Next Article in Journal
Transformer-Based Composite Language Models for Text Evaluation and Classification
Previous Article in Journal
Fixed-Point Approximation of Operators Satisfying (RCSC)—Condition in CAT(0) Spaces
Previous Article in Special Issue
Determining the Coefficients of the Thermoelastic System from Boundary Information
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Diameter Estimate in Geometric Flows

1
School of Mathematical Science, Yangzhou University, Yangzhou 225002, China
2
School of Mathematics and Statistics, Beijing Institute of Technology, Beijing 100081, China
*
Author to whom correspondence should be addressed.
Mathematics 2023, 11(22), 4659; https://doi.org/10.3390/math11224659
Submission received: 14 September 2023 / Revised: 29 October 2023 / Accepted: 14 November 2023 / Published: 16 November 2023
(This article belongs to the Special Issue Modern Analysis and Partial Differential Equations, 2nd Edition)

Abstract

:
We prove the upper and lower bounds of the diameter of a compact manifold ( M , g ( t ) ) with dim R M = n ( n 3 ) and a family of Riemannian metrics g ( t ) satisfying some geometric flows. Except for Ricci flow, these flows include List–Ricci flow, harmonic-Ricci flow, and Lorentzian mean curvature flow on an ambient Lorentzian manifold with non-negative sectional curvature.
MSC:
35K05; 35K08; 35K10; 53C21; 53E20

1. Introduction

The calculation and estimate of geometric quantities (e.g., volume, diameter, and curvature tensor [1]) play essential roles in the study of Riemannian geometry. It is also important and interesting to study the uniform properties of these geometric quantities under a family of Riemannian metric g ( t ) with t [ 0 , T ) for some T ( 0 , + ] . One of the most famous examples is Ricci flow
g i j t ( x , t ) = 2 R i j ( x , t ) ,
introduced by Hamilton [2]. It is a nonlinear weakly parabolic equation along which the Riemannian metric is evolved, and is a powerful theoretic tool to research geometric problems such as the Poincaré conjecture (see Perelman [3,4,5]). Up to now, we understand the κ non-collapsing property proved by Perelman [3], the improved non-collapsing property of Ricci flow proved by Jian [6], the κ non-inflated property proved by Zhang [7] (see also Chen-Wang [8]), and the uniform (logarithmic) Sobolev inequality along the Ricci flow proved by Zhang [9] and Ye [10]. It was also proved by Perelman that the diameter is uniformly bounded along the (normalized) Kähler–Ricci flow, which is the Ricci flow with the Kähler metric as its initial metric on Fano manifolds (see Sesum-Tian [11]). Recently, Jian-Song [12] proved that if the canonical line bundle K M of the Kähler manifold M is semi-ample, then the diameter is uniformly bounded for long-time solutions of the normalized Kähler–Ricci flow (see Jian-Song-Tian [13] for the most recent results in this direction). In [14], the author demonstrated a lower bound for the diameter along the Ricci flow with nonzero H 1 ( M n , R ) . For the Ricci flow (1) on a compact Riemannian manifold M with dim R M = n , Topping [15] showed that there is a uniform constant C = C ( g 0 , T ) such that if
diam ( M , g ( t ) ) C , t [ 0 , T ) ,
then, it yields that
diam ( M , g ( t ) ) C M | R | n 1 2 d μ ( t ) ,
where R denotes the scalar curvature of the Levi–Civita connection of g ( t ) . More details about the constant C can be found in [15]. Recently, Zhang [16] proved that the upper bound for the diameter of ( M , g ( t ) ) on compact Riemannian manifolds depends only on the L n 1 2 norm of the scalar curvature of g, the volume and the Sobolev constants (see [16]). Zhang [16] also deduced the lower bound for the diameter of ( M , g ( t ) ) , which depends only on the time t, the initial metric, and the L norm of the scalar curvature (see more details in Remark 1 and Theorem 1).
Motivated by [15,16], we investigate the geometric flow
g i j t ( x , t ) = 2 S i j ( x , t ) , ( x , t ) M × [ 0 , T ) ,
where T ( 0 , + ] and g ( 0 ) = g 0 is a Riemannian metric. Here, S i j ( x , t ) ’s denote the components of a symmetric 2-tensor S . We deduce the bound of the diameter of a compact manifold ( M , g ( t ) ) with dim R M = n ( n 3 ) and a family of Riemannian metrics g ( t ) satisfying the geometric flow (2) under certain assumptions. For later use, we need to define a tensor quantity D 2 associated to the tensor S (see [17], Definition 1.3).
Definition 1.
Let g ( t ) be a family of smooth Riemannian metrics satisfying the geometry flow (2) on M × [ 0 , T ) . Then, we define
D 2 ( S , X ) : = S t Δ g ( t ) S 2 | S | g ( t ) 2 + 4 ( i S i j ) X j 2 X i i S + 2 R i j X i X j 2 S i j X i X j , X X ( M ) ,
where S = i , j = 1 n g i j ( t ) S i j ( t ) . Here, ∇ and R i j denote the Levi–Civita connection and the Ricci curvature, respectively, of the Riemannian metric g ( t ) .
If
D 2 ( S , X ) 0 , X X ( M )
on [ 0 , T ) , then we say that D 2 ( S , · ) is non-negative.
Remark 1.
In fact, the quantity D 2 is the difference between two differential Harnack-type quantities for the symmetric tensor S [17]. Hence, the non-negativity of D 2 is equivalent to the corresponding differential Harnack-type inequality for the tensor S under the geometric flow.
We note that, if D 2 ( S , · ) and Ric S are non-negative, then there also hold a κ non-collapsing property, the so-called κ non-inflated property, and the uniform (logarithmic) Sobolev inequality along the geometric flow (2) (see [17,18,19,20,21,22,23] and the references therein). Thanks to [24,25], the many properties of the heat kernel also hold along the geometric flow (2) under the same assumptions, similar to those of the heat kernel along the Ricci flow (1). Given these preliminaries, we can consider the diameter estimate along the geometric flow (2).
Let g ( t ) be a family of smooth Riemannian metrics satisfying the geometry flow (2) on M × [ 0 , T ) . Then, we use the notations
d μ ( t ) : = the volume element of g ( t ) , and sometimes we also write d μ ( g ( x , t ) ) to emphasize the space variants , Vol g ( t ) ( M ) : = the volume of M with respect to g ( t ) , ( p , q , t ) : = the distance between p and q with respect to g ( t ) , B ( p , r , t ) : = the geodesic ball with center p and radius r with respect to g ( t ) , Vol g ( t ) ( B ( p , r , t ) ) : = the volume of B ( p , r , t ) with respect to g ( t ) .
Moreover, we will write the trace S as S t or S ( x , t ) to emphasize the time(-space) variant(s). Now, we state our main theorem.
Theorem 1.
Let g ( t ) be a family of smooth Riemannian metrics satisfying (2) on M × [ 0 , T ) with dim R M = n ( 3 ) . Assume that D 2 ( S , · ) defined in (3) and Ric S are non-negative. For the upper bound on the diameter of ( M , g ( t ) ) , we have
diam ( M , g ( t ) ) C Vol g ( t ) ( M ) + 1 + M S t + n 1 2 d μ ( t ) ,
where C denotes a constant depending only on n, A and B defined in (8). Here, S t + = max { 0 , S t } . For the lower bound on the diameter of ( M , g ( t ) ) , we have either diam ( M , g ( t ) ) t or
diam ( M , g ( t ) ) c 2 e 1 n α t β t S ( · , 0 ) e 2 n 0 t S ( · , s ) d s   ×   1 + 2 t n S ( · , 0 ) 1 2 Vol g 0 ( M ) 1 n .
Here, S ( x , t ) = min { 0 , S ( x , t ) } and c 2 are constants depending only on n. The constants α and β are positive constants which depend only on the infimum of the F functional defined by (7) for ( M , g 0 ) and the Sobolev constant of ( M , g 0 ) . Furthermore, if S ( · , 0 ) 0 , then we have α = 0 , which implies
diam ( M , g ( t ) ) c 2 e β n e 2 n 0 t S ( · , s ) d s Vol g 0 ( M ) 1 n .
Remark 2.
Our theorem will reduce to Zhang’s result [16] for the Ricci flow totally if the symmetric 2-tensor S is the Ricci curvature of g ( t ) .
Remark 3.
Our theorem can also be applied to the following geometric flows except for the Ricci flow. The geometric and physical meanings of these geometric flows and more details of the calculation of Ric S and D 2 ( S , · ) can be found in [17,21,26].
i.
List–Ricci flow (see List [27]).
g i j t ( x , t ) = 2 R i j ( x , t ) + 4 d ϕ d ϕ ( x , t ) , ϕ t ( x , t ) = Δ g ( x , t ) ϕ ( x , t ) ,
with ϕ C ( M × R , R ) . In this case, it follows that
Ric S = 2 d ϕ d ϕ 0
and
D 2 ( S , X ) = 4 | Δ ϕ X ϕ | 2 0 .
ii.
Harmonic-Ricci flow (see Müller [28]).
g i j t ( x , t ) = 2 R i j ( x , t ) + 2 α ( t ) ψ ψ ( x , t ) , ψ t ( x , t ) = τ g ( x , t ) ψ ( x , t ) ,
where M and N are compact manifolds with the Riemannian metrics g ( · , t ) and h, respectively,
ψ ( · , t ) : ( M , g ( · , t ) ) ( N , h )
are a family of smooth maps, τ g ψ is the intrinsic Laplacian of ψ, and α ( t ) is a smooth, positive and non-increasing function defined on R . In this case, there holds
Ric S = α d ψ d ψ 0
and
D 2 ( S , X ) = 2 α | τ g ψ X ψ | 2 α ˙ | ψ | 2 0 .
See Tadano [29] for a lower bound of the diameter for shrinking the Ricci-harmonic soliton.
iii.
Lorentzian mean curvature flow (see Holder [30] and Müller [17]). Let M be a compact space-like hyper-surface with dim R M = n in an ambient Lorentzian manifold L with dim R L = n + 1 , and let
F 0 : M L
denote a smooth immersion from M into L. We denote by
F ( · , t ) : M n L n + 1
a family of smooth immersions with F ( · , 0 ) = F 0 ( · ) and
F t ( p , t ) = H ( p , t ) ν ( p , t ) , ( p , t ) M × [ 0 , T ) ,
where ν ( p , t ) and H ( p , t ) are the future-oriented, time-like normal vector and the mean curvature of the hyper-surface M t = F ( M n , t ) at the point F ( p , t ) , respectively. It follows that
g i j t = 2 H A i j ,
where A = ( A i j ) denotes the second fundamental form on M t . In this setup, one has
S i j = H A i j , S = H 2 .
Mark the curvature tensor of L with a bar. If L has non-negative sectional curvature, then one can deduce that
D 2 ( S , X ) = 2 Ric ¯ ( H ν X , H ν X ) + 2 Rm ¯ ( X , ν ) ν , X + 2 | H A ( X , · ) | 2 0
and that
Ric ( X , X ) S ( X , X ) = Ric ¯ ( X , X ) + Rm ¯ ( X , ν ) ν , X + X i A i A j X j 0 , X X ( M ) .

2. Proof of Main Theorem

We need some preliminaries in order to prove Theorem 1. Let ( M , g ) be a compact Riemannian manifold with dim R M = n and the Riemannian metric g. Then, fixing a smooth function S C ( M , R ) , we can define the F entropy by
F ( g , h ) = M S + | h | g 2 e h d μ , h C ( M , R ) w i t h M e h d μ = 1 ,
where d μ is the volume element of the Riemannian metric g. When we take S to be the scalar curvature R of the Levi–Civita connection of the Riemannian metric g , the F entropy defined in (7) is exactly the one defined by Perelman [3].
Let v = e h 2 . Then, we have
F ( g , h ) = F * ( g , v ) = M 4 | v | g 2 + S v 2 d μ , M v 2 d μ = 1 .
We define
4 λ 0 ( g ) : = inf F * ( g , v ) : v C ( M , R ) , M v 2 d μ = 1 = inf F ( g , h ) : h C ( M , R ) , M e h d μ = 1 .
It follows from the standard theory partial differential equations that 4 λ 0 ( g ) is the first eigenvalue of the operator 4 Δ g + S . Let u 0 C ( M , R ) be a first positive eigenfunction of the operator 4 Δ g + S with
4 Δ g u 0 + S u 0 = 4 λ 0 ( g ) u 0 .
Then, h 0 = 2 log u 0 satisfies
4 λ 0 ( g ) = F ( g , h 0 )
with
2 Δ g h 0 + | h 0 | g 2 S = 4 λ 0 ( g ) .
Lemma 1
(Part of Lemma 3.1 in [21]). Let g ( x , t ) be a family of Riemannian metrics along the geometric flow (2) on M × [ 0 , T ) and let h ( x , t ) be a positive solution to the backward heat equation
t h ( x , t ) = Δ g ( x , t ) h + | h | g ( x , t ) 2 S ( x , t ) .
Then, we have
d F d t = M 2 | h i j + S i j | 2 + D 2 ( S , h ) e h d μ ( t )
where D 2 ( S , h ) is defined by (3).
In particular, the F entropy is non-decreasing in t if D 2 ( S , · ) is non-negative for all t [ 0 , T ) , from which we can obtain that λ 0 ( g ( t ) ) is non-decreasing of t.
Now, we can state the uniform Sobolev inequality along the geometric flow (2).
Lemma 2
(Theorem 1.5 in [21]). Let g ( t ) be a family of smooth Riemannian metrics satisfying (2) on M × [ 0 , T ) and assume that D 2 ( S , · ) defined in (3) is non-negative. Then, for each u W 1 , 2 ( M , g ( t ) ) , there holds
M | u | 2 n n 2 d μ ( t ) n 2 n A M 4 | u | g ( t ) 2 + S t u 2 d μ ( t ) + B M u 2 d μ ( t ) ,
where A and B are positive constants which depend only on M , g ( 0 ) , n , t , S ( · , 0 ) and the Sobolev constant of ( M , g ( 0 ) ) . In particular, if λ 0 ( g ( 0 ) ) > 0 , then B = 0 and A are independent of t.
The second ingredient of the proof can be considered as a quantified version of the non-collapsing theorem along the geometric flow (2) (see also Theorem 1.6 in [21]).
Lemma 3.
Let g ( t ) be a family of smooth Riemannian metrics satisfying (2) on M × [ 0 , T ) , and assume that D 2 ( S , · ) defined in (3) is non-negative. Then, it yields that
Vol g ( t ) ( | B ( x , | r , t ) ) 4 n 2 4 64 A + 4 A M 2 ( x , t , S t + , r ) + 4 B r 2 n 2 r n ,
where A and B are the constants in (8) and M 2 ( x , t , S t + , r ) is defined by (12).
Proof. 
Motivated by [31], define
u ( y ) = r ( x , y , t ) , y B ( x , r , t ) , 0 , o t h e r w i s e .
Substituting this u into (8), we obtain
M 4 | u | g ( t ) 2 d μ ( t ) = 4 Vol g ( t ) ( B ( x , r , t ) ) ,
M S t u 2 d μ ( t ) r 2 B ( x , r , t ) S t + d μ ( t ) = Vol g ( t ) ( B ( x , r , t ) ) r 2 Vol g ( t ) ( B ( x , r , t ) ) B ( x , r , t ) S t + d μ ( t ) M 2 ( x , t , S t + , r ) Vol g ( t ) ( B ( x , r , t ) ) ,
where M 2 ( x , t , S t + , r ) is a maximal type function defined by (see [15,16])
M 2 ( x , t , S t + , r ) = sup 0 < ρ r ρ 2 Vol g ( t ) ( B ( x , ρ , t ) ) B ( x , ρ , t ) S t + d μ ( t ) .
Note that
B ( x , ρ , t ) u 2 d μ ( t ) r 2 Vol g ( t ) ( B ( x , ρ , t ) ) .
Since u r 2 on B ( x , r 2 , t ) , from (8) and the Hölder inequality, we can deduce
r 2 4 Vol g ( t ) B x , r 2 , t B ( x , r , t ) u 2 d μ ( t ) Vol g ( t ) ( B ( x , r , t ) ) 2 n B ( x , r , t ) u 2 n n 2 n 2 n Vol g ( t ) ( B ( x , r , t ) ) 2 n A M ( 4 | u | g ( t ) 2 + S t u 2 ) d μ ( t ) + B M u 2 d μ ( t ) .
Substituting (10), (11) and (13) into (14) yields
4 A + A M 2 ( x , t , S t + , r ) + B r 2 Vol g ( t ) ( B ( x , r , t ) ) n + 2 n r 2 4 Vol g ( t ) B x , r 2 , t ,
i.e.,
Vol g ( t ) ( B ( x , r , t ) ) r 2 n n + 2 Vol g ( t ) B x , r 2 , t n n + 2 16 A + 4 A M 2 ( x , t , S t + , r ) + 4 B r 2 n n + 2 .
Note that, for any s ( 0 , r ] , Equation (15) still holds by replacing r with s. Since
M 2 ( x , t , S t + , s ) M 2 ( x , t , S t + , r ) , s 2 r 2 ,
we arrive at
Vol g ( t ) ( B ( x , s , t ) ) Vol g ( t ) B x , s 2 , t n n + 2 s 2 n n + 2 16 A + 4 A M 2 ( x , t , S t + , s ) + 4 B s 2 n n + 2 . Vol g ( t ) B x , s 2 , t n n + 2 s 2 n n + 2 16 A + 4 A M 2 ( x , t , S t + , r ) + 4 B r 2 n n + 2 .
Iterating (16) with s = r , r 2 , , r 2 k 1 for positive integers k gives
Vol g ( t ) ( B ( x , r , t ) ) 4 i = 1 k ( i 1 ) n n + 2 i Vol g ( t ) B x , r 2 k , t n n + 2 k r 2 16 A + 4 A M 2 ( x , t , S t + , r ) + 4 B r 2 1 i = 1 k n n + 2 i .
Letting k in (26) implies (10). □
We set
d ( t ) : = the diameter of M with respect to the Riemannian metric g ( t ) V ( t ) : = the volume of M with respect to the Riemannian metric g ( t )
for t ( 0 , T ] . For any x M , we write
κ = κ ( x , r ) : = Vol g ( t ) ( B ( x , r , t ) ) r n .
Also set
κ 0 : = min 4 n 2 4 ( 24 A + 4 B r 2 ) n 2 , ω n 2 ,
where ω n denotes the volume of the unit ball in R n . Now, we give the lower bound of M 2 ( x , t , S t + , r ) .
Lemma 4.
Let g ( t ) be a family of smooth Riemannian metrics satisfying (2) on M × [ 0 , T ) and assume that D 2 ( S , · ) defined in (3) is non-negative. If r d ( t ) 2 and κ κ 0 , then we have
M 2 ( x , t , S t + , r ) 2 .
In particular, if d ( t ) 2 and r 1 , then κ 0 can be taken as
κ 0 : = min 4 n 2 4 ( 24 A + 4 B ) n 2 , ω n 2 .
Proof. 
From (9), we have
16 A + 4 A M 2 ( x , t , S t + , r ) + 4 B r 2 4 n 2 κ 2 n .
This means, for r d ( t ) 2 ,
M 2 ( x , t , S t + , r ) 4 n 2 κ 2 n 4 B r 2 16 A 4 A .
Finally, Equation (18) follows from the definition of κ 0 and κ κ 0 . □
Based on the preliminaries as above, we can prove Theorem 1.
Proof of Theorem 1 .
For the upper bound of the diameter, let
N : = max n N : n d ( t ) 4 .
We choose two points a , b M with ( a , b , t ) = d ( t ) . Denote by γ a minimal geodesic connecting a and b. Choose p γ such that
( a , p , t ) = ( p , b , t ) = d ( t ) 2 .
We claim that, if
d ( t ) V ( t ) 4 n + 3 κ 0 ,
then, for at least N many positive integers i j in { 1 , 2 , , 2 N } , we have
Vol g ( t ) ( B ( p , i j , t ) ) Vol g ( t ) ( B ( p , i j 1 , t ) ) κ 0 4 n , j = 1 , , N .
Indeed, if the claim is not right, then there exist at least ( N + 1 ) integers k j in { 1 , 2 , , 2 N } such that
Vol g ( t ) ( B ( p , k j , t ) ) Vol g ( t ) ( B ( p , k j 1 , t ) ) κ 0 4 n , j = 1 , , N + 1 ,
which implies
d ( t ) κ 0 4 n 2 ( N + 1 ) κ 0 4 n i = 1 2 N Vol g ( t ) ( B ( p , i , t ) ) Vol g ( t ) ( B ( p , i 1 , t ) ) V ( t ) .
We can conclude that
d ( t ) V ( t ) 4 n + 2 κ 0 ,
which contradicts (22).
From now on, without loss of generality, assume that d ( t ) 2 and (22) hold. We pick N integers i 1 , i 2 , , i N with
{ i 1 , i 2 , , i N } { 1 , 2 , , 2 N }
and satisfying
Vol g ( t ) ( B ( p , i k , t ) ) Vol g ( t ) ( B ( p , i k 1 , t ) ) 4 n κ 0 , k = 1 , , N .
For i { i 1 , i 2 , , i N } , we set γ i : = γ B ( p , i , t ) B ( p , i 1 , t ) , and the middle point of γ i is denoted by p i . Then, there holds
B p i , 1 2 , t B ( p , i , t ) B ( p , i 1 , t ) ,
which implies that, for any x B p i , 1 4 , t , we have
B x , 1 4 , t B p i , 1 2 , t B ( p , i , t ) B ( p , i 1 , t ) .
Combining (23) and (24), we derive
Vol g ( t ) ( B ( x , 1 / 4 , t ) ) ( 1 / 4 ) n κ 0 .
Note that
lim ρ 0 Vol g ( t ) ( B ( x , ρ , t ) ) ρ n = ω n 2 κ 0 ,
where we use the definition of κ 0 in (19). From (25) and (26), there exists a positive number s = s ( x ) ( 0 , 1 / 4 ] such that
Vol g ( t ) ( B ( x , s ( x ) , t ) ) s ( x ) n = κ 0
and for 0 < ρ s ( x ) ,
Vol g ( t ) ( B ( x , ρ , t ) ) ρ n κ 0 .
From (18) and (27), we arrive at
M 2 ( x , t , S t + , s ( x ) ) 2 ,
which implies that there exists s 1 ( x ) ( 0 , s ( x ) ] such that
[ s 1 ( x ) ] 2 Vol g ( t ) ( B ( x , s 1 ( x ) , t ) ) B ( x , s 1 ( x ) , t ) S t + d μ ( t ) 1 .
From (28), we also have
Vol g ( t ) ( B ( x , s 1 ( x ) , t ) ) [ s 1 ( x ) ] n κ 0 .
Denote by σ the disjoint curves
i = i 1 , , i N γ i B p i , 1 4 , t .
Since N d ( t ) 8 , we know that
length g ( t ) ( σ ) d ( t ) 16 .
A family of balls { B ( x , s 1 ( x ) , t ) : x σ } covers σ . By Lemma 5.2 in [15], there exist a sequence of points { x : = 1 , 2 , } σ such that each of the balls B ( x , s ( x ) , t ) is disjoint from each other and these balls cover at least 1 3 of σ . Therefore, we can deduce
d ( t ) 16 length g ( t ) ( σ ) 96 | s 1 ( x ) | .
By (29) and the Hölder inequality, we obtain
Vol g ( t ) ( B ( x , s 1 ( x ) , t ) ) [ s 1 ( x ) ] 2 B ( x , s 1 ( x ) , t ) S t + d μ ( t ) B ( x , s 1 ( x ) , t ) S t + n 1 2 d μ ( t ) 2 n 1 Vol g ( t ) ( B ( x , s 1 ( x ) , t ) ) n 3 n 1 ,
which, combining (30), leads to
κ 0 s 1 ( x ) Vol g ( t ) ( B ( x , s 1 ( x ) , t ) ) [ s 1 ( x ) ] n 1 B ( x , s 1 ( x ) , t ) S t + n 1 2 d μ ( t ) .
Thanks to (31) and (32), we obtain
d ( t ) 96 κ 0 1 M S t + n 1 2 d μ ( t ) .
This inequality implies the desired upper bound (4) together with the assumptions d ( t ) 2 and (22).
For the lower bound for the diameter of ( M , g ( t ) ) , denote by G ( p , s ; x , t ) ( 0 s < t ) the fundamental solution to the conjugate heat equation
s f ( p , s ) + Δ g ( p , s ) f ( p , s ) S ( p , s ) f ( p , s ) = 0 ,
where G ( p , s ; x , t ) is seen as a function of ( p , s ) .
From Lemma 6.5 in [21], we know that
G ( p , s ; x , t ) c 1 J ( t ) ( t s ) n 2 e 2 ( p , x , t ) t s e 1 t s s t t v S ( z , v ) d v ,
where c 1 depends only on n. Here,
J ( t ) = exp [ α t β + t inf y M S ( y , 0 ) ] ,
where α and β are positive constants which depend only on the infimum of the F functional for ( M , g 0 ) and the Sobolev constant of ( M , g 0 ) . Furthermore, if S ( · , 0 ) 0 , then we have α = 0 . From (33), we can deduce
G ( p , s ; x , t ) c 1 J ( t ) ( t s ) n 2 e 2 ( p , x , t ) t s e s t S ( · , v ) d v .
Fix a time t 0 > 0 and a point x 0 M . Denote by r the diameter of ( M , g ( t 0 ) ) . Without loss of generality, we assume r < t 0 . Choosing p = x 0 , t = t 0 , s = t 0 r 2 in (35), for any x M , we have ( x 0 , x , t 0 ) r and arrive at
G ( x 0 , t 0 r 2 ; x , t 0 ) c 1 J ( t 0 ) r n e 1 0 t 0 S ( · , v ) d v .
From (3.20) in [21], we know that
S ( · , t ) 1 1 S ( · , 0 ) + 2 t n .
It follows from the adjoint property of the fundamental solution (see [32]) that, for fixed p and s, G ( p , s ; x , t ) is a function of ( x , t ) , and is the fundamental solution of the heat equation
t G ( p , s ; x , t ) Δ g ( x , t ) G ( p , s ; x , t ) = 0 .
This yields that
d d t M G ( p , s ; x , t ) d μ ( g ( x , t ) ) = M Δ g ( x , t ) G ( p , s ; x , t ) d μ ( g ( x , t ) ) M S ( x , t ) G ( p , s ; x , t ) d μ ( g ( x , t ) ) S ( · , t ) M G ( p , s ; x , t ) d μ ( g ( x , t ) ) ,
which, combining (37), yields
M G ( p , s ; x , t ) d μ ( g ( x , t ) ) 1 + 2 n S ( · , 0 ) ( t s ) n 2 .
It follows from (36) and (38) that
2 r 2 n S ( · , 0 ) + 1 n 2 M G ( x 0 , t 0 r 2 ; y , t 0 ) d μ ( g ( y , t 0 ) ) ( x 0 , y , t 0 ) r G ( x 0 , t 0 r 2 ; y , t 0 ) d μ ( g ( y , t 0 ) ) c 1 J ( t 0 ) r n e 1 0 t 0 S ( · , v ) d v ( x 0 , y , t 0 ) r d μ ( g ( y , t 0 ) ) .
Noting that r = diam ( M , g ( t 0 ) ) , we know
( x 0 , x , t 0 ) r d μ ( g ( x , t 0 ) ) = Vol g ( t 0 ) ( M ) .
In addition, we have
d d t Vol g ( t ) ( M ) = M S ( x , t ) d μ ( g ( x , t ) ) S ( · , t ) Vol g ( t ) ( M ) ,
which implies
Vol g ( t 0 ) ( M ) e 0 t 0 S ( · , t ) d t Vol g 0 ( M ) .
Combining (39)–(41) leads to
1 + 2 n S ( · , 0 ) r 2 n 2 r n c 1 J ( t 0 ) e 1 2 0 t 0 S ( · , s ) d s Vol g 0 ( M ) .
Since r = diam ( M , g ( t 0 ) ) and r < t 0 , by assumption, we can obtain
diam ( M , g ( t 0 ) ) c 2 e 1 n α t 0 β t 0 S ( · , 0 ) e 2 n 0 t 0 S ( · , s ) d s ×   1 + 2 t 0 n S ( · , 0 ) 1 2 Vol g 0 ( M ) 1 n .
If S ( · , 0 ) 0 , then we have α = 0 and S ( · , 0 ) = 0 , which implies that
diam ( M , g ( t 0 ) ) c 2 e β n e 2 n 0 t 0 S ( · , s ) d s Vol g 0 ( M ) 1 n .
Since t 0 is arbitrary, we can deduce (5) and (6). □

Author Contributions

Conceptualization, S.F. and T.Z. All authors have read and agreed to the published version of the manuscript.

Funding

Partially supported by the Natural Science Foundation of Jiangsu Province (BK20191435) and the National Natural Science Foundation of China (11771377).

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors are also grateful to the anonymous referees and the editor for their careful reading and helpful suggestions which greatly improved the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gladkov, S.O. To the question of Gauss’s curvature in n-dimensional Euclidean space. J. Math. Res. 2020, 12, 93–99. [Google Scholar] [CrossRef]
  2. Hamilton, R.S. Three-manifolds with positive Ricci curvature. J. Differ. Geom. 1982, 17, 255–306. [Google Scholar] [CrossRef]
  3. Perelman, G. The entropy formula for the Ricci flow and its geometric applications. arXiv 2002, arXiv:math/0211159. [Google Scholar]
  4. Perelman, G. Ricci flow with surgery on three-manifolds. arXiv 2003, arXiv:math/0303109. [Google Scholar]
  5. Perelman, G. Finite extinction time for the solutions to the Ricci flow on certain three-manifolds. arXiv 2003, arXiv:math/0307245. [Google Scholar] [CrossRef]
  6. Jian, W. On the improved no-local-collapsing theorem of Ricci flow. Peking Math. J. 2023, 6, 459–468. [Google Scholar] [CrossRef]
  7. Zhang, Q.S. Bounds on volume growth of geodesic balls under Ricci flow. Math. Res. Lett. 2012, 19, 245–253. [Google Scholar] [CrossRef]
  8. Chen, X.; Wang, B. On the conditions to extend Ricci flow (III). Int. Math. Res. Not. 2013, 2013, 2349–2367. [Google Scholar] [CrossRef]
  9. Zhang, Q.S. A uniform Sobolev inequality under Ricci flow. Int. Math. Res. Not. 2007, 2007, rnm056. [Google Scholar]
  10. Ye, R. The logarithmic Sobolev inequality along the Ricci flow. Commun. Math. Stat. 2015, 3, 1–36. [Google Scholar] [CrossRef]
  11. Sesum, N.; Tian, G. Bounding scalar curvature and diameter along the Kähler-Ricci flow (after Perelman). J. Inst. Math. Jussieu 2008, 7, 575–587. [Google Scholar] [CrossRef]
  12. Jian, W.; Song, J. Diameter estimates for long-time solutions of the Kähler-Ricci flow. Geom. Funct. Anal. 2022, 32, 1335–1356. [Google Scholar] [CrossRef]
  13. Jian, W.; Song, J.; Tian, G. A new proof of Perelman’s scalar curvature and diameter estimates for the Kähler-Ricci flow on Fano manifolds. arXiv 2023, arXiv:2310.07943. [Google Scholar]
  14. Ilmanen, T.; Knopf, D. A lower bound for the diameter of solutions to the Ricci flow with nonzero H1(Mn,R). Math. Res. Lett. 2003, 10, 161–168. [Google Scholar] [CrossRef]
  15. Topping, P. Diameter control under Ricci flow. Comm. Anal. Geom. 2005, 13, 1039–1055. [Google Scholar] [CrossRef]
  16. Zhang, Q.S. On the question of diameter bounds in Ricci flow. Ill. J. Math. 2014, 58, 113–123. [Google Scholar] [CrossRef]
  17. Müller, R. Monotone volume formulas for geometric flows. J. Reine Angew. Math. 2010, 643, 39–57. [Google Scholar] [CrossRef]
  18. Abolarinwa, A. Sobolev-type inequalities and heat kernel bounds along the geometric flow. Afr. Mat. 2016, 17, 169–186. [Google Scholar] [CrossRef]
  19. Băileşteanu, M.; Tran, H. Heat kernel estimates under the Ricci-harmonic map flow. Proc. Edinb. Math. Soc. 2017, 60, 831–857. [Google Scholar] [CrossRef]
  20. Fang, S.; Zheng, T. An upper bound of the heat kernel along the harmonic-Ricci flow. Manuscr. Math. 2016, 151, 1–18. [Google Scholar] [CrossRef]
  21. Fang, S.; Zheng, T. The (logarithmic) sobolev inequalities along geometric flow and applications. J. Math. Anal. Appl. 2016, 434, 729–764. [Google Scholar] [CrossRef]
  22. Liu, X.; Wang, K. A Gaussian upper bound of the conjugate heat equation along an extended Ricci flow. Pac. J. Math. 2017, 287, 465–484. [Google Scholar] [CrossRef]
  23. Zhu, A. Differential Harnack inequalities for the backward heat equation with potential under the harmonic-Ricci flow. J. Math. Anal. Appl. 2013, 406, 502–510. [Google Scholar] [CrossRef]
  24. Cao, X.; Guo, H.; Tran, H. Harnack estimates for conjugate heat kernel on evolving manifolds. Math. Z. 2015, 281, 201–214. [Google Scholar] [CrossRef]
  25. Chow, B.; Chu, S.; Glickenstein, D.; Guenther, C.; Isenberg, J.; Ivey, T.; Knopf, D.; Lu, P.; Luo, F.; Ni, L. The Ricci Flow: Techniques and Applications, Part III: Geometric-Analytic Aspects; Mathematical Surveys and Monographs; American Mathematical Society: Providence, RI, USA, 2010; Volume 163, pp. 2349–2367. [Google Scholar]
  26. Fang, S. Differential Harnack inequalities for heat equations with potentials under geometric flows. Arch. Math. 2013, 100, 179–189. [Google Scholar] [CrossRef]
  27. List, B. Evolution of an extended Ricci flow system. Comm. Anal. Geom. 2008, 16, 1007–1048. [Google Scholar] [CrossRef]
  28. Müller, R. Ricci flow coupled with harmonic map flow. Ann. Sci. Éc. Norm. Supér. 2012, 45, 101–142. [Google Scholar] [CrossRef]
  29. Tadano, H. A lower diameter bound for compact domain manifolds of shrinking Ricci-harmonic solitons. Kodai Math. J. 2015, 38, 302–309. [Google Scholar] [CrossRef]
  30. Holder, M. Contracting spacelike hypersurfaces by their inverse mean curvature. J. Austral. Math. Soc. Ser. A 2000, 68, 285–300. [Google Scholar] [CrossRef]
  31. Carron, G. Inégalités isopérimétriques de Faber-Krahn et conséquences. In Collection SMF Séminaires et Congres, Proceedings of the Actes de la Table Ronde de Géométrie Différentielle, Luminy, Marseille, France, 12–18 July 1992; Besse, A.L., Ed.; Société Mathématique de France: Paris, France, 1996; Volume 1, pp. 205–232. [Google Scholar]
  32. Guenther, C.M. The fundamental solution on manifolds with time-dependent metrics. J. Geom. Anal. 2002, 12, 425–436. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Fang, S.; Zheng, T. Diameter Estimate in Geometric Flows. Mathematics 2023, 11, 4659. https://doi.org/10.3390/math11224659

AMA Style

Fang S, Zheng T. Diameter Estimate in Geometric Flows. Mathematics. 2023; 11(22):4659. https://doi.org/10.3390/math11224659

Chicago/Turabian Style

Fang, Shouwen, and Tao Zheng. 2023. "Diameter Estimate in Geometric Flows" Mathematics 11, no. 22: 4659. https://doi.org/10.3390/math11224659

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop