This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Geometric analysis on manifolds with ends

Alexander Grigor’yan Department of Mathematics University of Bielefeld 33501, Bielefeld, Germany [email protected] Satoshi Ishiwata Department of Mathematical Sciences Yamagata University, Yamagata 990-8560, Japan [email protected]  and  Laurent Saloff-Coste Department of Mathematics Cornell University, Ithaca, NY, 14853-4201, USA [email protected]
Key words and phrases:
manifold with ends, heat kernel, Poincaré constant
2010 Mathematics Subject Classification:
Primary 58-02, Secondary 35K08, 58J65, 58J35
Partially supported by SFB 1283 of the German Research Council
Partially supported by JSPS, KAKENHI 17K05215
Partially supported by NSF grant DMS–1707589

Contents

1. Introduction

2. The state of the art

3. Manifold with ends with oscillating volume functions

References

1. Introduction

In this survey article, we discuss some recent progress on geometric analysis on manifold with ends. In the final section, we construct manifolds with ends with oscillating volume functions which may turn out to have a different heat kernel estimates from those provided by known results.

Throughout the history of geometric analysis, manifolds with ends have appeared in several contexts. For example, Cai [2], Kasue [19] and Li-Tam [22] et. al. studied manifolds with non-negative Ricci (sectional) curvature outside a compact set where manifolds with ends play an important role. It should be pointed out that there are other recent works on manifolds with ends. See, for instance, Carron [3], Doan [5], Duong, Li and Sikora [6], Hassel, Nix and Sikora [17], Hassel and Sikora [18].

Because of the bottleneck structure inherent to most manifolds with ends, geometric and analytic properties of manifolds with ends are very different from a manifold such as n\mathbb{R}^{n}. For example, in 1979, Kuz’menko and Molchanov [21] proved the following:

Theorem 1.1.

On M=3#3M=\mathbb{R}^{3}\#\mathbb{R}^{3}, the connected sum of two copies of 3\mathbb{R}^{3}, the weak Liouville property does not hold. Namely, there exists a non-trivial bounded harmonic function.

It is a well-known fact that the parabolic Harnack inequality ((PHI) in short) implies the weak Liouville property. See [14, Section 2.1] and [26, 5.4.5] for details. By a contraposition argument, the above theorem implies that (PHI) does not hold on 3#3\mathbb{R}^{3}\#\mathbb{R}^{3}.

Denote by p(t,x,y)p\!\left(t,x,y\right) the heat kernel of a non-compact weighted manifold (M,d,μ)\!(M,d,\mu), that is, the minimal positive fundamental solution of the heat equation tu=Δu\partial_{t}u=\Delta u, where Δ\Delta is the weighed Laplacian. In 1986, Li and Yau proved in [23] that

(LY)p(t,x,y)cV(x,t)ebd2(x,y)/t(LY)\qquad\qquad p(t,x,y)\asymp\frac{c}{V(x,\sqrt{t})}e^{-bd^{2}(x,y)/t}

holds on non-compact manifold with non-negative Ricci curvature. Here V(x,r):=μ(B(x,r))V(x,r):=\mu(B(x,r)), the measure of the open geodesic ball B(x,r)={yM:d(x,y)<r}B(x,r)=\{y\in M~{}:~{}d(x,y)<r\} and the sign \asymp means that both \leq and \geq hold but with different values of the positive constants CC and bb. We call this estimate a Li-Yau type bound and write (LY) in short. The following theorem is a combined result of [7], [25] based on previous contributions of Moser [24], Kusuoka–Stroock [20] et al.

Theorem 1.2.

On a geodesically complete, non-compact weighted manifold MM, the following conditions are equivalent:

(1) The Li-Yau type heat kernel estimates (LY).

(2) The parabolic Harnack inequality (PHI).

(3) The Poincaré inequality (PI)(PI): there exists C,κ>0C,\kappa>0 such that for any xMx\in M, r>0r>0 and fC(B(x,r)¯)f\in C^{\infty}\left(\overline{B(x,r)}\right),

(PI)B(x,r)|ffB(x,r)|2𝑑μCB(x,r)|f|2𝑑μ,(PI)~{}~{}~{}\int_{B(x,r)}|f-f_{B(x,r)}|^{2}d\mu\leq C\int_{B(x,r)}|\nabla f|^{2}d\mu,

where fB(x,r)=1V(x,r)B(x,r)f𝑑μf_{B(x,r)}=\frac{1}{V(x,r)}\int_{B(x,r)}fd\mu, and the volume doubling property (VD)(VD): there exists C>0C>0 such that for any xMx\in M, r>0r>0,

V(x,2r)CV(x,r).V(x,2r)\leq CV(x,r).

Combining the above two theorems, the connected sum M=M1#M2=3#3M=M_{1}\#M_{2}=\mathbb{R}^{3}\#\mathbb{R}^{3} satisfies neither (PI) nor (LY). Indeed, the function

f(x)={1xM1,1xM2f(x)=\left\{\begin{array}[]{cl}1&x\in M_{1},\\ -1&x\in M_{2}\end{array}\right.

implies that fB(o,r)=0f_{B(o,r)}=0 for a central reference point oMo\in M and

B(o,r)|ffB(o,r)|2𝑑μr3,B(o,r)|f|2𝑑μconst,\int_{B(o,r)}|f-f_{B(o,r)}|^{2}d\mu\simeq r^{3},\quad\int_{B(o,r)}|\nabla f|^{2}d\mu\simeq const,

which fails (PI). Here fgf\simeq g means

cfgCfcf\leq g\leq Cf

with some positive constants 0<cC0<c\leq C on a suitable range of functions ff, gg. Moreover, Benjamini, Chavel and Feldman [1] obtained in 1996 the following heat kernel estimate.

Theorem 1.3.

For n3n\geq 3, let M=M1#M2=n#nM=M_{1}\#M_{2}=\mathbb{R}^{n}\#\mathbb{R}^{n}. There exists ε>0\varepsilon>0 such that for xM1x\in M_{1}, yM2y\in M_{2} with |x||y|t|x|\simeq|y|\simeq\sqrt{t},

p(t,x,y)1tn+ε21tn/2.p(t,x,y)\leq\frac{1}{t^{\frac{n+\varepsilon}{2}}}\ll\frac{1}{t^{n/2}}.

This theorem asserts that the heat kernel between two different ends is significantly smaller than that on one end because of a bottleneck effect.

In view of the above facts, it is natural to ask on a manifold with ends the behavior of the heat kernel p(t,x,y)p(t,x,y) and the estimate of

Λ(B(x,r)):=supfC1(B(x,r)¯)fconstB(x,r)|ffB(x,r)|2𝑑μB(x,r)|f|2𝑑μ,\Lambda(B(x,r)):=\sup_{\begin{subarray}{c}f\in C^{1}(\overline{B(x,r)})\\ f\neq\mathrm{const}\end{subarray}}\frac{\int_{B(x,r)}|f-f_{B(x,r)}|^{2}d\mu}{\int_{B(x,r)}|\nabla f|^{2}d\mu}, (1.1)

which is called the Poincaré constant.

Notation. Throughout this article, the letters c,c,C,C,C′′c,c^{\prime},C,C^{\prime},C^{\prime\prime} denote positive constants whose values may be different at different instances. When the value of a constant is significant, it will be explicitly stated.

2. The state of the art

2.1. Setting

First of all, we begin with the definition of what we call a manifold with finitely many ends. For a fixed integer k2k\geq 2, let M1,,MkM_{1},...,M_{k} be a sequence of geodesically complete, non-compact weighted manifolds of the same dimension.

Definition 2.1.

We say that a weighted manifold MM is a manifold with kk ends M1,M2,MkM_{1},M_{2},\ldots M_{k} and write

M=M1##MkM=M_{1}\#...\#M_{k} (2.1)

if there is a compact set KMK\subset M so that MKM\setminus K consists of kk connected components E1,E2,,EkE_{1},E_{2},\ldots,E_{k} such that each EiE_{i} is isometric (as a weighted manifold) to MiKiM_{i}\setminus K_{i} for some compact set KiMiK_{i}\subset M_{i} (see Fig. 1). Each EiE_{i} (or MiM_{i}) will be referred to as an end of MM.

MiM_{i}EiE_{i}EiE_{i}KKKiK_{i}Ei\partial E_{i}M=M1##MkM=M_{1}\#\cdots\#M_{k}
Figure 1. Manifold with ends

Here we remark that the definition of end given above is different from the usual notion defined as a connected component of the ideal boundary.

We say that a manifold MM is parabolic if any positive superharmonic function on MM is constant, and non-parabolic otherwise. See [8] for details.

Throughout this article, we always assume that each end MiM_{i} satisfies (VD) and (PI). Moreover, if the end MiM_{i} is parabolic, then we also assume that MiM_{i} satisfies the relatively connected annuli condition defined as follows.

Definition 2.2 ((RCA)).

A weighted manifold MM satisfies relatively connected annuli condition ((RCA) in short) with respect to a reference point oMo\in M if there exists a positive constant A>1A>1 such that for any r>A2r>A^{2} and all x,yMx,y\in M with d(o,x)=d(o,y)=rd(o,x)=d(o,y)=r, there exists a continuous path from xx to yy staying in B(o,Ar)\B(o,A1r)B(o,Ar)\backslash B(o,A^{-1}r). See Fig. 2 and 3 for typical positive and negative examples.

\bullet~{}\bullet~{}\bullet
Figure 2. Manifold with (RCA)
2k2^{k}hole 11hole 22hole kk2244\bullet\qquad\bullet\qquad\bullet
Figure 3. Manifold without (RCA)

The assumption (RCA) seems technical but it makes it possible to obtain an optimal estimates of the first exit (hitting) probability and the Dirichlet heat kernel in the exterior of a compact set of a parabolic manifold satisfying (LY). See [12] and [13] for details.

2.2. Heat kernel estimates

2.2.1. Off-diagonal estimates

Let M=M1##MkM=M_{1}\#\cdots\#M_{k} be a manifold with kk ends. For t>0t>0, xEix\in E_{i} and yMy\in M, let pEiD(t,x,y)p^{D}_{E_{i}}(t,x,y) be the extended Dirichlet heat kernel on an end EiE_{i}, that is, the Dirichlet heat kernel in yEiy\in E_{i} and extension to 0 if yEiy\not\in E_{i}. Let τEi\tau_{E_{i}} be the first exit time of the Brownian motion from EiE_{i} and then x(τEi<t)\mathbb{P}_{x}(\tau_{E_{i}}<t) is the first exit probability starting from xx by time tt from EiE_{i}. We will use the following theorem to estimate the off-diagonal heat kernel estimates.

Theorem 2.3 (Grigor’yan and Saloff-Coste [15, Theorem 3.5]).

Let
M=M1##MkM=M_{1}\#\cdots\#M_{k} be a manifold with kk ends and fix a central reference point oKo\in K. For xEix\in E_{i}, yEjy\in E_{j} and t>1t>1,

p(t,x,y)\displaystyle p(t,x,y)\simeq pEiD(t,x,y)+p(t,o,o)x(τEi<t)y(τEj<t)\displaystyle p^{D}_{E_{i}}(t,x,y)+p(t,o,o)\mathbb{P}_{x}(\tau_{E_{i}}<t)\mathbb{P}_{y}(\tau_{E_{j}}<t)
+0tp(s,o,o)𝑑s(tx(τEi<t)y(τEj<t)+x(τEi<t)ty(τEj<t)).\displaystyle\!\!\!+\int_{0}^{t}p(s,o,o)ds\left(\partial_{t}\mathbb{P}_{x}(\tau_{E_{i}}<t)\mathbb{P}_{y}(\tau_{E_{j}}<t)+\mathbb{P}_{x}(\tau_{E_{i}}<t)\partial_{t}\mathbb{P}_{y}(\tau_{E_{j}}<t)\right). (2.2)

Under the assumption of (PI), (VD) on each end, and, in addition, (RCA) on each parabolic end, applying results in [12] and [13], the quantities pEiD(t,x,y)p^{D}_{E_{i}}(t,x,y), x(τEi<t)\mathbb{P}_{x}(\tau_{E_{i}}<t), y(τEj<t)\mathbb{P}_{y}(\tau_{E_{j}}<t), tx(τEi)\partial_{t}\mathbb{P}_{x}(\tau_{E_{i}}) and ty(τEj<t)\partial_{t}\mathbb{P}_{y}(\tau_{E_{j}}<t) can be estimated. Hence, estimating p(t,o,o)p(t,o,o) becomes the key missing estimate to obtain off-diagonal bounds on manifolds with ends.

2.2.2. Non-parabolic case

First we consider heat kernel estimates on M=M1##MkM=M_{1}\#\cdots\#M_{k}, where MM is non-parabolic, namely, at least one end MiM_{i} is non-parabolic.

For a fixed reference point oiKiMio_{i}\in K_{i}\subset M_{i}, let

Vi(r):=μ(B(oi,r))V_{i}(r):=\mu(B(o_{i},r))

and

hi(r):=1+(1rsdsVi(s))+.h_{i}(r):=1+\left(\int_{1}^{r}\frac{sds}{V_{i}(s)}\right)_{+}.

Here we remark that, under the assumption (LY), MiM_{i} is parabolic if and only if

limrhi(r)=.\lim_{r\rightarrow\infty}h_{i}(r)=\infty.

In 2009, Grigor’yan and Saloff-Coste [15] obtained the following (see also [16]).

Theorem 2.4.

Let M=M1##MkM=M_{1}\#\cdots\#M_{k} be a manifold with kk ends. Assume that each end MiM_{i} satisfies (PI) and (VD) and that each parabolic end satisfies (RCA). Assume also that MM is non-parabolic. Then for all t>0t>0,

p(t,o,o)1miniVi(t)hi2(t).p(t,o,o)\simeq\frac{1}{\min_{i}V_{i}(\sqrt{t})h_{i}^{2}(\sqrt{t})}.

If all ends M1,,MkM_{1},\ldots,M_{k} are non-parabolic, then all functions h1,,hkh_{1},\ldots,h_{k} are bounded. Hence, the above theorem implies that

p(t,o,o)1miniVi(t),p(t,o,o)\simeq\frac{1}{\min_{i}V_{i}(\sqrt{t})},

namely, the behavior of the heat kernel at the central reference point is determined by the smallest end!

As a typical example, let MM be M1#M2=n#nM_{1}\#M_{2}=\mathbb{R}^{n}\#\mathbb{R}^{n}, the connected sum of two copies of n\mathbb{R}^{n} with n3n\geq 3. Then the above theorem implies that

p(t,o,o)1tn/2.p(t,o,o)\simeq\frac{1}{t^{n/2}}.

Substituting this estimates into Theorem 2.3, we obtain that for xM1x\in M_{1}, yM2y\in M_{2},

p(t,x,y)1tn/2(1|x|n2+1|y|n2)ebd2(x,y)/t,p(t,x,y)\simeq\frac{1}{t^{n/2}}\left(\frac{1}{|x|^{n-2}}+\frac{1}{|y|^{n-2}}\right)e^{-bd^{2}(x,y)/t},

where |x|=1+d(x,K)|x|=1+d(x,K).

2.2.3. Parabolic case

Next, we consider the case of manifolds with ends, M=M1##MkM=M_{1}\#\cdots\#M_{k}, which are parabolic, that is, for which all ends M1,,MkM_{1},\ldots,M_{k} are parabolic. To prove an optimal heat kernel estimates, we need the following assumptions on each end.

Definition 2.5 (c.f. [10]).

An end MiM_{i} is called subcritical if

hi(r)Cr2Vi(r)(r>1)h_{i}(r)\leq C\frac{r^{2}}{V_{i}(r)}\quad~{}(\forall r>1)

and regular if there exist γ1,γ2>0\gamma_{1},\gamma_{2}>0 satisfying 2γ1+γ2<22\gamma_{1}+\gamma_{2}<2 such that

c(Rr)2γ2Vi(R)Vi(r)C(Rr)2+γ1(1<rR).c\left(\frac{R}{r}\right)^{2-\gamma_{2}}\leq\frac{V_{i}(R)}{V_{i}(r)}\leq C\left(\frac{R}{r}\right)^{2+\gamma_{1}}\quad(\forall 1<r\leq R). (2.3)

For example, a manifold MM with volume function V(r)=rα(logr)βV(r)=r^{\alpha}\left(\log r\right)^{\beta} is parabolic if and only if either α<2\alpha<2 or α=2\alpha=2 and β1\beta\leq 1. Moreover, MM is subcritical if α<2\alpha<2 and regular if α=2\alpha=2 and β1\beta\leq 1. We remark that if MiM_{i} satisfies (VD), then the reverse doubling property holds and that implies that for any subcritical end, there exists δ>0\delta>0 such that

Vi(r)Cr2δ(r>0).V_{i}(r)\leq Cr^{2-\delta}~{}(\forall r>0). (2.4)

For r>0r>0, let m=m(r)m=m(r) be a number so that

Vm(r)=maxiVi(r).V_{m}(r)=\max_{i}V_{i}(r). (2.5)

We can now state the following result.

Theorem 2.6 (Grigor’yan, Ishiwata, Saloff-Coste [10]).

Let M=M1##MkM=M_{1}\#\!\cdots\!\#M_{k} be a manifold with kk parabolic ends. Assume that each end MiM_{i} satisfies (PI), (VD), (RCA) and is either subcritical or regular. If there exist both of subcritical and regular ends, assume also that the constant δ\delta in (2.4) satisfies δ>γ2\delta>\gamma_{2}, namely, for any subcritical volume function Vi(r)V_{i}(r) and any regular volume function Vj(r)V_{j}(r),

Vi(r)Cr2δCr2γ2C′′Vj(r)(r>0).V_{i}(r)\leq Cr^{2-\delta}\leq C^{\prime}r^{2-\gamma_{2}}\leq C^{\prime\prime}V_{j}(r)~{}(\forall r>0).

Moreover, assume that there exists an end MmM_{m} such that for all i=1,,ki=1,\ldots,k and for all r>0r>0

Vm(r)cVi(r)andVm(r)hm2(r)CVi(r)hi2(r).V_{m}(r)\geq cV_{i}(r)~{}\mbox{and}~{}V_{m}(r)h_{m}^{2}(r)\leq CV_{i}(r)h_{i}^{2}(r). (2.6)

Then for t>0t>0

p(t,o,o)1Vm(t).p(t,o,o)\simeq\frac{1}{V_{m}(\sqrt{t})}. (2.7)

This means that the on-diagonal heat kernel estimates at the central reference point is determined by the largest end!

Remark 2.7.

In our approach, we require the existence of a fixed dominating end given by (2.6) for the optimal estimates in (2.7) to hold. Indeed, more generally, on a manifold with either regular or subcritical ends, we obtain for t>1t>1, (see [10] for the detail)

p(t,o,o)Cminihi2(t)miniVi(t)hi2(t).p(t,o,o)\leq C\frac{\min_{i}h_{i}^{2}(\sqrt{t})}{\min_{i}V_{i}(\sqrt{t})h_{i}^{2}(\sqrt{t})}. (2.8)

The assumption in (2.6) implies that for all r>1r>1

minihi2(r)miniVi(r)hi2(r)CVm(r),\frac{\min_{i}h_{i}^{2}(r)}{\min_{i}V_{i}(r)h_{i}^{2}(r)}\leq\frac{C}{V_{m}(r)}, (2.9)

which allows to apply [4, Theorem 7.2] for the matching lower bound. In Section 3, we construct manifolds with ends without a fixed dominating end and, in such cases, the estimates in (2.9) does not hold.

As illustrative examples, let M=M1##MkM=M_{1}\#\cdots\#M_{k} be a manifold with parabolic ends, where each end MiM_{i} satisfies (PI), (VD), (RCA). Let αi\alpha_{i} and βi\beta_{i} be sequences satisfying

(α1,β1)(α2,β2)(αk,βk)>(0,+)(\alpha_{1},\beta_{1})\geq(\alpha_{2},\beta_{2})\geq\cdots\geq(\alpha_{k},\beta_{k})>(0,+\infty)

in the sense of lexicographical order, namely (αi,βi)>(αj,βj)(\alpha_{i},\beta_{i})>(\alpha_{j},\beta_{j}) means that

αi>αjorαi=αjandβi>βj\alpha_{i}>\alpha_{j}~{}\mbox{or}~{}\alpha_{i}=\alpha_{j}~{}\mbox{and}~{}\beta_{i}>\beta_{j}

and we assume that

Vi(r)rαi(logr)βi,r>2.V_{i}(r)\simeq r^{\alpha_{i}}\left(\log r\right)^{\beta_{i}},\quad r>2.

Here we need (α1,β1)(2,1)(\alpha_{1},\beta_{1})\leq(2,1) so that all ends M1,,MkM_{1},\ldots,M_{k} are parabolic. Then the above theorem implies that

p(t,o,o)1tα1/2(logt)β1,t>2.p(t,o,o)\simeq\frac{1}{t^{\alpha_{1}/2}\left(\log t\right)^{\beta_{1}}},\quad t>2.

As an explicit example, suppose that k=2k=2 and (α1,β1)=(2,0)(\alpha_{1},\beta_{1})=(2,0) and (α2,β2)=(1,0)(\alpha_{2},\beta_{2})=(1,0). Substituting the above estimates into (2.2), we obtain for xE1x\in E_{1}, yE2y\in E_{2} and t>1t>1

p(t,x,y){1tebd2(x,y)/tif |x|>t1t(1+|y|tloget|x|)if |x|,|y|t1t(loget|x|)ebd2(x,y)/tif |x|t<|y|.p(t,x,y)\simeq\left\{\begin{array}[]{ll}\frac{1}{t}e^{-bd^{2}(x,y)/t}&\mbox{if }|x|>\sqrt{t}\\ \frac{1}{t}\left(1+\frac{|y|}{\sqrt{t}}\log\frac{e\sqrt{t}}{|x|}\right)&\mbox{if }|x|,|y|\leq\sqrt{t}\\ \frac{1}{t}\left(\log\frac{e\sqrt{t}}{|x|}\right)e^{-bd^{2}(x,y)/t}&\mbox{if }|x|\leq\sqrt{t}<|y|.\end{array}\right.
Remark 2.8.

Assume that all ends of a manifold M=M1##MkM=M_{1}\#\cdots\#M_{k} are subcritical. Then for xEix\in E_{i} and yEjy\in E_{j} with iji\neq j and t>1t>1,

p(t,x,y)CVm(t)ebd2(x,y)/tp(t,x,y)\asymp\frac{C}{V_{m}(\sqrt{t})}e^{-bd^{2}(x,y)/t}

(see [9, Theorem 2.3]).

2.3. Poincaré constant estimates

In this section, we consider the estimates of the Poincaré constant defined in (1.1). Recall that M=M1##MkM=M_{1}\#\cdots\#M_{k} is a manifold with ends M1,,MkM_{1},\ldots,M_{k}, where each end satisfies (VD) and (PI). Let oKo\in K be a central reference point. Our main interest is to obtain the Poincaré constant Λ(B(o,r))\Lambda(B(o,r)) at the central point oo. In fact, by the monotonicity of Λ\Lambda together with a Whitney covering argument (see [10]), for r>2|x|r>2|x|

Λ(B(x,r))Λ(B(o,r)).\Lambda(B(x,r))\simeq\Lambda(B(o,r)).

For r>0r>0, let n=n(r)n=n(r) be the number so that

Vn(r)\displaystyle V_{n}(r) =maximVi(r),\displaystyle=\max_{i\neq m}V_{i}(r),

where m=m(r)m=m(r) is the number of the largest end (see (2.5)). Then we obtain the following.

Theorem 2.9 (Grigor’yan, Ishiwata, Saloff-Coste [10]).

Let M=M1##MkM=M_{1}\#\!\cdots\!\#M_{k} be a manifold with kk non-parabolic ends. Assume that each end MiM_{i} satisfies (VD) and (PI). Then for sufficiently large r>1r>1

Λ(B(o,r))CVn(r).\Lambda(B(o,r))\leq CV_{n}(r).

Moreover, if for all r>0r>0

rV(r)CV(r),rV^{\prime}(r)\leq CV(r),

then for sufficiently large r>1r>1

Λ(B(o,r))Vn(r).\Lambda(B(o,r))\simeq V_{n}(r).

When MM has at least one parabolic end, we assume the following additional condition (see [10] for details).

Definition 2.10 ((COE)).

We say that a manifold M=#iIMiM=\#_{i\in I}M_{i} has critically ordered ends and write (COE) in short if there exist ε,δ,γ1,γ2>0\varepsilon,\delta,\gamma_{1},\gamma_{2}>0 such that

γ1<ε,γ1+γ2<δ<2,2γ1+γ2<2,\gamma_{1}<\varepsilon,~{}~{}\gamma_{1}+\gamma_{2}<\delta<2,~{}~{}2\gamma_{1}+\gamma_{2}<2,

and a decomposition

I=IsuperImiddleIsubI=I_{super}\sqcup I_{middle}\sqcup I_{sub}

such that the following conditions are satisfied:

  • (a)\left(a\right)

    For each iIsuperi\in I_{super} and all r1r\geq 1,

    Vi(r)cr2+ϵ.V_{i}(r)\geq cr^{2+\epsilon}\ .
  • (b)\left(b\right)

    For each iIsubi\in I_{sub}, ViV_{i} is subcritical (see Definition 2.5) and

    Vi(r)Cr2δ.V_{i}(r)\leq Cr^{2-\delta}\ .
  • (c)\left(c\right)

    For each iImiddlei\in I_{middle}, ViV_{i} is regular (see (2.3)). Moreover, for any pair i,jImiddlei,j\in I_{middle} we have either VicVjV_{i}\geq cV_{j} or VjcViV_{j}\geq cV_{i} (i.e., the ends in ImiddleI_{middle} can be ordered according to their volume growth uniformly over r[1,)r\in[1,\infty)) and VicVjV_{i}\geq cV_{j} implies that VihicVjhjV_{i}h_{i}\geq c^{\prime}V_{j}h_{j}. Besides, if MM is parabolic (i.e., all ends are parabolic) then ViCVjV_{i}\geq CV_{j} also implies Vihi2CVjhj2V_{i}h_{i}^{2}\leq C^{\prime}V_{j}h_{j}^{2}.

Theorem 2.11 ([10]).

Let M=M1##MkM=M_{1}\#\cdots\#M_{k} be a manifold with kk ends, where each end satisfies (VD) and (PI). Suppose that there exists at least one parabolic end and each parabolic end satisfies (RCA). If MM admits (COE), then for sufficiently large r>1r>1

Λ(B(o,r))CVn(r)hn(r).\Lambda(B(o,r))\leq CV_{n}(r)h_{n}(r).

If, in addition, each ViV_{i} satisfies that

rVi(r)CVi(r)(r>1),rV_{i}^{\prime}(r)\leq CV_{i}(r)~{}~{}(\forall r>1),

then, for sufficiently large r>1r>1

Λ(B(o,r))Vn(r)hn(r).\Lambda(B(o,r))\simeq V_{n}(r)h_{n}(r).

These results say that the Poincaré constant Λ(B(o,r))\Lambda(B(o,r)) is determined by the second largest end!

As an explicit example, let M=n#nM=\mathbb{R}^{n}\#\mathbb{R}^{n} with n2n\geq 2. Then Theorems 2.9 and 2.11 imply that

Λ(B(o,r)){rnn3,r2logrn=2.\Lambda(B(o,r))\simeq\left\{\begin{array}[]{ll}r^{n}&n\geq 3,\\ r^{2}\log r&n=2.\end{array}\right.

Let M=M1##MkM=M_{1}\#\cdots\#M_{k} be a manifold with ends. Assume that each end MiM_{i} satisfies (VD), (PI) and that each parabolic end satisfies (RCA). Suppose that for i=1,ki=1,\ldots k

Vi(r)rαi(logr)βi,V_{i}(r)\simeq r^{\alpha_{i}}\left(\log r\right)^{\beta_{i}},

where (α1,β1)(α2,β2)(αk,βk)(\alpha_{1},\beta_{1})\geq(\alpha_{2},\beta_{2})\geq\cdots\geq(\alpha_{k},\beta_{k}) as the lexicographical order. Then

Λ(B(o,r))\displaystyle\Lambda(B(o,r)) V2(r)h2(r)\displaystyle\simeq V_{2}(r)h_{2}(r)
{rα2(logr)β2if (α2,β2)>(2,1),r2logr(loglogr)2if (α2,β2)=(2,1),r2logrif (2,)<(α2,β2)<(2,1),r2if (α2,β2)<(2,).\displaystyle\simeq\left\{\begin{array}[]{ll}r^{\alpha_{2}}(\log r)^{\beta_{2}}&\mbox{if }(\alpha_{2},\beta_{2})>(2,1),\\ r^{2}\log r(\log\log r)^{2}&\mbox{if }(\alpha_{2},\beta_{2})=(2,1),\\ r^{2}\log r&\mbox{if }(2,-\infty)<(\alpha_{2},\beta_{2})<(2,1),\\ r^{2}&\mbox{if }(\alpha_{2},\beta_{2})<(2,-\infty).\end{array}\right.

3. Manifold with ends with oscillating volume functions

3.1. Preliminaries

The purpose of this section is to construct manifolds with ends for which the estimate in (2.8) might not give an optimal bound. To obtain such a manifold, we need a manifold with (VD) and (PI) together with oscillating volume function. First, let us recall the following theorem.

Theorem 3.1 (Grigor’yan and Saloff-Coste [14, Theorem 5.7]).

Let (M,μ)(M,\mu) be a complete non-compact wighted manifold with (PHI) and (RCA) at a reference point oMo\in M. If a positive valued smooth function W:[0,)W:[0,\infty)\rightarrow\mathbb{R} satisfies for all r>0r>0

sup[r,2r]W\displaystyle\sup_{[r,2r]}W Cinf[r,2r]W,\displaystyle\leq C\inf_{[r,2r]}W, (3.1)
0rW2(s)s𝑑s\displaystyle\int_{0}^{r}W^{2}(s)sds CW2(r)r2,\displaystyle\leq CW^{2}(r)r^{2}, (3.2)

then the weighted manifold (M,W2(d(o,))μ)(M,W^{2}(d(o,\cdot))\mu) also satisfies (PHI).

Let (M1,μ1)(M_{1},\mu_{1}) be the 22-dimensional Euclidean space 2\mathbb{R}^{2} with the Euclidean measure. We denote by (M2,μ2)(M_{2},\mu_{2}) a weighted manifold (2,W2(d(o,))μ1)(\mathbb{R}^{2},W^{2}(d(o,\cdot))\mu_{1}), where the positive valued smooth function W:[0,)W:[0,\infty)\rightarrow\mathbb{R} is defined as follows. For α>2\alpha>2, 0<β<20<\beta<2 define a function WW so that for all kk\in\mathbb{N}

0rW2(s)s𝑑s={r20<r<a1,akr<bk,(rbk)αbk2bkr<ck,r2logrckr<dk,(rdk)βdk2logdkdkr<ak+1,\int_{0}^{r}W^{2}(s)sds=\left\{\begin{array}[]{ll}r^{2}&0<r<a_{1},a_{k}\leq r<b_{k},\\ \left(\frac{r}{b_{k}}\right)^{\alpha}b_{k}^{2}&b_{k}\leq r<c_{k},\\ r^{2}\log r&c_{k}\leq r<d_{k},\\ \left(\frac{r}{d_{k}}\right)^{\beta}d_{k}^{2}\log d_{k}&d_{k}\leq r<a_{k+1},\end{array}\right. (3.3)

where the sequences akbk<ckdk<ak+1a_{k}\leq b_{k}<c_{k}\leq d_{k}<a_{k+1} satisfy a1>ea_{1}>e and

bk\displaystyle b_{k} =ck(logck)1α2,\displaystyle=\frac{c_{k}}{(\log c_{k})^{\frac{1}{\alpha-2}}}, (3.4)
ak+1\displaystyle a_{k+1} =dk(logdk)12β\displaystyle=d_{k}(\log d_{k})^{\frac{1}{2-\beta}} (3.5)

(see fig. 4). These sequences will be fixed later.

r2logrr^{2}\log rr2r^{2}rraka_{k}bkb_{k}ckc_{k}dkd_{k}ak+1a_{k+1}
Figure 4. Oscillating volume function (thick line)

Then the function WW satisfies that

W(r){10<r<a1,akr<bk,(rbk)α22bkr<ck,logrckr<dk,logdk(rdk)2β2dkr<ak+1,\displaystyle W(r)\simeq\left\{\begin{array}[]{ll}1&0<r<a_{1},a_{k}\leq r<b_{k},\\ \left(\frac{r}{b_{k}}\right)^{\frac{\alpha-2}{2}}&b_{k}\leq r<c_{k},\\ \sqrt{\log r}&c_{k}\leq r<d_{k},\\ \sqrt{\log d_{k}}\left(\frac{r}{d_{k}}\right)^{-\frac{2-\beta}{2}}&d_{k}\leq r<a_{k+1},\end{array}\right.
W(r){00<r<a1,ak<r<bk,rα42bkα22bk<r<ck,1logr1rck<r<dk,r4β2logdkdk2β2dk<r<ak+1.\displaystyle W^{\prime}(r)\simeq\left\{\begin{array}[]{ll}0&0<r<a_{1},a_{k}<r<b_{k},\\ r^{\frac{\alpha-4}{2}}b_{k}^{-\frac{\alpha-2}{2}}&b_{k}<r<c_{k},\\ \frac{1}{\sqrt{\log r}}\frac{1}{r}&c_{k}<r<d_{k},\\ -r^{-\frac{4-\beta}{2}}\sqrt{\log d_{k}}d_{k}^{\frac{2-\beta}{2}}&d_{k}<r<a_{k+1}.\end{array}\right.

Since

W(r)W(r){00<r<a1,ak<r<bk,1rbk<r<ck,1rlogrck<r<dk,1rdk<r<ak+1,\displaystyle\frac{W^{\prime}(r)}{W(r)}\simeq\left\{\begin{array}[]{ll}0&0<r<a_{1},a_{k}<r<b_{k},\\ \frac{1}{r}&b_{k}<r<c_{k},\\ \frac{1}{r\log r}&c_{k}<r<d_{k},\\ -\frac{1}{r}&d_{k}<r<a_{k+1},\end{array}\right. (3.10)

the condition in (3.1) holds. Indeed, the estimates in (3.10) imply that for any 0<r1<r20<r_{1}<r_{2}

Clogr2r1=r1r2Cdssr1r2W(s)W(s)𝑑s=logW(r2)W(r1)r1r2Cdss=Clogr2r1.-C\log\frac{r_{2}}{r_{1}}=-\int_{r_{1}}^{r_{2}}\frac{Cds}{s}\leq\int_{r_{1}}^{r_{2}}\frac{W^{\prime}(s)}{W(s)}ds=\log\frac{W(r_{2})}{W(r_{1})}\leq\int_{r_{1}}^{r_{2}}\frac{Cds}{s}=C\log\frac{r_{2}}{r_{1}}.

Then we obtain for any 0<r1r22r10<r_{1}\leq r_{2}\leq 2r_{1}

W(r1)W(r2).W(r_{1})\simeq W(r_{2}).

Taking r1r2,r32r1r_{1}\leq r_{2},r_{3}\leq 2r_{1} so that

W(r2)=sup[r1,2r1]W and W(r3)=inf[r1,2r1]W,W(r_{2})=\sup_{[r_{1},2r_{1}]}W\mbox{ and }W(r_{3})=\inf_{[r_{1},2r_{1}]}W,

we conclude the condition in (3.1).

Moreover, we see that

W2(r)r2{r20<r<a1,akr<bk,rαbk2αbkr<ck,r2logrckr<dk,rβdk2βlogdkdkr<ak+1,\displaystyle W^{2}(r)r^{2}\simeq\left\{\begin{array}[]{ll}r^{2}&0<r<a_{1},a_{k}\leq r<b_{k},\\ r^{\alpha}b_{k}^{2-\alpha}&b_{k}\leq r<c_{k},\\ r^{2}\log r&c_{k}\leq r<d_{k},\\ r^{\beta}d_{k}^{2-\beta}\log d_{k}&d_{k}\leq r<a_{k+1},\end{array}\right.

which satisfies the condition in (3.2). Applying Theorem 3.1, the weighted manifold (M2,μ2)=(2,W2(d(o,))μ1)(M_{2},\mu_{2})=(\mathbb{R}^{2},W^{2}(d(o,\cdot))\mu_{1}) satisfies (PHI) 111 We need a smooth modification of the function WW satisfying (3.3) to apply Theorem 3.1. However, we omit the smoothing argument for simplicity. .

Let Z:[0,)Z:[0,\infty)\rightarrow\mathbb{R} be a positive valued smooth function satisfying

Z(r)=1+2logr(r1).Z(r)=\sqrt{1+2\log r}\quad(r\geq 1).

Let (M3,μ3)(M_{3},\mu_{3}) be a weighted manifold (2,Z2(d(o,))μ1)(\mathbb{R}^{2},Z^{2}(d(o,\cdot))\mu_{1}). Then Theorem 3.1 implies also that (M3,μ3)(M_{3},\mu_{3}) satisfies (PHI).

For i=1,2,3i=1,2,3, we denote by Vi(r)V_{i}(r) the volume function on (Mi,μi)(M_{i},\mu_{i}) at oMi=2o\in M_{i}=\mathbb{R}^{2}. Then we obtain

V1(r)\displaystyle V_{1}(r) =πr2,\displaystyle=\pi r^{2},
V2(r)\displaystyle V_{2}(r) =2π0rW2(s)s𝑑s,\displaystyle=2\pi\int_{0}^{r}W^{2}(s)sds,
V3(r)\displaystyle V_{3}(r) =2π01Z2(s)s𝑑s+2πr2logr(r1).\displaystyle=2\pi\int_{0}^{1}Z^{2}(s)sds+2\pi r^{2}\log r~{}~{}(r\geq 1).

It is easy to obtain for sufficiently large r>1r>1

h1(r)logr and h3(r)loglogr.h_{1}(r)\simeq\log r~{}\mbox{ and }~{}h_{3}(r)\simeq\log\log r.

Now we estimate the function h2(r)h_{2}(r). Observe that

akbksdsV2(s)\displaystyle\int_{a_{k}}^{b_{k}}\frac{sds}{V_{2}(s)} =logbkak,\displaystyle=\log\frac{b_{k}}{a_{k}},
bkcksdsV2(s)\displaystyle\int_{b_{k}}^{c_{k}}\frac{sds}{V_{2}(s)} =bkckbkα2s1α𝑑s=1α2(11logck)1,\displaystyle=\int_{b_{k}}^{c_{k}}b_{k}^{\alpha-2}s^{1-\alpha}ds=\frac{1}{\alpha-2}\left(1-\frac{1}{\log c_{k}}\right)\simeq 1,
ckdksdsV2(s)\displaystyle\int_{c_{k}}^{d_{k}}\frac{sds}{V_{2}(s)} =ckdkdsslogs=log(logdklogck),\displaystyle=\int_{c_{k}}^{d_{k}}\frac{ds}{s\log s}=\log\left(\frac{\log d_{k}}{\log c_{k}}\right),
dkak+1sdsV2(s)\displaystyle\int_{d_{k}}^{a_{k+1}}\frac{sds}{V_{2}(s)} =dkak+1s1βdsdk2βlogdk=1(2β)(11logdk)1.\displaystyle=\int_{d_{k}}^{a_{k+1}}\frac{s^{1-\beta}ds}{d_{k}^{2-\beta}\log d_{k}}=\frac{1}{(2-\beta)}\left(1-\frac{1}{\log d_{k}}\right)\simeq 1.

Then we obtain

h2(an)=1+1a1sdsV2(s)+k=1n1akak+1sdsV2(s)k=1n(logbkak+loglogdklogck+1),h_{2}(a_{n})=1+\int_{1}^{a_{1}}\frac{sds}{V_{2}(s)}+\sum_{k=1}^{n-1}\int_{a_{k}}^{a_{k+1}}\frac{sds}{V_{2}(s)}\simeq\sum_{k=1}^{n}\left(\log\frac{b_{k}}{a_{k}}+\log\frac{\log d_{k}}{\log c_{k}}+1\right), (3.11)

which shows that the behavior of h2(r)h_{2}(r) depends on the choice of sequences akbk<ckdk<ak+1a_{k}\leq b_{k}<c_{k}\leq d_{k}<a_{k+1} satisfying (3.4) and (3.5).

3.2. Example 1

For the first case, let us choose sequences akbk<ckdk<ak+1a_{k}\leq b_{k}<c_{k}\leq d_{k}<a_{k+1} so that for any kk\in\mathbb{N}

akbk,ckdk.a_{k}\simeq b_{k},\quad c_{k}\simeq d_{k}. (3.12)

In this case, we obtain the following.

Lemma 3.2.

If a1a_{1} is large enough and the sequences akbk<ckdk<ak+1a_{k}\leq b_{k}<c_{k}\leq d_{k}<a_{k+1} satisfy (3.4), (3.5) and (3.12), then for sufficiently large r>1r>1,

h2(r)logrloglogr.h_{2}(r)\simeq\frac{\log r}{\log\log r}.
Proof.

By the estimate in (3.11), we obtain

h2(an)n.h_{2}(a_{n})\simeq n. (3.13)

Let us consider the behavior ana_{n}. By the assumption in (3.4), we always have

ckbk(logbk)1α2.c_{k}\simeq b_{k}\left(\log b_{k}\right)^{\frac{1}{\alpha-2}}. (3.14)

Assumptions in (3.4) and (3.5) imply that for any kk\in\mathbb{N}

ak+1\displaystyle a_{k+1} =dk(logdk)12βck(logck)12β\displaystyle=d_{k}\left(\log d_{k}\right)^{\frac{1}{2-\beta}}\simeq c_{k}\left(\log c_{k}\right)^{\frac{1}{2-\beta}}
bk(logbk)1α2(log(bk(logbk)1α2))12β\displaystyle\simeq b_{k}\left(\log b_{k}\right)^{\frac{1}{\alpha-2}}\left(\log\left(b_{k}\left(\log b_{k}\right)^{\frac{1}{\alpha-2}}\right)\right)^{\frac{1}{2-\beta}}
=bk(logbk)1α2(logbk+1α2loglogbk)12β\displaystyle=b_{k}\left(\log b_{k}\right)^{\frac{1}{\alpha-2}}\left(\log b_{k}+\frac{1}{\alpha-2}\log\log b_{k}\right)^{\frac{1}{2-\beta}}
bk(logbk)1α2+12β=bk(logbk)γak(logak)γ,\displaystyle\simeq b_{k}\left(\log b_{k}\right)^{\frac{1}{\alpha-2}+\frac{1}{2-\beta}}=b_{k}\left(\log b_{k}\right)^{\gamma}\simeq a_{k}\left(\log a_{k}\right)^{\gamma},

where γ=1α2+12β\gamma=\frac{1}{\alpha-2}+\frac{1}{2-\beta}. Taking a1a_{1} large enough, there exist positive constants c,Cc,C and γ>γ\gamma^{\prime}>\gamma such that for any nn\in\mathbb{N}

cnγnanCnγn.cn^{\gamma n}\leq a_{n}\leq Cn^{\gamma^{\prime}n}.

This implies that for sufficiently large nn\in\mathbb{N}

logannlogn,logloganlogn.\displaystyle\log a_{n}\simeq n\log n,\quad\log\log a_{n}\simeq\log n.

Then we obtain for sufficiently large nn\in\mathbb{N},

loganloglogannh2(an),\frac{\log a_{n}}{\log\log a_{n}}\simeq n\simeq h_{2}(a_{n}),

which concludes the lemma. ∎

Now we estimate the heat kernel on M=M1#M2M=M_{1}\#M_{2}. By the definition of V1(r)V_{1}(r), V2(r)V_{2}(r) and by Lemma 3.2, we obtain for sufficiently large r>1r>1

V1(r)\displaystyle V_{1}(r) =r2,\displaystyle=r^{2}, V2(r){r2akr<bkr2logrckr<dk,\displaystyle V_{2}(r)\simeq\left\{\begin{array}[]{ll}r^{2}&a_{k}\leq r<b_{k}\\ r^{2}\log r&c_{k}\leq r<d_{k},\end{array}\right.
h1(r)\displaystyle h_{1}(r) logr,\displaystyle\simeq\log r, h2(r)logrloglogr,\displaystyle h_{2}(r)\simeq\frac{\log r}{\log\log r},
V1(r)h12(r)\displaystyle V_{1}(r)h_{1}^{2}(r) r2(logr)2,\displaystyle\simeq r^{2}(\log r)^{2}, V2(r)h22(r){r2(logr)2(loglogr)2akr<bkr2(logr)3(loglogr)2ckr<dk.\displaystyle V_{2}(r)h_{2}^{2}(r)\simeq\left\{\begin{array}[]{ll}r^{2}\frac{(\log r)^{2}}{(\log\log r)^{2}}&a_{k}\leq r<b_{k}\\ r^{2}\frac{(\log r)^{3}}{(\log\log r)^{2}}&c_{k}\leq r<d_{k}.\end{array}\right.

According to the heat kernel upper estimate in (2.8), we obtain for ckt<dkc_{k}\leq\sqrt{t}<d_{k}

p(t,o,o)minihi2(t)miniVi(t)hi2(t)(logtloglogt)2t(logt)21t(loglogt)2.p(t,o,o)\leq\frac{\min_{i}h_{i}^{2}(\sqrt{t})}{\min_{i}V_{i}(\sqrt{t})h_{i}^{2}(\sqrt{t})}\simeq\frac{\left(\frac{\log\sqrt{t}}{\log\log\sqrt{t}}\right)^{2}}{t\left(\log\sqrt{t}\right)^{2}}\simeq\frac{1}{t(\log\log t)^{2}}.

Since V(r)maxiVi(r)=r2logrV(r)\simeq\max_{i}V_{i}(r)=r^{2}\log r in this interval, the above upper estimate is much larger than

1V(t)1tlogt,\frac{1}{V(\sqrt{t})}\simeq\frac{1}{t\log t},

which makes difficult to obtain matching lower bound by using [4, Theorem 7.2].

3.3. Example 2

Next, let us choose sequences akbk<ckdk<ak+1a_{k}\leq b_{k}<c_{k}\leq d_{k}<a_{k+1} so that for some δ>1\delta>1 and for any kk\in\mathbb{N}

akbk,ckδdk.a_{k}\simeq b_{k},\quad c_{k}^{\delta}\simeq d_{k}. (3.15)

In this case, we obtain the following.

Lemma 3.3.

If a1a_{1} is large enough and the sequences akbk<ckdk<ak+1a_{k}\leq b_{k}<c_{k}\leq d_{k}<a_{k+1} satisfy (3.4), (3.5) and (3.15), then for sufficiently large r>1r>1,

h2(r)loglogr.h_{2}(r)\simeq\log\log r.
Proof.

The estimate in (3.11) yields that

h2(an)n.h_{2}(a_{n})\simeq n.

By the estimates in (3.14) and (3.15), we obtain

dk\displaystyle d_{k} ckδ(bk(logbk)1α2)δakδ(logak)δα2.\displaystyle\simeq c_{k}^{\delta}\simeq\left(b_{k}\left(\log b_{k}\right)^{\frac{1}{\alpha-2}}\right)^{\delta}\simeq a_{k}^{\delta}\left(\log a_{k}\right)^{\frac{\delta}{\alpha-2}}.

Assumptions in (3.4) and (3.5) imply that

ak+1\displaystyle a_{k+1} =dk(logdk)12β\displaystyle=d_{k}\left(\log d_{k}\right)^{\frac{1}{2-\beta}}
akδ(logak)δα2(δlogak+δα2loglogak)12β\displaystyle\simeq a_{k}^{\delta}\left(\log a_{k}\right)^{\frac{\delta}{\alpha-2}}\left(\delta\log a_{k}+\frac{\delta}{\alpha-2}\log\log a_{k}\right)^{\frac{1}{2-\beta}}
akδ(logak)θ,\displaystyle\simeq a_{k}^{\delta}\left(\log a_{k}\right)^{\theta},

where θ=δα2+12β\theta=\frac{\delta}{\alpha-2}+\frac{1}{2-\beta}. Hence, there exist positive constants c,Cc,C and η>δ\eta>\delta such that for any kk\in\mathbb{N}

cakδak+1Cakη.ca_{k}^{\delta}\leq a_{k+1}\leq Ca_{k}^{\eta}.

This implies that for any nn\in\mathbb{N}

cδ(n1)1δ1a1δ(n1)anCη(n1)1η1a1η(n1).c^{\frac{\delta^{(n-1)}-1}{\delta-1}}a_{1}^{\delta^{(n-1)}}\leq a_{n}\leq C^{\frac{\eta^{(n-1)}-1}{\eta-1}}a_{1}^{\eta^{(n-1)}}.

Then we obtain

δn11δ1logc+δ(n1)loga1loganη(n1)1η1logC+η(n1)loga1.\displaystyle\frac{\delta^{n-1}-1}{\delta-1}\log c+\delta^{(n-1)}\log a_{1}\leq\log a_{n}\leq\frac{\eta^{(n-1)}-1}{\eta-1}\log C+\eta^{(n-1)}\log a_{1}.

Taking a1a_{1} large enough, we obtain for sufficiently large nn\in\mathbb{N}

loglogannh2(an),\log\log a_{n}\simeq n\simeq h_{2}(a_{n}),

which concludes the lemma. ∎

Now let us consider the estimate of the heat kernel on M=M2#M3M=M_{2}\#M_{3}. By the definition of V2(r)V_{2}(r), V3(r)V_{3}(r) and by Lemma 3.3, we obtain for sufficiently large r>1r>1

V2(r)\displaystyle V_{2}(r) {r2akr<bk,r2logrckr<dk,\displaystyle\simeq\left\{\begin{array}[]{ll}r^{2}&a_{k}\leq r<b_{k},\\ r^{2}\log r&c_{k}\leq r<d_{k},\end{array}\right. V3(r)r2logr,\displaystyle\!\!\!\!V_{3}(r)\simeq r^{2}\log r,
h2(r)\displaystyle h_{2}(r) loglogr,\displaystyle\simeq\log\log r, h3(r)loglogr,\displaystyle\!\!\!\!h_{3}(r)\simeq\log\log r,
V2(r)h22(r)\displaystyle V_{2}(r)h_{2}^{2}(r) {r2(loglogr)2akr<bk,r2logr(loglogr)2ckr<dk.\displaystyle\simeq\left\{\begin{array}[]{ll}r^{2}(\log\log r)^{2}&a_{k}\leq r<b_{k},\\ r^{2}\log r(\log\log r)^{2}&c_{k}\leq r<d_{k}.\end{array}\right. V3(r)h32(r)r2logr(loglogr)2.\displaystyle\!\!\!\!V_{3}(r)h_{3}^{2}(r)\simeq r^{2}\log r(\log\log r)^{2}.

Substituting above into (2.8), we obtain for sufficiently large kk\in\mathbb{N} and akt<bka_{k}\leq\sqrt{t}<b_{k}

p(t,o,o)C(loglogt)2t(loglogt)21t.p(t,o,o)\leq C\frac{(\log\log\sqrt{t})^{2}}{t(\log\log\sqrt{t})^{2}}\simeq\frac{1}{t}.

Since V(r)maxiVi(r)r2logrV(r)\simeq\max_{i}V_{i}(r)\simeq r^{2}\log r for all r>1r>1, the above upper estimate is much larger than

1V(t)1tlogt,\frac{1}{V(\sqrt{t})}\simeq\frac{1}{t\log t},

which makes difficult to obtain matching lower bound by using [4, Theorem 7.2].

We hope to prove matching heat kernel lower bounds in forthcoming work.

Acknowledgments

The second author would like to thank Professor Gilles Carron for telling him how to construct manifolds with oscillating volume function with (PHI).

References

  • [1] I. Benjamini, I. Chavel, E. A. Feldman, Heat kernel lower bounds on Riemannian manifolds using the old ideas of Nash. Proc. London Math. Soc., 72 (1996) 215–240.
  • [2] M. Cai, Ends of Riemannian manifolds with nonnegative Ricci curvature outside a compact set. Bull. A.M.S., 24 (1991) no. 2, 371-377.
  • [3] G. Carron, Riesz transforms on connected sums. Festival Yves Colin de Verdière. Ann. Inst. Fourier (Grenoble) 57 (2007), no. 7, 2329-–2343.
  • [4] T. Coulhon, A. Grigor’yan, On-diagonal lower bounds for heat kernels on no-compact manifolds and Markov chains. Duke Math. J., 89 (1997) no. 1, 133-199.
  • [5] H. Doan, Boundedness of maximal functions on nondoubling parabolic manifolds with ends. To appear in J. Aust. Math. Soc.
  • [6] X. T. Duong, J. Li and A. Sikora, Boundedness of maximal functions on non-doubling manifolds with ends. Proc. Centre Math. Appl. Austral. 45 (2012), 37–-47.
  • [7] A. Grigor’yan, The heat equation on noncompact Riemannian manifolds (in Russian). Mat. Sb., 182 (1991) no. 1, 55-87; English translation in Math. USSR-Sb., 72 (1992) no. 1, 47–77.
  • [8] A. Grigor’yan, Analytic and geometric background of recurrence and non-explosion of the Brownian motion on Riemannian manifolds. Bull. AMS, 36 (1999) 135–249.
  • [9] A. Grigor’yan, S. Ishiwata and L. Saloff-Coste, Heat kernel estimates on connected sums of parabolic manifolds, J. Math. Pures Appl. 113(2018), 155-194.
  • [10] A. Grigor’yan, S. Ishiwata and L. Saloff-Coste, Poincaré constants on manifolds with ends, in preparation.
  • [11] A. Grigor’yan and L. Saloff-Coste, Heat kernel on connected sums of Riemannian manifolds. Math. Research Letters, 6 (1999) no. 3-4, 307–321.
  • [12] A.Grigor’yan and L. Saloff-Coste, Dirichlet heat kernel in the exteior of a compact set. Comm. Pure Appl. Math, 55 (2002), 93–133.
  • [13] A. Grigor’yan and L. Saloff-Coste, Hitting probabilities for Brownian motion on Riemannian manifolds. J. Math. Pures Appl., 81 (2002) no. 2, 115–142.
  • [14] A. Grigor’yan and L. Saloff-Coste, Stability results for Harnack inequalities. Ann. Inst. Fourier, Grenoble, 55 (2005) no. 3, 825–890.
  • [15] A. Grigor’yan and L. Saloff-Coste, Heat kernel on manifolds with ends. Ann. Inst. Fourier, Grenoble, 59 (2009) no. 5, 1917–1997.
  • [16] A. Grigor’yan and L. Saloff-Coste, Surgery of the Faber-Krahn inequality and applications to heat kernel bounds, J. Nonlinear Analysis, 131 (2016) 243–272.
  • [17] A. Hassel, D. Nix and A. Sikora, Riesz transforms on a class of non-doubling manifolds II. arXiv:1912.06405.
  • [18] A. Hassel and A. Sikora, Riesz transforms on a class of non-doubling manifolds. Comm. Partial Differential Equations 44 (2019) no. 11, 1072–-1099.
  • [19] A. Kasue, Harmonic functions with growth conditions on a manifold of asymptotically nonnegative curvature I, Geometry and Analysis on Manifolds (Ed. by T. Sunada), Lecture Notes in Math., 1339, Springer-Verlag (1988), 158-181.
  • [20] S. Kusuoka, D. Stroock, Application of Malliavin calculus, Part III, J. Fac. Sci. Tokyo Univ., Sect. 1A, Math., 34 (1987) 391–442.
  • [21] Yu. T. Kuz’menko and S. A. Molchanov, Counterexamples of Liouville type theorems, (in Russian) Vestnik Moscov. Univ. Ser. I Mat. Mekh., (1976) no. 6, 39–43; Engl. transl. Moscow Univ. Math. Bull., 34 (1979) 35–39.
  • [22] P. Li and L-F Tam, Positive harmonic functions on complete manifolds with non-negative curvature outside a compact set, Ann. Math., 125 (1987), 171-207.
  • [23] P. Li and S.-T. Yau, On the parabolic kernel of the Schrödinger operator, Acta Math., 156 (1986) no. 3-4, 153–201.
  • [24] J. Moser, A Harnack inequality for parabolic differential equations, Comm. Pure Appl. Math., 17 (1964) 101–134. Correction: Comm. Pure Appl. Math. 20 (1967) 231-236.
  • [25] L. Saloff-Coste, A note on Poincaré, Sobolev, and Harnack inequalities. Internat. Math. Res. Notices (1992), no. 2, 27–38.
  • [26] L. Saloff-Coste, Aspects of Sobolev type inequalities. London Math. Soc. Lecture Notes Series 289, Cambridge Univ. Press, 2002.