On Asymmetric Distances
On Asymmetric Distances
Andrea C. G. Mennucci ?
May 28, 2007
Abstract
In this paper we discuss asymmetric metric spaces in an abstract setting, mim-
icking the usual theory of metric spaces. As a typical application, we consider
asymmetric metric spaces generated by functionals in Calculus of Variation.
K EYWORDS : ASYMMETRIC METRIC , GENERAL METRIC , QUASI METRIC , OSTENSIBLE METRIC , F INSLER METRIC ,
PATH METRIC , LENGTH , GEODESIC CURVE , H OPF –R INOW THEOREM
1 Introduction
“Besides, one insists that the distance function be symmetric, that is, d(x, x0 ) = d(x0 , x).
(This unpleasantly limits many applications [· · ·]).” M. Gromov ([12], Intr.)
The main purpose of this paper is to study an “asymmetric metric theory”; this
theory naturally generalizes the metric part of Finsler Geometry, much as symmetric
metric theory generalizes the metric part of Riemannian Geometry.
In order to do this, we will define the “asymmetric metric space” as a set M equipped
with a positive b : M × M → lR+ which satisfy the triangle inequality, but may fail
to be symmetric (see 2.1 for the formal definition). We will then state, (in sec. 2)
“asymmetric definitions” in this space, such as “forward local compactness”, “forward
completeness”, “forward boundedness”, and so on.
The invention of a (possibly) asymmetric distance is quite useful in some applied
fields, and mainly in Calculus of Variation, as we will exemplify in the next section
§1.ii, and discuss more in depth in sec. §5.i. It is then no surprise that this theory was
invented and studied many times in the past.
It was indeed shown in preceding studies that, even if a metric space does not have
the “smooth” character of a Riemannian structure, still many definitions and results can
be reformulated and carried on (1) : see for example in Busemann [5], or in Gromov’s
[12], or in Ambrosio et al [1, 2].
To exemplify this fact, in the next section we will present an example theorem that
generalizes the Hopf–Rinow theorem from finite dimensional Riemannian spaces to
suitable metric spaces; and a further generalization to the asymmetric case.
A different point of view is found in the theory of quasi metric (or ostensible metric)
where much emphasis is given to the topological aspects of the theory; indeed, a quasi-
metric will generate, in general, three different topologies (see §3.ii.1); this brings forth
many different and non-equivalent definitions of “completeness” (and of the other
? Scuola Normale Superiore Piazza dei Cavalieri 7, 56126 Pisa, Italy
(1) and Busemann [4] attributes the first such results in this respect to Cartan, in 1928
1
properties commonly stated in metric theory). In our definition of the “asymmetric
metric space” (M, b), we choose a different topology associated to the space than the
one used in “quasi metric spaces”: this makes it difficult to compare the results in the
two fields. We discuss further this issue in sec. §3.ii.
Our set of hypotheses is more resembling (and slightly generalizes) Busemann
theory of general metric spaces (as was defined in [4]); we will discuss differences and
similarities in section §3.i.
Summarizing, this presentation of the theory of asymmetric metrics is more specific
and more focused on “metric aspects” of what is seen in ostensible metric theory, and at
the same time is less “geometric oriented” than what is seen in Busemann work (indeed,
many of the examples presented are quite non-smooth).
Theorem 1.1 (symmetric Hopf-Rinow) Suppose that (M, d) is path-metric and lo-
cally compact; then the two following facts are equivalent
and they imply that each pair of points can be joined by a geodesic (i.e. a minimum of
lenγ).
The above is the “metric only” counterpart of the theorem of Hopf-Rinow in Rie-
mannian Geometry: indeed, if (M, g) is a finite-dimensional Riemannian manifold, and
d is the associated distance, then (M, d) is path-metric and locally compact.
Since there is a Hopf-Rinow theorem in Finsler Geometry, we would expect that
there would be a corresponding theorem for “asymmetric metric spaces”; and indeed
we can prove this result:
Theorem 1.2 Suppose that the asymmetric metric space (M, b) is path-metric and
forward-locally compact; then the following two facts are equivalent
• (M, b) is forward-complete
and they imply that any two points can be joined by a geodesic.
2
§1.ii asymmetric metric theories in Calculus of Variations
For the sake of this introduction section, let M be a smooth connected differential
manifold.
Consider an integrand F : T M → [0, ∞] that is Borel measurable and such that
v 7→ F (x, v) is convex and positively 1-homogeneous (that is, F (x, lv) = lF (x, v) for
l ≥ 0). Consider the length
Z 1
lenγ = F (γ(s), γ̇(s)) ds (1.3)
0
where γ : [0, 1] → M is a locally Lipschitz curve; this generates a “distance” b(x, y), as
the infimum of the length of curves γ connecting x to y; b is called a Finsler metric, since
is is easily seen to satisfy a triangular inequality; but in general b will be asymmetric.
Consider also the Hamilton-Jacobi equation
n
H(x, Du(x)) = 0, u(a) = 0 (1.4)
3
2 Asymmetric metric spaces
We review some basal concepts and results; those are similar to what is commonly found
in books on metric spaces. In all of this chapter, M will be a generic set.
with balls
B(x, ε)={y | d(x, y) < ε} = B + (x, ε) ∩ B − (x, ε)
def
• by using the triangular inequality, it is easily seen that the equivalence relation
x ∼ y ⇐⇒ d(x, y) < ∞ decomposes the space (M, b) in equivalence classes
that are all open (so two different classes are disconnected); so we may restrict
the study of the space to those equivalence classes;
• defining φ(t) = t/(1 + t) and φ(∞) = 1 and then b̃=φ ◦ b and d˜=φ ◦ d, we may
def def
build a space (M, b̃) that is topologically equivalent, and where the distance does
not assume the value +∞.
Unfortunately those remedies are not really useful in some cases: see 2.17.
4
§2.ii Lipschitz maps
Let (N, δ) be a symmetric metric space, and f : N → M , g : M → N . We could
define that f and g are “Lipschitz w.r.t. b”, or “Lipschitz w.r.t. d”. That is, in the first
case
b(f (x), f (y)) ≤ Cδ(x, y), δ(g(x), g(y)) ≤ Cb(x, y)
for some constant C > 0. In the second case
d(f (x), f (y)) ≤ Cδ(x, y), δ(g(x), g(y)) ≤ Cd(x, y)
for some constant C > 0.
The two concepts are equivalent for f , but not equivalent for g.
As an example, if we provide M 2 with the metric
§2.iii Length
We induce from b the length lenb γ of a curve γ : [α, β] → M , by using the total
variation (4)
n
X
lenb γ = sup
def
b γ(ti−1 ), γ(ti ) (2.5)
T i=1
where the sup is carried out over all finite subsets T = {t0 , · · · , tn } of [α, β] and
t0 ≤ · · · ≤ tn . If lenb γ < ∞, then γ is called rectifiable.
Remark 2.6 The above length is asymmetric: if γ̂(t)=γ(−t), then in general lenb γ
def
will be different from lenb γ̂; hence there does not exist a measure H on M such that the
length of a curve is the measure of its image, that is
H(image(γ)) = lenb γ ;
indeed image(γ) = image(γ̂) which is incompatible in the general case in which
lenb γ 6= lenb γ̂.
We define bg
bg (x, y) = inf lenb γ (2.7)
where the inf is taken in the class of all continuous curves γ connecting x to y.
Note that
lenb γ ≥ b(γ(α), γ(β)) (2.8)
so bg ≥ b. It is also easy to prove that bg satisfy a triangle inequality: so bg is an
asymmetric distance (possibly taking the value ∞; see 2.4 and 2.17 on this). bg is
called the geodesic (asymmetric) distance induced by b. We may endow M with
def
the distance dg (x, y)= bg (x, y) ∨ bg (y, x); then dg ≥ d, and the topology resulting in
(M, bg ) is finer than the one used in (M, b) (has more open sets and less compact sets);
the two topologies may easily differ, as in this simple example:
(4) according to Busemann [4], this definition goes back to Menger (1928)
5
Example 2.9 Consider
M = {x ∈ lR2 | − 1 ≤ x1 ≤ 1, x2 = 0} ∪
[
{x ∈ lR2 | − 1 ≤ x1 ≤ 1, x2 = (x1 − 1)/n}
n≥1
and b the Euclidean distance (see fig. 1). Then (M, b) is compact but (M, bg ) is not.
`(t)=lenb γ|[α,t]
def
(2.11)
Note that:
• if γ is Lipschitz of constant L, then, by direct substitution in (2.5),
so ` is Lipschitz. (5)
In this case the derivative d`
dt exists for almost all t, and it is called the metric
derivative of γ (as defined in [1, 2]).
• Supposing that ` is continuous (for simplicity), for t ≥ s,
since the length of γ restricted to [s, t] is `(t) − `(s), and γ is indeed a path
connecting γ(s) to γ(t);
• whereas the fact that ` is continuous does not imply that γ is continuous, as shown
in this example:
6
From this we induce the asymmetric distance as explained in the introduction §1.ii, to
get
|x − y| if x ≤ y or y < x ≤ 0
b(x, y) = − log(y) + log(x) if 0 < y < x (2.13.??)
+∞ if y ≤ 0 < x
then the topology τ + generated by forward balls is the usual euclidean topology on
[−1, 1]; whereas the topology τ − generated by backward balls makes [−1, 0] discon-
nected from (0, 1]; and τ = τ − .
γ : [−1, 1] → M defined by γ(t) = t has lγ (t) = t, but is not continuous.
Lemma 2.14 Suppose that ` is continuous; then γ is continuous if and only if its image
is compact. (Proof follows from lemma 3.7 and (2.12)).
§2.iv Definitions
We add some definitions
• We say that the (asymmetric) metric space (M, b) is a path-metric space, or that
b is intrinsic, if b = bg ;
• and we say that it admits geodesics, or that geodesics exist, if, for all x, y ∈ M ,
there is a geodesic connecting x to y.
We say that (M, b) is forward complete if any forward Cauchy sequence (xn )
converges to a point x (according to the topology τ : cf. 2.3 and §3.ii.1 on the notion of
convergence).
(These definitions agree with those used in Finsler Geometry (see ch. VI of [3]). See 3.4
for a different definition.)
7
• We say that (M, b) is forward-locally compact if ∀x ∃ε > 0 such that
def
D+ (x, y)={y | b(x, y) ≤ ε}
is compact.
is compact.
• We say that (M, b) is locally compact if ∀x ∃ε > 0 such that both D− (x, y) and
D+ (x, y) are compact.
Remark 2.17 In general, even if (M, b) is connected and b < ∞ at all points, it
may be the case that bg (x, y) = ∞ for some points; if we do not like this fact, we
may wish to try the remedies proposed in 2.4; in particular the equivalence relation
x ∼ y ⇐⇒ dg (x, y) < ∞ decomposes the space (M, bg ) in equivalence classes that
coincide with all connected components of (M, bg ) (since bg is itself path-metric, by
4.44).
Unfortunately the second remedy suggested in 2.17 is only a placebo when we are
interested in path-metric spaces and/or in studying geodesics: suppose b is path-metric
def
and we decide to set φ(t) = t/(1 + t) and define b̃=φ ◦ b: then b̃ is not path-metric; so if
we try to substitute b̃ by its generated path-metric distance b̃g we find out, by prop. 4.37,
that b̃g = bg = b, and we are back to square one. So we may sometimes be forced to
address the case when bg (x, y) = ∞ for some points.
8
• a “General metric space” is a space satisfying all requisites in 2.1, and moreover
We add some comments that highlight the differences between the approach in the
present paper and in [4]:
• the additional hypothesis (3.1) used in defining “General metric spaces” is not
really restrictive; in many applications it does hold (for example, when the space
is locally compact, as in [16]); still it is possible to find examples of path-metric
spaces such as 4.13 and 2.13 where it does not hold (even though this latter comes
from a Calculus of Variations problem of the form discussed in introduction);
• on the other hand, in the aforementioned papers it was (sometimes silently)
assumed that the metric b be intrinsic; (6) that is, no distinction was made between
b and bg ; this is restrictive, since in many simple situations we will consider the
two do differ.
A good reason to consider non-path-metric spaces is as follows: suppose (M 0 , b0 )
is an asymmetric metric space, and (M, b) is a subset of it, with b being the
restriction of b0 ; then it easy to devise examples where (M 0 , b0 ) is path-metric but
(M, b) is not (7) and vice versa. For example, in the forthcoming paper [8] a
family of (symmetric) metric space are studied that are not (and cannot possibly)
be path-metric; the reason being this theorem (adapted from 2.6 in [8]):
Theorem 3.2 Suppose that (M, d) is a complete path-metric space, and that
i : M → E is an isometric immersion in a uniformly convex Banach space E:
then i(M ) is convex, and any two points in M can be joined by a unique minimal
geodesic (unique up to reparametrization).
Then (the symmetric version of) theorem 4.24 was quite useful in establishing
existence of geodesics in [8].
So in most of this paper we will carefully distinguish the two metrics b and bg .
• In most of the cited works (with the exception of Phadke’s [17]) the space is
assumed to be locally compact (or even finitely compact).
According to the notes at end of the introduction of [6], some work was carried
on at that time to extend known results to forward locally compact spaces; but the
results in this paper do not appear in the announced papers.
(6) Note that the definition of “intrinsic metric” was not used at the time of [4]: it was in a sense replaced by
9
§3.ii Comparison with quasi metric spaces
§3.ii.1 Topology
As already pointed out, the definition 2.1 of the asymmetric metric coincides with
the definition of a quasi metric that is found in the literature, cf. Kelly [14], Reilly,
Subrahmanyam and Vamanamurthy [13] (8) Fletcher and Lindgren [11, (pp 176-181)],
Künzi [15]. The main difference between our theory of asymmetric metric spaces
and quasi metric spaces is in the choice of the associated topology.
Indeed, we have three topologies at hand:
• the topology τ , generated by the families of forward and backward balls, or
equivalently by the metric d defined in (2.2);
• the topology τ + generated by the families of forward balls;
• the topology τ − generated by the families of backward balls;
it may happen that these three topologies are different.
This problem has been studied in [14]: there Kelly introduces the notion of a
bitopological space (M, τ + , τ − ), and extends many definition and theorems, (such as
the Urysohn lemma, the Tietze’s extension theorem, the Baire category theorem (9) ) to
these spaces. Unfortunately Kelly does not include the topology τ in his studies.
In many applications τ = τ + = τ − : see 3.8 and section §5.i. We have chosen to
associate the topology τ to the “asymmetric metric space”. τ is a symmetric kind of
object: as a consequence, we have only one notion of “open set”, of “compact set”, of
“the sequence (xn ) converges to x”, and of “the functions f : N → M and g : M → N
are continuous”. Furthermore
Proposition 3.3 If xn → x (according to τ ) then the sequence (xn ) is both a forward
Cauchy sequence and a backward Cauchy sequence,
in accordance with the symmetric case.
10
Definition 3.4 • A sequence (xn ) ⊂ M is a “left b-Cauchy sequence” when
∀ε > 0 ∃x ∈ M and ∃k ∈ lN such that b(x, xm ) < ε whenever m ≥ k.
ii). Any “forward Cauchy sequence” (as defined in (2.15)) is a “left b-Cauchy se-
quence” (11) .
For those reasons, it is not easy to compare the results and examples in the above
papers, with the result and examples here presented.
From here on, we return to our setting of “asymmetric metric spaces”: we will
always use the topology τ .
Remark 3.5 Let ε > 0 and x, y, z be such that b(x, y) < ε and b(x, z) < ε. This does
not imply, in general, that b(y, z) < 2ε
z
x
?
y
i). In example 4.13.(iv) there exists a sequence such that b(xn , x) → 0 but xn 6→ x.
(Note that this example cannot be found in Busemann’s “General metric space”,
due to property (3.1)).
ii). Fix a sequence (xn ) ⊂ M . Suppose that ∀ε > 0 there exists a converging
sequence (yn ) such that b(yn , xn ) < ε. If b is symmetric and (M, b) is complete,
then (xn ) converges. If b is asymmetric and (M, b) is complete, then there is a
counter-example in 4.13.(vi).
(11) just choose n = N = k and x = xn in the definition of left b-Cauchy sequence
(12) Indeed, Kelly [14] had encountered this problem, which was a motivation of [13]
11
iii). In example 4.14.(iv) there is a ball B + (a, r) and points ai 6∈ B + (a, r) such that
B + (a, r) ⊂ i B + (ai , r/2): it is then difficult to find a useful definition of a
S
“precompact set”. (13)
The above may be proved by using this Lemma (that is similar to (2.3) and (2.6) in
Zaustinsky’s [19])
Proof. Define
f (r) = sup b(y, x)
x,y∈C, b(x,y)≤r
and then f is monotone. Since C is compact, then f < ∞ and limr→0 f (r) = 0:
otherwise we may find ε > 0 and xn , yn s.t. b(xn , yn ) → 0 while b(yn , xn ) > ε:
extracting converging subsequences, we obtain a contradiction.
R 2r
From f we can define an ω as required, for example ω(r) = 1r r f (s)ds (note
that ω ≥ f ).
and then τ = τ + = τ − .
• A is symmetrically bounded
(13) An useful definition of “precompact set” is found in [15], where, though, a different completeness
12
But note that in example 4.13.(v) there exists a set A ⊂ M that is forward bounded but
not backward bounded (and vice versa in 4.14.(i)).
Similarly
Proposition 4.2 Let (xn ) ⊂ M be a sequence; then the following are equivalent
• (xn ) is forward Cauchy and backward Cauchy
• (xn ) is symmetrically Cauchy
From that we obtain: if (M, b) is either forward or backward complete, then it is
symmetrically complete.
See examples 4.14, 4.15.
Proof. Suppose that (xn ) is symmetrically Cauchy: then ∀ε > 0 ∃N such that ∀n, m >
N , d(xn , xm ) < ε: then b(xn , xm ) < ε.
Suppose that (xn ) is forward Cauchy and is backward Cauchy: ∀ε > 0 ∃N 00 such
that ∀n > m > N 00 , b(xn , xm ) < ε, and ∃N 0 such that ∀n > m > N 0 , b(xm , xn ) < ε:
then we let N = N 0 ∨ N 00 , and ∀n, m > N , d(xn , xm ) < ε.
Suppose that (M, b) is forward complete; let (xn ) be symmetrically Cauchy: then it
is forward Cauchy, and then, since (M, b) is forward complete, there is an x such that
xn → x. Similarly if (M, b) is backward Cauchy.
Proposition 4.3 Suppose that (xn ) is forward Cauchy and there exist subsequence nk
and a point x such that limk xnk = x: then limn xn = x.
Proof. fix ε > 0; since (xn ) is forward Cauchy, ∃N such that ∀m, m0 with m0 ≥ m ≥
N , b(xm , xm0 ) ≤ ε; let H be such that nH ≥ N and ∀k ≥ H, d(x, xnk ) ≤ ε; for
n ≥ nH ,
so in conclusion d(xn , x) ≤ 2ε
Proposition 4.4 Suppose that the (asymmetric) metric space (M, b) is path-metric:
i). let ρ > 0 and
def
S + (a, ρ)={y | b(a, y) = ρ}
def
D+ (a, ρ)={y | b(a, y) ≤ ρ}
(which are closed, since b is continuous) then
13
ii). ∀θ ∈ (0, 1), ∀x, y ∈ M
Note that, if there exists a minimum point z ∈ D+ (x, ε) to (4.4.), then b(x, z) = ε,
i.e. z ∈ S + (x, ε). (14)
In the general case of a space (M, b), all the above applies to the metric space
(M, bg ) (that is path-metric, by 4.44).
On the other hand,
Proposition 4.5 if M is either forward or backward complete, and any one of the
following properties holds:
• ∃θ ∈ (0, 1) s.t. ∀x, y, property (4.4.??) holds,
• ∃θ ∈ (0, 1) s.t. ∀x, y, when ε = θb(x, y), property (4.4.) holds,
• ∃θ ∈ (0, 1) s.t. ∀x, y, when ε = θb(x, y), property (4.4.) holds,
then (M, b) is path-metric.
By using Zorn’s Lemma, the above may be strengthened to this statement
Proposition 4.6 if M is symmetrically complete, and any one of the following proper-
ties holds:
• ∀x, y, ∃θ ∈ (0, 1) s.t. property (4.4.??) holds,
• ∀x, y, there exists ε > 0, ε < b(x, y) s.t. property (4.4.) holds,
• ∀x, y, there exists ε > 0, ε < b(x, y) s.t. property (4.4.) holds,
then (M, b) is path-metric.
Regarding the completeness hypothesis, see example 4.11.
All of the above results related the existence of approximate intermediate points
to the fact that (M, b) be path-metric. If M admits geodesics, then, for any two
points x, y, there is always a curve joining them with optimal distance bg (x, y) (but not
necessarily equal to b(x, y)): this is again used to find exact intermediate points z at a
prescribed distance bg from x and y; this is at the base of the definition of “Menger
convexity”; unfortunately, there are, in the literature, two different definitions:
• M is “Menger (weakly) convex” if given two different points x, y ∈ M , a third
(different from x, y) point z exists such that bg (x, z) + bg (z, y) = bg (y, z);
(14) Indeed, b(x, z) ≥ b(x, y) − b(z, y) by triangular inequality, and b(x, y) − b(z, y) = ε by (4.4.);
14
• M is “Menger (strongly) convex” if given any two points x, y ∈ M , and any
0 ≤ λ ≤ bg (y, z), a third point z exists such that bg (x, z) = λ, bg (z, y) =
λ − bg (y, z);
the weaker definition being used for example in Busemann [4] in proposition (1.16), the
stronger in more modern treaties.
In the case when Menger convexity is intended in the weak form, the proof needs Zorn’s
Lemma. The above result is similar to proposition (1.16) in [4], (where though the
hypothesis was “M is Menger (weakly) convex and finitely compact”).
The above property (4.9.?) is stronger than the weak Menger convexity, and weaker than
the stronger. (15) Still, it is a condition that entails convexity, as seen in this example
Example 4.10 Let b(x, y) = |x − y|, M ⊂ lRn . If M is closed in lRn and (4.9.?) holds,
then M is convex (in the usual Euclidean sense).
§4.i.2 Examples
We consider several examples, for a better understanding of the above properties and
definitions.
This subset of lR2 complements the already seen example 4.7
Example 4.11 The set M = (lR × Q) ∪ ({0} × lR), equipped with b(x, y) = |x − y|,
admits approximate intermediate points, in the various forms seen in proposition 4.4;
and it is Menger weakly convex; and M is Lip-arc-connected; but (M, b) is not complete;
and (M, b) is not path-metric, since bg (x, y) = |x1 | + |y1 | + |x2 − y2 | when x2 6= y2 .
(15) Indeed “Menger weak convexity” may be equivalently stated as: ∀x, y ∈ M ∃θ ∈ (0, 1) s.t. bg (x, z) =
15
Example 4.12 Consider M ⊂ lRn to be an open set, and b to be the Euclidean distance;
then
with ∼ identifying all 0s into a class [0], and 1s into [1], i.e. (n, 0) ∼ (m, 0) and
(n, 1) ∼ (m, 1). Let ln = {n} × (0, 1). Let b be defined on segments as
def |x − y|(1 + en ) if x ≥ y and x, y ∈ ln
b(x, y)=
|x − y|(1 + e−n ) if x ≤ y and x, y ∈ ln
and
b([0], [1]) = b([1], [0]) = 1
and extend b (geodesically) using the rule
b(x, y) = inf b(x, y), b(x, [0]) + b([0], y), b(x, [1]) + b([1], y)
but
b([0], xn ) = (1 + en )/n → ∞
16
forw back
4 4
fbr(a,r) bbr(a,r)
a a
3 fbl(a,r) 3 bbl(a,r)
2 2
1 1
0 0
-1 -1
-2 -2
-3 -3
-4 -4
-4 -3 -2 -1 0 1 2 3 4 -4 -3 -2 -1 0 1 2 3 4
Figure 2: forward and backward balls, left and right extrema (M is vertical, r = 1/2, a
in abscissa)
v). Moreover the set A = {[0], xn | n} is backward bounded but not forward
bounded.
ey − ex
if x < y
b(x, y) = (4.14.?)
e−y − e−x if x > y
then b generates on lR the usual topology, and (M, b) is locally compact and path-metric
(this may be proved using 5.6 since b comes from a Finsler structure: see in sec. 5.3.1 in
[16]). The balls are the open intervals
i). We immediately note that B − (0, 1) = lR, that is, M is backward bounded (but is
not forward bounded).
17
forw back
4 4
fbr(a,r) bbr(a,r)
a a
3 fbl(a,r) 3 bbl(a,r)
2 2
1 1
0 0
-1 -1
-2 -2
-3 -3
-4 -4
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Figure 3: forward and backward balls, left and right extrema (M is vertical, a = 1/2, r
in abscissa)
ii). Let xn = n: then for m < n, b(xn , xm ) = e−m − e−n < e−m so this sequence
is backward-Cauchy: then M is not backward complete.
iii). (M, b) is forward complete (and then is symmetrically complete, by 4.2). Proof:
Suppose that (xn ) is an increasing forward-Cauchy sequence: then for m > n > N ,
b(xn , xm ) = exm − exn < ε that implies xm ≤ log(ε + exn ), so xm has a limit
xm → x (16) ; while if it were a decreasing forward-Cauchy sequence for m > n > N ,
b(xn , xm ) = e−xm − e−xn < ε that implies xm ≥ − log(ε + e−xn ), and again
xm has a limit. Suppose that xn is a generic forward-Cauchy sequence: from any
subsequence xnk of xn we may extract a monotonic sub-sub-sequence (xnk h )h : this
would be convergent to a point x; this point does not depend on the choice of subsequence:
indeed, b(xn , xnk h ) →h b(xn , x) < ε for n large.
iv). Consider the ball B + (0, ρ) of extrema (−R, R) with R(ρ) = log(ρ + 1), and
then the two balls
Proof: indeed the right extrema of B + (−R, r) is positive when log(r + e−R ) ≥ 0 that is
(r + e− log(ρ+1) ) ≥ 1 that is (r + 1/(ρ + 1)) ≥ 1
18
Example 4.15 Consider two copies (M, b) of the above space, join them at the origin,
reverse the metric on one: the resulting space M̃ , b̃ is symmetrically complete, but is
neither forward nor backward complete.
More precisely: let M+ = lR × {+}, M− = lR × {−},
M̃ = M+ ∪ M− / ∼
§4.ii Geodesics
Lemma 4.16 (change of variable) Let γ : [α, β] → M , ξ : [α1 , β1 ] → M be linked
by
γ =ξ◦ϕ
where ϕ : [α, β] → [α1 , β1 ] is monotone non decreasing, and continuous, and ϕ(α) =
α1 , ϕ(β) = β1 . Then
lenb γ = lenb ξ
Proof. For any partition T = {ti } of [α, β] we associate the partition S = {si = ϕ(ti )}
of [α1 , β1 ]: in this case,
n
X n
X
b γ(ti−1 ), γ(ti ) = b ξ(si−1 ), ξ(si ) (4.17)
i=1 i=1
Similarly, for any partition S = {si } of [α1 , β1 ] we choose ti ∈ ϕ−1 ({si }), and
associate the partition T = {ti } of [α, β]: then again (4.17) is verified regardless of the
choice of ti : indeed if t0i ∈ ϕ−1 ({si }) then γ(ti ) = γ(t0i ).
Remark 4.18 The above lemma holds also when ϕ is not bijective: this is very powerful,
since it simplifies the next lemma, and leads to a very short and simple proof of 4.24
and of 4.29.
The following is a long–known but very powerful result (that may be found in
section I in [4]; see also in Theorem 4.2.1 in [1], that, for the symmetric case, also
discusses the metric derivative issue):
Lemma 4.19 (reparametrization to arc parameter) For any rectifiable curve
γ : [α, β] → M of length L such that `γ is continuous, there exists an unique Lipschitz
curve ξ : [0, L] → M such that
19
In general we will say that a curve ξ is parametrized by arc parameter when `ξ (t) =
t. Combining (2.12) and (4.19.?), we can state that, for any curve ξ parametrized by
arc parameter, for 0 ≤ s < t ≤ L,
bg ξ(s), ξ(t) ≤ t − s
(4.20)
This leads to the following definition (due to Busemann [5], [6]):
Definition 4.21 (Segments) A curve γ : I → M (where I ⊂ lR is an interval) is a
segment if
∀s, t ∈ I, s < t , t − s = bg (γ(s), γ(t)) (4.21.?)
The “time parameter” is arbitrary: if γ(·) is a segment, then γ(· + t) is a segment as
well.
Unfortunately, a segment may fail to be continuous: this was seen in example 2.13
where the curve γ is a segment but is not continuous. The above example cannot be
found in Busemann’s general metric spaces, since it contradicts the extra hypothesis
(3.1). That example is one of the reasons why we stressed the necessity that curves be
continuous in the definition 2.10 of what a geodesic is.
Segments enjoy some nice properties:
• by 4.43, a segment is always parametrized by arc parameter, and the length of a
segment defined on I = [α, β] is β − α ;
• then a continuous segment γ is a minimal geodesic connecting γ(s) to γ(t),
∀s, t ∈ I, s < t.
Vice versa,
Lemma 4.22 (almost segment) Suppose γ : [0, L] → M is a curve parametrized by
arc parameter, connecting γ(0) = x to γ(L) = y; and suppose that
lenγ ≤ bg (x, y) + ε
for ε ≥ 0; then for any 0 ≤ s < t ≤ L,
t − s − ε ≤ bg γ(s), γ(t) ≤ t − s
(4.22.?)
(that is an approximate version of (4.21.?)); the rightmost inequality is (4.20); whereas
bg (x, y) ≤ bg x, γ(s) + bg γ(s), γ(t) + bg γ(s), y ≤
≤ s + bg γ(s), γ(t) + L − t .
again by (4.20).
So (choosing ε = 0) we obtain that a minimal geodesic is a continuous segment.
Segments can be joined, as follows:
Lemma 4.23 (joining) Suppose that γ : [0, L] → M , γ : [0, L0 ] → M , are geodesic
segments. Suppose that z = γ 0 (0) = γ(L): then the segments may be joined to form a
continuous curve ξ : [0, L + L0 ] → M , by defining
γ(t) for t ∈ [0, L]
ξ(t) =
γ 0 (t − L) for t ∈ [L, L + L0 ].
Then ξ is a geodesic segment if and only if
bg ξ(0), ξ(L + L0 ) = L + L0 .
20
§4.ii.1 Existence of geodesics
The following results rephrase (5),(6),(7) in section I in [6], but without assuming that
(M, b) be path–metric (by properly distinguishing b by bg ):
is compact (in the (M, b) topology (17) ). Then for any y such that bg (x, y) ≤ ρ, there is
a geodesic connecting x to y.
Note that it is not sufficient to assume that D+ (x, ρ) is compact (instead of D+g ):
see in example 4.34. It is instead enough to require that D+ (x, ρ) is compact for all
ρ > 0 (as noted also in remark 4.3.3. in [1]).
In the same spirit we state those corollaries
• As a consequence, if D+g (x, 2ρ) and D−g (x, 2ρ) are compact, and z, y ∈
D−g (x, ρ) ∩ D+g (x, ρ), then z, y may be joined by a geodesic.
We now prove the theorem; this needs some extra steps w.r.t. the proof in General
Metric Spaces (where (3.1) is assumed to hold).
Proof. Fix x, y ∈ M , ρ > 0, as above. We will write D+g instead of D+g (x, ρ) for
brevity. By lemma 3.7, let ω be the modulus of symmetrization of D+g (x, ρ); let
def
ω̃(r)= sup{r, ω(r)}: then
d(z, y) ≤ ω̃ b(z, y) (4.26)
for any z, y ∈ D+g .
Let L = bg (x, y); suppose y 6= x (otherwise γ ≡ x is the geodesic). Let γn :
def
[0, Ln ] → M be a sequence of continuous paths from x to y such that Ln =lenγn → L.
By using the above lemma 4.19, we assume without loss of generality that γn are
parametrized by arc parameter; by (4.26) above,
d γn (t), γn (s) ≤ ω̃(|t − s|) (4.27)
for all s, t ∈ [0, Ln ]. To confront different paths, we extend so that γn (t) = γn (Ln ) for
t > Ln .
Suppose also that L = bg (x, y) < ρ for simplicity; then definitively Ln ≤ ρ: by
(2.12), we know that all of γn is contained in D+g . Combining this argument and
(4.27) we can apply the Ascoli-Arzelà theorem: we know that there is a γ : [0, L] → M
(that again satisfies (4.27)) such that, up to a subsequence, there is uniform convergence
of γn → γ; this uniform convergence is w.r.t the distance d(x, y) = b(x, y) ∨ b(y, x).
(17) We recall that if D ⊂ M is compact in (M, bg ) then it is compact in (M, b); the opposite is not true, as
21
The functional γ 7→ lenb γ is lower semicontinuous w.r.t. uniform convergence, since it
is the supremum of the continuous functionals
n
X
b γ(ti−1 ), γ(ti )
i=1
since γ̃n are wholly contained in D+g , we can apply the above reasoning to say that
there is a continuous γ̃ such that γ̃n → γ̃ uniformly. To conclude the proof we need to
prove that γ̃(L) = y: since L = ρ ≤ tn ≤ Ln , then tn → L; moreover
This implies that R(y) ≥ ρ − b(x, y), and then R(y) ≥ R(x) − b(x, y); if R is
finite, the above entails d(x, y) ≥ |R(x) − R(y)|, since x, y are arbitrary.
In general (even when R(x) < ∞) it is possible to find examples where D+ (x, R(x))
is compact, and examples where it is not. In path–metric and forward-locally compact
spaces, instead, we can precisely describe the behaviour of R(x) as follows
22
Lemma 4.31 Suppose that (M, b) is path–metric and forward-locally compact. Choose
def
x ∈ M, ρ > 0 such that R(x) < ∞ and D+ (x, ρ) is compact; let δ = miny∈D+ (x,ρ) R(y)
(note that δ > 0); then R(x) = ρ + δ.
Proof. Since (M, b) is forward-locally compact, then δ > 0. Choose 0 < t < s < δ;
we want to prove that [
def
V= D+ (x, t)
y∈D + (x,ρ)
A corollary of the above lemma is that (in the above hypothesis) D+ (x, R(x)) is not
compact; so the above lemma is the quantitative version of the argument shown in (8) in
section I in [6].
Lemma 4.32 Suppose that (M, b) is path–metric, for simplicity (18) . Suppose that
0 < R(x) < ∞, let ρ = R(x); then for any y with b(x, y) = ρ there is a continuous
geodesic segment γ : [0, ρ) → M such that γ(0) = x and
Proof. Let z1 = x, and ρn = (1 − 2−n )ρ; we would like to define zn+1 as a minimum
point for the problem
min −n b(z, y) ; (4.32.??)
+ z∈D (zn ,2 ρ)
iteratively, using 4.4, we may prove those facts: b(x, zn ) ≤ ρn−1 and
where the r.h.s is compact: so the above problem has a minimum zn+1 that satisfy
b(zn , zn+1 ) = 2−n ρ, b(x, zn+1 ) ≤ ρn and
Using (4.32.) and 4.24, there exists a geodesic segment γn connecting zn to zn+1
(of length 2−n ρ). By using (4.32.??),(4.32.) and the triangle inequality for h > n
23
but on the other side
h−1
X h−1
X
b(zn , zh ) ≤ b(zj , zj+1 ) = ρ2−j
j=n j=n
we obtain that
b(zn , zh ) = (2−n+1 − 2−h+1 )ρ
so by the joining lemma 4.23 we can join all the segments γn to build the required
segment γ. By (4.32.), we obtain (4.32.?).
The example 2.13 shows that we cannot expect that γ(t) be continuous at ρ in general:
indeed the “identity curve” γ : [0, 1] → [0, 1] is a segment but is not continuous at 0.
We now use the above lemmas to prove the asymmetric Hopf–Rinow theorem (in a
more detailed version than what was seen in the introduction):
Theorem 4.33 (Hopf-Rinow) Suppose that (M, b) is path-metric and forward-locally
compact; then the following are equivalent
i). forward-bounded and closed sets are compact,
ii). (M, b) is forward complete,
iii). any continuous geodesic segment γ : I → M defined on I = [α, β) may be
completed to a continuous geodesic segment on [α, β],
and all imply that (M, b) admits geodesics.
In general, the implications i=⇒ii=⇒iii hold for any asymmetric metric space.
Proof. • Suppose that forward-bounded closed sets are compact. If xn is a forward-
Cauchy sequence, then there exists N s.t. b(xN , xm ) ≤ 1 for m > N , that is,
xm ∈ D+ (xN , 1) that is compact; then we can extract a converging subsequence,
and use lemma 3.7 to obtain the result.
• Suppose that (M, b) is forward complete, let γ be the segment; then as tn ↑ β,
the sequence γ(tn ) is forward–Cauchy, so there is an unique limit limn γ(tn ) that
we use to define γ(β).
• Suppose that any segment γ : I → M defined on I = [α, β) may be completed:
we will prove that forward–bounded closed sets are compact (that is, that the
radius of compactness R(x) ≡ ∞).
We proceed by contradiction, and suppose that there is a x s.t. ρ = R(x) < ∞:
we will show that D+ (x, ρ) is compact, contradicting 4.31. To this end let
{yn } ⊂ D+ (x, ρ). Let Ln = b(x, yn ).
If lim inf n Ln < ρ, we can extract a subsequence nk s.t. Lnk ≤ t < ρ, that is,
{ynk } ⊂ D+ (x, t) that is compact: so we can extract a converging subsequence.
Suppose now that limn Ln = ρ. For any yn s.t. Ln < ρ, we use 4.24 to
obtain the minimal geodesic segments γn : [0, Ln ] → M connecting x to yn ;
we extend γn constantly to [Ln , ρ] (as was done in eqn. (4.28)). When instead
Ln = ρ we use 4.32 to define γn : [0, ρ) → M , and define γn (ρ) = yn . We
use Ascoli–Arzelà theorem and a diagonal argument to find a subsequence nk
and a curve γ : [0, ρ) → M such that for each 0 < s < ρ, limk γnk (t) = γ(t)
24
uniformly for t ∈ [0, s]. Since the length functional is lower-semicontinuous,
we obtain that γ is a continuous geodesic segment on [0, ρ); by hypothesis, it can
be completed to a continuous segment on [0, ρ]. Let y = γ(ρ) Fix now ε > 0;
since γ is continuous, there is a t s.t. b(y, γ(s)) < ε for all s, t ≤ s ≤ ρ; fix
s = t ∨ ρ − ε; we apply the triangular inequality to prove that
b(y, yn ) ≤ b(y, γ(s)) + b(γ(s), γn (s)) + b(γn (s), yn ) ;
this is less than 3ε, definitively in n; so b(y, yn ) → 0, and then yn → y (by
prop. 3.6).
• Existence of geodesics is guaranteed by theorem 4.24.
We remark that the proof of the above equivalence cannot simply follow from the
proof for metric spaces §1.11 in [12], since that proof uses the property (ii) in sec. §3.iii;
neither it does follow from the proof in Finsler Geometry (see section VI of [3]), since
the latter uses the exponential map.
Hence we devised a different proof, that combines the idea of the diagonalization of
a sequence of subsequences, as in [12], and the idea of repeated application of (4.4.)
as in [3]; it is similar to the proof for general metric spaces in §I in [6], but it does
not use the hypothesis (3.1), and it exploits the special ingredients of the radius of
compactness Lemma and of 4.3.
25
A B
§4.iii Complements
§4.iii.1 Generating (path) metrics
For any asymmetric metric b, we may build many other asymmetric metrics b̃, as follows
Proposition 4.36 Let ϕ : lR+ → lR+ be continuous and concave, ϕ(x) = 0 only for
def
x = 0; let b̃=ϕ ◦ b: then b̃ is an asymmetric distance.
def
The above family b̃=ϕ ◦ b is quite large; hence the following question arises: what
happens if we induce b̃g from b̃: can we build a large family of path metrics on M , by
this simple procedure? The answer is no.
Proposition 4.37 Set everything as in the previous proposition. If moreover the deriva-
tive of ϕ exists and is finite at 0, then
ϕ0 (0)lenb γ = lenb̃ γ
and
bg (x, y)ϕ0 (0) = b̃g (x, y) .
Proof. Let ε > 0. In the definition (2.5) of lenb γ it is not restrictive to use only subsets
T of [α, β] such that b(γ(ti ), γ(ti+i )) ≤ ε ∀i ∈ {1, . . . , n − 1}.
Let now δ > 0; then there exists a ε > 0 such that
26
• C connects points: ∀x, y ∈ M there is at least one γ ∈ C, γ : [α, β] → M such
that γ(α) = x, γ(β) = y
def
• C is reversible: if γ ∈ C and γ̂(t)=γ(−t) then γ̂ ∈ C
• len is independent of reparametrization: if γ ∈ C then, ∀h > 0, k ∈ lR, its linear
reparametrization γ 0 (s) = γ(hs + k) is in C, and lenγ = lenγ 0 ,
• len is monotonic: if γ ∈ C then its restriction γ 0 = γ|[c,d] is in C, and lenγ 0 ≤
lenγ
• len is additive: if γ 0 , γ 00 ∈ C, γ 0 : [α0 , β 0 ] → M , γ : [α00 , β 00 ] → M , and
β 0 = α00 , then we may join the two curves and obtain γ: then γ is in C, and
Lemma 4.39 Suppose len and C satisfy all the hypotheses above. If len is lower semi
continuous w.r.t. uniform convergence, then lenγ = lenb γ for any γ ∈ C
The proof for the asymmetric case is not different from the symmetric case (cf. 1.6 [12]).
We provide here a detailed proof, for convenience of the reader.
Proof. Fix γ ∈ C. We insert len into the definition (2.5) of lenb γ:
n
X
lenb γ = sup inf lenξi (4.40)
T ξi
i=1
where the sup is done on all choices of T finite subset of [0, 1] and the inf is done
on the class ΞT of ξ ∈ C, ξ : [0, 1] → M , such that ξ(t) = γ(t) for t ∈ T .
Since we could choose ξ = γ, we immediately obtain that
27
Fix a T : by (4.41),
inf lenξ ≤ lenb γ
ξ∈ΞT
Let η > 0. Define ΞT,η as the class of ξ in ΞT such that lenξ ≤ lenb γ + η and, for
any interval [ti , ti+1 ] defined by T ,
so we can write
sup inf lenξ = lim inf lenξ
T ξ∈ΞT T →∞ ξ∈ΞT
satisfying all above hypotheses. len induces b which induces lenb γ. lenb γ is lower
b g
28
This immediately entails the result:
Corollary 4.44 The operation b 7→ bg is idempotent, that is, bg is path-metric, idem est
bg = (bg )g .
5 Applications
§5.i Asymmetric metric manifold
Suppose now that M is a differential manifold (20)
, with an atlas A and a topology τ M .
Definition 5.1 (A-Lip) We will say that a curve ξ : [α, β] → M is locally Lipschitz
w.r.t. A, if for any local chart ϕ : U → lRn in the atlas A, ϕ ◦ ξ is locally Lipschitz for
t ∈ ξ −1 (U ).
• F (x, v) = 0 iff v = 0.
29
Lemma 5.5 Assume 5.2 and let F be locally bounded from above. Then for any
z ∈ M there exists a compact neighborhood U of z, associated with local coordinates
ϕ : U → W ⊂ lRn , such that the push forward of b in local coordinates is bounded
from above and below by the Euclidean norm: that is, ∃c0 > c > 0
where t = 1 if ξ never exits from Z, otherwise it is the first time s such that
ξ(s) ∈ ∂Z: then b(x, y) ≥ c(ε ∧ |x̂ − ŷ|), and we choose a smaller U .
Corollary 5.6 Under the same hypotheses, τ b = τ M , i.e. b generates the same topology
that the atlas of M induces. Moreover both coincide with the topology generated by the
forward balls; or respectively, by the backward balls. (22)
This topology is locally compact: then
(23)
• xn → x iff b(xn , x) → 0 iff b(x, xn ) → 0 ; and
• (M, b) locally admits geodesics (by 4.24).
By (5.5.?), a curve ξ : [0, 1] → M is locally Lipschitz w.r.t A, as defined in 5.1, iff it
is locally Lipschitz w.r.t b, as defined in §2.ii.
Proof. i) is a consequence of 4.4. (24) By prop. 4.4, we need to check that ∀x, y ∈
M, ∀ε > 0 ∃z ∈ M such that b(x, z) < b(x, y)/2 + ε and b(z, y) < b(x, y)/2 + ε. Fix
x, y. Let ε > 0 and ξ connecting x to y such that lenL ξ < b(x, y) + 2ε; let
Z t
f (t) = ˙
F (ξ(s), ξ(s)) ds
0
(22) cf. 3.8 , or sec. 6.2C in [3] for the regular case
(23) cf. 6.2.5 in [3] for the regular case
(24) If v 7→ F (x, v) is convex, then this is an immediate consequence of the second point. Alternatively, we
have a proof that does not need that L(x, ·) is convex: indeed the operation of computing b implicitly replaces
F by its relaxed l.s.c.
30
then f is continuous, so let t0 be the time when f (t0 ) = f (1)/2, and z = ξ(t0 ); by
reparametrizing, ξ 0 (s) = ξ(st0 ), ξ 00 (s) = ξ(1 − (1 − s)(1 − t0 )); then since b(x, z) <
lenL ξ 0 = f (1)/2 < b(x, y)/2 + ε, and b(z, y) < lenL ξ 00 = f (1)/2 < b(y, z)/2 + ε.
ii). The proof is similar to the proof of lemma 4.39. We set C to be the class of
Lipschitz curves, and len=lenL . Fix γ, let L=lenL γ. If L is small, then the curve is
def def
Further applications of these ideas are in [16]; in particular, in the appendix of [16],
2
we consider a Lagrangian L : T M → lR+ (that is defined as L(x, v) = F (x, v) ) and
∗ +
its Legendre-Fenchel dual Hamiltonian H : T M → lR ; we show that, if H ∈ C 1,1
and p 7→ H(x, p) is positively 2-homogeneous and strictly convex, we can define an
exponential map and add it to the statement of the Hopf-Rinow theorem, to obtain a
statement exactly as in ch. VI of [3].
31
Acknowledgements
The author thanks Prof. S. Mitter, who has reviewed and corrected various versions of
this paper, and suggested many improvements; and Prof. L. Ambrosio for the discus-
sions.
Contents
1 Introduction 1
§1.i The Hopf–Rinow theorem . . . . . . . . . . . . . . . . . . . . . . . 2
§1.ii asymmetric metric theories in Calculus of Variations . . . . . . . . 3
5 Applications 29
§5.i Asymmetric metric manifold . . . . . . . . . . . . . . . . . . . . . 29
References
[1] L. Ambrosio and P. Tilli. Selected topics in ”analysis in metric spaces”. appunti.
edizioni Scuola Normale Superiore, Pisa, 2000.
[2] Luigi Ambrosio, Nicola Gigli, and Giuseppe Savarè. Gradient flows in metric
spaces and in the Wasserstein space of probability measures. Birkhüser, 2004.
32
[4] H. Busemann. Local metric geometry. Trans. Amer. Math. Soc., 56:200–274,
1944.
[5] H. Busemann. The geometry of geodesics, volume 6 of Pure and applied mathe-
matics. Academic Press (New York), 1955.
[8] A. Duci and A. Mennucci. Banach-like metrics and metrics of compact sets. in
preparation, 2007.
[14] J. C. Kelly. Bitopological spaces. Proc. London Math. Soc., 13(3):71–89, 1963.
33