Fluids, 18,
094101.
103. Sengupta, T. K., Sengupta, A., Sengupta, S., Bhole, A. & Shruti, K. S. (2016). Non-equilibrium
thermodynamics of Rayleigh–Taylor instability. International Journal Thermophysics, 37(4),
1–2. https://doi.org/10.1007/s10765-016-2045-1
104. Sengupta, T. K., Sengupta, A., Sharma, N., Sengupta, S., Bhole, A., & Shruti, K. S. (2016).
Roles of bulk viscosity on Rayleigh-Taylor instability: Non-equilibrium thermodynamics due
to spatio-temporal pressure fronts. Physics of Fluids, 28, 094102.
105. Sengupta, T. K., Singh, N., & Vijay, V. V. S. N. (2011). Universal instability modes in internal
and external flows. Computers & Fluids, 40, 221–235.
106. Sengupta, T. K., Sircar, S. K., & Dipankar, A. (2006). High accuracy compact schemes for
DNS and acoustics. Journal of Scientific Computing, 26(2), 151–193.
107. Sengupta, T. K., Vijay, V. V. S. N., & Bhaumilk, S. (2009). Further improvement and analysis of
CCD scheme: Dissipation discretization and de-aliasing properties. Journal of Computational
Physics, 228(17), 6150–6168.
108. Shokin, Y. I. (1983). The method of differential approximation. Berlin: Springer.
109. Stokes, G. G. (1845). On the theories of inertial friction of fluids in motion. Transaction of
Cambridge Philosphical Society, 8, 287–305.
110. Stoltz, S., Adams, N. A., & Kleiser, L. (2001). An approximate deconvolution model for large-
eddy simulation with application to incompressible wall-bounded flows. Physics of Fluids,
13(4), 997.
111. Swartz, B., & Wendroff, B. (1974). The relation between the Galerkin and collocation methods
using smooth splines. SIAM Journal on Numerical Analysis, 11(5), 994–996.
112. Tam, C. K. W. (1971). Directional acoustic radiation from a supersonic jet generated by shear
layer instability. Journal of Fluid Mechanics, 46, 757–768.
113. Tam, C. K. W., Viswanathan, K., Ahuja, K. K. & Panda, J. (2008). The sources of jet noise:
experimental evidence. Journal of Fluid Mechanics, 615, 253–292.
114. Tennekes, H., & Lumley, J. L. (1971). First course in turbulence. Cambridge, MA: MIT Press.
115. Touber, E. & Sandham, N. D. (2009). Large-eddy simulation of low-frequency unsteadiness in
a turbulent shock-induced separation bubble. Theoretical and Computational Fluid Dynamics,
23(2), 79–107.
116. Trefethen, L. N. (1982). Group velocity in finite difference schemes. SIAM Review, 24(2),
113–136.
117. Tucker, P. G. (2003). Differential equation-based wall distance computation for DES and
RANS. Journal of Computational Physics, 190, 229–248.
118. Van der Vorst, H. A. (1992). Bi-CGSTAB: A fast and smoothly converging variant of Bi-CG
for the solution of non-symmetric linear systems. SIAM Journal on Scientifc and Statistical
Computing, 12, 631–644.
119. Vasilyev, O. V., Lund, T. S., & Moin, P. (1998). A general class of commutative filters for
LES in complex geometries. Journal of Computational Physics, 146, 82.
120. Vichnevetsky, R., & Bowles, J. B. (1982). Fourier analysis of numerical approximations of
hyperbolic equations., SIAM studies of applied mathematics Philadelphia, USA.
121. Visbal, M. R., & Gaitonde, D. V. (1999). High-order-accurate methods for complex unsteady
subsonic flows. AIAA Journal, 37(10), 1231.
122. Visbal, M. R., & Gaitonde, D. V. (2002). On the use of higher-order finite-difference schemes
on curvilinear and deforming meshes. Journal of Computational Physics, 181, 155.
120 2 DNS of Navier–Stokes Equation

123. Wahlbin, L. B. (1974). A dissipative numerical method for the numerical solution of first
order hyperbolic equations. In C. de Boor (Ed.) Mathematical aspects of finite elements in
partial deferential equations (pp. 147–170). New York: Academic Press.
124. Wahlbin, L. B. (1975). A modified Galerkin procedure with cubics for hyperbolic problems.
Mathematics of Computation, 29, 978–984.
125. Wang, L. P., & Rosa, B. (2009). A spurious evolution of turbulence originated from round-off
error in pseudo-spectral simulation. Computers & Fluids, 38, 1943–1949.
126. Weinan, E., & Liu, Jian-Guo. (1996). Vorticity boundary condition and related issues for finite
difference schemes. Journal of Computational Physics, 124(2), 368–382.
127. Whitham, G. B. (1974). Linear and nonlinear waves. New York: Wiley-Intescience.
128. Williams, G. P. (1969). Numerical integration of the three dimensional Navier-Stokes equation
for incompressible flow. Journal of Fluid Mechanics, 37, 727–750.
129. Wong, A. K., & Reizes, J. A. (1984). An effective vorticity-vector potential formulation for
the numerical solution of three-dimensional duct flow problems. Journal of Computational
Physics, 55(1), 98–114.
130. Wong, A. K., & Reizes, J. A. (1986). The vector potential in the numerical solution of three-
dimensional fluid dynamics problems in multiply connected regions. Jorunal of Computa-
tional Physics, 62(1), 124–142.
131. Wu, X. H., Wu, J. Z., & Wu, J. M. (1995). Effective vorticity-velocity formulations for 3D
incompressible viscous flows. Journal of Computational Physics, 122, 68–82.
132. Zingg, D. W. (2000). Comparison of high-accuracy finite-difference schemes for linear wave
propagation. SIAM Journal on Scientific Computing, 22(2), 476–502.
Chapter 3
Receptivity and Instability

3.1 Linear Stability/Receptivity Theories: Classical


Approaches and Signal Problem

In this chapter, linear stability and receptivity analysis of the zero-pressure gra-
dient (ZPG) boundary layer, under the parallel flow assumption, is discussed. This
assumption implies that the equilibrium flow quantities do not grow in the streamwise
direction and requires solving the Orr-Sommerfeld equation (OSE) to study evolu-
tion of disturbance field in a linearized analysis. The concept of the spatio-temporal
wave-front (STWF) originates from the receptivity analysis with the OSE solved
for the response field. First, the simplified description of equilibrium flow in terms
of a similarity solution for ZPG boundary layer is presented. Following which the
OSE is derived for boundary layers, making use of the parallel flow approximation
[19, 53]. This equation have been solved for the ZPG boundary layer using ana-
lytical approaches in [28, 48, 71]. We instead introduce the compound matrix
method, a robust method for stiff differential equation useful for the OSE. Finally, the
receptivity analysis of the ZPG boundary layer flow is provided, with results taken
from [61, 62]. The unique feature of the materials in this chapter is the topic of
instability of mixed convection flows for which two theorems are enunciated for
an inviscid linear mechanism, based on materials extensively taken from Sengupta
et al., Physics of Fluids, 25, 094102 (2013).
The unique feature of the materials in this chapter is the topic of instability of
mixed convection flows for which two theorems are enunciated for an inviscid linear
mechanism, based on materials extensively taken from a publication.1

1 [Reproduced from Direct numerical simulation of transitional mixed convection flows: Viscous
and inviscid instability mechanism. Sengupta et al., Physics of Fluids, 25, 094102 (2013), with the
permission of AIP Publishing.]
© Springer Nature Singapore Pte Ltd. 2019 121
T. K. Sengupta and S. Bhaumik, DNS of Wall-Bounded Turbulent Flows,
https://doi.org/10.1007/978-981-13-0038-7_3
122 3 Receptivity and Instability

3.1.1 The Equilibrium Flow Equation

For flow past a flat plate the effects of shear stress due to fluid viscosity is lim-
ited to a very thin layer, commonly known as the boundary layer (or alternately
wall-bounded shear layer), whose thickness increases as one moves downstream
along the streamwise direction. This is the boundary layer assumption introduced
by Prandtl, which revolutionized our understanding of fluid mechanics and is the
bedrock of perturbation theory used in many branches of mathematics and physics.
For thin shear layers, the governing NSE can be simplified based on thin shear layer
assumption, culminating in the boundary layer equation [14]. Furthermore, if the
edge velocity at the free-stream Ue varies as x m , where x is the streamwise distance
from the plate or wedge leading edge, the boundary layer equations can be further
simplified by what is known as similarity transformation [14], that reduces this to
a nonlinear ordinary differential equations, called the Falkner–Skan equation and is
given as [14, 53],
  2 
d3 f df m + 1 d2 f
+ m 1 − + f =0 (3.1)
d η̄3 d η̄ 2 d η̄2

where, η̄ = y Ue /νx, ψ = (Ue νx)1/2 f (η̄) and u(x, y)/Ue (x) = d f /d η̄. Here, Ue ,
y, ψ, u and ν are the boundary layer edge velocity, wall-normal co-ordinate, stream
function, the streamwise velocity component and the kinematic viscosity, respec-
tively. To obtain the similarity profile, one has to solve Eq. (3.1) subject to the no-slip
and zero-normal velocity at the wall. The boundary condition at the free-stream is
obtained as u(x, y) → Ue (x) as y → ∞. In terms of f , these conditions translate
to

f (0) = f (0) = 0

f → 1 as η̄ → ∞ (3.2)

where,  the prime refers to first derivative with respect to the similarity variable, η̄.
For ZPG boundary layer Ue (x) = U∞ = Const.. This simplifies Eq. (3.1) to the the
well known Blasius equation [14] by noting that m = Uxe dU dx
e
= 0 for ZPG condition
and the governing equation is given as,

 1 
f + ff =0 (3.3)
2
  
In Fig. 3.1, f , f and f obtained from solving Eq. (3.3) are plotted as a function of
η̄. The governing equation is solved using uniformly distributed points up to a limit

of η̄max = 12 by fourth order, RK4 method [30]. In terms of physical significance, f ,
 
f and f are related to the streamwise component of velocity, its first and second

derivatives with respect to η̄, respectively. One notes from Fig. 3.1c that f (0) = 0,
which signifies that the inflection-point (i.e., the point where d 2 u/dy 2 = 0) for the
ZPG boundary layer is exactly at the wall. Hence, the ZPG Blasius profile does
3.1 Linear Stability/Receptivity Theories: Classical Approaches and Signal Problem 123

(a)
1
0.8
0.6
f′

0.4
0.2
0
0 2 4 η 6 8

(b)
0.3

0.2
f′′

0.1

0
0 2 4 η 6 8

(c)
0

-0.05
′′′
f

-0.1

0 2 4 η 6 8

Fig. 3.1 The solution of the Blasius equation plotted as a function of the similarity variable η̄ for
  
a f , b f and c f
124 3 Receptivity and Instability

not strictly satisfy the Rayleigh’s inflection-point theorem for inviscid instability
[19, 38].
The velocity profile of the mean flow at any streamwise location is characterized
by, (i) the edge velocity Ue , (ii) the displacement thickness δ ∗ and (iii) the momentum
thickness θ , at that particular location (see [14, 53] for details). The displacement
thickness and the momentum thickness for any velocity profile is defined as
 ∞ 
∗ u
δ = 1− dy
0 Ue
 ∞  
u u
θ= 1− dy (3.4)
0 U e U e

For Blasius profile given by Eq. (3.3), it can be shown that [14]

1.72x 0.664x
δ∗  √ and θ  √ (3.5)
Rex Rex

where, Rex = U∞ x/ν is the local Reynolds number. It can also be shown that
for Blasius profile, boundary layer thickness δ̄  3δ ∗ , where δ̄ is defined such that
u(x, δ̄)  0.99U∞ .

3.2 Linear Stability Equation

In classical instability theory, the governing equation is linearized, as one would


be looking for growth of infinitesimally small background disturbances. Thus, one
linearizes the governing NSE. However, solving a set of 3D linearized perturbation
equation is as difficult as solving the nonlinear equation itself. To circumvent this
problem, early researchers made additional assumption of a parallel flow, where the
growth of underlying shear layer is neglected. Defining the equilibrium flow, only in
terms of wall-normal co-ordinate (y), one can introduce Fourier-Laplace transform
for the perturbation quantities. This was also the basis of deriving Rayleigh’s stability
equation, as mentioned in Chap. 1. However, in this chapter and subsequently, we will
only be discussing about disturbance growth for viscous flows, and in the following
we describe the formulation of the OSE.
If U , V and W are the streamwise (x), spanwise (z) and wall–normal (y) compo-
nents of an equilibrium flow, then under parallel flow approximation,

U = U (y); V = 0; W = W (y)

which will be used in deriving the OSE in this section. The governing equations
for the evolution of the small perturbations in the shear layer are obtained from the
linearized NSE [53], with parallel flow approximation given in component form,
along with equation of continuity by,
3.2 Linear Stability Equation 125

∂u  ∂u  ∂u  dU ∂ p
+U +W + v =− + ν∇ 2 u  (3.6)
∂t ∂x ∂z dy ∂x

∂v ∂v ∂v ∂ p


+U +W =− + ν∇ 2 v (3.7)
∂t ∂x ∂z ∂y

∂w ∂w ∂w dW ∂ p


+U +W + v =− + ν∇ 2 w (3.8)
∂t ∂x ∂z dy ∂z

∂u  ∂v ∂w
+ + =0 (3.9)
∂x ∂y ∂z
   
where, u , v , w and p are the perturbation components of velocity and pressure,
respectively. This implies that the instantaneous quantity (q) is split as,

q(x, y, z, t) = Q(x, y, z) + εq  (x, y, z, t)

with ε as a small parameter representing the quantum of imposed disturbance. In a


linearized study, the response scales directly with ε, and is not considered important
to specify it explicitly. However, for receptivity analysis this has to be specified to
obtain the response of the dynamical system for input disturbance amplitude given
by ε.
Since, U and W components of the mean flow are invariant with either stream-
wise or spanwise coordinate, it is preferred to take the streamwise edge velocity Ue
and the displacement thickness δ ∗ as the velocity and length scales, respectively,
in non-dimensionalizing the perturbation equations. Furthermore, one can express
the velocity and pressure disturbances in terms of Fourier-Laplace transform due
to the adoption of parallel flow approximation, with x and z being treated as the
homogeneous directions. Thus,
    T
[u , v , w , p ] = [ f o ( ỹ), φ( ỹ), h o ( ỹ), πo ( ỹ)]T ei(α x̃+β z̃−ω0 t)

where, x̃ = x/δ ∗ , ỹ = y/δ ∗ and z̃ = z/δ ∗ . Substituting these in the linearized NSE
(Eqs. (3.6)–(3.9)) and further simplification, one obtains the OSE given by [28, 48,
53, 71],

˜ {αU ( ỹ) + βW ( ỹ) − ω0 }(D 2 − γ 2 )φ
(D 2 − γ 2 )2 φ = i Re

−α D 2 U ( ỹ) + β D 2 W ( ỹ)φ (3.10)

˜ = Ue δ ∗ /ν, and γ 2 = α 2 + β 2 . To study stability of the equi-


where, D = ddỹ , Re
librium flow, homogeneous boundary conditions at the wall and free-stream for the
OSE have to be satisfied. For example, at the wall, one uses the no-slip boundary
126 3 Receptivity and Instability

conditions and at the free-stream (i.e., as ỹ → ∞), one requires the disturbance
quantities to decay to zero. The implication of this set of homogeneous boundary
conditions is that the instability is triggered by disturbances within the shear layer,
which is progressively decaying with increase in height. Physically, this implies that
the imposed disturbances within the shear layer has finite energy and hence the dis-
turbances progressively decay with height. Thus, a boundary layer destabilized from
free stream would require different approach of study and would be explained in
greater details while discussing about receptivity of boundary layer. The following
homogeneous boundary conditions are used for the eigenvalue problem associated
with disturbance imposed from within shear layer (as performed experimentally by
near-wall excitation in [50]),

f o (0) = φ(0) = h o (0) = 0 (3.11)


f o ( ỹ), φ( ỹ), h o ( ỹ) → 0 as ỹ → ∞ (3.12)

The non-trivial solutions of Eq. (3.10), subject to homogeneous boundary con-


ditions given by Eqs. (3.11) and (3.12), exist only for particular combinations of
the parameters α, β, ω0 and Re.˜ This defines the dispersion relation of the resultant
eigenvalue problem to be satisfied at the wall as,

˜ =0
D̂(α, β, ω0 , Re) (3.13)

While discussing about receptivity of a boundary layer to wall excitation, we


will note that the dispersion relation originates, so as to satisfy the no-slip bound-
ary conditions at the wall. The common meaning of dispersion relation is to relate
spatial and temporal scales of the disturbance field of an equilibrium flow and can
arise from the governing differential equation of the disturbance field and/or the
associated boundary conditions. It is to be noted that the equilibrium flow can be
time-independent and/ or its space dependence be given by similarity solution. But,
no such assumption is made on the disturbance field. As noted here, the governing
equation for disturbance field in classical instability theory is derived from the NSE
with the small perturbation assumption. While the equilibrium boundary layer can
be obtained via many simplifications, even though the equation is nonlinear.
Historically, first few attempts to solve the above mentioned eigenvalue problem
were made by Heisenberg [28], Tollmien [71] and Schlichting [48]. In the dispersion
relation given by Eq. (3.13), α, β and ω0 can thus be complex quantities, but tradi-
tionally only two approaches have been adopted: (i) temporal amplification theory,
which considers complex ω0 with α and β as real and (ii) spatial amplification theory
that considers α and β to be complex, while keeping ω0 as real. A more elaborate
and detailed discussions on the finer aspects of both these theories and their mutual
relationship can be found in [53]. In the subsequent discussions, only 2D mean flow
subjected to 2D disturbances will be considered for which the governing OSE given
by Eq. (3.10) reduces to,
3.2 Linear Stability Equation 127
 
˜ (αU ( ỹ) − ω0 )(D 2 − α 2 )φ − α(D 2 U )φ
(D 2 − α 2 )2 φ = i Re (3.14)

For the solution of Eq. (3.14) satisfying the homogeneous boundary conditions given
by,
at ỹ = 0 : φ, φ  = 0 (3.15)

and as ỹ → ∞ : φ, φ  → 0 (3.16)

the associated dispersion relation for the 2D disturbance field is,

˜ =0
D̂(α, ω0 , Re) (3.17)

The solution for the 4th-order ODE given by Eq. (3.14) for general excitation can
be represented in terms of four fundamental solutions (φ1 , φ2 , φ3 , φ4 ) as,

φ = a 1 φ1 + a 2 φ2 + a 3 φ3 + a 4 φ4 (3.18)

It will be shown shortly that out of these four modes, two grow with height (say,
φ2 and φ4 ) and the other two decay with height, which we call as φ1 and φ3 . If we
consider wall excitation only, then it is implied that the disturbance field decay in
the free stream ( ỹ → ∞). At the free stream, U = 1 and all mean flow derivatives
with respect to ỹ are zero, simplifying Eq. (3.14) to,

˜
(D 2 − α 2 )2 φ = i Re(α − ω0 )(D 2 − α 2 )φ (3.19)

The special solution of Eq. (3.19) for ỹ → ∞, after removing exponentially grow-
ing terms (by enforcing a2 = a4 = 0) is given as,

φ∞ = a1 e−α ỹ + a3 e−Q ỹ (3.20)



where Q = α 2 + i Re(α ˜ − ω0 ), considering the cases where real parts of α and Q
are positive. It is to be emphasized that the complex wavenumbers, α (inviscid mode
due to independence of R̃e) and Q (viscous mode, that depends on Re) ˜ can have both
positive and negative real part. Apart from [53], other sources in literature do not
explain this aspect of differentiating between wall and free-stream excitations clearly.
As we go along, we will emphasize this aspect of instability and receptivity. Hence,
the general solution of the OSE, subject to the prescribed homogeneous boundary
conditions at the free stream given by Eq. (3.16) can only be in terms of φ1 and φ3
given by,

φ = a 1 φ1 + a 3 φ3 (3.21)
128 3 Receptivity and Instability

From Eq. (3.19), one can easily show the retained fundamental modes obey, φ1∞ =
e−α ỹ and φ3∞ = e−Q ỹ . It is noted that these two retained fundamental solutions
decay exponentially at two different widely disparate rates in the free stream, which
indicates that Eq. (3.14) constitute a stiff ODE. Solution of Eq. (3.14) can not be
obtained in a straight-forward manner due to this stiffness problem. This problem is
avoided by using compound matrix method (CMM), as described in [1, 43, 53] for
external problems.

3.3 Linear Receptivity Analysis of Parallel Boundary Layer

Here, a brief description of receptivity of the parallel shear layer is presented, while
more details of the method are reported in [57, 61, 62]. Schubauer and Skramstad
[50] in their experiment excited the boundary-layer by a vibrating ribbon near the
wall. Such excitation can be idealized by a localized harmonic exciter at the wall,
with perturbation streamwise and wall-normal velocity components given by,

u  (x̃, 0, t) = 0 (3.22)
v (x̃, 0, t) = δ(x̃) ei ω̄0 t H (t) (3.23)

where δ(x̃) is the Dirac-delta function in space, to show the localized nature of
excitation, and H (t) is the Heaviside function, that indicates the finite impulsive
start-up time of the harmonic excitation at the circular frequency ω̄0 , which is a
real quantity. In Fig. 3.2, an equivalent parallel boundary layer at the location of
a harmonic exciter is shown by dotted line. Whereas the actual boundary layer is
shown by the solid line, that starts with a zero thickness at the leading edge of the
plate. In the receptivity theory, two distinct approaches have been followed: (i) Signal
problem is the one where one assumes, a priori, that the response of the system will
be exactly at the frequency of excitation (ω̄0 ) and this approach was introduced in
[52] with the use of a Bromwich contour in the α-plane, and (ii) spatio-temporal
receptivity approach in which one does not fix the response at the forcing frequency
(ω̄0 ) and the spatio-temporal transfer function determines the response of the system
at the complex frequency ω0 , as was pioneered in [57] for the receptivity of Blasius
boundary layer. Both these approaches are also called the Bromwich contour integral
method (BCIM). Based on further results in [61, 62], the second approach is definitely
more meaningful for fluid-dynamical system prone to instability. All of these results
are described together in [53].
The boundary conditions at the wall, for the spatio-temporal approach are given
by Eqs. (3.22) and (3.23); these fix the wall-boundary conditions on φ as,

i
φ(0) = + π δ(ω0 − ω̄) (3.24)
(ω̄0 − ω0 )
φ  (0) = 0 (3.25)
3.3 Linear Receptivity Analysis of Parallel Boundary Layer 129

Fig. 3.2 Harmonic excitation of a parallel boundary layer corresponding to the location of the
exciter

Since the perturbations must decay at the free stream, Eq. (3.21) is applicable and
one can fix the values of the unknown constants, a1 and a3 , from Eqs. (3.24) and
(3.25) as

i
a1 φ1 (0) + a3 φ3 (0) = (3.26)
(ω̄0 − ω0 )
a1 φ1 (0) + a3 φ3 (0) = 0 (3.27)

From Eqs. (3.26) and (3.27), one obtains the general expression of φ( ỹ) from
Eq. (3.21) as
  
i φ3 (0)φ1 ( ỹ) − φ1 (0)φ3 ( ỹ)
φ( ỹ) = (3.28)
(ω̄0 − ω0 ) φ1 (0)φ3 (0) − φ3 (0)φ1 (0)

After obtaining φ( ỹ) from Eq. (3.28), one can get back the perturbation velocity
components by inverse Fourier-Laplace transform as
 2  
1
u  (x̃, ỹ, t) = −iαφ  ( ỹ)ei(α x̃−ω0 t) dα dω0 (3.29)
2π α Br ω0 Br
 2  
1
v (x̃, ỹ, t) = φ( ỹ)ei(α x̃−ω0 t) dα dω0 (3.30)
2π α Br ω0 Br

In Eqs. (3.29) and (3.30), the integration is performed along the Bromwich con-
tours in complex α- and ω0 -planes [53, 74], as shown in Fig. 3.3. The denominator in
the expression for φ( ỹ) in Eq. (3.28) is basically the dispersion relation obtained from
the satisfaction of wall boundary condition, D̂(α, ω0 , Re), ˜ which is also known as
the characteristic determinant of the eigenvalue problem given in Eqs. (3.14)–(3.17).
This establishes the connection between the receptivity and the corresponding sta-
bility problem. Hence, the eigenvalues of the OSE given by Eq. (3.14), which are the
zeros of the characteristic determinant D̂ given by Eq. (3.17), seen in the denomi-
nator of Eq. (3.28). Thus, the eigenvalues are nothing but the poles of the integrand
of φ defining the receptivity problem.
130 3 Receptivity and Instability

(a) α−plane αi

Pα4

Pα1
Pα2 Pα3
αr

αBr

Pα6
Strip of convergence
Pα5

ω0i
(b) ω0−plane

ω0,
Br

ω0r

Pω1
Zone of convergence
Pω2

Fig. 3.3 Bromwich contour along with the region of convergence shown in a α- and b ω0 -plane

Now, since the Laplace transform with respect to x̃ is basically a bilateral Laplace
transform (as x̃ varies from −∞ to +∞), and therefore a strip of convergence for the
Bromwich contour must exist in the α-plane, as shown in Fig. 3.3a [53, 74]. All the
poles lie outside this strip of convergence, and the ones above this strip correspond
to downstream propagating disturbances, while those below the strip of convergence
are for which the disturbances propagate upstream of the exciter.
In contrast, the Laplace transform with respect to t is called the unilateral
Laplace transform, as the integration in t is performed from 0 to +∞. Hence for
the Bromwich contour in ω0 -plane, the convergence of Bromwich contour integral
would be achieved, only if it is placed above all eigenvalues. This is done to satisfy
the causality principle [74]. It has been noted that in α-plane, all the eigenvalues lying
3.3 Linear Receptivity Analysis of Parallel Boundary Layer 131

above the Bromwich contour signify downstream propagating modes (with positive
group velocity, Vg ≥ 0), while the modes lying beneath it would imply upstream
propagating ones (with Vg ≤ 0). These are the criteria followed in [57, 61, 62] to fix
the Bromwich contour in the α-plane, while calculating the receptivity of a parallel
boundary layer in linearized framework. In general a Bromwich contour in α-plane,
can be taken parallel and below the αr -axis. There is also the added advantage for
this choice of contour in applying discrete fast Fourier transform (DFFT) along this
contour, as numerically evaluating the integral given in Eqs. (3.29) and (3.30).

3.3.1 Receptivity of Blasius Boundary Layer

Traditionally, stability analysis is carried out using parallel flow approximation of the
mean flow at any streamwise location of interest, as shown in Fig. 3.2. Parallel flow
approximation for stability analysis of the shear layers is valid only away from the
leading edge. Since at the leading order, boundary layer grows slowly with respect
to x̃ (as at the leading order δ̄, δ ∗ and θ varies with streamwise distance from the
leading edge as x̃ 1/2 ), where similarity transform applies. Without going through
details about solving the OSE by special method like CMM, here we present some
typical results with the physical parameters given.
First, the Blasius boundary layer is obtained by solving Eq. (3.3) for the range
0 ≤ η̄ ≤ 12 using 2400 uniformly distributed points. Subsequently, the OSE is
solved when the exciter is located where the Reynolds number based on displace-
ment thickness (δ ∗ ) is 1500 and the physical frequency is about 3.6 Hz. The solution
is shown in Fig. 3.4 at a non-dimensional time of 801.11 for the streamwise dis-
turbance velocity at a height of ỹ = 0.278. In calculating the physical frequency,
we have assumed the boundary layer has a displacement thickness of 5mm at the
exciter location. Considering the kinematic viscosity of air as 1.5 × 10−5 m2 /s., the
boundary layer edge velocity (Ue ) works out to be equal to 4.5 m/s. For this low

n
0.002 Local sol t = 801.11
0.001
u′

-0.001 Spatio-temporal
n
Asymptotic sol wave-front
-0.002
0 200 ~x 400 600

Fig. 3.4 Streamwise disturbance velocity plotted as function of x̃ at ỹ = 0.278, for the case of
˜ =
harmonic excitation of a parallel boundary layer. The location of the exciter corresponds to Re
1500 and non-dimensional frequency of excitation is ω̄0 = 0.025, with the result shown for the
viscous time scale of t = 801.11
132 3 Receptivity and Instability

frequency excitation, the solution of the OSE provides the streamwise disturbance
velocity component at t = 801.11 in Fig. 3.4, which is referred with respect to the
viscous time scale given by ν/Ue2 = 7.4074 × 10−7 s. If one computes NSE in non-
dimensional form with length scale given by δ ∗ and velocity scale by Ue , then the
corresponding convection time scale is δ ∗ /Ue = 1.11 × 10−3 s. Thus the solution
time of 801.11 shown in Fig. 3.4, is equivalent to the nondimensional time of DNS
of NSE for a non-dimensional time of 0.5342 only! Thus, the solution obtained by
the OSE is extremely finely-resolved time scale, while the DNS of NSE resolves the
time scale quite coarsely. This conversion of different formulations and scales used
for non-dimensionalization is extremely important and should be kept in mind for
future reference.
The solution shown in Fig. 3.4 at t = 801.11 for u  at ỹ = 0.278, is a typical
response for harmonic wall excitation (for 3.6 Hz) at a location where Re ˜ = 1500,
has three distinct components, with the origin for the abscissa is at the localized
delta function exciter. The first component of the solution is called the local solu-
tion, which is seen in the immediate neighbourhood of the exciter. Even though the
˜ cr = 519 for Blasius boundary
exciter is located at a super-critical position (as Re
layer), but the frequency is quite low, for which the linear spatial theory result [53]
indicates stability. This is also supported by the result obtained by linearized recep-
tivity analysis by BCIM. According to Abel’s theorem for any excitation problem
[53], the eigenvalues near the origin of the complex α-plane constitutes the asymp-
totic solution, as marked in the figure. This component of the solution corresponds
mostly to the least damped mode of the eigen-spectrum obtained by spatial stabil-
ity analysis for Re˜ = 1500 and ω̄0 = 0.025. We note the correspondence between
linear stability theory and corresponding linearized receptivity analysis. However,
Fig. 3.4 shows the third component, namely the STWF, which can only be obtained
by spatio-temporal BCIM. This element has been shown in [55, 61, 62], for Blasius
boundary excited at frequencies for which spatial theory indicates stability.
Decaying asymptotic solution and growing STWF separate from each other for
spatially stable cases, while for spatially unstable cases at early times, one cannot
distinguish between asymptotic solution and the STWF, as noted first in [57] for
Re˜ = 1000 and ω̄0 = 0.1. This distinction between spatial unstable and stable cases
are shown in frames (a) and (b) of Fig. 3.5, respectively, for viscous non-dimensional
time of t = 801.11. This is for the case of Re ˜ = 750, for ω̄0 = 0.10 (unstable) and
ω̄0 = 0.15 (stable). For the case shown in frame (a) of Fig. 3.5, the asymptotic
solution and the STWF is fused together, while in frame (b) the asymptotic solution
is damped. For this case, the STWF is noted at the leading edge of the disturbance
packet. In Chap. 5, we will see from DNS results that the STWF keeps growing
along with its higher propagation speed and for the spatially stable case, these two
components of the response field separate from each other.
3.3 Linear Receptivity Analysis of Parallel Boundary Layer 133

u′ 0.005 (a) t = 801.11

-0.005
0 200 ∼x 400 600

0.002
(b) t = 801.11
0.001
u′

-0.001

-0.002
0 200 ∼x 400 600

Fig. 3.5 Streamwise disturbance velocity plotted as function of x̃ at ỹ = 0.278, for the case of
˜ = 750
harmonic excitation of a parallel boundary layer. The location of the exciter corresponds to Re
and non-dimensional frequencies of excitation are a ω̄0 = 0.10 and b ω̄0 = 0.15, with the result
shown for the viscous time scale of t = 801.11

3.3.2 Near-Field of the Localized Excitation: The Local


Solution

Here, we explain the near-field of the response caused by localized delta function
excitation with the help of a qualitative explanation. The details of the mathematical
derivation can be found in [53], and more mathematically inclined readers can find
details there. In this reference, the near-field of the response caused by localized
wall excitation is shown to be due to the essential singularity of the bilateral Laplace
transform of v , i.e., φ. In Fig. 3.6, the straight-line Bromwich contour in the α-plane
is shown along with some representative discrete poles (eigenvalues). Let a circle of
radius R and center at the origin cut this Bromwich contour at the points A and B,
as shown Fig. 3.6. The perimeter of this circular arc above the Bromwich contour is
denoted by Γ and the area bounded by Γ and the Bromwich contour AB encompasses
N number of singularities at P j . One can form a closed contour of integration which
contains AB, Γ and small indented contours around the singularities connected to Γ
by straight lines, as shown in Fig. 3.6. The purpose of these indented contours is to
keep the singularities outside the closed domain. Thus, inside the closed contour, the
integrand is analytic and Cauchy’s integral theorem can be applied for integration of
Φ(α) = φei(α x̃−ω0 t) following the closed contour excluding all singularities. Noting
that ei(α x̃−ω0 t) does not contribute to additional poles or singularities, we can inves-
tigate various contribution arising from φ. As φ(α) is analytic for all points inside
134 3 Receptivity and Instability

αi

R
p2
C1 C3
p1

p3 p4
C4
C2 αr
C5
p5

-∞ +∞
A Bromwich contour B

Fig. 3.6 Bromwich contour in α-plane

the closed contour in the α-plane, and if the singularities shown in Fig. 3.6 denote
simple poles, then invoking Cauchy’s integral theorem [34] one finds that,
  
N
φ(α)dα = − φ(α)dα + 2πi × Residue at (P j ) (3.31)
AB Γ j=1

In the limit of R → ∞, the limits of integration on the left hand side become:
B, A → ±∞ -iαi (with αi indicating the distance between the real α-axis from
the Bromwich contour used), and the integral on the left hand side of Eq. (3.31)
is Bromwich contour integral in the α-plane (Br). Hence, the Bromwich contour
integral of φ(α) can be written as
  
N
φ(α)dα = − lim φ(α)dα + 2πi × Residue at (P j ) (3.32)
Br R→∞ Γ
j=1

where N is the total number of singularities in the α-plane. The first integral on the
right hand side disappears, if one can use the Jordan’s Lemma [34]. However, Jordan’s
lemma is valid only when the degree of the denominator of the integrand φ(α) is at
3.3 Linear Receptivity Analysis of Parallel Boundary Layer 135

least two orders higher than the degree of its numerator, i.e., |φ(α)| ≤ k1 /|α 2 | for
|α| → ∞ [3]. This condition has been specifically investigated in [53], for the limit
of |α| → ∞, by obtaining a closed form analytical solution of the OSE to show that
φ does not satisfy the condition to apply Jordan’s Lemma for |α| → ∞.
Using singular perturbation theory [6], the closed form analytic solution for the
OSE is shown as the inner solution, which is governed by the following reduced
equation as,

φiiv − 2β 2 φi + β 4 φi = 0 (3.33)

where, α = Reiθ = Rβ and φ = φi (Y ). Here, Y is a new independent rescaled


variable given by Y = ỹ/δ̂, with δ̂ is the thickness of the inner solution of OSE
governed by Eq. (3.33). This equation is obtained for the distinguished limit of
˜ and is equiv-
δ̂ = 1/R = 1/|α|. It is to be noted that Eq. (3.33) is independent of Re
˜ → 0) given by [53],
alent to the Stokes problem (in the limit of Re

∇ 4ψ = 0 (3.34)

One can solve Eq. (3.33) for the localized delta function excitation problem [53]
and the part solutions are given as,

φ = (1 + α ỹ)e−α ỹ for αr > 0 (3.35)


φ = (1 − α ỹ)eα ỹ for αr < 0 (3.36)

Clearly, these solutions do not satisfy the necessary condition for Jordan’s lemma
to be satisfied. Hence the first integral on the right hand side of Eq. (3.32) is not iden-
tically equal to zero. It has also been highlighted in [53] that this limit of the integral,
i.e., is the essential singularity of φ (α → ∞) provides non-trivial contribution. This
is very compatible with Tauber’s theorem [53, 74], which states that the near-field
solution (at x̃ → 0) is contributed by the point at infinity in the complex wavenumber
plane, i.e., by the circular arc with R → ∞. The above analytical solution of the OSE
helps us obtain the near-field response field.
Thus in the near field of the exciter, one observes a highly viscous flow and it is for
the same reason the near-field does not penetrate very far upstream and downstream
from the location of the exciter. Most books or monographs, other than [53], does
not even sketch the local solution or talk about its existence.
Performing the Bromwich contour integral, using Eqs. (3.35) and (3.36) over the
semi-circle Γ and collating terms, it can be shown that the solution at the near-field
of the exciter is given by [53],
  
e−i ω̄0 t iei Rz i ỹ
v (x̃, ỹ, t) = lim e−R x̃ cos R ỹ + 1 + R ỹ +
2π R→∞ z z
−i R z̄   Rz+i z 
ie i ỹ ie i ỹ
− 1 + R ỹ − − 1+
z̄ z̄ z z
136 3 Receptivity and Instability
  
ie−R z̄−i z̄ i ỹ
+i R ỹ + ỹ + 1− − i R ỹ + ỹ (3.37)
z̄ z̄

where, z = x̃ + i ỹ and z̄ = x̃ − i ỹ. At the wall, i.e., for ỹ = 0, Eq. (3.37) simplifies
to
  
 e−i ω̄0 t −R x̃ 2 sin R x̃ −R x̃ sin x̃
v (x̃, 0, t) = e − + 2e (3.38)
2π x̃ x̃

In the limit, R → ∞, the first and third terms of Eq. (3.38) do not contribute. But
the second term turns out to be the Dirichlet function, which is an approximation of
the Dirac-delta function, δ(x̃) [74]. This clearly shows that one recovers the applied
delta function at ỹ = 0, which is the imposed boundary condition. Therefore, the
delta function is totally supported by the point at infinity in the wavenumber space
(which is nothing but the circular arc of Fig. 3.6, i.e., the essential singularity of the
kernel of the contour integral).
The analysis in the present subsection clearly indicates that the localized delta
function excitation in the physical space is supported by the essential singularity
(α → ∞) in the image plane. This is made possible because φ( ỹ, α) does not satisfy
the condition required for the satisfaction of Jordan’s lemma. This is also consistent
with the Tauber’s theorem [53], as noted above. Thus, the solution very close to the
exciter is due to the essential singularity of the OSE and called the local solution.
The relevance of the eigenvalues of the OSE, located very close to the origin in the
α-plane, affects the response field far away from the exciter and manifest itself as
traveling waves with appropriate group velocities [53]. This part of the solution is
also given by the second term in the right hand side of Eq. (3.32), which is the total
contribution due to the residues calculated at all the poles.

3.3.3 The Spatio-Temporal Wave-Front (STWF)

To obtain the full receptivity solution to a linearized parallel shear-layer as depicted


in Fig. 3.2, excited by a time-harmonic localized wall excitation started at t = 0 and
located at x̃ = 0, one solves the OSE given by Eq. (3.14) subject to the boundary
conditions given by Eqs. (3.24) and (3.25). To overcome the stiffness problem one
can use the CMM to solve the receptivity problem by BCIM, whose details are in
[53, 57]. From the solution depicted in Figs. 3.4 and 3.5, one can see the genesis
of STWF and we will provide further details on this element of the response field,
while discussing DNS results in Chaps. 5 and 6. We briefly mention in passing here
that a renewed quest for finding STWF for any system by delta function excitation
in time, to look for tsunami-like disturbances caused by localized excitation has
yielded definitive results in [10]. The present receptivity studies are with respect to
time-periodic excitation, which shows the creation of all the three elements. What
3.3 Linear Receptivity Analysis of Parallel Boundary Layer 137

happens, when one excites the system instead by a localized impulse, i.e., by a wall-
normal excitation velocity at the origin at t = 0: v = δ(x)δ(t)? This question has
been addressed in [7, 10], and here one can see the presence of STWF only.

3.4 Stability and Transition of Mixed Convection Flows

So far we have discussed various aspects of instability and receptivity of pure hydro-
dynamic flows, whose linear dynamics can be followed by solving the fourth order
OSE. If one includes any one of the effects due to compressibility, heat transfer or
surface compliance, then the linear dynamics become more involved, as one needs
to solve instead the sixth order OSE [56]. This added complexity also enhances flow
instability, whether one is interested in linear analysis for small excitation amplitudes
or for large amplitude excitations for which one would be required to solve the full
NSE. Effects of heat transfer are studied from both linear and nonlinear instability
perspectives here. Focus is on mixed convection flow at low speed, for which heat
transfer can be modeled by Boussinesq approximation. One of the major difficul-
ties of studying mixed convection flows is a lack of available canonical equilibrium
flows, even for flow past a flat plate with heat transfer. For a constant external flow
past a horizontal flat plate, a similarity solution is available in [49]. This has been
studied in [56, 64] for instability and receptivity of this flow over horizontal flat plate,
while reference is made to flow past a vertical plate also, whose equilibrium flow is
obtained by non-similar approaches.
The spatial stability properties of mixed convection boundary layer, developing
due to a constant external flow over a horizontal plate with heat transfer are studied
here using linear and quasi-parallel flow assumptions, with the aim to find out if
there is a critical buoyancy parameter that highlights the importance of heat transfer
in destabilizing mixed convection flows, with buoyancy effects given by Boussinesq
approximation. The undisturbed flow is the one given by the similarity solution of
[49], which requires the wall temperature to vary as inverse square root of the distance
from the leading edge of the plate for the similarity to hold, when the boundary layer
edge velocity is held constant. Linear stability is investigated first by using the CMM,
which allows finding all modes in a chosen range of the complex wavenumber plane
by spatial stability analysis using grid search technique [53].
Mixed convection flows are important, as these are ubiquitous in nature and in
engineering devices. At the global scale of the atmosphere, geophysical fluid dynam-
ics depend critically on mixed convection flow instability properties. Similarly, heat-
ing and cooling in electronic devices at micro scale depend on mixed convection flow
transition. Instability and receptivity studies of such flows to different types of distur-
bance environments is of importance, but attention is focused here only on vortical
disturbance introduced at the wall for the receptivity study. The present approach of
using vortical excitation shows that weak vortical excitation create thermal fluctua-
tions (entropic disturbances).
138 3 Receptivity and Instability

Heat transfer effects at low speed are often modeled by buoyancy effects in one
of the momentum equations, due to which instabilities in mixed convection flows
differ considerably from instabilities of flows without heat transfer. A consequence
of heat transfer is to induce additional pressure gradient altering the equilibrium flow,
as compared to flows without heat transfer. In [13], it is noted that for natural and
mixed convection flows, instability arises due to the growth of small disturbances.
In this context, the similarity profile in [49] for flow past a horizontal plate becomes
important, as this can be used for mixed convection flows, for constant edge velocity.
For edge velocity varying as x m , generalization is possible for the equilibrium flow,
as given in [42].
Flow and heat transfer properties are more complex for mixed convection flows
past inclined or horizontal plates, due to the buoyancy forces inducing streamwise
pressure gradient altering equilibrium flow. Eckert and Soehngen [20] have shown
flow visualization pictures indicating transition to turbulence for the natural convec-
tion problem, originating as an instability of small disturbances. For natural con-
vection flow past inclined flat plates in [68], the authors have reported generation
of an array of longitudinal vortices. In [37, 78], the authors have experimentally
investigated instability of flow past an inclined plate and reported the presence of
two modes, depending on the inclination angle of the plate. For inclination angles
less than 14◦ with respect to the vertical, the authors reported wave-like instabil-
ity. Such wave-like instability has also been studied in [15]. For inclination angles
greater than 17◦ for the natural convection problem, it is experimentally observed
that the disturbance field is dominated by vortices, and is often termed vortex insta-
bility. This vortex mode of instability is present for horizontal, as well as inclined
plates. This has been variously studied in [25, 26, 40, 67, 75, 76], among many
other studies. Alternate viewpoint in [27], classifies instability of forced-convection
boundary layers over horizontal heated plates in terms of the two prototypical insta-
bilities: Rayleigh–Benard type, usually described for a closed convection system
heated from below, and the Tollmien–Schlichting type that is typical of isothermal
open flows, as in wall-bounded shear layers triggered by viscous actions.
Linear analysis has been traditionally performed using temporal theory, as in [41]
for mixed convection flow for isothermal vertical flat plate. Primary mean flow was
obtained by the local non-similarity method and the authors reported that for assisting
flows, the effect of buoyancy is to stabilize the flow. The critique of various non-
similar flow descriptions used in instability studies should be kept in view [13]. For an
inclined plate, the instability study has been performed in [15] and for the horizontal
isothermal plate in [16]. Results of these indicated that the flow along vertical and
inclined plates is more stable, when the buoyancy force aids external convection,
and the stability decreases, as the inclination angle approaches the horizontal. Also
noted for horizontal plates is the tendency of flow to become more unstable, when
buoyancy force is directed away from the surface.
3.4 Stability and Transition of Mixed Convection Flows 139

Despite the distinction between temporal and spatial methods, the neutral curve
is identical and is of vital interest for instability and transition for low frequency
excitation. In [29], results were reported using linear spatial theory with a parallel
flow approximation for free-convection flow past heated, inclined plates. The spatial
stability results of natural convection flow over inclined plates were reported in [72],
providing the eigenvalue spectrum.
When the Boussinesq approximation is adopted in NSE, effects of buoyancy
appears in terms of Gr/Re2 , where Gr is the Grashof number and Re is the Reynolds
number defined in terms of appropriate length, velocity and temperature scales.
However in [35, 69], authors have shown for boundary layers, that the equivalent
buoyancy parameter changes to K = Gr/Re5/2 . Experimental investigations in [25,
75] have also demonstrated that the onset of instability always occurs at the same
value of K , showing its importance as the relevant buoyancy parameter. Similarity
profiles derived in [42, 49] are also given in terms of K alone. One of the main aims
here is to identify a K cr , beyond which the transport property changes qualitatively
for a mixed convection flow past horizontal plate. In [17], the authors have noted
significant buoyancy effects for K x ≥ 0.05 and K x ≤ −0.03, for aiding and opposing
5/2
flows past a horizontal plate, where K x = Gr x /Rex .

3.5 Governing Equations

We consider laminar, 2D flow past a hot semi-infinite plate, with the free stream
velocity and temperature denoted by U∞ and T∞ , respectively. We focus our attention
on the top of the plate with a temperature distribution Tw (x), which is greater than T∞ ,
while assuming the leading edge of the plate as the stagnation point. The governing
equations are written in dimensional form (indicated by the quantities with asterisks),
along with the Boussinesq approximation, for the velocity and temperature fields [24]


∇∗ · V ∗ = 0 (3.39)


DV ∗ −

=−
→ 1
g j βt (T ∗ − T∞ ) − ∇ ∗ p ∗ + ν ∇ ∗2 V ∗ (3.40)
Dt ∗ ρ

DT ∗ ν q̄
= α ∇ ∗2 T ∗ + Φv + (3.41)
Dt ∗ Cp ρCv

which give the evolution of velocity and temperature fields with space and time. In
the Boussinesq assumption valid for small excursion of temperature, one retains the
density as constant, except that is present in the y-momentum equation, where density
variation is kept to produce buoyancy effects. Here DtD∗ represents the substantial
derivative, g j = [0, g, 0]T represents the gravity vector, βt is the volumetric thermal
expansion coefficient and α is the thermal diffusivity. In this analysis, both the viscous
140 3 Receptivity and Instability

dissipation (Φv ) and heat source terms (q̄) will not be considered in the energy
equation. To non-dimensionalize above equations, a length scale (L), a velocity
scale (U∞ ), a temperature scale (ΔTL = Tw (L) − T∞ ) and a pressure scale (ρU∞ 2
)
are adopted. Instead of prescribing L, a Reynolds number based on L is taken as 105
for the ensuing analysis. The non-dimensional form of these equations are,


∇· V =0 (3.42)


DV Gr 1 2− →
= θ ĵ − ∇ p + ∇ V (3.43)
Dt Re2 Re
Dθ 1
= ∇2θ (3.44)
Dt Re Pr

→ − →∗ ∗
−T∞
, Gr = gβt ΔT , Re = U∞ν L , Pr = αν and ĵ is the unit
3
where V = UV∞ , θ = T ΔT L ν2
LL

vector in the wall-normal direction. The Grashof number (Gr ) gives the ratio of
buoyancy to viscous force in the flow and the Richardson number (Ri) or Archimedes
Gr
number given by, Re 2 , shows the relative dominance of natural to forced convection.

Positive and negative signs of the Richardson number refer to assisting and opposing
flows. In mixed convection regime, Ri is of order one. The Prandtl number (Pr ) used
in the present study is 0.71, a value for air as the working medium. The instability
of boundary layer is also dependent upon buoyancy parameter defined as

Gr 1 −5
K = 5
= g βt [Tw (L) − T∞ ] (Lν) 2 U∞ 2 (3.45)
Re 2

The buoyancy parameter can be defined alternatively in terms of Ri as K = √RiRe . The


value of K = 0 indicates flow over a flat plate without any heat transfer, the Blasius
flow over a flat plate without heat transfer. K > 0 corresponds to assisting flows in
which the flow occurs above a heated flat plate, and K < 0 corresponds to opposing
flows over a cooled flat plate. K switches sign for flow below the corresponding
plates.

3.6 Equilibrium Boundary Layer Flows

For high Reynolds numbers, momentum and energy transport through gradients is
limited to a narrow region inside the boundary layer. The boundary layer approxima-
tion applied to the NSE yields the following equations for 2D incompressible flow
in Cartesian coordinates as,
∂U ∂V
+ =0 (3.46)
∂X ∂Y
3.6 Equilibrium Boundary Layer Flows 141

∂U ∂U ∂P ∂ 2U
U +V =− + (3.47)
∂X ∂Y ∂X ∂Y 2
∂P
0 = Kθ − (3.48)
∂Y

∂θ ∂θ 1 ∂ 2θ
U +V = (3.49)
∂X ∂Y Pr ∂Y 2
√ √ ∗
−T∞
where X = x ∗ /L, Y = y ∗ Re/L, U = u ∗ /U∞ , V = v∗ Re/U∞ , θ = TwT(L)−T ∞
and P = p ∗ /(ρ∞ U∞
2
). The above equations are solved subject to following boundary
conditions,
(i) at the wall, (Y = 0 and X > 0): U = V = 0 and
(ii) at the free stream, (Y → ∞): U = 1, p ∗ = p∞ and θ = 0.
The buoyancy parameter K appears in the Y -momentum equation as a wall-
normal pressure gradient. Integrating Y -momentum equation with respect to Y and
using the boundary condition at Y → ∞: P = 0, one gets the streamwise pressure

gradient as −K Y θ X dY , where the subscript in θ indicates a partial derivative.
Hence depending upon the sign of K , one can create either an adverse or a favorable
streamwise pressure gradient in the boundary layer. It is inappropriate to call these
adverse or favorable pressure gradients – as is customary for flows without heat
transfer based on stability property. Here, this will be only evident from the computed
stability solutions. For the present investigation of flow over a heated flat plate, K
is positive, with buoyancy resulting in a pressure gradient accelerating the flow in
the streamwise direction. This term, therefore, usually induces non-similarity in the
governing equations.

3.6.1 Schneider’s Similarity Solution

A similarity solution of the boundary layer equation derived above has been given
by Schneider [49]. Introducing the stream function (Ψ ) automatically satisfies the
continuity equation. Substituting the expression for streamwise pressure gradient in
the X -momentum equation and the energy equation, we get
 ∞
ΨY Ψ X Y − Ψ X ΨY Y − K θ X dY = ΨY Y Y (3.50)
Y

1
ΨY θ X − Ψ X θ Y = θY Y (3.51)
Pr
where subscripts X and Y denote partial derivatives with respect to X and Y , respec-
tively.
142 3 Receptivity and Instability

For the above system of equations to admit a similarity solution, the wall tempera-

ture distribution must be given by θw ∝ X −1/2 , where θw = TTww(x(L)−T
)−T∞

. This similarity
transform converts the above PDEs in X and Y , into an ODE for the similarity vari-
able defined by, η = Y X −1/2 . Transforming dependent variables as, Ψ = X 1/2 f (η)
and θ = θw Θ(η), yield the following governing equations

2 f  + f f  + K ηΘ = 0 (3.52)

2 
Θ + f Θ  + f Θ = 0 (3.53)
Pr
In the above, a prime indicates a derivative with respect to η, and these equations
have to be solved subject to the following boundary conditions at η = 0: f = f  =
0 and Θ = 1 and as η → ∞: f  = 1 and Θ → 0. Equation (3.53) is integrated
analytically once to obtain the following equation

2 
Θ + fΘ = 0 (3.54)
Pr
Thus, the similarity profile depends only on K , with dependence on Re implicit
through the definition of the Y -coordinate and η. It is seen from Eq. (3.54) and the
boundary condition at η = 0: f = 0 that irrespective of K , adiabatic condition exists
all over the plate, with heat transfer occurring singularly at the leading edge only.
The similarity solution for the problem under consideration is given for f  , f  ,
Θ and Θ  as functions of η in frames (a) to (d) of Figs. 3.7, 3.8 and 3.9, for different
values of K . Here f  represents the streamwise velocity, f  its second derivative, Θ
represents the temperature, and Θ  its wall-normal derivative. Figure 3.7 shows the
result for a low value of K = 1 × 10−6 . In Fig. 3.7a, the velocity profile monotoni-
cally grows following the boundary condition at η = 0 to η = ηmax , where it tends
to unity. Similarly from Fig. 3.7c, the temperature profile is seen to reach the free
stream condition, where it is zero according to the non-dimensionalization adopted.
In Fig. 3.7b, the inset provides a detailed view of the portion enclosed, helping detect
any existing inflection points. For this K , the result indicates no inflection point to be
present. Figure 3.8 shows the mean flow quantities plotted for K = 3 × 10−3 . The
general trend for the variation of mean flow quantities remains the same. But Fig. 3.8b
indicates an inflection point for η  9. The same trend is observed in Fig. 3.9, for
K = 9 × 10−2 , which has a greater heat transfer due to buoyancy effects. In addition,
we see an overshoot in the velocity profile in Fig. 3.9a. It has been noted in Fig. 3.8
of [49] that high values of K result in a velocity overshoot within the boundary layer.
These observations tell us that for low values of K , although the second derivative of
streamwise velocity remains zero at the wall and at the free stream, enhanced heat-
ing of the plate surface causes an inflection point at an intermediate height. From
Figs. 3.7, 3.8 and 3.9, it is seen that as K is increased, variation of mean parameters
progressively occurs within a lower range of η, which means that the mean flow quan-
tities and their gradients achieve free stream conditions, more rapidly with higher
3.6 Equilibrium Boundary Layer Flows 143

-
-
- -

Fig. 3.7 Variation of mean flow quantities obtained for K = 1 × 10−6

gradients. Increasing K , i.e., increasing the temperature difference between the plate
surface and the free stream, the boundary layer shrinks and the point of inflection
moves closer to the wall. A similar overshoot of velocity within the boundary layer
has been reported in [42] for wedge flow.
From Eq. (3.48), it is understood that the effect of heat transfer is to introduce a
wall-normal pressure variation within the boundary layer. This is contrary to flows
without heat transfer, where pressure is impressed upon the boundary layer by the
outer inviscid flow, which remains invariant with height within the boundary layer.
This aspect sets apart mixed convection boundary layers from isothermal flows. The
wall-normal pressure variation within the boundary layer modifies the streamwise
pressure gradient in the X -momentum equation given by
144 3 Receptivity and Instability

-
-
-
-
-

Fig. 3.8 Variation of mean flow quantities obtained for K = 3 × 10−3

 ∞
−K θ X dY
Y

We also have
θ = θw Θ = X −1/2 Θ, η = Y X −1/2

Here the temperature gradient is with respect to X (keeping Y fixed), i.e., given
∂θ
by | . With wall temperature varying as X −1/2 , this would be
∂X Y
 
∂ 1 ∂Θ ∂η
X −1/2 Θ(η) = − X −3/2 Θ + X −1/2
∂X 2 ∂η ∂ X

As
∂η η
=−
∂X 2X
3.6 Equilibrium Boundary Layer Flows 145

Fig. 3.9 Variation of mean flow quantities obtained for K = 9 × 10−2

Thus
∂θ 1
|Y = − 3/2 [Θ + η Θ  ]
∂X 2X
Therefore, the induced streamwise pressure gradient is given by
 ∞  ∞ 
K 
−K θ X dY = − 3/2 (Θ + η Θ )X 1/2

Y 2X η

  
K
=− Θdη + ηΘ − Θdη
2X

Using above relations and integrating the pressure gradient term, we get the following
form for the streamwise pressure gradient
146 3 Receptivity and Instability

Fig. 3.10 Buoyancy-induced pressure gradient term K ηΘ plotted as a function of η

 ∞  ∞
K K
−K θ X dY = − ηΘ =− ηΘ (3.55)
Y 2X η 2X

The boundary condition Θ → 0 for η → ∞ and Eq. (3.55) show that buoyancy
effects do not impose a pressure gradient at the wall and at the far stream. But
at all intermediate heights within the boundary layer, there is a height-dependent
streamwise pressure gradient. For heated plates, K has a positive value and hence
the pressure gradient within the mixed convection boundary layer is negative, which
accelerates the flow. The value K ηΘ determines the magnitude of this pressure
gradient at any given streamwise location, which is plotted as a function of η for
various buoyancy parameter (K ) values in Fig. 3.10, certain aspects of which are
noteworthy. The location of the maximum pressure gradient within the boundary
layer moves towards the plate, as K is increased. Also, its magnitude is of the same
order, as that of the corresponding K . The maximum value of favorable pressure
gradient along with the location at which it occurs (η f pg ) are given in Table 3.1.
3.6 Equilibrium Boundary Layer Flows 147

Table 3.1 Streamwise pressure gradient induced by buoyancy effects for a heated flat plate for
different buoyancy parameters
Case K η f pg Maximum pressure gradient
Parameter (K ηΘ)
1 1 × 10−6 2.62250 1.86443 × 10−6
2 1 × 10−3 2.61816 1.86144 × 10−3
3 3 × 10−3 2.61000 5.55667 × 10−3
4 3 × 10−2 2.52144 5.36419 × 10−2
5 9 × 10−2 2.38344 1.51597 × 10−2

In flows without heat transfer, it is well known that a favorable pressure gradient
stabilizes the flow, which is impressed upon the boundary layer by the outer inviscid
flow. Pressure remains invariant inside the shear layer in the wall-normal direction.
But for flows with heat transfer, we note the presence of a differential streamwise
pressure gradient within the boundary layer. The effect of this favorable pressure
gradient is to accelerate the flow differentially within the shear layer, which results
in a velocity overshoot, responsible for the presence of an inflection point. This is
clearly dependent on the value of K . The higher the value of K , the higher will be the
magnitude of the velocity overshoot. Also, Fig. 3.10 and Table 3.1 show that at higher
values of K , the maximum of the pressure gradient moves closer to the plate surface.
Figure 3.11 shows the contours of pressure gradient expression as given in Eq. (3.55)
for values of K under discussion in the (x, y)-plane (x = x ∗ /L and y = y ∗ /L). The
η level at which the maximum pressure gradient occurs is also marked by a dotted
line in Fig. 3.11.
The existence of an inflection point for the mean flow provides the necessary con-
dition for inviscid instability, according to the Rayleigh–Fj∅rtoft theorem. Therefore,
we are dealing with an equilibrium flow, which has an inherent tendency towards
inviscid temporal instability. This is in stark contrast to flows without heat transfer,
where phenomenologically one observed spatial growth of disturbances. Thus, one
notes the propensity of the flows with heat transfer to display tendencies of both
spatial and temporal linear instabilities.
Figure 3.12 shows variation of the inflection point with respect to K , indicating
the inflection point to move closer to the wall, as K is increased. It helps to conclude
that the mean flow becomes more and more unstable as the wall is heated, due to
the presence of an inflection point nearer to the wall. However, the quantification of
this should not be done using Rayleigh’s stability equation derived for flows without
heat transfer.
148 3 Receptivity and Instability

Fig. 3.11 (−K ηΘ/2X )− contours plotted in the (x, y)-plane for the indicated values of K . η f pg
is marked by a dotted line

3.6.2 Ambiguities of Spatial and Temporal Linear Theories:


Example of Mixed Convection Problem

Receptivity studies require computations of equilibrium flow, and whose response


to deterministic excitation is sought. Wall-excitation for a ZPG boundary layers by
DNS have been studied in [9, 55, 65]. The key features of this is that the same
methodology is used to compute equilibrium and disturbance field, for different
boundary conditions. When same methodology is used for mixed convection flows
past horizontal plate, with Boussinesq approximation to model heat transfer effects
3.6 Equilibrium Boundary Layer Flows 149

14

12

10
ηip

2
10-6 10 -5 10-4 10-3 10-2 10 -1 10 0
K

Fig. 3.12 Location of inflection point (ηi p ) plotted as a function of K

in [66], even the equilibrium flows could not be computed. For horizontal hot flat
plate with adiabatic wall conditions, the equilibrium flow is computed and its recep-
tivity correlates with linear spatial theory for lower buoyancy parameter. However,
receptivity of mixed convection flows to wall excitation for the following cases, did
not allow computing equilibrium flows by DNS: (i) Adiabatic horizontal flat plate
cooled significantly at the leading edge and (ii) strongly heated isothermal wedge
flow for an angle of 60◦ . The cold plate case is interesting, as we note in next section
that the linear spatial theory indicates enhanced stabilization with cooling for higher
magnitude of K i , instead for the cold plate, one notices disturbance growth outside
the shear layer. The authors in [66] re-investigated various mechanisms of instability
present for mixed convection flows, explaining relative roles of viscous and inviscid
mechanisms for strong heat transfer effects, within the Boussinesq approximation.
The main issue highlighted is: What happens to a flow, if both linear temporal and
spatial theory indicate instability, with growth/ decay rates completely different? This
ambiguity can only be resolved either by adopting linear spatio-temporal receptivity
analysis, or nonlinear, nonparallel analysis by DNS. We emphasize that DNS with
high accuracy scheme is preferred in such cases. New theorems of inviscid instability
are proposed [66], which are more generic than Rayleigh’s and Fjφrtoft’s theorems
for hydrodynamic instability without heat transfer. Thus, inviscid mechanism has
been stated for mixed convection flows by replacing Rayleigh’s equation [66].
150 3 Receptivity and Instability

Local instability studies have often been supplemented by linear and nonlin-
ear global studies, as in [18, 39], with flow instabilities classified as resonator and
amplifier type. It is noted that though linear local theories can account for amplifier
behavior, yet it is preferable to develop a fully nonlinear formulation involving the
presence of a front separating the base state region from the bifurcated state region
[18].
The resonator dynamics imply the ability of flow to self-sustain oscillations, which
is due to flow being globally unstable showing 3D stationary mode(s) [39]. For sep-
arated flows, this exhibits absolute instability, i.e., the disturbances do not convect
and grow with time. In contrast, amplifier dynamics show the flow displaying large
transient amplification of initial perturbation, which is often followed by convective
instability. For flow over a backward facing step, Blackburn et al. [11] noted that their
earlier computations reported absolute instability [4], while the flow is actually unsta-
ble at much lower Reynolds numbers to convective instability shown by investigating
directly the linear convective instability by means of transient-growth computations.
This highlights the central role played by accuracy of numerical methods used in
computing NSE, to demarcate convective and absolute instabilities correctly. For a
choice of time step, some wavenumbers can show spurious dispersion, while some
specific wavenumber may indicate absolute instability (with numerical group veloc-
ity being zero). While spectral element method is known for its accuracy, in [31]
a self-supported oscillation for a flow over 2D backward facing step was reported
for sub-critical Re, which was not found using a better resolved DNS [32]. This
problem shows the need to calibrate any method with model equation before actual
simulation. In [66], different local instability mechanisms in the mixed convection
flow over a flat plate and wedges are explained.
Receptivity is studied for those equilibrium flows, which are usually steady in
the mean, and often defined by similarity profile(s). For flows without heat transfer,
Blasius profile is for the canonical ZPG flow, whose instability has been studied
extensively [2, 12, 22, 47]. We also note that use of Blasius profile is misleading for
instability studies, as one excludes region near the leading edge of the flat plate, where
this is not an accurate representation of the equilibrium flow. It has been conclusively
shown in [60] that the leading edge is a site of disturbance energy, which creates
convected disturbances that remain outside the boundary-layer. This is known as
shear sheltering, which affects secondary instabilities and bypass transition over the
flat plate by free stream modes and these outer disturbances interfere with growing
disturbances inside the boundary layer during secondary stages. Apart from this, one
also needs to incorporate rapid growth of the boundary layer, near the leading edge of
the plate/ wedge. Such nonparallel and nonlinear effects change stability properties
of flow without heat transfer [58], and for flow with heat transfer having adiabatic
wall condition [53]. For these reasons, in [55, 58, 65], equilibrium flow has been
calculated by solving unsteady NSE, including the leading edge of the flat plate.
Identical approach is used in computing equilibrium and disturbance flows. We also
show that for study of mixed convection flows, there is truly no similarity profiles
that can be used as equilibrium flows.
3.7 Equilibrium Solution for Mixed Convection Flows: Isothermal Wall Case 151

3.7 Equilibrium Solution for Mixed Convection Flows:


Isothermal Wall Case

In [42] the formulation used in [49] has been generalized for flow past isothermal
wedge. This flow does not exhibit singular heat transfer at the leading edge. The
governing equation transforms to ordinary differential equation for the external flow
given by Ue  U∞ X n , with the new independent variable, η = Y X (n−1)/2 . The wall
temperature distribution is given by θw = X (5n−1)/2 . The choice of n = 1/5 provides
a wall-temperature distribution which is independent of X , which corresponds to
a wedge angle of 60◦ . Despite similarity, flow and heat transfer at the wall are
completely different in [42], as compared to that in [49]. However, as the boundary
layer edge velocity in [42] is a function of X , none of these velocity profiles directly
represents similar solution. Both these flows are considered to study spatial and
temporal viscous instabilities by solving NSE with different heat transfer at the wall,
and help identify the active instability mechanisms for mixed convection flows.
Heat transfer modeled by Boussinesq approximation for mixed convection flows
induces pressure gradient of the equilibrium flows. Such mean flows display flow
instabilities, including inviscid instability, similar to that given by Rayleigh’s equa-
tion for flows without heat transfer [19, 53]. This is shown here and new theorems
stated with necessary conditions for temporal instability by linear inviscid mecha-
nism. Also DNS of flows with heat transfer are provided, which show viscous and
inviscid mechanisms simultaneously. In such a scenario, it is essential that we show
predominance of one mechanism over the other.
The wedge flow given in [42] is a general equilibrium flow. In defining the gov-
erning equations for mixed convection flows, a buoyancy parameter is introduced as
G 0 in [42] and K in [49], these symbols will be used here as K i and K a , respectively,
with the subscript indicating isothermal and adiabatic conditions for flow over the
general wedge flow defined as,
Gr − 25
= g βt [Tw (x ∗ ) − T∞ ] (x ∗ ν) 2 Ue
1
K i or K a = 5
Re 2

where, Gr is the Grashof number. K a = 0 corresponds to flow without heat transfer,


which represents Falkner–Skan flow for the case of [42] and Blasius flow for the case
of [49]. The cases of K a > 0 correspond to assisting flows, occurring above a heated
plate and the cases with K a < 0 correspond to opposing flows, occurring over the
top surface of a cold plate.

3.7.1 Governing Equation for Flow over Isothermal Wall

The 2D steady Navier Stokes and energy equations for incompressible mixed con-
vection flow with Boussinesq approximation are written for a boundary layer, as in
152 3 Receptivity and Instability


−T∞
Eqs. (3.46)–(3.49) for Schneider’s similarity profile with θ = TwT(L)−T ∞
. The first
term on the right hand side of Eq. (3.48) is due to the free stream pressure gradient
acting on the boundary layer by the edge velocity Ue (X ). The second term on right
hand side of Eq. (3.47) represents the buoyancy-induced height dependent pressure
gradient inside the shear layer. Depending upon the sign of K a , this term can create
either an adverse or favorable pressure gradient upon the boundary layer. Equation
(3.48) shows these flows to display buoyancy effect by imposing a wall-normal vari-

ation of streamwise pressure gradient [53] by K a Y θ X dY . For flows with negative
K a , the cases in [49] have induced favorable pressure gradient. This discussion is
also valid for the isothermal wedge cases, with K a replaced by K i .
Equations (3.46)–(3.49) are solved for wedge flow subject to the boundary con-
ditions:

At the wall (Y = 0 and X > 0) : U = V = 0,

And at the free stream (Y → ∞) : U = Ue (X ) and θ = 0.

These are further transformed to ODEs using a independent variable, η =


Y X (n−1)/2 . We obtain equilibrium solution, following the procedures in [42, 49],
with the wall temperature given by θw = X (5n−1)/2 .
We introduce new variables for velocity, temperature and pressure as,

U = Ue f  (η), θ = X (5n−1)/2 Θ(η), P = X 2n q(η) (3.56)

Introduction of non-dimensional stream function, f (η), automatically satisfies


Eq. (3.46). Equations (3.47)–(3.49) are transformed in terms of the new variables as,

1 1
f  = n( f 2 − 1) − (n + 1) f f  + 2n K i q + (n − 1)K i η q  (3.57)
2 2

q = Θ (3.58)

1
Θ  = Pr [(5n − 1)Θ f  − (n + 1) f Θ  ] (3.59)
2
which are solved subject to the boundary conditions:

At η = 0 : f = f  = 0 and Θ = 1

As η → ∞ : f  = 1 and q = Θ = 0

These equations are for mixed convection flow over a wedge, with the wedge
angle: 2γ = 2nπ/(n + 1). Derivation of these are given in Appendix A. Schneider’s
case [49] is obtained by putting n = 0, which is for mixed convection flow over a
3.7 Equilibrium Solution for Mixed Convection Flows: Isothermal Wall Case 153

horizontal plate, with singular heat transfer at the leading edge. The wall temperature
here is obtained from Eq. (3.56) by θw ∼ X −1/2 .
Also, using wall boundary condition in the relation of Eq. (3.56), we note that
for n = 0.2, θw becomes constant with respect to X – a mixed convection flow over
an isothermal wedge with the half-wedge angle (γ ) as 30◦ in Fig. 3.13. For n = 0,
we have mixed convection flow over a flat plate, and by integrating Eq. (3.59), we
obtain, Θ  = − Pr2
f Θ. With the wall boundary condition, one notes that Θw = 0 at
all X , for any K a , as f = 0 for η = 0 for the case of [49]. Thus, all heat transfer
occurs singularly from the leading edge of the horizontal flat plate.

3.7.2 Mixed Convection Governing Equations and Boundary


Conditions for DNS

Study of nonlinear receptivity for mixed convection flows, starts with an equilibrium
flow obtained by solving the conservation equations in (ψ, ω)-formulation, along
with Boussinesq approximation. The physical plane is defined by non-dimensional
Cartesian co-ordinate, for which the same reference length (L) is used to obtain X
in Eqs. (3.46)–(3.49).
√ However, Y in the same set of equations and y ∗ are related as
Y = y ∗ Re for the boundary layer equation. Here, the equilibrium and disturbance
flows are obtained from DNS of the NSE and energy equation. This is performed
in the transformed (ξ, ζ )-plane. A schematic diagram of the problem is shown in
Fig. 3.13. In the transformed plane, vorticity transport, stream-function and energy
equations are obtained as,

y
Inflow
Ue = U∞ (cos γ + x ) Far-field boundary y = ymax
n
Outflow
Momentum boundary Thermal
U∞,θ∞ layer boundary layer
Exciter
γ xin θw = x(5n-1)/2 δ δθ
x
Inflow condn. x1 x2 xm wall xout

Fig. 3.13 Schematic diagram for mixed convection flow over a wedge. Identical computational
domains are used for computing equilibrium flow and its receptivity to wall–excitation. [Reproduced
from Direct numerical simulation of transitional mixed convection flows: Viscous and inviscid
instability mechanism. Sengupta et al., Physics of Fluids, 25, 094102 (2013), with the permission
of AIP Publishing.]
154 3 Receptivity and Instability
    
∂ω ∂ω ∂ω 1 ∂ h 2 ∂ω ∂ h 1 ∂ω
h1h2 + h2u + h1v = +
∂t ∂ξ ∂ζ Re L ∂ξ h 1 ∂ξ ∂ζ h 2 ∂ζ

+ K a Re L (h 2 θ ) (3.60)
∂ξ
   
∂ h 2 ∂ψ ∂ h 1 ∂ψ
+ = −h 1 h 2 ω (3.61)
∂ξ h 1 ∂ξ ∂ζ h 2 ∂ζ
    
∂θ ∂θ ∂θ 1 ∂ h 2 ∂θ ∂ h 1 ∂θ
h1h2 + h2u + h1v = + (3.62)
∂t ∂ξ ∂ζ Re L Pr ∂ξ h 1 ∂ξ ∂ζ h 2 ∂ζ

In the transformed plane, the contra-variant components of the velocity vector are
given by,
1 ∂ψ 1 ∂ψ
u= and v = −
h 2 ∂ζ h 1 ∂ξ
1
The scale factors of transformation h 1 and h 2 , are given by, h 1 = (xξ2 + yξ2 ) 2 and
1
h 2 = (xζ2 + yζ2 ) 2 . Here, ξ is in the direction along the wall and ζ is in the wall-normal
direction. Thus, if the transformations are given by, x = x(ξ ) and y = y(ζ ), then the
scale factors are, h 1 = xξ and h 2 = yζ . For the presented results, the parameters Re
and Pr are selected as 105 and 0.71, respectively.

3.7.2.1 Auxiliary Conditions

For the reported receptivity to wall excitation, Eqs. (3.60)–(3.62) are solved for the
equilibrium and disturbance flows. Dependent variables are split into equilibrium
(denoted by overbar) and disturbance component (indicated by subscript d). For
example, the vorticity is represented as ω = ω̄ + ωd . For the equilibrium flow, free
stream conditions are given at the top of the domain for all ξ given by,

∂ ψ̄
= Ue h 2 , ω̄ = 0 and θ̄ = 0 (3.63)
∂ζ

where Ue = U∞ cos γ for x ≤ 0 and Ue = U∞ (cos γ + x n ) for x > 0, with γ , as


the semi-wedge angle. The no-slip and impervious conditions are used at the wall
as the boundary conditions, which provides ψ and ω for the equilibrium flow. Thus,
for x ≥ 0, the boundary conditions at y = 0 are

ψ̄w = ψ̄o = constant (3.64)

1 ∂ 2 ψ̄
ω̄w = − (3.65)
h 22 ∂ζ 2
3.7 Equilibrium Solution for Mixed Convection Flows: Isothermal Wall Case 155

θ̄w = x (5n−1)/2 (3.66)

Ahead of the leading edge of the wedge (x < 0), Eqs. (3.65) and (3.66) remain
same and Eq. (3.64) changes as,

ψ̄ = ψ̄o + U∞ x sin γ (3.67)

The wall vorticity is obtained by noting ψ as a function of ζ and using the no-slip
condition to obtain the derivative in Eq. (3.65). The wall vorticity has been obtained
similarly in [53–55, 58, 65].
For equilibrium and disturbance field at the outflow, boundary conditions are used
as,
∂v
=0 (3.68)
∂x
∂ω ∂ω
+ Uc =0 (3.69)
∂t ∂x
∂θ ∂θ
+ Uc =0 (3.70)
∂t ∂x
The last two equations are the Sommerfeld outflow conditions for vorticity and
temperature, respectively. This type of boundary condition is quite common and used
in [12, 22, 55, 65]. This helps enforce the disturbance to convect smoothly out of the
domain with the velocity, Uc . The convection speed is taken as Ue , as also used in
[55, 60, 65]. One can also use filters to avoid numerical instability near the outflow,
as described in Chap. 2.
To start the simulation for equilibrium flow, properties are initialized in the interior
of the domain with potential flow solution, using the following initial conditions:

∂ ψ̄
= Ue h 2 (3.71)
∂ζ

ω̄ = 0 (3.72)

θ̄ = 0 (3.73)

The computational domain is taken as, −0.05 ≤ x ≤ 80; 0 ≤ y ≤ 2. In the wall-


normal direction, a stretched tangent hyperbolic grid is used, given by the following
for the jth point as  
tanh[b y (1 − ζ j )]
y j = ymax 1 −
tanh b y

Such stretching helps in controlling aliasing error [54]. Here, ymax = 2 and b y = 2
have been used to cluster grid and a total of 501 points are taken in the wall-normal
direction. In the streamwise direction, we have two segments: the first segment from
156 3 Receptivity and Instability

xin = −0.05 to xm = 10, contains stretched grid distributed by tangent hyperbolic


function, with the ith point obtained as,
 
tanh[bx (1 − ξi )]
xi = xin + (xm − xin ) 1 −
tanh bx

From xm = 10 to xout = 80, uniform distribution of points are used, so that 4501
points are used in the streamwise direction. The wall resolution used is, Δyw =
4.36 × 10−4 and the grid spacing in the streamwise direction at the exciter location
is taken as Δxexc = 0.002, which is more than adequate with the high accuracy
compact schemes used.
The diffusion operator (∇ 2 ) in stream function equation is discretized using sec-
ond order central differencing and is solved by Bi-CGSTAB algorithm [77]. First
derivatives in vorticity transport and energy equations are obtained using the OUCS3

(i) (ii)
1 0

0.8

0.6 -0.05
f′

−2
Ka = −10
f′′′

0.4 Ka = −10
−3

−2
0.2 -0.1 Ka = −10
−3
Ka = −10
0
0 1 2 3 4 5 6 0 1 2 3 4 5 6 7 8
η η
(iii) 0 (iv)
1
Ka = −10−2
0.8 Ka = −10−3
-0.1
0.6
θ′
θ

0.4 -0.2 −2
Ka = −10
−3
0.2 Ka = −10

0 -0.3
0 1 2 3 4 5 6 7 0 2 4 6 8 10
η η

Fig. 3.14 a Variation of mean flow quantities obtained from boundary layer solution for flow
past an adiabatically cooled horizontal flat plate buoyancy parameter K a = −0.001 and −0.01. b
Streamwise velocity component and temperature obtained from DNS for flow past an adiabatically
cooled plate K a = −0.001 and −0.01 plotted as a function of the similarity variable η at indicated
x-locations along with the corresponding similarity solution. [Reproduced from Direct numerical
simulation of transitional mixed convection flows: Viscous and inviscid instability mechanism.
Sengupta et al., Physics of Fluids, 25, 094102 (2013), with the permission of AIP Publishing.]
3.7 Equilibrium Solution for Mixed Convection Flows: Isothermal Wall Case 157

−3
(i) Ka = −10 ; t = 220
1

0.8
θ f′ Schneider’s soln
f′;θ

0.6 x = 9.27
x = 29.47
0.4 x = 49.65

0.2

0
0 1 2 3 η 4 5 6 7 8

−2
(ii) Ka = −10 ; t = 250
1

0.8
θ f′ Schneider’s soln
f′;θ

0.6 x = 9.27
x = 29.47
0.4 x = 49.65

0.2

0
0 1 2 3 η 4 5 6 7 8

Fig. 3.14 (continued)

scheme [63]. Optimized 3-stage Runge-Kutta method (ORK3) [45] has been used
for time integration. A very narrow buffer domain has been used at the outflow
(79.5 ≤ x ≤ 80), where second order 1D filter is used in both x- and y-directions,
that prevents reflections from the outflow boundary. Such buffer domain has been
used in [12, 33, 36, 59, 70] for the solution of NSE.
Solutions of Eqs. (3.57)–(3.59) are obtained for K a = −0.01 and −0.001, as
shown in Fig. 3.14a for the velocity profile ( f  ), curvature of the velocity profile ( f  ),
temperature (Θ) and heat flux (Θ  ). These cold plate cases have also been obtained
by solving NSE given in Eqs. (3.60)–(3.62), with the help of boundary and initial
conditions defined. Results from NSE are compared in Fig. 3.14a with the solution
obtained using the formulation of [49], for different streamwise stations. Results of
NSE match very well with Schneider’s profile. Temperature profiles show mismatch
due to the singularity at x = 0 for the wall temperature used in the formulation of
[49]. Similar results for hot plate cases have been shown in [53], which showed better
match between DNS and the solution of Eqs. (3.57)–(3.59).
158 3 Receptivity and Instability

(i) (ii)
1

0
0.8
3E-06
-0.02 2E-06
0.6

f ″′
1E-06
f′

f ″′
0
0.4 Adiabatic wall condition -0.04 -1E-06
with Ka = -0.05 0 0.01 η 0.02 0.03

0.2 -0.06

0
0 2 4 6 8 10 0 2 4 6 8 10
η η

(iii) (iv)
1 0

0.8

-0.1
0.6
Θ′
Θ

0.4
-0.2
0.2

0 -0.3
0 2 4 6 8 10 0 2 4 6 8 10
η η

Fig. 3.15 a Variation of mean flow quantities obtained from similarity solution for the case of flow
over adiabatic flat plate with buoyancy parameter K a = −0.05. b Variation of mean flow quantities
obtained from similarity solution for the case of flow over isothermal wedge with buoyancy param-
eter K i = 0.04. [Reproduced from Direct numerical simulation of transitional mixed convection
flows: Viscous and inviscid instability mechanism. Sengupta et al., Physics of Fluids, 25, 094102
(2013), with the permission of AIP Publishing.]

The DNS reproduces the boundary layer solution [49] to fair degree of accuracy,
in Fig. 3.14b. However for higher degree of cooling cases, the DNS does not even
produce steady solution. One such cases of adiabatic cold plate is shown in Fig. 3.15a
for K a = −0.05 for the solution of Eqs. (3.57)–(3.59). In frame (ii) of Fig. 3.15b, one
notices existence of a inflection point inside the boundary layer at η ≈ 1.5. Drawing
analogy with flow without heat transfer in [53], existence of inflection point was
identified as an indicator for inviscid instability following Rayleigh’s theorem [19,
53]. Another case of heated isothermal plate results are shown in Fig. 3.15b for
K i = 0.04, obtained by solving Eqs. (3.57)–(3.59). One notices an inflection point
distinctly at the outer inviscid part of the flow (η ≈ 6.3).
3.7 Equilibrium Solution for Mixed Convection Flows: Isothermal Wall Case 159

(i) (ii)
1 0

0.8 -0.05
4E-05

0.6 -0.1 2E-05

f ″′
f ′

f ″′
0
0.4 -0.15
Isothermal heated wedge -2E-05

case with Ki = 0.04 -4E-05


0.2 -0.2
5 6 7 8 9 10
η
0 -0.25
0 2 4 6 8 0 2 4 6 8
η η
(iii) (iv)
1 0

0.8
-0.1
Θ′

0.6
Θ

-0.2
0.4

0.2 -0.3

0
0 2 4 6 8 0 2 4 6 8
η η
Fig. 3.15 (continued)

Computed solution of Eqs. (3.60)–(3.62) are shown in Fig. 3.16, for the indicated
time frames corresponding to the case shown in Fig. 3.15a of adiabatic cold plate
with K a = −0.05. Vorticity contours shown up to t = 90, display accumulating
fluctuations near the leading edge, which is convected downstream along the edge of
the boundary layer. Fluctuations also build up progressively inside the boundary layer
simultaneously. Despite the inflection point being deep inside the boundary layer,
the fluctuations appear initially outside the shear layer. Such behavior is reminiscent
of disturbance energy growth mechanism explained in [60] for bypass transition via
vortex-induced instability. There the disturbances originated from the leading edge
for a ZPG flow without heat transfer. This justifies solving the full NSE including
the leading edge of the plate to investigate mechanisms of instability.
Solution of NSE, as the vorticity contours are shown for the isothermal hot wedge
case of Fig. 3.15b (K i = 0.04) in Fig. 3.17. Here also, one notices unsteady fluctua-
tions near the outer edges of the momentum boundary layer, with no such fluctuations
160 3 Receptivity and Instability

0.15 Ka = -0.05 (i) t = 30

0.1
y

-2
0.05 -5
-10
-20
-30
00 20 40 x 60

0.15 (ii) t = 50

0.1
y

-2
-5
0.05 -10
-20
00 20 40 x 60

0.15 (iii) t = 70

0.1
-2
y

-5
0.05 -10
-20
00 20 40 x 60

0.15 (iv) t = 90

0.1
-2
y

-5
0.05 -10
-20
00 20 40 x 60

Fig. 3.16 Vorticity (ω) contours plotted in (x, y)-plane at indicated times while computing equi-
librium flow cold adiabatic plate with buoyancy parameter K a = −0.05. [Reproduced from Direct
numerical simulation of transitional mixed convection flows: Viscous and inviscid instability mecha-
nism. Sengupta et al., Physics of Fluids, 25, 094102 (2013), with the permission of AIP Publishing.]

inside the shear layer, as seen in Fig. 3.16. It is noted in [66] that there are other qual-
itative differences between the cases in Figs. 3.17 and 3.16. First, the disturbances
do not originate from the leading edge. Secondly, fluctuations reveal significantly
taller vertical structures with higher intensity. Such differences are also perplexing,
because adiabatic plate with higher cooling from the leading edge induces favorable
pressure gradient. Despite this, the observed intense fluctuations, which do not allow
to compute even the equilibrium flows, prompted to study flow instability and their
propensity to very low amplitude small scale disturbances, which are due to numeri-
cal error, as noted also in Kaiktsis et al. [32]. The same numerical methods have been
3.7 Equilibrium Solution for Mixed Convection Flows: Isothermal Wall Case 161

0.2 (i) t = 20 Ki = 0.04

0.15
y
0.1
-0.05
0.05 -0.4
-4 -15
-50
00 20 x 40 60

0.2 (ii) t = 25
0.15
y

0.1
-0.05
-0.4 -0.4
0.05
-15
-50
00 20 x 40 60

0.2 (iii) t = 35
0.15
y

0.1

0.05 -0.4
-4 -15
-50
00 20 x 40 60

0.2 (iv) t = 45
0.15
y

0.1

0.05 -0.4
-4 -15
-50
00 -100
20 x 40 60

Fig. 3.17 Vorticity (ω) contours plotted in (x, y)-plane at indicated times while computing equilib-
rium flow of heated isothermal wedge with buoyancy parameter K i = 0.04. The dotted line in each
frame indicate the location in y direction where Fjφrtoft integrand Γ changes sign from positive to
negative. [Reproduced from Direct numerical simulation of transitional mixed convection flows:
Viscous and inviscid instability mechanism. Sengupta et al., Physics of Fluids, 25, 094102 (2013),
with the permission of AIP Publishing.]

used very successfully for receptivity study of ZPG boundary layer in [55, 65]. Same
methodologies have been used to study instability of adiabatic plate with heated
leading edge, without any problem [53, 73]. In these successful cases, equilibrium
flows have been obtained by DNS and whose receptivity to vortical wall excitation
studied. Such counter-intuitive behavior of the solution of NSE prompted the authors
in [66] to relook at the linear viscous flow instability mechanisms for both spatial
and temporal growths.
162 3 Receptivity and Instability

3.7.3 Linear Viscous Instability: Spatial and Temporal Routes

Linear viscous instabilities via spatial and temporal routes for the equilibrium flow
obtained as similarity solutions of mixed convection flows are studied next. Detailed
method of linear theory is described in [53, 73] for spatial analysis, and a brief account
of the same is provided here. As before. to obtain the equations for linear stability
analysis, consider the nondimensional NSE and energy equation with Boussinesq
approximation given by,
∂ ũ ∂ ṽ
+ =0 (3.74)
∂ x̃ ∂ ỹ
 
∂ ũ ∂ ũ ∂ ũ ∂ p̃ 1 ∂ 2 ũ ∂ 2 ũ
+ ũ + ṽ =− + + 2 (3.75)
∂ t˜ ∂ x̃ ∂ ỹ ∂ x̃ ˜ ∂ x̃ 2
Re ∂ ỹ
 
∂ ṽ ∂ ṽ ∂ ṽ ∂ p̃ 1 ∂ 2 ṽ ∂ 2 ṽ
+ ũ + ṽ = Ri θ̃ − + + (3.76)
∂ t˜ ∂ x̃ ∂ ỹ ∂ ỹ ˜ ∂ x̃ 2
Re ∂ ỹ 2
 2 
∂ θ̃ ∂ θ̃ ∂ θ̃ 1 ∂ θ̃ ∂ 2 θ̃
+ ũ + ṽ = + 2 (3.77)
∂ t˜ ∂ x̃ ∂ ỹ ˜ Pr ∂ x̃ 2
Re ∂ ỹ

where, Ri and Re ˜ are Richardson and Reynolds numbers, respectively. For linear
instability analysis, flow properties including velocity, pressure and temperature are
split into a mean and a fluctuating part, with variables represented as

z̃(x̃, ỹ, t˜) = Z(x̃, ỹ) + εz̄(x̃, ỹ, t˜) (3.78)

where, z̃ = [ũ, ṽ, p̃, θ̃]T represents the total quantities; Z = [U, V, P, T ]T rep-
resents the mean components and z̄ = [ū, v̄, p̄, θ̄ ]T represents fluctuating compo-
nents. Mean components are obtained after proper transformations of the mean flow
equations described before.
For 2D local linear instability studies, parallel flow assumption is invoked by
requiring U = U ( ỹ), V = 0, P = P( ỹ) and T = T ( ỹ). For non-dimensionalization,
suitable length, velocity and temperature scales chosen are δ ∗ (local displacement
thickness), shear layer edge velocity Ue and ΔT (x ∗ ) (difference between local plate
and free stream temperatures, i.e., ΔT (x ∗ ) = Tw (x ∗ ) − T∞ ). The non-dimensional
numbers in Eqs. (3.74)–(3.77) are defined as, Re ˜ Gr
˜ = Ue δ ∗ /ν and Ri = Re/ ˜ 2 with
Grashof number as Gr ˜ = gβt ΔT (x )δ /ν . Using Fourier-Laplace transform, lin-
∗ ∗ 3 2

ear analysis is carried out with the perturbation quantities given by,

[ū, v̄, p̄, θ̄ ] = [g̃( ỹ), φ( ỹ), π( ỹ), h( ỹ)] ei(k x̃−β t˜) dkdβ (3.79)

Here, k = kr eal + ikimg and β = β0 + iβi , are complex wavenumber and circular
frequency, respectively. For spatial instability studies one considers wavenumber k
3.7 Equilibrium Solution for Mixed Convection Flows: Isothermal Wall Case 163

to be complex, whereas β is treated as real. Spatial growth of disturbances is noted


for kimg being negative. For temporal instability, wavenumber k is real (k = kr eal ),
while β is complex. Positive values of βi signify temporal growth of disturbances. By
substituting Eq. (3.78) into Eqs. (3.74)–(3.77) and retaining O(ε) terms, linearized
equations are obtained. Further, substitution of Eq. (3.79) into linearized equations,
results in the ODEs, governing the disturbance amplitude functions (with ỹ as the
independent variable) which can be simplified to,

1 iv 
i(kU − β)(k 2 φ − φ  ) + ikU  φ = Ri k 2 h − (φ − 2k 2 φ + k 4 φ) (3.80)
˜
Re
1
i(kU − β)h + T  φ = (h  − k 2 h) (3.81)
˜ Pr
Re

Where prime indicate derivative with respect to ỹ. These are the sixth order OSE
for mixed convection flows. The system given by Eqs. (3.80) and (3.81) are solved
subject to six boundary conditions,

at ỹ = 0 : φ, φ  = 0; h = 0 (3.82)
and as ỹ → ∞ : φ, φ  , h → 0 (3.83)

General solution to the ODEs are given as

φ = a 1 φ1 + a 2 φ2 + a 3 φ3 + a 4 φ4 + a 5 φ5 + a 6 φ6 (3.84)
h = a1 h 1 + a2 h 2 + a3 h 3 + a4 h 4 + a5 h 5 + a6 h 6 (3.85)

Adopted boundary conditions imply disturbances which decay in the far stream ( ỹ →
∞). However in the free stream, h decouples from φ, as U ≈ 1 and U  , T  ≈ 0.
Thus in the free stream, energy equation Eq. (3.81), reduces to

˜ Pr (k − β) + k 2 ]h = 0
h  − [i Re (3.86)

whose solution at the free stream is given by the characteristic modes as

h ∞ = a5 e−S ỹ + a6 e S ỹ (3.87)

where S = k 2 + i Re ˜ Pr (k − β). To satisfy free stream boundary condition, we
must have a6 = 0 for real(S) > 0. It can be shown that as ỹ → ∞, h 1∞ = h 3∞ = 0.
At the free stream, U = 1, and all mean flow derivatives are zero, and linearized
momentum equation reduces to

˜
φ iv − [2k 2 + i Re(k

˜
− β)]φ + [k 4 + i Re(k ˜ k2h
− β)k 2 ]φ = Ri Re (3.88)
164 3 Receptivity and Instability

Even in the free stream, momentum equation is not decoupled from the thermal field,
with the latter providing forcing on the former. Solution of Eq. (3.88) is a sum of
homogeneous solution and a particular integral, which after removing exponentially
growing terms (by enforcing a2 = a4 = 0) gives

φ∞ = a1 e−k ỹ + a3 e−Q ỹ + a5 Γ˜ e−S ỹ (3.89)



where Q = k 2 + i Re(k ˜ − β) and Γ˜ = Ri Re ˜ k 2 /[S 4 − (k 2 + Q 2 )S 2 + k 2 Q 2 ],
with real parts of k, Q and S as positive. Hence, the general solution of the cou-
pled sixth order OSEs, subject to the prescribed homogeneous boundary conditions
at free stream given by Eq. (3.83) are

φ = a 1 φ1 + a 3 φ3 + a 5 φ5 (3.90)

h = a1 h 1 + a3 h 3 + a5 h 5 (3.91)

where h 1∞ = h 3∞ = 0, h 5∞ = e−S ỹ , and φ1∞ = e−k ỹ , φ3∞ = e−Q ỹ and φ5∞ =


Γ˜ e−S ỹ . As the three fundamental solutions decay exponentially at three different
widely separate rates, Eqs. (3.80) and (3.81) represent a set of stiff ODEs. One
adopts stiff solvers like the CMM [1, 44, 53] to solve the equations. As noted [53] in
CMM, one solves a set of auxiliary equations derived from the original equation, in
terms of second-compound variables, well-defined combinations of φ j , h j and their
higher derivatives. The second compound variables grow/ decay exponentially at
comparable rates ([53] provides details of CMM). The sixth order OSE is converted
into an initial value problem in CMM with twenty (20), second-compounds, with
initial conditions defined in free stream. The eigenvalues of Eqs. (3.80) and (3.81) are
found by integrating these twenty equations from the free stream to the wall, subject
to initial conditions described in [53, 73]. Satisfying the dispersion relation obtained
from the wall-boundary condition given by Eq. (3.82), one obtains the eigenvalues,
which represent complex wavenumber (k), for a particular combination of Re, ˜ β0
and K i (or K a ) for spatial stability analysis and complex frequency β in temporal
˜ kr eal and K i (or K a ). For disturbance components of
analysis for fixed values of Re,
temperature and velocity, wall-boundary condition in Eq. (3.82), can be written in
terms of fundamental solution components as given below (which in turn provides
the dispersion relation),

a 1 φ1 + a 3 φ3 + a 5 φ5 = 0 (3.92)
a1 φ1 + a3 φ3 + a5 φ5 = 0 (3.93)
a1 h 1 + a3 h 3 + a5 h 5 = 0 (3.94)

Characteristic determinant of the above linear system combined with definition of


second-compounds in [53] provides the dispersion relation

Dr + i Di = y3 = 0 at ỹ = 0 (3.95)
3.7 Equilibrium Solution for Mixed Convection Flows: Isothermal Wall Case 165

where [73],

φ1 φ 3 φ 5
y3 = φ1 φ3 φ5 . (3.96)
h1 h3 h5

3.7.4 Linear Spatial Viscous Theory for Mixed Convection


Flows

Linear spatial viscous instability of mixed convection flows over adiabatic cold plate
and isothermal heated wedge are studied first, where wavenumber k is complex
and frequency β is real. Obtained growth rate contours (kimg ) are plotted with the
corresponding neutral curves (kimg = 0) in Fig. 3.18a, b.

(a) Neutral curves for flow over adiabatic cold plate case (c) Neutral curves for adiabatic cold plate case
from spatial approach from linear temporal viscous analysis
0.15
A 0.4

Ka = -0.01 P2
0.3 Ka = -0.01
0.1
F = Const. ray
kreal
β0

P1 Ka = -0.05 0.2 Ka = -0.05

0.05
0.1
Ka = -0.001
Ka = -0.001
0
O 0
0 500 1000 1500 2000 2500 3000 3500 0 500 1000 1500 2000 2500 3000 3500
Reδ∗ Reδ*
(b) Neutral curves for the case of flow over heated (d) Neutral curves for heated isothermal wedge
0.15 isothermal wedge from spatial approach from linear temporal viscous analysis
0.4
Ki = 0.04
B
0.3 Ki = 0.04
0.1
kreal

Ki = 0.02 Q2
β0

5
0.2
K= Ki = Ki = 0.0
0.05 Ki = 0.001 i 0 . 02 0.0 1 01
Ki = 5
0.01 0.1
Q1
O
0 0
0 1500 3000 4500 6000 7500 0 1500 3000 4500 6000 7500
Reδ∗ Reδ*

Fig. 3.18 a, b Neutral curve obtained from linear spatial analysis shown in (Reδ ∗ , β0 )-plane buoy-
ancy parameter for: a flow over cold adiabatic plate and b flow over heated isothermal wedge.
Frames c, d show neutral curve obtained from linear temporal instability analysis shown in the
(Reδ ∗ , kr eal )-plane for indicated values of buoyancy parameter for the cases shown in frames (a)
and (b). [Reproduced from Direct numerical simulation of transitional mixed convection flows:
Viscous and inviscid instability mechanism. Sengupta et al., Physics of Fluids, 25, 094102 (2013),
with the permission of AIP Publishing.]
166 3 Receptivity and Instability

In Fig. 3.18a, b, neutral curves (kimg = 0 contour) are plotted in (Reδ∗ , β0 )-plane,
for flow over cold adiabatic plate and over a heated isothermal wedge for different

values of buoyancy parameter, respectively. The Reynolds number is Reδ∗ = Ueνδ and
β0 is the real part of β. These neutral curves are obtained for the similarity solution
in [49]. The flow is spatially unstable inside the neutral curve. Figure 3.18a shows
the case of flow over adiabatic cold plate, with the critical Reynolds number (Reδ∗ )cr
increasing with increased cooling. The highest β0 , above which all disturbances are
stable, is termed here as the critical circular frequency (β)cr . One notes that (β)cr
decreases with increase in cooling. The region inside the neutral curve becomes
wider for a given Reδ∗ when cooling is reduced.
The ray OA in Fig. 3.18a correspond to a constant non-dimensional physical fre-
quency defined as follows:

F = β/Reδ∗ = 2π ν f˜/Ue2

where f˜ is the excitation frequency in Hertz. When the flow is excited by a source
at constant frequency, one tracks the disturbance following the ray corresponding to
F = constant in this theory. The ray O A enters the neutral curve for K a = −0.01
at P1 in Fig. 3.18a, and P2 represents the point where it exits the neutral curve. In
Fig. 3.18b, neutral curves for flows over hot isothermal wedge have been plotted
in (Reδ∗ , kr eal )-plane, for different K i . In this case, (Reδ∗ )cr decreases and (β)cr
increases, with increasing value of K i . Thus, heating the plate destabilizes the flow
more by the linear local viscous spatial theory. Also, the neutral curve becomes wider
with increasing value of K i .

3.7.5 Linear Temporal Viscous Theory for Mixed Convection


Flows

For linear temporal viscous instability, wavenumber k is real (k = kr eal ), and the
frequency β is complex. Disturbance will grow with time, if βi is positive. Here, one
finds eigenvalues β = β0 + iβi , for chosen combinations of Reδ∗ , kr eal and buoyancy
parameter K a or K i . In Fig. 3.18c, d, the neutral curves (βi = 0) obtained from
temporal analysis are shown in (Reδ∗ , kr eal )-plane for adiabatic cold plate (frame (c))
and isothermal heated wedge (frame (d)). From Fig. 3.18c one notes that (Reδ∗ )cr
decreases with increased cooling, similar to corresponding results of spatial viscous
instability analysis shown in Fig. 3.18a. For isothermal heated wedge cases shown
in Fig. 3.18d, temporal analysis predicts (Reδ∗ )cr to increase with increased heating,
as is concluded from corresponding plots shown in Fig. 3.18b. One also notes from
different frames of Fig. 3.18 that for a chosen buoyancy parameter K a or K i , (Reδ∗ )cr
obtained from both spatial and temporal instability analysis are identical.
Temporal viscous growth rate is provided in Fig. 3.19, with ci (= βi /kr eal )-
contours plotted in (Reδ∗ , kr eal )-plane for the indicated cases. This parameter is used
3.7 Equilibrium Solution for Mixed Convection Flows: Isothermal Wall Case 167

(a) Ka = -0.001 Min = -0.71, (b) Ka = -0.05 Min = -0.71,


Max = 0.022 Max = 0.0058
0.4 0.4
-0.01 -0.05 0
0
-0.003 0. 0
0.3 0.3 0 03
-0.01
kreal

kreal
0 0.015
0.2 0.2 -0.02
0.02
-0.01 -0.03

0.1 0.1 -0.05


-0.1 -0.05
-0.1
0 0
1000 2000 3000 4000 0 1000 2000 3000 4000
Reδ* Reδ*
(c) Ki = 0.04 Min = -0.64 (d) Ki = 0.001 Min = -0.58
Max = 0.0255 Max = 0.003
0.4 0.4

-0.01 -0.022
0.3 0.3
-0.01
0
kreal

kreal

-0.01
0.2 0.2 0
-0.005
0.02
0
0.1 0.02 0.1 -0.022
0 -0.05
-0.1 -0.1
0 0
0 2000 4000 6000 0 2000 4000 6000
Reδ* Reδ*

Fig. 3.19 Temporal growth rate contours ci = βi /kr eal plotted in (Reδ ∗ , kr eal )-plane for indicated
mixed convection cases. [Reproduced from Direct numerical simulation of transitional mixed
convection flows: Viscous and inviscid instability mechanism. Sengupta et al., Physics of Fluids,
25, 094102 (2013), with the permission of AIP Publishing.]

for inviscid temporal instability studies. The maximum and minimum values of ci are
indicated in each frame. One notes that the adiabatic cold plate case with K a = −0.05
displays lesser temporal growth rate than K a = −0.001 case, i.e., cooling decreases
instability. For the isothermal heated wedge, one notes that heating enhances tem-
poral growth rate.
Thus the viscous temporal analysis for mixed convection flows over both adiabatic
cold plate and isothermal heated wedge show identical trend, as one notes for spatial
analysis. Both the viscous linear theories do not explain the instabilities noted in
Figs. 3.16 and 3.17 for DNS, with enhanced instability noted for heating and cooling
increased.
The mixed convection flow instabilities are not explained well by local analysis.
The local linear viscous theory has several drawbacks. The spatial theory posed as
signal problem creates output at the frequency of excitation, which has been shown
to be incorrect in [55, 65]. Linear theory based on parallel flow approximation is
168 3 Receptivity and Instability

not accurate near the leading edge of the plate. Results in Fig. 3.16, clearly show
onset of disturbance near the proximity of the leading edge, as noted earlier for
vortex-induced instability [60]. These linear analyses viewing instability either as
a spatial or a temporal growth problem has been shown as inadequate for Blasius
boundary layer [55, 65]. To incorporate the streamwise variation of the flow, one
can perform either global linear studies [5, 11] or use fully nonlinear formulation
[18, 39, 55, 65]. The full simulations provide results including nonlinear effects,
and help in relating the results with the local analysis for some cases. Results in
[66] has shown that there are many unexplained facets of local analysis for mixed
convection flows. Ambiguities of linear theories have been clearly seen with respect
to results in Figs. 3.16 and 3.17, even in not being able to compute the equilibrium
flow. Another aspect which has been completely overlooked is the inviscid instability
mechanisms in mixed convection flows, as compared to such well developed analy-
sis for flows without heat transfer. These suggest that a spatio-temporal receptivity
study of a mixed convection boundary layer to a deterministic excitation is essential
to fully understand various instability mechanisms based on solving NSE and few
representative cases are explored next.

3.8 Receptivity of Mixed Convection Flows

From Figs. 3.16 and 3.17, it is noted that in some mixed convection cases it is not
even possible to obtain equilibrium flows by solving the NSE, for those special cases
for which one obtains a steady similarity solutions [42, 49]. DNS produces steady
equilibrium flows, only when heat transfer is lower. In this section, we consider
few cases for which mean flow is obtained by solving NSE and their receptivity to
wall excitation is followed. These correspond to the cases of (i) adiabatic cold plate
with K a = −0.01 and −0.001 and (ii) the isothermal hot wedge with K i = 0.025.
Receptivity of such cases are studied for wall excitation by simultaneous blowing
and suction (SBS) strip. Such vortical excitation has been applied for flows without
heat transfer in [9, 22, 55]. The SBS strip provides monochromatic harmonic exci-
tation, and a brief description of the SBS strip is provided. The surface temperature
distribution is kept unaltered at the exciter, while the expressions for wall stream
function and wall vorticity are changed as,

ψw = ψ̄o + ψw p (3.97)
 
1 ∂ h 2 ∂ψ 1 ∂ 2ψ
ωw = − − (3.98)
h 1 h 2 ∂ξ h 1 ∂ξ h 22 ∂η2

Expression for wall vorticity contains additional ξ derivatives, because ∂∂vx = 0 at


the wall, when exciter is switched on. Wall perturbation stream function ψw p in Eq.
(3.97) is derived from the imposed wall velocity condition given by,
3.8 Receptivity of Mixed Convection Flows 169

u d = 0, vd = A(x) sin(β L t) (3.99)

where β L is the non-dimensional exciter frequency based on the reference length L.


The amplitude of exciter is obtain as A(x) = α1 Am (x), where α1 is an amplitude
control parameter. For an exciter located between x1 and x2 , the amplitude distribution
for the exciter is given by Am (x),
For x1 ≤ x ≤ xst
 5  4  
x − x1 x − x1 x − x1 3
Am = 15.1875 − 35.4375 + 20.25
xst − x1 xst − x1 xst − x1
(3.100)

and for xst ≤ x ≤ x2


 5  4  
x2 − x x2 − x x2 − x 3
Am = −15.1875 + 35.4375 − 20.25
x2 − xst x2 − xst x2 − xst
(3.101)

x1 +x2
where xst = 2
.

3.8.1 Receptivity of Cold Adiabatic Flat Plate Cases

A specific excitation applied to a cold adiabatic flat plate has been studied in [66],
with the frequency of excitation considered is F = 9.76 × 10−5 , and with an ampli-
tude control parameter of α1 = 0.001. Thus, the maximum excitation amplitude is
0.1% of the free stream speed. The cases studied are for K a = −0.001 and −0.01
with the exciter at x = 1.5 (where Reδ∗ = 670 for K a = −0.01), and we noted in
Fig. 3.18a, P1 corresponds to Reδ∗ = 744, while P2 is at Reδ∗ = 1256. It is evident
from Fig. 3.18a that the onset of growth for the lower cooling case occurs slightly
early, while the point at which the ray exits the neutral curve is almost the same. In
Fig. 3.20a, results for K a = −0.01 and −0.001 cases are shown side by side for the
disturbance streamwise velocity (u d ), at y = 0.0061, which demonstrate same ele-
ments of the disturbance field as reported in [55, 65], for flows without heat transfer.
The equilibrium flow is obtained before the excitation is switched on at t = 0 and
this is subtracted from the instantaneous solution for the excited flow to obtain the
disturbance field.
Results are shown for K a = −0.01 on the left column of Fig. 3.20a, showing a
nascent STWF visible at t = 85, ahead of the TS wave-packet visible at the left of
the panel. Features of the disturbance field are: (i) Fixed streamwise extent of the
TS wave-packet at all times and (ii) the growing STWF that convects downstream.
The STWF has been noted in [62] from the solution of OSE, implying its origin
by linear mechanism, is also noted here in the time evolution of u d up to t = 190.
170 3 Receptivity and Instability

0.05 (i) t = 85 0.05 (vi) t = 85


Ka = -0.01 ; y = 0.0061 -5 Ka = -0.001 ; y = 0.0061
0.025 α1 = 0.001 ; F = 9.76×10 0.025 α1 = 0.001 ; F = 9.76×10-5

ud

ud
0 0

-0.025 -0.025

-0.05 0 -0.050 20
x 40 60
20 40 60
x
0.05 (ii) t = 120 0.05 (vii) t = 120

0.025 0.025
ud

ud
0 0

-0.025 -0.025

-0.050 20 -0.050
x 40 60 20
x 40 60

(iii) t = 155 (viii) t = 155


0.05 0.4

0.025 ud0.2
ud

0 0

-0.025 -0.2

-0.050 -0.40 20
20
x 40 60 x 40 60

(iv) t = 190 0.4


(ix) t = 190
0.05

0.025 0.2
ud
ud

0 0

-0.025 -0.2

-0.050 -0.40 20
20
x 40 60 x 40 60

(v) t = 225 0.4


(x) t = 225
0.15
0.1
0.2
0.05
A
ud
ud

0 0
-0.05
-0.2
-0.1
-0.15 -0.4
0 20
x 40 60 0 20 x 40 60

Fig. 3.20 a Disturbance streamwise velocity u d at a height of y = 0.0061 plotted at indicated


times for flow past a horizontal adiabatic cold plate with K a = −0.001 (frames (i − v)) and −0.01
(frames (vi − x)) is excited by SBS strip placed at x = 1.5. Here, non-dimensional frequency and
amplitude of excitation are F = 9.76 × 10−5 and α1 = 0.001, respectively. The vertical dashed
lines in the frames indicate onset and decay of disturbance locations (as P1 and P2 in Fig. 3.18a) as
per local linear spatial instability theories. b u d is plotted at a height of y = 0.0061 for the indicated
times for flow past an isothermal heated wedge with K i = 0.025 excited by SBS strip placed at
x = 21.0. Here, F = 1.65 × 10−5 and α1 = 0.001, are used and the vertical dashed lines in the
frames indicate onset and decay of disturbance locations (as Q 1 and Q 2 in Fig. 3.18b) obtained
by linear spatial instability theory. Reproduced from [Direct numerical simulation of transitional
mixed convection flows: Viscous and inviscid instability mechanism. Sengupta et al., Physics of
Fluids, 25, 094102 (2013)], with the permission of AIP Publishing
3.8 Receptivity of Mixed Convection Flows 171

0.075 (i) t = 40 Ki = 0.025 ; y = 0.0061 -5


α1 = 0.001 ; F = 1.65×10
0.05
0.025

ud
0
-0.025
-0.05
-0.075 0 20 40 60
x
(ii) t = 52
0.75
0.5
0.25
ud

0
-0.25
-0.5
-0.75
0 20 40 60
x
(iii) t = 64
0.75
0.5
0.25
ud

0
-0.25
-0.5
-0.75
0 20 40 60
x
0.75 (iv) t = 72
0.5
0.25
ud

0
-0.25
-0.5
-0.75 0 20 40 60
x
0.75 (v) t = 80
0.5
0.25
ud

0
-0.25
-0.5
-0.75 0 20 40 60
x

Fig. 3.20 (continued)

STWF experiences nonlinear growth, as noted in the frame at t = 225. This also
shows the creation of an upstream disturbance packet, marked as A. In [55, 65],
dynamics and self-regeneration mechanism of STWF have been described for flows
without heat transfer. The lower cooling case of K a = −0.001 shown in the right
frames of Fig. 3.20a show that the instability is stronger and occurs earlier, implying
that the cooling stabilizes viscous modes, as noted in the last section, as the results
for K a = −0.001 show earlier onset of the STWF and larger amplitude in the top
frames. With higher growth, nonlinearity also sets in earlier, as noted at t = 155.
The y-axis is magnified for the lower K a case, to indicate its higher growth rate.
172 3 Receptivity and Instability

3.8.2 Receptivity of Hot Isothermal Wedge Case

Another case of receptivity is studied for hot isothermal wedge case, with the SBS
excitation frequency of F = 1.65 × 10−5 and α1 = 0.001. In Fig. 3.18b, this case has
been marked by the ray OB, with the exciter located at a position where Reδ∗ = 2190,
while point of entry inside the neutral curve at Q 1 corresponds to Reδ∗ = 2282 and
point of exit at Q 2 correspond to Reδ∗ = 4333. This is evidently at lower frequency
for the isothermal heated wedge, and the flow remains stable up to a longer streamwise
stretch. In Fig. 3.20b, one notes near-absence of TS wave-packet, while the STWF is
visible at an earlier time. Also, the growth rate for the STWF is higher and therefore
the onset of nonlinear distortion of the STWF occurs much earlier. Additionally,
one notices the presence of very strong secondary upstream STWF created by the
primary STWF leaving the computational domain. The low frequency disturbances
have stronger instability, and such response fields are qualitatively different from
moderate to higher frequency excitation cases for flows without heat transfer [7, 65].

3.8.3 DNS of Instability of Mixed Convection Flows: New


Theorems of Instability

So far we have noted contradictory observations for mixed convection flow insta-
bilities, both for heating and cooling at the wall. For example, adiabatic flow past a
horizontal plate indicate enhanced stability with cooling in Fig. 3.18a, c. However,
when the cooling is more intense, we are even unable to compute the equilibrium flow,
e.g., for K a = −0.05. But the lower cooling cases allow one to calculate equilibrium
flow and study its receptivity, and the observations from linear viscous instability
studies are compatible with DNS of the NSE. Same observation holds good for com-
puting mean flow and study receptivity of lesser heated isothermal wedges. For the
increased heating of K i = 0.04 case, the mean flow is not obtained by DNS. All these
evidences suggest that there can be additional instability mechanisms for mixed con-
vection flows, apart from that is obtained by solving the OSE by spatial or temporal
theories. From DNS shown in Figs. 3.16 and 3.17, we note in the course of failed
attempt in computing an equilibrium flow, disturbances appear naturally outside the
boundary layer, indicating temporal growth without significant convection. With pas-
sage of time, disturbances spread over larger streamwise stretch, without penetrating
inside the boundary layer. Such growths are reminiscent of inviscid instability mech-
anism arising from Rayleigh’s stability equation and the theorems stating necessary
conditions provided by Rayleigh and Fjφrtoft [19, 53] for flows without heat transfer.
The corresponding theorems for inviscid temporal instabilities for mixed convection
flow have been reported for the first time in [66], and is explained in the following.
The inviscid governing equations for disturbances in mixed convection flows
modeled by Boussinesq approximation are given in non-dimensional form as
3.8 Receptivity of Mixed Convection Flows 173

∂ ũ ∂ ṽ
+ = 0, (3.102)
∂ x̃ ∂ ỹ

∂ ũ ∂ ũ ∂ ũ ∂ p̃
+ ũ + ṽ =− , (3.103)
∂ t˜ ∂ x̃ ∂ ỹ ∂ x̃

∂ ṽ ∂ ṽ ∂ ṽ ∂ p̃
+ ũ + ṽ = Ri θ̃ − , (3.104)
∂ t˜ ∂ x̃ ∂ ỹ ∂ ỹ

∂ θ̃ ∂ θ̃ ∂ θ̃
+ ũ + ṽ = 0. (3.105)
∂t ˜ ∂ x̃ ∂ ỹ

Scales for length, velocity, pressure and temperature used to non–dimensionalize


these equations are the local displacement thickness (δ ∗ ), shear layer edge velocity
Ue , ρUe2 and ΔT (x ∗ ) = (Tw (x ∗ ) − T∞ ), respectively. The Richardson number Ri =
Gr˜
˜ 2
is retained in this linearized analysis for the inviscid mechanism, as the buoyancy
Re
effect is not a lower order perturbation. This is due to the fact that the boundary layer
grows as O(Re−1/2 ) and to have equal effects of free and forced convection in
mixed convection flows, we must have K i = Ri/Re1/2 ≈ O(1), which necessitates
retaining the buoyancy effects noted in wall-normal direction.
Once again, parallel flow approximation for the equilibrium flow is used, and
the perturbation quantities are expressed by Fourier-Laplace transform given in Eq.
(3.79). With these, Eqs. (3.102)–(3.105) are simplified to get a single equation in
terms of φ (Fourier-Laplace transform of v̄) as
 Ri dT 
d 2φ d ỹ d 2U
(U − c) + φ k 2
(c − U ) + − =0 (3.106)
d ỹ 2 (U − c) d ỹ 2

where, c = β/k is the complex phase speed for temporal instability. This is the
equivalent equation to study inviscid temporal instability of mixed convection flows.
For such studies, one takes k as real, (k = kr eal ) and β as a complex quantity. Hence,
c = cr + i ci , with cr and ci denoting the phase speed and temporal growth rate (if
ci > 0) of the disturbance field.
Rayleigh’s and Fjφrtoft’s theorems for flows without heat transfer are stated with-
out solving this equation explicitly, to state necessary conditions. We will obtain
robust conditions for temporal instability, by multiplying Eq. (3.106) with φ ∗ (com-
plex conjugate of φ) and integrating over all possible limit (0, ∞) to get

 2 

∗d φ2 Ri dT
d ỹ
d U
d ỹ 2
φ + |φ| 2
−k 2
+ − d ỹ = 0 (3.107)
0 d ỹ 2 (U − c) 2 (U − c)

This equation is simplified by using integration by parts and using boundary


conditions: φ(0) = 0 and φ → 0, as ỹ → ∞ to get
174 3 Receptivity and Instability

 ⎡ ⎤
  Ri dT
dφ 2 |φ|2 ⎣ d 2 U ∗ d ỹ ∗
2 2
+ k |φ| d ỹ + (U − c ) − (U − c ) ⎦ d ỹ = 0
2
dy |U − c|2 d ỹ 2 |U − c|2
(3.108)

where c∗ is the complex conjugate of c, i.e., c∗ = cr − ici . The imaginary part of


Eq. (3.108) is given as
 
ci |φ|2 −2Ri dT
d ỹ
(U − cr ) d 2U
+ dy = 0 (3.109)
|U − c|2 |U − c|2 d ỹ 2

For temporal instability, one must have ci = 0, and hence to have the integral to
be zero in Eq. (3.109), one must require the term within the third brackets to change
sign, somewhere in the flow domain. We define this as

−2Ri dT
d ỹ
(U − cr ) d 2U
Φ= + (3.110)
|U − c|2 d ỹ 2

For flows without heat transfer (Ri = 0), one gets back the well known Rayleigh’s
2
inflection point theorem, with Φ ≡ dd ỹU2 . Thus, we state an extension of the Rayleigh’s
theorem for inviscid temporal instability by [66]:
Mixed Convection Flow Theorem I - The necessary condition for inviscid insta-
bility of mixed convection flows described by the velocity and temperature profiles,
U ( ỹ) and T ( ỹ), the integrand Φ as defined in Eq. (3.110) should be zero somewhere
in the domain.
The consequence of Mixed Convection Flow Theorem I follows as: (i) The veloc-
ity and temperature fields have all viscous information of the parallel flow and can be
used without the need for any similarity profile, and hence DNS data can also be used
with a local parallel flow approximation to study the resultant inviscid instability; (ii)
Unlike flow cases without heat transfer, existence of an inflection point of velocity
profile alone does not define this temporal instability.
Existence of a point, where Φ changes sign depends on the complex phase speed
c, making the analysis more involved and can be performed parametrically. Noting
that Φ remains bounded for non-neutral cases, there is a critical layer for mixed
convection problem also for neutral disturbances where U ( ỹcr ) = cr .
Mixed Convection Flow Theorem I is obtained here from the imaginary part of
Eq. (3.108) and is identical to the procedure used in Rayleigh’s theorem. It was
subsequently modified by Fjφrtoft, who considered the real part of Eq. (3.108) also
in deriving another necessary condition.
Similarly, we obtain another necessary condition for mixed convection flows,
along the procedure used by Fjφrtoft for flows without heat transfer. This necessary
condition uses the real part of Eq. (3.108) for temporal instability of the mixed
convection flows. The real part of Eq. (3.108) is given by,
3.8 Receptivity of Mixed Convection Flows 175

 
|φ|2 −Ri dT
d ỹ
{(U − cr )2 − ci2 } d 2U
+ (U − cr ) d ỹ
|U − c|2 |U − c|2 d ỹ 2
 
dφ 2
+ + k 2 |φ|2 d ỹ = 0 (3.111)
d ỹ

In this equation, the second integral is always positive, so the integrand of the first
integral must be negative in some interval within the flow to have the net integral to
vanish. Let us indicate a height ỹ = ỹs , where Φ( y˜s ) = 0 and denote U ( ỹs ) = Us .
Multiplying Eq. (3.109) by (cr − Us ) and adding it with the first integral of Eq.
(3.111) and simplifying, we get
 
Ri dT
d ỹ d 2U
|φ| 2
{(Us − cr ) + − (U − Us ) } +
2
ci2 (U − Us ) d ỹ < 0
2
|U − c|2 d ỹ 2
(3.112)
We define the integrand within the third bracket as
 Ri dT 
d ỹ d 2U
ΓF = {(Us − cr ) +
2
ci2 − (U − Us ) } +
2
(U − Us ) (3.113)
|U − c|2 d ỹ 2

Defining Γ F as the Fjφrtoft’s integrand, we proceed and state the following the-
orem for inviscid temporal instability of mixed convection flow as [66]:
Mixed Convection Flow Theorem II - The necessary condition for the inviscid
instability for a mixed convection parallel flow described by velocity and temperature
profiles U ( ỹ) and T ( ỹ) is that the integrand Γ F (in Eq. (3.113)) should be negative
somewhere in the interior of the domain.
This is corresponding to Fjφrtoft’s theorem, and is valid for inviscid temporal
instability of mixed convection flows. As noted for Φ, Γ F is also a function of c,
which means that one can inspect parametrically the likelihood of such inviscid
instabilities by tracking different temporal growth rates (ci ) and associated time/
length scales (cr ).
The utilities of Mixed Convection Flow Theorems I and II, is to help explain
why in some cases of heat transfer, the equilibrium flow given by [42, 49] cannot be
realized from the time-accurate solution of NSE. We are specifically interested for
the cases in (i) Figs. 3.15a and 3.16 for the cold adiabatic plate case of K a = −0.05
and (ii) Figs. 3.15b and 3.17 for the hot isothermal wedge flow case of K i = 0.04.
The inviscid instability will be predicted by Mixed Convection Flow Theorems I and
II, when Φ = 0 and Γ F < 0 conditions are satisfied inside the flow field.
We have plotted Φ as a function of ỹ in Fig. 3.21a, for flow over an adiabatic
cold plate, with K a = −0.05. The similarity profiles obtained as functions of the
similarity variable η have been converted to functions of ỹ as ỹ = η/Cδ , where

Cδ = 0 (1 − f  ) dη. The variation of Φ with η are shown in frame (i) for the
fixed value of ci = 0.01. The three curves correspond to three different values of
176 3 Receptivity and Instability

Rayleigh integrand Φ for adiabatic cold plate, Ka = -0.05


(i) ci = 0.01
0.02 Edge of the
0.2 0.5 momentum
0.01 boundary layer
0

Φ
-0.50
0
-1
-0.01
0 0.1 0.10.2
y~
y0.30.2 0.4 0.5
0.3

-0.2
Φ

-0.4
cr = 0.001
cr = 0.01
-0.6 cr = 0.1

0 1 2 ~ 3 4
y

(ii) ci = 0.1

0
cr = 0.001 Edge of the
-0.1 cr = 0.01 momentum
cr = 0.1 boundary layer
-0.2
Φ

-0.3

-0.4

-0.5

-0.6
0 2 ~ 4
y

Fig. 3.21 a Rayleigh integrand Φ for flow over adiabatic plate parameter K a = −0.05 for
i ci = 0.01 and ii ci = 0.1. b Fjφrtoft integrand Γ F for flow over adiabatic plate with buoyancy
parameter K a = −0.05 for i ci = 0.01 and ii ci = 0.1. Reproduced from [Direct numerical simu-
lation of transitional mixed convection flows: Viscous and inviscid instability mechanism. Sengupta
et al., Physics of Fluids, 25, 094102 (2013)], with the permission of AIP Publishing

cr = 0.001, 0.01 and 0.1. One notes for the lower values of cr , Φ crosses zero near
the wall, as shown in the inset of this frame. For the highest frequency of cr = 0.1,
Φ flips sign at one value of η discontinuously. In the next frame (ii) of Fig. 3.21a,
similar curves are plotted for the higher growth rate of ci = 0.1. The zero-crossing
is noted for all the three values of cr shown in this frame. These imply that for all
the three time scales considered (varying by a factor of 100), are unstable for both
ci = 0.01 and 0.1. Such temporal growth is seen in the outer part of the shear layer,
near the leading edge of the plate in Fig. 3.16. This is a plausible explanation for
3.8 Receptivity of Mixed Convection Flows 177

(i) ci = 0.01 Fjφrtoft integrand ΓF for adiabatic plate case, Ka = -0.05

0
cr = 0.001
cr = 0.01
cr = 0.1
-0.1

Edge of the
ΓF

momentum
-0.2 y 0.3
boundary
0.4
layer

-0.3

-0.4
0 2 ~
y
4

(ii) ci = 0.1
cr = 0.001
0 cr = 0.01
cr = 0.1

-0.1
ΓF

Edge of the
momentum
-0.2 boundary layer

-0.3

0 2 ~
y
4

Fig. 3.21 (continued)

failing to obtain a mean flow for K a = −0.05, as due to inviscid temporal instability
for very high cooling rate, which however was shown to be stable for any cooling by
linear viscous instability theories.
Similarly, Γ F has been plotted as a function of ỹ in Fig. 3.21b, for K a = −0.05.
The necessary condition for inviscid instability according to Mixed Convection Flow
Theorem II is for Γ F becoming negative in the flow domain. The frames in Fig. 3.21b
are plotted again for ci = 0.01 and 0.1; with both the cases showing inviscid temporal
instability criterion to be satisfied. This satisfaction of both the necessary conditions
of Theorems I and II, clearly explain why an equilibrium flow is not computed for
this case. It is also equally important to note that when the flow is both temporally and
spatially unstable, the stronger one will dominate, as in the present case the spatial
instability is weaker while the temporal instability is stronger and is readily evident
while performing DNS of this case, with results shown in Fig. 3.16.
178 3 Receptivity and Instability

Rayleigh integrand Φ for flow over isothermal wedge, K i = 0.04

1.5 (i) ci = 0.01 Edge of


momentum
1 boundary layer
cr = 0.001
cr = 0.01
0.5 cr = 0.1

0
Φ

-0.5
1
0.5
0.0001
-1 00

Φ
Φ

-0.5
0 -1
-1.5 -1
-0.0001 -2
4.5 5 5.5 ~y 6 6.5 7
0 0.1 0.2
y~
y 0.3 0.4 0.5
-2
0 2 y~ 4 6

0
(ii) ci = 0.1
0.0001
-0.1
Φ

-0.0001
-0.2 5 5.5 ~y 6 6.5 7
Φ

-0.3
cr = 0.001
cr = 0.01
cr = 0.1
-0.4

0 2 y~ 4 6

Fig. 3.22 a Rayleigh integrand Φ for flow over isothermal wedge with buoyancy parameter K i =
0.04 for i ci = 0.01 and ii ci = 0.1. b Fjφrtoft integrand Γ F for flow over isothermal wedge with
buoyancy parameter K i = 0.04 for i ci = 0.01 and ii ci = 0.1. [Reproduced from Direct numerical
simulation of transitional mixed convection flows: Viscous and inviscid instability mechanism.
Sengupta et al., Physics of Fluids, 25, 094102 (2013), with the permission of AIP Publishing.]

Rayleigh integrand, Φ has been plotted for flow over an isothermal heated wedge
with K i = 0.04 in Fig. 3.22a for ci = 0.01 and 0.1 corresponding to cr = 0.001;
0.01; 0.1. Existence of multiple zero-crossing points, where Φ = 0, can be seen in
frame (i) of Fig. 3.22a. One notes from the figure that as cr decreases, the first zero-
crossing point moves towards the wall. But for all the values of cr , there also exists
a zero-crossing point around ỹ  5.2, which is closer to the edge of the momentum
boundary layer (approximately at ỹ = 4.54). Thus, the inner zero-crossing point
indicates temporal instability inside the shear layer, and affects the boundary layer
3.8 Receptivity of Mixed Convection Flows 179

Fjφrtoft integrand ΓF for flow over isothermal wedge, K i = 0.04


(i) ci = 0.01
1.5 2

cr = 0.001 1
1 cr = 0.01

ΓF
cr = 0.1 0

0.5 -1
~
ΓF

0 0.1 0.2 0.3 0.4


y
0

0.004
-0.5

ΓF
0

-1
-0.004
3 4
y~ 5

-1.5
0 1 2 ~ 3 4
y

(ii) ci = 0.1

cr = 0.001
0.4
cr = 0.01
cr = 0.1

0.3
ΓF

0.2 0.004
F
Γ

0.1 -0.004
3 4 ~
y 5

0
0 1 2 3 4
y~

Fig. 3.22 (continued)

temporally. The outer zero-crossing point is in the inviscid part of the flow and
indicates an inviscid mechanism of temporal instability. This indication of invis-
cid instability is clearly seen in the inset of frame (i) of Fig. 3.22a. For the higher
value of ci = 0.1 in frame (ii) of Fig. 3.22a, only an outer zero-crossing point exists
around ỹ  5.5. This indicates stronger temporal instability of mixed convection
flows outside the shear layer for all frequencies corresponding to cr < 1.
Fjφrtoft’s integrand, Γ F of Eq. (3.113) is plotted in Fig. 3.22b, as a function of
ỹ for K i = 0.04 (heated isothermal wedge). In calculating Γ F , the mean speed Us
corresponding to the outer zero-crossing point of Fig. 3.22b is used. For cr = 0.01,
we see Γ F < 0 in two ranges, one closer to the wall, while the other is near the edge
of the boundary layer. For higher value of cr = 0.1, Γ F < 0 occurs slightly away
from the wall, as compared to the lower values of cr . Also, Γ F < 0 at similar wide
range of ỹ, for all values of cr shown.
180 3 Receptivity and Instability

In Fig. 3.21a, only one zero-crossing point exists for the three cr ’s shown in the
figures for both the growth rates, unlike the case of isothermal heated wedge. So for
the adiabatic cold plate case, the temporal instabilities occur within the shear layer
by the inviscid mechanism and is dominated by viscous action by the equilibrium
velocity and temperature profiles obtained by the NSE. One notes that Γ F < 0 for
an extended range of ỹ within the shear layer, for all combinations of cr and ci in
Fig. 3.21b. The necessary condition of Γ F < 0 predicts temporally unstable distur-
bances to occur inside the boundary layer. The condition of Γ F < 0 is expected to
be stronger, than the condition given by Φ = 0, as the condition with Γ F provides a
necessary condition by using both the real and imaginary parts of Eq. (3.108).
For a combination of K i / K a and ci , there exists a critical value (cr )crit , such that
for cr > (cr )crit , no zero-crossing point of Φ exists in the flow. For a fixed frequency
excitation, cr is directly proportional to length scale. So critical cr represents a critical
length scale of disturbances lcrit = 2π(cβr )crit , for a given β. Scales larger than lcrit ,
do not suffer temporal instability. There exists positive ci for disturbances with cr <
(cr )crit , which satisfy the necessary condition for temporal instability.
The two inviscid instability theorems help explain the instability in Figs. 3.16
and 3.17, when full NSE is solved. The instability in Fig. 3.16 for the adiabatic
cold plate, displays high wavenumber/ frequency instability at the edge of the shear
layer originating from the leading edge. Additionally, relatively lower wavenumber/
frequency disturbances inside the boundary layer originates at the leading edge. These
are apparent from Fig. 3.21a, b, which show the occurrence of Φ = 0 (according
to Theorem I) closer to the wall for higher cr . As cr = β/k, this implies lower k
component disturbances near the wall, as seen in Fig. 3.16. Similarly, from Fig. 3.21b,
one notes Γ F < 0 occurring for a wide range of heights, with maximum negative
value occurring for ỹ  2, implying maximum inviscid instability at the edge of
the shear layer for all wavenumber components, as evident in Fig. 3.16. Comparing
these, one justifies the noted instability, and the length/ time scales for the case of
isothermal heated wedge, which occurs predominantly in the outer edge of the shear
layer. In Fig. 3.17, we have also drawn a dotted line above which Γ F is negative for
high wavenumbers, as noted in Fig. 3.22b.
In Fig. 3.23, (cr )crit is plotted in (K i , ci )- and (K a , ci )-planes in frames (a) and (b),
respectively. One distinguishes instabilities indicated in frames of Fig. 3.23. For the
flow over isothermal wedge, the line AB represents the (cr )crit values above which
all cr , including the critical value are temporally unstable. Thus the region below AB
indicates instability for all cr less than equal to 1, implying that for higher values of
K i , one gets higher growth rates for larger ranges of length scales. Similarly, for the
region above CD, (cr )crit = 0.0001 everywhere, which experiences instability only
for very small length scales, but with larger growth rates. Similar conclusions are
obtained for adiabatic cold plate shown in Fig. 3.23b, indicates enhanced temporal
instability with cooling rate increased.
3.8 Receptivity of Mixed Convection Flows 181

(a) Isothermal wedge case


D
0.015

Existence of Φ = 0 0.1
for (cr)critical ~ 0.0001
0.25 B
0.01
999
= 0.9
(c r) critical
ci

0.5

0.75
0.005 Existence of Φ = 0
for (cr)critical ~ 1
C

A
0.002 0.004 0.006 0.008 0.01
Ki
(b) Adiabatic cold plate case
0.02

0.75
0.015 Existence of Φ = 0
for (cr)critical ≤ 1 Existence of Φ = 0
for (cr)critical < 0.00001
0.007
ci

0.01

0.001
0.005 0.004
0.005 1E-05
0.003 0.002

-0.001 -0.0008 -0.0006 -0.0004 -0.0002


Ka

Fig. 3.23 Critical value (cr )crit plotted for cases of a flow over isothermal wedge in (K i , ci )-plane
and b for flow over adiabatic cold plate in (K a , ci )-plane. Here, (cr )crit represents the critical phase
speed such that for all cr > (cr )crit , no zero-crossing point of Φ exists in the flow. [Reproduced
from Direct numerical simulation of transitional mixed convection flows: Viscous and inviscid
instability mechanism. Sengupta et al., Physics of Fluids, 25, 094102 (2013), with the permission
of AIP Publishing.]

The necessary condition given by |φ|2 Γ F d ỹ < 0 is a consequence of Theorem


II, and is investigated for the cold adiabatic plate. As |φ|2 is strictly non-negative and
less than one, this condition is reduced to a more conservative inequality given by
IΓ = Γ F d ỹ ≤ 0. In Fig. 3.24, the contours of IΓ in (cr , ci )-plane for the indicated
values of K a are shown. For K a = −0.05, temporal instability is indicated by this
criterion for cr ≤ 0.892. Also for a fixed value of cr , all ci ’s are indicated as possible
in the plotted range. Strength of the instability is indicated by the minimum value
in the domain of interest. It is noted in all the frames, that the maximum instability
is obtained for cr → 0, i.e., for vanishing length scales. Such instabilities have been
noted in Fig. 3.16, for very small length scales near the outer edge, closer to the
leading edge. This criterion depends on the value of Us corresponding to Theorem I,
182 3 Receptivity and Instability

(a) Ka = -0.05 ; max = 0.38 & min = -2.45 (b) Ka = -0.03 ; max = 0.44 & min = -1.83

0.014 0.014

0.012 0.012

0.01 0.01
0 -0.6 0
0.008 -1.8 0.008 -0.2
-2.2 -1.4 -1 -0.6 -0.2
ci

ci
-1.4
0.006 0.006 -1

0.004 0.004

0.002 0.002
0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1
cr cr
(c) Ka = -0.01 ; max = 0.39 & min = -1.45

0.014

0.012

0.01
-1 -0.2
0.008 -0.6
ci

0
0.006-1.4

0.004

0.002
0.2 0.4 0.6 0.8 1
cr

Fig. 3.24 Contours of IΓ = Γ F d ỹ plotted in (cr , ci )-plane for flow past a adiabatic cold plate
with indicated values of buoyancy parameter K a . [Reproduced from Direct numerical simulation
of transitional mixed convection flows: Viscous and inviscid instability mechanism. Sengupta et al.,
Physics of Fluids, 25, 094102 (2013), with the permission of AIP Publishing.]

and is important to use the actual velocity profiles obtained from DNS. The snapshots
in Fig. 3.16 indicate disturbances in the outer layer, which spread in the streamwise
direction. Nonetheless, Theorems I and II are vital aids in explaining instabilities
that prevent one from even obtaining an equilibrium flow.
The role of inviscid instability mechanism in destabilizing the flow is emphasized,
by solving Eq. (3.106) for flows past adiabatic cold plate with (K a = −0.001, −0.05)
and isothermal heated wedge cases with (K i = 0.001, 0.05). Equation (3.106) is inte-
grated using fourth-order Runge-Kutta method from ỹmax = 12 to wall, for a partic-
ular choice of kr eal . Integration of Eq. (3.106) is performed by specifying φ∞ = e−k ỹ

and φ∞ = −ke−k ỹ at ỹ = ỹmax . Direct integration of Eq. (3.106) is possible, as it
does not have any stiffness problem that is inherent with the OSE. While integrat-
ing Eq. (3.106), ci = 0 is avoided, i.e., neutral disturbances, as it would give rise
to solution blow-up near the critical layer. Here, a legitimate eigenvalue would cor-
respond to satisfaction of homogeneous boundary condition at the wall, i.e., when
φr eal + iφimg = 0 at ỹ = 0. Thus in Fig. 3.25, the contours of φr eal = 0 (solid lines)
and φimg = 0 (broken lines) are plotted at the wall, in (cr , ci )-plane, for adiabatic
cold plate cases with indicated buoyancy parameters. Only two values are chosen:
3.8 Receptivity of Mixed Convection Flows 183

(a) Ka = -0.05, kreal = 0.2 (b) Ka = -0.001, kreal = 0.2


ϕreal = 0 ϕreal = 0
ϕimag = 0 ϕimag = 0
0.0004 0.0004

0.0002 0.0002
ci

ci
0 0

-0.0002 -0.0002

-0.0004 -0.0004

0.38 0.39 0.4 0.41 0.42 0.38 0.39 0.4 0.41 0.42
cr cr
(c) Ka = -0.05, kreal = 0.05 (d) Ka = -0.001, kreal = 0.05
0.001 ϕreal = 0 ϕreal = 0
ϕimag = 0 ϕimag = 0
0.0004

0.0005
0.0002
ci

ci

0 0

-0.0002
-0.0005

-0.0004

-0.001
0.38 0.39 0.4 0.41 0.42 0.38 0.39 0.4 0.41 0.42
cr cr

Fig. 3.25 Contours of φr eal = 0 (solid lines) and φimg = 0 (broken lines) plotted at the wall
in (cr , ci )-plane for flows past adiabatic cold plate with indicated buoyancy parameter K a and
disturbance wavenumber kr eal . [Reproduced from Direct numerical simulation of transitional
mixed convection flows: Viscous and inviscid instability mechanism. Sengupta et al., Physics of
Fluids, 25, 094102 (2013), with the permission of AIP Publishing.]

kr eal = 0.2 (frames (a) and (b)) and kr eal = 0.05 (frames (c) and (d)). In this figure,
plotted range are only for 0.38 ≤ cr ≤ 0.42, for ease of understanding. Intersection
of these two curves indicate an eigenvalue corresponding to inviscid instability mech-
anism. One notes from Fig. 3.25 that as one reduces the cooling rate, the number of
inviscid eigenvalues also reduces. The symmetry of the contours in Fig. 3.25, is due
to the fact that, if c = cr + ici is an eigenvalue of Eq. (3.106) with eigenfunction
φ, c∗ = cr − ici is also an eigenvalue of Eq. (3.106) with eigenfunction φ ∗ , where
φ ∗ is complex conjugate of φ. First ten most unstable inviscid eigenvalues are listed
in Table 3.2. One observes from Table 3.2 that for K a = −0.05, lower wavenumber
disturbances are more unstable than higher wavenumber disturbances. However for
lower cooling rate of K a = −0.001, the higher wavenumber disturbances exhibit
184 3 Receptivity and Instability

(a) Ki = 0.05 and kreal = 0.2 (b) Ki = 0.05 and kreal = 0.2 (c) Ki = 0.05 and kreal = 0.2
0.004 ϕreal = 0 0.004 ϕreal = 0 ϕreal = 0
ϕimg = 0 ϕimg = 0 ϕimg = 0
0.002
0.002 0.002
ci

ci

ci
0 0 0

-0.002 -0.002
-0.002

-0.004 -0.004
0.16 0.18 0.2 0.22 0.24 0.3 0.35 0.4 0.45 0.5 0.7 0.72 0.74 0.76 0.78 0.8
cr cr cr
(d) Ki = 0.001 and kreal = 0.2 (e) Ki = 0.001 and kreal = 0.2 (f) Ki = 0.001 and kreal = 0.2
0.004 ϕreal = 0 0.004 ϕreal = 0 0.004 ϕreal = 0
ϕimg = 0 ϕimg = 0 ϕimg = 0
0.002 0.002 0.002
ci

ci

ci
0 0 0

-0.002 -0.002 -0.002

-0.004 -0.004 -0.004


0.16 0.18 0.2 0.22 0.24 0.3 0.35 0.4 0.45 0.5 0.7 0.75 0.8
cr cr cr

Fig. 3.26 Contours of φr eal = 0 (solid lines) and φimg = 0 (broken lines) plotted at the wall in
(cr , ci )-plane for flows past isothermal hot plate with indicated buoyancy parameter K i . We have
chosen disturbance wavenumber kr eal = 0.2. [Reproduced from Direct numerical simulation of
transitional mixed convection flows: Viscous and inviscid instability mechanism. Sengupta et al.,
Physics of Fluids, 25, 094102 (2013), with the permission of AIP Publishing.]

Table 3.2 First ten most dominant unstable modes from inviscid instability analysis tabulated for
flows past adiabatic cold plate with K a = −0.05 and K a = −0.001. We have tabulated only for
disturbance wavenumber of kr eal = 0.2 and 0.05. [Reproduced from Direct numerical simulation
of transitional mixed convection flows: Viscous and inviscid instability mechanism. Sengupta et al.,
Physics of Fluids, 25, 094102 (2013), with the permission of AIP Publishing.]
K a = −0.05 K a = −0.001
Mode No. kr eal = 0.2 kr eal = 0.05 kr eal = 0.2 kr eal = 0.05
1 (0.192, 4.19e-4) (0.332, 4.24e-4) (0.263, 3.41e-4) (0.595 1.51e-4)
Other modes are
not clearly
distinguishable
2 (0.223, 3.84e-4) (0.337, 4.23e-4) (0.289, 2.96e-4)
3 (0.246, 3.57e-4) (0.342, 4.18e-4) (0.411, 2.56e-4)
4 (0.265, 3.29e-4) (0.321, 4.18e-4) (0.421, 2.52e-4)
5 (0.284, 2.94e-4) (0.306, 4.18e-4) (0.312, 2.49e-4)
6 (0.459, 2.58e-4) (0.316, 4.16e-4) (0.362, 2.31e-4)
7 (0.489, 2.58e-4) (0.348, 4.16e-4) (0.331, 2.23e-4)
8 (0.505, 2.55e-4) (0.327, 4.12e-4) (0.348, 2.23e-4)
9 (0.526, 2.48e-4) (0.295, 4.12e-4) (0.348, 2.23e-4)
10 (0.445, 2.45e-4) (0.311, 4.06e-4) (0.377, 2.16e-4)
3.8 Receptivity of Mixed Convection Flows 185

pronounced instability. Figure 3.25 together with Table 3.2, establish unambigu-
ously that cooling destabilizes the flow via the inviscid instability mechanism, which
cannot be explained by the OSE, but by Eq. (3.106). In Fig. 3.26, φr eal = 0 and
φimg = 0 contours are plotted at the wall for flows past isothermal heated wedge
with K i = 0.05 and K i = 0.001. As noted in DNS, inviscid instability mechanism
also shows that heating destabilizes the flow shown in Fig. 3.26 and Table 3.2. There-
fore, for mixed convection flow past a hot plate, destabilization of flow is indicated
simultaneously through viscous and inviscid mechanisms, with the latter dominating
over the former.
Having established the centrality of DNS for mixed convection flow, in the fol-
lowing we return to hydrodynamic instability and use DNS to highlight those aspects
of fluid flow, which depend upon the solution of the full NSE.

3.9 Nonlinear and Nonparallel Effects: Receptivity by DNS

The experimental verification of TS waves in Schubauer and Skramstad [50] heralded


renewed flow instability studies. This experiment identified the region where the
ZPG boundary layer is unstable for high to moderate frequencies. In Fig. 3.27,
the experimental neutral curve is identified by discrete symbols. The right half of
theoretically predicted thumb-shaped region (the neutral curve shown by continuous
line) in the (Reδ∗ , ω0 )-plane matches well with experiment [48]. As the theoretical
analysis is based on parallel flow assumption, it’s results require interpretation for
disturbance propagation in an actual growing boundary layer. It can be shown that a

BAC Linear parallel theory


0.2 Schubauer and Skramstad Experiments
R (F)
0.15
Q
Stable
ωo

0.1
P
Unstable
0.05
Stable
0 S 1000 T
0 2000 3000
Re δ*

Fig. 3.27 Theoretical neutral curve in (Reδ ∗ , ω0 )-plane. According to linear theory with parallel
flow assumption, three frequencies shown corresponding to OA, OB and OC are all stable. The
discrete symbols are the experimental data from Schubauer and Skramstad [50]
186 3 Receptivity and Instability

constant physical frequency excitation follows a ray starting from the origin, as shown
in Fig. 3.27. In this figure, three rays A, B and C correspond to the nondimensional
frequencies of excitation F = 3.0 × 10−4 , 3.5 × 10−4 and 2.5 × 10−4 , respectively.
It is noted in Fig. 3.27, that the experimental results of [50] reveal discrepancy at
(lower Reδ∗ - high ω0 ) combinations. Experimental critical Reδ∗ is significantly lower
from the linear instability theory value. In all the experimental cases, the boundary
layer was presumably excited at same physical location, although it is not recorded in
the paper. According to linear instability theory, the three frequencies corresponding
to A, B and C in Fig. 3.27, create disturbances those are stable, while the experimental
results indicate a finite length where the TS wave is unstable.
While the experiment qualitatively verified central features of instability theory,
there are features in the experiment that instability theory is incapable of explaining.
One realizes that the stability theory for a parallel flow model predicts the system
response with respect to small perturbations that is only valid at a distance far away
from the leading edge of the plate (x → ∞). The created waves (either spatially
stable or unstable) are associated with wavenumbers, very close to the origin of
the complex wavenumber plane. This was clearly explained in [51], with the help of
Abel’s and Tauber’s theorems given in [74]. These relate the poles (eigenvalues) near
the origin of the complex wavenumber plane with the asymptotic behaviour in the
physical plane. Similarly, the essential singularity of the Fourier-Laplace transform
in the wavenumber plane determines the local solution, in the vicinity of the exciter
in the physical plane as explained in Sect. 3.3.2.
In this section, nonparallel and nonlinear features of instability of Blasius bound-
ary layer is shown in detail, obtained with the help of high accuracy DNS. To calculate
the receptivity of a ZPG boundary layer, NSE has been solved in [22]. The exciter is
placed at a location close to the leading edge of the plate, where nonparallelism of
the flow is dominant. SBS exciter used in [22, 58] has amplitudes of the excitation
varying from a very small to large values (0.2–10% of U∞ ). The SBS exciter causes
vortical excitation at the wall. The frequency of excitation is varied from a very stable
(according to linear spatial theory) to moderate value, for the purpose of quantifying
the nonlinear and nonparallel effects to resolve conflicting claims in the literature
about the critical Reynolds number of the Blasius profile [22, 23, 46] with the help of
DNS, avoiding restrictive assumptions of signal problem [57] and recording effects
of nonparallelism and nonlinearity directly.
In Fig. 3.28, the input spectrum of the SBS strip exciter is shown for α1 = 0.05
and 0.10 for the exciter located between x1 = 0.22 (Reδ∗ = 255) and x2 = 0.264
(Reδ∗ = 284). The spectrum shows the wide range of wavenumbers excited by the
exciter with wide band. The response of the fluid dynamical system depends upon
its inherent transfer function.
3.9 Nonlinear and Nonparallel Effects: Receptivity by DNS 187

0.8000

0.6000 α1 = 0.05
α1 = 0.1
F(k)

0.4000

0.2000

0.0000
0 200 400 k 600 800 1000

Fig. 3.28 Fourier transform of the input disturbance by the SBS strip given by Eqs. (3.100) and
(3.101) for the amplitudes of excitation α1 = 0.05 and 0.1. The exciter is located from x1 = 0.22
to x2 = 0.264

3.9.1 Computational Domain and Grid Resolution

Particular focus here is on non-dimensional frequency of F = 3.0 × 10−4 on the


asymptotic solution. Effects of higher frequency excitations (for F ≥ 3.0 × 10−4 )
and various amplitudes of excitation cases are also investigated. Focus of present
section is solely on TS wave-packet and not on the STWF, for which one requires a
much longer domain. Results here employs a stretched grid in x- and y-directions
having 1600 and 600 points, respectively. The schematic diagram of the problem is
shown in Fig. 3.13, with γ = 0. All simulations here include the leading edge of
the plate as explained before. The reference velocity and length scales are chosen
to be U∞ and L, respectively. The Reynolds number for all simulations are for
Re L = U∞ν L = 105 and the computational domain extends from −0.05L to 10L in
the streamwise, and 0 to L in the wall-normal directions. For this computational
domain, L comes out to be approximately as 60δ ∗D , where δ ∗D is the displacement
thickness at the outflow of the computational domain (i.e., at x = 10L).
Tangent hyperbolic grid stretching is adopted in both streamwise and wall-normal
direction, as given by equations following Eq. (3.73), to cluster points near the
leading–edge and near the wall of the plate, respectively [21]. The grid mapping is
selected in such a way that the wall-resolution is given by: Δy0 = 2.455 × 10−4 and
the smallest grid size in the streamwise direction is given by: Δx0 = 9.218 × 10−4 .

3.9.2 The Equilibrium Flow

In Fig. 3.29, the similarity functions f  (η̄), f  (η̄) and f  (η̄) corresponding to the
equilibrium flow as obtained by DNS is plotted  in frames (a), (b) and (c), respectively,
as a function of the similarity variable η̄ = y Uνx∞ , at indicated streamwise locations.
The similarity functions f  (η̄), f  (η̄) and f  (η̄) obtained from directly integrating
188 3 Receptivity and Instability

1
(a)

0.8 x = 0.1
x = 0.2
0.6 x = 0.5
x = 0.8
f′ x = 1.0
0.4 x = 2.0
x = 5.0
0.2 x = 9.0
Theoritical

0 −
0 2 η 4 6 8

0.4
(b)

0.3 x = 0.1
x = 0.2
x = 0.5
f ′′

x = 0.8
0.2 x = 1.0
x = 2.0
x = 5.0
0.1 x = 9.0
Theoritical

0 −
0 2 η 4 6 8

(c)
0

-0.02 x = 0.1
x = 0.2
-0.04 x = 0.5
f ′′′

x = 0.8
-0.06 x = 1.0
x = 2.0
-0.08 x = 5.0
x = 9.0
-0.1 Theoritical

-0.12 −
2 η 4 6 8

Fig. 3.29 The similarity functions a f  (η̄), b f  (η̄) and c f  (η̄) to the equilibrium flow obtained
by DNS, plotted as a function of the similarity variable η̄, at indicated streamwise locations. The
corresponding solution of the Blasius equation is also shown for comparison

the Blasius equation f  + 21 f f  = 0, is also plotted in the respective frames for
comparison. The functions f  (η̄) and f  (η̄) show excellent match for all the values
of η and all the streamwise locations shown. The function f  , however, shows slight
departure from that predicted by the similarity solution for very small values of η.
This discrepancy is an outcome of the shortcomings of the similarity assumption.
One should note that for stability calculations, f  is very important, as it represents
2
the term ddyu2 , which appears in OSE given in Eq. (3.14). However, it is noted that
the results here are obtained by solving NSE and is not restricted by any limiting
assumptions.
3.9 Nonlinear and Nonparallel Effects: Receptivity by DNS 189

ud at t=8; α1= 0.1; F = 3.0 × 10


-4
0.06

0.04 local soln.

0.02 Asymptotic soln.


wave front
ud 0

-0.02

-0.04

-0.06
0 1 2 3 4
x
Fig. 3.30 Streamwise disturbance velocity u d obtained as a receptivity solution at t = 8, for y =
0.008, identifying the local, asymptotic and wave-front components due to a strip excitation near
the leading edge. The frequency of excitation of the SBS strip on the wall is F = 3.0 × 10−4

3.9.3 Typical Morphology of the Disturbance Field

A typical receptivity solution to localized wall excitation at a fixed frequency is


shown in Fig. 3.30 at t = 8, with the solution obtained by solving NSE for the fre-
quency corresponding to OA in Fig. 3.27. Here, t = 0 corresponds to the onset of
the excitation. According to linear theory, this solution should have been completely
damped, but the presented solution in Fig. 3.30, at a particular time indicate a stream-
wise stretch where the solution first grows and subsequently decays. The height at
which this solution is recorded corresponds to y = 0.008. This is a point very close
to the wall and all distances are non-dimensionalized.
The structure shown in Fig. 3.30 is generic and has distinct features, some of
which are also shared by the solution of the problem obtained using the OSE in [62].
Solutions in [52, 57] correspond to unstable cases and the leading STWF of Fig. 3.30
was not distinctly noted. However in [62], response field for spatially stable cases dis-
played a distinct STWF, different from the following TS waves, as seen in Fig. 3.30.
To understand the various components of the response shown in Fig. 3.30, the Fourier
transform of the response is shown plotted in Fig. 3.31. In this figure, the main peak
corresponds to the wave-packet representing the TS waves. As shown in [57, 62],
apart from the TS mode, there are only a few additional modes with very high spatial
damping rates and these are also noted among the different peaks of Fig. 3.31. The
exciter is located very near to the leading edge of the flat plate, and the computed
solution in Fig. 3.30 shows large peaks in the physical plane for the local solution and
the asymptotic solution emerges from it in a continuous manner. This local solution
corresponds to the marked distant peak in the legend of Fig. 3.31. The wave front
corresponds to the local maximum close to the origin in Fig. 3.31. We also note that
the solution of NSE is for the excitation, for which the linearized solution corre-
sponds to damped TS waves. When such an excitation is applied, the local and the
asymptotic solution remain stationary in space, at the location depicted in Fig. 3.30.
However, the STWF continuously propagate downstream.
190 3 Receptivity and Instability

A : peak due to wave front


1 B : peak due to the local solution
10
C : peak due to the local solution
D D : peak due to the asymptotic solution
E : 1st superharmonic of local solution
st
F : 1 superharmonic of asymptotoc solution
G : 2nd superharmonic of the asymptotic solution
10
0 C
A EF
Ud (k)
B

G
-1
10

10-2
50 100 150 200 250
k

Fig. 3.31 Fourier transform of the receptivity solution shown in Fig. 3.30, identifying the
wavenumbers of the local, asymptotic and wave front components in the legend

3.9.4 Receptivity to SBS Excitation: Nonlinear and


Nonparallel Effects

In this section, nonlinear and nonparallel effects on the receptivity of the Blasius
boundary layer is studied, when the flow is harmonically excited by SBS exciter with
F = 3 × 10−4 , α1 = 0.05. The exciter is located between x1 = 0.22 and x2 = 0.264.
In Fig. 3.32, the streamwise component of the disturbance velocity u d is plotted as a
function of streamwise co-ordinate x, at indicated heights for this case. In showing
the signal in the left side frames of Fig. 3.32, the origin refers to the leading edge
of the flat plate. From the solution of Eqs. (3.60) and (3.61) without the Boussinesq
term, one obtains instantaneous values of vorticity and velocity components. The
disturbance quantities are obtained (as displayed in Fig. 3.32) by subtracting the
mean from the total quantity. This corresponds to a spatially stable case in linear
theory.
The displayed variations of u d with x, at different heights over the plate indicate
an interesting phenomenon. This solution corresponds to the time t = 20 after the
onset of excitation. While the linear normal mode analysis, predicts this flow field
to be stable, one notices that barring the top and bottom-most left-side frames in
Fig. 3.32, the disturbances actually grow spatially at intermediate heights (frames
(c), (e) and (g)). All the frames clearly indicate the presence of multiple modes which
are also directly evident from the multiple peaks of the Fourier transform of these
signals shown in the right-side frames of Fig. 3.32. The Fourier amplitudes show the
dominant normal mode near the wall, while multiple dominant modes are seen in
the third and fourth frames showing the transform. Also, one notices the first super
harmonic of the asymptotic solution in the right side frames of Fig. 3.32, which is
clearly shown in Fig. 3.31.
3.9 Nonlinear and Nonparallel Effects: Receptivity by DNS 191

0.05 (a) km (b)


101
km = 70.65

Ud (k)
ud
0
0 10

10-1
-0.05
0 1 X 2 3 50 k 100 150
(c) km (d)
0.02 101
km = 70.65

Ud (k)
ud

0
0 10

-0.02 10
-1

0 1 X 2 3 50 k 100 150

(e) (f)
0.02 10
1
k1 k2 k1 = 68.81
Ud (k) k2 = 74.96
ud

0
0 10

-1
-0.02 10
0 1 X 2 3 50 k 100 150

(g) (h)
101 k1 k2
0.02 k1 = 67.58
Ud (k)

k2 = 73.53
ud

0
0 10

-0.02 10-1
0 1 X 2 3 50 k 100 150
(i) 1
km (j)
0.02 10
km = 70.65
Ud (k)
ud

0 100

-0.02 10
-1

0 1 X 2 3 50 k 100 150

Fig. 3.32 u d plotted as a function of x for t = 20 at a y = 0.00521, c 0.00662, e 0.00808, g


0.00958 and i 0.01137. The signals are for the SBS strip excitation amplitude of α1 = 0.05, and non-
dimensional frequency, F = 3.0 × 10−4 with exciter location between x = 0.22 and x = 0.264.
Fourier transform of the signals shown in right frames

Frames (c), (e) and (g) in Fig. 3.32, clearly show different streamwise stretches
at different heights, over which the disturbance actually grows with x. The Fourier
transform of the signal in frames (b), (d) and (j) display a dominant peak at k = km =
70.65. From the linear stability analysis at x = 0.3 (Reδ∗ = 298) corresponding to
the present excitation frequency, one should have a wave with wavenumber α =
0.225. Here, the wavenumber α is based on the local δ ∗ , whereas km is based on the
reference length scale L. The dominant √ wavenumber of the signal based on δ ∗ and
obtained from DNS is given as αm = 1.72√xδ∗ and at x = 0.3, αm = 0.21. One also
km Re

notices, the presence of two adjacent modes at k = k1 and k = k2 , for the signals at
192 3 Receptivity and Instability

y = 0.00808 and y = 0.00958. The lower mode at k = k1 corresponds to a mode


that moves downstream at a faster speed than the mode at k = k2 . It is also noted
that k1 + k2  km . The presence of the mode at k = k2 becomes weaker as the height
increases. However, there is another damped mode that dominates with increasing
height over the plate. It is noted that in earlier nonparallel studies (as referred to
in [22]), various investigators have tried to explain experimental data that indicated
instability at particular heights for the streamwise disturbance velocity components.
It has been noted in [57, 61, 62] that there are no particular physical reasons as to
why one has to adopt either spatial or temporal framework in studying boundary layer
stability. This is also readily apparent from the information in Figs. 3.32 and 3.33.
An interesting aspect of the computed results is noted with respect to time variation-
as shown in the left frames of Fig. 3.33 for signals collected at a distance of x = 0.3
from the leading edge, at the indicated heights. Despite the fact that the fluid dynamic
system is excited at a single constant frequency (ω̄0 ), the Fourier transform in the
right frames of Fig. 3.33 indicates presence of superharmonics at all heights and in
different proportions. From Fig. 3.27 one clearly notes the corresponding normal
modes to be more stable from linear stability point of view for these additional
superharmonics. Of specific interest is the data for y = 0.00662 and 0.00808, which
show the presence of strong second and third harmonics.
In Fig. 3.32 also, we noted most pronounced instability at these heights. Thus,
the disturbance growth at these heights must be related to nonlinear effect and is
experienced at intermediate heights only.
Apart from the fact that the growth is noted at intermediate heights, we also note
similarity of flow field at heights very close to the wall, and those in the outer edge
of the boundary layer. This aspect is depicted in Fig. 3.34, where spatial variation of
the streamwise disturbance velocity is shown for heights those are either very near
the wall or are in the outer part of the shear layer. Corresponding Fourier transform
of these signals are shown in the right frames of Fig. 3.34. At these heights, signals
display a range of x for which they remain neutral, while in the outer part, the signals
decay with streamwise distance.
Thus, Figs. 3.32, 3.33 and 3.34 show that this particular case represents nonlinear
receptivity at intermediate heights; neutral stability near the wall and in the outer
part of the boundary layer.
Presence of growing solution at intermediate heights is further investigated by
plotting the mean and disturbance component of streamwise velocity (obtained from
the DNS) with height at x = 0.3 (close to the exciter), in Fig. 3.35a. In the same figure,
the band of heights at which the direct simulation indicates growth is identified by
drawing two horizontal lines 1a and 1b. It is already noted that there are multiple
spatial modes (including the normal mode, identified as the TS mode in the present
figure) and superharmonics of the excitation frequency present in the simulated cases
(due to nonlinearity). Their effects are different at different heights, as shown in
Figs. 3.32, 3.33 and 3.34. For example in Fig. 3.33, one can clearly note the presence
of dominant superharmonics at y = 0.00662 and 0.00808. While these are also noted
at different heights, and play secondary roles there.
3.9 Nonlinear and Nonparallel Effects: Receptivity by DNS 193

(a) 103 (b)


1
ud / |ud| max 2

Ud (f)
10
0
101
-1
20 20.2 20.4 20.6 20.8 21 100
t 5 10 f 15 20
(c) 103 (d)
1
ud / |ud| max

102

Ud (f)
0
1
10
-1
20 20.2 20.4 20.6 20.8 21 100
t 5 10 f 15 20
(e) 10
3 (f)
1
ud / |ud| max

102
Ud (f)
0
101
-1
20 20.2 20.4 20.6 20.8 21 100
t 5 10 f 15 20
(g) 10 3 (h)
1
ud / |ud| max

2
Ud (f)

10
0
101
-1
20 20.2 20.4 20.6 20.8 21 100
t 5 10 f 15 20
(i) 103 (j)
1
ud / |ud| max

102
Ud (f)

0
101
-1
20 20.2 20.4 20.6 20.8 21 100
t 5 10 f 15 20

Fig. 3.33 u d /|u d |max plotted as a function of time for x = 0.3 at a y = 0.00385, c 0.00662,
e 0.00808, g 0.00958 and i 0.0144. The figure is shown for the SBS strip excitation amplitude
of α1 = 0.05 and non-dimensional frequency F = 3.0 × 10−4 with exciter location between x1 =
0.22 and x2 = 0.264. Fourier transforms plotted for the signals in right frames. Here, |u d |max is the
maximum absolute value of u d for 20 ≤ t ≤ 21 at the corresponding (x, y)-location

In discussing this aspect of Fig. 3.33, we reason that these are essentially nonlinear
and nonparallel effects. In Fig. 3.35a, the intermediate height of maximum TS wave-
packet growth is attempted to be linked with critical layers of the first and second
TS mode and their superharmonics. From the inviscid stability analysis [19, 28], the
ω̄
height where U (y) = α0jj is known to be the critical layer, where α j and ω̄0 j are
the wavenumber and circular frequency corresponding to the jth TS mode. These
critical layers for the TS and the second modes are identified in Fig. 3.35a by lines 2–5,
respectively. It is seen that the maximum growth height is located closer to the critical
194 3 Receptivity and Instability

0.1 (a) km (b)


1
10
km = 70.65

Ud (k)
ud
0 10
0

-1
-0.1 10
0 1 X 2 3 50 k 100 150
(c) km (d)
0.1 101
km = 70.65

Ud (k)
ud

0 100

-0.1
10-1
0 1 X 2 3 50 k 100 150

(e) km (f)
0.1 101
km = 70.65
Ud (k)
ud

0 100

-0.1
10-1
0 1 X 2 3 50 k 100 150

(g) km (h)
0.005 100
km = 7065
Ud (k)
ud

0 10-1
-0.005 -2
10
0 1 X 2 3 50 k 100 150

(i) km (j)
0.005 100
km = 70.65
Ud (k)

10-1
ud

0
10-2
-0.005
10-3
0 1 X 2 3 50 k 100 150

Fig. 3.34 u d plotted as a function of x for t = 20 at a y = 0.00124, c 0.00252, e 0.00385, g 0.03


and i 0.032. The ZPG flow past a flat-plate is excited by a SBS strip located between x1 = 0.22 and
x2 = 0.264 with α1 = 0.05 and the frequency F = 3.0 × 10−4 . Fourier transform of the signals
are shown in right frames

heights of the TS and second modes, corresponding to the first superharmonics,


at which the linearized convection terms disappear in the governing OSE for the
TS mode-fundamental frequency combination. However, this intermediate height is
closest to the point where u d variation with respect to y changes sign. This point is
further substantiated in Fig. 3.35b, where u d is plotted as a function of y at indicated
streamwise locations at time-instants, when corresponding u d at inner maxima attains
its maximum value. The range of y, where spatial growth of the TS wave-packet is
observed, is also marked in the figure. It can be clearly seen that this range of y lies
in a zone, where u d changes sign. The location where u d changes sign is also the
3.9 Nonlinear and Nonparallel Effects: Receptivity by DNS 195

location where absolute value of the disturbance normal velocity vd is maximum.


This can be explained from linear stability theory with parallel flow approximation, as
u d = φ  ( ỹ)ei(α x̃−ω0 t) dα and vd = φ( ỹ)ei(α x̃−ω0 t) dα. Hence, |φ( ỹ)| is maximum
at the height ỹ where, φ  ( ỹ) = 0. For actual nonparallel flow, though this analysis
does not exactly hold good, but even then vd attains its maxima near about the y-
location, where u d changes sign. This can be noted clearly from Fig. 3.35c, where vd
is plotted as a function of y at indicated x-locations at time instants, when its peak
value is maximum. The conclusions that one makes from the above discussion are
enumerated below.
(1) One obtains spatial growth of disturbance streamwise velocity component at the
intermediate heights even for a non-dimensional frequency of excitation, F, that does
not intersect the neutral curve.
(2) This range of intermediate heights is located around the point, where u d changes
sign and absolute value of vd is maximum.
(3) The dominant presence of second superharmonic of the excitation frequency ω̄0
is noted at this range of intermediate heights, where spatial growth of u d is noted.
This feature is shown in Fig. 3.33.
It should be pointed out that, such height–dependent variation of the neutral points
were also observed in [22], where “an ambiguous behaviour” of u d was noted to exists
as the height approaches “the location of the 180◦ phase shift of the u d eigenfunction”.
No detailed analysis of this height dependent behaviour was presented in [22].

3.9.5 Effects of Nonlinearity: Role of Amplitude of Excitation

Here, the role of the amplitude of excitation α1 is investigated for the exciter located
between 0.22 and 0.264. The non-dimensional excitation frequency is F = 3.0 ×
10−4 . The amplitude parameter α1 is varied from a very small value of 0.002 (i.e.,
excitation amplitude to be 0.2% of U∞ ) to a high value of 0.1 (excitation amplitude
is 10% of U∞ ). The variable xex is used subsequently to denote the center of the
exciter. In Sect. 3.9.6, various amplitudes of excitation results are compared for α1 =
0.002, 0.005 and 0.01. When α1 = 0.1, one notices the induction of micro-separation
bubbles on the wall, which is discussed in Sect. 3.9.7.

3.9.6 Comparison of Results for α1 = 0.002, 0.005 and 0.01

A more elaborate comparison is shown in Fig. 3.36 where û d corresponding to


the local and asymptotic solution are plotted as a function of x at ten indicated
heights for (a) α1 = 0.002, (b) α1 = 0.005 and (c) α1 = 0.01, with exciter located
between 0.22 and 0.264 and frequency of excitation is F = 3.0 × 10−4 . In Fig. 3.36
normalization of u d is done with respect to the α1 = 0.01 case as û d (x, y, t; α1 =
196 3 Receptivity and Instability

(a)
1) Maximum growth height is noted in between line 1a and 1b
2) Critical layer for the superharmonic of TS mode
0.014 3) Critical layer for for superharmonic of 2nd mode
4) Critical layer for TS mode
5) Critical layer for the 2nd mode
0.012

0.01
1b
0.008
y

0.006 1a
2
3
0.004 ud
4 U
0.002 5

0
0 0.2 0.4 0.6 0.8 1
U
(b)
x = 0.27
0.02 y range of x = 0.30
maximum growth x = 0.35
x = 0.40
x = 0.45
1a 1b
0.01 x = 0.50
x = 1.00
ud

0 0.02 0.04 0.06


y
(c) x = 0.27
0.006 y range of x = 0.30
maximum growth x = 0.35
0.004 x = 0.40
x = 0.45
0.002 1a x = 0.50
1b x = 1.00
vd

-0.002

-0.004

-0.006
0 0.02 y 0.04 0.06

Fig. 3.35 a Mean and disturbance components of streamwise velocity shown along with estimates
of critical layer for the second spatial mode. b u d and c vd plotted as function of y for indicated
streamwise locations at time-instants when corresponding u d at inner maxima attains its maximum
value. In all the three frames, the lines 1a and 1b indicates the range of heights where maximum
local spatial growth is noted. Presented data are for the SBS excitation with α1 = 0.05 and F =
3.0 × 10−4 with exciter location between 0.22 and 0.264

α¯1 ) = (0.01/α¯1 )u d (x, y, t; α1 = α¯1 ). The results shown are for the solution obtained
at t = 20. The ten heights shown are the 15, 20, 25, 30, 35, 40, 65, 75, 85, and 95th
point of the grid in the wall-normal (y) direction. From these plots, one can find
3.9 Nonlinear and Nonparallel Effects: Receptivity by DNS 197

0.01
(a) α1 = 0.002 0.01
(b) α1 = 0.005 0.01
(c) α1 = 0.01
0 0 0
ud

ud

ud
-0.01 y = 0.00384 -0.01 y = 0.00384 -0.01 y = 0.00384
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
x x x
0.01 0.01 0.01
0.005
α1 = 0.002 0.005
α1 = 0.005 0.005
α1 = 0.01
0 0 0
ud

ud

ud
-0.005 -0.005 -0.005
y = 0.00521 y = 0.00521 y = 0.00521
-0.01 -0.01 -0.01
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
x x x
0.005 0.005 0.005
α1 = 0.002 α1 = 0.005 α1 = 0.01
0 0 0
ud

ud

ud
y = 0.00662 y = 0.00662 y = 0.00662
-0.005 -0.005 -0.005
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
x x x
0.005 0.005 0.005
α1 = 0.002 α1 = 0.005 α1 = 0.01
0 0 0
ud

ud

ud
y = 0.00808 y = 0.00808 y = 0.00808
-0.005 -0.005 -0.005
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
x x x
0.005
α1 = 0.002 0.005
α1 = 0.005 0.005
α1 = 0.01
0 0 0
ud

ud

ud
y = 0.00958 y = 0.00958 y = 0.00958
-0.005 -0.005 -0.005
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
x x x
0.005
α1 = 0.002 0.005
α1 = 0.005 0.005
α1 = 0.01
0 0 0
ud

ud
d
u

y = 0.01137 y = 0.01137 y = 0.01137


-0.005 -0.005 -0.005
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
x x x

α1 = 0.002 α1 = 0.005 α1 = 0.01


0.002 0.002 0.002

0 0 0
d

d
ud

y = 0.0196 y = 0.0196 y = 0.0196


-0.002 -0.002 -0.002
0 0.5 1 x 1.5 2 2.5 3 0 0.5 1 x 1.5 2 2.5 3 0 0.5 1 x 1.5 2 2.5 3

0.002
α1 = 0.002 0.002
α1 = 0.005 0.002
α1 = 0.01

0 0 0
d

ud
u

y = 0.0235 y = 0.0235 y = 0.0235


-0.002 -0.002 -0.002
0 0.5 1 x 1.5 2 2.5 3 0 0.5 1
x 1.5 2 2.5 3 0 0.5 1
x 1.5 2 2.5 3

α1 = 0.002 α1 = 0.005 α1 = 0.01


0.001 0.001 0.001

0 0 0
d
d

d
u
u

-0.001
y = 0.0275 -0.001
y = 0.0275 -0.001
y = 0.0275

0 0.5 1 x 1.5 2 2.5 3 0 0.5 1


x 1.5 2 2.5 3 0 0.5 1
x 1.5 2 2.5 3

0.001 α1 = 0.002 0.001 α1 = 0.005 0.001 α1 = 0.01

0 0 0
d

d
d

u
u

y = 0.0318 y = 0.0318 y = 0.0318


-0.001 -0.001 -0.001
0 0.5 1 x 1.5 2 2.5 3 0 0.5 1
x 1.5 2 2.5 3 0 0.5 1
x 1.5 2 2.5 3

Fig. 3.36 Normalized disturbance streamwise velocity u d at t = 20 and indicated heights plotted
of x for a α1 = 0.002, b α1 = 0.005 and c α1 = 0.01. The normalization is done with respect
to α1 = 0.01 case. The exciter is located between 0.22 and 0.264 which excites the flow with
F = 3.0 × 10−4
198 3 Receptivity and Instability

the effects of nonlinearity and nonparallelism on the TS wave-packet, obtained from


DNS of the receptivity problem. One notices from Fig. 3.36 that at y = 0.00348 and
0.00521, u d for α1 = 0.002 displays monotonic decay with respect to x. However, the
α1 = 0.01 case displays near-neutral behaviour at these heights for 0.6 ≤ x ≤ 0.75
and α1 = 0.005 case shows reduced decay rate up to x ≤ 0.75. At the same time one
observes from Figs. 3.32 and 3.34, marginal instability up to x ≤ 1.0 for α1 = 0.05
for y = 0.00348 and 0.00521. Since, this marginal instability shows up for increased
amplitudes of excitation, hence one can definitely say that these effects are completely
due to nonlinearity of the governing equation. One also observes that the stretch of
the perturbed region also increases with the increase in amplitude of excitation.
One can also observe that the asymptotic solution displays a distinct spatial growth
phase at y = 0.00662, 0.00808 and 0.00958 for all α1 cases plotted here. For α1 =
0.002, this spatial growth of u d at y = 0.00958 is not distinctly identifiable from
Fig. 3.36. Spatial growth of u d correspond to TS wave-packet, at those heights, where
it is also noted for α1 = 0.05 in Fig. 3.32. However, the extent of the spatial growth
of the asymptotic solution reduces with the reduction in amplitude of the excitation.
It is the downstream end, where the growth concludes, and moves upstream with the
reduction in excitation amplitude. The upstream end of the onset of growth is located
almost at the same position for α1 cases studied.
The heights y = 0.0196, 0.0235, 0.0275 and 0.0318 are at the near proximity of the
edge of the boundary layer profile close to the exciter location. The features are similar
to the behaviour of u d , which are obtained very close to the wall (y ≤ 0.00521). One
notices a zone where disturbances evolve near-neutrally for α1 = 0.005 (0.6 ≤ x ≤
0.8). Mild growth for α1 = 0.01 in this streamwise extent can also be noted from
this figure. One also observes perceptible spatial growth of u d for α1 = 0.05 at these
heights from Fig. 3.34. However, any such streamwise zone of near-neutral behaviour
or growth of u d is clearly absent for α1 = 0.002, for which u d strictly decays at these
heights, are shown in Fig. 3.36.
In Fig. 3.37, time variation of u d is plotted at x = 0.35, for 20 ≤ t ≤ 21 and
indicated values of α1 . In this figure, u d is normalized with respect to its maximum
absolute value |u d |max . Figure 3.37 shows the dominant presence of second super-
harmonics of the excitation frequency, i.e., 3ω̄0 , even for α1 = 0.002 at y = 0.00662.
In contrast, time variation of u d at other heights for α1 = 0.002 display an almost
sinusoidal variation indicating negligible presence of superharmonics. It should be
noted that y = 0.00662 is the height, where maximum spatial growth of the asymp-
totic solution is noted in Fig. 3.36, even for α1 = 0.002. This feature of the dominant
presence of second superharmonic at around the height of y = 0.00662 is also noted
for α1 = 0.005 and α1 = 0.01 in Fig. 3.37b, c, respectively. This feature is also noted
for α1 = 0.05 and is already shown in Fig. 3.33. This was already noted for a height
very close to the point, where u d and vd variation with respect to y changes sign and
attains its maximum value, respectively, in Fig. 3.35. Similar features are also noted
here in Fig. 3.37 even for the smallest amplitude of excitation of α1 = 0.002.
3.9 Nonlinear and Nonparallel Effects: Receptivity by DNS 199

(a) α1=0.002 y = 0.00384 (b) α1=0.005 y = 0.00384 (c) α1=0.01 y = 0.00384


1 1 1
ud / |ud| max

ud / |ud| max

ud / |ud| max
0.5 0.5 0.5
0 0 0
-0.5 -0.5 -0.5
-1 -1 -1
20 20.2 20.4 t 20.6 20.8 21 20 20.2 20.4 t 20.6 20.8 21 20 20.2 20.4 t 20.6 20.8 21

y = 0.00521 y = 0.00521 y = 0.00521


1 1 1

ud / |ud| max

ud / |ud| max
ud / |ud| max

0.5 0.5 0.5


0 0 0
-0.5 -0.5 -0.5
-1 -1 -1
20 20.2 20.4 t 20.6 20.8 21 20 20.2 20.4 t 20.6 20.8 21 20 20.2 20.4 t 20.6 20.8 21

y = 0.00662 y = 0.00662 y = 0.00662


1 1 1

ud / |ud| max

ud / |ud| max
ud / |ud| max

0.5 0.5 0.5


0 0 0
-0.5 -0.5 -0.5
-1 -1 -1
20 20.2 20.4 t 20.6 20.8 21 20 20.2 20.4 t 20.6 20.8 21 20 20.2 20.4 t 20.6 20.8 21

y = 0.00808 y = 0.00808 y = 0.00808


1 1 1
ud / |ud| max

ud / |ud| max
ud / |ud| max

0.5 0.5 0.5


0 0 0
-0.5 -0.5 -0.5
-1 -1 -1
20 20.2 20.4 t 20.6 20.8 21 20 20.2 20.4 t 20.6 20.8 21 20 20.2 20.4 t 20.6 20.8 21

y = 0.00958 y = 0.00958 y = 0.00958


1 1 1
ud / |ud| max

ud / |ud| max
ud / |ud| max

0.5 0.5 0.5


0 0 0
-0.5 -0.5 -0.5
-1 -1 -1
20 20.2 20.4 t 20.6 20.8 21 20 20.2 20.4 t 20.6 20.8 21 20 20.2 20.4 t 20.6 20.8 21

y = 0.01137 y = 0.01137 y = 0.01137


1 1 1
ud / |ud| max

ud / |ud| max
ud / |ud| max

0.5 0.5 0.5


0 0 0
-0.5 -0.5 -0.5
-1 -1 -1
20 20.2 20.4 t 20.6 20.8 21 20 20.2 20.4 t 20.6 20.8 21 20 20.2 20.4 t 20.6 20.8 21

Fig. 3.37 u d /|u d |max plotted as a function of time for 20 ≤ t ≤ 21 and x = 0.35 at indicated
heights for a α1 = 0.002, b α1 = 0.005, c α1 = 0.01. The exciter is located between 0.22 and
0.264 and F = 3.0 × 10−4 . |u d |max is the maximum absolute value of u d for 20 ≤ t ≤ 21 at
corresponding (x, y)-location

3.9.7 Results for α1 = 0.1

To understand the role of the amplitude of excitation of the SBS strip, another case
is computed with identical excitation frequency and location, but with excitation
amplitude of α1 = 0.1, i.e., the excitation amplitude is 10% of U∞ . This excitation
amplitude causes induction of micro-separation near the exciter on the plate. Resul-
tant unsteady separation bubbles are confined to small height at the wall - as shown
in the stream function contours in Fig. 3.38 at t = 20. Only a few representative con-
tours have been shown plotted in this figure, which clearly show the micro-bubbles
near the exciter up to x ≤ 1.6 only. Effects of the present bubbles are noted in other
contours shown at other heights.
To understand the dynamics of this case better, instantaneous snapshot at t = 20
is shown for u d variation with x at three heights in the left frames of Fig. 3.39. In the
frames on the right hand side, corresponding Fourier transform is shown plotted to
help identify various components of the solution in the physical space. Presence of
separation bubbles causes fluid particles outside the bubbles to traverse in the wall-
normal direction more than in the streamwise direction, and this is reflected in the
lower value of u d for the higher α1 case, as can be ascertained by comparing Figs. 3.39
200 3 Receptivity and Instability

0.003
ψ - contour

0.002
y

0.001
Bubbles

0.000
0 1 2 3
X

Fig. 3.38 ψ-contours shown in the physical plane for t = 20 for the case of α1 = 0.10; F =
3.0 × 10−4 located between x1 = 0.22 and x2 = 0.264. Note the micro-bubbles near the exciter on
the plate

with 3.32 and 3.34. Also, these micro-bubbles interfere with the local solution that
is clearly seen in the Fourier transform of Fig. 3.39b, as compared to that shown
in Fig. 3.32, for the peak corresponding to the local solution near k = 145. For the
data shown at the second height (y = 0.000245), the local solution peak is more
pronounced as compared to the local solution peaks for y = 0.00521 and y = 0.035
in the right frames of Fig. 3.39. We conjecture that with increasing width of the
exciter and increasing amplitude of excitation further, separation bubbles will be
more pronounced, which could interfere destructively with the TS waves - as noted
in this case with low intensity. However, presence of TS waves is unmistakable for
α1 = 0.1. One also notices significant spatial growth of u d at y = 0.035 in Fig. 3.39,
than for u d at y = 0.032 for α1 = 0.05 in Fig. 3.34. This increase in the extent of
the spatial growth of u d at the edge of the shear layer with increase in the amplitude
of excitation was also noted for lower α1 cases.
The disturbance streamwise velocity u d at t = 15, normalized with respect to
α1 = 0.1 case, for y = 0.00252 (the first grid line above the plate) is plotted in
Fig. 3.40 to analyze effects of amplitude of excitation. In this figure for α1 = 0.01,
u d is multiplied by ten-times and the same is multiplied by two times for α1 = 0.05,
to compare with the case of α1 = 0.1. There are distinct differences in all the cases.
The lowest α1 case shows near-neutral variation of u d in 0.6 ≤ x ≤ 0.8. While the
higher α1 cases show distinct spatial growth of u d . The onset of spatial growth of u d
is located almost at the same position for α1 = 0.05 and α1 = 0.1 cases, while the
extent of spatial growth increases for α1 = 0.1. Also, u d for the higher α1 cases show
more dominant presence of superharmonics near the exciter, due to the induction of
unsteady separation micro-bubbles on the wall. The results in Fig. 3.41 are for the
same excitation parameters and at the same location of the exciter, but are recorded
at higher height for y = 0.00808. At this height, significant spatial growth is noted
for all the α1 cases considered till now. As noted for smaller α1 cases, here also one
3.9 Nonlinear and Nonparallel Effects: Receptivity by DNS 201

0.05 (a) 10
1
(b)

Ud(k)
100
ud 0

-0.05 10-1 50 100 150 200


0 1 X 2 3 k
(c) (d)
0.05 101

Ud(k)
100
ud

0
-1
10
-0.05
10-2 50 100 150 200
0 1 X 2 3 k
(e) 10
1
(f)
0.01
100

Ud(k)
10-1
ud

0
10-2
-0.01
10-3 50 100 150 200
0 1 X 2 3 k

Fig. 3.39 u d plotted for y = 0.00245; 0.00521 and 0.035, respectively, from top to bottom at
t = 20. Results presented are for the SBS strip excitation with α1 = 0.10; F f = 3.0 × 10−4 and
the exciter located between x1 = 0.22 and x2 = 0.264. b Fourier transform of the signals shown
in a

(a) α1=0.01
0.05
ud

-0.05
0 0.5 1 x 1.5 2

(b) α1 = 0.05
0.05
ud

-0.05
0 0.5 1
x 1.5 2

(c) 0.05
α1 = 0.1
ud

-0.05
0 0.5 1
x 1.5 2

Fig. 3.40 Normalized disturbance velocity uˆd plotted against x at t = 15 for amplitudes of exci-
tation at y = 0.00252, which is the first grid line above the plate
202 3 Receptivity and Instability

(a) α1=0.01
0.04
0.02
ud 0
-0.02
-0.04
0 0.5 1 x 1.5 2
(b) α1=0.05
0.04
0.02
ud

0
-0.02
-0.04
0 0.5 1 x 1.5 2
(c) α1=0.1
0.04
0.02
ud

0
-0.02
-0.04
0 0.5 1 x 1.5 2

Fig. 3.41 Normalized disturbance velocity uˆd (normalized with respect to α1 = 0.01 case) plotted
against x at t = 15 for amplitudes of excitation at y = 0.0081

Table 3.3 Onset and end of x-locations for the spatial growth of u d at y = 0.00662 and y =
0.00808 for different values of excitation amplitude α1
α1 Onset of spatial End of spatial Onset of spatial End of spatial
growth at growth at growth at growth at
y = 0.00662 y = 0.00662 y = 0.00808 y = 0.00808
0.002 0.312 0.548 0.586 0.851
0.005 0.311 0.766 0.591 0.938
0.01 0.318 0.843 0.577 0.971
0.05 0.323 0.941 0.556 1.035
0.1 0.325 1.148 0.572 1.148

observes: (1) The onset of spatial growth is located almost at the same position, even
for α1 = 0.1 and (2) extent of spatial growth increases with increase in α1 . The onset
and end-location of the spatial growth of u d at y = 0.00662 and y = 0.00808 are
tabulated in Table 3.3 for different values of α1 .
3.9 Nonlinear and Nonparallel Effects: Receptivity by DNS 203

1
(a) (b)

ud/|ud|max
2
10
0.5

Ud(f)
0 10
1

-0.5 0
10
-113 13.2 13.4 13.6 13.8 14 0 5 10 f 15 20
t
1
(c) (d)
ud/|ud|max

2
10
0.5

Ud(f)
0 101
-0.5 0
10
-1 0 5 10 f 15 20
13 13.2 13.4 t 13.6 13.8 14

1
(e) (f)
ud/|ud|max

102
0.5

Ud(f)
0 10
1

-0.5 0
10
-1 0 5 10 f 15 20
13 13.2 13.4 t 13.6 13.8 14

Fig. 3.42 a Time variation of u d /|u d |max plotted at x = 0.3 and y = 0.00124; 0.00662 and 0.034
for α1 = 0.1, F f = 3.0 × 10−4 . The exciter located between 0.22 and 0.264. In frame b, Fourier
transform of the signal shown in a. Here, |u d |max is the maximum absolute value of u d for 13 ≤
t ≤ 14 at the corresponding (x, y)-location

In the left frames of Fig. 3.42, the time series for u d /|u d |max is shown at x = 0.3
for the case of α1 = 0.1 at the indicated heights. Here, |u d |max is the maximum
absolute value of u d , during 13 ≤ t ≤ 14, at the heights, y = 0.00124, 0.00662 and
0.034. These time series are multi-periodic, as evident from the Fourier transform
shown in the corresponding right frames. Once again, the dominant presence of
second and third superharmonics are noticeable in the interior of the shear layer (at
around y = 0.00662), similar to the lower α1 cases. The dominant superharmonics
are not seen for y = 0.034 and above, for this α1 case.
On the left column of the Fig. 3.43, the time series for the disturbance vorticity
(ωd ) is shown at a distance of x = 1.34 from the leading edge of the plate, for
α1 = 0.1 and F = 3.0 × 10−4 at the indicated heights, which are very close to the
wall. The time series indicate variations to be different at different heights and it
is seen from the Fourier transforms shown on the right frames of Fig. 3.43. One
notices the first superharmonic of F very clearly and a weak second superharmonic
for heights closest to the wall. When the time series are recorded closer to the exciter
(x = 0.27), as shown in Fig. 3.44, then the results are seen to be significantly different
for the same heights over the plate. For the lowest height, the amplitude for the
fundamental is seen to be three times higher at this closer location from the exciter
and the first superharmonic is also of comparable magnitude at this distance, with
more harmonics seen in the spectrum. For the other two heights also, one notices
clearly the presence of higher harmonics.
204 3 Receptivity and Instability

α1 = 0.1, x = 1.34
104
Ff = 3.0 × 10
-4
10

ωd(f)
ωd 103 y = 0.00252
5
2
10
20.2 20.4 t 20.6 20.8 21 5 10 15 20 25 30 35
f

10 y = 0.0081
4
10

ωd(f)
ωd

0
3
10
-10
20 20.2 20.4 t 20.6 20.8 21 5 10 15 20 25 30 35
f
5 y = 0.0144
104
ωd(f)
0
ωd

103

-5
102
20 20.2 20.4 t 20.6 20.8 21 5 10 15 20 25 30 35
f

Fig. 3.43 Time variation of ωd plotted (left) for α1 = 0.1 and F f = 3.0 × 10−4 at x = 1.34 for
the three indicated heights close to the plate. Corresponding Fourier transforms shown on the right

3.9.8 Wall-Normal Variation of u d and vd

In this sub-section, the wall-normal variation of u d and vd are explored for F =


3.0 × 10−4 and the exciter located between 0.22 and 0.264. This is investigated in
Figs. 3.45 to 3.48, where normalized u d and vd are plotted as a function of ȳ = y/δ ∗
at x = 0.3 and 0.5, respectively, from t = 20.25 to t = 20.45, at a time interval of
0.02. All these time instants lie within one fundamental non-dimensional time-period
of excitation T = 2πω̄0
, which is 0.209 for these cases. In essence, Figs. 3.45, 3.46,
3.47 and 3.48 show the variation of u d (y) and vd (y) with respect to time, within one
fundamental time-period T , at two particular streamwise locations. Two x-stations
chosen represent locations close to the exciter (x = 0.3) and in the middle part of the
asymptotic solution (x = 0.5). The u d plotted in Figs. 3.45, 3.46 are normalized with
respect to the α1 = 0.01 case. Similar normalization is done for vd also in Figs. 3.47
to 3.48.
Figure 3.45 reveal that the inner maximum not only oscillates with respect to
y, but displays wider variation in its shape, as clearly noted at t = 20.31, 20.33,
20.41 and 20.43. At t = 20.31, 20.41 and 20.43 one notices the splitting of the inner
maximum to create more than one peaks inside the shear layer. Such splitting of inner
maximum is not predicted at all by the linear stability analysis. Existence of more than
one peaks inside the shear layer are observed for α1 = 0.002, the lowest amplitude
of excitation case considered here. Fluctuation of u d near the inner maximum is
3.9 Nonlinear and Nonparallel Effects: Receptivity by DNS 205

60 α1 = 0.1, F = 3.0 × 10-4, y = 0.00252


x = 0.27
40 104
ωd

ωd(f)
20
103
0
2
10
20 20.2 20.4 t 20.6 20.8 21 5 10 15 20 25 30 35
f
10
y = 0.0081
0 104
ωd

ωd(f)
-10 103

-20 102

-3020 101
20.2 20.4 t 20.6 20.8 21 5 10 15 20 25 30 35
f
1
y = 0.0144
0 10
ωd

100
-0.002
ωd(f)

-1
10
10-2
-0.004 -3
10

-0.00620 10-4
20.2 20.4 t 20.6 20.8 21 5 10 15 20 25 30 35
f

Fig. 3.44 Time variation of ωd plotted at x = 0.27 for α1 = 0.1 and F f = 3.0 × 10−4 (left) for
the heights shown in Fig. 3.43. Corresponding Fourier transforms are shown on the right

more prominent as one moves downstream, noted in Fig. 3.46. One also notices
the difference in u d for various amplitudes of excitation cases displaying effects
of nonlinearity. The asymptotic solution corresponding to α1 = 0.1 case creates
multiple unsteady separation bubbles on the plate, as shown in Fig. 3.38. This is
responsible for larger departure of u d variation with respect to y from the linear
theory, as well as from that obtained for other lower α1 -cases, evident in Fig. 3.46.
This is also to be noted that there is a striking similarity and match between the u d
profiles between α1 = 0.002, 0.005 and 0.01 cases, for all the time instants and the
streamwise locations shown in these figures. This implies that nonparallelism of the
base flow and the spatio-temporal nature of the perturbations, rather than nonlinearity
did play a dominant role in determining the u d profiles for these α1 -cases. However,
effects of nonlinearity built-up and modify the profiles for higher amplitude cases,
which ultimately induced micro-separation bubbles on the wall for α1 = 0.1 case.
However, the outer maximum, in comparison displays lesser variation with respect
to time for all the three streamwise locations considered here, as it is located at the
edge of the shear layer.
206 3 Receptivity and Instability

(a) t = 20.25 (b) t = 20.27 (c) t = 20.29


6 x = 0.3 6 6

5 C1 5 C1 5 C1
C2 C2 C2
C3 C3 C3
4 C4 4 C4 4 C4
C5 C5 C5

y/δ∗


y/δ

y/δ
3 3 3

2 2 2

1 1 1

0 0 0
0 0.01 0.02 0 0.01 0.02 0 0.01 0.02
ud ud ud
(d) t = 20.31 (e) t = 20.33 (f) t = 20.35
6 C1 6 6
C2 C1 C1
C3 C2 C2
5 5 5
C4 C3 C3
C5 C4 C4
4 4 C5 4 C5

y/δ∗


y/δ

y/δ
3 3 3

2 2 2

1 1 1

0 0 0
-0.005
ud 0 -0.015 -0.01
ud-0.005 0 -0.02
ud -0.01 0

(g) t = 20.37 (h) t = 20.39 (i) t = 20.41


6 6 6
C1
C1 C1 C2
5 C2 5 C2 5
C3
C3 C3 C4
4 C4 4 C4 4 C5
C5 C5

3 3 3
y/δ∗

y/δ∗


y/δ

2 2 2

1 1 1

0 0 0
-0.02 -0.01 0 -0.015 -0.01 -0.005 0 -0.005 0 0.005
ud ud ud
(j) t = 20.43 (k) t = 20.45
6 6
C1
C2 C1
5 C3 5 C2
C4 C3
C5 C4
4 4 C5

3 3
y/δ∗


y/δ

2 2

1 1

0 0
0 0.005 0 0.01
ud ud

Fig. 3.45 Normalized streamwise disturbance velocity u d plotted as a function of y/δ ∗ at x = 0.3
and indicated time instants. Here, C1, C2, C3, C4 and C5 represent the cases with α1 = 0.002,
0.005, 0.01, 0.05 and 0.1, respectively. The exciter is put between 0.22 and 0.264 with F = 3.0 ×
10−4
3.9 Nonlinear and Nonparallel Effects: Receptivity by DNS 207

(a) t = 20.25 (b) t = 20.27 (c) t = 20.29


6 6 6

C1 C1
5 5 C2 5 C1
C2
C3 C2
C3
C4 C3
C4
4 4 C5 4 C4
C5
C5

y/δ∗
3 3 3
y/δ

y/δ
2 2 2

1 1 1

0 0 0
-0.005 0 -0.004 -0.002 0 0.002 -0.002 0
ud ud ud 0.002 0.004

(d) t = 20.31 (e) t = 20.33 (f) t = 20.35


6 6 6

5 5 5
C1
C1 C2 C1
4 C2 4 C3 4 C2
C3 C4 C3
C4 C5 C4

y/δ∗
3 C5 3 3 C5
y/δ

y/δ

2 2 2

1 1 1

0 0 0
0 0.005 0
ud ud 0.005 0.01 0
ud 0.005 0.01

(g) t = 20.37 (h) t = 20.39 (i) t = 20.41


6 C1 6 6
C1 C1
C2
C2 C2
C3
C3 C3
5 C4 5 5
C4 C4
C5
C5 C5
4 4 4

3 3 3

y/δ∗
y/δ

y/δ

2 2 2

1 1 1

0 0 0
-0.01
ud -0.005 0 0.005 -0.01 -0.005 ud 0 -0.01 u -0.005
d
0

(j) t = 20.43 (k) t = 20.45


6 6
C1 C1
C2 C2
5 C3 5 C3
C4 C4
C5 C5
4 4

3 3


y/δ

y/δ

2 2

1 1

0 0
-0.01
ud -0.005 0 -0.01
ud -0.005 0

Fig. 3.46 Normalized streamwise disturbance velocity u d plotted as a function of y/δ ∗ at x = 0.5
and indicated time instants. Here, C1, C2, C3, C4 and C5 represent the cases with α1 = 0.002,
0.005, 0.01, 0.05 and 0.1, respectively. The exciter is between 0.22 and 0.264 with F = 3.0 × 10−4
208 3 Receptivity and Instability

(a) t = 20.25 (b) t = 20.27 (c) t = 20.29


C1
6 C2 6 6
C3
5 C4 5 5
C5

4 4 4

C1 C1
3 3 C2 3 C2
y/δ∗

y/δ∗

y/δ∗
C3 C3
C4 C4
2 2 C5 2 C5

1 1 1

0 0 0
0 -0.002
vd0.001 0.002 -0.002 -0.001
vd 0 -0.004 vd 0

(d) t = 20.31 (e) t = 20.33 (f) t = 20.35


6 6 6

5 5 5

4 4 4
C1 C1
C2 C2
3 3 3
C3 C3
y/δ∗

y/δ∗

y/δ∗
C4 C4
2 C5 2 C5 2

1 1 1

0 0 0
-0.005 vd 0 -0.005 vd 0 -0.002 0
vd
(g) t = 20.37 (h) t = 20.39 (i) t = 20.41
C1
6 C2 6 6
C3
C4
5 5 5
C5

4 4 4

y/δ∗

y/δ∗
y/δ

C1 C1
3 3 3
C2 C2
C3 C3
2 2 C4 2 C4
C5 C5

1 1 1

0 0 0
0 vd 0.001 0.002 0 vd 0.002 0.004 0 vd 0.005

(j) t = 20.43 (k) t = 20.45


6 6 C1
C2
C3
5 5 C4
C5
4 4
y/δ∗

y/δ∗

3 C1 3
C2
C3
2 C4 2
C5

1 1

0 0
0 vd 0.002 0.004 0 vd 0.002

Fig. 3.47 Normalized wall-normal disturbance velocity vd plotted as a function of y/δ ∗ at x = 0.3
and indicated time instants. Here, C1, C2, C3, C4 and C5 represent the cases with α1 = 0.002,
0.005, 0.01, 0.05 and 0.1, respectively. The exciter is between 0.22 and 0.264 with F = 3.0 × 10−4
3.9 Nonlinear and Nonparallel Effects: Receptivity by DNS 209

(a) t = 20.25 (b) t = 20.27 (c) t = 20.29


C1 C1 C1
6 C2 6 C2 6 C2
C3 C3 C3
C4 C4 C4
5 C5 5 5 C5
C5

4 4 4
y/δ∗

y/δ∗
3 3 3

y/δ
2 2 2

1 1 1

0 0 0
0
vd 0.002 0 0.002
vd 0 0.001
vd 0.002 0.003

(d) t = 20.31 C1 (e) t = 20.33 (f) t = 20.35


C2 C1
6 6 6 C2
C3 C3
C4 C4
5 C5 5 5 C5

4 4 4

C1
y/δ∗

y/δ∗
3 3 C2 3
y/δ

C3
C4
C5
2 2 2

1 1 1

0 0 0
0 vd 0.002 -0.002 vd 0 -0.004 -0.002 0
vd
(g) t = 20.37 (h) t = 20.39 (i) t = 20.41
6 6 6

5 5 5

4 4 4

C1 C1
3 3 3 C1
y/δ∗

y/δ∗

C2
y/δ

C2 C2
C3 C3 C3
2 C4 2 C4 2 C4
C5 C5 C5

1 1 1

0 0 0
-0.004 vd -0.002 0 -0.004 -0.002 vd 0 -0.002 vd 0 0.002

(j) t = 20.43 (k) t = 20.45


C1
6 6 C2
C3
C4
5 5 C5

4 4

C1
3 3
y/δ∗

C2
y/δ

C3
2 C4 2
C5

1 1

0 0
0 vd 0.002 0 vd 0.001 0.002

Fig. 3.48 Normalized wall-normal disturbance velocity vd plotted as a function of y/δ ∗ at x = 0.5
and indicated time instants. Here, C1, C2, C3, C4 and C5 represent the cases with α1 = 0.002,
0.005, 0.01, 0.05 and 0.1, respectively. The exciter is between 0.22 and 0.264 with F = 3.0 × 10−4
210 3 Receptivity and Instability

One observes also significant effects of nonlinearity for the profile of vd shown
in Figs. 3.47 and 3.48 for x = 0.3 and 0.5. Presence of multiple maxima can also be
noted for these cases at certain time instants as shown in Figs. 3.47f and 3.48a, e, k.
However, for cases with α1 = 0.002, 0.005 and 0.01, similar variation of vd is noted
for all time instants form these figures.

3.9.9 Further Evidences of Nonparallel, Nonlinear Effects


and Bypass Transition

An interesting fact arises when one compares u d and ωd , which show completely dif-
ferent kind of spatial behavior at two different heights from the plate. Spatial growth
of u d at intermediate heights has already been observed. This is further shown in
Fig. 3.49 by plotting u d for three different non-dimensional frequencies of excita-
tion, at y = 0.00662. The exciter is located between 0.22 and 0.264 which excites
the flow with α1 = 0.01. In the frames of Fig. 3.49, a vertical dotted line indicates
the location up to which the solution grows, beyond that line the solution decays.
One observes spatial growth of u d , whose extent tend to increase with decrease in
the non-dimensional frequency of excitation F.

F = 2.5 × 10
-4
(a) 0.004
0.002
ud

-0.002

-0.004
0 0.5 1 x 1.5 2 2.5

(b) 0.004 F = 3.0 × 10-4

0.002
ud

-0.002

-0.004
0 0.5 1 x 1.5 2 2.5

(c) 0.004 F = 3.5 × 10-4

0.002
ud

-0.002

-0.004
0 0.5 1 x 1.5 2 2.5

Fig. 3.49 u d plotted against x at t = 15 for y = 0.00662, α1 = 0.01 and the indicated frequencies
of excitation
3.9 Nonlinear and Nonparallel Effects: Receptivity by DNS 211

(a) 5 F = 2.5 × 10-4

ωd
0

-5 0 1 x 2
(b) 5 F = 3.0 × 10-4
ωd

-5 0 1 x 2
(c) 5 F = 3.5 × 10-4
ωd

-5 0 1 x 2

Fig. 3.50 ωd plotted as a function of x at t = 15 for y = 0.0066 with α1 = 0.01 for the displayed
frequencies

The interesting point emerges when ωd is plotted against x at the same height
and at same time, as shown in Fig. 3.50. For all the three frequencies, ωd decays
in contrast to the growing u d shown in Fig. 3.49. It is worth remembering that the
vorticity represents the rotational part of the flow field, while the velocity mainly
contains the translational part of the flow field.
However, a completely contrasting scenario is noted at y = 0.0112 in Figs. 3.51
and 3.52. At this height, ωd indicates growth, while u d is stable, as shown in Fig. 3.52.
Note the absence of local solution and the distinct growth of ωd in Fig. 3.51. In
Fig. 3.52, one can note that ωd shows monotonic decay for the two higher frequencies,
whereas there is a very small growth in that part of the signal shown in between the
two vertical lines for the lowest frequency of F = 2.5 × 10−4 .
In all the cases reported so far, ZPG boundary layer is investigated by exciting it
using a SBS harmonic exciter of finite width (from x1 = 0.22 to x2 = 0.264). Nonpar-
allel and nonlinear effects of receptivity specifically for high frequency disturbances
is recorded. Next, another case is studies, where the exciter width is increased fur-
ther, in both the directions, by considering the strip between x1 = 0.2 to x2 = 0.29.
Having extended it further towards the leading edge of the plate, more nonparallel
effects are introduced and by extending it in the downstream direction also, more
energy inside the shear layer is imparted. In Fig. 3.53, the stream function and vor-
ticity contours for these two cases of different width of the exciter are compared. In
212 3 Receptivity and Instability

(a) 1 F = 2.5 × 10-4

0.5
ωd
0

-0.5

-1 0 1 x 2

(b) 1 F = 3.0 × 10-4

0.5
ωd

-0.5

-1
0 1 x 2
(c) 1 F = 3.5 × 10 -4

0.5
ωd

-0.5

-1
0 1 x 2

Fig. 3.51 ωd plotted against x at t = 15 for the indicated frequencies of excitation at y = 0.0112
with α1 = 0.01

both the cases, the amplitude of excitation is taken as α1 = 0.1 and the frequency of
excitation given by F = 3.0 × 10−4 .
From the ψ-contours in (a), the presence of a few small bubbles in the immediate
downstream of the exciter, spaced at a gap similar to the wavelength of TS waves
is noted. These bubbles are responsible for setting up of wavy disturbance field.
Note that the wall-normal direction is stretched in all the frames shown in Fig. 3.53.
The vorticity contours exhibit a two-tier structure with the lower layer showing an
upstream propagating tendency, while the upper tier shows vertical direction of the
ejected vortices (those are slightly tilted towards downstream direction). For the case
of wider excitation strip in Fig. 3.53c, the stream function contours display bubbles,
over the streamwise stretch up to x = 3.3, as compared to the case of shorter exciter
which shows the visible undulation up to x = 1.7 only. Also the bubbles are of
significantly larger dimension in the wall-normal direction. These bubbles are not
steady and the ensuing unsteady downstream motion is what is termed as bypass
transition in [8, 60]. In [60], the unsteady separation on the surface of the plate was
initiated by a disturbance from the free stream by creating induced adverse pressure
gradient.
3.9 Nonlinear and Nonparallel Effects: Receptivity by DNS 213

F = 2.5 × 10
-4
(a) 0.005

ud
0

-0.005
0 0.5 1 x 1.5 2 2.5
F = 3.0 × 10
-4
(b) 0.005
ud

-0.005
0 0.5 1 x 1.5 2 2.5
F = 3.5 × 10
-4
(c) 0.005
ud

-0.005
0 0.5 1 x 1.5 2 2.5

Fig. 3.52 u d plotted against x at t = 15 for the indicated frequencies of excitation at y = 0.0112
with α1 = 0.01

3.9.10 Effects of the Location of Exciter

Next, the effects of the location of the exciter is described for F = 3.0 × 10−4 and
α1 = 0.01. Six different exciter locations are chosen for this purpose. All the exciters
are of identical width, wex = 0.05. In Table 3.4, the x-location of the center of the SBS
exciter, corresponding Reδ∗ , non-dimensional wavenumber αr , and spatial growth
rate αi at the exciter location, as obtained from the linear stability theory, are tabulated.
Since, αr and αi are defined with respect to the local displacement thickness δ ∗ , kr
and ki , (the non-dimensional wavenumber and spatial growth rate), with respect to
the reference length scale L, are also listed in Table 3.4.
One sees from Table 3.4, that the TS wave-packet caused by the exciter at xex =
0.1, should decay faster than any other exciter locations discussed in this section.
However, this does not happen for the receptivity cases computed here. In Fig. 3.54a,
b, wall-normal variation of |u d |/|u d |max and |vd |/|vd |max are plotted for all the exciter
locations at some representative times, after the complete establishment of the TS
wave-packet. The profiles are plotted at a location of x = 0.075 downstream of xex .
One notes from Fig. 3.54a, b that the height where u d = 0 and vd is maximum lies in
the range 0.005 ≤ y ≤ 0.01. Hence, one expects u d to display growth in this range
of heights as explained in Sect. 3.9.4.
For the subsequent discussions, the onset and end-location of the spatial growth or
near-neutral behavior of the u d corresponding to the TS wave-packet at a particular
height is referred as T Sonset and T Send , respectively. Hence, here T Sonset and T Send
214 3 Receptivity and Instability

(a) ψ contour for exciter between 0.22 and 0.264


0.01

y
0.005

0
0 0.5 x 1 1.5 2

(b) ω contour for exciter between 0.22 and 0.264


0.03

0.02
y

0.01

0
0 0.5 x 1 1.5 2
(c) ψ contour for exciter between 0.2 and 0.29
0.01
y

0.005

0
0 0.5 x 1 1.5 2

(d) ω contour for exciter between 0.2 and 0.29


0.03

0.02
y

0.01

0
0 0.5 x 1 1.5 2

Fig. 3.53 Stream function and vorticity contours for the two different strip exciter cases having
different width. a and b correspond to exciter located between 0.22 and 0.264; c and d correspond
to exciter located between 0.2 and 0.29

Table 3.4 Location of the exciter xex , corresponding Reδ ∗ , non-dimensional wavenumbers αr ; kr
and spatial growth rates αi ; ki at the exciter location as obtained from the linear stability theory are
tabulated. Here, αr and αi are defined with respect to local δ ∗ , whereas kr and ki is defined with
respect to L. For all the cases listed here, F = 3.0 × 10−4 and α1 = 0.01
xex Reδ ∗ αr αi kr ki
0.1 172.00 0.1377 0.037790 80.0581 21.9709
0.25 271.96 0.2079 0.019461 76.4451 7.1558
0.4 344.00 0.2559 0.009830 74.3895 2.8576
0.625 430.00 0.3149 0.003061 73.2326 0.7119
0.8 486.49 0.3546 0.002609 72.8895 0.5363
1.0 543.91 0.3955 0.006259 72.7142 1.1507
3.9 Nonlinear and Nonparallel Effects: Receptivity by DNS 215

(a)
xex = 0.1
1 xex = 0.25
xex = 0.4
xex = 0.625
|ud| / |ud| max
0.8
xex = 0.8
xex = 1.0
0.6

0.4

0.2

0
0 0.01 0.02 0.03 0.04 0.05
y

(b)
xex = 0.1
1 xex = 0.25
xex = 0.4
xex = 0.625
|vd| / |vd| max

0.8
xex = 0.8
xex = 1.0
0.6

0.4

0.2

0
0 0.01 0.02 0.03 0.04 0.05
y

Fig. 3.54 a |u d |/|u d |max and b |vd |/|vd |max plotted as a function of y for the indicated exciter
locations times after the complete establishment of the TS wave-packet. The profiles are plotted
at a location of x = 0.075 downstream of the center of the exciter. Here, F = 3.0 × 10−4 and
α1 = 0.01

are only the functions of height and xex . In Table 3.5, T Sonset and T Send for u d
at t = 20 are tabulated as a function of y for the indicated exciter locations. From
Table 3.5, one notes the existence of three distinct ranges of height. The first range
extends up to y  0.00956 from the wall, whereas the second range lies in 0.01113 ≤
y ≤ 0.01274. The third range lies at the top of the second range. For xex = 0.1 and
0.25, T Sonset and T Send occurs at identical locations for 0.00662 ≤ y ≤ 0.00956.
Other exciter locations also progressively display similar values of T Sonset and T Send ,
as the height approaches 0.00956. Monotonic spatial decay is observed for all the
216 3 Receptivity and Instability

Table 3.5 The onset and end of the spatial growth or the near-neutral behaviour of the TS wave-
packet listed for all heights, for the indicated locations of exciter. Here, F = 3.0 × 10−4 and α1 =
0.01. The blank cells indicate those heights, where no spatial growth is noticed
100 × y xex = 0.1 xex = 0.25 xex = 0.4 xex = 0.625 xex = 0.8 xex = 1.0
0.521 0.168 0.326 0.49 0.75 – –
0.875 0.878 0.855 0.889 – –
0.662 0.308 0.301 0.49 0.715 – –
0.874 0.888 0.908 0.908 – –
0.808 0.54 0.54 0.52 0.65 0.85 –
1.02 1.07 1.05 1.05 1.05 –
0.958 0.948 0.94 0.94 0.93 0.95 1.12
1.29 1.28 1.28 1.28 1.29 1.26
1.113 – – – – – –
– – – – – –
1.274 – – – – – –
– – – – – –
2.549 0.48 0.49 0.52 0.74 0.83 –
0.95 0.93 0.93 0.93 0.93 –
2.759 0.52 0.53 0.54 0.70 0.88 –
0.92 0.97 0.93 0.96 0.97 –
2.966 0.47 0.48 0.52 0.76 0.88 –
0.96 0.97 0.96 0.97 0.97 –
3.188 0.516 0.53 0.52 0.69 0.88 –
0.96 0.97 0.96 0.96 0.97 –

exciter locations in the second range of heights (0.1113 ≤ y ≤ 0.1274). The heights
corresponding to the third range are closer to the shear layer edge, where one observes
that first five exciter locations display similar locations for T Send , while the first
three cases display similar locations of T Sonset . However the xex = 1.0 case, does
not display any spatial growth of the TS wave-packet at these heights, because the
exciter itself is located downstream of corresponding T Send location for the other
exciter location indicated in the table.

3.9.11 Effects of Frequency of Excitation

Here, the effects of frequency of excitation are considered for the exciter of
width Δxex = 0.05 located at xex = 0.25 (Reδ∗ ex = 271.96), with amplitude of
excitation α1 = 0.01. Here, six different frequencies of excitation, with values
of F = 1.5 × 10−4 , 2.0 × 10−4 , 2.5 × 10−4 , 3.5 × 10−4 , 4.0 × 10−4 , 4.5 × 10−4 ,
4.0 × 10−4 and 5.5 × 10−4 are considered. Out of these six frequencies, only the
3.9 Nonlinear and Nonparallel Effects: Receptivity by DNS 217

Table 3.6 Non-dimensional frequency of excitation F, corresponding non-dimensional wavenum-


bers αr ; kr and spatial growth rates αi ; ki at the exciter location as obtained from the linear stability
theory along with the entry (x N 1 ) and exit (x N 2 ) point of the F = Const. line into the neutral curve
are tabulated. Here, αr and αi is defined with respect to local δ ∗ , whereas kr and ki is defined with
respect to reference length scale L. For all the cases listed here, the exciter is located at xex = 0.25
with α1 = 0.01
F f × 104 αr αi kr ki xN1 xN2
1.5 0.12055 0.02846 44.16325 10.42750 1.102 1.644
2.0 0.15191 0.02514 55.65174 9.20941 0.985 1.2775
2.5 0.18099 0.02207 66.30495 8.08661 – –
3.0 0.20879 0.01929 76.49143 7.06664 – –
3.5 0.23585 0.01687 86.40424 6.17933 – –
4.0 0.26246 0.01490 96.15182 5.45846 – –
4.5 0.28878 0.01347 105.79572 4.93468 – –
5.0 0.31491 0.01265 115.36928 4.63332 – –
5.5 0.34090 0.01249 124.89009 4.57481 – –

F = 1.5 × 10−4 and F = 2.0 × 10−4 line intersects the neutral curve, whereas the
straight line corresponding to the other frequencies stay outside the neutral curve.
In Table 3.6, the non-dimensional wavenumbers αr ; kr and spatial growth rates αi ;
ki at the exciter location, as obtained from the linear stability theory are tabulated,
along with the entry (x N 1 ) and exit (x N 2 ) point of the F = Const. line into the neu-
tral curve. The table includes the F = 3.0 × 10−4 case as well, for the purpose of
comparison, whose receptivity results have already been discussed.
As noted in Table 3.6, the cases of F = 1.5 × 10−4 and 2.0 × 10−4 lines intersect
the neutral curve at (x N 1 , x N 2 ) = (1.102, 1.644) and (x N 1 , x N 2 ) = (0.985, 1.2775),
respectively. Other frequency cases do not intersect the neutral curve at all. In
Fig. 3.55, the computed receptivity results are shown at y = 0.0254961 by plot-
ting u d at t = 20 as a function of x. The vertical lines drawn in frames (a) and
(b), indicate intersection points of the neutral curve with corresponding F = Const.
lines. In Table 3.7, the T Sonset and T Send of the TS wave-packet are listed for several
heights for indicated frequencies of excitation cases. In this table, the blank cells
indicate heights where no spatial growth is observed.
From this table one notes that the strict spatial growth of the amplitude of the TS
wave-packet for F = 1.5 × 10−4 , 2.0 × 10−4 and 2.5 × 10−4 takes place for three
different ranges of height. In the first range of heights, spatial growth of u d takes
place for these frequencies. This range of heights extend from the wall to y  0.016,
0.01273 and 0.0113 for F = 1.5 × 10−4 , 2.0 × 10−4 and 2.5 × 10−4 , respectively,
as noted in Table 3.7. Both T Sonset and T Send move downstream with increase in
height in this range of heights for these frequencies. Also the extent of the zone of
spatial growth (T Send − T Sonset ) is also seen to reduce with increase in height in
this range. Beyond this range, there exists another range of heights where no spatial
growth for u d is noted for these frequencies. This second range of heights are close
to the y-location, where respective u d changes sign. Above this range of heights, the
218 3 Receptivity and Instability

(a) F = 1.5 × 10
−4 (b) F = 2.0 × 10
−4
0.002 0.002
ud

ud
0 0

y = 0.254961E-01 y = 0.254961E-01
-0.002 -0.002
0 1 2 3 x 4 5 6 0 1 2 x 3 4 5
(c) F = 2.5 × 10−4 (d) F = 3.5 × 10−4
0.002 0.002
ud

ud
0 0

y = 0.254961E-01
-0.002 y = 0.254961E-01 -0.002
0 1 x 2 3 0 1 x 2 3
(e) F = 4.0 × 10−4 (f) F = 4.5 × 10−4
0.002 0.002
ud

ud

0 0

y = 0.254961E-01 y = 0.254961E-01
-0.002 -0.002
0 1 x 2 3 0 1 x 2 3
(g) F = 5.0 × 10−4 (h) F = 5.5 × 10−4
0.002 0.002
ud

ud

0 0

y = 0.254961E-01 y = 0.254961E-01
-0.002 -0.002
0 1 x 2 3 0 1 x 2 3

Fig. 3.55 u d plotted as a function of x for the indicated values of F at y = 0.02549 and t = 20.
Here, the exciter of width Δxex = 0.05 is located at xex = 0.25 with α1 = 0.01. The vertical straight
lines in a and b indicate the entry and exit point of the corresponding F f = Const. line into the
neutral curve

spatial growth of u d is again observed for these frequencies. This range starts from
y  0.02549, 0.0161 and 0.01439 for F = 1.5 × 10−4 , 2.0 × 10−4 and 2.5 × 10−4 ,
respectively.
For F ≥ 3.0 × 10−4 , there exists a range of heights close to the wall, where no
spatial growth is noted. However, spatial growth of u d is resumed above this range, for
these frequencies, as noted from Table 3.7. The terminal height of this second range of
spatial growth for 3.0 × 10−4 ≤ F ≤ 5.5 × 10−4 is noted to decrease with increase
in the frequency. For 3.0 × 10−4 ≤ F ≤ 4.0 × 10−4 , once again a range of heights
exists, where monotonic spatial decay of the amplitude of the TS wave-packet is
noted. One notes from Table 3.7 that the extent of this third range of heights increases
with increase in F from 3.0 × 10−4 to 4.0 × 10−4 . The spatial growth of the TS wave-
packet resumes above this third range of heights for 3.0 × 10−4 ≤ F ≤ 4.0 × 10−4 .
However, no spatial growth of the TS wave-packet can be observed for all the heights
above y  0.00662 for F ≥ 4.5 × 10−4 .
References 219

Table 3.7 The onset and end of the spatial growth or the near-neutral behaviour of the TS wave-
packet for all the heights shown in Fig. 3.55 for indicated F × 104 cases. The blank cells indicate
heights, where no spatial growth is observed
100 × y 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5
0.252 1.082 0.916 0.771 – – – – – –
2.607 1.701 1.198 – – – – – –
0.384 1.233 0.903 0.741 – – – – – –
2.606 1.694 1.181 – – – – – –
0.521 1.341 0.854 0.758 0.326 0.408 0.289 0.307 0.319 0.308
2.612 1.677 1.125 0.878 0.611 0.509 0.471 0.422 0.392
0.662 1.365 0.760 0.299 0.303 0.333 0.369 0.410 0.486 0.569
2.613 1.692 1.186 0.888 0.761 0.661 0.622 0.652 0.597
0.808 1.316 0.442 0.481 0.539 0.615 0.739 – – –
2.629 1.804 1.235 1.071 0.911 0.929 – – –
0.958 0.569 0.696 0.749 0.939 – – – – –
2.643 1.723 1.386 1.276 – – – – –
1.113 0.854 0.996 1.253 – – – – – –
2.691 1.918 1.699 – – – – – –
1.273 1.250 1.545 – – – – – – –
2.818 2.163 – – – – – – –
1.439 1.741 – 0.621 0.563 0.487 – – – –
2.960 – 0.873 0.760 0.645 – – – –
1.610 2.487 0.905 0.614 0.536 0.471 0.434 – – –
3.351 1.334 1.110 0.881 0.670 0.549 – – –
1.786 – 0.880 0.622 0.536 0.463 0.438 – – –
– 1.526 1.174 0.909 0.732 0.555 – – –
2.549 1.132 0.851 0.629 0.521 0.464 0.436 – – –
2.551 1.769 1.277 0.931 0.733 0.550 – – –

References

1. Allen, L., & Bridges, T. J. (2002). Numerical exterior algebra and the compound matrix method.
Numerische Mathematik, 92, 197–232.
2. Alizard, F., & Robinet, J. (2007). Spatially convective global modes in a boundary layer. Physics
of Fluids, 19, 114105.
3. Arfken, G. (1985). Mathematical methods for physicists (3rd ed.). Orlando: Academic Press.
4. Barkley, D., Gomes, M. G. M., & Henderson, R. D. (2002). Three-dimensional instability in
flow over a backward-facing step. Journal of Fluid Mechanics, 473, 167–190.
5. Barkley, D., Blackburn, H. M., & Sherwin, S. J. (2008). Direct optimal growth analysis for
timesteppers. International Journal for Numerical Methods in Fluids, 57, 1435–1458.
6. Bender, C. M., & Orszag, S. A. (1987). Advanced mathematical methods for scientists and
engineers. Singapore: McGraw Hill Book Co., International Edition.
7. Bhaumik, S. (2013). Direct numerical simulation of inhomogeneous transitional and turbulent
flows. Ph. D. thesis, I. I. T. Kanpur
220 3 Receptivity and Instability

8. Bhumkar, Y. G. (2011). High performance computing of bypass transition. Ph.D. thesis, I. I.


T. Kanpur
9. Bhaumik, S., & Sengupta, T. K. (2014). Precursor of transition to turbulence: Spatiotemporal
wave front. Physical Review E, 89(4), 043018.
10. Bhaumik, S., & Sengupta, T. K. (2017). Impulse response and spatio-temporal wave-packets:
The common feature of rogue waves, tsunami and transition to turbulence. Physics of Fluids,
29, 124103.
11. Blackburn, H. M., Barkley, D., & Sherwin, S. J. (2008). Convective instability and transient
growth in flow over a backward-facing step. Journal of Fluid Mechanics, 603, 271–304.
12. Brandt, L., & Henningson, D. S. (2002). Transition of streamwise streaks in zero-pressure-
gradient boundary layers. Journal of Fluid Mechanics, 472, 229–261.
13. Brewstar, R. A., & Gebhart, B. (1991). Instability and disturbance amplification in a mixed-
convection boundary layer. Journal of Fluid Mechanics, 229, 115–133.
14. Cebeci, T., & Bradshaw, P. (1977). Momentum transfer in boundary layers. Washington, DC:
Hemisphere Publishing Corporation.
15. Chen, T. S., & Moutsoglu, A. (1979). Wave instability of mixed convection flow on inclined
surfaces. Numerical Heat Transfer, 2, 497–509.
16. Chen, T. S., & Mucoglu, A. (1979). Wave instability of mixed convection flow over a horizontal
flat plate. International Journal of Heat and Mass Transfer, 22, 185–196.
17. Chen, T. S., Sparrow, E. M., & Mucoglu, A. (1977). Mixed convection in boundary layer flow
on a horizontal plate. ASME Journal of Heat Transfer, 99, 66–71.
18. Chomaz, J. M. (2005). Global instabilities in spatially developing flows: Non-normality and
nonlinearity. Annual Review of Fluid Mechanics, 37, 357–392.
19. Drazin, P. G., & Reid, W. H. (1981). Hydrodynamic stability. UK: Cambridge University Press.
20. Eckert, E. R. G. & Soehngen, E. (1951). Interferometric studies on the stability and transition
to turbulence in a free-convection boundary-layer. Proceedings of the General Discussion on
Heat Transfer, ASME and IME London (Vol. 321) (1951)
21. Eiseman, P. R. (1985). Grid generation for fluid mechanics computation. Annual Review of
Fluid Mechanics, 17, 487–522.
22. Fasel, H., & Konzelmann, U. (1990). Non-parallel stability of a flat-plate boundary layer using
the complete Navier-Stokes equations. Journal of Fluid Mechanics, 221, 311–347.
23. Gaster, M. (1974). On the effect of boundary-layer growth on flow stability. Journal of Fluid
Mechanics, 66(3), 465–480.
24. Gebhart, B., Jaluria, Y., Mahajan, R. L., & Sammakia, B. (1988). Buoyancy-induced flows and
transport. Washington, DC: Hemisphere Publications.
25. Gilpin, R. R., Imura, H., & Cheng, K. C. (1978). Experiments on the onset of longitudinal
vortices in horizontal Blasius flow heated from below. ASME Journal of Heat Transfer, 100,
71–77.
26. Haaland, S. E., & Sparrow, E. M. (1973). Vortex instability of natural convection flows on
inclined surfaces. International Journal of Heat Mass Transfer, 16, 2355–2367.
27. Hall, P., & Morris, H. (1992). On the instability of boundary layers on heated flat plates. Journal
of Fluid Mechanics, 245, 367–400.
28. Heisenberg, W. (1924). Über stabilität und turbulenz von flüssigkeitsströmen. Annalen der
Physik Leipzig, 379, 577–627 (Translated as ‘On stability and turbulence of fluid flows’. NACA
Tech. Memo. Wash. No 1291 1951)
29. Iyer, P. A., & Kelly, R. E. (1974). The instability of the laminar free convection flow induced
by a heated, inclined plate. International Journal of Hear Mass Transfer, 17, 517–525.
30. Jain, M. K., Iyengar, S. R. K. & Jain, R. K. (2003). Numerical methods for scientific and
engineering computation. New Delhi: New Age International
31. Kaiktsis, L., Karniadakis, G. M., & Orszag, S. A. (1991). Onset of three-dimensionality, equib-
ria and early transition in flow over a backward-facing step. Journal of Fluid Mechanics, 231,
501–528.
32. Kaiktsis, L., Karniadakis, G. M., & Orszag, S. A. (1996). Unsteadiness and convective insta-
bilities in two-dimensional flow over a backward-facing step. Journal of Fluid Mechanics, 321,
157–187.
References 221

33. Kloker, M., Konzelmann, U., & Fasel, H. (1993). Outflow boundary conditions for spatial
Navier-Stokes simulations of transitional boundary layers. AIAA Journal, 31, 620.
34. Kreyszig, E. (1999). Advanced engineering mathematics. Singapore: Wiley.
35. Leal, L. G. (1973). Steady separated flow in a linearly decelerated free stream. Journal of Fluid
Mechanics, 59, 513–535.
36. Liu, Z., & Liu, C. (1994). Fourth order finite difference and multigrid methods for modeling
instabilities in flat plate boundary layer-2D and 3D approaches. Computers and Fluids, 23,
955–982.
37. Lloyd, J. R., & Sparrow, E. M. (1970). On the instability of natural convection flow on inclined
plates. Journal of Fluid Mechanics, 42, 465–470.
38. Lord, R. (1880). On the stability or instability of certain fluid motions. Scientific Papers, 1,
361–371.
39. Marquet, O., Sipp, D., Chomaz, J. M., & Jacquin, L. (2008). Amplifier and resonator dynam-
ics of a low Reynolds-number recirculation bubble in a global framework. Journal of Fluid
Mechanics, 605, 429–443.
40. Moutsoglu, A., Chen, T. S., & Cheng, K. C. (1981). Vortex instability of mixed convection
flow over a horizontal flat plate. ASME Journal of Heat Transfer, 103, 257–261.
41. Mucoglu, A., & Chen, T. S. (1978). Wave instability of mixed convection flow along a vertical
flat plate. Numerical Heat Transfer, 1, 267–283.
42. Mureithi, E. W., & Denier, J. P. (2010). Absolute-convective instability of mixed forced-free
convection boundary layers. Fluid Dynamics Research, 372, 517–534.
43. Ng, B. S., & Reid, W. H. (1980). On the numerical solution of the Orr-Sommerfeld problem:
Asymptotic initial conditions for shooting method. Journal of Computational Physics, 38,
275–293.
44. Ng, B. S., & Reid, W. H. (1985). The compound matrix method for ordinary differential systems.
Journal of Computational Physics, 58, 209–228.
45. Rajpoot, M. K., Sengupta, T. K., & Dutt, P. K. (2010). Optimal time advancing dispersion
relation preserving schemes. Journal of Computational Physics, 229(10), 3623–3651.
46. Saric, W. S., & Nayfeh, A. H. (1975). Nonparallel stability of boundary-layer flows. Physics
of Fluids, 18(8), 945–950.
47. Schlatter, P., & Örl ü, R. (2012). Turbulent boundary layers at moderate Reynolds numbers.
Journal of Fluid Mechanics, 710, 5–34.
48. Schlichting, H. (1933). Zur entstehung der turbulenz bei der plattenströmung. Nach. Gesell. d.
Wiss. z. Gött., MPK,42, 181–208
49. Schneider, W. (1979). A similarity solution for combined forced and free convection flow over
a horizontal plate. International Journal of Heat and Mass Transfer, 22, 1401–1406.
50. Schubauer, G. B., & Skramstad, H. K. (1947). Laminar boundary layer oscillations and the
stability of laminar flow. Journal of Aerosol Science, 14(2), 69–78.
51. Sengupta T. K. (1990). Receptivity of a growing boundary layer to surface excitation. (Unpub-
lished manuscript).
52. Sengupta, T. K. (1991). Impulse response of laminar boundary layer and receptivity. In C.
Taylor (Ed.), Proceedings of the 7th International Conference Numerical Methods in Laminar
and Turbulent Layers. Stanford University
53. Sengupta, T. K. (2012). Instabilities of flows and transition to turbulence. Florida, USA: CRC
Press, Taylor & Francis Group.
54. Sengupta, T. K. (2013). High accuracy computing methods: Fluid flows and wave phenomenon.
USA: Cambridge University Press.
55. Sengupta, T. K., & Bhaumik, S. (2011). Onset of turbulence from the receptivity stage of fluid
flows. Physical Review Letters, 154501, 1–5.
56. Sengupta, T. K., & Venkatasubbaiah, K. (2006). Spatial stability for mixed convection boundary
layer over a heated horizontal plate. Studies in Applied Mathematics, 117, 265–298.
57. Sengupta, T. K., Ballav, M., & Nijhawan, S. (1994). Generation of Tollmien-Schlichting waves
by harmonic excitation. Physics of Fluids, 6(3), 1213–1222.
222 3 Receptivity and Instability

58. Sengupta, T. K., Bhaumik, S., Singh, V., & Shukl, S. (2009). Nonlinear and nonparallel recep-
tivity of zero-pressure gradient boundary layer. International Journal of Emerging Multidisci-
plinary Fluid Sciences, 1, 19–35.
59. Sengupta, T. K., Chattopadhyay, M., Wang, Z. Y., & Yeo, K. S. (2002). By-pass mechanism of
transition to turbulence. Journal of Fluids and Structures, 16, 15–29.
60. Sengupta, T. K., De, S., & Sarkar, S. (2003). Vortex-induced instability of an incompressible
wall-bounded shear layer. Journal of Fluid Mechanics, 493, 277–286.
61. Sengupta, T. K., Rao, A. K., & Venkatasubbaiah, K. (2006). Spatiotemporal growing wave
fronts in spatially stable boundary layers. Physical Review Letters, 96(22), 224504.
62. Sengupta, T. K., Rao, A. K., & Venkatasubbaiah, K. (2006). Spatiotemporal growth of dis-
turbances in a boundary layer and energy based receptivity analysis. Physics of Fluids, 18,
094101.
63. Sengupta, T. K., Sircar, S. K., & Dipankar, A. (2006). High accuracy compact schemes for
DNS and acoustics. Journal of Scientific Computing, 26(2), 151–193.
64. Sengupta, T. K., Unnikrishnnan, S., Bhaumik, S., Singh, P., & Usman, S. (2011). Linear spatial
stability analysis of mixed convection boundary layer over a heated plate. Program in Applied
Mathematics, 1(1), 71–89.
65. Sengupta, T. K., Bhaumik, S., & Bhumkar, Y. (2012). Direct numerical simulation of two-
dimensional wall-bounded turbulent flows from receptivity stage. Physical Review E, 85(2),
026308.
66. Sengupta, T. K., Bhaumik, S., & Bose, R. (2013). Direct numerical simulation of transitional
mixed convection flows: Viscous and inviscid instability mechanisms. Physics of Fluids, 25,
094102.
67. Shaukatullah, H., & Gebhart, B. (1978). An experimental investigation of natural convection
flow on an inclined surface. International Journal of Heat and Mass Transfer, 21, 1481–1490.
68. Sparrow, E. M., & Husar, R. B. (1969). Longitudinal vortices in natural convection flow on
inclined plates. Journal of Fluid Mechanics, 37, 251–255.
69. Sparrow, E. M., & Minkowycz, W. J. (1962). Buoyancy effects on horizontal boundary-layer
flow and heat transfer. International Journal of Heat and Mass Transfer, 5, 505–511.
70. Skote, M., Haritonidis, J. H., & Henningson, D. S. (2002). Varicose instabilities in turbulent
boundary layers. Physics of Fluids, 14, 2309–2323.
71. Tollmien, W. (1931). Über die enstehung der turbulenz. I, English translation. NACA TM 609
72. Tumin, A. (2003). The spatial stability of natural convection flow on inclined plates. ASME
Journal of Fluids Engineering, 125, 428–437.
73. Unnikrishnan, S. (2011). Linear stability analysis and nonlinear receptivity study of mixed
convection boundary layer developing over a heated flat plate. M. Tech. thesis (I.I.T. Kanpur,
2011)
74. Van der Pol, B., & Bremmer, H. (1959). Operational calculus based on two-sided Laplace
integral. Cambridge, UK: Cambridge University Press.
75. Wang, X. A. (1982). An experimental study of mixed, forced, and free convection heat transfer
from a horizontal flat plate to air. ASME Journal of Heat Transfer, 104, 139–144.
76. Wu, R. S., & Cheng, K. C. (1976). Thermal instability of Blasius flow along horizontal plates.
International Journal of Heat and Mass Transfer, 19, 907–913.
77. Zhang, S. L. (1997). GPBi-CG: Generalized product-type methods based on Bi-CG for solving
Non symmetric linear systems. SIAM Journal on Scientific Computing, 18(2), 537–551.
78. Zuercher, E. J., Jacobs, J. W., & Chen, C. F. (1998). Experimental study of the stability of
boundary-layer flow along a heated inclined plate. J. Fluid Mech., 367, 1–25.
Chapter 4
Nonlinear Theoretical
and Computational Analysis
of Fluid Flows

Morkovin [33] classified transition to turbulence in to two main types: (i) The clas-
sical primary instability route whose onset is marked along with the presence of TS
waves (as in ZPGBL) and (ii) the bypass routes, which encompass all other possible
transition scenarios that do not exhibit TS waves. Unfortunately, this is too simplistic
a classification scheme for the following reasons.
It is believed that canonical ZPG flow past a flat plate experience disturbance
growth via linear spatial viscous instability theory governed by the OSE. This pri-
mary instability is indicated by the appearance of TS wave, as was experimentally
verified by Schubauer and Skramstad [47]. Ever since, this has become the center-
piece of instability research for external flows past streamlined bodies. We often
loose sight of the fact that this experiment achieved success after repeated attempts,
starting with creation of a ultra-quiet wind tunnel, with extremely low noise. In this
tunnel, the investigators imposed monochromatic time harmonic excitation inside the
boundary layer forming over a ZPG flow to create TS wave. Acoustic disturbances
imposed from outside was not at all effective in creating TS wave. Thus, experimental
verification of transition process is far removed from the natural transition in many
respects. The major distinction comes from the fact that in natural setting, the bound-
ary layer is excited simultaneously by multiple frequencies by vortical, entropic and
acoustic disturbances. As a consequence, the response field will be a combination
of multiple TS waves convoluting each other. This has been demonstrated by con-
sidering multi-periodic excitation in [60] for Blasius boundary layer, where all the
constituent TS waves were unstable, yet the composite response field obtained from
solution of the OSE did not show the response that could be construed as TS wave
for a long time, essentially caused by wave cancellation due to phase shift among the
constituents. Thus, TS waves have been noted in extremely controlled experiments.
It is noted that in the presence of adverse pressure gradient, such spatial instabilities
are stronger in terms of lower critical Reynolds number and higher growth rate.
Flow transitions which do not display TS waves as a marker of instability, are
said to follow the bypass route. A similar point of view was also advanced in [6].
© Springer Nature Singapore Pte Ltd. 2019 223
T. K. Sengupta and S. Bhaumik, DNS of Wall-Bounded Turbulent Flows,
https://doi.org/10.1007/978-981-13-0038-7_4
224 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows

There are many such flows, where transition are caused without the presence of
monochromatic unstable TS waves. Prime examples of bypass transition are flows
past bluff bodies, Couette and pipes flows, leading edge contamination on swept
wings, two-dimensional and three-dimensional roughness effects, etc. In some of
these flows, unstable TS waves are not noted; instead disturbances are seen to grow
with time due to the presence of inflection point(s) for the velocity profile. Flow past
a bluff body is a typical example. In contrast, plane Poisseuille flow displays a critical
Reynolds number in excess of 5770, while the flow is noted experimentally to be
turbulent at significantly lower Reynolds numbers (around 1000 in [10]). This flow
is also not characterized by an inflectional velocity profile, yet it displays subcritical
transition. Another example of interest is the flow undergoing transition, right at the
leading edge of a swept wing of an aircraft. There are no definitive physical mech-
anisms identified in the system portrait of Fig. 3.1 in [33] for any of these bypass
transition examples. Instead, question marks have been raised related to the possibil-
ity of nonlinear, nonparallel or some unknown mechanisms as potential reasons. In
this chapter, we focus upon the physical mechanism for a few prototypical examples
of flow transition with the help of results from carefully designed receptivity experi-
ments and accurate flow computations. Next, we study cases where the excitation is
applied at the free stream without introducing any apparent time scale(s), unlike the
multi-periodic cases discussed in the previous chapter.
The idea that a distant vortex can induce a small longitudinal adverse pressure
gradient to destabilize a wall-bounded flow was mooted in [76], while studying
dependence of critical Reynolds number upon free-stream turbulence (FST). Monin
and Yaglom [32] in discussing this work noted that the change in critical Reynolds
number by the small longitudinal adverse pressure gradient is due to a sequence
of unsteady separations, presumably created by a train of vortices embedded in the
FST. The assumption implicit in this scenario is that the effect is connected with the
generation of fluctuations of longitudinal pressure gradient by these disturbances,
leading to the random formation of individual spots of unstable S-shaped velocity
profiles [32].
As mentioned above, that exciting a laminar flow by acoustic excitation from
the free stream in [47] was not successful, while vortical disturbances inside the
boundary layer triggered transition. In the experiment of Leib et al. [27], authors
used vibrating ribbon in the free stream to cause transition. It has also been shown in
[55] that free stream excitation can be effective also in the linear viscous instability
theory model given by the OSE.
To model the FST effects on boundary layer, experiments have been performed
in [28, 54, 56], to study the unit process of a convecting vortex moving at a con-
stant height and speed over a zero pressure gradient boundary layer. In this point
of view, FST can be thought of as an assembly of vortices moving aperiodically at
arbitrary heights. One of the key features of this experimental observation was that
the ensuing instability is subcritical with respect to instability given by the OSE
(displaying unstable TS waves) and this vortex-induced instability gives rise to
unsteady separation. As there are no linear theories available for studying vortex
dominated flows, a nonlinear theory was developed [56], which is explained in the
4 Nonlinear Theoretical and Computational Analysis of Fluid Flows 225

following for a general nonlinear instability for incompressible flows from NSE
without making any assumption, based on mechanical energy.

4.1 Nonlinear Instability Theory Based on Mechanical


Energy

Landahl and Mollo-Christensen [26] noted that to understand turbulence, one must
consider growth of total mechanical energy and not just simply the disturbance tur-
bulent kinetic energy (TKE). According to these authors, “it is possible to understand
such behaviour by studying the redistribution of the total mechanical energy of the
flow”. In contrast to popular practice of characterizing turbulence by discussing
about TKE, one must consider the roles of fluctuating pressure, as Morkovin [33]
suggested that unsteadiness during bypass transition can be due to shear noise term
in the Poisson equation for the static pressure. This prompted the authors in [56] to
develop an equation for the total mechanical energy from NSE, for the equilibrium
and the disturbance flow fields. Such equations for receptivity must also be capable
to explain classical linear theories as special cases. This has been demonstrated in
[59, 60] showing the existence of TS waves caused by spatially localized excitation
by this energy-based approach. In the following, we show the derivation of the equa-
tion for mechanical energy of incompressible fluid flow. NSE is written using vector
notation in rotational form, with the vector identity for the convective acceleration
term,

→ −
→ 1 −
→ −
→ −

( V · ∇) V = ∇(| V |2 ) − V × (∇ × V )
2
to yield the NSE as,


∂V 1 −
→ −
→ → ∇p −

+ ∇(| V |2 ) − V × −
ω =− + ν∇ 2 V (4.1)
∂t 2 ρ

Further, one can alter the diffusion term with the following identity,

→ −
→ −

∇ 2 V = ∇(∇ · V ) − ∇ × (∇ × V )

For incompressible flows, one gets the rotational form of NSE given by,

→  → 

∂V −
→ → p | V |2
− V ×−
ω = −∇ + − ν(∇ × −

ω) (4.2)
∂t ρ 2

Reverting back to Laplacian form of diffusion term and defining the total mechan-
ical energy as


p | V |2
E= +
ρ 2
226 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows

NSE is written with E as




∂V −
→ → −

− V ×−
ω = −∇E + ν∇ 2 V (4.3)
∂t
Taking a divergence and making incompressibility assumption for the equation
above, one gets the governing Poisson equation for E as,

→ →
∇2E = ∇ · ( V × −
ω) (4.4)

Further using the identity,



→ → −
→ −

∇ ·(V ×−
ω)=−

ω · (∇ × V ) − V · (∇ × −

ω)

one gets finally the governing equation for the total mechanical energy as


∇2E = −

ω ·−

ω − V · (∇ × −

ω) (4.5)

Thus, the distribution of E is directly related to enstrophy (the first term on the
right hand side). Without solving this equation, one can comment about the solution
from the sign of the right-hand side. It indicates the presence of a source or a sink
of E, as given by the property of Poisson equation for energy in [73]. A negative
sign signifies a local source. Thus, it may appear that the enstrophy stabilizes E
in Eq. (4.5). The development here is for nonlinear evolution of a flow, by tracing
destabilization of an equilibrium state. A more subtle picture emerges, if one divides
E into an equilibrium and a disturbance part with the help of the subscripts, m and
d in, E = Em + εEd . If one is interested in the growth of small perturbation applied
to equilibrium state, then the parameter ε represents a small value, and apply same
splitting scheme for all the quantities appearing in Eq. (4.5). The disturbance energy
equation for Ed is obtained by subtracting the equation for equilibrium quantities
from the equation for total quantities as,

→ −
→ −

∇ 2 Ed = 2 −

ωm·−

ω d + ε−

ωd ·−

ω d − V m · (∇ × −

ω d ) − V d · (∇ × −

ω m ) − ε V d · (∇ × −

ω d)
(4.6)

which can describe the onset of instability for an equilibrium flow. In [59, 60], exis-
tence of TS waves caused by spatially localized excitation for a boundary layer is
shown, by taking the equilibrium state as the Blasius profile and solving the linearized
version in Eq. (4.6). In the developed theory based on NSE without any assumption,
growth of primary disturbances are due to interactions of velocity and vorticity field
acting as source terms on the right-hand side of Eqs. (4.5) and (4.6). NSE is a con-
sequence of conservation of translation momentum, and E remains conserved, as
viscous term is absent in Eq. (4.5). Any instability that is observed, is caused by
growth of disturbance quantity, whose supply must be from the equilibrium flow or
by external agencies of boundary condition(s). Thus, the major issue is about how
4.1 Nonlinear Instability Theory Based on Mechanical Energy 227

the energy is initially exchanged from the equilibrium to the disturbance field and
this is clearly brought out by the first term on the right-hand side of Eq. (4.6), which
indicates an interaction between equilibrium and disturbance vorticity fields or the
disturbance enstrophy.
The above equation required no approximation on the equilibrium and disturbance
fields. Even for unsteady equilibrium flows, one can trace nonlinear instantaneous
instability. For the creation of surface gravity waves and side-band instability of
surface gravity waves by the mechanism in [2], unsteadiness/instability onset is
triggered by the boundary condition.

4.2 Vortex-Induced Instability: Application


of a Nonlinear Theory for Total Mechanical Energy

We focus on the topic of bypass transition by studying the unit process caused by a
single convecting vortex. The same physical mechanism is seen in other examples
of unsteady flow separation (as discussed in [12]), near-wall eddy formation in tur-
bulent boundary layer (as studied in [7]) In [43, 72], the authors have also discussed
formation of hairpin vortices in the near-wall region of turbulent flows. Other studies
[14, 39, 40, 60] considered the scenario where a vortex placed above a plane wall
caused the vortex to move and thereby create a thin unsteady boundary layer over
the wall. All these are examples of vortex-induced instability.

4.2.1 An Experimental Observation of Vortex-Induced


Instability

In [28, 54], a vortex with finite core size was created experimentally by a rotating
and translating circular cylinder, whose strength is Γ , and which is at a fixed height
from the plate (H ), and the sign of circulation can be easily controlled. The main
emphasis of the experiment was to control all the relevant parameters, so that the
resultant bypass transition can be reproduced. In Fig. 4.1, the schematic of the flow
is shown. It was seen that a slowly convecting vortex of anticlockwise circulation,
creates transition/unsteady separation ahead of it [28, 51, 56].
The experiment was performed in a recirculating water tunnel. One of the cases
is reproduced here to highlight the receptivity mechanism of the shear layer to a
convecting vortex in the free stream. In this experiment, the boundary layer was
formed on a flat plate, held vertically on its edge in the tunnel. A coherent bound
vortex is created by rotating a circular cylinder of diameter 15 mm, whose axis
was along the spanwise direction of the plate. The cylinder can be rotated in either
directions and was rotated at Ω = 5 r.p.s. in the anticlockwise direction, for the case
shown in Fig. 4.2. For flow visualization, food dye was released from six dye ports
located 88 mm downstream from the leading edge of the plate, as seen in Fig. 4.1.
228 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows

Rotating and translating cylinder


(a) x

c
Dyelines

U


(b) Flat plate
Leading edge of the plate (L.E.)
Dye port

Flap
U

H1
c

Fig. 4.1 Schematic of the experimental set-up. a Side view and b top view as seen in the tunnel.
Broken line boundary in b indicates the computational domain

The Reynolds number based on the diameter of the cylinder and free stream
speed (U∞ = 162 mm s−1 ) was 2600. The cylinder was convected at c = 0.15U∞ ,
at a height H = 90 mm. The ratio of surface speed to the relative speed of the free
stream and cylinder was 1.71. The noise level of the water channel is 1% at maximum
speed. It is known [1, 58] that a rotating cylinder ceases to shed coherent vortices
associated with K ármán vortex streets, when the surface speed is more than 1.5
times the free-stream speed. In this experiment, the bound vortex circulation is fixed
by controlling the rotation rate of the cylinder, and this controls the dynamics of the
flow.
In Fig. 4.2, flow visualization sequences indicate unsteady separation followed
by bypass transition in frames (b − h). The dye filaments are essentially parallel
at the onset (as in frames a − d ), showing overall two-dimensionality during this
stage. In frame (b), the dye filaments released very close to the plate is lifted up
due to the imposed disturbances, with negligible spanwise spreading. The location
where unsteadiness is seen is laminar in the absence of the convecting vortex, indi-
cating the subcritical nature of the instabilities. Only flow visualization was used, as
4.2 Vortex-Induced Instability: Application of a Nonlinear … 229

(a) Dye filaments U (b)

c Leading
edge of
flat
Rotating cylinder plate

(c) (d)

(e) (f)

(g) (h)

Fig. 4.2 Bypass transition created by a counter-rotating vortex for (U∞ = 16.26 cm s−1 , c =
0.154, H = 9 cm and Ω = +5 r.p.s.) for the experimental arrangement shown in Fig. 4.1

intrusive measurements changed the dynamics drastically. Therefore to quantify the


experimental observation, we have undertaken a numerical simulation of the prob-
lem described next. As we are interested in the onset stage of the instability, it is
sufficient to perform a 2D simulation of NSE.

4.3 Numerical Simulation of Vortex-Induced Instability

Vortex-induced instability problem is used here as a typical case to demonstrate


bypass transition caused by a convecting vortex in the free stream. In [56], this
problem was solved using (ψ, ω)-formulation due to its inherent advantages. The
divergence-free or solenoidality condition for velocity and vorticity is automatically
satisfied for 2D flow fields, for this formulation. However, for 3D disturbance compu-
tations, direct extension of (ψ, ω)-formulation by the vorticity (−→ω )-vector potential
230 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows

y +Γ
c
For Field Boundary C
B
U∞
Outflow
Boundary
Inlet H ymax
Edge of Shear Layer

δout
A Wall Boundary
xin O
D x
H xout

c

Fig. 4.3 Computational domain used in studying vortex-induced instability by an isolated convect-
ing (at a speed, c) vortex (of circulation, Γ ) in the free stream over a flat plate supporting a zero
pressure gradient boundary layer



( ψ ) formulation is not easy. The authors in [4, 5], developed a velocity-vorticity
formulation in a staggered grid, in solving the 3D receptivity problem from excita-
tion stage of laminar flow over a flat plate, to fully developed turbulent state. Here,
the same formulation is used for the 2D flow field shown in Fig. 4.3. In this case,
receptivity of ZPGBL to convected vortex in the free stream is investigated. This
problem requires time-dependent boundary forcing to excite physical instability of
the flow. The equilibrium flow is calculated by solving NSE, without the application
of external forcing. The origin of the co-ordinate system is placed at the leading
edge of the plate, whereas the computational domain starts slightly ahead of it. A
counter-clockwise rotating vortex is placed above the plate and this translates from
left to right at the constant height of H , with a constant speed, c. Even though
this free stream vortex convects far outside the shear layer, the presence of the flat
plate requires an image vortex below the plate, to ensure zero wall-normal velocity
boundary condition, as shown in the figure.

4.3.1 Velocity-Vorticity Formulations for 2D Flows

Vorticity transport equation is found very relevant for the analysis and solution of
viscous incompressible flows [48, 53, 65]. Attendant velocity field can be obtained
from the solution of the Poisson equation for the velocity components. In this for-
mulation, the unknowns are velocity components, u, v, and the out of plane vorticity


component, ω. For 2D flows, the non-dimensional governing equations for ( V , ω)-
formulation in Cartesian co-ordinate are given in [9, 19, 36, 81].
Here, we solve the governing equation in a transformed plane by high accuracy
compact scheme for discretization in finite difference framework. More details about
4.3 Numerical Simulation of Vortex-Induced Instability 231

the numerical methods, used grids etc. are given in [3, 49]. The governing vorticity
transport and velocity Poisson equations in the transformed (ξ, η)-plane are given as
  
∂ω 1 ∂ 1 ∂ ∂ h2 ∂ω
+ (uω) + (vω) = 1 1
∂ξ h1 ∂ξ
∂t h1 ∂ξ h2 ∂η Re h1 h2
 
∂ h1 ∂ω
+ ∂η h2 ∂η
(4.7)

    
∂ h2 ∂u ∂ h1 ∂u ∂ω
+ =− (4.8)
∂ξ h1 ∂ξ ∂η h2 ∂η ∂η

    
∂ h2 ∂v ∂ h1 ∂v ∂ω
+ = (4.9)
∂ξ h1 ∂ξ ∂η h2 ∂η ∂ξ

where h1 and h2 are the scale factors of the transformation given by, h1 = xξ2 + yξ2
 −

and h2 = xη2 + yη2 . The divergence of the velocity field, Dv = ∇ · V , in the com-
putational (ξ, η)-plane is given as,
 
1 ∂(h1 v) ∂(h2 u)
Dv = + (4.10)
h1 h2 ∂η ∂ξ

After solving Eq. (4.8) for u-component of velocity, v-component of velocity is


obtained by integrating the right hand side of Eq. (4.10) as equal to zero (solenoidality
condition) given as,
 η 
1 ∂(h2 u)
v(ξ, η) = v(ξ, 0) − dη (4.11)
h1 0 ∂ξ

This ensures satisfaction of the divergence-free condition of the velocity vec-


tor numerically in the computational plane. For the receptivity of the ZPG flow,
boundary condition on v-velocity at the far-field boundary is satisfied asymptoti-
cally and hence, Eq. (4.11) is suitable for external flows, which requires only the
prescription of v at the bottom boundary segment.

4.3.2 Boundary Conditions

The boundary conditions used in solving this problem are listed as following,
(i) At the inflow (segment AB of Fig. 4.3), uniform inlet velocity U∞ is imposed
along with the contribution imposed by the free stream vortex and its image system.
232 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows

Corresponding condition is applied on v-velocity component, as computed by the


free stream inviscid vortex and its image. The induced stream function created by
the finite core translating vortex in conjunction with the uniform flow is given by,
 
(y − H )(d /2)2 (y + H )(d /2)2 Γ x̄2 + (y + H )2
ψ∞ = U∞ y − (U∞ − c) + + Ln 2
x̄ + (y − H )
2 2 x̄ + (y + H )
2 2 4π x̄ + (y − H )2

where d is the core diameter of the convecting vortex of strength, Γ , convecting at


a constant height, H . The convection speed of the vortex is given by c, and hence
the displacement effect of the finite core vortex is determined by the relative speed,
(U∞ − c), and the circulation effect is given by the last term on the right hand
side. Also note that the time dependence of the boundary condition is given by the
translated co-ordinate, x̄ = x0 − ct, with x0 indicating the initial co-ordinate of the
convecting vortex.
(ii) At the wall, no-slip condition is imposed on both the components of velocity.
Wall vorticity is computed based on its kinematic definition.
(iii) In the segment AO of Fig. 4.3, vorticity and the v are prescribed to be zero, by
symmetry condition. The boundary condition used on u is also fixed by the condition
given by ∂u/∂y = 0.
(iv) At the far-field boundary: ω = 0, and u, v are as calculated from the Biot–
Savart law caused by the free stream vortex and its image.
(v) At the outflow, vorticity is calculated using radiative Sommerfeld boundary
condition given as
∂ω ∂ω
+ Uc =0
∂t ∂x
where the convective speed Uc is chosen to be U∞ . The condition used on u is
∂ 2 u/∂x2 = 0, whereas a condition given by, ∂v/∂x = ω + ∂u/∂y is used for v.

4.3.3 Numerical Methods and Grids

A staggered variable arrangement, as used in [20, 36], is adopted here. Staggered


grid arrangement is needed to reduce errors, which is smaller in staggered grid,
as compared to non-staggered grid arrangement, shown using this formulation in
[24]. In the transformed (ξ, η)-plane, the staggered grid is shown in Fig. 4.4, with
contravariant components of velocity marked at the mid-points of control surface
over which the component is normal. In Fig. 4.4, the vorticity is placed at the nodes.
In the numerical methods used, time advancement of ω is done by RK4 -scheme,
while convective derivatives ( ∂ω , ∂ω ) are discretized using OUCS3 scheme developed
∂ξ ∂η
in [57] and second derivatives are discretized using second-order central difference
scheme. To evaluate convective derivatives in the staggered grid, it is required to
interpolate u- and v-components at the locations of the vorticity; which is carried out
by an optimized compact mid-point interpolation scheme of [35].
4.3 Numerical Simulation of Vortex-Induced Instability 233

ω i,j+1
v i-1/2,j+1

u i,j+1/2
ω i-2,j ω i -1,j ω i,j ω i+1,j ω i+2,j
v i-3/2,j v i-1/2,j v i+1/2,j v i+3/2,j

ui -2,j-1/2 u i-1,j-1/2 u i,j-1/2 u i+1,j-1/2 u i+2,j-1/2


ω i,j-1
v i-1/2,j-1

u i,j-3/2
ω i,j-2
v i-1/2,j-2
u i,j-5/2
ω i,j-3
v i-1/2,j-3

ω u v


Fig. 4.4 Staggered grid system used in ( V , ω)-formulation for 2D problems. The velocity com-
ponents are at the mid-point of the elemental surface over which it is normal

A value of ε = 10−8 , is used for solving Poisson equations with the Bi-CGSTAB
method of [80]. Same numerical methods are used to obtain the equilibrium and
disturbance flow, as excited by the free stream vortex. Once the steady equilibrium
flow is established, the convecting vortex is activated to excite the flow.
After solving Eq. (4.7) for the vorticity, Eq. (4.8) is solved for the u-component of
velocity, while v-component of velocity is obtained by integrating the right hand side
of Eq. (4.10), as equal to zero (solenoidality condition). The non-dimensionalized
equations have been obtained with a length scale (defined later) and the free-stream
speed of the oncoming flow as the velocity scale. From these two scales, the time
scale is constructed and all computational results are in non-dimensional units.
To solve the governing equations in the computational domain of Fig. 4.3, the
parameters are similar to those in the experiments of [28], except for the strength of
the convecting vortex, which cannot be measured and is treated here as the parameter
of the problem. In solving the problem computationally, we define a length scale,
L, based on which the Reynolds number is 105 . The computational domain extends
over, −0.05 ≤ x∗ ≤ 20, i.e., the domain includes the leading edge of the plate. The
234 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows


height of the computational domain is taken as ymax = 1.0, with the wall resolution
−4
given by Δy = 3.688 × 10 , and this is half the wall resolution taken in [56]. For
the present computations, the strength of the convecting free stream vortex is taken
as Γ = 2.0 (as compared to 9 taken in [56]). This is important for experimental
and computational study of flow instability, where we need as small an excitation as
possible to study receptivity. The free stream convected vortex moves at constant fixed
height of 2, at a fixed speed. The Reynolds number based on displacement thickness
at the outflow of the undisturbed flow is 2432.44, and thus the domain is more than
five times, that was taken in [56]. The flow computed in the latter reference remained
sub-critical over the computational domain, and the present domain is significantly
longer.
The authors in [7, 38], have used numerical methods, which are O(ΔxΔt, ΔyΔt)
accurate. In contrast, the computations in [56] have used OUCS3 schemes for spatial
discretization of convection terms, which has more than seven times higher spectral
resolution as compared to second-order accurate schemes. The dispersion error of the
method is lower for the schemes in [56]. The present formulation and used staggered
compact scheme, further improves accuracy by reducing aliasing and dispersion
errors, as studied in [61]. It has been firmly established that in computing space-time
dependent problems, dispersion error is the largest source of error, as compared to
spatial and temporal discretization errors viewed in isolation. In [69], it has been
conclusively demonstrated that using a sixth order spatial compact scheme, along
with fourth order accurate temporal scheme for model convection equation, the error
scales as O(Δx4 ) in a non-uniform grid. Thus, the role of error dynamics, as described
in [49, 61, 75] cannot be overemphasized.
At the inflow and the top-lid of computational domain, one obtains flow variables
induced by Biot–Savart law due to the convecting vortex outside the computational
domain. At the outflow, the fully developed condition on the wall-normal component
of velocity is used. The above conditions are used to derive variables at all boundary
segments. The flow is started impulsively, with the initial location of the vortex


being ahead of the leading edge. In computing the flow by ( V , ω)-formulation,
computations have been performed with a time step of Δt = 8 × 10−5 .

4.3.4 Grid Generation

In order to resolve the points near the leading edge (which is the Goldstein singu-
larity), a non-uniform stretched grid is used in the streamwise direction that clusters
points at the leading edge. The tangent-hyperbolic function is used for grid cluster-
ing as described in [18], which has been shown to cause lower aliasing error [49].
Similar clustering of grid points have been performed in the wall-normal direction
as well, to accurately resolve the boundary layer and events very close to the wall
for receptivity problem, as one expects high gradients of the flow variables near the
wall. This grid-point clustering at the leading-edge and wall, causes the physical
4.3 Numerical Simulation of Vortex-Induced Instability 235

problem defined in the Cartesian co-ordinate (x, y) to be transformed to the uniform


computational (ξ, η)-coordinate system.
The specific form of grid transformation function in the streamwise direction is
given for, 0 ≤ ξ ≤ ξ1 and xin ≤ x ≤ xs by,
 
tanh[βx (1 − ξ )]
x(ξ ) = xin + (xs − xin ) 1 −
tanh βx

and for ξ1 ≤ ξ ≤ 1 (and xs ≤ x(ξ ) ≤ xout )


  
βx ξ − ξ1
x(ξ ) = xs + (xs − xin ) (4.12)
tanh βx ξ1

where, ξ1 = 1
and
1+A1   
xout − xs tanh βx
A1 =
xs − xin βx

Similarly, due to the grid clustering near the wall, the grid transformation function
in the wall-normal direction is given as
 
tanh[βy (1 − η)]
y(η) = ymax 1 − (4.13)
tanh βy

where, 0 ≤ η ≤ 1. Here, βx and βy are parameters which control the grid cluster-
ing in the streamwise and wall-normal direction, respectively. For most of the cases
reported, βx = βy = 2, have been used. Because of the transformation of the problem
from the physical (x, y)-plane to computational (ξ, η)-plane, one defines the scale
1 1
factors of transformation h1 and h2 by h1 = (xξ2 + yξ2 ) 2 and h2 = (xη2 + yη2 ) 2 , respec-
∂x ∂y
tively. Here, xη = ∂η = 0 and yξ = ∂ξ = 0, and the scale factors are simply given
as h1 = xξ and h2 = yη . The above grid transformation in the streamwise direction
given by Eq. (4.12) ensures the continuity of both h1 and dh dξ
1
at x = xs . The grid
used has 1001 points in the streamwise direction and 301 points in the wall-normal
direction.

4.3.5 Numerical Results and Discussion

The problem of vortex-induced instability experimentally and analytically studied in




[28, 56], has been solved here using ( V , ω)-formulation as a 2D problem. As noted,
the equilibrium solution is obtained first by solving NSE, without the convecting free
stream vortex. Once this is obtained, the free stream vortex is started with constant
236 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows

Fig. 4.5 Stream function contours plotted in the computational domain at the indicated times.
Arrowheads at the top show the streamwise location of the convecting vortex

convection speed (c = 0.3) from the initial location: (x = −5, y = 2). Presence of
the convecting free stream vortex above the flat plate creates an image vortex, as
shown in Fig. 4.3.
In Fig. 4.5, stream function contours are plotted at indicated times chosen to
show physically important events. For the frame at t = 20, the free stream vortex
is just downstream of the leading edge of the plate. As the vortex has circulation
4.3 Numerical Simulation of Vortex-Induced Instability 237

in the anti-clockwise direction, the action of the vortex is to lift up the boundary
layer ahead of the vortex. At the foot of the vortex (outside the boundary layer),
the induced velocity is maximum. Thus upstream of the vortex, one would notice a
favourable pressure gradient, while in the downstream direction, the induced velocity
reduces with x, i.e., an adverse pressure gradient is created. As the free stream vortex
convects at a constant height, its instantaneous stream-wise location is shown by
an arrowhead. With passage of time, for farther downstream location of the vortex,
the boundary layer experiences sustained adverse pressure gradient and in the frame
(b) of Fig. 4.5 one notices formation of unsteady separation bubbles near x = 4.
Downstream of these bubbles, the flow in the vicinity of the flat plate experiences
additional adverse pressure gradient. This is readily seen in frames (c) and (d) at t =
28 and 30, respectively, when one notices increased number of unsteady separation
bubbles, all of which convect downstream. That such a single convecting free stream
vortex, far out in the inviscid part of the flow, can cause bypass transition is noted
in frame (e) of Fig. 4.5. Unsteady separation bubbles, as a consequence of bypass
transition, on the wall were conjectured in [32] to be caused as a result of buffeting
of the boundary layer by FST vortices. The present study is for the unit-process
and provides a physical mechanism of a bypass transition process. In the present
study no model is required, with the effects governed by NSE. For example in [38],
the primary vortex forms as a consequence of unsteady flow evolution and does
not move at constant speed, which was modeled as a Batchelor vortex moving at
constant speed.
The unsteady separation and vortical structures near the wall are created due to
effects of free stream convecting vortex, as shown in Fig. 4.6. These are at the same
time instants, as shown for the stream function contours in Fig. 4.5. Despite the fact
that free stream acoustic excitation could not trigger disturbances in the experiments
of [47], in [27, 56] and here, the coupling between free stream vortical excitation
with viscous disturbances inside the boundary layer is established. An explanation
of this is provided in [55] with the help of spatial linear instability described by the
OSE. It is shown that boundary layers can support disturbances created by sources
inside or outside a shear layer by what is referred to as wall- and free stream-modes,
respectively. When free stream-modes are excited (as in here), those in turn cause
the wall-mode to be excited, by a coupling mechanism that ensures homogeneous
boundary conditions at the wall. The growth of the primary bubble and appearance of
subsequent separation bubbles are due to an instability, where the disturbance field is
enriched from the equilibrium flow. The conditions and mechanism by which these
instabilities appear is further discussed next.
238 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows

4.3.6 The Instability Mechanism

The experimental and accompanying computational results display the existence of


a receptivity mechanism inside the shear layer as a consequence of a single vortex
migrating in the free stream at a constant speed. The role of various parameters
responsible for this instability is by the redistribution of E in the flow. The equation


for E = p/ρ + 21 V 2 , for incompressible flows is the divergence of the rotational
form of the NSE given by Eq. (4.5). E depends on the enstrophy and its contribution

Fig. 4.6 Vorticity contours plotted in the computational domain at the indicated times, as in Fig. 4.5.
Same contour values are plotted in all the frames. Arrowheads at the top show the streamwise location
of the convecting vortex
4.3 Numerical Simulation of Vortex-Induced Instability 239

on the right hand side appears as positive, i.e., stabilizing. However, as one views the
evolution of Ed given in Eq. (4.6), one notices the corresponding term as a product
and leads to its growth when the primary and disturbance vorticities are of opposite
sign, indicating a transfer of energy from primary to disturbance flow. At the same
time, the third and fourth terms on the right hand side of Eq. (4.6) indicates that the
spatial variation of the vorticity field can interact with the velocity field to contribute
to instability, when this is a negative quantity. Here Eq. (4.6) has been used to explain
the complete nonlinear evolution of disturbances.
Here the velocity and vorticity fields at t = 20 are taken as representative equi-
librium flow. The sign of right hand side being positive or negative, indicates a sink
or source, respectively, in Eq. (4.6). Hence for the Ed equation, a negative right hand
side anywhere would indicate a source of Ed at that point in the flow field. In Fig. 4.7,
these distributed sources are plotted as negative contours. At t = 20, there are two
sites from where instability originates, one at the leading edge and the other on
the plate, downstream of leading edge. It is seen that the leading-edge instability is
the weaker of the two and the major one originates near x = 2.4, as was also seen
in the vorticity contours in Fig. 4.6. Disturbance energy structures from these two
regions remain distinct at t = 25. But from t = 28 onwards, these two sources of
Ed interact, as is evident from the bottom three frames in Fig. 4.7, where the spike
forming at the downstream site interacts with the vortical structure originating from
the leading edge.
In stream function and vorticity contour plots the spike is evident near x = 4.5 at
t = 28 in the form of a secondary bubble. Thus, it is important to include the leading
edge in the analysis, otherwise one would compute the unimpeded spike stage, as in
[38, 72]. However beyond t = 28, the instability originating from the leading edge
terminates before the downstream spike and subsequently the distance between the
instabilities further increases. The present analysis based on the right-hand side of
Eq. (4.6), more clearly reveals the physical nature of the problem compared to the
information from stream function and vorticity contours.
In the experimental study it is shown that a vortex with anti-clockwise circulation
creates instability ahead of it. The coupling between the convecting vortex outside the
shear layer and the generated unsteady vortical field inside is explained theoretically
by developing an equation for Ed . This vortex-induced instability is caused by the
adverse pressure gradient due to the translating vortex in the free stream. Effects of
the finite core translating vortex is shown in Fig. 4.8, indicating the displacement and
circulatory components of the input disturbances and associated induced pressure
gradient, at the edge of the shear layer. It is noted that unsteady separation occurs at
a lower value of adverse pressure gradients, indicated here by plotting the Falkner–
Skan parameter, m, plotted as functions of space and time in various frames of
Fig. 4.8. This figure should be contrasted with Fig. 4.8a, b in [48], where the induced
adverse pressure gradient by the free stream convecting vortex of strengths Γ = 14
and 30, indicated values of m, which were far in excess of what is needed for steady
separation. Thus for vortex-induced instability, it is not merely the strength of adverse
pressure gradient that triggers unsteady separation, but the time duration over which
the adverse pressure gradient acts matter more.
240 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows

Fig. 4.7 Contours of the right-hand side of the disturbance energy Eq. (4.6). The negative values
indicating disturbance energy sources are plotted as dark contours

4.4 Enstrophy Transport Equation: A New Approach to


Nonlinear Receptivity Theory

We are already familiar with the nonlinear receptivity/ instability theory based on
total mechanical energy. The governing equation for this is derived from NSE without
making any assumption, by taking, the divergence of this conservation equation
for translational momentum equation. Thus, this provides information about the
4.4 Enstrophy Transport Equation: A New Approach … 241

(i) =2
3
Displacement effect
Circulation effect
2

0
0 5 10 15 20 25 30 35
(ii)
c = 0.1
Pressure gradient

0.04 c = 0.4
0.02 c = 0.8

0
-0.02
-0.04
0 5 10 15 X 20 25 30 35
(iii) c=0
Pressure gradient

0.04 c = 0.1
c = 0.2
0.02 c = 0.4
c = 0.8
0
-0.02
-0.04
0 10 20 30 40 50 60
Non-dimensional time
(iv) c = 0.2 t=0
0.2 t = Tsep

m 0

-0.2

5 10 15 x 20 25 30

Fig. 4.8 (i) Disturbance stream function versus x caused by circulatory and displacement effects,
by a translating and rotating circular cylinder, while the disturbance source is at xc = 5 and H1 = 6
moving at the indicated propagation speed. (ii) Induced pressure gradient for the case of frame (i)
for the indicated propagation speed. (iii) The same pressure gradient shown as a function of time
for various convecting and non-convecting free stream vortex cases. (iv) Variation of Falkner–Skan
parameter m with x shown at two representative times for the case of c = 0.2. The time Tsep indicates
a large time for plotting m
242 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows

irrotational part of the conservation equation. In fluid mechanics, however, there are
no conservation equation for rotational momentum. This is partially addressed by
looking at the VTE which is obtained by taking a curl of NSE. Even when vorticity
field indirectly provides information about rotationality in a fluid flow, one needs
an appropriate measure of it, and one notes that the enstrophy to fulfill this role.
Interestingly, the turbulence literature alludes enstrophy to provide a measure of
dissipation only, as in [13], who studied the enstrophy transport equation (ETE) for
2D periodic flows and obtained the evolution of integrated enstrophy over the full
domain as  
d 1
||ω||2 = −ν||∇ω||22
2
(4.14)
dt 2

Here, the enstrophy is defined over the full periodic domain by ||ω||22 . Thus, one
notes the effects of diffusion to be strictly dissipative for periodic 2D flows viewed
globally. This shows that viscous diffusion destructs enstrophy in the full domain, for
2D periodic flows. However for the most generic cases of 3D inhomogeneous flows,
authors in [67] have developed a more general ETE, and is the subject of discussion
in the following. We also discuss the role of diffusive terms in fluid flows.
Investigation on the true role of diffusion has remained a problem, ever since the
time when it was considered to strictly stabilize fluid flow by damping disturbances,
an idea attributed to Kelvin, Helmholtz and Rayleigh in [48]. Such heuristic equating
of viscous diffusion with dissipation was the main reason in early instability studies,
which ignored diffusion, as noted in [17, 48]. Unfortunately, such inviscid studies
could not even explain instability of ZPGBL. But the same flow was subsequently
investigated by solving the OSE in [21, 46, 78]. We note that the governing equations
for E and Ed arises entirely from local and convective acceleration terms (in the
absence of body force).
In DNS, one discretizes all terms and numerical solution is obtained without
any models. However, diffusion is often equated with dissipation, as in [82], by
extrapolating Eq. (4.14), which is valid for homogeneous turbulent flows. However
if diffusion is viewed at any instant for any point in the flow, then the effects of
diffusion is not strictly dissipative, as has been explained in [67]. When one looks
at time-averaged TKE globally, effects of diffusion is again seen as dissipative in
[31, 77]. As shown in Eq. (4.34) of [31], time-average of diffusion term in NSE
manifests itself, as combination of (i) a strictly dissipation term and (ii) another
viscous transfer term. The viscous transfer term does not contribute, when integrated
over the flow domain, due to divergence theorem. This term is shown diffusive
for homogeneous turbulence [13]. It is further noted that the viscous transfer term is
negligible at high Reynolds numbers, except within the thin viscous layers, very near
solid surfaces. While on the other hand, the dissipative term is of crucial importance
to turbulence energetics everywhere. Similar observations are made in Sect. 3.3 of
[77], with respect to time-averaged TKE.
Denoting the instantaneous point property of enstrophy by Ω1 = − →
ω ·−→ω , Eq. (4.5)
can be written as,
4.4 Enstrophy Transport Equation: A New Approach … 243



∇ 2 E = Ω1 − V · (∇ × −

ω) (4.15)

This equation shows relevance of Ω1 and the diffusion operator to be central in


distributing E. In [82], a similar equation has been written for static pressure, p (see
Eq. (1.2) of the reference [82]) which in the present notations is given by
 
p
∇ 2
= (Ω1 − ε/ν)/2 (4.16)
ρ

where ε = 2ν sij sij and sij is the symmetric part of the strain tensor. This equation
is valid for both homogeneous and inhomogeneous flows. One notes that the second

→ −

term on the right hand side of Eq. (4.15) can be written as Vν · ∇ V , by drawing
analogy with the term ε/ν, on the right hand side of Eq. (4.16), even though the
right hand side of Eq. (4.15) purely originates from convection term. This confu-
sion prompted the authors in [15, 16, 82], to equate enstrophy with dissipation. To
understand the role of diffusion in creating rotationality for inhomogeneous flows,
an evolution equation is developed for enstrophy, as a point property and its higher
powers. This explains the roles of diffusion, dissipation and creation of rotationality
progressively to smaller scales.

4.4.1 Enstrophy Transport Equation

The role of Ω1 in rotational fluid flow is similar to kinetic energy describing the trans-
lational motion in fluid flow. While vorticity describes rotationality, a measure of it
is naturally obtained via enstrophy unambiguously, describing the energy expended
by the system in creating and sustaining rotationality. In all flows, physical insta-
bilities take an equilibrium state to another one and in the process, the energy of
the system is redistributed into rotational and translation degrees of freedom. Thus,
enstrophy is a natural dependent variable to study transitional and turbulent flows.
We explain instabilities and pattern formations, with the help of ETE derived from
the non-dimensional VTE in tensor notation given for 3D flows by

∂ωi ∂ωi ∂ui 1 ∂ 2 ωi


+ uj = ωj + (4.17)
∂t ∂xj ∂xj Re ∂xj ∂xj

where subscripts, i, j = 1, 2 and 3, represent Cartesian axes and repeated index


implies summation. Taking a dot product of Eq. (4.17) with ωi and using Ω1 = ωi ωi
to represent enstrophy, one obtains its transport equation as
  
∂Ω1 ∂Ω1 ∂ui 1 ∂ 2 Ω1 2 ∂ωi ∂ωi
+ uj − 2ωi ωj = − (4.18)
∂t ∂xj ∂xj Re ∂xj ∂xj Re ∂xj ∂xj
244 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows

The third term on the left hand side (LHS) of the above equation is due to vortex
stretching (corresponding to the first term on the right hand side (RHS) of Eq. (4.17)),
which is absent for 2D flows. The diffusion of ωi gives rise to RHS terms in Eq. (4.18).
The first term of RHS shows diffusion of Ω1 and the second term represents strictly
a loss or the dissipation term for the transport of Ω1 . The present study views Ω1 as a
point property in the flow and is different from the usage in [13], where the enstrophy
is defined over the full domain. The traditional approach utilizes the simplification
brought about for problems which are homogeneous and periodic. Focusing on 2D
flows with the spanwise component of vorticity (ω), ETE can be written in vector
form as,  
DΩ1 2 1 2
= ∇ Ω1 − (∇ω)2 (4.19)
Dt Re 2

The first term on RHS of Eq. (4.19) is missing from Eq. (4.14), due to periodicity
of the flow and strictly negative RHS shows the viscous action to dissipate energy
globally. In contrast for inhomogeneous flows, the first term on RHS of Eq. (4.19)
can be either positive or negative. Therefore, the diffusion term in ETE can create
or destroy rotationality, depending upon the sign of RHS in Eq. (4.19). Thus, the
diffusion term should not be identified strictly as dissipation for general flows. As
Ω1 > 0, positive RHS indicate the diffusion to cause instability. Negative RHS act as
a sink of Ω1 . This provides a mechanism of creating rotationality at different scales
by diffusion and is distinctly different form the concept of creating smaller scales
by vortex stretching, as the dominant mechanism of generating small eddies for 3D
flows. We note that the role of diffusion in creating new length scales is ubiquitous
for both 2D and 3D flows.

4.4.2 Enstrophy Cascade for General Inhomogeneous Flows

To further investigate effects of diffusion in Eq. (4.19) at multiple scales, one derives
transport equations for higher powers of Ω1 . Multiplying Eq. (4.19) with Ω1 and
n−1
defining Ωn = Ω12 , one obtains for Ω2 the following transport equation,
 
DΩ2 −1 1 2
= 2Re ∇ Ω2 − (∇Ω1 ) − Ω1 (∇ω)
2 2
(4.20)
Dt 2

Noting further that DΩDt


2
= 2Ω1 DΩ
Dt
1
, one can write Eq. (4.20) as the ETE, i.e., an
evolution equation for Ω1 . Multiplying the above equation with Ω2 and simplifying,
one can obtain transport equation for Ω3 , which can be used to write the ETE involv-
ing Ω1 , Ω2 and Ω3 . This process can be generalized to obtain the transport equation
for Ωn as,  
DΩn −1 1 2
= 2Re ∇ Ωn − (∇Ωn−1 ) − C
2
(4.21)
Dt 2
4.4 Enstrophy Transport Equation: A New Approach … 245

where,
⎛ ⎞

n−2 n−1
C= 2n−k−1 ⎝ Ωj ⎠ (∇Ωk )2 (4.22)
k=0 j=k+1

 
and Ω0 = ω with indicating summation over all k’s and indicating the product
of all the jth elements.
Also, the substantive derivative of Ωn can be written and simplified as
 
DΩn DΩn−1
= 2Ωn−1
Dt Dt
DΩn−2
= 2Ωn−1 (2Ωn−2 ) (4.23)
⎛ ⎞ Dt
n−1
⎝ DΩ1
=2 n−1
Ωj ⎠
j=1
Dt

which can be further simplified to yield,


DΩn −β DΩ1
= 2n−1 Ω1 (4.24)
Dt Dt

where β = 1 − 2n−1 . Using above relations, one can rewrite Eq. (4.21) as the ETE
given by
β  
DΩ1 Re−1 Ω1 1 2
= ∇ Ωn − (∇Ωn−1 )2 − C (4.25)
Dt 2n−2 2

One notes that while writing the transport equation for Ωn , the diffusion term
from the transport equation for Ωn−1 contributes two terms; one of which is strictly
dissipative (dependent on Ωn−2 ) and the other as a diffusion term for Ωn−1 . The
diffusion term involving Ωn−1 can be furthermore expressed into two terms involving
a strictly dissipative term involving Ωn−1 and another diffusive term involving Ωn−2 .
This process can cascade indefinitely in Eq. (4.25), for increasing n with the leading
term as a diffusion term and the rest are strictly dissipative. Higher order moments
of Ω1 will contribute more for higher wavenumbers, implying that the order of even
moments of Ω1 will be restricted by the energy supplied to the flow. For 3D flow
as well, the RHS of Eq. (4.25) is present as the forcing term. However in this case,
the vortex stretching term is retained. The ETE for 3D flow is same, as given by
Eq. (4.18). Following a similar approach as in deriving the transport equation for Ωn
for 2D flows, the transport equation for Ωn can be derived for 3D flows as,
246 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows

  n−1  
DΩn 1 2 ∂ui
= 2Re−1 ∇ Ωn − (∇Ωn−1 )2 − C + 2n Ωk ωi ωj (4.26)
Dt 2 ∂xj
k=1

where expression for C is same as in Eq. (4.22). Using Eq. (4.24), one can rewrite
the ETE for 3D flows as
β  
DΩ1 Re−1 Ω1 ∇ 2 Ωn ∂ui
= − (∇Ωn−1 ) − C + 2ωi ωj
2
(4.27)
Dt 2n−2 2 ∂xj

One also notes that the diffusion term gives rise to enstrophy cascade for both 2D
and 3D flows, for which the contribution at higher wavenumbers depends upon the
value of n decided by the energy supplied to the fluid dynamical system. However
in 3D flows, vortex stretching is also present, which provides an additional mecha-
nism of energy redistribution process. This indicates that in 3D flows, generation of
different scales of vorticity is due to enstrophy cascade via the stretching and diffu-
sion terms and the energy cascade is by the vortex stretching implicit in convection
process. In 2D flows, it is only the diffusion term, which gives rise to enstrophy (and
hence vorticity) at different scales.
It is emphasized that physically the role of diffusion for inhomogeneous flows is
not strictly dissipative, as is the case for homogeneous turbulent flows. By developing
the ETE, in terms of higher even moments of Ω1 , we identified the index n in this
equation, which is fixed from total energy imparted to drive a flow. This view of how
smallest scale is fixed is entirely different from the conventional logic used for the
dissipation of kinetic energy to heat, even for isothermal flows.
To understand the role of the developed ETE for inhomogeneous flows, in Fig. 4.9,
we note the growth rate of total enstrophy, (DΩ1 /Dt)-contours, for the problem of
vortex-induced instability, whose results are shown in Figs. 4.5 and 4.6 for the stream
function and vorticity contours. Corresponding Ed evolution is shown by contour-
plots of the RHS of Eq. (4.6) in Fig. 4.7. The times at which the total enstrophy
growth are shown in Fig. 4.9, are identical to the times shown earlier in Figs. 4.5, 4.6
and 4.7. The spatial scales shown in enstrophy growth rates are much more refined,
than those seen for Ed . Another aspect that draws our attention is the clear presence
of wall- and free stream-modes at early times, as explained in [48, 55]. At t = 20,
one notices growth of rotationality originating from leading edge of the plate (due to
Goldstein singularity), that remains at the edge of boundary layer very prominently.
While the wall-mode is in nascent stage, in the region: 1 ≤ x ≤ 2 of the flat plate.
Both these modes are noted more prominently in the frame at t = 25, and the growing
wall-mode causes a bulge in the growth region of the free stream-mode. In a short
time interval, one notices severe interaction between these two modes, as noted in
the frame for t = 28, which is noted more in the vorticity contours of Fig. 4.6, than
4.4 Enstrophy Transport Equation: A New Approach … 247

Fig. 4.9 The growth rate of total enstrophy Eq. (4.19) for the vortex-induced problem defined in
Fig. 4.3 for which the physical variables are shown in Figs. 4.5 and 4.6 for the identical time instants.
The marked regions are for the instability of enstrophy, i.e., where rotationality increases with time
at the indicated time instants
248 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows

in the disturbance mechanical energy of Fig. 4.7. This is apparent that the signature
of vortex-induced instability is more evident in the rotational part of the NSE, i.e.,
in ETE, than in the divergence or irrotational part of NSE for Ed . Also visible is the
presence of much finer wall-normal structures in Fig. 4.9, during early interactions
between wall- and free stream-mode. Such events are equally visible in the frame
at t = 30. In the bottom frames of Figs. 4.7 and 4.9, one can notice more energetic
events in Ed , as compared to those in (DΩ1 /Dt)-contours. The latter still grows in
the location over the plate, which is exactly beneath the convecting vortex.

4.5 Theory of Instability for Enstrophy: Creation


of Rotationality

Having described the nonlinear growth of E in Sect. 4.1, one can study nonlinear
growth/ decay of rotationality, quantified by Ω1 . We have the equation for the growth
rate of this in Eq. (4.18). If Ω1 is written as sum of the equilibrium and disturbance
values, Ω1 = Ωm + ε1 Ωd , then the linearized growth rate of Ωd can be evaluated
retaining the terms of order ε1 . If the primary variables are represented as, − →ω =

→ −
→ −
→ −
→ −

ω m + ε2 ω d , and V = V m + ε2 V d , then the growth or decay rate of Ω1 can be
written as,
DΩ1 ∂ −
→ −

= (Ωm + ε1 Ωd ) + ( V m + ε2 V d ) · ∇(Ωm + ε1 Ωd ) = RHS
Dt ∂t
From the above one can rewrite it for

DΩm ∂Ωd −
→ −
→ −

+ ε1 + ε1 V m · ∇Ωd + ε2 V d · ∇Ωm + ε1 ε2 V d · ∇Ωd = RHS
Dt ∂t
The order ε1 terms of the left hand side of the above is simply nothing but the
substantive derivative of (Ωm + ε1 Ωd ) and thus, one can write the evolution equation
for disturbance enstrophy as
 
DΩd ∂uid ∂uim ∂uim
= 2ωim ωjm + 2ωim ωjd + 2ωid ωjm
Dt ∂xj ∂xj ∂xj
  
1 ∂ Ωd
2
2 ∂ωim ∂ωid
+ − (4.28)
Re ∂xj ∂xj Re ∂xj ∂xj

Thus the growth or decay rate of Ωd is determined by the vortex stretching terms
given on the RHS of the above equation inside the curly brackets. These will not
4.5 Theory of Instability for Enstrophy: Creation of Rotationality 249

be present for 2D disturbance field. The contribution from the diffusion term to the
growth rate is obvious. However, the term that is identified as strictly dissipation
term in Eq. (4.18) for Ω1 , can contribute to growth of Ωd , as shown here by the
last term on RHS. We also note that unlike Ω1 (which is strictly positive definite),
Ωd can be either positive or negative, as this is given by − →ωm·− →ω d . It is noted that
this Ωd term also appears in the Poisson equation for Ed , Eq. (4.6). Also as Ωd can
be either positive or negative, its conditions for growth/ decay will be different for
different signs of the quantity, i.e., for positive Ωd , an instability would be indicated
when DΩd /Dt > 0 and for negative Ωd , its amplitude will grow when DΩd /Dt < 0.
These conditions are investigated next during vortex-induced instability.
In Fig. 4.10, this DΩd /Dt is plotted in the domain, at the same indicated times as
before. As noted above, we indicate two different conditions for linear growth rate
cases, depending upon the signs of Ωd . The enstrophy contours are drawn by solid
(for positive values) and dashed lines (for negative enstrophy) in the frames. Only
the growth rate regions are marked by flooded region for both the cases, in the frames
of Fig. 4.10. The confluence of the regions are indicative of modulus of Ωd growing.
Like the DΩ1 /Dt rate, here also the early-time growth is due to growth of wall-
and free stream-modes, with the latter dominating. The wall-mode growth for Ωd
is higher as compared to Ω1 . For the same reason, the wall- and free stream-modes
interact earlier and strongly, for the growth of Ωd . The region of negative Ωd growth
is associated with the wall-mode, as is easily evident in the figure. In Fig. 4.10, one
also notices distinct wall-normal streaks, ahead of the convecting free stream vortex
from t = 28 onwards, which was not the case for Ω1 . These vertical streaks are
seen to cover the flat plate ahead of the convecting vortex, near the outflow of the
domain. One also notices that for t ≥ 50, the regions near the leading edge of the
plate becomes quieter.
In Fig. 4.11, nonlinear Ωd growth rate is plotted in the domain, at the same indi-
cated times as before. This is obtained first by noting DΩ Dt
1
for the total enstrophy at
each and every point in the domain as a function of time, and then subtracting DΩ Dt
m

from it, to obtain the nonlinear growth rate. As before, two different conditions for
nonlinear growth rate cases are considered, depending upon the signs of the distur-
bance enstrophy (− →ωm·−→ω d ). The enstrophy contours are drawn by solid (for positive
values) and dashed lines (for negative Ωd ) in the frames. Only the growth, rate zones
are marked by flooded region for both the cases in the respective frames. The overlap
regions are indicative of modulus of Ωd growing with time. Comparing Figs. 4.10
and 4.11, one notices only very marginal differences between linear and nonlinear
growth regions. In nonlinear growth case, the rates are relatively lower and spread
over smaller overlap regions.
250 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows

Fig. 4.10 The growth rate of Ωd Eq. (4.19) for the vortex-induced instability problem defined in
Fig. 4.3, for which the physical variables are shown in Figs. 4.5 and 4.6 for the identical time instants.
On the left frames shown is the case with positive Ωd and on the right is the case of negative Ωd
shown at the indicated time instants
4.5 Theory of Instability for Enstrophy: Creation of Rotationality 251

Fig. 4.11 The nonlinear growth rate of Ωd for the vortex-induced instability defined in Fig. 4.3,
for which the physical variables are shown in Figs. 4.5 and 4.6 for the identical time instants. On
the left frames shown are the regions where Ωd is positive and on the right where Ωd is negative,
with overlap regions where rotationality increases with time at the indicated time instants
252 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows

4.6 Proper Orthogonal Decomposition

For high Reynolds number transitional and turbulent flows, POD analysis [25, 29]
can be viewed as providing nonlinear instability portrait, provided we can relate the
POD modes with instability modes. In simple terms, POD is nothing but the Galerkin
projection of space-time dependent variables describing system dynamics. Such a
projection helps describe the space and time dependence of the disturbance field.
Principally in the literature, there are two variants of POD. The classical one is based
on projection of a random field onto a deterministic basis (eigenfunctions) so that
the reduced order model captures maximum energy. While conceptually this is a
very appealing approach, but its practice in fluid mechanics is difficult in handling
pressure-velocity coupling terms when one is solving NSE, using primitive variable
formulation. This has been circumvented by the authors, in developing an enstrophy
based POD. This is based on the vorticity transport equation, which does not have the
pressure term. The enstrophy-based POD is preferred over those in [11, 23, 30, 37,
41, 71], where kinetic energy is used for POD analysis. In vortex dominated flows,
which are neither homogeneous nor periodic, enstrophy is a better descriptor of POD
over translational kinetic energy, as highlighted in [48, 67]. This approach has been
used in many earlier works [48, 56, 62–64, 68]. Authors in [64], used enstrophy
based POD approach to study both external and internal flows to show universality
of POD modes in terms of amplitude functions. In [34, 37, 70], the authors devised
a new POD mode which was obtained through a Galerkin projection on Reynolds-
averaged Navier–Strokes equation. In [8, 44], the authors performed reduced order
modeling using Koopman modes. In Sengupta et al. [63], the authors used enstrophy
based POD modes to analyze transitional flow past a flat plate. Thus, POD analysis
has been shown to be an useful tool in studying both internal and external flows of
different kinds.
POD technique was originally introduced by Kosambi [25] for a random field

→ −

vi ( X , t), where it was projected onto a set of deterministic vectors φi ( X ), so that
|(vi , φi )| is maximum, where the outer angular brackets signify time-averaging
2


and inner brackets signify an inner product. The computation of φi ( X ) can be posed
as an optimization problem in variational calculus,
 

→ − → −
→ −
→ −

Rij ( X , X ) φj ( X ) d 2 X = λ φi ( X ) (4.29)


→ − →
The kernel of this is the two-point correlation function, Rij = vi ( X )vj ( X ) of
the random field. It is noted [63] that classical Hilbert–Schmidt theory applies to flows
with finite energy, and, therefore, denumerable infinite orthogonal POD modes can
be computed. Hilbert–Schmidt theory is applicable here because instabilities in flow,
derive energy from the equilibrium flow, which itself has finite energy. Disturbance
vorticity field is thus, represented in POD formalism as

→ ∞ −

ω ( X , t) = Σm=1 am (t) φm ( X ) (4.30)
4.6 Proper Orthogonal Decomposition 253

where am (t) represents the amplitude function, which describes the spatio-temporal


variation of the modal amplitude and φm ( X ) is the corresponding spatial eigenfunc-
tion. Equation (4.29) is an eigenvalue problem in the integral form, which becomes
intractable even for moderate grid resolution. To overcome this difficulty, Sirovich
introduced the method of snapshots [71], which has an advantage of dealing with


smaller data sets in multiple dimensions. The eigenfunction φm ( X ) is expressed as a
linear combination of the instantaneous flow fields at distinct instants of time ti ’s as,

→ −

φm ( X ) = Σi=1
N
qmi ω ( X , ti ) (4.31)

where N is the number of snapshots used. This together with the expression for Rij
reduces Eq. (4.29) to an algebraic eigenvalue problem, [C̄]{q} = λ{q}. The entries
of the matrix [C̄] are obtained by integrating over the domain by using,
 
1 −
→ −
→ −

C̄ij = ω ( X , ti ) ω ( X , tj ) d 2 X (4.32)
N

with i, j = 1, 2 . . . N defined over the snapshots of the flow. The amplitude functions
am (t) are obtained from

 −
→ −
→ −

ω ( X , t) φm ( X ) d 2 X
am (t) =  −
→ −
→ (4.33)
φm2 ( X ) d 2 X

The eigenvalues λ and eigenvectors {q} of [C̄] are computed using a MATLAB
function based on QR decomposition. Once {q} is obtained, Eq. (4.31) is used to


find the eigenfunctions φm ( X ) and the corresponding amplitude functions, am (t)
are evaluated using Eq. (4.33). The kernel R(X, Y) is defined as
 t0 +T
1
R(X, Y) = < ω (X, t), ω (Y, t) > = ω (X, t)ω (Y, t)dt (4.34)
T t0

Such a decomposition of ω (X, t) into the various orthogonal modes that satisfies
Eq. (4.29) is called its POD. Equation (4.29) represents an eigenvalue problem, where
the eigenvalues λm represent the fraction of average enstrophy captured by the mth-
mode.

4.6.1 Some Useful Mathematical Relations

In this subsection, some useful mathematical relations regarding POD are provided.
These are as enumerated below:
254 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows

1. Multiplying Eq. (4.30) by φm (X), and integrating it over the full computational
domain S, by using the orthogonality condition, one obtains Eq. (4.33).
2. Using Eqs. (4.29), (4.30), and orthogonal properties of φm ’s and (4.33), it can be
shown that
 t0 +T
am (t)an (t)dt = λm δmn (4.35)
t0

This implies that that the amplitude functions am (t)’s are also orthogonal to each
other.
3. Multiplying Eq. (4.30) by the amplitude function an (t), and integrating over the
time interval (t0 , t0 + T ) and using Eq. (4.35), it can be shown that
 t0 +T
φm (X) = αm (t)ω (X, t)dt (4.36)
t0

where αm (t) = amλm(t) .


4. Inserting the decomposition of ω (X, t) and ω (Y, t), as given by Eq. (4.30) in the
definition of R(X, Y), given in Eq. (4.34), and using Eq. (4.35), it can be shown


R(X, Y) = λm φm (X)φm (Y) (4.37)
m=1

4. One can define the instantaneous enstrophy Ω1 (t) and the average enstrophy Ω̂1
over the time period T as

Ω1 (t) = ω (X, t)ω (X, t)d 2 X (4.38)
S

 t0 +T
1
Ω̂1 = Ω1 (t)dt (4.39)
T t0

From Eqs. (4.37), (4.38) and (4.39), one can show that,

Ω̂1 = R(X, X)d 2 X
S

 
= λm φm (X)φm (X)d 2 X
m=1 S


= λm (4.40)
m=1
4.6 Proper Orthogonal Decomposition 255

This justifies the statement made above that, the eigenvalues λm represent the
fraction of average enstrophy Ω̂1 captured by the mth POD mode.

4.6.2 Method of Snapshots

For a 2D flow field, to calculate POD eigenvalues λm , and eigenfunctions φm (X),


one has to solve the eigenvalue problem given by Eq. (4.29). If the DNS is carried
out using M and N points in ξ - and η-directions and POD is performed over Q
time-snapshots at equal time interval of Δt, then R(X, Y) given by Eq. (4.34) would
represent a L × L square symmetric matrix, where L = M × N . Each element of this
symmetric square matrix R is given as


Q
q q
Rmn = ωij ωkl Δt
q=1

where, m = (j − 1)M + i and n = (l − 1)M + k. Here, (i, k) and (j, l) denotes the
indices along ξ - and η-directions, respectively.
Since, M and N are generally orders of magnitude higher than Q, it is compu-
tationally very expensive to evaluate the eigenvalues of the matrix R. To avoid this
computational problem of evaluating the POD modes, the method of snapshots is
used, and described next in brief.
In the method of snapshots, one proposes formation of a correlation function
D(t, s) as

D(t, s) = ω (X, t)ω (X, s)d 2 X (4.41)
S

It is to be noted that for discrete computations D(t, s) represents a square sym-


metric matrix of size Q × Q with elements


M 
N
p q
Dpq = ωij ωij h1ij h2ij Δξ Δη
i=1 j=1

Hence, computing the eigenvalues of D(t, s) will be much more efficient and less
time-consuming than the direct method of solving the eigenvalues from Eq. (4.37).
This is in essence, the method of snapshots. Once, the eigenvalues λm , and the
amplitudes am (t) are obtained, one can calculate the POD spatial eigenfunctions
φm (X) from Eq. (4.34).
256 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows

4.6.3 TS Wave Instability over Zero-Pressure Gradient


Boundary Layer

Here, POD analysis is done for a ZPGBL for a semi-infinite flat plate excited by a SBS
exciter strip. For the present study, the domain along the streamwise direction is taken
as −0.1 ≤ x ≤ 22; while in the wall-normal direction, it is taken as 0 ≤ y ≤ 1. In the
presented simulations, Reynolds number based on reference length scale L is selected
again to be Re = 105 . The amplitude and non-dimensional frequency parameter of
the excitation are chosen as α1 = 0.01 and Ff = 1 × 10−4 , respectively.
For nonlinear receptivity and stability calculations, full NSE is solved for the total
quantities (without splitting into equilibrium and disturbance components) similar
to the cases discussed earlier. Disturbance quantities are obtained by subtracting the
equilibrium solution from the total solution.
In Fig. 4.12, time evolution of disturbance field is shown by plotting the vorticity
contours at six different time instants (t = 10, 15, 20, 30, 50 and 60). Development
of spatio-temporal disturbance field is clearly noticed in these frames. From t = 10
onwards, a wave-front can be seen to develop, which is prominently noted at t = 15.
This wave-front amplifies at a faster rate than the main trailing wave-packet. This

0.1 0.1
Developing wave-front
(a) t = 10 (b) t = 15
0.08 0.08
0.06 0.06
y

0.04 0.04
0.02 0.02
00 5 10 15 20 00 5 10 15 20
x x
0.1 0.1
(c) t = 20 (d) t = 30
y

0.05 0.05

00 5 10 15 20 00 5 10 15 20
x x
0.1 0.1
(e) t = 50 (f) t = 60
0.08
y

0.06
0.05
y

0.04
0.02
00 5 10 15 20 00 5 10 15 20
x x

Fig. 4.12 Vorticity contours at indicated times are shown plotted in frames (a) to (b) for ZPG flow
past a flat plate excited by SBS strip-exciter. The location of the midpoint of the exciter strip is
pointed by an arrow
4.6 Proper Orthogonal Decomposition 257

is further evident from the contours at t = 20, when one also notices a vortical
eruption associated with the wave-front. The mismatch of speed between the leading
wave-front and the following main wave-packet causes these to separate clearly at
later times. At t = 30, one notes two distinct coherent structures: the main wave-
packet and a separated, highly amplified fast moving STWF that after sufficient
magnification forms a vertically erupting separation bubble. This erupting bubble
perturbs the flow locally, creating secondary new bubbles in its vicinity. The wave-
packet after t = 20, continues to grow and move forward, but at a relatively lower,
exponential rate. The STWF leaves the computational domain at t = 45 cleanly
due to the application of Sommerfeld radiative boundary condition, as discussed
earlier. Growth and expansion of the region perturbed by the wave-packet almost
remains invariant after t = 50, which is evident from the contour plots at t = 50 and
60. Here, for the purpose of POD analysis of the solution incorporating nonlinear
and nonparallel effects, the full nonlinear equation is solved without neglecting the
growth of the underlying equilibrium flow.
POD of the disturbance vorticity field is performed using the time series taken
from t = 0 to t = 60, by storing flow-snapshots at a time interval of Δt = 0.01,
in a domain whose streamwise and wall-normal extent is −0.1 ≤ x ≤ 20 and
0 ≤ y ≤ 0.5, respectively. Input wall excitation imposes a non-dimensional time
scale of T = 0.63 on this fluid dynamical system, as the chosen non-dimensional
circular frequency of excitation is βL = Ff × Re = 10. Hence, the chosen time inter-
val of taking snapshots is more than adequate, as there are 63 snapshots within one
fundamental time interval of the perturbed flow-field.
In Fig. 4.13, the cumulative disturbance enstrophy content of the flow obtained
by POD analysis is shown, which helps one to identify the dominant modes play-
ing a vital role in determining response of the dynamical system by forcing. One
hundred and twenty modes shown in the figure collectively contribute 98.96% of
the total enstrophy. The first 30 modes contribute about 90% of the total enstrophy.

1
Cumulative Enstrophy Content

0.8

0.6

0.4

0.2

0
20 40 60 80 100 120
No of Modes

Fig. 4.13 Cumulative enstrophy content are shown plotted against corresponding mode number
for zero-pressure gradient flow over semi-infinite flat-plate with wall-excitation
258 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows

Table 4.1 Eigenvalues and POD mode index j Eigenvalues (λj ) Cumulative
cumulative enstrophy of the enstrophy content
POD modes k=j
(ej = k=1 λk )
1 0.2829E+00 0.2829
2 0.2808E+00 0.5638
3 0.1731E+00 0.7369
4 – 0.7369
5 0.2546E-01 0.7623
6 0.2532E-01 0.7878
7 0.1663E-01 0.8044
8 – 0.8044
9 0.1228E-01 0.8167
10 0.1227E-01 0.8290
11 0.7578E-02 0.8366
12 0.7534E-02 0.8441
13 0.6547E-02 0.8506
14 – 0.8506
15 0.4622E-02 0.8553
16 0.3907E-02 0.8592
17 0.3517E-02 0.8627
18 0.3189E-02 0.8659

Higher modes individually contribute very less but their collective contributions are
significant.
In Table 4.1, first eighteen POD modes with corresponding eigenvalues, and
cumulative enstrophy contribution up to that mode, have been listed. The numbering
of the modes here, follow the conventions used in [62, 64]. The regular POD modes
form pairs with almost equal eigenvalues, and closely resembling eigenfunctions.
Constituents of the pairs exhibit an approximate phase shift of 90◦ between corre-
sponding eigenfunctions, as well as, amplitude functions. Such lead pair of modes
have been postulated to be governed by Stuart-Landau equation [11, 30, 37, 74]. But
authors in [62] have shown from their DNS results for flow past a circular cylinder,
that in addition to these regular modes, solitary modes appear which do not form
pair. Noack et al. [37] have shown that the inclusion of these solitary mode effects are
necessary to explain the dynamics of flow past a circular cylinder. This shift mode
in their work was constructed as a ‘mean-field correction’ using a Gram–Schmidt
procedure. Why this is necessary, can be understood by looking at the disturbance
velocity components given by
 
ud (x, y, t) = [û ei(αx−ω0 t) + û∗ e−i(αx−ω0 t) ] d α d ω0 (4.42)
4.6 Proper Orthogonal Decomposition 259
 
vd (x, y, t) = [v̂ ei(αx−ω0 t) + v̂∗ e−i(αx−ω0 t) ] d α d ω0 , (4.43)

where quantities with asterisks denote complex conjugates. If one represents the
variables as a sum of mean field plus a disturbance and substitute these in NSE,
then one obtains the averaged governing equation (indicated by angular brackets) for
the mean field quantities to have gradient transport contribution terms of Reynolds
stress-like term given by,
  
< ud (x, y, t)vd (x, y, t) >= [û v̂∗ + û∗ v̂] + û v̂e2i(αx−ω0 t)

+ û∗ v̂∗ e−2i(αx−ω0 t) d α d ω0 (4.44)

One notes that in Eq. (4.44), the first term on the right hand side is phase inde-
pendent and should be included in the mean-field equations, irrespective of whether
one is talking about time or ensemble averaged equations. In [50] this was called
the wave-induced stress term, while studying flow past a wavy wall. This steady
streaming is always present in NSE for any flow field, in the presence of oscillatory
disturbances. This also explains, why one can have many such terms, coming from
different modes, with α and ω0 related via the dispersion relation of the problem
from governing equation and \ or boundary conditions. These type of terms will not
be present, when the modes appear in pairs. There is absolutely no restrictions on
the number of isolated modes and in the present case, third, seventh and thirteenth
are the isolated modes, as shown in Fig. 4.14. In Table 4.1, blanks have been left to
indicate the missing mode, next to the isolated modes. Note that these stress terms
are at most of order two in NSE, if the mean field equations are of zeroth order and
thus represent a correction to the mean field and is a nonlinear term.
In Figs. 4.14 and 4.15, time dependent amplitude functions aj (t) of the listed
eighteen modes in Table 4.1 and the Fast Fourier transform (FFT) are plotted against
nondimensional circular frequency (βL Δt). The pair-forming modes are plotted in
the same frame. Among the plotted modes in Fig. 4.15, the third, the seventh and the
thirteenth modes are anomalous modes of the first kind (T1 ), as defined in [62, 64].
These are similar to the shift modes of [37]. Time variation of a1 (t) and a2 (t) follows
the Stuart–Landau–Eckhaus model for instability modes. These type of modes which
follow Stuart–Landau–Eckhaus model are the regular POD modes (R) of [62].
There are certain salient features in the time evolution of first and second POD
modes. It is seen that for these modes, amplitude function grows exponentially up to
approximately t = 15. This exponential amplification factor is calculated to be about
0.19. From t = 15 to t = 19, this growth is saturated. But, thereafter the amplitude
again grows at a lower rate up to t = 43, and thereafter reaches a saturated value.
FFT of these two modes show the presence of only one dominant frequency, which
is exactly the frequency of excitation. The variation from t = 15 to t = 19 can be
attributed to the growth and detachment of the STWF from the wave-packet, as
can be seen from the vorticity contour plots at these times. The wave-front derives
260 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows

(i) 1 st mode ( ii ) 3
rd
mode
nd
2 mode
4
5
a 1,2

a3
0 3

-5
2

0 10 20 30 40 50 60 0 10 20 30 40 50 60
time time
th
5 mode ( iv ) 7
th
mode
4
( iii ) th
6 mode
1
2
a 5,6

a7
0
0
-2

-4
-1
0 10 20 30 40 50 60 0 10 20 30 40 50 60
time time
th th
9 mode 11 mode
2 (v) th ( vi ) th
10 mode 2 12 mode

1
a 11,12
a 9,10

0 0

-1

-2
-2
0 10 20 30 40 50 60 0 10 20 30 40 50 60
time time

2 15 th mode
( vii ) 13
th
mode ( viii ) th
2 16 mode
1
1
a 15,16
a 13

0
0

-1 -1

-2 -2
0 10 20 30 40 50 60 0 10 20 30 40 50 60
time time

2 ( ix ) 17 th mode
th
18 mode
1
a 17,18

-1

-2
0 10 20 30 40 50 60
time

Fig. 4.14 Time dependent amplitude function aj (t) of first eighteen POD modes for zero-pressure
gradient flow over semi-infinite flat-plate plotted as a function of time. The pair-forming modes are
shown together in a frame as indicated. The numbering of the modes follow the conventions used
in [62, 64]
4.6 Proper Orthogonal Decomposition 261

A st
1 mode rd
10
4
(i) nd 10
4
( ii ) 3 mode
2 mode
103 103
â 1,2

â3
-2
2 A = 7.324×10 2
10 10 C=3×A

C
1 1
10 10

0 0
10 10
0 0.2 0.4 0.6 0 0.2 0.4 0.6
βΔt βΔt
4 4
10 A th 10
5 mode th
( iii ) ( iv ) 7 mode
10
3 6 th mode
3
10
2
10
â 5,6

â7
2
B B=2 ×A 10 A
1 B
10
1
0 10
10

10-1 100
0 0.2 0.4 0.6 0 0.2 0.4 0.6
βΔt βΔt
4 4
10 B th
9 mode 10 th
11 mode
(v) A ( vi )
10
3 10 th mode 10
3 12 th mode

2 2
â 11,12

10 10
â 9,10

A B
1 1
10 10
0 0
10 10

10-1 10-1
0 0.2 0.4 0.6 0 0.2 0.4 0.6
βΔt βΔt
4
10
( vii ) 13 th mode 15 th mode
10
3
10
3 ( viii ) th
16 mode
2
A 10 A
102
â 15,16

E =5 ×A
â 13

B 101 A1 B
A1 = 0.041
10
1 C
10
0 C E
0 -1
10 10
-2
-1 10
10
0 0.2 0.4 0.6 0 0.2 0.4 0.6
βΔt βΔt
4
10 th
A1 17 mode
10
3 ( ix ) th
18 mode
2
10
A2 A A1 = 8.129 × 10
-3
â 17,18

-2
101 B A2 = 4.48 × 10
0
10 C E
-1
10
-2
10
0 0.2 0.4 0.6
βΔ t

Fig. 4.15 Fast Fourier transform of amplitude functions aj (t) of first eighteen POD modes are
shown plotted as a function of non-dimensional frequency βΔt
262 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows

0.1 ( i ) 1st mode ; min = -16.6 max = 16.7 0.1 ( ii ) 2nd mode ; min = -16.7 max = 16.7

0.05 0.05

0 0
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20

rd
0.1 ( iii ) 3 mode ; min = -1202 max = 439

0.05

0
0 2 4 6 8 10 12 14 16 18 20

th
0.1 ( iv ) 5 mode ; min = -16.21 max = 16.27 0.1 th
( v ) 6 mode ; min = -16.2 max = 16.2

0.05 0.05

0 0
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20

th
0.1 ( vi ) 7 mode ; min = -949.23 max = 347

0.05

0
0 2 4 6 8 10 12 14 16 18 20

0.1 th
( vii ) 9 mode ; min = -23 max = 23 0.1 ( viii ) 10th mode ; min = -22.9 max = 22.9

0.05 0.05

0 0
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20

Fig. 4.16 Contours of eigenfunctions φj (x, y) corresponding to the first ten POD modes for zero-
pressure gradient flow semi-infinite flat-plate with wall-excitation are shown plotted

substantial energy from the wave-packet for its growth, and subsequent detachment
in the mentioned time period causing a saturation of the amplification of these modes.
In Figs. 4.16 and 4.17, eigenfunctions φj (x, y) of the first eighteen POD modes
listed in Table 4.1 are plotted. The eigenfunction of first two modes in Fig. 4.16, rep-
resent previously described wave-packet, which are closely related to the generated
TS waves. The average wavenumber of the streamwise variation of these eigenfunc-
tions, in terms of the displacement thickness at the exciter is 0.16, whereas the same
is predicted from the linear parallel theory as 0.17. At the wall, these eigenfunctions
show amplification from Reδ∗ = 835 to Reδ∗ = 1598. This range of amplification
though decreases, as one moves away from the wall, which can be attributed to
nonparallel, nonlinear effects. Linear theory predicts amplification from Reδ∗ = 740
to Reδ∗ = 1235, for the chosen non-dimensional circular frequency of excitation
βL = Ff × ReL = 10.
4.6 Proper Orthogonal Decomposition 263

th
0.1 ( ix ) 11th mode ; min = -21.34 max = 21.66 0.1 ( x ) 12 mode ; min = -21.34 max = 21.66

0.05 0.05

0 0
0 5 10 15 20 0 5 10 15 20

th
0.1 ( xi ) 13 mode ; min = -287 max = 105

0.05

0
0 5 10 15 20

th th
0.1 ( xii ) 15 mode ; min = -80 max = 29 0.1 ( xiii ) 16 mode ; min = -121 max = 44

0.05 0.05

0 0
0 5 10 15 20 0 5 10 15 20

th th
0.1 ( xiv ) 17 mode ; min = -13 max = 35 0.1 ( xv ) 18 mode ; min = -105.2 max = 38

0.05 0.05

0 0
0 5 10 15 20 0 5 10 15 20

Fig. 4.17 Contours of eigenfunctions φj (x, y) corresponding to the first eighteen POD modes for
zero-pressure gradient flow semi-infinite flat-plate with wall-excitation are shown plotted

The third mode is an isolated one, responsible for 17% of total enstrophy of distur-
bance vorticity field. Figure 4.14 shows that the amplitude of this mode contains low
amplitude oscillations, overriding a slower time scale variation, which corresponds
to the frequency of excitation, as evident from the corresponding FFT in Fig. 4.15.
Up to t = 10, mean of this amplitude function has a constant value, after which it
grows almost linearly up to t = 43, and becomes almost constant thereafter. Ampli-
tude function of the other two isolated modes shown in Fig. 4.14 (the seventh and
the thirteenth modes) also show similar kind of time variation, namely an oscillating
part, over and above a slow time scale variation, that is constant up to t = 10 and
varies almost linearly up to t = 43 for the seventh mode, and up to t = 35, for the
thirteenth mode. The oscillatory part can be attribute to the second and third terms in
the right hand side of Eq. (4.44). After t = 43, mean of the amplitude function for the
seventh mode attains a constant value, similar to the third mode. All isolated modes
are characterized by the maximum contribution coming from time-independent part.
The eigenfunctions for these isolated modes, as shown in Figs. 4.16 and 4.17,
illustrate that these modes basically represent large scale vortices of opposite sign,
stacked one above the other. Unlike other modes, eigenfunctions of these modes
include the effect of the leading edge, as their maximum and minimum values occur
exactly at the leading edge. Eigenfunction of the third mode exhibits a three layered
structure, a zone of negative vorticity very near to the wall, covered by a positive
264 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows

vorticity zone, above which there is a negative vorticity layer on top. Eigenfunction of
the seventh mode contributes only 1.66% of total enstrophy, and also displays similar
structures with vortex centers shifted downstream with opposite signs of vorticity
stacked, one over the other.
It has been shown that these isolated modes depict nonlinear interactions by the
disturbance field on the mean flow arising due to averaged Reynolds stress term,
< u v >. FFT of the third mode in Fig. 4.15, exhibits a single distinct peak (marked
by A) at the exciter frequency βL . Also a smaller peak at the second super-harmonic,
3βL , is noted. However, dominant frequencies near the origin representing slow
time-scale variation, play a significant role for this mode, as explained earlier from
Eq. (4.44). This demonstrates two type of nonlinear interactions, the dominant one
between the mean flow-field and the disturbance field excited at almost zero frequency
(due to first term on right hand side of Eq. (4.44)) and another less dominant one at βL .
In the hierarchy of enstrophy content, the fifth and the sixth modes are the next
important ones, which form pair and exhibit wave-packet like features. The amplitude
function, as shown plotted in Fig. 4.14, initially grows exponentially at different rates
in different time intervals. However beyond t = 20, the amplitude starts decaying
up to t = 37, thereafter it grows and decays again. In [62, 64], this type of modes
have been termed as anomalous mode of second kind (T2 ), which also do not follow
Stuart–Landau–Eckhaus equation. These modes show time variation, that resembles
closely like a modulated wave-packet. The eleventh and the twelfth modes also are
of this type, as noted in Fig. 4.14. FFT of fifth-sixth and eleventh-twelfth mode pairs
show dominance of excitation frequency, though its first super-harmonic is also seen
to be present, with an amplitude that is three orders of magnitude lower than the
fundamental. For this reason, these mode pairs can be reconstructed by taking a
small band of frequency range around the dominant peak, as will be shown shortly.
Role of the first super-harmonic is very limited to small departure of oscillations
from pure sinusoidal variation.
Pair-forming fifth-sixth and eleventh-twelfth eigenfunctions are shown in
Figs. 4.16 and 4.17, which represent same wavenumber, as seen for the first-second
mode pair. Eigenfunction of fifth-sixth modes exhibit formation of two side-by-side
packets with a node at around x = 8.5 (Reδ∗ = 1585). Similarly, eleventh-twelfth
eigenfunctions show formation of three side-by-side packets with first node at around
x = 7.1 (Reδ∗ = 1499) and a second node at around x = 11 (Reδ∗ = 1803).
The ninth-tenth mode-pair depicts excitation of the first super-harmonic of βL ,
which is evident from the FFT plots in Fig. 4.15. Corresponding amplitude func-
tions in Fig. 4.14, show these modes to be present alongside first-second modes at all
times, with a frequency that is significantly higher. Eigenfunctions of these modes in
Fig. 4.16, show the corresponding wavenumber to be double the wavenumber of the
eigenfunctions of first-second, fifth-sixth and eleventh-twelfth modes. Time varia-
tion of amplitude function of these modes in Fig. 4.14, show that they qualitatively
follow variations similar to first-second modes. One notes that this variation is dif-
ferent from that is seen for bluff-body flow in [62], where a primary instability is
moderated by nonlinear self-interaction. For the cases of bluff body flows, no STWF
is formed, and there is a single dominant mode that follows the Stuart–Landau–
4.6 Proper Orthogonal Decomposition 265

Eckhaus equation. In Fig. 4.14, for first-second and ninth-tenth modes, punctuated
growth during t = 0 to 20 is due to the presence and the dynamics of the wave-
front not allowing the following wave-packet to grow, as dictated by the primary
linear spatial instability mechanism. However, this mode is similar to the fifth-sixth
regular modes in [62], which was the first super-harmonic of the first-second regular
mode pair shown in Fig. 4.20 of the reference. From Fig. 4.14, one notices similarity
between the amplitude envelope of first-second and ninth-tenth modes, except in
the time range 15 ≤ t ≤ 20. Difference between the envelopes of first-second and
ninth-tenth modes, during t = 15 to 20, is due to the generation of additional high
wavenumber oscillations near the wave-front, which is a secondary instability.
The fifteenth-sixteenth and seventeenth-eighteenth modes depict the STWF, as
seen in Fig. 4.14. Existence of STWF was established using Bromwich contour inte-
gral method of receptivity analysis and DNS results in Sengupta et al. [53, 59, 60].
Here, the same is explained in terms of POD modes. Amplitude and FFT of these
modes in Figs. 4.14 and 4.15, and eigenfunctions in Fig. 4.17, show that seventeenth-
eighteenth modes are the super-harmonic of fifteenth-sixteenth modes. Amplitudes
of theses modes are noticed to be almost zero up to t = 13, as the wave-front starts
developing after this time. Perceptible variation of the modal amplitudes is observed
from t = 25 to t = 45, the time interval when this wave-front forms and eventually
separates from the wave-packet leaving the computational domain.
Rempfer and Fasel [41, 42] have tried to relate regular modes with the coherent
structures in the flow. The jth coherent structure in [41, 42] have been defined as a
sum of (2j)th and (2j − 1)th pair-forming POD modes as

→ −
→ −

ρj ( X , t) = a2j (t) φ2j ( X ) + a2j−1 (t) φ2j−1 ( X ) (4.45)

Note that this splitting does not show that the pairs are orthogonal to each other. A
consistent splitting is given in Eqs. (5.1) and (5.2) in [62]. For example, for unstable
flows, vorticity perturbation can be expressed by Galerkin-type expansion in terms
of various instability modes as given in [17, 48],

 ∞

→ −
→ −

ω ( X , t) = [Aj (t)fj ( X ) + A∗j (t)fj ∗ ( X )] (4.46)
j=1

where quantities with asterisks again denote complex conjugates. Here, correspond-
ing to the jth instability mode, Aj (t) denote space-time dependent amplitude and


fj ( X ) describes the space-dependent eigenfunction, that satisfies prescribed bound-
ary conditions. As shown in [62], the instability modes  are related to POD modes
by defining a normalization factor εj = (λ2j + λ2j−1 )/ Nk=1 λk . By noting that pair-


forming POD modes have a phase shift of 90◦ , Aj (t) and fj ( X ) can be defined in
terms of the POD modes as,

Aj (t) = εj [a2j−1 (t) + i a2j (t)] (4.47)
266 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows

Table 4.2 Normalization √


Instability mode εj εj
factor εj and its square root
√ index j
εj for instability modes
1 0.563E+00 0.751E+00
2 0.173E+00 0.416E+00
3 0.508E-01 0.225E+00
4 0.166E-01 0.128E+00
5 0.244E-01 0.156E+00
6 0.151E-01 0.123E+00
7 0.650E-02 0.806E-01
8 0.850E-02 0.922E-01
9 0.670E-02 0.818E-01


→ 1 −
→ −

fj ( X ) = √ [φ2j−1 ( X ) − i φ2j ( X )] (4.48)
2 εj

The POD analysis is performed for disturbance enstrophy, and the eigenvalues
measure the same. We provide a note of caution that the disturbance enstrophy in the
POD context is not the same one as Ωd , that was used to describe enstrophy-based
instability theory. Thus, in representing the disturbance vorticity, the normalization
in Eqs. (4.47) and (4.48), is as indicated in terms of the square root of εj . In Table 4.2,
the normalization factor εj and its square root (used to relate instability modes with
POD modes) are tabulated for the first nine instability modes. Table 4.2 shows that ε4
is less than ε5 , and ε7 is less than both ε8 and ε9 . Reason behind this is clear from the
corresponding eigenvalues listed in Table 4.1. One can see from Table 4.1 that: (1) The
seventh and the thirteenth POD modes are isolated, anomalous modes of first kind; (2)
λ7 is less than (λ9 + λ10 ) and (3) λ13 is less than both (λ15 + λ16 ) and (λ17 + λ18 ).
Definition  of normalization factorsgives ε4 = λ7 / λk , ε5 = (λ9 + λ10 )/ λk ,
ε7 = λ13 / λk , ε8 = (λ15 + λ16 )/ λk and ε9 = (λ17 + λ18 )/ λk , which justify
the above observations. In fact, it is noted that whenever anomalous mode of first kind
appears (barring the first one), normalization factor εj of the respective instability
mode is less than several subsequent normalization factors.
In Figs. 4.18 and 4.19, modulus of the amplitude function |Aj (t)| and the space


dependent function |fj ( X )| of the instability modes are plotted. Comparing each
|Aj (t)| in Fig. 4.18 with corresponding a2j−1 (t) and a2j (t) plotted in Fig. 4.14, one
can see that for each instability mode, |Aj (t)| is essentially the upper envelope of


a2j−1 (t) and a2j (t). Similarly, if the modulus of space dependent functions |fj ( X )|

→ −

and (φ2j ( X ), φ2j−1 ( X )) are compared at any height as a function of x, it is noted

→ −
→ −

that |fj ( X )| defines the upper envelope of (φ2j ( X ), φ2j−1 ( X )).
4.6 Proper Orthogonal Decomposition 267

( i ) 1 st Instability Mode ( ii ) 2 nd Instability Mode


8
4
6
| A1 ( t ) |

| A2 ( t ) |
4 2
2

0 0
0 10 20 30 40 50 60 0 10 20 30 40 50 60
time(t) time(t)
rd th
( iii ) 3 Instability Mode 2 ( iv ) 4 Instability Mode
4 1
| A3 ( t ) |

| A4 ( t ) |
0
2
-1

0 -2
0 10 20 30 40 50 60 0 10 20 30 40 50 60
time(t) time(t)
th th
2 (v) 5 Instability Mode ( vi ) 6 Instability Mode

1.5 2
| A5 ( t ) |

| A6 ( t ) |

1
1
0.5

0 0
0 10 20 30 40 50 60 0 10 20 30 40 50 60
time(t) time(t)

( vii ) 7 th Instability Mode 3 ( viii ) 8 th Instability Mode


1
2
| A7 ( t ) |

| A8 ( t ) |

-1 1

-2 0
0 10 20 30 40 50 60 0 10 20 30 40 50 60
time(t) time(t)
th
( ix ) 9 Instability Mode
2
| A9 ( t ) |

0
0 10 20 30 40 50 60
time(t)

Fig. 4.18 Modulus of amplitude function |Aj (t)| of the instability modes shown plotted as a function
of time for zero-pressure gradient flow over semi-infinite flat-plate
268 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows

0.1 0.1
( i ) 1st Instability Mode ( ii ) 2nd Instability Mode

0.05 0.05

0 0
0 5 10 15 20 0 5 10 15 20

0.1 rd 0.1 th
( iii ) 3 Instability Mode ( iv ) 4 Instability Mode

0.05 0.05

0 0
0 5 10 15 20 0 5 10 15 20

0.1 th 0.1 th
( v ) 5 Instability Mode ( vi ) 6 Instability Mode

0.05 0.05

0 0
0 5 10 15 20 0 5 10 15 20

0.1 th 0.1 th
( vii ) 7 Instability Mode ( viii ) 8 Instability Mode

0.05 0.05

0 0
0 5 10 15 20 0 5 10 15 20

0.1 th
( ix ) 9 Instability Mode

0.05

0
0 5 10 15 20



Fig. 4.19 Modulus of space dependent function |fj ( X )| of the indicated instability modes are
shown plotted as a function of time for zero-pressure gradient flow over semi-infinite flat-plate

Here, an attempt is made to reconstruct the amplitude function of anomalous


modes of second kind by taking a band of frequencies about the most dominant
frequency (F0 ) noted in the FFT. The ordinate in Fig. 4.15 is shown in terms of
βL Δt. In Figs. 4.20 and 4.21, this type of reconstruction is shown for the fifth and
the eleventh modes, respectively. FFT of amplitude functions performed with 8192
points (hence Δt = (t1 − t0 )/8192) indicate F0 = 10.158 for the fifth mode and
F0 = 10.262 for the eleventh mode. As mentioned earlier, excitation frequency is
βL = 10 and thus the dominant frequencies are quite close to βL . About 170 data
samples for each time period corresponding to F0 have been taken. This shows that
the assumption of signal problem used for linear spatial stability theory is strictly
not correct and the dynamics should be obtained by spatio-temporal analyses.
Since, POD is performed from t0 = 0 to t1 = 60, the spectrum is obtained with
ΔF = 2π/(t1 − t0 ) = 0.105. From Figs. 4.20 and 4.21, it is seen that reconstruction
with ±12ΔF (ΔβL Δt = ±0.01841 as shown in Figs. 4.20 and 4.21) around F0 repro-
duces the original, perfectly in the time range ts ≤ t ≤ te . For the fifth mode, ts = 1.8
and te = 58.6, whereas for the eleventh mode, ts = 3.5 and te = 56.5. Mismatch near
4.6 Proper Orthogonal Decomposition 269

5 ( i ) Amplitude function of 5
th
POD mode

a5 ( t )
0

-5
0 10 20 30 40 50 60
time
( iii ) Reconstructed with a non-dimensional frequency band of
5 ( ii ) Reconstructed with a non-dimensional frequency 5 ΔβΔt= 6.136×10 about peak
-3
band of ΔβΔt= 3.0679×10-3 about peak
a5 ( t )

a5 ( t )
0 0

-5 -5
0 10 20 30 40 50 60 0 10 20 30 40 50 60
time time
( iv ) Reconstructed with a non-dimensional frequency band of ( v ) Reconstructed with a non-dimensional frequency band of
5 ΔβΔt= 9.204×10 about peak
-3 5 ΔβΔt= 0.01227 about peak
a5 ( t )

a5 ( t )

0 0

-5 -5
0 10 20 30 40 50 60 0 10 20 30 40 50 60
time time
( vi ) Reconstructed with a non-dimensional frequency band of ( vii ) Reconstructed with a non-dimensional frequency band of
5 ΔβΔt= 0.01534 about peak 5 ΔβΔt= 0.01841 about peak
a5 ( t )

a5 ( t )

0 0

-5 -5
0 10 20 30 40 50 60 0 10 20 30 40 50 60
time time

Fig. 4.20 Reconstructed signals of a5 (t) with indicated frequency band about the peak at βΔt =
7.234 × 10−2 shown in frames (ii) to (vii). The original signal a5 (t) is shown in the top frame

t0 and t1 are related to Abel’s and Tauber’s theorem that relates the unknown in phys-
ical and transformed spectral planes. Since calculations are performed with finite
resources, we do take a finite spectrum with approximate resolution and thereby
having inaccuracies near the origin and the far field. The Tauber’s theorem states that
signal near t = 0 is determined by the spectrum at βL → ∞ (see page 84 of [52] for
further details), which we do not have in the full range of βL Δt. However, the mis-
match near t1 is related to our calculation of FFT by assuming periodic continuation
outside the range of consideration in physical plane and this enforced periodicity
causes the mismatch at t1 . This type of mismatch near t1 is typical of any Fourier
spectral method.
Reconstruction of the anomalous modes of second kind in Figs. 4.20 and 4.21,
emphasizes the fact that the contributions come from the side-band of the main peak
noted in the FFT shown in Fig. 4.15. For both the cases, this central peak corresponds
270 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows

th
( i ) Amplitude function of 11 POD mode
2

a5 ( t )
0

-2
0 10 20 30 40 50 60
time

( ii ) Reconstructed with non-dimensional frequency band ( iii ) Reconstructed with non-dimensional frequency band
of ΔβΔt = 3.0679× 10-3 about peak of ΔβΔt = 6.136× 10-3 about peak
2 2
a11 ( t )

a11 ( t )
0 0

-2 -2
0 10 20 30 40 50 60 0 10 20 30 40 50 60
time time
( iv ) Reconstructed with non-dimensional frequency band ( v ) Reconstructed with non-dimensional frequency band
of ΔβΔt = 9.204× 10-3 about peak of ΔβΔt = 0.01227 about peak
2 2
a11 ( t )

a11 ( t )

0 0

-2 -2
0 10 20 30 40 50 60 0 10 20 30 40 50 60
time time
( vi ) Reconstructed with non-dimensional frequency band ( vii ) Reconstructed with non-dimensional frequency band
of ΔβΔt = 0.01534 about peak of ΔβΔt = 0.01841 about peak
2 2
a11 ( t )

a11 ( t )

0 0

-2 -2
0 10 20 30 40 50 60 0 10 20 30 40 50 60
time time

Fig. 4.21 Reconstructed signals of a11 (t) with indicated frequency band about the peak at βΔt =
7.234 × 10−2 are shown plotted in frames (ii) to (vii). The original signal a11 (t) is shown in the top
frame

to the frequencies mentioned above, those are slightly detuned from the excitation
frequency.
A contemporary system reduction strategy for fluid flows is the Koopman mode
analysis [44], based on the spectral analysis of the finite-dimensional representation
of the Koopman operator using Krylov subspace method, a variant of the com-
monly used Arnoldi iteration. This linear mapping is equivalent to the dynamic
mode decomposition (DMD) technique developed [22, 45, 79]. Koopman analysis
results in modes that are orthogonal in time (single frequency modes) while POD
gives spatially orthogonal, multi-time periodic modes, which capture the most ener-
getic/enstrophic structures of the flow. The results of Koopman mode analysis and its
variants in [8], show that this strategy although successful in the initial linear growth
stage, and probably in the limit cycle stage, is not so successful in describing the
4.6 Proper Orthogonal Decomposition 271

transient disturbance growth in which the frequency of large scale oscillations is no


longer approximately constant. POD modes being free from any restriction on the
frequency composition, are effective in capturing transient dynamics, as shown in
[62] and here. Also, we have shown here as to how one can relate the POD modes
with instability modes. Once this is achieved, one can perform space-time dependent
nonlinear analysis of instability.

References

1. Badr, H. M., Coutanceau, M., Dennis, S. C. R., & Menard, C. (1990). Unsteady flow past a
rotating circular cylinder at Reynolds numbers 1000 and 10,000. Journal of Fluid Mechanics,
220, 459–484.
2. Benjamin, T. B., & Feir, J. E. (1967). The disintegration of wave trains on deep water. Part 1.
Theory. Journal of Fluid Mechanics, 27(3), 417–430.
3. Bhaumik, S., & Sengupta, T. K. (2011). On the divergence-free condition of velocity in two-
dimensional velocity-vorticity formulation of incompressible Navier–Stokes equation. In 20th
AIAA CFD Conference, 27–30 June, Honululu, Hawaii, USA
4. Bhaumik, S., & Sengupta, T. K. (2014). Precursor of transition to turbulence: Spatiotemporal
wave front. Physical Review E, 89(4), 043018.
5. Bhaumik, S., & Sengupta, T. K. (2015). A new velocity-vorticity formulation for direct numer-
ical simulation of 3D transitional and turbulent flows. Journal of Computational Physics, 284,
230–260.
6. Breuer, K. S., & Kuraishi, T. (1993). Bypass transition in two and three dimensional boundary
layers. In AIAA 93-3050
7. Brinckman, K. W., & Walker, J. D. A. (2001). Instability in a viscous flow driven by streamwise
vortices. Journal of Fluid Mechanics, 432, 127–166.
8. Chen, K., Tu, J. H., & Rowley, C. (2012). Variants of dynamic mode decomposition: Boundary
condition, Koopman and Fourier analyses. Journal of Nonlinear Science, 22, 887–915.
9. Daube, O. (1992). Resolution of the 2D Navier–Stokes equations in velocity-vortcity form by
means of an influence matrix technique. Journal of Computational Physics, 103, 402–414.
10. Davies, S. J., & White, C. M. (1928). An experimental study of the flow of water in pipes of
rectangular section. Proceedings of the Royal Society of London. Series A, 119, 92.
11. Deane, A. E., Kevrekidis, I. G., Karniadakis, G. E., & Orszag, S. A. (1991). Low-dimensional
models for complex geometry flows: Application to grooved channels and circular cylinders.
Physics of Fluids, 3, 2337–2354.
12. Degani, A. T., Walker, J. D. A., & Smith, F. T. (1998). Unsteady separation past moving
surfaces. Journal of Fluid Mechanics, 375, 1–38.
13. Doering, C. R., & Gibbon, J. D. (1995). Applied analysis of the Navier–Stokes equations. UK:
Cambridge University Press.
14. Doligalski, T. L., Smith, C. R., & Walker, J. D. A. (1994). Vortex interaction with wall. Annual
Review of Fluid Mechanics, 26, 573–616.
15. Donzis, D. A., & Yeung, P. K. (2010). Resolution effects and scaling in numerical simulations
of passive scalar mixing in turbulence. Physica D, 239, 1278–87.
16. Donzis, D. A., Yeung, P. K., & Sreenivasan, K. R. (2008). Energy dissipation rate and enstrophy
in isotropic turbulence: resolution effects and scaling in direct numerical simulations. Physics
of Fluids, 20, 045108-1–16.
17. Drazin, P. G., & Reid, W. H. (1981). Hydrodynamic stability. UK: Cambridge University Press.
18. Eiseman, P. R. (1985). Grid generation for fluid mechanics computation. Annual Review of
Fluid Mechanics, 17, 487–522.
272 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows

19. Gatski, T. B., Grosch, C. E., & Rose, M. E. (1982). A numerical study of the 2-dimensional
Navier–Stokes equations in vorticity-velocity variables. Journal of Computational Physics, 48,
1–22.
20. Guj, G., & Stella, F. (1993). A vorticity-velocity method for the numerical solution of 3D
incompressible flows. Journal of Computational Physics, 106, 286–298.
21. Heisenberg, W. (1924). Über stabilität und turbulenz von flüssigkeitsströmen. Annalen der
Physik, 379, 577–627. (Translated as ‘On stability and turbulence of fluid flows’. NACA Tech.
Memo. Wash. No 1291 1951).
22. Hemati, M. S., Williams, M. O., & Rowley, C. W. (2014). Dynamic mode decomposition for
large and streaming datasets. Physics of Fluids, 26(11), 111701.
23. Holmes, P., Lumley, J. L., & Berkooz, G. (1996). Coherent structures, dynamical systems and
symmetry. Cambridge, UK: Cambridge University Press.
24. Huang, H., & Li, M. (1997). Finite-difference approximation for the velocity-vorticity formu-
lation on staggered and non-staggered grids. Computers & Fluids, 26(1), 59–82.
25. Kosambi, D. D. (1943). Statistics in function space. Journal of the Indian Mathematical Society,
7, 76–88.
26. Landahl, M. T., & Mollo-Christensen, E. (1992). Turbulence and random processes in fluid
mech. New York, USA: Cambridge University Press.
27. Leib, S. J., Wundrow, D. W., & Goldstein, M. E. (1999). Generation and growth of boundary
layer disturbances due to free-stream turbulence. In AIAA Conference Paper, AIAA-99-0408
28. Lim, T. T., Sengupta, T. K., & Chattopadhyay, M. (2004). A visual study of vortex-induced
sub-critical instability on a flat plate laminar boundary layer. Experiments in Fluids, 37, 47–55.
29. Loéve, M. (1978). Probability theory Vol. II (4th ed., Vol. 46)., Graduate Texts in Mathematics
New York, USA: Springer.
30. Ma, X., & Karniadakis, G. E. (2002). A low-dimensional model for simulating three-
dimensional cylinder flow. Journal of Fluid Mechanics, 458, 181–190.
31. Mathieu, J., & Scott, J. (2000). An introduction to turbulent flows. Cambridge, UK: Cambridge
University Press.
32. Monin, A. S., & Yaglom, A. M. (1971). Statistical fluid mechanics: mechanics of turbulence.
Cambridge, MA, USA: The MIT Press.
33. Morkovin, M. V. (1991). Panoramic view of changes in vorticity distribution in transition,
instabilities and turbulence. In D. C. Reda, H. L. Reed, & R. Kobyashi (Eds.), Transition to
turbulence (Vol. 114, pp. 1–12). USA: ASME FED Publication.
34. Morzynski, M., Afanasiev, K., & Thiele, F. (1999). Solution of the eigenvalue problems result-
ing from global nonparallel flow stability analysis. Computer Methods in Applied Mechanics
and Engineering, 169, 161–176.
35. Nagarajan, S., Lele, S. K., & Ferziger, J. H. (2003). A robust high-order compact method for
large eddy simulation. Journal of Computational Physics, 19, 392–419.
36. Napolitano, M., & Pascazio, G. (1991). A numerical method for the vorticity-velocity Navier–
Stokes equations in two and three dimensions. Computers & Fluids, 19, 489–495.
37. Noack, B. R., Afanasiev, K., Morzynski, M., Tadmor, G., & Thiele, F. (2003). A hierarchy of
low-dimensional models for the transient and post-transient cylinder wake. Journal of Fluid
Mechanics, 497, 335–363.
38. Obabko, A. V., & Cassel, K. W. (2002). Navier–Stokes solutions of unsteady separation induced
by a vortex. Journal of Fluid Mechanics, 465, 99–130.
39. Peridier, V. J., Smith, F. T., & Walker, J. D. A. (1991). Vortex-induced boundary-layer separa-
tion. Part 1. The unsteady limit problem. Re → ∞.. Journal of Fluid Mechanics, 232, 99–131.
40. Peridier, V. J., Smith, F. T., & Walker, J. D. A. (1991). Vortex-induced boundary-layer sepa-
ration. Part 2. Unsteady interacting boundary-layer theory. Journal of Fluid Mechanics, 232,
133–165.
41. Rempfer, D., & Fasel, H. (1994). Evolution of three-dimensional coherent structures in a flat-
plate boundary layer. Journal of Fluid Mechanics, 260, 351–375.
42. Rempfer, D., & Fasel, H. (1994). Dynamics of three-dimensional coherent structures in a
flat-plate boundary layer. Journal of Fluid Mechanics, 275, 257–283.
References 273

43. Robinson, S. K. (1991). Coherent motions in the turbulent boundary layer. Annual Review of
Fluid Mechanics, 23, 601–639.
44. Rowley, C., Mezi, I., Bagheri, S., Schlatter, P., & Henningson, D. S. (2009). Spectral analysis
of nonlinear flows. Journal of Fluid Mechanics, 641, 1–13.
45. Schmid, P. J. (2010). Dynamic mode decomposition of numerical and experimental data. Jour-
nal of Fluid Mechanics, 656, 5–28.
46. Schlichting, H. (1933). Zur entstehung der turbulenz bei der plattenströmung. Nach. Gesell. d.
Wiss. z. Gött., MPK, 42, 181–208.
47. Schubauer, G. B., & Skramstad, H. K. (1947). Laminar boundary layer oscillations and the
stability of laminar flow. Journal of Aerosol Science, 14, 69–78.
48. Sengupta, T. K. (2012). Instabilities of flows and transition to turbulence CRC press. Florida,
USA: Taylor and Francis Group.
49. Sengupta, T. K. (2013). High accuracy computing methods: Fluid flows and wave phenomenon.
USA: Cambridge University Press.
50. Sengupta, T. K., & Lekoudis, S. G. (1985). Calculation of turbulent boundary layer over moving
wavy surface. AIAA Journal, 23(4), 530–536.
51. Sengupta, T. K., & Dipankar, A. (2005). Subcritical instability on the attachment-line of an
infinite swept wing. Journal of Fluid Mechanics, 529, 147–171.
52. Sengupta, T. K., & Poinsot, T. (2010). Instabilities of flows: With and without heat transfer and
chemical reaction. Wien, New York: Springer.
53. Sengupta, T. K., & Bhaumik, S. (2011). Onset of turbulence from the receptivity stage of fluid
flows. Physical Review Letters, 154501, 1–5.
54. Sengupta, T. K., De, S., & Gupta, K. (2001). Effect of free-stream turbulence on flow over
airfoil at high incidences. Journal of Fluids and Structures, 15(5), 671–690.
55. Sengupta, T. K., Chattopadhyay, M., Wang, Z. Y., & Yeo, K. S. (2002). By-pass mechanism of
transition to turbulence. Journal of Fluids and Structures, 16, 15–29.
56. Sengupta, T. K., De, S., & Sarkar, S. (2003). Vortex-induced instability of an incompressible
wall-bounded shear layer. Journal of Fluid Mechanics, 493, 277–286.
57. Sengupta, T. K., Ganeriwal, G., & De, S. (2003). Analysis of central and upwind compact
schemes. Journal of Computational Physics, 192(2), 677–694.
58. Sengupta, T. K., Kasliwal, A., De, S., & Nair, M. (2003). Temporal flow instability for Magnus–
Robins effect at high rotation rates. Journal of Fluids and Structures, 17, 941–953.
59. Sengupta, T. K., Rao, A. K., & Venkatasubbaiah, K. (2006). Spatiotemporal growing wave
fronts in spatially stable boundary layers. Physical Review Letters, 96(22), 224504.
60. Sengupta, T. K., Rao, A. K., & Venkatasubbaiah, K. (2006). Spatiotemporal growth of dis-
turbances in a boundary layer and energy based receptivity analysis. Physics of Fluids, 18,
094101.
61. Sengupta, T. K., Dipankar, A., & Sagaut, P. (2007). Error dynamics: Beyond von Neumann
analysis. Journal of Computational Physics, 226(2), 1211–1218.
62. Sengupta, T. K., Singh, N., & Suman, V. K. (2010). Dynamical system approach to instability
of flow past a circular cylinder. Journal of Fluid Mechanics, 656, 82–115.
63. Sengupta, T. K., Bhaumik, S., & Bhumkar, Y. G. (2011). Nonlinear receptivity and instability
studies by POD. In AIAA Conference on Theoretical Fluid Mechanics, AIAA Paper No. 2011-
3293
64. Sengupta, T. K., Singh, N., & Vijay, V. V. S. N. (2011). Universal instability modes in internal
and external flows. Computers & Fluids, 40, 221–235.
65. Sengupta, T. K., Bhaumik, S., & Bhumkar, Y. (2012). Direct numerical simulation of two-
dimensional wall-bounded turbulent flows from receptivity stage. Physical Review E, 85(2),
026308.
66. Sengupta, T. K., Bhumkar, Y. G., & Sengupta, S. (2012). Dynamics and instability of a shielded
vortex in close proximity of a wall. Computers & Fluids, 70, 166–175.
67. Sengupta, T. K., Singh, H., Bhaumik, S., & Chowdhury, R. R. (2013). Diffusion in inhomoge-
neous flows: Unique equilibrium state in an internal flow. Computers & Fluids, 88, 440–451.
274 4 Nonlinear Theoretical and Computational Analysis of Fluid Flows

68. Sengupta, T. K., Haider, S. I., Parvathi, M. K., & Pallavi, G. (2015). Enstrophy-based proper
orthogonal decomposition for reduced-order modeling of flow a past cylinder. Physical Review
E, 91(4), 043303.
69. Sharma, N., Sengupta, A., Rajpoot, M., Samuel, R. J., & Sengupta, T. K. (2017). Hybrid sixth
order spatial discretization scheme for non-uniform Cartesian grids. Computers & Fluids,
157(3), 208–231.
70. Siegel, S. G., Seidel, J., Fagley, C., Luchtenburg, D. M., Cohen, K., & Mclaughlin, T. (2008).
Low-dimensional modelling of a transient cylinder wake using double proper orthogonal
decomposition. Journal of Fluid Mechanics, 610, 1–42.
71. Sirovich, L. (1987). Turbulence and the dynamics of coherent structures. Part (I) coherent struc-
tures, part (II) symmetries and trans-formations and part (III) dynamics and scaling. Quarterly
of Applied Mathematics, 45(3), 561–590.
72. Smith, C. R., Walker, J. D. A., Haidari, A. H., & Soburn, U. (1991). On the dynamics of near-
wall turbulence. Philosophical Transactions of the Royal Society of London A, 336, 131–175.
73. Sommerfeld, A. (1949). Partial differential equation in physics. New York: Academic Press.
74. Stuart, J. T. (1960). On the nonlinear mechanics of wave disturbances in stable and unstable
parallel flows. Part 1. The basic behaviour in plane Poiseuille flow. Journal of Fluid Mechanics,
9, 353–370.
75. Suman, V. K., Sengupta, T. K., Durga Prasad, C. J., Mohan, K. S., & Sanwalia, D. (2017).
Spectral analysis of finite difference schemes for convection diffusion equation. Computers &
Fluids, 150, 95–114.
76. Taylor, G. I. (1936). Statistical theory of turbulence. V. Effects of turbulence on boundary layer.
Proceedings of the Royal Society A, 156(888), 307–317.
77. Tennekes, H., & Lumley, J. L. (1971). First course in turbulence. Cambridge, MA: MIT Press.
78. Tollmien, W. (1931). The Production of Turbulence. NACA Report-TM 609.
79. Tu, J. H., Rowley, C. W., Luchtenburg, D. M., Brunton, S. L., & Kutz, J. N. (2014). On
dynamic mode decomposition: Theory and applications. Journal of Computational Dynamics,
1(2), 391–421.
80. Van der Vorst, H. A. (1992). Bi-CGSTAB: A fast and smoothly converging variant of Bi-CG
for the solution of non-symmetric linear systems. SIAM Journal on Scientific and Statistical
Computing, 12, 631–644.
81. Wu, X. H., Wu, J. Z., & Wu, J. M. (1995). Effective vorticity-velocity formulations for 3D
incompressible viscous flows. Journal of Computational Physics, 122, 68–82.
82. Yeung, P. K., Donzis, D. A., & Sreenivasan, K. R. (2012). Dissipation, enstrophy and pressure
statistics in turbulence simulations at high Reynolds numbers. Journal of Fluid Mechanics,
700, 5–15.
Chapter 5
Dynamics of the Spatio-Temporal
Wave-Front in 2D Framework

5.1 Introduction

A general consensus among fluid dynamicists is that flow transition from lami-
nar to turbulent state occurs due to its instability by imposed and/or background
omnipresent disturbances [7, 17]. These ideas have prompted researchers to study
the problem of flow transition from the perspective of the stability or receptivity of
equilibrium flows. For a ZPGBL, first attempts include analyses by linearized invis-
cid and viscous instability theories. For a parallel boundary layer the latter approach
gives rise to the OSE [7, 17]. As discussed in the previous chapters, the solution of
the OSE predicts the existence of spatially modulated wavy solutions, known as TS
waves [7]. First experimental detection of spatially evolving TS wave-packets were
reported by Schubauer and Skramstad [16], who essentially perturbed the ZPGBL
by vibrating a ribbon at a fixed frequency inside it. Following mathematical physics,
disturbance evolution in any continuum medium can occurs following temporal or
spatial or spatio-temporal routes. For example, there has been an effort [4], where
wave propagation problem in electromagnetic medium has been considered with the
proviso that the wave-train is preceded by a spatio-temporal wave-packet or STWF.
The success of the experiments by Schubauer and Skramstad [16] in detecting TS
waves, prompted researchers to predominantly consider the possibility that flow
transition in fluid flows (specifically for wall-bounded flows) follow spatial route
following the growth of TS waves only. Thus, in fluid mechanics no effects have
been made to find STWF for a long time.
The linear spatial theory does not satisfactorily explain all the aspects of the
experiment [16], and this has led to search for alternate explanations for route of
transition following spatio-temporal growth of disturbances. Wave-front dynam-
ics following spatio-temporal route have been investigated, as a consequence of
convective/absolute instability [11], adopting inviscid approach for a pulse excita-
tion. Considering non-localized finite amplitude excitation, Chomaz and co-authors
[5] have illustrated spatio-temporal front, as a consequence of initial perturbations
of finite extent and amplitude, which separates from undisturbed state. In [23,

© Springer Nature Singapore Pte Ltd. 2019 275


T. K. Sengupta and S. Bhaumik, DNS of Wall-Bounded Turbulent Flows,
https://doi.org/10.1007/978-981-13-0038-7_5
276 5 Dynamics of the Spatio-Temporal Wave-Front in 2D Framework

24], spatio-temporal receptivity analysis was presented for the OSE via Bromwich
contour integral method (BCIM). These aspects have already been discussed in the
previous chapters. In [23, 24], localized time-harmonic excitation with impulsive
onset was considered. These approaches of studying spatio-temporal evolution of
perturbation are different from those described in [5]. The finite start-up time excites
a large band of frequencies, even for the time-harmonic excitation case, which causes
the generation of STWF. The resultant STWF obtained by BCIM is distinctly differ-
ent from the one obtained in various experimental and theoretical approaches [10,
14], where wave-packets are created by pulse excitation and not by continuous exci-
tation, as in [16]. Reported results in [23, 24] show that STWF can grow even for
excitation parameters, which are stable according to spatial linear theory. Indeed, the
dynamics of STWF is distinctly different from the spatially modulating TS wave-
packet. Similar conclusions were also noted in [9] based on the obtained numerical
results, where the authors pointed out that “the nonlinear behavior of the leading
wave-packet would be very different from that of pure Tollmien-Schlichting waves”.
This chapter attempts to bridge this gap by providing analysis of the spatio-temporal
evolution of disturbances following localized excitations of various types.

5.2 From Linear Theory to Turbulence via Deterministic


Routes

The nonlinear receptivity results of the 2D ZPGBL is considered in the present


chapter, with the specific emphasis fixed on understanding the dynamics of the STWF.
As discussed in Chap. 3, when the wall bounded shear layer is excited harmonically
from the wall, obtained solution displays (i) a STWF, (ii) a local solution and (iii)
the wave-packet composed of Tollmien-Schlichting (TS) wave. If the boundary layer
is excited by a simultaneous blowing-suction strip at the wall, as in Fig. 5.1, so that
there is zero mass transfer at any instant of time, then the peak amplitude, frequency
of excitation and exciter width play major roles in deciding the response.
Different routes of disturbance evolution can be noted when 2D ZPG is excited
harmonically from the wall. Some of these distinct routes of disturbance evolution
are as follows:
1. Following this disturbance evolution route, the STWF is seen to be solely respon-
sible for transition to turbulence which starts off via a linear mechanism, as given
in [23], and its subsequent nonlinear growth triggers transition. These features
are noted for lower amplitude excitation, at moderate to high frequency cases.
2. If the excitation amplitude is increased for those frequency cases, where distur-
bance evolution follows previous route, the STWF is noted to directly enter into
non-linear phase of amplification, while inducing unsteady separations on the
wall to cause bypass transition.
5.2 From Linear Theory to Turbulence via Deterministic Routes 277

y
Fa - Boundary
U∞ Outflow
Boundary
Inlet ymax
Edge of Shear Layer
Exciter

Wall Boundary δout


x1 x2
xin O x
xout

Fig. 5.1 Schematic of 2D receptivity to SBS wall excitation

3. At lower frequencies, distinct presence of a spatially modulated TS wave-packet


is not noted. Instead, a spatio-temporally evolving wave-packet is noted, which at
subsequent time-instants intermittently issue secondary wave-packets, inducing
flow transition.
4. Depending on exciter location or amplitude, another route is noted which is inter-
mediate between the previous and the first route. Following this route, the spatio-
temporal wave-packet also interacts with the primary STWF at later times by
issuing subsequent secondary STWFs to cause transition of the flow. These dif-
ferent disturbance evolution routes are described in details, in subsequent section.

5.2.1 Governing Equations, Numerical Schemes


and Simulation Parameters

The schematic diagram of the receptivity problem is shown in Fig. 5.1. Here, all
the reported simulation results use, 4501 and 401 grid points along streamwise
and wall-normal directions, respectively, while the computational domain is from
xin = −0.05L to xout = 120L in the streamwise direction, and up to ymax = 1.5L in
the wall-normal direction. Numerical simulations performed with larger ymax show
it to have virtually no effects on the simulation results.
The Reynolds number based on the reference velocity U∞ (free-stream speed) and
length L is ReL = 105 , as also used in the previous chapters. A time-step of 8 × 10−5 is
used to integrate in time, using optimized three-stage Runge–Kutta (ORK3 ) method of
[15]. As illustrated in Chap. 2, ORK3 minimizes the dispersion, phase and attenuation
errors and allows one to adopt larger time-steps than the conventional RK4 method.
278 5 Dynamics of the Spatio-Temporal Wave-Front in 2D Framework

Grid clustering in the streamwise and wall-normal directions are performed as


given in Sect. 4.3.4. The minimum and maximum resolution along the stream-
wise direction are Δxmin = 2 × 10−3 and Δxmax = 2.7 × 10−2 , respectively. The
wall-normal resolution is given, as Δymin = 5.5 × 10−4 , which stretches up to
Δymax = 7.77 × 10−3 , at the far-field boundary. The Reynolds number based on
the displacement thickness δ ∗ at the outflow of the computational domain (i.e., at
x = xout ) is Reδ∗ (xout ) = 5958. Simulations are carried out in such a long domain in
the streamwise direction, as in [18, 19]. In [8] the longest simulations were reported
only up to Reδ∗ = 450. The width of the computational domain along the wall-normal
direction is 25 times the local displacement thickness at the outflow, in [18, 19].
Stream function and vorticity (ψ, ω)-formulation of 2D NSE is solved here in the
(ξ, η)-plane. These equations have been described previously in Chap. 3. These are
adopted to simulate the receptivity problem with better accuracy, and the ability to
satisfy the mass conservation exactly at all time instants. The convection terms in
the vorticity transport equation are discretized with OUCS3 scheme, while the self-
adjoint diffusion terms are discretized with second-order central difference scheme.
Simulations are performed in MPI parallelization framework using thirty-two cores,
using Schwarz domain decomposition technique. The computational domain is
decomposed along the streamwise direction only, with thirty points overlap between
successive domains. Such a large overlap among each successive domain is taken to
drastically minimize reflection and spurious wave generation from the inter-domain
boundaries.
One first computes the equilibrium flow governed by NSE in the whole domain
before switching on the excitation. At the inflow of the domain (AB) and the far-field
(BC) boundary of the domain, as shown in Fig. 5.1, the free stream conditions are
specified for both perturbed and unperturbed flow as

∂ψ
= h2 (5.1)
∂η

ω=0 (5.2)

This has to be emphasized that, the far-field boundary is practically very far away
from the wall, so the prescription of free-stream condition there does not affect the
receptivity solutions at all. This has also been confirmed numerically for two different
computational domains with different values of ymax for some representative cases.
For the equilibrium flow, the no-slip and zero-normal velocity conditions are
prescribed at the wall, which specifies the necessary boundary conditions for ψ and
ω on the wall by,
ψw = ψ0 = Constant (5.3)

1 ∂ 2ψ
ωw = − (5.4)
h22 ∂η2
5.2 From Linear Theory to Turbulence via Deterministic Routes 279

For the excited flow, the no-slip condition is still applicable on the wall. However,
the wall-normal velocity assumes a prescribed value vw (x, t), as fixed by the corre-
sponding specific exciter. The corresponding wall-boundary conditions on ψ and ω
are specified as
ψw = ψ0 + ψwp (5.5)
 
1 ∂ h2 ∂ψ 1 ∂ 2ψ
ωw = − − (5.6)
h1 h2 ∂ξ h1 ∂ξ h22 ∂η2

where, ψwp is the perturbation stream-function given as


 x
ψwp (x, t) = vw (x̂, t)d x̂ (5.7)
0

At the bottom boundary, ahead of the leading edge (the segment AO, as shown
in Fig. 5.1), the symmetry condition is specified for both perturbed as well as unper-
turbed flow. This specifies the relevant boundary conditions for ψ and ω as

ψ = ψ0 = Constant (5.8)

ω=0 (5.9)

At the outflow, vorticity is calculated from the convective Sommerfeld boundary


condition given by,
∂ω ∂ω
+ Uc = 0, (5.10)
∂t ∂x
which is time-advanced by using the same time-stepping method used for vorticity-
transport equation. The prescription of the radiative Sommerfeld boundary condition
allows the disturbances to smoothly convect out of the outflow boundary. Convective
speed of disturbances through the outflow, Uc , in the Sommerfeld radiative outflow
condition in Eq. (5.10), is taken as the free stream speed U∞ . However, it is seen
and verified by some numerical simulations that change in the value of Uc , does not
alter the solution in the interior of the computational domain. Only the solution at
a very narrow strip near the outflow boundary, is affected in such cases. Both mean
and perturbed stream-function ψ, at the outflow boundary is calculated from the
following condition on wall-normal component of velocity given as

∂v
=0 (5.11)
∂x
This is a so-called soft boundary condition that fixes the stream-function value.
Note that the computational domain is so large, that in the streamwise direction this
outflow boundary condition is equivalent to evaluating the vorticity at the outflow
from ω = − ∂u∂y
, an equivalent boundary layer approximation.
280 5 Dynamics of the Spatio-Temporal Wave-Front in 2D Framework

Other forms of boundary condition on ψ such as the one derived from

∂u ∂u
+ Uc =0
∂t ∂x
have also been tested, and found to produce identical results for both mean and
perturbation quantities, except in a very narrow zone, near the outflow boundary.
This boundary condition is not used for further 2D simulations.
As for the perturbed flow, the equilibrium solution is used as the initial condition,
as mentioned earlier. Equilibrium flow is computed till a steady flow-field is obtained
inside the entire domain. This is ensured by checking the time history of ∂ω/∂t.
In [2], the similarity functions f  (η̄), f  (η̄) and f  (η̄) obtained from the solution
of the Blasius equation are compared with the corresponding obtained equilibrium
solution at several streamwise locations, varying from very close to the leading edge
to xout . The match is good except at streamwise stations, which are close to the leading
edge of the plate. Next, the receptivity of the computed solution to monochromatic
deterministic excitation frequency is described.

5.3 Small Amplitude Disturbance at Moderate Frequency

Obtained equilibrium flow is perturbed time harmonically by the SBS exciter. The
amplitude function of the SBS exciter is defined in Chap. 3 (see Eq. (3.102)). Here,
the receptivity results of the ZPG boundary layer is discussed, where the amplitude
control parameter of the SBS strip exciter is α1 = 0.002 (see Eq. (3.101)), while the
non-dimensional frequency of the excitation is F = 1.0 × 10−4 . The nondimensional
frequency F, is related to the physical frequency f as F = 2π νf /U∞ 2
, where ν and
U∞ are kinematic viscosity and free-stream velocity, respectively. The amplitude of
the excitation is kept at 0.2% of the free-stream speed, U∞ . The exciter is placed at
xex = 1.5 (with corresponding Reδ∗ (xex ) = 666.13) and has a width of wex = 0.09.
One can represent the excitation frequency F = 1.0 × 10−4 , by a straight line in the
(Reδ∗ , β0 )-plane (see Fig. 3.18). This line intersects the lower and upper branches
of the neutral curve at Reδ∗ = 728.7 (x = 1.79) and Reδ∗ = 1233.88 (x = 5.15),
respectively. Therefore, the exciter location is stable according to linear spatial sta-
bility theory. As the exciter is switched on impulsively at t = 0, a STWF pierces out
of the TS wave-packet. For moderate frequency cases, although the TS wave-packet
does not play any role in triggering transition of the flow, this STWF is responsible
for transition, whose feature is explained next by studying its dynamics.
In Fig. 5.2, the evolution of disturbance at y = 0.0057 (which is the tenth line
from the wall) is shown by plotting streamwise disturbance velocity ud as a function
of x for the indicated time instants. For this case, STWF is marked in the top frame at
t = 20. It is also noted in Fig. 5.2, that the amplitude of the TS wave packet changes
slightly with time. The TS wave-packet is essentially a progressive wave, which
attenuates due to the spatial stability properties (modified by nonlinear, nonparallel
5.3 Small Amplitude Disturbance at Moderate Frequency 281

0.02 (i) t = 20
-4
0.01
α1 = 0.002, y = 0.0057, F = 1.0×10 and xex = 1.5
#
ud

-0.01
(#) Spatio-temporal wave front
-0.02
0 20 40
x 60 80 100

0.02 (ii) t = 100


0.01
ud

-0.01

-0.02
0 20 40
x 60 80 100

0.02 (iii) t = 150


0.01
ud

-0.01

-0.02
0 20 40
x 60 80 100

0.02 (iv) t = 200


0.01
ud

-0.01

-0.02
0 20 40
x 60 80 100

0.4 (v) t = 260


0.2
(*) TS wave-packet
*
ud

-0.2

-0.4
0 20 40
x 60 80 100

Fig. 5.2 ud plotted as a function of x at y = 0.0057 and indicated times for α1 = 0.002, F =
1.0 × 10−4 and xex = 1.5. The spatio-temporal wave front in the top frame and the TS wave-packet
in the last frame is indicated by # and asterisk symbol, respectively
282 5 Dynamics of the Spatio-Temporal Wave-Front in 2D Framework

effects), and remains virtually localized in space. In contrast, the STWF propagates
in space and time, while growing significantly to a larger amplitude, as compared to
the TS wave-packet. In [23, 24], it has been shown that the origin of the first STWF
is due to a linear mechanism, whose onset can be captured by the solution of the
OSE by BCIM. At t = 20 and 100, the wave-front is significantly smaller than the
TS wave-packet. At t = 150, maximum amplitude of STWF is of the same order,
as compared to the TS wave-packet. By t = 200, the amplitude of the STWF is
approximately four times the amplitude of the TS wave-packet, and due to this large
amplitude of the wave-front, the nonlinear effects play a dominant role thereafter.
With passage of time, nonlinear effects distort the fore-aft symmetry of STWF due to
induction of unsteady separation on the wall below it (see the frame at t = 260). At
t = 260, the maximum peak-to-peak variation of the wave-front is of the order of the
free-stream speed. Thus, it is apparent that for this moderate frequency excitation
case, the disturbance evolution is more dominated by the STWF, rather than the
TS wave-packet, contrary to the classical view-point of transition. According to the
classical view, the flow transition caused by low amplitude, moderate frequency
inputs are solely dominated by TS wave-packets, which in this case is found to be
stationary and does not significantly grow with time, as well.

5.3.1 Growth and Speed of STWF

The maximum amplitude of the STWF for streamwise disturbance velocity is traced
in Fig. 5.3a, where the variation of maximum amplitude (udm ) with time is shown
for the case of Fig. 5.2. Based on the physical processes of growth for this distur-
bance, six different stages can be identified (marked in the figure). The origin of
STWF can be traced to the linear mechanism governed by the OSE [23]. The stage I
denotes the initial spectral redistribution, which causes the maximum amplitude of
STWF to decrease due to dispersion of the wave-front, resulting in the increase of
the band-width of wavenumbers of STWF. Similar initial decrease in amplitude of
the wave-packet, due to dispersion, was also noted for pulse excitation case in [3].
During stage II , one observes growth of STWF, where the amplitude of the front-
grows exponentially with time. One notes that despite the monochromatic temporal
excitation at the wall, STWF is not monochromatic, with respect to both space and
time. Strong nonlinear effects start to affect its growth from stage III onwards, where
the amplitude grows at a reduced rate, due to self and multi-modal interaction. In
the context of flow past a cylinder [25, 26], self-and multi-modal interactions were
accounted for following Stuart–Landau–Eckhaus equation. Higher order non-linear
effects, secondary and higher order instabilities due to mean flow distortion by wave-
induced stresses [27] become dominant in stage I V , resulting in a super-exponential
growth, as seen in Fig. 5.3a. Eventually nonlinear saturation of amplitude is achieved
in stage V , where udm is of the order of the free-stream speed. This subsequently
results in the creation of fully developed 2D turbulence in stage V I .
5.3 Small Amplitude Disturbance at Moderate Frequency 283

(a) α1 = 0.002, F = 1.0×10−4,


y = 0.0057L and xex = 1.5L

10-1

10-2
udm

I II III IV V VI
10-3

10-4
0 50 100 150 200 250 300
t

(b) x-location of udm of STWF plotted as a


function of t

100 Appearance of a kink

80

60
Line of best fit :
x

x = 0.337 t + 4.01
40

20

50 100 150 200 250 300


t

Fig. 5.3 a Variation and b location of maximum peak-to-peak amplitude of the STWF for ud at
y = 0.0057 with time α1 = 0.002, F = 1.0 × 10−4 and xex = 1.5. Six stages of the growth of the
wave front are identified and marked in frame (a)

In Fig. 5.3b, the streamwise location of udm is shown as a function of time for the
height, y = 0.0057. One notes that STWF propagates roughly at a speed of 0.337U∞ ,
up to a time when its growth is significantly affected by the distortion of the mean
flow in the leading order. This deviation appears as a kink, as marked in Fig. 5.3b by
an arrowhead. The respective time instant lies at the end of stage IV of the growth
of STWF.
284 5 Dynamics of the Spatio-Temporal Wave-Front in 2D Framework

(a) ψ-contour at t = 200 (b) ω-contour at t = 200


0.2 wave-front 0.2 wave-front

0.15 0.15
y

y
0.1 0.1

0.05 0.05

0 0
0 20 40 60 x 80 100 120 0 20 40 60 x 80 100 120

(c) ψ-contour at t = 240 (d) ω-contour at t = 240


0.2 Unsteady seperation bubbles on the wall 0.2

0.15 0.15
y

y
0.1 0.1

0.05 0.05

0 0
0 20 40 60 x 80 100 120 0 20 40 60 x 80 100 120

(e) ψ-contour at t = 278 (f) ω-contour at t = 278


0.2 Unsteady seperation bubbles on the wall 0.2
Edge of shear layer

0.15 0.15
y

0.1 0.1

0.05 0.05

0 0
0 20 40 60 x 80 100 120 0 20 40 60 x 80 100 120

Fig. 5.4 The stream-function ψ- (left column) and vorticity ω-contours (right column) are plotted
at indicated times for α1 = 0.002, F = 1.0 × 10−4 and xex = 1.5

On a closer look, it is revealed that in stages V and V I , unsteady separation bubbles


form on the wall. This feature is shown in Fig. 5.4, with the help of ψ- and ω-contours
at t = 200 (in stage III ), 240 (at the beginning of stage I V ) and 278 (in stage V I , after
the creation of a fully developed turbulence). Each bubble creates adverse pressure
gradient upstream of it, which in turn creates another bubble and this cascading effect
is responsible for the rapid widening of the perturbed zone and the eventual transi-
tion. However as bubbles keep appearing, these also convect downstream, creating a
dynamical equilibrium, whereby no further upstream penetration of disturbances take
place. This type of saturation of growth rate for the STWF due to nonlinear interac-
tion in stage V I , is shown to be universal, for all input amplitude cases considered
and is discussed later. The unsteady separation bubble formation leading to transi-
tion is similar to the bypass route shown by free stream convecting vortex discussed
in Chap. 4. Once the bubbles form, ud amplitude saturates to the level of free stream
speed, as noted in Fig. 5.3. From the vorticity contours in the right column, one notes
that these unsteady separations on the wall cause highly unsteady vortical eruptions,
which pierces through the shear layer (see Fig. 5.4f).
5.3 Small Amplitude Disturbance at Moderate Frequency 285

(a) t = 30 km 2km (b) t = 80 km 2km (c) t = 100


-2 10
1
km 2k
10 -2 m
10 10
0

-3
~3km 10-1
10
Ud (k)

-3
10
10-2

Ud (k)
-4

Ud (k)
10 -3
-4 10
10
-5 -4
10 10

10-5
10-5 10
-6
10-6 0
100 101
k 102 100 101 k 102 10 101
k 102

(d) t = 150 k (e) t = 170 k (f) t = 194 km


m 2km m 2km
10
1
101 101

100 100 100


3km 3km
2km
-1 10-1 10-1
10 4km
10-2
10 -2 3km
10-2
Ud (k)
Ud (k)

4km

Ud (k)
-3 -3
10 10
10-3 5km
-4
10-4 10-4 6km
10
10-5 10-5
-5
10
0 1 2 10-6 0 1 2 10-6 0
10 10 10 10 10 10 10 101 102
k k k
(g) t = 215 (h) t = 225 (i) t = 234
102 102 102
km km
1 1
10 10 101
2km
10
0
2km 10
0
0
10
3km 4km
10-1 10-1
5km 10
-1
Ud (k)

Ud (k)

Ud (k)

10-2 10-2
-3
6km -3 10-2
10 10
-3
10
-4
7km & 10
-4 10
higher
10-5 0 1 2 10-5 0 1 2 10-4 0 1 2
10 10 10 10 10 10 10 10 10
k k k

Fig. 5.5 The spatial Laplace transform of the STWF at y = 0.0057 shown for indicated times. The
excitation parameters are Ff = 10−4 , xex = 1.5 and α1 = 0.002

5.3.2 Spatial Spectrum and Scale Selection of STWF

Scale selection and spectrum during the downstream propagation of STWF is illus-
trated in Fig. 5.5, in a log-log plot showing spatial Fourier–Laplace transform of ud
as a function of wavenumber k, plotted at indicated times instants. Frames (a–c)
show that as STWF propagates downstream, the dominant wavenumber km comes
down with the narrowing of the band of wavenumbers around it. The value of km
decreases from 18.01 at t = 20 to 6.5 at t = 194. At initial stages of the evolution of
the STWF, one also notes that there is a presence of the first superharmonic of the
dominant wavenumber km , but its amplitude is more than two orders of magnitude
lower than km (see frames (a–c)). This effectively rules out any nonlinear interaction
between these modes.
With progress in time, the amplitude of the first superharmonic also displays
growth. During stage III of the evolution of STWF, strong nonlinearity starts to
interfere with the growth of STWF, as evident from frames (d–f). One notes that not
286 5 Dynamics of the Spatio-Temporal Wave-Front in 2D Framework

t = 250
102

100

-2 −3
10 k
E11 (k)

10-4

-6
10

0 1 2
10 10 10
k
Fig. 5.6 E11 corresponding to the STWF plotted as a function of k, for SBS excitation case with
α1 = 0.002, Ff = 10−4 and xex = 1.5

only the growth rate of the amplitude at km increases rapidly, but also more and more
higher harmonics appear and at t = 194, presence of as many as six super-harmonics
of km can be noted. The rapid growth and induction of higher harmonics in stage III
has two obvious effects: first, it reduces the growth rate of the primary mode at km ,
and secondly, it distort the mean flow by wave-induced stresses [27]. The distortion
of the mean flow initiates secondary and higher order instabilities of STWF, causing
rapid growth in its overall amplitude. Growth of amplitude of the higher harmonics
are also accompanied by increase in the band-width resulting in a continuous spec-
tra at t = 234. In Fig. 5.6, energy spectra is plotted at t = 250, i.e., after the com-
plete breakdown of STWF, and there exists a range of wavenumber band, where the
spectra falls off following a k −3 -type variation which is typical of 2D turbulence [1,
6, 12, 13]. Here, E11 is obtained as

E11 = |Ud (k)|2

where Ud (k) is the Fourier–Laplace transform of ud with respect to the streamwise


wavenumber k.

5.3.3 Wall-Normal Variation of the Disturbance Velocity

In Fig. 5.7a–f, ud is plotted as a function of y/δ ∗ , at indicated times and streamwise


locations. Here, δ ∗ is the local displacement thickness. The indicated streamwise
5.3 Small Amplitude Disturbance at Moderate Frequency 287

(a) Stage I
(b) Stage II
0.0006 0.0004
t = 24, x = 11.365
t = 64, x = 26.428
0.0005 t = 34, x = 15.146
t = 71.5, x = 28.754
t = 59, x = 24.247 0.0003
0.0004 t = 79, x = 31.051

0.0003 0.0002

d
0.0002

u
0.0001
d
u

0.0001

0 0

-0.0001
0 2 4 6 8 10 -0.0001
0 2 4 6 8 10
y/ y/

(c) Stage III (d) Stage III


0.02
0.001 t = 84, x = 32.5 t = 149, x = 53.644
t = 94, x = 35.5 t = 179, x = 63.356
0.0008 t = 99, x = 37.0 0.015 t = 194, x = 68.765

0.0006
0.01
d
u

d
0.0004

u
0.005
0.0002

0 0

-0.0002
0 2 4 6 8 0 2 4 6 8 10
y/ y/

(e) Stage IV (f) Stages V & VI


t = 239, x = 86.822 (Stage V)
t = 214, x = 75.773 0.8 t = 259, x = 97.087 (Stage V)
t = 224, x = 78.768 0.7 t = 264, x = 99.529 (Stage VI)
0.2 t = 234, x = 83.507 0.6 t = 284, x = 112.847 (Stage VI)
0.5
0.4
0.1 0.3
0.2
d

d
u

0.1
0 0
-0.1
-0.2
-0.1
0 2 4 6 8 10 0 2 4 6 8 10
y/ y/

Fig. 5.7 ud corresponding to the STWF plotted as a function of y/δ ∗ at indicated time and location.
Here, δ ∗ is the local displacement thickness

locations correspond to where ud is maximum. Corresponding growth stages are


also noted in the respective frames. Figure 5.7 shows that during stages I , II and III
of the STWF, the y-variation of ud roughly follows the variation of the real part of
the eigenfunction φ  (see Chap. 3). Figure 5.7a–d also shows that ud becomes zero at
y/δ ∗  2. The inner and outer maxima of the wall-normal profile of ud also appears at
identical locations of y/δ ∗ for stages I and II , which starts to vary slowly from stage
III onwards. One observes considerable variations in ud during stages I V , V and V I
of the growth of STWF. Particularly significant is the presence of considerable high
perturbation shear stress τd = ∂ud /∂y at the wall during these stages. Such a high
value of τd at the wall is a signature of the presence of high intensity fluctuations
inside the boundary layer. One also observes that during stage V I , and later part
of stage V (t = 239 onwards), ud is of the order of the free-stream speed from
y/δ ∗  0.08 (very close to the wall) to 2.5 (near the shear-layer edge). This kind
of very high streamwise disturbance velocity inside the shear-layer promotes rapid
transport of momentum and is a hallmark of turbulent flows.
288 5 Dynamics of the Spatio-Temporal Wave-Front in 2D Framework

5.4 Dynamics of STWF for High Amplitude


Wall-Excitation and Turbulent Spot Regeneration
Mechanism

Dynamics of the STWF is different, when the excitation amplitude is higher. For
high amplitude wall-excitation cases, unsteady separation bubbles are induced on
the wall and the scenario is different than that is described in the previous section,
for small excitation amplitude cases. Here, the amplitude of excitation cases with
α1 = 0.01 and 0.05 are discussed. In Fig. 5.8, variation of the maximum amplitude
udm of the front is shown. For α1 = 0.01, the time duration of stages I (shown in the
inset), II and III are seen to be very small as compared to the α1 = 0.002 case. The
stage I V (the stage of secondary and higher order instabilities due to the distortion
of the mean flow by wave-induced stresses) for α1 = 0.01, takes over very rapidly
as ud saturates nonlinearly. The STWF for the case of α1 = 0.05 does not exhibit
the six stages of evolution as mentioned earlier, and induces bypass transition right
at the beginning by forming unsteady separation bubbles on the plate surface, which
causes unsteady vortical eruptions. This feature is shown in Fig. 5.9 with the help of
stream function ψ- and vorticity ω-contours at t = 24 after the onset of excitation.

(a) udm at y=0.0057 plotted as a function of t for Ff=1.0×10−4 and xex=1.5

10
-1
α1 = 0.002
α1 = 0.01
α1 = 0.05

10-2
0.002
α1=0.01
udm

0.00198
0.00196
-3
10 0.00194
0.00192
0.0019
19 20 21
-4
10
50 100 150 200 250 300
t

Fig. 5.8 a Variation of the maximum peak-to-peak amplitude udm corresponding to the STWF at
y = 0.0057 with time shown for indicated values of α1 . In the inset shown in frame a, the variation
of udm for α1 = 0.01 is shown for initial times
5.4 Dynamics of STWF for High Amplitude Wall-Excitation … 289

(b) t = 24, contour


(a) t = 24, contour Unsteady vortical eruption
due to wave-front
Exciter
0.04 Seperation bubbles 0.04
Separation bubble
under TS wave-packet under wave-front
y

y
0.02 0.02

0 0
2 4 6 8 10 12 2 4 6 8 10 12
x x
Fig. 5.9 ψ- and ω-contours plotted at indicated times for α1 = 0.05 with F = 1.0 × 10−4 and
xex = 1.5

One notices formation of separation bubbles not only at the location of the front, but
also due to the TS wave-packet, for this amplitude of excitation case.
In the previous section, we demonstrated that turbulence is a deterministic con-
sequence of a single STWF convecting downstream, i.e., once STWF is created,
its growth and nonlinear saturation leads to small scale unsteady separations on the
wall, which culminate into intermittent nature of the wall bounded flow. This could
lead one to conclude that this turbulence creation is a buffeting problem, i.e., sus-
tenance and creation of turbulence would require subsequent induction of another
STWF. It is relevant to ask, whether this latter process is due to intrinsic or extrinsic
dynamics. This is investigated here for α1 = 0.01 and 0.05. In Figs. 5.10 and 5.11,
the streamwise disturbance velocity component ud is shown plotted as a function
of x at y = 0.0057 for α1 = 0.01 and 0.05, respectively. The exciter is located at
xex = 1.5 (Reδ∗ (xex ) = 666.67), similar to α1 = 0.002 case discussed in the previous
section. In the top frame of Fig. 5.10 at t = 94, a single STWF is marked as A. In the
following frame at t = 164, nonlinear saturation of this STWF is noted. At t = 164
unsteady separations are created on the wall, underneath this front during this stage
of disturbance evolution, which not only widens the disturbance packet, but also
makes the flow intermittent.
The most important aspect of this propagating STWF during this stage is the
induction of another STWF, marked as B in the frame. The fact that this front is
created upstream of A is very significant, as this event negates the assumption used
in parabolized stability equation (PSE) approach, often used to study flow instability.
As time progresses, the amplitude of the front B also grows rapidly and its nonlinear
distortion is more rapid due to the presence of A, while the gap between the two
STWFs reduces with time, as shown in the frames at t = 194 and 244. Also in the
frame at t = 244, the induction of another STWF is noted, marked as C. In the
following frame at t = 324, one notices amalgamation of the STWFs marked as
A, B and C, while two new STWFs, marked as D and E, which are formed upstream
of the leading packet. These trailing packets grow with time, as the leading part of
the first packet leaves the computational domain. By t = 344, as shown in Fig. 5.10f,
the trailing fronts D and E grow and become part of the leading packet which finally
290 5 Dynamics of the Spatio-Temporal Wave-Front in 2D Framework

(a) t = 94
A
ud
0

0 20 40 x 60
x
80 100
(b) t = 164 A
B
ud

0 20 40 x 60
x
80 100
(c) t = 194 A
B
ud

0 20 40 x 60
x
80 100
(d) t = 244 B A
C
ud

0 20 40 x 60
x
80 100
(e) t = 324 C+B+A
E D
ud

0 20 40 x 60
x
80 100
(f) t = 344 Continuous perturbation
E D
ud

1
= 0.01, F = 1 10-4, y = 0.0057 and xex = 1.5
0 20 40 x 60
x
80 100
(g) t = 394
E D
F
ud

0 20 40 x 60
x
80 100

Fig. 5.10 ud corresponding to the STWF plotted as a function of x at y = 0.0057 and indicated
times for α1 = 0.01, F = 1.0 × 10−4 and xex = 0.5 in frames (a) to (g)

combines by t = 394. The upstream point of the turbulent zone remains in the vicinity
of x  40 for this amplitude case at t = 194, 244, 324. However, this is seen to move
forward up to x  60 at t = 394. Another wave-packet F is seen to slowly emerge
beyond t  360 (see Fig. 5.10g), which also eventually grows (results not shown
here). This process will go on, and will make the flow intermittent in nature. Similar
picture is also noted for even higher amplitude of excitation shown in Fig. 5.11, for
α1 = 0.05. Thus, the amplitude of excitation causes earlier intermittency of the flow,
which also penetrates upstream towards the TS wave-packet. Interestingly the TS
wave-packet, for the cases shown in Figs. 5.10 and 5.11, remains rooted at the same
spatial location, without affecting the transition process. These features show that a
time-harmonic excitation, at a fixed location, gives rise to self-regenerating sequence
of STWFs, which are responsible for the generation of turbulent flow field.
For all the amplitude cases of Ff = 10−4 and xex = 1.5, the TS wave-packet does
not play any role in determining the transition process, that is solely dictated by the
growth and subsequent breakdown of the STWFs. However, the TS wave-packet
also suffers marginal instability for this moderate frequency of excitation case. This
5.4 Dynamics of STWF for High Amplitude Wall-Excitation … 291

0.5 (a) t = 34 1 = 0.05, F = 1.0 10-4, y = 0.0057 and xex = 1.5


A
ud
0

-0.5
0 20 40 x 60 80 100
0.5 (b) t = 184 B A
D C
ud

-0.5
0 20 40 x 60 80 100
0.5 (c) t = 239 A+B+C+D
F E
ud

-0.5
0 20 40 x 60 80 100
0.5 (d) t = 254 A+B+C+D
G F E
ud

-0.5
0 20 40 x 60 80 100
0.5 (e) t = 289 B+C+D+E+F+G
H
ud

-0.5
0 20 40 x 60 80 100
0.5 (f) t = 305 B+C+D+E+F+G
H
ud

= 0.05, Ff = 1.0 10-4, y = 0.0057 and xex = 1.5


-0.5
0
1
20 40 x 60 80 100
0.5 (g) t = 555
L K
ud

-0.5
0 20 40 x 60 80 100

Fig. 5.11 ud corresponding to the STWF plotted as a function of x at y = 0.0057 and indicated
times for α1 = 0.05, F = 1.0 × 10−4 and xex = 0.5 in frames (a) to (g)

(a) t1 (b) t1
0.12
0.02 = 0.0005
= 0.01
= 0.001
0.1 = 0.05
= 0.002
ud

0.015 = 0.004
ud

0.08
0.01
0.06
0.005
0.04
0
0 50 100 150 t 200 250 300 0 50 100 150 t 200 250 300

Fig. 5.12 Maximum amplitude of ud at y = 0.0057 corresponding to the TS wave-packet plotted


as a function of time for indicated values of α1 with F = 1.0 × 10−4 and xex = 1.5
292 5 Dynamics of the Spatio-Temporal Wave-Front in 2D Framework

feature is noted in Fig. 5.12, where the TS wave-packet amplitude is plotted as a


function of time for indicated excitation amplitude cases. For α1 = 0.002, 0.004
and 0.01, two growth stages of the TS wave-packet are noted, i.e., a primary stage
(which ends at t = t1 as noted in Fig. 5.12) and a secondary stage which brings the TS
wave-packet to a higher saturated level of amplitude. The onset time of the secondary
growth stage is dependent on α1 . However, no such secondary growth is observed for
α1 = 0.05 case, as the TS wave-packet is nonlinearly distorted displaying a non-zero
mean, right from its onset, due to the induction of the unsteady separation bubbles
on the wall, as shown in Fig. 5.12b.

5.5 Low Frequency Excitation: Interaction of Near-Field


Solution and Primary STWF

For low frequency excitation cases, the evolution of disturbances follow a signifi-
cantly different route from moderate to high-frequency wall-excitation cases, dis-
cussed in previous sections. For latter cases, disturbances evolve such that the STWF
and the TS wave-packet remain distinct. The STWF grows in space and time whereas
the TS wave-packet remains, more or less, localized in space (close to the exciter
location) displaying no spatio-temporal evolution at later times. In contrast, when
the excitation frequency is lowered, the near-field solution evolves in both space and
time. Here, the near-field solution, refers to the part of the solution which corresponds
to the TS wave-packet.
For low-frequency excitation cases, the near-field solution evolves in space and
time, spawning additional subsequent wave-fronts. These fronts, interact with the
primary STWF at later stages, showing an entirely different mechanism of transi-
tion. In Fig. 5.13, ud is plotted at y = 0.0057 for α1 = 0.002 and Ff = 0.5 × 10−4 ,
with the exciter located at xex = 1.5. Initially, the primary STWF grows inducing
elongated fronts upstream, as noted at t = 230 and 250. At t = 250, the nonlinear
saturation of the primary STWF is noted. The incipient turbulent packet is marked as
A, whereas the perturbation wave-packet, which is induced upstream of it, is marked
as B. At later times, B also grows and saturate nonlinearly, which in turn induce the
disturbance wave-packet C upstream of it, at t = 290. At t = 340, A convects out of
the computational domain with C saturating nonlinearly. These type of spot regen-
eration mechanism, due to the nonlinear saturation of the primary STWF have also
been discussed in the previous section. Between t = 250 and t = 300, the near-field
solution evolves rapidly, which amplifies and extends downstream. The upstream
part of the near-field solution is rooted at the location of the exciter. It first shows
a streamwise bias, which amplifies and evolves into a sharp wave-front, as noted at
t = 290. This results in the subsequent emergence of a secondary wave-front (marked
by D), which is seen to convect downstream.
Nonlinear saturation of the amplitude of this secondary wave-front happens
around t = 310, for this case. Subsequently, it attempts to detach from the
5.5 Low Frequency Excitation: Interaction of Near-Field … 293

(a) t = 190.0
0.04
TS wave-packet
0.02 Wave-front

ud
0

-0.02

-0.04
0 20 40 x 60 80 100
(b) t = 230.0 Upstream induced Wave-front
0.1 perturbation
TS wave-packet
ud

-0.1
0 20 40 x 60 80 100
(c) t = 250.0
0.4
B A
0.2
ud

0
-0.2
-0.4
0 20 40 x 60 80 100
(d) t = 290.0
0.4
C B A
0.2
ud

0
-0.2
-0.4
0 20 40 x 60 80 100
(e) t = 310.0 A
C B
D
ud

0 20 40 x 60 80 100
(f) t = 340.0
F D C B
E
ud

0 20 40 x 60 80 100
(g) t = 360.0
E D C B
F
ud

0 20 40 x 60 80 100
(h) t = 370.0
C B
ud

0 20 40 x 60 80 100
(i) t = 400.0
Turbulent zone
ud

0 20 40 x 60 80 100

Fig. 5.13 ud plotted as a function of x at y = 0.0057 and indicated times for α1 = 0.002, F =
0.5 × 10−4 and xex = 1.5
294 5 Dynamics of the Spatio-Temporal Wave-Front in 2D Framework

parent near-field solution, as it convects and elongates downstream. However, due to


the induction of new wave-packets in the intermediate zone, in between the front D
and the near-field solution, these are never fully separated from each other. One such
induced intermediate wave-packet is shown marked at t = 340 as E. The wave-packet
E, also grows rapidly, and saturates nonlinearly, while elongating, and eventually
undergoing transition, and subsequently merging with D. The nonlinearly saturated
wave-packet E, also in turn, induces new perturbation wave-packets (marked as F
at t = 360) in the intermediate zone, between it and the near-field solution. These
newer wave-packets also undergo similar cycle, as its earlier predecessors. The train
of disturbance wave-packets, generated due to the spatio-temporal evolution of the
near-field solution, finally merges with the rear part of the disturbance packet C
(generated due to the cascading effects of the instability of the primary STWF). The
difference in length scale between these disturbances are clearly visible at t = 360.
At later times, it is difficult to distinguish between the perturbations generated from
these two different sources, and a continuous highly perturbed zone is obtained,
right from the near-field solution. In the present mechanism, more and more new
wave-packets keep on appearing in the front part of the TS wave-packet and sustains
the extent of the highly unsteady zone, which spans from x  30 to the end of the
computational domain. This route of disturbance evolution is completely different
from the moderate and high-frequency excitation cases.
Figure 5.14 shows the variation of the maximum peak-to-peak amplitude, udm
(corresponding to the near-field solution) for y = 0.0057 is plotted, as a function
of time, for the case shown in Fig. 5.13. For this case, four distinct stages of the
disturbance evolution can be noted. The first zone marked as S1 is the zone, where the
initial evolution of the near-field solution takes place, with its amplitude increasing
by linear mechanism. During S2 , the amplitude attains an almost constant value.
Nonlinear stage of growth S3 is noted subsequently, which brings the amplitude of
the near-field solution to the level of the free-stream velocity, leading to continuous
spawning of newer secondary wave-fronts, as discussed with respect to Fig. 5.13.
In Fig. 5.15, the spatial Laplace transform of ud corresponding to the near-field
solution (2.5 ≤ x ≤ 50) is plotted for the case of Fig. 5.13, as it evolves spatio-
temporally, initiating spawning secondary wave-fronts. This plot highlights the evo-
lution of different peaks and sub-peaks and the nonlinear instability mechanisms.
Existence of multiple peaks located at the wavenumbers k10 to k50 , are noted in
Fig. 5.15a. The peaks k20 to k50 are higher harmonics of the fundamental peak k10 .
One notes the existence of lower and higher side-band wavenumber components in
the neighbourhood of each peak marked in Fig. 5.15a. With progress in time, the
principal peak at k10 is seen to shift towards lower wavenumber, with additional sub-
peaks at higher wavenumbers appearing just adjacent to it, as marked in Fig. 5.15b.
These peaks are further traced in Fig. 5.16, where the amplitude of k10 and its sub-
peaks are plotted as functions of time. One notes form Fig. 5.16, that with increase in
time from t = 250 to 290, the amplitude of Ud (k) at k10 increases five times, while
the value of k10 decreases from 13.87 to 10.20. The additional sub-peaks k11 , k12 ,
k13 and k14 appear at t  270, which grows till t  283. The sub-peaks at higher
wavenumbers, such as k13 and k14 , grow at a significantly higher rate. For example,
5.5 Low Frequency Excitation: Interaction of Near-Field … 295

S3

10-1
S1 S2
udm

10-2

Initiation of continuous Spawning


of secondary wave-front
-3
10

0 50 100 150 200 250 300


t

Fig. 5.14 Variation of the maximum peak-to-peak amplitude of ud corresponding to the TS wave-
packet at y = 0.0057 plotted as a function of time for Ff = 0.5 × 10−4 , α1 = 0.002 and xex = 1.5

the sub-peak at k13 grows more than 25 times from its original value at t = 270 to
284, whereas the sub-peak at k14 grows more than 3000 times during t = 270 to
286. From t  285 to 310, the sub-peaks at k12 , k13 and k14 display almost constant
amplitude, while the amplitude of the sub-peak at k11 , first decreases mildly from
t = 286 to 296, followed by an increase up to t = 310, when it ceases to exist. The
amplitude and wavenumber of the peak at k11 register rapid drop at t = 294, which
continues to decrease further up to t = 310. These events cause the energy to trans-
fer to higher wavenumber sub-peaks, as these appear and amplify subsequently. The
amplitude of such a sub-peak at k17 (marked in Fig. 5.15b) is traced in Fig. 5.16a.
The sub-peak at k17 appears at t  282 and continues to grow up to t  310. In
Fig. 5.15c at t = 309, one notes the sub-peaks located from k11 to k17 , to have ampli-
tudes, which are almost similar to that of the principal peak at k10 . Appearance of
additional sub-peaks adjacent to k17 is also noted in this figure. This causes all the
main peaks shown in Fig. 5.15c to connect with each other via an array of increasing
sub-peaks for t > 310, causing the evolution of a continuous spectrum originating
from k = k10 . This initiates the spawning of secondary wave-fronts, as noted earlier
in Fig. 5.13, for the low-frequency excitation case.
296 5 Dynamics of the Spatio-Temporal Wave-Front in 2D Framework

α1= 0.002, Ff = 0.5×10− 4


102 (a) t = 250
k10 Main
101 Peaks
lower Higher
100 side-band side-band k20 Higher
Ud (k)

lower side-band
10-1 side-band k30
-2 k40
10 k50
-3
10
7 20 k 40 60 80

102 (b) t = 285 k k12


k 11
10 k13
101 k17 k20
0 k30
10 k40
Ud (k)

k50 k
-1 60
10
-2 sub-peaks
10
10-3
7 20 k 40 60 80

102 (c) t = 309


k17
101
100
Ud (k)

10-1 k10
-2
10
10-3
7 20 k 40 60 80

Fig. 5.15 Spatial Laplace transform of ud within the range 2.5 ≤ x ≤ 50 at y = 0.0057 plotted at
indicated times with Ff = 0.5 × 10−4 and α1 = 0.002. The exciter is located at xex = 1.5

5.5.1 Low Frequency Excitation: Dominant Role


of the Near-Field Solution

Another route of disturbance evolution corresponding to the low-frequency excitation


is noted, when amplitude is marginally increased, or the exciter location is changed.
This is elaborated here for the case, when following parameters for time-harmonic
wall-excitation is used: Ff = 0.5 × 10−4 , xex = 1.5 and α1 = 0.003. Therefore, in
comparison to the case described in Figs. 5.13, 5.14, 5.15 and 5.16, the current
case has slightly higher amplitude of excitation, while the frequency and the exciter
5.5 Low Frequency Excitation: Interaction of Near-Field … 297

30 (a) α1 = 0.002, Ff = 0.5×10


−4

k10 = 10.20
25
k10 = 8.86
20
Ud (k)

k10 = 10.03
15 k10 = 8.50 k17 = 13.65
k10 = 13.87
k10 = 13.66
10
k10 = 6.66
5
k17 = 13.86 k10 = 6.59

0
240 260 280 300 320 340
t

20 (b) 20 (d)
ceases to exist
15 15
Ud (k12 )
Ud (k11)

k11 = 13.24
k11 = 11.86
k12 = 13.76
k11 = 12.23 k12 = 10.76
10 10 k12 = 13.03
k11 = 10.65
k12 = 10.34
5 5

0 0
260 270 280 290 300 310 320 260 270 280 290 300 310 320
t t
20 (c) 20 (e)

15 15
Ud (k13)

Ud (k14)

k13 = 12.92 k14 = 13.29


10 k13 = 14.29 k13 = 11.28 10
k14 = 14.55 k14 = 11.92
k13 = 11.81

5 5
negligible
existence
0 0
260 280 300 320 260 270 280 290 300 310 320
t t

Fig. 5.16 Time variation of the amplitude of the main peak k10 and sub-peaks k11 to k17 as marked
in Fig. 5.14 are plotted

location is identical. The evolution of disturbance for this case is shown in Fig. 5.17,
where ud for y = 0.0057 is plotted, at indicated time instants. For this case, the
amplitude of the near-field solution grows continuously, right from the onset of
excitation. The primary STWF is seen initially to come out of the near-field solution,
as shown at t = 60, 90 and 110. At t = 90, the near-field solution is noted to amplify
and issue a secondary wave-front, with a sharp leading edge, similar to the case
described before. The secondary wave-front convects downstream faster than the
298 5 Dynamics of the Spatio-Temporal Wave-Front in 2D Framework

(a) t = 060.0
0.05 Wave-front

ud
0
= 0.003, xex = 1.5 and Ff = 0.5 10
-0.05 Exciter
0 20 40 x 60 80 100
(b) t = 090.0
0.2 Wave-front
ud

-0.2
0 20 40 x 60 80 100
(c) t = 110.0 Wave-front
0.2
ud

0
-0.2

0 20 40 x 60 80 100
(d) t = 120.0
0.5
ud

0
= 0.003, xex = 1.5 and Ff = 0.5 10
-0.5 0 20 40 x 60 80 100

0.5
(e) t = 150.0 Highly Unsteady
ud

-0.5 0 20 40 x 60 80 100

0.5
(f) t = 170.0 Highly Unsteady
ud

-0.5 0 20 40 x 60 80 100

0.5
(g) t = 200.0 Highly Unsteady
ud

-0.5 0 20 40 x 60 80 100

0.5
(h) t = 250.0 Highly Unsteady
ud

-0.5 0 20 40 x 60 80 100

Fig. 5.17 ud at y = 0.0057 plotted as a function of x at indicated times for α1 = 0.003, xex = 1.5
and Ff = 0.5 × 10−4
5.5 Low Frequency Excitation: Interaction of Near-Field … 299

(a) t = 060.0
0.05 Wave-front

ud
0
= 0.003, xex = 1.5 and Ff = 0.5 10
-0.05 Exciter
0 20 40 x 60 80 100
(b) t = 090.0
0.2 Wave-front
ud

-0.2
0 20 40 x 60 80 100
(c) t = 110.0 Wave-front
0.2
ud

0
-0.2

0 20 40 x 60 80 100
(d) t = 120.0
0.5
ud

0
= 0.003, xex = 1.5 and Ff = 0.5 10
-0.5 0 20 40 x 60 80 100

0.5
(e) t = 150.0 Highly Unsteady
ud

-0.5 0 20 40 x 60 80 100

0.5
(f) t = 170.0 Highly Unsteady
ud

-0.5 0 20 40 x 60 80 100

0.5
(g) t = 200.0 Highly Unsteady
ud

-0.5 0 20 40 x 60 80 100

0.5
(h) t = 250.0 Highly Unsteady
ud

-0.5 0 20 40 x 60 80 100

Fig. 5.18 ud at y = 0.0057 plotted as a function of x at indicated times for α1 = 0.003, xex = 1.75
and Ff = 0.5 × 10−4
300 5 Dynamics of the Spatio-Temporal Wave-Front in 2D Framework

corresponding primary STWF, and by t = 125, overtakes it. The secondary wave-
front grows very rapidly, and as a consequence, one observes a highly unsteady
and chaotic zone to form, out of the secondary wave-front, at t = 150. This zone is
marked by a horizontal arrow in subsequent frames of Fig. 5.17. The front of this
zone rapidly progresses downstream, while the upstream end remains fixed at x  40.
The rapidly advancing front of this zone, reaches the outflow of the computational
domain xout by t = 250 (see Fig. 5.17). At this time instant, the entire zone from
x  40 to xout is unsteady and chaotic, displaying features of 2D inhomogeneous
turbulent flow. However, the perturbation from the location of the exciter to x 
40 is composed of traveling waves with a definitive fundamental wavenumber. A
similar sequence of disturbance evolution is also noted, when the exciter location is
changed to xex = 1.75 from the previous location of 1.5, while keeping the excitation
amplitude at α1 = 0.003. This is shown in Fig. 5.18. Figure 5.18 shows that for this
case also, the primary STWF does not play any role in determining the dynamics
of the disturbance evolution, while the spatio-temporal evolution of the near-field
solution, and subsequent continuous spawning of wave-fronts are responsible for
eventual flow transition.

5.6 Dynamics of the STWF for Excited Flow Over


an Airfoil

To show the relevance of STWF, even for cases with varying pressure gradient, next
we study flow past the SHM1 airfoil, by solving NSE. The Reynolds number based on
the chord length is Re = 10.3 × 106 , which corresponds to the cruise condition for
Honda-jet aircraft, for which the airfoil is designed. The chosen airfoil is designed as
a natural laminar flow (NLF) airfoil, which keeps the flow laminar over a wider range
of streamwise stretch. More details of this airfoil and characteristics of flows over it
can be found in [20]. The design prediction of the transition location, for both the
top and bottom surfaces of this airfoil was performed based on the spatial growth of
the TS waves. Here, we compute the flow over SHM1 airfoil without any models for
either transition or turbulence. Two-dimensional NSE in (ψ, ω)-formulation using
orthogonal coordinate is solved here, similar to the previous flat-plate simulation
cases. However, for the present simulations the scale-factors are functions of both
azimuthal and wall-normal coordinates.
Figure 5.19 shows the stream function plot for the simulated results at zero degree
angle of attack. This figure shows that at zero angle of attack, the flow will accelerate
on the top surface almost up to the maximum thickness position (located aft of
0.30c). For Re = 10.3 × 106 , the boundary layer formed over this airfoil is very thin,
indicating flow separation over the airfoil at very small scales. To note separation
bubbles, zoomed view of the flow-field near the trailing edge portion is shown in the
bottom two frames of Fig. 5.19. These frames show the formation and downstream
convection of micro-bubbles, which are physical in origin. Satisfactorily capturing
these micro-bubbles are possible, because of the use of high-accuracy DRP schemes,
while solving the unsteady NSE. The presence of unsteady separation bubble on the
5.6 Dynamics of the STWF for Excited Flow Over an Airfoil 301

t = 4.50

t = 1.50

0.02

0.01
y

−0.01
0.85 0.9 x 0.95 1

t = 3.50

0.02

0.01
y

−0.01
0.85 0.9 x 0.95 1

Fig. 5.19 Stream function contours for flow past SHM1 airfoil shown at the indicated time instants
for Re = 10.3 × 106 , when the airfoil is kept at zero angle of attack (AOA). The top frame shows
the flow field around the full airfoil, while the bottom frames show the zoomed view of the flow
field near the trailing edge, showing the presence of small bubbles indicative of bypass transition

wall is characteristic of bypass transition, as also noted previously while describing


high-amplitude excitation cases.
In [17], Falkner–Skan pressure gradient parameter, m = Uxe ∂U ∂x
e
is plotted for this
flow field. For steady boundary layers when m goes below −0.0904, steady flow
separation is indicated. However, unsteady flows can sustain a higher adverse pressure
gradient (APG), without showing separation. It was noted for the flow past SHM1
airfoil [17], that it exhibits significantly larger swings for m, starting from x/c = 0.60,
while unsteady separations are only noted after x/c = 0.75 (see also Fig. 5.20). Flow-
field shown in Fig. 5.19, indicates that the flow undergoes bypass transition near the
trailing edge of the airfoil, where it encounters high APG. Such bypass transition is
not due to any explicit excitation on the airfoil surface. In contrast, ZPG boundary
layer discussed in the previous sections required definitive excitation to trigger flow-
transition. For flows experiencing an APG, flow is very susceptible to numerical
302 5 Dynamics of the Spatio-Temporal Wave-Front in 2D Framework

(i) t = 4.50
Top surface
Bottom surface
0.001

δ

0.0005

0 0.2 0.4 0.6


x

(ii) t = 4.50
Top surface
Bottom surface
0.2
m

−0.2 0.2 0.4 0.6


x
Fig. 5.20 Variation of azimuthal component of velocity (u) at indicated height of 1.242 × 10−6
from the top surface of the SHM1 airfoil at the indicated time instants

disturbances related to round-off error, and this along with other sources of numerical
errors, can trigger disturbance growth and eventual transition, as seen in Fig. 5.19,
near the trailing edge of the airfoil.
Figure 5.21 shows the sequence by which the top surface of the airfoil experi-
ences bypass transition. This airfoil has concavity on the top and bottom surfaces
near the trailing edge which manifests itself in flow instability, near the trailing edge
on both surfaces resulting in the local unsteady separation. The region over which
the disturbances are noted is seen to travel upstream. The upstream travel of distur-
bances generate new disturbance packets upstream, in exactly the same sequence,
bypass transition is noted to take place for a ZPG boundary layer in Sect. 5.4 by
high amplitude localized harmonic excitation at moderate to high frequencies. By
t = 4.5, the disturbed region on the top surface is noted to extend from x/c = 0.6 to
the trailing edge. These observations also corroborate that bypass transition is caused
by upstream propagating disturbances, as illustrated in Sect. 5.4 with respect to ZPG
boundary layer.
However, the transition location is far aft of what is experimentally observed
[17]. The reason for this difference is due to the fact that in computed flows, a
perfectly smooth geometry placed in a uniform flow is considered and the transition
is caused by the numerical disturbances acting, as the seed for the APG region over the
airfoil surface. Figure 5.22 shows the case, when SBS harmonic excitation is applied
5.6 Dynamics of the STWF for Excited Flow Over an Airfoil 303

0.04 t = 1.25
0.02
Bv
0
u

−0.02
−0.04 0 0.2 0.4 0.6 0.8
x
0.04 t = 1.28
0.02
Bv
0
u

−0.02
−0.04 0 0.2 0.4 0.6 0.8
x
0.04 t = 1.30
Bv
0.02
0
u

−0.02
−0.04 0 0.2 0.4 0.6 0.8
x
0.04 t = 1.50
0.02
0
u

−0.02
−0.04 0 0.2 0.4 0.6 0.8
x
0.04 t = 3.00
0.02
0
u

−0.02
−0.04 0 0.2 0.4 0.6 0.8
x
0.04 t = 4.50
0.02
0
u

−0.02
−0.04 0 0.2 0.4 0.6 0.8
x

Fig. 5.21 Variation of azimuthal component of velocity (u) at indicated height from the top surface
SHM1 airfoil at the indicated time instants

with a nondimensional frequency Ff = 1.11441 × 10−5 , on the top surface of the


airfoil. The location of the exciter is indicated by an arrowhead in Fig. 5.22. A little
before t = 1.75, a STWF is noted to be created downstream of the harmonic exciter.
With time, this grows and convects downstream. A little after t = 1.95, the STWF
merges with main disturbance packet created due to the bypass transition without
the imposed excitation. This enlarges the region over which transition is noted to be
present. The frame at t = 6.50 shows that the transitional location point moves to
x/c = 0.4, as also noted experimentally [17]. Therefore, to match the experimental
304 5 Dynamics of the Spatio-Temporal Wave-Front in 2D Framework

(i) t = 1.75
u 0.02

−0.02
0 0.2 0.4 x 0.6 0.8 1
(ii) t = 1.85
0.02
u

−0.02
0 0.2 0.4 x 0.6 0.8 1
(iii) t = 1.95
0.02
u

−0.02
0 0.2 0.4 x 0.6 0.8 1
(iv) t = 6.50
0.02
u

−0.02
0 0.2 0.4 x 0.6 0.8 1

Fig. 5.22 Variation of azimuthal component of velocity (u) at a height of 1.242 × 10−6 from the
top surface of the SHM1 airfoil at the indicated time instants. Wall excitation corresponds to an
SBS frequency of F = 1.11441 × 10−5 , with an amplitude of 0.001

drag coefficients, one must know the level of background disturbance present and
simulate the flow accordingly [21, 22].

References

1. Batchelor, G. K. (1969). Computation of the energy spectrum in homogeneous two-dimensional


decaying turbulence. Physics of Fluids,12, 233–239 [suppl. II].
2. Bhaumik, S. (2013). Direct Numerical Simulation of Inhomogeneous Transitional and Turbu-
lent Flows. Ph. D. Thesis, I. I. T. Kanpur, INDIA.
References 305

3. Bhaumik, S., & Sengupta, T. K. (2017). Impulse response and spatio-temporal wave-packets:
the common feature of rogue waves, tsunami and transition to turbulence. Physics of Fluids,
29, 124103.
4. Brillouin, L. (1960). Wave Propagation and Group Velocity. New York: Academic Press.
5. Chomaz, J.-M (2005). Global instabilities in spatially developing flows: non-normality and
nonlinearity. Annual Reviews Fluid Mechanics, 37, 357–392.
6. Davidson, P. A. (2004). Turbulence: An Introduction for Scientists and Engineers. UK: Oxford
University Press.
7. Drazin, P. G., & Reid, W. H. (1981). Hydrodynamic Stability. UK: Cambridge University Press.
8. Fasel, H., & Konzelmann, U. (1990). Non-parallel stability of a flat plate boundary layer using
the complete Navier–Stokes equation. Journal of Fluid Mechanics, 221, 331–347.
9. Fasel, H. F., Rist, U., & Konzelmann, U. (1990). Numerical investigation of the three-
dimensional development in boundary- layer transition. AIAA Journal, 28(1), 29–37.
10. Gaster, M. & Grant, I. (1975). An experimental investigation of the formation and development
of a wave packet in a laminar boundary layer. Proceedings of the Royal Society of London A:
Mathematical, Physical and Engineering Sciences, 347(1649), 253–269.
11. Huerre, P., & Monkewitz, P. A. (1985). Absolute and convective instabilities in free shear
layers. Journal of Fluid Mechanics, 159, 151.
12. Kraichnan, R. H. (1967). Inertial ranges in two-dimensional turbulence. Physics of Fluids, 67,
1417–1423.
13. Kraichnan, R., & Montgomery, D. (1980). Two-dimensional turbulence. Reports in Progress
in Physics, 43, 547–619.
14. Medeiros, M. A. F., & Gaster, M. (1999). The production of subharmonic waves in the nonlinear
evolution of wavepackets in boundary layers. Journal of Fluid Mechanics, 399, 301–318.
15. Rajpoot, M. K., Sengupta, T. K., & Dutt, P. K. (2010). Optimal time advancing dispersion
relation preserving schemes. Journal of Computational Physics, 229(10), 3623–3651.
16. Schubauer, G. B., & Skramstad, H. K. (1947). Laminar boundary layer oscillations and the
stability of laminar flow. Journal of the Aeronautical Science, 14(2), 69–78.
17. Sengupta, T. K. (2012). Instabilities of Flows and Transition to Turbulence. Taylor & Francis
Group, Florida, USA: CRC Press.
18. Sengupta, T. K., & Bhaumik, S. (2011). Onset of turbulence from the receptivity stage of fluid
flows. Physical Review Letters, 154501, 1–5.
19. Sengupta, T. K., Bhaumik, S., & Bhumkar, Y. (2012). Direct numerical simulation of two-
dimensional wall-bounded turbulent flows from receptivity stage. Physical Review E, 85(2),
026308.
20. Sengupta, T. K. & Bhumkar, Y. G. (2013). Direct numerical simulation of transition over a
NLF aerofoil: Methods and validation. Frontiers in Aerospace Engineering (FAE) 2(1).
21. Sengupta, T. K., Das, D., Mohanamuraly, P., Suman, V. K., & Biswas, A. (2009). Modelling
free stream turbulence based on wind tunnel and flight data for instability studies. International
Journal of Emerging Multidisciplinary Fluid Science, 1(3), 181–201.
22. Sengupta, T. K., De, S., & Gupta, K. (2001). Effect of free-stream turbulence on flow over
airfoil at high incidences. Journal of Fluids Structures, 15(5), 671–690.
23. Sengupta, T. K., Rao, A. K., & Venkatasubbaiah, K. (2006). Spatiotemporal growing wave
fronts in spatially stable boundary layers. Physical Review Letters, 96(22), 224504.
24. Sengupta, T. K., Rao, A. K., & Venkatasubbaiah, K. (2006). Spatiotemporal growth of dis-
turbances in a boundary layer and energy based receptivity analysis. Physics of Fluids, 18,
094101.
25. Sengupta, T. K., Singh, N., & Suman, V. K. (2010). Dynamical system approach to instability
of flow past a circular cylinder. Journal of Fluid Mechanics, 656, 82–115.
26. Sengupta, T. K., Singh, N., & Vijay, V. V. S. N. (2011). Universal instability modes in internal
and external flows. Computers & Fluids, 40, 221–235.
27. Tollmien, W. (1931). The Production of Turbulence. NACA Report-TM 609.
Chapter 6
3D Routes of Transition to Turbulence
by STWF

6.1 Introduction

In Chap. 5, we have discussed the dynamics of the STWF for 2D transition. We have
also shown the inadequacy of the linear spatial instability studies in determining the
evolution of disturbances. For monochromatic wall-excitation, the spatio-temporal
evolution of disturbance was noted [2] to depend on various factors like (a) excitation
frequency, (b) amplitude, (c) exciter location and its width and (d) nature of excitation
onset. In the present chapter, we would discuss about the 3D evolution of disturbances
and the associated process of transition to turbulence. We first start with the governing
equations, followed by numerical methods, problem definition and a brief description
of boundary conditions. We have chosen the velocity-vorticity formulation [4] of the
incompressible NSE for its inherent accuracy to compute the 3D excitation of a
nominally 2D ZPG boundary layer. Growth and evolution of disturbances, nature of
vortical structures in the transitional and turbulent zones, and integral properties of
the turbulent boundary layer (in terms of displacement and momentum thickness,
shape factor and skin friction coefficient) are described subsequently.

6.1.1 Governing Equations and Numerical Methods

Schematic diagram of the 3D receptivity problem for a 2D ZPG boundary layer and
the computational domain is shown in Fig. 6.1. A 2D ZPG boundary layer is excited
from the wall either through a Gaussian circular patch (GCP) exciter or a spanwise
modulated (SM) exciter. While simulating the receptivity problem, the leading edge
of the plate has been retained inside the computational domain to capture instability
arising from the leading edge, which is a source of disturbance creation [37]. The
origin of the reference co-ordinate system is located at the mid-point of the leading
edge of the plate. The computational domain is given as xin ≤ x ≤ xout along the
streamwise direction, with xin < 0; 0 ≤ y ≤ ymax along the wall-normal direction

© Springer Nature Singapore Pte Ltd. 2019 307


T. K. Sengupta and S. Bhaumik, DNS of Wall-Bounded Turbulent Flows,
https://doi.org/10.1007/978-981-13-0038-7_6
308 6 3D Routes of Transition to Turbulence by STWF

y
m
Far-field boundary

ax
y
Inflow

ow
Outfl

x
z
U
x out

e
Plat
Gaussian Bump Exciter
x ex
zm

Plate leading edge


ax

x in

Fig. 6.1 Schematic diagram of the 3D receptivity problem for a 2D ZPG boundary layer

and −z max /2 ≤ z ≤ z max /2, along the spanwise direction. For the GCP exciter, the
wall-excitation is provided through a circular patch centered around (xex , 0, 0) and
of radius r0 . For the SM exciter, wall-excitation is provided by an exciter strip which
is from x1 to x2 along streamwise direction and along the whole spanwise extent of
the computational domain.
The simulations are performed in the transformed (ξ, η, ζ )-plane, such that x =

→ − →
x(ξ ), y = y(η) and z = z(ζ ). The rotational variant of the ( V , Ω )-formulation of
the incompressible NSE is given by


∂Ω −

+∇ × H =0 (6.1)
∂t

→ − → − → −

where H = Ω × V + (1/Re L )∇ × Ω . Equation (6.1) in transformed coordinate
can be expressed as
 
∂Ωξ 1 ∂ Hζ 1 ∂ Hη
+ − =0 (6.2)
∂t h 2 ∂η h 3 ∂ζ
 
∂Ωη 1 ∂ Hξ 1 ∂ Hζ
+ − =0 (6.3)
∂t h 3 ∂ζ h 1 ∂ξ
 
∂Ωζ 1 ∂ Hη 1 ∂ Hξ
+ − =0 (6.4)
∂t h 1 ∂ξ h 2 ∂η
6.1 Introduction 309


→ −

where, Ω = (Ωξ , Ωη , Ωζ ) and V = (u, v, w). The scale factors h 1 , h 2 and h 3 are
∂y
given as h 1 = ∂∂ξx , h 2 = ∂η ∂z
and h 3 = ∂ζ . In Eqs. (6.2) to (6.4), the terms Hξ , Hη and
Hζ are given as
 
1 1 ∂Ωζ 1 ∂Ωη
Hξ = (wΩη − vΩζ ) + − (6.5)
Re L h 2 ∂η h 3 ∂ζ
 
1 1 ∂Ωξ 1 ∂Ωζ
Hη = (uΩζ − wΩξ ) + − (6.6)
Re L h 3 ∂ζ h 1 ∂ξ
 
1 1 ∂Ωη 1 ∂Ωξ
Hζ = (vΩξ − uΩη ) + − (6.7)
Re L h 1 ∂ξ h 2 ∂η

→ − →
For the ( V , Ω )-formulation, the attendant velocity vectors are obtained from the

→ −

velocity Poisson equation ∇ 2 V = −∇ × Ω , and in the transformed (ξ, η, ζ )-plane
these are given as,
 
∂Ωη ∂Ωζ
∇ξ2ηζ u = h 1 h 2 − h3h1 (6.8)
∂ζ ∂η
 
∂Ωζ ∂Ωξ
∇ξ2ηζ v = h 2 h 3 − h1h2 (6.9)
∂ξ ∂ζ
 
∂Ω ξ ∂Ω η
∇ξ2ηζ w = h 3 h 1 − h2h3 (6.10)
∂η ∂ξ

where the operator ∇ξ2ηζ is given as


     
∂ h2h3 ∂ ∂ h3h1 ∂ ∂ h1h2 ∂
h 1 h 2 h 3 ∇ξ2ηζ = + +
∂ξ h 1 ∂ξ ∂η h 2 ∂η ∂ζ h 3 ∂ζ

Note that the velocity field should also satisfy the divergence-free condition Dv =


∇ · V = 0 and in the transformed (ξ, η, ζ )-plane this is given as,

1 ∂u 1 ∂v 1 ∂w
Dv = + + =0 (6.11)
h 1 ∂ξ h 2 ∂η h 3 ∂ζ

In deriving these equations, the free-stream velocity U∞ and L are used as the
velocity and length scales. The Reynolds number based on L is Re L = U∞ L/ν =
105 , for all the simulations reported here. While solving the receptivity problem,
only the Poisson equations for u- and w-components of velocity given by Eqs. (6.8)
and (6.10) are solved. The v-component of the velocity is calculated by integrating
Eq. (6.11) from the wall as
 η 
h 2 ∂u h 2 ∂w
v(ξ, η, ζ ) = v(ξ, 0, ζ ) − + dη (6.12)
0 h 1 ∂ξ h 3 ∂ζ
310 6 3D Routes of Transition to Turbulence by STWF

Application of Eq. (6.12) not only identically satisfies the solenoidality condition
on velocity, but also nullifies the requirement of imposition of any boundary condition
on the v-component of the velocity, at the far-field boundary, as shown in Fig. 6.1.

6.1.2 Boundary Conditions

The boundary condition at the inflow of the computational domain for the components

→ −

of velocity V , and vorticity Ω , are given as

∂v ∂w ∂Ωξ
u = 1 and = = = Ωη = Ωζ = 0 (6.13)
∂ξ ∂ξ ∂ξ

At the far-field of the computational domain, one does not require any boundary
condition on the v-component of velocity, as Eq. (6.12) is used to compute it. For the
other five variables, the boundary conditions used in the far-field are given as

∂Ωη
u = 1 and w = Ωξ = = Ωζ = 0 (6.14)
∂η

Periodic condition on all the six variables (three components of velocity and three
components of vorticity) are used in both the spanwise boundaries. At the wall, time-
dependent wall-normal velocity, corresponding to the type of excitation is prescribed,
along with the no-slip boundary conditions on u- and w-components of velocity. The
boundary conditions on the six variables on the plate surface are given as

u = w = 0, v = vw (x, z, t), Ωη = 0,
   
∂w ∂vw ∂vw ∂u
Ωξ = − and Ωζ = − (6.15)
∂y ∂z ∂x ∂y

The sharp leading edge of the flat plate is assumed as the locus of stagnation points
for this flow and hence, at the bottom plane ahead of the leading edge, as shown in
Fig. 6.1, symmetry conditions are used on all the six variables. This prescribes the
boundary conditions on the six variables given as

∂u ∂w ∂Ωη
=v= = Ωξ = = Ωζ = 0 (6.16)
∂η ∂η ∂η

At the outflow boundary, the convective Sommerfeld boundary conditions are


applied on the variables u, Ωη and Ωζ as
6.1 Introduction 311

∂u ∂u
+ Uc =0 (6.17)
∂t ∂x
∂Ωη ∂Ωη
+ Uc =0 (6.18)
∂t ∂x
∂Ωζ ∂Ωζ
+ Uc =0 (6.19)
∂t ∂x
The boundary condition on Ωξ at the outflow boundary is derived from the solenoidal-
ity condition of vorticity as
 
∂Ωξ ∂Ωη ∂Ωζ
=− + (6.20)
∂x ∂y ∂z

The boundary conditions of small amplitude disturbance on the v- and w-components


of velocity at the outflow boundary are derived from the definition of the vorticity
component Ωη and Ωζ as

∂v ∂u
= + Ωζ (6.21)
∂x ∂y
∂w ∂u
= − Ωη . (6.22)
∂x ∂y

6.1.3 Initial Condition

For the 2D equilibrium flow, Ωξ , Ωη and w are identically zero. Thus, one has to
solve only the transport equation for Ωζ , while u- and v-components of velocity are
obtained by solving the simplified Poisson equation and integrating the continuity
equation for 2D flows, respectively. The equilibrium flow is simulated with the initial
condition of impulsive start, i.e.,

u = 1, v = 0 and Ωζ = 0 (6.23)

Once the 2D equilibrium flow is established, the 3D equilibrium flow is obtained


by specifying u, v and Ωζ variables at all the discrete spanwise stations, while
prescribing other variables, i.e. w, Ωξ and Ωη to be zero at all locations. With these
initial conditions, the 3D solver is run for approximately 1000 iterations (when all
the unsteady terms falls below machine zero), so that the flow field adjusts itself
to the 3D domain and boundary conditions. Subsequently, 3D periodic excitation is
initiated, whose receptivity are simulated by solving Eqs. (6.2) to (6.12), subject to
the boundary conditions described in Sect. 6.1.2.
312 6 3D Routes of Transition to Turbulence by STWF

6.1.4 Grid Generation

The grid is generated such that grid-points are clustered near the leading edge of the
plate and which becomes uniform after x = 5. The grid points are also clustered near
the wall of the plate to accurately resolve the boundary layer. Both the clustering in
the streamwise and wall-normal directions are performed using tangent hyperbolic
function, as it produces minimum aliasing error during computation [10]. Along the
spanwise direction uniform grid-points are used. The grid transformation function
along the x-direction are given as for, xin ≤ x ≤ xs (0 ≤ ξ ≤ ξ1 )
 
tanh[βx (1 − ξ )]
x(ξ ) = xin + (xs − xin ) 1 − (6.24)
tanh βx

while for xs ≤ x(ξ ) ≤ xout (ξ1 ≤ ξ ≤ 1),


  
βx ξ − ξ1
x(ξ ) = xs + (xs − xin ) (6.25)
tanh βx ξ1
  
xout −xs tanh βx
where ξ1 = 1
1+A1
and A1 = xs −xin βx
. The grid-transformation function
along the wall-normal direction is given as
 
tanh[β y (1 − η)]
y(η) = ymax 1 − (6.26)
tanh β y

where 0 ≤ η ≤ 1. Here, βx and β y are parameters that control the grid clustering
in the streamwise and wall-normal direction, respectively. Here, for all the cases,
βx = 1 and β y = 2, are used.
For some of the simulations described here, xin = −0.05, xout = 20 and ymax =
0.75 have been used, while z max = 2 for all the cases, except two cases, where

z max = 4 is taken. The reference length scale L is taken such that L = 41δout , where

δout is the displacement thickness at the outflow for the equilibrium flow of the com-
putational domain. Note the streamwise length of the domain for these 3D simulations
is much longer than the 2D simulation in [11].
For a typical simulation performed with xin = −0.05, xout = 20, ymax = 0.75,
and 1001 and 301 points along x- and y-directions, the minimum and max-
imum resolutions along the streamwise direction are Δxmin = 9.1 × 10−3 and
Δxmax = 2.1 × 10−2 , respectively, while the wall-normal resolution is given as
Δymin = 3.68 × 10−4 , which stretches up to Δymax = 5.18 × 10−3 at the far-field
boundary.
6.1 Introduction 313

6.1.5 Numerical Method and Solution Technique

While simulating the flow, optimized staggered compact schemes (OSCS) are used
for the purpose of both interpolation of the function and evaluation of first derivative
of the function [35]. The OSCS scheme has been described in Chap. 2. The second or
mixed derivative terms are evaluated by repeated application of the OSCS scheme for
the evaluation of first derivative. This way of evaluating second derivative is distinctly
different from that is use in discretizing diffusion term in self adjoint form [36], which
has been used for 2D transitional flow simulations [34, 35]. The O R K 3 scheme is
used to integrate the VTEs (Eqs. (6.2) to (6.4)) with a time-step of Δt = 8 × 10−5 .
To suppress numerical spanwise spurious oscillations, periodic sixth-order filter with
the filter coefficient α f = 0.45 is used in the spanwise direction. For the purpose of
de-aliasing, numerical fourth-order diffusion term is used with coefficient ε = 0.06,
in both the streamwise and wall-normal directions. Proper way of adding fourth-order
numerical diffusion in the vorticity transport equation is described next. Consider
Eq. (6.2), which can be further expressed as
 
∂Ωξ 1 ∂ ∂ ∂
+ (h 1 uΩξ ) + (h 2 vΩξ ) + (h 3 wΩξ )
∂t h 2 h 3 ∂ξ ∂η ∂ζ
 
1 ∂ ∂ ∂ 1 −

= (h 1 uΩξ ) + (h 2 uΩη ) + (h 3 uΩζ ) − [∇ × Ω ]ξ (6.27)
h 2 h 3 ∂ξ ∂η ∂ζ Re L

Terms included within the first square braces indicate nonlinear convection of Ωξ ,
whereas the first set of terms on the right hand side of Eq. (6.27) indicate the gen-


eration of Ωξ due to stretching of the other components of Ω . Therefore, one has
to add numerical diffusion in the respective convection terms for effective upwind-
ing, whereas the vortex stretching terms are to be discretized with central difference
schemes. Here, fourth-order numerical diffusion for Ωξ is added by combining the
following terms with the convection terms,
 
ε h 1 |u|Δi3 Ωξ h 2 |v|Δ3j Ωξ h 3 |w|Δ3k Ωξ
+ + (6.28)
h2h3 Δξ Δη Δζ

where | · | indicates the absolute value and Δ3p f is given as

Δ3p f = −3( f p+1 − f p ) + ( f p+2 − f p−1 )

the value of ε fixes the amount of diffusion added, and here ε = 0.06 is taken.
Equations (6.2) to (6.4) are solved in the computational domain shown in Fig. 6.1,
by domain decomposition technique and using MPI parallelization framework. For
solving the resulting tridiagonal matrix equations in ξ - and η-directions, domain-
decomposition technique of [38] is used, where an overlap of six points are taken.
In the spanwise direction, one has to retain periodicity of the problem, by solving
the periodic tridiagonal equations. As the technique of [38] is essentially for non-
314 6 3D Routes of Transition to Turbulence by STWF

periodic problem, it is inherently incapable of retaining symmetry of the problem,


a modified periodic version of the algorithm proposed in [27] to solve the periodic
tridiagonal equations in the spanwise directions is used. Most of the simulations are
carried out using 128 computing units (CU) with 16 CU in the streamwise, 4 CU in
the wall-normal and 2 CU in the spanwise direction. For some simulations a total of
512 CU are used, with 32 CU along streamwise, 8 CU along wall-normal and 2 CU
along spanwise direction.

6.2 Gaussian Circular Patch (GCP) Excitation

For the simulation results reported here, time-harmonic Gaussian type excitations
are provided in a circular patch. This is different from the experiments performed
for impulse response [12]. Present receptivity study is also different from other
theoretical effort [15]. The imposed wall-normal velocity component vw (x, z) on the
patch is given as

vw (x, z) = α1 Am (x, z) sin(ω̄0 t) (6.29)

where α1 is the amplitude control parameter, Am (x, z) is the amplitude function


with absolute value varying from zero to one, and ω̄0 is the non-dimensional circular
frequency given by ω̄0 = F f × Re L . The amplitude function Am (x, z) for the circular
Gaussian type excitation is given as
  
1 πr
Am (x, z) = 1 + cos (6.30)
2 rmax

for r ≤ rmax and Am (x, z) = 0 for r > rmax . Here, r = (x − x0 )2 + (z − z 0 )2 with
x0 and z 0 denoting the center of the circular patch. For the results reported here,
rmax = 0.09, x0 = 1.5 (Reδ∗ = 666) and z 0 = 0 are used. In Fig. 6.2a, the amplitude
function Am (r ) is plotted as a function of r/rmax . In Fig. 6.2b, a typical snapshot of
the receptivity result at t = 15 is shown, by plotting the streamwise component of
the disturbance velocity u d , in the (x, z)-plane, for F f = 5 × 10−5 and α1 = 0.01,
at y = 0.00189. The spatio-temporal front is marked, with the oblique TS wave-
packets shown clearly, along with the local solution. One notes that the amplitude of
the STWF is several orders of magnitude higher than the TS wave-packet, at this time.
In the left frames of Fig. 6.3, evolution of u d at z = 0, y = 0.00189 and indicated
times are shown, as a function of x. The Fourier transform of the signals, displayed
in the left frame, are shown in the right frames of Fig. 6.3. One observes that the
STWF exhibits significant growth in time, in comparison to the TS wave-packet, as
it propagates downstream. This is also evident from the spectrum of u d shown in the
right frames of Fig. 6.3. Induction of similar high velocity, very close to the wall,
gives rise to the evolution of streamwise elongated puffs, similar to what is described
in [9, 44], whose amplification gives rise to transition, via formation of turbulent
spots. The wave-front is also shown to induce an additional disturbance wave-packet
6.2 Gaussian Circular Patch (GCP) Excitation 315

1
Gaussian bump excitation
0.8
Am 0.6

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
r/rmax

Fig. 6.2 a The amplitude function Am (r ) plotted as a function of r/rmax . b u d plotted in (x, z)-
plane at y = 0.00189 and t = 15 for 2D ZPG boundary layer excited by a circular wall bump exciter
with non-dimensional frequency F f = 5 × 10−5

marked as A, just upstream of it. This induced wave-packet is also seen to suffer
marginal growth in time.
In Figs. 6.4, 6.5, 6.6, 6.7, 6.8 and 6.9 evolution of disturbances for the Gaussian
wall bump exciter are shown, for F f = 0.5 × 10−4 , 0.75 × 10−4 and 1.0 × 10−4 , by
plotting the contours of u d in the (x, z)-plane, at y = 0.00189 and indicated times.
One notes that with progress in time, a STWF evolves, whose shape resembles an
arrowhead. The width of these structures increase, as these propagate downstream.
Due to the enforcement of periodic conditions in the spanwise directions, soon neigh-
boring wave-fronts starts interacting with each other, resulting in the creation of high
wavenumber oscillations, at the spanwise boundaries of the computational domain.
Of the three frequencies considered here, the STWF for the lowest frequency grows
fastest, and hence it interacts most vigorously with the neighboring wave-fronts,
creating a turbulent spot at t = 27.30, as marked by P in Fig. 6.9a. At t = 30, one
also observes the wave-front for F f = 0.75 × 10−4 to induce a turbulent spot at
x  13.5 (Fig. 6.9b). However, its intensity is less than the spot shown in Fig. 6.9a.
316 6 3D Routes of Transition to Turbulence by STWF

(a) t = 15.00 Wave-front (b) t = 15.0


0.03 100

0.02 Local sol n wave-front


10-1
0.01 -2

Ud
10
ud

0
-3
Peak due to
-0.01
TS wave-packet
10 TS wave-packet
-0.02 10-4
0 5 x 10 15 20 100 k 101 102

0.03
(c) t = 20.00 100
(d) t = 20.0
0.02 10-1
0.01
A -2

Ud
10
ud

0
-3
10
-0.01
-0.02 10-4
0 5 x 10 15 20 100 k 101 102

0.03
(e) t = 25.00 100 (f) t = 25.0
0.02 10-1

0.01 A -2
Ud

10
ud

0
-3
10
-0.01
-0.02 10-4
0 5 x 10 15 20 100 k 101 102

(g) t = 27.30 (h) t = 27.3


0.03 100

0.02 10-1
0.01 A -2
Ud

10
ud

0
-3
10
-0.01
-0.02 10-4
0 5 x 10 15 20 100 k 101 102

Fig. 6.3 u d plotted as a function of x at z = 0, y = 0.00189 and indicated time-instants in the right
column for 2D ZPG boundary layer circular wall bump exciter with non-dimensional frequency
F f = 5 × 10−5 . The spectrum for the plotted u d in the left column is shown in the right

The wave-front created for F f = 1.0 × 10−4 , though interacts with the neighbor-
ing wave-fronts, but does not induce any turbulent spot by t = 30. Figures 6.4–6.9
clearly reveals that, it is only the STWF, which not only exhibits growth, but also
induces additional perturbations upstream of it, similar to the 2D receptivity problem
discussed in the previous chapter. One also notices in Figs. 6.8 and 6.9, that at t = 25
and 27.3, the magnitude of u d is of the order of the free-stream velocity. Considering
the fact that the height y = 0.00189, is only 3% of the boundary layer thickness at
x = 12, one concludes that such high streamwise perturbation velocity inside the
shear layer would not only give rise to elongated quasi-streamwise streaks, as found
in [9, 44], but also accelerate the transition process violently.
6.2 Gaussian Circular Patch (GCP) Excitation 317

Fig. 6.4 u d -contours at indicated time instant and y = 0.00189 shown in (x, z)-plane for a circular
wall bump exciter for indicated frequencies of excitation

Fig. 6.5 See caption of Fig. 6.4


318 6 3D Routes of Transition to Turbulence by STWF

Fig. 6.6 See caption of Fig. 6.4

Fig. 6.7 See caption of Fig. 6.4


6.2 Gaussian Circular Patch (GCP) Excitation 319

Fig. 6.8 See caption of Fig. 6.4

Fig. 6.9 See caption of Fig. 6.4


320 6 3D Routes of Transition to Turbulence by STWF

(a ) Ff = 0.5×10−4, t = 10 (b1) Ff = 1.0×10−4, t = 10


0.1 0.1
Ω -contour at (z = 0)-plane
0.08 ζ 0.08
0.06 0.06
y

y
0.04 0.04
0.02 0.02
0 0
0 2 4 6 8 10 12 14 16 18 0 2 4 6 x 8 10 12 14 16 18
x
(a2) Ff = 0.5×10−4, t = 15 Evolving (b2) Ff = 1.0×10−4, t = 15
0.1 0.1
n wave-front
0.08 Local sol and 0.08
0.06
TS wave-packet 0.06
y

y
0.04 0.04
0.02 0.02
0 0
0 2 4 6 8 10 12 14 16 18 0 2 4 6 x 8 10 12 14 16 18
x
(a3) Ff = 0.5×10−4, t = 20 (b3) Ff = 1.0×10−4, t = 20
0.1 0.1
0.08 0.08
0.06 0.06
y

0.04 y 0.04
0.02 0.02
0 0
0 2 4 6
x 8 10 12 14 16 18 0 2 4 6 x 8 10 12 14 16 18
−4 −4
(a4) Ff = 0.5×10 , t = 25 (b4) Ff = 1.0×10 , t = 25
0.1 0.1
0.08 0.08
0.06 0.06
y

0.04 0.04
0.02 0.02
0 0
0 2 4 6
x 8 10 12 14 16 18 0 2 4 6 x 8 10 12 14 16 18
(a5) Ff = 0.5×10−4, t = 27 (b5) Ff = 1.0×10−4, t = 30
0.1 0.1
0.08 0.08
0.06 0.06
y

0.04 0.04
0.02 0.02
0 0
0 2 4 6 x 8 10 12 14 16 18 0 2 4 6 x 8 10 12 14 16 18

Fig. 6.10 Ωζ -contours plotted in (x, y)-plane for z = 0 at indicated times for a circular wall bump
exciter with F f = 0.5 × 10−4 (left frames) and F f = 1.0 × 10−4 (right frames)

In Fig. 6.10, Ωζ -contours are plotted in the (x, y)-plane, and indicated times at
z = 0 plane, for F f = 0.5 × 10−4 (left frames) and F f = 1.0 × 10−4 (right frames).
This figure brings out the essential differences between 2D and 3D transition. For
the 3D transition, one notes that the dominant vortices underneath the wave-front
makes an angle with the streamwise direction. This is due to the 3D orientation of
the vortical structures, as also noted in [6, 8, 26, 40]. For 2D transition, the created
vortices are almost vertical on the plate surface, as noted in the previous chapter.
One also notes from Fig. 6.10a4, a5, that the high wavenumber fluctuations initiate
above the plate surface, close to the edge of the shear layer. In contrast, for all the
2D cases the breakdown is seen to be initiated very close to the plate surface.
In Fig. 6.11, u d is plotted as a function of y/δ ∗ , at (a) z = 0 and (b) z = 0.6625.
The time instants and the streamwise locations indicated in frames (a) and (b) of
Fig. 6.11 correspond to the maximum amplitude of the wave-front at z = 0 and
6.2 Gaussian Circular Patch (GCP) Excitation 321

(a) z = 0
0.1

0.05
ud

0
x = 3.58, t = 5.0
-0.05 x = 6.56, t = 10.0
x = 7.80, t = 12.0
-0.1 x = 9.55, t = 15.0
x = 12.53, t = 20.0
-0.15 x = 12.62, t = 25.0
x = 13.35, t = 27.0
-0.2
0 1 2 3 4 5
y/δ∗

(b) z = 0.6625
0.6
x = 3.23, t = 5.0
0.5 x = 4.68, t = 10.0
x = 5.65, t = 12.0
0.4 x = 6.50, t = 15.0
x = 8.91, t = 20.0
0.3 x = 11.30, t = 25.0
ud

x = 13.42, t = 27.0
0.2

0.1

-0.1
0 1 2 ∗ 3 4 5
y/δ

Fig. 6.11 u d plotted as a function of y/δ ∗ at a z = 0 and b z = 0.6625. The time instants and
the streamwise locations correspond to the maximum amplitude of the wave-front at z = 0 and
z = 0.6625, respectively for y = 0.00189

z = 0.6625, respectively. One notes that at the initial phases of evolution, the wall-
normal variation of u d resembles the eigenfunction φ  (y) of the OSE noted in Chap. 3.
At z = 0, as the wave-front moves downstream, significant amplification of u d is
noted, near the outer maxima of its wall-normal profile. This observation is in contrast
to the frame (b) where u d close to the inner maxima displays maximum growth. At
later times of the evolution of the wave-front, the wall-normal profile of u d displays
marked variation with the Orr-Sommerfeld eigenfunction, with u d attaining a value,
which is of the order of the free-stream speed at y/δ ∗  0.2. One also notes the
disturbances to rapidly penetrate higher wall-normal distances, as one approaches
the turbulent spot.
322 6 3D Routes of Transition to Turbulence by STWF

Fig. 6.12 Iso-surface of Q shown plotted at a t = 24.96 and b t = 27.28 for a circular wall bump
exciter for F f = 0.5 × 10−4

In Sengupta et al. [37], as well as in the previous chapter, the bypass transition
on a flat plate caused by a slowly convecting vortex of anti-clockwise circulation
is studied, which creates transition/unsteady separation ahead of it. The theoretical
explanation in [37], is based upon the time evolution of the disturbance mechanical
energy (E d ), defined as the difference between the total mechanical energy (E t )


and its equilibrium value (E m ), where E t = p/ρ∞ + | V |2 /2 and E m = pm /ρ∞ +

→ 2
| V m | /2. The governing equation for E d is given as

∇ 2 Ed = Q (6.31)
6.2 Gaussian Circular Patch (GCP) Excitation 323

Fig. 6.13 u d -contours plotted in the (x, z)-plane at indicated time for a z max = 2 and b z max = 4
when excited by a circular wall bump exciter for F f = 1.0 × 10−4

where

→ −
→ −

Q = 2−

ωm ·−

ω d + |−

ω d |2 − V m · (∇ × −

ω d ) − V d · (∇ × −

ω m ) − V d · (∇ × −

ω d)

Hence E d grows when Q < 0, and decays when Q > 0. This generic mechanism
is based on NSE, without any simplifying assumptions. In Fig. 6.12, the iso-surface
of Q = 1500 and Q = −1500 are plotted at t = 24.96 and 27.28, in frames (a) and
(b), respectively, for F f = 0.5 × 10−4 case. The maximum of the absolute value of
Q at both the time instants are of the order of 104 . The Q = 1500 and −1500 iso-
surfaces are depicted by the green and dark blue colors, respectively. One notes the
simultaneous presence of the positive and negative values of Q in a very confined
zone, gives rise to enhanced instability of the flow. With advance in time from t =
24.96 to 27.28, the fluctuation in the value of Q is seen not only to intensify, but also
to cover larger extent of the computational domain. At t = 27.28, the wall-normal
extent of these fluctuations are seen to pierce through the boundary layer at x  12.
The effects of the spanwise width of the computational domain is studied in
Figs. 6.13, 6.14, 6.15, 6.16, 6.17 and 6.18, where u d -contours are plotted in the
(x, z)-plane from t = 5 to 30, at an interval of five nondimensional time unit for
the computational domain with (a) z max = 2 and (b) z max = 4, when ZPG flow past
a flat-plate is excited by a GCP exciter with F f = 1.0 × 10−4 . One finds that in
324 6 3D Routes of Transition to Turbulence by STWF

Fig. 6.14 See the caption for Fig. 6.13

Fig. 6.15 See the caption for Fig. 6.13


6.2 Gaussian Circular Patch (GCP) Excitation 325

Fig. 6.16 See the caption for Fig. 6.13

Fig. 6.17 See the caption for Fig. 6.13


326 6 3D Routes of Transition to Turbulence by STWF

Fig. 6.18 See the caption for Fig. 6.13

the initial times up to t = 10, the evolution of disturbance is identical for both the
cases. At t = 10, the disturbance corresponding to z max = 2 case is seen to strike
the spanwise boundaries of the corresponding computational domain. Thereafter,
the neighboring zones’ disturbances, corresponding to the STWF, start to interact,
causing development of high wavenumber fluctuations at the spanwise boundaries as
noted in Figs. 6.15a and 6.16a. However, for the z max = 4 case, the spanwise extent
for the u d growth corresponding to the STWF increases. At t = 25, the STWF hits the
spanwise boundaries of the computational domain in this case. However for both the
cases, one finds that the structure of the wedge-shaped TS wave-packet is identical
in nature, for all the time instants shown in Figs. 6.13–6.18.
To further understand the nature of transition, in Fig. 6.19 amplitude of u d cor-
responding to STWF is plotted as a function of time, at four spanwise locations for
y = 0.00189, and excitation frequency, F f = 0.5 × 10−4 . One notes that the time
variation of the STWF amplitude is different from its 2D counterpart. For z = 0.65
and 0.9 (Fig. 6.19a) one notes that there exists a period of constant amplitude, after
the initial linear growth, because of streamwise and spanwise dispersion of STWF.
However, no such time duration exists for the spanwise location of z = 0.325. For
z = 0.325, dispersion causes the amplitude to slightly decay, after the initial linear
growth stage. This is followed by a period of continuous nonlinear growth, followed
by eventual nonlinear saturation of amplitude. At the mid-spanwise location (z = 0),
6.2 Gaussian Circular Patch (GCP) Excitation 327

Fig. 6.19 Amplitude of u d


(a)
corresponding to the
wave-front plotted as a
-1
function of time at indicated 10
spanwise locations of z = 0,
10-2
0.65 and 0.9 for
y = 0.00189. The frequency -3
10
of excitation is

udm
F f = 0.5 × 10−4 10-4
z = 0.325
z = 0.65
-5
10 z = 0.9

-6
10
0 5 10 15 20 25
t
(b)
0.014
0.012 z=0
0.01

0.008
udm

0.006

0.004

5 10 15 20 25
t

the growth of STWF is significantly different from the other three spanwise locations,
because of the enhanced spanwise domain.

6.3 Spanwise Modulated (SM) Excitation

In this section, effects of spanwise modulated excitation on 2D ZPG boundary layer


are studied. Here the amplitude function Am (x, z), as defined in Eq. (6.29), for the
imposed wall-normal velocity component vw (x, z), is given as
    
1 x − xm z
Am (x, z) = 1 + cos π sin 2π n
2 x2 − x1 z max
for x1 ≤ x ≤ x2 (6.32)

and Am (x, z) = 0 for x < x1 or x > x2 . Here xm = (x1 + x2 )/2, n = 4 and z max = 2.
The amplitude control parameter α1 , and the non-dimensional excitation frequency
328 6 3D Routes of Transition to Turbulence by STWF

Fig. 6.20 Perspective plot of u d at a fixed height shown for SM exciter. The frequency of excitation
is F = 0.5 × 10−4

F f , are taken as 0.01 and 1.0 × 10−4 , respectively. One notes that at the spanwise
locations, z = mz max /8: Am (x, z) = 0 where m = 0, ±1, ±2, ±3 and ± 4. As for
this case z max = 2, Am (x, z) is identically zero at z = 0, ±0.25, ±0.5, ±0.75 and
± 1. These spanwise locations are the nodes. Similarly, maximum excitation is
imposed at z = ±0.125, ±0.375, ±0.625 and ± 0.875, and which are the anti-
nodes or peaks. Out of these peak locations, excitation at z = −0.875, −0.375,
0.125 and 0.625 are in phase, and which have 180◦ phase difference with the imposed
excitation at other peaks at z = −0.625, −0.125, 0.375 and 0.875. In the following
discussion, the first set of spanwise locations are termed as “type-1 peak-locations”
whereas the second set of spanwise locations are termed as “type-2 peak-locations”
for the ease of explanation. It should be noted that such an excitation closely mim-
ics the experimental excitation of [21], where a spanwise modulated excitation is
obtained by applying spacers in periodic succession. The present type of excitation,
also creates streamwise fluctuating vortices, with the center at the nodes. Perspective
view of a typical disturbance evolution, corresponding to SM exciter, is shown in
Fig. 6.20. Here also, one notes the existence of three-component solution structure.
The local solution is seen very close to the exciter, and is followed by the TS wave-
packet corresponding to the frequency of excitation and its super-harmonics. For the
frequencies investigated, one also notices the presence of STWF for the displayed
cases at all spanwise stations. The amplitude and phase of STWF show significant
variation, as STWF propagates downstream. One notes that TS wave-packet does
not grow and cause transition; instead STWF is the main precursor of flow transition.
Detailed dynamics of the STWF is illustrated next.
In Fig. 6.21, the u d -contours are plotted in the (x, z)-plane for y = 0.00189 and
indicated time instants, for this case. One notes the evolution of a STWF, induc-
ing perturbations at t = 10, whose magnitude are higher than the corresponding
disturbances induced by the TS waves. At t = 10, one also notes the prominent
6.3 Spanwise Modulated (SM) Excitation 329

Fig. 6.21 u d -contours plotted in the (x, z)-plane at y = 0.00189 and indicated time, when 2D ZPG
boundary layer excited by spanwise modulated exciter strip for F f = 1.0 × 10−4
330 6 3D Routes of Transition to Turbulence by STWF

Fig. 6.22 The Ωξ -contours at indicated streamwise locations and t = 27.24 plotted in the (y, z)-
plane for spanwise modulated excitation for F f = 1.0 × 10−4

disturbances, due to the wave-front arranged in a staggered formation. Such stag-


gered formation is obtained due to the 180◦ phase difference between two succes-
sive peak locations. At t = 15, these staggered perturbations tend to merge with
each other, inducing high wavenumber fluctuations, in the range 10.5 ≤ x ≤ 13.5,
at t = 20. These fluctuations induce four turbulent spots, as noted at t = 25. These
turbulent spots are located at “type-2 peak-locations”, and are denoted by dark blue
spots in Fig. 6.21d, where u d  0.4. In contrast, a relatively elongated, but slightly
reduced perturbations are observed at the “type-1 peak-locations” at t = 25, trail-
ing the “type-2 peak-location” structures. Gradually, these disturbances penetrate
upstream to give rise to periodic spanwise-modulated, and streamwise-elongated
6.3 Spanwise Modulated (SM) Excitation 331

streaks at t = 30, while the above turbulent spots are seen to reach the outflow of
the computational domain. These streaks at t = 30 are located at the nodes, and are
possibly intensified, due to lift-up effect of the streamwise counter-rotating vortices,
which while lifting fluid with low velocity from the wall, and forcing high-speed
fluids towards the wall, are most effective in creating streamwise-oriented streaks,
as postulated in [5, 23, 24]. In [1, 25], a pair of counter-rotating streamwise vortices
have been reported, as being most effective for the generation and evolution of such
streamwise streak, whose stability analysis is performed in [32]. Similar streamwise
streaks are also shown in [44], to be the precursor to the formation of turbulent spots.
The spanwise modulation of the perturbation across the shear layer is shown
in Fig. 6.22, where Ωξ -contours are plotted in the (y, z)-plane, at t = 27.24, for
streamwise locations varying from 13.01 to 16.01. From Fig. 6.22, one notes the
presence of positive-negative vortex pairs, close to the wall, and negative-positive
vortex pairs, away from the wall. The positive-negative vortex pair directs fluids of
higher momentum towards the wall, whereas the negative-positive vortex pair pushes
fluids of lower momentum, away from the wall. One notes that for the peak locations
of “type-1”, the first pair is stronger, whereas for the “type-2 peak locations”, the
second pair is stronger. This causes the perturbations corresponding to the “type-1
locations” to amplify more vigorously, than the perturbations at the “type-2 peak
locations”. This circulation of fluid also causes maximum u d close to the wall to
occur at the positions of the nodes, as explained in [3]. This establishes the crucial
role streamwise vortices play, in determining transition for 3D flows, which has been
pointed out experimentally earlier in [21]. However, continuous fully developed
turbulence is not shown in the present simulations, due to short streamwise length
of the computational domain, which occurs at downstream first, and then induces
events upstream.
In Fig. 6.23, u d for y = 0.00189 is plotted as a function of x, for the spanwise
locations corresponding to the “type-1” and “type-2” peaks. One finds that the
amplitude variation of the TS wave-packet is identical for both types of peak loca-
tions, with 180◦ phase shift between these. However the evolution of the wave-front is
almost identical for both the types of the peak locations, at initial times up to t = 10.
It is also interesting to note that contrary to the corresponding TS wave-packet, the
wave-fronts generated at both types of the peak locations, are completely in phase at
initial times. The wave-fronts at “type-2 peak locations” develop higher wavenum-
ber fluctuations, than the “type-1” counterpart, as noted from Fig. 6.23c at t = 15.
As a result of this, the wave-front at the “type-2 location” nonlinearly saturates and
breaks down earlier, than the wave-front at “type-1” locations, as noted at t = 20
and 25.
Results also show that the wave-front at nodal spanwise locations have a phase
difference of approximately 180◦ , with the wave-fronts at both “type-1” and “type-2
peak locations”, which mutually are in phase at initial times. This feature is illus-
trated in Fig. 6.24a, b, where u d is plotted as a function of x for nodal, “type-1” and
“type-2 peak locations”. This along with the already noted feature that the TS waves
are 180◦ phase apart for “type-1” and “type-2 peak locations”, indicate that the
spanwise wavelength of the wave-front, at initial times, are half the spanwise wave-
332 6 3D Routes of Transition to Turbulence by STWF

(a) t = 05.0 z = 0.125 (type-1)


0.04 z = 0.375 (type-2)
z = 0.625 (type-1)
0.02 z = 0.875 (type-2)

ud
0
-0.02
0 5 x 10 15 20

(b) t = 10.0 z = 0.125


0.04 z = 0.375
z = 0.625
0.02 z = 0.875
ud

0
-0.02
0 5 x 10 15 20
(c) t = 15.0 z = 0.125
0.04 z = 0.375
z = 0.625
0.02 z = 0.875
ud

0
-0.02
0 5 x 10 15 20
(d) t = 20.0 z = 0.125
0.2 z = 0.375
z = 0.625
0.1
z = 0.875
ud

-0.1 0 5 10 15 20
x
(e) t = 25.0
0.4 z = 0.125
0.3 z = 0.375
z = 0.625
0.2
z = 0.875
ud

0.1
0
0 5 x 10 15 20
(f) t = 30.0
z = 0.125
0.2 z = 0.375
z = 0.625
0.1 z = 0.875
ud

0
0 5 x 10 15 20
(g) t = 33.0
z = 0.125
0.15
z = 0.375
0.1 z = 0.625
0.05 z = 0.875
ud

0
-0.05
0 5 x 10 15 20

Fig. 6.23 u d plotted at y = 0.00189, indicated times and spanwise locations for spanwise modu-
lated excitation for F f = 1.0 × 10−4 . All the indicated z-locations correspond to the peak positions
of the excitation
6.3 Spanwise Modulated (SM) Excitation 333

(a) t = 10.0 z = 0.25 (node)


0.04 z = 0.125 (type-1)
0.02 z = 0.375 (type-2)
ud 0
-0.02
0 5 x 10 15
(b) t = 15.0 z = 0.25 (node)
0.04 z = 0.125 (type-1)
0.02 z = 0.375 (type-2)
ud

0
-0.02
0 5 x 10 15
(c) Spatio-temporal wave-front amplitude
Turbulent spot forms

10-1
udm

z = 0.25 (node)
-2
z = 0.125 (type-1 peak)
10 z = 0.375 (type-2 peak)
5 10 15 20 25 30 35
t

Fig. 6.24 u d at y = 0.00189 and spanwise locations plotted for a t = 10 and b t = 15. Maximum
amplitude of the STWF u dm at y = 0.00189 plotted as a function of time for z = 0.25, 0.125 and
0.375. Here, spanwise modulated excitation is provided for F f = 1.0 × 10−4

length of the TS wave-packet. In other words, the fundamental spanwise wavenumber


of the wave-front, for this case, is twice the fundamental spanwise wavenumber of
the TS wave-packet. In Fig. 6.24c the maximum amplitude of the STWF for stream-
wise disturbance velocity, u dm , is plotted as a function of time, for z = 0.25 (nodal
location), 0.125 (type-1 peak location) and 0.375 (type-2 peak location). One notes
that for the wave-front at “type-1” and “type-2 peak location”, the amplitude of
the wave-front decays for 10 ≤ t ≤ 17, after initial phase of exponential growth.
This is because of the nonlinear self-and/ or multi-modal interactions, coupled with
dispersion. However, this stage is followed by a stage of rapid growth, because of sec-
ondary and higher order instabilities [14], which leads to the formation of turbulent
spots at t  24. However, for the wave-front at the nodal locations, the growth rate
of the wave-front is moderated during the time duration of 10 ≤ t ≤ 17, instead of
decaying, as noted for STWF at both “type-1” and “type-2” peak locations. For these
spanwise locations also, rapid growth of the wave-front takes place, after t = 15,
due to secondary and higher order instabilities.
334 6 3D Routes of Transition to Turbulence by STWF

6.4 Routes of Flow Transition: K - and H-Type Routes

Significant progress have been made for different aspects of early stages of 3D
transition by experiments and computations in recent times, yet the actual routes tra-
versed by flows from laminar to eventual turbulent stage are not completely under-
stood till recent times. To explain the unit processes and routes of 3D flow tran-
sition mechanisms, experimental efforts have been undertaken by researchers for
monochromatic deterministic excitation of wall bounded shear layers, as reported
in [13, 21, 22, 33]. In [21], a rectangular ribbon with spanwise spacers was
vibrated monochromatically at 1489Hz (corresponding non-dimensional frequency
F f = 2π ν f /U∞ 2
= 6.03 × 10−4 ), near the surface of a flat plate. It has been shown
in [21], that longitudinal vortices are associated with the nonlinear 3D wave motions.
The transition was shown to be characterized by the downstream growth of spanwise
modulation of the disturbance amplitude, with the formation of peaks and valleys
along the spanwise direction [17]. At downstream locations, spikes appear suddenly
in the peak spanwise locations, where the disturbance amplitudes reach the local
maxima. Flow transition has been caused by rapid amplification and multiplication
of these spikes, which has been attributed to high-frequency secondary instability
due to inflectional instantaneous velocity profiles and high-shear [20–22].
Excitations, such as used in the experiments of [21], caused the evolution of an
aligned pattern of Λ vortices in the transitional zone, termed subsequently as K -
type transition [17, 30]. The Λ-vortices have the appearance of horse-shoe shape
whose central portion is lifted up with respect to the legs of the hairpin vortex.
An alternate route of 3D transition by monochromatic excitation at a much lower
frequency of 120 Hz (corresponding non-dimensional frequency F f = 2π ν f /U∞ 2
=
−5
1.37 × 10 ) was reported in [19], where Λ vortices are found to be in staggered
arrangement; classified later as H - or N -type transition [17, 30].
Theoretical approaches attempted to explain the K - and H -type breakdown to
a resonant mechanism [7, 17, 18]. K -type transition route is described to happen,
where 2D disturbance wave interacts with two oblique 3D waves of identical fre-
quency [7, 17]. In contrast, H -type transition is noted, when a 2D disturbance wave
interacts with two oblique 3D waves, corresponding to half the frequency of the
2D wave, as theoretically predicted in [7, 45]. The computational efforts to induce
K - and H -type breakdown is attempted [31, 43] by simultaneously exciting waves
at fundamental frequency and its spanwise modulated counterpart. None of these
efforts register the role of STWF, as described here as the precursor of flow transi-
tion. This is because of the use of significantly shorter computational domain [31,
43]. In the previous chapters it has been noted that STWF are the spatio-temporal
eigenmodes of NSE, as observed from the solution of the OSE in spatio-temporal
framework [41]. It originates due to the onset of excitation, as a combination of
spatio-temporal eigenmodes with weightage, which depends on the exciter location
and excitation frequency. Therefore, properties of STWF are different from the pure
spatial eigenmodes of the OSE [39]. Here, we show by the solution of NSE that
transition of wall-bounded flows caused by the the growth of STWF follows both
6.4 Routes of Flow Transition: K - and H -Type Routes 335

Fig. 6.25 Perspective view of λ2 = −0.0025 iso-surface is shown in (x, z)-plane at t = 25 after
the onset of excitation corresponding to GCP exciter case for a F f = 1.0 × 10−4 and b F f =
0.5 × 10−4 . Flow is from left to right

K - and H - or N -types of routes for low amplitude, monochromatic, deterministic


wall-excitation, due to the growth of STWF. While the H -type transition is noted
for lower frequency excitation cases, K -type is seen to occur for monochromatic
excitations with higher frequency.
Transition process in 3D flows are dominated by highly unsteady vortical struc-
tures, with all three components important. Accurately representing these vortical
structures unambiguously is quite difficult. Here, we capture the unsteady vorti-
cal structures by negative λ2 iso-surfaces, following the method proposed in [16],
to capture the unsteady vortical structures in late stages of 3D transition process.
Here, λ2 is the second eigenvalue of the symmetric matrix Sik Sk j + Ωik Ωk j rep-
resenting rate of strain tensor, where Si j = (1/2)(∂u i /∂ x j + ∂u j /∂ xi ) and Ωi j =
(1/2)(∂u i /∂ x j − ∂u j /∂ xi ) are symmetric and anti-symmetric part of the velocity
gradient tensor, respectively.
In Fig. 6.25a, b, we have plotted the perspective of λ2 = −0.0025 iso-surface
(colored by streamwise velocity), for the GCP exciter cases with F f = 10−4 and
F f = 0.5 × 10−4 , respectively. In Fig. 6.25, trailing part of the STWF is focused to
show later stages of transition in the perspective plot. One notes the formation of
Λ-type vortices, in all cases with lifted Ω-like element of the hairpin vortices at the
center. In Fig. 6.25a, the darker spots denote the Ω-like element at the top of the
Λ-vortices.
In Fig. 6.25a, we note an aligned pattern of hairpin-vortices, while these are
arranged as Λ vortices in Fig. 6.25b, in a staggered arrangement. Therefore, the
case shown in Fig. 6.25b, has been conjectured to display subharmonic route of tran-
sition [17, 19, 30]. Unlike in [31, 43], here both the cases are due to monochromatic
excitation, with the only difference in the excitation frequency. Thus, the present
results clearly indicate that for moderate frequencies, one notices K -type transition,
as in Fig. 6.25a, while H -type transition occurs at significantly lower frequencies.
Similar differences in the arrangement of the vortices are also noted for the SM
exciter cases in Fig. 6.26a–e, for F f = 10−4 and 0.5 × 10−4 , respectively. While
a definitive staggered pattern of Λ-vortices are noted for F f = 0.5 × 10−4 case,
almost aligned arrangement of these vortices are seen for F = 10−4 case. In both the
336 6 3D Routes of Transition to Turbulence by STWF

Fig. 6.26 Perspective and top view (in (x, z)-plane) of λ2 = −0.015 iso-surface are shown for the
SM exciter case for (a, b) F f = 10−4 and (c, d, e) F f = 0.5 × 10−4 . Flow is from left to right

cases, existence of ring-like Ω-vortices connecting the legs of the vortices are noted.
These are also known as hairpin vortices [40]. One notes from Fig. 6.26a, c that,
while for higher frequency, the hairpin vortices are aligned along the x-direction,
these make an oblique angle for the lower frequency case. For F f = 10−4 case,
Ω-vortices connect the two spanwise neighboring hairpin vortices (see Figs. 6.25a
and 6.26a). For lower frequency case, Ω-vortices only connect the hairpin vortices,
whose lifted front parts are close to each other. In Fig. 6.26d, between the dominant
Λ-vortices, one notes the existence of some intermediate structures, which appear
due to the induction of pressure variation by the adjoining Λ-vortices [40]. At further
downstream locations, these structures become streamwise elongated, giving rise to
U-shaped vortices, noted in Fig. 6.26e, which penetrates the shear-layer.
It is to be pointed out that here both K - and H -type transition is noted for determin-
istic monochromatic excitation via the growth of STWF. So far, all the H -type transi-
tion routes demonstrated in experiments [19] and simulations [31, 43] are stated to be
via the amplification of the spatial TS wave-packet, where a fundamental frequency
along with its sub-harmonic is excited. This follows the theoretical view-point of
triad resonant interaction of a fundamental frequency f with its sub-harmonic com-
ponent f /2 [7, 45]. In contrast the presented results show flow transition to occur
6.4 Routes of Flow Transition: K - and H -Type Routes 337

via both K and H/N -type routes for monochromatic deterministic excitation, due to
the growth of STWF, without requiring any subharmonic frequency to be explicitly
excited.

6.5 Formation of Turbulent Spots and Fully Developed


Turbulent Flow

The formation of turbulent spots, and consequent fully developed turbulent flow for
the case of GCP exciter with F = 0.5 × 10−4 , is illustrated here. In Fig. 6.27a–d,
we show contours of Ωζ at the indicated time instants, along z = 0. At t = 30,
an intermittent turbulent zone spans from x  13 to 19. With time, one notes the
front of this highly perturbed zone to move downstream, at a speed comparable to
the free stream velocity. The trailing edge of STWF shows very minor movement
downstream up to t = 35, and beyond that time, this almost remains frozen at x  13.
The transitional flow, which spans from x  12 to 15 in frame (b), keeps elongating,
while causing vortical eruptions, which grow and merge with the turbulent part ahead
of it. One such set of vortical eruptions is marked as A in frames (b) and (c). One
also notes, thickening of the boundary layer gradually, from the laminar value at
x = 12 to 15 (which can be construed as the point of transition) and beyond for later
times. The intermittent zone is characterized by highly unsteady vortical eruptions.
Presence of unsteady vortical eruptions, and constant regeneration mechanism was
also noted for deterministically created 2D turbulent flow in Chap. 5. Here, the flow
is seen to be turbulent beyond x = 15, however, the intermittency is lower beyond
x  25.
To characterize the turbulent zone, we plot the wall-normal variation of u + (mean
streamwise velocity < U >, non-dimensionalized by wall-friction velocity u τ =

< τw > /ρ, where < τw > is the mean wall-shear and ρ is fluid density) as a
function of y + = u τ y/ν, in Fig. 6.28a, for indicated streamwise stations. According
to descriptions of Reynolds-averaged fully developed turbulent boundary layer in
[28, 42],
u+ = y+

in the viscous sub-layer (0 ≤ y + ≤ 10) and


1
u+ = ln(y + ) + a
κ1

in the inertial layer - above the viscous sub-layer and buffer layer (for 30 ≤ y + ≤
2000), where κ1 = 0.41, is the von Karman coefficient. This is the ‘logarithmic law
of the wall’in Fig. 6.28a, marked by dashed lines. To determine the mean stream-
wise velocity for all the streamwise stations, we have time-averaged the data from
t = 40 to 50. At x = 20 to 23, the match of u + at the outer part is not good, because
these stations are not in fully developed turbulent flow region. One obtains very
338 6 3D Routes of Transition to Turbulence by STWF

Fig. 6.27 ωz -contours at z = 0 station plotted in (x, y)-plane at a t = 30, b t = 35, c t = 40 and
d t = 50 for GCP exciter case for F = 0.5 × 10−4

good match with the above expressions for u + , at x = 25 and beyond. In Fig. 6.28b,
Reynolds stress variation with y + is shown at the indicated stations, which demon-
strates how this stress changes in the transitional flow. One notes that the Reynolds
stress is predominantly negative between 10 ≤ y + ≤ 2000, for x ≥ 25. The range,
10 ≤ y + ≤ 2000, corresponds to that part of the boundary layer, where predominant
negative values signify enhanced production of turbulent kinetic energy, as reported
in the literature. In Fig. 6.28c, the non-dimensional root mean square streamwise
6.5 Formation of Turbulent Spots and Fully Developed Turbulent Flow 339

−4
(a) GCP exciter with F=0.5×10
35 Time averaged between t=40 and 50
30 x = 20
x = 21
25
x = 23
20 x = 25
x = 27
+
u

15 x = 30
u+=y+
10 u+ = (ln y+)/κ + a
Viscous-sublayer where κ = 0.41
5 u+ = y+

101 y+ 102 103

(b) Reynolds stress


0
Reynolds stress

-1
x = 20
x = 21
x = 23
-2 x = 25
x = 27
x = 30
-3

1 2 3
10 y+ 10 10

(c) RMS of u’
x = 20
x = 21
5 x = 23
x = 25
4 x = 27
x = 30
u’rms/uτ

0 1 2 3
10 y
+ 10 10

Fig. 6.28 a u + plotted as a function of y + at indicated x-stations for GCP exciter case for F =
0.5 × 10−4 . b Non-dimensional Reynolds stress plotted as a function of y + at indicated x-stations
and c nondimensional streamwise component r ms velocity shown at the indicated streamwise
locations

velocity component is shown as function of y + . While in all the profiles one notices
an inner maximum, at downstream locations, one can also note the effects of con-
vecting vortices, via the presence of another maximum in the outer part.
In Fig. 6.29a–c, the skin friction coefficient C f variation for the section along
z = 0 are shown with the streamwise co-ordinate x, for the indicated times. For
−1/2
laminar flows, skin friction coefficient varies as C f = 0.664 × Rex , where Rex
is Reynolds number based on streamwise co-ordinate x, and shown in the frames
340 6 3D Routes of Transition to Turbulence by STWF

8.0E-03 (a) t = 35.0


GCP exciter with F = 0.5×10-4 −1/5
<CfT> = 0.074×Rex
6.0E-03

4.0E-03 Exciter
Cf
2.0E-03

0.0E+00 CfL = 0.664×Re−1/2


x
5 10 15 x 20 25 30 35 40

8.0E-03 (b) t = 41.0


<CfT> = 0.074×Re−1/5
x
6.0E-03

4.0E-03
Cf
2.0E-03
−1/2
0.0E+00 CfL = 0.664×Rex
5 10 15 x 20 25 30 35 40

8.0E-03 (c) t = 50.0


<CfT> = 0.074×Re−1/5
x
6.0E-03

4.0E-03
Cf
2.0E-03

0.0E+00 CfL = 0.664×Re−1/2


x
5 10 15 x 20 25 30 35 40

8.0E-03 (d) Time averaged between t=40 to 50.


−1/5
<CfT> = 0.074×Rex
6.0E-03

4.0E-03
Cf
2.0E-03

0.0E+00 −1/2
CfL = 0.664×Rex
5 10 15 x 20 25 30 35 40

0.03 (e) t= 40.0 at (z=0)-plane

0.02
y

0.01

0
14 14.2 14.4 14.6 x 14.8 15 15.2 15.4 15.6 15.8

Fig. 6.29 a–c Instantaneous and d time-averaged skin-friction coefficient C f along z = 0 spanwise
station plotted as a function of x. In d Time-averaged C f during the interval 40 ≤ t ≤ 50 are plotted
as function of x. For comparison, corresponding correlations for laminar and turbulent flows are
shown in all these frames. e Stream-trace is shown at t = 40 to indicate the recirculating region at
the onset of transition

as dashed line. For fully developed turbulent boundary layer, mean skin friction
−1/5
coefficient varies as < C f >= 0.74 × Rex , given in [28, 42]. This line is shown
in all the frames by dash-dotted line. The passage of the STWF is clearly evident in all
the frames, as highly intermittent flow, with the value of C f fluctuating significantly.
When we plot the time-averaged skin-friction coefficient (averaged between t = 40
and 50) as a function of x, we note that in the central part of the turbulent spot, the
time-averaged C f displays very good match with the turbulent boundary layer value.
6.5 Formation of Turbulent Spots and Fully Developed Turbulent Flow 341

0.2
(a) t = 30 3.5 HL = 2.59 (f) t = 30
3
0.15 δ*
θ 2.5

δ,θ
0.1
*

H
2 HT ~ 1.4
0.05
1.5

0 10 20 30 40 50 1 10 20 30 40 50
Rex×10−5 Rex×10−5

0.2 (g) t = 35
(b) t = 35 3.5 HL = 2.59
* 3
0.15 δ
θ 2.5
δ,θ

0.1
*

H
2 HT ~ 1.4
0.05
1.5

0 10 20 30 40 50 1 10 20 30 40 50
Rex×10−5 Rex×10−5

0.2 (h) t = 40
(c) t = 40 3.5 HL = 2.59
* 3
0.15 δ
θ 2.5
δ,θ

0.1
*

2 HT ~ 1.4
0.05
1.5

0 10 20 30 40 50 1 10 20 30 40 50
Rex×10−5 Rex×10−5

0.2
(d) t = 45 3.5 HL = 2.59 (i) t = 45
3
0.15 δ*
θ 2.5
δ,θ

0.1
*

2 HT ~ 1.4
0.05
1.5

0 10 20 30 40 50 1 10 20 30 40 50
Rex×10−5 Rex×10−5

0.2
(e) t = 50 3.5 HL = 2.59 (j) t = 50
0.15 3
δ∗
2.5
δ,θ

0.1 θ
*

2 HT ~ 1.4
0.05
1.5

0 10 20 30 40 50 1 10 20 30 40 50
Rex×10−5 Rex×10−5

Fig. 6.30 a–e Displacement thickness δ ∗ , and momentum thickness θ, at z = 0 section plotted at
indicated time instants. f–j Shape factor H at z = 0 section plotted at indicated time instants

One notes from Fig. 6.29, that near the location x  14, the averaged C f shows a
dip, and this is due to the onset of unsteady separation associated with the transition.
This is shown in the stream trace plot shown in Fig. 6.29c.
In Fig. 6.30, corresponding influence on other integrated quantities are shown in
terms of displacement and momentum thickness (δ ∗ and θ ), at the indicated times
on the left column. The right column of this figure shows the variation of the shape
factor (H = δ ∗ /θ ), as a function of Reynolds number based on current length. In these
frames, the dashed lines indicate the typical values of laminar (top) and turbulent
flows (bottom). Once again, one can notice an excellent match with the known
experimental and theoretical trends for these quantities.
342 6 3D Routes of Transition to Turbulence by STWF

(a) Time averaged spectral density


3
10

<kx E11> 102


5/3

z=0
10
1 z = 0.38
z = 0.48

0
10
0 1 2
10 kx 10 10

(b)
0
10
<kx5/3 E22>

-1
10

-2
z=0
10 z = 0.38
z = 0.48

10-3
100 kx 101 102

(c)
2
10
1
10
<kx5/3 E33>

0
10
-1
10
z=0
10-2 z = 0.38
z = 0.48
-3
10
0 1 2
10 kx 10 10

Fig. 6.31 a–c Time averaged compensated streamwise spectral density for streamwise, wall-normal
and spanwise velocity components plotted as a function of streamwise wavenumber k x for z = 0,
0.38 and 0.48

In Fig. 6.31, compensated energy spectra, as defined in Saddoughi and Veeravalli


[29] by E 11 , E 22 and E 33 , are shown along three spanwise locations given by z = 0.0,
0.38 and 0.48, plotted in different frames. These are obtained by squaring the Fourier–
Laplace transform of u d , vd and wd , respectively. For example, E 11 is obtained as,

E 11 = |Ud (k x )|2 (6.33)

where Ud (k x ) is the Fourier–Laplace transform of u d , with respect to the streamwise


wavenumber, k x . Similarly E 22 and E 33 can also be defined. Variation of the com-
pensated energy spectra shows deviation from the homogeneous isotropic turbulence
6.5 Formation of Turbulent Spots and Fully Developed Turbulent Flow 343

values given due to Kolmogorov [28]. From this figure, one notes the existence of an
−5/3
intermediate wavenumber region, where spectral densities vary as k1 – a variation
in the inertial sub-range, predicted for 3D isotropic homogeneous turbulence. More
importantly, displayed computed spectra show similarity with the experimental data
for inhomogeneous flows in [29]. In this latter reference, local isotropy of the bound-
ary layer about the streamwise wavenumber (k x ) is explored experimentally. Thus,
the STWF is noted to take the flow all the way from receptivity to an equilibrium flow,
obtained for local isotropic turbulence, explored experimentally for inhomogeneous
flows by Saddoughi and Veeravalli [29].

References

1. Andersson, P., Berggren, M., & Henningson, D. S. (1999). Optimal disturbances and bypass
transition in boundary layers. Physics of Fluids, 11, 134–150.
2. Ashpis, D. E., & Reshotko, E. (1990). The vibrating ribbon problem revisited. Journal of Fluid
Mechanics, 213, 531–547.
3. Bhaumik, S. (2013). Direct Numerical Simulation of Inhomogeneous Transitional and Turbu-
lent Flows. Ph. D. Thesis, I. I. T. Kanpur, INDIA.
4. Bhaumik, S., & Sengupta, T. K. (2015). A new velocity-vorticity formulation for direct numer-
ical simulation of 3D transitional and turbulent flows. Journal of Computational Physics, 284,
230–260.
5. Brandt, L., & Henningson, D. S. (2002). Transition of streamwise streaks in zero-pressure-
gradient boundary layers. Journal of Fluid Mechanics, 472, 229–261.
6. Chen, L., & Liu, C. (2011). Numerical study on mechanisms of second sweep and positive
spikes in transitional flow on a flat plate. Computers and Fluids, 40, 28–41.
7. Craik, A. D. D. (1971). Non-linear resonant instability in boundary layers. Journal of Fluid
Mechanics, 50, 393–413.
8. Duguet, Y., Schlatter, P., Henningson, D. S., & Eckhardt, B. (2012). Self-sustained localized
structures in a boundary-layer flow. Physical Review Letters, 108, 044501.
9. Durbin, P. A., & Wu, X. (2007). Transition beneath vortical disturbances. Journal of Fluid
Mechanics, 39, 107–128.
10. Eiseman, P. R. (1985). Grid generation for fluid mechanics computation. Annual Review of
Fluid Mechanics, 17, 487–522.
11. Fasel, H., & Konzelmann, U. (1990). Non-parallel stability of a flat-plate boundary layer using
the complete Navier-Stokes equations. Journal of Fluid Mechanics, 221, 311–347.
12. Gaster, M., & Grant, I. (1975). An experimental investigation of the formation and development
of a wave packet in a laminar boundary layer. Proceedings of the Royal Society of London Series
A. Mathematical and Physical Sciences, 347(1649), 253–269.
13. Hama, F. R., & Nutant, J. (1963). Detailed flow-field observations in the transition process in a
thick boundary layer. In Proceedings of the 1963 Heat Transfer and Fluid Mechanics Institute,
(pp. 77–93). Stanford: Stanford University Press.
14. Herbert, Th. (1988). Secondary instability of boundary layers. Annual Review of Fluid Mechan-
ics, 20, 487–526.
15. Hill, D. C. (1995). Adjoint systems and their role in the receptivity problem for boundary
layers. Journal of Fluid Mechanics, 292, 183–204.
16. Jeong, J., & Hussain, F. (1995). On the identification of a vortex. Journal of Fluid Mechanics,
285, 69–94.
17. Kachanov, Y. S. (1994). Physical mechanisms of laminar-boundary-layer transition. Annual
Review of Fluid Mechanics, 26, 411–482.
344 6 3D Routes of Transition to Turbulence by STWF

18. Kachanov, Y. S. (1987). On the resonant nature of the breakdown of a laminar boundary layer.
Journal of Fluid Mechanics, 184, 43–74.
19. Kachanov, Y. S., & Levchenko, V. Y. (1984). The resonant interaction of disturbances at laminar-
turbulent transition in a boundary layer. Journal of Fluid Mechanics, 138, 209–247.
20. Klebanoff, P. S., & Tidstrom, K. D. (1959). Evolution of amplified waves leading to transition
in a boundary layer with zero pressure gradient. In N.A.S.A. Technical Note, D–195.
21. Klebanoff, P. S., Tidstrom, K. D., & Sargent, L. M. (1962). The three-dimensional nature of
boundary-layer instability. Journal of Fluid Mechanics, 12, 1–34.
22. Kovasznay, L. S. G., Komoda, H., & Vasudeva, B. R. (1962). Detailed flow-field in transition.
In Proceedings of the Heat Transfer and Fluid Mechanics Institute, Palo Alto, California:
Stanford University Press.
23. Landahl, M. T. (1975). Wave breakdown and turbulence. SIAM Journal on Applied Mathemat-
ics, 28, 735.
24. Landahl, M. T. (1980). A note on an algebraic instability of inviscid parallel shear flows. Journal
of Fluid Mechanics, 98, 243–251.
25. Luchini, P. (2000). Reynolds-number independent instability of the boundary layer over a flat
surface. Part 2: Optimal perturbations. Journal of Fluid Mechanics, 404, 289–309.
26. Lu, P., Wang, Z., Chen, L., & Liu, C. (2012). Numerical study on U-shaped vortex formation
in late boundary layer transition. Computers and Fluids, 55, 36–47.
27. Mattor, N., Williams, T. J., & Hewett, D. W. (1995). Algorithm for solving tridiagonal matrix
problems in parallel. Parallel Computing, 21, 1769–1782.
28. Pope, S. B. (2000). Turbulent flows. UK: Cambridge University Press.
29. Saddoughi, S. G., & Veeravalli, S. V. (1994). Local isotropy in turbulent boundary layers at
high Reynolds number. Journal of Fluid Mechanics, 268, 333–372.
30. Saric, W. S., & Thomas, A. S. W. (1984). Experiments on the subharmonic route to turbulence
in boundary layers. In T. Tatsumi (Ed.), Turbulence and Chaotic Phenomena in Fluids. Elsevier,
USA: North Holland.
31. Sayadi, T., Hamman, C. W., & Moin, P. (2013). Direct numerical simulation of complete H-
type and K-type transitions with implications for the dynamics of turbulent boundary layers.
Journal of Fluid Mechanics, 724, 480–509.
32. Schoppa, W., & Hussain, F. (2002). Coherent structure generation in near-wall turbulence.
Journal of Fluid Mechanics, 453, 57–108.
33. Schubauer, G. B., & Skramstad, H. K. (1947). Laminar boundary layer oscillations and the
stability of laminar flow. Journal of the Aeronautical Sciences, 14, 69–78.
34. Sengupta, T. K., Bhaumik, S., & Bose, R. (2013). Direct numerical simulation of transitional
mixed convection flows: Viscous and inviscid instability mechanisms. Physics of Fluids, 25(9),
094102. (1994-present).
35. Sengupta, T. K., Bhaumik, S., & Bhumkar, Y. G. (2012). Direct numerical simulation of
two-dimensional wall-bounded turbulent flows from receptivity stage. Physical Review E, 85,
026308.
36. Sengupta, T. K., Bhaumik, S., & Usman, S. (2011). A new compact difference scheme for
second derivative in non-uniform grid expressed in self-adjoint form. Journal of Computational
Physics, 230(5), 1822–1848.
37. Sengupta, T. K., De, S., & Sarkar, S. (2003). Vortex-induced instability of incompressible
wall-bounded shear layer. Journal of Fluid Mechanics, 493, 277–286.
38. Sengupta, T. K., Dipankar, A., & Rao, A. K. (2007). A new compact scheme for parallel
computing using domain decomposition. Journal of Computational Physics, 220, 654–677.
39. Sengupta, T. K., Lele, S. K., Sreenivasan, K. R., & Davidson, P. A. (2015). Advances in
Computation, Modeling and Control of Transitional and Turbulent Flows. In IUTAM Symposia
Proceedings, Singapore: World Scientific Publishing Company.
40. Singer, B. A., & Joslin, R. D. (1994). Metamorphosis of a hairpin vortex into a young turbulent
spot. Physics of Fluids, 6, 3724–3736.
41. Sengupta, T. K., Rao, A. K., & Venkatasubbaiah, K. (2006). Spatio-temporal growing wave
fronts in spatially stable boundary layers. Physical Review Letters, 96(22), 224504.
References 345

42. Tennekes, H., & Lumley, J. L. (1972). A first course in turbulence. Cambridge, MA: MIT Press.
43. Würz, W., Sartorius, D., Kloker, M., Borodulin, V. I., & Kachanov, Y. S. (2012). Detuned
resonances of Tollmien-Schlichting waves in an airfoil boundary layer: Experiment, theory,
and direct numerical simulation. Physics of Fluids, 24, 094103.
44. Wu, X., Jacobs, R. G., Hunt, J. C. R., & Durbin, P. A. (1999). Simulation of boundary layer
transition induced by periodically passing wakes. Journal of Fluid Mechanics, 399, 109–153.
45. Zelman, M. B., & Maslennikova, I. I. (1993). Tollmien-Schlichting-wave resonant mechanism
for subharmonic-type transition. Journal of Fluid Mechanics, 252, 449–478.
Appendix A

A.1 Boundary Layer Equation for Mixed Convention


Problem

The physical problem considered is a 2D flow of a fluid over a semi-infinite inclined


flat plate with the leading edge as the stagnation point. The free-stream velocity and
temperature is U∞ and Tin f t y , respectively. The plate is inclined at an angle β with
the direction of the flow as shown schematically in Fig. A.1. This implies that the
velocity at the edge of the boundary layer is different from Uin f t y and therefore,
is a function of the streamwise coordinate x ∗ . Let it be denoted as Ue∗ (x ∗ ). The
surface of the plate is maintained at a temperature Tw∗ (x ∗ ). The plate temperature is
greater or less than T∞ depending upon the plate under consideration is hot or cold.
Here, all the dimensional quantities are represented with an asterisk as superscript,
whereas those without the superscript asterisk symbol denote the non-dimensional
quantities. The governing equations for this case is given by the incompressible
Navier–Stokes equation subjected to boundary-layer approximation [1, 3], where
Boussinesq approximation is used to account for the buoyancy effects due to heat
transfer. One also has to consider an additional energy equation as the governing
equation for the temperature field. These equations in the dimensional form are
given as,

∂u ∗ ∂v∗
+ =0 (A.1)
∂x∗ ∂ y∗
∂u ∗ ∂u ∗ ∂ p∗ ∂ 2u∗
u ∗ ∗ + v∗ ∗ = − ∗ + ν ∗ 2 (A.2)
∂x ∂y ∂x ∂y

∂p
− gβT (T − T∞ ) = 0 (A.3)
∂ y∗
∂T ∗ ∂T ∗ ∂2T ∗
u ∗ ∗ + v ∗ ∗ = αT ∗ 2 (A.4)
∂x ∂y ∂y

© Springer Nature Singapore Pte Ltd. 2019 347


T. K. Sengupta and S. Bhaumik, DNS of Wall-Bounded Turbulent Flows,
https://doi.org/10.1007/978-981-13-0038-7
348 Appendix A

Fig. A.1 Schematic diagram for the mixed-convection flow over a semi-infinite inclined flat plate.
The plate is inclined with the free-stream velocity direction at an angle β. Here, δ and δT are the
hydrodynamic and thermal boundary layer thickness, respectively

where ν, g, βT and αT denote the kinematic viscosity, gravitational acceleration,


volumetric thermal expansion coefficient and the thermal diffusivity, respectively.
Note that while deriving Eqs. (A.1)–(A.4), we have not considered viscous dissipation
of kinetic energy and also neglected any possible presence of source for internal heat
generation. The relevant boundary conditions are given as

u ∗ (x ∗ , y ∗ ) = v∗ (x ∗ , y ∗ ) = 0 and T ∗ (x ∗ , y ∗ ) = Tw∗ (x ∗ )at y ∗ = 0 (A.5)


u ∗ (x ∗ , y ∗ ) → Ue∗ (x ∗ ) and T ∗ (x ∗ , y ∗ ) → T∞ asy ∗ → ∞ (A.6)

To nondimensionalize Eqs. (A.1)–(A.4), a length scale (L), a velocity scale (U∞ ),


a temperature scale (ΔTL = Tw∗ (L) − T∞ ) and a pressure scale (ρU∞
2
) are adopted.
The nondimensionalized form of Eqs. (A.1)–(A.4) are given as,

∂u ∂v
+ =0 (A.7)
∂x ∂y
∂u ∂u ∂p 1 ∂ 2u
u +v =− + (A.8)
∂x ∂y ∂x Re ∂ y 2
∂p Gr
− θ =0 (A.9)
∂y Re2
∂θ ∂θ 1 1 ∂ 2θ
u +v = (A.10)
∂x ∂y Re Pr ∂ y 2

where, θ = (T ∗ − T∞ )/ΔTL , Gr = gβT ΔTL L 3 /ν 2 is the Grashof number, Re =


U∞ L/ν is the Reynolds number and Pr = ν/αT is the Prandtl number. The Grashof
number (Gr ) gives the ratio of buoyancy and viscous forces present in the fluid and
the Richardson number (Ri), given by Ri = Re Gr
2 , indicates the relative dominance

of natural to forced convection. For such cases, Ri ≥ 0 or Ri ≤ 0 refer to assisting


and opposing flows. In the mixed convection regime Ri is of order one. In general,
Pr = 0.71 is used, which is the Prandtl number for air as the working medium. Fol-
Appendix A 349

lowing Eqs. (A.5)–(A.6), corresponding boundary conditions for the nondimensional


Eqs. (A.7)–(A.10) are given as

u(x, y) = v(x, y) = 0 and θ (x, 0) = Θw (x) at y = 0 (A.11)


u(x, y) → Ue (x) and θ (x, y) → 0 as y → ∞ (A.12)

where Θw (x) = (Tw∗ (x ∗ ) − Tin f t y )/ΔTL . For flows with high Reynolds number, the
boundary layer thickness is orders of magnitude smaller than the dimensions of the
plate. Therefore, it is customary to scale up the boundary layer thickness, so that both
x- and y-scales are of similar orders of magnitude. This can be accomplished by √fol-
lowing coordinate
√ transformation from (x, y) to (X, Y ). Let, X = x, Y = y Re,
U = u; V = v Re andP = p. Following these coordinate and variable transfor-
mations, Eqs. (A.7)–(A.10) changes to

∂U ∂V
+ =0 (A.13)
∂X ∂Y
∂U ∂V ∂P ∂ 2U
U +V =− + (A.14)
∂X ∂Y ∂X ∂Y 2
∂P
− Kθ = 0 (A.15)
∂Y
∂θ ∂θ 1 ∂ 2θ
U +V = (A.16)
∂X ∂Y Pr ∂Y 2

where K = Gr /Re5/2 . The corresponding boundary conditions are

U (X, 0) = V (X, 0) = 0, and θ (X, 0) = w (X ) as


Y → ∞, U (X, Y ) → Ue (X )and θ (X, Y ) → 0 (A.17)

One notes from Eqs. (A.14) and (A.15) that at the free-stream the quantity P + 21 Ue2 is
function of the wall-normal coordinate Y only. This prompts one to define a modified
pressure as

1
P̃ = P + Ue (X )2 (A.18)
2
Now, one can also define the streamfunction ψ such that

∂ψ
U= (A.19)
∂Y
∂ψ
V =− (A.20)
∂X
350 Appendix A

Equations (A.19) and (A.20) automatically satisfies the continuity equation (A.13).
Substituting Eqs. (A.18)–(A.20) into (A.13)–(A.16) one gets,

∂ψ ∂ 2 ψ ∂ψ ∂ 2 ψ ∂ P̃ dUe ∂ 3ψ
− = − + U e + (A.21)
∂Y ∂ X ∂Y ∂ X ∂Y 2 ∂X dX ∂Y 3
∂ P̃
− Kθ = 0 (A.22)
∂Y
 
∂ψ ∂θ ∂ψ ∂θ 1 ∂ 2θ
− = (A.23)
∂Y ∂ X ∂ X ∂Y Pr ∂Y 2

The appropriate boundary conditions for Eqs. (A.21)–(A.23) are given as

∂ψ
at Y = 0 ψ(X, Y ) = 0, = 0 and θ (X, 0) = Θw
∂Y
∂ψ
→ Ue (X ) and θ → 0 as Y → ∞
∂Y

A.2 Similarity Transformation

The similarity transformation [2, 4] is performed following the coordinate √ transfor-


mation from
√ (X, Y )- to (ξ, η)-coordinate system where ξ = X and η = Y Ue / X .
Let ψ = Ue X f (ξ, η), P̃ = Ue2 q(ξ, η), θ = Θw g(ξ, η). Let us further assume that
the Ue (X ) and Θw (X ) varies as X n and X r , respectively [4]. Therefore,

X dUe X d w
= n and =r
Ue dX w dX

These transformations lead to the following relationships

  
∂ψ 1 Ue 
= (n + 1) f + η(n − 1) f + 2X f ξ
∂X 2 X
  
∂ 2ψ Ue   1 
= X f ξ + n f + η(n − 1) f
∂ X ∂Y X 2

∂ψ Ue 
= Ue X f  = Ue f 
∂Y X
  3/2
∂ 2ψ Ue  Ue
= U e X f = √ f 
∂Y 2 X X
 3/2 
∂ 3ψ Ue  U 2
e 
= U e X f = f
∂Y 3 X X
Appendix A 351
 
∂ P̃ X 1 
= 2 Xqξ + 2nq + η(n − 1)q
∂X Ue 2
5/2
∂ P̃ Ue
= 1/2 q 
 ∂Y X 
∂θ Θw 1
= rg + η(n − 1)g  + Xgξ
∂X X 2

∂θ Ue
= Θw g 
∂Y X
∂ 2θ Ue
= Θw g  (A.24)
∂Y 2 X

In Eq. (A.24), the subscript ξ and the prime ( ) denote the partial derivatives with
respect to ξ and η, respectively. Substituting Eq. (A.24) in Eqs. (A.21)–(A.23), one
gets

   
 n+1  2 1 
f + f f + n(1 − f ) − 2nq + η(n − 1)q
2 2
= ξ( f  f ξ − f ξ f  + qξ ) (A.25)

5/2
Ue
q − K g = 0 (A.26)
ξ 1/2 Θw
 
1  n+1
g − r f g + f g  = ξ( f  gξ − f ξ g  ) (A.27)
Pr 2

For self-similarity, Eqs. (A.25)–(A.27) would be independent of ξ . Therefore all the


partial derivatives with respect to ξ would be zero. Following Eq. (A.26), one also
requires that

5/2
Ue
= Const. (A.28)
ξ 1/2 Θw

As Ue X n and Θw X r , Eq. (A.28) indicates that

5n − 1
r=
2
Following these restrictions, one obtains the self-similar equations for the bound-
ary layer profiles for mixed convection flow past a flat-plate as
352 Appendix A
     
n+1 1
f  + f f  + n(1 − f 2 ) − 2nq + η(n − 1)q  = 0 (A.29)
2 2
q  − K g = 0 (A.30)
   
1  5n − 1  n+1
g − f g+ f g  = 0 (A.31)
Pr 2 2

The appropriate boundary conditions for these nonlinear ordinary differential equa-
tions are given following Eqs. (A.25)–(A.27) as

f (0) = f  (0) = 0, and g = 1 (A.32)


f  → 1, and g → 0 as η → ∞ (A.33)

If the edge velocity is constant i.e., Ue = Const. = U∞ then n = 0 (r = −1/2), and


Eqs. (A.29)–(A.31) simplifies to [2]

1  1
f  +
f f + K ηg = 0 (A.34)
2 2
1  1 
g + ( f g + f g) = 0 (A.35)
Pr 2
Equations (A.34) and (A.35) and the boundary condition f (0) = 0 show that at
η = 0, g  = 0 i.e., the flat plate surface has an adiabatic wall condition. This is true
for all streamwise locations x > 0 except at the leading edge (x = 0), where all heat
transfer occurs singularly while the wall temperature varies as x −1/2 for both hot and
cold plate cases.

References

1. Schlichting, H. (1933). Zur entstehung der turbulenz bei der plattenströmung.


Nach. Gesell. d. Wiss. z. Gött., MPK, 42, 181–208.
2. Schneider, W. (1979). A similarity solution for combined forced and free con-
vection flow over a horizontal plate. International Journal of Heat and Mass
Transfer, 22, 1401–1406.
3. White, F. M. (2008). Fluid mechanics (6th ed.) New York: The McGraw Hill
Companies.
4. Mureithi, E. W., & Denier, J. P. (2010). Absolute-convective instability of
mixed forced-free convection boundary layers. Fluid Dynamic Research, 42(5),
055506.
Index

A Boussinesq approximation, 1, 17, 137, 139,


Abel’s theorems, 186 153, 347
Absolute instability, 275 Boussinesq approximation, in Navier-Stokes
Adaptive filter 2D, 86, 94, 95 equation, 162
Adverse pressure gradient boundary layer, Bromwich Contour Integral Method
224, 239, 284 (BCIM), 9, 18, 132, 136, 265
Aliasing error, 49, 77, 86, 95, 114, 234, 312 Bromwich contours, 129
Amplitude functions, 259, 268 Buffer layer, 31, 337
Anomalous modes of first kind, 266 Bulk viscosity, 21
Anomalous modes of second kind, 268, 269 Buoyancy parameter, 139, 140
Anti-diffusion, 37, 44, 51 buoyancy parameter, critical, 137
Archimedes number, 140 Bypass transition, 67, 150, 210, 212, 229,
Asymptotic solution, 132 237, 284, 288, 301, 322
Bypass transition, micro-separation, 199

B C
Bi-CGSTAB Cauchy equations, 20
algorithm, 156 Cauchy’s integral theorem, 133
iterative scheme, 98 C D2 scheme, 35, 56, 71, 103
method, 233 C D4 scheme, 35
solver, 100 Central interior filters, 78
Bilateral Laplace transform, 130 CFL number, 63
Biot–Savart law, 232, 234 Characteristic determinant, 129, 164
Blasius boundary layer, 223 Combined Compact Differencing (CCD)
Blasius equation, 122, 188, 280 schemes, 48
Bluff-body flow, 9, 264 Compact schemes first derivative, 39, 75,
Boundary closure schemes, 43, 44, 53, 79 234, 313
Boundary filters, 78 Compact schemes for second derivative, 44
Boundary layer disturbance, inner maxi- Complex transfer function, 81
mum, 204 Compound Matrix Method (CMM), 128,
Boundary layer equation, 122 136, 164
boundary layer, displacement thickness, Conservative form of VTE, 28
125 Consistency condition of compact schemes,
boundary layer, momentum thickness, 40, 45, 90
124 Continuity equation, 25, 28
boundary layer thickness, 124 Continuous spectrum, 286, 295
© Springer Nature Singapore Pte Ltd. 2019 353
T. K. Sengupta and S. Bhaumik, DNS of Wall-Bounded Turbulent Flows,
https://doi.org/10.1007/978-981-13-0038-7
354 Index

Contravariant velocity components, 27 Enstrophy based POD, 252


Control-mass system, 19 Enstrophy cascade, 29, 244, 246
Convecting vortex, 92, 233, 234, 322, 339 Enstrophy transport equation, 242, 243
Correlation function, 252, 255 Entropic disturbances, 8
Critical layers, 193 Equilibrium flow, 311
Critical Reynolds number, 6, 98, 223 Equivalent amplification factor, 79
Cumulative enstrophy, 259 Error equation
-form, 61
-form, 61
D Error metrics, 95, 102
Deep water waves, 34 Error propagation equation, 65, 76
Derived variable formulation, 24, 27, 83 Essential singularity, 186
Diracdelta function, 128
Dirichlet function, 136
Discrete Fast Fourier Transform (DFFT), F
131 Falkner–Skan equation, 122
Dispersion error, 31, 75, 100, 234 Falkner–Skan pressure gradient parameter,
Dispersion relation, 13, 76, 126, 127, 164, 239, 301
259 Fast Fourier Transform (FFT), 259
Dispersion Relation Preserving (DRP) meth- Favorable pressure gradient, location, 146
ods, 61, 67, 75 Filters, 1D, 78, 83, 86
Displacement thickness δ ∗ , 4, 84, 234, 262, Filters for DNS and LES, 76
278, 286, 312, 341 Filters upwinded, 1D, 83
Disturbance energy equation, 226 Finite volume method, 68
Disturbance enstrophy, 227, 257, 266 Fjφrtoft criterion, 174
Disturbances, 2D field, 126 Fjφrtoft integrand ( F ), 174, 176, 178
disturbance, inviscid governing equa- Flow over cold adiabatic plate, 166
tions, 172 Flow over iso-thermal wall, 151
Disturbances vortical, entropic,acoustic, Flow past a cylinder, 282
223, 224 Flows past a natural laminar aerofoil, 95
Divergence–free or solenoidal field, 19, 229 Fourier-Laplace transform, 125, 342
DNS equation, mixed convection flow, 153 Fourier-Laplace transform, bilateral, 133
DNS, mixed convection flows, 138, 149, 179 Fourier-Laplace transform, unilateral, 129
DNS of flows with heat transfer, 151 Free stream convecting vortex, 237, 284
DNS solution, use of buffer domain, 157 Free stream excitation, 224
DNS spatial discretization, OUCS3 scheme, FST vortices, 237
157
DNS time discretization, ORK3 scheme, 157
Downstream propagating disturbances, 130 G
DRP properties, 53, 68 Galerkin methods, 68, 69
Dynamics, 150 Galerkin projection, 252
amplifier type dynamics, 150 Gaussian Circular Patch (GCP) exciter, 307,
resonator type dynamics, 150 314
Goldstein singularity, 234
Grashof number, 139, 162
E Grid staggering, 103
Effects of nonlinearity and nonparallelism, Group velocity, 14, 131
195
Eigenfunction, 252, 262
Eigenvalue analysis, grid search technique, H
137 Heated isothermal wedge, 149, 166
Eigenvalue analysis, mixed convection flow, Helmholtz decomposition, 26
163 High amplitude wall-excitation, 288
Eigenvalue problem, hydrodynamics, 126 Hilbert–Schmidt theory, 252
Index 355

Homogeneous isotropic turbulent flows, 342 L


Hybrid Fourier–Laplace transform, 69 Lambda (λ2 ) iso-surfaces, 335
Laplace transform, 12
bilateral, 86, 130
I Laplacian and rotational forms of VTE, 105
Inertial subrange, 30 Leading edge contamination, 224
Inflection points, 3, 142, 147, 158, 224 Lid driven cavity, 104
inflection point, location, 149 Linear basis functions FEM, 69
Instability, 8, 229 Linear instability of mixed convection flows,
bluff body flows, 6, 224, 264 162
free stream modes, 6, 248 Linearized perturbation equation, parallel
nonlinear effects, 282 flow approximation, 124
nonparallel effects, 257 Linear spatial viscous instability, 161
wall mode, 6 Linear stability equation, 124
Instability, absolute, 150 Linear temporal viscous instability, 165
Instability, convective, 150 Localized wall excitation, 136, 189
instability, Tollmien-Schlichting type, Local or near-field solution, 135, 186, 189
138 local solution, inner layer, 135
instability, vortex type, 138 Local solution, 276, 314
instability, wave-like, 138 Logarithmic law of the wall, 337
Intermittent zone, 337 Low-Dispersion Runge-Kutta (LDDRK)
Interpolation in staggered grids, 52 schemes, 75
Inverse energy cascade, 29 Low frequency excitation, 292, 296
Inverse Fourier-Laplace transform, 129
Inviscid governing equation, disturbance
M
field, 175
Mechanical pressure, 20, 21
Inviscid governing equation, Rayleigh’s
Method of snapshots POD, 255
equation, 178
Micro-bubbles, 300
Inviscid instability mechanism, 172
Micro-separation, 195, 199
Inviscid instability theorems, 180
Mixed convection boundary layer, 137, 146,
Inviscid temporal instability, 147, 151, 173,
153
174
Mixed convection flow, 137, 150, 165
Rayleigh’s and Fjφrtoft’s theorem, 149, Mixed convection flow equation with
172 Boussinesq approximation, 153
Rayleigh’s inflection-point theorem, 3 Mixed convection flow, isothermal flat plate,
theorem I, 175, 180 151
theorem II, 175, 181 Mixed convection flow, over adiabatic flat
Inviscid temporal instability of mixed con- plate, 141
vection flow, 173 Mixed convection flow over a wedge, 152
Iso-thermal heated wedge, 165, 172, 182 Mixed convection flow theorems I, II, 173,
Iso-thermal wall case, 151 174
Modified wavenumber, 38, 48, 87
Molecular viscosity, 20
J Momentum conservation equation, 19
Jordan’s lemma, 134 Momentum thickness, 307
Multi-time periodic modes, 270

K
Kelvin–Helmholtz instability, 10, 14 N
Kinetic energy, 2, 252, 348 Navier–Stokes Equation (NSE), 5, 278
Kolmogorov scale, 29, 30 Navier–Stokes equation (NSE)


Koopman modes, 252, 270 ( p, V )− formulations, 101
Kronecker delta function, 20 rotational form, 225, 238
356 Index

Navier-Stokes Equation (NSE), Mixed con- POD amplitude functions, 259


vection flow, 162 POD analysis, 252, 266
Navier-Stokes Equation (NSE), without heat POD modes, 252, 255, 265
transfer DNS, 185 anomalous mode of second kind (T2 ),
NCCD scheme, 51 264
Near-field response field, 135 first kind T1 modes, 259
Neutral curve, 280 regular or R modes, 258
Newtonian fluids, 20 shift mode, 258, 259
Non-dimensional phase speed, 64 Poisson equation of velocity, 28, 100, 231
Nonlinear and nonparallel effects, 186, 190 Prandtl number, 2, 140, 348
Nonlinear effects, 186 Primitive variable formulation, 23, 103
nonlinear receptivity, ZPG boundary
layer, 192
nonparallel effects, 186, 192, 195 Q
Nonlinear instability for incompressible Quadratic basis functions FEM, 69, 70
flows, 225 QUICK scheme, 68
Nonlinear receptivity theory, 240 q−waves, 65, 67, 75
Non-periodic problems, 43, 47, 49, 80
Numerical amplification factor, 61, 72, 75
Numerical discretization, 34
R
Numerical filtering, 76
Rayleigh integrand (), 176, 178
Numerical phase speed, 64
Rayleigh’s inflection-point theorem, 124
Nyquist limit, 39, 53, 81, 88
Receptivity by DNS, 185
Receptivity, effects of amplitude of excita-
O tion, 200
One dimensional convection equation, 32 Receptivity, effects of frequency of excita-
Optimized compact mid–point interpolation tion, 216
scheme, 232 Receptivity, effects of location of exciter,
Optimized Staggered Compact Schemes 213
(OSCS), 313 Receptivity of Blasius boundary layer, 128,
Optimized three-stage Runge-Kutta method, 131
157 Receptivity of cold adiabatic flat Plate, 169
Optimized upwind compact scheme Receptivity of mixed convection flows, 168
OUCS3, 232 Receptivity theory, 121
sixth order Orr-Sommerfeld equation, receptivity of Blasius boundary layer,
137, 164 131
O R K 3 method, 277 receptivity of cold adiabatic flat plate
Orr–Sommerfeld modes cases, 169
inviscid mode, 127 receptivity of hot isothermal wedge case,
viscous mode, 127 172
Orr-Sommerfeld Equation (OSE), fourth receptivity of mixed convection flows,
order, 121, 125 168
Orr-Sommerfeld Equation (OSE), sixth Receptivity to SBS excitation, 190
order, 137, 163, 164 Receptivity to wall excitation, 5
Reduced order modeling, 252
Resonator and amplifier dynamics, 150
P Reynolds number, 2–4, 30–32, 83, 95, 97,
Padé schemes, 39 98, 104, 139, 162, 224, 228, 233, 234,
Padé type filters, 77, 86 242, 252, 256, 277, 278, 300, 309,
Parallel flow approximation, 131 339, 341, 348, 349
Phase speed, 33, 61, 64 Richardson number, 2, 140, 162, 348
Physical dispersion relationship, 31 R K 4 method, 68, 277
Plane Poisseuille flow, 224 Role of upwinding, 37
Index 357

Rotating and translating circular cylinder, Stokes’ problem, 135


227 Stream function equation, 27
Rotational form of VTE, 24, 102 Stream function–vorticity formulation, 26
Roughness effects, 224 Streamwise pressure gradient induced, 147
Routes of flow transition, 334 Stuart–Landau–Eckhaus equation, 264, 265,
K –type transition, 334 282
H– or N-type transition, 334 STWF, genesis of, 136
STWF, mixed convection flow, 168
STWF, secondary upstream, 172
S STWF, self-regeneration mechanism, 171
SBS harmonic exciter, 211 SUPG method, 70, 71, 74, 75
SBS strip exciter, 83, 256, 280
Scale factors of the transformation, 231
T
Schneider’s similarity solution, 141
Tangent hyperbolic grid, 155, 187
Schwarz domain decomposition technique,
Tauber’s theorem, 135, 186, 269
278
Temporal amplification theory, 126
Second coefficient of viscosity, 20
Temporal viscous theory for mixed convec-
Self–Adjoint Compact Scheme (SACD), 56,
tion flows, 166
60
Theorem for inviscid temporal instability,
SGS or Leonard stresses, 77
175
Shear sheltering, 150
Thermal diffusivity, 139
Signal problem, 128, 186, 268
Thermodynamic pressure, 20, 21
Signal speed, 33
Thomas’ tridiagonal matrix algorithm, 40
Similarity solutions, 168
Tollmien-Schlichting instability analysis,
Similarity transform, 142
126
Similarity transformations, Blasius profile, Tollmien-Schlichting (TS) wave, 4, 276
122 Total enstrophy, 246, 263
Similarity transformations, Schneider’s pro- Total mechanical energy, 225, 322
file, 141 Transfer function, 81, 90, 91
Similarity transformations, Wedge flow pro- Transitional flows, 18, 97
file, 151 Transition of mixed convection flows, 137
Simultaneous Blowing and Suction (SBS) Tridiagonal compact scheme, 40, 41, 45
excitation, 168 TS wave-packet, 187, 198, 216, 276, 277,
Singular perturbation theory, 135 331
Skin friction coefficient C f , 307, 339 Turbulent Kinetic Energy (TKE), 30, 225
SLDC problem, 98 Turbulent spot, 321, 330, 337
Solenoidality condition, 25, 102, 233 Two–dimensional filter, 89
Solenoidality error, 100, 108 Two–dimensional higher order filters, 86
Sommerfeld boundary conditions, 232, 279, Two–dimensional turbulence, 5, 286
310
Space-time discretization, 8, 62, 67
Spanwise modulated exciter, 327, 329, 334 U
Spatial amplification theory, 126 Unsteady Bernoulli’s equation, 11
Spatio-temporal receptivity approach, 128 Unsteady separation bubbles, 83, 237, 288
Spatio-Temporal Wave-Front (STWF), 5, Unsteady separation bubbles, bypass transi-
83, 121, 132, 136, 171, 276 tion, 199, 205
Spectral amplification factor G, 62 Unsteady separations, 224, 276, 301
Spectral resolution, 38, 42, 234
Spot regeneration mechanism, 289, 292
Staggered grid arrangement, 51, 232 V
Stiff ordinary differential equation, 122 Vector potential, 26, 27
Stokes’ equation, 226 Velocity of energy propagation, 33
Stokes’ hypothesis, 21 Velocity Poisson equation, 231, 309
358 Index

Velocity potential, 10, 12 2 type, 337


Velocity–vorticity formulation, 25, 101, 102, Vorticity, 25
230 Vorticity Transport Equation (VTE), 24, 230
Velocity–vorticity formulations for 2D
flows, 97, 102, 230
Vibrating ribbon experiment, 9 W
Viscous dissipation term, energy equation,
Wall–normal variation of u d and vd , 204
140
Wave cancellation, 223
Viscous sub-layer, 337
Wave front, leading STWF, 189
Volumetric thermal expansion coefficient,
Wave–induced stresses, 282, 288
139
Wedge flow, isothermal case, 151
vortex-induced instability, 168
Von Neumann analysis, 65
Vortex-induced instability, 6, 229, 239, 246
Vortex stretching, 244, 245 Z
Vortical eruptions, 337 Zero-crossing point, 178
Vortical excitation, 237 inner zero-crossing point, 178
Vortices, 224, 227, 228 outer zero-crossing point, 179
- type, 335 ZPG boundary layer, 211

You might also like