This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
(1 + \yncn JM 0 and _. It follows from (4.16)-(4.17) that 2. Let [0, oo) by (¥>t)*(d"tU(„)) =Vt(v) (dvP\y), 0,ti
V
fdVp)dt,
(3.29)
J ¥ >~ 1 (*)
r J-oo
( f V
fdVp)dt.
^v-1(t)
(3.30)
'
There are many important applications of Theorem 3.3.2. For example, one can use it to extend Gromov's estimates on the Filling radius of Riemannian spaces to Finsler spaces [Grl]. Below is a simple application of Theorem 3.3.2 Proposition 3.3.3 Let S be the indicatrix in a Minkowski space (V, F). Let F denote the induced Finsler metric on S. Then 2-"nw„_ 1 < Voli?(S) < (1 + A) n nw„_ 1 , where A is given in (3.28) and w„_i = Vol(B n
(3.31)
).
Proof. Let p(x) := F(x) denote the distance function from the origin. By definition, the volume of the unit ball B in (V, F) is equal to Vol(B ). Let dVp denote the volume form of the induced Finsler metric F on S(r). Note that Volp(S(r)) = r"- 1 Vol J? (S). It follows from (3.29) that Vol(B") = VoliKB)
<
2 n c„ /
Volp(S(t))dt
Jo
= =
2ncn f Jo
tn-1Vo\F(S)dt
fL^Volp(S).
This gives Vol^(S) > 2- n nVol(B n )/c n = 2- n nVol(B n
1
).
The proof of the inequality on the right hand side of (3.31) is similar, so is omitted. Q.E.D.
Co-Area
50
Formula
When the Minkowski norm is not reversible, the volume Volj?(S) has no universal upper bound. Example 3.3.1 FE on R 2
For 0 < e < 1, consider the following Minkowski norm Fe(u, v) := v u2 + v2 — eu.
The indicatrix SE = F~l(l)
is an ellipse,
Parametrize Se by 1
u=-
£
„
jCos6>+g, 1 - e2 1 - e2
l
v—
•
n
= smfl, VI - e2 r
so that FE{u,v) = y ' ( I - £ - I ) 2 s i „ 2 ^ + - J - j + This gives
Voi^(s£) = fj J^y^e
^ s i n ,
+ ^dB >v T ^ t
- 4 OO,
05 e —> 1 . Compare [Ma2j. Let (V, i*1) be a Minkowski space. F induces a Riemannian metric g on V - {0} b y g(u,v) :=gy(u,v),
u,v€TyV
= V.
Let (yl) be a global coordinate system in V associated with a basis {k>i}™=1. The Riemannian volume form of g is given by dVa =
1
y/det{gij(y))dy
---dyn,
where gi:j(y) = g y (b;,b.,). Let g denote the induced Riemannian metric on S := F~~x(\). Identifying V \ {0} with (0, oo) x S in a natural way, i.e., y = (F(y),y/F(y)). Since
Co-Area
gv(y,w) form.
Formula
51
= 0 for any w £ TyS C V, g can be decomposed to the following g = dt2® t2g.
(3.32)
Denote by Vol r (B) and Volr(S) the Riemannian volume of (B, g) and (S, g), respectively. By (3.32), f1
1 V o l r ( B ) = / tn~1Vo\r{S)dt=-Volr(S). Jo n
(3.33)
Recall that for the induced Finsler metric F on S, the volume Vol^(S) has a universal lower bound, but no universal upper bound. In addition, if F is reversible, then Vol^-(S) has a universal upper bound. For the induced Riemannian metric g on S, we have the following P r o p o s i t i o n 3.3.4 Let (V, F) be an n-dimensional reversible Minkowski space. Then the Riemannian volume of the indicatrix S satisfies Vol r (B) < Vol(B n _ 1 ),
Volr(S) < Vol(S n _ 1 ).
Equality holds if and only if F is Euclidean . Proof. Let {bi}" = 1 be a basis for V and {0l}™=1 the dual basis for V*. Let B" := {(), Ftfhi)
< l},
B*" := {(fc), F*(CiOl) < l } .
By Santalo inequality [MePa] which requires the symmetry of B n and B* n . Vol(Bn)Vol(B*n) < Vol(B"). Equality holds if and only if B™ is an ellipsoid. The Legendre transformation I: B n ->• B* n given by yl -^-d has Jacobian det(gij(y)).
/
•=9ij(y)yj
Thus
det(9ij(y))dy = f
d£ = Vol(B*n).
52
Co-Area
Formula
Observe that VoL-(B)
=
/
Jdet(gij(y))dy
<
[J ^ det(9ij (y))dy\
-
[vol(B* n )Vol(B n )l"
<
Vol(B n ).
1/2
Vol(B") 1 / 2
Equality holds if and only if det(gij(y)) = constant and B n is an ellipsoid. This completes the proof. Q.E.D. The idea of the proof is borrowed from [Du], where C. Duran established a sharp upper bound on the volume of the Sasaki metric on the unit tangent sphere bundle. Here we are only concerned with the volume of the induced Riemannian metric on the indicatrix. When F is not reversible, then Volr(S) can be greater than Vol(§ n ~ ). Consider the case when dimV = 2. Let c(t) = u(t)hi + v(t)b2,
a
be an arbitrary parametrization for the indicatrix S = i ? _ 1 ( l ) . For any vector v = Xc(t) + (c(t) G Tc{t)V, .
,
.
x2
u'(t)v"(t)-u"(t)v'(t)
~
.on„.
This gives a formula for the Riemannian perimeter L r (S) with respect to g, rb
MS) = I Ja
E x a m p l e 3.3.2
U'{t)v"(t) u(t)v'(t) -
u"(t)v'(t) dt. u'{t)v(t)
Consider the following Randers norm on R 2 , FE(u, v) := V u2 + v2 — eu,
where 0 < e < 1. Parametrize S£ = F'1^) 1
, x
i - e
2•cos(t) w
by c(t) =
£
+ -i -
(u(t),v(t)),
1
e
2j
',
v=
yr
^sin(t).
Co-Area
Formula
53
By (3.34), we obtain
The Riemannian perimeter L r (S £ ) of S e with respect to gE is r27T
L r (S £ )
=
/ Jo
^dt
\/l + + ££COs(t) A/1
= 2r/2{ 7o >
27T
y
i1
+ , yjl
i /1l — £COs(t)
V
-ecos(t)
* u + £ COs(i)
Vl+£cos(t)J
Chapter 4
Isoperimetric Inequalities
The classical isoperimetric inequality for the standard unit sphere S™ C K. states as follows: for any domain fl C § with regular boundary d£l,
S> Vol(S n ) -
hsn(s) S W
'
s-
V
°K»)
Vol(S n )'
where /is»(s) denotes the ratio Vol(9Bs)/Vol(S™) for a geodesic ball B s C S with Vol(Bs) = s • Vol(S n ). The function /ign is called the isoperimetric profile of § . The isoperimetric profile can be defined for Finsler m spaces. In this chapter, we will discuss the relationship between the isoperimetric profile and the Sobolev constants (first eigenvalue) of a Finsler m space. The classical isoperimetric inequality is used by P. Levy to discuss his concentration theory [Le]. Levy's concentration theory is developed further to Riemannian spaces and more general metric measure spaces [GrMi][Gr4]. In his study on the concentration of metric and measure spaces, M. Gromov introduces many new geometric invariants. Among them are the two basic quantities: the expansion distance and the observable diameter [Gr4]. In this chapter, we will discuss the relationship between the isoperimetric profile and these two quantities. 55
56
4.1
Isoperimetric
Inequalities
Isoperimetric Profiles
Let (M, F, d/j,) be a Finsler m space. We define the isoperimetric profile hM • [0,1] ->• [0, oo) by (4.1)
hM{s):=mi
iWy
where the infimum is taken over all regular domains Q C M such that /i(fi) = s • n(M), and v denotes the induced measure on dfl (cf. (2.23)). Note that hM(s) = hM(l - s) > 0,
Vse(0,l)
and hM(0) = 0 =
hM(l).
(j,(£l) = s • fi(M
Therefore we make the following Definition 4.1.1 A continuous function h : (0,1] isoperimetric function if it satisfies h(s) = h(l-s)
> 0 , V s e (0,1)
and h(0) = 0 = h{l).
R_i_ is called an
Isoperimetric
Profiles
57
Given an arbitrary isoperimetric function h, we will construct an open interval (-L, L) and a function Oh{t) on (-L, L) such that /
crh(u)du = s,
ah(t) = h(s).
(4.2)
First, we try to find a solution s = s(t) to the following O.D.E. ds dt s(0)
= h(s)
(4.3)
2'
s 1I
s = s(t) 1
F
-L
- t *
Let r)(s) :=
h(u)
du.
(4.4)
77 is an increasing C 1 function with a symmetric range (—L, L). Let
sW:=^W Then s(t) is a solution to (4.3). Let
ah(t)~h(S(t))
= h(V-l(t)).
Since the range of s(i) is the open interval (0,1), we have lim a hit) = 0. t—>±L
(4.5)
Isoperimetric
58
Inequalities
Further, s(t) = J
°h(t) = /i( /
ah{u)du,
ah{u)duj
= /i( /
(4.6)
ah{u)duj.
This gives (4.2). We can view the Finsler m space ((—L,L),\ • \,ah(t)dt) dimensional model for h.
4.2
(4.7)
as a one-
Sobolev Constants and First Eigenvalue
There are many important geometric invariants for Finsler m spaces. Let (M, F, dfi) be a closed Finsler m space. For 1 < p, q < oo, define
A Pl ,(Af):=
y
inf.
u6C°°(M) i. n f
- '
—
—
—'
u
A p
^R(^)/M! - l ^):
\p,q(M) is called the (p,q)-Sobolev constant. Another geometric invariant is the the first eigenvalue of defined by
Ai M :=
inf
fM\F*(du)\2dn . /M' \ ;' *
(M,F,dfi),
, .
x
4.8
Clearly, Ai(M) = [A 2 , 2 (M)] 2 . The Sobolev constants XPiq(M) are closely related to the isoperimetric constants (cf. [Be] and [Ga] in the Riemannian case). In what follows, we shall establish some relationships between the Sobolev constants and the isoperimetric constants for Finsler m spaces.
Sobolev Constants
and First
Eigenvalue
59
For an isoperimetric function h and 1 < p, q < oo, define
(/oVM'M*)^)17' XPtq(h) := inf — ^ 0
^-—,
(4.9)
(infA6R/>(s)-A|W)
where the infimum is taken over all bounded piecewise C 1 -functions
•
\,ah(t)dtj.
By symmetrization based on the one-dimensional model, we can prove the following Theorem 4.2.1 ([GeSh]) Let (M,d,fi) be a closed Finsler m space. Suppose that fiM > h for some isoperimetric function h. Then A P ,,(M) > A M (/i).
(4.10)
In particular, Ai(M) > [A2,2(/i)]2. Proof. Under the assumption hM > h, for any regular domain fi c M, fi(M) ~
\n(M)/'
Now we process to prove (4.10). Take an arbitrary function u G C°°(M). Since u can be approximated by Morse functions, we may assume that u is a Morse function on M. Decompose u into the difference of the positive and negative parts, u = u+ — u_. For a > 0, let fl+(a) := {x G M,
u+(x) > a}.
Define t+ : [0,supw + ) —> (—L,L) by
fV"'"""^ Jt+
f"1'
60
Isoperimetric Inequalities
The function t+ is an increasing piecewise Cl function. By assumption h-M > h, we have that for a > 0, M(M)
>
/i
=
/i( / Wt + (a)
-
V V u(M) M(M)
=
)
ah{t)dt) ' CTfc(t+(a)).
(4.12)
The last equality follows from (4.7). Define ip+ : (—L,L) —> R + by setting
t'+(aY We know that (4.11),
f°° f /
F(VM+)
= F*(du+).
wn—^)du
By the co-area formula (3.25) and
= ^+( fl )) = ^(M) /
**(')*• (4-13)
Differentiating (4.13), one obtains /
*
dv =
tiM)oh(t+{a))t+(a).
(4.14)
The Holder inequality yields
„(«,+ (»)) = j
lF*(du+)]'?(b/
Jan+(a) [F*(du+)] « < ( f
[F*(du+)]^y(
[F*(du+)]-ldv)'~\
f
(4.15)
It follows from (4.12) - (4.15) that / [F*(du+)}qdn JM
=
[ ( f JO \Jdn+(a)
[F*{du+)]q-ldu)da ' ' F
d
1
{hn+(a) *( ^r ^)
1-1/9
Sobolev Constants and First
Eigenvalue
61
(Th{t+{a))
> n(M) f Jo
1-1/9.
L
(^(t + (a))t;(o)) J
fx
1
= ^( M )y o
^
=
/x(M) /
v+(*+(a))
=
n(M)J
(a) ] 9 ^(
f
+( ffl )) f +(°) da (7/,(t + (o))t' + (a)da
\d
That is, (4.16)
||F*(du+)||,>/i(M)«||dV+|
Let A be an arbitrary number. By (4.14) and the co-area formula again, we obtain / |u+-A|"d/i Jn+(o)
=
f°|a-A|"[/ l Jo Jdn+(a)F
=
n{M)
) dv\da idu+) i
\a-\\v<jh(t+{a))t'+{a)da Jo r°° i
=
/x(M) / Jo
ip
\tp+(t+{a)) - X\ crh{t+{a))t'+{a)da ' '
= KM) I
\
Jt+{0)
That is, \W+ - A|| LP(n+ (0)) = KM)*
\\
Similarly, for the negative part w_ and a > 0, let n_(a) := {x £ M, u _ ( i ) > a}. Define t_ : [0,oo) -> ( - L , L ) by
L
ah{t)dt
--mr-
Define
62
Isoperimetric
Inequalities
\\F*(du-)\\q>^My/*\\d^\\g.
(4.17)
Note that t~(a) < t+(a), Va > 0. Set
\\u-x\\p =
f,(M)1/"y-x\\p,
\\F%du)\\q > ^(M^Wd.pW,. Therefore
"^''"-'"iPIrs K*h-
(4 18)
-
Take the substitution s = s(t) and let
(/fi iv'wi'^wdt)1^
(/0: mswhwds)1'9
( i n f W - L | v ( t ) - \\Voh{t)dtf'P
( i n f ^ R / o 1 | # s ) - A|Pd S ) 1 / P (4.19)
(4.18) and (4.19) imply A Pl ,(M) > A Pi ,(h). Q.E.D. For an isoperimetric function h, define MM)
:= i a f - p n
^
r
,
(4.20)
/I(M(«)/M(M)JM(M)
where the infimum is taken over all regular domains fi in M. lh(M) is called the h-isoperimetric constant of M.
Sobolev Constants
and First
Eigenvalue
63
Take an arbitrary domain Q,
h(s)n(M)
> Ih(M).
This gives /i(M)
>lh(M)h(s).
Thus hM(s) >
Ih(M)h(s).
(4.21)
Consider the following isoperimetric function ha(s) := min(s, 1 — s)i - i
1 < a < oo.
Here we assume that ^ = 0 if a = oo. It can be shown that if — > - — -, a
then \p,q(ha)
a — q
p'
> 0. In particular, A2>2(^Q) > 0 for any 1 < a < oo. Let Ia(M)=Ih„(M).
We call / ^ ( M ) the Cheeger constant. The isoperimetric profile of a Riemannian manifold can be estimated under a lower Ricci curvature and an upper diameter bound. This is done by M. Gromov using Levy's method. Gromov applies his estimates on the isoperimetric profiles to obtain some estimates for the eigenvalues of the Laplacian on a Riemannian manifold. See [Gr4] [Gr5] for details.
64
4.3
Isoperimetric
Inequalities
Concentration of Finsler m Spaces
Let R°° be an infinite-dimensional Hilbert space and S°° the unit sphere in R . Consider a finite-dimensional unit sphere S in R , which is the intersection of S°° with a finite-dimensional subspace in R . Let us take a look at this sphere S. The diameter of S is always equal to 7r and the sectional curvature of S is always equal to one (see Section 6.1 for the definition of sectional curvature). A natural question arises: How can we detect the dimension of S and how to tell the difference between two finitedimensional unit spheres in R ? We have a satisfied answer to this problem using the Levy concentration theory of mm spaces [Le]. The Levy concentration theory of mm spaces has been developed further by M. Gromov [Gr4]. He introduce several geometric quantities for mm spaces. Among them is the so-called expansion distance. Definition 4.3.1 Let (M,d,n) be an mm space with fx{M) < oo. For 0 < s < \ , the expansion distance ExDist(M;e) is defined to be the infimal p such that the following statement holds
If a subset A C X satisfies
H(A) > efi{M),
then
n(up(AJ) > (1 - eMM). Here Up(A) denotes the p-neighborhood of A in M,
UP(A) := lx e A, d{A,x) < p\.
Concentration
of Finsler m Spaces
65
The expansion distance tells us how the mass concentrates, with respect to the underlying metric. We say that a sequence of mm spaces (Mi, di,fH) concentrates to a point if ExDist(Xi; e) -> 0 as i -> oo for any e € (0,1/2). A trivial inequality is that ExDist(M;e) < Diam(M). Thus if the diameter of (M i? du /x») converges to zero, then (M i5 di^i) concentrates to a point. There are sequences of mm spaces which do not collapse to a point, but concentrate to a point. E x a m p l e 4.8.1 Let S* denote the standard unit spheres in R* C E°° with the induced Riemannian metric and Riemannian volume form. We have ExDist(S^e)«^, yji
where C(e) depends only on e £ (0,1/2). Thus the sequence {S*} concentrates to a point as i -+ oo, while, {S*} does not collapse to a point.
dim —» oo
8
66
Isoperimetric
Inequalities
The above concentration phenomenon was first discovered by Paul Levy [Le]. Levy explained this by using the sharp isoperimetric inequality for S . Levy's result has been generalized to Riemannian spaces by M. Gromov and V. D. Milman [GrMi]. They proved that for any closed Riemannian spaces {M,g), ExDist(M;e) < -
C
& -
where \\(M) denotes the first eigenvalue of X (see also [Gr4]. The same method can be carried over to Finsler spaces, due to Lemma 3.2.3. Theorem 4.3.2 0 < e < 1/2,
Let (M,F,d/j.)
be a closed Finsler m space. For any
ExDist(M;£) < 2\ — - 1
,
Proof. Fix p > 0 and 0 < e < 1/2. Let A c M be any closed subset and UP{A) the p-neighborhood of A. Define
u{x). im
- ^
where Vp := ii(Up(A)) and V0 := n(A). Let V = fi(M). By Lemma 3.2.3, we see that f -7T1
x € UJA) - A py
F*(du) = F(Vu) - I 7^va \
0
'
X€AU(M-UP(A)):
This
[ [F*(du)}2dn= [
_ J _ d M = i.
Let A € R be an arbitrary number. Note that / JUP(A)
|« - \\2d(i > J \u- \\2dfj, = JA
X2n(A).
Concentration of Finsler m Spaces
67
Thus / \u-X\2dfi JM
f \u-X\2dfi+ JM~Up P
=
[ Jup
\u-\\2dn
* ( A - V ^ ) 2 ( y - ^ + A ^° (V - VP)V0 2 (VP-V0)(V-VP + V0 -P
> We obtain X (M\ <
/ M ^ W ^ M
(VP-Vo)(V-VP + Vo)-2
<
M(M)
-MX€RfM\u-\\*dp-
{V-Vp)Vo
P
•
In other words ' (Vp - V0)(V - Vp + Vq)
>*f
-v,m
L
vsm-
(422)
'
For any p < ExDist(M;£), there is a subset A satisfying V0 > eV,
VP<(1-
e)V.
Then by (4.22), we obtain l
P<2\/^"1 2e
y/%{M)
This implies that ExDist(M;e) < 2W — - 1 2e vOi(M) Q.E.D. The expansion distance is also controlled by the isoperimetric profile. Theorem 4.3.3 Let (M, F, d(i) be a closed Finsler m space. Suppose that hM > h, where h is an isoperimetric function. Then for any 0 < e < j , ,i-e
ExDist(M;£) < / Js
1
TT^duh(u)
( 4 - 23 )
68
Isoperimetric
Inequalities
Proof. Let A C M be a closed subset and Vp := /i(fp(^)),
V0 = M ( 4 ) ,
V-
ti(M).
Define ip : [0,pi) ->• (~L,L) by rTTdu-
V(P) == /
where pi := smpxeM d(A, x). Applying (3.25) to the special function
( Atdt, Jo
we obtain the
(4.24)
where At := v{ip~l(t)) denotes the induced measure of ip~1(t). Note that
It follows from HM > h that This implies /•v,/v x / TT~\du
=
'Vo/
V(P) - P>
Vp
(4.25)
Let du. Assume that
(J,(A)
> e/j,(M), i.e.. Vo ^
> e.
(4.26)
We assert that n(UPo{A)) > (1 - e)fi(M), i.e., ^ > l -
£
.
(4.27)
Observable
Diameter
69
Note that if p0 > p\, then
>-£ = !>!-.. Now we assume that p0 < p\. In this case, by (4.25) and (4.26), we have /
Je
TT^du
TT^du
- /
M«)
Jv0/V
=
TT^du-
V(P°) ^P°=
h
W
h
Ku)
This implies (4.27). Therefore, we can conclude that fl-e
r ExDist(M;e) < p0 = Je
i
h{uY -—rdu. Q.E.D.
h[U
By (4.21) and (4.23), we see that for any isoperimetric function h, 1 fl~e 1 ExDist(M;e) < ——— / -r^-du. Ih(Af) Je h(u) In particular, for ha(s) = min(.s,1 — s ) 1 - ^ , 1 < a < oo, E
X
Dist(M;,)<^,
where
C7a = ( ^ [ ( 3 ) 1 / a - e V a ] i
-21n(2e)
ifl
if a = oo
The expansion distance of Finsler m spaces can be estimated from below if they satisfy certain curvature bound. See Theorem 16.4.1 below.
4.4
Observable Diameter
To study the Levy concentration of mm spaces, Gromov [Gr4] also introduce the notion of observable diameter. Let (M,d,p,) be an mm space. Denote by Lip x (M) the set of all Lipschitz functions / : M —» R satisfying -d(xi,x2)
< / ( x i ) - f(x2) < d{x2, xi),
Vxi,x2
€ M.
70
Isoperimetric
Inequalities
By (3.21), we know that for any compact subset A C M, the distance function p+{x) := d(A,x) belongs to Lip x (M), so does p~(x) := — d(x, A). For 0 < e < 1, define Diam(/»/i,e) := inf {Diam(J) : I c R, M / _ 1 0 O ) > (1 -
<0A»(AO}
Th° observable diameter is defined by ObsDiam(M;£) :=
sup
Diam(/»/i,e).
/€LiPl(M)
The expansion distance is equivalent to the observable diameter according to the following lemma. Lemma 4.4.1 Let X = (M,F,dp.) be a compact reversible Finsler m space with p,(M) < oo. Then for any 0 < e < 1/2, ExDist(M;£) < ObsDiam(M;£).
(4.28)
ObsDiam(M; e) < 2 • ExDist(M; e/2).
(4.29)
For any 0 < e < 1,
Proof. First we prove (4.28). Let p < ExDist(M;£). There exists a subset A c M such that H(A) > ep(M),
ii{up(A))
< (1 - £)MM).
Observable Diameter
71
Define J
, . _ f dist{A,x) W - \ p
x£Up(A) xeM-Up(A).
Let I = [a, b] be an arbitrary interval such that M
(/-1(/))>(1-£)M(M).
(4.30)
We assert that a < 0 and b> p. Hence Diam(7) = b — a > p. lfb
Kr\l))
< I*(UP(AJ) < (1 - e)MM),
that contradicts (4.30). If a > 0, then / ^ ( I ) C M - £/ a (A). Thus M / _ 1 ( / ) ) < M(M) - /*(tf 0 (A)) < p(M) - M(A) < (1 -
e)n{M),
that contradicts (4.30). Therefore Diam(Z) = b — a > p. This implies Diam(/*/i;e) = inf Diam(J) > p. Thus ObsDiam(M;e) > p. Letting p -s- ExDist(M;e) yields (4.28). We now prove (4.29). Let p < ObsDiam(M;£:). There is a Lipschitz function / G Lip^X) such that Diam(/,/Lx;£) > p. Let a be the supremum of r such that p{f-l{-^r))<£-p{M). Let b be the infimum of R such that Mr1(JR,oo))<|MM). Then for any r < a and R > b, M/
_ 1
M ) = MM) - n(f-l(-™,r))
~ M r ^ o o ) ) > (1 -
e)p(M).
72
Isoperimetric
Inequalities
Thus Diam([r, R]) > Diam(/*/^;£r) > p. This implies b — a > p. Let A = f-l(-oo,a)
and B = /- 1 (6,oo). Then M (A)
> |M(M),
p(B) >
£
-p{M).
Let
For any x G [7^/2(^4) rr
\
^ r,
an
d 2 £ A with d(,z, x) < p/2, ,/
\
x
P
f(x)
^
<
Cl + b
b —d a
+ - _ = _ _ .
Thus z G A For a; G Up/2{B) and z e B with d(.z,x) < p/2, f[x) > f(z)-d(x,z)
> &- £ > 6 - —
=
— .
By the above argument, we conclude that Up/2(A) c A, Since AnB
Up/2{A) c B.
= <8,
p(yp/2(A))
+PL(UP/2(B))
< v(A) + p(B) = p(AnB)
< p(M).
We may assume that p(up/2(A))
<\p{M)<{l-e)p{M).
This implies ExDist(M;e/2) > p/2. Letting p -> ObsDiam(M;e), we obtain (4.29).
Q.E.D.
Observable
Diameter
73
It follows from (4.28) and (4.29) that for any 0 < e < 1/2, ExDist(M; £ ) < 2 m ExDist(M; £ /2 m ). and for 0 < £ < 1, ObsDiam(M;e) < 2 m ObsDiam(M;£/2 m ).
Corollary 4.4.2 0 < £ < 1,
Let (M, F, dfi) be a closed Finsler m space. For and
ObsDiam(M;e) < 4 J - - 1 e
* Ai(M)'
Chapter 5
Geodesies and Connection
In a metric space, minimizing curves are of special interest to us. Locally minimizing curves in a Finsler space are determined by a system of second ordinary differential equations (geodesic equations). With the geodesic coefficients, we introduce the Chern connection and the notion of covariant derivative.
5.1
Geodesies
In this section, we shall derive the Euler-Lagrange equations for locally minimizing curves in a Finsler space (M,F). Let c : [a, b] 4 M be a constant speed piecewise C°° curve F(c) = A = constant. By definition, there is a partition of [a, 6],
a = t0 < • • • < tk = b,
such that c on each [£i_i,£i] is C°°. Fix the above partition and consider a piecewise C°° map H : (—e,e) x [a, b] —> M such that
(a) H is C° on (—£,e) x [a, 6]; (b) H is C°° on each (—£,£) x [U-i,ti], % = 1, • • •, k; (c) c{t) = H(0,t), a
Geodesies and
76
Connection
cu{b)
The vector field
n * ) = ^ ( t ) ^ |l e W( : = ^ ( 0 , * )
dx ' du is called the variationfieldof H. The length of cu(t) := H{u, t) is given by
L(u):= J F(cu(t))dt = J2J*
F(^{u,t))dt
Observe that
i=l
=
^;{[F2]x>-[F2)xiy>-[F2]y>yli*}vkdt
/
+ibF-93^vk i=i r6
(5.1) where g^y) := |[F 2 ] y V (y) and G'(y) := \gU(y){[F2}xW(y)yk
- [F2U(y)}.
(5.2)
Geodesies
77
Let
^••-Fikv^l+2Gi{c)}i-Mt)-
(5 3)
-
n(t) is called the geodesic curvature of c at c(t). We can express (5.1) in index-free form L'(0)
=
g c ( « , ^ ) ^ + A- 1 g ( ; ( 6 ) ( C (6),^(&))-g ( ; ( a )(c(a),l/(a))
-Xj
fc-i
+
A_1
E
{Sc(. r ) (c(*D- ^(*i)) " 8c(t+) ( < ^ + ) , n * i ) ) }, (5-4)
where A = F(c(t)) is a constant by assumption. Assume that c has minimal length. Then 1/(0) = 0 for any piecewise C°° variation H of c fixing endpoints. First, we take an arbitrary piecewise C°° variation H oic with -ff (M, ti) = c(ti) (hence V(U) = 0), i = 0, • • •, k. By (5.4), we obtain L'(0) = -\f
g d ( K , 7 ) d t = 0.
This implies /c(t) = 0 . Now for any 1 < i 0 < fc — 1 and v G T c ( t . j M , we take a piecewise C°° variation H of c, that fixes two endpoints of c with v
(ti„)=v,
H(u,ti)
= c(ti),
i^i0.
By (5.4), we obtain L'(0) = A - ^ g ^ + ^ c ^ t ) ^ ) - g ^ - ^ c f o - J . " ) } = 0 . By Lemma 1.2.4, we conclude that t{t:o) =
c(tf).
That is, c is C 1 at each tj. In local coordinates, K = 0 is equivalent to the following system c* + 2G*(c) = 0.
(5.5)
78
Geodesies and
Connection
(5.5) is a system of second order ordinary differential equations. Thus c must be C°° at each t\. We have proved the following Proposition 5.1.1 Let c be a constant speed piecewise C°° curve in a Finsler space (M, F). If c has minimal length, then c is a C°° curve with vanishing geodesic curvature n = 0. In virtue of Proposition 5.1.1, we make the following Definition 5.1.2 A C°° curve in a Finsler space (M,F) is called a geodesic if it has constant speed and its geodesic curvature K = 0. First Variation Formula: Let c : [a, b] —» M be a unit speed geodesic and a : (—e, e) —* M a C 1 curve with
(5-6)
Equation (5.6) is called the first variation formula.
The local functions Gl in (5.2) can be expressed by G\y) = \9U(y){2^(y)-^(y)}y^.
(5.7)
We call G% the geodesic coefficients. The geodesic coefficients G1 give rise to a globally defined vector field on TM \ {0}
G : = ^ - 2 ^ ) A
(5.8)
Geodesies
79
G is C°° on TM \ {0} and C 1 at zero tangent vectors in TM. We call G the spray induced by F. A curve c is a geodesic if and only if it is the projection of an integral curve of G. See [Sh9] for more detailed discussion on sprays. From the definition, the geodesic coefficients Gl (y) satisfy the following homogeneity condition G\\y)
= X^G'iy),
A > 0.
(5.9)
But, Gl{y) are not quadratic in y G TXM in general. Definition 5.1.3 A Finsler metric is called a Berwald metric if in standard local coordinate systems (xl,yl), the geodesic coefficients Gl{y) are quadratic in y e TXM for all x G M, that is, there are local functions Tljk(x) on M such that
Berwald metrics are special Finsler metrics. We are going to show that Riemannian metrics are special Berwald metrics. Example 5.1.1 Let F(y) = \fgij{x)yly^ manifold M. By (5.7), we have
be a Riemannian metric on a
where (gi:i(x)) := (gijix))"1. Clearly, Gl{y) are quadratic in y e TXM. Hence F is a Berwald metric. tf There are many non- Riemannian Berwald metrics [Sz]. Below is a simple example. Example 5.1.2 Consider a Randers metric F = a + 0 on a manifold M, where a(y) = ^Ja,ij{x)ylyi is a Riemannian metric and f3(y) — bi(x)yl is a 1-form with \\(3\\x =
sup px(y) = JaV^biWbjix) a«(l/) = l
V
< 1.
80
Geodesies and
Connection
According to Example 5.1.1, the geodesic coefficients of a can be expressed in the following form &{y) =
\rjk{x)yiyk,
where Tl-k{x) = TkAx) are local functions of x G M. Define b^j by bi{jdxj := dbi - bjTJikdxk.
(5.10)
Let r
ij : = 2 [bi\j + bJ\i) >
S
iJ := g \biti ~ bJ\ V •
The geodesic coefficients G l of F are given by G* = & + Py{ + Q\ where P(V)-Q\y)
=
^^{niVW
: =
-2a(y)brarvSplyl)
aairsriyl.
Assume that (3 is parallel with respect to a, i.e., biy = 0. Then T{j — U — Sij.
Thus P = 0 and Qi = 0. This implies that G\y) = G\y) are quadratic in y G TXM for all x G M. By definition, F = a + ft is a Berwald metric [Hale]. In this case, the geodesies of F coincide with that of a as point sets. By an elementary algebraic argument, one can show that if F is a Berwald metric, then b^ = 0. See [Ma5] [Hale] [Ki2] [SSAY] for details.
5.2
Chern Connection
In 1943, S.S. Chern studied the equivalence problem for Finsler spaces using the Cartan's exterior differentiation method [Chl][Ch2][BaCh]. He discovered a very simple connection. Later on, H. Rund independently
Chern
Connection
81
introduced this connection in a different setting. Thus, Chern's connection was also called the Rund connection in literatures. In this book, we will introduce Chern's connection using the exterior differential method. Let M be an n-manifold and 7r : TM —» M the natural projection. Denote by ir*TM the pull-back tangent bundle over the manifold TM\{0}. The vectors in ir*TM are denoted by (y,v), where y,v £ TXM. Take a local coordinate system (xl) in M. The local natural frame {^fr} for TM determines a local natural frame {di} for n*TM, d d
'i» : = = (»'^?i*)'
y£T*M-
This gives rise to a linear isomorphism between n*TM\y and TXM for every y € TXM. Given a Finsler metric F on M, F(y) — F(yt-^\x)
is a function of
(j/*) € R n at each point x € M. Let
S«(y):=^ 2 ]yV(!/)>
^(!/):=J[F\Vyi(!,).
(5.11)
We obtain two tensors g and C on n*TM defined respectively by g(u,v)
:= QijiyW'V^
C(U, V,W)
:= CiMirViW",
(5.12) (5.13)
where 17 = Uldi\y, V = V3dj\y and W7 = W/fc9fc|y. We call g the fundamental tensor and C the Cartan tensor of F . The Cartan torsion was actually appeared in P. Finsler's thesis [Fi]. But it was E. Cartan who first gave a geometric interpretation of this quantity [Ca]. Let Sy'-dy'
+ NiMdx*,
(5.14)
where N](y) := w(y). Then U*TM := span|dec*|,
V*TM := span j ^ j
(5.15)
Geodesies and
82
Connection
is a well-defined subbundle of T*(TM \ {0}). We obtain a decomposition forT*(TM\{0}), T*(TM \ {0}) = WTM © V*TM. Let {j~r, A } denote the local frame dual to {dxl,5yz},
that is
N
h^h- ^w
(516)
Then UTM := s p a n { ^ - } ,
VTM := s P an{ A }
are well-defined subbundles of T(TM \ {0}). We obtain a decomposition for T ( T M \ {0}), T(TM \ {0}) = WTM © VTM. Observe that ^2]=2cto(2/)yfc^eV*TM. Thus, d[F2] — 2FdF vanishes on horizontal tangent vectors. We obtain the following Lemma 5.2.1 F is horizontally constant, i.e., for any horizontal vector X = X i j | T eUTM X{F) = dF(X) = 0.
(5.17)
Let {e;}7=1 be an arbitrary local frame for ix*TM and {o/,wra+z}™=1 denote the corresponding local coframe for WTM © V*TM. The correspondence is determined by di o dxi «->
If ei = di, then a/ = dx and w
n+!
(5-18)
= J y \ Set
ylei = {y,y)
(5.19)
and 9ij •= 9(ei,ej),
Cijk:=C(ei,ej:ek).
(5.20)
Chern
Connection
83
We obtain a set of local function yl, gtj and djk on TM \ {0}. Theorem 5.2.2 (Chern) There is a unique set of local 1-forms {w/} on TM \ {0} such that djJ
= CJJ Aco/ k
(5.21) k
dgij
= gkjuji +gkjuji
n+i
= dtf + yiu/.
u
n+k
+ 2Cijkoj
,
(5.22)
(5.23)
Proof. Without loss of generality, we may take a natural local frame {e; = di}™=1 for TT*TM so that the corresponding local coframe for T*(TM\{0}) is {uil = dxl,u)n+l = 5yl}™=l. In this case, 9ij(y) = ^[F2}viyj{y),
Cijk(y) =
-[F2]yiyjyk(y).
Let Lijk
:= ^ g f V - 2CijklGl
- CmN\ - CukN] - CtjlNlk,
(5.24)
where
cijki{y) = -g^(y) = ^ W w O / ) We claim that the following set of 1-forms are the unique set of solutions to (5.21) and (5.22), cu/ := T)k dxk, where
^••=S^-9U^-
(5-25)
Observe that duj' - UJJ A to/ = -dxj A T)kdxk = - ^ { r j f c - rlkj\dxj
A dxk = 0.
Thus (5.21) holds. Rewrite (5.7) as follows
^-\{^-a-BW-
<«•»>
Geodesies and
84
Connection
Differentiating (5.26) with respect to yi yields 9klN
* -2\d^
+
^r--Mfy ~2CjklG
•
(5 27)
-
m
Differentiating (5.27) with respect to y , we obtain 1
rfc
[ d9ji dgmi dgjm -i dCjim k m dxi dxl J dxk V 2\dx —2CjkimG — 2CjkiNm — 2CkimNj — Ljim.
=
9kiLjm
(5.28)
From (5.28), we obtain dgji dxr
gkiT>;m + gkjT*rn + 2C3kiN*m.
(5.29)
This implies to (5.22). The proof of uniqueness is omitted. By the homogeneity of F, we have Cijkv1 = Cijkyj
= Cijkyk = 0
(5.30)
LijkV1 = LijkyJ
- Lijkyk
(5.31)
and - 0.
Thus yi k y
J - dyJdyk
y9L kl
i ~~dy3-i-
Thus 5yl = dyl + y]T)kdxk
= dyl +
This gives (5.23).
y'u/. Q.E.D.
From the above argument, we obtain a new tensor £ on w*TM defined by C(U, V, W) := LwMU'ViW",
(5.32)
where U = 11%^, V = V'dj\y and W = Wkdk\y- We call £ the Landsberg tensor. The Cartan tensor and Landsberg tensor play an important role in Finsler geometry.
Chern
Connection
85
Using the set of local 1-forms {w/} in Theorem 5.2.2, we define a map V : T{TM) x C°°(iv*TM) -> ix*TM by
V^U := [dU\X) + U'ujiX)}
® ei:
or simply
V^ — jd^ + C/^/J^ei. V is a linear connection on n*TM. We call it the Chern connection. For a vector field X = (X1, • • • ,Xn) on an open subset U C R71, the directional derivative D„X in a direction v £ TXR™ = R™ is defined by DVX := (dX'iv),
• • •, d X » ) = < / | | .
We can extend the notion of directional derivative to vector fields on a Finsler space. Let (M, F) be a Finsler space. At each point x £ M, define a map D : TXM x C^iTM)
-> T^M
by DyC/ := {dCT(y) + ^ ' ( z ) ^ ) } ^ ! *
(5-33)
where y 6 T X M and U G C°°{TM). Here A/)(y) are local functions on T M defined in (5.15). We call DyU(x) the covariant derivative of t/ at a; in the direction y. Set Dyll(x) = 0 for y = 0. D has the following properties: (a) D„(17 + V) =D y ?7 + DyVr; (b) Dv(fU) = dfx(y)U + f(x)DyU; (c) DXvU = AD„17, A > 0. The family D := {Dy}y£TM is called the connection of F. connection in a usual sense. If, in addition, D is linear, i.e., (d) Dy+vU = DyU + DVU,
D is not a
86
Geodesies and Connection
then D is called an affine connection on TM (or M). The following proposition is obvious. Proposition 5.2.3 The connection D of a Finsler metric F is affine if and only if F is a Berwald metric. Proof. In a standard local coordinate system (xl, yl) in TM, N;(y)
8G\ = —j(y), dyi
N^yW
=2G\y).
Thus Gl(y) are quadratic in y G TXM if and only if NUy) are linear in y G TXM. This proves the proposition. Q.E.D. For Berwald metrics, the connection D is an affine connection on the tangent bundle TM. We call it the Levi-Civita connection. E x a m p l e 5.2.1 Consider a Riemannian metric F(y) = \Jgij{x)ylyi a manifold M. The geodesic coefficients Gl(y) are given by
G\y) = 9Hx){^(*)-l^)}y>yk-
on
(5.34)
Thus Gl{y) are quadratic in y G TXM for all x G M and D is an affine connection. The Levi-Civita connection D satisfies the torsion-free condition BuV-DvU
= [U,V]
(5.35)
and d[g(U,V)](y)=g(pvU,v)+g(u,Dyv),
(5.36)
where y G TXM and U, V G C°°(TM). It can be easily proved that the Levi-Civita connection is the unique linear connection satisfying (5.35) and (5.36). ' jj Now we discuss t h e relationship between t h e Levi-Civita connection D on TM and the Chern connection V on ir*TM for Berwald metrics. For a Berwald metric, in a local coordinate system (xl) in M , the Christoffel symbols F)k(x) •= a ^ f \ (y) are functions of x only, so that NiAy)=T)k{x)yk.
Chern
Connection
87
The Levi-Civita connection is given by DU={dUi
WT)kdxk}»-^i,
+
where U = Ui-^l e C°°(TM). Take an arbitrary local frame {b;}™^ for TM and let {8l}f=l denote the local dual frame for T*M. Write DU = {dUi +
6ji}®bi,
where U = £/*b;. {#/} are called the Levi-Civita connection forms with respect to {bi}™=1. Let 7T : TM —»• M denote the natural projection. We can lift {bj}™=1 to a local frame {ej}™=1 for n*TM, where ei\v •= {y,hi\x),
y€TxM.
The Chern connection forms {u> •*} with respect to {ej}™=1 are given by
Let yl denote the local functions on TM determined by (y,y) = yzei = The local coframe {v*,un+i}?=1 o/=7T*6>\
(y,ylbi).
for T*(TM \ {0}) are given by ojn+i = dyi + yjTr*eji.
Consider a tensor S = S) b; ® 9j on M. Define Q = Qlei by Q\y):=yjSi(x), Then Q is a well-defined tensor on
y= TT*TM.
y%.
On M we define S*,fc by
dS} + Sk6ki-Sl6k=:S^k6k.
(5.37)
While on TM \ {0}, we define Q\k and Q*fc by dQ i + Qkujki =: Qffcwfc + Q > n + f c .
(5.38)
88
Geodesies and
Connection
Observe that
dQ' + Q V
=
y^^dS^
+ Sldy' +
y^^S^)
=
y V [dS) + Sty]
=
yjn* [dS] + Sk6k* - SiOj"] +
=
yin*(S)lkek)
+ Si [un+k - yiir'O/ Slkun+k
Siujn+k.
+
We obtain
5.3
Q\k
=
VJS)\k
(5.39)
Q\
=
Sl
(5.40)
Covariant Derivatives
Let (M,F) be a Finsler space and c(t), a < t < b, a geodesic. For a vector field V = Vl{t)-^\c{t) along c, define
D^(t):={^(*) + V^(c(*))}^| DiV(t) is called the covariant derivative of V{t) along c. A vector V(t) is said to be parallel along c, if D6V(t) = 0. Let U = ^ W ^ T U W and V = ^(*)gfrU(i) be parallel vector fields along c. They satisfy the following equations dJJm —
dVm = -WN™(c):
—
= -VlNr(c).
(5.41)
Consider the function gc(U,V)=9jlUjVl. By (5.41), we obtain Jt [mUjVl]
= { ^ c
m
- gjmNr
-
9jmNr
-
ACjlmGm}wVl.
Covariant
Derivatives
89
Contracting (5.29) with ym yields Vm^
= QmiNJ1 + gimN?
+ 4CjlmGm.
(5.42)
It follows from (5.42) that
|Mt/,v)
Ik""1
Thus Sc{U, V) = constant.
Let c(t), a < t < b, be a geodesic. Define a map Pc : TC^M
—> Tc(p)M
by Pc(v) := V(b),
v G T c ( o ) M,
where V(t) is the parallel vector field along c with F(a) = v. Pc is called the parallel translation along c. The above argument proves the following Lemma 5.3.1 For any geodesic c(t), a < t < b, in a Finsler space (M,F), the parallel translation Pc preserves the inner products gc along c, gc(6) (-Pc(w), Pc{v)j = g c(o ) (u, vj,
u, v £ T c(a) M.
In general, the parallel translation does not preserve the Minkowski norms. Y. Ichijo [ic] proved that on a Berwald space, the parallel translation along any geodesic preserves the Minkowski norms. Thus Berwald spaces can be viewed as Finsler spaces modeled on a single Minkowski space. Lemma 5.3.2 ([ic]) Let (M,F) be a Berwald space and c(t), a < t < b, a geodesic. Then the parallel translation Pc preserves the Minkowski norm,
Fc(b) (P c (v)) = Fc(a)(v),
v G Tc{a)M.
90
Geodesies and
Proof. Let V = Vt(t)-^\c(t)
Connection
be parallel along c. V satisfies dVl dt
VjN}(c).
(5.43)
Consider the function F(t) := F(V(t)). By assumption, N]{yy
= T)k{x)r?yk
= NJ(v)yi
By Lemma 5.2.1 and (5.43), we obtain b
[t)
~
dt
[C dx*
+
=
dF[^rjk(c)Vk
Thus F(t) = constant
dt dyt
- V^)k{c)ck)^-\
= 0. Q.E.D.
On a Berwald space, the canonical connection D is an affine connection. According to Szabo [Sz], if a Finsler metric F is Berwaldian, then there is a Riemannian metric g whose Levi-Civita connection coincides with the canonical connection. There are many non-Riemannian Berwald metrics. See Example 5.1.2 above. 5.4
Geodesic Flow
Let (M, F) be a Finsler space and
the spray of F. For a vector y £ TM, denote by 4>t{y) the integral curve of G with (fio(y) = V- From the definition, we see that the curve c(t) :— 4>t{y) is a geodesic with c(0) = y. We call <j>t the geodesic flow of F. It follows from Lemma 5.2.1 that d F(My))]=dF(GMy))=o. dt
Geodesic Flow
91
Thus F(
(5.44)
This means that the geodesic flow <j>t preserves the Finsler metric. Recall the Hilbert form on TM \ {0} that is given by (5.45) where gij(y) •= \[F2]yiyi{y). Lemma 5.4.1
The Hilbert form has the following property.
For any t,
| [ ( & ) ' W ] = ± d [(&)*[**]]•
(5-46)
Hence {(j>t)*duj — cLo.
Proof. Let (ipl(t),ipl(t)) be the standard local coordinates of y(t) := 4>t{y) in TM. y(t) is an integral curve of G, hence
d
-§{t) =
(5.47)
-2G\y{t)).
By (5.47), 1, dpi ^ (
^
+
Note that ldxiaX
+
V dvJ dy
J t=o
'
dtldxi**
4
dyi
V
\ t=o
V
We immediately obtain d di L
J u=o
±{[F2]x>yiyk
-2[F2]yiy>Gk(y)}dxi
+^[F2}rdyi
(5.48)
It follows from (5.2) that 2[F\iykGk
= [F2]xkyiyk
- \F2}
(5.49)
Geodesies and
92
Connection
Substituting (5.49) into (5.48) yields dt L" '
J t=o
Since 4>t+s -= (t> 4>t°
![<*H=4s(<«*")] = M<«''f2iHd r, ,
%i
, l
,r d t , , ,„
1
M
Thus {4>t)*du) = ((j>0)*du — LJ.
Q.E.D. T h e Finsler metric F induces a Riemannian metric of Sasaki t y p e on TM \ {0}, 9 •= 9ij{y)dxl
® dx3 + gij{y)Syl
where Sy1 = dyl + Nj(y)dyJ. T h e Riemannian metric g determines a nondegenerate 2n-form dVg on TM \ {0} by dVs = det (gij)dx1
• • • dxndy1
• • • dyn.
(5.50)
Observe t h a t du =
', ykdx
A dx1 — gijdx1 A dy3 = —g^dx1 A 6y3.
Thus, we can express dVg as follows n(n + l) 1
dVg = ( - 1 )
2
— dw A • • • A da>.
dVg is just the volume form we have mentioned in (2.16). By Lemma 5.4.1, we obtain the following P r o p o s i t i o n 5.4.2 ([Dal]) The geodesic flow cf>t preserves the volume on TM \ {0}.
Riemannian
Let i : SM -» TM \ {0} be the natural embedding. By (5.44), <j>t : SM -» S M . Let CJ := i*u). It follows from (5.44) and (5.46) t h a t
" W*\
dt
o.
Geodesic Flow
93
In other words, (fit preserves u>. Let g :— i*g denote the induced Riemannian metric on SM. Then the induced Riemannian volume form dVg on SM is given by n(n + l)
dVh = (-1)
2
,
1
-.
rrtj Add) A • • • Add}.
(n-1)! Proposition 5.4.3
([Dal]) The geodesic flow (fit preserves dVg on SM.
It is worth making further investigation on the geodesic flow on a Finsler space. The geodesic flow (fit on a reversible Finsler space (M, F) is of Anosov type if there exists a decomposition of T(SM) into
T(SM)=ES®EU®E°, where E° = R • G and Es ^ 0, Eu ^ 0 are invariant under the flow in the sense that
Further, there are constants A, B > 0 such that ||(&).(A-)|| < A e - s t | | X | | , ||(0_ t ).(X)|| < Ae- B t ||X||,
VX e S s , Vt > 0, VX G B", Vt > 0.
P. Foulon [Fol] proved that the geodesic flow on a closed reversible Finsler space of negative curvature must be of Anosov type. Recently, Foulon [Fo2] has done some work on contact Anosov flows on closed manifolds. In particular, he shows that for a smooth contact Anosov flow on a closed three manifold, the measure of maximal entropy is in the Lebesgue class if and only if the flow is up to finite covers conjugate to the geodesic flow of a metric of constant negative curvature on a closed surface. This shows that the ratio between the measure theoretic entropy and the topological entropy of a contact Anosov flow is strictly smaller than one on any closed three manifold which is not a Seifert bundle.
Chapter 6
Riemann Curvature
Riemann curvature is the central concept in Riemannian geometry. This quantity was first introduced by B. Riemann in 1854 for Riemannian metrics. Since then, Riemannian geometry has become one of the important branches in modern mathematics. In 1926, L. Berwald extended the Riemann curvature to Finsler metrics [Bwl] [Bw2]. In this chapter, we will introduce and discuss this very important quantity in Finsler geometry from various points of view.
6.1
Birth of the Riemann Curvature
The Riemann curvature of a Finsler space is a family of linear transformations on tangent spaces. It can be defined via the variations of geodesies. This approach is different from Riemann's and Berwald's approaches. Let (M,F) be a Finsler space. Consider a geodesic c(t), a < t < b. A C°° map H : (—s,e) x [a, b] —• M is called a geodesic variation of c if H{0,t)
=c{t)
and for each u € (—£,£), the curve cu(t)
:=H(u,t)
is a geodesic. We will show that for any geodesic variation, its variation field J(£) := ^ ( 0 , i ) satisfies a special system of second order ordinary differential equations. 95
96
Riemann
Curvature
Lemma 6.1.1 Let (M, F) be a Finsler space. There is a family of transformations ILy : TXM -» TXM, y £ TXM \ {0}, such that for any geodesic variation H of a geodesic c, the variation vector field J(t) := ^ - ( 0 , £) along c satisfies the following equation D a DeJ + R c ( J ) = 0 .
(6.1)
Proof. By assumption, each cu(t) = H(u,t) is a geodesic. Thus dt2
(6.2)
\-dTJ
For simplicity, let rp
rp%
d dx* '
dH dt '
"-"£"£•
Equation (6.2) becomes dT + 2Gl(T) = 0. dt
(6.3)
Note that dT du
d /dW du \ dt )
=U
dU{ dt\
du )
dt
Differentiating (6.3) with respect to u yields dt2
dxk
dt
dyi
Observe that d_ G\T) du . d dGl (T) dt L dyi
dxk(
!+
2{Zni
rph
dt
V QTk
dG (T) + dt dxkdy:>
rpk " Gl
d2Qi
dyidyk
•(T)
0 ^, f c
dxkdy^{T)~2Gk{T)dJd^{T)- dyidyk By the above identities, one obtains
Birth of the Riemann i ( dG dG* k I dx
-ukR\(T)~,
Curvature
i „,„• d22G G^ dx^dyk
97
„„„• d2Gi idyidyk
dGldG^ d • \ A dyi dyk 1dx*
d
where k[V}
:
"
dxk
V
d&dyk
+
G
dyidyk
dyi dyk'
{
'
For every vector y £ TXM \ {0}, define a linear transformation R^ = R\{y)-^-r®dxk\x
: TXM ->• T X M.
We obtain D T D T J7 + RT(i7) = 0 . Restricting the above equation to c, we obtain an equation for J(t) := U(0,t), DcDcJ + Rc(J) = 0 .
(6.5)
This completes the proof.
Q.E.D.
The geodesic variations give rise to a family of transformations R = {R„ : TXM -> TXM | yeTxM\{0},
x € M}.
We call R the Riemann curvature. It follows from (6.4) that Hy(y) = 0.
(6.6)
Moreover, the R^ is self-adjoint with respect to g v , (u),vj
= gy[u,B.y(v)j,
u,veTxM.
(6.7)
This fact will be proved in Section 8.1 below. By granting (6.7), we obtain gtf(R1/(u),y)=g„(«,Rv(y))=0.
(6.8)
98
Riemann
Curvature
Let P C TXM be a tangent plane. For a vector y G P\ {0}, define gv(Ry(u),u) Zy{y,y)%y{u,u)-zv(y,u)%y(y,uy where u G P such that P = sp&n{y,u}. By (6.6) and (6.8), we can easily show that K(P,y) is independent of u G P with P = span{?/,u}. The number K(P, y) is called the flag curvature of the flag (P, y) in TXM. When n = dim M = 2, K(y) := K(TXM, j/),
y£TxM\
{0},
is a scalar function T M \ {0}. We call K(y) the Gauss curvature. For a vector y £ TXM \ {0}, there are infinitely many tangent planes P C T X M containing y. The flag curvature K(p, y) depends on the tangent planes P containing y. A Finsler metric F on a manifold M is said to be of scalar curvature K(j/) if K(P, j/) = K(y) is independent of the tangent plane P containing y for all y G TXM. This is equivalent to the following Ry(u) = K{y) {g y (y, y) u - gy{y, u)y},
y,u e T . M \ {0}.
By definition, every two-dimensional Finsler metric is of scalar curvature. Define n
Ric(y):=^R\(y).
(6.10)
i=i
Ric is a scalar function on TM \ {0} with the following homogeneity: Ric(Ay) = A 2 Ric(y), We call Ric the Ricci curvature.
A > 0.
For the sake of convenience, let
We call R(y) the Ricci scalar. By the homogeneity of JR, we have
Geodesic Fields
6.2
99
Geodesic Fields
In this section, we will give an alternative formula for the Riemann curvature. This formula is not good for computing the Riemann curvature, but it is very useful in comparison geometry. Lemma 6.2.1 Let Y be a geodesic field on an open subset U in a Finsler space (M,F). Let g := gy denote the Riemannian metric on U. Then F := \fg is Y-related to F in the following sense N](Y) = NJ(Y).
(6.11)
In particular, Y is a geodesic field of F.
Proof. Equation (5.27) gives
Note that
It follows from (6.12) and (6.13) that
-2gil{Y)Cjkl{Y)Gk(Y) =
N)(Y).
This gives (6.11). By (6.11), we obtain 2Gl{Y) = Nlj(Y)Yj
= N}(Y)Yj
= 2Gi{Y).
Thus Y also satisfies • dYi Y3-z-^ + 2G,(Y) = 0. Thus Y is also a geodesic field of F.
Q.E.D.
Let Y be a geodesic field on an open subset U C M and g := gy denote the induced Riemannian metric on U. It follows from Lemma 6.2.1 that the
Riemann
100
Curvature
connection D of F and the Levi-Civita connection of F := \fg are related by DYU = t>YU.
(6.14)
Further, we have the following Proposition 6.2.2 Let Y be a geodesic field on an open subset U in a Finsler space (M,F) and g := gy. Let R and R denote the Riemann curvature of F and F = ^fg respectively. Then for y = Yx s TXM R^ = Ay.
(6.15)
' Observe that k [
dx
dYj
= U^Y
'
dYj
-"We* dxk{
(6-16)
'
and dNt
•(
=
d
[ -
i
-•
-i
Yi-£-[Nl(Y)]+2G\Y)rkl{Y)
= Y^(Y)+ril(Y)Yi— =
dYl
yi
+ 2Gl(Y)rkl(Y)
^ j ( y ) - 2G'(r)n,(^) + 2G'(y)r^(y).
This implies y i
"^
( y )
"
2G
'(y)r^(y) =
YJ
-d^W
~ tG\Y)rkl{Y).
(6.17)
Plugging (6.16) and (6.17) into (6.4) yields R\(Y)
= ^fc(F). Q.E.D.
Projectively Related Finsler
Metrics
101
For a Riemannian metric F(y) = \/g(y, y) on a manifold M, the Riemann curvature Rj, : TXM —> TXM is quadratic in y G TXM. Moreover, R y is self-adjoint with respect to g, g(Ky(u), v) = g(u, R » ) ,
u,veTxM.
(6.18)
The proof of (6.18) is given in standard text books on Riemannian geometry. We will prove it later in Section 8.1. Granting that (6.18) holds for Riemannian metrics, we consider a Finsler space (M,F). For a vector y G TXM \ {0}, extend y to a geodesic field Y in a neighborhood U of x. Let R ^ : TZM -> TZM, w e TZM \ {0}, denote the Riemann curvature of g := gy. By (6.18), g(TLY{u),
u) =g(u, Ry(w)),
u,vGTzM.
Restricting the above equation to x yields gy^Ryiu), vj =gy(u, R j » ) ,
u,v€TxM
(6.19)
Thus if (6.19) holds for Riemannian metrics, then it holds for Finsler metrics. 6.3
Projectively Related Finsler Metrics
Two Finsler metrics on a manifold are said to be pointwise projectively related if they have the same geodesies as point sets. The Riemann curvature is determined by geodesies. A natural question arises: What happens to the Riemann curvature if two Finsler metrics on a manifold have the same geodesies as point sets? In this section, we will discuss this problem. See [Dg][BuKe][Ma4] for related discussion. Given a Finsler space (M, F) and a Finsler metric F on M. We view F as a scalar function on TM. Let {ujl,u>n+l} be a local coframe for T*(TM\{0}) corresponding to a local frame {ei}™=1 for IT*TM. See (5.18) for the correspondence. Define dF = F^UJ1 + F.iWn+\
d.F[t - F b V = FW<J + FWjwn-
Riemann
102
Curvature
In a standard local coordinate system (xl,yl) 6y\ F\k and F\k.i are given by F\k — Fxk — NkFyk,
in TM, wz = dx% and u)n+l =
F\k.t = {F\k)yi.
l
Let G (y) and G(y) denote the geodesic coefficients of F and F in a common standard coordinate system (x1, yl) in TM. They are denned by (5.2). From the geodesic equations, one can easily see that F and F have the same geodesies as point sets if and only if Gi{y)=Gi{y)
+ P{y)y\
where P{y) is a scalar function satisfying P(Xy) = XP(y),
A > 0.
The following important result is due to A. Rapcsak [Ra]. Lemma 6.3.1 (Rapcsak) For two Finsler metrics F and F on a manifold M, the following conditions are equivalent (a) F and F have the same geodesies as point sets; (b) There is a scalar function P{y) on TM such that Gi(y)=Gi(y)+P{y)yi. In this case, P is given by
P(y) = %£. Kyi
2F
(c) F satisfies Flk.iyk - F„ = 0. Proof: By a direct computation, one can verify that & = Gi + Pyi + Qi, where 2F ' Q* =
\Fga{F\k.iVk-F\i}-
(6.20)
Projectively Related Finsler
Metrics
103
Thus F and F are pointwise projectively related if and only if Ql = 0, i.e., (6.20) holds. Q.E.D. Example 6.3.1 ([Hale]) Let (M,F) be a Finsler space. Consider a 1form P(y) = bi(x)y1 on M. Assume that ||/3|| := sup (3{y) < 1. F(y)=l
Then F := F + (3 is again a Finsler metric on M. Observe that
_
fc
%
k
, SA^' fc
Note that «V
y
dykdyl
J
dyl
l
'
Thus a
k
n
{dh
dbk\
k
By Lemma 5.2.1, F|fc = F x * - JV'F,,, = 0. We obtain
Thus F = F + /? is pointwise projective to F if and only if (3 is close.
jj
Given two Finsler metrics F and F on an n-dimensional manifold M. Assume that F and F are pointwise projectively related, hence by Lemma 6.3.1 &{y) = G\y) +
P{y)y\
where P(y) = ^f.
(6.21)
104
Riemann
Curvature
Let Rlk{y) and Rlk(y) denote the coefficients of the Riemann curvature of F and F respectively. By (6.4), we immediately obtain R\(y)
= R\(y)
+ z(v)S{ + rk{y)y\
(6.22)
where E = P2 - P\d
rk = 3(P,fe - PP.k) + S.k.
Equation (6.22) is proved in [MaWe]. In index-free form we can rewrite (6.22) as follows Ry(u) = Hy(u)+E{y)I
+ Ty(u)y,
u&TxM,
where / denotes the identity map and TV{U) := Tk(y)uh,
u =•
uk-^\x.
Consider a Finsler metric F on an open domain Q. C R n . According to Lemma 6.3.1, the following conditions are equivalent (a) The geodesies of F are straight lines in O; (b) The geodesic coefficients G1 of F is in the following form Gi(y)=P(y)yi. In this case, P is given by 2F • (c) F satisfies Fxkyiy
=
Fxi.
Assume that F is pointwise projectively flat. Then the Riemann curvature is in the following form Hy(u)=E(y)u
+ Ty(u)y,
ueTxM.
By (6.8), we obtain 0 = gy(RJ/(u),y) =
E(y)gy(y,u)+Tv(u)F\y).
This gives T
v(u) =
-j^sv{y,u).
Projectively
Related Finsler Metrics
105
Let
By the above formulas, we obtain R„(u) = K(j/){g„(y, j/) u - g„(j/,u) y } , Thus F is of scalar curvature.
y,u € TXM \ {0}.
Chapter 7
Non-Riemannian Curvatures
When one looks at a Finsler space, he not only sees the shape of the space, but also the "color" of the space. In this chapter, we will introduce and discuss the Cartan torsion and the Chern curvature, etc. Roughly speaking, the Cartan torsion describes the "color" of the space at a point and the Chern curvature tensor and its portion (the Landsberg curvature) describe the rate of changes of the color over the space. These quantities all vanish for Riemannian spaces. Thus Finsler spaces are much more "colorful" than Riemannian spaces.
7.1
Cartan Torsion
The Cartan torsion is a non-Euclidean quantity of Minkowski spaces. Since every tangent space on a Finsler space is a Minkowski space, the Cartan torsion is defined for a Finsler space. Let (V,F) be an n-dimensional Minkowski space. Recall that the induced inner product gy in V is independent of y G V \ {0} if and only if F is Euclidean. For a vector y € V \ {0}, define Cy(u,v,w)
: = -^|gy+ttu(u,*;)J| t = 0 ,
u,v,w € V.
We have
C (u t w) =
" ' ''
\~dLdt [p2{y+ru+sv+H
L=t=o-
Thus for each y € V \ {0}, Cy is a symmetric multi-linear form on V. The 107
Non-Riemannian
108
Curvatures
homogeneity of F implies Cy(y,v,w)
=0,
v,weV.
The family C = {CJ/}j/6v\{o} is called the Cartan torsion. Let {bj™ =1 be a basis for V. F(y) = F^h,)
is a function of (yl) £ R™.
Let Cijk(y) = - [ F 2 ] y V y * ( y ) .
9ij(v) = ^\F\^{y), We have gy(u,v) = gij(y)ulvJ,
Cy(u,v. w) := Ci:jk{y)utv:,wk,
(7.1)
where u = Mlbj, v = v-'b,- and w = uikh'KThe Cartan torsion is related to the Cartan tensor in (5.13) by Cy(u,v,w)
=C(U.,V,W),
u,v,wGTxM,
where U = (y, u), V = (y, v), W = (y, w) e ir*TM. Define n
M«) : = £
n
^ y (w) c *( b *» b J. u ) = Y, 9lj(y)Cljk(y)uk,
ij=l
u = u'bi,
r/'=l
(7.2) where (glJ(y)) •= (gij(y)) 1. The family I = {ly}yev\{o} is called the mean Cartan torsion. Note that in dimension two, the mean Cartan torsion completely determines the Cartan torsion. Iy can be also expressed as • d
I
M u ) = u%-Q-i l n Vdetfefc(2/)),
" = w'bj.
(7.3)
Thus I = 0 if and only if det(gij(y)) = constant. Using the maximal principles of the Laplacian on S = F _ 1 ( l ) , Deicke [De] proved the following important result Theorem 7.1.1 if 1 = 0.
(Deike) A Minkowski norm F is Euclidean if and only
Cartan
Torsion
109
See also [Bk] and Chapter 14 in [BCSl] for a proof. We should point out that Theorem 7.1.1 does not hold for singular Minkowski spaces [jiSh]. There are infinitely many singular Minkowski norms with vanishing mean Cartan torsion. Define
sup
^
sup
|C„(tj,i;,i;)| rr^.
(7.4)
L
(7.5)
Proposition 7.1.2 Let F = a + j3 be a Randers norm on V, where a is an Euclidean norm and p is a linear functional with \\0\\:=
sup
|/3(J/)|<1.
a(y) = l
The Cartan torsion is uniformly bounded. More precisely,
< ^^/l-yr^^p
IIIII
yen < ^ y T v r = W -
(7.6)
(7-7)
Proof. Fix a basis {bi}™=1 for V. Let P{y)=biV\
ly(u) = Ii{y)u\
y = y'bi, u = ifbi.
By (1.7) and (7.3), we obtain Ti
(y)
=
5TT ln
\fdet(9jk(y))
n + 1 (Fyi 2 I F 2F By (7.4), the norm of Iy is given by
ayi i a )
Non-Riemannian
110
Curvatures
It follows from (1.8), (7.8) and (7.9) that
M-^fMifl'-®!-
<™»
Since (3{y) < \\0\\a(y), we can write (3(y) = ||/3|| cos# for some 0 < 0 < 2n. Assume that F(y) = 1. Then a(i/) = l - / % ) = l-||/?||a(j/)cos0. This gives w
l + 6cos0
Plugging it into (7.10) yields l|I 11
n + 1 / ||/?Psin 2 fl
<
n + l
r
7 = = f
» - - 2 - V l + ll^llcosfl - ^ T V ! " VI " 11/311 •
This gives (7.6). To prove (7.7), we first reduce the problem to the two-dimensional case. Let y0,v0 with F{y0) = 1 and gya(v0,v0) = 1 such that l|C||
=Cyo(v0,v0,v0).
Let V = span< y0, v0 \ and F := F\v. We have 1 d3 r 1 C yo (v 0 ,w 0 ,v 0 ) = -—-^F 2 (y 0 -|-st; 0 )j| s = o = Cj,0(t;0, u 0 ,v 0 ). Let a := a|y and (3 = (i\V. We have 11,311= sup P(y)< a(y) = l
sup /3(2/) = p | | . a(y) = l
Let I(y 0 ) denote the main scalar of F at y0. It is easy to see that ||C|| = max |I|. Taking a special orthonormal basis {bi,b2} for (V,a) such that J3(y) = ||/3||u for y = ubi + vb^. With respect to this basis, F(y) = \lv? +v2 + \\j3\\ u,
y = ubx
+vb2.
Cartan
Torsion
111
It follows from (7.6) that
m^V1-^
(7.11)
Thus |C|| < ||C|| = max |I| < -y=\H
v/i~rPip<4v/w^ N/2
Q.E.D. In dimension two, the bound (7.7) is just the bound (7.6). The bound for the Cartan torsion in dimension two is suggested by B. Lackey. See Exercise 11.2.6 in [BCSl]. The norm of the (mean) Cartan torsion can be as large as one wishes. For example, the norm of the Cartan torsion of the following Minkowski norms F\ on R n approaches oo as A —» oo.
Fx(y) := \ E x a m p l e 7.1.1
£(^)2 + A E(^ 1 ) 4 i=l i=l
Let F(u,v) := lu4 + 3cu2v2 + v4\ ' ,
where 0 < c < 2. F is a Minkowski norm on R 2 (Example 1.2.3). For any y — (u,v) with u > 0, Iv
2v/3(9c 2 -4)(i> 4 -w 4 ) 3 (2c u 4 + (4 - 3c2) u2v2 + 2c vA
Take y = (w, v) with v = J-^u
3/2 '
> 0 , we obtain
,T , (9c2 - 4) 2 Iv = 1 yl 64c
>• oo.
. a s c -> 0"1".
Non-Riemannian
112
7.2
Curvatures
Chern Curvature
Besides the Cartan curvature, there are other quantities which always vanish on Riemannian spaces. In this section, we are going to discuss two non-Riemannian quantities — the Chern curvature and the Landsberg curvature. Let (M, F) be a Finsler space. In a standard local coordinate system (xl,yz) in TM, the local functions gij(y) = \[F2}yiyj(y) and Cijk(y) := \[F2]yiyjyk(y) on TM define the inner product g y and the Cartan torsion Cy on TXM respectively. See (7.1). Let L^k be the set of local functions defined in (5.24). Lijk{v)
= V1-^-
- 2CijklGl
where Cijki := j[F2]yiyjykyi(y). connection are given by jk
where Lzkl quantity:
- CljkN! - CukN] - CijhNl
(7.12)
The Christoffel symbols of the Chern
dyWyk
'
jk
'
:= g%:>Ljki- Differentiating Tl-k with respect to yl yields a new
P
.
(v)
.
jki\yi-
dT
U
&
dUjk
dyidykdyl
Qyl
dyl '
y
'
For a vector y e TXM \ {0}, define Ly(u,v,w)
: =
Lijk(y)uivJwkJ
Py(u,v,w):
=
Pjikl{y)u^vkwl
(7.14) —\x,
(7.15)
where u = u%-^\x,v = v1-£r\x,w = wk-^\x € TXM. L := {Ly}yeTM\{0} is called the Landsberg curvature [Lal][La2] and P := {Py}yeTM\{o} 1S called the Chern curvature [Chl][Ch2]. The Landsberg curvature is related to the Landsberg tensor in (5.32) by Ly(u,v,w)
= C(U,V,W),
u,v,weTxM,
where U = (y, u), V = (y, v), W = (y, w) e w*TM.
Chern Curvature
113
Let c(t) b e a n arbitrary geodesic in (M,F). Take arbitrary parallel vector fields U(t), V(t),W(t) along c. According to (7.12), Lc(t) (U(t),
V(t), W(tj)
= ~ [Ci(t) (U(t), V(t), W(t))].
(7.16)
T h u s the Landsberg curvature measures the rate of changes of the C a r t a n torsion along geodesies. A Finsler metric is called a Landsberg metric if L = 0. By (7.16), we see t h a t on a Landsberg space, the C a r t a n torsion is constant along geodesies. P r o p o s i t i o n 7.2.1 vature by
The Landsberg curvature
Ly{u,v,w)
is related to the Chern
=
-gy(u,
Py(y,v,w)j
=
g y ( p i / ( w , v , w ) , yj.
cur-
(7.17) (7.18)
Proof. By t h e homogeneity of G \ we obtain „.spi
— ,,«
y ^S u - y
sk _
i
d
-
u
dyi
I
Sri
\y
L
\
ji
sk) -
h
—_ji
ki-
^ u-
Thus 9ijysPs\i
= -9ijL\i
=
-LiM.
This gives (7.17). Differentiating (5.28) with y1, t h e n contacting the resulting identity with l y , we obtain y 9kl
~dy1~
=
y
IhF
+2CiikNl -2Lijm
+
AGi kmG
^
+ 2CikmN?
+ 2CikmNk
dLf1 y —^ dyl
l oLjim y'
^•^ijm
o - i( \lf L'jlmj
Ijijml
Thus ylgklPjkml
-
= ~
l y
g
k l i^
dy
= L ijm-
-^i,
Non-Riemannian
114
Curvatures
This gives (7.18).
Q.E.D.
Berwald metrics can be characterized by the Chern curvature. Proposition 7.2.2 j / P = 0.
A Finsler metric F is a Berwald metric if and only
Proof. Assume that F is Berwaldian. By Definition 5.1.3, Gz(y) are quadratic in y e TXM. Then dNk3 QymQyi
d3G' - 0. dyWykdyl
Differentiating (5.27) with respect to ym and yl, then contracting the resulting equation with yl, we obtain 0 = - 2 i / ' -l ^ + 4CijkmGk dx
+ 2CjkmNk
+ 2CijkN^
+
2CikmNk.
By (7.12), we see that Lijm = 0. Thus jk
dyidyk
and p *
Ih. = —a l 1 dy dyWkdy Thus Gz(y) are quadratic in y e TXM and F is a Berwald metric. Assume that P = 0. Thus L = 0 by (7.17). Equation (7.13) implies 3 kl
dT
)k
3 kl
dyl
Thus T%jk(y) are independent of y £ TXM and G\y) = \N]{y)yi
=
\r)k(x)y3yk
are quadratic in y G TXM. We conclude that F is a Berwald metric Q.E.D. From (7.17) and Proposition 7.2.2, we immediately conclude that every Berwald space is a Landsberg space. It is an open problem in Finsler geometry whether or not there is a Landsberg metric which is not a Berwald metric. So far no example has been found.
Chern
Curvature
115
Example 7.2.1 Consider the Funk metric F on a strongly convex domain S I c R " . F is a nonnegative function on TO. = Q x R™. According to (1.21), F satisfies Fxi=FFyi.
(7.19)
By (7.19), one can easily show that p
..k
p
and the geodesic coefficients G* of F are given by G\y) = \F(y)y\
(7.20)
Hence
^ =^ y
+^--
(7.2i)
We are going to compute the Landsberg curvature using (7.12). By (7.19), we obtain
Using (7.20) and (7.21), we obtain yl _
1
pi
dC
ijk „
CijkiGl = -Frf-g± 2 V dyl
=
_1
2
--FCijk
and 1„ ,„ 1_,„ 1 CijkNi = -Fyiy Cijk + -FdiCijk -FCiik. 2 Plugging them into (7.12) yields ^ijk
—
n
^ijk-
In index-free form, L„(u, v, w) =
F(v) ^Cyiu,
v, w).
(7.22)
Let c(t) be a unit speed geodesic in fi with c(0) = y € TXQ, and C/(t) a parallel vector field along c. From (7.22), we see that the function C(t):=Ct(t){U{t),U(t),U{t))
Non-Riemannian
116
Curvatures
satisfies C'(t) + \c{t) = 0.
(7.23)
The general solution of (7.23) is C(i)=C(0)exp(-i*). The maximal interval of c is (—5, oo), where
F(y) In [1 + F(y)
n-y)
Thus C(£) is a bounded function on (—6,00). It is not clear that if C is bounded on Q, when O is not the unit ball in R n . jj Let {t>i}™=1 b e a n a r b i t r a r y basis for TXM. W e define t h e m e a n of hy by n u
3y( )
•=
^ ^ ij = l
J
(
y
)
L
« (
u
'
b i
'
b
j ) '
where gtj(y) = g 3/ (b i , by). The family J = {Jy}y€TM\{o} is called the mean Landsberg curvature. A Finsler metric is called a weak Landsberg metric if J = 0. In dimension two, J completely determines L. Let (M, F) be a Finsler space and c an arbitrary geodesic. Take an arbitrary parallel vector field V(t) along c. It follows from (7.16) that Ji{t)(v(t))=jt[l6{t)(y(t))].
(7.24)
Thus on a weak Landsberg space, the mean Cartan torsion I is constant along any geodesic. There is an induced Riemannian metric g of Sasaki type on TM \ {0}, 9 = gij(y)dxl ® dx3 + gtjSy1 ® 5yJ. T. Aikou proved that if L = 0, then all the slit tangent spaces TXM \ {0} are totally geodesic in TM \ {0} [Ai]. We can show that if J = 0, then all the slit tangent spaces TXM \ {0} are minimal in TM \ {0}.
S-Curvature
7.3
117
S-Curvature
In this section, we will introduce and discuss an important non-Riemannian curvature for Finsler m spaces. It measures the rate of changes of Minkowski tangent spaces over a Finsler m space. Let (M, F, dfj,) be a Finsler m space. Take an arbitrary basis {bj}™=1 for TXM and its dual basis {^}^ = 1 for T*M. Express dy, = a{x) 61 A • • • A 0n, For a vector y £ TXM \ {0}, we define
Jdet(gij(y)) r ( y ) : = In-¥
-,
(7.25)
a
where gij{y) '.= g y (b{,bj). T is called the distortion of (M,F,djj,). The distortion r has the following homogeneity property r(Xy) = r(y),
X > 0, y G TXM \ {0}.
The vertical derivative of T is nothing but the mean Cartan torsion, namely, ft[r(y
+ tv)]\t=0
= Iy(v),
v£TxM.
(7.26)
First, let us give a proof for (7.26). In local coordinates T(y) = ln h/^Mi L
a
(7.27) J
where gij(y) = gtXgfrU, afrU) and d\i = a(x)dx1 • • • dxn. Observe that ^
[in y/det(gij(y))\
=
l
-g^{y)^{y)
= g^(y)Cijk(y)
=
Ik(y).
This gives (7.26). According to Deike's theorem (Theorem 7.1.1), C = 0 if and only if 1 = 0. Therefore, the following conditions are equivalent (a) (b) (c) (d)
r(y) = constant; I = 0; C = 0; F is Euclidean.
Non-Riemannian
118
Curvatures
To measure the rate of changes of the distortion along geodesies, we define S(y) :=
It
-(^)L'
(7.28)
where c(t) is the geodesic with c(0) = y. S is called the S-curvature [Sh9]. It is also called the mean covariation in [Sh2] and mean tangent curvature in [Sh4]. Let Y be a non-zero geodesic field on an open subset U C M. By (7.28), we have S(Y) =
(7.29)
Y[r(Y)]
The S-curvature S satisfies the following homogeneity condition S(Ay) = XS(y),
A > 0.
In general, S(y) is not linear in y. Differentiating S(y) twice with respect to y gives new quantity. For a vector y e TxM\{0}, define Ej, : TxMxTxM -> Rby Ey(u,v) :--
1 d2 2dsdt
S(y + su + tv)
u, v £ TXM.
(7.30)
We call E = {Ey}veTM\{o} the E-curvature. We will show that the Ecurvature is independent of the volume form, although the S-curvature does. We first derive local formulas for the S-curvature and E-curvature. In local coordinates, r(y) is given by (7.27). Observe that in a standard local coordinate system (xl,yl) in TM, _d_ In d dxk . d llnJdetig^) dykr
*M = 2°"lB 9ijQ
t j fe-
(7.31) (7.32)
lt follows from (5.27) that 29
WV
(7.33)
S- Curvature
119
Let c(t) be the geodesic with c(0) = y. By (7.31)-(7.33), we obtain
S(c): = =
=
|[r(c(t)) |(lnydet(3ij(c)))-|(lna(c))
^
t
?
Nm(c)-
—
^
- 2 S «(c)C« fc (c)G*(c) -
^ ^ r ( c )
— (c)
This gives
s(y) = ^ ( « ) - ^ ^ r W -
a-34)
Thus
We obtain a local formula for the E-curvature 1
E*(«.«)
1
f)3(~im
= 2SwV(„)t,V = a ^ ^ ^ d / ) ^ ,
(7.35)
where u = u ^ U ^ = ^ g ^ U G TXM. From (7.35), we see that the E-curvature is independent of the volume form. It is purely a geometric quantity of Finsler metrics. By (7.30), we see that S(y) is linear in y G TXM if and only if E = 0 on TXM \ {0}. In particular, if F is a Berwald metric, then S(y) is linear in y G TXM for all x [Sh2]. In fact, S = 0 for Berwald metrics if we consider the S-curvature of the Busemann-Hausdorff measure Vo\p. More precisely, we have Proposition 7.3.1 For any Berwald space (M,F), S = 0 with respect to the Busemann-Hausdorff volume form dVp. Proof. Let c : (—e,e) —> M be an arbitrary geodesic. Take an arbitrary parallel frame {b,(t)}™=1 along c. Let {#2(i)}™=1 denote the dual coframe
Non-Riemannian
120
Curvatures
along c. By (2.7), the Busemann-Hausdorff volume form dVp of F is given by dVF\c(t)=aF(t)el{t)A---A9n{t), where aF{t) : Vol{(y')eR" | F(i/
s(c(i)) =
it r (
oW)
)
By Lemmas 5.3.1 and 5.3.2, Sc(t){bi(t),bj(t)j
= g £ ( 0 ) (b i (0),b J -(0)J = constant,
F ^ b ; ^ ) ) = F ^ l b i ( 0 ) ) = constant. Thus det [gc(t)(bi(t),bj(i)jj = det [g y (b*, b., J1, Vol{(^) e Rn | F ( V b ^ ) ) < l } = V o l j ^ ) G R n , | i ^ j / ' b i ) < l } . The last equality implies that <XF(£) = constant. Thus S(c(£)) = 0. Since c is arbitrary, we conclude that S = 0. Q.E.D. Now we use (7.34) to prove the following Example 7.3.1 Let F = a + (3 be a Randers metric on a manifold M, where a(y) = y/aij{x)yiyi a n d / % ) = fei(a;)yi with ||/?j| := sup Q ( j / ) = 1 /3(j/) < 1. WE will find a sufficient and necessary condition on a and /3 for S = 0. In particular, we will show that if /3 is a Killing form of constant length, then S = 0. In a standard local coordinate system {xl,yl) in TM, define b^j by
S- Curvature
121
where 6l := dx1 and 6^ := TJikdxk denote the Levi-Civita connection forms of a. Let r
H : = g V^
+ bj]i
Sij
)'
:=
2 V i U ~~
bjli
)'
%
The geodesic coefficients G of F are related to the coefficients Gl of a by G l = & + Py{ + Q\ where P
(V)-
W^-AnjV'v3-2a(y)brar"sply1} 2F{y) aairsriyl.
=
Ql(y) : = Observe that dQ —— = a
1
aijyjalrsriyl
+ aairsri
= a lSjiyjyl
+ aalrsri
= 0.
Thus Nl = Nl +
(n+±)P.
Put dVp = <j(x)dxl • • • dxn,
dVa =
We have
Nl
By (7.34), we obtain a formula for S,
S(y) = (n + l){P(y) - d[ln v T = W ] (l/)}, where P(V) := 2 ^ y ( r y y V " 2a( 2 /)6 r a'-% ; 2 /).
(7.36)
Non-Riemannian Curvatures
122
Now we are going to find a sufficient and necessary condition on /? under which S = 0. For the sake of simplicity, we choose an orthonormal basis for TXM such t h a t aij — Sij. Let bkb fell
lnyr^w
l-IW
Observe t h a t S = 0 if and only if - 2abpspiyl
rijyW
P)AlVl.
= 2(a +
This is equivalent t o the following equations
(7.37) (7.38)
bpSpi + Ai = 0.
We claim t h a t (7.37) implies (7.38). First, contracting (7.37) with bi and bj yields 2bibi\jbj
bibiijbj = - 2 1 | Thus
biAi = - /
l
!LL = o.
i - Ml2
Using this fact, we obtain kbj\i
=
"ZbiTij - bibi\j
=
2\\f3\\2Aj + {l-\\P\\2)Aj
= {l +
\\l3\\2)Aj.
Finally, we obtain bpSpi
=
-(bpbp\i - bpb^p)
= ^{-(l-il/3|| 2 )^-(l-P| 2 )^} = -^. This gives (7.38). We conclude t h a t S = 0 if and only if (7.37) holds. We can rewrite (7.37) as follows. (1 - ||/3|| 2 ) (biy
+ 6 j H ) + 2bkbklibj
+ 2bkbkijbi
= 0.
(7.39)
Note t h a t if 0 is a Killing form (r^- = 0) with constant length (bibiy = 0), then (7.39) holds. Hence S = 0 and E = 0. (t
S-Curvature
123
Below is a specific family of Randers metrics on S 3 with S = 0. Example 7.3.2 Let {C^C^C 3 } De the canonical left-invariant co-frame on the Lie group Sp(l) = S 3 satisfying dC1=2<2AC3,
dC2 = 2C 3 AC\
dC3 = 2C 1 AC 2 -
(7-40)
Consider F = a + (3, where + A 2 [ C W + A2[C3Q/)]2,
a(y) := \ M C W
P(y) ••= e C\v),
where K > 0, A > 0 and e > 0. As Assume that K
Then F is a special Randers metric. Let 61 := {6l ,92,03}
62 := AC2,
K(1,
03 := AC3.
is an orthonormal co-frame for TS 3 with respect to a. Put
where &i = - ,
6 2 = 0,
63=0.
K
It follows from (7.40) that d0i = P A /, where 0 l are the Levi-Civita connection forms of a, given by Of + 9/ = 0 and
*• = $«•.
«.'--£»•.
t-&-%)*•
By definition, 6j|j are defined by dbi-bjOi3'=:6i|j0*.
A direct computation gives j.
°2|3 -
£
-^2 ,
».
»3|2 ~
£
^2
and all other components bi\j — 0. Thus /? is a Killing form of constant length. By Example 7.3.1, we conclude that S = 0, hence E = 0. (t
Non-Riemannian
124
Curvatures
E x a m p l e 7.3.3 Let F denote the Funk metric on a strongly convex domain CI in R™. The geodesic coefficients are given by Gi =
2
i -Fy y .
Observe that Nl = l{Fy%
= \{Fy^
+ FS\) =
^ - F
According to Example 2.2.4, the Busemann-Hausdorff volume form dVp = a(x)dx1 • • • dxn has constant coefficient, CTF{X) =
constant.
By (7.34), we obtain S(y) = ^F(y). The angular form hy(u,v)
(7.41)
:= hij{y)ulv:' on TXM is given by ys
yl
9^Y{y)gitF~(y)'
hij := F Fyiyi = gij -
h = {hy} is called the angular m.etric. By (7.35), we obtain 1„ n+1 Eij — o ^ i / V ~
',
-^j/V •
In index-free form, Ey(u,v)
= ^±^hy(u,v).
(7.42)
Chapter 8
Structure Equations
In previous chapters, we introduced the Riemann curvature and many nonRiemannian quantities. We discovered some relationship among these quantities. In this chapter, we are going to use the exterior differential methods to find more relationships among these quantities. At the end, we compute the Riemann curvature of a special class of Randers metrics.
8.1
Structure Equations of Finsler Spaces
In this section, we first introduce the curvature forms by differentiating the Chern connection forms. Then we derive some identities for the curvature coefficients by differentiating the fundamental tensor. Let (M, F) be a Finsler space. In Section 5.2, we introduced the fundamental tensor g, the Cartan tensor C and the Landsberg tensor £ on TT*TM. See (5.12), (5.13) and (5.32) for definitions. Let {ei}™=1 be an arbitrary local frame for TT*TM and {u/,w n+l }™ =1 the corresponding local coframe for T*(TM \ {0}). See (5.18) for the correspondence. Let (y,y) = yleiand
yl, gij, Cijk and L^k are local functions on TM. By Theorem 5.2.2, there is a set of 1-forms {wJJ}™J=1 on TM \ {0}. The Chern connection V on 125
Structure
126 TT*TM
Equations
is expressed by
VU = {dET + C^'u/Jgiej, where U = Ulei G C°°(Tr*TM). T h e Chern connection is uniquely determined by (5.21) a n d (5.22). For convenience, we rewrite (5.21) and (5.22) again.
=
UJJ ALO/
=
k
(8.1) k
n+k
gkjw +gikuj +2Cijkio
.
(8.2)
Let Q/ := doj/-
ujjk A w / .
(8.3)
l
{ilj }ij=1 is a set of local 2-forms on TM \ {0}. T h u s we can express fi •* in t h e following form ft/
= ±R/klu;k
Acol + P/kluk
A ujn+l + \ Q
3
\ ^
k
A «/*+',
(8.4)
where R/ki
+ R/ik = 0,
Q/ki
+ Qjlik = o.
(8.5)
Differentiating (8.1), we obtain
wj' A nj = o. This gives R3\i
+ -Rfc'y + R-i%jk = 0^i'fci
=
Pkju
(8-6) (8-7)
Equation (8.4) simplifies to ft/ - ^ i ? / ^ f c A w ' + P / ^ A o;"+'.
(8.8)
Structure
Equations of Finsler
Spaces
127
Define R(U,V)W:
=
V(U,V)W:
=
R/kl(y)UkVlWjei: Pjikl(y)UkVlWi
<x,
where U = Ukek,V = VleuW = Wjej 6 ir*TM. We call K and V the Riemannian curvature tensor and the Chern curvature tensor respectively.
Let Qi:=dujn+i-u>n+j
ALJJ.
(8.9)
According to Theorem 5.2.2, wn+l are given by un+i=dyi
+ yiwJi.
(8.10)
Differentiating (8.10) yields
By (8.8), f2l can be expressed by Qi = ±I?klu>kAujl
+
Piklu>kA<jn+l,
where R^r^y'R/u,
P'u-ytPSu.
Put R\
:= ffwl/' = ^ i J / f c , j / ' .
(8.11)
We claim that Rlk are just the coefficients of the Riemann curvature in (6.4) and Pjlkl are just the coefficients of the Chern curvature in (7.13). Hence Plkl = —g^Ljki are just the coefficients of the Landsberg curvature in (7.12). Let (xl,yl) be a standard local coordinate system in TM. Take w* = dx\
Lun+i = Sy\
Structure
128
Equations
T h e n by (5.25), d2G
Plugging t h e m into (8.3), we obtain
j k l
dxk
dxl
dym
l
dym
k
+T™lrmk-T™kYiml P<
=
* * dyl
j kl
(8.12)
* & dyidykdyl
_
|
g L
*
dyl
'
{
(813) '
Equation (8.13) is just (7.13). By the homogeneity of F, we obtain
k
P\i
~~
dxk
=
-L*kl.
V
dxidyk
+
dyidyk
dyi dyk'
(
' (8.15)
Equation (8.11) is just (6.4). T h u s Rlk coefficients of the Riemann curvature.
defined in this section are t h e
Now we go back to the general setting. Define Cijk\i and Cijki by dCijk — Cijku:i
— Cuk^j
— Cijiuik
=: Cijk\i(*> + CijkioJ71 •
In virtue of (5.24), we have Lijk = Cijk\iyl.
(8.16)
Differentiating (8.2) and using (8.9), we obtain ft* * V + 9ipttjP + 2ClJpnp
+ 2 ( C m k i 0 k + Ciji.kun+k)
Aoj n + l = 0. (8.17)
From (8.17), we obtain 9pjRi ki + 9ipRj ki + ZCijpR 9PJPi \i + 9iPP? ki + 2CijpP Ciji.k = Cijk-i-
p kl
= 0,
* + 2Cimk
(8.18) = 0,
(8.19) (8.20)
Structure Equations of Finsler Spaces
We first derive some important identities for Rf Rj k '•— Rj
129
ki-
Let
kiV •
Contracting (8.18) with yi yields VJ9vjRiVkl
= -9iPRPkl-
(8.21)
By (8.21), we obtain yj9jpRpki
-yl9iPRPki-
=
This gives ylgipRpkl
= 0.
(8.22)
Contracting (8.18) with yl yields 9viRi\
+ 9iPRjPk
+ 1CijpR\
= 0.
(8.23)
Contracting (8.21) with yl yields gipR\
= -ytgpjR^.
(8.24)
It follows from (8.5) and (8.6) t h a t
Rpk = RiW = -RkpJ
-Ripikyl = Rk\
-*V
It follows from (8.22) and (8.24) t h a t 9iPRPk
-y39pjRPk
=
yj9Pi(Rpik-Rkp)
= =
-yJ9pjRki
=
9kPRPi.
We obtain 9iPRPk
= gkpR'i.
(8.25)
Now we derive some important identities for -PA;three times t o the combination (9PJPAI
+ 9rPPPkl)
- {9PkPPu
+ gPJPkPu)
+ {9piPkji
Applying (8.19)
+
gpkP^i),
130
Structure
Equations
we obtain 9ipPj kl
~
~~^ijpP kl ~ (-'jkpP it " (->ikpi ji -Ciji\k
- Cjkl\i - Ciki\j-
(8.26)
By (8.26), we can easily obtain the following Liu = -gipPPki,
v3-27)
Uk^yig^P^
(8.28)
and VJ9jpPPkl
8.2
= Likiy1 = 0.
(8.29)
S t r u c t u r e E q u a t i o n s of R i e m a n n i a n M e t r i c s
In this section, we will deal with Riemannian metrics in a traditional way. All t h e quantities are defined on the base manifold. At t h e end, we view Riemannian metrics as special Finsler metrics and deal with t h e m as in t h e previous section. We will give a link between these two different approaches. Let (M, g) be an n-dimensional Riemannian space. In a local coordinate system {xl) in M, g — gij(x)dxl ® dx^ and the Christoffel symbols
?fcW •
dyidy*
(V)
29
[
}
\ dx*
{
'
+
dxk
[
are functions of x only, so t h a t N]{y)=T)k{x)y\
V=
d_ ykk7-k\ dx
T h e Levi-Civita connection is given by
d Dy[/ = {d[/i(2/) + ^r} f c y f c }^!
'
dx>
[
}
)
Structure Equations of Riemannian
Metrics
where U = U{-^ £ C°°{TM) and y = yk-£^\xtion is uniquely determined by DVV-DVU
The Levi-Civita connec-
= [U,V},
W[g(U, V)} = g(DwU, V) + g(U, DWV). where U,V,W
131
(8.30) (8.31)
eC°°(TM).
Let {bj}" = 1 be a local frame for TM and {#l}f=1 the dual coframe for T*M. Express the Levi-Civita connection D by Dbj = 0 / ® b i . The set of 2-forms {#/} are called the Levi-Civita connection forms. Let 9ij '•= ff(bi,bj). Equations (8.30) and (8.31) are equivalent to d6i = 6j A 6/, dgij=gkAk
+ gikOjk.
(8.32) (8.33)
Set 6 / = dfl/ - 6k A ^ = i i ? / w 0 f c A 0',
(8.34)
Rhi + */«* = °-
(8-35)
where
Differentiating (8.32) and (8.33) yields 0j A 0 / = 0, and 9kjQik + gikQk
= 0.
These two identities imply Rjiki + Rkiij + Riijk = 0,
(8-36)
9kjR{ kl + 9ikRj kl — °-
(8.37)
Structure
132
Equations
Define R(u,v)w
:= R/klukvlwj
bi;
where u = ukhk,v = vlhi and w = w^bj. We call R(u,v)w curvature tensor. From the definition, we have R(U, V)W = DuDyW where U,V,W
- DvDuW
-
the Riemann
DlUiV]W,
GC°°.
Equations (8.35), (8.36) and (8.37) can be written R(u,v)w + R{v,u)w = 0,
(8.38)
R(w, v)w + R(v, w)u + R(w, u)v = 0,
(8.39)
g(R(u,v)w,
(8.40)
z) + g(w,R(u,v)z)
=0.
It follows from (8.38) and (8.40) that g\R(u,y)y,uj
= -g[y,R(u,y)u)
= g(y,R(y,u)vj
= g(R(y,
u)u,yj.
This gives g[R(u,y)y,uj
g[R(y,u)u,yj
g(u, u)g(y, y) - g(y, u)g(y, u)
g(u, u)g(y, y) - g(y, u)g(y, u)'
Thus the flag curvature K(P, y) is independent of y e P for any tangent plane P C TXM. In this case, we denote by K(P) := K(P,y). We call K ( P ) the sectional curvature of the section P C TXM. In a local coordinate system (xl) in M, 6-z = Tl,kdxk. into (8.34) yields /BY*
3Y^
H kl
\.'d^k+^lLmk)~\~dx^+jkml)-
* ~
Using Gl = \^)kyjyk, pi
R kiyy
j
we can express i?. 8 kl y j y l as follows
j i_odGl 2
Plugging them
~ w~
y
i
d2Gi
lh^
«™
+ 2G
d2Qi
9GidGj
WW'WW'
{
'
Structure Equations of Riemannian
Metrics
133
Comparing (8.41) with the coefficients i?lfc(j/) of the Riemann curvature in (6.4), we obtain = R/uyty1,
R\{y)
(8.42)
that is, Rj,(u) = R(u, y)y.
y,u<= TXM.
Observe that
d2R\
d2Ri, 1
dyidy
dyidyk
-{Ri\qyq + Rp\ivp - Rkliqyq -
dyi
^ kj "I" nj
— 3Rj
kl ~ nk
R;lkyp]
lj ~ -"-j Ik
kl.
We obtain _l
d2R\ -( dyldy^y
(
'
We view a Riemann metric as a special Finsler metric. A natural question arises: What are the relationship between the Riemannian curvature tensor on the base manifold and that on the slit tangent bundle? We lift {bi}™=1 to a local frame {el}f=1 for ir*TM, where eilj, := (y,bj). l
Let y be defined by {y,y)
=yle{.
Let ujn+i := dy1 + j/V'6)/.
J := TT*0\ Then <
:= n*e/
are the Chern connection forms of F with respect to {ei}™=1. We have n/ : =
=
(ko/ - w / A uj
^(dsj-ofsoi)
Structure
134
=
Equations
**(±R>klVMl) R/klLokAojl
Thus the coefficients of the Riemannian curvature tensor are functions of x e M only and the Chern curvature vanishes. However, if we take an arbitrary local frame {ej}™=1 for ir*TM, then the coefficients of the Riemsnnian metric curvature tensor might depend on y € TXM.
8.3
Riemann Curvature of Randers Metrics
In general, it is much more difficult to compute the Riemann curvature. In what follows, we are going to give a formula for the Riemann curvature of a Randers metric F = a + (3, where )3 is a Killing form with constant length. Let F = a + (3 be a Randers metric on a manifold M. In a standard local coordinate system (xl,yl) in TM, a and 3 are expressed by «(j/) = \Jaij(x)yiy^
p(y) = bi(x)y'.
Define b^j by
dbt - b^i =: biUej, where 9l := dxl and 6/ := Tfkdxk denote the Levi-Civita connection forms. Let r
H '•= j [bi\J
+ b3\i) '
S
iJ '•= 2 ( 6 i | i ~
bj l
\)
'
The geodesic coefficients Gl of F are related to the coefficients Gz of a by Gi = Gi + Pyi + Q\ where P(V)--
=
Ql(y) : =
2F\y){TijyiyJ aairsrlyl.
-2a(y)brarpsply1}
Riemann
Curvature of Randers
Metrics
135
From now on, we assume that f3 is a Killing form (r^ = 0) with constant length {bibi\j = 0), that is, (3 satisfies r
H = 2 (bi\J + bJ\i) = °.
haijbjlk
= 0.
(8.44)
Equations (8.44) implies that brarpspi = brarpbp\i = 0. Thus P = 0 and G"s are simplified to Gi = & + Qi
(8.45)
with Q1 =
aatrbr\iyl.
Note that Q"s define a tensor Q = Qldi on T M \ {0}. Let R%k and i?'fc denote the coefficients of the Riemann curvature of a and F respectively. Plugging (8.45) into (6.4) yields R\
= R\ + {2Q% - y^Q'^y,
- ( Q ' ) ^ ( ^ ' ) y * + 2Q^Q%jyk},
(8.46)
where Q1,- denote the covariant derivatives of Ql with respect to a. See (5.38). By Lemma 5.2.1), we know that ay = 0. According to (5.39), we have Q\j = a(airbAlyl\
=
aairbmjyl.
For the sake of simplicity, we take a local orthonormal frame {b;}™=1 on M with respect to a so that a(y) =
'jr.WY,
P(y) = J2biy{'
v = yihi-
\ The local frame {bj}™=1 determines a local frame {ej = (y, bj)}™=1 for +i •K*TM and the corresponding local coframe {LJ\ w " } ^ = 1 for T*(rM\{0}). The formula (8.46) still holds the coefficients of the Riemann curvature with respect to {ej}f=1, although it is stated originally in a standard local coordinate system in TM. We have Qi = abi]pyp
Structure
136
Ql\k J
Equations
= a*i|p| f c 2/ p l
= y3(abiMjyp)yk
y (Q u)yk
(Q%i = a-1{bilpyp) (Q* W
= a~1(bilpiJyJyp)
yk +
abi]k\pyp
yj + abi{j
= """HfypV") $jk " a-3(bilpyp)
yjyk + a~lbAJ yk + a^b^
yj
Plugging them into (8.46) yields R\
=
Rik +
2abilplkyp-abilklpyp-a-1(biMgypycl)yk
-a2bilmbmlk
+ 3(bilpyp) (bk]qyq) + (bi]mbmlpyp)
yk. (8.47)
We have the following Ricci identity for the covariant derivatives of (3, h\j\k — h\k\j — bmRi^k-
Using the Ricci identity, we obtain bi\v\kyp = (bi\k\P + bmRimpkS)yv bi\P\qypyq
= =
= bmpyp
-
bmRimkpyp,
-bplilqypyq -(bPlq\iypyq
bmRpmiq)ypy(l
+
= —bm,R"iPlugging them into (8.47), we obtain R\
=
Rik+abiWpyp~2abmRirlPyp
+
-a2bi]mbm]k
+ (bilmbmlpyp)
+ 3(bilpyp)(bk]gy'!)
a'lbmRmiyk yk.
(8.48)
Let R and R denote the Ricci scalar of F and a respectively. Since a is a Riemannian metric, 77 — 1
D m _ _ n t _ __ • rl i ip -rLm ip ~ n
-
p
flymyv-
From (8.48), we immediately obtain an equation for the Ricci curvature. R = R + abmRymyPyp
+ ^-^{a2(bmlp)2
+ 2(6 m | p 2 / p ) 2 }.
(8.49)
Chapter 9
Finsler Spaces of Constant Curvature
As we know, every Riemannian metric of constant curvature A is locally isometric to a canonical Riemannian metric of constant curvature A. However, for Finsler metrics, this is no longer true. For each A, there are infinitely many non-isometric Finsler metrics of constant curvature A. In this chapter, we will discuss some basic properties and well-known examples of Finsler metrics with constant curvature. 9.1
Finsler Metrics of Constant Curvature
The purpose of this section is to derive some equations for Finsler metrics of constant curvature A. We will show that Landsberg metrics of non-zero constant curvature must be Riemannian. Let F be a Finsler metric on a manifold M. Differentiating (8.1) yields the following Bianchi identities:
dn/ = - n / A
=
*j km\l ~ Pj lm\k + Pj kt^
Im. ~~ Pj lt^
km'
(9-1)
Contracting (9.1) with yi yields P-m ki — Plki-m. + Lzkm\i — L ;m|fe + L ktL
lm
— LlltL
km.
(9.2)
Contracting (9.2) with yl yields Rl kiV1 = R\.m
~ R\m 137
+ L\Mlyl.
(9.3)
Finsler Spaces of Constant Curvature
138
By (8.6) and (9.3), we obtain R%km = Rm klV ~ Rk mlV pi
—
Jx
km
_ pi _9pi IX mk ^£x km-
This gives R\i = \(R\.i-R\.k)-
(9-4)
Theorem 9.1.1 Let (M, F) be a Finsler space of constant curvature A ^ 0. Suppose that J = 0. Then F is Riemannian. Proof. By assumption, Rik = \(F28i-gkpy^yi).
(9.5)
Plugging (9.5) into (9.4) yields Rlki=*{9ji6k-9jk6i}.
(9.6)
Plugging (9.6) into (9.2), we obtain Rfki
= ^[9jisk ~ 9jkSlj + LljkV
- Ll^k
+i? f c f L t J , - L ^ L V
(9.7)
Using (9.7), we can rewrite (9.1) as follows 2A [Cjlm5k - Cjkm^lJ
=
Pj\m\l~
+L u-mL jk ~~ Llkt.mL
Rj\m\k +R/ktL
lm~ Pjl[tL
km
jl + L ltL jk.m — L ktL jl.m l
+ L JHkm - L''jk\l-m-
(9-8)
Thus, for Finsler metrics of constant curvature A, the Cartan torsion, the Chern curvature and the Landsberg curvature satisfy (9.8). Contracting (9.8) with ysgiS yields XCjimgksys
- XCjkmgisV3 = Ljkmll
- Ljim\k.
(9.9)
Then contracting (9.9) with yl, we obtain Ljkm\iyl
+ XF2Cjkm
= 0.
(9.10)
Finsler Metrics of Constant Curvature
139
Contracting (9.10) with gim yields Jk{lyl + \F2Ik
= 0.
(9.11)
Thus if Jk = 0, then Ik — 0. By Deike's theorem, we conclude that F is Riemannian. Q.E.D. Theorem 9.1.1 improves a result by S. Numata in [Nu] where he that L = 0 instead. Let c(t) be a unit speed geodesic in a Finsler space (M, F). Let V = V(t) be an arbitrary parallel vector field along c. Consider C(t) :=Cm(V(t),V(t),V(t)),
L(t)
I(t):=Ii{t)(V(t)),
:=L6{t)(V(t),V(t),V(t)). J(t):=J6{t)(V(t)).
It follows from (7.16) and (7.24) that L(t) = C'(t),
3(t) = I'(t).
By (9.10) and (9.11), we obtain the following ODE: L'(t) + XC(t) = 0 J'{t) + AI(i) = 0. Thus C(t) and I(t) satisfy C"{t) + AC(i) = 0
(9.12)
I"{t) + Xl(t) = 0.
(9.13)
The general solutions of (9.12) and (9.13) are ;iven by C(t) I(t)
= =
s A (t)L(0)+si(t)C(0), s A (t)J(0) + s'A(t)I(0),
(9.14) (9.15)
where s\(t) is the unique solution of the following equation y"(t) + Xy(t) = 0,
y(0) = 0 ,
y'(0) = 1.
Assume that F is complete, i.e., every geodesic is defined on the whole line (—00,00). Let us take a look at the following cases:
140
Finsler Spaces of Constant
Curvature
Case 1: A = —1. In this case, (9.14) gives C{t) = sinh(*)L(0) + cosh(t)C(0). Assume that C is bounded. Then L(0) = 0 = C(0). Since c is arbitrary, we conclude that C = 0 and F is Riemannian. By a similar argument using (9.15), we can show that if I is bounded, then 1 = 0. Hence F is Riemannian by Deike's theorem. Case 2: A = 0. In this case, (9.14) gives C(i) =*L(0) + C(0). Assume that C is bounded. Then L(0) = 0. Since c is arbitrary, we conclude that L = 0 and F is a Landsberg metric. By a similar argument using (9.15), we can show that if I is bounded, then J = 0. Hence F is a weak Landsberg metric. Theorem 9.1.2 ([AZ]) Let (M, F) be a complete Finsler space of constant curvature K = A. Assume that the (mean) Cartan torsion is bounded. (a) If X = 0, then F is a (weak) Landsberg metric; (b) If A < 0, then F is a Riemannian metric. When M is compact, the Finsler metric is complete with bounded Cartan torsion. Thus the conclusions in Theorem 9.1.2 hold. When (M,F) is compact with vanishing flag curvature. In this case, all geometric quantities are bounded. By further argument, we can actually prove that the Chern curvature vanishes. It is known that every Finsler metric with vanishing Chern curvature and flag curvature must be locally Minkowskian [BCSl]. We conclude that every Finsler metric on a compact manifold with vanishing flag curvature must be locally Minkowskian. See also [BCS2][Pa] for some rigidity results on Finsler surfaces. The geometric and topological structures of Finsler spaces of positive constant curvature will be discussed in Section 18.3 below.
Examples
9.2
141
Examples
In this section, we will discuss several important examples of Finsler metrics of constant curvature. These examples show that the classification problem seems to be not solved within my life time.
Example 9.2.1 Let F be the Funk metric on a strongly convex domain fl C R™. According to Example 7.2.1, the geodesic coefficients Gl of F are given by &{y) = \F{y)y\
(9.16)
Thus geodesies are straight lines in 0 and F is projectively flat. Plugging (9.16) into (6.4) and using (7.19), we obtain Rik =
-\{F2(y)Sik-9ki(y)ylyi}.
In index-free form, Ry(u) = —-|gy(y,y) u-gy(y,u)
yj,
u <E TXM.
Thus for any tangent plane P = span{?/,w} C TXM, the flag curvature K of F satisfies K(P ) 3 /) = - ^ .
Now we take a look at a special case when 0, = B is the unit ball in £". From the definition, the Funk metric F is determined by |
.
2
/
|
T
F(y)\ We obtain F{
) = V\y\2-(\x\2\y\2-^y)2) 1 — \x\
2
+ (x>y)_
(9.17)
Finsler Spaces of Constant
142
Curvature
E x a m p l e 9.2.2 Let F denote the Funk metric on a strongly convex domain ft C R". The Klein metric F on ft is defined by F(y)-=\{F(y) + F(-y)}-
(9-is)
By (7.19), one can easily prove that F satisfies the following equations Fxkyiyk
= Fxi
and the geodesic coefficients Gl of F are given by &{y)=l-{F{y)-F{-y))yi.
(9.19)
Thus geodesies are straight lines in ft and F is projectively flat. Plugging (9.16) into (6.4) and using (9.19), we obtain
Rik =
-{F2(y)Sik-gkl(y)ylyi}.
In index-free form, Rj/(w) = -[iy(y,y)
u-£y(y,u)y},
u&TxM.
Thus for P = span{y, u}, the flag curvature K of F satisfies K(P,y) = -l. By (9.17), we obtain the Klein metric F(y) on the unit ball B
C IR":
p, , = V\y\2 - ( N 2 H 2 - (x,y)2) {y) i-kl2 B E x a m p l e 9.2.3 ([Sh8]) Minkowski spaces are the most trivial Finsler spaces with K = 0. We may construct non-trivial Finsler metrics with K = 0 as follows. Let F denote the Funk metric on a strongly convex domain ft in R™. Assume that a Finsler metric F on an open subset U C ft satisfies the following system Fxk=(FF)yk.
(9.20)
Examples
143
Observe that Fxkyiyk
= {FF)ykyiyk
=
(FF)yi
This implies that Fxkyk IF
_ (FF)ykyk
_ 2FF
IF
2F
F.
By the above identities and (5.2), we obtain that & = F(y) y\
(9.21)
Thus, the geodesies of F are straight lines and F is projectively flat. Plugging (9.21) into (6.4) and using (7.19), we obtain that R\
= 0.
Thus the flag curvature K of F vanishes. When fi = B is the unit ball in K , the Funk metric F can be expressed by elementary functions in (9.17). However, no explicit elementary expression for F has been found in this case. (j Now we give some examples of Finsler metrics on t h e n-sphere with positive constant curvature. E x a m p l e 9.2.4 ([Brl][Br2]) Let V 3 be a three-dimensional real vector space and V 3 ® C its complex vector space. Take a basis {bi,b2,bs} for V 3 and define a quadratic Q on V 3 x C by Q(u,v) = e i a u V + e i / 3 u V + e " i Q u V , where u = u ^ v = v{hi and a, (3 £ R. For X G V 3 \ {0}, let [X] := {tX, t > 0}. Then S2 := {[X], X £ V 3 \ {0}} is diffeomorphic to the standard unit sphere S in the Euclidean space K . For a vector Y e V 3 , denote by [X, Y] € T{X]S2 the tangent vector to the curve c(i) := [X + tY] at t = 0. Define F : TS 2 -> R by
where 7^[] denotes the real part of a complex number. Clearly, F is welldefined. Assume that \/3\ < a < -|. Then F is a Finsler metric on S 2 . Bryant has verified that F has constant curvature K = 1.
Finsler Spaces of Constant
144
Curvature
Now we shall give an explicit expression for a special class of Bryant metrics on S 2 with (3 = a. Take an arbitrary vector y G TXR2, let X:=(X,1)GR3,
y:=(y,0)GR3.
We have Q(X,X)
=
e lQ |x| 2 + e - i a
Q(X,Y)
=
e i Q (x,y)
Q(Y,Y)
=
eia\y\2
Hence Q{X, X)Q{Y, Y) - Q{X, Y)2 = e2a 'flxfly I2 - (x, y} 2 ) + |y| 2 . The real part of -i Q{X, Y)/Q{X, X) is r _ . Q(X,Y)i L l Q(X,X)\
=
sin(2a)(x,y) |x| + 2cos(2a)|x| 2 + f 4
For a complex number z, the real part of y/z is given by
KM - ^ ^ For Z =
Q ( x , x ) Q ( y , y ) - Q ( x , y ) 2 _ e2« *(|x| 2 |y| 2 - (x,y) 2 ) + |y| : Q(x,x) 2 2 2 a i | x | 4 + 2|x| 2 + e - 2 a i '
= 1
'
TZ(7) ^ j
=
V(lx| 2 |y| 2 - (x,y) 2 ) 2 + 2cos(2 Q )(|x| 2 |y| 2 - (x,y) 2 )ly| 2 + |y| 4 |x| 4 + 2cos(2a)|x| 2 + l lxl2ly|2-(x;y)2+cos(2Q)iyl2 |x| 4 + 2cos(2a)|x| 2 + l
2
/ sin(2a)(x,y) N2 4 2 V|x| + 2cos(2a)|x| + 1 / '
We obtain F([X, Y]) = Jn{Z)
+ |Z|
+
sin(2a)(x, y) |x| 4 + 2cos(2a)|x| 2 + l '
Define Fa(y) := F([X,Y]),
y G TXR2,
Randers Metrics of Constant
Curvature
145
where X = (x, 1), Y = (y, 0) € R 3 . We obtain a Finsler metric Fa on R 2 . This is the pull-back of the Finsler metric F on S 2 by
V(x) := f ,
X
, . *
V
Clearly, we can generalize Fa to a Finsler metric defined on R n without any modification. One can show that the generalized metric Fa on R™ also has constant curvature K = 1. Note that FQ is just the spherical Riemannian metric on R n , y/|yl 2 T+V I ( | x | 2 | v l 2 X- ( x , y ) 2 ) rl .+ ,J, |x|a: "
F0(y) = VIJI
( 9 - 23 )
All F a are pointwise projective to the Euclidean metric.
9.3
Randers Metrics of Constant Curvature
The Funk metric F in (9.17) is a Randers metric. In fact the Funk metric on any shifted ellipsoid Q C R™ is a Randers metric. We would like to find more Randers metrics of constant curvature. According to Section 8.3, the Riemann curvature a Randers metric F = a + (3 with a Killing form (3 of constant length takes a relatively simple form (8.47). In this section we will investigate this type of Randers metrics. Let F = a + j3 be a Randers metric on M, where (3 is a Killing form of constant length. At a point, take a local orthonormal frame {t>i}™=1 on M so that <x(y)
. E^1)2'
y = yibi.
Piy) = biy\
By assumption,
hi + hi = °'
'°ihi
= °-
( 9 - 24 )
Suppose that F is of constant curvature A, i.e, the coefficients of Riemann curvature with respect to {bi}" = 1 are in the following form:
R\
=
\F2(6ik-^Fyk)
Finsler Spaces of Constant
146
Curvature
\{{a2+p2)8lk-yiyk-l3bkyi}
=
+\a(2(36ik-bkyi")-\a-1Pyiyk Compare it with (8.47), we obtain R\
=
A{(a 2 + / ? 2 ) * f c - y y - / ? 6 f c i / i } +a2bilmbm]k
bmRmiyk
- 3(bilpyp)(bklqyg)
- (bi]mbm\Pyp)y
(9.25)
a2(biWpyP-2bmRimkpyp)
+
-Xf3yiyk + Xa2
=
(2(36ik-bkyi).
According to (8.42) and (8.43), Rlk and R/ki
R\ = Rhrfy',
j kl Rh
(9.26)
are related by
1 r d2R\ 3 I dy'dy1
d2R\ dyidyk .
}
Differentiating Rlk in (9.25) with respect to yi and yl, we obtain R/ki
=
X\6ikSji - Siidjk) + X\bjbi6ik - bjbkSii) +bi\pbp\k5ji - bi\pbp\i6jk + b^bj^ - bi\kbj^ — 26fc|;6j|j(9.27)
Contracting (9.27) with bi and yl, and using (9.24), we obtain bmR^y1
= X(bkyj - (35jk)
bmRmk = X (a2bk - / ? / ) .
Plugging them into (9.26) yields k\k\p = X (bkSip - biSkpJ. By taking the Hessian of M(3\\2 = \{bi)2,
\M
\k\i
=
bi\kbm + bibi\k\i
=
h\kbi\i + Xbi(bk6u - bidkA
=
b^bm + X ^
- \\(3\\2Skiy
(9.28)
Randers Metrics of Constant
Curvature
147
we obtain 6P|ft6P|/ = A(||/3|| 2 J f c ; -6 f c 6 i ).
(9.29)
(&P|*)2=A(n-l)P||2.
(9.30)
From (9.29), we obtain
Thus A > 0. If A = 0, then (3 is parallel and R\ = 0. Plugging (9.29) into (9.25) yields R\
=
A{ [(1 - ||/3|| 2 )a 2 + /32] Slk + a%bk - (1 - ||/?|| 2 )yV -Why*
+ hyk))
- 3(bilpyP)(bklqyq).
(9.31)
Conversely, one can easily verify that (9.28) and (9.31) imply (9.26). Thus under the assumption (9.24), F is of constant curvature A if and only if (9.28) and (9.31) are satisfied. Proposition 9.3.1 Let F = a + (3 be a Randers metric where (3 is a Killing form of constant length. F is of constant flag curvature A if and only if a and (3 satisfy (9.29) and (9.31). In this case, either A > 0 or A = 0 (hence a is flat and (3 is parallel). H. Yasuda and H. Shimada [YaSh] proved a better result than Proposition 9.3.1. They proved that a Randers metric F = a + (3 is of constant flag curvature A > 0 if and only if (a) P is a Killing form of constant length with respect to a, (b) a and (3 satisfy (9.29) and (9.31). However, their proof is much more complicated, so it is omitted here. See [Ma3] for another proof of Yasuda-Shimada's theorem. Recently, D. Bao and Z. Shen have found a specific family of Randers metrics on S 3 with constant flag curvature K = 1. The metrics on S 3 are described in the following example. Example 9.3.1 ([BaSh2]) We view S 3 as a compact Lie group. C^C^C 3 be the standard right invariant 1-forms on S 3 satisfying dC1 =2C 2 AC 3 ,
dC2 = 2 < 3 A C \
dC3 = 2C1AC2-
Let
148
Finsler Spaces of Constant
Curvature
The metric
«i(y) := V I C W + l C W + l W is the standard Riemannian metric of constant curvature 1. For k > 1, define
ak(y) := xA 2 [CW + M<2(»)]2 + M C W and
Then f* := «fc + &
(9.32)
is a Finsler metric on S 3 . Bao-Shen show that Fk is of constant curvature K = 1 for any k > 1. We give a brief argument below which is different from that in [BaSh2]. First, we show that (3k is a Killing form of constant length. Take an orthonormal coframe for T*S 3 , ^:=k(\
62:=y/k(2,
03 =
^(3,
so that
ak(y) 0k(y)
= V[01(y)? + [02(y)]2 + {93(y)?, = b1el(y) + b262(y) + b3e3(y),
where &i =
,
&2 = 0,
b3 = Q.
Clearly, (3k has constant length with respect to ak, that is,
\\pk\U = ^PThe Levi-Civita connection forms 6 -l are given by 6 -1 + 6^ = 0 and
e2l=e\
e3l = -e2,
e2
(H*-
Randers Metrics of Constant
Curvature
149
By definition, b^j are defined by dbi - bjOf =: bi{j9j. A direct computation gives &i|i=0,
62|i=0,
6i|2=0,
6212=0,
&i|3 = 0,
62|3 =
,/£2 — k ,
i/£ 2 — k &3|1 = 0,
&3|2 =
7
,
&3|3 = 0.
Thus h\j + bj\i = 0. We conclude that (3k is a Killing form with respect to akBy definition, b^k are defined by dbi\j - bk\j9i
- bi\k0j
=: bi\j\k9
•
A direct computation gives 0l|2|fc =
T
G
2k,
Ox|3|fc =
d3k,
02|3|fc = U.
Other components are determined by bt|j|fc + ^j|t|fc = 0- O n e c a n easily verify that ak and /3fc satisfy (9.29) with A = 1. Now we compute the Riemann curvature of ak- By definition, Rf kl are defined by
deZ-efAe^^R/^Ae1, where R/ki
+ R/ik = 0. A direct computation gives R2 12
=
1)
R2 13
^ 3 13
—
1>
- ^ 3 12 — 0 '
^ 3 23 — .
—
3,
R3
=
13
0,
— 0,
i t 2 13 = 0,
-^3 23
R3
—
12
0'
— 0.
150
Finsler Spaces of Constant
Other components of R? Rjki
are determined by
kl
+ R/ki
Curvature
=
0,
Rjki
+ Rkij
+ R-iljk = °-
One can easily verify t h a t ak and (3k satisfy (9.31) with A = 1. By Proposition 9.3.1, we conclude t h a t Fk = ak + (3k has constant flag curvature K = 1. According t o Example 7.3.2, Fk satisfies S = 0,
E = 0.
T h e volume form of ak is given by dVQk = k2(l
A C2 A C3 =
k2dVai.
By (2.10), t h e volume forms of Fk and ak are related by
dVFk = (l - H & H ^ ) 2 ^ = (l - (l - l))2k2dVai
= dVai.
This gives Vol Ffe (S 3 ) = V o l a i ( S 3 ) = Vol(S 3 ). T h a t is, the volume of (S3,Fk) is a constant. This implies t h a t for any point p e S 3 , there is a unique point q € S 3 with d(p,q) = n such t h a t all geodesies c(t), 0 < t < oo, with c(0) = p is minimizing on [0,7r] with c(t) = q. T h e details will be given in Section 18.3 later. Let (x, y, z, u, v, w) be t h e standard coordinate system in T R 3 = R 3 x R 3 . Let ip± : R 3 - * S3^. denote t h e diffeomorphisms defined by rpe{x, y, z) :=
( x , y, z, e), +y +z +1 >
1
2
yjx
2
(x, y, z) G R 3 ,
v
where e = ± 1 . We can express FK on R 3 . Pulling back £' by ipe onto R 3 , we obtain —edx — zdy + ydz
x
~
x2 + y2 + z2 + 1 zdx — edy — xdz
=
x2 + y2 + z2 + 1 —ydx + xdy — edz x2 + y2 + z2 + 1
2 C
^3 ~
Randers Metrics of Constant Curvature
151
Plugging them into aK and (3K, we obtain ah
=
\Jk2{su + zv — yw)2 + k(zu — ev — xw)2 + k(yu — xv + ew)2 1 + x2 + y2 + z2 1 + xz + yz + z2
One can extend above construction to odd-dimensional spheres S 2 n + 1 , because of the special fibration structure S 2 n + 1 -4 C P " = S ^ + V S 1 . fl
A Finsler metric F on an n-dimensional manifold M is of constant Ricci curvature A if Ric(y) = (n-1)A J F , 2 ( 2 ;),
y G TM.
Finsler metrics of constant Ricci curvature are also called Einstein
metrics.
Let us take a close look at Randers metrics of constant Ricci curvature A. Let F...= a + (3 be a Randers metric. We still assume that /? is a Killing form of constant length, so that the Ricci curvature of F is in a relatively simple form (8.49). Take a local orthonormal frame {bi}™=1 for TM so that
X^)2-
a(y) = \
P(y) = biyi,
y=
yibieTxM.
i=i
By Assumption, bi\j + bj\i = 0,
bibi\j = 0,
bibj\i = 0.
According to (8.49), the Ricci scalar R and R of F and a are related by R = R + abmRymyvVv
+ ^j{a2(bmlp)2
+ 2(6 m | p 2 /") 2 }.
(9.33)
Assume that F is of constant Ricci curvature A. Then \{a + 0)2 =R + abmRymyPyP
+ _ L - | a 2 ( 6 m | p ) 2 + 2(6 m | p 2 / p ) 2 }.
Finsler Spaces of Constant
152
Curvature
We obtain bmRymyPyp
= 2X0,
R+ ^j{a2(bmlp)2
(9.34) + 2(bmlpyP)2}
=X(a2+p2).
(9.35)
^ V + — [ { M 6 m | P ) 2 + bm\ibm\j} = 2\(6ij + bibj).
(9.36)
Equation (9.35) is equivalent to the following
Contracting (9.36) with bi and bj and using (9.34), we obtain (bm]p)2 = (n-l)\\\(3\\2.
(9.37)
From (9.37), we see that A > 0. Plugging (9.37) into (9.35) or (9.36), we obtain R = \{(1-
||/3|| 2 )a 2 + P2} - ^(bm\Pypf-
(9-38)
or equivalently, Ryiyj = 2A{(1 - ||/3|i 2 )^ + bibj} - ^jbmlibmlj.
(9.39)
Note that if A = 0, then bm]p = 0 by (9.37) and R = 0 by (9.38). Thus a is Ricci-flat and /? is parallel. Conversely, assume that the Ricci scalar R of a satisfies (9.38), one can easily show that (9.34) and (9.35) hold. We have proved the following Proposition 9.3.2 Let F = a + j3 be a Randers metric where P is a Killing form of constant length. F is of constant Ricci curvature A if and only if the Ricci scalar R of a satisfies (9.38). In this case, either A > 0 or A = 0 (hence a is Ricci-flat and (3 is parallel).
Chapter 10
Second Variation Formula
In previous chapters, we have introduced several important curvatures for Finsler metrics. We have also discussed some relationships among these quantities. In this chapter, we are going to introduce another nonRiemannian quantity — the T-curvature. This quantity has a close relationship with the distance function.
10.1
T-Curvature
Let (M,F) be a Finsler space. Given a vector y 6 TXM \ {0}. Extend y to a geodesic field Y in a neighborhood U of x. Let D denote the Levi-Civita connection of the induced Riemannian metric g = gy on U. Define T„(«) := g„(D„V,y) -gx(i>„V,y),
v G TXM,
where V is a vector field with Vx = v. T = {Ty}yeTM\{o} T-curvature. T is also called the tangent curvature in [Sh4]. From the definition of T we have Tv(Xv) = X2Ty(v),
(10.1) is called the
A> 0
and T v (l/) = 0. Now we are going to derive a formula for Ty(v) 153
in a local coordinate
Second Variation
154
Formula
system. First, contracting (5.28) with yl gives , / „ rk
-
1 d 9 j l
i d9ml
f
d
9jm\
r,r
,„ ,
rk
Observe that dfjij _ dgij
dY
# = igf<m*w)dx
k'
Since Y = Y%-^-z is geodesic, Y satisfies BY1 Y -— + 2Gi(Y)=0. k
(10.3)
We obtain vVidM _ = ±Y*^(Y), k k
(10.4)
Y
(10.5)
dx
dx
= Yk^(Y)-ACi3k(Y)Gk(Y).
"&
Let f} fe denote the Christoffel symbols of g. By (10.2), (10.4) and (10.5), we obtain v^ 9kl rk m > ~ =
1 dg l
( i 2\dx™
.
d9ml
+
\{B%{Y)
dxi
dffj" 1 vi dxi V
+
°ixT{Y)- l&W}*1
+
2C
^n(Y)Gk(Y)
Ylgkl(Y)Tkm(Y)
=
In particular, the following holds ylgkl(x)tkm(x)
= ylgkl{y)Tkjm{y).
(10.6)
Thus gx(pvV,y)
= ylgkl(x){dVk(v) =
+
rkrn(x)^vm}
Vl9ki(v){dVk(v)+r*m(y)v*vm}.
(10.7)
On the other hand, K„(pvV,y)
= ylgki(y){dVk(v)
+ Tkm{v)v^vm).
(10.8)
T-Curvature
155
From (10.7) and (10.8), we obtain T »
:= ylgkl{y){vkjm{v)
- Tkm{y))^vm.
(10.9)
From (10.9), we see that Ty(v) is independent of the extensions of y and v. Proposition 10.1.1
P = 0 if and only i / T = 0.
Proof. Suppose P = 0. By (7.13), we see that Tjm(y) y. (10.9) immediately implies T = 0. Suppose that T = 0. In local coordinates,
are independent of
yl9ki(y){rUv) - r* m (y)}^ m = o.
(10.10)
This implies that Vjk(v) are independent of v. Hence P = 0. The proof is left for the reader. Q.E.D. The T-curvature is defined for Finsler metrics. There is another interesting quantity which depends only on the spray of the Finsler metric. Let F b4 a Finsler metric and
be a spray of F. Let Nl(u,v)
:= vjNUu) - ujNUv) = vj^-^(u)
-
uj^—(v).
For a pair of vectors (u,v) G TXM x TXM, define N(u,„) := One can verify that N : TxMxTxM it is anti-symmetric,
N\u,v)^-\x. -» TXM is a well-defined map. Further,
N(u, v) +N(v,u)
= 0,
and satisfies the following homogeneity N(Xu,v) = XN(u,v), Proposition 10.1.2
N(u,Xv) = AN(u,u),
P = 0 if and only if N = 0.
The proof is left for the reader.
A > 0.
156
Second Variation Formula
10.2
Second Variation of Length
In this section, we will derive the second variation formula for a geodesic. The geometric meanings of the Riemann curvature and the T-curvature lie in this important formula. Let (M,F) be a Finsler space. Take an arbitrary local frame {e*}f=1 for n*TM. Let { w > n + , } ? = 1 denote the local coframe for T*(TM \ {0}), corresponding to {e;}. The correspondence is given by (5.18). For vectors X, Y on TM \ {0} and a section U = Wej of T T T M , define Q{X, Y)U := Ujn/(X,
Y) e*.
Let V denote the Chern connection on n*TM. By (8.3), we have v
x V Y U ~ v f vx&
~ V[x,Y}U = n(X, Y)U,
where U e C°°{TV*TM) and X, Y e C°°(T(TM \ {0})). Let {ei,ei}f =1 be the local frame for T*(TM \ {0}) that is dual to { w ^ w " ^ } ^ . We have n(e fc) e«)ej
=
R/kl
a
(10.11)
n(e fc ,e,)ej
=
P/ f c , *.
(10.12)
For a vector o n T M \ { 0 } , X = Xiei + Yiei, we always denote X =
Xiei.
Let V = Viei + Uiei,
f = Tlei + Slei
be vector fields on TM \ {0} . Fix a vector y e TXM \ {0}. Assume that T% = yl,
S% = 0,
where yl are determined by (y,y) = yzei.
Second Variation of Length
157
By (8.24) and (8.29), we obtain g(n(V,f)V,f)\y
V^Vkyiylg(n{ek,el)ej,el)
=
+V^Ukyiylg(n(ek,el)ej,ei) =
V>Vk ylgipR3pklyl
=
V*Vk y'gipR;,
- V*Uk + VWk
yigtpPjplkyl y'g^P^
k
=
-VJV gjpR\.
We obtain an important identity = -VWk gjpR\.
g(n(V,f)V,f)\y
(10.13)
Let {b,;}™=1 be the basis for TXM determined by ei\y = G/,bi). Let V := V^i € TXM. We rewrite (10.13) as follows g(n(V,f)V,f)\y
= -Sy(Ky(V),v)
(10.14)
Let c : [a, b] —> M be a geodesic with F(c) = A > 0. Consider a piecewise C°° variation of c if : (-£,£) x [a,b] ->• M More precisely, there is a partition a = to < • • • < tk = b such that (a) # is C° on (-£,£); (b) H is C°° each (-£,e) x \ti,ti+i) and (c) # ( 0 , i ) =c(t), a
x
dH
The map if : (-£,£) x [a, 6] - 4 T M
Second Variation
158
Formula
is not C° on (—£,e) x [a, b], but it is C°° on each (—e,e) x [£i,£i+i]. i? is not C° on (e,e) x [a, 6] unless H is C°°. Let -
V:
dW d d2H{ { dt dx dt2 dH* d d2^ „s n d 9a;,* + dsdt
dH dt dH = — 5s =
d dyi d dyl
Then -
dH*
T:
d
-
dW V:=
-dTdi-
=-df "
Let
v(t)--=v^£Mt) = f{0,t). Then f(0, t) = (c(t), c(t)),
V(Q, t) = (c(t),
V{t)).
Let V denote the Chern connection on ir*TM. The torsion-free condition (5.21) is equivalent to that V ^ y - VyX
= {X,Y}.
(10.15)
Without loss of generality, we may assume that T and V can extended to smooth vector fields on TM \ {0}. Then [T, V] is independent of any extensions. Observe
\fv]-(d2Hi
d2Hl
1
dsdt) dxi
'
J
\dtds
\
d
(d3Hl
d3Ri
\dt2ds
dsdt2)
\
d
dyi
By (10.15), VyT=VfV.
Now we consider the following length function
(10.16)
Second Variation of Length
159
We obtain L'(s)
= J
g{f,f)-V2g(Vvf,f)dt
= J g{f,f)-l'2g(vfV,f).
(10.17)
Differentiating (10.17) and using (10.16) yield pb
L"(s)
g(f,f)-1f2{g(V^tV,f)+g(^tV,Vfv)}dt
= J
g{t,f)-*l2g(vtV,t)2dt
- J = J
g{f,f)-ll2g(si(V,f)V,f)dt
+ J g(f,f)~1/2{g(VfVyV,f) 2
g(f,f)-3'2g(yfV,f)
-J
dt +g(vfV,Vfv)}
dt
Now we deal with the above identity along c term by term. Since c is a geodesic, A,
^
9H* ,n , d =
d2H\n
.d
c\t)-^- 2Giv(c(t))^= c{t) —l . v v 'dxl " dyl Sx
Thus T| s = 0 is horizontal. Equation (10.14) gives g(Q(V,f)V,
f)|,=o =
-Sc(t)(^nt){V(t)),V(t)).
Observe that g(vfVyV,f) First, we study VfT.
= f [fl(VvV-,f)] - 5 ( v ^ , V f T ) .
(10.18)
160
Second Variation
Formula
We obtain V t f | . = 0 = {c + 2Gi{c)}di\c = 0. Now we fix t with a < t < b. Let at(s) := H(s,t). Similarly,
We obtain
V^V-|fl=0
=
{al + &iTjk{c)dkt]di\t
=
[a1 + 2Gi(&t)}di\6
+ {Fjk(c) - Fjk(at)}&ia?
d^.
Assume that at has constant speed with F(&t(s)) =: Ct- Then the geodesic curvature Kt of at (s) at s = 0 is given by
- = ^ { ^ ) + ^(0))}£-
o-«(0)-
This gives 5(v^V-,f)|i=o
= (C t ) 2 gc W («t,c(t)) - T fi(t) (<7 t (0)).
Note that the above function is piecewise C°°. Thus
J g(f,fr1/2g(vtVyV,f)\^0dt A" 1 (Cbf
=
g£(6) («6, C(6)J - (C„) 2 g c ( a ) ( « a , c ( a ) J
-A" 1 T c ( f c ) ( ^ ( 0 ) ) - T , ( a ) ( c r a ( 0 ) ) . Observe that
We obtain
V t V-|, =0 = {^
+
™$(c)}ft|
Second Variation of Length
161
This gives
g (ytv, v f v) | s=0 = gc(t) (DC^W, DdK(t)), 5(vtV-
, f ) | s = 0 = gc(t)(DcV(t), c(t)).
Let F x ( t ) := K(i) -
X~2gc{t)(V(t),c(t))c(t).
We have gc(t)(De ( t )^ X (t),D ( ; ( t ) K X (t)) = gc(t)(pnt)V{t),B m V{t))
-
A-2gi(t)(vmV{t),c(t)).
We obtain the following Second Variation Formula: L"(0)
=
A-1/
{gi(D^,D^-L)-gc(Rc(^x),^i-)}^
+A- 1 [F 2 (y(6))g ( i ( 6 ) (K 6 (0), C (6)) -F 2 (V(a)) g ( i ( a ) (rv a (0),c(a))] +A" 1 [T £ ( a ) (V(a)) - T 6(6) (V(6))].
(10.19)
The second variation formula is derived in [AbPa]. But the T-curvature is not clearly defined. The second variation formula for variations with fixed endpoints is derived in [BCSl] in a different way. Let c(t), a < t < b, be a unit speed geodesic in (M, F). Take a special geodesic variation H(s, t) of a geodesic c(t) such that H(s, a) = c(a) and
L"(0)
=
/
{gc(DeVx,D^x)-K(P,c)ge(v±,Vx)}dt
Second Variation
162
Formula
-T£(6)(
(10.20)
where P(t) = span{c(£), V(t)} is a family of tar gent planes along c. Thus the behavior of the geodesies issuing from a point depends on the flag curvature and the T-curvature. 10.3
Synge Theorem
The second variation formula has many applications. In this section, we will apply this formula to prove that any even-dimensional compact Finsler space with K > 0 has simple fundamental group m = {0} or Z 2 (Synge Theorem). First we prove the following Lemma 10.3.1 Let (M,F) be an even-dimensional, oriented Finsler space and c be a closed geodesic. There is at least one parallel vector field W along c, which is not tangent to c. Proof. Let c : [0,r] -> M denote the closed geodesic with c(0) = x = c(r) and Pc denote the parallel translation along c. By Lemma 5.3.1, Pc : TXM —> TXM is a linear isometry with respect to gy, where y — c(0) = c(r), i.e., g„ (P c («), P c (v)) = sy(u, v),
u,ve TXM.
(10.21)
Let
/ : = f e W gy{y,v) = 0.}. It follows from (10-21) that Pc : y1- -> y1- is a linear isometry. Since dimj/ x = n — 1 is odd, there is at least one vector w G y x such that Pc(w) — w. Let W(t) be the parallel vector field along c with W(0) =w = W{r). Then W(0) = Pc(w) = W(r). Thus W must be a smooth parallel vector field along c.
Q.E.D.
Lemma 10.3.1 has important application. Theorem 10.3.2 (Synge, 1926) Let (M, F) be an even-dimensional closed oriented Finsler manifold with K > 0. Then M is simply connected.
Synge
Theorem
163
Proof. Suppose that TTI(M) contains a non-trivial element a. By Proposition 12.3.1 below, there is a closed geodesic c G a with L(c) = \a\. Assume that c(i), 0 < t < \a\, is parametrized by arclength. It follows from Lemma 10.3.1 that there is a parallel vector field W{t) along c. Take a variation H : {-£,£) x [0, |Q|] -> M of c such that
W(t) =
^-(0,t).
Applying the second variation formula to
and using the fact that Wc(0) = Wc(|a|); we obtain
L"(0) = -Ja
gm (Rc(t) (W(tj), W(t))dt < 0.
Thus cs(t) := H(s,t) is shorter than c for small s > 0. Since cs is homotopy equivalent to c, we have \a\ < L{cs) < L(c) = \a\. It is a contradiction, since c is assumed to be the shortest closed curve in a. Q.E.D.
Chapter 11
Geodesies and Exponential Map
Geodesies are the most important objects in a Finsler space. In this chapter, we will define the exponential map and use it to discuss some basic properties of geodesies. We will show that the geodesic completeness is equivalent to the metric completeness (Hopf-Rinow theorem).
11.1
Exponential Map
In this section, we will define the exponential map and discuss its regularity. The exponential map plays a very important role in comparison geometry. Let (M,F) be a Finsler space and ir : TM —> M denote the natural projection. For a vector y e TXM \ {0}, let cy(t) denote the integral curve of the induced spray G in TM \ {0}. By definition, cy(t)
:=irocy(t)
is a geodesic in M with cy(Q) = y. In a standard local coordinate system (xl,yl) in TM, geodesies are characterized by the following equations + 2Gi{c) = 0.
(11.1)
By the O.D.E. theory, cy smoothly depends on y € TM \ {0}. The homogeneity of G implies that for any A > 0 and y G TM \ {0}, c\y(t) = cy(\t),
Vf>0.
Let U(M) denote the set of all tangent vectors y € TM such that cy is 165
166
Geodesies and Exponential
Map
defined on [0,1 + e) for some e > 0. Define exp : U(M) ->• M by exp(y) := cy(l). We call exp the exponential map. Since cy smoothly depends on y € TM \ {0}, we conclude that exp is C°° on U{M) \ {0}. For a point x £ M, let expx := exp \UXM • UXM -> M. expx is C°° on UXM - {0}, where UXM := UM n T X M. We call exp x the exponential map at a;.
For any vector y £ UXM, we identify TyUxM with T X M in a natural way. Then the differential of exp x at a vector y G UXM is a linear map ^(expJij, : T^M -> T Z M,
z=
expx(y).
Thus, d(exp)| 0 :TXM
-*TXM
is an endomorphism, provided that exp^ is differentiable at the origin of TXM. The problem is whether or not exp^. is C°° at the origin. This problem was answered by Whitehead. Theorem 11.1.1 (Whitehead [Wh]) Let (M,F) be a Finsler space. The exponential map exp is C1 on the zero sections of •K : TM —> M and for
Jacobi Fields
167
any x G M, d(exp x )| 0 : TXM —> TXM is the identity map at the origin 0 G TXM. The proof is technical, so is omitted. See [BCSl] for details. For any precompact open subset A c M, let Ur(A) denote the rneighborhood Ur(A) of A in TM, Ur{A) := { G TXM, F(y) < r, x G
A}
C TM.
For a point a; G M, let B r (:r) := Ur({x}), i.e., B r (z) := {y G TXM, Fx{y) < r } C T X M. By Theorem 11.1.1, we immediately obtain the following Corollary 11.1.2 Let (M,F) be a Finsler space. For any precompact open subset A C M, there is a positive number r0 = r0(A) > 0 such that for any z G A, exp z : TZM -> M is a C1 -diffeomorphism on B 2 (r) C TZM onto its image for all 0 < r < r0. In Section 11.3, we will show that expx[Bx(r)] coincides with the metric ball B(x,r). 11.2
Jacobi Fields
The Riemann curvature is defined via geodesic variations. Thus we will use geodesic variations to study the relationship between geodesies and the Riemann curvature. Let (M, F) be a Finsler space. Let c : [a, b] -¥ M be a geodesic and H : (—e,e) x [a,b] —• M a geodesic variation of c, namely, each cs(t) := H(s,t) is a geodesic. Let J(t) := Q?L(Q,t). By Lemma 6.1.1, we know that J(t) satisfies DeDe J + Rc( J) = 0.
(11.2)
Equation (11.2) is called the Jacobi equation of c. Conversely, for every vector field J(t) along c satisfying the Jacobi equation (11.2), there is a geodesic variation H of c whose variation field is equal to J{t). Thus we call a vector field J{t) along c satisfying (11.2) a Jacobi field.
Geodesies and Exponential
168
Lemma 11.2.1 F(c) = A. Then
Map
(Gauss) Let J(t) be a geodesic along a geodesic c with
gc(t) (•/(*). c(t)) = A2(a + bt),
g d ( t ) (p6J(t),
c(t)) = A26.
(11.3)
where Jx(t)
:=J{t)-{a
+ bt)c(t)
is also a Jacobi field along c such that J±(t) to c(t) with respect to g^t)-
and DcJ x (i) are orthogonal
Proof. Observe that |[gc(t)(j«,c(*))]
=
gc W (Dc7(t),c(t))
_[g.(t)(j(i),c(*))]
=
gc(t)(DcDc7(t),c(t))
=
-gi{t)(^c(t)(J(t)),
=
-gc(t)(^(*), Rc(t)(c(<)))=0.
C(t))
Thus (11.3) holds for some constants a,b. Plugging J{t) = J±(t) + {at + b)c(t) into (11.2), we obtain D£DeJx(i)+Re(J±)=0. Thus J x ( t ) is also a Jacobi field along c. By (11.3), gc(t) (jHt),
c(*)) = 0,
ge(t) ( D , ^ ( i ) , c(t)) = 0. Q.E.D.
Fix a unit vector y e TXM and let c(i) := expa.(iy), 0 < £ < a. Consider a special geodesic variation H(s, t) := expx[t(y + sv)}),
0 < t < a, |s| < e.
Jacobi Fields
169
By Lemma 6.1.1, J(t) := ^ ( 0 , t ) = d(expx)\ty(tv)
(11.4)
is a Jacobi field along c. Namely, J(t) satisfies (11.2) on (0, a). Since exp x is C 1 at the origin, J(t) is C° at t = 0. We can extend c to a geodesic c on (-S, a) for some 6 > 0. Take a number r close to 0 with 0 < r < a. There is a unique Jacobi field J along c satisfying J{r) = J ( r ) ,
D £ J » = D, J(r).
By the uniqueness, J(t) = J(i),
0 < * < a.
Thus J(i) can be extended to a Jacobi field J(i) along c. By this observation, we can prove the following L e m m a 11.2.2 The Jacobi field J{t) in (11.4) JS C°° along c{t) = expx(ty), 0 < t < a. It satisfies the following initial conditions: J(0) = 0,
Dd J(0) = v.
Proof. First observe that J(0)=d(expx)|o(0)=0. By the above argument, we may extend J to a Jacobi field J along a geodesic c(t), —6 < t < a. Write j(t) = tW(t),
-6
where W(t) is a C°° vector field along c and W(t) = d{expx)\ty(v),
0
Observe that BrJ(t) = Note that W(t) = d(expx)\ty(v),
W(t)+tBcW(t). t > 0, satisfies
W{0) = lim. d(expx)\ty(y)
= d(expx)l0{v)
= v.
Geodesies and Exponential
170
Map
We obtain Di J(0) = W(0) = v. Q.E.D. Let c : (—£, e) —> M be a unit speed geodesic with c(0) = j / G T p M. For a vector i> € T P M, let J be the Jacobi field along c satisfying J(0) = 0,
D d J(0) = v.
Consider the following function f(t):=Si(t)(J(t),Jt)). Define Rc ( t ) : Tc{t)M -> T c ( t ) M by R* ( t ) (n*)):=Dc( t ) [Rc(t)(V r (t))], where V(£) is a parallel vector field along c. Similarly, we can define Rc(t)By the Jacobi equation (11.2), we obtain /'
=
2g6(D,J,j),
/"
=
2ge(DeJ,Dej) - 2g, ( R , ( J), j ) ,
/(3)
=
-8ge(Rc(J),Dcj)-2ge(Re(J),j),
/(4)
=
-8ge(Rc(DeJ),Dej)+8ge(Rc(J),Ri(J)) -12g£(Ri(J),Dcj) -2ge(R,(J), j ) .
We obtain the following P r o p o s i t i o n 11.2.3
T/ie aftoae Jacobi field J(t) satisfies
gc(t) (-/(*), . 7 ( 0 ) = gy(«, ^
Sd(t)(D6J{t),J(t))
- \sy ( R J M v)tA + o(* 4 ),
=Sy(v,v)t-lgy(Ry{v),vy
+ o(t:i).
(11.5)
(11.6)
At the beginning, we have mentioned that exp is C°° on U(A) \ {0}. Now we discuss the regularity of exp on the set of zero sections in TM.
Minimality
of Geodesies
171
We first consider the case when F is a Berwald metric, i.e., the geodesic coefficients Gl(y) = ^Yljk{x)y3yk are quadratic functions in y G TXM for all x G M. Geodesies are characterized by dV
„; . s dcj dck
n
By the O.D.E. theory, cy(t) smoothly depend on y G TM (for small t). Thus exp is C°° at y — 0 G TM. In particular, exp x is C°° at y = 0 for all x e A. Assume that exp x is C2 at y = 0 for all a;. Let f(y) := exp(y). By definition, for any r > 0 and any precompact open subset A, there is a number e > 0 such that c(t) = f{ty), 0 < £ < l + £ , is a geodesic for every y G TXM, where x € A, with F(j/) < r. Substituting c into the geodesic equation (11-1) and letting t = 0 yields
In other words, Gl(y) are quadratic in y G TXM. By definition, such a Finsler metric is a Berwald metric. We have proved the following Theorem 11.2.4 (Akbar-Zadeh [AZ]) Let (M,F) be a Finsler space. The exponential map exp is C2 at zero sections if and only if F is a Berwald metric.
11.3
Minimality of Geodesies
In Section 5.1, we define a geodesic to be a critical curve of the arc-length function of curves joining two fixed points. In this section, we shall prove that geodesies are locally minimizing curves. We have the following Proposition 11.3.1 Let (M,F) be a Finsler space. Suppose that expx : TXM —*• M is a C1 diffeomorphism on B x (r) C TXM. Then for any y G SXM, the radial geodesic c{t) = expx{ty),
0 < t < r,
Geodesies and Exponential
172
Map
has minimal length r among all piecewise C1 curves in M joining x to z = c(r). Moreover, for any 0 < t < r, the exponential map restricted to Bx(t) C TXM and Sx(t) = dBx(t) are diffeomorphisms onto the metric t-ball B(x,t) C M and S(x,t) = dB(x,t) respectively, exp, : Bx(t) -> B(x, t),
exp x : S x (t) -> S{x, t).
(11.7)
Proof By Corollary 11.1.2, we may assume that for some r and e > 0, exp x is a C1 -difFeomorphism from B x (r + e) C TXM onto its image. Take an arbitrary piecewise C 1 curve a : [0,1] -> M,
= X = C(0),
cr(l) = z = c{r).
First we assume that a is contained in exp^B^fr + e)\. Then a(s) = expx[t(s)y(s)]
y(s) e SXM,
where t = t(s) is a piecewise C°° function with t(0) = 0, t(l) = r and 2/ : (0,1] -»• S X M is a piecewise C°° map with y(l) = y. It is easy to show that lim
y(s)
=
s->0+
Hence, we may set y(0) :=
0 < s < 1, 0 < t < r ,
such that a(s) =
H(t(s),y(s)).
Minimality
of Geodesies
Let
a* "
of Then
a = y— + v. as
For each s € [0,1], Vs(t) := V(t, s) is a Jacobi field along cs{t) := exp x (*i/(s)),
0 < t < v.
V(t) satisfies
V.(o) = o,
D6svs(o) = y(s)^-y(s).
By the Gauss Lemma (Lemma 11.2.1), Vs(t).Lcs(t) with respect to Sc.{t)Thus SY(Y,V)=Q.
(11.8)
Recall (1.14) g„(l/, w) < F(y)F(w)
\/y ^ 0 .
(11.9)
Taking w = a(s) in (11.9) and using (11.8), we obtain F(&) > SY(Y,&)
= gy(Y,Y)^s+Sy(Y,V)
= £.
Hence, L(a)=
f F(d)ds> Jo
f ^-ds = t(l)-t(0)=r Jo as
= L(c).
174
Geodesies and Exponential
Map
This shows that if a is contained in expx[Bx(t+e)}: then its length is greater than or equal to that of c. Now we suppose that a is not completely contained in expx[Bx(r + e)]. Let 0 < s0 < 1 be the smallest number such that a(s0) G expx[Sx(r)]. Then by the above argument, the length of cr|[o,s„] is n ° t less than r. We obtain L(a) = L(a\[0iSo]) + Z-(a|[So,i]) > L(a\[0iSo]) >r = L(c). Thus c has minimal length among all piecewise C 1 curves joining x to z. Now it becomes obvious that (11.7) holds for 0 < t < r. Q.E.D. Corollary 11.3.2 Let (M,F) be a Finsler space and c{t), a < t < b be a geodesic. Then for any t0 (= (a, b), there is a small interval [t0 — e, t0 + e\ C (a, b) on which c has the smallest length among all curves issuing from c(t0 - e) to c{t0 + E). Proof. Take a sufficiently small e > 0 such that exp^. is a diffeomorphism from Bx(r) C TXM onto B(x, r) C M for x — c(t0 - e) and r > 2e. This is guaranteed by Corollary 11.1.2. Note that for y = c(t0 — e), exp, (ty) = c{t0 -e + t),
Vt G [0,2e].
By Proposition 11.3.1, we have d(c{t0 - e), c(t0 -
e
+ tj)=
tF(y),
Vi e [0,2e].
Thus c is minimizing on [t0 — e, t0 + e}.
Q.E.D.
Now we study distance functions in a Finsler space (M,F). the distance function from x £ M, dx(z) := d(x,z),
Consider
z e M.
According to Theorem 11.1.1, exp x is C°° on TXM — {x} and C 1 at y = 0 G TXM. By the local Minimality of geodesies (Proposition 11.3.1), we see that for z nearby x, dx(z)
=Fx(exp-l(z)).
Thus dx is C°° nearby x and C 1 at x. A natural question arises: Is dx always C°° at xl This question is answered in the following
Completeness
of Finsler Spaces
175
Proposition 11.3.3 Let (M,F) be a Finsler space. Suppose that d\ is C2 at x, then Fx is Euclidean at x. Thus dx is C2 for all x G M if and only if F is Riemannian. Proof. Fix an arbitrary basis {b;}™=1 for TXM. Let
—dlotp-^u).
Note that h(u) > h(Q) = 0. By assumption, h is C2 at 0. Thus
M«) = | ^ 7 ( 0 ) « V + o ( N » ) . On the other hand, for u = u*b; G TXM, <£>_1(fat) = c(t) is a C2 curve with c(0) = u. d(x,c(t)) = Jhitu)
= t\l--^-(0)^
ui + o{t).
By (1.17), we obtain ^ . ,
,.
r^o+
d{x,y-l(tu))
t
,.
•/i(rtt)
r->o+
Thus F x is Euclidean at a;.
11.4
r
/l
d 2 /i
,„, . .
V 2 axlox:> Q.E.D.
Completeness of Finsler Spaces
A Finsler metric F on a manifold M is said to be positively (resp. negatively) complete if every geodesic defined on [0, a) (resp. (—a, 0]) can be extended to a geodesic denned on [0, oo) (resp. (-co, 0]). Thus F is positively complete if and only if exp is defined on the whole TM. F is positively complete if and only if F(y) := F(—y) is negatively complete. F is said to be complete if and only if F is both positively and negatively complete. A sequence of points {xi\ in a metric space {M,d) is called a forward Cauchy sequence if for any e > 0, there exists a number N such that for
Geodesies and Exponential
176
Map
any j > i > N, d(xi,Xj)
< e.
A sequence {xi} is said to be forward convergent (resp. backward complete) if there is a point x G M such that lim d(xi,x) = 0
(resp. lim d(x,Xi) = 0 ) .
d is said to be positively complete (resp. negatively compete) if any forward Cauchy sequence (resp. backward Cauchy sequence) is forward convergent (resp. backward complete).
Lemma 11.4.1 Let (M,F) be a connected Finsler space. Suppose that F is positively complete, then for any pair of points, x,z £ M, there exists a globally minimizing geodesic from x to z. Proof. Let d = dp be the metric induced by F. By Proposition 11.3.1, we know that for sufficiently small r > 0, expx[Bx{r)} = B{x, r),
e x p J S ^ r ) ] = S(x, r).
(11.10)
Moreover, all geodesies from a; to a point in Sx(r) are minimizing. Let r0 = d(x, z). If z £ B(x, r), then we are done by the above fact. So we suppose that z £ B(x, r). There is a point m G S(x, r) such that r + d(m, z) = d(x, z) = r0. Write m = expx(ry)
for some y G SXM.
Completeness
of Finsler Spaces
177
We claim that the geodesic c(t) := expx(ty), 0 < t < r0, is a minimizing geodesic. In this case, c(r0) — z since geodesies do not have branched points. Let / be the set of t G [0, ra] such that t + d(c(t),z) = r0. In order to prove that c is minimizing, it suffices to prove that / = [0,r o ]. For any 0 < t < r, r0
+ d{c{t), z)
+ d(c(t), c(r)) + d(c(r), z) = r + d{c{r), z) = ra.
This implies t + d(c(t),z) = r0. Thus [0, r] C / . It is easy to see that / is closed. We claim that / is also open. Fix an arbitrary value t0 S I with t0 > r. For any t sufficiently close to t0 with 0 < t < t0, r0
+ d{c(t), z)
+ d(c{t), c(t0)) + d{c(t0), z) < t0 + d(c(t0), z) = r0.
This implies that t + d(c(t),z) = rQ. Thus [t0 — £,t0] C i for a sufficiently small e > 0. Suppose that t0 < r0. Take a sufficiently small 5 > 0. There is a point m' G S(c(t0),S) such that d(c(t0),m')
+ d(m',z) =
d(c(t0),z).
Observe that r0
<
d(x,m') + d(m',z)
<
d(x,c(t0))+d{c(t0),m')
=
t0 + d(c(t0),z)
+ d(m',z)
(H-H)
= r0.
This implies that d(x,m')
= d(x,c(t0)) +
d(c(t0),m').
We conclude that m' = c(t0 + 6) since geodesies do not have branched points. Plugging m' — c(t0 + 6) into (11.11), we obtain t0 + 6 + d(c(t0 + 5),z)
=r0.
178
Geodesies and Exponential
Therefore [t0,t0 + 6] C I. I=[0,ro].
Map
Since / is close and open, we conclude that Q.E.D.
Using Lemma 11.4.1, one can easily prove the following Hopf-Rinow theorem. Theorem 11.4.2 (Hopf-Rinow) Let (M,F) be a Finsler space and dp be the induced metric. The following two conditions are equivalent. (a) The metric space (M, dp) is positively complete, (b) The Finsler space (M, F) is positively complete. The proof is omitted here. See [Mai] [BCSl].
Chapter 12
Conjugate Radius and Injectivity Radius
For a positively complete Finsler space (M, F), the exponential map at any point x e M, expj.
:TXM^-M,
is an onto map. In this chapter, we discuss the singularity of exp x and the largest radius r for which exp x is a diffeomorphism on Bx(r) C TXM. Throughout this chapter, we always assume that Finsler spaces are positively complete. 12.1
Conjugate Radius
To study the singularity of the exponential map, we introduce the notion of conjugate value for unit tangent vectors. Definition 12.1.1 Let (M,F) be a positively complete Finsler space. For a unit vector y 6 SXM, we define cy > 0 to be the first number r > 0 such that there is a Jacobi field J(t) along c(t) = expx(ty), 0 < t < r, satisfying J ( 0 ) = 0 = J(r). cy is called the conjugate value ofy. cx :=
inf
c„,
yeS*M
"
Put cM := inf cx. xeM
cx and CM are called the conjugate radius of x and the conjugate radius of M respectively. 179
180
Conjugate Radius and Injectivity
Radius
We will show that the function c : y G SM —>• c y £ (0, oo] is lowercontinuous. Let yi G SM be a sequence of unit vectors converging to y G S X M. Assume that r :— liminf^oo cyi < oo. Then cyi < oo for large i. Since exp x is singular at c yi y; G T X M, exp x must be singular at ry by the continuity of d(expx) on TXM. This implies that c^ < r. Namely, cv < liminf cy.. Thus c y is lower-continuous in y G SM. To study the conjugate value of a tangent vector, we introduce the notion of index form. Let c(t), 0 < t < r, be a unit speed geodesic. For vector fields U = U(t) and V = V(t) along c, define XC(U,V):=J
{gc(Dctf,DcV0-g6(Rc(t0,f)}dt.
(12.1)
Xc is called the index form along c. Since R y is self-adjoint with respect to g y , Xc is symmetric, i.e., Ic(U,V)=Xc{V,U). Lemma 12.1.2 (Index Lemma) Let c(i),0 < t < r, be a geodesic with c(0) = y G TXM and J(t) be a Jacobi field along c with J(0) = 0. Suppose that 0 < r < cy. For any piecewise C°° vector field V(t),0
V(r) = J(r),
the following holds XC{J,J)
Jv(r),
where Jv denotes the Jacobi field along c with Jv(0) = 0,
D6Jv(0) = v.
$ is a linear map. By assumption, $(v) ^ 0 for w G T x M \ { 0 } . We conclude that $ is an isomorphism. Let {b;}™=1 be a basis for TciQ}M. Since $ is
Conjugate
Radius
181
non-singular, we can extend {b,}™=1 to a Jacobi frame along c, Ji(t)'•= Jbi(t),
i=l---n,
such that J(t) — Yl?=iaiJi(t)Since V(0) = 0, there are piecewise C 1 functions fi(t) on [0,r] such that n
Since J(r) = V(r), we get that a* = /j(r) and n
J(t) = J2fi(r)Mt)i=l
Let
A(t):=Y,f!(t)Mt). i=l
Integration by parts yields I C (V, F) = I c ( J , J) + f Jo
&6{A,A)dt.
Thus Xc(V,V)>Ic{J,J). Equality holds if and only if A(i) = 0, i.e., / f (t) = constant, This implies that J(£) = V(i).
i = 1, • • •, n, Q.E.D.
Lemma 12.1.3 Let 0 < r < cy and c(t) = expx(ty), 0 < t < r. For any piecewise C°° vector field V(t) =£• 0 along c with V{Q) = 0 = V(r), Ic(V,V)>0 and the equality holds if and only if r = cy and V is a Jacobi field along c.
Conjugate Radius and Injectivity
182
Radius
Proof. First we assume that r < cy. The Jacobi field J(t) along c with J(0) = 0 = J{r) must vanish. By Lemma 12.1.2, we know that Zc(V, V) > 1C(J, J) = 2 C (0,0) = 0. Now we assume r = cy. Let r» := cy — \/i and Vi(t), 0 < t < ri, be a sequence of piecewise C°° vector fields along Ci(t) = expx(ty),0 < t < ri such that Vj(0) = 0 = Vi{r{) and limj^oo Vi = V. By the above argument, ICi{VuVi)>Q. Letting i —> oo yields
Let Z7(t) be an arbitrary piecewise C°° vector field along c with U(Q) = 0 = J7(r). For any A G R, IC(V + XU,V + XU) = XC(V, V) + 2X1C(U, V) + X2XC(U, U) > 0. This implies XC(U, V) = 0. Take a partition 0 = to < t\ < • • • < tk-i < tk = r such that V(t) is C°° on each [ti_i,tj], i = 1, • • • ,k. Observe that for any piecewise C°° vector field U along c with the same regularity as V,
lc(U,V) = J2sc(u,Div)
' -/
UJu,
DtDdV+R6(V))}dt.
(12.2)
First we choose an arbitrary U(t) with U(ti) = 0, i = 0, • • •, k. We obtain
J {ge(V, D6DiV + R6(V)}}dt = 0. We conclude that V(t) is a Jacobi field on each [tj_i, ^ ] . Then for piecewise C°° vector field C/(i) along c with t/(0) = 0 = U(r), (12.2) simplifies to fc-i
1C(U,V) =
J2^(U,BCV) i=o
t
- -
k
= E Bc(«0 ( ^ i ) - D c^ t ") - DfiV(t+)) = 0. i=l
Injectivity
Radiiis
183
Since U(ti) are arbitrary, we obtain D6V(t-)
=
DtV(t+).
That is, V(t) is C 1 at £j, i = 1, • • •,fc— 1. Since V is a Jacobi field on each [ti_i,ti], we conclude that V is C°° along the whole c. Q.E.D.
12.2
Injectivity Radius
According to Proposition 11.3.1, for sufficiently small r > 0, the exponential map exp x is a diffeomorphism from the tangent r-ball Bx(r) c TXM onto the metric r-ball B(x,r) C M. In this section, we will discuss the largest possible domain around the origin in TXM on which the exponential map exp;,. is a diffeomorphism. Definition 12.2.1 Let (M,F) be a positively complete Finsler space and x G M. For a vector y € SXM, we define iy to be the supremum of r > 0 •such that c(t) := expx(ty) is minimizing on [0,r]. Set ix :=
inf yeSxM
i„,
i M := inf ix.
u
xeM
We call ix the injectivity radius at x and 'IM the injectivity radius of M. Let
Vx := J expx(ty),
0 < t < i„, y e S X M| C M
and CX:=M
- Vx
Cx and T>x are called the cut-locus and the cut-domain of x respectivelyLet Vx := [ty,
0
ye S X M } C TXM.
T>x is called the tangent cut-domain at x. The exponential map expx is an onto diffeomorphism.
-.Vx-^Vx
184
Conjugate Radius and Injectivity
Radius
The function i : SXM ->• [0, oo] is continuous. Thus for any compact subset A C M, \A •= inf ix > 0. xeA Moreover, the cut-locus Cx has null measure. See Section 8.4 in [BCSl] for the details.
Lemma 12.2.2 Let c : [0, oo) —> (M,F) be a unit speed geodesic with c(0) = y € SXM. If there is another geodesic a issuing from x to c(r) with L(a) < r, then \y < r. Proof. Suppose that for some r < \y. Let a = length(cr). a must be minimizing with the length of r. Since a does not coincide with c, we have v := &(a) ^ u := c{r). By Lemma 1.2.3 Sv(v,u)
= l.
(12.3)
Take a small 0 < e < iy — r, and consider the following distance function, p(z) := d(a(a-
e),z),
z G M.
The first variation formula gives d p(c(t)) dtV
= gv{v,U) < 1.
Injectivity
Radius
185
Thus, for sufficiently small e > 0 and e' > 0, there is a minimizing geodesic r from a(a — e) to c(r + e') with length < e + c'. This implies d(x, c(r + e')) < d(x, a (a - e)) + length(r) < r + e'. This is a contradiction. Lemma 12.2.3
Q.E.D.
At each point x G M, iy < cy for all y € SXM.
Hence
Proof. Let c(t) := expx(ty), 0 < t < cy + e. There is a nonzero Jacobi field J{t), 0 < t < cy, such that J(0) = 0 = J(cy). Fix an arbitrary small e > 0. Extend J to a piecewise C°° vector field V{t), 0 < t < cy + e, by assigning V(t) — 0 for cy < t < cy + e. Choosing a smaller e > 0 if necessary, we know that there is a unique Jacobi field J£(t), cy — e < t < cy + e with the prescribed values Je(cy - e) = V(cy - e) = J(cy - e), Js{cy + e) = ^(Cy + e) = 0 . Define a piecewise C°° vector field U(i), 0 < t < cy + e as follows.
nm = / v{t) l j
'
\
= J{t)
J e (t)
if
o < i < cy - £
if C y - e < t < c y + e .
Let 7 := c|[0,Cl/_E] and T := c| [CB _ £ , Cy+£] . Note that 17(U,U)=11(V,V). In view of Lemma 12.1.3, 1T(U,U) =lT(JeiJc)
<1T(V,V).
186
Conjugate Radius and Injectivity
Radius
We obtain IC(U, U) = l^U, U) + 1T{U, U) < Zy(V, V) + 1T(V, V) = 0. Let U^{t) :=U(t) ~ zm(c{t),U{t))
c(t).
Then fV
J C ( [ / \ C ^ ) =1(U,U)~
'
[sc(t)(c(t),U(t)y2 dt < 0 .
We can construct a piecewise C°° variation H(s,t) variation H(s,0) = c(0),
of c such that the
H(s, cy + e) = c(cy + e)
and U(t) =
—(0,t).
Consider the length function of cs(t) := H(s, t), 0 < t < cy + e,
L(s) := J " F(c,(t)yt. By the second variation formula, L"(0)=1C(U±,U-L)
<0.
Thus for small s, d(x, c(cy + £•)) < L(cs) < L(c) = cy + e. Thus there is another minimizing geodesic a issuing from x = c(0) to c(c y + e) with L(a) < L(c) = cy + e. By Lemma 12.2.2, we conclude that ly
+ £.
Since e is arbitrary, we conclude that ij, < cy.
Q.E.D.
Injectivity
Radius
187
Lemma 12.2.4 Let c : [0, oo) —» M be a unit speed geodesic with c(0) = y G SXM. Suppose that \y < cy. Then there is another minimizing geodesic a issuing from x to c(i y ). Proof. There is a sequence of minimizing unit speed geodesies Oi issuing from x to c(iy + £;) with limj^ooEi = 0. Let yi := (7,(0) G SXM. Without loss of generality, we may assume that y, —>• y* and ri := d(x,c(iy +£»)) -> \y. We claim that y* ^ y. Observe that expx(riVi)
-»
expx(iyy*).
Since dexpx is nonsingular at iyy, we see that y* ^ y. Let a : [0, iy] —> M denote the geodesic with
(12.4)
188
Conjugate Radius and Injectivity
Radius
By the Schwartz inequality, - g „ ( u , v) < F(u)F(v)
= 1,
-g„(u, v) < F(u)F(v)
= 1.
Thus Su(u,w)
=
Su(u,
Sv(v,w)
=
~gv{v,v)
u) - gu(u,v) -gv(u,
= - 1 - Su(u,v) < 0
v) = - 1 -gv(v,u)
< 0.
Take a smooth curve 7 : (—£,£) —> VW with c(0) = w G T Z M. Since exp x is a local difFeomorphism on B Cl («) C T X M and expjixy) = ^ = exp x (i x y*), there two curves 7,7* : (—£,£) —>• BCa.(x) satisfying exp x [7(s)] = expx[7*(s)] = c(s) with 7(0) = ixy and 7*(0) = ixy*. Consider the following variations cs and c*s, ca(t) = e x p j ^ s ) ] ,
<7s(t) = e x p j r y * ^ ) ] .
Geodesic Loops and Closed Geodesies
Let L(s) := L{cs) and L*(s) := L(as). -^(0)=gu(u,w)<0, as
189
By the first variation formula, as
^-(0)=Sv{v,w)<0.
Thus both curves cs and as are shorter than c and
Geodesic Loops and Closed Geodesies
In this section, we show that geodesic loops exist in a Finsler space (M, F) with non-trivial fundamental group. In topology, the fundamental group 7Ti(M, a;) is the space of homotopy equivalent classes of continuous loops at x. Each class a £ ni(M,x) contains at least one piecewise C°° loop at x. For a class a e ni(M,x), the geometric length \a\ is defined by \a\ = inf L(e), where the infimum is taken over all piecewise C°° loops c a t i . A piecewise C°° loop c G a is said to be shortest if L(c) = \a\. By the same argument as in §5.1, one can show that if c is a shortest piecewise C°° curve and parametrized by arc-length, then it must be a C°° geodesic loop at x. By Theorem 11.1.1, we know that exp x : Bix(a;) C TXM -» B(x,ix) C M is a diffeomorphism which is C 1 at the origin and C°° away from the origin. Thus for any non-trivial a £ wi(M,x), \a\ > 2i x . Let 7Ti (M) denote the space of free homotopy equivalent classes of closed curves in M. For a class a € 7Ti(M), if a piecewise C°° closed curve c & a
Conjugate Radius and Injectivity
190
Radius
satisfies L(c) — \a\, then it must be closed C°° geodesic in a. Thus |or| >
2'IM-
We have the following existence theorem for shortest geodesies. Proposition 12.3.1
(Hilbert) Let (M,F)
be a Finsler space.
(i) Suppose that (M, F) is positively complete. Then every non trivial class a S 7Ti(M, x) contains a shortest geodesic loop at x; (ii) Suppose that M is compact without boundary. Then every nontrivial class a S 7Ti (M) contains a shortest closed geodesic. Proof. We shall only prove (i). The proof of (ii) is similar. Let Cj 6 a be a sequence of piecewise C°° loops such that |a| = lim
L(a).
i—>oo
Clearly, C; is contained in a compact subset K in M. Since K is compact, \K > 0. For any pair x\,x-i G K with d(x1,x2) < \K, there is exactly one minimizing geodesic segment c joining X\ and x-i- Thus there is an integer N > 0 such that for each i, there are at most iV points xf = Ci on ct with
d{xlx^)
s = l,---,N,
where cf denotes the portion of ct from x? to xf+1. Let
By taking a subsequence if necessary, one may assume that Ci consists of exactly N minimizing geodesic segments cf, s = 1, • • •, N, for all i For a fixed 1 < s < N, the endpoints xf,x*+1 sub-converge to xs,xs+1. Let as s s+1 denote the minimizing geodesic segment from x to x . We must have
lim L(ci) = L(as).
Geodesic Loops and Closed Geodesies
191
Let a := U^=1as. a is a piecewise C°° geodesic loop at x and N
L(a) = J2 lim L(o-l) = Urn L f o ) = \a\. s=l
Thus
Q.E.D.
Lemma 12.3.2 Let (M, F) be a closed reversible Finsler space. Suppose that \M < CM- Then there is a closed geodesic c with L(c) = 2\MProof. Assume that \M < CM- Since M is compact, there is a point x G M such that i x = IM- It follows from Lemma 12.2.5 that there is a geodesic loop at x, c : [0, 2ix] - • M,
c(0)=c(2i x ).
Let z = c(ix) G Cx. It is obvious that i 2 = i x = i ^ . By Lemma 12.2.5 again, c(0) + c(2i x ) = 0. Thus c is a closed geodesic.
Q.E.D.
There are many Finsler spaces satisfying the assumption in Lemma 12.3.2. We will show the following facts in Chapter 12.3. (i) if K < 0, then c ^ = oo. Hence IM < CM — °o; (ii) if K > 1 and Diam(M) < 7r, then \M < CMWe know that for a generic Finsler metric on a closed manifold, the initial vectors to closed geodesies are dense in the unit tangent bundle. Finsler metrics with finitely many closed geodesies are rather exceptional. The first interesting work was done by Katok [Kat] who found some nonRiemannian Finsler metrics on the n-sphere S" with only finitely many closed geodesies. Later on, W. Ziller [Zi] made a close examination on Katok's examples. He actually generalized the construction of Katok and found some other interesting examples. In particular, he proves that there are Finsler metrics on S" and S"~x with only n closed geodesies, where n is even. These Finsler metrics exist in any neighborhood of the standard metric.
192
Conjugate Radius and Injectivity
Radius
Katok's construction also shows that there is a big difference between non-reversible Finsler metrics and reversible ones. For example, on S n , one can construct non-reversible Finsler metrics (close to the standard metric) with only n closed geodesies of lengths close to 2ir and the length of all other closed geodesies larger than any prescribed number. (Note: any Finsler metric on S n sufficiently C2 close to the standard metric has at least n closed geodesic [Zi]). However, for reversible Finsler metrics (including Riemannian metric s), Lusternik-Schnirelmann theory implies that for reversible Finsler metrics close to the standard metric, there are at least g(n) = 2n — s — 1 closed geodesies with lengths close to 2w, where n = 2fc + s < 2 f c + 1 . Notice that ( 3 n - l ) / 2
Chapter 13
Basic Comparison Theorems
In this chapter, we will discuss the geometric meanings of the Riemann curvature and its mean — the Ricci curvature. In particular, we will prove that the exponential map is nonsingular if the flag curvature is non-positive (Cartan-Hadamard Theorem) and the diameter must be bounded from above if the Ricci curvature is strictly positive (Bonnet-Myers Theorem).
13.1
Flag Curvature Bounded Above
Let (M,F) be a Finsler space. For a vector y € TXM \ {0}, the Riemann curvature Ry : TXM -» TXM is a self-adjoint linear transformation with respect to gy. For a tangent plane P C TXM with y G P, the flag curvature K(P, y) is given by gj,(Rj,(u),u) A K(P,y) = - ' Zy(y, v)Sy(u, u) - gy{y, u)gy{y, u)'
where u G P such that P — span{j/,u}. We say that the flag curvature is bounded from above by A, denoted by K < A, if for any flags (P,y),
K(P,y)<\. This is equivalent to the following Sy\Ky(u),uJ
< A|gj / (y,2/)g !/ (u,u) -
Similarly, we define the lower bound K > A. 193
Ey(u,y)gy{y,u)j.
194
Basic Comparison Theorems
For a number A G R, let
sx(t) := <
sin^tl
if A > 0
t sinh(v^It)
if A = 0 if A < 0,
-A
and c\(t) := s'x(t). Both s\(t) and C\(t) satisfy the following equation y" + Ay = 0. Let t^ > 0 be the first positive zero of s\(t).
t* := .
TT/VX,
More precisely,
if A > 0
. oo, if A < 0. We have the following comparison result in one-dimension. Lemma 13.1.1 Let r be a number with 0 < r < t\. f eCl[0,r] with f{Q) = 0 ,
For any function
[ {iff - V}* > [M.)' [ {«)» - W } * - /
0 < t < r.
The proof of Lemma 13.1.1 is elementary, so is omitted. Now we are ready to prove the following important theorem in Finsler geometry. Theorem 13.1.2 (Cartan-Hadamard [Au]) Let (M,F) be a positively complete Finsler space. Suppose that the flag curvature satisfies K < A. Then the conjugate radius satisfies cy > t\ for any y G SM. In particular, if K < 0, then 0 ^ = 0 0 for any y G SM. Hence exp x : TXM -t M is non-singular for any x G M. Proof. Fix a unit vector y G SXM. Let c{t) := expx{ty),
t>0.
Let J(t) be a non-trivial Jacobi field along c with J(0) = 0 and J(cy) = 0. By Lemma 11.2.1, we have J(t) = J^{t) + atc{t),
Flag Curvature Bounded
Above
195
where J"L(i) is a Jacobi field which is g^tj-orthogonal to c{t). We must have a = 0 and J(t) = J _L (t). Namely, J(t) is gc(t)-orthogonal to c(t). Let /(t) := y/snt){J(t),J(t)),
t > 0.
Observe that for 0 < t < cv, gc(Dc J, J )
•r\2
(/')
Assume that c^ < t\. Integration by parts gives gc(t)(D i J(t),J(t))
=
|
=
J
jt[gc(DcJ,j)]dT {gc (Da J, Dc j ) - gc ( R c ( J ) , J ) }dT
> /{(/') a -A/ 2 }dr
=
^ SA(t) ;w
/(t)2
This implies that for any 0 < t < cy, ± r f(t) ] _ fe(Dc J(t), (*)) dU SA(*) /(*)a
s^)
> 0.
SA(*)
Thus the quotient \ / g c ( t ) (•/(*), •/(*)) ^ SA(i)
/(*) SA(t)
is non-decreasing. Since J(0) = 0, we can express J{t) = tW(t),
0
where W(t) is a C°° vector field along c with W(0) = BtJ(0) implies
= v. This
196
Basic Comparison
Theorems
Therefore \Jsc(t)(J(t),J(t)) > 8X(t)yJev(v,vj> 0,
0
(13.1)
By assumption, cy < t A , we conclude that J{cy) ^ 0. This is a contradiction. Q.E.D. Let y £ SXM and c{t) — expx(ty), t > 0, be a unit speed geodesic. According to Lemma 11.2.1, for an arbitrary vector v € TXM, Jv(t) :— d(expx)\ty(tv) can be expressed in the form Jv{t) = Jv±{t) + where vL = v — gv(y,v)y. argument,
gy{y,v)tc{t),
Hence Jvx(t)±c(t)
3d(t)(j^(i),J^(i))
for all t > 0. By the above
>[sx(t)]2Sy(vL,vL),
0
Thus Sc(.t)(jv(t),Jv{t))
= g ( i ( t ) ( j ^ ( t ) , J ^ ( t ) ) + [g y (y,n)] 2 t 2
= M t ) ] 2 g „ M + [g,(2/,^)]2(t2 - Mt)] 2 ). Note that for w = d(expx)\ty{v)
£ TC^M,
d{expx)\ty{v)
:=
-Jv{t).
Thus Sd(t) (d(expx)\ty{v),
d{expx)\ty(v)j
> { ^ ) \ M H ^ V , v ) ? ^ ^ .
(13.2)
When K < A = 0, s\(t) = t. Inequality (13.2) becomes Zc(t)(d(expx)\ty{v),d(expx)\ty(v)}
> gy(v,v),
Thus expx is non-singular at any point ty € TXM.
v e TXM.
(13.3)
Flag Curvature Bounded
Above
197
Theorem 13.1.3 (Cartan-Hadamard) Let (M,F) be positively complete with K < 0. For any x G M, exp x : TXM —» M is a covering map. Moreover, for any y € SXM, (13.3). holds. By further argument, one can show that exp x in Theorem 13.1.3 is a covering map. See [BCSl] for a proof.
A Finsler space (M, F) is called a Hadamard space if it is positively complete, simply connected with K < 0. We have just proved that for a Hadamard space, the exponential map exp x : TXM —• M is non-singular for all x € M. In [Egl][Eg2], Egloff studied uniform Hadamard spaces and uniform negatively curved Finsler spaces. A Finsler space is said to be uniform with uniformity constant C if for any x € M and u, v € TXM, C~lgu <&v< CSu.
(13.4)
Hence C _ 1 F 2 ( u ) < g„(u,u) < CF2{u). We have the following Theorem 13.1.4 Let (M,F) be a complete reversible uniform Finsler space satisfying (13.4)- Assume that M is simply connected and F satisfies K < 0. Then exp x : (TXM,FX) —> M is distance quasi-nondecreasing. More precisely, Fx(V2-Vi)
< v / Cd(exp x (i/i),exp a! (ift ! )),
Vyi,y2 e TXM.
(13.5)
Proof. Let a : [0,1] —> M be a minimizing geodesic joining from exp x (j/i) to expx(?/2)- Since exp x is a covering map and M is simply connected, exp x is a diffeomorphism. There is a curve a : [0,1] —> TXM joining from y\ to 2/2, such that expx(a(t))=o-{t),
0
198
Basic Comparison
Theorems
expx(?/2)
K <0
exPz(yi) Observe that F(a(t))
=
F(d(expx)(a(t)))
Thus ?
x(V2 ~-yi)
<
[ Fx{d{t))dt Jo
< VC f F(a(t))dt Jo
This proves (13.5).
Q.E.D.
Speaking of Finsler spaces with negative curvature, we should mention Foulon's work on geodesic flows. In [Fol], Foulon proved that the geodesic flow of a compact reversible Finsler space with K < 0 is of Anosov type. See also [Egl] for further studies on the geodesic flow of reversible Finsler spaces.
13.2
Positive Flag Curvature
Now we discuss the geometric properties of even-dimensional Finsler spaces with positive curvature. As we know, the conjugate radius is always less
Positive Flag
Curvature
199
than or equal to the injectivity radius (Lemma 12.2.3). We will show that they are equal for even-dimensional oriented Finsler spaces with positive curvature. Theorem 13.2.1 (Klingenberg) Let (M, F) be an even-dimensional closed oriented Finsler space with K > 0. Then iu = CMProof. We assume that iM < CM- By Lemma 12.3.2, there is a closed geodesic c with L(c) = 2i M . Let c(t), 0 < t < 2\M be parametrized by arc-length. By Lemma 10.3.1, there is a parallel vector field W{t) along c and not tangent to c(t). Choose a variation H(s,t) of c(t) as in the proof of the Synge Theorem. The length function /•2i M
QH
satisfies L"(0) = - j
M
gc(Rc(W), W)dt < 0.
Thus for small s > 0, the curve cs(t) := H(s, t) is shorter than c, L{cs) < L(c). Let zs G cs such that d(cs(0),zs)
= supd(cs{0),zs)
There is a unique minimizing geodesic segment as : [O.rv,] -> M from cs(0) to za = cs(rs). It follows from the first variation formula that
Sy,{ys,cs(rs)J =0, where ys = &s(rs).
200
Basic Comparison
Theorems
c(ix)
We may assume that zSi converges to z — C(IM) for some sequence Si —> 0 + . Further, we can assume that lim^oo &Si(rSi) —>• v . Thus aSi converges to a minimizing geodesic segment a from c(0) to z. We still have &y0(yo,c{iMyj
= 0,
where y0 — &(r0). We obtain two minimizing geodesic from c(0) to z which do not form a smooth curve at z. This contradicts Lemma 12.2.5. Q.E.D. Positively complete Finsler spaces with nonnegative flag curvature have not been completely understood. For Riemannian spaces with nonnegative sectional curvature, we have the following Cheeger-Gromoll's soul theorem.
Theorem 13.2.2 ([ChGr2]) Let {M,g) be a complete Riemannian space with K > 0. There is a totally geodesic closed submanifold S C M such that M is diffeomorphic to the normal bundle of S in M.
The proof is based on the Toponogov comparison theorem [To] and quite long, so is omitted. See also [ChEb].
Ricci Curvature Bounded
13.3
Below
201
Ricci Curvature Bounded Below
Let (M, F) be an n-dimensional Finsler space. The Riemann curvature in a direction y G TXM is a linear transformation Hy : TXM —> TXM. The Ricci curvature is denned as the trace of the Riemann curvature. Let {bi}" = 1 be an arbitrary basis for TXM and express Ry(u) = R\{y)uk
u = ukhk G TXM.
bu
By definition, 71
mC(y) = j2Rii(y)i=l
Assume that {bj}™=1 is an orthonormal basis with respect to g y such that bn = y/F{y). Let Pi := spa.n{hi,y},
i = 1, • • •, n - 1.
Since Rj,(2/) = 0, we have K(P i ,i,) = - ^ g ] / ( R y ( b i ) ) b i ) . Thus ra-l
2
Ric(y) = F ( 2 / ) ^ K ( P i ) 2 / ) . By the above formula, we see that K > A implies Ric(j/) > (n - l)AP 2 (y),
0 / Vy G TM.
(13.6)
We say that Ric > (n - 1)A if (13.6) holds. Theorem 13.3.1 (Bonnet-Myers[Au]) Lei (M,F) be an n-dimensional positively complete Finsler space. Suppose that Ric > n — 1. T/ien for any unit vector y G SM, t/te conjugate value satisfies cy < n. Proof. Suppose that cy > ir for some y G SM. Let c(t) = exp;c(tj/), 0 < t < n. Take a parallel gc-orthonormal frame {jEi(t)}o
0 < t < n.
202
Basic Comparison
Theorems
By Lemma 12.1.3, we have n-l
Y,I*(Vi,Vi) > 0. i=l
Ric > n — 1 On the other hand, n-l
Y,MVi,Vi)
„.
= <
|(n-l)cos2(t)-Ric(c(t))sin2(t)]d*
/ (n-l)
I
{cos 2 (f)-sin 2 (i)}
This is a contradiction. Thus cy < TT for any j/ S SM.
Q.E.D.
Let ( M . F ) be as in Theorem 13.3.1. By Lemma 12.2.3, \y
ye
SXM.
Note that the diameter at x € M satisfies Diam z (M) := supd(x,z)
=
zeM
sup
iy.
yeS*M
Thus the diameter Diam x (M) of M around x must be bounded by 7r, i.e., Diam x (M) < TT. Since the induced Finsler metric on the universal cover TT : M —> M also satisfies the same Ricci curvature bound, we still have Diam £ (M) < TT, where TT(X) = x. Thus M is compact and the fundamental group TT\(M) is finite.
Green-Dazord
Theorem
203
Positively complete Finsler spaces with nonnegative Ricci curvature are not completely understood yet. One of the most important theorems is the Cheeger-Gromoll splitting theorem for Riemannian spaces. Theorem 13.3.2 ([ChGrl]) Let (M,g) be a complete Riemannian space with Ric > 0. Suppose that there is a line c : (—oo, oo) —> M in M, i.e.,
\t-t'\,
l(c(t),c(t'))
M'eR.
Then M is isometric to the product space J V x R . Many results on Riemannian spaces with nonnegative Ricci curvature are obtained using the Cheeger-Gromoll splitting theorem. Hence they are not true for Finsler spaces. 13.4
Green-Dazord Theorem
By Proposition 5.4.3, we know that the geodesic flow
The equality holds if and only if F has constant curvature K = ( ^ J . Proof. Let y E SXM. The curve c(t) :=
3
is a geodesic with c(0) = y. Take an arbitrary basis {b;}" = 1 with b„ = y. Let {Ei(t)} be a gc-orthonormal frame along c with En(t) = c(t). According to Lemma 12.1.2, the following vector fields Vi(t):=sm(-t\Ei(t), satisfy
2c(v<,v;)>o,
0
204
Basic Comparison
Theorems
and equality holds if and only if Vi(t) are Jacobi fields along c. Observe that n-l " —1
Integrating $ over SM yields /
JSM
^ ) ^ = ^V0l,(SM)-/rsin2(^)[/ 2r J0 V r / L JSM
^ f l d V ^ d t . n-l J
Since 4>t preserves the volume form dVg, we have /
Bic(4>t(v))dVg = /
Ric(l/)d^-
Thus (-)2Volg(SM)-
[
^ ^ d V ^ - l
*dV6>0.
(13.9)
We obtains (13.7). Assume that the equality in (13.7) holds. (13.9) implies that $(y) — 0 for any y £ SM. Thus for any set of vector fields Vi{t) = sm(-rrt/r)Ei(t) along any geodesic c(t) = expx(ty), 0 < t < r, where Ei(t) are parallel vector fields along c, 2 c ( ^ , K ) = 0,
t = l,---,n-l.
By Lemma 12.1.3, r = cy and each Vi{t) is a Jacobi field along c. That is, 0 = DiD6Vi(t) + Rc(Vi(t)) = { - ( ^ ) 2 E i ( i ) + R c ( ^ ( t ) ) } s i n 2 This implies R6{Ei(t))=(^>j
Et{t),
i=
!,-••,n-l.
In particular, at t = 0, Ry(bi) = bj,
i = !,-••, n - l .
(^).
Green-Dazord
Theorem
205
Since y is arbitrary and bj's are arbitrary, we conclude that F has constant curvature K = (f)
.
Q.E.D.
Chapter 14
Geometry of Hypersurfaces
The purpose of this chapter is to introduce the notions of Hessian and Laplacian on functions. For the level hypersurfaces of a distance function, we the relationship between the Hessian of the distance function and the normal curvature of its level hypersurfaces, that between the Laplacian of the distance and the mean curvature of its level hypersurfaces.
14.1
Hessian and Laplacian
Let (M, F) be a Finsler space and / a C 2 function on M. The Hessian of / is a map D 2 / : TM ->• R defined by
DV(.):=|;(/°c)[=o,
ytTxM,
where c : (—e,e) —> M is the geodesic with c(0) = y G TXM. coordinates,
d2f df J i ( x ) y V -2-^-(x)G*(y). dx dx^ dx From the definition, we have D2/(A2/) = A 2 D 2 /( 2/ ),
A>0.
In general, D 2 /(y) is not quadratic in y £ TXM. 207
In local
(14.1)
Geometry of
208
Hypersurfaces
The Hessian of a C°° distance function p on an open subset U C M has special geometric meanings. First, the gradient Vp is a unit vector field on U. It induces a C°° Riemannian metric on U, 2/ G ™ -
Hy) := y/svP(v,v), By Lemma 3.2.1,
F ( V p ) = l = F(Vp). Thus Vp = Vp is also the gradient of p with respect to F. However, the Hessians D 2 p and D 2 p of p with respect to F and F are not equal. Lemma 14.1.1 Let (M,F) tion p on an open subset U,
be a Finsler space. For a C°° distance func-
B2p(y) = t)2p(y) - Tvp(y),
y € TXU,
(14.2)
where D 2 p denotes the Hessian of p with respect to F = ,/gvp. Proof. Let r*fc (a;) denote the Christoffel symbols of the Levi-Civita connection D. By (14.1), the Hessian of D 2 p is given by t)2p{y) = ®? K.(x)yiyj v ' dx*dxi ' where
-
2^r-{x)Gi(y), dx'K ' K '
2&(y):=rjk(x)yiyk
D2p(y) - D2p(y) = 2{Gi(y) - <*!%)} J ^ * ) . dx
By definition (3.15) and (10.1), we have T Vp (2/)
=
g v p ( D y r - L \ y , Vp)
=
dp(pvV-Dvv)
= =
2{G\y)-G\y)}^-{x) t)2p(y) - D 2 p(y),
where Y is an arbitrary extension of y.
Q.E.D.
Now we introduce the Laplacian on a Finsler m space. Let (M, F, dp) be a Finsler m space. For a Ck (k > 1) vector field X on (M,
Hessian and Laplacian
209
divergence div(X) is a Ck x function on M. On the other hand, for a Ch (k > 2) function / on (M,F), the gradient V / is Ck~x on U := {df ^ 0} and C° at M - U. Therefore, div(V/) is a C fc_2 function on U. Define A / on U by A / := div(V/). Let d/i = a ^ d a : 1 • • dxn and V / = VV^frpressed by
(14.3) The Laplacian of / is ex-
For any 0 e C£°(W), div(0V/) = 0 A / + # ( V / ) . Applying the divergence formula (2.6) to the above identity yields /
cf> A / d M = - /
#(V/)dM,
(14.4)
This leads to the definition of A / on the whole M in the distribution sense. / (AfiWn
:= - f dJ>(Vf)dn,
JM
V
(14.5)
JM
We call A the Laplacian on functions. Compare [Sh4] and [BePa]. Assume that df / 0 on an open subset U C M. The gradient of / is given by W = ^(d/)£-,
(14.6)
where A%(j}) are given in a standard local coordinate system (xl, %) in T*M by
The Laplacian is locally expressed by
1
d (/ ,^. .^, .(, d„ ,/ . *„ ,) ^3 )/ .
(14.7)
Geometry of
210
Hypersurfaces
Thus the Laplacian is a fully non-linear elliptic operator. The vtiriational problem of the canonical energy functional also gives rise to the Laplacian. Let H 1 denote the Hilbert space of all L2 functions on M such that df £ L2. Denote by H* the space of functions a G H 1 with JM ud\i = 0, if dM = 0 and with u\gM = 0 if dM ^ 0. The canonical energy functional £ on H* is defined by fM[F*(du)]2dn
S(u)
U 2
IM
^
where F* denotes the dual Finsler metric, namely, F*(£) := s u p ^ ) ^ £(M). For functions u, tp £ Hj,
Thus, for any u £ H* with JM u2dfi = 1, du£(
/ u
V
(14.8)
where A = £(u). From (14.5) and (14.8) it follows that a function « e H j satisfies d u £ = 0 with A = £{u) if and only if AM
+
AM
= 0.
In this case, A and u are called an eigenvalue and an eigenfunction of (M,F,dn), respectively. It is shown in [GeSh] that any eigenfunction / is C 1 -" for some 0 < OL
MM):=
M
. /»'7'f "•* .
infA e R J M |M - Aj2d^t We claim that \\ = Ai(M) is the smallest eigenvalue of (M, F,d^). UGC~(M)
(.49) ;
Hessian and
Laplacian
211
For any function u £ C°°(M), / JM
\u- X\2dp = / u2dp-2X JM
udp + X2 /
dp.
JM
JM
Thus inf /
\u-X\2dp
= /
XeR
JM
u - p(M)
1
u\ dp
JM
Since C°°{M) is dense in H*, the above identity implies .nf fM[F*(du)]2dp uec°°(M) i n f A 6 R / M | u - A | 2 ( i ^
.nf «eHj,(M)
JM[F*(du)}2dp JMu2dp
Thus AX(M) is the minimum of the energy functional E, that is, the smallest eigenvalue of M. Lemma 14.1.2 Let (M, F, dp) be a Finsler m space and p a 0°° distance function on an open subset. Then Ap = A p - S ( V p ) ,
(14.10)
where A denotes the Laplacian of g := gvpProof: Let dp, and dVg denote the volume forms of F and g, respectively. In a local coordinate system (xl), write dp = a{x)dxl A • • • A dxn,
dVg = &(x)dxl A • • • A dxn.
Let Vp = VVgfr- By Lemma 3.2.1, we know that Vp is also the gradient of p with respect to F and g, i.e., Vp = Vp. For the Riemannian metric g,
a(x) =
SJdet(9x(£-i\x,£-\x)).
By (7.25) and (7.29), the S-curvature of F is given by
S(Vp) = Vpfln' 71 Equation (14.3) gives
1 d /
•\
SVV
*>">£{>*>)-&+H**)-
Geometry of
212
Hypersurfaces
Thus Ap
= -
^ + Vp(ln.)
0V> + Vp(lna)
- Vp(ln-)
da:* Ap-S(Vp).
=
Q.E.D. The relation between the Hessian and the Laplacian becomes simpler when the Finsler metric is Riemannian. Suppose that F is Riemannian, i.e., F(y) = y/gij(x)yiy^. Note that gij, g*li — gl\ r*-, etc. are functions of a; only. By (14.1) and (14.7), we have
where a := y/det (gij). Note d
1
dxi\
d
/\„_\ „ki 9ki (lna)=i ffa
J
2
dxi
Thus 5
9jk-
(i„ ffH , a fi„J\5a;' 9a>?
^
„ki^9jk , ludg i kidgki „kid9jk ki_
nkkl,
„ r tt
,
.
Plugging (14.13) into (14.12) yields
A/
= ^{^-^}-
(14 14
-^
Let {bij-JLj be an orthonormal basis for TXM. (14.14) can be expressed by n
A/ = ^D2/(bs).
(14.15)
i=l
In the above sense, the Laplacian A / is the mean value of the Hessian D 2 / .
Normal
14.2
Curvature
213
Normal Curvature
The purpose of this section is to introduce the notion of normal curvature for submanifolds in a Finsler space. We will use it to study the level hypersurfaces of a distance function on a Finsler space. Let (M, F) be a Finsler space and ip : N -» M be a submanifold. Let F := ip*F denote the induced Finsler metric on N. For simplicity, we denote d
AG/) := -D fi c(0). From the definition, A(y) has the following homogeneous property A(Xy) = X2A(y),
A > 0.
A(y) is called the normal curvature of iV in the direction of y G T^iV. Let (x a ) be a local coordinate system in N and (a;x) a local coordinate system in M. Let y = c(0) = dip(y), where ?/ = c(0). We have
Dtc(0) = {^l ( 0) + 2 G - ( t ,)}A|,
Then the normal curvature A(i/) in the direction y = d
Assume that JV is a hypersurface in M. Let n denote a normal vector ol N at x € N. Define (n,A(2/)),
2/GT.AT.
hn(y) is called the normal curvature of iV in the direction y with respect to n. There are exactly two normal vectors n, n' G TXM. In general, n' ^ —n and A n =£ — A n '.
214
Geometry of
Hypersurfaces
Let n = n%-£^\x € TXM. A„(y) is expressed by
An(») := - n ^ ( n ) { J ^ ( z ) y V - 2^G"(y) + 2G*(y)}. (14.16)
Note that A(y) is not orthogonal to N with respect to g n . Thus A„(y) does not capture all the data of A(j/). Example 14.2.1 Consider the indicatrix S = F _ 1 ( l ) in a Minkowski space (V,F). Let
g„(y,w)=0}cV.
Define g and C foy ff(u,v) := gy(w,v),
5(C(u,«),«;) := Cj,(u,t>,ti;),
(14.17)
where u,v,w S T y S = W y . Fia; an arbitrary basis {bj}™=1 /or V. TTie vector-valued function tp = iplhi satisfies the following Varga equation [Va] i>2
where M£b are some local functions onS. It is easy to see that n — ip, since F2(ip) = 1. Namely,
Normal
Curvature
215
Observe that
An{y) = - n ^,. (n ){j!|l,( S ),V- 2 |g^(,)} = -n^(n){ - M< 6 g - 9ab^ - *%;G"{y)} =
-T^9ij{^)
=
9(2/, y)-
g{y,y)
We obtain K{y)=g{y,y),
VyeTxS.
(14.18) tl
Let A'' be a level hypersurface of a C°° distance function p on an open subset U C M. By (3.15), dp(v) = gv P (Vp, w),
v G TM.
(14.19)
This implies that
gVp(V/),i/) = o,
« s m
Thus n := Vp|;v is a normal vector to N. Proposition 14.2.1 Let p be a C°° distance function on an open subset U C (M,F) and N := p~l(s) C U. With respect to the normal vector n := Vpx at x G N, the normal curvature A n satisfies D2p(y) = An(y),
Vy e TXN.
(14.20)
Geometry of
216
Hypersurfaces
Proof. Let ip : N —>• U be the natural embedding and F = ip*F the induced Finsler metric. For a vector y = dtp(y) £ TXN, where y € TXN, let c(t) be the geodesic in (N,F) with ^f (0) = y. Then the curve c(t) := ip o c(t) satisfies c(0) = y. Since p(c(t)) = constant, differentiating it twice with respect to t yields „ .n -(xWyy 3 y dxldxiy '
a = —^r(x) ; h(x)y,yayby + -^(x)^-(x)G (y), n dxlK 'dxadxbK dxlK JdxaK '
y
'
y (14.21)
'
where we have used the geodesic equation (5.5) of c. On the other hand, by the definition of D 2 p,
° 2p(2/) = d^h{x)ylyJ
~ 2Z{x)Gi{y)-
(14 22)
-
Plugging (14.21) into (14.22) yields BMV) = - ^ ( x ) { ^ (
2
) ^ - 2?£&>(y)
+
2CP(y)}.
(14.23)
In virtue of (14.16) and (14.23), we conclude that D2p(y) = dp(A(y))
= g V p ( v P , A(t/)) = An(y),
This proves (14.20).
y e TXN. Q.E.D.
Corollary 14.2.2 Let p be a C°° distance function on an open subset U C (M,F) and g := gvp- Let A n anrf A n denote the normal curvature of N := p~l(s) in (U,F) and (U,g) with respect to n := Vp x = V'px at x G N, respectively. Then A„fo) = A „ ( ! / ) - T n ( j / ) ,
VyeTxN
(14.24)
Proof. By Lemma 3.2.1, we know that p is also a distance function of g. Thus (14.20) holds for g, i.e., t)2p(y) = A„(y), Then (14.24) follows from (14.2).
Vy £ TXN.
(14.25) Q.E.D.
Let N c (M, F ) be an embedded hypersurface. N is said to be convex at a: € N if there is an open neighborhood W of i such that all geodesies in U, issuing from x and tangent to AT, lie on one side of N. In addition, if these geodesies do not intersect AT except for x, then N is said to be strictly
Normal
Curvature
217
convex at x. Note that the indicatrix of a Minkowski space is strictly convex. More general, we have the following Theorem 14.2.3 Let N C (M, F) be an embedded hypersurface and A n the normal curvature of N at x £ N with respect to a normal vector n. / / N is convex at x G N, then either A n > 0 or A n < 0. / / A n > 0 or An < 0, then N is strictly convex at x. Proof. Suppose that N is convex at x. Let U be an open neighborhood of x, in which all geodesies issuing from x and tangent to N at x, lie on one side of N. Choose a smaller U if necessary, we may assume that there is a C°° distance function p on U such that N n U = p~1(0). Further we may assume that all these geodesies lie in U+ := {x : p(x) > 0}.
Let y € TXN and c(t) be the geodesic in U with c(0) = y. By assumption, p o c(i) > 0 for all small t and p o c(0) = p(a;) = 0. Thus for n = Vp x
An(2/) = D2p(y) = ^ [ p o C ( i ) ] | t = o > o . This proves the first statement. Now we assume that A n > 6 > 0 for n = Vp x . For any normal geodesic c in M with c(0) = t / € T^iV, -^ [p ° c(t)] | t =o = D2p(y) = An(?/) > 6 > 0. Thus, there is a small number e > 0 such that P°c(t) > -6t2,
V t e (-£,£).
Geometry of
218
Hypersurfaces
Clearly, e can be taken independent of the unit vectors y GTXN. Therefore all geodesies in a small neighborhood U with c(0) = y G TXN lie on one side of N, which intersect N only at x. This proves the second statement. Q.E.D.
14.3
Mean Curvature
In this section, we are going to introduce the notion of mean curvature for hypersurfaces in a Finsler m space. Since every Finsler metric induces the Busemann-Hausdorff volume form, the mean curvature is of course defined for hypersurfaces in a Finsler space. Let (M, F, dp) be an n-dimensional Finsler m space. Express dp by dp = 4>^dVF,
(14.26)
where >M is a positive C°° function on M and dVp denotes the BusemannHausdorff volume form of F. Consider a hypersurface N C M. Let p be a C°° distance function on an open subset U C M such that p~1(s) = N n U for some s. Let dvt denote the volume form on JV( := P~l{t) induced by dp, and dv := dvs. Let c(t) be an integral curve of Vp with c(0) G Ns. We have p(c(t)) = t, hence c(s + e) € A^s+£ for small £ > 0. By definition, the flow <\>e of Vp satisfies
Ns:->Ns+£.
The pull-back (n — l)-form (^>e)*
V:r G TV,
(14.27)
Note 0(a;,O) = l,
VrrGA^.
(14.28)
Mean
Curvature
219
Set nn:=|-(lne(i,£)) OS \
.
(14.29)
I |e=0
n „ is called the mean curvature ol N at x with respect to n := Vp x . Proposition 14.3.1 Let p be a C°° distance function on an open subset U C (M, F). Let II n be the mean curvature of N := p~l(s) at x with respect to n := Vp x . Then Ap(x) = n n ,
Vz € N.
Proof. Take a special local coordinate system in M,
T'-^k + r-k with y 1 = 1 and Ya = 0 . Put dp, = er(£, x a )dt A dx2 A • • • A cten. The induced volume form dvt on Nt := p _ 1 (i) is given by dvt = a(t, xa)dx2 A • • • A dx71. Observe that <j>s • (s, xa) -> (s + £, z a ) so that {<j>e)*dvs+E = =
a{s + e,xa)dx2 ••• Adxn a(s + e,xa) L a 2 J, n ——, a —— K o(s,x )dx ---!\dx a(s,x ) '
(14.30)
Geometry of
220
=
Hypersurfaces
a(s + e, xa) , ; \—dv s . a(s, xa)
Thus e(x,e)
=
a(g + £
.
'af},
V* =
G N.
This gives Un = -f\]ne(x,e)] del
= ^-\\na(t,xa)} J |e=o at L
. J |t=s
(14.31)
On the other hand, by (14.3), we have Ap(x)^^[lna(t,xa)]^=s. Now (14.30) follows from (14.31) and (14.32).
(14.32) Q.E.D.
Corollary 14.3.2 Let p be a C°° distance function on an open subset U C (M,F) and g := gvp- Let Hn and Yln denote the mean curvature of N := p~l{s) in (U,F) and (K,g) with respect to n = VP\N, respectively. Then n n = fl n - S(n).
(14.33)
Proof. Let A denote the Laplacian of g. By Lemma 3.2.1, we know that p is also a distance function of g. Then Proposition 14.3.1 holds for g, that is, Ap = n n . By (14.10), (14.30) and (14.34), one obtains (14.33).
(14.34) Q.E.D.
We have defined the normal curvature and the mean curvatures. We are going to show that for hypersurfaces in a Berwald space with the BusemannHausdorff measure, the mean curvature is the mean value of the normal curvature. Proposition 14.3.3 Let N be a hypersurface in an n-dimensional Berwald space (M, F). Then for a normal vector n at x S N, n„ = ^ A n ( b i ) ,
(14.35)
Shape Operator
221
where {b;}™=1 is an orthonormal basis for TXN with respect to g n . Proof. There is a C°° distance function p in a neighborhood U of x, such that P _ 1 (0) =UC\N and n := Vp x is the normal vector at x G iV. Let D 2 p and Ap denote the Hessian and Laplacian of p with respect to g := gv p , respectively. By Propositions 14.2.1 and 14.3.1, we have An nn
= =
D 2 P, Ap.
Let c be the geodesic in M passing through x € N with c(0) = n. For small i>0, p o c(t) = t. Thus D 2 p(n) = ^ [ p o c ( t ) |t=o
0.
Take an orthonormal basis {bi}™=1 with b„ = n for (TxM,gn). from (14.15) that
It follows
n-l
n n = Ap(x) = ^
2
A
D p(bi) = Yl
i=l
n( b *)-
(14-36)
i=l
Since F is a Berwald metric, T = 0. By Proposition 7.3.1, S = 0. Then (14.35) follows from (14.24), (14.33) and (14.36). Q.E.D. 14.4
Shape Operator
In order to study the relationship between the normal curvature and the Riemann curvature, we introduce the notion of shape operator and derive the Riccati equation for the shape operators along a geodesic which is orthogonal to the level surfaces of a distance function. Let (M, F) be a Finsler space and N an embedded hypersurface in M. Let p be a C°° distance function in a neighborhood U of N such that p~~l{s) =UnN. Let D denote the Levi-Civita connection of g := gv P - For the normal vector n := Vp|jv, g(n, n) = 1. Thus 0 = w g(n, n) = Tg ( f ^ n , n ) ,
Vw£ TxN.
222
Geometry of
Hypersurfaces
This implies that D^n G TXN,
Vu> G TXN.
Define the shape operator S : TXN -> T^AT by S(io) := L\,n,
w G TXAT.
(14.37)
We claim that the shape operator is self-adjoint with respect to g n , i.e., gn(S(u),v)
= g n (u, S(v)),
Vu, v G TXN.
(14.38)
To prove (14.38), take two vector fields U and V in a neighborhood of x in M such that Ux = u,Vx = v. U and V can be chosen so that they are tangent to N. Thus T>uV\x-t>vU\x
= [U,V]\x = 0.
Then (14.38) follows from the following observation, g„[S(u),vJ
=
g(l)un,v)
=
-9 ( n , D y V ) |
=
-g ( n , D v C / ) |
= s (u,D„nJ =
gn(w,S(f)J.
The normal curvature of N in (U, g) is determined by the shape operator. More precisely, Lemma 14.4.1 Let A denote the normal curvature of N in (U,g) with respect to n. Then A„(u>) = g„ (s(w),«;),
Vu; G TJV.
(14.39)
Proof. Let £ : (—e, e) —>• N be a geodesic curve in (U,g) with £(0) = w. Observe that An(u>)
=
D2p(w)
- s ; ['(««>)] i -
Shape
Operator
223
|s=0
s[»(v*«<")
a=0
g n (s(iu), tuj. This proves the lemma.
Q.E.D.
Let Nt := p _ 1 (i) be a family of hypersurfaces in U. Fix an integral curve c(t) of Vp with poc(t) — t, Vi € [a, 6]. For small t, c(t) is orthogonal to Nt at c(t) with respect to gc(t)-
Let S t : Tc(t)JVt -> Tc{t)Nt denote the shape operator of Nt at c(t). Define S t : Tc(f)7Vt -^ Tc(t)Nt by S t (f/) := D fi(t) [st(U)] - St ( D c ( t ) t / J ,
(14.40)
where U = U(t) is an arbitrary vector field along c. When U is parallel, ,1
S t (I/):=D« ( t ) [S t (E0j. We have the following Riccati equation,
224
Geometry of
Hypcrsurfaces
Lemma 14.4.2 S* + S? + Rc (t ) = 0.
(14.41)
Proof. Fix a number t0. Take a C°° curve £(s), — £ < s < e in Nto with £(0) = v € Tc(to)A^to. Take a variation of c, H : (—£,£) x [a, 6] —> M, such that each curve c s (t) = H(s,t) is a geodesic with c s (t 0 ) = f(s) € iVt0. Let V(M):=^-(M),
y(S,t):=^-(5,i).
Note that Y(s, t) = Vp\}j(s,t) is normal to Nt and J(t) := V(0, t) is a Jacobi field along c with J(t0) = v. Let D and D denote the connection of F and F = y/gVp, respectively. By (6.14), Dvp = DvpRestricting it to c yields D6J(t)=T)6J(t). Observe that DeJ(i 0 ) = t>tJ{t0) = I)vY(0,to)
= S4».
Thus %„)(«)
=
-DcD £ J(t 0 )
=
-Da[st(J(t))l L
=
=
- S t »
J |t=t„ - S to ( D 6 . 7 ( * 0 ) )
- S , » - S ? » .
This gives (14.41).
Q.E.D.
Consider a family of hypersurfaces Nt in a Finsler space (M, F) and a geodesic c(t) such that (1) c(t) € JVt; (2) c(t) is a normal vector to Nt.
Shape
Operator
225
Let J(t) be a Jacobi field along c such that J(t) and DcJ{t) are Sc(ty orthogonal to c(t). By the above arguments, we see that the shape operator is given by St(J(t))=D6J(t).
(14.42)
It follows from (14.39) and (14.42) that h(J(t)) = gm (p6J(t), J(t)).
(14.43)
Now we study the shape operators of the regular metric i-sphere S(p, t) in the cut-domain Vp,
S(p,t)=VpnS(p,t). Let p(x) := d{p, x). p is a C°° distance function on T>p. We can view S(p, t) as a level hypersurface of p, S(p,t)=p-\t). The gradient Vp restricted to S(p,t) is a normal vector field along S(p,t). Fix a unit vector y G S P M. Let c(t) := expp(iy), 0 < t < iy. For a vector v G TpM, let J(t) be the Jacobi field along c satisfying 7(0) = 0,
Dc7(0) = v.
Assume that gy(y,v) = 0. By the Gauss Lemma (Lemma 11.2.1), we know that J{t) and DcJ(t) are gc-orthogonal to c(t) for all 0 < t < e, Sc(t) (c(t),J(t))
=0 = gc W (c(t), D 6 7(£)).
Let St be the shape operator of S(p,t) at c(t). By (14.42), S t (7(i)) = D 6 7(i).
(14.44)
Let V = V(t) be a parallel vector field along c with V(0) = v. Then V(t) is gc(t)-° r th°g o n a l to c(t) for all t G [0, e), gi(t)(c(t),y(i))=o.
226
Geometry of
Hypersurfaces
L e m m a 14.4.3 Let V{t) be a parallel vector field along c with V(0) = v. Assume that v is gy-orthogonal to y. The shape operator St of S(p,t) at c(t) has the following Taylor expansion St(V(t)) = -tV{t) - ±R(t)t + o(t), where R(i) is a parallel vector field along c with R(0) =
(14.45) Hy(u).
Proof. Let {Ei(t)}™=1 be a gc-orthonormal frame along c with En(t) = c{t). Let J be a Jacobi field along c with J(0) = 0 and DcJ(O) = v. Put J(t) = 7* (*)£((*) and K6(t)(Ej(t))=Rij(t)Ei(t). J satisfies Ji(t) + R)(t)J>(t) = 0, with J^O) = 0 and jl{0) = v\ This implies J\t)
= vH - ^ ' . ( O ) ^ ' * 3 + o(t3). 6
(14.46)
Let S}(t)£ t (*) := S t ( ^ ( i ) ) . It follows from (14.44) that for t > 0, S}(t) [vh - ^R{(0)vkt3
+ o(i 3 )] =v'-
±R)(0)v*t2 + o(t2).
Thus
S)(t) = This implies (14.45).
-t{6}-lR)(0)t2+o(t2)}. Q.E.D.
Now we take a look at the normal curvature and the mean curvature of the regular metric ^-sphere S(p,t) in the cut-domain Vp. Fix a unit vector y e S P M and let c(t) := exp p (iy), 0 < t < iy. Denote by At and lit the normal curvature and the mean curvature of S(p,t) at c(t) in (Vp,g) with respect to the normal vector c(£), where g := gyp is the Riemannian metric
Shape Operator
227
induced by the distance function p(x) = d(p, x) from p. Let v G TpM with gy(V' v) = 0 and V(t) denote the parallel vector field along c with V(0) = v. We know that V(t) is always tangent to S(p,t) at c(t) for any 0 < t < iy. From (14.39) and (14.45), we obtain MV) = ^SyM-lzyfaiv^vjt 77
nt
1
+ oit),
(14.47)
1
= — - - R i c ( i / ) t + o(i).
(14.48)
Define Ty : TXM -^ TXM by Tm(V(t))=Ty(v)+Ty(v)t
+ o(t).
By (14.24), we obtain the following Proposition 14.4.4 Let V = V(t) be a parallel vector field along a geodesic c(t) = expx(ty), 0 < t < e. The normal curvature At of S(p, t) at c(t) has the following Taylor expansion, MV(t))
= \sy(v,v)
-T »
- I [ R » + 3 t » ] i + o(t).
(14.49)
Now we consider a Finsler m space (M, F, dp,). For a vector y £ TXM, define S(y) by S(c(t)) = S(y) + S(y)t + o(t), where c(t) denotes the geodesic with c(0) = y. By (14.33), we obtain the following Proposition 14.4.5 Let (M, F, dp,) be a Finsler m space. curvature Ht of S(p, t) at c(t) have the following forms nt = ^
- S(y) - i [Ric(y) + 3S(y)] t + o(t).
The mean
(14.50)
Chapter 15
Geometry of Metric Spheres
Metric spheres in a Finsler space are the level hypersurface of a distance function from a point. In Section 14.4, we studied the geometry of small metric spheres. In this chapter, we will study metric spheres at large scale. We will show that the normal (resp. mean) curvature is under control by the flag (resp. Ricci) curvature and the T-curvature (resp. S-curvature).
15.1
Estimates on the Normal Curvature
Let (M, F) be a positively complete Finsler space and p € M. The regular metric i-sphere S(p, t) is defined by
S(p,t) =
S(P,t)nvp,
where Vp denotes the cut-domain at p. The distance function p(x) := d(p,x) is C°° on Vp \ {p} such that S{p,t) = p~l{t) and S(p,t) = p~l{t) n Vp. The gradient Vp is a C°° unit geodesic field on T>v \ {p}, hence it induces a Riemannian metric g :— gvp on Vp \ [p}. Let Vp denotes the gradient of p with respect to g. According to Lemma 3.2.1, Vp = Vp is also a unit geodesic field of F := \/g. Further, n = ^p\§
K{P,y)<\. 229
230
Geometry of Metric
Spheres
This is equivalent to saying that (y),v)
< A {gy(y,y)gy(v,v)
-
Sy(y,v)sy(y,v)},
where y,v £ TM \ {0}. Similarly, we define the bound K > A.
Theorem 15.1.1 Let (M,F) be a positively complete Finsler space and p £ M. Let At denote the normal curvature of S(p, t) in (Dp,g) with respect to the outward-pointing normal vector. (a) Suppose that K < A. Then i
s
'x(t) -
(b) Suppose that K > A. Then SAW S
A(*)
Proof. Fix a unit vector y £ SpM and let c(t) := expp(ty),
0 < t < iy.
Take an arbitrary vector v S TpM with gy(y, v) = 0. Let J(t) be the Jacobi field along c with .7(0) = 0,
DcJ(0) = v G TpM.
By the Gauss Lemma 11.2.1, J(t) and DcJ(t) are tangent to S(p,t) at c(t) for any i with 0 < t < iy.
Estimates
on the Normal
Curvature
231
We first prove (a) under the assumption that K < A. By (14.43), K(J(t))=Sc(t)(p6J(t),J{t)). Let m
MJ(t))
_ Sc(t)(Dc^(t), •/(*))
' ' ~ 9cW(J(t),J(t))-
g6{t)(J(t),J(t))
•
Using the Jacobi equation (11.2), we obtain gc(Rc(J),J)
- gawj)
, ge(DeJ,DeJ)gc(.7,J)-gc(DcJ,J)2
,2
^air
*•
+
,., .,
(15J)
By the Schwartz inequality, g e ( D e J , DiJ)Si(J,
J) - g c ( D c J , J ) 2 > 0,
we obtain < } > ' > - \ -
(15.2)
To estimate >, we introduce a new function 0 A ( * ) : = ^ .
(15.3)
4>x satisfies the following ODE,
(15.4)
Let / ( t ) : = e /WW,(«P[ # ) ^ x ( ( ) ].
It follows from (15.2) and (15.4) that
/'(*) = e / ( * ( t ) + ^ ( t » * [ 0 ' ( t ) + ct>2(t) - 4>'x(t) - 4>\{t)] > 0. Thus / is non-decreasing. By Proposition 11.2.3, we have 1 t
lgy(R(,),) 3 gy{v,V)
Thus f(t) > lim f(t) = 0. t-+o+
232
Geometry of Metric
Spheres
This implies that
g,(D,J,J),s'Aft) -Sx(ty Si(j,J)
u < t < 1
i lb - b >
y
Now we prove (b). Let E{t) be a parallel unit vector field along c with respect to gc(t). Let 4>{t) :=
At(E(t)).
By (14.39), we have
= -S6(R6(E),E)-ei(st(E),St(E)).
(15-7)
By the Schwartz inequality g6(st(E),Ey
<Si(st(E),St(E)),
we obtain
/(*):= e/(* + * A)
= eI(*+M"
^'(t) + 4>2{t) - 4>'x{t) - cj>\(t)} < 0.
(15.8)
Since <j>(t) has the Taylor expansion (15.5), from (15.8), we obtain / ( t ) < t l i m + / ( t ) = 0. Thus
0 < t < iy. Q.E.D.
Estimates
on the Normal
233
Curvature
Remark 15.1.2 Let t)2p denote the Hessian of p with respect to the Riemannian metric g = gvp on the cut-domain T>p. Let Ay P denote the normal curvature of S{p,t) in (Vp,g) with respect to the outward-pointing normal vector Vp. According to (14-25), we have t)2p{v) = D2p(vx)
= Avpiv^),
Vv € TXM,
(15.9)
where v1- := v — 3(Vp x ,^)Vp x . The estimates on Avp in Theorem 15.1.1 also give estimates on D2p. For instance, if K < A, then t)2P(v) > ^ M g ( ^ , ^ ) , S
v e
rxM,
(15.10)
AU>)
Recall that the T-curvature satisfies T A y = AT„,
T„(y) = 0 ,
A > 0.
We say that the T-curvature satisfies the bound T > — 5 if
T„(«) > -<j{gv(«,«) - [gv(u, ^ y ) ] 2 } ^ ^ ' where y, u € T M \ {0}. Similarly, we define the bound T < 6. Equation (14.24) and Theorem 15.1.1 imply T h e o r e m 15.1.3 Let (M,F) be a positively complete Finsler space and p € M. Let At denote the normal curvature of S(p,t) in {VPlF) with respect to the outward-pointing normal vector n. Then (i) if K < A and T < 6,
A ^ si(t)
'Mii5M s -
(ii) ifK>\
andT>
(I5n)
-6,
A
'^{lti +i K
< 1512 >
Geometry of Metric
234
Spheres
Remark 15.1.4 Let D 2 p denote the Hessian of p with respect to F. (14-2), we have D2p = D2p-TVp.
By
(15.13)
In virtue of (15.9) and (15.13), we see that the estimates on At in Theorem 15.1.1 also give estimates on D 2 p. For instance, z/K < A and T < S, then V2p(v) > { ^ 4 where vL := v 15.2
" s\gvP(vx,
vL),
W G TXM
(15.14)
g^Pa!(Vpx,v)Vpx.
Convexity of Metric Balls
Let (M, F) be a complete reversible Finsler space. It is proved by J. H. C. Whitehead [Wh] that at any point p G M, there is a small r > 0 such that B(p, r) is strictly convex, that is, for any minimizing geodesic c(t), a < t < b with c(a),c(b) G B(p,r), the whole curve c is contained in B{p,r).
We can estimate the size of the convex balls in M under certain curvature bounds. Let A, 5 € R. Define r0 — r0(A, S) to be the first zero of s'x(t) - 6sx{t) = 0. Let n = n(p, A, S) := min (r0,
ip).
Convexity of Metric
Balls
235
Let c(t),0 < t < a, be an arbitrary minimizing unit speed geodesic with c(0),c(a) e B{p,rx/2). Note that a < d(p, c(0)) + d(p, c(a)) < rx. We claim that c is contained in B(p, ri). Let t be an arbitrary number with 0 < t < a. lit< a/2, d(p, c{t)) < d{p,c(0)) + d(c(0),c(t)) < n / 2 +
t
If r > a/2, then d(p, c{t)) < d{p, c(a)) + d(c(a), c(t)) < rx/2 + (a - t) < r±. We conclude that c(t) 6 B(p,r{) for any 0 < t < a. We will show that c is actually contained in B{p,r\/2). Theorem 15.2.1 Let (M, F) be a complete reversible Finsler space. Suppose that the flag curvature and the tangent curvature satisfy the bounds K < A,
T < S.
Then the metric ball B(p,r) is strictly convex for any r < r\/2. Proof. Let c : [a, 6] —> M be a minimizing geodesic with c{a), c(b) £ B(p, r). We may assume that c does not pass through p. By assumption, c C B(p,r\). Let p(x) = d(p,x). By (15.14) the function f(t) := p o c(t) satisfies
f"{t) = D2p(c(t)) > { g Z | | _ 6}gVp(c{t)\c{t)^) > 0, where c(i)±:=c(i)-gVp(Vp,c(t))V/9. We conclude that /(i)<max(/(a),/(&)),
Vt £ (a, b).
This proves the theorem.
Q.E.D.
Assume that A < 0. We have lim ^ t^oo
= v^A. S\(t)
Geometry of Metric
236
Spheres
If \/^A - 6 >0, then s\(t) - 6sx(t) ^ 0 for all t > 0. By this observation, we obtain the following Corollary 15.2.2 Let (M, F) be a complete reversible Finsler space. Let (M, F) be a positively complete simply connected Finsler space. Suppose that M is simply connected and the flag curvature and the S-curvature satisfy the bounds K < A,
T < S,
where A < 0, S > 0 with y/—X — S > 0. any metric ball B(p,r) convex.
15.3
is strictly
Estimates on the Mean Curvature
In the previous section, we have obtained some estimates on the normal curvature of metric spheres under certain bounds on the flag curvature. The mean curvature of a metric sphere can viewed as the mean of the normal curvature. In this section, we will show that the mean curvature is under control by the Ricci curvature and the S-curvature. Let (M, F, dfi) be a positively complete Finsler m space and p G M. The distance function p(x) := d(p,x) induces a Riemannian metric g :— gv P on the cut-domain Vp C M. Theorem 15.3.1 Let (M,F) be an n-dimensional positively complete Finsler space. Let tlt denote the mean curvature of S(p, t) in (Vp, g) with respect to the outward-pointing normal vector. (i) Suppose that K < A. Then
A(t) Ut>(n-l)^y-.
(15.15)
(ii) Suppose that Ric > (n — 1)A. Then n
t
< ( n - l S) ^ .
(15.16)
A(C)
Proof, (i) follows from Theorem 15.1.1 (i). Now we prove (ii). Let St be the shape operator of the metric sphere S(p,t). Fix a unit tangent
Estimates
on the Mean
Curvature
237
vector y € SpM and let c(t) := expp(ty), t > 0, and {£;(£) }™=1 be a gc(t)orthonormal frame along c with En{t) = c(t). It follows from (15.7) that ,
n~1
n—1
^ [ £ ^ ( W ) ' ^ ) ] =-Ric(c)-^g,(s t (£ i ).St(^))i=l
(15-17)
i=l
By Proposition 14.3.3, the mean curvature lit is the mean value of the normal curvature A t , ra-l
nt = £At(£i(t))Let 4>{t) := ~Tl^t-
(15.18)
By (14.39), we have nn -—1 l
1
i=\
Observe that n—1
„
n—1
2
{^g6(st(£0.^i)} < ( n - l ) E ^ ( S ^ ) > S ^ ) ) i=l
i=l
Namely, n —1
1
^W ^ ^£gc(st(£i),St(£0)-
(15-19)
t=i
By (15.17) and (15.19), we obtain if < - A
Take a Jacobi field J(i) along c such that J(0) = 0 and DcJ(O) = v, where v satisfies gy{y,v) = 0. By the Gauss Lemma (Lemma 11.2.1), we know that both J(t) and DcJ(t) are g^tj-orthogonal to c(t). By (14.43), MJ(t))
=
scW(i>cJ{t),J(t)).
Geometry of Metric
238
Spheres
Using the Taylor expansions in Proposition 11.2.3, we obtain
#*) = 7 - ^r^mc^t t
+ °W-
(15-2°)
J(n — 1)
By the standard argument as in the proof of Theorem 15.1.l(ii), one obtains
Ht)<
S'A(*)
SAW
This gives (ii).
Q.E.D.
Recall that the S-curvature is positively homogeneous of degree one, S(Xy) = XS(y),
A >0.
We say that the S-curvature satisfies the bound S > — (n — 1)6 if S(y)>-(n-l)6F(y),
Vj/eTM\{0}.
Similarly, we define the lower bound S < (n — 1)6. By (14.33) and Theorem 15.3.1, we immediately obtain the following Theorem 15.3.2 Let (M,F,dfi) be an n-dimensional positively complete Finsler m space. Let Ht denote the mean curvature of S(p,t) in (DP,F) with respect to the outward-pointing normal vector. (i) Suppose that K < A,
S < (n - 1)6.
Then nt>(n-l)&&-(n-l)5. s\{t)
(15.21)
(ii) Suppose that Ric>(n-1)A,
S>-(n-l)6.
Then Ut<(n-l)^§-
+ (n-l)6. S
A(<)
(15.22)
Estimates
on the Mean
Curvature
239
In virtue of (14.30) and (15.22), we obtain the following Theorem 15.3.3 Let (M,F,dp.) be an n-dimensional Finsler m space. Suppose that the Ricci curvature and the S-curvature satisfy the bounds,
Ric>(n-1)A,
S>-(n-l)<5.
1 hen the distance function p(x) := d(p, x) satisfies
Ap < ( n - 1 ) ^ 4 + ( n ~ 1)*SA(P)
In virtue of (14.30) and (15.21), we obtain the following Theorem 15.3.4 Let (M,F,du) be an n-dimensional Finsler m space. Suppose that the flag curvature and the S-curvature satisfy the bounds K < A,
S < (n - 1)6.
Then
Ap>(n-1)^4-("-!)<*•
(15-23)
Below is an application of Theorem 15.3.4. Theorem 15.3.5 Let (M, F, d^i) be an n-dimensional positively complete simply connected Finsler m space. Suppose that for constants A < 0 and 5 > 0 with \J—\ — 5 > 0, the flag curvature and the S-curvature satisfy the bounds K
S<(n-1)<$
Then for any bounded regular domain il C M,
Proof. Let p(x) = d(p, x), where p $ Cl. Then p is C°° on fi.
(15.24)
240
Geometry of Metric
Spheres
By definition, Ap = div(Vp). By Theorem 2.4.2, we obtain
Jn
Apdfi=
Jen
g n (n, Vp) dv < v(dQ),
where n is the outer normal along dfl. By (15.23), we have
Ap> {n - l)(\/^\ - 6) > 0 . Therefore (n - lJCV^A - 6)ii{tl) < f Apdfi< Jn This proves the theorem.
i/(9fi). Q.E.D.
Remark 15.3.6 By Theorem 15.3.5, one can obtain an upper bound on the first eigenvalue of a Finsler space satisfying (15.24).
Metric Spheres in a Convex
15.4
Domain
241
Metric Spheres in a Convex Domain
In this section, we shall estimate the mean curvature of a metric sphere from below under certain upper bounds on the Ricci curvature and the Scurvature. The metric spheres under our consideration lie inside a convex domain at a point. For a unit vector y S S P M, the convex value iy of y is defined to the largest 0 < r < iy such that the shape operator St of th regular metric sphere S(p,t) has nonnegative eigenvalues at exp p (ty) € S(p,t) for all 0 < t < r. Let
Vp := {expp(ty), 0 < t < \y}. Vp is called the convex domain of p. From the definition, we know that t>p C Vp. For t > 0, let S(p,t):=S(p,t)nf)p. By definition, S{p,t) C S(p,t). Let y G SPM and c(t) := expp(ty), 0 < t < i y . Take an arbitrary g£(t)orthonormal basis {Ei(t)}"Zi for TxS(p,t). By definition, the symmetric matrix (hij) has nonnegative eigenvalues hij::=gc(st(£i),£j). Thus 2_j hikhki < \2_jhii) i,fc=l
•
i=\
This implies that n—l
n— 1
^
£gc(st(£i),St(£i))< []Tg<;(St(£U£i)] . i=l
i=l
Assume that Ric < (n - 1)A. Let n t denote the mean curvature of S(p,t) at c(t) in (Vp,g) and n— 1
(15.25)
242
Geometry of Metric
Spheres
It follows from (15.17) and (15.25) that l)4>2.
<j>' >-X-(nLet 1 1>x(t) •=
S
(„-I)A(*)
n — 1 s( n - l ) A (*)'
Then
1^ = - A - ( n - l ) V £ . It follows from (15.20) that n — 1 We may assume that n > 2 (the case when n = 2 is covered by Theorem 15.3.1(i)). Then for sufficiently small t0 > 0,
Consider the following function
/(t):=e/ ( * + ^ ) d t [0(t)-^(t)". We obtain
/'(*) = e / ( * + ^ ) * [<^) + ^2(i) - ^(i) - V*(*)] > 0. Thus /(*) > /(*o),
to < t < i„.
This implies
3Jt° 4>(t0) - *l>\(to)
> 0.
Since t 0 > 0 can be arbitrary small, we have proved the following Theorem 15.4.1 Let (M,F) be an n-dimensional positively complete Finsler space. Let tit denote the mean curvature of S(p,t). Suppose that the Ricci curvature satisfies the bound Ric < (n - 1)A.
Metric Spheres in a Convex
Domain
243
Then S(n-l)\(t)
The above theorem is a slight modification of Ding's estimate on the Laplacian of the distance function on a Riemannian space [Di] (see also [Xi]). Remark 15.4.2
Suppose that Ric < (n - 1)A,
S < (n - 1)6.
It follows from Theorem 15.4.1 that the Laplacian of p = d(p, •) on the convex domain T>v and the mean curvature II,, of S(p, p) satisfy Ap = IL, >
S(
"~ 1 ) A ,
- (n - 1)<J.
(15.26)
S(n-1)A(P)
Remark 15.4.3 Let Ap denote the Laplacian of p with respect to g. According to (14-34), we know that the estimates on n n in Theorem 15.3.1 also give estimates on Ap.
Chapter 16
Volume Comparison Theorems
In the previous chapter, we estimate the mean curvature of metric spheres from above. By a similar argument, we can estimate the mean curvature of the level hypersurfaces of a distance function. In this chapter, we will show that there is a close relationship between the growth of the mean curvature and the growth of the volume of metric spheres. Then we establish a volume comparison theorem for the volume of metric balls under a lower Ricci curvature bound and a lower S-curvature bound. There are applications of this volume comparison theorem. Throughout this chapter, we always set w n :=Vol(B n ), 16.1
a n _! := VoKS"- 1 ).
Volume of Metric Balls
In this section, we will estimate the volume of metric balls in a Finsler m space. Let (M, F, d/j.) be an n-dimensional Finsler m space. We first take a look at the Busemann-Hausdorff volume form dVp and its induced volume form dAp on hypersurfaces. For a vector y G S p M, let {bj}™=1 be a basis for TpM such that bi=y,
gy(2/,bi)=0,
i=
2,---,n.
Extend {bi}™=1 to a global frame on TVM in a natural way. We see that bj's are tangent to SPM for 2 < i < n. Let {9l}™=1 denote the basis for T*M dual to { b i } ^ . Express dVF at p by dVp = aF(p)61 245
A - - - A 6>n,
Volume Comparison
246
Theorems
where , , aF{p) =
Vol(B") -r rVolU^eR" F(ylbi)
Extending {^2}"=1 to a global coframe on TpM in a natural way, we obtain a volume form dVp on TpM. The induced volume form dAp by dVp on SpM is given by dAp=
A---A9n.
We have /
dAp =
an-\.
J SPM
The volume form dfi can be expressed as dfi = <j> dVF-
At point p € M, let dvv denote the induced volume form by dp,p on SPM. We have dvp = 4>(p)dAp. Thus / JS„M
di/p = (j)(p) dAp =
(16.1)
For a point p € M and £ > 0, let Ej, := {y G S P M | i„ > i } C S P M. S P is an open subset in S P M. Note that S* = S P M for 0 < t < ix and Ep'cSp",
t"
Let Vp denote the cut-domain at p and S(p, t) the metric sphere of radius t around p. The regular part S{p,t):=S{p,t)nVp
Volume of Metric
Balls
247
is a C°° hypersurface. We call it the regular metric sphere of radius t around p. Define (pt : Dp -> S(p, t) by
y € Dp, t > 0.
?t is a diffeomorphism. Let dA t denote the induced volume form by dVF on S(P,t). Then efi/t =
V&K
(16-2)
and TH(V) = 0,
yeSpM^.
We have
rlt(y)=tn-1[l
+ o(t)].
(16.3)
We prove (16.3) as follows. Fix a vector y € SPM and let c(t):=
0
Take an arbitrary basis {bi}™=1 for TpM as above and define Ei(t) := (exp p )»| tl/ (bi),
t>0,
i=
l,---,n.
Ei(tys are C°° vector fields along c. Let {#*(£)}™=1 denote the basis for Tc{t)M dual to {£*(£) }? =1 and dn\c{t)=a(t)01(t)A---A8n(t),
where o-(t) :=
f ; -rr r-. V o l [ ( y » ) e R n | F{y'Ei(t)J < 1)
We have d ^ =a(t)62{t)
A--- Affn(t).
Since {£;(i)}™=1 are C°° at t = 0, we conclude that
(vtTe'mbj) = ei(t)((
Volume Comparison
248
Theorems
Thus
(
{iptYdvt = ^ ) W ' [ « 2 ( f ) A . - A f ( t ) ] = tn-l(x(t) e2 A • • • A en =
i
=
i
v(0)
o-(O) 6>2 A • • • A 6n
. n - l *(*) dvr, Or(O) - P
Thus
a(t) »7t(y) =
(°)
Vol | V ) G R F
( H < *} ,„-a
VoljV) G R" I F(y'£i(t)) < l}
Thus (16.3) holds. For a small number e > 0, define ^W-Vt+s0^1^).
xeS(p,t).
(16.4)
For a point a; € 5(p, £), there is an open neighborhood U of x in 5(p, t) and a small number 6 > 0 such that >e is defined on W for all 0 < £ < 6.
(p, t + e)
As in (.14.27), define Q(x,s) by (^ e )*(^t+ E U,(x)J = 6(a;,e)di/ t U,
x G 5(p,t).
Volume of Metric
Balls
249
By (16.2) and (16.4) we have rv
N
Vt+e{y) W)
, ,
(16.5)
Let n t denote the mean curvature of S(p,t) at x with respect to the outward-pointing normal vector. In virtue of (14.29) and (16.5), we have
n* = de
_d_ (\nr) (y)). t
ln©(a;,£) E=O
(16.6)
dt
Let
x(t):=[e"sA(t)]n_1. By Theorem 15.3.2 and (16.6), we obtain nt = | ( l n ^ ) ) < | ( l n
X
(t)).
(16.7)
This implies d (Vt(V)\ /Vt(y) 0. dt V V xy(t) (t) J Thus the quotient h(t) := r)t(y)/x(t)
(16.8)
is non-increasing.
Now we study the induced volume u(S(p,t)) sphere S(p,t).
of the regular metric
By (16.2), we have v(S{p,tj) V
= / '
(16.9)
Vt{y)disp.
JS„M
Applying the co-area formula to the distance function p{x) :=• d(p,x), we obtain n(B(p,tj)
=n(B(p,t)nVp}
=J
v(S(p,s))ds.
Thus JB(p,t)) V
= f '
JO
\ f
]
-JS„M
ris{y)dup\ds. J
(16.10)
250
Volume Comparison
Theorems
It follows from (16.3) that v(S(p,t)) V
T)t(y)dvp = 0(p)a n _ 1 * n - 1 fl + o(t)
= f '
(16.11)
L
JSPM
Now we are ready to prove the following T h e o r e m 16.1.1 ([Sh2]) Let (M,F,dfj.) be an n-dimensional complete Finsler m space. Assume that Ric>(n-1)A,
S>-(n-l)6.
positively
(16.12)
Then the quotients v(s(p,t)) i ^ r j - and e*'s A (i)| /„*
»(B{p,t)) ^ >—_ \e^sx(s)ds)
(16.13)
are non-increasing. In particular, we have M(B(P,<))
< i ( ? K /
[eSssx(s)Y'lds.
Proof. From (16.8) we see that r)t{y)/x(t) is non-increasing.
vt2(y)x{ti) < vtAy)x(t2), o
(w.u)
Integrating (16.14) over SPM with respect to dvp
A(t2)x(h)
< i4(*i)x(*2),
0 < h < t2.
(16.15)
Inequality (16.15) implies that the first quotient in (16.13) is non-increasing. Rewrite (16.15) as follows A(t")x(t')
< A(t')x(t"),
0 < t! < t".
(16.16)
Integrating (16.16) with respect to t" over [ii,t 2 ], then integrating the resulting inequality with respect to t' over [0,*i], we obtain the following inequality / JO
X(t)dt
/ Jt!
A(t)dt < / .10
A(t)dt / Jti
X{t)dt.
(16.17)
Volume of Metric
Balls
251
Inequality (16.17) implies that the second quotient in (16.13) is non-increasing. This completes the proof. Q.E.D. The above comparison theorem on the Busemann-Hausdorff volume is due to Bishop [BiCr] and Gromov [GLP][Gr4] in the Riemannian case. Integrating (16.15) with respect to t\ over [0,i2] yields e5u*x(u)
i/(s(p,i))
(16.18) Jo< [e^s^w)]™" 1 du
M(-B(P,*))
In other words,
jt[ln,(B{p,t))}
By (14.50) and (16.6), we obtain a Taylor expansion for 7]t(y) which is better than (16.3) Vt (y)
= i " " 1 [l - S(y)t - i (Ric(2/) + 3S(y) - 3S 2 (y))i 2 + o(t2)].
(16.19)
Since Vp contains the metric ball B(p, ix), S(p, t) = S(p, t), 0 < t < ix. By (16.9) and (16.10), we obtain a Taylor expansion for v(S(p,t))
which is
better than (16.11),
i/(s(p,i))
=
r(p)+3h(p)+2 r(p) + 3h(p)
where
—I On-l
s(p)
:
h(p):
=
Kic(y)dup
JSSM
/
—f O'n-l
2
—t - -^~ T^tz + o(t )}, 6(n + 2) n+l
1
r{p) :
2 tz + o(t )},
JsvM
S(y)disp,
S(y)-S2(y)
\dvv
252
Volume Comparison
Theorems
Note that if F is reversible, i.e., F(~y) = F(y), then S(-y) = -S(y). s(p) = 0.
16.2
Thus
Volume of Tubular Neighborhoods
In this section, we will estimate the volume of tubular neighborhoods of a hypersurface under lower bounds on the Ricci curvature and the Scurvature. Let (M, F, dp) be an n-dimensional Finsler m space. We still assume that Ric > (n - 1)A,
S > -{n - 1)8.
Consider a hypersurface N C M and a normal field n along N. For the sake of simplicity, we assume that N is closed. Take a small neighborhood U of N. N divides U into two parts. Let U+ denote the part toward which n points and U- the other part. Choosing a smaller U if necessary, we may assume that the following function p(x) :=
f d(N, x) \ - d(x, N)
x G U+ x e U+
is a C°° distance function on U of N with p _ 1 (0) = N and n = Vp|jvp induces a Riemannian metric g := gy p on U. Let n t denote the mean curvature of Nt := p~l(t) in (U,F) with respect to V'p\Nt and n n denote the mean curvature of N in (W, g) with respect to n. Assume that n„<(n-l)r.
(16.20)
We will show that the mean curvature n t of Nt in (U, F) satisfies nt<(n-l)ft{\n[eSt(rSx(t)+s'x(t))]}.
(16.21)
We prove (16.21) as follows. Define ipt : N ->• M by ipt(z) := exp z (tn 2 ),
z e N.
For small t > 0,
Volume of Tubular
Neighborhoods
253
respect to the normal vector Vp — Vp, respectively. Let *{t) :=
^ U ,
Since g is a Riemannian metric, the mean curvature lit is the mean value of the normal curvature At. See Proposition 14.3.3. Let St denote the shape operator of Nt - /0 -1 (i) at c(t) e Nt. By (14.39), we have n-l
:—[Yi&c(St(Ei),Ei)1
where {Ei(t)} is a frame along c. As we see in the proof of Theorem 15.3.1, the Riccati equation (14.41) implies 4>' < - A -
(16.22)
Let 0 o ( t ) : = ^ [ l n ( T S A ( t ) + s'A(i)) . Note that & = "A - €
(16-23)
Let
f(t) := e/o W')+*o(.»<*. j ^ ( t ) _ 0 o ( i ) ] _ By (16.22) and (16.23), we have
/ ' ( * ) = e /o w*)+*°(»))
t > 0.
That is, nt<(n-l)0o(t).
(16.24)
254
Volume Comparison
Theorems
By (14.33) and (16.24), we have
nt
=
n t - S(c(*)) < (n - l)4>o + (n- 1)8
=
(n-l)jt{\n[e^(rsx(t)+s'x(t))}}.
(16.25) Q.E.D.
This proves (16.21).
Now we study the growth of the volume of the tubular neighborhoods of AT, Ut(N) := ix e U+
d(N,x)
For a point z £ N, let dz denote the largest number r > 0 such that p(
0
Define r)t(z) by (
=m{z)(dv\z),
(16.26)
where dvt denotes the induced volume form on Nt by d\x. Set rjt(z) = 0 for t>dz.
We have vt{Nt) = /
JN
nt(z)dv.
Volume of Tubular
Neighborhoods
255
and li(ut(N))=J
vs(Ns)ds
=J
[J
ns(z)di
For a point x :=
By a similar argument as for (16.5), we obtain e{x,e)
= %±Z&, Vt(z)
x = M*)-
By (14.29), we obtain nt =
|[lne(,,£)]|£=o = | [ l n ^ ) ] .
(16.27)
It follows from (16.25) and (16.27) that dr. , ,i „ . , _ d -[\nVt(z)]=Ut<(n-l)~{\n[eSi{rsx(t)
+
S'x(t))]}.
Thus the following ratio Vt(y) eSt(rsx(t)+sx(t)) is non-increasing. By the same argument as above for the volume of metric balls, we obtain the following Theorem 16.2.1 Let (M, F) be an n-dimensional positively complete Finsler space. Let N c M be a closed hypersurface and n a normal vector field on N. Suppose Ric>(n-1)A,
S>-(n-l)<5, .
ft„<(n-l)r.
Then the following ratio (16.28) , ( J V ) / 0 , [ e * » ( r s A ( U ) + s' A («))] n
du
256
Volume Comparison
Theorems
is non-increasing on [0, oo). In particular, when N is the boundary of M, M (M)
< u{N) J
\e6u (r s A (u) + s' A (u))] " " ' d u ,
(16.29)
where d = sup : r 6 M d(N, x). The above result in the Riemannian case is due to Heintze-Karcher [HeKa].
16.3
Gromov Simplicial Norms
In order to find some relations between the geometry and the topology of Riemannian manifolds, M. Gromov introduced many new topological quantities. Among them are the so-called Gromov simplicial volume [Gr3]. In this section, we will give a brief discussion on the Gromov simplicial volume and its relationship with the Ricci curvature and the S-curvature. Let M be a topological space. Denote by Cfc(M) the fc-th complex of real singular chains c = J ^ r^ai, where <7j : Afc —> M are fc-dimensional simplices and r; are real numbers which are all, but finite, zero. Let Hfc(M) denote the real singular homology of M. The natural L 1 -norm || • ||i on Ck{M) is defined by llcll = X ] l r i l '
for
c = ^Vi
i
i
This L 1 -norm induces a pseudo-norm || • || in Hfc(M). We call || • || the Gromov simplicial norm. For a class z e Hfc(M), ||z|| is given by \\z\\ := inf |fc||i. z=[c]
Assume that M is an n-dimensional closed oriented manifold. For the fundamental class [M] 6 H n (M), we set
||M||:=||[M]||. The constant ||M|| is called the Gromov simplicial volume.
Gromov Simplicial
Norms
257
Assume that (M, F) is an oriented reversible Finsler space. For a C°° immersed submanifold a : N —• M, define Voider) :=
Vo\a.F(N).
This notion can be extended to all singular chains as follows. For a singular simplex a : Afc ->• M, define VolF(er) as the infimal of VolG(Afc), where G is a reversible Finsler metric on Afc such that a : (Ak,G) -> (M,F) is distance-decreasing. Set VO1F(
:=^2\ri\Vo\F(ai). i
For a homology class z G Hfc(M), we set Vol F (z) := inf Vol F (c). z=[c]
Then for the fundamental class [M] G H„(M), Vol F (M) = Vol F ([M]). Gromov proved the following Lemma 16.3.1 ([Gr3]) Let (M,F) be an n-dimensional oriented closed reversible Finsler space. For any class z G Hj.(M),
where B(x,r) and S(x,r) denote the metric ball and the regular metric sphere of radius r around x in the universal cover ir : M = M respectively and F := n*F. In particular,
\\M\\ < nlmin r > 0 sup
[v^f(f^TvolF(M).
igM LVol F (B(x,r))J
By the volume comparison theorem (Theorem 16.1.1), we can easily prove the following Theorem 16.3.2 Let (M,F) be an n-dimensional oriented closed reversible Finsler space. Suppose that the Ricci curvature and the S-curvature
Volume Comparison
258
Theorems
satisfy the bounds: Ric>-(ra-l),
S>-(n-l).
(16.30)
Then for any homology class z € Hfc(M), ||,z||
\\M\\ < n\(n-
Proof. Let F — ir*F denote the lifted Finsler metric on the universal cover 7r : M —> M. The Ricci curvature Ric and the S-curvature of F still satisfy the same bounds as in (16.30). By (16.18), we have Ap(S(x,r))
[ersmh(r)]'
^
Volp{B(*,r))-
J T U ^ Q
71-1
dt
Note that -j n— 1
er sinh(7-) l lim — \ J 0 er sinh(i)
,
= 2(n - 1). ai
It follows from Lemma 16.3.1 that for any z G Hfc(M), Ikll < k\(n -
l)k2kVo\F(z). Q.E.D.
Let (M, F) be an n-dimensional closed oriented reversible Finsler space. An open subset U C M is said to be amenable if for any connected component i : W C U —> M, the image ,(TTI(W,
xj) C it\{M,x),
xeW
is an amenable subgroup of TTI(M). Hence if TT\ (M) is amenable, then every open subsets in M are amenable. M. Gromov proved that if the manifold is covered by amenable open subsets {Ui} such that every point is contained at most n open subsets Ui, then the simplicial volume vanishes.
Estimates
on the Expansion
Distance
259
Assume that the Ricci curvature and the S-curvature of F satisfy the bounds (16.30). Following the line in [Gr3], one can construct an (n — 1)dimensional polyhedron pn~l and a map / : M —> P n _ 1 such that / _ 1 (W) of any star neighborhood U in pn~x is amenable, when the volume of F is sufficiently small. Since P™~x is covered by its star neighborhoods of vertices and each point in P™ _1 is contained in at most n star neighborhoods, we conclude that M can be covered by amenable open subsets {Ui} (the pre-images of star neighborhoods) such that every point x is contained in at most n = d i m M open subsets Ui. This implies ||M|| = 0 . More precisely, Theorem 16.3.3 Let (M, F) be an n-dimensional versible Finsler space satisfying Ric>-(n-l),
closed oriented re-
S>-(n-l).
(16.31)
There is a number e(n) > 0 such that if for some 0 < v < 1, VolP(B{x,p0))
< e(n) vn,
Vx G M
(16.32)
then there exist an (n — 1) -dimensional polyhedron Pn~l and a continuous map f : M —> P n _ 1 such that for any star neighborhood U C Pn~l, the pre-image f~l(U) is almost nilpotent. Further Diam(/" 1 (W)) < C{n)v. Thus \\M\\ = 0 . According to Theorem 16.3.3, if M is an n-dimensional oriented closed manifold with ||M|| ^ 0 and F is a reversible Finsler metric on M satisfying the curvature bounds (16.31), then there is a positive number e(n) depending on n such that Vol F (M) > e(n).
16.4
Estimates on the Expansion Distance
In many cases, metric measure spaces concentrate to a point if the dimension goes to infinity. A natural question arises: under what conditions the
260
Volume Comparison
Theorems
dimension of metric measure spaces in a concentrating sequence must go to infinity? We shall give an answer to this problem for Finsler m spaces. Let (M, F, dp,) be an n-dimensional compact Finsler m space. Let dVp denote the Busemann-Hausdorff volume form of F and S denote the Scurvature of (F, dVp). Put dp = 4>ixdVp.
The S-curvature SM of (F, dp) is related to S by SM = S - % .
(16.33)
We have Theorem 16.4.1 ([GuSh]) Given numbers A < 0 , J > 0 , d > 0 and r > 0. Let X = (M, F, dp) be an n-dimensional compact reversible Finsler m space. Suppose the Finsler metric F satisfies Ric > (n - 1)A,
S > - ( n - 1)8,
Diam(M) < d
(16.34)
and the volume form dp — (p^dVp satisfies ^|<(n-l)T,
(16.35)
where un denotes the volume of the standard unit ball in M. . Then for any 0<e< l/{l + enkd), ExDist(M; e) > e^d
[(1 - e )± - e" d £ *] [ ^ W ]
" > 0,
(16.36)
where K : = %/—A + 5 + r. Proof. From the definition of expansion distance, we know that if for some p > 0, there exists a compact subset A
-e)V,
then ExDist(M; e) > p.
Estimates
on the Expansion
Distance
261
Let d := Diam(M),
V := /x(M),
v :=
VO1F(M) U)„
By assumption and (16.33), the S-curvature SM of (F,dfi) is bounded S
M
> S - ^ > - ( n - l ) ( < 5 + r).
0M
Let p € M such that 0M(p) = inf 0^. Then by assumption, V = n{M) > ^ ( p ) V o l F ( M ) = w n 0 M (p)u n .
(16.37)
Take a small ball A = B(p, r) such that M(A) = Ve. Then £ / p ( A ) = B ( P ) r + p) = B( P ) d p ),
p>0,
where dp := min(d,r + p). By Theorem 16.1.1, we have (16.38)
H(A) = Ve< 0 M (p)V(r), and
i(yp(AJ)
<
V(dP) •^e, V(r)
(16.39)
where V(r):=a7i_1|r[e^tsA(t)]n
'dt.
From (16.38), we obtain Ve = ,i(A) < ^ ( p ) V ( r ) < ^ ( p ) w „ e n K d r n ,
(16.40)
where K := V^A + 6 + r. From (16.37) and (16.40), we obtain a lower bound on r: Ve ( \ - \
kd 1/n
-
ve-
e .
(16.41)
Volume Comparison
262
Theorems
Inequality (16.41) implies that for any p > 0 V(dp V(r)
<
^djdpy \ r J
<
enKd\l +
<enKd(l
\
+ ^) rrJ>
v-1ekde-1/np\n.
(16.42)
Let p0 be the number satisfying enKd\l + v"1ekde-1/np0\nVe
=
V(l-e).
(16.43)
p0 is given by Po = e
-2nd
-Vol F (M)-ii
;i-e)»-eM£
Assume that 0 < s < 1/(1 + enkd). Then p0 > 0. It follows from (16.39) and (16.42) that for any 0 < p < p0,
This implies that ExDist(M;£) > p. Letting p —> p0, we obtain ExDist(M;e) > p0 Q.E.D.
Corollary 16.4.2 Given numbers A < 0 , 5 > 0,d > 0 and r > 0, let Xi = (Mi,Fi,dp,i) be a sequence of reversible Finsler m spaces satisfying (16.34) and (16.35) for A, 5, d and T. Suppose that for any 0 < e < ^, lim ExDist(Xi;e)
= 0,
then either n{ = dim Mi - > o o o r there is a subsequence %k such that riik = n and lim VolF. k~+oo
k
(Mik)=Q.
Estimates
on the Expansion
Distance
263
Remark 16.4.3 According to [Gr3], for a compact oriented manifold M, the Gromov simplicial volume \\M\\ does not vanish if M admits a Riemannian metric of negative curvature. But \\M\\ = 0 if TTI(M) is almost nilpotent. According to Theorem 16.3.3, for any n-dimensional closed oriented reversible Finsler space (M,F) with \\M\\ ^ 0, if
Ric>-(n-l),
S>-(n-l),
then Vol F (M) > e{n). Thus if all Mi in Corollary 16.4-2 satisfy \\Mi\\ ^ 0, then the second case does not occur, i.e., the dimensions of Mi must approach infinity.
Chapter 17
Morse T h e o r y of Loop Spaces
The Morse theory plays an important role in differential topology. In order to study the relationship between the topology and the Riemann curvature, we consider the canonical energy functional on the loop space at a point. Since the loop space at a point is infinite-dimensional, the Morse theory can not be directly applied to this case. This problem can be solved by approximations.
17.1
A Review on the Morse Theory
In this section we shall give a quick review on the Morse theory. Technical details will be omitted. One is referred to [Mi] for details. Let (M, F) be an n-dimensional Finsler manifold and / : M —> R a C 2 function. A point x £ M is called a critical point of / if dfx = 0. A number a € R is called a critical value of / if f~l(a) contains a critical point of / . It is known that the set of critical values has zero measure in R (the Sard Theorem). In a local coordinate system, the Hessian of / at a point x is given by
At a critical point x, J ^ (x) = 0, hence
We see that D 2 / is a quadratic form on TXM. 265
A critical point x is said
266
Morse Theory of Loop Spaces
to be non-degenerate if D 2 / is non-degenerate on TXM. The index of / at a critical point x is defined to be the maximal dimension of subspaces in TXM, on which D 2 / is negative definite. At a non-degenerate critical point x0 of / , there exists a local coordinate system (xl) such that f(x) = f{x0) -x\
x2k + x2k+l + • • • + xl,
where k = \ndXo(f). Thus, non-degenerate critical points are isolated. A C°° function / on M is called a Morse function if it has only non-degenerate critical points. Proposition 17.1.1
Let f be a Morse function on a manifold M.
(i) If f~l{a,b] is compact, anda,b are regular values, then f~1[a,b\ has the homotopy type of a finite CW complex obtained from f~l(a) by attaching one cell of dimension A for each critical point of index A. (ii) If f~l[a, c] is compact for all c < b and a is a regular value, then f~1[a,b) has the homotopy type of a (possibly infinite) CW complex obtained from f~1(a) by attaching one cell of dimension A for each critical point of index A. Proposition 17.1.1 is very important. However, functions arising from geometry, might have degenerate critical points. To overcome this problem, we approximate them by Morse functions. Given a C°° function / on an open subset W e i , let x G U be a critical point of / . There is an open neighborhood Ux of x and e > 0 such that if h is a Morse function on U satisfying ||D2/-D2/i||co<£ then for any critical point z £ U of h indz(h) >
mdx(f).
Proposition 17.1.2 Let f be a C°° function on a manifold M. Supposethat / _ 1 [ a , b] is compact for some regular values a,b G R. Then f can be approximated by a C°° Morse function h on M such that h~l[a,b] — f~l\a,b] and h = f near the boundary d(f~1[a,b}). Further, ifindx(f) > A for all critical points x of f on f~l[a,b\, then mdx{h) > A for all critical points x of h on /i _ 1 [a,6].
A Review on the Morse
Theory
267
We now quote some important facts from the homotopy theory. Let Y be a Hausdorff topological space and {Xi}?Z0 be an increasing closed subsets in Y such that Y = \J°Z0 Xi. Suppose that each Xi has the homotopy type of a CW complex with cells of dimension > A. Then Y also has the homotopy type of a CW complex with cells of dimension > A. Let X C Y be Hausdorff topological spaces. Suppose that Y has the homotopy type of a CW complex obtained from X by attaching cells of dimension > A. Then, for any k < A — 1, every map
irk(Y,X)=0,
k<X-l.
In a special case when A = 2, every curve c : [0,1] —> Y with c(0), c(l) € X can be deformed to a curve c* : [0,1] —> X. With this in mind, one can prove the following path-deformation theorem. Proposition 17.1.3 Let f : M —> R be a C°° function (possibly with degenerate critical points). Suppose that f~l(—oo,6] is compact for all S < b. Then for any C°° curve c : [0,1] —> f"1(—oo,b) and any e > 0, there is a homotopy cs of c with
c0 = c,
ca(0)=c(0),
c,(l)=c(l),
0<s
such that C\ is a C°° curve and
f o ci < max(d, m) + £,
where d := max ( / o c(0), / o c(l) J and m denotes the largest critical value of f with index one or zero.
268
Morse Theory of Loop Spaces
Proof. Let a — max(m, d). Assume that a + e < b. By the Sard Lemma, one can find two regular values a*,b* of / such that a < a* < a + e < b* < b and c C / _1 (—00,6*) . It follows from Proposition 17.1.2 that / can be approximated by a C°° function g : M —» R such that g-1{a*,b*] =
r1[a*,b*}
and g has only non-degenerate critical points of index > 2 in g^x[a*,&*]. Therefore, f~l[a*,b*] has the homotopy type of a CW complex obtained from f~l(a*) by attaching cells each of dimension > 2. Thus, there is a homotopy c s , 0 < s < 1, of c with CQ = c, fixing two endpoints, such that ci C / _ 1 ( a * ) , i.e., / o c\ < a* < a + e. Ci might not be a C°° curve, but one can always deform C\ to a C°° curve cj with / o c\ < a + e. Q.E.D.
17.2
Indexes of Geodesic Loops
Before we study the energy functional on a loop space, we discuss some basic properties of the index form along a geodesic.
Indexes of Geodesic Loops
269
Let (M, F) be a Finsler space and c : [a, b] -» M a geodesic. Recall that the index form Xc along c is defined by Ic(U,V):=J
{gc(Dc£/,D^)-gc(Rc(£/),K)}dt,
(17.1)
where [/, V are piecewise C°° vector fields along c. Let Wc denote the space of all piecewise C°° vector fields V along c with V(a) = V(b) — 0. Xc is a symmetric quadratic form on W c . Define ind(Xc) to be the maximal dimension of subspaces of W c on which Xc is negative definite. Set ind(c) := ind(Xc).
(17.2)
We call ind(c) the index of c. The index of a geodesic is independent of positively oriented reparametrizations. That is, for any geodesic c(t), a < t < b, and a new parametrization, t(s) = c + ds, where d > 0, the resulting geodesic c(s) := c o t(s) ind(c) = ind(c). The geodesic c is said to be degenerate if Xc is degenerate, i.e., there is a non-zero vector field J G W c such that
xc(j,v) = o,
vewc.
We see that such a vector field J must be a C°° Jacobi field along c with J(a) = 0 = J(b). Thus c is degenerate if and only if c(b) is conjugate to c(a) along c. Fix any partition a = to < • • •
U
contains no conjugate points to c(£j) along a. Py Lemma 12.1.2, for any piecewise C°° vector field V ^ 0 along Cj with V(£j) = V(Q. Let J7"c denote the subspace of piecewise C°° Jacobi fields J € W c along c such that J is C°° on each [ij, U+\\. Each J|[ti,ti + i] is uniquely determined by J{ti) and D £ J(ii). Thus dimj7c < oo.
Morse Theory of Loop Spaces
270
Let Xc denote the restriction of Ic on Jc. Define the index ind(Xc) of the quadratic form Ic in a usual way. We have L e m m a 17.2.1 For any C°° geodesic c(t), a < t < b and any partition a = to < • • • < tk = b as above. ind(c) = ind(Xc) = ind(Xc)Moreover, c is degenerate if and only ifXc is degenerate on Jc. The proof of Lemma 17.2.1 will be completed in the following two lemmas. We denote by Vc the set of piecewise C°° vector fields V along c satisfying V(U) - 0, 0 < i < k. For any V € W c , let J (= Jc such that J(ti) = V(U). Then V - J G Vc. Since c(ti+i) is not conjugate to c(U) along a = c\[ti,ti+1], Vc n Jc = {0}. Then we obtain the following direct decomposition WC = VC® Jc. L e m m a 17.2.2
For any V <EVC and J G Jc, Xe(V, J) = 0.
Proof. Observe that Tc(V,J)
=
/
Ja
{s6(pcV,D6fj-gc(Rc(V),
j)}dt
iMv^J)}dt
fc-1
t+
^gd(^,Dcj)|t'_+1=0. i=0
Q.E.D. Now we prove the second lemma. L e m m a 17.2.3
Xc|yc is positive definite.
Proof. By the choice of the partition of [0,1], any smooth Jacobi field J along Ci := c|[ tiiti+1 j with J(ti) = 0, J(ti+\) = 0 must be trivial. According to Lemma 12.1.2, for any piecewise C°° vector field V ^ 0 along a with V{U) = 0 and V(ti+l) = 0, ICi(V,V)>0
Energy Functional
on a Loop Space
271
Thus for any 0 ^ V G Vc, fc-i
W t O = J>c,(v,iO>o. Q.E.D. By Lemmas 17.2.2 and 17.2.3, we see that the index of Xc on W c is determined by that of Xc =XC\ j - c and Xc is degenerate on Wc if and only if Ic is degenerate on Jc. This proves Lemma 17.2.1.
17.3
Energy Functional on a Loop Space
Let (M, F) be complete. For a point p G M, we denote by Cl(p) the set of piecewise C°° curves c : [0,1] —> M with c(0) = p = c(l). There is a canonical functional -E defined on fi(p) by
£(c) = f[F(c))2dt. Jo
E is is called the canonical energy functional on Q,(p). The topology of fl{p) is the open-compact topology so that E is continuous. Fix a loop c G f2(p) and a partition 0 = to < • • • < tk = 1- Consider a curve c s , —e < s < s passing through c in Q, i.e., c s Gfi(p),
c0=c,
\s\<e.
(17.3)
We assume that cs is C°° in the sense that H(s,t) :— cs(t) is a piecewise C°° map on (— e,e) x [0,1] such that if is C°° on each (—£,£) x \ti,ti+i). Let V:=^-(-,t),
T:=-(s,t).
(17.4)
Note that F(i) := V(0, t) is a piecewise C°° vector field along c with V(0) = V(l) = 0 and T(Q,t) = c(t). Thus V(t) G VVC. We can view W c as the tangent space of fi(p) at c. Define the differential dcE : W c —> R by
4£;(^):=4-Nc,)l| as L
J i s =o
,
T/ = ^ | s = 0 . as
(17.5)
Morse Theory of Loop Spaces
272
We have
s=0
I'V[9T(T,T)} In
L
dt\ 'J
ls=0
2 J [T[gT{V,T)] - gT(V,DTT)}dt\s=o -2 f
g c (V,D c c)ctt
fc-i
2
+ E
{SC(*D (^(*o, «(*D)
- s6(t+) (n*o. c(*+))}.
From the above identity, we see that if c is a C°° geodesic, then dcE{V) = 0,
^ G Wc.
By definition, a loop c G O(p) is called a critical loop of 2? if dcE = 0. L e m m a 17.3.1 A curve c G fl(p) is a critical loop of E if and only if c is a C°° geodesic. Proof. Suppose that c G 0,{p) is a critical loop of E. First, take a piecewise defined function ip on [0,1] such that (a) (p is C°° on each [t{, ij+i], i = 0, • • •, k — 1; (b)
= 0.
Thus D c c = 0, i.e., c is a geodesic on each [it,t i + 1 ]. Fix i 0 with 0 < i0 < k. For an arbitrary vector v G T c ( ti )M, take a piecewise C°° vector field V(t) G W c such that K(t io ) = v and V(*i) = 0 for all i with i ^ i0. Then d c £?(7) = 2 { g 4 ( t - } (Wl c ( i - ) ) - g c ( t + } (y, c(t+)) } = 0. Since v is arbitrary, by Lemma 1.2.4, we conclude that c(tf) Therefore c is a C°° geodesic on [0,1] from O.D.E. theory.
=
c(t~). Q.E.D.
Approximation
of Loop Spaces
273
Let c be a C°° geodesic loop at p. Formally, we define the Hessian d\E of E at c by d2cE(V) = ^
[E(cs)] \S=0,
V £ Wc,
where cs is a curve in 0(p) whose variation field along c is V. By the second variation formula, we have pi
ds2
E(cs)]\s=o
= 2j
{gd(D6V,D6v)
-gi(Ri(V),v)}dt
=
21c(V,V).
From the above identity, we see that the index indcE of E is equal to that of Ic, respectively. indcE = ind(Xc) = ind(c).
17.4
(17.6)
Approximation of Loop Spaces
To apply the Morse theory to the energy functional E on a loop space tt(p), we approximate fi(p) by finite-dimensional manifolds. For a number 6 > 0, let fi'(p) := -ET^-oo.J] = j c £ fi(p), E(c) < o"}. Since B(p, vo) is precompact, there is a positive number i0 > 0 such that the injectivity radius satisfies ix>i0,
Vx e
B(p,V6).
Fix a partition 0 = io < • • • < tk — 1 such that ti+i-ti<-f,
i2 o
0
(17.7)
We denote by Cls(to, • • •, ifc) the set of loops c G fls(p) such that c|[ ti]ti+1 j is a minimizing geodesic. For a loop c e fi(p), £<5 c
() = E /
i*^)] 2 *
Morse Theory of Loop Spaces
274
fc-i
d\c(ti),c(ti+i)j
= y Denote by Es the restriction of E to ils(to, • • •, ifc). By assumption, for any cens(t0,---,tk), E(c) <6, i.e., ^d(c(ti),c(ti+1))2
£(c) = £ ^ — — r J - < 6 -
(17 8)
-
Let N denote the set of (x\, • • •, x^-i) € M x • - • x M such that
gd(^
*
i+1
(179)
**
where XQ — p, Xk = q- Hence, for any (xi, • • •, Xk-\) £ N, d{xi,xi+1) < VA*i+i -U)S
(17.10)
For each 0 < i < k — 1, there is a unique minimizing geodesic Ci from Xi to Xi+\. They form a piecewise C°° geodesic c = lJi=o c* e f&*(*0) • • • ,tk)Define a map $ : fi*(£o, • • •, *fc) -> AT by * ( c ) : = (c(ti),---,c(t f c _i)). By the above argument, we can see that <3? is one-to-one and onto. Since N is a C°° manifold without boundary, we can endow a manifold structure on Qs(t0, • • •, tk) by $ so that $ is a diffeomorphism. Lemma 17.4.1 There is a deformation retraction offl(p) onto Cl (to, • • • ,tk) such that Od (p) is mapped onto Qs(to, • • •, tk) n ft6 (p) for any 5' < 5. Proof. Let c G 0 5 (p). Observe that for any t £ [ti,ti+i], d(c(U),c(tf)
<
L(c\[tiA)
< /"' + 1 F(c)dt
Approximation of Loop Spaces
275
T h u s for any 0 < s < 1, there is a unique minimizing geodesic &l(t), U
t.(.)W-{"•»>• » * * ' * * < • > [ c(i),
if ij(s) < i < tj+i.
Observe t h a t for 0 < s < 1 /•ti(s) r
Jt.
/<^hl2
I \~df)\
l
d(c(U),c(U(s)))
~
(ti+1 - u)s [i:;{s) mdt]2 (U+i - U)s rU{s) / F(c)2dt
< This implies E(*s(c))<E(c).
(17.11)
Notice t h a t $ o = identity and \I>i : ils(p) —> Jl<5(*o, • • •, ifc) such t h a t ^ i = identity on S7* (to, - - •, *&) - Thus \P S is
J(t):=^L{0,t) is a piecewise C°° Jacobi field along c with J ( 0 ) = J ( l ) = 0 such t h a t J is C°° on each [ti: U+i]. T h u s we can identify the tangent space TcQs(to, • • •, £/t) with Jc. Consider Es as a C°° function on the finite-dimensional manifold Qs(to,- • • ,tk)Let c G f2*(p, Q') be a C°° geodesic loop at p. T h e differential dcEs : Jc —> R is given by dcEs(J):=-f\Es(c,j\\ ds L
JI s =o
,
JeJc,
276
Morse Theory of Loop Spaces
where cs G £1 (to, • • • ,tk) with CQ = c and its variation field along c is J. We have fe-i
dcEs(J)
= 2^
{ g . ( ( r ) (Vuc(t-))
- g. (t + } (V,,c(tf)) } =
dcE(J). (17.12)
By Lemma 1.2.4 and (17.12), we see that
dcEs(J)=0,
JeJc
if and only if c(tj) = c(i+). Thus c is a critical point of E5 in fls(to, • • •, tk) if and only if c is a C°° geodesic. The Hessian d%Es on Jc is given by d2cEs(J) = 21C(J, J) = 2XC(J, J) = ^ £ ( J ) ,
J G Jc
(17.13)
where Z c denotes the restriction of Ic onto J c . Define indc_E in a usual way. By Lemma 17.2.1 and (17.13), we see that ind c £* = ind(Zc) = ind(c) = ind(Xc) = ind c £.
(17.14)
Theorem 17.4.2 Suppose that all geodesies in il(p) are non-degenerate. Then £l(p) has the homotopy type of a CW complex with one cell of dimension A for each geodesic c of index A. Proof. Let 5k —> oo. Construct £lSk (to, • • •, tk) as above. Let E6k denote the restriction of E to QSk (to, • • •, tk). By Lemma 17.4.1, there is a deformation from £lSk(p) onto Clih(to,- • • ,tk). Hence, these two spaces are homotopy equivalent to each other. Since c G flSk (p) is a critical point of E if and only if c is a critical loop of ESk, by (17.4) (17.6) and (17.14), indcE6k = ind(c)
(17.15)
For any critical loop c G Q,Sk, since c is non-degenerate by assumption, the index form Ic is non-degenerate on W c . According to Lemma 17.2.1, Ic is non-degenerate on JCl hence d2cE5 is non-degenerate on Tci1Sk(to, • • • ,tk). We conclude that ESk is a Morse function on QSk(to, • • • ,tk)- Notice that {Es-)-1(-oo,S\
=
ils-(t0,---,tk)nils(P)
is always compact for any 0 < 6 < 6k- Applying Proposition 17.1.1 to Edk, we conclude that
Approximation
of Loop Spaces
277
has the homotopy type of a CW complex with one cell of dimension A for each geodesic c G fi*fc(p) of index A. Since fi (p) increases in fl(p) and
n(p) = Un**(p), fc
we can conclude that £l(p) has the homotopy type of a CW complex with one cell of dimension A for each geodesic c G Q(p) of index A. Q.E.D. Theorem 17.4.2 is beautiful, but in reality, geodesic loops in fl(p) might be degenerate. By approximation and Proposition 17.1.2, we obtain the following Theorem 17.4.3
Suppose that all indices of geodesies c G 0,(p) satisfy ind(c) > A.
Then £l(p) has the homotopy type of a CW complex with cells each of dimension > A. Proof. The proof is similar to that of Theorem 17.4.2. Take 5k -> oo. By (17.15), we know that for any geodesic c G QSk(to, • • •, tk) indcESk = ind c £ = ind(c) > A. By Proposition 17.1.2, we conclude that QSk(to, • • • ,ifc) has the homotopy type of a CW complex with cells each of dimension > A. Notice that
n(p) = Un*»(p). k
Thus 0.(p) has the same homotopy type.
Q.E.D.
Chapter 18
Vanishing Theorems for Homotopy Groups
In the previous chapter, we discuss the Morse theory of the canonical energy functional on a loop space of a Finsler space. In this chapter, we will show that under certain curvature conditions, the indices of C°° geodesic loops are bounded from below. Then we prove some vanishing theorems for the homotopy type of Finsler spaces.
18.1
Intermediate Curvatures
By definition, the Ricci curvature is the trace of the Riemann curvature. In order to reveal more relationship between the topological quantities and the geometric quantities, we introduce the intermediate curvatures between the Riemann curvature and the Ricci curvature. Let (M,F) be a Finsler space. For a [k + l)-dimensional subspace V C TXM, define the Ricci curvature Ricy on V to be the trace of the Riemann curvature restricted to V. Ricy is given by fc+i
Ric v (j/) = ^ g y ( ^ R y ( b 1 ) , b ^ ,
yeV,
i=l
where { b i } ^ 1 is an arbitrary orthonormal basis for (V, gy). Set Ricy(O) = 0. We call Ricy the Ricci curvature on V. Ricy is positively homogeneous of degree two on V, Ric v (Aj/) = A 2 Ric v (?/), 279
A > 0, y £ V.
280
Vanishing Theorems for Homotopy
Groups
From the definition, Ric(y) = Ric Tl M(y),
y &TXM.
For a tangent plane V = P
Put TJ-
• f
• (
RiC
v(y)
Ric^fci :=
int mi ^ „ . , , w dimv=fc+iyev F2(y) where the infimum is taken over all (fc + l)-dimensional subspace V C TXM and y G V \ {0}. We have Ric(i) < • • • < ^Ric(fc) < • • • < ^ — j R i c ^ . j ) . Note that Ric ( 1 ) = inf K(P,y), v ;
18.2
(P,y)
Ric ( n _i) =
inf
Ric(y).
F(y)=i
Vanishing Theorem for Homotopy Groups
Now we are ready to study the homotopy type of a Finsler space under certain curvature bounds. Let (M, F) be an n-dimensional Finsler space and E denote the canonical energy functional on a loop space Cl(p) at p £ M. We will show that if Ric( fc) > k for some 1 < k < n - 1, then for any geodesic loop c e fi(p) with L(c) > •n satisfies ind(c) > n — k. On the other hand, we will show that for any simply connected closed reversible Finsler space with 1 < K < 4, every geodesic loop c G fi(p) has length L{c) > 7r. Therefore M is a homotopy sphere. Let (M,F) be a Finsler space and p G M. Consider a geodesic loop c : [0,1] —> M with c(0) = p = c(l). Fix a non-zero piecewise C°° function
Vanishing Theorem for Homotopy
Groups
281
/ : [0,1] -> R with /(0) = / ( l ) = 0. Let Sc(f)
:= IfW;
W parallel along c and
W±c\.
Here W±c means that W(t) is orthogonal to c(t) for all 0 < t < 1 with respect to gc(t)- Denote by E the restriction of E to Sc{f). For a vector field V = fW € 5 C (/), ^ E ( V ) = I C (V, V)=2C
j
{/'(i) 2 - r 2 K ( p ( i ) , c(i))/(*) 2 }cft,
where K(P(i), c(t)) denotes the flag curvature of -P(i) := span{c(£), W"(t)}, C = Sc(W(t),W(t)) and r — F(c(t)) are constants. We always have ind(c) = indcj5 > ind c i?. For further discussion, we need the following elementary lemma. Lemma 18.2.1 For any r > n, there is a piecewise C°° function f : [0,1] ->• R+ with /(0) = / ( l ) = 0 s«c/i that
J f(t)2dt = l,
J
{f'(t)2~r2f(t)2}dt<0.
Proof. Choose a sufficiently small s > 0. Define a piecewise C°° function
{
sin(rt),
if 0 < t <
s i n ( e ) [ l - ( r t - 7 r + e)/2e]
if ^
0
^
< t < *±£
if 2L±£ < i < 1 r — —
For a sufficiently small s > 0, /
{ / ' ( t ) 2 - r 2 / W 2 } ^ = ^sin(£)[-£cos(£) - - £ s i n ( e ) + -sin(e)j < 0.
Normalizing / ,18.2.2 we obtain the desired function. Q.E.D. Proposition Suppose that Ric( fc) > k for some 1 < k < n — 1. Then for any geodesic c G fi(p) wit/i length L(c) > n, ind(c) > n — k.
282
Vanishing Theorems for Homotopy
Groups
Proof. Let r = L(c) > TT and / : [0,1] -» R be as in Lemma 18.2.1. For V = fWe Sc(f) with Si(W, W) = 1, d2cE{V) = 2 ^ {f(tf
r2K(p(t),c(t))f(t)2}dt
-
where P(t) = span{W(t), c(t)} and r = F(c(t)). ( , )c on Sc{f) by
Define an inner product
„i
(fWufWih = J f(t)2em(w1(t),w2(t))dt =
Sc(t)(w1(t),W2(t)).
(18.1)
There is an orthonormal set {/Wj}™",1 for Sc(f) with respect to ( , ) c such that d2cE(fWi) = \i. By (18.1), {Wi(t)}"~i is also an orthonormal set with respect to g<;(t). We may assume that Ai < • • • < A n _i. Let V(t) = span{c(£), Wn-k(t),
• • • , W n _ i ( t ) } C Tc{t)M.
Observe that
£
A, =
i=n—k
£ <%E(fWi) i=n—k
=
21
{kf{tf-r2K\cv(t){c{t))f{t)2}dt
2k J {f'(t)2-r2f(t)2}dt<0. to This implies that at least Xn-k < 0. Thus d2E is negative definite on an (n — fc)-dimensional subspace of Sc(f) <
yV(t)=span{w1(t),---,Wn-k{t)Y We conclude that ind(c) >n — k.
Q.E.D.
According to Proposition 18.2.2, if Ric(fej > k, then the index of geodesic loops c with length L(c) > TT has ind(c) > n — k. However, there
Vanishing Theorem for Homotopy
Groups
283
might be shorter geodesic loops in a Finsler manifold, whose indices are not under control. For a point p e M, let sys(M,p) := inf L(c), c
where the infimum is taken over all geodesic loops c at p. Set sys(M) := inf p 6 M sys(M,p). Clearly, sys(M,p) > 2i p , where i p denotes the injectivity radius at p. By Proposition 18.2.2 and Theorem 17.4.3, we immediately obtain the following Theorem 18.2.3 Let (M, F) be a connected Finsler space. Suppose that for some 1 < k < n — 1, R-ic(fc) > k,
sys(M) > ir.
Then tt(p) has the homotopy type of a CW complex with cells each of dimension > n — k. In particular, 1Ti(M,p) =7Ti_i(f2(p),*) = 0
for 1 < i < n — k. According to Theorem 13.1.2, an upper bound on the flag curvature implies a lower bound on the conjugate radius. More precisely, if the flag curvature satisfies K < A for some A > 0, then CM > TT/VX ( = oo if A = 0). However, the upper bound on flag curvature implies no lower bound on the injectivity radius. A typical example is the torus T 2 = S 1 x S 1 with the standard flat Riemannian product metric. If we shrink one factor, the Gauss curvature always vanishes, but the injectivity radius approaches zero. Theorem 18.2.4 Let (M, F) be a closed simply connected reversible Finsler manifold of dimension n > 3. Suppose that Ric (n _2) > n - 2, Then \M = CM •
1 c M > -j^.
284
Vanishing Theorems for Homotopy
Groups
Proof. We prove it by contradiction. Suppose t h a t \M ¥" CM- T h e n i M < CM by Lemma 12.2.3. It follows from Lemma 12.3.2 t h a t there is a closed geodesic c with L(c) = 2 i M < 2 c M Assume t h a t c : [0,1] —> M is parametrized proportional to arc-length with c(0) = p = c ( l ) . Since M is simply connected, there is a homotopy cs such t h a t Ci = c and CQ = p. We may assume t h a t cs is piecewise C ° ° . Consider t h e canonical energy functional E on t h e loop space £l(p). We know t h a t t h e critical points of E are exactly the smooth geodesic loops at p. It follows from Proposition 18.2.2 t h a t if y/E(
sup
L(cs).
0<s
Approximate fi*(p) by a finite-dimensional manifold ils(t0, • • • ,tk), where 0 = to < • • • < tk = 1 is a partition of [0,1] t h a t is defined as in (17.7). There is a deformation from fl5(p) onto ft6(to, • • • ,tk) (Lemma 17.4.1). We denote by E5 t h e restriction of E to Qs(to, • • •, tk)- A geodesic loop c £ 0,s(p) is a critical point of E if and only if c e fid (to, • • •, tk) is a critical point of Es. T h e index of E at c £ Q(p) is exactly t h e index of Es at cr G SI 0 (io, • • •, tk)- Take a number <5' with 8 > 6' > sup L(c s ) 0<s
such t h a t 5' is not a critical value of .E"5. According to Proposition 17.1.3 there exists a curve c* in Qs(to, • • •, tk) from CQ = p to c\ = c such t h a t L(c*s) = ^Es(c*s)
< max(2i M ,TT) + 2e < 2 c M - 2e,
where e is a small positive number. We make the following A s s e r t a t i o n . Each c* can be lifted to a loop in B C M _ e ( p ) C TpM
by exp .
Vanishing Theorem for Homotopy
Groups
285
We grant this assertation for a moment. Then c\—c can be lifted to a loop in TpM by exp p . This is impossible, because that the pre-image of c under exp p in the ball B CM _ £ (p) C TpM is a straight line issuing from the origin. Proof of Assertation. Let / denote the set of s € [0,1] for which c* can be lifted to a loop in B CM _ £ (p) by exp p . (i) Suppose that c*o can be lifted to a loop cSo in B CM _ £ (p) by exp . Since exp p is a local difFeomorphism on B CM (p), for s sufficiently close to sot c* can be lifted to a curve c* in B CM _ £ (p). Therefore I is open. (ii) Let Si -4 s0 be such that c*. can be lifted to a loop cSi in B C M _ e (p). Note that L(cSi) < 2CM — 2e. It follows from the Ascoli theorem that cSi sub-convergent to a loop cSo in BCM_e(p) C BCM(p). The limit curve cSo is the lift of cSg. Since L(cSo) < 2cjy — 2e, cSg must be contained in B CM _ e (p). Thus I is closed. (i) and (ii) imply that / = [0,1]. This proves the assertation. Q.E.D. Theorem 18.2.5 Let (M, F) be a complete simply connected reversible Finsler manifold of dimension n > 3. Suppose that for some 1 < k < n — 2, 1 R-lC(fc) > k,
CM >
-7T.
Then £l(p) has the homotopy type of a CW complex with cells each of dimension > n — k. In particular, Wi(M,p) = 0 , I < i < n — k. Proof. By Theorem 18.2.4, we see that 1 l M = CM
> -7T.
Thus every geodesic loop at p has length greater than 7r. It follows from Proposition 18.2.2. that every geodesic loop c at p has index ind(c) >n — k. By Theorem 18.2.3, we can conclude that £l(p) has the homotopy type of a CW complex with cells each of dimension > n — k. Q.E.D. According to the Bonnet-Myers theorem (Theorem 13.3.1), if M is a positively complete simply connected Finsler surface with K > 1, then Diam(M) < it and M is diffeomorphic to S 2 . By this observation, Theorems 13.1.2 and 18.2.5, we immediately obtain the following
286
Vanishing Theorems for Homotopy
Groups
Theorem 18.2.6 (Homotopy Sphere Theorem). Let (M,F) be a complete simply connected reversible Finsler manifold of dimension n > 2. Suppose that one of the following conditions is satisfied (a) Ric(fc) > k, cM > |TT, (b) Ric (fc) >k, K < 4, where k = [j]. Then M is a homotopy sphere. J. Kern proved a diffeomorphism sphere theorem for almost Riemannian Finsler spaces of pinched curvature. See [Kel][Ke2] for details. Using Toponogov's comparison theorem [To] for Riemannian spaces, we can prove the following homeomorphism sphere theorem in a much more direct way. Theorem 18.2.7 If (M,g) is a simply connected closed Riemannian manifold with 1 < K < 4, then M is homeomorphic to a sphere. This celebrated theorem is the result of efforts by several people: Rauch, Berger [Bgl] [Bg2], Toponogov, and Klingenberg [Kll][K12]. See [ChEb] [Bg3] [Pe3] for further developments.
18.3
Finsler Spaces of Positive Constant Curvature
There are infinitely many non-reversible Finsler metrics on the n-sphere with positive constant curvature. These examples are given in Section 9.2. In this section, we will discuss some basic geometric and topological properties of Finsler spaces of positive constant curvature. First, we prove the following Theorem 18.3.1 Let (M,F) be a positively complete Finsler space with K = 1. Then for any point p e M, there is a point q € M such that every geodesic c(t). 0 < t < oo, with c(0) = p passes through q at t = n, i.e., C(TT)
= q.
Proof. Let exp p : TpM -> M denote the exponential map at p. By Theorem 13.1.2, exp p is non-singular on B ^ p ) C TpM. Take any C°° curve f : (—e, E) —>• S P M and consider the geodesic variation H : (— e, e) x [0, n] —• M defined by
H(s,t):=expM(s)\-
Finsler Spaces of Positive Constant
Curvature
287
For each fixed s, t h e variation field BH J.(t):=—(s,t),
0
is a Jacobi field along t h e unit speed geodesic cs(t) := exp,j££(s)], 0 < t < n. Js is C ° ° on (0,7r]. By Lemma 11.2.2, Js is C°° along cs(t), t > 0, satisfying Js(0)=0,
DesJs(0)=^(S).
Further, J s ( i ) is perpendicular t o cs(t) with respect to gc,(t)- Since K = 1, the Jacobi field J 5 satisfies n6sD6ajs(t)
+ js(t) = o.
Let JB S denote t h e parallel vector field along c s with ES(Q) = £(s). We obtain t h a t Js(t):=
(sint)Es(t)
In particular, we have BH
-^-(s,n)
OS
= J S (TT) = (sin7r)i? s (7r) = 0 .
Thus exp x (7r£(s)) = H(s,n)
= {point},
\s\ < e.
Since £(s) is an arbitrary curve in SPM, we conclude t h a t there is a point q 6 M such t h a t expp[7r£] =q, This proves t h e theorem.
V£ <E S P M . Q.E.D.
Now let us look a t t h e Bao-Shen metrics Fk = a.k + Pk on S 3 . See Example 9.3.1 above. Fk has t h e following properties (a) K = 1; (b) S = 0;
(c) VolFt(S3) = Vol(S3).
288
Vanishing Theorems for Homotopy
Groups
Fix an arbitrary point p G S 3 . By Theorem 18.3.1 and (a), there is a point q G S 3 such that exp p (7r0 = q,
V£ G S P S 3 .
By the volume comparison theorem, for any r > 0, Vol F f c (B(p,r))
(18.2)
where a^(r) denotes the volume of the metric ball of radius r in S . From the proof of Theorem 16.1.1, we see that the equality in (18.2) holds if and only if B(p,r) C Vp, that is, the injectivity radius i p > r. By the Bonnet-Myers theorem (Theorem 13.3.1), we know that Diam < n. Thus B{p, TT) = M. By (c). we obtain that V o l F f c ( B ( p , 7 r ) ) = V o l F f c ( S 3 ) = V o l ( S 3 ) = <7 3 (7T).
We conclude that i p > IT. Therefore, d(p, q) — TT; all geodesies issuing from p pass through q; and geodesies from p to q are minimizing. Now we consider a complete reversible simply connected Finsler space ( M , F ) with K = 1. By the Bonnet-Myers theorem (Theorem 13.3.1), Diam(M) < TT. By the Cartan-Hadamard theorem (Theorem 13.1.2), CM > TT. By Theorem 18.2.4, \M = CM- We conclude that IM = CM = Diam(M) = -K. Fix a point p G M. By Theorem 18.3.1, there exists a point q G M such that 9
= expp(7r£),
V£eSpM,
all geodesies in M are closed of length 27r. We have proved the following Theorem 18.3.2 ([Shl])Le£ (M,F) be a complete simply connected reversible Finsler space of constant curvature K = 1. Then M is diffeomorphic to S n and iM = CM = Diam(M) ~ IT. Further, all geodesies are closed with perimeter 2TT.
Finsler Spaces of Positive Constant Curvature
289
So far we have not found any non-Riemannian reversible Finsler metrics on S" with K = 1. We know that any Landsberg metric defined on an open subset U C Rra must be Riemannian if it has constant flag curvature K = A / 0 (Theorem 9.1.1). This is a local result. We conjecture that every reversible Finsler metrics on S n with K = 1 and S = 0 must be Riemannian.
Chapter 19
Spaces of Finsler Spaces
So far we have considered "individual" Finsler spaces. Even for this purpose it can be very helpful to view the space as a member of a larger collection of Finsler metrics. A simple example of this is the collection of Minkowski spaces. Here we present some of the fundamental work of Gromov [GLP][Gr4] on collections of Finsler spaces. In particular, we will discuss the so-called precompactness and compactness theorems. We expect that there would be further developments in future.
19.1
Gromov-Hausdorff Distance
In the beginning of 80's, M. Gromov [GLP] introduced a generalized Hausdorff distance between two metric spaces. His notion leads to great developments in Riemann geometry in 80's and 90's. We begin with the classical Hausdorff distance on the collection of subsets in a metric space. Then we generalize it to a "distance function" on the collection of compact metric spaces. For the sake of simplicity, we always assume that metrics are reversible throughout this chapter. Let (X, d) be a metric space. For subsets A,B Hausdorff distance dn between A and B in X by d${A,B)
C X, we define the
:= inf {e, A c Ue{B) and B c UC{A)\
where U£(A) := {x € X, d(x,A) < e] denotes the £-neighborhood of A in X. 291
Spaces of Finsler Spaces
292
Let A and B be compact metric spaces. Endow X — A II B with the obvious metric dx on A and B and dx(a,b)
:=maxJDiam(^),Diam(B)|,
a£A,b€B.
(19.1)
T h e natural embeddings of A and i? into X = AllB are isometries. Thus, we can always view A and B as subsets in a metric space. For arbitrary metric spaces A and B, the Gromov-Hausdorff distance dcH between A and B is defined by
dGH(A,B)
:=infd%(f(A),f(B)),
where the infimum is taken over all metric spaces X and isometric embeddings / : A —• X and g : B -» X . By the natural embeddings of A and i? into X = All B with the above metric dx in (19.1), we obtain t h a t dGH(A,B)
< m a x J D i a m ( A ) , D i a m ( B ) | < oo.
Let M. denote the set of all isometric classes of compact metric spaces. Proposition 19.1.1 metric on the isometric is complete.
([Pe2]) The Gromov-Hausdorff distance dcH is a classes of compact metric spaces. Moreover, {M,doH)
T h e proof is quite technical, but elementary. According to Proposition 19.1.1, for compact metric spaces A and B, dcii(A,B) = 0 if and only if A is isometric to B. This is not true for non-compact metric spaces. For example, for the set Q of all rational numbers,
dG* ([o,i],Qn[o,i])=o. To have a better understanding on the Gromov-Hausdorff distance, we introduce the notion of c-nets. Let (X, d) be a metric space. For e > 0, a subset A of X is called an e-net of X if d{x, A) := inf d(x, a) < e,
Vx e X.
We denote an e-net in X by Nx. Note t h a t the set Q n [0,1] is an £-net in [0,1] for any e > 0. But Q n [0,1] is not isometric to [0,1]. From the definition, we see t h a t doH{Nx,X) < e. T h u s we can estim a t e t h e Gromov-Hausdorff distance between two metric spaces by their e-nets. T h e following lemma is very useful.
Precompactness
Theorem
293
Lemma 19.1.2 Let A,B be compact metric spaces. If there are e-nets N* = {xiHU C A, N^ = {zi}^=1 C B such that \dA(xi,xi)-dB(zi,zj)\ ThendGH{A,B)
< e.
< 3e.
Proof. We endow the disjoint union X = A II B with the obvious metric dx on A and B and dx(x,z)
:= m i n j d ^ a : , a:,) + d B (z, zt) +e\,
x £ A,z £ B.
We see that A C C/3e(5) and B C f73e(A). Hence d Gff (A, B) < 3e. Q.E.D.
19.2
Precompactness Theorem
The space of all isometric classes of compact metric spaces equipped with the Gromov-Hausdorff distance is a complete metric space. However this metric space is not compact. In applications, for a given class of metric spaces, we want to know if it is (pre-)compact.
Let (X, d) be a compact metric space. For an £ > 0, define Cov(X, e) as the minimal number of closed £-balls needed to cover X and Cap(X, e) as the maximal number of disjoint £-balls in X. Clearly, Cov(X,2£) < Cap(X,£). Further, if dGH(Xi,X^)
< S, then for any e > 0,
COV(XI,E)
>
Cov(X 2 ,e + 2<5),
Cap(Xi,e)
>
C&p(X2,e + 25).
Spaces of Finsler Spaces
Theorem 19.2.1 ([GLP][Gr4]). Let C C M be a subset. The following are equivalent: (1) C is precompact. (2) There is a function N : (0, a] —• (0, oo) such that Cap(X,e) < N(e),
VX € C, e £ (0, a).
(3) There is a function N : (0, a/2] —• (0, oo) such that Cov(X,£) < N(e),
VX e C , e€ (0,a/2).
See [GLP] [Pe2] for the details. Given n e Z , A G R, S > 0 and d > 0, let ^-"(n, A, <5, d) denote the space of all isometric classes of n-dimensional closed reversible Finsler spaces satisfying the following bounds Ric>(n-1)A,
|S|<(n-l)<5,
Diam < d.
We have the following Theorem 19.2.2
([Sh2]) The subspace F{n, A, S, d) is precompact in (M, dan)-
Proof. Let X € T(n, X,S,d). Take a maximal set of disjoint £-balls B(xi, s), i = 1, • • • m, in X. Assume that B(xio,e) has the minimal volume among these £-balls. Then 771
mWo\F{B{xio,e))
<
^VoMBfo.e)) i=l
Precompactness
Theorem
295 III
= Vol F (|J B(Xi,ej) t=i
<
Vol F (M)
=
Vo\F(B(xio,d)).
Then by the volume comparison theorem, n-l
Vol F (B io ,d) Jo [e 6t s A (i) dt Cap(X,e) = m< 'i,? u < " ^ ^=T~ =: JV(£) V o l H ^ i . ^ ) ) - ; - ^ ^ - ^ By Theorem 19.2.1, we conclude that J-{n, A, 5, d) is precompact in (M, dcii)Q.E.D. Remark 19.2.3 Let lZ(n, A, d) denote the subspace of all isometric classes of n-dimensional closed Riemannian spaces (M, g) satisfying the following bounds Ric > (n - 1)8,
Diam(M) < d.
Then lZ(n,\,d) is precompact in {Ai,doH)- This is the so-called Gromov precompactness theorem in Riemann geometry. See [GLP][Gr4].
Example 19.2.1 Let (V,F) be a Minkowski space and S = F~X{1) denote the indicatrix of F. Let g and C denote the induced Riemannian metric and the Cartan torsion on S. See (14.17). From the definition of C, we have g(c{C(u,«),«;),
z) = g(c(u, v), C(w, z)).
Define g(c{v,v),C{v,v)} ||C|| := sup vers
g(v,v)
Assume that for some 6 < 1, the Cartan torsion satisfies ||C|| < 6.
(19.2)
296
Spaces of Finsler
Spaces
Let R„(w) denote the Riemann curvature of g. According to [Kil][Kaw], Rv(u) = g(v, v)u - g(u, v)v + C(v, C(u, v)) - C(u, C(v, v)).
(19.3)
For any orthonormal set {u, v}, the sectional curvature K of P = span{w, v) satisfies K(P)
=
g (R„ («),«)
=
g(c(u, v),C(u,«))
-
^ ( C ( M , U), C(V, V ) )
>
l-j(c(tt,u),C(i;,T)))
>
1 — J 2 > 0.
+ 1
This implies the diameter of (S, g) is bounded from above.
Diam(S,£) <
7T
VT^s*'
Let 7V((n, 5) denote the collection of the Riemannian metrics g on S induced by a Minkowski norm satisfying (19.2). By the Gromov precompactness theorem, AA(n,8) is precompact in the Gromov-Hausdorff topology. (j Since the subclass T(n,X,5,d) is precompact in (M.,dQn), a natural question arises: how many homotopy types among Finsler spaces in T(n, A, 6, d) ? Unfortunately, there are infinitely many homotopy types among Finsler spaces in F{n,\,5,d). If we pick up only Finsler spaces in T{n,\5,d), whose metric balls of a fixed radius having simple topology, then the topological type of the space can be controlled. A contractibility function p(e) : [0, r) —> [0, oo) is a function which satisfies (a) p(0) = 0, (b) p(e) > e, (c) p{e) -» 0 as e -* 0, (d) p is non-decreasing. A metric space (X, d) is said to be LGC(p) for some contractibility function p(e) : [0, r] ->• [0, oo) if for every e e [0, r] and x e X, the e-ball B{x,e) is contractible inside B(x,p(e)).
Precompactness
Theorem
297
Given a function N(e) : [0,a] -> [0,oo) and a contractibility function p(e) : [0,r) —• [0,co). Let C(N(e),p(e)) denote the subset of all isometric classes of compact metric spaces X G M such that X is LGC(p) and Cov(X, e) < iV(e) for all e G [0, a]. We have the following theorem due to P. Petersen. Theorem 19.2.4 ([Pel][Pe2]) Suppose that lim£_>o enN{e) < 00. Then the subclass C(N(e), p(e)) is compact in (M, dan) and contains only finitely many homotopy types. Below is a consequence of our volume comparison theorem. Given a contractibility function p(e) : [0,r] —>• [0,oo). Let J-{n, X,S, d,p(e)) denote the subclass, of J-(n, X,6, d) of Finsler spaces which in addition are also GLC(p). Proposition 19.2.5 topy types.
T{n, A, 6, d, p(e)) contains only finitely many homo-
Our next task is to find bounds on the geometric quantities under which a Finsler space is LGC(p) for some contractibility function depending only these bounds. In [GrPe], Grove-Petersen proved the following result: given any n, A, D and v, there exist two numbers r > 0 and R > 1 such that if a compact Riemann n-manifold (M, g) satisfies the bounds K > A,
Diam < D,
Vol > v,
then (M, g) is LGC(p) for p(e) = Re,
0 < e < r.
The graduate textbook by P. Petersen [Pe3] is an excellent book for related results on Riemannian spaces. For a Finsler space, the lower sectional curvature bound is replaced by a lower flag curvature bound. However, an additional bound on certain non-Riemannian curvatures is required in order to obtain a similar contractibility function.
Bibliography
M. Abate and G. Patrizio, Finsler metrics-a global approach, with applications to geometric function theory, Lecture Notes in Mathematics 1591, SpringerVerlag, Berlin, 1994. T. Aikou, Some remarks on the geometry of tangent bundles of Finsler spaces, Tensor, N. S. 52(1993), 234-242. P. Antonelli, R. Ingarden and M. Matsumoto, The theory of sprays and Finsler spaces with applications in physics and biology, Kluwer Academic Publishers, 1993. J.C. Alvarez Paiva, Some Problems in Finsler Geometry, preprint. J.C. Alvarez Paiva and E. Fernandes, Fourier Transforms and the HolmesThompson volume, International Mathematical Research Notices 19 (1999),1031-1042. L. Auslander, On curvature in Finsler geometry, Trans. Amer. Math. Soc. 79(1955), 378-388. H. Akbar-Zadeh, Sur les espaces de Finsler a. courbures sectionnelles constantes, Bull. Acad. Roy. Bel. CI, Sci, 5e Serie - Tome LXXXIV (1988) 281-322. D. Bao and S. S. Chern, On a notable connection in Finsler geometry, Houston J. Math. 19(1)(1993), 135-180. D. Bao and Z. Shen, On the volume of unit tangent spheres in a Finsler space, Results in Math. 26(1994), 1-17. D. Bao and Z. Shen, Finsler metrics of constant curvature on the Lie group S3, preprint. D. Bao, S. S. Chern and Z. Shen, An Introduction to Riemann-Finsler Geometry, Springer-Verlag, 2000. D. Bao, S.S. Chern and Z. Shen, Rigidity issues on Finsler surfaces, Rev. Roumaine Math. Pures Appl. 42(1997), 707-735. D. Bao, S.S. Chern and Z. Shen (ed.), Finsler Geometry, Contemporary Math. 196(1996). G. Bellettini and M. Paolini, Anisotropic motion by mean curvature in the context 299
300
Bibliography
of Finsler geometry, Hokkaido M a t h . J. 25(1996), 537-566. P. Berard, Lectures on spectral geometry, Lecture Notes in M a t h . , vol. 1 2 0 7 , Springer-Verlag, Berlin a n d New York, 1986. M. Berger, Les varietes Riemanniennes j-pincees, A n n . Scuola N o r m . S u p . P i s a (III) 14(2) (1960). M. Berger, Sur les varietes a courbure positive de diametre minimum, Comment. M a t h . Helv. 3 5 ( 1 ) (1961). M. Berger, Riemannian geometry during the second half of the twentieth century, I H E S preprint, 1997. R. Bishop a n d R. C r i t t e n d e n , Geometry of manifolds, Academic Press, 1964. R. B r y a n t , Finsler structures on the 2-sphere satisfying K = 1, Finsler Geometry, C o n t e m p o r a r y M a t h e m a t i c s 1 9 6 , A m e r . M a t h . S o c , Providence, RI, 1996, 27-42. R. B r y a n t , Projectively flat Finsler 2-spheres of constant curvature, Selecta M a t h . , New Series, 3(1997), 161-204. F . Brickell, A new proof of Deicke's theorem on homogeneous functions, P r o c . of A M S 16(1965), 190-191. H. B u s e m a n n a n d J. P. Kelly, Projective geometry and projective metrics, Academic P r e s s , New York, 1953. H. B u s e m a n n , Intrinsic area, A n n . of M a t h . , 4 8 ( 1 9 4 7 ) , 234-267. H. B u s e m a n n , The foundations of Minkowskian geometry, C o m m e n t . M a t h . Helvet. 24(1950), 156-187. H. B u s e m a n n , The geometry of geodesies, Academic Press, New York, 1955. H. B u s e m a n n a n d W . Mayer, On the foundations of the calculus of variations, T r a n s . A M S . 4 9 ( 1 9 4 1 ) , 173-198. L. Berwald, Untersuchung der Krummung allgemeiner metrischer Rdume auf Grund des in ihnen herrschenden Parallelismus, M a t h . Z. 2 5 ( 1 9 2 6 ) , 40-73. L. Berwald, Parallelubertragung in allgemeinen Raumen, A t t i Congr. I n t e r n . M a t . Bologna 4(1928), 263-270. E. C a r t a n , Les espaces de Finsler, Actualites 79, Paris, 1934. J. Cheeger a n d D. Ebin, Comparison theorems in Riemannian geometry, N o r t h Holland P u b l i s h i n g Company, 1975. J. Cheeger a n d D. GromoU, The splitting theorem for manifolds of nonegative Ricci curvature, J. Differential G e o m e t r y 6(1971), 119-128. J. Cheeger a n d D. GromoU, On the structure of complete manifolds of nonnegative curvature, A n n . of M a t h . 9 6 , 413-443. S. S. C h e r n , On the Euclidean connections in a Finsler space, P r o c . N a t i o n a l Acad. S o c , 29(1943), 33-37; or Selected P a p e r s , vol. I , 107-111, Springer 1989. S. S. C h e r n , On Finsler geometry, C. R. Acad. Sc. Paris 3 1 4 ( 1 9 9 2 ) , 757-761. P. Dazord, Proprietes globales des geodesiques des espaces de Finsler, These, Universite de Lyon, 1969. P. Dazord, Tores Finsleriens sans points conjugues, Bull. Soc. M a t h . France, 9 9 ( 1 9 7 1 ) , 171-192.
Bibliography
301
A. Deicke, Uber die Finsler-Rdume mil Ai = 0, Arch. Math. 4(1953), 45-51. Q. Ding, A new Laplacian comparison theorem and the estimate of eigenvalues, Chin. Ann. of Math. 15B: 1(1994), 35-42. J. Douglas, The general geometry of paths, Ann. of Math. 29(1927-28), 143-168. C. Duran, A volume comparison theorem for Finsler spaces, Proc. of AMS 126(1998), 3079-3082. E.D. Dzhafarov and H. Colonius, Fechnerian metrics in unidimensional and multidimensional stimulus, Psychological Bulletin and Review, 6(1999), 239268. D. Egloff, Some new developments in Finsler geometry, Ph. D. thesis, Aus dem Institut fur Mathematik, Universitat Freiburg (Schweiz), 1995. D. Egloff, Uniform Finsler Hadamard manifolds, Ann. Inst. Henri Poincare, 66(1997), 323-357. P. Finsler, Uber Kurven und Flachen in allgemeinen Raumen, (Dissertation, Gottingan, 1918), Birkhauser Verlag, Basel, 1951. P. Foulon, Estimation de I'entropie des systemes lagrangiens sans points conjuges, Ann. Inst. Henri Poincare 57(2)(1992), 117-146. P. Foulon, Entropy rigigity of Anosov flows in dimension 3, preprint. P. Funk, Uber Geometrien bei denen die Geraden die Kiirzesten sind, Math. Ann. 101(1929), 226-237. P. Funk, Uber zweidim,ensionale Finslersche Raume, insbesondere iiber solche mit geradlinigen Extremalen und positiver konstanter Kriimmung, Math. Z. 40(1936), 86-93. S. Gallot, Theoremes de comparaison entre varietes et entre spectres et applications, Actes du convegno-studio du C. N. R. (Rome 1986), soumis a Asterisque. Y. Ge and Z. Shen, Eigenvalues and eig enfunctions of metric measure manifolds, Proc. London Math., to appear. W. Gu and Z. Shen, Levy concentration of metric measure manifolds, Fundamental Theories of Physics, 109(1999), 169-178. M. Gromov, J. Lafontaine and P. Pansu, Structures metriques pour les varieties Riemanniennes, Cedic-Fernand Nathan, Paris (1981). L. W. Green, Auf Wiedersehensfldchen (in English), Ann. of Math. 78(1963), 289-299. M. Gromov, Filling Riemann manifolds, J. Diff. Geom. 18(1983), 1-147. M. Gromov, Sign and geometric meaning of curvature, Rend. Sem. Mat. Fis. Milano 61(1991), 9-123. M. Gromov, Volume and bounded cohomology, IHES Publ. Math. 56(1983), 213307. M. Gromov, Metric structures for Riemannian and non-Riemannian spaces, Birkhauser, 1999. M. Gromov, Paul Levy isoperimetric inequality, IHES preprint (1980). Published in [Gr4]. M. Gromov, Dimension, non-linear spectra and width, Lecture Notes in Mathe-
302
Bibliography
matics, 1317(1988), 132-185. M. Gromov and V. D. Milman, A topological application of the isoperimetric inequality, Amer. J. Math. 105(1983), 843-854. K. Grove and P. Petersen V, Bounding homotopy types by geometry, Ann. of Math. 128(1988), 195-206. M. Hashiguchi and Y. Ichijyo, On some special (a, (3) metrics, Rep. Fac. Sci. Kagoshima Univ. 8(1975), 39-46. E. Heintze and H. Karcher, A general comparison theorem with applications to volume estimates for submanifolds, Ann. Sci. Ec. Norm. Super. 11(1978), 451-470. H. Hrimiuc and H. Shimada, On the L-duality between Finsler and Hamilton manifolds, Nonlinear World 3(1996), 613-641. Y. Ichijyo, Finsler spaces modeled on a Minkowski space, J. Math. Kyoto Univ. 16(1976), 639-652. M. Ji and Z. Shen, On strongly convex indicatrices in Minkowski geometry, Canadian Math. Bull., to appear. A. B. Katok, Ergodic properties of degenerate integrable Hamiltonian systems, Izv. Akad. Nauk SSSR 37(1973), [Russian]; Math. USSR-Izv. 7(1973), 535-571. A. Kawaguchi, On the theory of non-linear connections II, Theory of Minkowski spaces and of non-linear connections in a Finsler space, Tensor, N. S. 6(1956), 165-199. J. Kern, Fastriemannsche Finslerschen metriken, Manuscripta Math. 4(1971), 285-303. J. Kern, Das pinchingproblem in fastriemannschen Finslerschen mannigfaltigkeiten, Manuscripta Math. 4(1971), 341-350. S. Kikuchi, Theory of Minkowski space and of non-linear connections in Finsler space, Tensor, N.S. 12(1962), 47-60. S. Kikuchi, On the condition that a space with (a, f3)-metric be locally Minkowskian, Tensor, N. S. 33(1979), 242-246. W. Klingenberg, Contributions to Riemannian geometry in the large, Ann. Math. 69(1959). W. Klingenburg, Uber Riemannische Mannigfaltigkeiten mit positiver Krummung, Comment. Math. Helv. 35(1961). G. Landsberg, Uber die Totalkriimmung, Jahresberichte der deut Math. Ver. 16(1907), 36-46. G. Landsberg, Uber die Krummung in der variationsrechung, Math. Ann. 65(1908), 313-349. P. Levy, Problemes Concretes d'Analyse Fonctionelle, Gauthier-Villars, Paris, 1951. M. Matsumoto, Foundations of Finsler Geometry and special Finsler Spaces, Kaiseisha Press, Japan 1986. M. Matsumoto, Theory of curves in tangent planes of two-dimension Finsler spaces, Tensor, N. S. 37(1982), 35-42. M. Matsumoto, Randers spaces of constant curvature, Rep. Math. Phys. 28(1989),
Bibliography
303
249-261. M. Matsumoto, Projective changes of Finsler metrics and projectively flat Finsler spaces, Tensor, N. S. 34(1980), 303-315. M. Matsumoto, On Finsler spaces with Randers metric and special forms of important tensors, J. Math. Kyoto Univ. 14(1974), 477-498. M. Matsumoto and X. Wei, Projective changes of Finsler spaces of constant curvature, Publ. Math. Debrecen 44(1994), 175-181. John Milnor, Morse Theory, Princeton University Press, 1963. M. Meyer and A. Pajer, On Santalo's inequality, Geometric aspects of functional analysis, Springer Lecture Notes in Mathematics 1376, 261-263. S. Numata, On Landsberg spaces of scalar curvature, J. Korea Math. Soc. 12(1975), 97-100. K. Okubo, Some theorems on Ss(K) metric spaces, Rep. on Math. Phys., 16(1979), 401-408. T. Okada, On models of projectively flat Finsler spaces of constant negative curvature, Tensor, N. S. 40(1983), 117-123. G. Paternain, Finsler structures on surfaces with negative Euler characteristic, Houston J. Math. 23(1997), 421-426. P. Petersen, A finiteness theorem for metric spaces, J. Differential Geom. 31(1990), 387-395. P. Petersen, Gromov-Hausdorff convergence of metric spaces, Proc. Symp. Pure Math. 54, No. 3 (1993), 489-504. P. Petersen, Riemannian geometry, Graduate texts in mathematics, 171, Springer, 1997. A. Rapcsak, Uber die bahntreuen Abbildungen inetrisher Raume, Publ. Math. Debrecen, 8(1961), 285-290. W. Rinow, Die Innere Geometrie der Metrischen Raume, Springer, New York, 1961. H. Rund, The Differential Geometry of Finsler Spaces, Springer, 1959. Z. Shen, Finsler spaces of constant positive curvature, In: Finsler Geometry, Contemporary Math. 196(1996), 83-92. Z. Shen, Volume comparison and its applications in Riemann-Finsler geometry, Advances in Math. 128(1997), 306-328. Z. Shen, On Finsler geometry of submanifolds, Math. Ann. 311(1998), 549-576. Z. Shen, Curvature, distance and volume in Finsler Geometry, preprint (1997). Z. Shen, Conjugate radius and positive scalar curvature, Math. Z. (to appear) Z. Shen, On projectively related Einstein metrics in Riemann-Finsler geometry, Math. Ann. (to appear). Z. Shen, On R-quadratic Finsler spaces, Publicationes Math. Debrecen, (to appear). Z. Shen, Funk metrics and R-flat sprays, preprint (2000). Z. Shen, Differential Geometry of Spray and Finsler Spaces, Kluwer Academic Publishers, 2001. M. Spivak, A comprehensive introduction to differential geometry, Vol III, Publish
304
Bibliography
or Perish, Inc. Berkeley, 1979. C. Shibata, H. Shimada, M. Azuma and H. Yosuda, On Finsler spaces with Randers' metric, Tensor, N. S. 31(1977), 219-226. Z. Szabo, Positive definite Berwald spaces (Structure theorems on Berwald spaces), Tensor, N. S. 35(1981), 25-39. A.C. Thompson, Minkowski Geometry, Encyclopedia of Math, and Its Applications, Vol. 63, Cambridge Univ. Press, Cambridge, 1996. V. Toponogov, Riemannian spaces with curvature bounded below, Uspehi Mat. Nauk. 14(1959). O. Varga, Die Krummung der Eichfldche des Minkowskischen Raumes und die geometrische Deutung des einen Krummungs-tensors des Finslerschen Raumes, jour Abh. Mth. Sem. Univ. Hamburg 20(1955), 41-51. J. H. C. Whitehead, Convex regions in the geometry of paths, Quart. J. Math. Oxford Ser. 3 (1932), 33-42. Y. L. Xin, Harmonic maps of bounded symmetric domains, Math. Ann. 303(1995), 417-433. H. Yasuda and H. Shimada, On Randers spaces of scalar curvature, Rep. on Math. Phys. 11(1977), 347-360. W. Ziller, Geometry of the Katok examples, Ergod. Th & Dynam. Sys. 3(1982), 135-157 .
Index
amenable, 258 angular form, 10 angular metric, 124 Busemann-Hausdorff volume
Riemann, 96, 97 cut-domain, 183 cut-locus, 183 , 22
defining function, 8 degenerate geodesic, 269 distance function, 1, 45 distortion, 117 divergence, 21
Cartan tensor, 81 Cartan torsion, 108 Cheeger constant, 63 Chern connection, 85 Chern curvature, 112 Chern curvature tensor, 127 complete, 139, 175 conjugate radius, 179 conjugate value, 179 connection, 85 contractibility function, 296 convex domain, 241 convex hypersurface, 216 convex value, 241 covariant derivative, 85 critical loop, 272 curvature sectional, 132 Chern, 112 Gauss, 98 Landsberg, 112 mean, 219 normal, 213 Ricci, 98, 201
E-curvature, 118 eigenfunction, 210 eigenvalue, 210 Einstein metric, 151 energy functional, 210, 271 Euclidean metric, 2 Euclidean norm, 2 Euclidean space, 2 Euclidean volume, 20 Euclidean volume form, 20 expansion distance, 64 exponential map, 166 Fchner function, 14 Finsler volume form, 31 induced, 31 Finsler m space, 19 Finsler metric, 12 Finsler space, 12 305
306
first eigenvalue, 58 flag curvature, 98 flat metric, 2 flat metric space, 6 fundamental tensor, 81 Funk metric, 3, 15, 115 Gauss curvature, 98 geodesic coefficients, 78 geodesic curvature, 77 geodesic flow, 90 geodesic variation, 95 gradient, 41 Gromov simplicial norm, 256 Gromov simplicial volume, 256 Gromov-Hausdorff distance, 292 Hadamard space, 197 Hausdorff distance, 291 Hausdorff measure, 22 Hessian, 207, 212, 221 Hilbert form, 26, 91 Holmes-Thompson volume form, 27 index form, 180, 269 index of a function, 266 injectivity radius, 183 isoperimetric constant, 62 isoperimetric function, 56 isoperimetric profile, 56 Jacobi equation, 167 Jacobi field, 96 Klein metric, 5, 16, 142 Landsberg tensor, 84 Landsberg curvature, 112 Landsberg metric, 113 Laplacian, 209, 212, 220, 221, 243 Legendre transformation, 37, 51 length structure, 5 Levi-Civita connection, 87, 131 Levi-Civita connection forms, 87
Index Levi-Civita connection, 86 loop space, 271 mean covariation, 118 mean curvature, 219, 243 mean Landsberg curvature, 116 mean tangent curvature, 118 metric, 1 Einstein, 151 Finsler, 12 Funk, 15 Klein, 5, 16 Minkowski, 6 path, 5 Riemannian, 13 metric measure space, 19 metric space, 1 Minkowski metric, 6 Minkowski norm, 6 Minkowski space, 6 Morse function, 266 negatively complete, 175 non-degenerate critical point, 266 non-degenerate geodesic, 269 normal curvature, 213 observable diameter, 70 parallel translation, 89 path metric, 5 positively complete, 175 psychometric function, 14 Randers metric, 14 Randers norm, 7 regular metric sphere, 247 reversible Finsler metric, 12 reversible metric, 1 Riccati equation, 223 Ricci curvature, 98, 201 Ricci scalar, 98 Riemann curvature, 96, 97 Riemann curvature tensor, 132
Riemannian metric, 13 Riemannian curvature tensor, 127 Riemannian volume form, 20 S-curvature, 118 sectional curvature, 132 shape operator, 222 Sobolev constant, 58 spray, 79 strictly convex, 8 strictly convex hypersurface, 217 strongly convex, 8 T-curvature, 153 Tangent curvature, 153 tangent cut-domain, 183 Varga equation, 214 volume form, 19 weak Landsberg metric, 116
LECTURES ON FINSLER GEOMETRY by Z h o n g m i n Shen (Indiana University-Purdue
University
Indianapolis, USA) In 1854, B Riemann introduced the notion of curvature for spaces with a family of inner products. There was no significant progress in the general case until 1918, when P Finsler studied the variation problem in regular metric spaces. Around 1926, L Berwald extended Riemann's notion of curvature to regular metric spaces and introduced an important nonRiemannian curvature using his connection for regular metrics. Since then, Finsler geometry has developed steadily. In his Paris address in 1900, D Hilbert formulated 23 problems, the 4th and 23rd problems being in Finsler's category. Finsler geometry has broader applications in many areas of science and will continue to develop through the efforts of many geometers around the world. Usually, the methods employed in Finsler geometry involve very complicated tensor computations. Sometimes this discourages beginners. Viewing Finsler spaces as regular metric spaces, the author discusses the problems from the modern metric geometry point of view. The book begins with the basics on Finsler spaces, including the notions of geodesies and curvatures, then deals with basic comparison theorems on metrics and measures and their applications to the Levy concentration theory of regular metric measure spaces and Gromov's Hausdorff convergence
theory.
ISBN 981-02-4530-0
www. worldscientific.com 4619 he
9 "789810"245306"