A Course in Metric Geometry
A Course in Metric Geometry
Dmitri Burago
Yuri Burago
Sergei Ivanov
Preface vii
Chapter 3. Constructions 59
§3.1. Locality, Gluing and Maximal Metrics 59
§3.2. Polyhedral Spaces 67
§3.3. Isometries and Quotients 74
iii
iv Contents
vii
viii Preface
away from the gist of the matter. For instance, our introduction to Riemann-
ian geometry begins with metrics on planar regions, and we even avoid the
notion of a manifold. Of course, manifolds do show up in more advanced sec-
tions. Some exercises and remarks assume more mathematical background
than the rest of our exposition; they are optional, and a reader unfamiliar
with some notions can just ignore them. For instance, solid background in
differential geometry of curves and surfaces in R3 is not a mandatory prereq-
uisite for this book. However, we would hope that the reader possesses some
knowledge of differential geometry, and from time to time we draw analogies
from or suggest exercises based on it. We also make a special emphasis on
motivations and visualizations. A reader not interested in them will be able
to skip certain sections. The first chapter is a clinic in metric topology; we
recommend that the reader with a reasonable idea of metric spaces just skip
it and use it for reference: it may be boring to read it. The last chapters
are more advanced and dry than the first four.
Figures. There are several figures in the book, which are added just to
make it look nicer. If we included all necessary figures, there would be at
least five of them for each page.
• It is a must that the reader systematically studying this book makes
a figure for every proposition, theorem, and construction!
Exercises. Exercises form a vital part of our exposition. This does not
mean that the reader should solve all the exercises; it is very individual.
The difficulty of exercises varies from trivial to rather tricky, and their
importance goes all the way up from funny examples to statements that
are extensively used later in the book. This is often indicated in the text.
It is a very helpful strategy to perceive every proposition and theorem as an
exercise. You should try to prove each on your own, possibly after having
a brief glance at our argument to get a hint. Just reading our proof is the
last resort.
situation. By now the scope of the theory of length spaces has grown quite
far from its cradle (which was a theory of convex surfaces), including most
of classical Riemannian geometry and many areas beyond it. At the same
time, geometry of length spaces perhaps remains one of the most “hands-
on” mathematical techniques. This combination of reasons urged us to write
this “beginners’ course in geometry from a length structure viewpoint”.
Metric Spaces
The purpose of the major part of the chapter is to set up notation and to
refresh the reader’s knowledge of metric spaces and related topics in point-
set topology. Section 1.7 contains minimal information about Hausdorff
measure and dimension.
It may be a good idea to skip this chapter and use it only for reference,
or to look through it briefly to make sure that all examples are clear and
exercises are obvious.
1.1. Definitions
Definition 1.1.1. Let X be an arbitrary set. A function d : X × X →
R ∪ {∞} is a metric on X if the following conditions are satisfied for all
x, y, z ∈ X.
(1) Positiveness: d(x, y) > 0 if x 6= y, and d(x, x) = 0.
(2) Symmetry: d(x, y) = d(y, x).
(3) Triangle inequality: d(x, z) ≤ d(x, y) + d(y, z).
A metric space is a set with a metric on it. In a formal language, a metric
space is a pair (X, d) where d is a metric on X. Elements of X are called
points of the metric space; d(x, y) is referred to as the distance between
points x and y.
When the metric in question is clear from the context, we also denote
the distance between x and y by |xy|.
Unless different metrics on the same set X are considered, we will omit
an explicit reference to the metric and write “a metric space X” instead of
“a metric space (X, d).”
1
2 1. Metric Spaces
Semi-metrics.
Definition 1.1.4. A function d : X × X → R+ ∪ {+∞} is called a semi-
metric if it satisfies all properties from Definition 1.1.1 of a metric except
the requirement that d(x, y) = 0 implies x = y. This means that we allow
zero distance between different points.
1.2. Examples
Various examples of metric spaces will appear everywhere in the course. In
this section we only describe several important ones to begin with. For many
of them, verification of the properties from Definition 1.1.1 is trivial and is
left for the reader.
Example 1.2.1. One can define a metric on an arbitrary set X by
(
0 if x = y,
|xy| =
1 if x 6= y.
This example is not particularly interesting but it can serve as the initial
point for many constructions.
Example 1.2.2. The real line, R, is canonically equipped with the distance
|xy| = |x − y|, and thus can be considered as a metric space. There is an
immense variety of other metrics on R; for instance, consider dlog (x, y) =
log |x − y|.
Example 1.2.3. The Euclidean plane, R2 , with its standard distance,
is another familiar metric space. The distance can be expressed by the
Pythagorean formula,
p
|xy| = |x − y| = (x1 − y1 )2 + (x2 − y2 )2
where (x1 , x2 ) and (y1 , y2 ) are coordinates of points x and y. The triangle
inequality for this metric is known from elementary Euclidean geometry.
Alternatively, it can be derived from the Cauchy inequality.
Example 1.2.4 (direct products). Let X and Y be two metric spaces. We
define a metric on their direct product X × Y by the formula
p
|(x1 , y1 )(x2 , y2 )| = |x1 x2 |2 + |y1 y2 |2 .
In particular, R × R = R2 .
Exercise 1.2.5. Derive the triangle inequality for direct products from the
triangle inequality on the Euclidean plane.
Example 1.2.6. Recall that the coordinate n-space Rn is the vector space
of all n-tuples (x1 , . . . , xn ) of real numbers, with component-wise addition
and multiplication by scalars. It is naturally identified with the multiple
direct product R × · · · × R (n times). This defines the standard Euclidean
distance,
p
|xy| = (x1 − y1 )2 + · · · + (xn − yn )2
where x = (x1 , . . . , xn ) and y = (y1 , . . . , yn ).
4 1. Metric Spaces
Restricting the distance is the simplest but not the only way to define a
metric on a subset. In many cases it is more natural to consider an intrinsic
metric, which is generally not equal to the one restricted from the ambient
space. The notion of intrinsic metric will be explained further in the course,
but its intuitive meaning can be illustrated by the following example of the
intrinsic metric on a circle.
Example 1.2.9. The unit circle, S 1 , is the set of points in the plane lying at
distance 1 from the origin. Being a subset of the plane, the circle carries the
restricted Euclidean metric on it. We define an alternative metric by setting
the distance between two points as the length of the shorter arc between
them. For example, the arc-length distance between two opposite points of
the circle is equal to π. The distance between adjacent vertices of a regular
n-gon (inscribed into the circle) is equal to 2π/n.
Exercise 1.2.10. (a) Prove that any circle arc of length less or equal to π,
equipped with the above metric, is isometric to a straight line segment.
(b) Prove that the entire circle with this metric is not isometric to any
subset of the plane (regarded with the restriction of Euclidean distance onto
this subset).
It is easy to see that a normed space with the above distance is a metric
space. The norm is recovered from the metric as the distance from the
origin.
The Euclidean space Rn described in Example 1.2.6 is a normed space
whose norm is expressed by
q
|(x1 , . . . , xn )| = x21 + · · · + x2n .
There are other natural norms in Rn .
Example 1.2.12. The space Rn1 is the coordinate space Rn with a norm
k · k1 defined by
k(x1 , . . . , xn )k1 = |x1 | + · · · + |xn |
(where | · | is just the absolute value of real numbers).
Example 1.2.13. Similarly, the space Rn∞ is Rn with a norm k · k∞ where
k(x1 , . . . , xn )k∞ = max{|x1 |, . . . , |xn |}.
Exercise 1.2.14. Prove that
(a) R21 and R2∞ are isometric;
(b) Rn1 and Rn∞ are not isometric for any n > 2.
Example 1.2.15. Let X be an arbitrary set. The space ℓ∞ (X) is the set
of all bounded functions f : X → R. This is naturally a vector space with
respect to pointwise addition and multiplication by scalars. The standard
norm k · k∞ on ℓ∞ (X) is defined by
kf k∞ = sup |f (x)|.
x∈X
Exercise 1.2.16. Show that Rn∞ = ℓ∞ (X) for a suitable set X. Hint: an
n-tuple (x1 , . . . , xn ) is formally a map, isn’t it?
1.2.3. Spheres.
Example 1.2.25. The n-sphere S n is the set of unit vectors in Rn+1 , i.e.,
S n = {x ∈ Rn+1 : |x| = 1}. The angular metric on S n is defined by
d(x, y) = arccos hx, yi .
In other words, the spherical distance is defined as the Euclidean angle
between unit vectors. It equals the length of the shorter arc of a great circle
connecting x and y in the sphere. Another formula for this metric is
|x − y|
d(x, y) = 2 arcsin .
2
The metric on the circle described in Example 1.2.9 is a partial case of this
example.
Exercise 1.3.2. Let x1 and x2 be points of some metric space, and let r1
and r2 be positive numbers. Show that
(a) if |x1 x2 | ≥ r1 + r2 , then the balls Br1 (x1 ) and Br2 (x2 ) are disjoint;
(b) if |x1 x2 | ≤ r1 − r2 , then Br2 (x2 ) ⊂ Br1 (x1 );
(c) the converse statements to (a) and (b) are not always true (give
counterexamples).
Exercise 1.3.4. Prove that a metric product carries the standard product
topology.
The following two exercises provide useful tools for proving completeness
of some spaces.
Exercise 1.5.6. Let {xn } be a Cauchy sequence in a metric space. Prove
that
(a) If {xn } has a converging subsequence, then it converges itself.
(b) For any sequence {εn } of positive numbers there exists a subsequence
{yn } of {xn } such that |yn yn+1 | < εn for all n.
Exercise P1.5.7. Let {xn }∞
n=1 be a sequence in a metric space such that
the series ∞ |x x
n=1 n n+1 | has a finite sum. Prove that {xn } is a Cauchy
sequence.
Exercise 1.5.8 (fixed-point theorem). Let X be a complete space, 0 < λ <
1, and let f : X → X be a map such that |f (x)f (y)| ≤ λ|xy| for all x, y ∈ X.
Prove that there exists a unique point x0 ∈ X such that f (x0 ) = x0 .
Hint: Obtain x0 as the limit of a sequence {xn } where x1 is an arbitrary
point and xn+1 = f (xn ) for all n ≥ 1.
Proof of Theorem 1.5.10. Let X denote the set of all Cauchy sequences
in X. Introduce the distance in X by the formula
d({xn }, {yn }) = lim |xn yn |.
n→∞
It is easy to check that, if {xn } and {yn } are Cauchy sequences, then {|xn yn |}
is either a Cauchy sequence of real numbers or |xn yn | = ∞ for all large
enough n. Therefore the above limit always exists. Clearly d is a semi-
metric on X. Define X̃ = X/d (see Proposition 1.1.5 and a remark after
it).
There is a natural map from X to X̃, namely let a point x ∈ X be
mapped to a point of X̃ represented by the constant sequence {x}∞ n=1 .
Since this map is distance-preserving, we can identify X with its image
in X̃ (formally, change the definition of X̃ so that points of X replace their
images). This way X becomes a subset of X̃. It is dense because a point of
X̃ represented by a sequence {xn } is the limit of this sequence (thought of
as the sequence in X ⊂ X̃).
The uniqueness part of the theorem follows from Proposition 1.5.9
applied to the inclusion maps from X to X̃ and X̃ ′ . ¤
1.6. Compact Spaces 13
Baire’s theorem.
Definition 1.5.12. A set Y in a topological space X is nowhere dense if
the closure of Y has empty interior.
Proof of the theorem. Let X be a complete metric space and {Yi }∞ i=1 be
a countable family of nowhere dense sets. We have to S show that any open
set U ⊂ X contains a point which does not belong to ∞ i=1 Yi . Since Y1 is
nowhere dense, there is a (closed) ball B1 ⊂ U which does not intersect Y1 .
Since Y2 is nowhere dense, there is a closed ball B2 ⊂ B1 which does not
intersect Y2 . And so on. This way we obtain a sequence B1 ⊃ B2 ⊃ . . . of
closed balls where each ball Bi has no common points with the respective
set Yi . We may choose the radii of the balls Bi so that they converge to
zero. Then the centers of the balls form a Cauchy sequence. The limit of
this sequence belongs to all balls and therefore does not belong to any of
the sets Yi . ¤
Exercise 1.6.8. Prove that a metric space (with possibly infinite distances)
is compact if and only if it is a union of a finite number of compact subsets
each of which carries a finite metric.
Proof. We may assume that the metric of X is finite and none of the sets
Uα covers the whole space. Then one can define a function f : X → R by
Since {Uα } is an open covering, f (x) is well-defined and positive for all
x ∈ X. Clearly f is nonexpanding function and hence continuous. Therefore
it attains a (positive) minimum r0 . Define ρ = r0 /2. ¤
Exercise 1.6.13. Prove that a locally Lipschitz map from a compact space
is Lipschitz.
16 1. Metric Spaces
Proof. Suppose the contrary, i.e., let p ∈ X \ f (X). Since f (X) is compact
and hence closed, there exists an ε > 0 such that Bε (p) ∩ f (X) = ∅.
Let n be the maximal possible cardinality of an ε-separated set in X (see
Exercise 1.6.4) and let S ⊂ X be an ε-separated set of cardinality n. Since f
is distance-preserving, the set f (S) is also ε-separated. On the other hand,
dist(p, f (S)) ≥ dist(p, f (X)) ≥ ε and therefore f (S) ∪ {p} is an ε-separated
set of cardinality n + 1. Contradiction. ¤
Theorem 1.6.15. Let X be a compact metric space. Then
(1) Any nonexpanding surjective map f : X → X is an isometry.
(2) If a map f : X → X is such that |f (x)f (y)| ≥ |xy| for all x, y ∈ X,
then f is an isometry.
Proof. 1. Suppose the contrary, i.e., that |f (p)f (q)| < |pq| for some points
p, q ∈ X. Fix p and q and pick an ε > 0 such that |f (p)f (q)| < |pq| − 5ε.
Let n be a natural number such that there exists at least one ε-net in
X of cardinality n. Consider the set N ⊂ X n of all n-tuples of points of X
that form ε-nets in X. This set is closed in X n and therefore it is compact.
Define a function D : X n → R by
n
X
D(x1 , . . . , xn ) = |xi xj |.
i,j=1
The second
P condition is referred to as σ-additivity. Note that the
expression µ(Ai ) is either a finite sum or a series; its value is well defined
and independent of the order of terms since the terms are nonnegative.
Exercise 1.7.4. Let µ be a measure. Prove the following statements:
(a) Let {Ai }∞
i=1 be a sequence of measurable sets such that Ai ⊂ Ai+1
S for
all i. Then the sequence {µ(Ai )} is nondecreasing and lim µ(Ai ) = µ( Ai ).
(b) Let {Ai }∞
i=1 be a sequence of measurable sets such that Ai ⊃ Ai+1
for all i, and assume that µ(A1T ) < ∞. Then the sequence {µ(Ai )} is
nonincreasing and lim µ(Ai ) = µ( Ai ).
(c) The assumption µ(A1 ) < ∞ in (b) is essential.
measures of such covers. Speaking about “the total measure of a cover” one
means here that certain measure is already assigned to simple sets.
To define Hausdorff n-dimensional measure on a metric space, one could
proceed along the same lines: cover a set by metric balls such that all their
radii are less than ε. For each ball, consider a Euclidean ball of the same
radius and add their volumes for all balls from the cover: this will be the
total measure of the cover. Taking its infimum over all covers and passing
to the limit as ε approaches zero, one gets a version of Hausdorff measure of
the set. Instead of adding volumes of Euclidean balls of the same radii, one
could simply add the radii of balls from the cover raised to the power n: the
result is the same up to a constant multiplier. It turns out that arbitrary
sets and diameters are technically more convenient to use than metric balls
and radii.
Now we pass to formal definitions.
Now we can define the normalization constant C(n) from the definition of
Hausdorff measure. Namely, choose C(n) so that µn (I n ) = 1. The existence
of such a constant follows from Theorem 1.7.12. Theorem 1.7.5 then implies
1.7. Hausdorff Measure and Dimension 21
Proof. We may assume that every ball B ∈ B contains at least one point
of X and exclude the balls with radius greater than 1. Then all these balls
are contained in the 2-neighborhood of X which is bounded and hence has
finite volume. We construct a sequence {Bi }∞ i=1 of balls by induction. If
B1 , . . . , Bm are already constructed, we choose the next ball Bm+1 as follows.
Let Bm denote the set of balls from the collection that do not intersect any
of B1 , . . . , Bm . If Bm is empty, then B1 ∪ · · · ∪ Bm covers the entire set X
and the proof is finished (this follows from the condition that every point
is covered by balls of arbitrarily small radii). If Bm is not empty, choose
Bm+1 to be any element of Bm with
1
(1.1) diam(Bm+1 ) > sup{diam(B) : B ∈ Bm }.
2
The balls Bi are disjoint by the construction. We will now show that they
cover X up to a set of zero measure. Fix an ε > 0. Since the P∞balls are disjoint
and are contained in a set of finiteP volume, we have i=0 µn (Bi ) <S ∞.
Hence there is an index m such that ∞ i=m+1 µ n (Bi ) < ε. Let x ∈ X \ i Bi
and let B be any ball from the collection that contains x S and does not
intersect the balls B1 , . . . , Bm . Note that B must intersect i Bi because
otherwise B ∈ Bm for all m which contradicts that µn (Bi ) → 0. Let k
be the minimal index such that B ∩ Bk 6= ∅. Then B ∈ Bk−1 and hence
diam(Bk ) > 12 diam(B) by (1.1). It follows that the distance from x to the
center of Bk is not greater than 5 times the radius of Bk . Hence x belongs
to the ball with the same center as Bk and radius 5 times larger. We denote
this ball by 5Bk .
S
We have just provedSthat every S∞x ∈ X \ i Bi belongs to a ball 5Bk for
some k > m. Thus X \ i Bi ⊂ i=m+1 (5Bi ); hence
[ ∞
X ∞
X
n
µn (X \ Bi ) ≤ µn (5Bi ) = 5 µn (Bi ) < 5n ε.
i i=m+1 i=m+1
22 1. Metric Spaces
S
Since ε is arbitrary, it follows that µn (X \ i Bi ) = 0. ¤
Corollary 1.7.15. The normalization constant for the n-dimensional Haus-
dorff measure equals the volume of the Euclidean n-ball of diameter 1.
Proof. Let Cn denote the constant from the formulation. Then the volume
of a Euclidean n-ball equals Cn dn where d is its diameter. Let µ′n be the
n-dimensional Hausdorff measure with normalization constant Cn . We have
to prove that µ′n = µn , i.e., that µ′n (I n ) = 1.
1. µ′n (I n ) ≤ 1. To prove this, apply Theorem 1.7.14 to the set B of all
closed Euclidean balls contained in I n . ThisSyields a countable collection
{Bi } of such balls such
P that the set Y = I nP\ Bi has zero measure. Hence
µn (I ) ≤ µn (Y ) + Cn diam(Bi ) = 0 + mn (Bi ) ≤ mn (I n ) = 1.
′ n ′ n
1.7.4. Hausdorff dimension. The next theorem tells us how the Haus-
dorff measure of a fixed set depends on dimension. Briefly, the measure is
zero or infinite for all dimensions except at most one. More precisely, there
is a “critical dimension” below which the measure is infinity and above
which the measure is zero. This dimension is an important characteristic
of a metric space, called the Hausdorff dimension. Warning: at the critical
dimension, all three possibilities (the measure is zero, positive number or
+∞) may take place.
Theorem 1.7.16. For a metric space X there exists a d0 ∈ [0, +∞] such
that µd (X) = 0 for all d > d0 and µd (X) = ∞ for all d < d0 .
Length Spaces
A
B
PSfrag replacements
A
B
Figure 2.1: “A crow flies” along the segment AB; for a pedestrian it probably
takes longer.
25
26 2. Length Spaces
Did you notice that you have used the following fact?
Exercise 2.1.4. Prove that admissible paths of finite length are continuous
with respect to (X, dL ).
This exercise deals with the topology determined by the metric dL rather
than the initial topology of the space X. There really are examples where
these two topologies differ; such examples will appear later in this book.
Exercise 2.1.5. Prove that the topology determined by dL can be only
finer than that of X: any open set in X is open in (X, dL ) as well.
2.1. Length Structures 29
Not every metric can arise as a length metric. Even if (X, d) is a length
space and A ⊂ X, the restriction of d to A is not necessarily intrinsic. For
example, consider a circle in the plane.
Moreover, not every metrizable topology can be induced by intrinsic
metrics:
Exercise 2.1.7. 1. Prove that the set of rational numbers is not homeo-
morphic to a length space.
2. Prove that the union of the graph {(x, y) : y = sin(1/x), x > 0} and
the y-axis (with its topology inherited from R2 ) is not homeomorphic to a
length space.
There can be more delicate reasons why a topological space may be not
homeomorphic to a length space:
Exercise 2.1.8. Consider the union of segments
∞
[
[(0, 0), (cos 1/i, sin 1/i)] ∪ [(0, 0), (1, 0)]
i=1
in the Euclidean plane, depicted in Figure 2.2.
Figure 2.2: The space (with the topology inherited from R2 ) is not homeomorphic
to a length space.
Exercise 2.1.9. Prove that a length space is locally path connected: every
neighborhood of any point contains a smaller neighborhood which is path-
connected.
Admissible paths are all (piecewise smooth) paths contained in the region. If
the region is convex, this length structure induces usual Euclidean distance.
One may think of this region as an island, and the distance is measured by
a creature who cannot swim. Drawing balls in intrinsic metrics arising this
way may be quite fun; see Figure 2.3. Is this metric strictly intrinsic? What
if we consider the closure of the region?
The reader can generalize this example for a subspace of a space with
length structure. Certainly, only sensible choices for a subspace lead to rea-
sonable examples. For instance, restricting the Euclidean length structure
in R2 to a circle leads to the angular metric. More generally, one obtains
spherical geometry by restricting the usual Euclidean length structure to a
round sphere. On the other hand, restricting the standard length structure
of R to the set of rational points we obtain a space with each accessibility
component consisting of just one point.
Example 2.2.3 (induced length structure). The formal contents of this
example is comprised in the following definition. Let f : X → Y be a
continuous map from a topological space X to a space Y endowed with a
length structure. One defines the induced length structure in X as follows.
A path in X is admissible if its composition with f is admissible in Y . The
length of an admissible path in X is set to the length of its composition with
f with respect to the length structure in Y .
(In fact, this construction may not define a length structure in X because
the new length function may fail to satisfy the fourth condition from section
2.1.1. We use the term “induced length structure” only if this is indeed a
length structure.)
32 2. Length Spaces
At first glance, the above definition may sound like a tautology. However,
the properties of an induced metric may drastically differ from the properties
of a metric we began with. For instance, the leading example of an induced
metric when f is a surface (that is an immersion f : Ω ⊂ R2 → R3 of a
two-dimensional region into R3 ) has served as the main motivating example
in metric geometry for over a century. For a reader who is already familiar
with Riemannian metrics, we mention that it is also true (though hard to
believe and not easy to prove) that every Riemannian length structure on
Rn can be induced by a map f : Rn → Rn (which makes lots of folds and is
rarely smooth).
Example 2.2.4 (“Crossing a swamp”: conformal length). The space is the
Euclidean plane, and admissible paths are all (piecewise smooth) paths. Let
f : R2 → R be a positively-valued continuous (or even L∞ ) function. Define
the length of a path γ : [a, b] → R2 by
Z b
L(γ) = f (γ(t)) · |γ ′ (t)|dt.
a
This length structure can be thought of as a weighted Euclidean distance.
For instance, a traveler who measures the length (=time needed to cover)
of a certain route would apparently assign big values to f in a territory
that is difficult to traverse (for instance, a swamp or a mountain trail).
From the mathematical viewpoint, this is the first example of a Riemannian
length structure, which will be discussed further in Chapter 5; the word
“conformal” in the title of this subsection reflects the fact that such types
of Riemannian structures are called conformally flat.
Example 2.2.5 (Finslerian length). Thinking of the previous example as
a length structure for a traveler who assigns weights to different parts of
his/her path, one notices that an important feature of real travel is not
reflected here. Namely, the difficulty of traversing a region depends not
only on the region itself but also on the direction of the route; for instance,
choosing a direction in which most ravines are oriented might essentially
simplify the trip. To incorporate this additional information, one introduces
a function f in two variables and applies it to both γ and its velocity γ ′ .
The expression for the length reads:
Z b
L(γ) = f (γ(t), γ ′ (t)) dt.
a
(A physics-oriented reader recognizes that this structure can be interpreted
as action.) In order for this expression to be invariant under bijective
reparameterizations of paths, one has to require that f satisfies f (x, kv) =
|k|f (x, v) for all scalars k, points x and vectors v (check this as an exercise for
change of variable in a definite integral). Usually a stronger requirement is
2.3. Length Structures Induced by Metrics 33
imposed on f , namely for every point x the function f (x, ·) must be a norm.
A motivation for this will be explained in section 2.4.2. Length structures
obtained from this type of constructions are called Finslerian, or Finsler.
Remark 2.2.6. A reader who seriously tests our definitions against needs
of travelers and mountaineers will notice that some features are still missing.
Namely the fact that walking downhill may be easier than climbing uphill
cannot be reflected in a length structure. Since the distance in a metric
space must be symmetric, we had to require that the length is invariant
under all changes of variable including the orientation-reversing ones like
t 7→ −t. One could modify the definitions to allow nonsymmetric length
structures and metrics. Everything in this chapter can be adapted to such
generalized settings; however, this would not make sense in the rest of the
book.
Indeed, some of the main examples of length structures are those induced
by metric structures. For admissible paths one may use just all continuous
paths; for some of them the length may be infinite. In some cases a better
choice is the class of Lipschitz paths, that is, the class of maps γ : [a, b] → X
such that dX (γ(t), γ(t′ )) ≤ C|t − t′ |, for all t, t′ ∈ [a, b]; C is a positive
constant.
How do we define the length of a path in Euclidean space? We approx-
imate the path by broken lines and define the length as the limit of their
lengths. For each of these broken lines, its vertices belong to (the image
of) the path and they are well ordered with respect to the parameter of
the path. Since all that we actually use are distances between neighboring
vertices in this list, we can mimic this definition in a general metric space
in the most straightforward way (compare also with 2.4.12).
over all the partitions Y is called the length of γ (with respect to the metric
d) and denoted Ld (γ). A curve is said to be rectifiable if its length is finite.
The length structure induced by the metric d is defined as follows: all
continuous paths (parameterized by closed intervals) are admissible, and the
length is given by the function Ld .
This definitions can be formally applied to any metric space, but one
gets sensible examples only by a wise choice of a metric space to begin with:
for instance, if we start with a discrete space, there are no nonconstant
continuous paths at all. If it is clear from the context which metric d induces
the length L, we usually drop d in the notation Ld .
The usual “Euclidean” definition uses passing to a limit as the edges of
broken lines approach zero. The following exercise shows that there is no
difference here:
Exercise 2.3.3. Prove that the definition of length is compatible with the
one used in differential geometry. Namely if (V, | · |) is a finite-dimensional
normed vector space and γ : [a, b] → V is a differentiable map, then
Rb
L(γ) = a |γ ′ (t)| dt.
2.3. Length Structures Induced by Metrics 35
Proof. (i) Indeed, the triangle inequality implies that Σ(Y ) ≥ d(γ(a), γ(b))
for all Y ’s, and thus the inequality persists under passing to the limit.
(ii) First notice that if Y ′ is obtained from Y by adding one point, then
Σ(Y ) ≥ Σ(Y ) (by the same triangle inequality). Adding c to a partition Y
of [a, b] and then splitting it into two partitions of [a, c] and [c, b] completes
the argument.
We repeat again that the readers should try to consider such lemmas
as exercises and try to prove them on their own. If reading the proof was
needed, then drawing a figure with all notations is a must!
(iii) We prove the continuity of L in d, a < d ≤ b, from the left (the
other cases are analogous and are left to the reader). Take ε > 0 and
consider a partition Y such that L(γ) − Σ(Y ) < ε. One may suppose that
yj−1 < d = yj . Then
L(γ, yj−1 , d) − d(γ(yj−1 ), γ(d)) < ε,
and the same inequality takes place for each c such that yj−1 ≤ c ≤ d.
(iv) Let paths γj converge pointwise to γ. Take ε > 0 and fix a partition
Y for γ such that L(γ) − Σ(Y ) < ε. Now consider the sums Σj (Y ) for paths
γj corresponding to the same partition Y . Choose j to be so large that the
inequality d(γj (yi ), γ(yi )) < ε holds for all yi ∈ Y . Then
L(γ) ≤ Σ(Y ) + ε ≤ Σj (Y ) + ε + (N + 1)ε ≤ L(γj ) + (N + 2)ε.
Since ε is arbitrary, this implies (iv). ¤
Remark 2.3.5. In general, functional L is not continuous. A stairs-like
example is shown in Figure 2.4.
36 2. Length Spaces
PSfrag replacements
A
BA
where db = dLd .
Exercise 2.3.6. Prove that the intrinsic metric induced by the restriction
of Euclidean distance to the circle x2 + y 2 = 1 is the angular metric.
Note that the topology of the induced intrinsic metric may be very poorly
connected with the original topology of the space, as can be seen from the
following examples-exercises:
Exercise 2.3.7. Find the induced intrinsic metric for the metric
p
d((x1 , y1 ), (x2 , y2 )) = |x1 − x2 | + |y1 − y2 |
on R2 . What is the topology of the resulting length space?
Answer: A continuum of disjoint real lines, each with its standard metric.
Exercise 2.3.8. Consider the union of segments
∞
[
U= [(0, 1), (1/n, 0)] ∪ [(0, 1), (0, 0)] ⊂ R2 .
n=1
The sequence of points {(1/n, 0)} converges to (0, 0) in the topology inher-
ited by U from R2 . Since all pairwise distances between these points in the
induced intrinsic metric are at least 2, this sequence diverges with respect
to the intrinsic metric. Prove these statements.
Exercise 2.3.9. Connecting the point (0, 1) ∈ R2 with all points of the
standard Cantor set in the segment [0, 1] = [0, 1] × {0} ⊂ R2 , one obtains a
2.3. Length Structures Induced by Metrics 37
compact connected set. Show that in the induced intrinsic metric this set is
noncompact although it is still connected.
Exercise 2.3.10. Begin with a simple nonrectifiable curve in R2 ⊂ R3 and
build a cone over its image by choosing a point (vertex of the cone) in
R3 and connecting the vertex with every point in the image of the curve
by a segment. The intrinsic distance in this cone induced by Euclidean
distance in R3 is finite for every two points: one can go from one point to
the vertex of the cone along a straight segment and then get to the other
point along another segment. Show that removing the vertex of the cone
makes it disconnected in the topology of induced intrinsic metric, while it is
still connected in usual topology.
You can begin with a simple curve whose restriction to any nontrivial
interval is nonrectifiable (prove that such a curve does exist). In the original
topology this cone is still homeomorphic to a disc. Prove that in the induced
intrinsic metric this cone is homeomorphic to the bouquet of a continuum
of intervals (that is, the disjoint union of segments glued at one point).
Exercise 2.3.11. For two vectors V, W ∈ R2 , set
¯ ¯ p
d(V, W ) = ¯|V | − |W |¯ + min(|V |, |W |) · ∡(V, W ),
where ∡(V, W ) denotes the angle between V and W . Prove that
(i) the topology determined by d is the standard Euclidean one;
(ii) the induced intrinsic metric db is
(¯ ¯
¯|V | − |W |¯ if ∡(V, W ) = 0,
b
d(V, W ) = ¯ ¯
¯|V | + |W |¯ otherwise;
One can consider the length Ldb induced by the new metric d, b and this
length in turn determines a “second-stage” intrinsic metric. The reader may
imagine how much confusion between various lengths and metrics might
arise. However, this is not the case, as the length induced by db is the same
as induced by d.
Proposition 2.3.12. Let (X, d) be a metric space and db be the intrinsic
metric induced by d.
(1) If γ is a rectifiable curve in (X, d), then Ldb(γ) = Ld (γ).
(2) The intrinsic metric induced by db coincides with d.b In other words,
inducing a length metric is an idempotent operation.
Proof. The fact that the length of every curve in (X, d) is not less than the
distance between its endpoints implies that db ≥ d. It follows immediately
38 2. Length Spaces
that Ldb(γ) ≥ Ld (γ). To prove the inverse inequality, let [a, b] be the domain
of γ and let Y = {yi } be an arbitrary partition of [a, b]. Observe that
b
d(γ(y i ), γ(yi+1 )) ≤ Ld (γ, yi , yi+1 ) because the left-hand value is the infimum
of lengths one of which is written on the right-hand side. Therefore
X
Σdb(Y ) = b
d(γ(y i ), γ(yi+1 )) ≤ Ld (γ).
Proof. Let L be the length function that defines d and Ld be the length
induced by d. Observe that Ld (γ) ≤ L(γ) for any admissible curve γ of finite
length (repeat the respective argument from the proof of Proposition 2.3.12).
Obviously a smaller length function determines a smaller metric; thus db ≤ d.
On the other hand, we already know that db ≥ d; hence db = d. ¤
Proof. The inequality Ld (γ) ≤ L(γ) holds for any length structure—see
Propositions 2.3.12 and 2.4.1. Let us prove the opposite inequality. By
property 2 of length structure, the function L(t) = L(γ|[a,t] ) is uniformly
continuous in [a, b] for each rectifiable curve γ : [a, b] → X. Hence for
every ε > 0, there exists a partition a = t0 ≤ t1 ≤ · · · ≤ tk+1 = b such
that dL (γ(ti ), γ(ti+1 )) < ε for every integer i between 0 and k. According
to the definition of dL , for each i = 0, 1, . . . , k there exists a curve σi :
[ti , ti+1 ] → X with endpoints σi (ti ) = γ(ti ), σi (ti+1 ) = γ(ti+1 ) such that
L(σi ) ≤ d(γ(ti ), γ(ti+1 )) + ε/k. For the concatenation hε of the curves σi we
have
k
X k
X
L(hε ) = L(σi ) ≤ dL (γ(ti ), γ(ti+1 )) + ε ≤ Ld (γ) + ε.
i=0 i=0
From the triangle inequality one readily sees that d(γ(t), hε (t)) ≤ 3ε for
every t ∈ [a, b]. It follows that hε (t) → γ(t) since the topology determined
by d is finer than the initial one. Now the lower semi-continuity for L implies
L(γ) ≤ lim inf L(hε ) ≤ Ld (γ),
ε→0
Proof. Repeat the same arguments as in the proof of the previous lemma
for a path γ connecting x and y, and such that L(γ) − d(x, y) ≤ ε. ¤
Exercise 2.4.11. Let d be an intrinsic metric. Show that, if r1 + r2 >
d(x, y), then the balls Br1 (x) and Br2 (y) have a nonempty intersection.
42 2. Length Spaces
For intrinsic metrics the last formula in the corollary should be replaced
Pk−1
by i=1 d(xi , xi+1 ) − d(x, y) ≤ ε.
Exercise 2.4.13. If x and y are two points in a length space (X, d) and
r < d(x, y), then dist(y, Br (x)) = d(x, y) − r. Prove this.
Exercise 2.4.14. Let X be a length space, Y a metric space, and let a map
f : X → Y be locally Lipschitz with a Lipschitz constant C. Prove that f
is Lipschitz with the same constant.
Exercise 2.4.15. Let (X, d) be a length space and A a connected open
subset of X. Then d induces on A the (finite-valued) intrinsic metric dA .
Moreover each point p ∈ A has a neighborhood U ⊂ A such that for any
points p, q ∈ U we have d(p, q) = dA (p, q).
This theorem has the following immediate corollary, which can be used
as an alternative criterion for a metric to be intrinsic:
Corollary 2.4.17. A complete metric space (X, d) is a length space iff,
given a positive ε and two points x, y ∈ X, there exists a finite sequence
of points x1 = x1 , x2 , . . . , xk = y such that every two neighboring points in
this sequence are ε-close (i.e., d(xi , xi+1 ) ≤ ε for all i = 1, . . . , k − 1) and
Pk−1
i=1 d(xi , xi+1 ) < d(x, y) + ε.
2.4. Characterization of Intrinsic Metrics 43
This corollary says that a metric is intrinsic if and only if, given two
points and a positive ε, one can reach one of the points starting from the
other one and hopping with jumps shorter than ε and with the total length
of the jumps not exceeding the distance between the points plus ε.
This inequality implies that the map γ, defined on the set of dyadic
rationals, is Lipschitz. Since X is complete and the set of dyadic rationals
is dense in [0, 1], this map can be extended to the entire interval [0, 1] (cf.
Proposition 1.5.9). Thus we obtained a path γ : [0, 1] → X connecting x
and y. Then (2.1) implies that L(γ) = d(x, y). ¤
Exercise 2.4.19. Let X be a compact topological space and let {dn }∞ n=1
be a sequence of intrinsic metrics on X that uniformly converge to a metric
d (recall that metrics are functions on X × X, so the notion of uniform
convergence applies here). Prove that d is intrinsic too.
44 2. Length Spaces
The term “curve” is used for both unparameterized curves and their
parameterizations (i.e., paths). In most cases, “curve” is formally a synonym
for “path”. However, the former is more appropriate when parameterization-
independent properties are considered.
This means that a simple path is allowed to stop at a point for a while,
but having left a point it never comes back. Roughly speaking, the image
of a simple path is a curve without ”self-intersections”. The following very
easy exercise justifies the correctness of the definition:
Exercise 2.5.6. If two simple curves have the same image, then they are
equivalent up to a change of variable t 7→ −t.
Our next goal is to choose our favorite parameterization for every curve.
These parameterizations are analogous to motion with unit speed in physics
or differential geometry:
Proof. The idea of construction of ϕ is trivial: let γ̄(τ ) be the point on (the
image of) γ such that the length of the interval of γ between its origin and
that point is equal to τ . This is formalized as follows. Define ϕ(t) = L(γ, a, t)
for all t ∈ [a, b]. Then the function ϕ is nondecreasing and continuous (by the
continuity property of length). The set of its values is the interval [0, L(γ)].
Now for every τ ∈ [0, L(γ)] pick a t ∈ [a, b] such that ϕ(t) = τ and define
γ̄(τ ) = γ(t). This definition does not depend on the choice of t. Indeed, if
ϕ(t) = ϕ(t′ ), then γ(t) = γ(t′ ) because L(γ, t, t′ ) = ϕ(t′ ) − ϕ(t) = 0.
Thus we have defined a map γ̄ : [0, L(γ)] → X. The relation γ = γ̄ ◦ ϕ
follows immediately from the definition. It remains to verify that γ̄ is
continuous and parameterized by arc length. For the former, let τ1 = ϕ(t1 )
and τ2 = ϕ(t2 ). Then γ̄(τ1 ) and γ̄(τ2 ) are the endpoints of the path γ|[t1 ,t2 ] .
The length of this path is L(γ, t1 , t2 ) = ϕ(t2 ) − ϕ(t1 ) = τ2 − τ1 . Since
the distance between endpoints is no greater than the length, we obtain
that d(γ̄(τ1 ), γ̄(τ2 )) ≤ |τ1 − τ2 |. This means that γ̄ is a nonexpanding
and hence continuous map. Furthermore γ|[t1 ,t2 ] is a re-parameterization
of γ̄|[τ1 ,τ2 ] and hence L(γ̄, τ1 , τ2 ) = L(γ, t1 , t2 ) = τ2 − τ1 . Thus γ̄ is a natural
parameterization. ¤
Exercise 2.5.10. Prove that a natural parameterization of a curve is unique
up to a translation [a, b] → [a + c, b + c] : t 7→ t + c of the variable.
Exercise 2.5.11. Given v > 0, any rectifiable curve admits a parameteri-
zation with constant speed v.
To prove that the sequence {γi (t)} converges for every t ∈ [0, 1], we will
show that this is a Cauchy sequence.
Given ε > 0, choose tj ∈ S such that |t − tj | < ε/C and then N ∈ N
such that d(γi (tj ), γk (tj )) < ε for all i, k > N . For these choices
d(γi (t), γk (t)) ≤ d(γi (t), γi (tj )) + d(γi (tj ), γk (tj )) + d(γk (tj ), γk (t)) ≤ 3ε.
This proves that γi (t) is a Cauchy sequence and hence we can define
γ(t) = limj→∞ γj (ti ).
Passing to the limit in (2.2) we get
(2.3) d(γ(t), γ(t′ )) ≤ C|t − t′ |,
and thus γ is a continuous map.
Let us show that γi converges to γ uniformly. Given ε > 0, choose N >
4C
ε and let M be such that d(γ(k/N ), γi (k/N )) < ε/2 for all k = 0, 1, . . . , N
and all i > M . This choice is possible since γi converges to γ point-wise.
Combining (2.2) with (2.3), for every 0 ≤ t ≤ 1 and k/N ≤ t ≤ (k + 1)/N
we have
d(γ(t), γi (t)) ≤ C|t − k/N | + ε/2 + C|t − k/N | ≤ ε
for all i > M . This concludes the proof. ¤
Shortest paths in length spaces possess some nice properties that do not
hold in general metric spaces. One such property is the following
Proposition 2.5.17. If shortest paths γi in a length space (X, d) converge
to a path γ as i → ∞, then γ is also a shortest path.
L(γi ) → d(x, y), where x, y are the endpoints of γ. By the lower semi-
continuity of length,
L(γ) ≤ lim L(γi ) = d(x, y).
i→∞
¤
Exercise 2.5.18. Give an example showing that a limit of shortest paths
in a metric space may fail to be a shortest path.
Proof. Let Linf denote the infimum of lengths of rectifiable curves con-
necting x and y. Then there exists a sequence {γi } of such curves with
L(γi ) → Linf . According to Theorem 2.5.14, the sequence {γi } contains a
converging subsequence. Without loss of generality we may assume that
{γi } itself converges to a curve γ. Then γ has the same endpoints and,
by the lower semi-continuity of length, L(γ) ≤ lim L(γi ) = Linf . Thus
L(γ) = Linf . ¤
Corollary 2.5.20. Let (X, d) be a boundedly compact metric space. Then
for every two points x, y ∈ X connected by a rectifiable curve there exists a
shortest path between x and y.
Proof of the theorem. Implications (ii) =⇒ (i) =⇒ (iii) =⇒ (iv) are left
as easy exercises. We will prove that (iv) implies (ii). The proof uses the
same general scheme as the one of Proposition 2.5.22. The details are slightly
more complicated because we cannot utilize completeness at this point. We
can use only the property (iv) and this requires a more delicate argument.
Since X is locally compact, sufficiently small closed balls B r (p) are
compact. Reasoning by contradiction (that is, assuming that there are
noncompact closed balls), define
R = sup{r : B r (p) is a compact set }
and assume that R < ∞. The argument consists of two steps.
1. First, we prove that the open ball BR (p) is pre-compact. To do
this, it suffices to show that every sequence {xi } in this ball contains a
converging subsequence (whose limit does not necessary belong to this ball).
Set ri = d(p, xi ). One may assume that ri → R as i → ∞; otherwise a
subsequence of {xi } is contained in a ball B r (p) for some r < R, and then
there is a converging subsequence because this smaller ball is compact (by
the choice of R).
Let γi : [0, ri ] → X be a (naturally parameterized) shortest path connec-
tion p to xi . Such a shortest path exists because xi belongs to a compact
2.6. Length and Hausdorff Measure 53
ball centered at p (see the proof of Corollary 2.5.20). We can choose a sub-
sequence of {γi } such that the restrictions of the paths to [0, r1 ] converge
(along this subsequence). From this subsequence, we choose a further sub-
sequence of paths whose restrictions to [0, r2 ] converge, and so on. Then
the Cantor diagonal procedure (that is, picking the nth element from the
nth subsequence for n = 1, 2, . . . ) yields a sequence {γin } such that for
every t ∈ [0, R) the sequence {γin (t)} converges in X. (More precisely,
points γin (t) are well-defined for all large enough n and form a converging
sequence.)
Define γ(t) = lim γin (t); then γ : [0, R) → X is a nonexpanding map and
a shortest path (see Proposition 2.5.17). By the assertion (iv), there is a
continuous extension γ : [0, R] → X. One easily sees (exercise!) that the
points xin (i.e., the endpoints of our converging curves {γin }) converge to
γ(R).
2. Since the open ball BR (p) is pre-compact, the closed ball B R (p) is
compact. (Recall that a closed ball in a length space is the closure of the
respective open ball.) Now we show that a ball BR+ε (p) is pre-compact
for some ε > 0. Since X is locally compact, for every x ∈ B R (p) there
is an r(x) > 0 such that the ball Br(x) (x) is pre-compact. Choose a finite
subcover B r(xi ) (xi ) out of the cover of B R (x) by these balls. The union of
these balls is pre-compact and contains the ball BR+ε (p) for ε = min ri > 0.
This contradicts the choice of R. ¤
for a small δ > 0. Then diam Si′ ≤ diam Si + 2δ/2i and hence w1 ({Si′ }) ≤
w1 ({Si }) + 2δ. Since δ is arbitrary, it follows that the measure can be ap-
proximated by 1-weights of open coverings. (Exercise: extend the argument
to work for measures of all dimensions.)
2. Let X be a connected topological space and {Si } an open covering
of X. Then for every two points x, y ∈ X there exists a finite sequence
Si1 , . . . , Sin of sets such that x ∈ Si1 , y ∈ Sin , and Sik ∩ Sik+1 6= ∅ for all k,
1 ≤ k ≤ n − 1. To prove this, fix a point x ∈ X and consider the set Y of
all points y ∈ X for which such a sequence exists. It is clear that for every
set Si one has either Si ⊂ Y or Si ⊂ X \ Y . Therefore both Y and X \ Y
are open sets. Since X is connected, it follows that Y = X and therefore
any point y ∈ Y is “accessible” by a sequence of sets as described above.
3. Let {Si } be an open covering of X, x and y two points of X, and
Si1 , . . . , Sin a sequence from Step 2. For k = 1, . . . , n − 1 pick a point
xk ∈ Sik ∩ Sik+1 , and define x0 = x, xn = y. Then |xk−1 xk | ≤ diam Sik for
all k = 1, . . . , n because both xk−1 and xk are in Sik . Hence
X n
X n
X
diam Si ≥ diam Sik ≥ |xk−1 xk | ≥ |x0 xn | = |xy|.
k=1 k=1
Due to the observation made in step 1, it follows that µ1 (X) ≥ |xy|. Since
x and y are arbitrary points, this means that µ1 (X) ≥ diam X. ¤
Theorem 2.6.2. Let X be a metric space, γ : [a, b] → X a simple curve.
Then L(γ) = µ1 (γ([a, b])).
Proof. Let S = γ([a, b]) and L = L(γ). Without loss of generality assume
that γ is parameterized by arc length: γ : [0, L] → X. Then for every
natural number N , S is covered by N intervals of γ each of length L/N ,
L L
namely by the sets γ([i N , (i + 1) N ]), i = 0, 1, . . . , N − 1. The diameters of
these sets are no greater than L/N . Hence the sum of the diameters is no
greater than L, whereas the diameters themselves approach 0 as N → ∞.
This shows that µ1 (S) ≤ L(γ).
On the other hand, for a partition a = t0 ≤ t1 ≤ · · · ≤ tn = b of [a, b],
let Si = γ([ti , ti+1 ]) for i = 1, . . . , n − 1. The sets Si are disjoint modulo
a finite number of points γ(ti ). Since the one-dimensional
P measure of a
single point is obviously zero, one has µ1 (S) = µ1 (Si ). By Lemma 2.6.1,
µ1 (Si ) ≥ diam Si ≥ |γ(ti )γ(ti+1 )|, and hence
X X
µ1 (S) = µ1 (Si ) ≥ |γ(ti )γ(ti+1 )|
for any partition {ti }. This implies that µ1 (S) ≥ L(γ). ¤
Remark 2.6.3. If γ is not simple, the same argument shows that L(γ) ≥
µ1 (γ([a, b])).
2.7. Length and Lipschitz Speed 55
Of course, the speed of a curve may not exist. However, it exists almost
everywhere (i.e., except a set of zero measure) for a wide class of curves,
namely for every Lipschitz curve. (Note that every rectifiable curve admits
Lipschitz parameterization. For example, all natural parameterizations are
Lipschitz with unit Lipschitz constant.) Moreover, the length of a Lipschitz
curve equals the (Lebesgue) integral of its speed.
As a first step, we prove the following
Theorem 2.7.4. Let (X, d) be a metric space, γ : [a, b] → X a rectifiable
curve. Then for almost all t ∈ [a, b] (i.e., for all t except a set of zero
56 2. Length Spaces
One can let ε or ε′ in the above formulas be zero. This gives the following
Corollary 2.7.5. If γ is as in Theorem 2.7.4, then for almost all t ∈ [a, b]
either
L(γ|[t,t+ε] )
lim inf = 0,
ε→0 |ε|
or
d(γ(t), γ(t + ε))
lim =1
ε→0 L(γ|[t,t+ε] )
(if ε < 0, the interval [t, t + ε] in the denominator of the last formula should
be interpreted as [t + ε, t]).
Proof of Theorem 2.7.4. Suppose the contrary. For every α > 0 let Zα
denote the set of all t ∈ [a, b] such that
L(γ|[t−ε,t+ε′ ] )
lim inf >α
′
ε,ε →0+ ε + ε′
and
d(γ(t − ε), γ(t + ε′ ))
lim inf < 1 − α.
′
ε,ε →0+ L(γ|[t−ε,t+ε′ ] )
Then Sµ1 (Zα ) > S 0 for all sufficiently small α. Indeed, otherwise the set
Z0 = α>0 Zα = n∈N Z1/n would have zero measure, and this is equivalent
to the statement of the theorem. Fix an α > 0 such that µ1 (Zα ) > 0. For
brevity, we denote Z = Zα and µ = µ1 (Z). Choose ε0 so small that for
any partition {yi }Ni=1 (a = y0 ≤ y1 ≤ · · · ≤ yN = b) of [a, b] such that
maxi (yi − yi−1 ) < ε0 , one has
N
X
L(γ) − d(γ(yi−1 ), γ(yi )) < µα2 /2.
i=1
Such an ε0 exists by Exercise 2.3.2. Consider the set B of all intervals of
the form [t − ε, t + ε′ ] such that t ∈ Z, ε + ε′ < ε0 , L(γ|[t−ε,t+ε′ ] ) > α(ε + ε′ )
and
d(γ(t − ε), γ(t + ε′ )) < (1 − α)L(γ|[t−ε,t+ε′ ] ).
By the definition of Z = Zα , every point t ∈ Z is contained in arbitrarily
short elements of B. Applying Vitali’s covering theorem (Theorem 1.7.14)
2.7. Length and Lipschitz Speed 57
we can extract from B a countable collection {[ti − εi , ti + ε′i ]}∞ i=1 of disjoint
intervals that covers Z up to a set of zero measure. In particular,
X∞
¡[ ¢
(εi + ε′i ) = µ1 [ti − εi , ti + ε′i ] ≥ µ1 (Z) = µ.
i=1
Hence for a sufficiently large M ,
M
X
(εi + ε′i ) > µ/2.
i=1
because d(γ(t), γ(t + ε)) ≤ L(γ|[t,t+ε] ). In the second case, it follows that
d(γ(t), γ(t + ε))
vγ (t) = lim = f ′ (t).
ε→0+ |ε|
Thus vγ (t) exists and equals f ′ (t) in both cases. The theorem follows. ¤
Exercise 2.7.7. Give an example of a nonconstant curve γ in R2 for which
vγ = 0 almost everywhere.
Exercise 2.7.8. Let X be a metric space and γ : I → X a curve. For a
t ∈ [a, b] define
dilt (γ) = lim sup dil(γ|[t−ε,t+ε] ).
ε→0+
Constructions
59
60 3. Constructions
3.1.2. Gluing. Imagine that we glue together two length spaces, or even
some points in the same length space, as we do to glue a Möbius strip or a
ring out of a paper rectangle. How do we measure distances in this space?
For two points “on different sides of the gluing line”, the first thing that
occurs to one’s mind would be to look for a shortest path that consists of
two pieces: first it goes in one space to the gluing edge and then from this
edge it continues in the other space. In other words, one thinks of a path
with a “gap”, but the gap is eliminated by gluing. As the matter of fact,
3.1. Locality, Gluing and Maximal Metrics 61
this is not at all so simple. It may very well happen that a shortest path
has to cross the gluing edge several or even infinitely many times.
Before passing to rigorous definitions, let us consider several examples.
Example 3.1.6. Begin with a strip R × [0, 1] ⊂ R2 and glue together its
edges by identifying (x, 1) with (x+100, 0) for all x. We obtain a topological
cylinder.
Question. Is the distance between (0, 1/2) and (1000, 1/2) (after the
identifications are made) greater than 899?
Answer. No, it is even less than 11! Indeed, consider the path whose
itinerary is
|(0, 1/2) → (0, 1)| = 1/2, |(0, 1) → (100, 0)| = 0 (“free”),
|(100, 0) → (100, 1)| = 1, |(100, 1) → (200, 0)| = 0 (“free”),
...,
|(900, 0) → (900, 1)| = 1, |(900, 1) → (1000, 0)| = 0 (“free”),
|(1000, 0) → (1000, 1/2)| = 1/2.
The length of this path is 10.
Exercise 3.1.7. Is this a shortest path?
Exercise 3.1.8. Consider the region E = {(x, y) : 0 ≤ y ≤ e−x , x ≥ 0}.
(This is the region enclosed between the graph of e−x , x ≥ 0, and the x-axis.)
We glue together the infinite edges of this region by identifying (x, 0) with
(x + 1, ex+1 ) for all x ≥ 0. Show that the diameter of the resulting space is
finite!
Hint: Consider the path consisting of the segments {[(n, 0), (n, e−n )],
n = 1, 2, . . .}. Observe that this is indeed a continuous path and its length
is equal to
X∞ Z ∞
−m
e < e−x dx = 1.
m=1 0
Certainly, both the arguments and the questions themselves are informal:
we have not given a definition of gluing yet. To show that this notion is not
so simple and indeed needs a rigorous definition, look at the following more
striking example:
Example 3.1.10. Begin with R2 and glue each point (x, y) with (−y, 2x).
Question 1. If we forget about metric and do this identification topolog-
ically, what is the quotient space?
62 3. Constructions
Question 2. How can we find the distance between two points after this
identification?
Answer. Very simple: the distance is identically zero.
Exercise 3.1.11. Prove that, given a positive ε and two arbitrary points in
the space from the above example, there is a path between the points whose
length is less than ε.
Hint: Our gluing rule can be re-written as
³y ´ µ ¶ µ ¶ µ ¶
−x −y −y x x −y
(x, y) → , −x → , → , → , → ....
2 2 2 4 2 4 4
These examples suggest the following strategy of measuring the metric
resulting from “gluing some points together” in terms of the intact space
(before identifications are made). We consider finite sequences of paths such
that the terminating point of each of them will be glued with the starting
point of the next one. To measure the distance between two points we
require that the first path starts at one of the points and the last path ends
at the other. The infimum of lengths of such paths gives us the new distance.
In other words, to measure the distance between x and y, we connect x and
y by a finite sequence of points (a “dotted line”). If we added the distances
between neighboring points in this sequence and took the infimum of such
sums, this would be just the original distance between x and y. What we
do instead is we add only the distances between neighboring points that are
nonequivalent. Intuitively, we may jump from a point to an equivalent point
for free.
Definition 3.1.12. Let (X, d) be a metric space and let R be an equivalence
relation on X. The quotient semi-metric dR is defined as
k
X
dR (x, y) = inf{ d(pi , qi ) : p0 = x, qk = y, k ∈ N}
i=1
where the infimum is taken over all choices of {pi } and {qi } such that qi is
R-equivalent to pi+1 for all i = 1, . . . , k − 1. As usual, we associate with a
semi-metric space (X, dR ) a metric space (X/dR , dR ) by identifying points
with zero dR -distances (and keep the same notation dR for this metric; see
Section 1.1). This space is called the quotient metric space. One also says
that it results from gluing the space (X, d) along the relation R.
Exercise 3.1.13. Prove that dR is indeed a semi-metric, i.e., it is nonneg-
ative, symmetric and satisfies the triangle inequality.
Now let us begin with the simplest example of two length spaces (X, dX )
and (Y, dY ) and a bijection I : X ′ → Y ′ between subsets X ′ ⊂ X and
64 3. Constructions
Exercise 3.1.20. Prove that the (equivalence class of) a vertex of Q in the
quotient metric has a neighborhood isometric to a Euclidean region.
This exercise explains why this quotient space is called a flat torus.
Example 3.1.21. Let G be a group of isometries (rigid motions) of R2 .
Introduce an equivalence relation R defined as follows: xRy iff there exists
a g ∈ G such that x = gy (i.e., x and y belong to the same orbit).
Exercise 3.1.22. Prove that the quotient metric between (the equivalence
classes of) two points x and y is equal to inf g∈G |x − g(y)|, where | · | stands
for Euclidean distance.
Proof. One may assume that the metrics dα are defined on the whole space
X by setting dα (x, y) = ∞ if x ∈ / Xα or y ∈ / Xα . To prove the existence of
dm , apply Lemma 3.1.23 to the function b(x, y) = inf α dα (x, y). To prove
that dm is intrinsic if dα are, consider the intrinsic metric dbm induced by dm .
Since all dα are intrinsic and dm ≤ dα , it follows that dbm ≤ dα , so dbm belongs
to D. Then dbm = dm because dm is maximal. ¤
Note that the function bR in the theorem is the minimum of two length
semi-metrics: the original metric d and the one which equals 0 for R-
equivalent points and ∞ otherwise. Thus we are in conditions of Corollary
3.1.24, which gives us another proof for the fact that dR is intrinsic.
3.2.2. Metric graphs. We presume that the reader is familiar with topo-
logical graphs at least as a nice way to draw a few cities connected by several
roads. Metric graphs are just one-dimensional polyhedral spaces. Regard-
less the fact that the definition of a polyhedral space is given in the previous
section, we repeat it here to better visualize its meaning in this simplest
case. This definition generalizes the example of a “cobweb” from the first
chapter.
By a metric segment (of length a) we mean a metric space isometric to
a segment [0, a].
Definition 3.2.9. A (metrized) graph is the result of gluing of a disjoint
collection of metric segments {Ei } and points {vj } (regarded with the length
metric of disjoint union) along an equivalence relation R defined on the union
of the set {vj } and the set of the endpoints of the segments.
the length of the segment (it may be different from the distance between
its endpoints in the graph). The cardinality of the endpoints of segments in
an equivalence class representing a vertex is called its degree. Most of the
graphs considered in this book will be locally finite and therefore with finite
degree of every vertex (one exception is a monster-graph in an example in
this section).
The most natural way to define a graph is to take a set V of vertices,
indicate which pairs of vertices are connected by edges, and specify lengths
of these edges. To recover the original definition from these data, take a
collection of segments {Ei } corresponding to the desired edges and having
the given lengths. Then, if the edge corresponding to Ei should connect
vertices a and b, let one endpoint of Ei be equivalent to a and another be
equivalent to b. This generates an equivalence relation R used to glue the
graph from V and {Ei }.
In particular, every topological graph can be turned to a metric graph
by assigning lengths to its edges. However, one should remember that (for
infinite graphs) this may change the topology of the space. See also Exercise
3.2.12 below.
Note that a pair of vertices may be connected by two or more different
edges. Definition 3.2.9 even allows a graph to have loops, that is, edges
connecting a point to itself. One can always obtain a graph without
loops and multiple edges by dividing edges into smaller ones (compare with
Exercise 3.2.5).
To get better understanding of the definitions, we suggest the following
Exercise 3.2.10. Unwind the definition of gluing and verify that for finite
graphs our definition agrees with the usual one: the distance from x to y is
the shortest way of reaching y from x by means of traversing finitely many
segments such that the terminating point of each segment is identified with
the beginning of the next one. Is this true for infinite graphs? What if we
replace “the shortest” by “the infimum”?
On the other hand, the identification X → X/dR can never affect the
interiors of segments:
Exercise 3.2.13. Prove that points in the interiors of edges do not get
identified and thus the disjoint union of these interiors is embedded into the
graph.
In other words, every edge is a simple curve connecting two vertices and
not passing through other vertices, and the interiors of different edges have
no common points.
Exercise 3.2.14. Prove that the length of every edge as a curve in the
graph equals the length of the segment from which the edge was obtained.
Exercise 3.2.15. Let X be a length space and V ⊂ X a finite set.
Suppose that X \ V is covered by a finite collection of rectifiable curves,
each connecting two points of V , and these curves have no self-intersections
and do not intersect one another (except at endpoints). Prove that X is a
metric graph.
Exercise 3.2.16. Show that the statement of the previous exercise fails
without the assumption that the collection of curves is finite.
Hint: Every length space is a union of points, but not all length spaces
are 0-dimensional polyhedra.
Although our definitions sound quite scientific, the object is very familiar
and our everyday intuition is a good guide through simple cases of graphs.
For instance, to find intrinsic distances and shortest curves in the frame of
a cube is a good exercise for a 10-year old kid. Graphs may even seem dull
as being too discrete and combinatorial for a geometrically-thinking reader.
To relieve this feeling we suggest the following approach: one can try to
approximate an arbitrary length space by graphs whose vertices are dense
enough in the space; see Chapter 7 for more details. For the most skeptical
reader, the following example shows that every metric space can be realized
as the set of vertices of a graph, with the distance between points being
equal to that in the graph. Alas, this graph is apparently a hardly useful
monster.
Example 3.2.17. Begin with a general metric space (X, d); consider a
disjoint union of segments Ix,y parameterized by pairs x, y ∈ X. Equip each
segment Ix,y with the metric of the segment [0, d(x, y)]. These are edges.
The identifications are most natural: one identifies the right ends of two
segments Ix,y and Ix′ ,y′ if y = y and the left ends if x = x′ . We leave as an
exercise to verify that the set X can be canonically identified with the set of
vertices of the graph and that the intrinsic metric of this graph restricted to
3.2. Polyhedral Spaces 73
the set of its vertices is the original metric d in X. (To get better intuition
behind this construction, one can think of it this way: for each pair of points
x, y ∈ X, one takes a segment of length d(x, y) and attaches one of its ends
to x and the other one to y.)
space is nontrivial, this is often expressed by saying that the space possesses
symmetries. There is a remarkable class of very symmetrical spaces, namely
homogeneous spaces.
Definition 3.3.2. A length space X is said to be homogeneous if for every
x, y ∈ X there exists an isometry I : X → X such that I(x) = y.
Example 3.3.3. Euclidean spaces and (round) spheres are homogeneous
spaces. The cylinder x2 + y 2 = 1 in R3 and the torus from Example 3.1.19
are also homogeneous. On the other hand, the cone x2 + y 2 = z 2 and the
paraboloid x2 + y 2 = z are not homogeneous spaces.
Can you find the isometry groups for these spaces?
There is an important notion, which is relevant to introduce here.
Definition 3.3.4. A metric spaces X is said to be locally isometric to a
homogeneous space Y if each point in X possesses a neighborhood isometric
to an open set in Y . Spaces locally isometric to a Euclidean space are said
to be flat.
Flat spaces cannot be distinguished from Euclidean space by local mea-
surements: they look the same. For instance, the cylinder x2 + y 2 = 1 in R3
is flat. To realize that this is a cylinder, a two-dimensional creature would
have to undertake a long “Magellan” trip around the cylinder. The cone
x2 + y 2 = z 2 , z ≥ 0, is not flat, as it looks very different at its apex than the
Euclidean plane.
Question: What would you suggest for a two-dimensional creature living
in the cone to tell that this is not a two-plane? What about a two-
dimensional creature living in the sphere? Did we have to travel around the
Earth to figure out that it is not planar? Notice that the experiments with
a horizon are illegal in this set-up: they use measurements in the ambient
space, and we are allowed to use intrinsic measurements only. The answer
to these questions is not that easy, but the reader should be able to give
them after reading this textbook.
d(x, y) = inf{d(x, y) : x ∈ x, y ∈ y}
k
X
d(pi , qi )
i=0
p̃0 = p0 = p, q̃0 = q0 ,
p̃1 = g1 (p1 ), q̃1 = g1 (q1 ),
p̃2 = g1 (g2 (p2 )), q̃2 = g1 (g2 (q2 )),
...
p̃k = g1 ◦ g2 ◦ · · · ◦ gk (pk ), q̃k = g1 ◦ g2 ◦ · · · ◦ gk (qk ).
3.3. Isometries and Quotients 77
On the other hand, by the choice of the isometries gi , q̃i = p̃i+1 . Thus
k
X k
X
d(p, q̃k ) ≤ (d(p̃i , q̃i ) + d(q̃i , p̃i+1 )) = d(pi , qi ).
i=0 i=0
Recalling that
q̃k = g1 ◦ g2 ◦ · · · ◦ gk (qk ) = g1 ◦ g2 ◦ · · · ◦ gk (q)
and thus q̃k belongs to the orbit O(q) of q, we conclude that the distance
between the orbits of p and q is less than or equal to the distance between the
equivalence classes of p and q in the quotient metric. The opposite inequality
is obvious. Indeed, if q ′ ∈ O(q) and p′ ∈ O(p), then p′ and q ′ are equivalent
to p and q respectively, and hence one chooses the path p0 = q0 = p, p1 = p′ ,
q1 = q ′ and p2 = q2 = q with only one nonzero jump, and the length of this
path is equal to d(p′ , q ′ ).
¤
Exercise 3.3.7. Give a direct proof (that is, without referring to Definition
3.1.12 of gluing and the notion of maximal metric) that the metric d defined
this way is an intrinsic metric.
Hint: Here is a proof for the case when X is locally compact and
complete and all orbits are compact. To get a general proof, one has to
use approximations.
Given a, b ∈ X/G, choose their representatives x ∈ a, y ∈ b (as a and
b are equivalence classes) such that d(a, b) = d(x, y). This is possible since
we assumed that all orbits are compact. Let z be a mid-point for x, y, i.e.
d(x, z) = d(y, z) = 21 d(x, y). Obviously d(a, O(z)) ≤ d(x, z) = 12 d(x, y) =
1 1
2 d(a, b). Similarly d(b, O(z)) ≤ 2 d(a, b). The triangle inequality shows that
both inequalities turn out to be equalities, so O(z) is a mid-point for a, b. ¤
Example 3.3.8. For the subgroup G of isometries of R2 acting by integer
translations (g(x, y) = (x + k, y + l), k, l are integers), the quotient space
R2 /G is a torus (isometric to the flat torus from Example 3.1.19).
Example 3.3.9. For the subgroup G of isometries of R2 consisting of
two elements, the identity and the symmetry −id : (x, y) → (−x, −y), the
quotient space is isometric to a cone x2 + y 2 = cz 2 , z ≥ 0. Can you find this
coefficient c? What about G being a finite group of rotations?
78 3. Constructions
These conditions on Y in the theorem look natural and hold for most
topological spaces except “pathological” ones. From now on, we consider
only topological spaces satisfying these conditions.
One of the most important features of universal coverings is their relation
to fundamental groups. Let f : X → Y be a covering map. Fix y0 ∈ Y and
choose an x0 ∈ f −1 (y0 ). Then every path γ in Y starting at x0 has a unique
lift γ̃ in X starting at x0 (recall that γ̃ being a lift of γ means that f ◦ γ̃ = γ).
In particular, a loop with endpoints at y0 is lifted to a curve starting at x0
and ending at some point of f −1 (y0 ). The standard “covering homotopy”
lemma tells that lifts of homotopic paths are homotopic (by a homotopy
of paths we mean a homotopy with fixed endpoints). In particular, lifts
of homotopic paths have the same endpoints; i.e., if two loops γ1 and γ2
with endpoints at y0 are homotopic, then their lifts γ̃1 and γ̃2 starting at x0
end at the same point y ∈ f −1 (y0 ). If f is a universal covering, then the
converse statement is also true: every two paths in X connecting a given
pair of points are homotopic and hence they are lifts of homotopic paths
in Y .
Thus we have a 1-1 correspondence between classes of homotopic loops
in Y with endpoints at y0 (that is, the fundamental group π1 (Y, y0 )) and
the elements of f −1 (y0 ) ⊂ X.
Now let Y be a length space and X be equipped with the length metric
lifted from Y . Then every deck transformation is an isometry from X to
itself (prove this!). In the case of a universal covering, this yields an action
of π1 (Y ) on X by isometries.
This fact (that the fundamental group of a length space acts on its
universal covering space by isometries) is very important. Some of its
numerous applications can be found in this book.
Proof. The proof is trivial but it is a good test for understanding the
definitions.
Let f : X → X/G be the projection. If x ∈ X and U is a neighborhood
of x from the definition of totally discontinuous action, then V = f (U ) is an
evenly covered neighborhood of f (x). Indeed, f −1 (V ) is the disjoint union
of the sets gU over g ∈ G, and f |gU is a homeomorphism from gU to V .
Obviously every element of G (considered as a homeomorphism from X
to itself) is a deck transformation. If y ∈ Y and x, x′ ∈ f −1 (y), then there is
an element g ∈ G such that gx = x′ . This means that G acts transitively on
sheets; hence f is a regular covering. On the other hand, there is at most one
deck transformation which maps x to x′ ; therefore every deck transformation
is represented by an element of G. ¤
Proof. The proof is similar to the construction for lifts of paths in the
theory of covering spaces. Assume that γ is parameterized by arc-length.
Consider the subintervals [0, t) of [0, a] such that γ|[0,t) can be lifted to X.
This means that there is a path γ̃t : [0, t) → Y such that γ|[0,t) = f ◦ γ̃
together with γ̃(0) = x0 . The set of such subintervals is not empty since the
restriction of f to some neighborhood of x0 is a homeomorphism onto its
image. Let [0, t0 ) be the union of all such subintervals, i.e., the maximum
interval that admits a lifting. We have a path γ̃t0 : [0, t0 ) → Y such that
γ|[0,t) = f ◦ γ̃t0 . Choose a sequence {ti } such that ti < t0 and ti → t0 as
i → ∞. Then γ̃t0 (ti ) is a Cauchy sequence and, since X is complete, it tends
to a point p. Clearly the choice of p is independent of the choice of {ti }.
Since the points f ◦ γ̃t0 (ti ) = γ(ti ) converge to γ(t0 ) as i → ∞, one can set
p to be the lift of γ(t0 ) and thus γ is lifted on the closed interval [0, t0 ]. If t0
is not equal to a, then γ could be lifted on a larger interval [0, t0 + ε) since
f is a local homeomorphism on a neighborhood of p. Since this contradicts
the maximality of [0, t0 ), we conclude that t0 = a. ¤
3. Give a definition of local arcwise isometry and prove that local arcwise
isometries are just arcwise isometries (since length structures are local).
model of a polyhedron using scissors, thick paper and glue, and then flatten
this model by stepping on it.
It is much more surprising that every surface homeomorphic to a sphere
admits an arcwise isometry onto a round sphere. This map is not at all
as nice as the one that can be used for polyhedra: for a generic surface, it
has everywhere dense folds and singular points. Although the proof of the
existence of such arcwise isometries for surfaces is beyond the frame of this
course, we sketch two absolutely elementary arguments proving that every
2-dimensional polyhedral space can be folded onto a polygon.
Proposition 3.5.5. Every (locally-)finite 2-dimensional polyhedral space
admits an arcwise isometry onto a planar polygon. In other words, every
2-dimensional polyhedral metric can be induced by a map to R2 .
Proof. We will carry out the argument for strictly intrinsic spaces; the
second part is left as an exercise. By Theorem 2.4.16, it is sufficient to
show that there is a midpoint between two given points z1 = (x1 , y1 ) and
z2 = (x2 , y2 ) in Z. For the midpoints xm and ym between x1 , x2 and y1 , y2 ,
resp., one sees that
The product metric d constructed this way enjoys the following property:
consider a fiber Sy = {(x, y) : y = const} together with the restriction of the
metric d. Then the projection map of Sy to X is an isometry. (In particular,
the restriction of d is an intrinsic metric.) This follows immediately from the
definition of the product metric. Evidently, the same holds for the “vertical”
fibers Sx .
In addition, the group of isometries of (Z, d) is at least as rich as the
product of the isometry groups of X and Y .
3.6. Products and Cones 89
Remark 3.6.3. One should not think that our formula for product metrics
is the only possible definition of a product metric on Z enjoying these nice
properties. For instance, by setting d((x, y), (x′ , y ′ )) = dX (x, x′ ) + dY (y, y ′ )
one gets an intrinsic metric that possesses all the properties listed above.
Prove this!
More generally, every norm k , k on R2 such that its restrictions to the
rays {x0 , y > 0} and {x > 0, y0 } are monotone give rise to a construction of
product metrics by assigning d(z, z ′ ) = kdX (x, x′ ), dY (y, y ′ )k. Moreover, if
one axiomatizes the desirable properties of product metrics, then all possible
constructions arise this way. The reason we used the Pythagorean theorem
and thus the standard Euclidean norm for k , k in our preferred definition
will become clear later (as a glimpse ahead, let us mention that this is the
only definition that respects the boundedness of curvature).
Proof. We may assume that both α and β are parameterized by arc length.
Introduce F : [a, b] × [c, d] → Z given by F (t, s) = (α(t), β(s)). This map is
an isometry:
d2 (F (t, s), F (t′ , s′ )) = d2X (g(t), g(t′ )) + d2Y (h(t), h(t′ )) = (t − t′ )2 + (s − s′ )2 .
Applying Lemma 3.6.9 finishes the proof. ¤
How should one equip a cone with a metric? To get an idea, imagine
that X is a subset of the unit sphere S 2 = {(x, y, z) : x2 + y 2 + z 2 = 1} ⊂ R3
equipped with the spherical (angular) metric. To build a cone over X, we
draw a ray from the origin through every point x ∈ X. Thus a point a in
the cone can be described by as (x, r), where x is a point in X that belongs
to the ray Oa and r = |aO| the distance from the origin; the latter is a
nonnegative number. Thinking of a and x as vectors, we can write a = rx.
How can we express the Euclidean distance between two points a = (x, t)
and b = (y, s) in terms of x, y, t, s? Consider the triangle △Oab. We have
|Oa| = t, |Ob| = s, and the angle ∡aOb between the sides equals the angular
distance d(x, y) in X. Hence by the cosine formula
p
|ab| = t2 + s2 − 2ts cos(d(x, y)).
We will use this formula to define cone metrics in general:
Proof. Positiveness and symmetry for dc are trivial. Let us prove the
triangle inequality. Consider three points y1 = (x1 , r1 ), y2 = (x2 , r2 ) and
y3 = (x3 , r3 ) in Con(X). Denote α = d(x1 , x2 ) and β = d(x2 , x3 ). Construct
three points y 1 , y 2 , y 3 ∈ R2 so that their distances from the origin O equal
r1 , r2 and r3 respectively, ∡y 1 Oy 2 = α, ∡y 2 Oy 3 = β, and the rays Oy 1 and
Oy 3 are in different half-planes with respect to Oy 2 . Then |y 1 y 2 | = dc (y1 , y2 )
and |y 2 y 3 | = dc (y2 , y3 ). Now we have two cases: α + β ≤ π and α + β > π.
If α+β ≤ π, then ∡y 1 Oy 3 = α+β ≥ d(x1 , x3 ); hence |y 1 y 3 | ≥ dc (y1 , y3 ).
Then the triangle inequality for y1 , y2 and y3 follows from the triangle
inequality in R2 :
dc (y1 , y2 ) + dc (y2 , y3 ) = |y 1 y 2 | + |y 2 y 3 | ≥ |y 1 y 3 | ≥ dc (y1 , y3 ).
If α + β > π, the planar triangle inequality does not help, but there is a
better estimate for |y 1 y 2 | + |y 2 y 3 |. Indeed, since the broken line y 1 y 2 y 3 lies
outside the sector y 1 Oy 3 , we have |y 1 y 2 | + |y 2 y 3 | ≥ |y 1 O| + |Oy 3 |. Then
dc (y1 , y2 ) + dc (y2 , y3 ) ≥ |y 1 O| + |Oy 3 | = r1 + r3 ≥ dc (y1 , y3 )
(the last inequality follows from the definition of dc ). ¤
92 3. Constructions
Cone over a large space. Now we drop the assumption that diam(X) ≤
π. The formula (3.1) is not suitable for defining a metric on Con(X) if
diam(X) > π; for example, the triangle inequality for dc may fail. How
does one define the cone metric in the general case? The guidelines are the
following: we want the formula (3.1) to hold for “small” distances in X, and
we want a cone over a length space be a length space.
If X is a length space, the existence and uniqueness of such a metric
is guaranteed by Lemma 3.1.2. In fact, this metric has a simple explicit
description:
Definition 3.6.16. Let (X, d) be a metric space. The cone distance dc (a, b)
between points a = (x, t) and b = (y, s) in Con(X) is defined as
(p
t2 + s2 − 2ts cos(d(x, y)), d(x, y) ≤ π,
dc (a, b) =
t + s, d(x, y) ≥ π.
Proof. First suppose that d is strictly intrinsic at distances less than π. Let
x, y ∈ X, a, b ∈ Con(X), a = (x, t), b = (y, s). If d(x, y) < π, apply Lemma
3.6.15 to a shortest path γ connecting x and y. It follows that there is a curve
of length dc (a, b) connecting a and b. If d(x, y) ≥ π, then dc (a, b) = t + s and
there is a curve of length t + s connecting a and b, namely, the union of the
two segments connecting a and b to the origin. Thus dc is strictly intrinsic.
Conversely, suppose that the metric dc is strictly intrinsic. For any two
points x, y ∈ X with d(x, y) < π apply the result of Exercise 3.6.14 to a
shortest path γ̃ connecting the points a = (x, 1) and b = (y, 1) in the cone.
Since L(γ) = dc (a, b) < 2, γ̃ does not pass through the origin and hence
has a well-defined (and continuous) projection γ in X. The inequality from
Exercise 3.6.14 implies that L(γ) = d(x, y).
The proof for non-strictly intrinsic metrics is similar and is left to the
reader. ¤
Exercise 3.6.19. Let X be a line segment of length α, 0 < α < 2π. Prove
that Con(X) is isometric to the planar sector of angular measure α with its
intrinsic metric.
This exercise is a partial case of the next proposition which says that
every polyhedral space locally looks like a cone. For simplicity we restrict
ourselves to 2-dimensional polyhedral spaces. In higher dimensions, one
should consider cones over spherical polyhedral spaces.
after points at zero distance from each other are identified. This new space
with the resulting metric will also be called a warped product.
Exercise 3.6.23. Let Con(X) be the cone over a length space X, and
diam X < π. Prove that Con(X) = [0, ∞) ×f X, where f (t) = t.
Exercise 3.6.24. Let Σ(X) be the spherical suspension over a length space
X. Show that Σ(X) = [0, π] ×f X, where f (t) = sin t.
In this notation,
∡(α, β) = lim θ(s, t).
s,t→0,s→0
If α and β are shortest paths parameterized by arc length, then d(p, α(s)) =
s, d(p, β(t)) = t, and θ(s, t) is determined by the distance d(α(s), β(t)):
s2 + t2 − d(α(s), β(t))2
θ(s, t) = arccos .
2st
Then the definition of angle (for unit-speed curves) can be rewritten as
follows: the angle ∡(α, β) exists and equals θ0 ∈ [0, π] if and only if
The notion of angle will be mainly used for shortest paths, but let us first
analyze what we get for different paths in usual Euclidean space. It turns
out that the existence of angle is closely connected with differentiability:
There are lots of examples of length spaces where the angle ∡(α, β) does
not exist even for two shortest curves α, β. Perhaps the simplest example
of this type can be made out of two copies of R by gluing them together at
the origin and at all points of the form 2−n with integer n.
Another example is given in the following exercise:
Proposition 3.6.30. 1. Every shortest path forms zero angle with itself.
2. If two shortest segments [a, b] and [b, c] are such that their concatena-
tion (“[abc]”) is also a shortest path, then the angle between [b, a] and [b, c]
is π.
Remark 3.6.31. In this course we will mainly deal with length spaces with
certain curvature bounds. In such spaces the angle between two shortest
paths is always well-defined. For general spaces one may consider upper
angles, which always exist. Namely,
Definition 3.6.32. The upper angle ∡U (α, β) is defined as
e
∡U (α, β) = lim sup ∡(α(s), p, β(t)).
s,t→0
Remark 3.6.33. Historically the notion of angle played a very important
role in the development of metric geometry. It is now well understood that
the usage of angles can be eliminated (without essential complications) from
most arguments dealing with singular (non-Riemannian) spaces. On the
other hand, involving angles often helps to visualize arguments and evokes
Riemannian analogies.
3.6.6. Space of directions. As we mentioned above, the notion of angle
does not play a very important role in modern metric geometry. Neverthe-
less, we will use it to define the cornerstone notion of the space of directions.
In a sense, this notion replaces the concept of tangent space in the theory
of smooth manifolds, and the advanced theory of Alexandrov spaces begins
with this notion.
Let us fix a point p and consider the space of all curves starting from
p. Our goal is to choose a subspace of “nicely behaved curves” and put an
angular semi-metric on this space.
To do this we need a version of the triangle inequality for angles.
Important remark: Please pay very close attention to the proof of
this theorem. This is the first example of an argument based on comparing
certain configurations of points in a length space with in some sense anal-
ogous configurations in the Euclidean plane. This type of arguments is as
basic and widely used in metric geometry as ε-δ arguments in analysis.
Theorem 3.6.34. Consider three curves γ1 , γ2 and γ3 starting at p. Assume
that the angles α1 = ∡(γ2 , γ3 ), α2 = ∡(γ1 , γ3 ) exist. If the angle α3 =
∡(γ1 , γ2 ) also exists, then it satisfies the following triangle inequality:
α3 ≤ α1 + α2 .
Remark 3.6.35. One of the assumptions of this theorem is the existence
of the angles α1 and α2 . In fact, the same inequality holds for upper angles,
which always exist.
3.6. Products and Cones 99
Now we consider only curves having direction and observe that the
property of having the same direction is an equivalence relation. A class
of equivalent curves is called a direction. One notices that this space of
directions is a metric space with respect to the upper angle.
Unfortunately, this seemingly very general notion has not proven useful
(as perhaps it is indeed too general). As a glimpse ahead, let us informally
explain what we will use instead of this space. We will mainly study such
spaces where the angle between shortest paths always exists. We will restrict
ourselves to such curves that are equivalent to a shortest curve. In other
words, we care only for directions arising from shortest paths. The space of
such directions will be a metric space with respect to angle (vs. upper-angle
in the general case). Alas, this is not yet the space of directions, as this
space may be noncomplete, and only its completion is called the space of
directions or directional space.
Chapter 4
Spaces
of Bounded Curvature
4.1. Definitions
4.1.1. Introduction. General length spaces can be extremely nasty, and
there are almost no nontrivial results which do not impose additional re-
strictions on spaces in question. Among such geometric restrictions, which
serve as sort of regularity assumptions, the main role is played by curvature
bounds. Loosely speaking, curvature bounds guarantee a certain degree of
convexity or concavity for distance functions. We begin with the two most
important curvature bounds: spaces of nonpositive curvature and spaces of
nonnegative curvature. These are two classes of spaces which enjoy very dif-
ferent and distinct features, and whose analysis however involves very similar
machinery. While other classes of curvature bounds are also very important,
most of their main features can be traced in the two main classes. Spaces
with a curvature bound (either above or below) will be called Alexandrov
spaces.
At the first glance, spaces of bounded curvature may have neither
topologic nor metric resemblance with Euclidean spaces. Nevertheless, their
definition is based on comparisons with the Euclidean world and (in certain
respects) they are less monstrous than, for instance, normed spaces. For
instance, the notion of angle makes perfect sense in spaces of bounded
curvature, while there are very serious reasons why this notion cannot be
generalized to normed spaces.
We give several equivalent definitions of nonpositively (resp. nonnega-
tively) curved spaces. These definitions formalize that:
101
102 4. Spaces of Bounded Curvature
• Distance functions for such spaces are not less convex (resp. not
less concave) than for the Euclidean plane;
• Geodesics emanating from one point diverge at least as fast as (resp.
not faster) than in the Euclidean plane;
• Triangles are not thicker (resp. not thinner) than Euclidean trian-
gles with the same side lengths.
Very vaguely, these spaces are not smaller (resp. not larger) than the
Euclidean space: for instance, they admit more nonexpanding (resp. ex-
panding) maps.
The sphere is an example of a positively curved space where the above
properties are seen easily. Spherical triangles look “fat” compared to their
Euclidean counterparts (for example, the angles in spherical triangles are
greater than those in Euclidean ones). Spherical geodesics tend to “attract”
one another rather than diverge linearly (more precisely, the distance be-
tween two geodesics emanating from one point in the sphere is a concave
function). These are typical features of spaces having positive curvature.
More generally, every convex surface in R3 (that is, a boundary of a
convex body) is a nonnegatively curved space, and every smooth saddle
surface (locally looking like a hyperbolic paraboloid) is a nonpositively
curved space. For a reader with a background in Riemannian geometry, we
mention in advance that a Riemannian manifold is a nonnegatively (resp.
nonpositively) curved length space if and only if its sectional curvatures are
nonnegative (resp. nonpositive).
We will use the name Alexandrov space for all spaces with a curvature
bound, and in particular for spaces of nonpositive and nonnegative curva-
ture.
Proof for Example 4.1.3. Denote by O the common point of the three
rays. Every shortest path in R(3) is either a segment in one of the rays, or a
concatenation of two segments in two different rays. Let γ : [0, T ] → R(3) be
a shortest path and p ∈ R(3) . If two of the points γ(0), γ(T ) and p belong
to the same ray, then the statement is trivial because γ and p are contained
in a union of two rays and this union is isometric to R. So we consider only
the case when the three points p, a = γ(0), and b = γ(T ) belong to different
rays. For every x ∈ [Oa] one has |px| = |pa| − |ax|. For the function g
from Definition 4.1.2, this means that g(t) = g(0) − t if γ(t) ∈ [Oa]. For
the function g0 , one has g0 (t) ≥ g0 (0) − t by the triangle inequality. Since
g0 (0) = g(0), the desired inequality g(t) ≤ g0 (t) follows for γ(t) ∈ [Oa]. The
case γ(t) ∈ [Ob] is similar. ¤
Proof for Example 4.1.4. The cone over a circle is flat outside the vertex;
every sub-cone over a segment of length α ≤ max{L/2, π} is convex and
isometric to a planar sector of angular measure α. (Recall the discussion
of cones in Section 3.6.2.) For a shortest path γ : [0, T ] → K and a point
p ∈ K, consider a triangle △ composed of three shortest paths between the
points p, a = γ(0) and b = γ(T ). There are two possibilities:
• The triangle △ bounds a region not containing O, or one of the
points a, b, p coincides with O.
• The triangle △ bounds a region containing O, or some of its sides
pass through O.
In the first case, the bounded region is isometric to (the region bounded
by) a triangle in the plane, and functions g and g0 from Definition 4.1.2
coincide.
We consider the second case separately for L < 2π and L > 2π. (The
case L = 2π is trivial because the cone is isometric to R2 .)
1. Suppose that L < 2π. Then one can represent the cone K as a
trihedral cone in R3 with edges passing through the vertices of △. (The
three-hedral cone is regarded with its induced length metric; “represent”
means that the two cones are isometric.) To do this, cut K into three
sectors by the rays Oa, Ob and Op, and place these sectors in R3 so as to
form a three-hedral angle (this is possible due to the triangle inequality for
angles).
Now the sides of △ are straight segments in R3 . These three segments
are contained in some plane P ⊂ R3 , and one can use the same triangle,
regarded as a subset of P , to define the function g0 from Definition 4.1.2.
In other words, g0 (t) equals the distance in R3 from p to γ(t). Since the
4.1. Definitions 105
b b
PSfrag replacements
a
O
b c
p c
Oa a
distances in the cone’s intrinsic metric are greater than or equal to those in
the ambient space R3 , it follows that g ≥ g0 . Therefore the cone is a space
of nonnegative curvature.
2. Suppose that L > 2π. The triangles △abO, △apO and △bpO are flat,
i.e., isometric to planar ones. Consider the triangles △abO and △bpO and
place their isometric copies △abO and △bpO in the plane at different sides
of the common side Ob. (If a shortest path [ab] or [bp] passes through O, its
isometric copy degenerates to a segment.) Observe that ∡aOb + ∡bOp > π
and ∡aOp ≤ ∡aOc; hence |ap| ≤ |ap|. Let us rotate the triangle △abO
around b until |ap| becomes equal to |ap| (see the left part of Figure 4.1).
This shows that isometric copies of △abO and △bpO lie without overlapping
in a planar triangle △abp whose sides are equal to those of △abp. This
arguments work for any pair of triangles △aOb, △bOp, △pOa; hence their
isometric copies lie in △abp without overlapping, as shown in the right part
of Figure 4.1.
Then it is clear that all distances between points in the sides of △abp
are less than or equal to distances between corresponding points of the
comparison triangle △abp, i.e., g ≤ g0 . Therefore the cone is a space of
nonpositive curvature. ¤
Remark 4.1.6. From the above proof one can see that the converse state-
ments are also true: if a cone over a circle is nonnegatively (resp. nonpos-
itively) curved, then the length of the circle is not greater (resp. not less)
than 2π.
The following exercise tells one about curvature of a cone over a segment
(note that a cone over a sufficiently short segment is just a planar sector).
We will use the result of this exercise later in this chapter.
Exercise 4.1.7. Let X be a cone over a segment of length L. Prove that
106 4. Spaces of Bounded Curvature
Proof for Example 4.1.5. First note that straight lines in a normed vec-
tor space are always shortest paths, because the length of a straight segment
equals the distance between its endpoints. Note that there may be other
shortest paths as well.
The unit sphere in X is the Euclidean square with vertices (1, 0), (0, 1),
(−1, 0), (0, −1). Let p = (0, 0). First consider the shortest path connecting
the points (1, 0) and (0, 1), parameterized by the interval [−1, 1]. For this
shortest path, g(t) ≡ 1 while g0 (t) = |t|, so g > g0 except endpoints, contrary
to the definition of nonpositive curvature. Then consider the shortest path
connecting points ( 21 , 12 ) and ( 21 , − 12 ) and parameterized by the interval [0, 1].
Here g(0) = g0 (0) = 1 and g(1) = g0 (1) = 1. However g( 21 ) = 12 while
√
g0 ( 12 ) = 23 > 21 . So in this case g( 21 ) < g0 ( 21 ), contrary to the definition of
nonnegative curvature.
The definitions of nonpositive and nonnegative curvature are local, but
one readily sees that the same construction can be repeated in an arbitrary
neighborhood of the origin since Euclidean homotheties are homotheties
w.r.t. the norm as well. ¤
4.1.4. Distance comparison for triangles. Probably the reader has not
yet gained much insight into the geometry of nonpositively (nonnegatively)
curved spaces from our analysis of distance functions. Let us re-formulate
our definition in more geometric terms. Traditionally the assertion of these
definitions is abbreviated as the CAT(0)-condition, and spaces fitting these
definitions are called CAT(0)-spaces. Here CAT stands for the comparison
of Cartan–Alexandrov–Toponogov, and (0) means that we impose zero as
the curvature bound, reflecting the fact that we compare our space with
the Euclidean plane. Comparing with other spaces (such as spheres) one
defines other CAT(k)-spaces, k ∈ R. This abbreviation is usually used for
upper curvature bounds only, whereas in case of lower curvature bounds one
just speaks of Alexandrov spaces of curvature bounded below by k. Note
that the term “Alexandrov space” can often be confusing since it does not
indicate whether k is the upper bound or the lower bound, and this must
be specified.
By a triangle in X we mean a collection of three points, a, b and c
(vertices) connected by three shortest paths (sides). For brevity we denote
these shortest segments by [ab], [bc], [ca] and their lengths by |ab|, |bc|, |ca|,
respectively. Recall that the vertices alone may not define a triangle uniquely
since there may be several different shortest paths between the same pair of
4.1. Definitions 107
vertices. By ∡abc we denote the angle between the shortest paths [ba] and
[bc] at b (if this angle is well-defined).
For each triangle △abc in X, we construct a triangle △abc in the
Euclidean plane with the same lengths of sides, i.e.
|ab| = |ab|, |bc| = |bc|, |ac| = |ac|.
Definition 4.1.8. Such a triangle △abc is called a comparison triangle for
the triangle △abc.
C
K<0
K=0
K > 0A C A C A C
every triangle △abc contained in this neighborhood, the angles ∡bac, ∡cba
and ∡abc are well defined and satisfy the inequalities
e
∡bac ≤ ∡bac e
∡abc ≤ ∡abc e
∡bca ≤ ∡bca
e
(recall that ∡abc e
denotes the comparison angle, i.e., ∡abc = ∡abc where
△abc is a comparison triangle, cf. Definition 3.6.25).
A length space X is a space of nonnegative curvature if every point of X
has a neighborhood such that, for every triangle △abc contained in this
neighborhood, the angles ∡bac, ∡cba and ∡abc are correctly defined and
satisfy the inequalities
e
∡bac ≥ ∡bac e
∡abc ≥ ∡abc e
∡bca ≥ ∡bca,
and, in addition, the following holds: for any two shortest path [pq] and [rs]
where r is a inner point of [pq], one has ∡prs + ∡srq = π.
4.2. Examples
With the three definitions in mind, we are ready to give more examples
of Alexandrov spaces. Apparently the examples that historically motivated
the notion of curvature are convex and saddle surfaces. Unfortunately, we
need to develop some machinery to prove that these examples satisfy our
definition; however we suggest the reader keep them in mind.
Here we consider examples of Alexandrov spaces whose justification does
not involve any techniques, namely a few trivial examples and polyhedral
spaces in low dimensions (1 and 2).
Example 4.2.1. Euclidean spaces are obviously Alexandrov spaces (both
nonpositively and nonnegatively curved at the same time).
Example 4.2.2. A convex set in an Alexandrov space is obviously an
Alexandrov space (with the same sign of curvature).
Example 4.2.3. An open set in an Alexandrov space, regarded with the
induced length metric, is an Alexandrov space (with the same sign of
curvature). This is so because our definitions are local, and the induced
length metric of an open set locally coincides with the metric restricted
from the ambient length space.
Example 4.2.4. A “fan” made of several segments glued together at one
end is a space of nonpositive curvature. The proof could be the same as for
Example 4.1.3. But now, using the angle definition, we are able to prove
110 4. Spaces of Bounded Curvature
this in a few words: every triangle in our space either has three zero angles
(in which case the angle condition obviously holds) or is degenerate (that
is, contained in one of its sides). In the latter case its comparison triangle
is degenerate as well.
Since every point in a locally-finite graph has a neighborhood that is
either a segment or a bunch of segments attached to one point, all locally-
finite connected graphs are nonpositively curved spaces.
Proof. We will show that every triangle in our space satisfies the angle
condition. Note that every shortest path with endpoints in the plane is
entirely contained in the plane. And a shortest path starting at a point in
the line can leave the line only through the origin O. Now one can see that
the angle condition for a triangle is trivial if all its vertices belong to the
plane or to the line, or if two vertices are in the line and one is in the plane.
In the only nontrivial case when vertices a, b reside in the plane and c in the
line, the triangle △abc consists of the flat triangle △abO and the “tail” Oc.
It is then an elementary exercise to check that the angles of this triangle are
less than the corresponding angles of a comparison triangle. ¤
Example 4.2.6. Consider several copies of the plane R2 and identify their
origins. This gives a nonpositively curved space. For two copies of R2 , this
space topologically looks like the cone x2 + y 2 = z 2 .
Warning: This cone together with its intrinsic metric induced from R3
is not an Alexandrov space!
The proof is essentially the same as in the previous example.
The above three examples are partial cases of the following general
statement.
Example 4.2.11. “Notebook” example 2.2.7 from Section 2.2 (several half-
planes glued together along their edges) is a nonpositively curved space.
It turns out that upper bounds for curvature are less restrictive and
therefore there are fewer different types of examples of nonnegatively curved
spaces. In particular, all two-dimensional spaces of nonnegative curvature
are topological manifolds, possibly with boundary. We will see later that
all convex surfaces are nonnegatively curved spaces. Here we prove this for
polyhedral surfaces. More generally, we can carry out a complete analysis
of which two-dimensional polyhedral spaces are Alexandrov spaces. It is
advisable to refresh the basics of polyhedral metrics given in 3.2. We need
the following important lemma, whose proof is already contained in the
proof for Example 4.1.4 above.
Lemma 4.2.13. The cone over a circle is a nonpositively curved space iff the
length of the circle is at least 2π. The cone over a circle is a nonnegatively
curved space iff the length of the circle is less than or equal to 2π.
112 4. Spaces of Bounded Curvature
is a circle, Lemma 4.2.13 implies that its length is not greater than 2π. If
the link is a segment, its length is not greater than π by Exercise 4.1.7.
2. Now we pass to the case of nonpositive curvature. We have to prove
the following:
The cone K over a graph Γ is a space of nonpositive curvature if and
only if the length of each nontrivial loop in Γ is not less than 2π.
First assume that K is nonpositively curved, and let us show that Γ
does not contain nontrivial loops of length less than 2π. Let γ be a shortest
nontrivial loop in Γ. Then, for every two points x, y ∈ γ, one of the two
intervals of γ between x and y is a shortest path in Γ. It follows that γ is a
convex set in Γ; hence the sub-cone over γ is a convex set in K (by Lemma
3.6.15). Therefore this sub-cone is nonpositively curved as long as the cone
is; hence L(γ) ≥ 2π by Lemma 4.2.13. Since γ is a shortest loop, all loops
have length no less than 2π.
Now assume that all nontrivial loops in Γ are not shorter than 2π, and
let us prove that K is nonpositively curved. Consider a triangle △abc in K.
First suppose that the sides of the triangle do not pass through O. Consider
the projection of △abc to Γ. This projection is a triangle △a′ b′ c′ whose
sides are shortest paths in Γ (recall the discussion of shortest path in cones
in Section 3.6.2). Consider two cases.
(a) Suppose that △a′ b′ c′ (as a subset of Γ) does not contain a simple
loop (a simple loop is a set homeomorphic to the circle). Then it is easy
to see that all three sides of △a′ b′ c′ have a common point d ∈ Γ; more
precisely, there are shortest paths [a′ d], [b′ d] and [c′ d] such that the triangle
△a′ b′ c′ , as a subset of Γ, coincides with the “fan” [a′ d] ∪ [b′ d] ∪ [c′ d]. This
fan is a convex set in Γ (because every pair of its points belong to a shortest
path contained in the fan). Hence the original triangle △abc is contained in
the cone over the fan which is a convex set in K consisting of three sectors
glued together along a ray. Then absolutely the same argument as for the
“notebook” Example 4.2.11 finishes the proof.
(b) Now suppose that △a′ b′ c′ does contain a simple loop. Then the
perimeter L = |a′ b′ | + |b′ c′ | + |a′ c′ | is no less than 2π. Consider the cone K1
over the circle S of length L. Fix a length-preserving map g from S onto
the triangle △a′ b′ c′ ; namely, split S into three arcs of lengths |a′ b′ |, |b′ c′ |
and |a′ c′ |, and let g map these arcs onto the respective sides of the triangle.
This map g induces the map g : K1 → K sending a point (x, t) ∈ K1 to
the point (g(x), t) ∈ K. The map g is an arcwise isometry (and hence is
nonexpanding); and the triangle △abc is the image of some triangle △a′′ b′′ c′′
in K1 (again, recall the structure of shortest paths in cones discussed in
Section 3.6.2).
114 4. Spaces of Bounded Curvature
Proof. We use only the following fact: if two sides of a planar triangle are
fixed, then the angle between them is a monotone (increasing) function of
116 4. Spaces of Bounded Curvature
the third side. In other words, if |xy| = |x′ y ′ | and |yz| = |y ′ z ′ | for two planar
triangles △xyz and △x′ y ′ z ′ , then ∡xyz > ∡x′ y ′ z ′ if and only if |xz| > |x′ z ′ |,
and vice versa.
Take a point c1 on the ray ad so that d is between a and c1 , and
|dc| = |dc1 |. Suppose that ∡adb + ∡bdc > π; then ∡bdc1 < ∡bdc. Hence
|bc1 | < |bc| = |b′ c′ | from the triangles △bdc and △bdc1 . Now applying the
same observation to the triangles △abc1 and △a′ b′ c′ (for which |ab| = |a′ b′ |
and |ac1 | = |a′ c′ |), we obtain that ∡bac1 < ∡b′ a′ c′ . Therefore |bd| < |b′ d′ |
(from the triangles △bad and △b′ ad′ ).
The case ∡adb+∡bdc < π is similar, up to reversing the inequalities. ¤
Remark 4.3.4. The lemma is true (and the same proof works) if the
triangles are placed in a sphere or a hyperbolic plane (the latter is defined
in Chapter 5) instead of the Euclidean plane. We will use this remark later
when we define more general curvature bounds.
Theorem 4.3.5. All the definitions (distance 4.1.2, triangle 4.1.9, angle
4.1.15, and monotonicity 4.3.1) are equivalent.
Remark 4.3.6. Our definitions are local. Their equivalence means that if
one of the definitions holds in a region U , then others hold in some region
V , possibly smaller than U . Nevertheless we will refer to all such regions as
normal regions.
Proof of Theorem 4.3.5. The proofs for spaces of nonpositive and non-
negative curvature are similar up to reversing the inequalities. To avoid rep-
etition, we prove the equivalence for nonpositively curved spaces and then
indicate the necessary modifications for the case of nonnegative curvature.
(1) The distance and triangle conditions are obviously equivalent.
(2) Assume the triangle condition holds, and let us show that then
the monotonicity condition holds as well. Consider a hinge of two
shortest paths α = [p, a], [p, b] and a point a1 in α starting at p.
Consider comparison triangles △pab and △pa1 b for the triangles
△pab and △pa1 b. Let ã be a point in the side pa1 such that
|pã| = |pa|. Then the triangle condition implies |ãb| ≥ |ab| = |ab|.
This means that ∡a1 pb ≥ ∡apb, which is the monotonicity of
angles.
(3) The monotonicity condition 4.3.1 implies the angle condition 4.1.15.
To prove this, consider a triangle △abc. Let its sides [ba], [bc] be
the shortest paths α, β, α(0) = β(0) = b. Monotonicity of angles
implies that
∡abc ≡ ∡(α, β) = lim θ(t, t) ≤ θ(|ab|, |bc|)
t→0
4.3. Equivalence of definitions 117
such that |qp1 | = |qp1 |, |qr1 | = |qr1 | the inequality |p1 r1 | ≤ |p1 r1 | (resp.
|p1 r1 | ≥ |p1 r1 |) holds.
PSfrag replacements ci c
a c
aai ci
ai b
b bi b a
bi bi ai
c
ci
Example 4.3.9. Let [ab] and [bc] be edges of a cube (we consider the surface
of the cube but not the interior). Consider segments [ai bi ] and [bi ci ] parallel
to [ab] and [bc] resp., at distances 1/i and placed in different faces of the
cube; see Figure 4.4.
The hinges ([ai bi ], [bi ci ]) converge to the hinge ([ab], [bc]) as i → ∞.
However ∡abc = π/2 while ∡ai bi ci = π for all i.
Example 4.3.10. Consider the coordinate plane R2 without the coordinate
quadrant {x > 0, y > 0}. Let a = (1, 0), ai = (1, −1/i), b = (0, 0),
bi = (−1/i, −1/i), c = (0, 1), ci = (−1/i, 1). Then the hinges ([ai bi ], [bi ci ])
converge to the hinge ([ab], [bc]), but ∡abc = π while limi→∞ ∡ai bi ci =
1/2π.
Note that the space in Example 4.3.9 has nonnegative curvature while
the one in Example 4.3.10 has nonpositive curvature. This illustrates the
following property of semi-continuity of angles
4.4. Analysis of Distance Functions 119
Proof. Let α and αi denote the angles ∡bac and ∡bi ai ci respectively. For
a small positive x, let b′ ∈ [a, b], c′ ∈ [a, c] and b′i ∈ [ai , bi ], c′i ∈ [ai , ci ] be
points at the distance x from a and ai respectively. Denote by θ(x) and
θi (x) the comparison angles ∡b e ′ ac′ and ∡b e ′ ai c′ , respectively. Notice that
i i
′ ′ ′ ′
|ai bi | = |ab|, |ai ci | = |ac| and |bi ci | → |bc| as i → ∞ (for fixed x), and hence
limi→∞ θi (x) = θ(x).
By the definition of the angle, α = limx→0 θ(x) and αi = limx→0 θi (x).
If X is nonpositively curved, then θ and θi are nondecreasing functions by
the monotonicity condition. Therefore θi (x) ≥ αi for all x, whence
θ(x) = lim θi (x) ≥ lim sup αi .
i→∞ i→∞
The next exercise is an analog of the fact that a convex function possesses
a “supporting linear function” at every point.
Exercise 4.4.6. Prove that a nonnegative nonexpanding function g : R → R
is E-convex (resp. E-concave) if and only if for every x ∈ R there exists a
function f ∈ E such that f (x) = g(x) and f ≤ g (resp. f ≥ g) everywhere.
Notice that in particular all nonnegative nonexpanding E-convex func-
tions are convex in the usual sense.
The class E is the set of nonnegative solutions of the differential equation
g ′′ (t)g(t) = 1 − (g ′ (t))2 . Analogously to the case of convex functions, a
smooth nonnegative nonexpanding function g is E-convex (resp. E-concave)
if and only if
1 − (g ′ (t))2 1 − (g ′ (t))2
g ′′ (t) ≥ (resp. g ′′ (t) ≤ ).
g(t) g(t)
Exercise 4.4.7. Prove this.
Convex functions, possibly nonsmooth, enjoy nice properties: they have
right and left derivatives everywhere; they are differentiable except at no
more than countably many points, and their derivatives have positive (resp.
negative) jumps at points of nonsmoothness. The same is true for E-convex
and E-concave functions:
Exercise 4.4.8. Let g : R → R be a nonnegative nonexpanding E-convex
(resp. E-concave) function. Prove that
1. g is continuous.
2. g has right and left derivatives everywhere, and the left derivative is
not greater (resp. not less) than the right one.
3. The set of points where g is not differentiable (i.e., where the right
and left derivatives are not equal) is finite or countable.
4. The derivative of g is continuous on the set where it is defined.
Hint: g(t)2 − t2 is a convex (resp. concave) function.
We are going to show that similar formulas are valid for spaces of
nonpositive and nonnegative curvature.
Later on we use the following notation. Let X be a length space,
γ : [0, T ] → X a unit-speed shortest path, a = γ(0), d = γ(T ), and
p ∈ X \ {a}. For each t ∈ [0, T ], set l(t) = |pγ(t)| and fix a shortest
path σt connecting γ(t) to p.
The following proposition is a general fact; it is valid for any length space
without any curvature restrictions.
Proposition 4.5.2. If there exist the angle α = ∡pad between the shortest
paths γ and [ap] = σ0 , then
l(t) − l(0)
(4.1) lim sup ≤ − cos α.
t→+0 t
Remark 4.5.3. The reader who likes the traditional notation for infini-
tesimal quantities will probably prefer the following form of the inequality
(4.1):
l(t) ≤ l(0) − t cos α + o(t), t → +0.
Remark 4.5.4. Since the left-hand side of (4.1) does not depend on σ0 ,
one can write
l(t) − l(0)
lim sup ≤ inf (− cos α),
t→0 t σ0
or, equivalently,
l(t) − l(0)
lim sup ≤ − cos αmin
t→0 t
where αmin is the infimum of angles between γ and all possible shortest
paths from a to p. (See also Corollary 4.5.7 below.)
Proof. Denote |ab| = y and |bc| = z (see Figure 4.5). By the cosine rule we
have
t2 + y 2 − z 2 y−z y+z t
cos α = = + .
2ty t 2y 2y
Then
¯ ¯ ¯ ¯
¯ y − z ¯ ¯y − z y + z t y − z ¯
¯cos α − ¯=¯ + − ¯
¯ t ¯ ¯ t 2y 2y t ¯
¯ ¯ ¯ ¯
¯y − z ¯ ¯y + z ¯ t t t t
¯
≤¯ ¯ ¯
·¯ − 1¯¯ + ≤1· + ≤
t ¯ 2y 2y 2y 2y y
¯ ¯ ¯ ¯
¯y − z ¯ ¯y + z ¯ t by the triangle inequality.
because ¯ t ¯ ≤ 1 and ¯ 2y − 1¯ ≤ 2y ¤
First let us show how the theorem follows from (4.3). Applying Lemma
4.5.5 to a comparison triangle for △aci bi yields
2
e i bi + ti .
l(0) = |pa| ≤ |pbi | + |bi a| ≤ |pbi | + |bi ci | − ti cos ∡ac
|bi ci |
Since |pbi | + |bi ci | = l(ti ), it follows that
l(ti ) − l(0) e i bi − ti = cos ∡ac
e i bi − ti .
≥ cos ∡ac
ti |bi ci | r
Then by (4.3),
l(ti ) − l(0) e i bi ) ≥ cos(π − α) = − cos α,
lim inf ≥ lim inf (cos ∡ac
i→∞ ti i→∞
and the theorem follows.
The proof for (4.3) is different for nonpositively and nonnegatively
curved spaces.
1. Let X be a space of nonnegative curvature. Then
e i bi ≤ ∡aci bi = π − ∡bi ci d
∡ac
by the angle condition. Then (4.3) follows because lim inf i→∞ ∡bi ci d ≥ α
by semi-continuity of angles (Theorem 4.3.11).
2. Let X be a space of nonpositive curvature. Denote by b the point in
[ap] = σ0 such that |ab| = r. Then ∡babi ≤ ∡babe i , and ∡bab
e i → 0 as i → ∞
because |bi b| → 0 while |ab| and |abi | stay bounded away from zero. Hence
∡babi → 0 as i → ∞.
By the triangle inequality for angles it follows that ∡ci abi → α as i → ∞.
Then
e i abi ≥ lim inf ∡ci abi = α
lim inf ∡c
i→∞ i→∞
4.5. The First Variation Formula 125
e i abi + ∡ac
by the angle condition. On the other hand, ∡c e i bi → π as i → ∞
e i abi + ∡ac
because ∡c e i bi + ∡ab
e i ci = π and ∡ab
e i ci → 0. Thus
e i bi = π − lim inf ∡c
lim sup ∡ac e i abi ≤ π − α.
i→∞ i→∞
Theorem 4.5.6 obviously implies the following rule for differentiating the
length: if {σt }t∈[0,T ] is a continuous family of shortest paths connecting p
to points γ(t), then there exists the right derivative dl/dt|t=0 = − cos α.
Moreover, one can differentiate the distance from p to γ(t), even if shortest
paths connecting p to γ(t) are not unique. Namely, the following holds:
three by re-scaling. We still assume that all metrics in question are strictly
intrinsic (more general definitions can be found in Chapters 9 and 10).
Historically, the notion of curvature comes from differential geometry in
the form of Gaussian curvature for two-dimensional surfaces or Riemannian
manifolds, and sectional curvatures in higher dimensions. (At this point,
it does not matter what these terms mean; they are just classical objects
of differential geometry. We will consider Riemannian manifolds and their
curvatures in Chapters 5 and 6. For now, it suffices to mention that Gauss-
ian and sectional curvatures are real-valued functions defined by means of
certain differential expressions.) A Riemannian manifold is a nonpositively
(resp. nonnegatively) curved Alexandrov space if and only if its sectional cur-
vatures are nonpositive (resp. nonnegative) everywhere. These two classes
of Riemannian manifolds play an important role in Riemannian geometry,
and this is one of the motivations for studying Alexandrov spaces.
Similarly, Alexandrov spaces of curvature ≥ k and ≤ k (that we are
about to define) include all Riemannian manifolds whose sectional curva-
tures are ≥ k (resp. ≤ k) everywhere.
To stress the connection between our previous classes (of nonnegative
and nonpositive curvature) and new ones (not defined yet), we mention the
following facts that will be proved later:
1. If X is a length space of nonnegative curvature (or, more generally,
of curvature bounded below), then its spaces of directions are spaces of
curvature ≥ 1.
2. If X is a length space of nonpositive curvature (or, more generally, of
curvature bounded above), then its spaces of directions have curvature ≤ 1.
• R2 , if k = 0;
128 4. Spaces of Bounded Curvature
√
• the Euclidean sphere of radius 1/ k (with its intrinsic metric), if
k > 0;
• the hyperbolic plane of curvature k, that is, the
√ standard Loba-
chevsky plane with the metric multiplied by 1/ −k, if k < 0.
Note that the k-plane is bounded (i.e., has finite diameter) if k > 0, and
not bounded if k ≤ 0. Denote the diameter of the k-plane by Rk , i.e.,
( √
π/ k, k > 0,
Rk =
∞, k ≤ 0.
All the other definitions can be re-formulated in the similar way. Except
a few inessential points, all statements and proofs related to the equivalence
of the definitions and simplest properties remain the same as for k = 0. Later
we will see that local properties of spaces of curvature ≤ k and of curvature
≥ k (well, not all local properties but all that we are interested in) do not
depend on k. But there are important global properties (properties “in the
large”) that do depend on k. More precisely, spaces of curvature ≥ k where
k > 0 and spaces of curvature ≤ k where k < 0 definitely have additional
interesting nonlocal properties.
√
In fact, it is not necessary to consider triangles of perimeter 2π/ k:
Exercise 4.6.7. Let k > 0, and X is a length space such that the distance
condition from Definition 4.6.2 (either for curvature ≥ k or for√curvature
≤ k) holds for all triangles whose perimeters are less than 2π/ k. Prove
that X is a space of curvature respectively ≥ k or ≤ k in the large.
case of curvature ≤ k, the distance condition for triangles
Hint: In the √
of perimeter 2π/ k simply reduces to nothing. √In the case of curvature ≥ k,
approximate a given triangle of perimeter 2π/ k by triangles with smaller
perimeters and obtain the desired inequalities by passing to a limit.
The reader probably has noticed that the second part of the above
exercise is much harder than the first one. The reason is, of course, that
in the case k1 > 0 the definition of curvature bounded by k2 involves more
triangles than that of curvature bounded by k1 . Namely the latter
√ does not
tell us anything about triangles of perimeter greater than 2π/ k1 while the
former
√ requires considering triangles of perimeters up to the greater value
2π/ k2 (or infinity if k1 < 0).
In fact, this difficulty does not exist: if a space has curvature√ ≥ k in
the large, then all triangles have perimeters no greater than 2π/ k. Well,
this rule has some exceptions among one-dimensional spaces (intervals and
circles). We will give a precise formulation and prove this fact in Chapter 10.
Globalization theorems. There are two theorems saying that, under some
conditions, local curvature bounds imply global ones. These theorems are
very important; actually they are central facts of the whole theory. Though
we prove these theorems in later chapters, it is worthwhile to formulate them
here.
1. Globalization theorem for nonpositive curvature (Theorem 9.2.9):
every complete simply connected space of curvature ≤ k, where k ≤ 0, is a
space of curvature ≤ k in the large.
4.7. Curvature of Cones 131
PSfrag replacements
a
b B
c
d
O C
A D
a′
b′ B’
c′
X
d′ A’
C’
D’
X
Figure 4.6: Cone is a space of K ≥ 0 if and only if the space itself has K ≥ 1.
4.7. Curvature of Cones 133
Smooth Length
Structures
This chapter deals with certain length spaces whose definition and analysis
involve smooth structures (calculus of variations). We begin with a basic
discussion of Finsler and Riemannian manifolds, followed by a metric intro-
duction to hyperbolic geometry. Then we consider some other interesting
examples. Here is a brief plan of the chapter along with some guidance
through it.
Section 5.1 describes Riemannian metrics in coordinates and promotes
an important concept that a Riemannian metric locally is almost Euclidean.
For the sake of simplicity of notations we mainly stick to two-dimensional
regions; generalizations to higher-dimensional manifolds are obvious. In
Section 5.2 we exploit an observation that shortest paths in a Riemannian
metric satisfy a particular second-order differential equation. This equation
enables us to draw geodesics with given initial positions and velocity vectors.
Using this, we can build normal coordinate systems. They are analogous
to polar, spherical, and Cartesian coordinates: one family of coordinate lines
is formed by geodesics parameterized by arc length, and the other family
consists of equidistant curves orthogonal to the curves from the first family.
There are two ways of regarding normal coordinates. These two ap-
proaches correspond to two dual viewpoints in mechanics and optics: prop-
agation of rays and wavefronts. Namely, one can first begin with a family
of geodesics (for instances, emanating from one point—thus obtaining nor-
mal coordinates centered at the point). These geodesics can be perceived
as rays, and their orthogonal trajectories, which happen to be equidistant
135
136 5. Smooth Length Structures
is called the derivative of ϕ. One can check that it is a linear map on each
Tp Ω.
that γ ′ (t) = dγ(t)/dt = V (γ(t)) for all t. The existence and uniqueness of an
integral curve with a fixed initial condition (γ(0) = p) is a principal theorem
in the theory of ordinary differential equations.
Exercise 5.1.1. Show that the Finsler length structure defined by (5.1) is
indeed a lower semi-continuous length structure, and therefore it is induced
by its intrinsic metric.
Exercise 5.1.3. Show that if one introduces new coordinates (u, v) such
that x = x(u, v), y = y(u, v) the rule of converting E to the new coordinates
is given by the following formula:
∂x ∂x ∂y ∂y
Enew = E( )2 + 2F + G( )2 .
∂u ∂u ∂u ∂u
Find analogous formulas for F and G.
Note: Recall that a coordinate system is formally a map ϕ : Ω1 → Ω.
Changing coordinates to (u, v) such that x = x(u, v), y = y(u, v), where
x(u, v) and y(u, v) are two given functions (it is just convenient to denote
them by the same letters as coordinates) formally means that we consider
a map ψ : Ω2 → Ω1 , ψ(u, v) = (x(u, v), y(u, v)), and our new coordinate
system is a map ϕ ◦ ψ : Ω2 → Ω.
If you are not happy with these abstract definitions, you may think of
a Riemannian manifold as a smooth surface in a Euclidean space. With
this simplification, one does not even sacrifice generality: according to the
famous Nash’s Embedding Theorem, every Riemannian manifold is isometric
to a smooth embedded surface in a Euclidean space of some (sufficiently
large) dimension. Two-dimensional examples of this kind are discussed in
the next subsection.
It is still not clear yet that the shortest paths are smooth. This issue is
discussed in the next section. We finish this section with a brief comment
on the relationship between intrinsic and extrinsic geometries of a surface.
By intrinsic geometry we understand the properties that depend only
on the Riemannian metric. In case of an embedded surface the Riemannian
metric is induced by an embedding r.
It is easy to give examples of different embeddings that induce the same
Riemannian structure.
Exercise 5.1.9. Verify that the embeddings ϕ, ψ : [0, 1] × [0, 1] → R3 given
by
ϕ(u, v) = (u, v, 0),
ψ(u, v) = (sin u, cos u, v)
induce the same Riemannian metric.
To show that two Riemannian metrics are not isometric, one looks for a
metric property that can tell them apart. Here is an example:
Lemma 5.1.12. Given a point p, one can choose coordinates such that the
metric coefficients at p are E = G = 1, F = 0. More precisely, one can
choose a coordinate system ϕ : Ω′ ⊂ R2 → Ω, ϕ(x0 , y0 ) = p such that the
metric coefficients at p are E(x0 , y0 ) = G(x0 , y0 ) = 1, F (x0 , y0 ) = 0 in this
coordinate system.
Thus one can choose a coordinate system such that, at one given point
p, the metric coefficients look the same as the Euclidean metric in Cartesian
coordinates. This means that the coordinate vectors form an orthonormal
basis of (Tp Ω, h, ip ). In other words, the derivative d(x0 ,y0 ) ϕ of the coordinate
map is a linear isometry
d(x0 ,y0 ) ϕ : (T(x0 ,y0 ) R2 , Euclidean Scalar Product) → (Tp Ω, h, ip ).
between their velocity vectors γ1′ = γ1′ (0) and γ2′ = γ2′ (0) at t = 0.
Exercise 5.1.15. Prove Lemmas 5.1.13 and 5.1.14.
Remark. The arguments we know are absolutely straightforward but some-
what tedious.
∂
∂ρ = (−r sin ρ, r cos ρ) (where the right-hand side expressions are given in
Cartesian coordinates). Hence
(5.7) E(r, ρ) = 1, F (r, ρ) = 0, G(r, ρ) = r2 .
Now let us consider the sphere of radius R with the (degenerate) coor-
dinates given by the map
(ϕ, ρ) −→ (x = R sin(ϕ/R) cos ρ, y = R sin(ϕ/R) sin ρ, z = R cos(ϕ/R)).
This is almost the usual spherical coordinate system, and the only difference
is that the ϕ-coordinate is rescaled to have the ϕ-lines parameterized by arc
length. A trivial computation yields:
(5.8) E = 1, F = 0, G = R2 sin2 (ϕ/R).
Similarly to the previous case of the Euclidean plane, only the rotational
symmetry (ϕ, ρ) −→ (ϕ, ρ + const) of the metric is obvious directly from
the formulas. Of course, we know that a sphere is a perfectly symmetric
space from Euclidean considerations: there is a rigid motion of R3 that
maps the sphere to itself and sends any given point to any other given
point. It is, however, not at all transparent from the intrinsic viewpoint
when we look at the formula (5.8); if we defined a sphere as a surface whose
Riemannian metric is given by (5.8), producing formulas for all isometries
would be quite a task (compare with the previous exercise; a reader who
likes spherical geometry may enjoy converting to spherical coordinates a
3-dimensional rotation about a line other than the z-axis).
Now we define the hyperbolic plane of curvature k (where k < 0!) as a
plane with Riemannian metric given by the following formulas for its metric
coefficients:
1 √
(5.9) E = 1, F = 0, G = sinh2 ( −k r).
−k
Other definitions of the hyperbolic plane are given in Section 5.3. It will be
shown later that the hyperbolic planes are as homogeneous as the spheres
and the Euclidean plane.
150 5. Smooth Length Structures
Exercise 5.1.19. Using (5.8) and (5.9), compute the length of a circle
of radius r in the sphere of a radius R and in the hyperbolic plane of a
curvature k.
This definition may look strange at first glance. To motivate the choice of
this differential equation, in the next chapter we will use variational methods
to derive it as an equation that is satisfied by every smooth shortest path.
A reader can notice that we actually will not use the result obtained by a
variational argument. Instead, we will give a simple proof based on certain
properties of the equation that the geodesics are the locally shortest curves.
It will also become obvious later that this (coordinate) equation defines the
same geometric object if we change the coordinate system.
The reader familiar with classical mechanics may also note that these
equations describe a free particle whose kinetic energy is given by E ẋ2 +
2F ẋẏ + Gẏ 2 (the quadratic form of the metric).
Exercise 5.2.2. Prove that every geodesic (i.e., every solution of equations
(5.10) and (5.11)) except a constant map is a curve parameterized propor-
tionally to arc-length.
5.2. Exponential Map 151
d d
Hint: Multiply equations (5.10) and (5.11) by dt x and dt y, respectively,
and then sum them up. After that compare the result with the equation
µ µ ¶2 µ ¶2 ¶
d dx dx dy dy
E + 2F +G = 0.
dt dt dt dt dt
Exercise 5.2.3. Show by a direct computation that if a curve is a geodesic
in a coordinate system, then it is a geodesic with respect to every coordinate
system.
Hint: Use Exercise 5.1.3.
We are going to show that the converse is also true, namely, that in
normal coordinates the x-lines and the y-lines are mutually orthogonal (not
only at x = 0). This fact is called the Gauss Lemma:
Lemma 5.2.8. In normal coordinates the metric coefficient F = hX, Y i is
identically zero.
such that x(t) = x0 . Then the segment σ|[0,b ] is contained in our coordinate
1
system, and we can estimate its length by integration:
Z b1 p Z b1 Z b1
′
L(σ) ≥ ′ 2 ′ 2
(x ) + G(y ) dt ≥ |x | dt ≥ x′ dt = x0 .
0 0 0
Hence the length of any path from p to q is at least b. Analyzing the equality
case in our inequalities, the reader can easily verify that it can happen only
if σ = γ. ¤
Exercise 5.2.10. Show that a by-product of the argument above is the fact
that all shortest paths are smooth!
5.3.1. Motivations.
4. The last but still important reason is that historically the discovery
and study of hyperbolic geometry had tremendous impact on geometry and
mathematical ideology in general.
cross-sections of the sphere). The spherical angle between two lines is de-
fined as the Euclidean angle between the Euclidean planes containing these
lines. Finally the “spherical distance” is defined to be the Euclidean angle
at the origin. Note that the “ spherical straight lines” do not look straight
from the viewpoint of the ambient Euclidean space, and as well the spherical
distance is different from the distance in the ambient space. A reader may
say: “Aha, but these are just geodesics, angles and distances in the induced
intrinsic metric!” This is true, and this fact had its impact on the history of
hyperbolic geometry. However, we are just lucky to have such a nice model
for spherical geometry—not all models can be explained in such a natural
way.
After this general digression, let us turn back to hyperbolic geometry.
There is a very short axiomatic definition of the hyperbolic planes:
consider Euclid’s set of axioms and substitute the Fifth Postulate (stating
that “two lines parallel to the same line and having a point in common must
coincide”) by its negation. Certainly, Euclid’s formulations are not rigorous
enough by the modern standards, so one may use their version updated by
Hilbert. A space satisfying this new set of axioms is a hyperbolic plane.
There is, however, a long historical path behind this definition.
The story goes back to Euclid, who demonstrated ingenious intuition by
writing that after a few unsuccessful attempts to derive the fifth postulate
from other axioms, he gave up since he felt it could lead him too far.
Indeed, numerous attempts to prove the Fifth Postulate (based on the other
postulates) led quite far! There were many “proofs”, which usually drew a
“contradiction”, but a contradiction with the everyday intuition rather than
the other axioms. For instance, one can prove that, if the Fifth Postulate
were incorrect, then there is a number such that the area of every triangle
is smaller than this number. Is not it a “contradiction”, for we “know” that
one can build as huge a triangle as she wishes? The only problem is that
this “knowledge” does not follow from Euclid’s axioms.
Perhaps the first mathematician who systematically studied the system
of axioms with the negation of the Fifth Postulate was K. F. Gauss. (The
mathematical world described by this collection of axioms is called “non-
Euclidean geometry”.) Gauss seriously suspected that there may be no
contradiction in this axiomatic system, although many properties of non-
Euclidean geometry are very different from what the common sense based
on physical experience would prompt us to expect. In this respect Gauss
went as far as Lobachevsky and Bolyai, and the reason why the latter two are
known as “the inventors of non-Euclidean geometry” is that Gauss had not
published his work. Surprisingly enough, it was just “too much knowledge”
that kept Gauss from publishing his discovery: he realized that one had to
158 5. Smooth Length Structures
a metric by a constant factor). Let us begin with the hyperbolic plane with
k = −1, which will be referred to as the hyperbolic plane.
It may seem that the hyperbolic plane defined via the Poincaré model is
rather nonhomogeneous: it even has two types of lines: Euclidean semi-
circles and rays. When we approach the x-axis of the Poincaré model,
5.3. Hyperbolic Plane 161
hyperbolic distances are huge compared to Euclidean ones; and they become
small as we move towards y → ∞. However, intrinsically the hyperbolic
plane is quite homogeneous; this becomes obvious by looking at hyperbolic
rigid motions. Indeed, given two points, it is easy to find a hyperbolic rigid
motion that maps one of the points to the other. Notice that one can do
this even without using inversions! They become handy when we want to
map a given ray to another given ray (to prove that hyperbolic geometry is
isotropic).
Note that the metric coefficients are given by exactly the same expression
as in the formula (5.9) for k = −1. Hence we proved that the definition of
the hyperbolic plane via the Poincaré model is equivalent to the one given
in Section 5.1 by (5.9).
This example might prompt a completely wrong idea that this procedure
always leads to an isometric space. Actually, this happens only in excep-
tional cases. Indeed, let Sc be the length space whose points are points of
the unit 2-dimensional sphere S 2 , and the distance function dc is obtained
√
from the distance function of S 2 by multiplication by c. Then Sc is not
√
isometric to S1 = S 2 : Sc is isometric to a sphere of the radius c. Thus
we obtain the family of the spheres. The number c−1 is called the curvature
of the sphere (and it is equal to the Gaussian curvature of its Riemannian
metric; see the next chapter).
Analogously, for each positive c we consider a space Hc obtained from
√
H by multiplying all distances by c. The formula for its length structure
164 5. Smooth Length Structures
reads: ¯ Z ¯
b
1 ¯¯ dγ(t) ¯¯
√
L(γ, a, b) = c ¯ dt ¯ dt.
a y
We will see that all spaces Hc are different (pairwise nonisometric). The
number −1/c2 is called the curvature of Hc (and we will also see later that
it is equal to its Gaussian curvature).
Remark 5.3.8. Let us fix a number r, and let Bk be a ball of radius r in
Hk . Then the spaces Bk “converge” to B0 as k → 0, and thus on any fixed
scale the hyperbolic metric with k → 0 looks more and more like a Euclidean
one (one can formalize this by choosing two normal coordinate systems and
comparing distances between points with the same coordinates, as in the
proof of Lemma 5.1.16).1
Remark 5.3.9. In fact, already the choice of the rigid motions of the
hyperbolic plane in the Poincaré model determines the Riemannian length
structure up to a multiplicative constant. Indeed, let us try to choose
expressions for the metric coefficients. Since the group of isometries contains
the translations along the x-axis, the expressions for metric coefficients
cannot depend on the x-coordinate, and thus they are functions of y. Since
Euclidean homotheties are hyperbolic isometries, the metric coefficients
must be homogeneous of order −2; that is, they have to satisfy f (cy) =
f (y)/c2 for all positive c. Hence each metric coefficient has to be of the
form const · y −2 . Now one notices that the derivatives of the inversions that
fix a given point generate all rotations around this point, and hence the
hyperbolic scalar product has to be a multiple of the Euclidean one. This
uniquely determines E = G = const · y −2 , F = 0.
Rigid motions and hyperbolic lines. We want to make sure that our
definitions of geometric notions introduced via the Poincaré model are con-
sistent. First, let us verify that the hyperbolic rigid motions are isometries.
Since all transformation called “the hyperbolic rigid motions” are bijective,
it suffices to check that they preserve the hyperbolic length structure.
Parallel translations along the x-axis obviously leave the integrand in
(5.13) unchanged, for they do not change the magnitude of γ ′ , nor the y-
coordinate. The same applies to the symmetries in vertical lines.
1It is a historical fact that Gauss tried to experimentally verify whether the physical space
is non-Euclidean. He did this by measuring angles in a huge triangle (with the tops of three
mountains as vertices, and using light rays as the sides of his triangle). He observed that, with
available precision, the sum of the angles was π. This allowed him only to conclude that, even
if that was a hyperbolic triangle, the curvature had to be very small. As a matter of fact, the
measurements of Gauss did not have nearly enough precision to detect effects caused by general
relativity, or otherwise he indeed would have noticed that the sum of the angles is different from
π, indicating that physical space is indeed curved!
5.3. Hyperbolic Plane 165
For a curve γ(t) = (x(t), y(t)), the length of its composition with I is
Z b ¯ dI(γ(t)) ¯
¯ ¯
L(I(γ, a, b)) = Iy (γ(t))−1 ¯ ¯ dt
a dt
Z bµ ¶−1 sµ ¶2 µ ¶2
y d x d x
= + dt = L(γ, a, b).
a x2 + y 2 dt x2 + y 2 dt x2 + y 2
Exercise 5.3.10. Re-prove these statements using a complex representation
az + b
T (z) = , a, b, c, d ∈ R, ad − bc = 1,
cz + d
for a hyperbolic rigid motion T .
Exercise 5.3.11. Let us denote by Ta,b,c,d a hyperbolic rigid motion given
by a complex formula
az + b
Ta,b,c,d (z) = , a, b, c, d ∈ R, ad − bc = 1.
cz + d
Verify that the composition of transformations Ta,b,c,d and Ta′ ,b′ ,c′ ,d′ corre-
sponds to the product of matrices with entries a, b, c, d and a′ , b′ , c′ , d′ :
Ta,b,c,d ◦ Ta′ ,b′ ,c′ ,d′ = Taa′ +bc′ ,ab′ +bd′ ,ca′ +dc′ ,cb′ +dd′ .
Hence the group of orientation preserving hyperbolic rigid motions is iso-
morphic to SL2 (R)—the group of 2 × 2-matrices with determinant 1.
166 5. Smooth Length Structures
Proof. The proof becomes obvious if we use the model in the disc. Indeed,
both vertices of the rays can be mapped to the center of the disc (by
homogeneity). Now the rays are represented by two semi-open Euclidean
segments (two radii of the disc), and there is a rotation that maps one of
them to the other. ¤
We will use the richness of the group of hyperbolic rigid motions to show
that the hyperbolic lines are shortest paths. Notice that we will prove not
only that hyperbolic lines are geodesics (and hence locally shortest paths):
we will show that every segment of a hyperbolic line is the shortest path
between its endpoints!
First of all, let us prove a particular case of this statement:
Lemma 5.3.19. A segment that belongs to a diameter of the disc in the
disc model is the unique shortest path between its points.
Now, given two points p and q (in the disc model), by Lemma 5.3.14 we
can map them by a rigid motion I onto two points I(p), I(q) lying on the
same diameter of the disc. Since I is an isometry of the hyperbolic plane, the
image of a shortest path between p and q is a shortest path between I(p)
and I(q). According to Lemma 5.3.19, the unique shortest path between
I(p) and I(q) is the Euclidean segment [I(p), I(q)]. Therefore its pre-image,
which is an arc of the circle passing through p and q and orthogonal to the
boundary of the disc, is the unique shortest path between p and q.
Along the way we verified one of the principal axioms of both Euclidean
and hyperbolic geometries: there is exactly one line passing through any pair
of distinct points. Of course, this statement becomes obvious in the Poincaré
model: move the two points to the same vertical line of the Poincaré model
by a (hyperbolic) rigid motion. Now it is clear that two distinct points in the
same vertical line cannot belong to the same circle centered at the x-axis,
and hence the only line that passes through the points is the vertical ray.
Exercise 5.3.20. Draw an example of two intersecting lines that do not
intersect another line (a counter-example to the Fifth Postulate).
168 5. Smooth Length Structures
Exercise 5.3.21. Show that, for every two intersecting lines, there exists
a line that is perpendicular to one of the lines and does not intersect the
other one.
whose closures have one endpoint in Γ the same. The property of lines to
be asymptotic is an equivalence relation (prove this). Nonintersecting and
not asymptotic lines are said to be hyperparallels.
It is clear (from the disc model) that, for an oriented line l and a point
p∈/ l, there is exactly one line lp+ passing through p and asymptotic to l. If
we reverse the orientation of l, we also get one line lp− that passes through a
and is asymptotic to l with reversed orientation. All lines passing through
p “between” lp+ and lp− are hyperparallel to l (i.e., do not intersect it and
are not parallel to it). We suggest the reader make a sketch to visualize this
observation.
Exercise 5.3.25. Prove that in the Poincaré model the horocycles are
represented by the circles tangent to the x-axis and the horizontal lines.
Prove that in the disc model the horocycles are represented by the circles
tangent to the boundary circle.
For a fixed t, one can think of the function dt (p) = d(p, γ(t)) − t as the
distance function to γ(t) normalized by subtracting t to get zero at γ(0).
Hence a level curve Sτ,t = {p ∈ H : d(p, γ(t)) − t = −τ } of dt is a circle
centered at γ(t) and passing through a point γ(τ ) (orthogonally to γ ′ (τ )).
The reader probably has already realized that a web formed by a family
of asymptotic geodesics and the family of horocycles orthogonal to them can
always be represented by coordinate lines in a normal coordinate system. In
particular, horocycles form an equidistant family.
at a reference point y), and a set of limit points of the image of X. This
set of limit points, denoted by X(∞), is the Busemann boundary at infinity
of X, and its elements are called Busemann functions. These are distance-
like Lipschitz-one functions that arise as limits (uniform on compacts) of
sequences of normalized distance functions dn (z) = d(z, xn ) − d(xn , y).
There is another notion of the boundary at infinity (there are at least
five different and useful notions of boundaries at infinity). The ray boundary
consists of equivalence classes of rays starting from a reference point (and
two rays are equivalent if they are asymptotic, that is, stay within bounded
distance from each other). This set is equipped with the topology of uniform
convergence on open sets. For the hyperbolic plane, ray boundary and
Busemann boundary coincide (check this!).
Theorem 5.3.32.
∡a + ∡b + ∡c − π = k · Area(T ),
Proof. It is enough to consider two cases: k = 1 and k = −1, and then use
a dilation.
Case 1: k = 1. A spherical triangle can be represented as the inter-
section of three semi-spheres. Our argument will consist of an elementary
application of the inclusion-exclusion formula.
Let a′ , b′ , c′ be the points of the unit sphere opposite to a, b, c,
respectively. For each pair of opposite points, say a, a′ , one of the two
bi-gons B(aa′ ), B ′ (aa′ ) with the vertices a and a′ covers the triangle △abc,
and the other one covers the antipodal triangle △a′ b′ c′ . Notice that the area
of each of these bi-gons is equal to twice the angle α of △abc at a.
All the six bi-gons cover the entire sphere with multiplicity 1 (each point
is covered by exactly one of them) except that the triangles △abc and △a′ b′ c′
are covered three times each. By the inclusion-exclusion formula we get
4π = Area (S 2 )
= 2(Area (B(aa′ )) + Area (B(bb′ )) + Area (B(cc′ ))) − 4Area (T ).
Together with the observation that the area of each of the bi-gons is twice
the corresponding angle of △abc this proves case 1.
Case 2: k = −1. Observe that if a triangle (maybe an ideal one) is cut
into two triangles and the conclusion of Theorem 5.3.32 holds for two of the
three triangles, then it also holds for the third one. Every triangle can be
obtained by cutting an ideal triangle off another ideal triangle. Hence it is
enough to prove 5.3.32 for ideal triangles.
Consider an ideal triangle △abc, a ∈ Γ. Using the Poincaré model, one
can map it (by a hyperbolic rigid motion) to such an ideal triangle that two
of its sides belong to two vertical rays, and two vertices lie on the circle
x2 + y 2 = 1 (and the third one is mapped to the “infinitely remote point
y = ∞). An elementary (Euclidean) geometric consideration shows that if
the angles of this triangle are α and β (α is at the left vertical side and β is at
the right one), then its vertices have Cartesian coordinates (− cos α, sin α)
and (cos β, sin β).
Now the argument is based on a straightforward integration. Indeed,
the hyperbolic area of the triangle T is equal to
Z cos β Z ∞
dy
Area (T ) = dx √ 2
= π − α − β.
− cos α 1−x2 y
Exercise 5.3.34. Show that the area of a hyperbolic triangle does not
exceed π.
Remark 5.3.35. There is a remarkable trick that allows one to prove many
statements in hyperbolic geometry by their spherical analogs. For example,
in the proof of the Gauss-Bonnet Formula we could not apply the inclusion-
exclusion formula in Case 2 since the hyperbolic plane has infinite area.
However, this trick allows one to apply the result of Case 1 to claim that
the Gauss-Bonnet Formula is correct in general. For instance, one can also
use this trick to show that the altitudes of a hyperbolic triangle intersect in
one point.
Here is the idea of the trick. There are many ways to formalize the
claim that spheres, the Euclidean plane and hyperbolic planes form a one-
parameter family. For instance, formulas (5.9), (5.8), and (5.7) √ for the metric
coefficients can be represented as one formula G(r) = (Re sin( k r))2 using
complex numbers. Now if an identity that can be expressed by a formula
via metric coefficients is valid for all positive values of k, then by analyticity
it is true for all k (for a nonzero analytic function has only isolated zeroes).
All details are left to the interested reader.
Large triangles. First of all, the Gauss-Bonnet Formula implies that the
area of a triangle is at most π. Since the area of a circle of radius 1 is greater
than π, the radius of a circle inscribed in a triangle is less than 1 (of course,
this is a very rough upper bound; can you give a better estimate?) This
5.3. Hyperbolic Plane 175
means that if one chooses three points in the hyperbolic plane (arbitrarily
far apart), the three segments connecting the points “almost stick together
in the center of the triangle”: they intersect the same disc of radius 1! Hence
each side of a hyperbolic triangle is contained in the 2-neighborhood of the
union of the two other sides. When seen from far away, a big hyperbolic
triangle looks like a union of three rays emanating from the center of the
triangle: the triangle is contained in the 1-neighborhood of this union.
Loosely speaking, the triangle looks very “slim”, as if its sides have been
“sucked inside”.
Using these considerations, it is easy to observe the following remarkable
phenomenon. Consider a triangle ∆abc, and let p be a point on the
side [bc]. Then at least one of the triangles ∆abp and ∆acp is almost
degenerate, in the sense that at least one of the following inequalities holds:
d(a, b) + d(b, p) ≤ d(a, p) + 4 or d(a, c) + d(c, p) ≤ d(a, p) + 4. This property
will be used to define Gromov hyperbolic (δ-hyperbolic) spaces.
Note that, unlike Euclidean geometry, there is a criterion for hyperbolic
triangles to be congruent “by three angles”: two triangles are congruent
(one of them can be obtained from the other by a rigid motion) if each
angle of one triangle is equal to the corresponding angle of the other one.
On the other hand, funny enough, but there are no rectangles in hyperbolic
geometry!
every shortest path of the new metric stays within a bounded distance from
a hyperbolic line. This suggests that the large-scale structure of geodesics
in the hyperbolic plane is very stable; this phenomenon is closely related to
structural stability of hyperbolic (Anosov) flows.
tiles. However, the only regular tiles that can be used to pave the Euclidean
plane are triangles, squares and hexagons, whereas the hyperbolic plane can
be paved by regular n-gons for all n.
i = 1, . . . , k, such that
X
γ ′ (t) = vi (t)Vi (γ(t)).
This relation means that γ is allowed to move along any linear combination
of Vi ’s. One can think of Vi ’s as available controls to direct a moving point;
then vi ’s would be control functions. For two vector fields V1 and V2 that
are linearly independent at every point, one can introduce a distribution H
of the planes spanned by V1 and V2 , thus reducing this construction to the
previous one. A difference arises if the vector fields happen to be linearly
dependent at some points.
Exercise 5.4.2. Show that, for V1 (x, y, z) = (1, 0, 0) and V2 (x, y, z) =
(0, 1, x), every two points are connected by an admissible path.
Lie bracket. Here we briefly recall the notion of Lie bracket; one can find
details and proofs in many books (see for example [BC]).
Two main objects associated with a smooth vector field V are the cor-
responding derivation of smooth functions f → V f and the one-parameter
group of diffeomorphisms ϕtV generated by V . Recall that, for each t, ϕtV is
a diffeomorphism, and for a fixed p,
d t
ϕ (p) = V (ϕtV (p)).
dt V
182 5. Smooth Length Structures
In other words, for each x, γ(t) = ϕtV (p) is an integral curve of V that passes
through p at t = 0. (A curve γ is an integral curve of a vector field V if
d
dt γ(t) = V (γ(t)). )
Here are three equivalent definitions of Lie bracket [V, W ] of vector fields
V , W.
Definition 5.4.3. Lie bracket [V, W ] is a vector field defined by the condi-
tion: [V, W ]f = V W f − W V f for every smooth function f .
Of course, one has to prove that there exists a unique vector field [V, W ]
that differentiates functions according to the formula in the definition. The
definition immediately implies that Lie bracket is linear in each argument
and skew-symmetric. It is easy to check that the following property holds:
[f V, W ] = f [V, W ] − (W f )V , where f is a smooth function.
The next definition is especially important in our geometric context:
d ¯¯ √ √ √ √
Definition 5.4.4. [V, W ](p) = ¯ (ϕV t ◦ ϕWt ◦ ϕV− t ◦ ϕ− t
W (p)).
dt t=0+
To better understand its geometric meaning, one can use a coordinate
chart (to be able to “subtract points”) and rewrite it in a less invariant form
as
1
[V, W ](p) = lim 2 (ϕtV ◦ ϕtW ◦ ϕ−t −t
V ◦ ϕW (p) − p).
t→0 t
We suggest the reader make a sketch with a “quadrilateral” formed by four
segments of integral curves of V and W traversed back and forth. The
“quadrilateral” has its first vertex at p, then its first side follows the integral
curve of W backwards for time t, and so on. This “quadrilateral” does
not quite “close up” (unless we are dealing with coordinate vector fields,
in which case we actually get a closed “quadrilateral”): when one follows
all four sides of the “quadrilateral”, he gets to a point p′ close to p; this
displacement p′ − p is of second order in t; dividing it by t2 and sending
t → 0 gives the Lie bracket.
The last definition stresses similarity between Lie brackets and usual
derivatives:
Definition 5.4.5.
1 d¯
[V, W ](p) = lim (dϕ−tV W (ϕtV (p)) − W (p)) = ¯t=0 dϕ−t
V W (γ(t)),
t→0 t dt
where γ is the integral curve of v with γ(0) = p.
Proof. Let us prove that, for every p ∈ Ω, the set of points that can be
connected with p by admissible paths is open.
Let ϕt1 , ϕt2 be the one-parameter groups of diffeomorphisms generated
d t
by V1 , V2 : dt ϕi (p) = Vi (ϕti (p)), i = 1, 2. Recall that
d ¯¯ √ √ √ √
[V1 , V2 ](p) = ¯ (ϕ1 t ◦ ϕ2 t ◦ ϕ− 1
t
◦ ϕ−
2
t
(p)).
dt t=0+
Consider a map F : R3 → Ω:
√ √ √ √
|τ | sign(τ ) |τ | − |τ | −sign(τ ) |τ |
(5.17) F (u, v, τ ) = ϕu1 ◦ ϕv2 ◦ ϕ1 ◦ ϕ2 ◦ ϕ1 ◦ ϕ2 (p).
This map is at least C 1 -smooth (check this!) and
∂ ∂ ∂
∂u F (0) = V1 (0), ∂v F (0) = V2 (0), ∂τ F (0) = [V1 , V2 ](0),
where 0 = (0, 0, 0).
Hence by the Inverse Function Theorem, the image of F contains a
neighborhood of p. Now it remains to observe that a point F (u, v, τ ) is
connected with p by the following admissible path γ : [0, 6] → Ω, which
alternates moving along V1 and V2 (that is, it is built out of segments of
integral curves of V1 and V2 ):
√
−sign(τ ) |τ |t
γ(t) = ϕ2 (p), t ∈ [0, 1],
√ √
− |τ |(t−1) −sign(τ ) |τ |
γ(t) = ϕ1 ◦ ϕ2 (p), t ∈ [1, 2],
√ √ √
sign(τ ) |τ |(t−2) − |τ | −sign(τ ) |τ |
γ(t) = ϕ2 ◦ ϕ1 ◦ ϕ2 (p), t ∈ [2, 3],
√ √ √ √
|τ |(t−3) sign(τ ) |τ | − |τ | −sign(τ ) |τ |
γ(t) = ϕ1 ◦ϕ ◦ϕ ◦ϕ (p), t ∈ [3, 4],
√ 2 √ 1 √ 2 √
v(t−4) |τ | sign(τ ) |τ | − |τ | −sign(τ ) |τ |
γ(t) = ϕ2 ◦ ϕ1 ◦ ϕ2 ◦ ϕ1 ◦ ϕ2 (p), t ∈ [4, 5],
184 5. Smooth Length Structures
√ √ √ √
|τ | sign(τ ) |τ | − |τ | −sign(τ ) |τ |
γ(t) = ϕu(t−5) ◦ϕv2 ◦ϕ1 ◦ϕ2 ◦ϕ1 ◦ϕ2 (p), t ∈ [5, 6].
Hence every point in the image of F can be connected with p; since the image
of F contains a neighborhood of p, its connectivity component is open.
Since connectivity components are open and either coincide or do not
intersect, the theorem immediately follows from connectedness of Ω. ¤
Exercise 5.4.7. Show that under the assumptions of Theorem 5.4.6 every
two points in Ω can be connected by a smooth admissible path (as opposed
to a piecewise smooth one).
Ball-Box Theorem. Now we want to get some insight into infinitesimal be-
havior of a sub-Riemannian metric generated by a distribution Hp in Ω. We
will restrict ourselves to the case dim Ω = 3; unlike Chow’s Theorem, whose
general version does not involve new ideas and requires only more tedious
and cumbersome notations, even the formulation of the Ball-Box Theorem
5.4. Sub-Riemannian Metric Structures 185
Proof. I. First we prove that there is a positive c such that Prc ⊂ Br for all
sufficiently small r. To prove this inclusion, it suffices to show that every
point from Prc can be connected with the origin by an admissible path of
length at most r. Notice that W = [V1 , V2 ](0, 0, 0) does not lie in the xy-
plane (by Chow’s condition).
We will use the map F (for p = (0, 0, 0)) defined in (5.17) and an
admissible curve γ connecting (0, 0, 0) and F (u, v, τ ) constructed in the proof
of the Chow–Rashevsky Theorem 5.4.6. Let us estimate the length of γ. Let
M be such that |V1 (p)| ≤ M , |V2 (p)| ≤ M for all p in a neighborhood p of
the origin. Then the length of γ is bounded by M |u| + M |v| + 4M |τ |.
Indeed, the velocity
p of γ pfor the p intervals p t ∈ [0, 1], t ∈ [1, 2], t ∈ [2, 3],
t ∈ [3, 4] is − |τ |V2 , − |τ |V1p, |τ |V1 , |τ |V2 respectively. Hence its
speed for t ≤ 4 is bounded by |τ |M , and thus the length of γ|[0,4] is at
p
most 4M |τ |. Analogously for t ∈ [4, 5] and t ∈ [5, 6], the velocity of γ
is vV1 and uV2 respectively, and hence the length of these segments of γ
is at most M |u| + M |v|. Thus we conclude that the Carnot–Carathéodory
p
distance from the origin to F (u, v, τ ) is at most M (|u| + |v| + 4 |τ |). Thus
it remains to show that (there exists a positive c such that) c
p Pr is contained
in the image under F of the set {(u, v, τ ) : |u| + |v| + 4 |τ | ≤ r/M }. This
image in its turn contains the image of the rectangle
{(u, v, τ ) : |u| ≤ δr, |v| ≤ δr, |τ | ≤ δ 2 r2 },
186 5. Smooth Length Structures
where δ = 1/6M .
Now all is left is to recall that
∂ ∂ ∂
F (0, 0, 0) = V1 (0, 0, 0), F (0, 0, 0) = V2 (0, 0, 0), F (p) = W.
∂u ∂v ∂τ
Since V1 (0, 0, 0), V2 (0, 0, 0), W are linearly independent, there is a constant
c > 0 such that
¡ ¢
Prc ⊂ F [−δr, δr] × [−δr, δr] × [−δ 2 r2 , δ 2 r2 ] ,
provided that r is small enough. This proves part I.
II. Now we need to show that there is a positive C such that Br ⊂ PrC
for all sufficiently small r; in other words, we need to estimate Carnot–
Carathéodory distance from below. Let γ be an admissible path connecting
the origin and (u, v, τ ). We want to estimate from below the (Euclidean)
length L of γ. It is enough to show that
p
(5.18) CL ≥ max(|u|, |v|, |τ |).
Recalling that we need to deal only with points with |τ | ≥ 2(1 + A)(u2 + v 2 ),
we see that for such points
¯Z ¯
¯ ¯
(5.19) ¯ ω ¯ ≥ 1 |τ | − Aτ 2 ≥ 1 |τ |
¯ ¯ 2 3
α
5.4. Sub-Riemannian Metric Structures 187
Note that admissible curves cannot go vertically, and hence the integrand
in this formula never vanishes.
We will look at this example in various clothing, which will give different
insights into the geometry of sub-Riemannian length structures.
The most important feature of this situation is that every smooth curve
in the base can be lifted to R3 as an admissible path. Once the lift of
one point is fixed, the lifting of the whole curve is unique. Indeed, the
restriction of the differential dp π of the projection to the base to Hp is a
linear isomorphism; hence for a vector field X tangent to the base B, there
is a unique lift of X, that is, a vector field X̃ that lies in the distribution
H and dπ(X̃) = X. From here, it is easy to see that given a smooth path
γ : R → B, and a point p with π(p) = γ(0), there exists a unique lift of
γ, that is, an admissible curve γ̃ with γ̃(0) = p and π(γ̃) = γ. Of course,
all admissible curves come this way. Thus if one wants to connect two
points p, q ∈ R3 by an admissible curves, one can present it by drawing
its projection connecting the projections p0 , q0 of the points to the base.
Notice, however, that if one connects p0 and q0 by some path γ : [0, 1] → B,
its lift γ̃ with γ̃(0) = p will connect p with some point in the fiber containing
q, γ̃(1) ∈ q0 × F , though there is no reason to expect that it will be just q.
This suggests the following simple but useful definition.
connected if the holonomy group at a point (and hence at every point) acts
transitively (verify this assertion!)
This may be somewhat surprising, for the Lie bracket used to define Z
involves derivatives; this is similar to the situation with the curvature tensor
(see Chapter 6, 6.3.3), which is defined using vector fields, but at the end it
happens to depend only on their values at one point. As a matter of fact,
the curvature tensor is a particular case of this construction.
The geometric meaning of the curvature form ω is given by the following
important exercise:
Exercise 5.4.16. Let γ be a loop at p: γ(0) = γ(1) = p. Assume that γ
bounds a simple region Γ, whose orientation agrees with that of γ. Then
the transformation Gγ is a translation z → z + A, where
Z
A= ω.
Γ
Hence we obtain the same conclusion as at the end of the previous section:
the shortest curve connecting p and q projects to the shortest curve bounding
a region of area z.
Note that Theorem 5.5.5 uniquely defines the volume of any measurable
set in a Riemannian manifold. On the other hand, in its proof (including
the proof of the change of variable formula) we only used the two basic
properties mentioned after Definition 5.5.1. Hence these two properties
uniquely determine the volume, and we obtain the following
Theorem 5.5.6. Let V be a function associating to every Riemannian
manifold M a Borel measure over it. Suppose that V is monotone with
respect to the metric and yields the standard Euclidean volume in Rn . Then
V coincides with the Riemannian volume.
measure does not depend on the choice of Riemannian structure (in fact,
both notions can be defined in terms of local coordinates).
Rademacher’s theorem implies that for a Lipschitz map f : M → N the
JacobianR Jac f (x) is defined for almost all x ∈ M . Therefore the Lebesgue
integral M Jac f is defined. This integral is called the volume (or area) of f .
If f is injective, it equals the volume of f (M ) in N . If f covers some parts
of N multiple times, the multiplicity should be taken into account:
Theorem 5.5.8 ([Fe], Theorem 3.2.3). If M and N are n-dimensional
Riemannian manifolds and f : M → N is a Lipschitz map, then
Z Z
Jac f (x) d VolM (x) = #(f −1 (y)) d VolN (y)
M N
where # denotes the cardinality.
The integral on the right-hand side is, of course, the Lebesgue integral of
the function y 7→ #(f −1 (y)) which is always measurable. Since the function
takes only integer values and the value infinity, its integral can be written
as
Z X
#(f −1 (y)) d VolM (y) = k · VolN ({y ∈ N : #(f −1 (y)) = k})
N k∈N∪{∞}
5.5.3. Finslerian volumes. Theorem 5.5.6 tells us that there is only one
reasonable notion of volume for Riemannian manifolds. This is not the case
with Finsler metrics. One can define Finslerian volume in different ways
and obtain essentially different results. Some examples are given below in
this section.
For now, we assume that some Finslerian volume functional is fixed.
That is, every Finsler manifold is equipped with a Borel measure (depending
on its Finsler structure) that we denote by Vol. We require that Vol satisfy
the properties listed after Definition 5.5.1, namely, the Euclidean compat-
ibility and monotonicity with respect to metric. Note that monotonicity
implies that (smooth) isometries preserve the volume.
Let k · k be a norm in Rn . As a particular case of a Finsler manifold, the
normed space (Rn , k · k) carries a Finslerian volume, Volk·k . Let | · | be the
Euclidean norm. Since k · k and | · | are bi-Lipschitz equivalent (Theorem
1.4.11), there exist positive constants c and C such that c|x| ≤ kxk ≤ C|x|
for all x ∈ Rn . The monotonicity of volume then implies that cn mn ≤
Volk·k ≤ C n mn . (To prove this, consider the identity map as a map from
(Rn , k · k) to (Rn , c| · |) and from (Rn , C| · |) to (Rn , k · k).) In particular, the
k · k-volume of a unit cube is finite and positive. We denote this quantity
Volk·k ([0, 1]n ) by ν(k · k).
The volume Volk·k is preserved by parallel translations since they are
isometries of (Rn , k · k). By Lebesgue’s theorem (1.7.5) it follows that the
Volk·k
measure ν(k·k) coincides with the Lebesgue measure mn . Thus
A plan of proof. Having a measure over every normed space allows one
to define the Jacobian of a map from one Finsler manifold to another,
similarly to Definition 5.5.2. Define Finslerian volume by the formula from
Proposition 5.5.11; then the change of variable formula (Theorem 5.5.4)
follows easily. The conditions 2 and 3 imply that a linear nonexpanding map
does not increase the measure. Observe that derivatives of a nonexpanding
maps between Finsler manifolds are nonexpanding linear maps between
their tangent spaces (equipped with norms restricted from the Finslerian
structures). Then the change of variable formula implies that the Finslerian
volume is well-defined (i.e., does not depend on the choice of a coordinate
system) and monotone with respect to metric. ¤
While there are many different natural Finsler volume functionals, they
are nevertheless “not too different”. Namely the ratio of any two of them is
bounded by a constant depending only on dimension:
Theorem 5.5.18. Let Vol and Vol′ be two n-dimensional Euclidean-compa-
tible monotone Finsler volume functionals. Then Vol(Ω) ≤ n3n/2 Vol′ (Ω) for
any measurable set Ω is any Finsler manifold.
This proposition, along with the fact that at least one regular value
exists, allows us to introduce the following
Definition 5.6.4. Let f : M → N be a smooth map. Then deg2 (f ; y),
where y ∈ N is an arbitrary regular value of f , is called the degree modulo 2
of f and denoted deg2 (f ).
Proposition 5.6.5. If smooth maps f1 and f2 from M to N are homotopic,
then deg2 (f1 ) = deg2 (f2 ).
Proof. First, reduce the proposition to the case of no boundary just like
Proposition 5.6.6. Then suppose f is not surjective. If f is smooth,
consider a y ∈ N \ f (M ). Since f −1 (y) = ∅, y is a regular value and
deg2 (f ) = deg2 (f ; y) = 0. In the general case, replace f by a smooth
approximation f1 . If f1 is sufficiently close to f , it is homotopic to f and
still nonsurjective; hence deg2 (f ) = deg2 (f1 ) = 0. ¤
Remark 5.6.8. If M and N are oriented manifolds, one can define the
integer-valued degree of a map f : M → N , denoted deg(f ). Namely, if
y ∈ N is a regular value of f , let
X
deg(f ; y) = ε(x)
x∈f −1 (y)
where (
+1, if dx f is orientation-preserving,
ε(x) =
−1, otherwise.
All properties of deg2 that we formulated hold for deg with obvious modifi-
cations. It is clear that deg2 (f ) = deg(f ) mod 2.
5.6. Besikovitch Inequality 205
Proof. During this proof all distances, Jacobians, etc on I n are taken with
respect to g. For each i = 1, . . . , n define a function fi : I n → R by
fi (x) = min{di , dist(x, Fi0 )}
and let f : I n → Rn be the function whose coordinate functions are fi , i.e.,
f (x) = (f1 (x), . . . , fn (x)). Observe the following trivial facts.
(1) The functions fi are nonexpanding.
(2) fi (x) ∈ [0, di ] for all x ∈ I n .This means that f maps I n to the
parallelotope P = [0, d1 ] × [0, d2 ] × . . . [0, dn ].
(3) fi (x) = 0 if x ∈ Fi0 and fi (x) = di if x ∈ Fi1 . In other words, f
maps each face of I n to the corresponding face of the parallelotope
P.
By Proposition 5.5.10 the first of the above facts implies that Jac f ≤ 1
almost everywhere. By Corollary 5.5.9 it follows that Volg (I n ) ≥ µn (f (I n )).
We will show that f (I n ) = P .
The second of the above facts allows us to consider f as a map from
I n to P . The third one implies that f (∂I n ) ⊂ ∂P ; hence the notion of
degree applies. We will show that the degree of f (modulo 2) as a map
from I n to P equals 1. Consider the linear map A : Rn → Rn defined by
A(x1 , . . . , xn ) = (x1 /d1 , . . . , xn /dn ). Its restriction A|P is a homeomorphism
from P to I n . Hence it is sufficient to prove that A ◦ f has degree 1 as a
map from I n to I n (then apply Proposition 5.6.7). We will prove this by
showing that A ◦ f is homotopic to the identity via a homotopy {ϕt }t∈[0,1]
such that ϕt (∂I n ) ⊂ ∂I n for all t (see Proposition 5.6.6). Let {ϕt } be a linear
homotopy between A◦f and the identity, i.e., ϕt (x) = (1−t)A(f (x)+tx. Let
x ∈ ∂I n and let F be a face of I n to which x belongs. Then y := A(f (x)) ∈ F
since A ◦ f |I n maps each face of I n to itself. It follows immediately that
ϕt (x) = tx + (1 − t)y ∈ F ⊂ ∂I n for all t.
Thus ϕt (∂I n ) ⊂ ∂I n for all t. Therefore deg2 (f ) = deg2 (A ◦ f ) =
deg2 (IdI n ) = 1 as we claimed above. In particular, the image of f is the
206 5. Smooth Length Structures
entire P . Then
n
Y
Volg (I n ) ≥ Vol(f (I n )) = Vol(P ) = di .
i=1
¤
Exercise 5.6.10. Prove the following generalizations of Theorem 5.6.9.
1. Let M be an n-dimensional
Qn Riemannian manifold with ∂M = ∂I n .
Prove that Vol(M ) ≥ i=1 distM (Fi0 , Fi1 ).
2. Let M be an n-dimensional Riemannian manifold and f : ∂M → ∂I n
be
Qn a continuous map with nonzero degree modulo 2. Then Vol(M ) ≥
−1 (Fi0 ), f −1 (Fi1 )).
i=1 distM (f
Exercise 5.6.11. Let M be a Riemannian manifold homeomorphic to
the sphere S 2 . Suppose that there are four points a, b, c, d ∈ M with
the following distances between them: |ab| = |bc| = |cd| = |da| = 1,
|ac| = |bd| = 3/2. Prove that the area of M is at least 1/2.
Hint: Prove that the distance between a shortest path from a to b and
one from c to d is not less than 1/2, and similarly for shortest paths from b
to c and from d to a. Then apply the Besikovitch inequality to each of the
two quadrilaterals into which these four paths divide the sphere.
Exercise 5.6.12. Let M be a Riemannian manifold homeomorphic to the
projective plane RP2 . Suppose that the length of any noncontractible loop
in M is at least 1. Prove that the area of M is at least 1/16.
Hint: Cut M along a shortest noncontractible loop. The resulting space
is homeomorphic to a disc. Divide its boundary into four arcs of equal
lengths and prove that the distance between “opposite” arcs is not less
than 1/4. Then apply the Besikovitch inequality to this disc (which is
homeomorphic to I 2 ).
Remark 5.6.13. In fact, the constant 1/16 above can be replaced by 2/π
(this is known as the theorem of Pu, see [Gro1]). The constant 2/π is
optimal and is achieved for the metric on RP2 space obtained from the
sphere of radius 1/π (by identifying opposite points).
Exercise 5.6.14. Let M be a Riemannian manifold homeomorphic to the
torus T 2 = S 1 × S 1 . Suppose that the length of any noncontractible loop in
M is at least 1. Prove that the area of M is at least 1/4.
Remark √ 5.6.15. The optimal constant in the statement of the above
exercise is 3/2 (Loewner’s theorem; see [Gro1]). Can you find an example
for which this is achieved?
Finally, let us mention that there is no “inverse Besikovitch inequality”
(i.e., no similar upper bound for the volume):
5.6. Besikovitch Inequality 207
Exercise 5.6.16. Prove that for every C > 0 and every integer n ≥ 2 there
exists a Riemannian structure g in the n-cube I n such that the Hausdorff
distance (in (I n , distg )) between Fi0 and Fi1 is less than 1 for all i, but
Volg (I n ) > C.
or, equivalently,
Vol(M, g)
c1 (M ) = inf
g sys1 (M, g)n
where the infimum is taken over all Riemannian metrics g on M . One then
asks, for a given topological type of M , what is the value of this constant,
or at least whether it is positive or not.
Though the precise values of c1 (M ) are currently known only in a few
simplest cases (all of which are two-dimensional), a vast class of manifolds
M with c1 (M ) > 0 has been found (see [Gro1]).
The notion of systole can be generalized to higher dimensions in a
variety of ways. One possible approach is that the k-dimensional systole
of a Riemannian manifold M is the infimum of k-volumes of homologically
nontrivial closed k-dimensional “films” in M . Here one has to specify what
is meant by a “film”. It is not good to consider only k-submanifolds as
films; one of the reasons is that not all homology classes can be represented
by submanifolds. Instead, one takes the infimum over all singular Lipschitz
chains.
Let us explain this in more detail. Let τ be a k-dimensional Lipschitz
singular simplex, that is, a Lipschitz map of the standard k-simplex into M .
For such a simplex, the k-dimensional volume Vol(τ ) is well-defined (see
208 5. Smooth Length Structures
Section 5.5.2). This allows one to introduce the notion of volume for k-
PN
dimensional Lipschitz chains. Namely for c = , where xi ∈ R
i=1 xi τiP
N
and τi are Lipschitz singular simplices, define Vol(c) = i=1 |xi | Vol(τi ).
The boundary of a chain, cycles and homology groups are defined as usual.
Finally one defines
sysk (M, g) = inf{Vol(c) : c is a k-dimensional cycle not homological to 0}
and
Vol(M, g)
ck (M ) = inf ,
g sys (M, g)n/k
k
and asks whether ck (M ) is positive or not.
It is a bit surprising that the answer is absolutely different for k = 1
and k > 1. While c1 (M ) > 0 in many cases, it has been proved recently
([KS] and [Bab]) that one always has sysk (M ) = 0 for 1 < k < dim M .
This property is called systolic softness. Actually in [Bab] a more general
fact on so-called inter-systoles is proved; furthermore, M is allowed to be a
polyhedral space instead of a manifold.
Chapter 6
Curvature
of Riemannian Metrics
The main conclusion of this chapter is the fact that a Riemannian manifold of
sectional curvature bounded by k is a length space with the same curvature
bound in the sense of Chapter 4. Hence the reader who is ready to take this
statement on faith (or who is familiar with Riemannian geometry) can omit
this chapter. It is included mainly to make this book more self-contained.
In our introduction to Riemannian geometry we basically stop where
substantial Riemannian geometry begins. There are many textbooks con-
taining various accounts of methods of differential and Riemannian geome-
try. In this course we regard Riemannian manifolds (as well as other length
spaces whose definitions rely on calculus of variations) just as one of the
important (or even leading) sources of examples. From this viewpoint, all
differential methods can retire after they produce certain basic informa-
tion to feed into the synthetic machinery. A shortest path leading from
infinitesimal definitions to local metric properties is surprisingly short, and
we intentionally make it somewhat longer to provide better motivations.
Basically all we need at the end is to convert a Riemannian metric to nor-
mal coordinates, and then prove the (local) Cartan-Alexandrov-Toponogov
Comparison Theorems 6.5.6.
For the sake of simplicity of exposition we restrict our introduction to
Riemannian geometry mainly to the case of two-dimensional manifolds; this
also allows us to give explicit formulas avoiding cumbersome indices and
to eliminate linear algebra. Higher-dimensional generalizations are mostly
straightforward and left to the reader (as an important exercise). We point
209
210 6. Curvature of Riemannian Metrics
reader looking for geometric intuition and applications can take the results
of these sections for the definition of sectional curvature and completely skip
covariant differentiation. The latter is, however, needed for computations
in actual examples, as well as to show the correctness of such “synthetic”
definitions of Riemannian curvature.
Denote Tε = dγ dTε
dt and note that dε (t, ε) = V (t). Now differentiating under
ε
but for our “scalar product” h, i this formula is just incorrect! Recall that
we use h, i to denote a bilinear form Qp (·, ·) on (tangent) vectors. The
coefficients of this form depend on p, and hence one has to use the Chain
Rule to differentiate it. There is an invariant and very convenient way
of doing this by introducing covariant derivatives (see Section 6.2). As a
matter of fact, already the expressions V̇ (t), Ṫ (t) themselves depend on
the choice of a coordinate system! Covariant derivatives are designed to
substitute this “noninvariant” coordinate-wise derivative by an “invariant”
(coordinate-independent) operation.
We will proceed here with a coordinate computation now. A geometrical
meaning of what is going on here will become clear in Section 6.2.
Recall that
Substitute the last expression in (6.1) and take into account that Eε =
dE ∂E ∂E
dε = ∂x ẋ + ∂y ẏ (the Chain Rule) and simultaneously for Fε , Gε . After
some intermediate calculations we get:
Z bD E b DZ E
(6.2) 0 = T (t), V̇ (t) dt = hT, V i |ba
− ( Ṫ , V + mA + nB) dt =
a a
Z bµ 2 2
¶
d x d y d2 x d2 y
= hT, V i |ba − m(A + E 2 + 2 F ) + n(B + F 2 + G 2 ) dt,
a dt dt dt dt
where
1 ∂E ∂E ∂F 1 ∂G
A = − ẋ2 − ẋẏ − ẏ 2 ( − ),
2 ∂x ∂y ∂y 2 ∂x
∂F 1 ∂E ∂G 1 2 ∂G
B = −ẋ2 ( + ) − ẋẏ − ẏ .
∂x 2 ∂y ∂x 2 ∂y
Note that hT, V i |ba = 0 since we supposed that all curves γε have the
same endpoints. Recall that this identity is true for all choices of functions
m, n, with the only restriction that they vanish at a and b. Using a standard
analytical argument one shows that this implies that both expressions
A + E ẍ + F ÿ and B + F ẍ + Gÿ
must be identically zero! Indeed, if for instance the first expression was
nonzero at some point t0 , one chooses n = 0 and a nonnegative function m
such that m(t0 ) = 1 and m vanishes outside a small neighborhood of t0 . It
is easy to see that (6.2) is not satisfied for this choice of m, n.
Hence we see that our path γ satisfies just the system of two nonlinear
second-order differential equations (5.10) and (5.11) that we used to define
geodesics:
1 ∂E ∂E ¡ ∂F 1 ∂G ¢
E ẍ + F ÿ = ẋ2 + ẋẏ + ẏ 2 − ,
2 ∂x ∂y ∂y 2 ∂x
¡ ∂F 1 ∂E ¢ ∂G 1 2 ∂G
F ẍ + Gÿ = ẋ2 − + ẋẏ + ẏ .
∂x 2 ∂y ∂x 2 ∂y
Exercise 6.1.1. Repeat the same computation for a conformal metric in Ω
(given by a function λ(p)) for a shortest curve γ parameterized by Euclidean
arc length. Show that the Euclidean
¿ curvature k(t)À of γ satisfies the relation
∆λ(γ(t))
k(t) = T (t), .
λ(γ(t))
This equation says that a shortest path must bend towards the direction
where λ increases. Although this may sound like nonsense at first glance
(for a shortest path should try to go through small values of λ), the following
consideration relieves this feeling. Imagine that you walk around a swamp
(with λ being very big at the center of the swamp). Suppose that the swamp
214 6. Curvature of Riemannian Metrics
is on your left. Notice that you also keep turning to your left, that is towards
the swamp.
Exercise 6.2.2. Two vector fields V = (v1 (x, y), v2 (x, y)) and W =
(w1 (x, y), w2 (x, y)) on R2 . commute if and only if their components v1 ,
v2 , w1 , w2 satisfy certain differential relation. Find this relation and prove
this.
Hint: You may use the previous exercise applied to coordinate functions.
µ µ ¶ µ ¶¶
DV 1
dE dF 1 dF dG
(6.4) = v̇1 + v1 + v2 , v̇2 + v1 + v2
dt 2dt dt 2 dt dt
¿ À
DV
(recall that E = G = 1, F = 0 at p!). Then of course ,W =
D E dt
V̇ , W + S(V, W ), and hence we get a nice product rule:
¿ À ¿ À
d D D
hV, W i = V, W + V, W .
dt dt dt
Z b¿ À
d
(6.5) dr(T (t)), dr((V (t)) dt = 0.
a dt
Now we can use integration by parts, for we can apply the product rule
to the Euclidean scalar product! We get
Z b
d b d
0 = L(ε)|ε=0 = hdr(T ), dr(V )i |a − hdr(T (t)), dr(V (t))i dt.
dε a dt
Notice that not every vector in R3 can be represented as dr(V (t)) for
some V : dr(V (t)) has to be tangent to our surface. Now arguing as in
d
Subsection 6.1.1 one concludes that dt dr(T ) has to be orthogonal to all
vectors tangent to the surface at r(γ(t)). In other words, the acceleration
of a shortest path parameterized by arc length remains orthogonal to the
surface at all times. If thinking of a geodesic as a free motion of a particle
confined to the surfaces, one can recognize a well-known mechanical principle
that the centrifugal acceleration is orthogonal to the surface.
The only disadvantage of this argument is that it does not have
much meaning in terms of intrinsic geometry of the surface. Imagine two-
dimensional creatures living in the surface. Vectors in R3 other than those
tangent to the surface have no meaning in their physical world. There were
two objects that stuck out of two dimensions in our argument, namely,
d d
dt dr(T (t)) and dt dr(V (t)). Both of them are of the same nature: they came
from differentiating a tangent vector field, and this derivative might have
a nontrivial component orthogonal to the surface. Notice that both vec-
tors come into our formulas multiplied by a vector tangent to the surface,
and hence they can be replaced by their orthogonal projections to the plane
tangent to the surface at r(γ(t)). This suggests the following strategy of
defining an “intrinsic differentiation” of tangent vector fields: begin with
220 6. Curvature of Riemannian Metrics
the usual derivative in the ambient space, and then get rid of its component
orthogonal to the surfaces by projecting it to the tangent plane:
D d
V (t) = dr−1 Proj dr(Tγ(t) Ω) dr(V (t)).
dt dt
If one wants to forget about the embedding r and think of Ω as a subset of
R3 , this formula takes a simpler form:
D dV (t)
V (t) = Proj Tγ(t) Ω .
dt dt
D
It is not clear a priori that, if defined this way, dt would not change
3
if we consider another isometric embedding Ω into R , that is, another
embedding that induces the same Riemannian metric on Ω. To prove that
D
dt depends only on the intrinsic geometry (Riemannian metric) of Ω, one
D
can easily check that dt satisfies axioms (1)–(6) of covariant derivative, and
then apply Lemma 6.2.4 (do this as an exercise!).
6.2.4. Parallel transport. Though our exposition will not rely on the
notion of a parallel transport of vectors, we find it relevant to give its
definition and list its main properties.
When we were introducing the concept of a tangent vector, we empha-
sized that there is no way to define the notion of “the same direction at
different points” in a consistent way. However, if we consider a region Ω
with a Riemannian metric, a choice of a path γ between two points p = γ(a)
and q = γ(b) induces a natural correspondence between Tp Ω and Tq Ω. It is
enough to assume that γ is piecewise smooth.
We say that a vector field along γ is parallel if it satisfies the following
first-order linear differential equation: DV (t)/dt = 0. Every initial value
V (a) = W uniquely determines a parallel vector field V , and thus one can
6.3. Geodesic and Gaussian Curvatures 221
This formula is called the first variation formula. Now suppose that curves
d
γε have the same endpoints. Then hV, T i |ba = 0. In this case dε L(ε)|ε=0 = 0
for any choice of V (since γ0 is a shortest path). Arguing as in Subsection
6.1.1, we obtain a very simple differential equation:
D
(6.8) T (t) = 0, or in alternative notation ∇T T = 0.
dt
Definition 6.3.1. A path γ is called a geodesic if its velocity vector field T
satisfies the equation (6.8).
This is very similar to the Euclidean case: a straight line is a curve traced
by a motion with constant velocity. Here for a velocity to be “constant”
means that its covariant derivative is zero.
Owing to invariant notations, equation (6.8) makes sense in any dimen-
sion.
Note that every geodesic
D is® parameterized proportionally to arc-length
d
(because dt hT, T i = 2 dt T, T = 0).
Exercise 6.3.2. Prove that in dimension two Definitions 5.2.1 and 6.3.1 are
equivalent.
Exercise 6.3.3. Re-prove the results of subsections 5.2.2 and 5.2.1 using
the new definition of geodesics.
move), we obtain
dρ
= hV, T i ,
dτ
where T is the tangent vector of γτ at its end σ(τ ). In other words, the
derivative of the distance from p to a point moving at unit speed is equal to
the negative cosine of the angle between the velocity vector of the moving
point and the direction from the moving point towards p (since the latter is
just −T ).
Exercise 6.3.4. Let σ(t) be a smooth closed curve, and let σ(t0 ) be the
point of σ closest to p. Show that every shortest path connecting p and
σ(t0 ) is orthogonal to (the velocity of) σ at t0 .
Let us now look at the case when a curve γ is not necessarily a geodesic.
D
Notice that the vector ∇T T = dt T (t) is always orthogonal to T (t). Indeed,
¿ À
d D
0= hT (t), T (t)i = 2 T (t), T (t) .
dt dt
D
The number | dt T (t)| = kg (t) is called the geodesic curvature of γ; in
dimension 2 one can assign it a sign + or − by choosing an orientation.
D
In higher dimensions one often calls S(t) = dt T (t) the curvature vector. As
a matter of fact, this is nothing but the value of the second fundamental
form of our curve (regarded as a one-dimensional submanifold) applied to
T : compare this definition with formula (6.6). This give a new geometric
insight into the meaning of the second fundamental form: it describes how
the metric of a submanifold changes under variations of the submanifold.
For a variation with fixed endpoints (more generally, if the variation field is
orthogonal to the curve at it ends), the double substitution term hV, T i |ba
disappears, and (6.7) takes the following form:
Z b
d
(6.10) L(ε)|ε=0 = h−S(t), V (t)i dt.
dε a
Exercise 6.3.6. Carry out these computations for curves in R2 and R3 with
the Euclidean metric. Compare S with a well-known notion of curvature in
differential geometry. Study the length of equidistants of a closed planar
curve by shrinking it towards its inward normal.
6.3.3. Curvature tensor. Choose a coordinate system (x, y), and let
X and Y be (as usual) the coordinate fields. (The reader who follows
higher-dimensional generalizations can take two first coordinate fields of
a coordinate system (x, y, · · · , ).) Then of course X and Y commute as
differential operators on functions: XY f = Y Xf for every smooth function
D
f . However, unlike the Euclidean case, in general the operators ∇X = dx
D
and ∇Y = dy do not commute: for a vector field Z, ∇X ∇Y Z 6= ∇Y ∇X Z.
The difference R(X, Y )Z = ∇Y ∇X Z − ∇X ∇Y Z is called the curvature
operator. A real-valued expression hR(X, Y )Z, W i, which depends on four
vector arguments X, Y, Z, W , is called the curvature tensor. The definition
of R suggests that it depends on four vector fields (and the first two of them
are supposed to be coordinate vector fields for some coordinate system). It
is clear that R(X, Y )Z is a linear function in each of its three arguments.
The following exercises shows that R(X, Y )Z at a point p depends only on
the values of X, Y , and Z at p ! The proofs of the exercises consist of a
straightforward computation based on axioms of the covariant derivative:
Exercise 6.3.7. Show that, for a smooth function f :
(1) ∇X ∇Y (f Z) − ∇Y ∇X (f Z) = f · (∇X ∇Y Z − ∇Y ∇X Z),
(2) ∇X ∇(f Y ) Z − ∇(f Y ) ∇X = f · (∇X ∇Y Z − ∇Y ∇X Z).
Exercise 6.3.8. Show that:
(1) R(X, Y )Z = −R(Y, X)Z; in particular, R(X, X)Z = 0,
(2) hR(X, Y )Z, W i = hR(Z, W )X, Y i.
Exercise 6.3.9. Combining the two previous exercises, show that R(X, Y )Z
= 0 at p if at least one of the fields X, Y, Z vanishes at p. Conclude that
R(X, Y )Z at p depends only on the values of X, Y , and Z at p!
where the last identity follows from axiom (6) of covariant differentiation:
∇X Y = ∇Y X for commuting (coordinate) vector fields X, Y . We will denote
D D D2
∇X ∇X = dx dx by dx2 .
Thus we get the following Jacobi equation, which is a second-order linear
differential equation for Y :
D2
(6.12) Y = −R(X, Y )X.
dx2
Notice that Y is nothing but a variational vector field for a family
of geodesics (the x-lines); it is of no importance that these lines form a
nondegenerate coordinate system, and the same argument can be repeated
verbatim for a general variation. Let us give the corresponding definition:
Definition 6.4.1. A vector field Y (t) along a geodesic γ(t) is called a Jacobi
vector field if it satisfies equation (6.12).
Exercise 6.4.2. Show that every Jacobi field along γ can be represented
as a variational vector field by including γ into a family of geodesics.
Hint: Since a Jacobi vector field Y satisfies a second-order linear differ-
ential equation, it is uniquely determined by its Cauchy data, that is, its
D
value Y (t0 ) and its derivative dt Y (t0 ) at a point t0 . On the other hand, it
is easy to include γ into a geodesic variation whose variation vector field Y1
has the same value and the same derivative at t0 as Y . This implies that
X = Y , for Y has to satisfy the same equation (6.12), since it is a variation
field of a family of geodesics.
D D
Let us notice now that actually dt N = 0, for dt N is orthogonal to both
T and N , and we are dealing with the two-dimensional case! Hence in the
two-dimensional case equation (6.14) turns into
d2 g
(6.15) = −gK.
dt2
The reason why we postponed this observation till now and derived the more
complicated equation (6.14) is explained in the following remark:
Important remark. Now we have arrived at the point where there
is an essential difference between two-dimensional surfaces and the higher-
dimensional case. In the two-dimensional case the direction of Y is deter-
mined by the fact that Y is orthogonal to T . Hence we were able to transform
the Jacobi equation (6.12) into one scalar equation for the magnitude of Y .
In higher dimensions the Jacobi equation (6.12) involves both the magnitude
and the direction of Y . The higher-dimensional situation is complicated by
the following phenomenon: geodesics from our family “twist around γ”. It
is important to notice that the equation for the length of Y that we obtained
for the two-dimensional case is incorrect in higher dimensions! However, the
main geometrical corollary, the Rauch Comparison Theorem 6.5.1, is still
true. One of the proofs consists of a computation based on decomposing y
with respect to a basis of parallel vector fields along γ, that is, such vector
fields that their covariant derivatives along γ are zero. One such vector field
along γ is T , and in our two-dimensional argument N just happened to be
another parallel vector field along γ. The next section gives an alternative
approach via equidistant variations. This approach can be easily generalized
to higher dimensions, and it can be used instead of the Rauch Comparison
Theorem to prove the Cartan-Alexandrov-Toponogov Comparison Theorem.
Note also that many geometric corollaries can be extracted directly from the
d2 g
inequality ≥ −gK implied by (6.14).
dx2
Equidistant variations. The previous section described the Gaussian cur-
vature as a quantitative way of measuring how a family of geodesics diverges.
There is an alternative approach, which avoids using the Jacobi equation.
It has certain advantages for Riemannian geometry (and in particular it
avoids the complication in higher dimensions that we ran into when deriv-
ing (6.15)). On the other hand, this approach can hardly be used for length
spaces other than manifolds.
The idea of this approach is that, instead of studying a bunch of
geodesics, we will concentrate our attention on the family of lines orthogonal
to them. In a normal coordinate system this is just the family of the y-lines.
The distance between two nearby geodesic lines is approximately the length
of the segment of the orthogonal line enclosed between them. This suggests
6.4. Geometric Meaning of Gaussian Curvature 231
that we study how the length along an orthogonal line changes within its
family. A convenient way of doing so is to consider this family of orthogonal
lines as a variation of a curve (in higher dimensions one studies variations
of hypersurfaces).
Let γε be a family of curves whose variation vector field V is a unit vector
field orthogonal to the curves of the family: hV, V i = 1, and hV, T i = 0
d
(where as usual T = dt γ, and we drop the arguments t, ε to simplify
notation). Such variation is said to be equidistant. The reason is explained
in the following exercise.
Exercise 6.4.4. Prove that, for all sufficiently small ε and for any fixed t0 ,
inf d(γε (t0 ), γ0 (t))} = ε.
t
In other words, points of γε lie at the same distance ε from (the image of)
γ0 .
Hint: Use the fact that the lines σ(ε) = γε (t) are geodesics (see below),
and argue as in the proof of the Gauss Lemma.
One should not think of the curves γε as geodesics, but rather as lines
orthogonal to a family of geodesics. Indeed, statement (ii) of Lemma 6.4.6
will imply that the orthogonal curves σ(ε) = γε (t) are geodesics.
Let us assume that the line γ = γ0 is parameterized by arc length.
According to the formula of the first variation (6.10),
Z b Z b
d
L(γε , a, b) = h−S, V i dt = −kg (t) dt,
dε a a
where S = ∇T T , and kg = h∇T T, V i is the geodesic curvature of γ at t (and
its sign is chosen with respect to the orientation prescribed by V ). Hence
a curve in our family shrinks (if kg > 0) or expands (kg < 0) at a rate
proportional to its curvature.
Exercise 6.4.5. Show that for a variation with a variation field f V , where
V is a unit normal to the curve and f is a scalar function,
Z b
d
L(γε , a, b) = −f (t)kg (t) dt.
dε a
Proof. (i) This is true just because V is a unit vector field (Lemma 6.4.3).
(ii) Since hV, T i = 0, we have
d
0= hV, T i = h∇V V, T i + hV, ∇V T i
dε
1d
= h∇V V, T i + hV, ∇T V i = h∇V V, T i + hV, V i = h∇V V, T i .
2 dt
Together with (i) this implies that the vector ∇V V is zero, for it is orthogonal
to both V and T .
(iii) Differentiating h∇T V, V i = 0 (which follows from (i)), we get
d
0= h∇T V, V i = h∇T ∇V T, V i + h∇T V, ∇T V i .
dt
Together with ∇T V = ∇V T (since T and V are commuting vector fields)
this implies (iii).
(iv) Since |kg | = |∇T T |, the equality h∇T T, ∇T T i = kg2 is trivial.
Differentiating hV, T i = 0, we get
d
0= hT, V i = h∇T T, V i + hT, ∇T V i .
dt
Since V and T are unit vectors, ∇T T is proportional to V , and ∇T V is
proportional to T , this implies that |∇T T | = |∇T V |, which proves (iv). ¤
Remark. Although it may seem that the argument for (ii) uses the fact
that dim(Ω) = 2 in the same way as in the computation used to derive
(6.15), it is not the case. We leave as an exercise to show that the integral
curves of a variation field V of an equidistant variation of a hypersurface are
geodesics. To do this, one just shows that ∇V V is orthogonal to V and to
all vectors tangent to the hypersurfaces. This is a reason why this approach
is sometimes more convenient in higher dimensions.
Now we want to look at the second derivative of length; that is, we want
to know how kg changes as we change the parameter ε. We have
dkg d
= h∇T T, V i = h∇V ∇T T, V i + h∇T T, ∇V V i = h∇V ∇T T, V i ,
dε dε
since ∇V V = 0 by Lemma 6.4.6, (ii).
Now we want to change the order of differentiations in h∇V ∇T T, V i,
and thus the curvature operator shows up. We get
dkg
= h∇T ∇V T, V i + hR(T, V )T, V i .
dε
6.4. Geometric Meaning of Gaussian Curvature 233
d2 g(t)
(6.16) = −K(t)g(t),
dt2
where K(t) = K(γ(t)) is the Gaussian curvature along γ. Recall that this
equation is valid only as long as g(t) does not vanish; vanishing of g(t)
geometrically means that nearby geodesics from the family meet γ (at least
up to the second order).
For a normal coordinate system centered at p we have g(0) = G(0, y) = 0,
for the coordinate vector Y vanishes at p. Notice that dg/dt(0) = 1. This
is a more delicate statement: it follows from Lemma 5.1.13 (two x-lines
emanating from p diverge in the first order at the same rate as two Euclidean
rays emanating from one point and forming the same angle). Hence, by
(6.16), d2 g/dt2 (0) = 0 and therefore g(t) = t + o(t2 ). Plugging this again in
(6.16), we get
d2 g d3 g
(t) = −Kt + o(t2 ) and hence (0) = −K.
dt2 dt3
6.4. Geometric Meaning of Gaussian Curvature 235
Let us notice again that the Gaussian curvature measures just the third-
order distortion of metric.
The construction used to prove this lemma is very useful, and it will
be applied (with minor modifications) in this chapter several times! Notice
also that expp does not distort distances from p, and hence it maps spheres
centered at the origin of Tp Ω to spheres centered at p.
surfaces that are locally isometric to the Euclidean plane, a sphere, and a
hyperbolic plane. Indeed, if k = 0, the Jacobi equation (6.16) tells us that
d2
dt2
g(t) = 0, and hence g(t) is a linear function. Indeed, Euclidean geodesics
are straight lines, and the distance between straight lines is a linear function
(as long as it does not vanish!) If k > 0, one explicitly √ solves the Jacobi
√
equation (6.16), obtaining its general solution g(t) =√A sin k t + B cos k t.
√ k t. It is easy to check
In particular, if g(0) = 0, ġ(0) = 1, we get g(t) = sin
that if one rotates a meridian of a sphere of radius k around its pole at the
unit angular speed, the √speed of a point of this meridian lying at a distance
t from the pole is sin K √ t; in other words, nearby meridians diverge at a
speed proportional to sin√ k t. If k < √ 0, the general solution of equation
(6.16) is g(t) = A sinh −k t + B cosh −k t. This formula describes the
divergence of a bunch of lines in the hyperbolic plane of curvature k. Let us
formalize these considerations:
We consider the case k = 0; the two other cases are similar. The
argument is very similar to the proof of Lemma 6.4.12. Choose a normal
coordinate system (x, y) centered at p. Let γ(t)p= (t, y) be a coordinate
line emanating from p. Denote, as usual, g(t) = G(t, y). As in the proof
of Lemma 6.4.12, g(0) = 0 and dg dt (0) = 1. Solving the Jacobi equation
(6.16) subject to the initial conditions g(0) = 0, g ′ (0) = 1, we immediately
get g(t) = t, and hence G(x, y) = y 2 . Comparing these expressions for
metric coefficients E, F, G with the metric coefficients of the Euclidean
metric in polar coordinates (r, ρ) (formula (5.7)), we see that the map
(x, y) → (r = x, ρ = y) is an isometry to a neighborhood of the origin
in the Euclidean plane in polar coordinates. ¤
Exercise 6.4.14. Using formulas (5.8) and (5.9), repeat the same argument
to show that a surface of constant curvature is locally isometric to a sphere
or a hyperbolic plane.
Remark. Notice that we proved that the Gaussian curvature of the hyper-
bolic plane of curvature k is k; moreover, our argument also implies that the
hyperbolic planes are homogeneous: notice that we constructed an isometry
mapping any given point in a region of constant negative curvature to the
origin of the polar coordinates used in formula (5.9)!
6.5. Comparison Theorems 237
Proof. Let g1 (t) = |Y1 (t)|, g0 (t) = |Y0 (t)|. Then g1 (t) satisfies the equation
d2 g1
(t) = −K1 (t)g1 (t) subject to g1 (0) = 0, ġ1 (0) = 1,
dt2
and g0 (t) satisfies the similar equation.
We want to prove that g1 (t) ≤ g0 (t) for all 0 ≤ t ≤ T .
The idea is to consider a function ϕ(t) = gg10 (t) (t)
and prove that it
is (non-strictly) monotone increasing on the interval ]0, T [. Then, since
limt→0 ϕ(t) = 1 by the L’Hopital rule, it will follow that ϕ(t) ≥ 1 and hence
g0 (t) ≥ g1 .
To prove the monotonicity of ϕ, it suffices to verify that ϕ̇(t) ≥ 0 for
all t. We have
ġ0 (t)g1 (t) − g0 (t)ġ1 (t)
ϕ̇(t) = .
g1 (t)2
Denote the numerator of the last formula by ψ(t). Since the denominator
g1 (t)2 is positive, we have to prove that ψ(t) ≥ 0. Observe that ψ(0) = 0
because g0 (0) = g1 (0) = 0. So again it suffices to prove that ψ̇(t) ≥ 0 for
all t. From the equations for g0 and g1 one gets
ψ̇ = g̈0 g1 − g0 g̈1 = (K1 − K0 )g0 g1 .
Thus ψ̇ > 0 wherever g0 ≥ 0 (recall that g1 > 0 on ]0, T [ by assumption of
the Theorem).
Thus, it remains to show that g0 does not change the sign on ]0, T [.
Suppose the contrary, and let t0 ∈ ]0, T [ be the first point where g0 vanishes.
Then restrict the above argument to the interval ]0, t0 ]: since g0 (t) ≥ 0 for
all t ∈ ]0, t0 ], one has ψ̇(t) ≥ 0 for all t ∈ ]0, t0 ], therefore g0 ≥ g1 on
238 6. Curvature of Riemannian Metrics
Exercise 6.5.3. Use Exercise 5.1.19 to show that the length of a circle of
radius r centered at p is equal to πr2 (1 − 61 K(p)r2 + o(r2 )).
The proof of Lemma 6.4.13 was based on the observation that, for a
region Ω of constant curvature k, σ happens to be an isometry. We will
generalize this as the following property of σ, which perhaps best explains
the geometric meaning of Gaussian curvature:
Theorem 6.5.4. If the Gaussian curvature of Ω satisfies k ≤ K(s) for all
s ∈ Ω, then σ is a nonexpanding map in some neighborhood of q; that is,
there exists an ε > 0 such that
d(σ(s1 ), σ(s2 )) ≤ d(s1 , s2 ) for all s1 , s2 ∈ Bε (q).
If the Gaussian curvature of Ω satisfies k ≥ K(s) for all s ∈ Ω, then σ
is a noncontracting map in some neighborhood of q; that is,
d(σ(s1 ), σ(s2 )) ≥ d(s1 , s2 ) for all s1 , s2 ∈ Bε (q).
Most of our discussions in this course are about relations between various
properties and characteristics of a metric space. In other words, we study one
metric space at a time. However it may be useful to consider every particular
space as a representative of a class of similar objects, which could be called
the “space of metric spaces”. There is a similar idea in the foundation of the
mathematical analysis: instead of talking about a single number, one studies
the entire real line. This brings new notions like continuity, derivatives,
etc., and they appear to be powerful tools that give new information about
the original objects, numbers. For example, a common method of proving
inequalities is finding maxima and minima, and these notions make no sense
without considering a set of numbers rather than a single number.
Another example of using this “global” approach is given by the theory
of convex sets in Rn . Introducing Hausdorff distance (discussed in Sub-
section 7.3.1 below) turns the set of all compact convex sets into a metric
space. This opens a way to apply “analytic” techniques to convex sets just
like numbers. For example, one can use maxima and minima because the
space of compact convex sets is boundedly compact just like R (cf. Theo-
rem 7.3.8). Also, a simple argument shows that polyhedra are dense in this
space, and this allows us to extend certain statements “by continuity” from
polyhedra to arbitrary convex sets.
In this chapter we extend this approach further and introduce a distance
between abstract metric spaces. In fact, we will define several distances
suitable for different purposes. Let us make some remarks in advance.
First, in most cases the distance itself is not essential; what matters is the
topology that it defines. Concerning the “space of metric spaces”, we will
mainly study converging sequences and similar notions, not exact numerical
241
242 7. Space of Metric Spaces
values of the distance. Second, to make use of the topological structure one
usually needs a collection of quantities that are continuous, properties that
are preserved through passing to the limit, etc. The finer the topology is, the
more such quantities and properties exist. However, the topology being finer
means that there are fewer converging sequences and fewer compact sets,
and this makes the topology less useful. For this reason different topologies
may be needed. This is illustrated by functional analysis where every natural
class of functions comes with its own topology—recall C 0 , C 1 , L1 , L2 and
so on.
To ignite curiosity, we begin with a mathematical fairy tale, or a science
fiction story—for this preparation level. It was a famous open problem
whether every group of polynomial growth is virtually nilpotent (since this
is a fairy tale, we do not need to define either of the notions). Here is an
idea of Gromov’s solution of this problem. Gromov suggested considering
a sequence of metric spaces (G, hd) where G is our finitely generated group
with a word metric d, and h is a positive number tending to zero. Using
his criterion for pre-compactness of certain spaces of metric spaces, Gromov
specified a subsequence converging to some length space, along with an
action of the group on this space (a reader may think of the example (Z2 , hd),
where d is the word metric for the standard choice of two generators,
as a sequence of finer and finer grids converging to R2 with the norm
||(x, y)|| = |x| + |y|). The limit space happened to be a Finsler manifold.
Thus Gromov was able to reduce the conjecture to the case of an isometry
group of a manifold, and in this case the answer was to be affirmative.
7.1. Examples
We start with several examples, without formal definitions and proofs. Not
all these examples are covered by notions discussed later in this course; some
of them will be used for motivations and comparisons only.
PSfrag replacements
p
Figure 7.1: Spaces with dilated metrics converge to the tangent cone.
asymptotic cone that does this “near infinity”. Given a (closed) convex set
X ⊂ Rn and a point p ∈ X, the asymptotic cone is the limit of homothetic
sets λX as λ → 0 (the homotheties are centered at p). As in the case of
tangent cone, there is no problem with a definition of the limit. T Since {λX}
is a nested family of sets, the “limit” is just their intersection, λ>0 λX (see
Figure 7.2). It is easy to see that the asymptotic cone is just the union of
all rays emanating from p and contained in X. The shape of the asymptotic
cone does not depend on the choice of p in the sense that asymptotic cones
with different choices for p are parallel translations of one another.
PSfrag replacements
p p
Figure 7.2: Spaces with dilated metrics converge to the asymptotic cone.
Loosely speaking, the asymptotic cone is what one sees observing the
set from far away (assuming that moving the observation point stretches
the visual image homothetically). A similar association applies to tangent
cones: the tangent cone at p is what an observer sees looking at p under
a microscope. The surface of the Earth looks planar to us; indeed, the
tangent cone of the sphere is a plane (more precisely, that of a solid ball is
a half-space).
We will extend the above definition of the asymptotic cone to general
metric spaces in the next chapter (Section 8.2).
PSfrag replacements A
A
B
The reader already familiar with Hausdorff distance may notice that it
provides some way to formalize the above examples: the distance between
these sets is small because each of them is contained in a small neighborhood
of the other in R3 . However, if we consider two surfaces as length spaces
with close intrinsic metrics, and it is not at all clear that these metrics can
be realized by close embedding (even if they can be realized by embeddings
248 7. Space of Metric Spaces
Let M(X) denote the set of closed subsets of X equipped with Hausdorff
distance. The above proposition tells us that the M(X) is a metric space.
Moreover every element of the quotient 2X/dH can be represented by a closed
set and therefore 2X/dH is naturally identified with M(X).
7.3. Gromov–Hausdorff Distance 253
Proof. Let {Sn }∞ n=1 be a Cauchy sequence in M(X). Let S denote the set
of all points x ∈ X such that for any neighborhood U of x one has U ∩Sn 6= ∅
for infinitely many n. We will prove that Sn → S. Fix ε > 0 and let n0
be such that dH (Sn , Sm ) < ε for all m, n ≥ n0 . It suffices to show that
dH (S, Sn ) < 2ε for any n ≥ n0 .
1. dist(x, Sn ) < 2ε for every x ∈ S. There exists an m ≥ n0 such
that Bε (x) ∩ Sm 6= ∅. In other words, there is a point y ∈ Sm such that
|xy| < ε. Since dH (Sm , Sn ) < ε, one has dist(y, Sn ) < ε and therefore
dist(x, Sn ) ≤ |xy| + dist(y, Sn ) < 2ε.
2. dist(x, S) < 2ε for every x ∈ Sn . Let n1 = n and for every integer
k > 1 chose an index nk such that nk > nk+1 and dH (Sp , Sq ) < ε/2k for
all p, q ≥ nk . Then define a sequence of points {xk }, where xk ∈ Snk , as
follows. Let x1 = x, and xk+1 be a point of Snk+1 such that |xk xk+1 | < ε/2k
for all k. Such a point can be found because dH (Snk , Snk+1 ) < ε/2k . Since
P ∞
k=1 |xk xk+1 | < 2ε < ∞, the sequence {xk } is a Cauchy sequence
P and hence
it converges to a point y ∈ X. Then |xy| = lim |xxn | ≤ |xk xk+1 | < 2ε.
Since y ∈ S by construction, it follows that dist(x, S) < 2ε. ¤
Theorem 7.3.8 (Blaschke). If X is compact, then M(X) is compact.
SA ∈ 2S defined by
SA = {x ∈ S : dist(x, A) ≤ ε}.
Since S is an ε-net in X, for every y ∈ A there exists an x ∈ S such that
|xy| ≤ ε. Since dist(x, A) ≤ |xy| ≤ ε, this point x belongs to SA . Therefore
dist(y, SA ) ≤ ε for all y ∈ A. Since dist(x, A) ≤ ε for any x ∈ SA (by the
definition of SA ), it follows that dH (A, SA ) ≤ ε. Since A is arbitrary, this
proves that 2S is an ε-net in M(X). ¤
Note that X ′ and Y ′ in the above definition are regarded with the
restriction of the metric of the ambient space Z (as opposed to the induced
intrinsic metric). For example, if X is a sphere with its standard Riemannian
metric, one cannot take Z = R3 and X ′ = S 2 ⊂ R3 because X and X ′ would
be only path-isometric but not isometric.
It is trivial that the Gromov–Hausdorff distance between isometric
spaces is zero. Later we will show that dGH is a metric in the same sense as
dL ; i.e., it is a metric on the space of the isometry classes of compact met-
ric spaces (Theorem 7.3.30). Naturally one says that a sequence {Xn }∞ n=1
of (compact) metric spaces converges in the Gromov–Hausdorff sense to a
(compact) metric space X if dGH (Xn , X) → 0. For noncompact spaces
7.3. Gromov–Hausdorff Distance 255
Proof. 1. For any r > dGH (X, Y ), there exists a correspondence R with
dis R < 2r. Indeed, since dGH (X, Y ) < r, we may assume that X and Y are
subspaces of some metric space Z and dH (X, Y ) < r in Z. Define
R = {(x, y) : x ∈ X, y ∈ Y, d(x, y) < r}
where d is the metric of Z. That R is a correspondence follows from the
fact that dH (X, Y ) < r. The desired estimate dis R < 2r follows from the
triangle inequality: if (x, y) ∈ R and (x′ , y ′ ) ∈ R, then
|d(x, x′ ) − d(y, y ′ )| ≤ d(x, y) + d(x′ , y ′ ) < 2r.
The next corollary gives us more techniques for handling the Gromov–
Hausdorff distances. While it does not give explicit expressions for the
distance, it provides another quantity which differs from the distance by
no more than two times. Note that an estimate of this type is sufficient
to study the topology (on the space of metric spaces) determined by the
Gromov–Hausdorff distance.
Corollary 7.3.28. Let X and Y be two metric spaces and ε > 0. Then
1. If dGH (X, Y ) < ε, then there exists a 2ε-isometry from X to Y .
2. If there exists an ε-isometry from X to Y , then dGH (X, Y ) < 2ε.
Hence dis R ≤ 3r, and Theorem 7.3.25 implies dGH (X, Y ) ≤ 32 r < 2r. ¤
Remark 7.3.29. It is important that we do not require continuity of ε-
isometries. Even if two spaces are very close with respect to the Gromov–
Hausdorff distance, it can happen that there are no continuous maps with
small distortion—recall the spheres with small handles in Subsection 7.1.4.
Proof. We have already proven all statements of this theorem except the
claim that dGH (X, Y ) = 0 implies that X and Y are isometric. Let X and Y
be two compact spaces such that dGH (X, Y ) = 0. By Corollary 7.3.28, there
exists a sequence of maps fn : X → Y such that dis fn → 0. Fix a countable
dense set S ⊂ X. Using the Cantor diagonal procedure, one can choose a
subsequence {fnk } of {fn } such that for every x ∈ S the sequence {fnk (x)}
converges in Y . Without loss of generality we may assume that this holds for
{fn } itself. Then one can define a map f : S → Y as the limit of fn , namely,
set f (x) = lim fn (x) for every x ∈ S. Since |d(fn (x), fn (y)) − d(x, y)| ≤
dis fn → 0, we have d(f (x), f (y)) = lim d(fn (x), fn (y)) = d(x, y) for all
x, y ∈ S. In other words, f is a distance-preserving map from S to Y . Then
f can be extended to a distance-preserving map from the entire X to Y
(by Proposition 1.5.9). Now we can finish the proof in the same way as in
Theorem 7.2.4. Namely there is a similar distance-preserving map from Y
to X, and it follows that X and Y are isometric. ¤
The reader may wonder why we pay attention to the trivial case of finite
spaces. One of the reasons is that finite spaces form a dense set in the
Gromov–Hausdorff space:
Example 7.4.9. Every compact metric space X is a limit of finite spaces.
Indeed, take a sequence εn → 0 of positive numbers and choose a finite
εn -net Sn in X for every n. Then Sn −→
GH X, simply because dGH (X, Sn ) ≤
dH (X, Sn ) ≤ εn .
262 7. Space of Metric Spaces
Moreover these ε-nets can be chosen so that, for all sufficiently large n,
Sn have the same cardinality as S.
(2) For every ε > 0 there exists a natural number N = N (ε) such that
every X ∈ X contains an ε-net consisting of no more than N points.
Exercise 7.4.14. Prove that the first condition of the above definition is
redundant (i.e., is implied by the second one) if all elements of X are length
spaces.
The result of the above exercise does not mean that any class of compact
Riemannian n-manifolds with uniformly bounded volumes is pre-compact
in the Gromov–Hausdorff topology. The catch is in the phrase “sufficiently
small”— the actual bound for being “small” depends on M .
Proof. Let X be the length space in question. Pick small positive numbers
ε and δ (where δ is much smaller than ε), and choose a finite δ-net S in X.
7.5. Convergence of Length Spaces 267
Then consider the following graph G: the set of vertices of G is S, two points
x, y ∈ S are connected by an edge if and only if |xy| < ε, and the length of
this edge equals |xy|.
Let us show that G is an ε-approximation for X (cf. Definition 7.4.10) if δ
is small enough, say, δ < 14 ε2/diam(X). We consider S both as a subset of X
and a subset of G. Obviously S is an ε-net in both spaces, and |xy|G ≥ |xy|
for all x, y ∈ S where | · |G denotes the distance in G. It remains to show
that |xy|G ≤ |xy| + ε.
Let γ be a shortest path in X connecting x and y. Choose n points
x1 , . . . , xn , where n ≤ 2L(γ)/ε, dividing γ into intervals of lengths no
greater than ε/2. For every i = 1, . . . , n, there is a point yi ∈ S such
that |xi yi | ≤ δ. In addition, set x0 = y0 = x and xn+1 = yn+1 = y. Note
that |yi yi+1 | ≤ |xi xi+1 | + 2δ < ε for all i = 0, . . . , n. In particular, so yi and
yi+1 are connected by an edge in G provided that δ < ε/4. Then
n
X n
X
|xy|G ≤ |yi yi+1 | ≤ |yi yi+1 | + 2δn = |xy| + 2δn.
i=0 i=0
4 diam(X)
|xy|G ≤ |xy| + δ · < |xy| + ε
ε
if δ < 14 ε2/diam(X).
Thus we have a finite graph which is an ε-approximation for X. Passing
ε to zero yields a sequence of graphs converging to X. ¤
Exercise 7.5.6. Prove that every compact length space X can be repre-
sented as a Gromov–Hausdorff limit of finite graphs isometrically embedded
in X, i.e., of topological graphs embedded in X and equipped with their
induced length metrics.
Hint: Utilize the same construction as in the above proof, but draw
the edges in X (as shortest paths) adding new vertices when these shortest
paths intersect one another. To get rid of weird cases when one has to add
infinitely many vertices, show that shortest paths can be chosen so that the
intersection of any two of them is the empty set, or a single point, or an
interval of both paths.
Note that the number of vertices and edges of graphs constructed in the
proof of Proposition 7.5.5 tends to infinity as ε approaches zero. In other
words, the graphs get more and more complex. This is not a defect of this
particular construction but a consequence of a general obstacle:
268 7. Space of Metric Spaces
Most of the properties given in the above group of exercises are only
valid in the two-dimensional case. Below are some counterexamples to their
higher-dimensional counterparts.
Exercise 7.5.16. Let B be the standard three-dimensional ball. Show that,
for any given ε > 0, there exists a length space B ′ homeomorphic to B such
that dGH (B, B ′ ) < ε and ∂B ′ is an ε-net in B ′ . (Compare with Exercise
7.5.12.)
Hint: Push the boundary inside the ball so that it becomes an ε-net but
the intrinsic distances in the ball do not change much.
Exercise 7.5.17. Construct a sequence {Xn } of length spaces homeomor-
phic to S 3 and converging to the three-dimensional ball. (Compare with
Exercise 7.5.13.)
Hint: Consider doublings of the spaces B ′ from the previous exercise (i.e.,
the result of gluing together two isometric copies of B ′ along the boundary.)
Exercise 7.5.18. Construct a sequence {Xn } of length spaces homeomor-
phic to S 3 such that Xn −→ 3 3
GH S but no homeomorphism from Xn to S has
distortion less than 1/10. (Compare with Exercise 7.5.15.)
Chapter 8
Large-scale Geometry
To get an idea of what we are going to talk about in this chapter, imagine
a device measuring distances with a precision, say, one mile. Such a tool
is useless for an engineer investigating the shape of a car, but it is more
than excellent for learning geometry of the solar system. In this chapter
we consider metric properties for which such a measuring device is perfectly
good. In other words, we are not going to distinguish two metrics if the
difference between them is bounded by a constant. More precisely, the
properties that we will discuss are the same for spaces lying within finite
Gromov–Hausdorff distance from one another. For example, a Euclidean
space and a lattice in it (regarded with the restriction of the Euclidean
metric) look the same from this point of view. Of course, no local properties
survive through such a transformation of a space but many global and
asymptotic ones remain.
In some cases, we will admit even less precise “measurement of dis-
tances”. For example, consider a measuring instrument which may give the
result ten times greater or smaller than the actual distance, plus the same
one-mile error. Quite surprisingly, this instrument allows one, for example,
to tell the Euclidean plane from the hyperbolic one. (A question of this sort,
concerning the physical universe, is a famous problem in modern cosmogony
and physics. Alas, our real instruments are not that good.)
271
272 8. Large-scale Geometry
Exercise 8.1.2. Prove that for compact metric spaces the convergence of
pointed spaces is equivalent to the ordinary Gromov–Hausdorff convergence
in the following sense:
1. (Xn , pn ) −→
GH (X, p) implies that Xn −→
GH X.
2. If Xn −→
GH X and p ∈ X, then one can choose a point pn in every Xn
so that (Xn , pn ) −→
GH (X, p).
The first two requirements in the above definition imply that the image
f (Br (pn )) is contained in the ball of radius r + ε centered at p. This and
8.1. Noncompact Gromov–Hausdorff Limits 273
the third requirement imply (by Corollary 7.3.28) that the ball Br (pn ) in
Xn lies within the Gromov–Hausdorff distance of order ε from a subset of
X between the balls of radii r − ε and r + ε centered at p (here “between”
means that the set contains one ball and is contained in the other).
If X is a length space, this remains true for the r-ball centered at p
instead of the unknown subset; in other words, for every r > 0 the r-balls in
Xn centered at pn converge (with respect to the Gromov–Hausdorff distance)
to the r-ball in X centered at p. In the general case (for nonlength spaces)
such a simplification is not possible (see exercises below).
Exercise 8.1.3. Let (Xn , pn ) −→
GH (X, p) and let X be a length space. Prove
that Br (pn ) −→
GH Br (p) for every r > 0.
Exercise 8.1.4. Show that the statement of the previous exercise fails
without the assumption that X is a length space. To do that, construct
a sequence {Xn } of compact metric spaces converging to a compact metric
space X such that no sequence of closed unit balls in Xn converges to a
closed unit ball in X.
This property that the balls converge does not yet imply convergence of
pointed spaces. The first requirement in Definition 8.1.1 includes additional
information that puts the points pn and p into a special position. Roughly
speaking, not only should the balls converge, but also their distinguished
central points should converge at the same time.
Exercise 8.1.5. Construct a compact metric space X and two points
p, q ∈ X such that for every r > 0 the balls Br (p) and Br (q) in X are
isometric, but there is no isometry map from X to itself that maps p to q.
Hint: There are examples among finite spaces.
Given such X, p and q, let Xn = X and pn = q for all n ≥ 1. Prove
that the sequence {(Xn , pn )} of pointed spaces does not converge to (X, p)
despite the fact that Br (pn ) −→
GH Br (p) for every r > 0.
Proof. We only outline the proof here; the details are similar to those of the
proof of Theorem 7.3.30. Given an r > 0 and ε > 0, the map whose existence
is guaranteed by Definition 8.1.1 can be converted into a correspondence Rr,ε
between sets of Yr,ε ⊂ X and Yr,ε ′ ⊂ X ′ such that: Y ′
r,ε and Yr,ε contain the
balls of radius r − ε and are contained in the balls of radius r + ε centered at
p and p′ , respectively; p and p′ are in correspondence with each other; and
dis Rr,ε < ε.
Choosing one corresponding point for every point of Yε yields a map
′ that maps p to p′ and has distortion < ε. Using the Cantor
fr,ε : Yr,ε → Yr,ε
diagonal procedure first for ε → 0 and then for r → ∞, one can end up with
a distance-preserving map from a dense subset of X to X ′ , which extends to
a distance-preserving map f : X → X ′ with f (p) = p′ . Since f is distance-
preserving, it maps every ball Br (p) in X to the corresponding ball Br (p′ )
in X ′ .
In addition, the images of maps fr,ε are ε-nets in the respective subsets of
X ′ , and this implies that the balls Br (p′ ) in X ′ are compact as well (compare
with Exercise 7.3.31). Hence a similar distance-preserving map f ′ : X ′ → X
exists. Due to compactness of balls, this implies that the restriction of f to
Br (p) is an isometry onto Br (p′ ) for every r > 0. Hence f is an isometry
onto X ′ . ¤
In the sequel, we always assume that the pointed spaces under consid-
eration are boundedly compact. As the next exercise shows, this property
is inherited by limit spaces.
Exercise 8.1.8. Suppose that (Xn , pn ) −→
GH (X, p) where the spaces Xn are
boundedly compact and X is complete. Prove that X is boundedly compact.
Proof. Similar to that of Theorem 7.4.15. Again, one has to apply the
Cantor diagonal procedure twice, for ε → 0 and for r → ∞. ¤
Note that a cone is not necessarily a cone over a metric space as defined
in Subsection 3.6.2.
8.2.1. Tangent cone. The tangent cone is a local notion which does not
belong to large-scale geometry. Nevertheless we define it here because the
definition is in some sense similar to that of the asymptotic cone.
Definition 8.2.2. Let X be a (boundedly compact) metric space, p ∈ X. A
Gromov–Hausdorff limit of pointed spaces (λX, p) as λ → ∞, if one exists,
is called the Gromov–Hausdorff tangent cone of X at p.
As usual, the “limit as λ → ∞” can be interpreted as the limit through
any sequence of values for λ tending to infinity (which should exist and be
the same for all sequences).
Note that the tangent cone is a pointed metric space. Its distinguished
point (the natural ancestor of p) is called the origin or the apex of the cone.
The tangent cone is indeed a cone in the sense that it is isometric to any
dilatation of itself (via a pointed isometry). The tangent cone is a local
invariant: it is determined by any small neighborhood of the point. More
precisely, if U is a neighborhood of p in X, then the tangent cones of U and
X at p are isometric. This follows immediately from the definition.
The tangent cone is indeed a cone in the sense of Definition 8.2.1. This
follows from the following simple fact:
Exercise 8.2.3. Let {(Xn , pn )}∞n=1 be a sequence of pointed metric spaces,
(Xn , pn ) −→
GH (X, p). Prove that (λXn , pn ) −→
GH (λX, p) for any λ > 0.
Exercise 8.2.5. Prove that the tangent cone of a convex set, as defined in
Subsection 7.1.1, is also the Gromov–Hausdorff tangent cone.
Proposition 8.2.8. The asymptotic cone does not depend on the choice of
the reference point p.
Exercise 8.2.9. Prove that the asymptotic cone of a convex set, as defined
in Subsection 7.1.2, is also its Gromov–Hausdorff asymptotic cone.
Exercise 8.2.10. Let X and Y be metric spaces and dGH (X, Y ) < ∞.
Prove that, if X has an asymptotic cone, then Y has one too, and the two
cones are isometric.
In particular, if a metric space X lies within a finite Gromov–Hausdorff
distance from some cone Y , then Y is an asymptotic cone of X.
Exercise 8.2.11. Prove that the grid described in Subsection 7.1.3 has
asymptotic cone isometric to R21 .
The next exercise generalizes the previous one. There are further gen-
eralizations in Subsection 8.5.1.
the norm on Rn whose unit ball is the convex hull of S. Prove that there
exists a constant C (depending on S) such that
kx − yk ≤ d(x, y) ≤ kx − yk + C
for all x, y ∈ Zn .
In particular, the Gromov–Hausdorff distance between (Zn , d) and the
normed space (Rn , k · k) is finite (not greater than C), and hence (Rn , k · k)
is the asymptotic cone of (Zn , d).
Exercise 8.2.13. Prove that the hyperbolic plane H2 does not have an
asymptotic cone.
1
Hint: The sequence of rescaled balls n Bn (p), n ∈ N, p ∈ H2 , is not
uniformly totally bounded.
In plain words, the reason why the hyperbolic plane does not have an
asymptotic cone is that its metric balls grow too fast when the radius goes
to infinity. There are other constructions encoding asymptotic properties
of such “fast growing” spaces and in other cases when Gromov–Hausdorff
asymptotic cones are not applicable. One possible approach is to use a
weaker type of limit than the Gromov–Hausdorff one. We do not discuss
such generalized definitions in this book. Let us only mention that the
“generalized” asymptotic cone of the hyperbolic plane is an infinite (and
not locally finite) tree.
Another construction serving the same purposes is the ideal boundary
of a space; cf. Subsection 5.3.3 for the case of hyperbolic plane. While the
asymptotic cone is “the tangent cone at infinity” in some sense, the ideal
boundary is a sort of “space of directions” at infinity.
8.3. Quasi-isometries
8.3.1. Definitions and first examples. Quasi-isometries are a large-
scale analog of bi-Lipschitz maps. Two metric spaces are quasi-isometric if
they are bi-Lipschitz equivalent up to a finite Gromov–Hausdorff distance.
This is formally described as follows:
Proposition 8.3.4. For any metric spaces X and Y , the following three
assertions are equivalent:
(i) X and Y are quasi-isometric;
(ii) there is a quasi-isometry f : X → Y whose image f (X) is a net in
Y;
(iii) X and Y contain bi-Lipschitz homeomorphic separated nets. ¤
Corollary 8.3.5. Being quasi-isometric is an equivalence relation.
Exercise 8.3.8. Prove that the orbit of a point is a net if and only if the
action is co-bounded.
This can be interpreted as follows: one identifies G with the orbit Gx and
uses the metric d restricted on Gx. The metric dG defined by the above
formula will be referred to as an orbit metric.
8.3. Quasi-isometries 281
Remark 8.3.14. One can apply the same construction to other coverings,
not only to the universal one; however the covering must be regular. (This
is necessary in order to represent Y as the quotient space of the group of
deck transformations.) In this case, the group of deck transformation is a
factor-group of π1 (Y ).
We will soon see that all possible orbit metrics of co-compact actions on a
given (finitely generated) group are bi-Lipschitz equivalent to one another.
In other words, the quasi-isometry class of a metric on a group depends
only on the group, not on a length space and an action used to define the
metric. This yields the following remarkable corollary: if two compact length
spaces have isomorphic fundamental groups, then their universal coverings
are quasi-isometric. Indeed, each universal covering has finite Gromov–
Hausdorff distance from the fundamental group equipped with an orbit
metric, and two orbit metrics are bi-Lipschitz equivalent (as we will prove).
Since all metrics on a fixed group are bi-Lipschitz equivalent, we can talk
about quasi-isometries between abstract finitely generated groups without
referring to a particular metric. Namely, two groups are quasi-isometric if
they are quasi-isometric when equipped with arbitrary word metrics.
Example 8.3.15. The free groups F2 and F3 with two and three generators
(equipped, say, with word metrics) are quasi-isometric. (Compare this with
the fact that Zm and Zn are not quasi-isometric if m 6= n, as we have seen
in Example 8.3.6.)
To prove this, recall that F2 is the fundamental group of a bouquet
of two circles (say, of unit length). This bouquet X admits a two-sheeted
covering by a graph Y consisting of three circles S1 , S2 , S3 connected as a
chain: S1 has one common point with S2 , S2 has one common point with
S3 , and S1 ∩ S2 = ∅. Metrically, Y is a circle of length 2 glued with two
circles of unit length; the smaller circles are attached to opposite points of
the larger one. Constructing the covering map from Y to X is left as an
exercise.
Since Y is a covering space for X, the universal covering of Y coincides
with the universal covering of X. This covering space is quasi-isometric to
both fundamental groups π1 (X) ∼ = F2 and π1 (Y ) ∼
= F3 . (The latter group is
F3 because Y is homotopy equivalent to a bouquet of three circles. To see
this, just contract the edges.) Thus F2 and F3 are quasi-isometric.
Proof. Let |·|1 and |·|2 be the distance functions of the identity in two word
metrics defined by generating sets S1 and S2 , respectively. Recall that the
sets S1 and S2 must be finite. Let g1 . . . gn be a shortest word in generators
from S1 representing a given element g ∈ G. Then |g|1 = n and
|g|2 = |g1 . . . gn |2 ≤ |g1 |2 + · · · + |gn |2 ≤ C1 n = C1 |g|1 ,
where C1 = maxh∈S1 |h|2 . (Note that we have not yet used that | · |2 is a
word metric.) Similarly, |g|1 ≤ C2 |g|2 for some constant C2 not depending
on g. Thus | · |1 and | · |2 are bi-Lipschitz equivalent. ¤
Theorem 8.3.19. Let G be a finitely generated group and d be an orbit
metric of a free co-compact action of G by isometries on a length space X.
Then d is bi-Lipschitz equivalent to a word metric.
In particular, all such orbit metrics on G are bi-Lipschitz equivalent to
one another.
Corollary 8.3.20. All length spaces X admitting a free totally discontinu-
ous co-compact action of a given group G are quasi-isometric to one another,
and are quasi-isometric to the group G equipped with any word metric.
In particular, if Y is a length space and X its universal covering with the
metric lifted from Y , then X is quasi-isometric to the group π1 (Y ) equipped
with any word metric.
Proof of the theorem. Let | · | denote the distance from the identity in
the metric d. If | · |w is a word metric, then | · | ≤ C| · |w for some constant
C, similarly to the previous proof. Thus it suffices to prove that | · |w ≤ C| · |
for some constant C and some word metric | · |w .
Let x ∈ G be the point whose orbit is used to define the metric d (i.e.,
d(g, h) = dX (gx, hx) for all g, h ∈ G). Since the action is co-compact,
X is locally compact and complete and the orbit Gx is a separated net.
Therefore every metric ball in X contains only finitely many elements of the
orbit. Hence every ball in (G, d) is a finite set.
Let D be a number such that the orbit Gx is a D-net in X, and let r
be so large that r > 2D + 1 and the r-ball in (G, d) centered at the identity
contains a set of generators. We may assume that this ball itself is chosen
as the set of generators S defining the word metric | · |w . (We have an
option to choose a word metric to be compared with d; then the result will
284 8. Large-scale Geometry
follow for any word metric because all word metrics are already proven to
be equivalent.)
Pick a g ∈ G and let γ be a shortest path in X connecting x to gx. We
are going to show that |g|w ≤ C|g| = C ·L(γ) for a constant C not depending
on g. Divide γ into intervals of length 1 or less by points x1 , . . . , xn , where
n ≤ L(γ) ≤ n + 1. For every i = 1, . . . , n, find a yi ∈ Gx such that
dX (xi , yi ) ≤ D. In addition, set y0 = x and yn+1 = Gx. Let gi ∈ G be such
−1
that gi x = yi (note that g0 = e and gn+1 = g), and set hi = gi gi−1 . Then
hi xi−1 = yi ; hence |hi | = dX (xi−1 , xi ) ≤ 2D + 1. Therefore hi belongs to our
set of generators (recall that it is just the r-ball in G, and r > 2D + 1). On
the other hand, the product (word) h1 . . . hn+1 equals g; hence |g|w ≤ n + 1.
Thus |g|w ≤ L(γ) + 1 = |g| + 1. Since the distances |g| are separated
away from zero (i.e., there is an ε > 0 such that |g| ≥ ε for all ε > 0),
this estimate can be written in the desired form |g|w ≤ C|g|, namely,
|g|w ≤ (1 + 1/ε)|g|. ¤
Remark 8.3.21. One can remove the condition that the action is free from
the theorem. The only problem is that an orbit metric is no longer a metric;
it is only a semi-metric. As a consequence, it is not bi-Lipschitz equivalent
to word metrics, but it is still quasi-isometric to them. The same proof
works with obvious modifications.
The condition that a given metric is an orbit metric of some action seems
hard to verify and somewhat restrictive (it is not a ”large-scale” one, after
all). There are various possible replacements for it; one of such conditions
is given in the following exercise.
Exercise 8.3.22. Prove the following statement which is slightly more
general than the above theorem. Let d be a left-invariant metric on a finitely
generated group G such that every ball of d is a finite set, and there is a
constant C > 0 satisfying the following: for any x, y ∈ G there is a z ∈ G
with d(x, z) < 12 d(x, y)+C and d(y, z) < 21 d(x, y)+C. Then d is bi-Lipschitz
equivalent to the word metric.
Exercise 8.4.5. Show that if the inequality (8.3) is satisfied for all triangles
△a1 a2 a3 and all choices of p ∈ [a2 a3 ] for a strictly intrinsic metric d on X,
then (X, d) is Gromov hyperbolic.
To get more insight into the geometric meaning of Definition 8.4.6, let
us look what it means for a 4-point space:
Lemma 8.4.10. Let X be a 4-point 0-hyperbolic space. Then X is isometric
to the set of leaves of a tree (with at most six vertices). Here the tree carries
288 8. Large-scale Geometry
Definition 8.4.18. A path γ (in a length space (X, d)) is called C-quasi-
geodesic if L(γ)[s,t] ≤ C ·d(γ(s), γ(t)) for all s, t in the domain of γ. In other
words, the length of every segment of γ is at most C times longer than the
distance between its endpoints.
Proof of the Morse Lemma. The proof consists of several steps, and we
advise the reader to try to use them as exercises before reading our proofs.
Though the proof contains ugly estimates, they are needed only to formally
show that certain distances play no role for our large-scale consideration.
8.4. Gromov Hyperbolic Spaces 291
Now we are ready to prove the Morse Lemma (Theorem 8.4.20). Con-
sider a C-quasi-geodesic γ parameterized by arc length and connecting
p = γ(0) and q = γ(T ). Let 2R = maxt∈[0,T ] d(γ(t), [pq]), and choose τ
with d(γ(τ ), [pq]) = 2R.
Reasoning by contradiction, assume that R > k 2 δ, where k > 20C + 8
is a natural number (so in particular R > 400δ).
Let [t′ , t′′ ] be the biggest interval containing τ and such that
d(γ|[t′ , t′′ ], [pq]) ≥ R (in particular, d(γ(t′ ), [pq]) = d(γ(t′′ ), [pq]) = R).
It is obvious that |t′′ − t′ | (which is equal to L(γ, t′ , t′′ )) is at least 2R.
Choose t0 = t′ > t1 > t2 · · · > tn = t′′ so that R/2 ≤ ti − ti−1 ≤ R. Then
|t′′ − t′ | ≥ nR/2. Denote b = b0 = γ(t′ ) = γ(t0 ), bi = γ(ti ). Let bi1 be a point
closest to bi in [pq].
By Lemma 8.4.24 one has d(bi1 , bi+1 1 ) ≤ 6δ (recall that R ≫ δ). Thus
d(γ(t′ ), γ(t′′ )) ≤ 4R + 6nδ, and therefore
|t′′ − t′ | nR
C≥ ′ ′′
≥ .
d(γ(t ), γ(t )) 8R + 12nδ
Since R > 400Cδ, the last inequalities imply n < 9C.
8.4. Gromov Hyperbolic Spaces 293
Now set a = b01 ; then d(a, γ(t′ )) = R. Obviously γ|[t′ t′′ ] lies outside a
ball of radius R centered at a. Let c be the point in a segment [aγ(t′′ )] at
distance R from a. Since d(bn1 , γ(t′′ ) = R) and d(b01 , bn1 ) ≤ 4nδ ≤ 36Cδ, by
the triangle inequality
(8.7) d(c, γ(t′′ )) ≤ 36Cδ < d(γ(t′ ), γ(t′′ ))
and
1
(8.8) d(b, c) ≥ d(γ(t′ ), γ(t′′ )) − 36Cδ > d(γ(t′ ), γ(t′′ )).
2
(Note that these inequalities are very crude; but all we want is to show that
[cγ(t′′ )] is short and plays no role in our large-scale considerations.)
Then a path consisting of γ|[t′ t′′ ] continued by a segment [γ(t′′ )c] also
lies outside the ball of radius R centered at a. Applying Lemma 8.4.22 to
this path, we can estimate the length of γ|[t′ t′′ ] as
k
|t′′ − t′ | ≥ d(b, c) − d(γ(t′′ ), c).
10
Combining this with (8.7) and (8.8), we get
k
|t′′ − t′ | ≥ (
− 1) d(γ(t′ ), γ(t′′ )).
20
Since k/20 − 1 > C, this contradicts the assumption that γ is a C-quasi-
geodesic. ¤
of Γ. Recall that the latter has Γ as its set of vertices, and two vertices
γ1 and γ2 are connected by an edge (of length 1) if and only if γ1 s = γ2
for some s ∈ S. Then the word distance between two elements of Γ is the
natural intrinsic distance between them in the Cayley graph. Of course,
the Gromov-Hausdorff distance between Γ and C(Γ) is finite, and thus they
share all large-scale properties. We will usually prefer to deal with C(Γ)
since it is a length space.
Definition 8.4.25. A group Γ is said to be hyperbolic (with respect to some
finite generating set S) if the Cayley graph C(Γ) (equivalently, just Γ) is a
Gromov hyperbolic metric space.
Recall that any two word metrics are bi-Lipschitz equivalent (Proposi-
tion 8.3.18): if dS , dT are associated to generating sets S, T , then there is a
positive C such that
1
dS (x, y) ≤ dT (x, y) ≤ CdS (x, y).
C
Indeed, every element in S can be represented as a product of elements of
T , and vice versa—then we can choose for C the maximum length of such
representations.
Therefore, by Theorem 8.4.16, if Γ is hyperbolic with respect to some
word metric, then it is hyperbolic with respect to every word metric, and
we can speak of a hyperbolic group without specifying a metric. In other
words, to be hyperbolic is an algebraic property of a group, and it does not
depend on which metric we choose.
The first and dull example of a hyperbolic group is Z. Of course, if
Γ is hyperbolic, then so is Γ × G for a finite group G (just because the
Gromov-Hausdorff distance between Γ and Γ × G is finite—check this!)
Hence Z × G is also hyperbolic for any finite G. These examples, however,
lack the most interesting features possessed by all other hyperbolic groups.
However, already a free (nonabelian) group Fk with k ≥ 2 generators is a
characteristic representative of the hyperbolic world.
Exercise 8.4.26. Prove that a tree is a 0-hyperbolic space, and hence Fk
is a hyperbolic group.
Every hyperbolic group “on the large scale looks like a tree” (see Propo-
sition 8.4.14). This property happens to be characteristic for hyperbolic
groups, and the latter can also be defined as follows:
Definition 8.4.27. A finitely-generated group Γ is said to be hyperbolic if
every subcone at infinity for C(Γ) is a topological tree.
It is clear from Proposition 8.4.14 that this definition follows from
Definition 8.4.25; we omit the proof of the converse in our exposition.
8.4. Gromov Hyperbolic Spaces 295
We will not prove here that this definition is equivalent to the two
previous definitions (while showing that hyperbolic groups satisfy a linear
isoperimetric inequality is not difficult and can be suggested to the reader as
an exercise, a proof of the converse is somewhat involved). Let us, however,
discuss isoperimetric inequality for groups in a more geometrical context.
Recall that every finitely presented group is the fundamental group of a
certain finite two-dimensional cell complex. Indeed, begin with the bouquet
of k circles. Choose an orientation for each of the circles and label the
circles by pairs of generators (gi , gi−1 ). Then glue a two-cell for each relation
following the word representing this relation. This means that the boundary
circle of the cell is glued to a curve traversing the circles in the bouquet in the
same order as their labels follow in the word representing the relation (and
choosing in which direction we follow a circle labeled by (gi , gi−1 ) depending
on whether we met gi or its inverse). The Van Kampen theorem tells us
that the fundamental group of this cell complex K is Γ. If each circle comes
with an intrinsic metric, and each two-cell is represented by a polygon glued
along isometries of its sides, K turns into a length space. Its universal cover
K̃ is quasi-isometric to Γ, and hence it is Gromov hyperbolic (note that
showing that K̃ is Gromov hyperbolic, for instance by introducing a metric
of negative curvature on K̃, is in its turn a method of arguing that Γ is
hyperbolic; see Exercise 8.4.32).
Let us fix a vertex (1-cell) p in K̃. Then a word ω in generators
g1 , g1−1 , . . . , gk , gk−1 determines a path γ in K̃ starting from p and following
the edges of the 1-dimensional skeleton of K̃ with labels (lifted from K)
corresponding to letters in ω. The value of ω is the identity if and only if
this path γ is closed (returns back to p).
What is the meaning of a simple modification of ω in terms of γ?
Removing a subword identical to one of the relations means contracting
a loop of γ going around one 2-cell; inserting a relation means adding such
a loop; finally, crossing out a generator and its inverse next to each other
corresponds to contracting a trivial loop (when γ follows an edge and then
immediately comes back). Therefore a sequence of simple modifications can
be realized by a homotopy of γ, and the number of 2-cells swept by γ in the
course of this homotopy is no more than the number of simple modifications
in this sequence. Recall that contracting a closed curve to a point sweeps a
topological disc. Now if γ is closed (that is, the value of ω is the identity), the
number of simple modifications required to transform ω into an empty word
is the same as the combinatorial area (the number of 2-cells) in a topological
disc bounded by γ. Hence we see that a linear isoperimetric inequality for
298 8. Large-scale Geometry
We conclude this section with a few remarks, which can perhaps ignite
the reader’s curiosity and push her to systematic study of this subject. It
is well known that the word problem (determining whether the value of
a given word is the identity) is algorithmically undecidable for a general
finitely presented group. Using a linear isoperimetric inequality, it is easy to
see that this problem is always decidable for hyperbolic groups; moreover, a
more delicate analysis involving the Morse Lemma shows that it can always
be decided in linear time. As a matter of fact, hyperbolic groups belong
to the class of automatic groups, the groups whose multiplication can be
checked by a finite automaton. The reader familiar with this notion can try
to re-prove this (elementary, however tricky and very elegant) result of W.
Thurston.
vector space V . The reader may keep in mind the picture of ZN as the set
of integer points in RN , but remember that the Euclidean structure is irrele-
vant. (In formal algebraic language, the ambient vector space V is obtained
from G by tensor multiplication, V = G ⊗ R.)
d(nv)
Proposition 8.5.1. For any v ∈ G, the limit limn→∞ n exists (n is
assumed natural).
Proof. The statement follows from the following lemma, which is a baby
version of the subadditive ergodic theorem:
Lemma 8.5.2. Let {xn }∞n=1 be a sequence of nonnegative real numbers such
that xm+n ≤ xm + xn for all m, n. Then the limit limn→∞ n1 xn exists.
To derive the proposition from the lemma, let xn = d(nv). The condition
xm+n ≤ xm + xn follows from the triangle inequality. ¤
Proof. First observe that the first statement implies the second one. In-
deed, one has to prove that (G, λd) −→ GH (V, k · k) as λ → 0 in the sense
of pointed space convergence (the distinguished point in 0). In fact, the
maps fλ : (G, λd) → (V, k · k) given by fλ (v) = λv satisfy the requirements
of the definition of pointed space convergence (Definition 8.1.1). The only
nontrivial part is that for every R > 0 the distortion of fλ within the ball of
radius R goes to zero (as λ → 0), and this follows from the first statement
of the theorem.
It remains to prove the first statement. It is trivial that kvk ≤ d(v) for all
v ∈ G. On the other hand, by Theorem 8.3.19 we have d ≤ Ck · kE for some
constant C > 0. Fix an ε > 0 and let S be a finite ε-net of rational vectors
(i.e., S ⊂ QG) in the unit ball of (V, k · k). There exists a large natural
number M such that M v ∈ G and moreover d(M v) ≤ (1 + ε)M kvk ≤
(1 + ε)M for all v °∈ S (cf. Proposition
° 8.5.1). For every w ∈ G there exists
a v ∈ S such that °w/kwk − v ° < ε.
Let k be the integer such that k ≤ kwk < k+1, then kw−kvk ≤ εkwk+1,
then
d(M w) ≤ d(kM v) + d(M w − kM v) ≤ (1 + ε)kM + CM (εkwk + 1),
hence d(M w)/kM wk ≤ 1 + ε + Cε + M/kwk. Note that this is an upper
bound on d/k · k for all vectors divisible by M in G. There is a constant
C1 = C1 (M ) such that every v ∈ G has a vector divisible by M in C1 -
neighborhood of it. It follows easily that d(v)/kvk ≤ 1 + ε(C + 2) for all
v ∈ G outside the ball of some radius R = R(M, ε). Since ε is arbitrary, the
first statement of the theorem follows. ¤
Remark 8.5.5. Let d be an orbit metric of a co-compact totally discontin-
uous action of G on a length space X. Recall that the same action defines
a variety of orbit metrics depending on the choice of the orbit. However,
8.5. Periodic Metrics 301
Remark 8.5.6. The first statement of the theorem remains valid without
the assumption that k · k is a norm. (The proof is essentially the same.) The
second statement requires an obvious correction if k·k is not a norm. Namely
one has to apply the standard factorization to make (V, k · k) a metric space.
The result is a normed vector space of a smaller dimension.
Remark 8.5.9. The statement of the last exercise implies that the Gromov–
Hausdorff distance between (G, d) and its asymptotic cone (V, k · k) is finite.
This holds not only for word metrics on G ≃ ZN but for any length space
X admitting a co-compact action of ZN by isometries ([Bur]). The proof
of this fact in full generality is beyond the scope of this book. In the next
subsection we prove it in the special case X = R2 and G = Z2 (see Corollary
8.5.13).
The first thing to prove is that the limit in the definition exists.
Proposition 8.5.18. The asymptotic volume Ω(d) exists and does not
depend on x ∈ Rn . Moreover, in the notation introduced above
Ω(d) = Vold (I n ) · µn (D)
where I n = [0, 1]n and µn is the standard Lebesgue (or Hausdorff ) measure
in Rn . In particular, Ω(d) is finite and positive.
Proof. Let k · k be the stable norm of d, D its unit ball and Q be an affine
cube such that n1 Q ⊂ D ⊂ Q (such a Q exists due to Lemma 5.5.19). Since
D ⊂ Q, one has kv − v ′ k ≥ 2 whenever v and v ′ belong to opposite faces
306 8. Large-scale Geometry
of Q. The idea of the proof is the following: we let x = 0 replace the ball
Br (x) ≈ rD in the definition of Ω(d) by a smaller set nr Q. Then the volume
of nr Q is estimated below by means of the Besikovitch inequality.
Now we pass to a formal argument. Fix an ε > 0 and let r be so large
that
(1 − ε)kx − yk ≤ d(x, y) ≤ (1 + ε)kx − yk
whenever kx − yk ≥ r/n or d(x, y) ≥ r/n (see Theorem 8.5.4 and the
beginning of this section). Then the distance (in d) between the opposite
faces of nr Q is not less than 2(1 − ε)r/n. By the Besikovitch inequality it
follows that ³r ´
Vold Q ≥ (1 − ε)n (2r/n)n .
n
On the other hand, since nr Q ⊂ rD, we have kxk ≤ r and hence d(0, x) ≤
(1 + ε)r for all x ∈ nr Q. This means that our affine cube nr Q is contained in
the ball B(1+ε)r (0). Thus
Vold (B(1+ε)r (0)) Vold ( nr Q)
Ω(d) = lim ≥ lim inf ≥ (2(1 − ε)/n)n/(1 + ε)n .
r→∞ (1 + ε)n rn r→∞ (1 + ε)n r n
Spaces of Curvature
Bounded Above
307
308 9. Spaces of Curvature Bounded Above
and this indicates that the differential technique in the Riemannian proofs
is not that essential. On the contrary, convexity arguments often play the
key role.
modified for the case of curvature ≤ k, and the four resulting definitions are
equivalent (Exercise 4.6.3).
An unpleasant part of our definition is the requirement that any two
points (in a neighborhood U ) can be connected by a shortest path. While
we mainly restrict ourselves to locally compact spaces where shortest paths
always exist, not locally compact spaces can arise in certain constructions.
In order to handle them, it is natural to have a more general version
of Definition 9.1.1 where midpoints are replaced by “almost midpoints”
(similarly to the “intrinsic versus strictly intrinsic” case in Subsection 2.4.3).
Proof. Let Br (p) be a normal ball, a, b ∈ Br (p). Since r < Rk /2, a ball
Brk (p) in the k-plane is a convex set. By the comparison property for a
9.1. Definitions and Local Properties 313
comparison triangle △pab, one has |pd| < r for any point d ∈ [ab]. Hence
|pd| ≤ |pd| < r for any d ∈ [a, b], and this means that [ab] ⊂ Br (p). ¤
Proof. (1) Suppose there are two shortest paths, [ab]1 and [ab]2 . For every
point c ∈ [ab]1 one can consider the triangle △abc and its comparison triangle
△abc. The latter one is degenerate: c ∈ ab. Let d be a point of [ab]2 such
that |ac| = |ad|. Then it has to be |cd| ≤ |cd| = 0, so c = d and [ab]1 = [ab]2
since c is arbitrary.
Since a limit of shortest paths is a shortest path, (2) follows from (1).
The proof of (3) is the same as for Proposition 9.1.16. And (4) follows
from the fact that a comparison triangle for △abc is degenerate with ∡bac =
π.
(5) Let γ : [0, L] → U be a geodesic but not a shortest path. Let
t ∈ [0, L] be the maximal value of the parameter such that the restriction
of γ to [0, t] is still a shortest path. Such a maximal t exists because the
restriction to [0, t] is a shortest path if and only if |γ(0)γ(t)| = t, and the
latter condition defines a closed set. Since γ is a geodesic, its restriction
to [t, t + ε] is a shortest path for a sufficiently small ε > 0. Applying the
statement (4) to the shortest paths γ|[0,t] and γ[t,t+ε] yields a contradiction
with the maximality of t. ¤
Simple examples, like flat tori or hyperbolic space forms, show that, in
general, the local curvature conditions do not imply analogous properties
in the large. This is related to the fact that these spaces contain a closed
geodesic. Tori and surfaces of higher genus are not simply connected, so
closed geodesics can be found by minimizing the length in a free homotopy
class of noncontractible closed curve. The space in Example 9.1.13 is simply
connected (it is homeomorphic to the sphere) but its convexity radius can
be made arbitrarily small while the curvature bound stays the same. We
will see later that this effect (small convexity radius in a simply connected
space) is possible only for k > 0.
3. (ii) =⇒ (iii). Suppose that |bx| = |bx| for some x ∈ [a, c] \ {a, c}.
First we show that the equality |by| = |by| holds for all y ∈ [a, c]. Since
|bx| = |bx|, the triangles bxa and bxc are comparison triangles for △bxa
and △bxc. Therefore ∡bxa ≤ ∡bxa and ∡bxc ≤ ∡bxc. On the other
hand, ∡bxa + ∡bxc ≥ ∡axc = π. Since ∡bxa + ∡bxc = π, it follows that
∡bxa = ∡bxa and ∡bxc = ∡bxc, so the assertion (i) holds for triangles
△bxa and △bxc. Since we have already proved that (i) implies (ii), we can
conclude that |by| = |by| for all y ∈ [ax] ∪ [cx].
Now applying the first variation formula (for k = 0 it is Theorem 4.5.6;
its generalization to arbitrary k is trivial) to the distance from b along the
geodesic [ac], we obtain that α = α and β = β. The remaining equality
γ = γ follows by a combination of the implication (i) =⇒ (ii) and the
already proved part of the implication (ii) =⇒ (iii).
4. Thus the assertions (i)–(iii) are equivalent. It remains to prove that
they imply (iv). The desired totally geodesic surface can be obtained by
sweeping it by a family of shortest paths connecting b and [ac].
Let x, y ∈ [ac]. Construct corresponding points x, y in [ac]. We assume
that y ∈ [x, c]. We have
β ≥ ∡abx + ∡xby + ∡ybc ≥ ∡abx + ∡xby + ∡ybc ≥ β = β.
Hence all these inequalities turn out to be equalities, in particular, ∡xby =
∡xby. Hence the assertions (i)–(iii) hold for △xby implying that the dis-
tances between points of [bx] and points of [by] are the same as between the
respective points in [bx] and [by] in the k-plane. Therefore the union of the
shortest paths {[bx]}x∈[a,c] is a desired “full” triangle isometric to △abc. ¤
Exercise 9.1.20. Let △ be a triangle contained in a normal region of a
space of curvature ≤ 0. Let a, b, c denote the lengths of the sides of △ and
α, β, γ the respective (opposite) angles. Prove the following inequalities:
α + β + γ ≤ π,
c ≥ a2 + b2 − 2ab cos γ,
2
c ≤ b cos α + a cos β.
Show that the equality in any of these inequalities implies that the assertions
from the above proposition holds.
Hint: Observe that these inequalities turn to equalities for a comparison
triangle in R2 ; then apply the angle condition.
Then there are points c1 , c2 on the shortest paths [ac], [bc], resp. contained
in C. Decompose △abc into three triangles △ac1 c2 , △bac2 , △cc1 c2 . Place
their comparison triangles in the plane “in the natural” way (see Figure 9.1).
Usual angle comparison arguments show that the angles at c1 and c2 in the
polyhedron ac1 cc2 b are concave (i.e., not less than π). Loosely speaking, one
can consider this polyhedron like a triangle abc having two “concave sides”,
ac1 c and bc2 c. “Straightening this triangle” (compare with Alexandov’s
Lemma 4.3.3) we see that the angles of △abc are not greater than the angles
of its comparison triangle. ¤
In other words, every geodesic can be extended in both sides “to infin-
ity”. Note that no curvature restriction is assumed here.
Proof. By Theorem 2.4.16 it is sufficient to prove that for every two points
a, b ∈ Br (a0 ) and every ε > 0 there is an ε-midpoint. Recall that Σp is
the completion of its subset Σ′p consisting of points represented by shortest
paths. Without loss of generality, one can assume that points a and b
belong to Σ′p . So they can be represented by shortest paths α and β,
α(0) = β(0) = p. Take an ε > 0 and choose points x = α(t), y = β(t)
so close to p that 0 ≤ ∡xpy − ∡xpy ≤ ε, where ∡xpy is the angle in a
comparison triangle △xpy. Let z be a midpoint between x and y. Our
assumption guarantees that d(a, b) < π, where d is the distance in Σ′p .
Hence z 6= p for sufficiently small ε. Place comparison triangles △xpz and
△ypz in the k-plane in different half-planes separated by the line pz and,
as usual, compare the resulting quadrilateral with a comparison triangle for
the triangle △xpy. Since |xy| ≤ |x, y|, we obtain ∡xpz + ∡ypz ≤ ∡xpy + ε.
With the obvious equality ∡xpz = ∡ypz, it gives
1 1
∡xpz ≤ ∡xpz ≤ (∡xpy + 2ε), ∡xpz ≤ ∡ypz ≤ (∡xpy + 2ε)
2 2
if points the x and y are sufficiently close to p.
Let c denote the point of Σ′p represented by a shortest path pz. Now the
last two inequalities can be re-written as
1 1
d(a, c) ≤ (d(a, b) + 2ε), d(b, c) ≤ (d(a, b) + 2ε).
2 2
By the triangle inequality for angles we have
These three inequalities immediately imply that |d(a, c) − 12 d(a, b)| ≤ 2ε,
|d(b, c) − 21 d(a, b)| ≤ 2ε. Since ε is arbitrary, this proves the midpoint
property, so the metric of Σ′p is locally intrinsic. ¤
Tangent cone. Denote by Kp the cone over the direction space Σp (see
Subsection 3.6.2 for the definition of the cone over a length space). Every
point w ∈ Kp except its origin o is represented by a pair (ξ, r), where ξ ∈ Σp
and r = |ow|. The cone Kp is called the tangent cone at p. It is clear that
for a smooth surface in R3 all tangent cones are planes.
Proof. The idea of the proof is very simple. One takes a triangle in Kp and
considers a sufficiently small homothetic one (with respect to a homothety
centered at the origin). The map expp allows us to associate a triangle in X
to a small triangle in Kp . These two triangles have almost the same lengths
of the respective sides (to the first order), so the curvature comparison
condition translates from the triangle in X to the one in Kp .
To make it more clear, let us assume for now that Kp is locally compact
and that each point of Kp is represented by a shortest path. Let △a1 a2 a3
be a triangle in Kp and a4 the midpoint of the side [a1 a3 ]. Denote by γi ,
i = 1, 2, 3, 4, the shortest paths representing points ai and by aλi the point of
the shortest path γi such that |paλi | = λ|Oai |. Here O is the origin of Kp and
λ is a small positive number. By the very definitions of the angular metric
and the cone’s metric, we have λ−1 |aλi aλj | ∼ |aj aj | as λ → 0 (in the sense that
9.1. Definitions and Local Properties 323
the ratio of the two quantities converges to 1). For a small λ, consider the
comparison triangle △λ for △aλ1 aλ2 aλ3 in the k-plane and the dilated triangle
λ△λ . The latter one has sides that converge to those of △a1 a2 a3 and lies
in the λ−1 k-plane (whose curvature goes to zero as λ → 0). Therefore the
curvature comparison conditions for △aλ1 aλ2 aλ3 and a suitable (close to aλ4 )
point of [aλ1 aλ3 ] turn in the limit (as λ → 0) to the same condition (but with
zero curvature bound!) for △a1 a2 a3 and a4 . This proves the theorem with
our special assumptions.
To get a proof in the general case, use the “generalized” Definition 9.1.2.
The same argument works except that a4 should be an ε-midpoint instead
of the real midpoint (which may not exist), and each point ai should be
replaced by a sufficiently close point represented by a shortest path. ¤
Combining the above theorem with Theorem 4.7.1 about cones over
length spaces yields the following
Corollary 9.1.45. Under the conditions of the theorem, the space of direc-
tions Σp at every point p ∈ X is a space of curvature ≤ 1.
simply connected). Recall (see Remark 9.1.18) that such a space is locally
simply connected and therefore has a universal covering; i.e., there exist
a simply connected topological space X̃ (the so-called universal covering
space) and a covering map f : X̃ → X.
The metric of X is (canonically) lifted to X̃ so that the covering map
becomes a local isometry. Since being nonnegatively curved is a local
property and f is a local isometry, X̃ is nonpositively curved as long as
X is. And a covering space of a complete space is complete (cf. Exercise
3.4.8). Thus X̃ is a Hadamard space.
One can apply general properties of Hadamard spaces to the universal
covering space X̃ and then derive information about the original space X.
This section and Section 9.3 contain a number of statements whose proofs
work this way.
The next lemma says that, given a geodesic segment γ connecting points
p and q, for every point q ′ near q there exists a unique geodesic connecting
p and q ′ and passing near γ. Note that there may be other geodesics
connecting p and q ′ but passing “far away” from γ; moreover the constructed
geodesic may be not a shortest path even if γ is. A reader familiar with
Riemannian geometry will notice that this lemma is an counterpart of the
fact that geodesics in a nonpositively curved Riemannian manifold have no
conjugate points.
Lemma 9.2.4. Let X be a complete locally compact space of nonpositive
curvature, γ : [0, 1] → X be a constant-speed geodesic with endpoints p =
γ(0), q = γ(1). Let r > 0 be so small that the convexity radius of γ([0, 1])
(cf. subsection 9.1.3 for the definition) is greater than 10r.
Then for every q ′ ∈ Br (q) there exists a unique constant-speed geodesic
α : [0, 1] → X such that α(0) = p, α(1) = q ′ , and the uniform distance
between γ and α is less than r, i.e., |γ(t)α(t)| < r for all t.
Remark 9.2.5. It is easy to see that the geodesic α depends continuously
on q ′ (prove this as an exercise).
The idea of the proof is simple. Consider a class of “broken lines” (that
is, curves composed of several shortest paths) with a fixed number of vertices
connecting p and q ′ , having short edges and passing in a neighborhood of γ.
9.2. Hadamard Spaces 327
We will show that a broken line of minimal length in this class is the desired
geodesic.
enough” length space (more precisely, one with locally unique shortest paths)
is a covering map. Since X is nonnegatively curved, it is “good enough” in
this sense, so it remains to show that Xp (with the length metric lifted from
X) is complete.
By Hopf–Rinow Theorem 2.5.28, it is sufficient to prove that, for some
P ∈ Xp , any constant-speed geodesics Γ : [0, 1) → Xp emanating from P
can be extended to the closed interval [0, 1].
Let P be the constant geodesic (speed 0) resting at p. Then every
geodesic γ : [0, 1) → X with γ(0) = p has a unique lift Γ : [0, 1) → Xp with
Γ(0) = P . (Being a lift means that γ = eg xpp ◦ Γ.) This lift Γ is defined in
a natural way: for every t ∈ [0, 1), the point Γ(t) in the space of geodesics
is nothing but the (reparameterized) restriction γ|[0,t] . The uniqueness of a
lift follows from the fact that eg
xpp is a local homeomorphism.
It is easy to extend such a lift Γ to [0, 1]. Extend γ to [0, 1] (this is
possible since X is complete) and take the lift of the extended curve. In
other words, define Γ(1) = γ|[0,1] . Since every geodesic in Xp is a lift of
some geodesic in X (namely, of its own image under eg xpp ), we have proved
the desired sufficient condition for Xp to be complete. The theorem then
follows as explained above. ¤
9.2.2. Globalization.
Proof of the theorem. We will prove that the angle condition holds for
all triangles. (This implies that all other curvature conditions hold globally,
because the “global” versions of the definitions of bounded curvature are
equivalent just like the local ones.)
1. The key observation is the following: if △abc is a triangle in X, d is
a point in its side [ac], and the angle condition holds for the triangles △abd
and △cbd, then it holds for △abc.
Indeed, place comparison triangles △abd and △cbd for △abd and △cbd
in the k-plane in different half-planes with respect to their common side bd.
From the angle condition for △abd and △cbd we have ∡bda ≤ ∡bda and
∡bdc ≤ ∡bdc; hence ∡bda + ∡bdc ≥ ∡bda + ∡bdc ≥ π. Therefore the angle
at d of the quadrilateral abcd is not less than π. Using Alexandrov’s Lemma
4.3.3 we obtain that ∡bad ≤ ∡bad ≤ b1 a1 c1 , where △b1 a1 c1 is a comparison
triangle for △bac. Similarly ∡bcd ≤ ∡b1 c1 a1 . Finally,
∡abc ≤ ∡abd + ∡cbd ≤ ∡abd + ∡cbd = ∡abc ≤ ∡a1 b1 c1
(the last inequality follows from the fact that |ac| ≤ |ad| + |cd| = |a1 c1 |).
2. We say that a triangle △abc is slim if its edges [ab] and [ac] are very
close to each other; more precisely, the (uniform) distance between those
sides is less than r/10 where r is the convexity radius of [ab] ∪ [ac] (so every
r-ball centered in a point of [ab] ∪ [ac] is a normal region).
c
PSfrag replacements
a
ba
c b
Let us “cut” such a slim triangle △abc into small triangles as shown in
Figure 9.2. Each small triangle is contained in a normal ball and therefore
satisfies the angle condition. Then the angle condition for △abc follows by
induction on the number of small triangles from the fact proved in the first
step.
3. Thus the angle condition is satisfied for all slim triangles. Now
consider an arbitrary triangle △abc. Since shortest paths in X are unique
9.2. Hadamard Spaces 331
the closure of Y . Then dist(z, Y ) ≤ |zz ′ | ≤ 12 (|xx′ | + |yy ′ |). Here the second
inequality follows from Proposition 9.2.13. ¤
Corollary 9.2.16. Let X be a Hadamard space and i : X → X an isometry
of X. Then the “displacement function” δ(x) = d(x, i(x)) is convex.
Theorem 9.2.19. For every point p in a Hadamard space, the function d2p
is 1-convex.
Conversely, if X is a complete space with strictly intrinsic metric and
d2p is 1-convex for every p ∈ X, then X is a Hadamard space.
Proof. In the Euclidean plane, the function d2p has the form t 7→ t2 + const
along a unit-speed straight line. Since in a Hadamard space the distance
function is “more convex” than in the plane (this is the distance condition,
our first definition of nonpositive curvature), d2p is “more convex” than t2 .
Here is the same argument filled in with formulas. By Proposition 9.2.18,
it is sufficient to verify that
Proof. The uniqueness part is easy. If f (x) = f (y) = min f and x 6= y, let
z be a midpoint between x and y, and apply Proposition 9.2.18. This yields
f (z) < 21 (f (x) + f (y)) = min f , contradiction.
Now let us prove that a point of minimum exists. Let m = inf f and
{xi }∞
i=1 be a minimizing sequence for f , i.e., f (xi ) → m as i → ∞. We are
going to show that {xi } is a Cauchy sequence.
334 9. Spaces of Curvature Bounded Above
Fix an ε > 0 and let i0 be so large that f (xi ) < m + ε for all i > i0 .
Then for all i, j > i0 and z ∈ X we have
f (xi ) + f (xj )
f (z) ≥ m > − ε.
2
Taking a midpoint between xi and xj for z and subtracting the inequality
from
p Proposition 9.2.18, we obtain that λ|xi xj |2 /4 < ε. Thus |xi xj | <
4ε/λ for all i, j > i0 . Since λ is fixed and ε is arbitrarily small, it follows
that {xi } is a Cauchy sequence.
Now define x = lim xi . Then f (x) = lim f (xi ) = m, so x is the desired
minimum point. ¤
9.2.4. Parallel rays and lines. Recall that a line in a length space X is
a (unit-speed) geodesic γ : R → X such that any closed subinterval of γ
is a shortest path. A ray is a geodesic γ : R+ → X possessing the same
property. In Hadamard spaces, every geodesic segment is a shortest path,
so lines and rays are just complete geodesics of infinite length.
As in Euclidean and hyperbolic spaces, one can introduce a notion of
parallel rays and lines in a Hadamard space.
Definition 9.2.26. Two rays or two lines are said to be parallel if the
uniform distance between them is finite. Parallel rays are called asymptotic
as well.
In the Euclidean plane two rays or lines are parallel if and only if they
are parallel in ordinary (“scholar”) sense. It is easy to see that if two lines
in the Lobachevsky plane are parallel, then they coincide (but parallel rays
starting at points a, b in the Lobachevsky plane do exist for every pair of
points a, b). We will see soon that in the general case the situation is very
similar.
Proposition 9.2.28. Let X be a Hadamard space and p ∈ X. Then for
every ray γ in X there exists a unique ray starting at p and parallel to γ.
Proof. We prove this fact only for locally compact spaces. Let γt denote the
shortest path connecting p to γ(t). Since the distance between geodesics is a
convex function (Proposition 9.2.13), this shortest path is contained in the
(closed) r-neighborhood of γ where r = |pγ(0)|. Since the space is locally
compact, one can choose a sequence {ti }, ti → ∞, such that {γti } converges
to some ray α : R+ → X. This ray stays within the r-neighborhood of γ
and hence is parallel to γ.
To prove the uniqueness, suppose that there are two rays α and β starting
at p and parallel to γ. Since being parallel is an equivalence relation, the
rays α and β are parallel, so the function t 7→ |α(t)β(t)| is bounded. On the
other hand, this function is convex and equals zero at t = 0; therefore it is
zero for all t. This means that α = β. ¤
With the notion of parallel rays, one defines the ideal boundary of a
Hadamard space similarly to the ideal boundary of the Lobachevsky plane.
Namely, the points of the ideal boundary are equivalence classes of parallel
rays. The ideal boundary of a Hadamard space X is denoted by X(∞). It is
common to say that a ray “connects” its initial point with the point of the
336 9. Spaces of Curvature Bounded Above
Proof. Draw the “diagonal” [ac] in the quadrangle Q. The angle condition
for nonpositive curvature implies that the sums of angles of △abc and △adc
are not greater than π. On the other hand, the triangle inequality for angles
implies that the sum of all six angles in △abc and △adc is not less than the
sum of angles of Q and hence is no less than 2π. Therefore all mentioned
inequalities turn to equalities; in particular, the sum of angles of Q equals 2π,
and in both triangles △abc and △acd the sum of angles equals π. Therefore
(by Proposition 9.1.19) each of these triangles bounds a totally geodesic flat
surface (a “solid” triangle). Denote these surfaces by T1 and T2 , respectively.
We are going to show that T1 ∪ T2 is a desired quadrangle.
By a similar argument, the triangles △abd and △cbd bound flat solid
triangles T3 and T4 . To finish the proof, it is sufficient to show that
T1 ∪ T2 = T3 ∪ T4 , or, equivalently, that the shortest path [ac] lies in the
surface T3 ∪ T4 . Observe that both surfaces T1 ∪ T2 and T3 ∪ T4 are arcwise-
isometric to a quadrangle abcd whose sides and angles equal the respective
sides and angles of Q. Then the length of [ac] equals the diagonal ac, and
this diagonal corresponds to a curve of the same length in T1 ∪ T2 . Thus
there is a curve in T3 ∪ T4 whose length equals |ac|. This means that this
9.2. Hadamard Spaces 337
curve is a shortest path and, since shortest paths in a Hadamard space are
unique, coincides with [ac]. ¤
so the point β(t) is nearest to α(t) in β, and the point α(t) is nearest to β(t)
in α. Then by the first variation formula (Theorem 4.5.6) all four angles
that a shortest path [α(t)β(t)] forms with half-lines of α and β are no less
than π/2. Hence the quadrangle α(0)β(0)β(t)α(t) satisfies the assumption
of Lemma 9.2.30, so its angles equal π/2 and it bounds a flat surface. The
union of such rectangles over all t ∈ R is the desired flat strip. ¤
Therefore |β(t)p| > t + t0 /2 for all large enough t. We are going arrive
at a contradiction by showing that there is a path from β(t) to p (passing
through a point of γ) whose length is less than t + t0 /2.
By Theorem 9.2.29 the lines γ and α bound some flat strip S1 , and
lines γ and β bound some flat strip S2 . Consider the union S = S1 ∪ S2
of these strips. It is arcwise-isometric to a planar strip whose width is the
sum of widths of S1 and S2 (more precisely, S is the image of this planar
strip under an arcwise isometry). Furthermore, the segment between points
corresponding to p = α(0) and β(0) in the planar strip is orthogonal to its
boundary lines. Reasoning as above, we conclude that
t − |β(t)p|s = |β(t)β(0)|s − |β(t)p|s −−−−→ 0
t→+∞
Let us recall the basic facts from Subsection 3.4.2 about fundamental
groups and universal coverings. Let X be a length space and f : X̃ → X
be the universal covering map. Then the fundamental group π1 (X) acts on
X̃ by isometries; more precisely, π1 (X) is isomorphic to the group of deck
transformations, that is, of maps from X̃ to itself commuting with f .
In other words, the isometry group Iso(X̃) contains a subgroup iso-
morphic to π1 (X); moreover this subgroup acts freely (i.e., every nontrivial
9.3. Fundamental Group of a Nonpositively Curved Space 339
element of this subgroup is a map without fixed points) and totally discon-
tinuously.
If X is a complete nonpositively curved space, then X̃ is a Hadamard
space. Thus, instead of fundamental groups of such spaces X, one can study
groups acting by isometries on a Hadamard space. In this case the action is
assumed to be free and totally discontinuous.
Since deck transformations have no fixed points, the theorem yields the
following
First of all we want to show that β̃ = α̃. Suppose it is not the case.
Take a point x of the geodesic α̃, and consider the quadrangle Q with the
vertices x, x1 = ax, x2 = bx and x3 = abx = bax.
Since a, b are isometries, we have equalities
(a1 , a2 , . . . , aN ) = (x1 , y1 , z1 , x2 , y2 , z2 , . . . , xN , yN , zN ).
N
X
K((v1 , v2 , . . . , vN ), (v1 , v2 , . . . , vN )) = mi hvi , vi i ,
i=1
seem very useful, it already can deliver certain information. For instance,
the Liouville theorem (invariance of the Liouville measure) for billiard flows
follows immediately from the Liouville theorem for geodesic flows.
Let us illustrate Arnold’s suggestion by a simple example of the billiard
table in the complement of a disc in a two-torus (or the Euclidean plane).
Starting with two copies of the torus with (open) discs removed, and gluing
them along the boundary circles of the discs, one obtains a Riemannian
manifold (a surface of genus 2) with a metric singularity along the gluing
circle. This manifold is flat everywhere except at this circle. One can think
of this circle as carrying singular negative curvature, and indeed this is a
nonpositively curved length space. Smoothing this metric by changing it in
an (arbitrarily small) collar around the circle of gluing, one can obtain a
smooth nonpositively curved Riemannian metric, which is flat everywhere
except in this collar. To every segment of a billiard trajectory, one can
(canonically) assign a geodesic in this metric. Collisions with the disc would
correspond to intersections with the circle of gluing, where the geodesic
leaves one copy of the torus and goes to the other one.
Exercise 9.4.2. Repeat the same construction in dimension three: consider
two copies of R3 with a unit ball removed, and glue them along the boundary
spheres. Show that this is a space of curvature bounded above by 1, and it
is not a nonpositively curved space!
Hint: In this space, the gluing locus (which is a sphere) is a totally
geodesic subspace.
Show that the presence of positive curvature in this example persists un-
der smoothenings that do not change the metric outside some neighborhood
of the gluing locus.
This exercise shows that there are serious problems with using this
construction in higher dimensions. Moreover, even in dimension two many
geodesics do not correspond to billiard trajectories. They can be described
as coming from “fake” trajectories hitting the disc at zero angle, following
an arc of its boundary circle (possibly even making several rounds around
it) and then leaving it along a tangent line. Dynamically, such geodesics
carry “the main portion of entropy” and they cannot be disregarded. On
the other hand, it is difficult to tell actual trajectories from the fake ones
when analyzing the geodesic flow on this surface.
Main construction. The purpose of this section is to give an informal
and elementary account of how Arnold’s idea can nonetheless be formalized
(and the difficulties mentioned above can be partially avoided) by using
Alexandrov spaces. Several important open problems in the area were
recently solved by this method.
9.4. Example: Semi-dispersing Billiards 345
geodesics, which go through the disc. However, there are fewer of them than
before and it is easier to separate them.
It might seem more natural to glue along the boundaries of Wni rather
than along the whole Bni . For instance, one would do so thinking of this
gluing as “reflecting in a mirror” or by analogy with the usual development
of a polygonal billiard. However, gluing along the boundaries will not give
us a nonpositively curved space in any dimension higher than 2 (see Exercise
9.4.2).
One may wonder how the interiors of Bi ’s may play any role here, as they
are “behind the walls” and billiard trajectories never get there. For instance,
instead of convex walls in a manifold without boundary, one could begin with
a manifold with several “convex” boundary components (with a nonnegative
definite second fundamental forms with respect to the inner normal). Even
for one boundary component, there are examples where it is impossible to
“fill in” the boundary by a nonpositively curved manifold. Moreover, our
main dynamical result does fail for an example of this sort. Thus, it is
indeed important that the walls are not only locally convex surfaces, and we
essentially use the fact that they are filled by convex bodies.
Sketch of Proof for Theorem 9.4.5 for two walls. Consider a trajec-
tory γ making N collisions with the (only possible) sequence of walls
K = {1, 2, 1, . . . , 2, 1}. Reasoning by contradiction, assume that N > 3P +1,
where P is the local bound on the number of collisions (whose existence is
guaranteed by Theorem 9.4.4). Again consider the space MK , but now we
will “close it up” by gluing M0 ∈ MK and MN ∈ MK along the copies of
B1 . Denote the resulting space by M̃ . We cannot use Reshetnyak’s Theo-
rem 9.1.21 directly to conclude that M̃ is a nonpositively curved space any
more, since we identify points in the same space and we do not glue two
spaces along a convex set.
We recall that a space has nonpositive curvature if and only if every
point possesses a neighborhood such that, for every triangle contained in
the neighborhood, its angles are no bigger than the corresponding angles
of the comparison triangle in the Euclidean plane. However, using the
correspondence between geodesics and billiard trajectories, one can conclude
(reasoning exactly as in the proof of the local estimates on the number of
collisions) that each side of a small triangle cannot intersect interiors of
more than P copies of the billiard table. Since N > 3P + 1, for every
small triangle for which we want to verify the angle comparison property,
we can undo one of the gluings without tearing the sides of the triangle. This
ungluing may only increase triangle’s angles, and now we find ourselves in
a nonpositively curved space (which is actually just MK ), and thus we get
the desired comparison for the angles of the triangle.
To conclude the proof, it remains to notice that the development of γ
in M̃ is a geodesic connecting two points in the same copy of B1 . This
is a contradiction since every geodesic in a simply connected nonpositively
curved space is the only shortest path between its endpoints; on the other
hand, there is a shortest path between the same points going inside this
copy of B1 . ¤
Chapter 10
Spaces of Curvature
Bounded Below
351
352 10. Spaces of Curvature Bounded Below
the triangle comparison conditions hold “in the large”. This yields a number
of results about global geometry of Alexandrov spaces (recall from Chapter 9
that one can say much more about a Hadamard space than about a general
space of curvature bounded above). Furthermore, Toponogov’s Theorem
implies that the class of Alexandrov spaces of curvature ≥ k is closed with
respect to Gromov–Hausdorff convergence and this allows us to study this
class “as a whole” (recall the discussion in the beginning of Chapter 7).
Second, (finite-dimensional) Alexandrov spaces have nice local structure.
By dimension we mean Hausdorff dimension (cf. Subsection 1.7.4) but in
fact all known notions of dimension are equivalent for an Alexandrov space
(in particular, Hausdorff dimension equals topological dimension). Some
of the local properties of finite-dimensional Alexandrov spaces are worth
mentioning right now. Hausdorff dimension of such a space is always an
integer; such a space is a manifold (in fact, almost a Riemannian one)
everywhere except a tiny set of singular points; the space of directions of an
n-dimensional Alexandrov space is an (n − 1)-dimensional Alexandrov space
of curvature ≥ 1 and in fact is isometric to S n−1 at almost every point.
Though these formulations are nice and clear, known proofs are long
and technical. We do not get into these details until the end of the chapter
(Sections 10.8 and 10.9). Some knowledge of the local structure is, however,
required in earlier sections; in such cases we refer to the appropriate results
from the last sections.
Proof of Proposition 10.1.1. First suppose that (10.1) holds for every
quadruple. To verify the triangle condition (from Definition 4.1.9) for a
triangle △abc and d ∈ [ac], apply the quadruple condition to (d; a, b, c).
e
Since ∡adc e
= π, it follows that ∡bdc e
+ ∡bda ≤ π. Then Lemma 4.3.3
implies the desired inequality |db| ≥ |db|.
Now let X be a space of curvature ≥ k. Consider a quadruple (a; b, c, d)
and a point a′ in a shortest path [ab]. Then
e ′ d + ∡da
∡ba e ′ c + ∡ca
e ′ b ≤ ∡ba′ d + ∡da′ c + ∡ca′ b
≤ (∡ba′ d + ∡da′ a) + (∡aa′ c + ∡ca′ b) = 2π.
We used the angle condition (Definition 4.1.15), the triangle inequality for
angles, and the fact that the sum of adjacent angles equals π (Lemma 4.3.7).
Now let a′ converge to a; then (10.1) follows by continuity of comparison
angles. ¤
√
If k > 0, all pairs (x, π/ k) should be identified (because the distance
between them is zero). These pairs represent a point o′ ∈ Conk (X) √ which
could be taken as the origin instead of o (the map (x, r) 7→ (x, π/ k − r)
is an isometry of Conk (X)). In this case the k-cone is also called the k-
spherical cone or the k-suspension. The origins o and o′ are referred to
as its poles. The standard spherical suspension defined in Subsection 3.6.3
corresponds to k = 1.
If k < 0, k-cones are also called (−k)-hyperbolic cones. A hyperbolic
cone is a k-cone for k = −1. If k = 0, k-cones are ordinary (“Euclidean”)
cones.
By the very definition, Conk (S 1 ) is isometric to the k-plane. Similarly,
Conk (S n ) is the standard (n+1)-dimensional
√ space form of curvature k, i.e.,
an n-dimensional sphere of radius 1/ k, a rescaled hyperbolic space, or Rn ,
depending on the sign of k.
The primary use of cones in this chapter is to consider cones over spaces
of directions. Note that a space X can be recovered as the space of directions
of the cone Conk (X) at the origin.
The results of Section 4.7 about curvature bounds of cones can be
simplified for Alexandrov spaces in view of Toponogov’s Theorem 10.3.1
(saying that the triangle comparison conditions hold in the large) and
Corollary 10.4.2 (saying that the perimeter of every triangle in a space of
curvature ≥ 1 is no greater than 2π).
Theorem 10.2.3. Let X be a complete metric space and k ∈ R.
1. If X is an Alexandrov space of curvature ≥ 1, then Conk (X) is an
Alexandrov space of curvature ≥ k.
2. If Conk (X) is an Alexandrov space of curvature ≥ k and, in addition,
|xy| + |yz| + |xz| ≤ 2π for all x, y, z ∈ X, then X is an Alexandrov space
of curvature ≥ 1 or a space consisting of two points at distance π from each
other.
The proof of this theorem is in no way different from that of Theorem
4.7.1. Note that in the second part of the theorem we do not assume in
advance that X is a length space. Due to the other assumption, this property
follows from the fact that Conk (X) is a length space. Nor do we assume
that X is connected—unlike the case of curvature bounded above, a cone
over X cannot have a lower curvature bound if X is not connected (except
the trivial case of a two-point space).
Remark 10.3.2. Recall that “in the large” means that the curvature
comparison conditions are satisfied for all triangles for which comparison
triangles exist and are unique. The latter requirement on triangles is
essential only if k > 0 and is approximately
√ equivalent to the statement
that the perimeter is not greater than 2π/ k. In Subsection 4.1.15 we will
√ of curvature ≥ k cannot contain a
show that, in fact, an Alexandrov space
triangle of perimeter greater than 2π/ k.
Proof of Theorem 10.3.1. The global versions of the definitions are equiv-
alent just like the local ones. We will show that the “angle condition” 4.1.15
is satisfied for every triangle. Note that angles between shortest paths do
exist and the sum of adjacent angles is equal to π because these properties
are local.
We present a proof for the case k ≤ 0 and then explain how to modify
the argument for the case k > 0.
Step 1. Suppose that the theorem is false. Then there exists a triangle
△pqr satisfying the following:
e
(i) The angle condition fails for its angle at q, i.e., ∡pqr < ∡pqr.
10.3. Toponogov’s Theorem 361
Hence the triangles △adx1 , and △bdx2 in Figure 10.3 have no common
interior points and
∡x1 dx2 = ∡adb − ∡adx1 − ∡bdx2 ≥ ε.
Since |ax3 | = |ax1 | and |bx3 | = |bx2 |, it then follows that △abx3 has no
common interior points with △adx1 , and △bdx2 .
d
PSfrag replacements
x1
x1 gaga x2
x2
x3 al be
a al be
b x3
d
α
β
γ γa b
e
∡axb − ∡axb = ∡ax3 b − ∡axb > ε.
ε
On the other hand, |ax| + |bx| ≤ |ad| + |db| − 10 |dx| if |dx| is small enough.
Step 5. Consider the set S of all triangles △ayb (with a and b fixed)
such that
e
(i) ∡ayb − ∡ayb ≥ ε, and
(ii) max{|ay|, |by|} ≤ 0.26 R.
We supposed that this set was nonempty. Since our space is locally
compact, there is a triangle △abd ∈ S with
However Step 4 shows that one can find a point x (close to d) such that
△axb ∈ S and |ax| + |xb| < |ad| + |db|. This contradiction proves the
theorem for the case k ≤ 0.
Step 6. If k > 0, the same scheme works with certain modifications.
First, one has to check carefully that the involved comparison triangles are
well-defined. Second, in the inequalities that follow from the Gauss–Bonnet
formula in Step 4, one cannot drop the integral curvature of the triangles
△x1 x2 x3 and dx1 x2 . As a result, one loses the condition (i) in Step 5 when
△adb is replaced by △axb. To work this around, replace the condition (i)
in Step 5 by a more complicated one, namely,
e 1
(10.4) (∡ayb − ∡ayb) − c(|ay| + |yb|) ≥ ε,
2
where c is a sufficiently small positive constant. Then the loss in the first
term of (10.4) is compensated by the gain in the second one. The details
are left to the reader as an exercise .
The reader who has checked all these details might have noticed that √
the argument works only for triangles of perimeter strictly less than 2π/ k.
√ out the case of a triangle △abc with perimeter
So we have to work √ |ab| +
|bc| + |ac| = 2π/ k, and prove that ∡abc = π if max{|ab|, |bc|} < π/ k. Fix
a positive ε < k. Observe that a space of curvature ≥ k is also a space of
curvature ≥ k − ε. Applying the already proven part of the theorem to X as
a space of curvature ≥ k − ε, we obtain that ∡abc ≥ ∡ e k−ε abc. As ε → 0, a
comparison triangle for △abc in the (k − ε)-plane converges to a comparison
e k−ε abc → ∡
triangle in the k-plane. Therefore ∡ e k abc = π as ε → 0, implying
that ∡abc ≥ π. ¤
364 10. Spaces of Curvature Bounded Below
Proof. Suppose
√ the contrary, and let a, b ∈ X√ be two points such that
|ab| > π/ k. We may assume that |ab| = (π+ε)/ k where√0 < ε < π/4. Let
c be the midpoint of a shortest path [ab] and U be an (ε/3 k)-neighborhood
of c.
First we show that U contains a point which does not belong to [ab].
Indeed, suppose the contrary. For every point x ∈ X consider a shortest
path γ connecting c to x. Our assumption implies that this shortest path
coincides with a subinterval of [ab] in a neighborhood of c. Since geodesics do
not branch (Exercise 10.1.2), it follows that x belongs to the unique geodesic
containing [ac]. Therefore X is covered by two minimal geodesics (that
is, geodesics whose intervals are shortest paths) starting at c and passing
through a and b respectively. Depending on whether these geodesics are
finite or infinite and whether their endpoints coincide, we conclude that X
is one of the one-dimensional exceptional spaces.
Choose√an x ∈ U \ [ab], and
√ let y be a nearest point to x in [ab]. Then
|ay| > π/2 k and |by| > π/2 k. The first variation formula implies (see
Corollary 4.5.7) that ∡xya = ∡xyb = π/2. Consider a comparison√triangle
△xya for △xya in the k-plane which is the sphere of radius 1/ k. √By
Toponogov’s Theorem we have ∡xya ≤ ∡xya = π/2. Since |ya| > π/2 k,
it follows that |xa| < |ya|. (To see this, place a at the north pole of the
sphere; then the inequality ∡xya ≤ π/2 implies that x lies above the great
circle passing through y orthogonally to the meridian [ax]. Since y is strictly
10.4. Curvature and Diameter 365
below the equator, y is the lowest point of this great circle; hence y is strictly
lower than x, i.e., |xa| < |ya|.) Thus |xa| < |ya| and similarly |xb| < |yb|.
Hence |ya| + |yb| > |xa| + |xb| ≥ |ab|, contrary to the fact y belongs to a
shortest path [ab]. ¤
Then prove the following facts. First, for every x, y ∈ Y , shortest paths [px]
and [qx] are unique and ∡pxy = ∡qxy = ∡pyx = ∡qyx = π/2. Then the
triangles △pxy and △qxy (as unions of shortest paths) are isometric to their
comparison triangles in the sphere. Then Y is convex in X and therefore Y
is a length space of curvature ≥ 1. Finally, every point x ∈ X belongs to
some shortest path connecting p and q, and therefore to the union [py] ∪ [qy]
for some y ∈ Y . These facts imply that X is the spherical cone over Y .
Another proof is based on the splitting theorem from the next section;
see Remark 10.5.8.
366 10. Spaces of Curvature Bounded Below
Lemma 10.5.4. Let γ be a line in X and p ∈ X. Then for every two points
x, y ∈ γ the angles at x and y of a triangle △pxy are equal to the respective
e
angles of its comparison triangle △pxy.
Proof. Suppose that there are two lines, γ1 and γ2 , passing through p
and parallel to γ. Consider three corresponding parallel lines γ 1 , γ, and
γ 2 in the plane R2 . We assume that γ lies between the two other lines.
Let points a ∈ γ 1 , b ∈ γ 2 correspond to points a ∈ γ1 , b ∈ γ2 such that
|pa| = |pb| and let the segment [ab] intersect the line γ at point c. Then
|ab| ≤ |ac| + cb| = |ac| + |cb| = |ab|. Applying Lemma 10.5.4 to the point a
and the line γ2 , one sees that |ab| → ∞ as |pa| = |pb| → ∞. The latter is
impossible because |ab| is a constant.
The second assertion of the lemma is proved similarly. The only differ-
ence is that one should consider four parallel lines in the plane. ¤
Proof of the theorem. Step 1. Let γ be a straight line. First of all, for
every point p ∈ X \ γ, we construct a straight line γp passing through p and
parallel to γ. To do this, denote ci = γ(ti ) and consider shortest paths [pci ]
where ti → ∞ as i → ∞. There is a subsequence of these paths converging
to a geodesic ray γp+ : [0, ∞) → X. Similarly, taking points di = γ(−ti ),
368 10. Spaces of Curvature Bounded Below
ti → ∞, one can find a geodesic ray γp− : [0, −∞)) → X. We are going to
show that these two rays together form a straight line γp parallel to γ.
The comparison angles ∡c e i pdi obviously approach π as i → ∞. Since
e i pdi , this implies that angles ∡ci pdi converge to π as well. Hence
∡ci pdi ≥ ∡c
the angle between γp+ and γp− at p equals π.
Now let us prove that γp is a line. Fix a t > 0 and let b+ −
i and bi be
the points in the shortest paths [pci ] and [pdi ] at the distance t from p.
These points converge to the points b+ = γp (t), b− = γp (−t). Consider a
+ −
comparison triangle △ci adi and points bi , bi in its sides at distances t of
+ −
a. It is obvious that |b+ b− | = limi→∞ |b+ −
i bi | ≥ limi→∞ |bi bi | = 2t. This
means that each segment of γp is a shortest path; hence γp is a line.
Now we prove that γ and γp are parallel. Consider two lines γ, γp at the
distance |pp′ | in a plane R2 where p′ is the projection of p to γ. Let p ∈ γ p
and p′ be the projection of p to γ. Then the limit process used above and
Corollary 10.5.5 imply that for every two points a ∈ γ, b ∈ γa and points
a ∈ γ, b ∈ γa , lying in the corresponding rays and such that |pa| = |pa|,
|p′ b| = |p′ b|, the equality |ab| = |ab| holds. This means that lines γ and γp
are parallel.
Note that we have proved that for every sequence {ti }, ti → +∞,
shortest paths [pγ(ti )] converge to a half-line of a geodesic parallel to γ
if they converge at all. Since such a geodesic is unique (Lemma 10.5.6),
it follows that these shortest paths converge to this half-line γp+ for every
sequence {ti }, ti → +∞. Moreover the angles between these paths and γp+
go to zero.
Step 2. Applying Lemma 10.5.6, we see now that X is split into parallel
lines, and we have the relation “one is the projection of the other” between
points. We are going to show that this relation is symmetric and transitive.
To prove that this relation is symmetric, consider two parallel lines γ1
and γ2 and points a1 ∈ γ1 , a2 ∈ γ2 such that a2 is the projection of a1 to γ2 .
Consider triangles a1 a2 γ(t) as |t| → ∞. Since ∡a1 a2 γ(t) = ∡a e 1 a2 γ2 (t) =
π/2 and |a2 γ2 (t)| → ∞, we have ∡a2 a1 γ2 (t) = ∡a e 2 a1 γ2 (t) → π/2. Therefore
[a1 a2 ] is orthogonal to γ1 , or, equivalently, a1 is the projection of a2 to γ1 .
To prove that the projection relation is transitive, consider three parallel
lines γ1 , γ2 , γ3 and let a2 and a3 be the projections of a point a1 ∈ γ1 to γ2
and γ3 , respectively. Then
|a2 γ1 (t)| − |a3 γ1 (t)| → 0, t → +∞,
because |a2 γ1 (t)|−|a1 γ1 (t)| → 0 and |a3 γ1 (t)|−|a1 γ1 (t)| → 0. It follows that
e 2 a3 γ(t) → π/2, and therefore [a2 a3 ] is orthogonal to γ + = lim[a3 , γ1 (t)].
∡a 3
Hence a3 is the projection of a2 to γ3 .
10.6. Dimension and Volume 369
Step 3. Now it is not difficult to finish the proof of the theorem. For a
point p ∈ γ denote by N (p) the set of projections of p to all lines parallel
to γ. This set is convex in X and hence is a length space of nonnegative
curvature. Every two such sets N (p) and N (q) are canonically isometric via
an isometry defined as follows: if x ∈ N (p), then the image of x is N (q) ∩ γx ,
where γx is the line parallel to γ and passing through x.
Fixing p = γ(0), consider the set N (p)×R1 and the map f : N (p)×R1 →
M such that f (x, t) is the point of N (γ(t)) which corresponds to x under
the canonical isometry between N (p) and N (γ(t)). It is clear that f is an
isometry. ¤
Proof. Every open set U contains a ball Br (p) for some p ∈ X and r > 0,
and dimH (Br (p)) ≤ dimH (U ) ≤ dimH (X) because Br (p) ⊂ U ⊂ X. It
is sufficient to prove that dimH (Br (p)) = dimH (X). We will show that
dimH (Br (p)) = dimH (BR (p)) for every R > r. The theorem follows from
this, because X is can be represented as a union of countably many balls of
the form BR (p); for example, take R = r + 1, r + 2, . . . . If the dimensions
of these balls equal dimH (Br (p)), it follows that dimH (X) = dimH (Br (p));
cf. Proposition 1.7.19.
Denote A = Br (p) and B = BR (p). The inequality dimH (A) ≤ dimH (B)
is trivial because A ⊂ B. To prove that dimH (B) ≤ dimH (A), we construct
a map f : B → A such that for some c > 0 one has |f (x)f (y)| ≥ c|xy| for
all x, y ∈ A. If such a map f exists, its inverse f −1 : f (B) → B is Lipschitz;
therefore by Proposition 1.7.19 we have
dimH (B) ≤ dimH (f (B)) ≤ dimH (A).
This map f is used so many times in this chapter that it deserves a special
description:
Lemma 10.6.2. Let X be a complete locally compact space of curvature
≥ k, p ∈ X, and 0 < λ < 1. For every x ∈ X, let f (x) be a point in some
(arbitrarily chosen) shortest path [px] such that |pf (x)| = λ|px|. Then
(1) If k = 0, then |f (x)f (y)| ≥ λ · |xy| for all x, y ∈ X.
(2) If k < 0, then for every R > 0 there is a positive number c(k, λ, R)
such that |f (x)f (y)| ≥ c(k, λ, R) · |xy| for all x, y ∈ BR (p).
sinh(−kλR)
In fact, one can take c(k, λ, R) = sinh(−kR) .
Proof. Since X is complete and locally compact, all balls in X are pre-
compact. By Theorem 10.6.1 all these balls have the same Hausdorff
dimension. Since X can be covered by a countable collection of balls (e.g.,
by balls of integer radii centered at a fixed point), dimH (X) equals the
dimension of these balls; cf. Proposition 1.7.19. ¤
VR0
µn (BR (p)) ≤ (R/r)n µn (Br (p)) = µn (Br (p))
Vr0
Note that for k 6= 0 the same argument proves an inequality of the form
µn (BR (p)) ≤ C(r, R, k)µn (Br (p)) for some (nonoptimal) constant C(r, R, k).
This is sufficient for many purposes.
To prove the theorem in full generality, we need the following lemma
which is similar to the Gromov–Bishop inequality but involves spheres
instead of balls.
10.6. Dimension and Volume 373
If k > 0, the Bishop inequality and the upper bound for the diameter
(Theorem 10.4.1) imply that the volume of
√ X is no greater than the volume
of the n-dimensional sphere of radius 1/ k. In other words, the following
statement holds.
Corollary 10.6.9. If X is an n-dimensional Alexandrov space of curvature
≥ k where k > 0, then
µn (S n )
µn (X) ≤ .
k n/2
Theorem 10.6.8 immediately follows from the next proposition. The
proposition says that the ball Br (p) is “not greater” than the r-ball in Mkn
in the sense that there is a noncontracting map from one to the other.
Proposition 10.6.10. Let X be an n-dimensional Alexandrov space of
curvature ≥ k, and p ∈ X. Then there exists a map from f : X → Mkn
such that |f (x)f (y)| ≥ |xy| for all x, y ∈ X (i.e., f is noncontracting) and
|f (p)f (x)| = |px| for all x ∈ X.
Proof. According to Theorem 7.4.15, it suffices to prove that for every ε > 0
there is an N = N (ε, n, k, D) > 0 such that a space from this class cannot
contain an ε-separated set of more than N points.
Let X be an n-dimensional Alexandrov space of curvature ≥ k with
diam(X) ≤ D. By Proposition 10.6.10, there is a noncontracting map from
X to the ball of radius D in the space form Mkn . Let N be the maximal
possible cardinality of an ε-separated set in the D-ball in Mkn . Then X
cannot contain an ε-separated set of more than N points, because the image
of an ε-separated set under a noncontracting map is ε-separated. ¤
Remark 10.7.4. There is another proof which appears simpler when all
necessary components from the next sections are substituted. Instead of
Proposition 10.6.10, one needs the Gromov–Bishop inequality (a simple
version with nonsharp constants is sufficient), and the following fact implied
by the Gromov–Bishop inequality and Theorem 10.8.3: the volume µn of an
n-dimensional bounded Alexandrov space is positive and finite.
Now let X be as above. Then by the Gromov–Bishop inequality we have
µn (Bε (p)) ≥ c(n, k, D)µn (BD (p)) = µn (X)
for every p ∈ X (note that B D (p) = X since diam(X) ≤ D). Therefore X
cannot contain more than N = c(n, k, D)−1 disjoint ε-balls, or, equivalently,
cannot contain a (2ε)-separated set of more than N points. The proposition
follows.
Remark 10.7.5. The method described in the previous remark also proves
the following statement: the class of Riemannian n-manifolds of Ricci curva-
ture ≥ k and of diameter no greater than D is precompact in the Gromov–
Hausdorff space (because the Gromov–Bishop inequality works with lower
bounds for the Ricci curvature). However, no reasonable description is
known for the closure of this class in the Gromov–Hausdorff space.
Remark 10.7.6. A limit of n-dimensional spaces may have dimension
strictly less than n. For example, for every compact nonnegatively curved
378 10. Spaces of Curvature Bounded Below
These three theorems are proved in the end of Subsection 10.8.2; see
Corollaries 10.8.20, 10.8.21 and 10.8.23.
Theorem 10.8.3 is the first step in the study of the local topological
structure of an Alexandrov space. According to it, a space is divided into the
set of topologically regular points (those having Euclidean neighborhoods),
and a nowhere dense set of topological singularities. To obtain more detailed
information, one can study the set of singular points; some results about it
are discussed later in Section 10.10.
Another way to obtain a classification of points into good and bad ones
is based on tangent cones. Namely, a point is said to be regular if the
tangent cone at it is Euclidean, i.e., isometric to Rn , or, equivalently, the
space of directions is isometric to the standard sphere S n−1 .
Theorem 10.8.4. Let X be an n-dimensional Alexandrov space and p ∈ X.
Then the following three assertions are equivalent.
10.8. Local Properties 379
(1) p is regular.
(2) For every ε > 0 there is a neighborhood U of p at a Lipschitz
distance less than ε from some open set in Rn . (See Section 7.2 for
the definition of Lipschitz distance.)
(3) The Gromov–Hausdorff tangent cone of X at p is isometric to Rn .
In other words, the dilated balls 1ε Bε (p) converge in the Gromov–
Hausdorff sense to the unit ball of Rn as ε → 0.
This theorem sounds very similar to the respective fact (Theorem 9.1.44)
about spaces of curvature bounded above. However the proof for the case
of curvature bounded below is not that easy. Even the fact that the
metric of Σp (X) is intrinsic appears difficult to prove. We prove Theorem
10.8.6 using Gromov–Hausdorff convergence. The core facts are: Σp (X) is
compact, and the tangent cone Kp (X) is also the Gromov–Hausdorff tangent
cone of X at p. Then the desired properties of Σp (X) follow from the
respective properties of Kp (X), which in turn follow from general properties
of Gromov–Hausdorff limits.
Note that compactness of the space of directions does not follow from
(local) compactness of a space itself, if one does not assume that the di-
mension is finite. For example, consider a direct metric product of infinitely
many spaces Xi where Xi is a sphere of radius 1/i and i ranges over all
positive integers. This product is a compact length space of nonnegative
curvature; however its space of direction at any point is not compact (it
contains an infinite (π/2)-separated set and in fact is isometric to a sphere
in an infinite-dimensional Hilbert space).
One does not have to send ε to zero. For our purposes it is sufficient to fix
1
once and forever a small ε, say, ε < ε0 = 100m . If ε satisfies this inequality,
we omit it and write simply “m-strainer” and “m-strained point”. Obviously
the set of (m, ε)-strained points is open for any fixed m and ε.
Definition 10.8.11. Let X be an Alexandrov space. The strainer number
of X is the supremum of numbers m such that there exists an m-strainer
in X.
A strainer number at a point x ∈ X is the supremum of numbers m such
that every neighborhood of x contains an m-strained point.
We will show in this section that the strainer number equals the Haus-
dorff dimension of the space (in particular, the latter is an integer or infinity).
Right now, observe that X admits an (1, ε)-strainer for every ε > 0 unless
X is a single point. To prove this, pick any two points a, b and let p be an
almost midpoint (more precisely, a δ-midpoint with δ depending on |ab| and
ε); then (a, b) is the desired (1, ε)-strainer for p.
The following proposition tells us that the notion of a strained point is
local, and moreover can be formulated in terms of the space of directions.
Proposition 10.8.12. 1. If {(ai , bi )} is an (m, ε)-strainer for p ∈ X, and
a′i ∈ [ai p], b′i ∈ [bi p] for i = 1, . . . , m, then {(a′i , b′i )} is an (m, ε)-strainer for
p as well. Even if shortest paths do not exist, the same is true for points a′i
and b′i taken in suitable almost shortest paths. In particular, there exists an
(m, ε)-strainer {(a′i , b′i )} with arbitrarily small distances |pa′i | and |pb′i |.
382 10. Spaces of Curvature Bounded Below
Proof. The first statement follows from the monotonicity of angles. The
second one is obtained as the limit of the first one as a′i and b′i converge to p
along the respective shortest paths. To prove the third statement, observe
that {(a′i , b′i )} is an (m, ε)-strainer if a′i and b′i are sufficiently close to p in
the respective geodesics [pai ] and [pbi ]. ¤
The following two lemmas are simple but important technical facts that
are used everywhere in this section. The first one is an analog of the fact
that the sum of adjacent angles equals π, but with a strainer and comparison
angles instead of a geodesic and angles between directions.
Lemma 10.8.13. Let (a, b) be a (1, ε)-strainer for p, and q ∈ X be such
that
ε
|pq| < min{|pa|, |pb|}.
4
Then
e
|∡apq e
+ ∡bpq − π| < ε.
In particular, if shortest paths [pa], [pb] and [pq] exist, the angles between
them are close to the comparison angles:
e
0 < ∡apq − ∡apq < 2ε,
e
0 < ∡bpq − ∡bpq < 2ε.
e
hence |ab| > |ab|. The inequality |pq| < 4ε |pa| implies ∡paq < arcsin 4ε < 2ε ;
then
e e ε e
∡aqp = π − ∡paq − ∡apq > π − − ∡apq,
2
e
and similarly ∡bqp > π − ε/2 − ∡bpq. Thus
∡aqp + ∡bqp > 2π − ε − (∡apqe e
+ ∡bpq) >π
(i.e., the angle at q of the quadrangle apbq is greater than π). Since
|ab| > |ab|, it follows that
e
∡aqb > ∡aqb = 2π − ∡aqp − ∡bqp,
e
or, equivalently, ∡aqb e
+ ∡aqp e
+ ∡bqp > 2π, contrary to the quadruple
condition for (q; a, b, p).
The second statement of the lemma (about angles between shortest
paths) follows from the first one. Indeed, the quadruple condition implies
that ∡apq + ∡bpq ≤ 2π − ∡apb < π + ε. Therefore
e
(∡apq − ∡apq) e
+ (∡bpq − ∡bpq) < (π + ε) − (π − ε) = 2ε.
e
Since both terms ∡apq − ∡apq e
and ∡bpq − ∡bpq are positive, it follows that
they are bounded above by 2ε. ¤
Lemma 10.8.14. Let p, a1 , a2 , b1 , b2 ∈ X be such that (a1 , b1 ) and (a2 , b2 )
are (1, ε)-strainers for p,
ε
|a2 b2 | < min{|pa1 |, |pb1 |}
4
and ¯ ¯
¯|a1 a2 | − |a1 b2 |¯ < ε|a2 b2 |.
Then {(a1 , b1 ), (a2 , b2 )} is a (2, ε)-strainer for p.
PSfrag replacements
a1 a2
a2 b1
p
b1 a1
b2
p b2
1
and choose i0 for which |δi0 | = δ. Clearly m ∆n ≤ δ ≤ ∆n . Now let
xn+1 be a point in the union of shortest paths [xn ai0 ] ∪ [xn bi0 ] such that
|xn+1 ai0 | = yi0 ; i.e., the i0 -th coordinate of f (xn+1 ) takes the desired value.
(If shortest paths do not exist, use “almost shortest” paths instead.) We are
going to show that other coordinates of f (xn+1 ) are only slightly different
from those of f (xn ).
First, observe that |xn xn+1 | < 2δ. If |xn ai0 | > |xn+1 ai0 |, then xn+1 ∈
[xn ai0 ] and we have |xn xn+1 | = |xn ai0 | − |xn+1 ai0 | = δ. Otherwise xn+1 ∈
e i xn xn+1 ≥ ∡a
[xn bi0 ]; then ∡a e i xn bi > π − ε by monotonicity of angles,
0 0 0
and the inequality |xn xn+1 | < 2δ follows from (10.7) and the relation
|xn+1 ai0 | = |xn ai0 | + δ.
Recall that {(ai , bi )} is an (m, ε)-strainer for xn as long as xn ∈ U ,
and by Proposition 10.8.12 it remains an (m, ε)-strainer if one replaces ai0
or bi0 by xn+1 (depending on whether xn belongs to [xn ai0 ] or [xn bi0 ]).
Hence |∡ae i xn xn+1 − π | < 11ε for every i 6= i0 (cf. Remark 10.8.10). Since
2
|ai xn | > 1 and |xn xn+1 | < 2δ, we have ∡x e n ai xn+1 < 4δ < ε; hence
e i xn+1 xn − | < 12ε if i 6= i0 . Then (10.7) implies that
|∡a π
2
¯ ¯
¯|ai xn | − |ai xn+1 |¯ < |xn xn+1 | · sin(12ε) < 24δε
and therefore |δi,n+1 | < |δi,n | + 24εδ for all i 6= i0 . Thus
∆n+1 < ∆n − δ + 24(m − 1)εδ < ∆n − 21 δ < (1 − 1
2m )∆n
1
(here we use that ε ≤ 100m ). Note that we have verified one of our inductive
1 n
assumptions: ∆n+1 < ∆n < ∆0 . Moreover, ∆n < (1 − 2m ) ∆0 , so {∆n } is
a vanishing sequence and f (xn ) → y as n → ∞, provided that xn ∈ U for
all n.
It remains to prove that the sequence {xn } does not leave U and is a
Cauchy sequence. Recall that
1 n
|xn xn+1 | < 2∆n < 2∆0 (1 − 2m )
while xn ∈ U ; therefore
n
X
|pxn+1 | = |x0 xn+1 | ≤ |xj xj+1 | < ∆0 · C(m)
j=0
P
where C(m) = ∞ 1 j
j=0 (1 − 2m ) < ∞. We may choose our target y ∈ R
m
so close to f (p) that ∆0 = ky − f (p)k1 < diam(U )/C(m); then |pxn+1 | <
diam(U
P∞ ), and by induction the whole sequence {xn } is contained in U . Since
n=1 n xn+1 | < diam(U ) < ∞, the sequence {xn } is Cauchy one and its
|x
limit belongs in U . This limit is the desired point x such that f (x) = y. ¤
Corollary 10.8.16. The strainer number of an Alexandrov space is not
greater than its Hausdorff dimension.
10.8. Local Properties 387
Proof. Let {(ai , bi )} be an (m, ε)-strainer for p. Let us move these points
(including p) so as to obtain an (m, ε′ )-strainer. We will abuse notation and
denote the new points by the same letters p, ai , bi . The improvement of the
strainer is done in two stages. First, the angles ∡a e i pbi are made very close
π
to π. Then other angles are made very close to 2 .
To make an angle ∡a e i pbi close to π, do the following. Consider the
0 0
distance coordinates f : U → Rm associated with {(ai , bi )} where U is a
very small neighborhood of p. Since f (U ) is an open set in Rm , one can find
two points x = (x1 , . . . , xm ) and y = (y1 , . . . , ym ) in f (U ) such that xi = yi
for all i 6= i0 , and xi0 6= yi0 . Choose a and b in U such that x = f (a) and
y = f (b); then |ai a| = |ai b| for all i 6= i0 . Replace ai0 by a, bi0 by b, and
p by a midpoint p′ between a and b (or an almost midpoint p′ such that
e ′ b > π − ε′ ). Lemma 10.8.14 applied to points p′ , ai , bi , a, b implies that
∡ap
the new collection of points is an (m, ε)-strainer for the new point p = p′ .
Applying this construction for i0 = 1, . . . , m in turn, one obtains an
e i pbi > π − ε′ . (One may wonder why the relation,
(m, ε)-strainer in which ∡a
e 1 pb1 > π − ε′ , does not break down while we fix other angles. The
say, ∡a
answer is easy: the subsequent neighborhoods in which we choose points are
much smaller than the first one.)
Now apply the same procedure one more time. Since all (ai , bi ) are
now (1, ε′ )-strainers, Lemma 10.8.14 implies that the resulting collection of
points is an (m, ε′ )-strainer. ¤
point in a shortest path [px] such that |pf (x)| = Rr |px|. The metric of X
is not strictly intrinsic a priori, but one can modify this construction using
“almost shortest” paths instead of shortest ones. This way an inequality
r
|f (x)f (y)| > 2R |xy| can be ensured for points x, y ranging over any given
finite set, in particular, any finite subset S ′ ⊂ S. Thus the ball Br (p)
r
contains an 2R -separated net f (S ′ ) of an arbitrarily large number of points
and therefore it is not precompact. This contradiction shows that every ball
in X is precompact; hence X is locally compact. ¤
Exercise 10.8.22. Prove the statement of the above corollary without the
local compactness assumption.
Hint: The only case not covered yet is the following: dimH (X) = ∞ but
m < ∞. Since m < ∞, there is at least one precompact open set (namely, a
neighborhood of an m-strained point). Show that the proof of dimensional
homogeneity (namely the construction of a homothety) works in this case.
Proof. The fact that the set of n-strained points is an n-manifold follows
immediately from Theorem 10.8.18. The first statement of the corollary is
equivalent to saying that the local strainer number at every point p ∈ X
equals n. Let m be the local strainer number at p. Then some neighborhood
of U of p is bi-Lipschitz homeomorphic to Rm and hence dimH (U ) = m.
Then m = dimH (X) = n by dimensional homogeneity. ¤
Proof. The proofs for compact and pointed spaces are similar. Let {Xi } be
a sequence of compact Alexandrov spaces of curvature ≥ k, dimH (Xi ) ≤ n
for all i, Xi −→
GH X as i → ∞. By Proposition 10.7.1, X is a (compact)
space of curvature ≥ k. Suppose that dimH (X) > n. Since dimH (X) equals
the strainer number of X (Corollary 10.8.21), there is an (n + 1)-strained
point p ∈ X. Fix an (n + 1)-strainer {(aj , bj )}n+1
j=1 for p. As soon as Xi gets
sufficiently close to X, one can find points p , a′j , b′j whose distances from
′
one another are almost equal to the respective distances between points p,
aj , bj in X. Then the comparison angles (∡) e involving these points in Xi are
almost equal to the respective comparison angles in X, because the angles ∡ e
are continuous functions of distances. In particular, the inequalities from the
definition of a strained point are satisfied for p′ , a′j , b′j . Thus Xi contains an
(n + 1)-strainer, and therefore dimH (Xi ) ≥ n + 1, for all large enough i. ¤
Now continue the proof of the proposition. We will show that for every
ε > 0 the space of directions cannot contain an ε-separated set of more
than C/εn points where C is some constant (almost the same one as in the
lemma). Let S = {xi }N i=1 be an ε-separated set in Σp (X). We may assume
that the points xi are representable by shortest paths γi emanating from p.
e i (t)pγj (t) → ∡(γi , γj ) = |xi xj | as t → 0.
By the definition of angle, ∡γ
Therefore one can find such a small r > 0 that ∡γ e i (t)pγj (t) > 1 |xi xj | > ε/2,
2
and hence |γi γj | > 2r sin(ε/4) > rε/4, for all i, j (i 6= j). In other words, we
obtain an (rε/4)-separated set in an r-ball. Applying Lemma 10.9.2 yields
a desired bound for the number of points in the set. Since ε is arbitrary, it
follows that Σp (X) is compact. ¤
Recall that the tangent cone Kp (X) over the space of directions Σp (X)
consists of the origin o and points represented by (ξ, r) where x ∈ Σp (X)
and r > 0. The distance from (ξ, r) to the origin equals r, and the distance
e 0 a1 oa2 = ∡(ξ1 , ξ2 ).
from a1 = (ξ1 , r1 ) and a2 = (ξ2 , r2 ) is defined so that ∡
Theorem 10.9.3. Let X be a finite-dimensional Alexandrov space and
p ∈ X. Then the Gromov–Hausdorff tangent cone of X at p exists and
is isometric to Kp (X).
392 10. Spaces of Curvature Bounded Below
Proof. Let B denote the unit ball centered at the origin o of the cone
Kp (X). We have to show that the rescaled balls 1r Br (p) converge to B in
the Gromov–Hausdorff metric as r → 0. Fix an ε > 0. By one of the
criteria of Gromov–Hausdorff convergence (namely, by Proposition 7.4.11),
1
¯ ε-nets {xi }¯ in B and {yi } in r Br (p), where r is small
it suffices to find finite
enough, such that ¯|xi xj | − |yi yj |¯ < ε for all i, j.
Pick a finite ε-net {xi } in B whose elements are representable by shortest
paths. Let γi denote a shortest path representing xi , parameterized with the
constant speed |oxi |. Assume that r is so small that the point pi = γi (r) is
defined for every i, and
¯ ¯
¯ |pi pj | ¯
¯ ¯
¯ r − |xi xj |¯ < ε
Remark 10.9.4. Note that the above argument not only proves that Kp (X)
is the Gromov–Hausdorff limit of rescaled balls, but also gives explicit small-
distortion maps from the balls to the cone. Namely the distortion of the
“logarithm” map goes to zero with the radius of the ball. Reformulating
this in terms of angles instead of distances, one gets the following useful
fact: for every ε > 0 there is an r > 0 such that |∡xpy e − ∡xpy| < ε
whenever |px| and |py| are less than r.
Lemma 10.9.14. For every integer n ≥ 2 and every ε > 0 there is δ > 0
such that the following holds.
Let Y be an Alexandrov space of curvature ≥ 1 and of dimension no
greater than n − 1. Suppose that Y contains n pairs of points {(xi , yi )}ni=1
such that |xi yi | > π−δ for all i, and the distances |xi xj |, |xi yj |, |yi yj | (i 6= j)
are δ-close to π/2. Then dGH (Y, S n−1 ) < ε.
Proof. Suppose the contrary. Then there is a sequence of spaces {Ym }∞m=1
satisfying the assumptions of the lemma for δ = δm where δm → 0 and
such that dGH (Ym , S n−1 ) ≥ ε. Due to the Compactness Theorem 10.7.2
we may assume that {Ym } converges in the Gromov–Hausdorff topology to
an Alexandrov space Y of curvature ≥ 1 and of dimension no greater than
n − 1. We are going to show that Y is isometric to S n−1 , contrary to our
assumption.
Recall (Theorem 7.3.25) that there is a correspondence between Ym
and Y with distortion no greater than 2dGH (Ym , Y ). Thus an “almost
orthogonal” collection of points in Ym corresponds to an “almost orthogonal”
collection in Y (with δm replaced by δm + dGH (Ym , Y ) which goes to zero
anyway). Extracting a converging subsequence of these collections in Y ,
one obtains an orthogonal collection. By Lemma 10.9.10, it follows that
Y ≃ S n−1 . ¤
Exercise 10.9.15. Prove that for every n ∈ N and ε > 0 there is a δ > 0
such that for any n-dimensional Alexandrov space X of curvature ≥ 1 the
following holds.
396 10. Spaces of Curvature Bounded Below
(the second approximate equality follows from the fact that in the sphere
there is a real orthogonal collection near {(ξ i , η i )}. ¤
Another proof. One can prove the theorem using regular points instead of
Lemma 10.9.14.
P Namely, the above argument shows that the approximate
equality cos2 |ξξi | ≈ 1 holds if Σx (X) is isometric to S n−1 , i.e., x is a
regular point. Then the inequalities (1 + ε)−1 |xy| ≤ |f (x)f (y)| ≤ (1 + ε)|xy|
follow for a regular point x. Since regular points are dense, these inequalities
follow for all x, y. ¤
Note that the second proof makes it possible to write an explicit bound
for δ in terms of ε, while the first one only proves that a δ exists (see how
Lemma 10.9.14 was proved).
Sometimes the following reformulation of Theorem 10.9.16 is useful.
Remark 10.9.18. Points p ∈ X with dGH (Σp (X), S n−1 ) < ε are referred to
as ε-regular points. The set Xε of ε-regular points features some interesting
properties. Obviously every point of Xε is (n, 2ε)-strained and hence has a
Euclidean neighborhood. For the same reason, the interior of Xε is dense in
X (see Corollary 10.8.23).
Furthermore, though it sounds a little surprising, the set Xε is convex
in the sense that a shortest path connecting two points of Xε does not leave
Xε . This follows from the fact that the tangent cone is constant along any
shortest path except its endpoints; in other words, if x and y belong to the
interior of some shortest path, then Kx (X) is isometric to Ky (X) [Pet].
Hint: Using a modified argument from the proof of the Bishop inequality,
show that the (n−1)-dimensional measure of Σp (X) is close to that of S n−1 .
Then apply the last part of Exercise 10.9.15.
≥ k says that for every sufficiently short segment of γ and every point p
sufficiently close to it, the function g must be no less concave than gk .
The same comparison property may hold for curves γ that are not
geodesics. Curves possessing this property are called k-quasigeodesics. A
quasigeodesic is a k-quasigeodesic in an Alexandrov space of curvature ≥ k.
Quasigeodesics have many nice properties; in particular all the above
mentioned missing properties of geodesics do hold for quasigeodesics.3
In addition, the class of quasigeodesics is closed under noncollapsing
Gromov–Hausdorff convergence.
Let us mention one more property. If X is a space of curvature ≥ 1,
two points a, b ∈ X are called antipodes if |ax| + |xb| ≤ π for all x ∈ X. If
two quasigeodesics have a common endpoint p and their directions at p are
antipodes, then the concatenation of these paths (taken with the opposite
orientation) is a quasigeodesic too. This allows one to build quasigeodesics
“from pieces”.
A geodesic in an extremal set (in particular, a geodesic in the boundary)
is a quasigeodesic in the ambient Alexandrov space [PP].
There are several important applications of quasigeodesics. One of them
is the proof of the following gluing theorem (see [Pet1] and the preprint
mentioned above).
Theorem 10.10.12. Let X and Y be Alexandrov spaces of the same dimen-
sions and of curvature ≥ k. Suppose that their boundaries are isometric.
Then the space Z obtained by gluing X and Y along an isometry of their
boundaries is a space of curvature ≥ k.
3See the preprint by G. Perelman and A. Petrunin, Quasigeodesics and gradient curves in
Alexandrov spaces.
404 10. Spaces of Curvature Bounded Below
10.10.7. Curvature bounded from both sides. One may wonder what
can be said about a space of curvature bounded both above and below. The
answer is simple though it required hard work to prove it: loosely speaking,
such a space is a Riemannian manifold whose metric tensor has slightly
lower smoothness than usually considered in Riemannian geometry. More
precisely, the following theorem holds.
Theorem 10.10.13 (I. Nikolaev, see [BN]). Let (X, d) be a length space
with both side upper and lower curvature bounds (of curvature ≥ k and ≤ K).
If (X, d) has no boundary, then X is a manifold and possesses a C 3 -smooth
atlas such that in its charts the metric d can be defined by a Riemannian
metric tensor whose coefficients gij are in the class Wp2 for every p > 1.
If X has a boundary, then the interior of X is an almost Riemannian
manifold in the same sense as above; however the boundary may be non-
smooth.
If k = K, then (X, d) is usual space of constant curvature.
Recall that Wp2 means the space of functions whose second Sobolev
derivatives belong to Lp . (In particular gij ∈ C 1,α for all α < 1.)
The theorem allows us to define formal Christoffel symbols and curva-
ture tensor via derivatives of gij . These formal values are defined almost
everywhere; nevertheless they have a reasonable geometric meaning. For
instance, parallel transport defined geometrically coincides (for almost all
paths) with one given by Christoffel symbols. And formal sectional curva-
tures have geometric meaning as well.
Also I. Nikolaev proved that such a space can be approximated (in
the sense of uniform convergence) by Riemannian manifolds of the same
dimension and with curvature bounds converging to that of (X, d).
Bibliography
405
406 Bibliography
[BB] Yu. Burago, S. Buyalo, Metrics of upper bounded curvature on 2-polyhedra. II, St.
Petersburg Math. J. 10 (1999), 619–650.
[BGP] Yu. Burago, M. Gromov, G. Perelman, A. D. Alexandrov spaces with curvatures
bounded below, Russian Math. Surveys 47 (1992), 1–58.
[BZ] Yu. Burago and V. Zalgaller, Geometric Inequalities, A Series of Comprehensive
Studies in Mathematics 285, Springer-Verlag, Berlin, 1988.
[Car] M. P. do Carmo, Riemannian Geometry, Birkhäuser, Boston, 1992.
[CE] J. Cheeger and D. Ebin, Comparison Theorems in Riemannian Geometry, North
Holland, New York, 1975.
[CFG] J. Cheeger, K. Fukaya, and M. Gromov Nilpotent structures and invariant metrics
on collapsed manifolds, J. Amer. Math. Soc. 5 , no. 2,(1992) 327–372.
[CG1] J. Cheeger and D. Gromoll, On the structure of complete manifolds of non-negative
curvature, Ann. of Math. 96 (1972), 413-443.
[CG] J. Cheeger and D. Gromoll, The splitting theorem for manifolds of nonnegative Ricci
curvature, J.Differential Geometry 6 (1971-72), 119-128.
[Cha] I. Chavel, Riemannian Geometry — a modern introduction, Cambridge Univ. Press,
Cambridge, 1993.
[Ch] , C. H. Chen Warped products of metric spaces of curvature bounded from above,
Trans. Amer. Math. Soc. 351 (1999),4727-4740.
[E] P. B. Eberlein, Geometry of Nonpositively Curved Manifolds, The University of
Chicago Press, Chicago and London, 1996.
[Fe] H. Federer, Geometric Measure Theory, Springer-Verlag, Berlin, 1969.
[GHL] S. Gallot, D. Hulin, and J. Lafontaine, Riemannian geometry , Springer-Verleg,
Berlin, 1990, xiv+284.
[Gro1] M. Gromov, Filling Riemannian manifolds, J. Differential Geometry 18 (1983),
1–147.
[Gro2] M. Gromov, Systols and intersystolic Inequalities, Actes de la table rande de
géométrie differentielle en l’honeur de Marcel Berger. Collection SMF 1 (1996), 291–
362.
[Gro3] M. Gromov, Metric Structures for Riemannian and Non-Riemannian Spaces,
Progress in Mathematics 152, Birkhäuser, Boston, 1999.
[Grove] K. Grove, Metric differential geometry, Lecture Notes in Math. 1263 (1987), 171–
227.
[GP] K. Grove and P. Petersen (Editors), Comparison geometry, Papers from the Special
Year in Differential Geometry held in Berkeley, CA. 1993–94. Research Institute
Publications, 30 (1997), Cambridge University Press, Cambridge, x+262 pp.
[HM] S. Hartman and J. Mikusiński,, The theory of Lebesgue measure and integration,
International series of monographs on Pure and Applied Mathematics, vol. 15, 1961.
[HS] J. Hass and P. Scott, Bounded 3-manifolds admits negatively curved metric with
concave boundary, J. Differential Geom. 40 (1994), no. 3, 449-459
[J] J. Jost, Nonpositive curvature: Geometric and analytic Aspects, Lectures in Mathe-
matics ETH Birkhäuser, 1997.
[Ka] H. Karcher, Riemannian comparison constructions, in Global Differential Geometry
(S. S. Chern ed.), MAA Studies in Mathematics 27 (1989), 170–222.
[Kl] B. Kleiner, The local structure of length spaces with curvature bounded above, Math.
Z. 231 (1999), no. 3, 409–456.
Bibliography 407
[KS] M. Katz and A. Suciu, Systolic freedom of loopspace, (2000), 1-13, preprint.
[Mas] W. S. Massey, Algebraic topology: an introduction, Graduate Texts in Mathematics,
vil. 56. Springer, New York, 1987
[Mi] J. Milnor, Topology from the differentiable viewpoint, Univ. Press of Virginia, 1965.
[Mo] F. Morgan, Geometric measure theory: a beginner’s guide, Academic Press, 1988.
[Mun] Munkres, J. R. Topology: a first course, Prentice-Hall, Inc., Englewood Cliffs, N.J.,
1975. xvi+413
[Mun1] Munkres, J. R. Analysis on manifolds, Addison-Wesley Publishing Company,
Advanced Book Program, Redwood City, CA, 1991, xiv+366 pp.
[OS] Y. Otsu and T. Shioya The Riemannian structure of Alexandrov spaces J. Differential
Geom. 39 (1994), no. 3, 629–658.
[Ot] Y. Otsu, Differential geometric aspects of Alexandrov spaces, in Comparison Geome-
try , Math. Sci. Res. Inst. Publ 30 Cambridge Univ. Press, Cambridge (1997), 135–148.
[Per1] G. Perelman, The Beginning of the Morse Theory on Alexandrov spaces, St. Pe-
tersburg Math. Journ. 5 (1994), no. 1, 205-214.
[Pet] A. Petrunin, Parallel transportation for Alexandrov spaces with curvature bounded
below, Geom. Funct. Analysis 8 (1998), no. 1, 123-148.
[Pet1] A. Petrunin, Applications of quasigeodesics and gradiant curves, in Comparison
Geometry, Math. Sci. Res. Inst. Publ 30, Cambridge Univ. Press, Cambridge (1997),
203-219
[Pl] C. Plaut, Metric spaces of curvature ≥ k, in “A handbook of Geometric Topology”
(to appear)
[PP] G. Perelman and A. Petrunin, Extremal subset in Alexandrov spaces and a generalised
Liberman theorem, St. Petersburg Math. Journ. 5 (1994), no. 1, 215-227.
[PP1] G. Perelman and A. Petrunin, Quasigeodesics and gradient curves in Alexandrov
spaces, Preprint (1994), 1-23.
[Resh] Yu. G. Reshetnyak, Nonexpending maps into spaces of curvature not greater K,
Siberian. matem. j. 9, no. 4 (1968), 918-927 (in Russian).
[Sp] M. Spivak, Calculus on manifolds. A modern approach to classical theorems of
advanced calculus, W. A. Benjamin, Inc., New York-Amsterdam 1965, xii+144 pp.
[St] L.Stoyanov, An estimate from above of the number of periodic orbites for semi-
dispersedbilliards Comm. Math. Phys. 24 (1989), no. 2, 217-227.
[Yam] T. Yamaguchi, A convergence theorem in the geometry of Alexandrov spaces, Actes
de la Table Ronde de Geometrie Differentielle (Luminy, 1992), 601–642, Semin. Congr.,
1, Soc. Math. France, Paris, 1996.
Index
409
410 Index