0% found this document useful (0 votes)
23 views61 pages

Algebraic Topology A Toolkit 1st Edition Kevin P. Knudson PDF Available

Algebraic Topology: A Toolkit by Kevin P. Knudson is a streamlined introduction to essential elements of algebraic topology, designed for a one or two-semester course. The book covers topics such as homology, cohomology, and homotopy theory, and includes exercises and projects for deeper exploration. It aims to provide a solid foundation for further study in the field, catering to both emerging and established mathematicians.

Uploaded by

moerisabie5736
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)
23 views61 pages

Algebraic Topology A Toolkit 1st Edition Kevin P. Knudson PDF Available

Algebraic Topology: A Toolkit by Kevin P. Knudson is a streamlined introduction to essential elements of algebraic topology, designed for a one or two-semester course. The book covers topics such as homology, cohomology, and homotopy theory, and includes exercises and projects for deeper exploration. It aims to provide a solid foundation for further study in the field, catering to both emerging and established mathematicians.

Uploaded by

moerisabie5736
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/ 61

Algebraic Topology A Toolkit 1st Edition Kevin P.

Knudson pdf available

Now available at ebookfinal.com


( 4.7/5.0 ★ | 498 downloads )

https://ebookfinal.com/download/algebraic-topology-a-toolkit-1st-
edition-kevin-p-knudson/
Algebraic Topology A Toolkit 1st Edition Kevin P. Knudson
Pdf Download

EBOOK

Available Formats

■ PDF eBook Study Guide Ebook

EXCLUSIVE 2025 ACADEMIC EDITION – LIMITED RELEASE

Available Instantly Access Library


We believe these products will be a great fit for you. Click
the link to download now, or visit ebookfinal
to discover even more!

Algebraic Topology 1st Edition C. R. F. Maunder

https://ebookfinal.com/download/algebraic-topology-1st-edition-c-r-f-
maunder/

An Alpine Expedition Through Algebraic Topology Christian


Ausoni

https://ebookfinal.com/download/an-alpine-expedition-through-
algebraic-topology-christian-ausoni/

Foundations Of Algebraic Topology 1, Second Printing


Edition Samuel Eilenberg

https://ebookfinal.com/download/foundations-of-algebraic-
topology-1-second-printing-edition-samuel-eilenberg/

A History of Algebraic and Differential Topology 1900 1960


1, Reprint of the 1989 edition Edition Jean Dieudonné

https://ebookfinal.com/download/a-history-of-algebraic-and-
differential-topology-1900-1960-1-reprint-of-the-1989-edition-edition-
jean-dieudonne/
Dynamic Mechanical Analysis A Practical Introduction
Second Edition Kevin P. Menard (Author)

https://ebookfinal.com/download/dynamic-mechanical-analysis-a-
practical-introduction-second-edition-kevin-p-menard-author/

Historicising Gender and Sexuality 1st Edition Kevin P.


Murphy

https://ebookfinal.com/download/historicising-gender-and-
sexuality-1st-edition-kevin-p-murphy/

Monoidal Topology A Categorical Approach to Order Metric


and Topology 1st Edition Dirk Hofmann

https://ebookfinal.com/download/monoidal-topology-a-categorical-
approach-to-order-metric-and-topology-1st-edition-dirk-hofmann/

Geometry with an Introduction to Cosmic Topology 2018th


Edition Michael P. Hitchman

https://ebookfinal.com/download/geometry-with-an-introduction-to-
cosmic-topology-2018th-edition-michael-p-hitchman/

Surface Topology Third Edition Peter A Firby

https://ebookfinal.com/download/surface-topology-third-edition-peter-
a-firby/
Algebraic Topology A Toolkit 1st Edition Kevin P.
Knudson Digital Instant Download
Author(s): Kevin P. Knudson
ISBN(s): 9783111014814, 3111014819
Edition: 1
File Details: PDF, 5.06 MB
Year: 2024
Language: english
Kevin P. Knudson
Algebraic Topology
Also of Interest
Differential Equations, Fourier Series, and Hilbert Spaces
Lecture Notes at the University of Siena
Raffaele Chiappinelli, 2023
ISBN 978-3-11-129485-8, e-ISBN (PDF) 978-3-11-130252-2

Differential Equations
A first course on ODE and a brief introduction to PDE
Shair Ahmad, Antonio Ambrosetti, 2023
ISBN 978-3-11-118524-8, e-ISBN (PDF) 978-3-11-118567-5

Dynamics of Discrete Group Action


The Legacy of David R. Adams
Boris N. Apanasov, 2024
ISBN 978-3-11-078403-9, e-ISBN (PDF) 978-3-11-078410-7
in: Advances in Analysis and Geometry
ISSN 2511-0438, e-ISSN 2511-0543

Topics in Complex Analysis


Dan Romik, 2023
ISBN 978-3-11-079678-0, e-ISBN (PDF) 978-3-11-079681-0

Advanced Mathematics
An Invitation in Preparation for Graduate School
Patrick Guidotti, 2022
ISBN 978-3-11-078085-7, e-ISBN (PDF) 978-3-11-078092-5
Kevin P. Knudson

Algebraic Topology


A Toolkit
Mathematics Subject Classification 2020
Primary: 55-01; Secondary: 55N10, 55N31, 55P10, 55R20

Author
Prof. Kevin P. Knudson
University of Florida
Department of Mathematics
PO Box 118105
Gainesville FL 32611
USA
[email protected]

ISBN 978-3-11-101481-4
e-ISBN (PDF) 978-3-11-101485-2
e-ISBN (EPUB) 978-3-11-101486-9

Library of Congress Control Number: 2024938448

Bibliographic information published by the Deutsche Nationalbibliothek


The Deutsche Nationalbibliothek lists this publication in the Deutsche Nationalbibliografie;
detailed bibliographic data are available on the Internet at http://dnb.dnb.de.

© 2024 Walter de Gruyter GmbH, Berlin/Boston


Cover image: Kevin P. Knudson
Typesetting: VTeX UAB, Lithuania
Printing and binding: CPI books GmbH, Leck

www.degruyter.com

This one goes out to the one I love...
Preface
When I entered my PhD program in the fall of 1991, I was pretty sure I would study dif-
ferential geometry, even though I had spent a lot of time as an undergraduate learning
general topology and a lot of algebra. My first semester did not clarify things for me–the
standard first course in measure theory, more algebra, a topics course in algebraic ge-
ometry that was a bit advanced for me at the time. My mathematical trajectory changed
forever that spring, however, when I was introduced to algebraic topology. Finally, here
was a branch of mathematics that appealed to my topological interests and my affinity
for algebraic calculation. I was hooked.
Our professor, Richard Hain (who would eventually become my PhD advisor), told
very good stories. His geometric intuition was (and still is) otherworldly, but he brought
it down to earth. I still have the course notes I took that semester, and they formed the
basis for the notes I have used when I teach algebraic topology. They have expanded to
include more material and topics that are especially interesting to me, and this book is
the LATEX’ed result.
My goal here is to present a streamlined introduction to what I view as the essential
elements of algebraic topology, a toolkit as it were. This text is not intended to serve
as an encyclopedic reference work, but to be a solid foundation and an invitation to
further study. An ambitious instructor could, with judicious omissions, cover it in one
semester. I prefer a more leisurely two-semester amble, but that is partially a function
of my department’s course requirements.
Readers familiar with standard texts in algebraic topology will notice some influ-
ences. That first course for me was based on the book of Greenberg & Harper [1]. It has
much to recommend it, but it is quite terse (perhaps too much so). The current standard
text is Hatcher’s excellent book [2], and I have used it successfully in the past. Indeed,
his direct calculation of the cohomology ring of projective space (rather than appeal-
ing to Poincaré duality) appears here with little modification; the extra technical work
required to do it is actually quite illuminating. Hatcher’s book is rather encyclopedic,
however, and as mentioned above, I am aiming for a more pared-down version. Much
of the material on cellular homology and spectral sequences is modeled on the place
where I learned it, the book of Griffiths & Morgan [3]. Spectral sequences are an impor-
tant tool that every topologist needs in their kit, and that is why they are included here.
I will say that this presentation is analogous to what we show undergraduates in the
introductory calculus sequence: a focus on how to use these techniques over the details
of how they work. Other sources I consulted include Spanier’s classic text [4], Munkres’s
algebraic topology book [5], Steenrod’s treatise on fiber bundles [6], and Weibel’s book
on homological algebra [7].
The book contains four chapters. Chapter 1 focuses on preliminary material: contin-
uous maps and homotopy, lifting properties and fibrations, basic category theory, and
the fundamental group. Chapter 2 is about homology in all its forms–simplicial, singular,

https://doi.org/10.1515/9783111014852-201
VIII � Preface

cellular–and the properties thereof. Manifolds and their homology are covered in de-
tail. This chapter also includes a section about persistent homology and topological data
analysis. This area has gained prominence over the last couple of decades, bringing what
was once a very theoretical branch of mathematics into a world of applications. Chap-
ter 3 is about cohomology and duality, including the cup product and cohomology ring.
Poincaré duality appears here, of course. Finally, Chapter 4 introduces homotopy groups
and the basics of the homotopy theory of cell complexes. The Leray–Serre spectral se-
quence is introduced and used to compute many interesting examples of cohomology
rings and homotopy groups.
Here is one topic that I have omitted that may be controversial: a detailed study
of covering spaces. While this theory is beautiful and interesting, every time I teach
the course I find myself regretting spending so much time on it. It is true that covering
spaces are needed from time to time later in the text, but the limited discussion here is
sufficient. I have included an exploration of covering spaces as a project at the end of
the first chapter, however, if an instructor wishes to cover it.
Each chapter concludes with one or two projects for students to explore topics in
greater depth. Exercises are included throughout as well.
I hope the reader finds this book to be a clear and compelling introduction to a
subject that continues to find new applications. If mathematicians, both emerging and
established, find it useful, then I have been successful.

Gainesville, Florida, July 2024 K.P.K.


Acknowledgments
This is my third book, but it never seems to get easier. As anyone who has done this
knows, there are periods of intense enthusiasm interspersed with nagging feelings of
doubt. But here it is, so enthusiasm must win out in the end.
I am indebted to many excellent mathematicians from whom I have learned so
much over the years. These include Peter Fletcher, Richard Hain, John Harer, Herbert
Edelsbrunner, Henry King, Neža Mramor, Ulrich Bauer, Robert Ghrist, Peter Bubenik,
Chuck Weibel, Gunnar Carlsson, Andrei Suslin, Eric Friedlander, Mark Walker, Nick
Scoville, Vidit Nanda, and on and on. My apologies to anyone I may have omitted.
Many thanks are due to the fine folks at De Gruyter, especially to Steve Elliot for
recruiting me to do this, and to Ms. Ute Skambraks for her excellent editorial assistance.
I am also grateful to a pair of anonymous referees for their comments on a draft of this
text. Any errors or omissions are entirely my responsibility, not theirs.
Finally, I am always grateful for the love and support of my wife Ellen and our son
Gus, the best people I know.

https://doi.org/10.1515/9783111014852-202
Contents
Preface � VII

Acknowledgments � IX

1 Preliminaries � 1
1.1 Topological spaces � 1
1.1.1 Some notions from general topology � 1
1.1.2 Simplicial complexes � 5
1.1.3 Cell complexes � 9
1.1.4 Manifolds � 11
1.1.5 Dramatis personæ � 13
1.2 Continuous maps and homotopy � 19
1.2.1 Homotopy of maps � 19
1.2.2 Homotopy equivalence of spaces � 22
1.2.3 Homotopy extension � 25
1.3 Lifting properties and fibrations � 30
1.3.1 Liftings of continuous maps � 30
1.3.2 Locally trivial fiber bundles � 32
1.4 Basic category theory � 36
1.5 The fundamental group � 39
1.6 The Seifert–Van Kampen Theorem � 45
Project: covering spaces and the fundamental group � 51
Project: SL2 (ℤ) as an amalgamated free product � 55

2 Homology � 61
2.1 What is homology? � 61
2.1.1 Chain complexes � 61
2.1.2 Homology of chain complexes � 62
2.2 Simplicial homology � 65
2.2.1 Simplical chains � 66
2.2.2 Calculations � 68
2.3 Singular homology � 73
2.3.1 Singular chains � 74
2.3.2 Relative homology � 78
2.3.3 Excision � 82
2.3.4 Homology of spheres revisited � 88
2.3.5 The Mayer–Vietoris sequence � 91
2.3.6 Homology with coefficients � 95
2.4 Cellular homology � 96
2.4.1 Cellular chains � 98
XII � Contents

2.4.2 Calculations � 101


2.5 Comparison of homology theories � 104
2.5.1 Simplicial vs. singular � 104
2.5.2 Cellular vs. singular � 105
2.5.3 Eilenberg–Steenrod axioms � 108
2.6 Homology of manifolds � 109
2.6.1 Orientability � 109
2.6.2 The top homology group and the fundamental class � 112
2.6.3 Euler characteristic � 119
2.7 Persistent homology and topological data analysis � 122
2.7.1 Filtrations, barcodes, and persistence diagrams � 123
2.7.2 Stability � 132
2.7.3 Persistence modules � 137
2.7.4 Vietoris–Rips complexes � 139
Project: the Jordan–Brouwer Separation Theorem � 142
Project: the Nerve Theorem � 143
Project: example persistent homology calculations � 145

3 Cohomology and duality � 147


3.1 Cohomology of spaces � 147
3.1.1 Motivation from calculus � 147
3.1.2 Cochains � 152
3.1.3 Calculations � 159
3.2 Cup products and the cohomology ring � 165
3.2.1 Definition of the cup product � 166
3.2.2 Calculations of cohomology rings � 170
3.3 Cohomology of manifolds and Poincaré duality � 182
3.3.1 Cap products and the duality theorem � 182
3.3.2 Interaction with cup products � 192
3.3.3 Consequences of duality � 193
3.4 Universal coefficient and Künneth theorems � 196
3.4.1 Universal coefficient theorems � 196
3.4.2 The Künneth formula � 199
Project: the geometry of Poincaré Duality � 203

4 Homotopy and spectral sequences � 205


4.1 Homotopy groups � 205
4.1.1 Definitions and basic calculations � 205
4.1.2 Whitehead’s theorem � 210
4.1.3 Hurewicz theorem � 211
4.1.4 The long exact sequence of a fibration � 214
4.2 The Leray–Serre spectral sequence � 221
Contents � XIII

4.2.1 Cohomology of fibrations � 221


4.2.2 Calculations � 227
Project: obstruction theory and the Hopf index formula � 237

Bibliography � 243

Index � 245
1 Preliminaries
1.1 Topological spaces
We assume that the reader already has a familiarity with the definitions and basic prop-
erties of topological spaces, including the notions of continuity, connectivity, compact-
ness, separation axioms, etc. For completeness, we give a brief summary here.

1.1.1 Some notions from general topology

Definition 1.1.1. A topological space is a set X equipped with a collection 𝒯 of subsets


called open sets such that the collection 𝒯 is closed under arbitrary unions and finite
intersections. The sets X and 0 are required to belong to 𝒯 .

Definition 1.1.2. A function f : X → Y between topological spaces is continuous if for


each open set V ⊂ Y the inverse image f −1 (V ) is open in X.

Example 1.1.3. Consider the set ℝ of real numbers with its standard topology having
basis the usual open intervals. The definition of continuity for a function f : ℝ → ℝ that
one learns in a first analysis course reads as follows. A function f : ℝ → ℝ is continuous
at a if for every ε > 0 there exists a δ > 0 such that if |x − a| < δ then |f (x) − f (a)| < ε.
Note that this is equivalent to Definition 1.1.2 above.

Definition 1.1.4. A metric on a set X is a function d : X × X → ℝ satisfying the following.


– d(x, y) ≥ 0 for all x, y ∈ X, with equality if and only if x = y;
– d(x, y) = d(y, x) for all x, y ∈ X;
– (triangle inequality) for all x, y, z ∈ X, we have d(x, z) ≤ d(x, y) + d(y, z).

Given a metric d on X, x ∈ X, and ε > 0, define the ε-neighborhood of x to be the set


Bε (x) = {y ∈ X : d(x, y) < ε}. The sets Bε (x), x ∈ X, form the basis of a topology on X and
we call (X, d) a metric space.

Example 1.1.5. The euclidean space ℝn is a metric space with d(x, y) being the standard
euclidean distance between points.

Definition 1.1.6. A topological space X is Hausdorff if for every pair of points x ≠ y in


X there exist disjoint open sets U and V such that x ∈ U and y ∈ V .

Example 1.1.7. The euclidean space ℝn is Hausdorff: given points x ≠ y, let ε = ‖x − y‖/2
and take U = Bε (x), V = Bε (y).

Definition 1.1.8. A space X is connected if it is impossible to write X = A ∪ B, where A


and B are nonempty disjoint open sets.

https://doi.org/10.1515/9783111014852-001
2 � 1 Preliminaries

Example 1.1.9. The real line ℝ is connected, as is any interval in ℝ. In general, ℝn is


connected for all n.

By an open cover of a space X we mean a collection {Uλ }λ∈Λ of open subsets of X


whose union is all of X. A subcover is a subcollection of the Uλ that also covers X.

Definition 1.1.10. A topological space X is compact if every open cover of X has a finite
subcover.

Example 1.1.11. In ℝ a closed interval [c, d] is compact. In general, any closed and
bounded set in a euclidean space ℝn is compact (this is the Heine–Borel Theorem).

Quotient spaces
A common construction in topology is to begin with a space and make identifications
among some elements in the space to create a new object. This is often done via some
equivalence relation on the original space and since there is a canonical map to the set of
equivalence classes, we want to ensure that the topology on the new space is compatible
with this map.

Definition 1.1.12. A surjective function p : X → Y is a quotient map if a subset V ⊂ Y is


open if and only if p−1 (V ) is open in X.

Note that a quotient map is continuous by definition; it is the condition that V is


open in Y only if its inverse image is open in X that is new here. This condition imposes a
topology on Y . What one should really be thinking of here is that we have an equivalence
relation ∼ on X defined by x1 ∼ x2 ⇔ p(x1 ) = p(x2 ) and then the quotient map p defines
a topology on the set Y of equivalence classes.
The quotient topology is characterized by the following universal mapping property.

Proposition 1.1.13. Suppose p : X → Y is a quotient map and suppose f : X → Z is a


continuous map with f (x1 ) = f (x2 ) whenever p(x1 ) = p(x2 ). Then there exists a unique
continuous map f : Y → Z such that f = f ∘ p.

Proof. If y ∈ Y , write y = p(x) for some x ∈ X and then set f (y) = f (x). The map f is
well-defined, since if p(x ′ ) = y, then f (p(x ′ )) = f (p(x)) and hence f (y) is independent of
the choice of lift. The map f is unique: Suppose g : Y → Z satisfies f = g ∘ p. Then if
y ∈ Y , choose a lift x and then

g(y) = g(p(x)) = f (x) = f (p(x)) = f (y).

−1
Finally, to see that the map f is continuous, suppose U ⊆ Z is open. We must show f (U)
−1 −1
is open in Y . But f (U) is open in Y if and only if p−1 (f (U)) is open in X. The latter set
is simply f −1 (U), and this is open in X since f is continuous.
1.1 Topological spaces � 3

Examples
We first list some standard spaces. The n-dimensional euclidean space ℝn consists of
n-tuples of real numbers (x1 , . . . , xn ). The topology on ℝn is the usual one, in which an
ε-ball around a point x consist of all points y with ‖x − y‖ < ε, where ‖ ⋅ ‖ denotes the
standard euclidean norm. The n-disc, Dn , is the set {x = (x1 , . . . , xn ) ∈ ℝn |‖x‖ ≤ 1}, with
the subspace topology. The n-sphere, S n is the boundary of the disc Dn+1 ; it is the set
{x = (x1 , . . . , xn+1 )|‖x‖ = 1}. Note that S 0 consists of two points {±1}.
(1) Dn is homeomorphic to the quotient space S n−1 × [0, 1]/(S n−1 × {0}). Here, if A is
a subspace of X, the notation X/A denotes the space obtained by “collapsing A to a
point.” Formally, this is the set of equivalence classes for the relation ∼, where a ∼ a′
for all a, a′ ∈ A, x ∼ x for x ∈ X \ A. The space S n−1 × [0, 1] is a “cylinder” and the
subspace S n−1 ×{0} is the base of that cylinder. Collapsing that base to a point creates
a cone over S n−1 × {1}; viewing this from above one sees a disc. See Figure 1.1.


S n−1 × I

Dn
n−1
S
S n−1 × I/S n−1 × {0}
× {0}

Figure 1.1: The disc as a quotient of S n−1 × [0, 1].

That is the intuition anyway. To prove this formally, we proceed as follows. First put
the quotient topology on S n−1 × [0, 1]/S n−1 × {0}; call the quotient map p. Now define
φ : S n−1 ×[0, 1] → Dn by φ(x, t) = tx. Observe that this is well-defined since S n−1 ⊂ ℝn
and ‖tx‖ = t ≤ 1 for all x ∈ S n−1 . This function is clearly continuous and surjective.
Moreover, φ(x, t) = 0 if and only if (x, t) ∈ S n−1 × {0}. By the universal mapping
property, we get a unique map φ : S n−1 × [0, 1]/S n−1 × {0} → Dn with φ = φ ∘ p. One
checks easily that φ is a homeomorphism (see the exercises for this section).
(2) Dn /S n−1 is homeomorphic to S n . This is easy to visualize when n = 2. In this case, we
have a disc and we are collapsing the boundary circle to a single point. The result is
a sphere: imagine a plastic trash bag with a drawstring; pulling the drawstring tight
seals the bag. See Figure 1.2.
To prove this, we will use the previous example. Define ψ : S n−1 × [0, 1] → S n by
ψ(x, t) = (cos(πt), sin(πt)x) ∈ ℝ × ℝn = ℝn+1 . Note that since ‖x‖ = 1, ‖ψ(x, t)‖ = 1
and so the image of ψ lies in S n . Note that ψ(S n−1 × {0}) = (1, 0, 0, . . . , 0). So by the
universal mapping property, we get a continuous map ψ : S n−1 × [0, 1]/S n−1 × {0} =
Dn → S n . Moreover, S n−1 ⊂ Dn is the image of S n−1 × {1} under ψ. Also, ψ(S n−1 ×
4 � 1 Preliminaries

S n−1 Sn

Dn ψ(S n−1 )

Figure 1.2: The sphere as a quotient of the disc.

{1}) = (−1, 0, 0, . . . , 0). Using the universal property again, ψ induces a continuous
ψ : Dn /S n−1 → S n . The map ψ is a homeomorphism.
(3) The torus T. This space is the surface of a donut and is defined as the product space
S 1 ×S 1 . However, it is more often thought of as a quotient of the unit square as follows.
Given the square [0, 1] × [0, 1], identify points on the boundary in the following way:
(a, 0) ∼ (a, 1) and (0, b) ∼ (1, b) (see Figure 1.3). There are two ways to see that the
resulting space is S 1 × S 1 . First, products and quotients commute (sometimes): if
p1 : X1 → Y1 and p2 : X2 → Y2 are quotient maps and Y1 and X2 are locally compact,
then p1 × p2 : X1 × X2 → Y1 × Y2 is a quotient map. In this case, we have pj : [0, 1] → S 1
as above and then p1 × p2 : [0, 1] × [0, 1] → S 1 × S 1 is a quotient. Alternatively, we
could note that the composition of quotient maps is a quotient; again see Figure 1.3.

I ×I T

Figure 1.3: The torus as a quotient of the unit square.

(4) Adjunction spaces. Suppose X and Y are spaces and that f : A → Y is a continuous
map, where A is a subspace of X. The adjunction space is the quotient

X ∪f Y = X ⨿ Y / ∼,

where ∼ is the relation a ∼ f (a) for a ∈ A. Intuitively, this is the space obtained
by gluing X to Y along the image of the subspace A. In particular, if A consists of
a single point, then this space is often denoted by X ∨ Y and called the wedge of X
and Y . For example, S 1 ∨ S 1 is two circles joined at a single point to form a figure
1.1 Topological spaces � 5

eight. A common construction used in the sequel is to take X = Dn with A = S n−1 ,


the boundary sphere. Then the space Dn ∪f Y is the space Y with an n-cell attached
along the attaching map f : S n−1 → Y .

Exercises

Exercise 1.1.1. Prove that a continuous bijection from a compact space to a Hausdorff
space is a homeomorphism.

Exercise 1.1.2. Prove that the map φ : S n−1 × [0, 1]/S n−1 × {0} → Dn is a homeomorphism.

Exercise 1.1.3. Prove that the map ψ : Dn /S n−1 → S n is a homeomorphism.

Exercise 1.1.4. Prove that if f : X → Y is a quotient map and Z is locally compact, then
f × idZ : X × Z → Y × Z is a quotient map. Deduce that if p1 : X1 → Y1 and p2 : X2 → Y2
are quotient maps with Y1 and X2 locally compact, then p1 × p2 : X1 × X2 → Y1 × Y2 is a
quotient map.

Exercise 1.1.5. Prove that the composition of quotient maps is a quotient map.

Exercise 1.1.6. Prove that the sphere S n is the quotient space obtained from two disjoint
copies of Dn by gluing them along their boundary spheres S n−1 .

1.1.2 Simplicial complexes

Simplicial complexes are fundamental objects in algebraic topology. Most of the spaces
we are interested in may be realized in this way, and simplicial complexes have nice
combinatorial descriptions that make them easy to implement on a computer. We begin
with a very concrete description of these objects.
Recall that a collection of points v0 , . . . , vn in a euclidean space ℝk is in general po-
sition if for a set of real scalars t0 , . . . , tn the equations ∑ ti = 0 and ∑ ti vi = 0 imply that
ti = 0 for all i = 0, . . . , n. Geometrically this means that no three of the vi lie on the same
line, no four lie in the same plane, etc.

Definition 1.1.14. Let n be a nonnegative integer and let v0 , . . . , vn be points in general


position in some euclidean space ℝk . The n-simplex σ spanned by v0 , . . . , vn is the convex
hull of these points. This consists of the following set:

n n
σ = {x 󵄨󵄨󵄨 x = ∑ ti vi , ∑ ti = 1, ti ≥ 0}.
󵄨
i=0 i=0

The real numbers t0 , . . . , tn are called the barycentric coordinates of the point x. The
points v0 , . . . , vn are called the vertices of the simplex. We write σ = ⟨v0 , . . . , vn ⟩.

A 1-simplex is a line segment; a 2-simplex is a triangle; a 3-simplex is a tetrahedron;


etc. In general, an n-simplex has (n + 1) vertices and is an n-dimensional object.
6 � 1 Preliminaries

Example 1.1.15. The standard n-simplex, Δn is the n-simplex in ℝn+1 with vertices the
standard basis vectors e1 , . . . , en+1 . See Figure 1.4.

(0, 0, 1)
Δ2
(0, 1)
Δ1
(0, 1, 0)

(1, 0, 0)
(1, 0)

Figure 1.4: The standard n-simplex Δn for n = 1, 2.

Remark 1.1.16. Note that we can realize Δn in ℝn as the convex hull of {0, e1 , . . . , en }. This
is sometimes more convenient. However, the resulting n-simplex does not have sides of
equal length, whereas the standard Δn ⊂ ℝn+1 does.

Definition 1.1.17. Let σ be an n-simplex with vertices v0 , . . . , vn . A face τ of σ is a


k-simplex spanned by a subset of the vertices of σ; that is, τ = ⟨vi0 , vi1 , . . . , vik ⟩. We
denote this relationship by τ < σ.

Loosely speaking, a simplicial complex is obtained by gluing together simplices


along common faces. Formally, we have the following.

Definition 1.1.18. A simplicial complex K in ℝn is a collection of simplices such that


(1) Every face of a simplex of K is in K; and
(2) The intersection of any two simplices of K is in K.

A subset L ⊆ K is a subcomplex if L is itself a simplicial complex.

Note that this definition prohibits oddities such as gluing a vertex of a triangle to a
point on a side of another. See Figure 1.5 for some examples of simplicial complexes.

Figure 1.5: Some simplicial complexes.

Definition 1.1.19. Let K be a simplicial complex. The p-skeleton of K, denoted K (p) , con-
sists of all k-simplices of K with k ≤ p. It is a subcomplex of K.
1.1 Topological spaces � 7

For example, K (0) consists of all vertices of K, K (1) is a graph whose vertex set is K (0) ,
and so on.

Definition 1.1.20. Suppose σ = ⟨v0 , . . . , vn ⟩ is a simplex and let v be a vertex disjoint


from σ. The join, v ∗ σ, is the (n + 1)-simplex v ∗ σ = ⟨v, v0 , . . . , vn ⟩.

Example 1.1.21. Consider the standard n-simplex Δn . The (n − 1)-skeleton consists of all
faces of dimension ≤ n−1. This is the set of n+1 codimension-one faces, each identifiable
with Δn−1 glued together along their common faces. We often denote this by 𝜕Δn and call
it the boundary of Δn . See Figure 1.6.

𝜕Δ1 𝜕Δ2 𝜕Δ3


Figure 1.6: The boundary of Δn , n = 1, 2, 3, as realized in ℝn .

Definition 1.1.22. Let K be a simplicial complex in ℝn . Denote by |K| the union of the
simplices in K. The space |K|, equipped with the subspace topology inherited from ℝn ,
is called the polytope or underlying space of K.

Note that if K consists of finitely many simplices, |K| is compact.

Simplicial maps
Because we want our constructions to respect whatever structures we have in place, we
need to specify the types of maps we allow between objects.

Definition 1.1.23. Suppose K and L are simplicial complexes. A map f : K → L is sim-


plicial if for every simplex σ in K, f (σ) is a simplex in L.

Note that if f is a simplicial map, then f (K (0) ) ⊆ L(0) ; that is, f takes vertices to
vertices. It then takes edges to edges, and so on. In particular, f is determined by what it
does to vertices. More is true, however.

Theorem 1.1.24. Let K and L be simplicial complexes and suppose f : K (0) → L(0) is a
map such that if v0 , . . . , vk span a simplex σ in K, then f (v0 ), . . . , f (vk ) span a simplex in
L. Then there is a continuous map g : |K| → |L| extending f such that if x = ∑ ti vi ∈ σ, we
have g(x) = ∑ ti f (vi ).

Proof. This is more or less clear. We extend the barycentric coordinates on each simplex
of K to all of K by noting that if x ∈ |K|, then x is either in K (0) or it lies in the interior of
exactly one simplex σ = ⟨w0 , . . . , wr ⟩ of K. In that case, x = ∑ ti wi with ti > 0 for each i
and ∑ ti = 1. This allows us to give coordinates for x relative to all the vertices by setting
8 � 1 Preliminaries

tv (x) = 0 if v ≠ wj and tv (x) = tj if v = wj . We then define g as in the statement of the


theorem. The continuity of g is left as an exercise.

The map g is called the simplicial map induced by f .

Barycentric subdivision
A useful construction when working with simplicial complexes is the barycentric subdi-
vision.

Definition 1.1.25. Suppose σ = ⟨v0 , . . . , vk ⟩ is a k-simplex. The barycenter σ̂ of σ is the


point in the interior of σ with barycentric coordinates

1 k
σ̂ = ∑v .
k + 1 i=0 i

Geometrically, the barycenter is the centroid of the simplex. So the barycenter of an


edge is its midpoint, of a 2-simplex is the intersection of the three edge bisectors, etc. The
barycenter of a vertex is the vertex itself. We now define the barycentric subdivision of
a simplex σ = ⟨v0 , . . . , vk ⟩ inductively by assuming the (k − 1)-skeleton has been subdi-
vided. Let σ̂ be the barycenter of σ and then for each (k − 1)-simplex τ in the subdivided
(k − 1)-skeleton build the k-simplex σ̂ ∗ τ. See Figure 1.7.

Figure 1.7: The barycentric subdivision of a 2-simplex.

Now, given a simplicial complex K, the barycentric subdivision of K, denoted sd(K),


is obtained by taking the barycentric subdivision of each simplex in K. See Figure 1.8.

Figure 1.8: The barycentric subdivision of a simplicial complex.


Another Random Scribd Document
with Unrelated Content
rate, approximately equivalent to the working capacity of the plant.
Sales of accessible timber usually do not exceed 5 years in length.
However, in the case of inaccessible tracts requiring a large
investment for transportation facilities an exception is made and
periods of from 15 to 20 years may be granted.
Readjustment of Stumpage Rates. In all sales exceeding 5 years
in length provision is made to have the stumpage rates readjusted
by the Forester at the end of three or five year intervals to meet
changing market and manufacturing conditions.

Figure 68. An excellent illustration showing the difference


between unrestricted logging as practised by lumbermen, and
conservative logging as practised by the Forest Service. In the
foreground is the unrestricted logging which strips the soil of
every stick of timber both large and small; in the background is
the Forest Service logging area which preserves the young
growth to insure a future supply of timber for the West.
Bitterroot National Forest, Montana.
Refunds. Deposits to cover or secure advance cutting or to
accompany bids apply on the first payment if a sale is awarded to
the depositor; otherwise they will be refunded. Refunds are also
made to the purchaser if the last payment is in excess of the value
of the timber that is cut.

THE DISPOSAL OF TIMBER TO HOMESTEAD


SETTLERS AND UNDER FREE USE
Besides selling the timber and other forest products outright, as
has just been described, some timber is sold to settlers at cost and
much timber is given away to the local people under the free use
policy.
Sales to Homestead Settlers and Farmers. Sales to homestead
settlers and farmers are made without advertisement in any amount
desired, at the price fixed annually for each National Forest region of
similar conditions by the Secretary, as equivalent to the actual cost
of making and administering such sales. Only material to be used by
the purchaser for domestic purposes exclusively on homesteads or
farms is sold in this way. Such uses include the construction or repair
of farm buildings, fences, and other improvements and fuel. Such
sales are restricted to mature dead and down timber which may be
cut without injury to the forest.
Free Use. Free use of timber is granted primarily to aid in the
protection and silvicultural improvement of the Forests. Hence the
material taken is, except in unusual cases, restricted to dead, insect
infested and diseased timber, and thinnings. Green material may be
taken in exceptional cases where its refusal would clearly cause
unwarranted hardship. The use of such material is granted freely:
(1) To bona fide settlers, miners, residents, prospectors, for fire
wood, fencing, building, mining, prospecting, and other domestic
purposes; and to any one in case its removal is necessary for the
welfare of the Forest; (2) for the construction of telephone lines
when necessary for the protection of forests from fire; (3) to certain
branches of the Federal Government. Free use is not granted for
commercial purposes or of use in any business, including sawmills,
hotels, stores, companies or corporations. Such persons are required
to purchase their timber.

Figure 69. View showing the Forest


Service method of piling the brush and
débris after logging, and also how stump
heights are kept down to prevent waste.
New Mexico.
Figure 70. A tie-cutting operation on a
National Forest. These piles of railroad
ties are being inspected, stamped, and
counted by Forest rangers. From this
point the ties are "skidded" to the banks
of a stream to be floated to the shipping
point. Near Evanston, Wyoming.

The aggregate amount of free use material granted annually to


any user must not exceed $20 in value, except in cases of unusual
need or of dead or insect infested timber, the removal of which
would be a benefit to the forest, or in the case of any timber which
should be removed and whose sale under contract cannot be
effected. In these cases the amount may be extended to $100.
Supervisors have authority to grant free use permits up to $100,
District Foresters up to $500, and larger amounts must have the
approval of the Forester.
Free use material is appraised in the same manner and in
accordance with the same principles as timber purchased under sale
agreements. The valuation of such material is at the same rate as
that prevailing for similar grades of stumpage in current sales in the
same locality.
The magnitude of the free use business may be appreciated
from the fact that during the fiscal year 1917 there were 41,427
individuals or companies who received timber under this policy. The
total amount thus given away was 113,073,000 board feet valued at
over $150,000.
Permits for this use are required for green material, but dead
timber may be taken without a permit. Supervisors designate as
free-use areas certain portions or all of any National Forest and
settlers, miners, residents, and prospectors may cut and remove
from such areas free of charge under Forest Service regulations any
timber needed for their own use for firewood, fencing, buildings,
mining, prospecting, or other domestic purposes.
Material cut under free-use regulations must not be removed
from the cutting area until scaled or measured by a Forest officer. In
some cases this requirement is waived when by it the needs of the
users are met with greater dispatch and the cost of administration is
thereby reduced. The free-use applicant is required to utilize the
trees cut in accordance with local Forest Service practice and he is
required to avoid unnecessary damage to young growth and
standing timber.

TIMBER SETTLEMENT AND ADMINISTRATIVE


USE
When timber on National Forest land is cut, damaged, killed, or
destroyed in connection with the enjoyment of a right-of-way or
other special use, it is not necessary to advertise it for sale, but
payment therefor is required at not less than the minimum rate
established by the Secretary of Agriculture. Timber removed in this
way is usually scaled, measured, or counted and the procedure is
identical with that of a timber sale. But where timber is destroyed or
where it is not worked up in measurable form or where the cutting is
done in such a way that scaling is impracticable, settlement is
required on the basis of an estimate.

Figure 71. Brush piles on a cut-over


area before burning. Forest Service
methods aim to clean up the forest after
logging so that forest fires have less
inflammable material to feed on.
Bitterroot National Forest, Montana.
Figure 72. At a time of the year when
there is least danger from fire the brush
piles are burned. Missoula National Forest,
Montana.

In 1912 a new branch of the Southern Pacific Railroad was built


across a portion of the Lassen National Forest in California. The
company was going to use some of the timber, but most of it was to
be destroyed or disposed of in the easiest manner. Scaling was
impossible, so the company paid for the timber—about $10,000—on
the basis of a careful estimate made by the writer, then Forest
Examiner.
The charge for all such timber is made on the basis of the
current stumpage rates for timber of like quality and accessibility
included in sales for all classes of material which have to be cut or
destroyed and which are commonly salable on the Forest.
Timber is often used by the Forest Service itself in the
administration of the National Forests. The Forester, District
Foresters, and the Supervisors are authorized to sell or dispose of
under free use or otherwise, within the amount each one is
authorized to sell, any timber upon the National Forests when such
removal is actually necessary to protect the Forest from ravages or
destruction, or when the use or removal of the timber is necessary in
the construction of roads, trails, cabins, and other improvements on
the National Forests or in experiments conducted by the Forest
Service.

THE RENTAL OF NATIONAL FOREST RANGE


LANDS
The forage crop on the National Forests is for the use of the
sheep and cattle of the western stockmen and it is procured by
means of grazing permits which are issued and charged for upon a
per capita basis. The primary objects of the administration of
government grazing lands are: the protection and conservative use
of all National Forest land adapted to grazing; the permanent good
of the live stock industry through the proper care and use of grazing
lands; and the protection of the settler and home builder against
unfair competition in the use of the range.
Importance of the Live Stock Industry. The grazing business,
more than any other feature of National Forest management, is
immensely practical, because it is immediately concerned with
human interests. This industry furnishes not only meat, but leather,
wool, and many by-products.
That the National Forests play a big part in the maintenance of
this industry there can be little doubt, for it has been estimated
recently that 30 per cent. of the sheep and 20 per cent. of the cattle
of the far Western States are grazed in the National Forests. The
Forests contain by far the largest part of the summer range lands in
the far Western States and hence are of paramount importance. The
winter grazing lands in the West are so much greater in area than
the summer lands, that for this reason also National Forest range
lands are in great demand.
Permits Issued in 1917. During the fiscal year 1917 more than
31,000 permits to graze cattle, hogs, or horses, and over 5,500
permits to graze sheep or goats were issued. These permits
provided for 2,054,384 cattle, 7,586,034 sheep, about 100,000
horses, about 50,000 goats, and about 3,000 hogs. The total
receipts for 1917 were over $1,500,000. The gross receipts to the
owners of the stock probably exceeded $50,000,000 and the capital
invested in the stock no doubt amounted to over $200,000,000.
An idea of the growth of the grazing business may be gotten
from the Forest Service statistics for the fiscal years 1908 and 1917.
The increase in the number of permits and the volume of the
business is due primarily to a better administration and better
regulation of grazing interests and more specifically to the increase
in the carrying capacity of government lands by wise and restricted
use. Between these two fiscal years there was no appreciable
increase in the total area of the Forests which would account for the
increased business. In 1908 there were issued 19,845 permits for
1,382,221 cattle, horses and hogs; in 1917 there were issued 31,136
permits for 2,054,384 animals. In 1908 there were issued 4,282
permits for 7,087,111 sheep and goats; in 1917 5,502 permits were
issued for 7,586,034 sheep and goats. The number of cattle and
horses grazed has increased therefore by 50 per cent. and the
number of sheep and goats by 7 per cent. The total receipts have
increased from $962,829.40 in 1908 to $1,549,794.76 in 1917.
Kinds of Range, Grazing Seasons, and Methods of Handling
Stock. For the proper understanding of the grazing business on the
National Forests it is necessary to know something about the
different kinds of range, the length of grazing seasons, and the
methods of handling different classes of stock. Sheep and goat
range differs materially from cattle and horse range and the proper
distribution of stock over a National Forest cannot be effected unless
this difference is recognized. Sheep and goat range usually consists
of low shrubs or brush and is known collectively as "browse"; cattle
and horses subsist mainly upon grass, flowering plants and herbs.
Sheep feel more at home on high mountain slopes, while cattle and
horses range usually on the lower slopes and in the valleys, and
especially in the broad meadows, around lakes and along streams.
Sheep are more apt to find feed in the forests, that is under the
trees; cattle prefer the open; they usually avoid the forest, preferring
to keep out on the open meadows and grassy slopes.
Naturally some ranges have feed at some seasons of the year
and other ranges at other seasons. Some of the National Forests in
California extend from an elevation of a few hundred feet in the
foothills of the great valleys to an elevation of more than 10,000 feet
at the crest of the Sierra Nevada Mountains. The lower foothills
afford excellent feed soon after the beginning of the fall rains in
November and, due to the very mild winter which this region enjoys,
there is excellent feed in February and March. This is known as
winter range. The medium high slopes of the mountains have a later
growing season and the sheep and cattle reach there about June
and stay until August or September. Still higher up the forage
matures later and the grazing season extends from August until
November. At these elevations the snowbanks usually lie until July
and the growing season is very short, for the new snow usually
buries the vegetation about the first of November. Thus stockmen
have what they call "winter range," "summer range," and "fall
range," depending upon what seasons of the year the forage crop
can be utilized. The National Forests on the whole contain very little
winter range, hence stockmen must move their stock in the fall to
private lands at lower elevations either where the climate is
considerably warmer or where there is very little snowfall. A large
part of the western winter grazing lands are in regions of light
snowfall, such as at the lower elevations in Utah, Nevada, Wyoming,
and Colorado. Here the stock feeds on dry grass. Stockmen who
cannot get winter range lands must feed their stock at ranches.
The characteristic habits of sheep and cattle require that they be
handled differently on the range. Sheep are herded in bands while
cattle are handled in scattered groups. The new and approved
method of handling sheep called the "burro system" calls for a burro
with the sheep to pack the herder's blankets and provisions. The
herder camps where night overtakes him. The herder and his band
keep moving over the allotted range from one camp to another until
he has covered the whole range. After leaving his last camp he is
ready to begin all over again, since the feed near the camp where he
began has had two to three weeks' time to grow a new crop. Cattle
usually run loose singly or in groups on their allotted range. Usually
a range rider is camped on the range to keep the cattle from
straying to other ranges. He salts the cattle to keep them on their
own range, takes care of cattle that have gotten sick, and takes care
of the stock in other ways.
Grazing Districts and Grazing Units. The Secretary of Agriculture
not only has the authority to regulate grazing and prescribe the
schedule of grazing fees to be charged but he also regulates the
number and class of stock which are allowed to graze on each
National Forest annually.
The ranges within the National Forests are used by the kind of
stock for which they are best adapted except when this would not be
consistent with the welfare of local residents or the proper protection
of the Forests. For convenience in administration Forests are divided
into grazing districts. A typical Forest is divided into from 4 to 6
districts which may be natural grazing units, natural administrative
units (coinciding with the Ranger districts), or parts of the Forest
used by different classes of stock or parts of the Forest having
different lengths of grazing seasons. Each grazing district is also
subdivided into smaller divisions, units, or allotments. These are
usually natural divisions defined by topographic boundaries, such as
ridges, mountains, streams, etc., or more or less artificial divisions
determined by the class of stock which uses them. For example,
cattle and horses ordinarily graze in the valleys along the streams,
while sheep and goats graze the crests of ridges and the slopes of
mountains and will cross none but shallow streams. Each range
division or unit is usually given a well-known local name, such as
"Duck Lake Unit" or "Clover Valley Unit." One or more stockmen may
be allotted to such a unit, depending upon the size of the unit and
the number of animals it can feed. If only one stockman uses it, it
becomes an individual allotment. Usually a sheep owner with several
large bands of sheep is allotted one large unit adapted to sheep
grazing, while a large unit adapted to cattle and horses may be
allotted to one large cattle owner or to two or more smaller owners.
The manner in which sheep and goats are handled makes individual
allotments both practicable and desirable.
The boundaries of range allotments are usually well defined. In
the case of sheep they are marked with cloth posters. In most
Forests range allotments are fairly well settled. Each stockman gets
with his permit each spring a small map showing his own range and
the surrounding ranges.
Who Are Entitled to Grazing Privileges. The Secretary of
Agriculture has the authority to permit, regulate, or prohibit grazing
on the National Forests. Under his direction the Forest Service allows
the use of the forage crop as fully as the proper care and protection
of the National Forests and the water supply permit. The grazing use
of the National Forest lands is therefore only a personal and non-
transferable privilege. This privilege is a temporary one, allowable
under the law only when it does not interfere with the purposes for
which the National Forests were created. It is non-transferable
because it is based upon the possession of certain qualifications
peculiar to the permittee. To understand these qualifications it is
necessary to briefly look into the history of the grazing of live stock
on the western grazing lands.

Figure 73. Counting sheep as they


leave the corral. Sheep and cattle are
pastured on National Forests at so many
cents per head, hence they must be
counted before they enter in the spring.
Wasatch National Forest, Utah.
Figure 74. Logging National Forest
timber. Santa Fe National Forest, New
Mexico.

By long use of the public lands of the United States for grazing
purposes, long before the National Forests were created, stock
owners have been allowed to graze their stock upon such lands
under certain conditions of occupancy, residence, and ownership of
improved lands and water rights. This use, continuing through a long
period of years, has, in the absence of congressional legislation,
been commonly accepted in many communities, even receiving the
recognition of certain of the courts. It was allowed under "unwritten
law," as it were, only by the passive consent of the United States,
but by force of the presidential proclamation creating National
Forests, such passive consent ceased, being superseded by definite
regulations by the Secretary of Agriculture prescribed under the
authority of Congress. Therefore grazing stock on the Forests, as it
was done before the Forests were created, is trespass against the
United States. Due to the fact that local stockmen have used certain
public ranges year after year by the passive consent of the United
States, these stockmen are recognized in these localities as having
preference rights or equities in the use of range lands. These
equities form the basis upon which grazing privileges are allowed.
Grazing permits are issued only to persons entitled to share in
the use of the range within the National Forests by reason of their
fulfilling certain conditions or requirements. Prior use and occupancy
of National Forest lands for grazing purposes is the first and
foremost requirement. Local residence and ownership of improved
ranch property within or near the Forest and dependence upon
government range are also conditions that may entitle a stockman to
grazing privileges. The Forest Service also recognizes those
stockmen who have acquired by purchase or inheritance stock
grazed upon National Forest lands under permit and improved ranch
property used in connection with the stock, provided circumstances
warrant the renewal of the permit issued to the former owner. The
regular use of a range during its open season for several successive
years before the creation of the National Forest and under grazing
permit thereafter is what is meant by "prior use" or "regular
occupancy." The longer the period or use the greater the preference
right. No one can acquire this right to the use of National Forest
range, nor can it be bought or sold, but stockmen may acquire a
preference in the allotment of grazing privileges. This preference
right does not entitle him to continued use of a certain part of a
Forest, but only to preference over other applicants less entitled to
consideration in the use of the ranges open to the class of stock
which he wishes to graze. Certain stockmen may be given
preference in ranges secured by prior use and occupancy
supplemented by heavy investments in improved property and water
rights.
Citizens of the United States are given preference in the use of
the National Forests, but persons who are not citizens may be
allowed grazing permits provided they are bona fide residents and
owners of improved ranch property either within or adjacent to a
National Forest. Regular occupants of the range who own and reside
upon improved ranch property in or near National Forests are given
first consideration, but will be limited to a number which will not
exclude regular occupants who reside or whose stock are wintered
at a greater distance from the National Forests. With this provision
applicants for grazing permits are given preference in the following
order:

Class A. Persons owning and residing upon improved ranch


property within or near a National Forest who are
dependent upon National Forests for range and who do
not own more than a limited number of stock (known as
the protective limit).
Class B. Regular users of National Forests range who do not
own improved ranch property within or near a National
Forest, and persons owning such ranch property but who
own numbers of stock in excess of the established limit.
Class C. Persons who are not regular users of the National
Forest range and who do not own improved ranch
property within or near a National Forest. Such persons
are not granted permits upon Forests which are fully
occupied by classes A and B. Classes B and C are not
allowed to increase the number of stock grazed under
permit except by the purchase of other permitted stock.

From this classification it is very evident that the small local


stockmen who own approximately from 30 to 300 head of cattle and
from 500 to 2,000 head of sheep and who own and reside upon the
ranches near the Forests are given the preference in the allotment of
grazing privileges.
Grazing Permits. Various kinds of grazing permits are required
each year on the National Forests. These are known as ordinary
grazing permits, on-and-off permits, private land permits, and
crossing permits.
All persons must secure permits before grazing any stock on a
National Forest except for the few head in actual use by prospectors,
campers, ranchers, stockmen, and travelers who use saddle, pack
and work animals, and milch cows in connection with permitted
operations on the National Forests. Under these conditions 10 head
are allowed to graze without permit.
Persons owning stock which regularly graze on ranges partially
included within a National Forest, or upon range which includes
private land may be granted permits for such portions of their stock
as the circumstances appear to justify. This regulation provides for
cases where only a part of a natural range unit is National Forest
land, and where the economical use of the entire unit can be
secured only by the utilization of the Forest land in connection with
the other land. The regulation contemplates a movement of the
stock governed by natural conditions, between the Forest range and
the adjoining outside range, or between Forest land and
intermingled private land. This is called an on-and-off permit.
Permits on account of private lands are issued to persons who
own, or who have leased from the owners, unfenced lands within
any National Forest which are so situated and of such a character
that they may be used by other permitted stock to an extent
rendering the exchange advantageous to the Government. The
permits allow the permittees to graze upon National Forest land, free
of charge, the number of stock which the private lands will support,
by waiving the right to the exclusive use of the private land and
allowing it to remain open to other stock grazed on National Forest
land under permit.
The regular grazing permit carries with it the privilege of driving
the permitted stock over National Forest lands to and from the
allotted ranges at the beginning and end of the grazing season and
from the range to the most accessible shearing, dipping, and
shipping points during the term of the permit. But crossing permits
are necessary for crossing stock over National Forest lands to points
beyond the National Forest, for crossing stock to private lands within
a National Forest, or for crossing stock to reach dipping vats or
railroad shipping points. Rangers sometimes are detailed to
accompany the stock and see that there is no delay or trespassing.
No charge is made for crossing permits, but it is absolutely
necessary that persons crossing stock comply with the regulations
governing the National Forests and with the quarantine regulations
prescribed by the Secretary of Agriculture and the state authorities.
Grazing Fees. The full grazing fee is charged on all animals
under 6 months of age which are not the natural increase of stock
upon which the fees are paid. Animals under 6 months which are the
natural increase of permitted stock are not charged for. A reasonable
fee is charged for grazing all kinds of live stock on National Forests.
The rates are based upon the yearlong rate for cattle, which is from
60 cents to $1.50 per head, depending upon conditions on the
Forest. The yearlong rates for horses are 25 per cent. more and the
yearlong rate for swine 25 per cent. less than the rate for cattle. The
rate for sheep is 25 per cent. of the yearlong rate for cattle. The
rates for all kinds of stock for periods shorter than yearlong are
computed in proportion to the length of the season during which the
stock use National Forest lands. All grazing fees are payable in
advance.
When notice of the grazing allowance, periods, and rates for the
year has been received by the Supervisor he gives public notice of a
date on or before which all applications for grazing must be
presented to him. These public notices are posted in conspicuous
places, usually in the post offices. Applications for grazing permits
are submitted on blank forms furnished by the Supervisor. As soon
as an applicant for a grazing permit is notified by the Supervisor that
his application has been approved, he must remit the amount due
for grazing fees to the District Fiscal Agent and upon receipt of
notice by the Supervisor that payment has been made a permit is
issued allowing the stock to enter the Forest and remain during the
period specified. All grazing fees are payable in advance and the
stock is not allowed to enter the National Forest unless payment has
been made.
Stock Associations. The thirty or more grazing regulations
effective on the National Forests are for the primary purpose of
making the National Forest range lands as useful as possible to the
people consistent with their protection and perpetuation. It is clearly
impossible to meet the wishes and needs of each individual user, but
it is often entirely possible to meet the wishes of the majority of
users if made known through an organization. The organization of
stock associations is encouraged by the Forest Service and the
opinions and wishes of their advisory boards are recognized when
they represent general rather than individual or personal interests. It
is often possible through these organizations to construct range
improvements such as corrals, drift fences, roads, trails, and sources
of water supply for the common good of the members of the
organization and paid for by them.
Protective and Maximum Limits. In order to secure an equitable
distribution of grazing privileges, the District Forester establishes
protective limits covering the number of stock for which the permits
of Class A owners will be exempt from reduction in the renewal of
their permits. Permits for numbers in excess of the protective limits
will be subject to necessary reductions and will not be subject to
increase in number except through purchase of stock or ranches of
other permittees.

Figure 75. Sheep grazing on the


Montezuma National Forest at the foot of
Mt. Wilson, Colorado. Over 7,500,000
sheep and goats grazed on the National
Forests during the fiscal year 1917.

Figure 76. Grazing cattle on a National


Forest in Colorado. Permits were issued
during 1917 to graze over 2,000,000
cattle, horses, and swine on the National
Forests.

Protective limits are established to protect permittees from


reduction in the number of stock which they are allowed to graze
under permit below a point where the business becomes too small to
be handled at a profit or to contribute its proper share toward the
maintenance of a home. The average number of stock which a
settler must graze in order to utilize the products of his farm and
derive a reasonable profit is determined upon each Forest or, if
necessary, upon each grazing district thereof, and serves as the basis
for the protective limit. Protective limits have been established for
various Forests running from 25 to 300 head of cattle and from 500
to 2,000 head of sheep and goats.
Increases above the protective limit are allowed only to
purchasers of stock and ranches of permit holders and any such
increase must not exceed the maximum limit. Class A permittees
owning a less number of stock than the protective limit are allowed
to increase their number gradually. Whenever it is found necessary
to reduce the number of stock allowed in any National Forest, Class
C stock is excluded before the other classes are reduced. The
reduction on a sliding scale is then applied to Class B owners. Class
A owners are exempt from reduction. When new stock owners are
allowed the use of National Forest range upon a Forest already fully
stocked, reductions in the number of permitted stock of Class B and
C owners is made in order to make room for the new man. Thus it is
seen that the matter of protective limits is actually a protection to
the small stock owner; he is protected from the monopoly of the
range by big corporations.
When necessary to prevent monopoly of the range by large
stock owners, the District Forester establishes maximum limits in the
number of stock for which a permit may be issued to any one
person, firm or corporation.
Prohibition of Grazing. It often becomes necessary to prohibit all
grazing on an area within a National Forest or at least to materially
reduce the amount of stock which is allowed to graze on a given
area. Sheep may be excluded from a timber-sale area for a certain
number of years after cutting or until the reproduction has become
well established. Where planting operations are being carried on it is
usually necessary to exclude all classes of stock. If investigations
show that grazing is responsible for the lack of reproduction over a
considerable area, the area or a portion of it may be withdrawn from
range use until young growth has become established again. The
watersheds of streams supplying water for irrigation, municipal or
domestic purposes may be closed to grazing of any or all kinds of
domestic stock when necessary to prevent erosion and floods or
diminution in water supply. Camping grounds required for the
accommodation of the public may be closed to the grazing of
permitted stock. Limited areas which are the natural breeding or
feeding grounds of game animals or birds may be closed to grazing.
Areas within National Forests infested seriously by poisonous plants
may be closed to grazing.
Protection of Grazing Interests. The protection of National Forest
grazing interests is secured by the prevention of overgrazing, by the
prevention of damage to roads, trails, or water sources, by the
proper bedding of sheep and goats, by the proper disposition of
carcasses, by salting the stock and by the proper observation of the
national and state live stock and quarantine laws.
When an owner, who has a permit, is ready to drive in his stock
upon the National Forest he must notify the nearest Forest officer
concerning the number to be driven in. If called upon to do so he
must provide for having his stock counted before entering a National
Forest. Each permittee must repair all damage to roads or trails
caused by the presence of his stock. Sheep and goats are not
allowed to be bedded more than three nights in succession in the
same place (except during the lambing season) and must not be
bedded within 300 yards of any running or living spring. The
carcasses of all animals which die on the National Forests from
contagious or infectious diseases must be burned and are not
permitted to lie in the close vicinity of water. In order to facilitate the
handling of stock and prevent their straying off their range, they
must be salted at regular intervals and at regular places.
In order to facilitate the moving of stock by stockmen from their
home ranches to their grazing allotments and to minimize the
damage of grazing animals to the Forests, stock driveways are
established over regular routes of travel.

SPECIAL USES
All uses of National Forest lands and resources permitted by the
Secretary of Agriculture, except those specifically provided for in the
regulations covering water power, timber sales, timber settlement,
the free use of timber, and grazing, are designated "special uses."
Among these are the use or occupancy of lands for residences,
farms, apiaries, dairies, schools, churches, stores, mills, factories,
hotels, sanitariums, summer resorts, telephone and telegraph lines,
roads and railways; the occupancy of lands for dams, reservoirs and
conduits not used for power purposes; and the use of stone, sand,
and gravel. No charge is made for a large number of these permits,
some of which are the following: (1) agricultural use by applicants
having preference rights under the Act of June 11, 1906; (2)
schools, churches, and cemeteries; (3) cabins for the use of miners,
prospectors, trappers, and stockmen in connection with grazing
permits; (4) saw mills sawing principally National Forest timber; (5)
conduits, and reservoirs for irrigation or mining or for municipal
water supply; (6) roads and trails (which must be free public
highways); (7) telephone lines and telegraph lines with free use of
poles and connections for the Forest Service.
The occupancy and use of National Forest land or resources
under a special use permit (except those given free of charge) are
conditioned upon the payment of a charge and are based upon
certain rates. Agricultural use of land is given to permittees at a
charge of from 25 cents to $1.00 an acre. Not over 160 acres are
allowed to any one permittee. Cabins cost from $3.00 to $5.00; hay
cutting from 20 to 50 cents an acre; hotels and roadhouses from
$10.00 to $50.00; pastures from 4 to 25 cents per acre; residences
covering from one to three acres cost from $5.00 to $25.00; resorts
from $10.00 to $50.00; stores from $5.00 to $50.00 for two acres or
less; and other uses in proportion.
Perhaps the use that is purchased most of all on the National
Forests is that for residences and summer homes. On many of the
Forests they are already in great demand. A large proportion of the
population of the far Western States seek the cool and invigorating
air of the mountains in the early summer because the heat of the
valleys, especially in California, is almost unbearable.
There are many desirable pieces of land on the National Forests
that are being reserved by the Forest Service especially for this
purpose for the people of the neighboring towns. For example, on
the Angeles National Forest in California the Supervisor had about
250 suitable sites surveyed in one picturesque canyon and in six
months 226 of them were under special use permits as summer
homes. A large reservoir—Huntington Lake—was constructed on the
Sierra National Forest in California as the result of a dam constructed
by a hydro-electric power company. Immediately there was a keen
demand among the residents of San Joaquin Valley for summer
homes on the shores of the lake. In a few years it is expected there
will be a permanent summer colony of from 2,000 to 3,000 people.
The Forest Service has already authorized an expenditure of $1,500
in order to furnish an adequate supply of domestic water for the
colony.

CLAIMS AND SETTLEMENT


Claims can be initiated upon National Forest lands under (1) the
Act of June 11, 1906, (2) under the mining laws, and (3) under the
coal land laws. In connection with these claims it is the duty of the
Forest Service to examine them, but the determination of questions
involving title is within the jurisdiction of the Secretary of the
Interior.
It is the purpose of the Forest Service to protect the lands of the
United States within the National Forests from acquisition by those
who do not seek them for purposes recognized by law. When it is
apparent that an entry or a claim is not initiated in good faith and in
compliance with the spirit of the law under which it was asserted,
but is believed from the facts to be a subterfuge to acquire title to
timber land, or to control range privileges, water, a water-power site,
or rights of way; or if it otherwise interferes with the interests of the
National Forests in any way, the Forest Service recommends a
contest, even if the technical requirements of the law appear to have
been fulfilled. It is bad faith, for instance, to hold a mining or
agricultural claim primarily for the timber thereon or to acquire a site
valuable for water power development.
The National Forest Homestead Act. At the present time there is
very little, if any, fraud connected with the Forest Homestead Act
because the land is classified before it is opened to entry. The
greater part of the work dealing with fraudulent claims is a relic of
the old régime. Before the Forests were established many
Homestead and Timber and Stone entries were made for the
purpose of securing valuable timber. A large number of persons
resorted to settlement in order to secure the preference right. It was
the common custom in those days for land cruisers to locate men on
heavily timbered land either before or immediately after survey and
before the filing of the plats and the opening of the land to entry. A
cabin would be built upon the land and some unsubstantial
improvements made. When the National Forests were created they
contained great numbers of these squatters' cabins. Many were
abandoned but others attempted to secure title. Under the old
Timber and Stone Act timber could be secured for $2.50 per acre,
but the National Forests are not subject to entry under this act. So
as a last resort the squatters tried to prove up on the land under the
Homestead law. When the Forests were created the Service found a
great many of these fraudulent claims on their books, many of which
were being brought up annually for patent. Between December,
1908, and June 30, 1913, a total of 498 entries for National Forest
land were canceled in a single administrative district. These entries
represented fraudulent efforts to secure title to 85,906 acres of
National Forest land for speculative purposes, involving nearly a
billion feet of merchantable timber. During the fiscal year 1913 alone
300,000,000 board feet of merchantable timber in one district was
retained in public ownership primarily because the Forest officers
brought out the facts. The lands in all cases were covered with
heavy stands of timber, very small portions of the land had been
cleared, the claimant's residence on the land was not in compliance
with the law, seldom was any crop raised on the land, and the
claimant in other ways did not carry out the intent of the law.
The Act of June 11, 1906, known as the National Forest
Homestead Act, provides for the acquisition by qualified entrymen of
agricultural lands within National Forests. The Act is in effect an
extension of the general provisions of the Homestead laws to the
agricultural lands within the National Forests, with the essential
difference that the land must be classified by the Secretary of
Agriculture as chiefly valuable for agriculture.
This Act authorizes the Secretary of Agriculture in his discretion
to examine and ascertain, upon application or otherwise, the
location and extent of lands both surveyed and unsurveyed in the
National Forests, chiefly valuable for agriculture, which may be
occupied for agricultural purposes without injury to the National
Forests or public interests. He is authorized to list and describe such
lands by metes and bounds or otherwise and to file such lists and
descriptions with the Secretary of the Interior for opening to entry in
accordance with the provisions of the Act. Agricultural lands listed by
the Secretary of Agriculture are opened by the Secretary of the
Interior to homestead entry in tracts not exceeding 160 acres at the
expiration of 60 days from the filing of the lists in the local Land
Office. Notice of the filing of the list is posted in the local Land Office
and is published for a period of not less than four weeks in a local
newspaper. The Act provides that the person upon whose application
the land is examined and listed, if a qualified entryman, shall have
the preference right of entry. To exercise this preference right,
application to enter must be filed in the local Land Office within 60
days after the filing of the list in that office. The entryman can
perfect his title to the land within a certain period of years by
fulfilling certain conditions of residence and cultivation.
By the Act of June 6, 1912, known as the "Three Year
Homestead Act," the period of residence necessary to be shown in
order to entitle a person to patent under the Homestead laws is
reduced from 5 to 3 years and the period within which a homestead
entry may be completed is reduced from 7 to 5 years. The new law
requires the claimant to cultivate not less than 1/16 of the area of
his entry beginning with the second year of entry and not less than
1/8 beginning with the third year and until final proof, except that in
the case of the enlarged Homestead laws, double the areas given
are required. On a 160-acre claim, therefore, it is required that 1/8
or 20 acres be under cultivation. A mere breaking of the soil does
not meet the requirements of the statute, but such breaking of the
soil must be accompanied by planting and sowing of seed and tillage
for a crop other than native grasses. The period within which the
cultivation should be made is reckoned from the date of the entry.
The Secretary of the Interior, however, is authorized upon a
satisfactory showing therefor to reduce the required area of
cultivation on account of financial disabilities or misfortunes of the
entryman or on account of special physical and climatic conditions of
the land which make cultivation difficult. The entryman must
establish an actual residence upon the land entered, 6 months after
the date of the entry. After the establishment of residence the
entryman is permitted to be absent from the land for one continuous
period of not more than 5 months in each year following. He must
also file at the local Land Office notice of the beginning of such
intended absence.
The Mining Laws. Mineral deposits within National Forests are
open to development exactly as on unreserved public land. A
prospector can go anywhere he chooses and stake a claim wherever
he finds any evidences of valuable minerals. The only restriction is
that mining claims must be bona fide ones and not taken up for the
purpose of acquiring valuable timber or a town or a water power
site, or to monopolize the water supply of a stock range. Prospectors
may obtain a certain amount of National Forest timber free of charge
to be used in developing their claims. More than 500 mining claims
are patented within the National Forests every fiscal year.
A good example of mining claims located for fraudulent purposes
were those located on the rim and sides of the Grand Canyon in
Arizona to prevent the people from gaining free access to the canyon
and make them pay to enter it. These claims were shown to be
fraudulent since no deposits of any kind were ever found on them.
They were canceled by the higher courts and the land reverted to
the people.
Coal-Land Laws. Coal lands are mineral lands and as such are
subject to entry the same as other mineral lands in the National
Forests.

ADMINISTRATIVE USE OF NATIONAL FOREST


LANDS
Lands within National Forests may be selected for administrative
uses such as Supervisor's and Ranger's headquarters, gardens,
pastures, corrals, planting or nursery sites or rights-of-way. These
administrative sites are necessary for the present and probable
future requirements of the Forest Service for fire protection and the
transaction of business on the National Forests.

WATER POWER, TELEPHONE, TELEGRAPH,


AND POWER TRANSMISSION LINES
Along the streams within the National Forests are many sites
suitable for power development. These are open to occupancy for
such purposes and have the advantage of being on streams whose
headwaters are protected. The aggregate capacity of the water
power sites on the National Forests is estimated at 12,000,000
horsepower.
Figure 77. North Clear Creek Falls, Rio
Grande National Forest, Colorado. The
National Forests contain about one-third
of all the potential water-power resources
of the United States.

Figure 78. The power plant of the


Colorado Power Company, on the Grand
River, Holy Cross National Forest,
Colorado. Every fiscal year there is a
substantial increase in water power
development on the National Forests.
The Government does not permit the monopolization of power in
any region or allow sites to be held for speculative purposes. The
objects of the regulations are to secure prompt and full development
and to obtain a reasonable compensation for the use of the land
occupied and the beneficial protection given the watershed.
Permits for power development on the National Forests usually
run for a term of 50 years and may be renewed at their expiration
upon compliance with the regulations then existing. Such permits,
while granting liberal terms to applicants, contain ample provision for
the protection of the public interests.
Applications for power permits are filed with the District Forester
of the Forest Service District in which the desired site is located.
Preliminary permits are issued to protect an applicant's priority
against subsequent applicants until he has had an opportunity to
study the proper location and design of the project and to obtain the
data necessary for the final application. Operation is allowed under
the final permit only. The permittee is required to pay an annual
rental charge under the preliminary and final power permits and
definite periods are specified for the filing of the final application,
beginning of construction and of operation. The rental charges are
nominal in amount, the maximum being about 1/16 of a cent per
kilowatt hour. The amount of annual payment for transmission lines
is $5.00 for each mile or fraction thereof if National Forest land is
crossed by the line. No rental charges are made for small power
projects (under 100 horsepower capacity), or for transmission lines
used in connection therewith, or for transmission lines which are
part of a power project under permit or for any power project in
which power is to be used by a municipal corporation for municipal
purposes.
The Secretary of Agriculture has authority to permit the use of
rights-of-way through the National Forests for conduits, reservoirs,
power plants, telephone and telegraph lines to be used for irrigation,
mining, and domestic purposes and for the production and
transmission of electric power. No rental charges are made for the
telephone and telegraph rights-of-way, but the applicant must agree
to furnish such facilities to Forest officers and to permit such
reasonable use of its poles or lines as may be determined or agreed
upon between the applicant and the District Forester.

Figure 79. This is only one of the


thousands of streams in the National
Forests of the West capable of generating
electric power. It has been estimated that
over 40 per cent. of the water power
resources of the western states are
included in the National Forests. Photo by
the author.
Figure 80. View in the famous orange
belt of San Bernardino County, California.
These orchards depend absolutely upon
irrigation. The watersheds from which the
necessary water comes are in the
National Forests and are protected by the
Forest Service. Some of the smaller
watersheds in these mountains are said
to irrigate orchards valued at
$10,000,000.
Welcome to our website – the ideal destination for book lovers and
knowledge seekers. With a mission to inspire endlessly, we offer a
vast collection of books, ranging from classic literary works to
specialized publications, self-development books, and children's
literature. Each book is a new journey of discovery, expanding
knowledge and enriching the soul of the reade

Our website is not just a platform for buying books, but a bridge
connecting readers to the timeless values of culture and wisdom. With
an elegant, user-friendly interface and an intelligent search system,
we are committed to providing a quick and convenient shopping
experience. Additionally, our special promotions and home delivery
services ensure that you save time and fully enjoy the joy of reading.

Let us accompany you on the journey of exploring knowledge and


personal growth!

ebookfinal.com

You might also like