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

Heat kernel asymptotics on the free step-two Carnot group with 33 generators

Hong-Quan Li, Sheng-Chen Mao, Ye Zhang

Abstract. In this work, we establish the uniform heat kernel asymptotics as well as sharp bounds for its derivatives on the free step-two Carnot group with 33 generators. As a by-product, on this highly non-trivial toy model, we completely solve the Gaveau–Brockett problem, in other words, we obtain the expression of the squared Carnot–Carathéodory distance, as explicitly as one can possibly hope for. Furthermore, the precise estimates of the heat kernel, and its small-time asymptotic behaviors are deduced.

Mathematics Subject Classification (2020): 58J37; 35B40; 35H10; 35B45; 35K08; 43A80; 58J35; 43A85

Key words and phrases: Asymptotic behavior; Carnot–Carathéodory distance; Free step-two Carnot group; Heat kernel; Precise estimates

1 Introduction

The heat semi-group and its kernel play a crucial role in various areas of Mathematics. See for example [39, 10, 46, 53, 1, 17, 47, 3, 23, 5] and huge references therein.

We restrict our attention to the heat kernel in the setting of Carnot groups. Let =2,3,\ell=2,3,\ldots. According to [53, p. 44] or [12, Definition 2.2.3], a connected and simply connected Lie group 𝔾\mathbb{G} is said to be step-\ell Carnot group (or stratified group) if its left-invariant Lie algebra 𝔤\mathfrak{g} admits a direct sum decomposition

𝔤=𝔤1𝔤,[𝔤1,𝔤j1]=𝔤j, 2j,[𝔤1,𝔤]={0},\displaystyle\mathfrak{g}=\mathfrak{g}_{1}\oplus\ldots\oplus\mathfrak{g}_{\ell},\quad[\mathfrak{g}_{1},\mathfrak{g}_{j-1}]=\mathfrak{g}_{j},\ 2\leq j\leq\ell,\quad[\mathfrak{g}_{1},\mathfrak{g}_{\ell}]=\{0\},

where [,][\cdot,\cdot] denotes the Lie bracket on 𝔤\mathfrak{g}. We say that 𝔤1\mathfrak{g}_{1} is the first slice in the stratification above. We identify 𝔾\mathbb{G} with 𝔤\mathfrak{g} via the exponential map, and fix a (bi-invariant) Haar measure dgdg on 𝔾\mathbb{G}, namely the lift of Lebesgue measure on 𝔤\mathfrak{g} via exp\exp. Fix a basis 𝕏={X1,,Xn}{\mathbb{X}}=\{{\mathrm{X}}_{1},\cdots,{\mathrm{X}}_{n}\} for 𝔤1\mathfrak{g}_{1} and consider the sub-Laplacian Δ=j=1nXj2\Delta=\sum_{j=1}^{n}{\mathrm{X}}_{j}^{2}. Let php_{h} (h>0h>0) denote the associated heat kernel, i.e. the fundamental solution of hΔ\frac{\partial}{\partial h}-\Delta. It is well-known that 0<phC(+×𝔾)0<p_{h}\in C^{\infty}(\mathbb{R}^{+}\times\mathbb{G}), and (see for example [53])

ehΔf(g)=fph(g)=𝔾f(g)ph(g1g)𝑑g.e^{h\,\Delta}f(g)=f*p_{h}(g)=\int_{\mathbb{G}}f(g_{*})\,p_{h}(g_{*}^{-1}\,g)\,dg_{*}.

We use d:=d𝕏d:=d_{{\mathbb{X}}} to denote the Carnot–Carathéodory distance defined by 𝕏{\mathbb{X}}, which inherits the left-invariant property (see e.g. [53, III. 4]), i.e.

d(gg1,gg2)=d(g1,g2),g,g1,g2𝔾.d(g\,g_{1},g\,g_{2})=d(g_{1},g_{2}),\qquad\forall\,g,g_{1},g_{2}\in\mathbb{G}.

We shall therefore put d(g):=d(o,g)d(g):=d(o,g), where oo is the identity of 𝔾\mathbb{G}. Let QQ denote the homogeneous dimension of 𝔾\mathbb{G}, namely, Q=j=1jdim𝔤jQ=\sum_{j=1}^{\ell}j\,{\rm dim}\,\mathfrak{g}_{j}.

There is a natural family of dilations on 𝔤\mathfrak{g} defined for r>0r>0 as follows:

δr(i=1vi):=i=1rivi,with vi𝔤i.\displaystyle\delta_{r}\left(\sum_{i=1}^{\ell}v_{i}\right):=\sum_{i=1}^{\ell}r^{i}v_{i},\quad\mbox{with $v_{i}\in\mathfrak{g}_{i}$}.

This induces the definition of dilation on 𝔾\mathbb{G}, which we still denote by δr\delta_{r}. The following scaling properties are well-known

d(δh(g))=hd(g),ph(g)=hQ2p1(δ1h(g)),h>0,g𝔾.\displaystyle d(\delta_{h}(g))=h\,d(g),\quad p_{h}(g)=h^{-\frac{Q}{2}}p_{1}(\delta_{\frac{1}{\sqrt{h}}}(g)),\quad\,h>0,\ g\in\mathbb{G}. (1.1)

Let =(X1,,Xn)\nabla=({\mathrm{X}}_{1},\ldots,{\mathrm{X}}_{n}) denote the horizontal gradient. The classical Li–Yau estimates for the heat kernel and their improvements (as well as their wide application) can be found in some much more general situations than Carnot groups. See for instance [20, 45, 52, 53, 16, 49, 50, 18, 14] and the references therein. In particular, it holds for any h>0h>0 and all g𝔾g\in\mathbb{G} that:

ph(g)ChQ2(1+d(g)h)Q1ed(g)24h,\displaystyle p_{h}(g)\leq C\,h^{-\frac{Q}{2}}\,\left(1+\frac{d(g)}{\sqrt{h}}\right)^{Q-1}\,e^{-\frac{d(g)^{2}}{4h}}, (1.2)
ph(g)C(ϖ)hQ2ed(g)24(1ϖ)h,0<ϖ<1,\displaystyle p_{h}(g)\geq C(\varpi)\,h^{-\frac{Q}{2}}\,e^{-\frac{d(g)^{2}}{4\,(1-\varpi)h}},\qquad 0<\varpi<1, (1.3)
|ph(g)|ChQ+12(1+d(g)h)3Q+1ed(g)24h,\displaystyle|\nabla p_{h}(g)|\leq C\,h^{-\frac{Q+1}{2}}\,\left(1+\frac{d(g)}{\sqrt{h}}\right)^{3Q+1}\,e^{-\frac{d(g)^{2}}{4h}}, (1.4)

where the absolute constant C>0C>0 is independent of (h,g)(h,g), and C(ϖ)C(\varpi) dependent only on ϖ\varpi. Moreover, we have for all h>0h>0 and g𝔾g\in\mathbb{G} that

|jhjkph(g)|C(k,j,ϖ)hjk2Q2ed(g)24(1+ϖ)h,k,j=1,2,, 0<ϖ1.\displaystyle\left|\frac{\partial^{j}}{\partial h^{j}}\nabla^{k}p_{h}(g)\right|\leq C(k,j,\varpi)\,h^{-j-\frac{k}{2}-\frac{Q}{2}}\,e^{-\frac{d(g)^{2}}{4\,(1+\varpi)h}},\qquad k,j=1,2,\ldots,\ 0<\varpi\leq 1. (1.5)

By the fact that hph=Δph\frac{\partial}{\partial h}p_{h}=\Delta p_{h}, the last claim is equivalent to

|lph(g)|C(l,ϖ)hl2Q2ed(g)24(1+ϖ)h,l=1,2,, 0<ϖ1.\displaystyle\left|\nabla^{l}p_{h}(g)\right|\leq C(l,\varpi)\,h^{-\frac{l}{2}-\frac{Q}{2}}\,e^{-\frac{d(g)^{2}}{4\,(1+\varpi)h}},\qquad l=1,2,\ldots,\ 0<\varpi\leq 1. (1.6)

A direct consequence of (1.2) and (1.3) is the following well-known Varadhan’s formula, which is actually valid in a very general circumstance than Carnot groups (cf. eg. [28], [2] and the references therein):

limh0+4hlnph(g)=d(g)2,g𝔾.\displaystyle\lim_{h\to 0^{+}}4\,h\ln{p_{h}(g)}=-d(g)^{2},\qquad\forall\,g\in\mathbb{G}. (1.7)

In this work, we investigate much more complicated problems: uniform asymptotic behaviours at infinity for php_{h} in the sense of

ph(g)=hQ2Θ(g,h)ed(g)24h(1+o(1)),asd(g)h+,\displaystyle p_{h}(g)=h^{-\frac{Q}{2}}\,\Theta(g,h)\,e^{-\frac{d(g)^{2}}{4h}}\,\left(1+o(1)\right),\qquad\mbox{as}\quad\frac{d(g)}{\sqrt{h}}\to+\infty, (1.8)

as well as sharp estimates for its derivatives

|lnph(g)|Cd(g)h,h>0,g𝔾,\displaystyle|\nabla\ln{p_{h}(g)}|\leq C\,\frac{d(g)}{h},\qquad h>0,\ g\in\mathbb{G}, (1.9)
|lph(g)|C(l)hl2(1+d(g)h)lph(g),l=2,3,,h>0,g𝔾.\displaystyle\left|\nabla^{l}p_{h}(g)\right|\leq C(l)\,h^{-\frac{l}{2}}\,\left(1+\frac{d(g)}{\sqrt{h}}\right)^{l}\,p_{h}(g),\qquad l=2,3,\ldots,\ h>0,\ g\in\mathbb{G}. (1.10)

where Θ(g,h)\Theta(g,h) has at most polynomial growth w.r.t. d(g)/hd(g)/\sqrt{h}. Here and in the sequel, we use the notation f=o(w)f=o(w) if limfw=0\lim\frac{f}{w}=0. Notice that by the classical Li–Yau estimates, a direct consequence of (1.8) is the precise bounds for the heat kernel. Also recall that these stronger estimates play a crucial role in [26, 29, 4, 24, 13, 34] for instance.

Another by-product of (1.8) is the complete description of short-time behaviour,

ph(g)=Θh(g)ed(g)24h(1+og(1)),as h0+,go,\displaystyle p_{h}(g)=\Theta_{h}(g)\,e^{-\frac{d(g)^{2}}{4h}}\left(1+o_{g}(1)\right),\qquad\mbox{as $h\to 0^{+}$,}\ g\neq o, (1.11)

where Θh(g)>0\Theta_{h}(g)>0 is frequently of type C(g)hσ(g)C(g)\,h^{-\sigma(g)}. Sometimes, it allows us to determine the cut locus for the identity Cuto\mathrm{Cut}_{o} on the underlying group, namely the set of points where the squared distance is not smooth.

We point out that the satisfactory explicit formula of the heat kernel for Carnot groups with step 3\ell\geq 3 is still missing. There are some attempts to study the short-time behaviour of php_{h} on Carnot groups, by means of the generalized Fourier transform and the Trotter product formula. However, as far as the authors know, it is unclear whether we can recover correctly the well-known Varadhan’s formulas (1.7) along the way, even in the simplest Carnot group (namely the Heisenberg group of dimension 3) without using the explicit formula of the heat kernel. For example, C. Séguin and A. Mansouri have conducted such investigation in [48], but the “remainder terms” in their main results (cf. e.g. (2), (3) as well as the estimate following (41) in [48]) are not real remainders since they are of polynomial decay, which is much larger than the exponential decay of the heat kernel by Varadhan’s formula (1.7). Also notice that the deduction from (38) to (39) in [48] is incorrect due to the same reason.

From now on, we focus on step-two Carnot groups, in other words, =2\ell=2. In such case, php_{h} can be written as the partial Fourier transform w.r.t. the center of the underlying group. See for example [19] or [7]. In particular, there exist two key functions, scotss\cot s in the phase and ssins\frac{s}{\sin s} in the amplitude.

Recall that determining the cut locus and the distance between two points on the underlying space are two fundamental problems in (Riemannian) geometry and optimal control. It seems that the second one is much more difficult than the first one. As pointed out in [9, § 6.5.4], the distance and the cut locus of a Riemannian manifold cannot in general be explicitly computed. To our best knowledge, few has been known even in the setting of step-two Carnot groups endowed with a left invariant Riemannian metric. As for the corresponding sub-Riemannian case, there are masses of works, but only limited to nonisotropic Heisenberg groups and H-type groups, cf. e.g. [21, 8, 51, 44]. Their method is quite standard, namely finding all the possible candidates (which is usually to solve some differential equations with some boundary conditions) and picking out the shortest one.

Recently, the first author introduces an original, new and very powerful method in [33] to attack these problems, say the operator convexity (of the key function scots-s\cot{s}). More precisely, by combining it with the method of stationary phase and Varadhan’s formulas, the distance and the cut locus are characterized in an enormous kind of step-two Carnot groups, say GM-groups (cf. [33, Corollary 2.3 and Theorem 2.7], also [38, Theorem 4, Corollary 9] for various equivalent characterizations of GM-groups via basic geometric properties). Furthermore, for a given step-two Carnot group 𝔾\mathbb{G}, by using the known sub-Riemannian exponential mapping in addition, up to a set of measure zero, all normal geodesics from the origin to any given point g𝔾g\in\mathbb{G} are characterized, by [33, Theorem 2.4], which implies more qualitative results of d(g)2d(g)^{2}.

It follows from [33, § 11] and [38, Corollary 13] that the simplest step-two non-GM group is the free step-two Carnot group with 33 generators, N3,2N_{3,2} (see below for a definition). In such case, the long-standing open problem of Gaveau–Brockett is completely solved in [33, § 11] and [38, § 7] (cf. also Theorem 2.2 and Corollary 2.3 below) by means of two quite different methods, which rely heavily on the Hamilton–Jacobi theory. Also notice that the results of [33, 38] are very dependent on the properties of scotss\cot{s}, but the other key function ssins\frac{s}{\sin{s}} is not used at all.

Unlike in the setting of GM-groups, the expression of d2d^{2} on N3,2N_{3,2} cannot be well explained directly by the integral expression for the heat kernel. Motivated by this new phenomenon, the first and third authors naturally turned to searching for a more essential explanation by studying the small-time behavior for the heat kernel in such case. In particular, we discovered the expression (2.30) below where the properties of Bessel functions have been used, since (ssins)1=π2s12J12(s)(\frac{s}{\sin{s}})^{-1}=\sqrt{\frac{\pi}{2}}\,s^{-\frac{1}{2}}J_{\frac{1}{2}}(s). See Section 4 below for more details. Furthermore, by using the recursion formulas for the Bessel functions and the Fourier transform of the Gaussian functions, the initial expression in the form of oscillatory integral for the heat kernel has been reformulated into a much more suitable one in the form of Laplace integral with positive integrand (cf. [37, (15), (16) and Proposition 4]). Following main ideas in [33] (with some new methods introduced), the new expression of the heat kernel allows us to understand some basic geometric problems well in the setting of step-two Carnot groups. For instance, at least in principle, the extremely difficult problem of determining the distance from any given point to the origin is reduced to an elementary calculus, see [37, Algorithm 1] for more details. We point out that the methods in [37] can be adapted to study the corresponding Riemannian geometric problems on any step-two Carnot group.

One goal of this article is to re-obtain the expression of d2d^{2} on N3,2N_{3,2} via the heat kernel, without using the Hamilton–Jacobi theory. Indeed it can be considered as an illuminative model as well as application of [37]. Notice that in this article, we do not care about any other geometric problems except the expression for d2d^{2}. Hence some reductions such as (2.21) will be used. Moreover, for this article to be self-contained, Theorem 2.1 and Corollary 2.3, as well as their proofs, namely main parts in Sections 3 and 5 (up to a slight modification), are extracted from [33, 38].

On the other hand, by adapting the method introduced in [33], refining and extending the argument in the proof of [32, Theorem 1.3] (also of [36, Theorem 3], the first and third authors study (1.8)-(1.10) in [35] on a large class of step-two Carnot groups which are simultaneously GM and Métivier groups.

Notice that N3,2N_{3,2} is neither a GM-group nor a Métivier group. The main goal of the work is to consider N3,2N_{3,2} as a toy model in order to study (1.8)-(1.10). We hope it provides a routine for (at least concrete) step-two Carnot groups. Although some new original and powerful methods are provided in [33, 37, 35], we will find that they are not enough and it is still very challenging to complete our mission on N3,2N_{3,2}. More explanation about our major obstacles can be found in Subsection 2.5.

As an application of results obtained in this work, we will establish the gradient estimate for the heat semi-group on N3,2N_{3,2} in a forthcoming paper. Moreover, the method presented here will be adapted to study the heat kernel associated to the full Laplacian (which corresponds to the canonical left-invariant Riemannian structure) on N3,2N_{3,2} as well. Precise estimates of the heat kernel for the Riemannian case will be provided in a future work.

1.1 Notation

We denote the sets of integers and nonnegative integers by {\mathbb{Z}} and {\mathbb{N}} respectively, and put :={0}{\mathbb{Z}}^{*}:={\mathbb{Z}}\setminus\{0\}, :={0}{\mathbb{N}}^{*}:={\mathbb{N}}\setminus\{0\}. Given a multi-index α=(α1,α2,α3)3\alpha=(\alpha_{1},\alpha_{2},\alpha_{3})\in\mathbb{N}^{3} and x=(x1,x2,x3)x=(x_{1},x_{2},x_{3}), λ=(λ1,λ2,λ3)3\lambda=(\lambda_{1},\lambda_{2},\lambda_{3})\in\mathbb{R}^{3}, we write xα:=x1α1x2α2x3α3\partial^{\alpha}_{x}:=\partial^{\alpha_{1}}_{x_{1}}\partial^{\alpha_{2}}_{x_{2}}\partial^{\alpha_{3}}_{x_{3}}, λα:=λ1α1λ2α2λ3α3\lambda^{\alpha}:=\lambda_{1}^{\alpha_{1}}\lambda_{2}^{\alpha_{2}}\lambda_{3}^{\alpha_{3}}, and |α|:=α1+α2+α3|\alpha|:=\alpha_{1}+\alpha_{2}+\alpha_{3} in the usual way; we also use the notation that λ=(λ1,λ)×2\lambda=(\lambda_{1},\lambda^{\prime})\in{\mathbb{R}}\times{\mathbb{R}}^{2}. The real part and imaginary part of a complex number zz are denoted by (z)\Re(z) and (z)\Im(z) respectively. Supposing that qq\in{\mathbb{N}}^{*}, we let 𝕀q\mathbb{I}_{q} represent the q×qq\times q identity matrix, and Oq{\rm O}_{q} the orthogonal group of order qq.

The symbols CC and cc are used throughout to denote implicit positive constants which may vary from one line to the next. And when necessary we will specify with a subscript which parameters the values of CC and cc depend on.

The usual asymptotic notation is employed. Let ww be a non-negative real-valued function. By f=O(w)f=O(w) (resp. fwf\lesssim w if ff is also real-valued) we mean |f|Cw|f|\leq Cw (resp. fCwf\leq Cw). Correspondingly fwf\gtrsim w if fCwf\geq C\,w. Moreover, we will use the counterparts of such notation for Hermitian (in particular, real symmetric) operators or matrices. If the implicit constant depends on parameter α0\alpha_{0}, we will write α0\alpha_{0} in the subscript of O,o,,O,o,\lesssim,\sim, etc.

The statement fwf\ll w, f=o(w)f=o(w) or wfw\gg f is shorthand for f(x)/w(x)0f(x)/w(x)\to 0 as xx tends to some x0x_{0} (could be \infty); it will be clear that which variable is taken to the limit. Especially, if f1f\equiv 1 we also write w+w\to+\infty. The statement f=w(1+o(1))f=w(1+o(1)) means fw=o(w)f-w=o(w).

For example, we use “if A1A\lesssim 1 and B+B\to+\infty then E1E\sim 1” to represent that “for every ζ0>0\zeta_{0}>0, there exists a constant C(ζ0)C(\zeta_{0}) large enough and a constant C(ζ0)1C^{\prime}(\zeta_{0})\geq 1 such that when Aζ0A\leq\zeta_{0} and BC(ζ0)B\geq C(\zeta_{0}), we have C(ζ0)1EC(ζ0)C^{\prime}(\zeta_{0})^{-1}\leq E\leq C^{\prime}(\zeta_{0})”.

Finally, all the vectors appearing in this work are regarded as column vectors unless otherwise stated. However, we may write a column vector tt in q\mathbb{R}^{q} with scalar coordinates t1,,tqt_{1},\ldots,t_{q}, simply as (t1,,tq)(t_{1},\ldots,t_{q}). Moreover, for a function F:=F(ν,ν)F:=F(\nu,\nu^{\prime}) depending on (ν,ν)q1×q2(\nu,\nu^{\prime})\in\mathbb{R}^{q_{1}}\times\mathbb{R}^{q_{2}}, we will use the following notation to denote the ν\nu-gradient (resp. ν\nu-Hessian matrix) at the point (ν0,ν0)(\nu_{0},\nu^{\prime}_{0}):

ν0F(ν0,ν0),(resp.Hessν0F(ν0,ν0)).\nabla_{\nu_{0}}F(\nu_{0},\nu^{\prime}_{0}),\quad(\mbox{resp.}\ \mathrm{Hess}_{\nu_{0}}F(\nu_{0},\nu^{\prime}_{0})).

2 Description of the setting and statements of results

Recall the free step-two Carnot group with 3 generators N3,2N_{3,2} is given by 3×3\mathbb{R}^{3}\times\mathbb{R}^{3} with the group structure

(x,t)(x,t)=(x+x,t+t12x×x),(x,t)\cdot(x^{\prime},t^{\prime})=\left(x+x^{\prime},t+t^{\prime}-\frac{1}{2}\,x\times x^{\prime}\right),

where “×\times” denotes the cross product on 3\mathbb{R}^{3}, i.e.,

x×x=(x2x3x3x2,x3x1x1x3,x1x2x2x1).x\times x^{\prime}=(x_{2}x_{3}^{\prime}-x_{3}x_{2}^{\prime},x_{3}x_{1}^{\prime}-x_{1}x_{3}^{\prime},x_{1}x_{2}^{\prime}-x_{2}x_{1}^{\prime}).

Here we choose the same definition of N3,2N_{3,2} as the one in [33, § 11] rather than the one in [38]. In fact, they differ in a negative sign before the term x×xx\times x^{\prime} and it will not affect the expression of the heat kernel. The corresponding left-invariant vector fields and sub-Laplacian are given by

X1:=x112x3t2+12x2t3,X2:=x2+12x3t112x1t3,X3:=x312x2t1+12x1t2,Δ:=j=13Xj2.\begin{gathered}{\mathrm{X}}_{1}:=\frac{\partial}{\partial x_{1}}-\frac{1}{2}x_{3}\frac{\partial}{\partial t_{2}}+\frac{1}{2}x_{2}\frac{\partial}{\partial t_{3}},\quad{\mathrm{X}}_{2}:=\frac{\partial}{\partial x_{2}}+\frac{1}{2}x_{3}\frac{\partial}{\partial t_{1}}-\frac{1}{2}x_{1}\frac{\partial}{\partial t_{3}},\\ {\mathrm{X}}_{3}:=\frac{\partial}{\partial x_{3}}-\frac{1}{2}x_{2}\frac{\partial}{\partial t_{1}}+\frac{1}{2}x_{1}\frac{\partial}{\partial t_{2}},\qquad\Delta:=\sum_{j=1}^{3}{\mathrm{X}}_{j}^{2}.\end{gathered} (2.1)

It is well-known that the associated heat kernel php_{h} (h>0h>0), i.e. the fundamental solution of hΔ\frac{\partial}{\partial h}-\Delta, has the following integral formulas (cf. e.g. [21, 19] or [40])

ph(x,t)=𝐂h92p(xh,th),h>0,(x,t)N3,2,\displaystyle p_{h}(x,t)=\frac{\mathbf{C}}{h^{\frac{9}{2}}}\,p\left(\frac{x}{\sqrt{h}},\frac{t}{h}\right),\quad\forall\,h>0,\,(x,t)\in N_{3,2}, (2.2)

with some positive constant 𝐂\mathbf{C} and

p(x,t)=3𝐕(λ)e14ϕ~((x,t);λ)𝑑λ,\displaystyle p(x,t)=\,\int_{\mathbb{R}^{3}}\mathbf{V}(\lambda)\,e^{-\frac{1}{4}\widetilde{\phi}((x,t);\lambda)}\,d\lambda, (2.3)

where

𝐕(λ):=|λ|sinh|λ|=j=1+(1+λλj2π2)1\displaystyle\mathbf{V}(\lambda):=\frac{|\lambda|}{\sinh{|\lambda|}}=\prod_{j=1}^{+\infty}\left(1+\frac{\lambda\cdot\lambda}{j^{2}\pi^{2}}\right)^{-1} (2.4)

by [22, 1.431.2] and

ϕ~((x,t);λ):=|x|2+|λ|coth|λ|1|λ|2(|λ|2|x|2(λx)2)4itλ.\displaystyle\widetilde{\phi}((x,t);\lambda):=|x|^{2}+\frac{|\lambda|\coth{|\lambda|}-1}{|\lambda|^{2}}\,(|\lambda|^{2}|x|^{2}-(\lambda\cdot x)^{2})-4it\cdot\lambda. (2.5)

The corresponding reference function in our setting is defined by (cf. [33] and [35])

ϕ((x,t);τ):=ϕ~((x,t);iτ)=|x|21|τ|cot|τ||τ|2(|τ|2|x|2(τx)2)+4tτ.\displaystyle\phi((x,t);\tau):=\widetilde{\phi}((x,t);i\tau)=|x|^{2}-\frac{1-|\tau|\cot{|\tau|}}{|\tau|^{2}}\,(|\tau|^{2}|x|^{2}-(\tau\cdot x)^{2})+4t\cdot\tau. (2.6)

Finally, the following equivalence between the Carnot–Carathéodory distance and a homogeneous norm (see for example [53, III.4]) will be used again and again:

d(x,t)2|x|2+|t|,(x,t)N3,2.\displaystyle d(x,t)^{2}\sim|x|^{2}+|t|,\qquad\forall\,(x,t)\in N_{3,2}. (2.7)

2.1 Reductions

Our main target of this work is to establish the uniform heat kernel asymptotics at infinity on N3,2N_{3,2} and the following simple observation will simplify our asymptotic problem to a large extent. In fact, using the orthogonal invariant property and the symmetry w.r.t. the origin of the heat kernel, we reduce the original problem (which is 66-dimensional) to a 33-dimensional one. Moreover, when deriving the explicit expression of the Carnot–Carathéodory distance, we can even reduce to a 22-dimensional problem; see the beginning of Subsection 2.3 for more details.

To be more precise, through the change of variables λOλ\lambda\mapsto O\lambda and λλ\lambda\mapsto-\lambda, (2.3)-(2.5) imply that

p(Ox,Ot)=p(x,t),OO3,p(x,t)=p(x,t).p(O\,x,O\,t)=p(x,t),\quad\forall\,O\in\mathrm{O}_{3},\qquad p(x,t)=p(x,-t). (2.8)

From this, together with the smoothness of pp and the benefit that our heat kernel asymtotics will be uniform enough, without loss of generality, we may assume that

x=|x|e1=|x|(1,0,0)0 and t=(t1,t2,0), with t1,t2>0.\displaystyle x=|x|\,e_{1}=|x|(1,0,0)\neq 0\quad\mbox{ and }\quad t=(t_{1},t_{2},0),\mbox{ with }t_{1},t_{2}>0. (2.9)

Here and in the sequel, e1e_{1} will simultaneously denote the vectors in Euclidean spaces (possibly with different dimensions) with 1 in the first coordinate and zeros elsewhere. Further reduction will be provided in Assumption (A) below.

2.2 Two key analytic diffeomorphisms

Set in the sequel

>,+2:={(u1,u2)2;u2>2πu1>0},\displaystyle\mathbb{R}^{2}_{>,+}:=\left\{(u_{1},u_{2})\in\mathbb{R}^{2};\,u_{2}>\frac{2}{\sqrt{\pi}}\sqrt{u_{1}}>0\right\}, (2.10)
<,+2:={(u1,u2);u1>0,0<u2<2πu1},\displaystyle\mathbb{R}_{<,+}^{2}:=\left\{(u_{1},u_{2});\,u_{1}>0,0<u_{2}<\frac{2}{\sqrt{\pi}}\sqrt{u_{1}}\right\}, (2.11)

Ω+,1:={v:=(v1,v2)2;v1,v2>0,v12+v22<π2},\displaystyle\Omega_{+,1}:=\{v:=(v_{1},v_{2})\in\mathbb{R}^{2};\,v_{1},v_{2}>0,v_{1}^{2}+v_{2}^{2}<\pi^{2}\}, (2.12)
ψ(ρ):=1ρcotρρ2,ρ{kπ;k},\displaystyle\psi(\rho):=\frac{1-\rho\,\cot{\rho}}{\rho^{2}},\qquad\rho\not\in\left\{k\,\pi;\,k\in{\mathbb{Z}}^{*}\right\}, (2.13)

K3(v1,v2):=2ψ(r)+ψ(r)rv22 with r:=|v|=v12+v22,\displaystyle\mathrm{K}_{3}(v_{1},v_{2}):=2\psi(r)+\frac{\psi^{\prime}(r)}{r}v_{2}^{2}\ \mbox{ with }\ r:=|v|=\sqrt{v_{1}^{2}+v_{2}^{2}}, (2.14)
2π<4.4933<ϑ1<4.4935<32π such that tanϑ1=ϑ1,\displaystyle\sqrt{2}\pi<4.4933<\vartheta_{1}<4.4935<\frac{3}{2}\pi\ \mbox{ such that }\tan{\vartheta_{1}}=\vartheta_{1}, (2.15)
Ω,4:={(v1,v2);v2<0,π<v1<r<ϑ1,K3(v1,v2)<0},\displaystyle\Omega_{-,4}:=\left\{(v_{1},v_{2});\,v_{2}<0,\pi<v_{1}<r<\vartheta_{1},\mathrm{K}_{3}(v_{1},v_{2})<0\right\}, (2.16)

and

Λ(v1,v2):=v2[ψ(r)rv2v+2ψ(r)e2],0<r=|v|π<ϑ1.\displaystyle\Lambda(v_{1},v_{2}):=v_{2}\left[\frac{\psi^{\prime}(r)}{r}\,v_{2}\,v+2\,\psi(r)\,e_{2}\right],\qquad 0<r=|v|\neq\pi<\vartheta_{1}. (2.17)

Notice that if (u1,u2)=Λ(v1,v2)(u_{1},u_{2})=\Lambda(v_{1},v_{2}), then Λ(±v1,±v2)=(±u1,±u2)\Lambda(\pm v_{1},\pm v_{2})=(\pm u_{1},\pm u_{2}). This analytic function actually comes from the equation of critical point: θϕ((e1,41(u1,u2,0));θ)=0\nabla_{\theta}\phi((e_{1},4^{-1}(u_{1},u_{2},0));\theta)=0. The interested readers can find the background as well as the geometrical meaning of θ\theta in [33, § 2 and § 11].

The starting point of this article is the following diffeomorphisms (see Figure 1 for the sketch map):

\begin{overpic}[width=497.92322pt,height=213.39566pt]{DiffN32.jpg} \put(32.0,6.0){$\Omega_{-,4}$} \put(12.0,22.0){$\Omega_{+,1}$} \put(68.0,25.0){$\mathbb{R}^{2}_{>,+}$} \put(80.0,13.0){$\mathbb{R}^{2}_{<,+}$} \put(47.0,19.5){$\Lambda$} \put(46.0,18.0){$\to$} \put(41.0,13.0){$v_{1}$} \put(1.5,37.0){$v_{2}$} \put(99.0,6.0){$u_{1}$} \put(57.0,35.0){$u_{2}$} \put(90.0,27.0){$\searrow$} \put(80.0,30.0){$u_{2}=\frac{2}{\sqrt{\pi}}\sqrt{u_{1}}$} \put(-0.5,6.0){$\mathrm{K}_{3}=\frac{\psi^{\prime}(r)}{r}v_{2}^{2}+2\psi(r)=0\to$} \put(39.5,14.0){$\downarrow$} \put(39.5,16.0){$\vartheta_{1}$} \end{overpic}
Figure 1: Sketch map of the analytic-diffeomorphism Λ\Lambda
Theorem 2.1.

The map Λ\Lambda is an analytic-diffeomorphism from:
(1) Ω+,1\Omega_{+,1} onto >,+2\mathbb{R}^{2}_{>,+}. Moreover, let

u=(u~,0) with u~>,+2,gu=(e1,41u)N3,2,and θ=(Λ1(u~),0).u=(\widetilde{u},0)\mbox{\ with \ }\widetilde{u}\in\mathbb{R}^{2}_{>,+},\quad g_{u}=(e_{1},4^{-1}u)\in N_{3,2},\ \mbox{and }\theta=(\Lambda^{-1}(\widetilde{u}),0).

Then the Hessian matrix of ϕ(gu;)-\phi(g_{u};\cdot) at θ\theta is positive definite.
(2) Ω,4\Omega_{-,4} onto <,+2\mathbb{R}^{2}_{<,+}. Furthermore, using the above notation with u~\widetilde{u} belonging to <,+2\mathbb{R}^{2}_{<,+} instead of >,+2\mathbb{R}^{2}_{>,+}, the Hessian matrix of ϕ(gu;)\phi(g_{u};\cdot) at θ\theta is nonsingular, and has exactly two positive eigenvalues.

Its proof is based on the operator convexity and Hadamard’s Theorem. See Section 3 below for the details. We point out that the two diffeomorphisms first appeared in [33]. In fact, the first claim is a direct consequence of [33, Propositions 11.1 and 11.2], while the second one is exactly [33, Theorem 11.2].

In light of Theorem 2.1, we will make the following assumption repeatedly, unless otherwise specified:

Assumption (A): {θ:=(θ1,θ2,0)with(θ1,θ2)Ω+,1Ω,4,ϵ:=ϑ1|θ|;u:=(u1,u2,0):=(Λ(θ1,θ2),0)(so u1,u2>0 and πu224u1);u:=41u,gu:=(e1,u)N3,2;g=(x,t):=δ|x|(gu)=(|x|e1,|x|2u)=(|x|e1,t1,t2,0)with |x|>0;𝐬¯=(𝐬¯1,𝐬¯2):=(1,0)2,𝐰¯:=θ2ψ(|θ|)>0.\displaystyle{\text{\bf Assumption (A): }}\begin{cases}\theta:=(\theta_{1},\theta_{2},0)\ \mbox{with}\ (\theta_{1},\theta_{2})\in\Omega_{+,1}\cup\Omega_{-,4},\,\epsilon:=\vartheta_{1}-|\theta|;\\[2.84526pt] u:=(u_{1},u_{2},0):=(\Lambda(\theta_{1},\theta_{2}),0)\\ \qquad\mbox{(so $u_{1},u_{2}>0$ and $\pi\,u_{2}^{2}\neq 4\,u_{1}$)};\\[2.84526pt] u_{*}:=4^{-1}u,\ g_{u}:=(e_{1},u_{*})\in N_{3,2};\\[2.84526pt] g=(x,t):=\delta_{|x|}(g_{u})=(|x|e_{1},\ |x|^{2}u_{*})=(|x|e_{1},t_{1},t_{2},0)\\ \qquad\mbox{with $|x|>0$};\\[2.84526pt] \mathbf{\overline{s}}=(\mathbf{\overline{s}}_{1},\mathbf{\overline{s}}_{2}):=(-1,0)\in{\mathbb{R}}^{2},\,\mathbf{\overline{w}}:=\theta_{2}\,\psi(|\theta|)>0.\end{cases} (2.18)

With this assumption, (2.3) and the reference function (2.6) become

p(g)=3|λ|sinh|λ|exp{|x|24[1+|λ|coth|λ|1|λ|2(λ22+λ32)iuλ]}𝑑λ,\displaystyle p(g)=\int_{\mathbb{R}^{3}}\frac{|\lambda|}{\sinh{|\lambda|}}\,\exp\left\{-\frac{|x|^{2}}{4}\left[1+\frac{|\lambda|\coth{|\lambda|}-1}{|\lambda|^{2}}(\lambda_{2}^{2}+\lambda_{3}^{2})-i\,u\cdot\lambda\right]\right\}\,d\lambda, (2.19)

and

ϕ(g;τ)=|x|2[1ψ(|τ|)(τ22+τ32)+uτ],τ=(τ1,τ2,τ3)3,\displaystyle\phi(g;\tau)=|x|^{2}\,\left[1-\psi(|\tau|)\,(\tau_{2}^{2}+\tau_{3}^{2})+u\cdot\tau\right],\quad\tau=(\tau_{1},\tau_{2},\tau_{3})\in\mathbb{R}^{3}, (2.20)

respectively.

At this moment we will not formulate our main result, namely the uniform heat kernel asymptotics at infinity in the sense of (1.8), since there are too many notations to be introduced which may lead to confusion. Instead, we shall postpone the precise statement until Section 11.1 and present some byproducts in this section. The first one is the following:

2.3 Explicit expression for the squared Carnot–Carathéodory distance

Using the scaling property of the heat kernel (cf. (2.2)) and (1.7), (2.8) implies its counterpart for dd:

d(Ox,Ot)2=d(x,t)2,OO3,\displaystyle d(Ox,Ot)^{2}=d(x,t)^{2},\quad\forall\,O\in\mathrm{O}_{3}, d(x,t)2=d(x,t)2.\displaystyle\quad d(x,t)^{2}=d(x,-t)^{2}. (2.21)

Combining with the scaling property of dd, it suffices to determine d(g)d(g) for special gg, such as (0,e1)(0,e_{1}) and (e1,(t1,t2,0))(e_{1},(t_{1},t_{2},0)) with t1,t20t_{1},t_{2}\geq 0.

In the sequel, we introduce some positive functions on (0,+)(0,\ +\infty),

h(r):=r2+rsinrcosr2sin2r(=ψ(r)r3sin2r,r{kπ;k}),\displaystyle h(r):=r^{2}+r\sin{r}\cos{r}-2\sin^{2}{r}\ (=\psi^{\prime}(r)\,r^{3}\sin^{2}r,\,r\notin\{k\pi;\,k\in\mathbb{N}^{*}\}), (2.22)
φ1(r):=r2sin2rrsinrcosr,φ2(r):=r(r2sin2r)h(r),φ3:=φ1φ2.\displaystyle\varphi_{1}(r):=\frac{r^{2}-\sin^{2}{r}}{r-\sin{r}\cos{r}},\quad\varphi_{2}(r):=\frac{r\,(r^{2}-\sin^{2}{r})}{h(r)},\quad\varphi_{3}:=\sqrt{\varphi_{1}\varphi_{2}}. (2.23)

Indeed, the positivity of hh follows from Corollary 3.6 below, and the others are clear.

Theorem 2.2.

Let θ,u,gu,g\theta,u,g_{u},g be given as in the Assumption (A) (cf. (2.18)). Then

d(gu)2=θ12|θ|2+θ22sin2|θ|=θ22ψ(|θ|)+uθ+1=φ1(|θ|)(u1θ1|θ|+u2θ2|θ|)+1=φ2(|θ|)u1|θ|θ1+1=φ3(|θ|)u1(u1+u2θ2θ1)+1.\displaystyle\begin{aligned} d(g_{u})^{2}&=\frac{\theta_{1}^{2}}{|\theta|^{2}}+\frac{\theta_{2}^{2}}{\sin^{2}{|\theta|}}=-\theta_{2}^{2}\,\psi(|\theta|)+u\cdot\theta+1=\,\varphi_{1}(|\theta|)\left(u_{1}\,\frac{\theta_{1}}{|\theta|}+u_{2}\,\frac{\theta_{2}}{|\theta|}\right)+1\\ &=\varphi_{2}(|\theta|)\,u_{1}\,\frac{|\theta|}{\theta_{1}}+1=\varphi_{3}(|\theta|)\sqrt{u_{1}\,\left(u_{1}+u_{2}\,\frac{\theta_{2}}{\theta_{1}}\right)}+1.\end{aligned} (2.24)

In particular, we have

d(gu)2=ϕ(gu;θ),d(g)2=ϕ(g;θ).\displaystyle d(g_{u})^{2}=\phi(g_{u};\theta),\qquad d(g)^{2}=\phi(g;\theta). (2.25)

For the other cases, an argument of limit implies that:

Corollary 2.3.

It holds that:

  1. (i)

    For g=(0,e1)g_{*}=(0,e_{1}), we have d(g)2=4πd(g_{*})^{2}=4\pi.

  2. (ii)

    If α>0\alpha>0 and gα:=(e1,41(α2π,2πα,0))g_{\alpha}:=(e_{1},4^{-1}(\frac{\alpha^{2}}{\pi},\frac{2}{\pi}\alpha,0)) such that (α2π,2πα)>,+2(\frac{\alpha^{2}}{\pi},\frac{2}{\pi}\alpha)\in\partial\mathbb{R}^{2}_{>,+}. Then d(gα)2=1+α2d(g_{\alpha})^{2}=1+\alpha^{2}.

  3. (iii)

    Let β>0\beta>0 and g(β):=(e1,41(β,0,0))g(\beta):=(e_{1},4^{-1}(\beta,0,0)). Then

    d(g(β))2=φ3(r)β+1,\displaystyle d(g(\beta))^{2}=\varphi_{3}(r)\,\beta+1,

    where rr is the unique solution of the following equation

    2ψ(r)r2+2rψ(r)ψ(r)=β,π<r<ϑ1.\displaystyle-2\,\psi(r)\,\sqrt{r^{2}+2\,r\,\frac{\psi(r)}{\psi^{\prime}(r)}}=\beta,\quad\pi<r<\vartheta_{1}. (2.26)
  4. (iv)

    Set μ(ρ):=2ρsin(2ρ)2sin2ρ\mu(\rho):=\frac{2\rho-\sin(2\rho)}{2\sin^{2}\rho}, π<ρ<π-\pi<\rho<\pi. For g(γ):=(e1,41(0,γ,0))g(\gamma)^{*}:=(e_{1},4^{-1}(0,\gamma,0)) with γ>0\gamma>0, we have

    d(g(γ))2=(rsinr)2,where r is the unique solution of μ(r)=γ.d(g(\gamma)^{*})^{2}=\left(\frac{r}{\sin{r}}\right)^{2},\quad\mbox{where $r$ is the unique solution of $\mu(r)=\gamma$.}

We give two remarks on these results as follows:

Remark 2.4.

(1) For (i), the distance between oo and g=(0,t)g=(0,t) has been computed in [15] in the setting of free step-two Carnot groups with k3k\geq 3 generators. (iii) have been obtained in [41]. (ii) (resp. (iv)) can be found in (resp. deduced from) [33, Corollary 11.1] (resp. [33, Theorem 11.1]). As for our main results, i.e. (2.25), the case where gu>,+2g_{u}\in\mathbb{R}^{2}_{>,+} is given by [33, Theorem 11.1]. For the opposite case gu<,+2g_{u}\in\mathbb{R}^{2}_{<,+}, there exist two different proofs (as mentioned in Introduction), one is based on [33, Corollary 2.4] (cf. [33, §11.2]), the other is based on [33, Theorem 2.4] (cf. [37, §7.4]).

(2) The LHS of (2.26) is exactly the function 4P(r)\frac{4}{P(r)} with PP defined as in [41, (3.3)]. Then from [41, Lemma 3.5] it follows that PP is a strictly increasing diffeomorphism between (π,ϑ1)(\pi,\ \vartheta_{1}) and (0,+)(0,\ +\infty), which justifies the uniqueness of the solution rr in (π,ϑ1)(\pi,\ \vartheta_{1}).

(3) For (iv), the function μ\mu is actually a differomorphism from (π,π)(-\pi,\pi) to \mathbb{R}. See [21, Lemme 3, p. 112] for more details. In this case, the expression of the square distance is the same with the point (e1,41γ)(e_{1},4^{-1}\gamma) on the Heisenberg group of real dimension 33.

Remark 2.5.

In our setting N3,2N_{3,2}, Cuto={(x,t);x and t are linearly dependent}{\rm Cut}_{o}=\{(x,t);\,\mbox{$x$ and $t$ are linearly dependent}\}. See for example [38, § 7.5] for an elementary explanation. We will not use this fact in the proof, but it helps us understand better some of our difficulties encountered in the uniform heat kernel asymptotics.

2.4 Precise estimates for the heat kernel

Another main byproduct of our uniform heat kernel asymptotics at infinity is the following sharp upper and lower bounds:

Theorem 2.6.

Under the Assumption (A) (cf. (2.18)), we have

p(x,t)(1+d(g))21+ϵd(g)1+ϵd(g)+ϵt212|x|12(d(g)2|x|2)14ed(g)24.\displaystyle p(x,t)\sim(1+d(g))^{-2}\frac{1+\epsilon\,d(g)}{1+\epsilon\,d(g)+\epsilon\,t_{2}^{\frac{1}{2}}\,|x|^{\frac{1}{2}}\,(d(g)^{2}-|x|^{2})^{\frac{1}{4}}}\,e^{-\frac{d(g)^{2}}{4}}. (2.27)

Here we use the formula (2.27) in order to match the precise estimates for the heat kernel in our setting with the known results, cf. eg. [32, 36, 35]. However, a much more explicit expression for d(g)1d(g)\gg 1 can be found in Corollary 11.4 below.

Remark 2.7.

(1) From the above explicit expression of d2d^{2} on N3,2N_{3,2}, it is easy to get that d(g)2|x|2d(g)^{2}\geq|x|^{2} for any g=(x,t)N3,2g=(x,t)\in N_{3,2}, with the equality holding if and only if g=(x,0)g=(x,0). Indeed, the result is still valid on any step-two Carnot group (cf. [33, §2]).

(2) Let 𝔾\mathbb{G} be a Métivier group with dim𝔤2=m{\rm dim}\,\mathfrak{g}_{2}=m, the standard Laplace’s method implies that (cf. also [33, §4.2]) its heat kernel with time 11 at (x,0)(x,0) is |x|me|x|2/4\sim|x|^{-m}e^{-|x|^{2}/4} as |x|+|x|\to+\infty. Remark that our result (2.27) with g=(|x|e1,0)g=(|x|\,e_{1},0) and |x|+|x|\to+\infty is totally different from the classical one. Indeed in our case, the Laplace’s method is no longer applicable, since it follows from (2.5) that the set of minimal points of ϕ~((|x|e1,0);)\widetilde{\phi}((|x|\,e_{1},0);\cdot) equals {(λ1,0,0);λ1}\{(\lambda_{1},0,0);\,\lambda_{1}\in\mathbb{R}\}. From a geometric point of view, it says that there exists on N3,2N_{3,2} a non-trivial abnormal set (or geodesics), which is exactly {(x,0);x3}{o}\{(x,0);x\in\mathbb{R}^{3}\}\setminus\{o\}.

(3) Compare this result with the one in the setting of Heisenberg groups or even generalized H-type groups (cf. [30, 31, 36, 35]), (2.27) is much more complicated since the group law is more complex. Naturally, we believe that there will be some more complicated terms for precise bounds of the heat kernel on concrete step-two groups.

2.5 Idea of the proof

Let’s start by explaining how to prove d(gu)2=ϕ(gu;θ)d(g_{u})^{2}=\phi(g_{u};\theta), namely the first equality in (2.25). From (2.19) and Varandhan’s formula, it suffices to investigate the asymptotic behavior of

3|λ|sinh|λ|exp{14h[1+|λ|coth|λ|1|λ|2(λ22+λ32)iuλ]}𝑑λ,h0+.\displaystyle\int_{\mathbb{R}^{3}}\frac{|\lambda|}{\sinh{|\lambda|}}\,\exp\left\{-\frac{1}{4h}\left[1+\frac{|\lambda|\coth{|\lambda|}-1}{|\lambda|^{2}}(\lambda_{2}^{2}+\lambda_{3}^{2})-i\,u\cdot\lambda\right]\right\}\,d\lambda,\ h\rightarrow 0^{+}. (2.28)

It is a typical oscillatory integral. The usual processing method is to use the method of stationary phase, and it allows us to guess the correct answer (cf. Corollary 11.7)

|θ|sin|θ|eϕ(gu;θ)4h(8πh)32j=13rj12(1+o(1)),\displaystyle\frac{|\theta|}{\sin{|\theta|}}e^{-\frac{\phi(g_{u};\theta)}{4h}}\,(8\pi h)^{\frac{3}{2}}\,\prod_{j=1}^{3}\mathrm{r}_{j}^{-\frac{1}{2}}\,(1+o(1)), (2.29)

where r1,r2,r3\mathrm{r}_{1},\mathrm{r}_{2},\mathrm{r}_{3} are the eigenvalues of Hessθϕ(gu;θ)-{\mathrm{Hess}}_{\theta}\,\phi(g_{u};\theta), and we adopt the convention (r)1/2=ir1/2(-r)^{-1/2}=-i\,r^{-1/2} for r>0r>0. Indeed, it is a special case of [33, §4] provided gu>,+2g_{u}\in\mathbb{R}^{2}_{>,+}. More precisely, it suffices to deform the contour from 3\mathbb{R}^{3} to 3+iθ\mathbb{R}^{3}+i\theta, then apply the method of stationary phase at the nondegenerate critical point iθi\theta, since all assumptions of the method can be verified to be satisfied. However, for the opposite case gu<,+2g_{u}\in\mathbb{R}^{2}_{<,+}, the situation becomes very tricky, and the two main obstacles we encountered were as follows. The first one is how to choose a suitable integration path passing through the point iθi\theta so that all assumptions hold for the method of stationary phase. The second one is how to treat the “residue problem”, since |θ|>π|\theta|>\pi in such case and the integrand in (2.28) has singularities as {λ+iτ;|λ|2+2iλτ|τ|2=k2π2,k}\{\lambda+i\tau;\,|\lambda|^{2}+2i\lambda\cdot\tau-|\tau|^{2}=-k^{2}\pi^{2},k\in\mathbb{N}^{*}\}. It is actually the main difference between GM-groups and non-GM groups.

This motivates us to find a more appropriate expression. For that, we drop hh, and set in the sequel

𝐏(X,T):=3𝒱(λ)e14Γ~((X,T);λ)𝑑λ,X2,T3,\displaystyle\mathbf{P}(X,T):=\int_{\mathbb{R}^{3}}\,\mathcal{V}(\lambda)\,e^{-\frac{1}{4}\widetilde{\Gamma}((X,T);\lambda)}\,d\lambda,\qquad X\in\mathbb{R}^{2},\,T\in\mathbb{R}^{3},

with

𝒱(λ):=|λ|3|λ|cosh|λ|sinh|λ|,Γ~((X,T);λ):=|λ|2|λ|coth|λ|1|X|24iTλ.\displaystyle\mathcal{V}(\lambda):=\frac{|\lambda|^{3}}{|\lambda|\cosh{|\lambda|}-\sinh{|\lambda|}},\quad\widetilde{\Gamma}((X,T);\lambda):=\frac{|\lambda|^{2}}{|\lambda|\coth{|\lambda|}-1}|X|^{2}-4i\,T\cdot\lambda.

Remark that Γ~\widetilde{\Gamma} admits the following scaling property: Γ~((hX,h2T);λ)=h2Γ~((X,T);λ)\widetilde{\Gamma}((hX,h^{2}T);\lambda)=h^{2}\widetilde{\Gamma}((X,T);\lambda) for all h>0h>0 and (X,T)(X,T). By means of properties of Bessel functions, we can establish the following much more useful expression (compared with [37, (15)] which is valid for any step-two Carnot groups):

Proposition 2.8.

Let θ,u,g\theta,u,g and 𝐰¯\mathbf{\overline{w}} be as in Assumption (A) (cf. (2.18)). Then we have

p(g)=14π|x|2𝐰¯2e|x|242𝒫(s)𝑑s,p(g)=\frac{1}{4\pi}\,|x|^{2}\,\mathbf{\overline{w}}^{2}\,e^{-\frac{|x|^{2}}{4}}\int_{\mathbb{R}^{2}}\mathcal{P}(s)\,ds, (2.30)

where

𝒫(s)=𝒫(x,u;s):=𝐏(s|x|𝐰¯,14|x|2(u+2𝐰¯s1e2+2𝐰¯s2e3)).\displaystyle\mathcal{P}(s)=\mathcal{P}(x,u;s):=\mathbf{P}\left(s\,|x|\,\mathbf{\overline{w}},\frac{1}{4}|x|^{2}(u+2\mathbf{\overline{w}}\,s_{1}\,e_{2}+2\mathbf{\overline{w}}\,s_{2}\,e_{3})\right). (2.31)

Its proof is provided in Subsection 4.1. We emphasize that we have used here the new coordinates |x|𝐰¯s|x|\,\mathbf{\overline{w}}\,s in view of the scaling property of Γ~\widetilde{\Gamma}. The reason for our choice of 𝐰¯\mathbf{\overline{w}} herein can be found in Theorem 2.10 below. The key point is that 𝐏\mathbf{P} can be considered formally as the heat kernel at time 11 on an imagined but non-existent Heisenberg-type group (2,3)\mathbb{H}(2,3), with dimension 33 in the center and dimension 22 on the first slice in the stratification. To see this more explicitly, as in [33, Theorem 2.2 and Corollary 2.3] (cf. also [35, Theorem 4.2]), we introduce the squared “intrinsic distance” associated to 𝐏\mathbf{P}:

𝐃(X,T)2:=supτ3,|τ|<ϑ1Γ~((X,T);iτ),(X,T)2×3.\displaystyle\mathbf{D}(X,T)^{2}:=\sup_{\tau\in\mathbb{R}^{3},\ |\tau|<\vartheta_{1}}\widetilde{\Gamma}((X,T);i\tau),\qquad(X,T)\in\mathbb{R}^{2}\times\mathbb{R}^{3}.

See Proposition 4.2 below for more properties of 𝐃(X,T)2\mathbf{D}(X,T)^{2}, especially the scaling property.

Let I0I_{0} denote the modified Bessel function of order 0 (cf. [22, § 8.431.3]), i.e.,

I0(r):=1π0πercosγ𝑑γ=12πππercosγ𝑑γ,\displaystyle I_{0}(r):=\frac{1}{\pi}\int_{0}^{\pi}e^{r\cos\gamma}\,d\gamma=\frac{1}{2\pi}\int_{-\pi}^{\pi}e^{r\cos\gamma}\,d\gamma, (2.32)

and set in the sequel the even functions

Υ(r):=r21rcotr,𝔮(r):=r2Υ(r)sinrΥ(r)Υ′′(r),r(ϑ1,ϑ1).\displaystyle\Upsilon(r):=\frac{r^{2}}{1-r\,\cot{r}},\qquad\mathfrak{q}(r):=\frac{r^{2}\,\Upsilon(r)}{-\sin r\,\Upsilon^{\prime}(r)\,\sqrt{-\Upsilon^{\prime\prime}(r)}},\qquad r\in(-\vartheta_{1},\vartheta_{1}). (2.33)

The following proposition says that 𝐏\mathbf{P} satisfies the uniform bounds [31, (1.5)] as well as the uniforms asymptotics behaviors [31, Théorèmes 1.4-1.5] with n=1n=1 and m=3m=3 (of course, with (x,t;|x|,d(x,t))(x,t;|x|,d(x,t)) therein replaced by our (X,T;|X|,𝐃(X,T))(X,T;|X|,\mathbf{D}(X,T)), and some slight, natural modifications):

Proposition 2.9.

𝐏\mathbf{P} admits the following properties:

  1. (i)

    The positivity, namely 𝐏(X,T)>0\mathbf{P}(X,T)>0.

  2. (ii)

    P(1;3;,)P(1;3;\cdot,\cdot)-type uniform bounds in the sense of [31, Théorème 1.1], namely,

    𝐏(X,T)1(1+𝐃(X,T))2(1+|X|𝐃(X,T))12e𝐃(X,T)24,(X,T)2×3.\mathbf{P}(X,T)\sim\frac{1}{(1+\mathbf{D}(X,T))^{2}\,(1+|X|\,\mathbf{D}(X,T))^{\frac{1}{2}}}\,e^{-\frac{\mathbf{D}(X,T)^{2}}{4}},\quad\forall(X,T)\in\mathbb{R}^{2}\times\mathbb{R}^{3}. (2.34)
  3. (iii)

    It holds uniformly for any X0X\neq 0 and all 𝐃(X,T)\mathbf{D}(X,T) large enough that

    𝐏(X,T)=2ϑ116π2𝔮(|τ|)e𝐃(X,T)24eϑ1|X|22(ϑ1|τ|)I0(ϑ1|X|22(ϑ1|τ|))(ϑ1|τ|)12|X|2(1+o(1)),\mathbf{P}(X,T)=\sqrt{2\vartheta_{1}}16\pi^{2}\,\mathfrak{q}(|\tau^{*}|)\,e^{-\frac{\mathbf{D}(X,T)^{2}}{4}}e^{-\frac{\vartheta_{1}|X|^{2}}{2(\vartheta_{1}-|\tau^{*}|)}}\,I_{0}\left(\frac{\vartheta_{1}|X|^{2}}{2(\vartheta_{1}-|\tau^{*}|)}\right)\,\frac{(\vartheta_{1}-|\tau^{*}|)^{-\frac{1}{2}}}{|X|^{2}}\,(1+o(1)),

    where τ=τ(X,T)\tau^{*}=\tau^{*}(X,T) is the unique critical point of Γ~((X,T),i())\widetilde{\Gamma}((X,T),i\,(\cdot)) in B3(0,ϑ1)B_{{\mathbb{R}}^{3}}(0,\vartheta_{1}).

To show Proposition 2.9, it is enough to adapt the method in [37] for (i), and the method in [35] for (ii) and (iii). Hence its proof is postponed to Appendix A.

Return to p(δh1/2(gu))p(\delta_{h^{-1/2}}(g_{u})) (h0+h\to 0^{+}), namely (2.28). Using the scaling property of 𝐃\mathbf{D} (cf. Proposition 4.2 (iii) below), Propositions 2.8 and 2.9 say that it is now a typical Laplace-type integral with positive integrand. Then inspired by the main idea of the standard Laplace’s method, it is natural to study the minimum of 𝐃(𝐰¯s,14(u+2𝐰¯s1e2+2𝐰¯s2e3))2\mathbf{D}\!\left(\mathbf{\overline{w}}\,s,\frac{1}{4}(u+2\mathbf{\overline{w}}\,s_{1}e_{2}+2\mathbf{\overline{w}}\,s_{2}e_{3})\right)^{2}. The following theorem will play an important role.

Theorem 2.10.

Let θ,u\theta,u and 𝐰¯\mathbf{\overline{w}} be as in Assumption (A) (cf. (2.18)). Set

𝒟(u;s):=𝐃(𝐰¯s,14(u+2𝐰¯s1e2+2𝐰¯s2e3))2,s2.\displaystyle\mathcal{D}(u;s):=\mathbf{D}\!\left(\mathbf{\overline{w}}s,\frac{1}{4}(u+2\mathbf{\overline{w}}s_{1}e_{2}+2\mathbf{\overline{w}}\,s_{2}e_{3})\right)^{2},\quad s\in\mathbb{R}^{2}. (2.35)

Then 𝐬¯=(1,0)\mathbf{\overline{s}}=(-1,0) is the unique minimum point of 𝒟(u;)\mathcal{D}(u;\cdot). Moreover, with φ1,φ2\varphi_{1},\varphi_{2} and φ3\varphi_{3} defined by (2.23), we have:

𝒟(u;𝐬¯)\displaystyle\mathcal{D}(u;\mathbf{\overline{s}}) =θ22ψ(|θ|)+uθ=θ22|θ|2[(|θ|sin|θ|)21]=φ1(|θ|)|θ|uθ\displaystyle=-\theta_{2}^{2}\,\psi(|\theta|)+u\cdot\theta=\frac{\theta_{2}^{2}}{|\theta|^{2}}\left[\left(\frac{|\theta|}{\sin{|\theta|}}\right)^{2}-1\right]=\frac{\varphi_{1}(|\theta|)}{|\theta|}u\cdot\theta (2.36)
=φ2(|θ|)u1|θ|θ1=φ3(|θ|)u1(u1+u2θ2θ1).\displaystyle=\varphi_{2}(|\theta|)\,u_{1}\,\frac{|\theta|}{\theta_{1}}=\varphi_{3}(|\theta|)\sqrt{u_{1}\!\left(u_{1}+u_{2}\,\frac{\theta_{2}}{\theta_{1}}\right)}.

Its proof will be postponed to Section 4. Combining this with Proposition 2.9 (ii), it is easy to show that d(gu)2=1+𝒟(u;𝐬¯)d(g_{u})^{2}=1+\mathcal{D}(u;\mathbf{\overline{s}}), namely Theorem 2.2.

However, the Laplace’s method is far from sufficient for the main goal of this article. Roughly speaking, the phase function in our Laplace-type integral (cf. Propositions 2.8 and 2.9), namely |x|2𝒟(u;s)/4|x|^{2}\mathcal{D}(u;s)/4, is not C1C^{1} when s=0s=0 (see Remark 4.3 below), which makes things difficult when 𝐰¯\mathbf{\overline{w}} small. In fact, as far as the authors know, there is no suitable method to deal with this situation. Furthermore, let 𝔪\mathfrak{m} denote its minimum (i.e. 𝔪=|x|2𝒟(u;𝐬¯)/4\mathfrak{m}=|x|^{2}\mathcal{D}(u;\mathbf{\overline{s}})/4), 𝔏1,𝔏2>0\mathfrak{L}_{1},\mathfrak{L}_{2}>0 two eigenvalues of Hess𝐬¯|x|24𝒟(u;𝐬¯)\mathrm{Hess}_{\mathbf{\overline{s}}}\,\frac{|x|^{2}}{4}\mathcal{D}(u;\mathbf{\overline{s}}) with 𝔏1𝔏2\mathfrak{L}_{1}\gtrsim\mathfrak{L}_{2} (cf. Subsection 7.3 below). It is worthwhile to point out that for d(g)+d(g)\to+\infty, the Laplace’s method is sufficient only for the case where both 𝔪,𝔏1\mathfrak{m},\mathfrak{L}_{1} and 𝔏2\mathfrak{L}_{2} are large enough.

For instance, in the case 𝔪+\mathfrak{m}\to+\infty, we first use Proposition 2.9 (iii) to simplify the integrand in (2.30). The most delicate situation is that |θ|1|\theta|\geq 1 with 𝔏11\mathfrak{L}_{1}\gg 1, and notice that 𝔏2\mathfrak{L}_{2} can be bounded. To obtain the uniform asymptotic behavior of p(g)p(g) in such case, we will choose suitable coordinates (in fact the modified polar coordinates up to a scaling), by which the phase function can be divided into the angular part and the radical part in the main region. We will use Laplace’s method to deal with the radical part and the method in [30] to cope with the angular part. In the opposite case where 𝔪\mathfrak{m} is bounded, we’ll make use of a completely different approach.

Furthermore, to establish the sharp estimates for the derivatives of the heat kernel, unlike in the setting of GM-Métivier groups, new techniques need to be introduced.

The remainder of this article is organized as follows. Theorem 2.1 is proved in Section 3. The proofs of Proposition 2.8 and Theorem 2.10 are given in Section 4. We obtain the explicit expression of the squared distance as a consequence of Laplace’s method in Section 5. The uniform asymptotic behaviour at infinity for the heat kernel at time 1 are divided into cases. We establish the first asymptotic in Section 6 for the case where |θ|3|\theta|\leq 3 and θ2|x|+\theta_{2}|x|\to+\infty. After some preparations in Section 7, we will attack the most difficult and complicated situation in Section 8. Section 9 is devoted to the case where 𝔪+\mathfrak{m}\to+\infty with 𝔏11\mathfrak{L}_{1}\lesssim 1, while Section 10 is for the remaining case. The summaries of our main results as well as the proof of Theorem 2.6 are collected in Section 11. Finally in Section 12 we establish the sharp bounds for derivatives of the heat kernel. In Appendix A we give the proof for our Proposition 2.9.

3 Proof of Theorem 2.1

In this section, we establish the two key analytic-diffeomorphisms provided in Theorem 2.1. Recall that this part (up to some modification) is extracted from [33].

As we have seen in Theorems 2.1-2.2 and Corollary 2.3, the function ψ\psi plays an important role. Let us begin with

3.1 Fine properties of the function ψ\psi and its derivatives

Notice that:

ψ(r):=1rcotrr2=j=1+(jπ)2(11rjπ+11+rjπ)=2j=1+1(jπ)2r2.\displaystyle\psi(r):=\frac{1-r\cot{r}}{r^{2}}=\sum_{j=1}^{+\infty}(j\pi)^{-2}\left(\frac{1}{1-\frac{r}{j\,\pi}}+\frac{1}{1+\frac{r}{j\,\pi}}\right)=2\sum_{j=1}^{+\infty}\frac{1}{(j\,\pi)^{2}-r^{2}}. (3.1)

In the second and last “==” we have used the series expansion of the function rcotrr\cot r; see for example [22, §1.421.3, p. 44] with slight modification. That is,

rcotr=1+r2j=1+(jπ)2(11rjπ+11+rjπ)=1+1π1πr21ρr𝑑ν(ρ),\displaystyle-r\cot{r}=-1+r^{2}\sum_{j=1}^{+\infty}(j\pi)^{-2}\left(\frac{1}{1-\frac{r}{j\,\pi}}+\frac{1}{1+\frac{r}{j\,\pi}}\right)=-1+\int_{-\frac{1}{\pi}}^{\frac{1}{\pi}}\frac{r^{2}}{1-\rho r}\,d\nu(\rho), (3.2)

where the positive finite measure ν\nu on [1π,1π][-\frac{1}{\pi},\,\frac{1}{\pi}] is given by

ν=j=1+(1jπ)2(δ1jπ+δ1jπ)with δa the usual Dirac measure at point a.\displaystyle\nu=\sum_{j=1}^{+\infty}\left(\frac{1}{j\,\pi}\right)^{2}\left(\delta_{\frac{1}{j\,\pi}}+\delta_{-\frac{1}{j\,\pi}}\right)\quad\mbox{with $\delta_{a}$ the usual Dirac measure at point $a$.}

Now we collect some properties for ψ\psi on [0,π)[0,\pi).

Lemma 3.1.

We have

ψ′′(r)>ψ(r)rlimr0ψ(r)r>0,0<r<π,\displaystyle\psi^{\prime\prime}(r)>\frac{\psi^{\prime}(r)}{r}\geq\lim_{r\to 0}\frac{\psi^{\prime}(r)}{r}>0,\quad 0<r<\pi, (3.3)
ψ(r)ψ′′(r)>2ψ(r)2,0r<π.\displaystyle\psi(r)\,\psi^{\prime\prime}(r)>2\psi^{\prime}(r)^{2},\quad 0\leq r<\pi. (3.4)
Proof.

A simple and direct proof for (3.3) is to use the fact that (see the second equality in (3.1), or [22, §1.411.7, p. 42] with slight modification of notation)

ψ(r)=j=1+bjr2j2,with bj>0,j.\displaystyle\psi(r)=\sum_{j=1}^{+\infty}b_{j}r^{2j-2},\quad\mbox{with }\,b_{j}>0,\,\forall\,j\in\mathbb{N}^{*}.

Aiming at (3.4), it follows from (3.1) that:

ψ(r)=1π1πdν(ρ)1rρ,ψ(r)=1π1πρ(1rρ)2𝑑ν(ρ),ψ′′(r)=21π1πρ2(1rρ)3𝑑ν(ρ),\displaystyle\psi(r)=\int_{-\frac{1}{\pi}}^{\frac{1}{\pi}}\frac{d\nu(\rho)}{1-r\rho},\quad\psi^{\prime}(r)=\int_{-\frac{1}{\pi}}^{\frac{1}{\pi}}\frac{\rho}{(1-r\rho)^{2}}\,d\nu(\rho),\quad\psi^{\prime\prime}(r)=2\int_{-\frac{1}{\pi}}^{\frac{1}{\pi}}\frac{\rho^{2}}{(1-r\rho)^{3}}\,d\nu(\rho),

then Hölder’s inequality shows that

2ψ(r)2<2(1π1π|ρ|(1rρ)2𝑑ν(ρ))2<ψ(r)ψ′′(r).\displaystyle 2\psi^{\prime}(r)^{2}<2\left(\int_{-\frac{1}{\pi}}^{\frac{1}{\pi}}\frac{|\rho|}{(1-r\rho)^{2}}\,d\nu(\rho)\right)^{2}<\psi(r)\,\psi^{\prime\prime}(r).

This concludes the proof of (3.4) and hence the lemma. ∎

The following two lemmas will play an important part in the proof for the first analytic-diffeomorphism of Theorem 2.1. Lemma 3.3 is also vital to the heat kernel asymptotics in Section 6.

Lemma 3.2.

It holds that

ψ(r)>ψ(r)r, 0r<π.\displaystyle\psi(r)>\sqrt{\frac{\psi^{\prime}(r)}{r}},\qquad\forall\,0\,\leq r<\pi. (3.5)
Proof.

By the second equality in (3.1), we get

ψ(r)=2j=1+1(jπ)2r2,ψ(r)r=4j=1+[(jπ)2r2]2,\displaystyle\psi(r)=2\sum_{j=1}^{+\infty}\frac{1}{(j\,\pi)^{2}-r^{2}},\qquad\frac{\psi^{\prime}(r)}{r}=4\sum_{j=1}^{+\infty}\left[(j\,\pi)^{2}-r^{2}\right]^{-2}, (3.6)

which implies immediately the required estimate. ∎

Lemma 3.3.

Let 0<ζ010<\zeta_{0}\leq 1. Then

Hessτ(τ22ψ(|τ|))ζ0(τ221)0,for allτ=(τ1,τ2)2with |τ|πζ0.\displaystyle\mathrm{Hess}_{\tau}\left(\tau_{2}^{2}\,\psi(|\tau|)\right)\sim_{\zeta_{0}}\begin{pmatrix}\tau_{2}^{2}&\quad\\ \quad&1\end{pmatrix}\geq 0,\quad\mbox{for all}\ \tau=(\tau_{1},\tau_{2})\in\mathbb{R}^{2}\ \mbox{with $|\tau|\leq\pi-\zeta_{0}$. }
Proof.

In fact, writing r:=|τ|r:=|\tau|, then a direct computation gives

Hessτ(τ22ψ(|τ|))=\displaystyle\mathrm{Hess}_{\tau}\left(\tau_{2}^{2}\,\psi(|\tau|)\right)= τ22[ψ(r)r𝕀2+(ψ′′(r)ψ(r)r)τr(τr)T]+2ψ(r)τr(0,τ2)\displaystyle\,\tau_{2}^{2}\,\left[\frac{\psi^{\prime}(r)}{r}\mathbb{I}_{2}+\left(\psi^{\prime\prime}(r)-\frac{\psi^{\prime}(r)}{r}\right)\frac{\tau}{r}\left(\frac{\tau}{r}\right)^{\mathrm{T}}\right]+2\psi^{\prime}(r)\frac{\tau}{r}(0,\tau_{2})
+2ψ(r)(0τ2)(τr)T+2ψ(r)(0001).\displaystyle+2\psi^{\prime}(r)\left(\begin{array}[]{c}0\\ \tau_{2}\\ \end{array}\right)\left(\frac{\tau}{r}\right)^{\mathrm{T}}+2\psi(r)\left(\begin{array}[]{cc}0&0\\ 0&1\\ \end{array}\right).

Next, it is easy to prove that for ξ=(ξ1,ξ2)2\xi=(\xi_{1},\xi_{2})\in\mathbb{R}^{2} satisfying |ξ|=1|\xi|=1,

11+π2(τ22+ξ22)τ22ξ12+ξ22τ22+ξ22,τ22π.\frac{1}{1+\pi^{2}}\big{(}\tau_{2}^{2}+\xi_{2}^{2}\big{)}\leq\tau_{2}^{2}\xi_{1}^{2}+\xi_{2}^{2}\leq\tau_{2}^{2}+\xi_{2}^{2},\quad\tau_{2}^{2}\leq\pi.

Hence it remains to show that for π|τ|ζ0\pi-|\tau|\geq\zeta_{0}, we have

τ22ξ12+ξ22\displaystyle\tau_{2}^{2}\xi_{1}^{2}+\xi_{2}^{2} τ22+ξ22ζ0Hessτ(τ22ψ(|τ|))ξ,ξ\displaystyle\sim\tau_{2}^{2}+\xi_{2}^{2}\sim_{\zeta_{0}}\langle\mathrm{Hess}_{\tau}\left(\tau_{2}^{2}\,\psi(|\tau|)\right)\,\xi,\xi\rangle (3.7)
=[ψ(r)r+(ψ′′(r)ψ(r)r)r02]τ22+4ψ(r)r0ξ2τ2+2ψ(r)ξ22,\displaystyle=\left[\frac{\psi^{\prime}(r)}{r}+\left(\psi^{\prime\prime}(r)-\frac{\psi^{\prime}(r)}{r}\right)r_{0}^{2}\right]\tau_{2}^{2}+4\psi^{\prime}(r)\,r_{0}\,\xi_{2}\,\tau_{2}+2\psi(r)\,\xi_{2}^{2},

where r0=ξτ|τ|[1,1]r_{0}=\xi\cdot\frac{\tau}{|\tau|}\in[-1,1]. To prove this, we write RHS of (3.7) as

(τ2ξ2)(ψ(r)r+(ψ′′(r)ψ(r)r)r022ψ(r)r02ψ(r)r02ψ(r))(τ2ξ2).\displaystyle\begin{pmatrix}\tau_{2}&\xi_{2}\end{pmatrix}\begin{pmatrix}\frac{\psi^{\prime}(r)}{r}+\left(\psi^{\prime\prime}(r)-\frac{\psi^{\prime}(r)}{r}\right)r_{0}^{2}&2\psi^{\prime}(r)\,r_{0}\\ 2\psi^{\prime}(r)\,r_{0}&2\psi(r)\end{pmatrix}\begin{pmatrix}\tau_{2}\\ \xi_{2}\end{pmatrix}. (3.8)

Note that we have uniformly for all 0rπζ00\leq r\leq\pi-\zeta_{0} that the diagonal entries of the last matrix

ψ(r)r+(ψ′′(r)ψ(r)r)r02ζ01,ψ(r)ζ01\displaystyle\frac{\psi^{\prime}(r)}{r}+\left(\psi^{\prime\prime}(r)-\frac{\psi^{\prime}(r)}{r}\right)r_{0}^{2}\sim_{\zeta_{0}}1,\quad\psi(r)\sim_{\zeta_{0}}1 (3.9)

by (3.6), (3.3) respectively, and the determinant

2{ψ(r)ψ(r)r(1r02)+(ψ(r)ψ′′(r)2ψ(r)2)r02}ζ01\displaystyle 2\left\{\psi(r)\frac{\psi^{\prime}(r)}{r}(1-r_{0}^{2})+\left(\psi(r)\,\psi^{\prime\prime}(r)-2\psi^{\prime}(r)^{2}\right)r_{0}^{2}\right\}\sim_{\zeta_{0}}1 (3.10)

by Lemma 3.1, which implies its eigenvalues ζ01\sim_{\zeta_{0}}1, thereby proving (3.7). ∎

Recall that ϑ1\vartheta_{1} denotes the unique solution of tanr=r\tan{r}=r in the interval (π, 1.5π)(\pi,\,1.5\,\pi). We will use repeatedly the following simple observation:

Lemma 3.4.

We have for any 0<rπ<ϑ10<r\neq\pi<\vartheta_{1} that:

Υ(r)=1ψ(r),Υ(r)=ψ(r)ψ(r)2,Υ′′(r)=ψ(r)ψ′′(r)+2ψ(r)2ψ(r)3.\displaystyle\Upsilon(r)=\frac{1}{\psi(r)},\quad\Upsilon^{\prime}(r)=-\frac{\psi^{\prime}(r)}{\psi(r)^{2}},\quad\Upsilon^{\prime\prime}(r)=\frac{-\psi(r)\psi^{\prime\prime}(r)+2\psi^{\prime}(r)^{2}}{\psi(r)^{3}}. (3.11)

Moreover,

ψ(r)r(πr)2,|ψ(r)|ϑ1r|πr|,for all 0<rπ<ϑ1\psi^{\prime}(r)\sim\frac{r}{(\pi-r)^{2}},\qquad|\psi(r)|\sim\frac{\vartheta_{1}-r}{|\pi-r|},\quad\mbox{for all $0<r\neq\pi<\vartheta_{1}$. } (3.12)
Proof.

The equations in (3.11) are trivial. For (3.12), the first claim (resp. the second one) is a direct consequence of the second equality in (3.6) (resp. (3.1), the fact that ψ(ϑ1)=0\psi(\vartheta_{1})=0 and ψ(ϑ1)>0\psi^{\prime}(\vartheta_{1})>0). ∎

Now we turn to the property of ψ\psi on (π,ϑ1)(\pi,\vartheta_{1}), which plays crucial roles in the proof of the second analytic-diffeomorphism in Theorem 2.1.

Lemma 3.5.

It holds for any r(0,+){jπ;j}r\in(0,\ +\infty)\setminus\{j\pi;\,j\in{\mathbb{N}}^{*}\} that:

K1:=ψ(r)r=4j=1+[(jπ)2r2]2>0.\displaystyle\mathrm{K}_{1}:=\frac{\psi^{\prime}(r)}{r}=4\sum_{j=1}^{+\infty}\left[(j\,\pi)^{2}-r^{2}\right]^{-2}>0. (3.13)

Moreover, we have for any π<r<ϑ1\pi<r<\vartheta_{1} that:

0>1rcotrr2=ψ(r)=2[1r2π2j=2+1(jπ)2r2],\displaystyle 0>\frac{1-r\cot{r}}{r^{2}}=\psi(r)=-2\left[\frac{1}{r^{2}-\pi^{2}}-\sum_{j=2}^{+\infty}\frac{1}{(j\,\pi)^{2}-r^{2}}\right], (3.14)
K2:=r1(ψ(r)r)=16[1[r2π2]3j=2+1[(jπ)2r2]3]<0,\displaystyle\mathrm{K}_{2}:=r^{-1}\left(\frac{\psi^{\prime}(r)}{r}\right)^{\prime}=-16\left[\frac{1}{[r^{2}-\pi^{2}]^{3}}-\sum_{j=2}^{+\infty}\frac{1}{[(j\,\pi)^{2}-r^{2}]^{3}}\right]<0, (3.15)
2ψ(r)K24K12<0,5K1+K2r2<0.\displaystyle 2\psi(r)\,\mathrm{K}_{2}-4\mathrm{K}_{1}^{2}<0,\qquad 5\mathrm{K}_{1}+\mathrm{K}_{2}\,r^{2}<0. (3.16)
Proof.

Using (3.6), we get (3.13) and (3.14) without difficulties. The inequality (3.15) follows from (3.14) and the observation that (jπ)2r2>r2π2(j\,\pi)^{2}-r^{2}>r^{2}-\pi^{2} for each j2j\geq 2.

From (3.14) and (3.15), we see that

2ψ(r)K2<2×2×161r2π2×1(r2π2)3=4×(4(r2π2)2)2<4K12,2\psi(r)\,\mathrm{K}_{2}<2\times 2\times 16\frac{1}{r^{2}-\pi^{2}}\times\frac{1}{(r^{2}-\pi^{2})^{3}}=4\times\left(\frac{4}{(r^{2}-\pi^{2})^{2}}\right)^{2}<4\,\mathrm{K}_{1}^{2},

where we have used (3.13) in the last inequality, which proves the first assertion of (3.16). To prove the second one, it suffices to employ (3.14) and the following:

5K1+K2r2=4ψ(r)r+ψ′′(r)=2ψ(r)r2csc2r1r2,0<rπ<ϑ1,\displaystyle 5\mathrm{K}_{1}+\mathrm{K}_{2}\,r^{2}=4\,\frac{\psi^{\prime}(r)}{r}+\psi^{\prime\prime}(r)=2\,\psi(r)\,\frac{r^{2}\csc^{2}r-1}{r^{2}},\quad 0<r\neq\pi<\vartheta_{1}, (3.17)

which is due to the trivial facts that

ψ(r)=r2csc2r+rcotr2r3,ψ′′(r)=2r3cotrcsc2r+r2csc2r+rcotr3r4.\psi^{\prime}(r)=\frac{r^{2}\csc^{2}r+r\cot r-2}{r^{3}},\quad\psi^{\prime\prime}(r)=-2\,\frac{r^{3}\cot r\csc^{2}r+r^{2}\csc^{2}r+r\cot r-3}{r^{4}}.

This finishes the proof of Lemma 3.5. ∎

Recall that (cf. (2.22)) h(r)=r2+rsinrcosr2sin2rh(r)=r^{2}+r\sin{r}\cos{r}-2\sin^{2}{r} for r>0r>0. Notice that h(jπ)=(jπ)2h(j\,\pi)=(j\,\pi)^{2} (jj\in\mathbb{N}^{*}) and h(r)=ψ(r)r3sin2rh(r)=\psi^{\prime}(r)\,r^{3}\sin^{2}r for r{jπ;j}r\not\in\{j\,\pi;\,j\in\mathbb{Z}^{*}\}. A direct consequence of (3.13) is the following:

Corollary 3.6.

We have h(r)>0h(r)>0 for all r>0r>0.

Now, we give the

3.2 Proof of the two key analytic-diffeomorphisms

Proof.

There are four steps.

Step 1. The Jacobian determinant of Λ\Lambda is positive on Ω+,1\Omega_{+,1} and negative on Ω,4\Omega_{-,4}, respectively. Notice that (cf. (2.17)) for 0<r=|(v1,v2)|π<ϑ10<r=|(v_{1},v_{2})|\neq\pi<\vartheta_{1} we have

Λ(v1,v2)=v[v22ψ(|v|)]=(ψ(r)rv1v22,v2[ψ(r)rv22+2ψ(r)]).\displaystyle\Lambda(v_{1},v_{2})=\nabla_{v}[v_{2}^{2}\,\psi(|v|)]=\left(\frac{\psi^{\prime}(r)}{r}\,v_{1}v_{2}^{2},\ v_{2}\left[\frac{\psi^{\prime}(r)}{r}\,v_{2}^{2}+2\,\psi(r)\right]\right).

Then the Jacobian matrix of Λ\Lambda at v=(v1,v2)v=(v_{1},v_{2}), saying JΛ(v,r)\mathrm{J}_{\Lambda}(v,r), equals by a routine calculation

(v22[ψ(r)r+(ψ(r)r)v12r]v1v2[2ψ(r)r+(ψ(r)r)v22r]v1v2[2ψ(r)r+(ψ(r)r)v22r]2ψ(r)+v22[5ψ(r)r+(ψ(r)r)v22r]).\displaystyle\left(\begin{array}[]{cc}v_{2}^{2}\left[\frac{\psi^{\prime}(r)}{r}+\left(\frac{\psi^{\prime}(r)}{r}\right)^{\prime}\frac{v_{1}^{2}}{r}\right]&\qquad v_{1}v_{2}\left[2\frac{\psi^{\prime}(r)}{r}+\left(\frac{\psi^{\prime}(r)}{r}\right)^{\prime}\frac{v_{2}^{2}}{r}\right]\\ \mbox{}&\qquad\mbox{}\\ v_{1}v_{2}\left[2\frac{\psi^{\prime}(r)}{r}+\left(\frac{\psi^{\prime}(r)}{r}\right)^{\prime}\frac{v_{2}^{2}}{r}\right]&\qquad 2\psi(r)+v_{2}^{2}\left[5\frac{\psi^{\prime}(r)}{r}+\left(\frac{\psi^{\prime}(r)}{r}\right)^{\prime}\frac{v_{2}^{2}}{r}\right]\\ \end{array}\right). (3.21)

Notice that JΛ(v,r)=Hessv(v22ψ(|v|))\mathrm{J}_{\Lambda}(v,r)=\mathrm{Hess}_{v}\left(v_{2}^{2}\,\psi(|v|)\right), which is positive definite on Ω+,1\Omega_{+,1} by Lemma 3.3. This in particular gives the first claim. For the second one, recalling (see (3.13), (3.15), and (2.14))

K1:=ψ(r)r,K2:=r1(ψ(r)r),K3:=2ψ(r)+K1v22,\displaystyle\mathrm{K}_{1}:=\frac{\psi^{\prime}(r)}{r},\quad\mathrm{K}_{2}:=r^{-1}\left(\frac{\psi^{\prime}(r)}{r}\right)^{\prime},\quad\mathrm{K}_{3}:=2\psi(r)+\mathrm{K}_{1}v_{2}^{2},

we find

detJΛ(v,r)=v22{(K1+K2v12)[K3+v22(4K1+K2v22)]v12(2K1+K2v22)2}:=v22K,\det\mathrm{J}_{\Lambda}(v,r)=v_{2}^{2}\left\{(\mathrm{K}_{1}+\mathrm{K}_{2}v_{1}^{2})\,[\mathrm{K}_{3}+v_{2}^{2}(4\mathrm{K}_{1}+\mathrm{K}_{2}v_{2}^{2})]-v_{1}^{2}(2\mathrm{K}_{1}+\mathrm{K}_{2}v_{2}^{2})^{2}\right\}:=v_{2}^{2}\,\mathrm{K},

with the observation that

K(v1,v2)\displaystyle\mathrm{K}(v_{1},v_{2}) =(2ψ(r)+K1v22)(K1+v12K2)+4K12v22+K1K2v244K12v12\displaystyle=(2\psi(r)+\mathrm{K}_{1}v_{2}^{2})\cdot(\mathrm{K}_{1}+v_{1}^{2}\mathrm{K}_{2})+4\mathrm{K}_{1}^{2}v_{2}^{2}+\mathrm{K}_{1}\mathrm{K}_{2}v_{2}^{4}-4\mathrm{K}_{1}^{2}v_{1}^{2}
=2ψ(r)K1+v12[2ψ(r)K24K12]+v22K1(5K1+K2r2).\displaystyle=2\psi(r)\mathrm{K}_{1}+v_{1}^{2}\left[2\psi(r)\mathrm{K}_{2}-4\mathrm{K}_{1}^{2}\right]+v_{2}^{2}\,\mathrm{K}_{1}\left(5\mathrm{K}_{1}+\mathrm{K}_{2}\,r^{2}\right). (3.22)

Hence from (3.13), (3.14), and (3.16), we conclude that detJΛ(v,r)\det\mathrm{J}_{\Lambda}(v,r) is negative on Ω,4\Omega_{-,4}.

Step 2. The analytic map Λ\Lambda is a diffeomorphism from Ω+,1\Omega_{+,1} onto >,+2\mathbb{R}^{2}_{>,+}. First we claim that Λ\Lambda is from Ω+,1\Omega_{+,1} into >,+2\mathbb{R}^{2}_{>,+}, namely,

v2(2ψ(r)+ψ(r)v22r)>2πψ(r)v22rv1,v1,v2>0 with r=v12+v22<π.\displaystyle v_{2}\left(2\psi(r)+\psi^{\prime}(r)\frac{v_{2}^{2}}{r}\right)>\frac{2}{\sqrt{\pi}}\sqrt{\psi^{\prime}(r)\frac{v_{2}^{2}}{r}v_{1}},\quad v_{1},v_{2}>0\mbox{ with }r=\sqrt{v_{1}^{2}+v_{2}^{2}}<\pi.

This is a direct consequence of the inequality ψ(r)>ψ(r)r\psi(r)>\sqrt{\frac{\psi^{\prime}(r)}{r}} for 0r<π0\leq r<\pi; see (3.5).

Notice that Ω+,1\Omega_{+,1} and >,+2\mathbb{R}^{2}_{>,+} are both connected and simply connected and we have shown the Jacobian determinant of Λ\Lambda vanishes nowhere in Ω+,1\Omega_{+,1}. Then by Hadamard’s theorem (see for example [27, §  6.2]), it remains to prove that Λ\Lambda is proper, that is, whenever {v(j)}j=1+Ω+,1\{v^{(j)}\}_{j=1}^{+\infty}\subseteq\Omega_{+,1} satisfies v(j)=(v1(j),v2(j))Ω+,1v^{(j)}=(v^{(j)}_{1},v^{(j)}_{2})\to\partial\Omega_{+,1} then Λ(v(j))>,+2\Lambda(v^{(j)})\to\partial\mathbb{R}^{2}_{>,+}. To see this, by contradiction, assume that Λ\Lambda is not proper. Then we can find a sequence {v(j)}j=1+Ω+,1\{v^{(j)}\}_{j=1}^{+\infty}\subseteq\Omega_{+,1} such that v(j)Ω+,1v^{(j)}\to\partial\Omega_{+,1} but {Λ(v(j))}j=1+\{\Lambda(v^{(j)})\}_{j=1}^{+\infty} stays in a compact set in >,+2\mathbb{R}^{2}_{>,+}. After picking subsequences, we can assume further that v(j)v(0)Ω+,1v^{(j)}\to v^{(0)}\in\partial\Omega_{+,1} and Λ(v(j))(a1,a2)>,+2\Lambda(v^{(j)})\to(a_{1},a_{2})\in\mathbb{R}^{2}_{>,+}. Recalling the definition of ψ\psi (see (3.1)), we always obtain a contradiction in each of the following 4 possible cases:

  1. (i)

    If v(0)[0,π)×{0}v^{(0)}\in[0,\pi)\times\{0\}, then Λ(v(j))0>,+2\Lambda(v^{(j)})\to 0\in\partial\mathbb{R}^{2}_{>,+}.

  2. (ii)

    If v(0){v;|v|=π,v10,v2>0}v^{(0)}\in\{v;\,|v|=\pi,v_{1}\geq 0,v_{2}>0\}, by the fact that (cf. (3.6))

    limrπ(πr)ψ(r)=1π,limrπ(πr)2ψ(r)=1π,\displaystyle\lim_{r\to\pi^{-}}(\pi-r)\,\psi(r)=\frac{1}{\pi},\qquad\lim_{r\to\pi^{-}}(\pi-r)^{2}\,\psi^{\prime}(r)=\frac{1}{\pi}, (3.23)

    we have Λ(v(j))>,+2\Lambda(v^{(j)})\to\infty\in\partial\mathbb{R}^{2}_{>,+}.

  3. (iii)

    If v(0){0}×(0,π)v^{(0)}\in\{0\}\times(0,\pi), then Λ(v(j)){0}×(0,+)>,+2\Lambda(v^{(j)})\to\{0\}\times(0,+\infty)\subseteq\partial\mathbb{R}^{2}_{>,+}.

  4. (iv)

    Assume v(0)=(π,0)Ω+,1v^{(0)}=(\pi,0)\in\partial\Omega_{+,1}. Set Λ(v(j))=(u1(j),u2(j))\Lambda(v^{(j)})=(u^{(j)}_{1},u^{(j)}_{2}). Then we have

    a1=limj+u1(j)=limj+v1(j)|v(j)|(v2(j))2ψ(|v(j)|)=limj+1π(v2(j)π|v(j)|)2,\displaystyle a_{1}=\lim_{j\to+\infty}u^{(j)}_{1}=\lim_{j\to+\infty}\frac{v^{(j)}_{1}}{|v^{(j)}|}\left(v^{(j)}_{2}\right)^{2}\psi^{\prime}(|v^{(j)}|)=\lim_{j\to+\infty}\frac{1}{\pi}\left(\frac{v^{(j)}_{2}}{\pi-|v^{(j)}|}\right)^{2},

    where we have used (3.23) in the last equality. Using (3.23) again, it implies that

    limj+u2(j)\displaystyle\lim_{j\to+\infty}u^{(j)}_{2} =limj+[2v2(j)ψ(|v(j)|)+v2(j)|v(j)|(v2(j))2ψ(|v(j)|)]\displaystyle=\lim_{j\to+\infty}\left[2v^{(j)}_{2}\psi(|v^{(j)}|)+\frac{v^{(j)}_{2}}{|v^{(j)}|}\left(v^{(j)}_{2}\right)^{2}\psi^{\prime}(|v^{(j)}|)\right]
    =limj+2πv2(j)π|v(j)|=2πa1,\displaystyle=\lim_{j\to+\infty}\frac{2}{\pi}\frac{v^{(j)}_{2}}{\pi-|v^{(j)}|}=\frac{2}{\sqrt{\pi}}\sqrt{a_{1}},

    and hence (a1,2πa1)>,+2(a_{1},\frac{2}{\sqrt{\pi}}\sqrt{a_{1}})\in\partial\mathbb{R}^{2}_{>,+}.

Step 3. Λ\Lambda is a diffeomorphism from Ω,4\Omega_{-,4} onto <,+2\mathbb{R}^{2}_{<,+}. First, we need to show Λ\Lambda is from Ω,4\Omega_{-,4} into <,+2{\mathbb{R}}^{2}_{<,+}, i.e,

2ψ(r)+ψ(r)rv22>2πψ(r)rv1,(v1,v2)Ω,4.\displaystyle 2\psi(r)+\frac{\psi^{\prime}(r)}{r}v_{2}^{2}>-\frac{2}{\sqrt{\pi}}\sqrt{\frac{\psi^{\prime}(r)}{r}v_{1}},\quad\forall\,(v_{1},v_{2})\in\Omega_{-,4}.

Using (3.13), it remains to prove that ψ(r)>ψ(r)r\psi(r)>-\sqrt{\frac{\psi^{\prime}(r)}{r}}, which is obvious by (3.14) and (3.13).

Using an argument similar to that in Step 2, it is easily seen that the smooth function Λ\Lambda is proper. Moreover, we have obtained in Step 1 that the Jacobian determinant of Λ\Lambda vanishes nowhere on Ω,4\Omega_{-,4}. From these and Hadamard’s theorem the desired conclusion follows, since both Ω,4\Omega_{-,4} and <,+2\mathbb{R}^{2}_{<,+} are connected and simply connected.

Step 4. The 3×33\times 3 matrix Hessθϕ(gu;θ){\mathrm{Hess}}_{\theta}\,\phi(g_{u};\theta) is nonsingular, negative definite on Ω+,1\Omega_{+,1}, and has exactly two positive eigenvalues on Ω,4\Omega_{-,4}. Indeed, setting θ~=(θ1,θ2):=Λ1(u~)\tilde{\theta}=\left(\theta_{1},\theta_{2}\right):=\Lambda^{-1}(\tilde{u}), then a direct calculation gives

Hessθϕ(gu;θ)=(JΛ(θ~,|θ|)𝕆2×1𝕆1×22ψ(|θ|)+θ22ψ(|θ|)|θ|=K3(θ~)).{\mathrm{Hess}}_{\theta}\,\phi(g_{u};\theta)=-\left(\begin{array}[]{cc}\mathrm{J}_{\Lambda}(\widetilde{\theta},|\theta|)&\mathbb{O}_{2\times 1}\\[2.84526pt] \mathbb{O}_{1\times 2}&2\psi(|\theta|)+\theta_{2}^{2}\frac{\psi^{\prime}(|\theta|)}{|\theta|}=\mathrm{K}_{3}(\tilde{\theta})\end{array}\right). (3.24)

Note that we have K3(θ~)>0\mathrm{K}_{3}(\tilde{\theta})>0 on Ω+,1\Omega_{+,1} by (3.13), and K3(θ~)<0\mathrm{K}_{3}(\tilde{\theta})<0 on Ω,4\Omega_{-,4} (cf. (2.16)). Moreover, from Step 1 we know the matrix JΛ(θ~,|θ|)\mathrm{J}_{\Lambda}(\widetilde{\theta},|\theta|) is positive definite on Ω+,1\Omega_{+,1} and detJΛ(θ~,|θ|)=θ22K(θ~)<0\mathrm{det\,J}_{\Lambda}(\widetilde{\theta},|\theta|)=\theta_{2}^{2}\,\mathrm{K}(\tilde{\theta})<0 on Ω,4\Omega_{-,4}. Then the claim of Step 4 follows easily and the proof of Theorem 2.1 is completed. ∎

The above proof and the fact that Hessθϕ(g;θ)=|x|2Hessθϕ(gu;θ)\mathrm{Hess}_{\theta}\,\phi(g;\theta)=|x|^{2}\,\mathrm{Hess}_{\theta}\,\phi(g_{u};\theta) give the following lemma immediately.

Lemma 3.7.

Let gg and θ\theta be given as in Assumption (A) (cf. (2.18)). We have

det(Hessθϕ(g;θ))=θ22K(θ1,θ2)K3(θ1,θ2)|x|6.\det(-\mathrm{Hess}_{\theta}\,\phi(g;\theta))=\theta_{2}^{2}\,\mathrm{K}(\theta_{1},\theta_{2})\,\mathrm{K}_{3}(\theta_{1},\theta_{2})\,|x|^{6}. (3.25)
Remark 3.8.

Suppose Assumption (A) holds. Then from (2.17) it follows that

u1θ1θ22|π|θ||2;\displaystyle u_{1}\sim\theta_{1}\,\theta_{2}^{2}\,|\pi-|\theta|\,|^{-2}; (3.26)
u2|θ2|3|π|θ||2+|θ2||π|θ||1,as|θ|<π;\displaystyle u_{2}\sim|\theta_{2}|^{3}\,|\pi-|\theta|\,|^{-2}+|\theta_{2}|\,|\pi-|\theta|\,|^{-1},\quad{\rm as}\,\,|\theta|<\pi; (3.27)
θ11andϵ|π|θ||2|ψ(|θ|)|=u1θ1+u2|θ2|θ22(π|θ|)2+u2|θ2|,as|θ|>π,\displaystyle\theta_{1}\sim 1\,\,{\rm and}\,\,\frac{\epsilon}{|\pi-|\theta|\,|}\sim 2\,|\psi(|\theta|)|=\frac{u_{1}}{\theta_{1}}+\frac{u_{2}}{|\theta_{2}|}\sim\frac{\theta_{2}^{2}}{(\pi-|\theta|)^{2}}+\frac{u_{2}}{|\theta_{2}|},\quad{\rm as}\,\,|\theta|>\pi, (3.28)

where we have used (3.12) with r=|θ|r=|\theta|. These estimates will be used in the proof of Lemma 7.6 below.

Recall that K3\mathrm{K}_{3} and Ω,4\Omega_{-,4} are defined by (2.14) and (2.16), respectively. The following characterization of Ω,4\Omega_{-,4} will be used in the proof of Theorem 2.10 (cf. Subsection 4.3 below):

Lemma 3.9.

We have

Ω,4={(v1,v2);v2<0<v1,K3(v1,v2)<0,πr=v12+v22<ϑ1}.\displaystyle\Omega_{-,4}=\left\{(v_{1},v_{2});\,v_{2}<0<v_{1},\ \mathrm{K}_{3}(v_{1},v_{2})<0,\ \pi\neq r=\sqrt{v_{1}^{2}+v_{2}^{2}}<\vartheta_{1}\right\}. (3.29)
Proof.

Indeed, Ω,4\Omega_{-,4} is obviously contained in RHS. Conversely, given a point (v1,v2)(v_{1},v_{2}) in RHS, since it deduces from (3.6) that

0>K3(v1,v2)\displaystyle 0>\mathrm{K}_{3}\left(v_{1},v_{2}\right) =4[j=1+v22((jπ)2r2)2+j=2+((jπ)2r2)11r2π2]\displaystyle=4\left[\sum_{j=1}^{+\infty}v_{2}^{2}\left((j\pi)^{2}-r^{2}\right)^{-2}+\sum_{j=2}^{+\infty}\left((j\pi)^{2}-r^{2}\right)^{-1}-\frac{1}{r^{2}-\pi^{2}}\right]
>4[v22(π2r2)21r2π2]=4π2v12(r2π2)2,\displaystyle>4\left[v_{2}^{2}\,(\pi^{2}-r^{2})^{-2}-\frac{1}{r^{2}-\pi^{2}}\right]=4\,\frac{\pi^{2}-v_{1}^{2}}{\left(r^{2}-\pi^{2}\right)^{2}},

then r>v1>πr>v_{1}>\pi whenever v1>0>v2v_{1}>0>v_{2}, which implies that (v1,v2)Ω,4(v_{1},v_{2})\in\Omega_{-,4}. ∎

The expression of the function K\mathrm{K} (cf. (3.22)) in fact can be simplified as the following lemma shows, which will be used in Section 11.

Lemma 3.10.

Suppose that v=(v1,v2)2v=(v_{1},v_{2})\in{\mathbb{R}}^{2} with 0<|v|π<ϑ10<|v|\neq\pi<\vartheta_{1}. Then

K(v1,v2)=2Υ(|v|)3|v|[Υ′′(|v|)|v|v12+Υ(|v|)sin2(|v|)v22].\displaystyle\mathrm{K}(v_{1},v_{2})=\frac{2}{\Upsilon(|v|)^{3}\,|v|}\left[\frac{-\Upsilon^{\prime\prime}(|v|)}{|v|}v_{1}^{2}+\frac{-\Upsilon^{\prime}(|v|)}{\sin^{2}(|v|)}v_{2}^{2}\right]. (3.30)
Proof.

Write r:=|v|r:=|v|. Then from (3.22) and a routine computation it follows that

K(v1,v2)\displaystyle\mathrm{K}(v_{1},v_{2}) =2ψ(r)K1+v12[2ψ(r)K24K12]+v22K1(5K1+K2r2)\displaystyle=2\psi(r)\mathrm{K}_{1}+v_{1}^{2}\left[2\psi(r)\mathrm{K}_{2}-4\mathrm{K}_{1}^{2}\right]+v_{2}^{2}\,\mathrm{K}_{1}\left(5\mathrm{K}_{1}+\mathrm{K}_{2}\,r^{2}\right)
=2v12[ψ(r)r2K1+ψ(r)K22K12]+v22K1[2ψ(r)r2+5K1+K2r2]\displaystyle=2v_{1}^{2}\left[\frac{\psi(r)}{r^{2}}\mathrm{K}_{1}+\psi(r)\mathrm{K}_{2}-2\mathrm{K}_{1}^{2}\right]+v_{2}^{2}\,\mathrm{K}_{1}\left[2\frac{\psi(r)}{r^{2}}+5\mathrm{K}_{1}+\mathrm{K}_{2}\,r^{2}\right]
\xlongequal(3.17)2v12[ψ(r)(K1r2+K2)2ψ(r)2r2]+v22ψ(r)r[2ψ(r)r2+2ψ(r)r2(r2csc2r1)]\displaystyle\xlongequal{\eqref{psi_sum}}2v_{1}^{2}\left[\psi(r)\,\left(\frac{\mathrm{K}_{1}}{r^{2}}+\mathrm{K}_{2}\right)-2\frac{\psi^{\prime}(r)^{2}}{r^{2}}\right]+v_{2}^{2}\,\frac{\psi^{\prime}(r)}{r}\left[2\frac{\psi(r)}{r^{2}}+2\frac{\psi(r)}{r^{2}}(r^{2}\csc^{2}r-1)\right]
=2v12[ψ(r)ψ′′(r)2ψ(r)2r2]+2v22[ψ(r)ψ(r)rsin2r]\displaystyle=2v_{1}^{2}\left[\frac{\psi(r)\psi^{\prime\prime}(r)-2\psi^{\prime}(r)^{2}}{r^{2}}\right]+2v_{2}^{2}\left[\frac{\psi(r)\psi^{\prime}(r)}{r\sin^{2}r}\right]
\xlongequal(3.11)2Υ(r)3r[Υ′′(r)rv12+Υ(r)sin2(r)v22].\displaystyle\xlongequal{\eqref{relUp}}\frac{2}{\Upsilon(r)^{3}\,r}\left[\frac{-\Upsilon^{\prime\prime}(r)}{r}v_{1}^{2}+\frac{-\Upsilon^{\prime}(r)}{\sin^{2}(r)}v_{2}^{2}\right].

This is the desired result. ∎

4 Proof of Proposition 2.8 and Theorem 2.10

In this section, we will show the more useful integral expression for the heat kernel (2.30), which plays a role in overcoming our major obstacle in Section 8 below. Furthermore, we give some basic properties for the “intrinsic distance” 𝐃\mathbf{D}, as well as the proof of Theorem 2.10. Let us begin with the

4.1 Proof of Proposition 2.8

Proof.

Let ϑk(k)\vartheta_{k}\,(k\in\mathbb{N}^{*}) denote the unique solution of tanr=r\tan{r}=r in (kπ,(k+12)π)(k\pi,(k+\frac{1}{2})\pi). The following equality, which is a counterpart of (2.4) for r/sinhrr/\sinh{r}, comes from the property of the Bessel function J32J_{\frac{3}{2}} (cf. [22, § 8.544, § 8.464.3])

r3rcoshrsinhr\displaystyle\frac{r^{3}}{r\cosh{r}-\sinh{r}} =3k=1+(1+r2ϑk2)1.\displaystyle=3\,\prod_{k=1}^{+\infty}\left(1+\frac{r^{2}}{\vartheta_{k}^{2}}\right)^{-1}. (4.1)

Then taking logarithmic derivative of the equality above, a direct calculation gives:

r2rcothr1\displaystyle\frac{r^{2}}{r\coth{r}-1} =3+2k=1+r2ϑk2+r2=:Υ~(r).\displaystyle=3+2\,\sum_{k=1}^{+\infty}\frac{r^{2}}{\vartheta_{k}^{2}+r^{2}}=:\widetilde{\Upsilon}(r). (4.2)

Next, let AA be a real, positive definite q×qq\times q matrix, and YqY\in\mathbb{C}^{q}. Recall the well-known formula (cf. for instance [25, Theorem 7.6.1])

qe12As,s+iY,s𝑑s=(2π)q21detAe12A1Y,Y.\displaystyle\int_{\mathbb{R}^{q}}e^{-\frac{1}{2}\langle A\,s,\,s\rangle+i\langle Y,\,s\rangle}ds=(2\pi)^{\frac{q}{2}}\,\frac{1}{\sqrt{\det{A}}}\,e^{-\frac{1}{2}\langle A^{-1}\,Y,\,Y\rangle}. (4.3)

Now applying this formula with q=2q=2, Y=|x|2𝐰¯2(λ2,λ3)Y=\frac{|x|^{2}\,\mathbf{\overline{w}}}{2}(\lambda_{2},\lambda_{3}) and A=12|λ|2|λ|coth|λ|1|x|2𝐰¯2𝕀2A=\frac{1}{2}\frac{|\lambda|^{2}}{|\lambda|\coth|\lambda|-1}|x|^{2}\,\mathbf{\overline{w}}^{2}\,{\mathbb{I}}_{2} to (2.19), we obtain the desired result

p(g)=14π|x|2𝐰¯2e|x|242𝐏(s|x|𝐰¯,14|x|2(u+2𝐰¯s1e2+2𝐰¯s2e3))𝑑s,\displaystyle p(g)=\frac{1}{4\pi}\,|x|^{2}\,\mathbf{\overline{w}}^{2}\,e^{-\frac{|x|^{2}}{4}}\int_{\mathbb{R}^{2}}\mathbf{P}\left(s|x|\,\mathbf{\overline{w}},\frac{1}{4}|x|^{2}(u+2\mathbf{\overline{w}}\,s_{1}\,e_{2}+2\mathbf{\overline{w}}\,s_{2}\,e_{3})\right)\,ds,

where

𝐏(X,T)=3𝒱(λ)e14Γ~((X,T);λ)𝑑λ,\displaystyle\mathbf{P}(X,T)=\int_{\mathbb{R}^{3}}\,\mathcal{V}(\lambda)\,e^{-\frac{1}{4}\widetilde{\Gamma}((X,T);\lambda)}\,d\lambda, (4.4)

with

𝒱(λ)=|λ|3|λ|cosh|λ|sinh|λ|=3k=1+(1+λλϑk2)1,\displaystyle\mathcal{V}(\lambda)=\frac{|\lambda|^{3}}{|\lambda|\cosh{|\lambda|}-\sinh{|\lambda|}}=3\,\prod_{k=1}^{+\infty}\left(1+\frac{\lambda\cdot\lambda}{\vartheta_{k}^{2}}\right)^{-1}, (4.5)
Γ~((X,T);λ)=Υ~(|λ|)|X|24iTλ=3|X|2+k=1+2|λ|2ϑk2+|λ|2|X|24iTλ.\displaystyle\widetilde{\Gamma}((X,T);\lambda)=\widetilde{\Upsilon}(|\lambda|)\,|X|^{2}-4i\,T\cdot\lambda=3|X|^{2}+\sum_{k=1}^{+\infty}\frac{2|\lambda|^{2}}{\vartheta_{k}^{2}+|\lambda|^{2}}|X|^{2}-4i\,T\cdot\lambda. (4.6)

Notice that (4.4)-(4.6) have been defined in Subsection 2.5. ∎

4.2 The “intrinsic distance” 𝐃\mathbf{D} associated to 𝐏\mathbf{P}

We are in a position to provide some basic properties of the squared “intrinsic distance” associated to 𝐏\mathbf{P},

𝐃(X,T)2:=supτB3(0,ϑ1)Γ((X,T);τ),(X,T)2×3,\displaystyle\mathbf{D}(X,T)^{2}:=\sup_{\tau\in B_{\mathbb{R}^{3}}(0,\vartheta_{1})}\Gamma((X,T);\tau),\qquad(X,T)\in\mathbb{R}^{2}\times\mathbb{R}^{3},

where the smooth function Γ((X,T);)\Gamma((X,T);\cdot) is defined by

Γ((X,T);τ):=Γ~((X,T);iτ)=Υ(|τ|)|X|2+4Tτ,τB3(0,ϑ1),\displaystyle\Gamma((X,T);\tau):=\widetilde{\Gamma}((X,T);i\tau)=\Upsilon(|\tau|)|X|^{2}+4T\cdot\tau,\qquad\tau\in B_{\mathbb{R}^{3}}(0,\vartheta_{1}), (4.7)

with

Υ(r):=r21rcotr(=1ψ(r),r±π)=32k=1+r2ϑk2r2,ϑ1<r<ϑ1,\displaystyle\Upsilon(r):=\frac{r^{2}}{1-r\cot{r}}(=\frac{1}{\psi(r)},\ r\neq\pm\pi)=3-2\sum_{k=1}^{+\infty}\frac{r^{2}}{\vartheta_{k}^{2}-r^{2}},\,\,-\vartheta_{1}<r<\vartheta_{1}, (4.8)

where we have used (4.2) in the last equality with a complexification. Hence Υ-\Upsilon is operator convex on (ϑ1,ϑ1)(-\vartheta_{1},\ \vartheta_{1}) in the sense of [11].

From the above series expansion, one has immediately

Υ(r)=4k=1+ϑk2r(ϑk2r2)2,Υ′′(r)=4k=1+ϑk2(ϑk2+3r2)(ϑk2r2)3(<0),Υ′′′(r)=48k=1+ϑk2(ϑk2+r2)r(ϑk2r2)4,r(ϑ1,ϑ1).\begin{gathered}\Upsilon^{\prime}(r)=-4\sum_{k=1}^{+\infty}\frac{\vartheta_{k}^{2}\,r}{(\vartheta_{k}^{2}-r^{2})^{2}},\quad\Upsilon^{\prime\prime}(r)=-4\sum_{k=1}^{+\infty}\frac{\vartheta_{k}^{2}\,(\vartheta_{k}^{2}+3r^{2})}{(\vartheta_{k}^{2}-r^{2})^{3}}\ (<0),\\ \Upsilon^{\prime\prime\prime}(r)=-48\sum_{k=1}^{+\infty}\frac{\vartheta_{k}^{2}\,(\vartheta_{k}^{2}+r^{2})\,r}{(\vartheta_{k}^{2}-r^{2})^{4}},\quad r\in(-\vartheta_{1},\vartheta_{1}).\end{gathered} (4.9)

In particular, we get the following counterpart of the key function μ(r)=(2rsin2r)/(2sin2r)\mu(r)=(2r-\sin{2r})/(2\sin^{2}{r}) (π<r<π-\pi<r<\pi) in the setting of Heisenberg group as well as generalized Heisenberg-type groups (cf. [21, Lemma 3, p. 112], [8, Theorem 1.36] and [35, Section 7]):

Lemma 4.1.

The function rΥ(r)r\mapsto-\Upsilon^{\prime}(r) is an odd increasing diffeomorphism between (ϑ1,ϑ1)(-\vartheta_{1},\vartheta_{1}) and \mathbb{R}.

We denote the smooth inverse function of Υ()-\Upsilon^{\prime}(\cdot) by 𝒵()\mathcal{Z}(\cdot), that is,

Υ(𝒵(ρ)):=ρ,ρ.-\Upsilon^{\prime}(\mathcal{Z}(\rho)):=\rho,\quad\rho\in\mathbb{R}. (4.10)

And we define a smooth function Φ()\Phi(\cdot) on the real line by setting

Φ(ρ):=Υ(𝒵(ρ))+ρ𝒵(ρ),ρ.\displaystyle\Phi(\rho):=\Upsilon(\mathcal{Z}(\rho))+\rho\,\mathcal{Z}(\rho),\quad\rho\in\mathbb{R}. (4.11)

The following result (especially the assertions (ii) and (iii) below) says that 𝐃\mathbf{D} can be imagined as a Carnot–Carathéodory distance on 2×3\mathbb{R}^{2}\times\mathbb{R}^{3}. Indeed it comes from the same spirit of the characterization for the squared Carnot–Carathéodory distance on GM-groups, even on GM-Métivier groups (cf. [33] and [35]). So one can consider that we are in a special case of GM-groups.

Proposition 4.2.

It holds that:

  1. (i)

    The smooth function Γ((X,T);)\Gamma((X,T);\cdot) is concave in B3(0,ϑ1)B_{\mathbb{R}^{3}}(0,\vartheta_{1}).

  2. (ii)

    The function 𝐃(X,T)2\mathbf{D}(X,T)^{2} has the following explicit expression:

    𝐃(X,T)2={4ϑ1|T|,if |X|=0,Υ(|τ|)|X|2+4Tτ=Φ(4|T||X|2)|X|2,if |X|0,\displaystyle\mathbf{D}(X,T)^{2}=\begin{cases}4\vartheta_{1}|T|,&\mbox{if \ }|X|=0,\\ \Upsilon(|\tau^{*}|)|X|^{2}+4T\cdot\tau^{*}=\Phi\left(\frac{4|T|}{|X|^{2}}\right)|X|^{2},&\mbox{if \ }|X|\neq 0,\end{cases} (4.12)

    where τ:=τ(X,T)\tau^{*}:=\tau^{*}(X,T) denotes the unique critical point of Γ((X,T);)\Gamma((X,T);\cdot) in B3(0,ϑ1)B_{\mathbb{R}^{3}}(0,\vartheta_{1}), i.e.,

    Υ(|τ|)|τ||X|2τ+4T=0,soτ=𝒵(4|T||X|2)T^,\displaystyle\frac{\Upsilon^{\prime}(|\tau^{*}|)}{|\tau^{*}|}|X|^{2}\,\tau^{*}+4T=0,\quad{\rm so}\quad\tau^{*}=\mathcal{Z}\left(\frac{4|T|}{|X|^{2}}\right)\,\widehat{T}, (4.13)

    with the convention T^=0\widehat{T}=0 for T=0T=0, and T^=T/|T|\widehat{T}=T/|T| otherwise.

  3. (iii)

    𝐃(X,T)2>0\mathbf{D}(X,T)^{2}>0 for any (X,T)(0,0)(X,T)\neq(0,0). Moreover, the following properties hold for all (X,T)2×3(X,T)\in\mathbb{R}^{2}\times\mathbb{R}^{3} :

    𝐃(X,T)2|X|2+|T|,𝐃(Xh,Th)2=1h𝐃(X,T)2,h>0.\mathbf{D}(X,T)^{2}\sim|X|^{2}+|T|,\qquad\mathbf{D}\left(\frac{X}{\sqrt{h}},\frac{T}{h}\right)^{2}=\frac{1}{h}\,\mathbf{D}(X,T)^{2},\qquad\forall\,h>0. (4.14)
  4. (iv)

    𝐃2\mathbf{D}^{2} is continuous on 2×3\mathbb{R}^{2}\times\mathbb{R}^{3}.

Proof.

For the reader’s convenience, we provide a direct proof.

First, the concavity can be directly verified by showing that HessτΓ((X,T),τ)0\mathrm{Hess}_{\tau}\Gamma((X,T),\tau)\leq 0 from the series expression (4.8). Now we prove (ii). Recall that for a smooth concave function, the maximizer is equivalent to the critical point (cf. [42, IMPORTANT, p.146]). Through a simple computation we have

𝐃(X,T)2={4ϑ1|T|,if |X|=0,Υ(|τ|)|X|2+4Tτ,if |X|0,\displaystyle\mathbf{D}(X,T)^{2}=\begin{cases}4\vartheta_{1}|T|,&\mbox{if \ }|X|=0,\\ \Upsilon(|\tau^{*}|)|X|^{2}+4T\cdot\tau^{*},&\mbox{if \ }|X|\neq 0,\end{cases}

where τ\tau^{*} is defined by (4.13) (this is clearly consistent with the one defined in Proposition 2.9 (iii)). From this and (4.11) we obtain (ii).

Next we show (iii). It follows from (4.8) that Υ(r)3\Upsilon(r)\leq 3. Hence (4.7) and the definition of 𝐃2\mathbf{D}^{2} yield that 𝐃(X,T)23|X|2+4ϑ1|T|\mathbf{D}(X,T)^{2}\leq 3|X|^{2}+4\vartheta_{1}\,|T|. Next, notice that Γ((X,T);0)=3|X|2\Gamma((X,T);0)=3|X|^{2}, and Γ((X,T);πT^)=4π|T|\Gamma((X,T);\pi\,\widehat{T})=4\pi|T| since Υ(π)=0\Upsilon(\pi)=0 (cf. the first equality in (4.8)). In conclusion, we get that 𝐃(X,T)2|X|2+|T|\mathbf{D}(X,T)^{2}\sim|X|^{2}+|T|, which implies immediately the first claim. On the other hand, the scaling property 𝐃(X/h,T/h)2=𝐃(X,T)2/h\mathbf{D}(X/\sqrt{h},T/h)^{2}=\mathbf{D}(X,T)^{2}/h is clear from the definition of 𝐃2\mathbf{D}^{2}.

We are left with the proof of (iv). Notice that the continuity of 𝐃2\mathbf{D}^{2} at (X0,T0)(X_{0},T_{0}) where X00X_{0}\neq 0 follows from (4.12) and the smoothness of Φ\Phi. And the continuity at (X0,T0)=(0,0)(X_{0},T_{0})=(0,0) follows easily from the first formula of (4.14). For the point (X0,T0)=(0,T0)(X_{0},T_{0})=(0,T_{0}) where T00T_{0}\neq 0, initially by (4.12) we have 𝐃(0,T0)2=4ϑ1|T0|\mathbf{D}(0,T_{0})^{2}=4\vartheta_{1}|T_{0}|. Supposing (X,T)(0,T0)(X,T)\to(0,T_{0}), we consider two cases. If X=0X=0, then by the first equality in (4.12) we have 𝐃(X,T)2=4ϑ1|T|𝐃(X0,T0)2\mathbf{D}(X,T)^{2}=4\vartheta_{1}|T|\to\mathbf{D}(X_{0},T_{0})^{2}. If X0X\neq 0, by (4.13) we have |X|2=4|T|/Υ(|τ|)|X|^{2}=-4|T|/\Upsilon^{\prime}(|\tau^{*}|) and |τ|=𝒵(4|T|/|X|2)ϑ1|\tau^{*}|=\mathcal{Z}(4|T|/|X|^{2})\to\vartheta_{1}^{-} (from Lemma 4.1 and the fact that |T|/|X|2+|T|/|X|^{2}\to+\infty). Then noticing that Υ(r)ϑ1/(ϑ1r)\Upsilon(r)\sim-\vartheta_{1}/(\vartheta_{1}-r) and Υ(r)ϑ1/(ϑ1r)2\Upsilon^{\prime}(r)\sim-\vartheta_{1}/(\vartheta_{1}-r)^{2} for 0<ϑ1r0<\vartheta_{1}-r small enough, by (4.12) we obtain

𝐃(X,T)2=4|T|(|τ|Υ(|τ|)Υ(|τ|))4ϑ1|T0|=𝐃(X0,T0)2\mathbf{D}(X,T)^{2}=4|T|\left(|\tau^{*}|-\frac{\Upsilon(|\tau^{*}|)}{\Upsilon^{\prime}(|\tau^{*}|)}\right)\to 4\vartheta_{1}|T_{0}|=\mathbf{D}(X_{0},T_{0})^{2}

as well. ∎

Remark 4.3.

Combining (4.12) with (7.7) below we can see that 𝐃2\mathbf{D}^{2} is not C1C^{1} at {(0,T);T0}\{(0,T);\,T\neq 0\}. Moreover, 𝐃2\mathbf{D}^{2} is not C1C^{1} at (0,0)(0,0) by the first estimate of (4.14). In conclusion, 𝐃2\mathbf{D}^{2} is smooth on (2{0})×3(\mathbb{R}^{2}\setminus\{0\})\times\mathbb{R}^{3} and is not C1C^{1} on {0}×3\{0\}\times\mathbb{R}^{3}.

Before providing the proof of Theorem 2.10, we recall that (cf. (2.23)) φ1(r)=r2sin2rrsinrcosr>0\varphi_{1}(r)=\frac{r^{2}-\sin^{2}{r}}{r-\sin{r}\cos{r}}>0 for r>0r>0, and state the following simple observation:

Lemma 4.4.

φ1>0\varphi_{1}>0 is strictly increasing on (0,+)(0,\ +\infty).

Proof.

A simple computation shows

φ1(r)=8(rcosrsinr)2(2rsin(2r))2>0,r(0,+){ϑk;k},\varphi_{1}^{\prime}(r)=8\ \frac{(r\cos{r}-\sin{r})^{2}}{(2r-\sin(2r))^{2}}>0,\quad\forall\,r\in(0,\ +\infty)\setminus\{\vartheta_{k};k\in{\mathbb{N}}^{*}\},

which implies the monotonicity of φ1\varphi_{1}. ∎

4.3 Proof of Theorem 2.10

Proof.

Let us begin with the proof of (2.36). Recall that the Assumption (A) (cf. (2.18)) says that 𝐰¯=θ2ψ(|θ|)>0\mathbf{\overline{w}}=\theta_{2}\,\psi(|\theta|)>0, 𝐬¯=(𝐬¯1,𝐬¯2)=(1,0)\mathbf{\overline{s}}=(\mathbf{\overline{s}}_{1},\mathbf{\overline{s}}_{2})=(-1,0), and u=(u1,u2,0)u=(u_{1},u_{2},0) with

u1=ψ(|θ|)|θ|θ22θ1>0,u2=θ2[ψ(|θ|)|θ|θ22+2ψ(|θ|)]=θ2K3(θ1,θ2)>0.u_{1}=\frac{\psi^{\prime}(|\theta|)}{|\theta|}\,\theta_{2}^{2}\,\theta_{1}>0,\qquad u_{2}=\theta_{2}\,\left[\frac{\psi^{\prime}(|\theta|)}{|\theta|}\,\theta_{2}^{2}+2\,\psi(|\theta|)\right]=\theta_{2}\,\mathrm{K}_{3}(\theta_{1},\theta_{2})>0.

It follows from Υ(r)=1/ψ(r)\Upsilon(r)=1/\psi(r) (0<rπ<ϑ10<r\neq\pi<\vartheta_{1}) that

Υ(|θ|)|θ|𝐰¯2|𝐬¯|2θ+4×14(u+2𝐰¯𝐬¯1e2+2𝐰¯𝐬¯2e3)=0.\displaystyle\frac{\Upsilon^{\prime}(|\theta|)}{|\theta|}\,\mathbf{\overline{w}}^{2}\,|\mathbf{\overline{s}}|^{2}\,\theta+4\times\frac{1}{4}\Big{(}u+2\,\mathbf{\overline{w}}\,\mathbf{\overline{s}}_{1}e_{2}+2\,\mathbf{\overline{w}}\,\mathbf{\overline{s}}_{2}e_{3}\Big{)}=0. (4.15)

Combining this with (4.13) and (4.12), we obtain that

𝒟(u;𝐬¯)=Υ(|θ|)𝐰¯2|𝐬¯|2+uθ+2𝐰¯𝐬¯1θ2=θ22ψ(|θ|)+uθ,\displaystyle\mathcal{D}(u;\mathbf{\overline{s}})=\Upsilon(|\theta|)\mathbf{\overline{w}}^{2}|\mathbf{\overline{s}}|^{2}+u\cdot\theta+2\mathbf{\overline{w}}\,\mathbf{\overline{s}}_{1}\theta_{2}=-\theta_{2}^{2}\,\psi(|\theta|)+u\cdot\theta,

which is the first equality in (2.36).

Next, observe that uθ=θ22(|θ|ψ(|θ|)+2ψ(|θ|))u\cdot\theta=\theta_{2}^{2}\,(|\theta|\,\psi^{\prime}(|\theta|)+2\,\psi(|\theta|)). Then to obtain the second equality in (2.36), it suffices to use the first one and

ψ(r)+rψ(r)=(rψ(r))=(1rcotrr)=1sin2r1r2.\psi(r)+r\psi^{\prime}(r)=(r\psi(r))^{\prime}=\left(\frac{1-r\cot{r}}{r}\right)^{\prime}=\frac{1}{\sin^{2}{r}}-\frac{1}{r^{2}}.

To show the third one, we make use of the second one and the following basic equality

φ1(r)=φ0(r)r2ψ(r)+2rψ(r),withφ0(r):=(rsinr)21.\varphi_{1}(r)=\frac{\varphi_{0}(r)}{r^{2}\psi^{\prime}(r)+2r\psi(r)},\quad\mbox{with}\quad\varphi_{0}(r):=\left(\frac{r}{\sin{r}}\right)^{2}-1.

For the fourth one, it suffices to use the second one, the fact that u1/θ1=ψ(|θ|)θ22/|θ|u_{1}/\theta_{1}=\psi^{\prime}(|\theta|)\,\theta_{2}^{2}/|\theta| and φ2(r)=φ0(r)/(r2ψ(r))\varphi_{2}(r)=\varphi_{0}(r)/(r^{2}\,\psi^{\prime}(r)). The last one follows from the fourth one, the fact that

u1(u1+u2θ2θ1)=u12|θ|2ψ(|θ|)+2|θ|ψ(|θ|)ψ(|θ|)θ12=(u1|θ|θ1)2φ0(|θ|)φ1(|θ|)|θ|2ψ(|θ|)=(u1|θ|θ1)2φ2(|θ|)φ1(|θ|)u_{1}\,(u_{1}+u_{2}\frac{\theta_{2}}{\theta_{1}})=u_{1}^{2}\frac{|\theta|^{2}\,\psi^{\prime}(|\theta|)+2\,|\theta|\,\psi(|\theta|)}{\psi^{\prime}(|\theta|)\,\theta_{1}^{2}}=\left(u_{1}\frac{|\theta|}{\theta_{1}}\right)^{2}\frac{\frac{\varphi_{0}(|\theta|)}{\varphi_{1}(|\theta|)}}{|\theta|^{2}\,\psi^{\prime}(|\theta|)}=\left(u_{1}\frac{|\theta|}{\theta_{1}}\right)^{2}\frac{\varphi_{2}(|\theta|)}{\varphi_{1}(|\theta|)}

and φ3=φ1φ2\varphi_{3}=\sqrt{\varphi_{1}\,\varphi_{2}}.

Turning to the first assertion. Fix uu as in Assumption (A). Notice that from Proposition 4.2 (iii)-(iv), the continuous function 𝒟(u;)\mathcal{D}(u;\cdot) must attain its minimum at some point s2s^{*}\in\mathbb{R}^{2}. Then we only need to show that s=𝐬¯s^{*}=\mathbf{\overline{s}}. This proceeds in three steps as follows:

Step 1: It holds that s=(s1,s2)0s^{*}=(s_{1}^{*},s_{2}^{*})\neq 0.

We argue by contradiction. If s=0s^{*}=0, it follows from (4.12) that 𝒟(u;0)=𝐃(0,u/4)2=ϑ1|u|\mathcal{D}(u;0)=\mathbf{D}(0,u/4)^{2}=\vartheta_{1}\,|u|. Next, recalling tanϑ1=ϑ1\tan{\vartheta_{1}}=\vartheta_{1}, a direct calculation yields that

φ1(ϑ1)=ϑ12sin2ϑ1ϑ1sinϑ1cosϑ1=tan2ϑ1sin2ϑ1tanϑ1sinϑ1cosϑ1=tanϑ1=ϑ1.\varphi_{1}(\vartheta_{1})=\frac{\vartheta_{1}^{2}-\sin^{2}{\vartheta_{1}}}{\vartheta_{1}-\sin{\vartheta_{1}}\,\cos{\vartheta_{1}}}=\frac{\tan^{2}{\vartheta_{1}}-\sin^{2}{\vartheta_{1}}}{\tan{\vartheta_{1}}-\sin{\vartheta_{1}}\,\cos{\vartheta_{1}}}=\tan{\vartheta_{1}}=\vartheta_{1}.

Thus the fact that φ1\varphi_{1} is strictly increasing on (0,+)(0,+\infty) (cf. Lemma 4.4) implies that 𝒟(u;0)=φ1(ϑ1)|u|>φ1(|θ|)|θ|uθ=𝒟(u;𝐬¯)\mathcal{D}(u;0)=\varphi_{1}(\vartheta_{1})\,|u|>\frac{\varphi_{1}(|\theta|)}{|\theta|}u\cdot\theta=\mathcal{D}(u;\mathbf{\overline{s}}). This leads to a contradiction.

It follows from (4.12) and (4.13) that 𝐃2\mathbf{D}^{2} is smooth on (2{0})×3(\mathbb{R}^{2}\setminus\{0\})\times\mathbb{R}^{3}. Then 𝒟(u;)C(2{0})\mathcal{D}(u;\cdot)\in C^{\infty}(\mathbb{R}^{2}\setminus\{0\}). Since its minimum point s0s^{*}\neq 0, using (4.12) again, we obtain

0=s1𝒟(u;s)=2𝐰¯(Υ(|τ|)𝐰¯s1+τ2),0=s2𝒟(u;s)=2𝐰¯(Υ(|τ|)𝐰¯s2+τ3),\displaystyle 0=\partial_{s_{1}}\mathcal{D}(u;s^{*})=2\,\mathbf{\overline{w}}\,(\Upsilon(|\tau_{*}|)\,\mathbf{\overline{w}}\,s^{*}_{1}+\tau_{*2}),\quad 0=\partial_{s_{2}}\mathcal{D}(u;s^{*})=2\,\mathbf{\overline{w}}\,(\Upsilon(|\tau_{*}|)\,\mathbf{\overline{w}}\,s^{*}_{2}+\tau_{*3}), (4.16)

with τ=(τ1,τ2,τ3)=τ(s)\tau_{*}=(\tau_{*1},\tau_{*2},\tau_{*3})=\tau(s^{*}), where the smooth mapping τ(s)=(τ1(s),τ2(s),τ3(s))\tau(s)=(\tau_{1}(s),\tau_{2}(s),\tau_{3}(s)) on 2{0}\mathbb{R}^{2}\setminus\{0\} is defined by

Υ(|τ(s)|)|τ(s)|𝐰¯2|s|2τ(s)+u+2𝐰¯s1e2+2𝐰¯s2e3=0.\displaystyle\frac{\Upsilon^{\prime}(|\tau(s)|)}{|\tau(s)|}\,\mathbf{\overline{w}}^{2}\,|s|^{2}\,\tau(s)+u+2\,\mathbf{\overline{w}}\,s_{1}e_{2}+2\,\mathbf{\overline{w}}\,s_{2}e_{3}=0. (4.17)

In particular,

Υ(|τ(s)|)|τ(s)|𝐰¯2|s|2τ3(s)+2𝐰¯s2=0.\displaystyle\frac{\Upsilon^{\prime}(|\tau(s)|)}{|\tau(s)|}\,\mathbf{\overline{w}}^{2}\,|s|^{2}\,\tau_{3}(s)+2\,\mathbf{\overline{w}}\,s_{2}=0. (4.18)
Step 2: It holds that |τ|π|\tau_{*}|\neq\pi.

We argue by contradiction again. Assume that |τ|=π|\tau_{*}|=\pi. Observing that Υ(|τ|)=Υ(π)=0\Upsilon(|\tau_{*}|)=\Upsilon(\pi)=0, then (4.16) implies that τ2=τ3=0\tau_{*2}=\tau_{*3}=0. Inserting this into (4.17) with s=ss=s^{*} and noticing Υ(π)=π\Upsilon^{\prime}(\pi)=-\pi (which can be checked by (3.11) and (3.23) for example), we get that s2=0s^{*}_{2}=0 and πu22=4u1\pi u_{2}^{2}=4u_{1}, which contradicts with our Assumption (A) (cf. (2.18)).

Step 3: We have τ=θ\tau_{*}=\theta and s=𝐬¯s^{*}=\mathbf{\overline{s}}.

Since s0s^{*}\neq 0 and 0<|τ|π<ϑ10<|\tau_{*}|\neq\pi<\vartheta_{1}, it follows from (4.16) and Υ=1/ψ\Upsilon=1/\psi that

𝐰¯s1=ψ(|τ|)τ2,𝐰¯s2=ψ(|τ|)τ3,𝐰¯2|s|2=ψ(|τ|)2((τ2)2+(τ3)2).\displaystyle\mathbf{\overline{w}}\,s^{*}_{1}=-\psi(|\tau_{*}|)\,\tau_{*2},\quad\mathbf{\overline{w}}\,s^{*}_{2}=-\psi(|\tau_{*}|)\,\tau_{*3},\quad\mathbf{\overline{w}}^{2}\,|s^{*}|^{2}=\psi(|\tau_{*}|)^{2}\,\Big{(}(\tau_{*2})^{2}+(\tau_{*3})^{2}\Big{)}. (4.19)

Substituting this into (4.17) with s=ss=s^{*}, together with Υ=ψ/ψ2\Upsilon^{\prime}=-\psi^{\prime}/\psi^{2}, we can write

u=ψ(|τ|)|τ|((τ2)2+(τ3)2)τ+2(τ2e2+τ3e3)ψ(|τ|).\displaystyle u=\frac{\psi^{\prime}(|\tau_{*}|)}{|\tau_{*}|}\Big{(}(\tau_{*2})^{2}+(\tau_{*3})^{2}\Big{)}\,\tau_{*}+2\,(\tau_{*2}\ e_{2}+\tau_{*3}\ e_{3})\,\psi(|\tau_{*}|).

The equation together with the fact that ψ(r)/r>0\psi^{\prime}(r)/r>0 for 0<rπ<ϑ10<r\neq\pi<\vartheta_{1} (cf. (3.13)), we get that τ1>0\tau_{*1}>0, τ20\tau_{*2}\neq 0, τ3=0\tau_{*3}=0 (otherwise we would yield from u3=0u_{3}=0 that ψ(|τ|)|τ|((τ2)2+(τ3)2)+2ψ(|τ|)=0\frac{\psi^{\prime}(|\tau_{*}|)}{|\tau_{*}|}\Big{(}(\tau_{*2})^{2}+(\tau_{*3})^{2}\Big{)}+2\,\psi(|\tau_{*}|)=0, which leads to u2=0u_{2}=0, a contradiction!), and Λ(τ1,τ2)=(u1,u2)\Lambda(\tau_{*1},\tau_{*2})=(u_{1},u_{2}). In particular, recalling the definition of K3\mathrm{K}_{3} (cf. (2.14)), we have u2=τ2K3(τ1,τ2)>0u_{2}=\tau_{*2}\,\mathrm{K}_{3}(\tau_{*1},\tau_{*2})>0.

At this point, (4.19) leads to s1=τ2𝐰¯ψ(|τ|)s_{1}^{*}=-\frac{\tau_{*2}}{\mathbf{\overline{w}}}\psi(|\tau_{*}|), s2=0s_{2}^{*}=0 and 𝐰¯2|s|2=ψ(|τ|)2τ22\mathbf{\overline{w}}^{2}|s^{*}|^{2}=\psi(|\tau_{*}|)^{2}\tau_{*2}^{2}. Taking gradient w.r.t. ss on both sides of (4.18), we obtain sτ3(s)=(0,2𝐰¯|s|2|τ|Υ(|τ|))\nabla_{s^{*}}\tau_{3}(s^{*})=(0,-\frac{2}{\mathbf{\overline{w}}|s^{*}|^{2}}\frac{|\tau_{*}|}{\Upsilon^{\prime}(|\tau_{*}|)}). Then using the expression of s2𝒟(u;s)\partial_{s_{2}}\mathcal{D}(u;s^{*}) (cf. (4.16)), a direct computation gives s1s22𝒟(u;s)=0\partial_{s_{1}s_{2}}^{2}\mathcal{D}(u;s^{*})=0 and

s22𝒟(u;s)=2𝐰¯2Υ(|τ|)+2𝐰¯s2τ3(s)=2𝐰¯2|τ|(τ2)2ψ(|τ|)ψ(|τ|)K3(τ1,τ2),\displaystyle\partial_{s_{2}}^{2}\mathcal{D}(u;s^{*})=2\,\mathbf{\overline{w}}^{2}\,\Upsilon(|\tau_{*}|)+2\,\mathbf{\overline{w}}\,\partial_{s_{2}}\tau_{3}(s^{*})=\frac{2\,\mathbf{\overline{w}}^{2}\,|\tau_{*}|}{(\tau_{*2})^{2}\,\psi(|\tau_{*}|)\,\psi^{\prime}(|\tau_{*}|)}\,\mathrm{K}_{3}(\tau_{*1},\tau_{*2}),

which is non-negative, by the fact that the Hessian matrix at a minimum point is positive semi-definite.

To summarize, we get that

τ=(τ1,τ2,0),0<|τ|π<ϑ1,τ1>0,τ2K3(τ1,τ2)>0,\displaystyle\tau_{*}=(\tau_{*1},\tau_{*2},0),\quad 0<|\tau_{*}|\neq\pi<\vartheta_{1},\quad\tau_{*1}>0,\quad\tau_{*2}\,\mathrm{K}_{3}(\tau_{*1},\tau_{*2})>0,
ψ(|τ|)K3(τ1,τ2)>0,Λ(τ1,τ2)=(u1,u2).\displaystyle\psi(|\tau_{*}|)\,\mathrm{K}_{3}(\tau_{*1},\tau_{*2})>0,\quad\Lambda(\tau_{*1},\tau_{*2})=(u_{1},u_{2}).

Using the characterization of Ω,4\Omega_{-,4} (cf. (3.29)), one can easily verify that (τ1,τ2)Ω+,1Ω,4(\tau_{*1},\tau_{*2})\in\Omega_{+,1}\cup\Omega_{-,4}. Then by Theorem 2.1 we see that τ=θ\tau_{*}=\theta. Therefore, s1=τ2𝐰¯ψ(|τ|)=1s_{1}^{*}=-\frac{\tau_{*2}}{\mathbf{\overline{w}}}\psi(|\tau_{*}|)=-1 and s2=0s_{2}^{*}=0, namely, s=𝐬¯s^{*}=\mathbf{\overline{s}}, which finishes the proof of Theorem 2.10. ∎

Recall that 𝐰¯=θ2ψ(|θ|)\mathbf{\overline{w}}=\theta_{2}\,\psi(|\theta|) and u2=θ2K3(θ1,θ2)u_{2}=\theta_{2}\,\mathrm{K}_{3}(\theta_{1},\theta_{2}). The above proof also gives the following:

Corollary 4.5.

Let θ,u\theta,u be as in Assumption (A) (cf. (2.18)). Then

s1s22𝒟(u;𝐬¯)=0,ands22𝒟(u;𝐬¯)=2ψ(|θ|)|θ|ψ(|θ|)K3(θ1,θ2)=2u2𝐰¯|θ|θ22ψ(|θ|)>0.\displaystyle\partial_{s_{1}s_{2}}^{2}\mathcal{D}(u;\mathbf{\overline{s}})=0,\ \mbox{and}\quad\partial_{s_{2}}^{2}\mathcal{D}(u;\mathbf{\overline{s}})=2\,\frac{\psi(|\theta|)\,|\theta|}{\psi^{\prime}(|\theta|)}\,\mathrm{K}_{3}(\theta_{1},\theta_{2})=2\,\frac{u_{2}\,\mathbf{\overline{w}}\,|\theta|}{\theta_{2}^{2}\,\psi^{\prime}(|\theta|)}>0. (4.20)

5 The complete answer to the Gaveau–Brockett problem on N3,2N_{3,2}

Now we are in a position to determine the expression for the sub-Riemannian distance from the origin to any given point. As indicated in Introduction and Remark 2.4, the main result (i.e., Theorem 2.2) has been proved in [33, 38], via two different methods. Here our third method use only the heat kernel. Although we can apply our uniform heat kernel asymptotic at infinity, namely Theorem 8.1 below (combining with Remark 8.3), to obtain Theorem 2.2, it turns out that a simpler argument in the proof below works as well. Indeed the argument is closely related to one of basic ideas in [33, 37]. Also note that at least in this very broad framework of step-two groups, this method looks much more practical and elementary than the classical one.

5.1 Proof of Theorem 2.2

Proof.

Combining the scaling property (2.2) with (2.30), we obtain that

ph(gu)=𝐂𝐰¯24πh112e14h2𝐏(s𝐰¯h,14h(u+2𝐰¯s1e2+2𝐰¯s2e3))𝑑s.p_{h}(g_{u})=\frac{\mathbf{C}\,\mathbf{\overline{w}}^{2}}{4\pi h^{\frac{11}{2}}}e^{-\frac{1}{4h}}\int_{\mathbb{R}^{2}}\mathbf{P}\left(\frac{s\,\mathbf{\overline{w}}}{\sqrt{h}},\frac{1}{4h}(u+2\mathbf{\overline{w}}\,s_{1}\,e_{2}+2\mathbf{\overline{w}}\,s_{2}\,e_{3})\right)\,ds.

Then it follows from (ii) of Proposition 2.9 and the second equation of (4.14) that

ph(gu)𝐰¯2h42e𝒟(u;s)+14h(h+𝒟(u;s))(h2+𝐰¯2|s|2𝒟(u;s))14𝑑s.p_{h}(g_{u})\sim\frac{\mathbf{\overline{w}}^{2}}{h^{4}}\int_{\mathbb{R}^{2}}\frac{e^{-\frac{\mathcal{D}(u;s)+1}{4h}}}{(h+\mathcal{D}(u;s))(h^{2}+\mathbf{\overline{w}}^{2}\,|s|^{2}\,\mathcal{D}(u;s))^{\frac{1}{4}}}\,ds. (5.1)

Recall that 𝒟(u;s)max{𝒟(u;𝐬¯),c(𝐰¯2|s|2+u1)}\mathcal{D}(u;s)\geq\max\{\mathcal{D}(u;\mathbf{\overline{s}}),c\,(\mathbf{\overline{w}}^{2}|s|^{2}+u_{1})\} for some constant c>0c>0 (cf. Theorem 2.10 and the first equation in (4.14)). Then

ph(gu)𝐰¯h4e𝒟(u;𝐬¯)+14h2ds(𝐰¯2|s|2+u1)|s|u112h4e𝒟(u;𝐬¯)+14h.p_{h}(g_{u})\lesssim\frac{\mathbf{\overline{w}}}{h^{4}}e^{-\frac{\mathcal{D}(u;\mathbf{\overline{s}})+1}{4h}}\int_{\mathbb{R}^{2}}\frac{ds}{(\mathbf{\overline{w}}^{2}|s|^{2}+u_{1})|s|}\lesssim{u}^{-\frac{1}{2}}_{1}\,h^{-4}\,e^{-\frac{\mathcal{D}(u;\mathbf{\overline{s}})+1}{4h}}.

On the other hand, for any given 1>ς0>01>\varsigma_{0}>0, by the continuity of 𝒟(u;)\mathcal{D}(u;\cdot), there exists a ball Bu,ς0B_{u,\varsigma_{0}} with center 𝐬¯\mathbf{\overline{s}} and radius <1<1 where 𝒟(u;s)𝒟(u;𝐬¯)+ς0\mathcal{D}(u;s)\leq\mathcal{D}(u;\mathbf{\overline{s}})+\varsigma_{0}. Hence

ph(gu)u,ς0𝐰¯2h4e𝒟(u;𝐬¯)+ς0+14h(h+𝒟(u;𝐬¯)+1)1[h2+4𝐰¯2(𝒟(u;𝐬¯)+1)]14.p_{h}(g_{u})\gtrsim_{u,\varsigma_{0}}\frac{\mathbf{\overline{w}}^{2}}{h^{4}}e^{-\frac{\mathcal{D}(u;\mathbf{\overline{s}})+\varsigma_{0}+1}{4h}}(h+\mathcal{D}(u;\mathbf{\overline{s}})+1)^{-1}\left[h^{2}+4\mathbf{\overline{w}}^{2}\,(\mathcal{D}(u;\mathbf{\overline{s}})+1)\right]^{-\frac{1}{4}}.

It follows from Varadhan’s formulas that 𝒟(u;𝐬¯)+1d(gu)2𝒟(u;𝐬¯)+ς0+1\mathcal{D}(u;\mathbf{\overline{s}})+1\leq d(g_{u})^{2}\leq\mathcal{D}(u;\mathbf{\overline{s}})+\varsigma_{0}+1, which implies immediately the desired result. ∎

Up to a slight modification, the proof below is extracted from [33, 38].

5.2 Proof of Corollary 2.3

Proof.

We first show (ii). Fix an α>0\alpha>0 and let gα=(e1,41(α2π,2πα,0))g_{\alpha}=(e_{1},4^{-1}(\frac{\alpha^{2}}{\pi},\frac{2}{\pi}\alpha,0)). From the proof of Theorem 2.1, we can pick a sequence {v(j)=(v1(j),v2(j))}Ω+,1\{v^{(j)}=(v_{1}^{(j)},v_{2}^{(j)})\}\subseteq\Omega_{+,1} such that v(j)(π,0)Ω+,1v^{(j)}\to(\pi,0)\in\partial\Omega_{+,1} and

v2(j)π|v(j)|α,u(j)=Λ(v(j))(α2π,2πα),asj+.\frac{v_{2}^{(j)}}{\pi-|v^{(j)}|}\to\alpha,\quad u^{(j)}=\Lambda(v^{(j)})\to\left(\frac{\alpha^{2}}{\pi},\frac{2}{\pi}\alpha\right),\quad\mbox{as}\,\,j\to+\infty.

Hence gu(j)gαg_{u^{(j)}}\to g_{\alpha} as j+j\to+\infty. Since d()2d(\cdot)^{2} is continuous, it yields from Theorem 2.2 that

d(gα)2=limj+d(gu(j))2=φ1(π)α2π×ππ+1=1+α2.d(g_{\alpha})^{2}=\lim_{j\to+\infty}d(g_{u^{(j)}})^{2}=\varphi_{1}(\pi)\,\frac{\alpha^{2}}{\pi}\times\frac{\pi}{\pi}+1=1+\alpha^{2}.

To show (iii), as before we first fix a β>0\beta>0. Let 0<κ<2πβ0<\kappa<\frac{2}{\sqrt{\pi}}\sqrt{\beta}, u(κ):=(β,κ)u(\kappa):=(\beta,\kappa), and v(κ)=(v1(κ),v2(κ)):=Λ1(uκ)v(\kappa)=(v_{1}(\kappa),v_{2}(\kappa)):=\Lambda^{-1}(u_{\kappa}). Since {u(κ)}<,+2\{u(\kappa)\}\subseteq\mathbb{R}^{2}_{<,+}, by Theorem 2.1 (2) and (2.17) we have {v(κ)}Ω,4\{v(\kappa)\}\subseteq\Omega_{-,4} and

v2(κ)<0<π<v1(κ)<|v(κ)|<ϑ1,\displaystyle v_{2}(\kappa)<0<\pi<v_{1}(\kappa)<|v(\kappa)|<\vartheta_{1},
β=ψ(|v(κ)|)|v(κ)|v1(κ)v22(κ),κ=v2(κ)(ψ(|v(κ)|)|v(κ)|v22(κ)+2ψ(|v(κ)|)).\displaystyle\beta=\frac{\psi^{\prime}(|v(\kappa)|)}{|v(\kappa)|}\,v_{1}(\kappa)\,v^{2}_{2}(\kappa),\qquad\kappa=v_{2}(\kappa)\left(\frac{\psi^{\prime}(|v(\kappa)|)}{|v(\kappa)|}\,v^{2}_{2}(\kappa)+2\,\psi(|v(\kappa)|)\right). (5.2)

By the compactness of B2(0,ϑ1)¯\overline{B_{\mathbb{R}^{2}}(0,\vartheta_{1})}, up to subsequences, we may take κj0+\kappa_{j}\to 0^{+} as j+j\to+\infty such that the corresponding v(κj)v(0):=(v1(0),v2(0))v(\kappa_{j})\to v(0):=(v_{1}(0),v_{2}(0)). Obviously, v2(0)0v_{2}(0)\leq 0 and πv1(0)r:=|v(0)|ϑ1\pi\leq v_{1}(0)\leq r:=|v(0)|\leq\vartheta_{1}.

We claim that v2(0)0v_{2}(0)\neq 0, so v(0){(π,0),(ϑ1,0)}v(0)\notin\{(\pi,0),(\vartheta_{1},0)\} and π<r<ϑ1\pi<r<\vartheta_{1} (by (5.2) and the fact that ψ(ϑ1)=0,ψ(ϑ1)>0\psi(\vartheta_{1})=0,\,\psi^{\prime}(\vartheta_{1})>0). Were this not the case, it would follow that v2(κj)0v_{2}(\kappa_{j})\to 0^{-}. Then the first equation in (5.2) implies that

limj+ψ(|v(κj)|)|v(κj)|=+,so |v(κj)|π+and v1(κj)π+,\lim_{j\to+\infty}\frac{\psi^{\prime}(|v(\kappa_{j})|)}{|v(\kappa_{j})|}=+\infty,\quad\mbox{so }\ |v(\kappa_{j})|\to\pi^{+}\ \mbox{and }\ v_{1}(\kappa_{j})\to\pi^{+},

since π<v1(κj)<|v(κj)|<ϑ1\pi<v_{1}(\kappa_{j})<|v(\kappa_{j})|<\vartheta_{1} and ψ(ρ)+\psi^{\prime}(\rho)\to+\infty (π<ρ<ϑ1\pi<\rho<\vartheta_{1}) only if ρπ+\rho\to\pi^{+} (cf. (3.6)). Notice that the limit (3.23) remains valid as rπ+r\to\pi^{+}, then using (5.2) again we infer that

limj+1π(v2(κj)|v(κj)|π)2=β,0=2πlimj+v2(κj)|v(κj)|π=2πβ>0.\displaystyle\lim_{j\to+\infty}\frac{1}{\pi}\left(\frac{v_{2}(\kappa_{j})}{|v(\kappa_{j})|-\pi}\right)^{2}=\beta,\quad 0=-\frac{2}{\pi}\lim_{j\to+\infty}\frac{v_{2}(\kappa_{j})}{|v(\kappa_{j})|-\pi}=\frac{2}{\sqrt{\pi}}\sqrt{\beta}>0.

This leads to a contradiction.

In conclusion, by the continuity of d()2d(\cdot)^{2} and the last equality in (2.24), we obtain that d(g(β))2=φ3(r)β+1d(g(\beta))^{2}=\varphi_{3}(r)\,\beta+1, where π<r<ϑ1\pi<r<\vartheta_{1} satisfies

β=ψ(r)rv1(0)v2(0)2,ψ(r)rv2(0)2+2ψ(r)=0.\displaystyle\beta=\frac{\psi^{\prime}(r)}{r}\,v_{1}(0)\,v_{2}(0)^{2},\qquad\frac{\psi^{\prime}(r)}{r}\,v_{2}(0)^{2}+2\,\psi(r)=0. (5.3)

That is

β=2ψ(r)r2v2(0)2=2ψ(r)r2+2rψ(r)ψ(r),\displaystyle\beta=-2\,\psi(r)\sqrt{r^{2}-v_{2}(0)^{2}}=-2\,\psi(r)\,\sqrt{r^{2}+2\,r\,\frac{\psi(r)}{\psi^{\prime}(r)}},

which, together with Remark 2.4 (1), implies (iii).

Noticing that ρ2ψ(ρ)+2ρψ(ρ)=(ρ2ψ(ρ))=μ(ρ)\rho^{2}\,\psi^{\prime}(\rho)+2\rho\,\psi(\rho)=(\rho^{2}\,\psi(\rho))^{\prime}=\mu(\rho), then a similar and simpler argument gives (iv).

Finally we are left with the proof of (i). Applying the continuity of d(g)2d(g)^{2} again with (2.21), it follows from the scaling property (cf. (1.1)) and (iv) that

d(g)2=d(0,e2)2=limε0+d(εe1,e2)2=limε0+ε2d(g(4/ε2))2=limε0+(r(ε)sinr(ε))2ε2,d(g_{*})^{2}=d(0,e_{2})^{2}=\lim_{\varepsilon\to 0^{+}}d(\varepsilon e_{1},e_{2})^{2}=\lim_{\varepsilon\to 0^{+}}\varepsilon^{2}\,d(g(4/\varepsilon^{2})^{*})^{2}=\lim_{\varepsilon\to 0^{+}}\left(\frac{r(\varepsilon)}{\sin{r(\varepsilon)}}\right)^{2}\varepsilon^{2},

where r(ε)r(\varepsilon) is the unique solution of μ(r(ε))=4/ε2\mu(r(\varepsilon))=4/\varepsilon^{2}. From this and the facts that

limρπ(ρπ)2μ(ρ)=π,limρπ(ρπ)2(ρsinρ)2=π2,\lim_{\rho\to\pi^{-}}(\rho-\pi)^{2}\mu(\rho)=\pi,\qquad\lim_{\rho\to\pi^{-}}(\rho-\pi)^{2}\left(\frac{\rho}{\sin{\rho}}\right)^{2}=\pi^{2},

the assertion (i) follows easily. ∎

6 Uniform asymptotics for the simplest case: |θ|3|\theta|\leq 3 and θ2|x|+\theta_{2}|x|\to+\infty

Recall that d(g)2=ϕ(g;θ)d(g)^{2}=\phi(g;\theta). In this section we establish the following theorem:

Theorem 6.1.

Let |θ|3|\theta|\leq 3 and θ2|x|ζ0\theta_{2}\,|x|\geq\zeta_{0} with ζ01\zeta_{0}\gg 1. Then

p(g)=(8π)32ed(g)24𝐕(iθ)1det(Hessθϕ(g;θ))(1+oζ0(1)).p(g)=(8\pi)^{\frac{3}{2}}\,e^{-\frac{d(g)^{2}}{4}}\,\mathbf{V}(i\theta)\,\frac{1}{\sqrt{\det(-\mathrm{Hess}_{\theta}\,\phi(g;\theta))}}\,(1+o_{\zeta_{0}}(1)). (6.1)

Essentially, Theorem 6.1 is a special case of [33, Theorem 2.2] (up to a slight modification), which is based on the method of stationary phase and the operator convexity. For the reader’s convenience, we provide a direct proof. Let us begin with the following crucial lemma:

Lemma 6.2.

Let |θ|3|\theta|\leq 3. Then there exists a constant c>0c>0 such that

(ϕ(g;θiλ)ϕ(g;θ))cλTHessθϕ(g;θ)λ1+|λ|2,λ3.\Re(\phi(g;\theta-i\lambda)-\phi(g;\theta))\geq c\,\frac{-\lambda^{\mathrm{T}}\,\mathrm{Hess}_{\theta}\,\phi(g;\theta)\lambda}{1+|\lambda|^{2}},\quad\forall\,\lambda\in\mathbb{R}^{3}.
Proof.

In fact, from the definition of the reference function (cf. (2.20)) and (3.1), we have

ϕ(g;θiλ)=|x|22j=1+(θ2iλ2)2+(iλ3)2j2π2(|θ|22iλθ|λ|2)|x|2+4t(θiλ).\phi(g;\theta-i\lambda)=|x|^{2}-2\sum_{j=1}^{+\infty}\frac{(\theta_{2}-i\lambda_{2})^{2}+(i\lambda_{3})^{2}}{j^{2}\pi^{2}-(|\theta|^{2}-2i\lambda\cdot\theta-|\lambda|^{2})}|x|^{2}+4t\cdot(\theta-i\lambda).

Hence (ϕ(g;θiλ)ϕ(g;θ))/|x|2\Re(\phi(g;\theta-i\lambda)-\phi(g;\theta))/|x|^{2} equals

2j=1+[λ22+λ32θ22+2iλ2θ2j2π2|θ|2+|λ|2+2iλθ+θ22j2π2|θ|2]\displaystyle 2\sum_{j=1}^{+\infty}\Re\left[\frac{\lambda_{2}^{2}+\lambda_{3}^{2}-\theta_{2}^{2}+2i\lambda_{2}\theta_{2}}{j^{2}\pi^{2}-|\theta|^{2}+|\lambda|^{2}+2i\lambda\cdot\theta}+\frac{\theta_{2}^{2}}{j^{2}\pi^{2}-|\theta|^{2}}\right]
=2j=1+[(λ22+λ32θ22+2iλ2θ2)(j2π2|θ|2)+θ22(j2π2|θ|2+|λ|2+2iλθ)(j2π2|θ|2+|λ|2+2iλθ)(j2π2|θ|2)]\displaystyle\qquad=2\sum_{j=1}^{+\infty}\Re\left[\frac{(\lambda_{2}^{2}+\lambda_{3}^{2}-\theta_{2}^{2}+2i\lambda_{2}\theta_{2})(j^{2}\pi^{2}-|\theta|^{2})+\theta_{2}^{2}(j^{2}\pi^{2}-|\theta|^{2}+|\lambda|^{2}+2i\lambda\cdot\theta)}{(j^{2}\pi^{2}-|\theta|^{2}+|\lambda|^{2}+2i\lambda\cdot\theta)(j^{2}\pi^{2}-|\theta|^{2})}\right]
=2j=1+{θ22(j2π2|θ|2+|λ|2)|λ|2+λ22|λ|2(j2π2|θ|2)[(j2π2|θ|2+|λ|2)2+4(λθ)2](j2π2|θ|2)\displaystyle\qquad=2\sum_{j=1}^{+\infty}\left\{\frac{\theta_{2}^{2}(j^{2}\pi^{2}-|\theta|^{2}+|\lambda|^{2})\,|\lambda|^{2}+\lambda_{2}^{2}\,|\lambda|^{2}\,(j^{2}\pi^{2}-|\theta|^{2})}{[(j^{2}\pi^{2}-|\theta|^{2}+|\lambda|^{2})^{2}+4(\lambda\cdot\theta)^{2}](j^{2}\pi^{2}-|\theta|^{2})}\right.
+λ32(j2π2|θ|2)(j2π2|θ|2+|λ|2)+[λ2(j2π2|θ|2)+2θ2(λθ)]2[(j2π2|θ|2+|λ|2)2+4(λθ)2](j2π2|θ|2)}.\displaystyle\qquad\qquad+\left.\frac{\lambda_{3}^{2}(j^{2}\pi^{2}-|\theta|^{2})(j^{2}\pi^{2}-|\theta|^{2}+|\lambda|^{2})+[\lambda_{2}(j^{2}\pi^{2}-|\theta|^{2})+2\theta_{2}(\lambda\cdot\theta)]^{2}}{[(j^{2}\pi^{2}-|\theta|^{2}+|\lambda|^{2})^{2}+4(\lambda\cdot\theta)^{2}](j^{2}\pi^{2}-|\theta|^{2})}\right\}.

Notice that every term in the series is non-negative, and thus its value is greater than the single term j0=2j_{0}=2. Thus,

(ϕ(g;θiλ)ϕ(g;θ))\displaystyle\Re(\phi(g;\theta-i\lambda)-\phi(g;\theta))
θ22(1+|λ|2)|λ|2+λ22|λ|2+λ32(1+|λ|2)+[λ2(j02π2|θ|2)+2θ2(λθ)]2(1+|λ|2)2|x|2.\displaystyle\qquad\gtrsim\frac{\theta_{2}^{2}(1+|\lambda|^{2})|\lambda|^{2}+\lambda_{2}^{2}|\lambda|^{2}+\lambda_{3}^{2}(1+|\lambda|^{2})+[\lambda_{2}(j_{0}^{2}\pi^{2}-|\theta|^{2})+2\theta_{2}(\lambda\cdot\theta)]^{2}}{(1+|\lambda|^{2})^{2}}\,|x|^{2}.

Next, it follows from j02π2|θ|21j_{0}^{2}\pi^{2}-|\theta|^{2}\sim 1 and the following elementary inequality

(a+b)2a21+cb2c,a,b,c>0,(a+b)^{2}\geq\frac{a^{2}}{1+c}-\frac{b^{2}}{c},\qquad a,b\in\mathbb{R},\ c>0,

that θ22(1+|λ|2)|λ|2+[λ2(j02π2|θ|2)+2θ2(λθ)]2θ22(1+|λ|2)|λ|2+λ22\theta_{2}^{2}(1+|\lambda|^{2})|\lambda|^{2}+[\lambda_{2}(j_{0}^{2}\pi^{2}-|\theta|^{2})+2\theta_{2}(\lambda\cdot\theta)]^{2}\gtrsim\theta_{2}^{2}(1+|\lambda|^{2})|\lambda|^{2}+\lambda_{2}^{2}. Consequently,

(ϕ(g;θiλ)ϕ(g;θ))θ22λ12+(λ22+λ32)1+|λ|2|x|2.\displaystyle\Re(\phi(g;\theta-i\lambda)-\phi(g;\theta))\gtrsim\frac{\theta_{2}^{2}\lambda_{1}^{2}+(\lambda_{2}^{2}+\lambda_{3}^{2})}{1+|\lambda|^{2}}\,|x|^{2}.

Therefore, to finish the proof of Lemma 6.2, it suffices to show that for all |θ|3|\theta|\leq 3,

Hessθϕ(g;θ)(θ22|x|2000|x|2000|x|2).\displaystyle-\mathrm{Hess}_{\theta}\,\phi(g;\theta)\sim\begin{pmatrix}\theta_{2}^{2}|x|^{2}&0&0\\ 0&|x|^{2}&0\\ 0&0&|x|^{2}\\ \end{pmatrix}. (6.2)

In fact, by (3.1) and (3.13), K3(θ1,θ2)1\mathrm{K}_{3}(\theta_{1},\theta_{2})\sim 1. Then from this, (3.24), Lemma 3.3, and the fact that Hessθϕ(g;θ)=|x|2Hessθϕ(gu;θ)\mathrm{Hess}_{\theta}\,\phi(g;\theta)=|x|^{2}\,\mathrm{Hess}_{\theta}\,\phi(g_{u};\theta), the estimate (6.2) follows. ∎

Next, using the trivial inequality

|1+(iτ+λ)(iτ+λ)|(1|τ|2)(1+|λ|2),τ,λ3with|τ|<1,\big{|}1+(i\tau+\lambda)\cdot(i\tau+\lambda)\big{|}\geq(1-|\tau|^{2})\,(1+|\lambda|^{2}),\quad\tau,\lambda\in\mathbb{R}^{3}\ \mbox{with}\ |\tau|<1,

it follows from (2.4) that

Lemma 6.3.

It holds that

|𝐕(λ+iθ)|𝐕(iθ)𝐕(λ)=|θ|sin|θ|𝐕(λ)|θ|sin|θ|,|θ|<π,λ3.\displaystyle\big{|}\mathbf{V}(\lambda+i\theta)\big{|}\leq\mathbf{V}(i\theta)\,\mathbf{V}(\lambda)=\frac{|\theta|}{\sin{|\theta|}}\,\mathbf{V}(\lambda)\leq\frac{|\theta|}{\sin{|\theta|}},\quad|\theta|<\pi,\ \lambda\in\mathbb{R}^{3}. (6.3)

We are in a position to give the

6.1 Proof of Theorem 6.1

Proof.

Under our assumptions, first it deduces from (6.2) and (2.17) that θ\theta is a nondegenerate critical point of ϕ(g;)\phi(g;\cdot) in B3(0,π)B_{\mathbb{R}^{3}}(0,\pi), i.e. the open ball in 3\mathbb{R}^{3} with the center 0 and the radius π\pi.

Next, notice that the integrand in (2.19) is holomorphic on 3+iB3(0,3)¯={λ+iτ;λ3,τB3(0,3)¯}3\mathbb{R}^{3}+i\,\overline{B_{\mathbb{R}^{3}}(0,3)}=\{\lambda+i\tau;\,\lambda\in\mathbb{R}^{3},\ \tau\in\overline{B_{\mathbb{R}^{3}}(0,3)}\}\subset\mathbb{C}^{3}, and decays exponentially from Lemma 6.2 and (6.3). Hence we can deform the contour from 3\mathbb{R}^{3} to 3+iθ\mathbb{R}^{3}+i\theta in (2.19), and get

p(g)=ed(g)243𝐕(λ+iθ)exp{14(ϕ(g;θiλ)ϕ(g;θ))}𝑑λ.\displaystyle p(g)=e^{-\frac{d(g)^{2}}{4}}\int_{\mathbb{R}^{3}}\mathbf{V}(\lambda+i\theta)\,\exp\left\{-\frac{1}{4}\Big{(}\phi(g;\theta-i\lambda)-\phi(g;\theta)\Big{)}\right\}\,d\lambda. (6.4)

Since θ2|x|1\theta_{2}\,|x|\gg 1 (so |x|1|x|\gg 1), we split 3{\mathbb{R}}^{3} into the following three regions:

1:={λ3;|λ1|(θ2|x|)34,|λ|(θ2|x|)14|x|1},2:={λ3;|λ|1}1,and3:={λ3;|λ|>1}.\begin{gathered}\diamondsuit_{1}:=\{\lambda\in\mathbb{R}^{3};\,|\lambda_{1}|\leq(\theta_{2}|x|)^{-\frac{3}{4}},\,|\lambda^{\prime}|\leq(\theta_{2}|x|)^{\frac{1}{4}}\,|x|^{-1}\},\\[2.84526pt] \diamondsuit_{2}:=\{\lambda\in\mathbb{R}^{3};\,|\lambda|\leq 1\}\setminus\diamondsuit_{1},\quad\mbox{and}\quad\diamondsuit_{3}:=\{\lambda\in\mathbb{R}^{3};\,|\lambda|>1\}.\end{gathered}

Then p(g)=eϕ(g;θ)4l=13𝐉lp(g)=e^{-\frac{\phi(g;\theta)}{4}}\sum_{l=1}^{3}\mathbf{J}_{l}, where

𝐉l:=l𝐕(λ+iθ)exp{14(ϕ(g;θiλ)ϕ(g;θ))}𝑑λ,l=1,2,3.\displaystyle\mathbf{J}_{l}:=\int_{\diamondsuit_{l}}\mathbf{V}(\lambda+i\theta)\,\exp\left\{-\frac{1}{4}\Big{(}\phi(g;\theta-i\lambda)-\phi(g;\theta)\Big{)}\right\}\,d\lambda,\qquad l=1,2,3.

We begin with the estimate of the leading term 𝐉1\mathbf{J}_{1}. It can be checked that

𝐕(λ+iθ)=𝐕(iθ)(1+O((θ2|x|)34)),\displaystyle\mathbf{V}(\lambda+i\theta)=\mathbf{V}(i\theta)\,\big{(}1+O((\theta_{2}|x|)^{-\frac{3}{4}})\big{)},
ϕ(g;θiλ)ϕ(g;θ)=12λTHessθϕ(g;θ)λ+O(|λ||x|2(|λ|2+θ22|λ|2))=12λTHessθϕ(g;θ)λ+O((θ2|x|)14),\displaystyle\begin{aligned} \phi(g;\theta-i\lambda)-\phi(g;\theta)&=-\frac{1}{2}\lambda^{\mathrm{T}}\mathrm{Hess}_{\theta}\,\phi(g;\theta)\lambda+O\big{(}|\lambda|\,|x|^{2}(|\lambda^{\prime}|^{2}+\theta_{2}^{2}\,|\lambda|^{2})\big{)}\\ &=-\frac{1}{2}\lambda^{\mathrm{T}}\mathrm{Hess}_{\theta}\,\phi(g;\theta)\lambda+O((\theta_{2}|x|)^{-\frac{1}{4}}),\end{aligned}

for all λ1\lambda\in\diamondsuit_{1}. As a result, the standard Laplace’s method shows that:

𝐉1\displaystyle\mathbf{J}_{1} =𝐕(iθ)1e18λTHessθϕ(g;θ)λ𝑑λ(1+O((θ2|x|)14)),\displaystyle=\mathbf{V}(i\theta)\int_{\diamondsuit_{1}}e^{\frac{1}{8}\lambda^{\mathrm{T}}\mathrm{Hess}_{\theta}\,\phi(g;\theta)\lambda}\,d\lambda\,(1+O((\theta_{2}|x|)^{-\frac{1}{4}})),
=𝐕(iθ)(3(1)c)(1+O((θ2|x|)14)),\displaystyle=\mathbf{V}(i\theta)\left(\int_{\mathbb{R}^{3}}-\int_{(\diamondsuit_{1})^{c}}\right)(1+O((\theta_{2}|x|)^{-\frac{1}{4}})),
=𝐕(iθ)(8π)32det(Hessθϕ(g;θ))(1+O((θ2|x|)14)).\displaystyle=\mathbf{V}(i\theta)\,\frac{(8\pi)^{\frac{3}{2}}}{\sqrt{\det(-\mathrm{Hess}_{\theta}\,\phi(g;\theta))}}\,(1+O((\theta_{2}|x|)^{-\frac{1}{4}})).

For 𝐉2\mathbf{J}_{2}, it follows from Lemma 6.2 that

(ϕ(g;θiλ)ϕ(g;θ))λTHessθϕ(g;θ)λ(θ2|x|)12,λ2,\Re(\phi(g;\theta-i\lambda)-\phi(g;\theta))\gtrsim-\lambda^{\mathrm{T}}\mathrm{Hess}_{\theta}\,\phi(g;\theta)\lambda\gtrsim(\theta_{2}|x|)^{\frac{1}{2}},\quad\forall\,\lambda\in\diamondsuit_{2},

which, together with (6.3), implies that

|𝐉2|𝐕(iθ)ec(θ2|x|)123ecλTHessθϕ(g;θ)λ𝑑λ=o(𝐉1).\displaystyle|\mathbf{J}_{2}|\leq\mathbf{V}(i\theta)\,e^{-c\,(\theta_{2}|x|)^{\frac{1}{2}}}\int_{\mathbb{R}^{3}}e^{c\,\lambda^{\mathrm{T}}\mathrm{Hess}_{\theta}\,\phi(g;\theta)\lambda}\,d\lambda=o(\mathbf{J}_{1}).

We are left with 𝐉3\mathbf{J}_{3}. Set λ^=λ/|λ|\widehat{\lambda}=\lambda/|\lambda| for λ0\lambda\neq 0. Notice that Lemma 6.2 again yields

(ϕ(g;θiλ)ϕ(g;θ))λ^THessθϕ(g;θ)λ^(θ2|x|)2,λ3.\Re(\phi(g;\theta-i\lambda)-\phi(g;\theta))\gtrsim-\widehat{\lambda}^{\mathrm{T}}\,\mathrm{Hess}_{\theta}\,\phi(g;\theta)\widehat{\lambda}\gtrsim(\theta_{2}|x|)^{2},\quad\forall\,\lambda\in\diamondsuit_{3}.

Moreover, (6.3) says that |𝐕(iθ+λ)|100𝐕(iθ)e|λ|/2|\mathbf{V}(i\theta+\lambda)|\leq 100\,\mathbf{V}(i\theta)\,e^{-|\lambda|/2}. Then using the polar coordinates λ=rγ\lambda=r\,\gamma with γ𝕊2\gamma\in\mathbb{S}^{2}, we have that:

𝐕(iθ)1|𝐉3|\displaystyle\mathbf{V}(i\theta)^{-1}\,|\mathbf{J}_{3}| ec(θ2|x|)23e|λ|/2ecλ^THessθϕ(g;θ)λ^𝑑λ\displaystyle\lesssim e^{-c\,(\theta_{2}|x|)^{2}}\,\int_{\diamondsuit_{3}}e^{-|\lambda|/2}\,e^{c\,\widehat{\lambda}^{\mathrm{T}}\mathrm{Hess}_{\theta}\,\phi(g;\theta)\widehat{\lambda}}d\lambda
ec(θ2|x|)21+r2er/2𝑑r𝕊2ecγTHessθϕ(g;θ)γ𝑑γ\displaystyle\lesssim e^{-c\,(\theta_{2}|x|)^{2}}\int_{1}^{+\infty}r^{2}e^{-r/2}dr\int_{\mathbb{S}^{2}}e^{c\,\gamma^{\mathrm{T}}\mathrm{Hess}_{\theta}\,\phi(g;\theta)\gamma}d\gamma
ec(θ2|x|)2𝕊2ec(γ22+γ32)|x|2𝑑γ,\displaystyle\lesssim e^{-c\,(\theta_{2}|x|)^{2}}\int_{\mathbb{S}^{2}}e^{-c\,(\gamma_{2}^{2}+\gamma_{3}^{2})\,|x|^{2}}d\gamma,

because of (6.2). Next, using the spherical coordinates γ=(cosρ1,sinρ1cosρ2,sinρ1sinρ2)\gamma=(\cos{\rho_{1}},\,\sin{\rho_{1}}\cos{\rho_{2}},\,\sin{\rho_{1}}\sin{\rho_{2}}), the last integral equals

02π𝑑ρ20πec|x|2sin2ρ1sinρ1dρ1\displaystyle\int_{0}^{2\pi}d\rho_{2}\int_{0}^{\pi}e^{-c\,|x|^{2}\sin^{2}{\rho_{1}}}\sin{\rho_{1}}d\rho_{1} =4π0π2ec|x|2sin2ρ1sinρ1dρ1\displaystyle=4\pi\int_{0}^{\frac{\pi}{2}}e^{-c\,|x|^{2}\sin^{2}{\rho_{1}}}\sin{\rho_{1}}d\rho_{1}
0π2ρ1ec|x|2ρ12𝑑ρ11|x|2,\displaystyle\lesssim\int_{0}^{\frac{\pi}{2}}\rho_{1}e^{-c\,|x|^{2}\rho_{1}^{2}}d\rho_{1}\lesssim\frac{1}{|x|^{2}},

where the penultimate “\lesssim” from the fact that sinρρ\sin{\rho}\sim\rho on [0,π2][0,\frac{\pi}{2}]. Hence from (6.2), we get also |𝐉3|=o(𝐉1)|\mathbf{J}_{3}|=o(\mathbf{J}_{1}) under our assumption.

Combining all the estimates obtained above, we complete the proof of Theorem 6.1. ∎

Remark 6.4.

From the above proof, one can easily see that the condition |θ|3|\theta|\leq 3 in Theorem 6.1 can be weakened to |θ|α0|\theta|\leq\alpha_{0} with 3α0<π3\leq\alpha_{0}<\pi and in such case oζ0(1)o_{\zeta_{0}}(1) should be modified to oζ0,α0(1)o_{\zeta_{0},\,\alpha_{0}}(1) as well. Also note that in this case the choice of ζ0\zeta_{0} depends on α0\alpha_{0}.

7 Preparations for asymptotics in difficult cases

In this section, we study more properties of the functions introduced in previous sections, which will be used to develop the uniform asymptotics of our heat kernel.

The first lemma is concerned with the modified Bessel function of order 0 with an additional parameter ρ\rho, which is defined by

I0(ρ;r):=12πρρercosγ𝑑γ,r, 0<ρπ.\displaystyle I_{0}(\rho;r):=\frac{1}{2\pi}\int_{-\rho}^{\rho}e^{r\cos{\gamma}}\,d\gamma,\qquad r\in\mathbb{R},\ 0<\rho\leq\pi. (7.1)

Notice that I0(π;r)I_{0}(\pi\,;r) is exactly I0(r)I_{0}(r) (cf. (2.32)).

Lemma 7.1.

It holds that:

I0(r)=er2πr(1+O(r1)),asr+,\displaystyle I_{0}(r)=\frac{e^{r}}{\sqrt{2\pi r}}\,(1+O(r^{-1})),\qquad{\rm as}\,\,r\to+\infty, (7.2)
I0(r)er(1+r)12,r>0,\displaystyle I_{0}(r)\sim e^{r}\,(1+r)^{-\frac{1}{2}},\quad\forall\,r>0, (7.3)
I0(ρ;r)=er2πr(1+o(1))=I0(r)(1+o(1)),asrρ2+.\displaystyle I_{0}(\rho;r)=\frac{e^{r}}{\sqrt{2\pi r}}\,(1+o(1))=I_{0}(r)\,(1+o(1)),\qquad{\rm as}\,\,r\rho^{2}\to+\infty. (7.4)
Proof.

The asymptotics (7.2) and (7.4) follow directly from Laplace’s method; see also [22, § 8.451.5, p. 920] for (7.2). To show (7.3), it suffices to observe that I0(r)1I_{0}(r)\sim 1 whenever rr is bounded. ∎

Some basic properties of the functions Υ\Upsilon^{\prime} (cf. (4.9)), 𝒵\mathcal{Z} (cf. (4.10)) and Φ\Phi (cf. (4.11)) originated from Υ\Upsilon (cf. (2.33) or (4.8)) are collected in the following subsection:

7.1 Properties of Υ,𝒵\Upsilon^{\prime},\mathcal{Z} and Φ\Phi

Lemma 7.2.

The following conclusions hold:

  1. (i)

    Φ\Phi is even, positive and strictly increasing on [0,+)[0,\ +\infty).

  2. (ii)

    It holds uniformly for all 0<r<ϑ10<r<\vartheta_{1} that

    Υ(r)r(ϑ1r)2,Υ′′(r)1(ϑ1r)3,Υ′′′(r)r(ϑ1r)4.-\Upsilon^{\prime}(r)\sim\frac{r}{(\vartheta_{1}-r)^{2}},\quad-\Upsilon^{\prime\prime}(r)\sim\frac{1}{(\vartheta_{1}-r)^{3}},\quad-\Upsilon^{\prime\prime\prime}(r)\sim\frac{r}{(\vartheta_{1}-r)^{4}}. (7.5)
  3. (iii)

    We have uniformly for all ρ0\rho\geq 0 that

    0𝒵(ρ)<ϑ1,|𝒵(ρ)|(ϑ1𝒵(ρ))3,|𝒵′′(ρ)|(ϑ1𝒵(ρ))5.0\leq\mathcal{Z}(\rho)<\vartheta_{1},\quad|\mathcal{Z}^{\prime}(\rho)|\lesssim(\vartheta_{1}-\mathcal{Z}(\rho))^{3},\quad|\mathcal{Z}^{\prime\prime}(\rho)|\lesssim(\vartheta_{1}-\mathcal{Z}(\rho))^{5}. (7.6)
  4. (iv)

    As ρ+\rho\to+\infty, we have

    Φ(ρ)=ϑ1ρ2ϑ1ρ+2+O(ρ12),\displaystyle\Phi(\rho)=\vartheta_{1}\,\rho-2\sqrt{\vartheta_{1}\,\rho}+2+O(\rho^{-\frac{1}{2}}), (7.7)
    𝒵(ρ)=ϑ1ϑ1ρ+O(ρ32),𝒵(ρ)=O(ρ32),𝒵′′(ρ)=O(ρ52).\displaystyle\mathcal{Z}(\rho)=\vartheta_{1}-\sqrt{\frac{\vartheta_{1}}{\rho}}+O(\rho^{-\frac{3}{2}}),\quad\mathcal{Z}^{\prime}(\rho)=O(\rho^{-\frac{3}{2}}),\quad\mathcal{Z}^{\prime\prime}(\rho)=O(\rho^{-\frac{5}{2}}). (7.8)

    Moreover,

    Φ(ρ)1+ρ(ϑ1𝒵(ρ))2,ρ0.\displaystyle\Phi(\rho)\sim 1+\rho\sim(\vartheta_{1}-\mathcal{Z}(\rho))^{-2},\qquad\rho\geq 0. (7.9)
Proof.

Recall that the odd function 𝒵\mathcal{Z} satisfies 𝒵(0)=0\mathcal{Z}(0)=0 and 𝒵(ρ)<ϑ1\mathcal{Z}(\rho)<\vartheta_{1}. From the definition of Υ\Upsilon (cf. (4.8)), we see that Φ(ρ)=Υ(𝒵(ρ))+ρ𝒵(ρ)\Phi(\rho)=\Upsilon(\mathcal{Z}(\rho))+\rho\mathcal{Z}(\rho) is even, and Φ(0)=Υ(0)=3>0\Phi(0)=\Upsilon(0)=3>0. Taking derivatives, it is clear that

Φ(ρ)=𝒵(ρ),𝒵(ρ)=1Υ′′(𝒵(ρ))>0,𝒵′′(ρ)=Υ′′′(𝒵(ρ))(Υ′′(𝒵(ρ)))3,ρ,\displaystyle\Phi^{\prime}(\rho)=\mathcal{Z}(\rho),\quad\mathcal{Z}^{\prime}(\rho)=-\frac{1}{\Upsilon^{\prime\prime}(\mathcal{Z}(\rho))}>0,\quad\mathcal{Z}^{\prime\prime}(\rho)=-\frac{\Upsilon^{\prime\prime\prime}(\mathcal{Z}(\rho))}{(\Upsilon^{\prime\prime}(\mathcal{Z}(\rho)))^{3}},\quad\forall\,\rho\in\mathbb{R}, (7.10)

which implies item (i).

Notice that item (ii) (resp. (iii)) is a simple consequence of (4.9) (resp. item (ii) and (7.10)).

Now we turn to the proof of item (iv). For rr near the point ϑ1\vartheta_{1}, we first observe that Υ(r)=r2sinr/(sinrrcosr)\Upsilon(r)=r^{2}\sin{r}/(\sin{r}-r\cos{r}). Then using the Taylor expansion at ϑ1\vartheta_{1} (recalling tanϑ1=ϑ1\tan\vartheta_{1}=\vartheta_{1}):

r2sinr\displaystyle r^{2}\sin{r} =ϑ12sinϑ1+3ϑ1sinϑ1(rϑ1)+O(|rϑ1|2),\displaystyle=\vartheta_{1}^{2}\sin{\vartheta_{1}}+3\vartheta_{1}\sin{\vartheta_{1}}(r-\vartheta_{1})+O(|r-\vartheta_{1}|^{2}),
sinrrcosr\displaystyle\sin{r}-r\cos{r} =ϑ1sinϑ1(rϑ1)+sinϑ1(rϑ1)2+O(|rϑ1|3),\displaystyle=\vartheta_{1}\sin{\vartheta_{1}}(r-\vartheta_{1})+\sin{\vartheta_{1}}(r-\vartheta_{1})^{2}+O(|r-\vartheta_{1}|^{3}),

a direct calculation shows that:

Υ(r)=ϑ1rϑ1+2+(r),\displaystyle\Upsilon(r)=\frac{\vartheta_{1}}{r-\vartheta_{1}}+2+\square(r), (7.11)

where (r)\square(r) is an analytic function satisfying (ϑ1)=0\square(\vartheta_{1})=0. Taking derivative, we obtain that:

Υ(r)=ϑ1(rϑ1)2+O(1),Υ′′(r)=2ϑ1(rϑ1)3+O(1),Υ′′′(r)=6ϑ1(rϑ1)4+O(1).\begin{gathered}\Upsilon^{\prime}(r)=-\frac{\vartheta_{1}}{(r-\vartheta_{1})^{2}}+O(1),\quad\Upsilon^{\prime\prime}(r)=\frac{2\vartheta_{1}}{(r-\vartheta_{1})^{3}}+O(1),\\ \Upsilon^{\prime\prime\prime}(r)=-\frac{6\vartheta_{1}}{(r-\vartheta_{1})^{4}}+O(1).\end{gathered} (7.12)

As a result, whenever ρ+\rho\to+\infty, the first asymptotic above gives

ρ=Υ(𝒵(ρ))=ϑ1(ϑ1𝒵(ρ))2+O(1),soϑ1𝒵(ρ)=O(ρ12).\displaystyle\rho=-\Upsilon^{\prime}(\mathcal{Z}(\rho))=\frac{\vartheta_{1}}{(\vartheta_{1}-\mathcal{Z}(\rho))^{2}}+O(1),\quad\mbox{so}\quad\vartheta_{1}-\mathcal{Z}(\rho)=O(\rho^{-\frac{1}{2}}). (7.13)

Taking square root, we get that

ρ=ϑ1ϑ1𝒵(ρ)+O(ϑ1𝒵(ρ))=ϑ1ϑ1𝒵(ρ)+O(ρ12),\displaystyle\sqrt{\rho}=\frac{\sqrt{\vartheta_{1}}}{\vartheta_{1}-\mathcal{Z}(\rho)}+O(\vartheta_{1}-\mathcal{Z}(\rho))=\frac{\sqrt{\vartheta_{1}}}{\vartheta_{1}-\mathcal{Z}(\rho)}+O(\rho^{-\frac{1}{2}}),

which implies the first equation of (7.8). So we get the other two estimates of (7.8) from (7.6). On the other hand, using (7.11), it turns out that:

Φ(ρ)\displaystyle\Phi(\rho) =Υ(𝒵(ρ))+ρ𝒵(ρ)=ϑ1𝒵(ρ)ϑ1+2+ϑ1ρϑ1ρ+O(ρ12)\displaystyle=\Upsilon(\mathcal{Z}(\rho))+\rho\mathcal{Z}(\rho)=\frac{\vartheta_{1}}{\mathcal{Z}(\rho)-\vartheta_{1}}+2+\vartheta_{1}\rho-\sqrt{\vartheta_{1}\rho}+O(\rho^{-\frac{1}{2}})
=ϑ1ρ(1+O(ρ1))+2+ϑ1ρϑ1ρ+O(ρ12)=ϑ1ρ2ϑ1ρ+2+O(ρ12),\displaystyle=-\sqrt{\vartheta_{1}\rho}\,(1+O(\rho^{-1}))+2+\vartheta_{1}\rho-\sqrt{\vartheta_{1}\rho}+O(\rho^{-\frac{1}{2}})=\vartheta_{1}\rho-2\sqrt{\vartheta_{1}\rho}+2+O(\rho^{-\frac{1}{2}}),

which is exactly (7.7).

Finally, (7.9) follows from items (i) and (iv). ∎

Recall the even function 𝔮\mathfrak{q} is defined by (cf. (2.33))

𝔮(r)=r2Υ(r)sinrΥ(r)Υ′′(r),ϑ1<r<ϑ1.\displaystyle\mathfrak{q}(r)=\frac{r^{2}\,\Upsilon(r)}{-\sin{r}\,\Upsilon^{\prime}(r)\,\sqrt{-\Upsilon^{\prime\prime}(r)}},\quad-\vartheta_{1}<r<\vartheta_{1}.

The following simple observation will be used in the proof of Propositions 2.9 and 8.2, as well as of Theorem 9.1 below:

Lemma 7.3.

The even function 𝔮\mathfrak{q} is positive and smooth on (ϑ1,ϑ1)(-\vartheta_{1},\ \vartheta_{1}). Moreover, we have

𝔮(r)=2ϑ1ϑ12sinϑ1(ϑ1r)52(1+O(ϑ1r)),asrϑ1,\mathfrak{q}(r)=\frac{\sqrt{2\vartheta_{1}}\,\vartheta_{1}}{-2\sin\vartheta_{1}}(\vartheta_{1}-r)^{\frac{5}{2}}\,(1+O(\vartheta_{1}-r)),\quad{\rm as}\,\,r\to\vartheta_{1}^{-}, (7.14)

and

𝔮(r)(ϑ1|r|)52,|𝔮(r)|(ϑ1|r|)32,ϑ1<r<ϑ1.\mathfrak{q}(r)\sim(\vartheta_{1}-|r|)^{\frac{5}{2}},\qquad|\mathfrak{q}^{\prime}(r)|\lesssim(\vartheta_{1}-|r|)^{\frac{3}{2}},\qquad-\vartheta_{1}<r<\vartheta_{1}. (7.15)
Proof.

The proof is easy. Using (4.8) and (4.9), it suffices to observe that Υ\Upsilon is monotonically decreasing on [0,ϑ1)[0,\ \vartheta_{1}), Υ(0)=3\Upsilon(0)=3, and π\pi (resp. 0) is a simple zero of Υ\Upsilon (resp. Υ\Upsilon^{\prime}). Notice that we can use directly (7.11) and (7.12) to deduce (7.14). ∎

In Sections 8 and 9 below, we need to utilize more information of 𝒟\mathcal{D} defined by (2.35) near its unique minimum point. To do this, we introduce

7.2 The auxiliary function 𝐇\mathbf{H}

Let u,𝐰¯u,\mathbf{\overline{w}} be defined as in Assumption (A) (cf. (2.18)). We set for s=(s1,s2)2s=(s_{1},s_{2})\in\mathbb{R}^{2}

𝐀~(s)\displaystyle\widetilde{\mathbf{A}}(s) =𝐀~(u;s):=u12+(u2+2𝐰¯s1)2+4𝐰¯2s22=|u|2+4𝐰¯2|s|2+4u2𝐰¯s1,\displaystyle=\widetilde{\mathbf{A}}(u;s):=\sqrt{u_{1}^{2}+(u_{2}+2\mathbf{\overline{w}}\,s_{1})^{2}+4\mathbf{\overline{w}}^{2}s_{2}^{2}}=\sqrt{|u|^{2}+4\mathbf{\overline{w}}^{2}|s|^{2}+4u_{2}\mathbf{\overline{w}}\,s_{1}}, (7.16)

namely, the modulus of the vector u+2𝐰¯s1e2+2𝐰¯s2e3u+2\mathbf{\overline{w}}s_{1}\,e_{2}+2\mathbf{\overline{w}}\,s_{2}\,e_{3} in (2.35), and write

𝐔~(s)=𝐔~(u;s):=𝐰¯2|s|2𝐀~(s),s0.\displaystyle\widetilde{\mathbf{U}}(s)=\widetilde{\mathbf{U}}(u;s):=\mathbf{\overline{w}}^{-2}|s|^{-2}\widetilde{\mathbf{A}}(s),\qquad s\neq 0. (7.17)

Then (4.12) implies the following more suitable expression of 𝒟\mathcal{D} than (2.35):

𝒟(u;s)=𝐰¯2|s|2Φ(𝐔~(s)),s0.\displaystyle\mathcal{D}(u;s)=\mathbf{\overline{w}}^{2}|s|^{2}\,\Phi(\widetilde{\mathbf{U}}(s)),\quad\forall\,s\neq 0. (7.18)

In what follows, it will be convenient to work in the modified polar coordinates

s=(s1,s2)=(wcosγ,wsinγ),w>0,π<γ<π,for s0.s=(s_{1},s_{2})=(-\mathrm{w}\cos{\gamma},-\mathrm{w}\sin{\gamma}),\quad\mathrm{w}>0,\quad-\pi<\gamma<\pi,\ \mbox{for $s\neq 0$}. (7.19)

Correspondingly, given a function ff on 2{\mathbb{R}}^{2} we shall write fp(w,γ):=f(s)f_{\mathrm{p}}(\mathrm{w},\gamma):=f(s). And for brevity, given a function g(u;s)g(u;s) (resp. g(u;w,γ)g(u;\mathrm{w},\gamma)) on 2\mathbb{R}^{2} (resp. (0,)×(π,π)(0,\infty)\times(-\pi,\pi)) with the parameter uu, we will write it simply g(s)g(s) (resp. g(w,γ)g(\mathrm{w},\gamma)) when there is no confusion. For example, the notation above implies that

𝒟(u;s)=𝒟(u;wcosγ,wsinγ)=:𝒟p(u;w,γ)=𝒟p(w,γ).\mathcal{D}(u;s)=\mathcal{D}(u;-\mathrm{w}\cos{\gamma},-\mathrm{w}\sin{\gamma})=:\mathcal{D}_{\mathrm{p}}(u;\mathrm{w},\gamma)=\mathcal{D}_{\mathrm{p}}(\mathrm{w},\gamma).

Since the unique minimizer of 𝒟p()\mathcal{D}_{\mathrm{p}}(\cdot) (which is exactly (1,0)(1,0)) belongs to (0,)×{0}(0,\infty)\times\{0\}, we will see later that it is natural to consider the restriction of 𝒟p()\mathcal{D}_{\mathrm{p}}(\cdot) on (0,)×{0}(0,\infty)\times\{0\}. To be more precise, we set

𝐇(w):=𝒟p(w,0),w>0.\displaystyle\mathbf{H}(\mathrm{w}):=\mathcal{D}_{\mathrm{p}}(\mathrm{w},0),\qquad\mathrm{w}>0. (7.20)

Similarly, after introducing the counterparts of 𝐀~\widetilde{\mathbf{A}} and 𝐔~\widetilde{\mathbf{U}}:

𝐀(w):=𝐀~(w,0)=|u|2+4𝐰¯2w24u2𝐰¯w,𝐔(w):=𝐔~(w,0)=𝐰¯2w2𝐀(w),\begin{gathered}\mathbf{A}(\mathrm{w}):=\widetilde{\mathbf{A}}(-\mathrm{w},0)=\sqrt{|u|^{2}+4\mathbf{\overline{w}}^{2}\,\mathrm{w}^{2}-4u_{2}\,\mathbf{\overline{w}}\,\mathrm{w}},\\ \mathbf{U}(\mathrm{w}):=\widetilde{\mathbf{U}}(-\mathrm{w},0)=\mathbf{\overline{w}}^{-2}\,\mathrm{w}^{-2}\mathbf{A}(\mathrm{w}),\end{gathered} (7.21)

we obtain

𝐇(w)=𝐰¯2w2Φ(𝐔(w)),w>0.\displaystyle\mathbf{H}(\mathrm{w})=\mathbf{\overline{w}}^{2}\,\mathrm{w}^{2}\,\Phi(\mathbf{U}(\mathrm{w})),\qquad\forall\,\mathrm{w}>0. (7.22)

From the definition, 𝐇\mathbf{H} is smooth on (0,+)(0,+\infty) and 11 is its minimizer, so 𝐇(1)=0\mathbf{H}^{\prime}(1)=0. For the convenience of readers, we list here some basic facts, which can be deduced immediately from their definitions by recalling (4.15), (4.8), (4.10), (4.11) and (7.10) :

{𝐀(1)=θ22ψ(|θ|)>0,𝐔(1)=𝐀(1)𝐰¯2=ψ(|θ|)ψ2(|θ|)=Υ(|θ|),𝒵(𝐔(1))=𝒵(Υ(|θ|))=|θ|,𝒵(𝐔(1))=1Υ′′(|θ|),Φ(𝐔(1))=Υ(|θ|)|θ|Υ(|θ|)=Υ(|θ|)+|θ|𝐔(1).\left\{\begin{gathered}\mathbf{A}(1)=\theta_{2}^{2}\psi^{\prime}(|\theta|)>0,\quad\mathbf{U}(1)=\frac{\mathbf{A}(1)}{\mathbf{\overline{w}}^{2}}=\frac{\psi^{\prime}(|\theta|)}{\psi^{2}(|\theta|)}=-\Upsilon^{\prime}(|\theta|),\\ \mathcal{Z}(\mathbf{U}(1))=\mathcal{Z}(-\Upsilon^{\prime}(|\theta|))=|\theta|,\quad\mathcal{Z}^{\prime}(\mathbf{U}(1))=-\frac{1}{\Upsilon^{\prime\prime}(|\theta|)},\\ \Phi(\mathbf{U}(1))=\Upsilon(|\theta|)-|\theta|\Upsilon^{\prime}(|\theta|)=\Upsilon(|\theta|)+|\theta|\mathbf{U}(1).\end{gathered}\right. (7.23)

In Section 8 below, we will fully utilize the Taylor expansion of order 22 of 𝐇\mathbf{H} at its minimizer 11 and the fact that

The critical point 11 is nondegenerate.

To do this, we start to calculate 𝐇′′(1)\mathbf{H}^{\prime\prime}(1) explicitly (equality (7.32)). Indeed, a simple calculation shows that:

𝐀=2𝐰¯2𝐰¯wu2𝐀,𝐀′′=4u12𝐰¯2𝐀3,𝐀′′′=12u12𝐰¯2𝐀4𝐀,\displaystyle\mathbf{A}^{\prime}=2\mathbf{\overline{w}}\,\frac{2\mathbf{\overline{w}}\,\mathrm{w}-u_{2}}{\mathbf{A}},\quad\mathbf{A}^{\prime\prime}=\frac{4u_{1}^{2}\,\mathbf{\overline{w}}^{2}}{\mathbf{A}^{3}},\quad\mathbf{A}^{\prime\prime\prime}=-\frac{12u_{1}^{2}\,\mathbf{\overline{w}}^{2}}{\mathbf{A}^{4}}\,\mathbf{A}^{\prime}, (7.24)
𝐔=𝐰¯2(2w3𝐀+w2𝐀),𝐔′′=𝐰¯2(6w4𝐀4w3𝐀+w2𝐀′′),\displaystyle\mathbf{U}^{\prime}=\mathbf{\overline{w}}^{-2}(-2\mathrm{w}^{-3}\mathbf{A}+\mathrm{w}^{-2}\mathbf{A}^{\prime}),\quad\mathbf{U}^{\prime\prime}=\mathbf{\overline{w}}^{-2}(6\mathrm{w}^{-4}\mathbf{A}-4\mathrm{w}^{-3}\mathbf{A}^{\prime}+\mathrm{w}^{-2}\mathbf{A}^{\prime\prime}), (7.25)
𝐔′′′=𝐰¯2(24w5𝐀+18w4𝐀6w3𝐀′′+w2𝐀′′′)=6w1𝐔′′6w2𝐔+𝐰¯2w2𝐀′′′.\displaystyle\begin{aligned} \mathbf{U}^{\prime\prime\prime}&=\mathbf{\overline{w}}^{-2}(-24\mathrm{w}^{-5}\mathbf{A}+18\mathrm{w}^{-4}\mathbf{A}^{\prime}-6\mathrm{w}^{-3}\mathbf{A}^{\prime\prime}+\mathrm{w}^{-2}\mathbf{A}^{\prime\prime\prime})\\[2.84526pt] &=-6\,\mathrm{w}^{-1}\,\mathbf{U}^{\prime\prime}-6\,\mathrm{w}^{-2}\,\mathbf{U}^{\prime}+\mathbf{\overline{w}}^{-2}\,\mathrm{w}^{-2}\,\mathbf{A}^{\prime\prime\prime}.\end{aligned} (7.26)

Moreover, using the chain rule and the fact that Φ(r)=𝒵(r)\Phi^{\prime}(r)=\mathcal{Z}(r) (cf. (7.10)), we get that:

𝐇=𝐰¯2[2wΦ(𝐔)+w2Φ(𝐔)𝐔]=𝐰¯2[2wΦ(𝐔)+w2𝒵(𝐔)𝐔],\displaystyle\mathbf{H}^{\prime}=\mathbf{\overline{w}}^{2}\,[2\mathrm{w}\,\Phi(\mathbf{U})+\mathrm{w}^{2}\Phi^{\prime}(\mathbf{U})\mathbf{U}^{\prime}]=\mathbf{\overline{w}}^{2}\,[2\mathrm{w}\,\Phi(\mathbf{U})+\mathrm{w}^{2}\mathcal{Z}(\mathbf{U})\mathbf{U}^{\prime}], (7.27)
𝐇′′=𝐰¯2[2Φ(𝐔)+4w𝒵(𝐔)𝐔+w2𝒵(𝐔)(𝐔)2+w2𝒵(𝐔)𝐔′′],\displaystyle\mathbf{H}^{\prime\prime}=\mathbf{\overline{w}}^{2}\,[2\Phi(\mathbf{U})+4\mathrm{w}\,\mathcal{Z}(\mathbf{U})\mathbf{U}^{\prime}+\mathrm{w}^{2}\mathcal{Z}^{\prime}(\mathbf{U})(\mathbf{U}^{\prime})^{2}+\mathrm{w}^{2}\mathcal{Z}(\mathbf{U})\mathbf{U}^{\prime\prime}], (7.28)
𝐇′′′=𝒵(𝐔)𝐀′′′+𝐰¯2𝒵(𝐔)[6w(𝐔)2+3w2𝐔𝐔′′]+𝐰¯2w2𝒵′′(𝐔)(𝐔)3,\displaystyle\mathbf{H}^{\prime\prime\prime}=\mathcal{Z}(\mathbf{U})\mathbf{A}^{\prime\prime\prime}+\mathbf{\overline{w}}^{2}\mathcal{Z}^{\prime}(\mathbf{U})[6\mathrm{w}\,(\mathbf{U}^{\prime})^{2}+3\mathrm{w}^{2}\,\mathbf{U}^{\prime}\mathbf{U}^{\prime\prime}]+\mathbf{\overline{w}}^{2}\mathrm{w}^{2}\mathcal{Z}^{\prime\prime}(\mathbf{U})(\mathbf{U}^{\prime})^{3}, (7.29)

where we have used (7.26) in the last equality.

In particular, at the minimizer w=1\mathrm{w}=1, by the fact that 𝒵(𝐔(1))=|θ|\mathcal{Z}(\mathbf{U}(1))=|\theta| and 𝐇(1)=0\mathbf{H}^{\prime}(1)=0, it follows from (7.27) that

𝐔(1)=2Φ(𝐔(1))|θ|=2Υ(|θ|)2Υ(|θ|)|θ|\mathbf{U}^{\prime}(1)=-2\frac{\Phi(\mathbf{U}(1))}{|\theta|}=2\Upsilon^{\prime}(|\theta|)-2\frac{\Upsilon(|\theta|)}{|\theta|} (7.30)

by the last equation in (7.23). Using the first equation of (7.25), we obtain

𝐀(1)=𝐰¯2𝐔(1)+2𝐀(1).\mathbf{A}^{\prime}(1)=\mathbf{\overline{w}}^{2}\,\mathbf{U}^{\prime}(1)+2\mathbf{A}(1).

This together with the fact that 𝐀′′(1)=4u12𝐰¯2/𝐀(1)3\mathbf{A}^{\prime\prime}(1)=4\,u_{1}^{2}\,\mathbf{\overline{w}}^{2}/\mathbf{A}(1)^{3} (via the second equality in (7.24)) and 𝐔(1)=𝐰¯2𝐀(1)\mathbf{U}(1)=\mathbf{\overline{w}}^{-2}\mathbf{A}(1) (cf. (7.23)) allows us to yield

𝐔′′(1)=2𝐔(1)4𝐔(1)+4u12𝐀(1)3\mathbf{U}^{\prime\prime}(1)=-2\mathbf{U}(1)-4\mathbf{U}^{\prime}(1)+\frac{4u_{1}^{2}}{\mathbf{A}(1)^{3}}

from the second equation of (7.25). Inserting this into (7.28), together again with the fact that 𝒵(𝐔(1))=|θ|\mathcal{Z}(\mathbf{U}(1))=|\theta| and 𝒵(𝐔(1))=1/Υ′′(|θ|)\mathcal{Z}^{\prime}(\mathbf{U}(1))=-1/\Upsilon^{\prime\prime}(|\theta|) (cf. (7.23)), we have

𝐇′′(1)=[2Φ(𝐔(1))2|θ|𝐔(1)(𝐔(1))2Υ′′(|θ|)+4|θ|u12𝐀(1)3]𝐰¯2=[𝔭(|θ|)+4|θ|u12𝐀(1)3]𝐰¯2,\displaystyle\mathbf{H}^{\prime\prime}(1)=\left[2\Phi(\mathbf{U}(1))-2|\theta|\mathbf{U}(1)-\frac{(\mathbf{U}^{\prime}(1))^{2}}{\Upsilon^{\prime\prime}(|\theta|)}+\frac{4|\theta|u_{1}^{2}}{\mathbf{A}(1)^{3}}\right]\mathbf{\overline{w}}^{2}=\left[\mathfrak{p}(|\theta|)+\frac{4|\theta|u_{1}^{2}}{\mathbf{A}(1)^{3}}\right]\mathbf{\overline{w}}^{2},

where we have used (7.30) and the last equality of (7.23) in the last “=”, and set that

𝔭(r):=2Υ(r)4(Υ(r)rΥ(r))2r2Υ′′(r)=4Υ(r)2Υ′′(r)sin2r,0<r<ϑ1.\mathfrak{p}(r):=2\,\Upsilon(r)-4\,\frac{(\Upsilon(r)-r\Upsilon^{\prime}(r))^{2}}{r^{2}\,\Upsilon^{\prime\prime}(r)}=4\frac{\Upsilon(r)^{2}}{-\Upsilon^{\prime\prime}(r)\sin^{2}r},\quad 0<r<\vartheta_{1}. (7.31)

The second “=” in (7.31) can be verified directly. In fact, from (3.11), we see that:

𝔭(r)\displaystyle\mathfrak{p}(r) =2r2[ψψ′′+2(ψ)2]2(ψ+rψ)2r2Υ′′ψ4=22ψ+r2(ψ′′+4ψr)r2Υ′′ψ3\displaystyle=2\frac{r^{2}\left[-\psi\,\psi^{\prime\prime}+2\,(\psi^{\prime})^{2}\right]-2\,(\psi+r\,\psi^{\prime})^{2}}{r^{2}\,\Upsilon^{\prime\prime}\,\psi^{4}}=2\frac{2\,\psi+r^{2}\,(\psi^{\prime\prime}+4\frac{\psi^{\prime}}{r})}{-r^{2}\,\Upsilon^{\prime\prime}\,\psi^{3}}
=22ψ+2ψ(r2csc2r1)r2Υ′′ψ3=4Υ(r)2Υ′′(r)sin2r,\displaystyle=2\frac{2\,\psi+2\,\psi\,(r^{2}\,\csc^{2}{r}-1)}{-r^{2}\,\Upsilon^{\prime\prime}\,\psi^{3}}=4\frac{\Upsilon(r)^{2}}{-\Upsilon^{\prime\prime}(r)\sin^{2}r},

where the penultimate equality follows from (3.17). Remark that π\pi, the unique zero of Υ\Upsilon on (0,ϑ1)(0,\ \vartheta_{1}), is simple. By (4.9), it is obvious that 𝔭>0\mathfrak{p}>0, which shows that 𝐇′′(1)>0\mathbf{H}^{\prime\prime}(1)>0. In conclusion, by the fact that 𝐀(1)=θ22ψ(|θ|)\mathbf{A}(1)=\theta_{2}^{2}\,\psi^{\prime}(|\theta|), u1=θ1θ22ψ(|θ|)/|θ|u_{1}=\theta_{1}\,\theta_{2}^{2}\,\psi^{\prime}(|\theta|)/|\theta| and 𝐰¯=θ2ψ(|θ|)\mathbf{\overline{w}}=\theta_{2}\,\psi(|\theta|), we yield the following:

Lemma 7.4.

Let θ,u\theta,u be as in Assumption (A) (cf. (2.18)). Then

𝐇′′(1)=s12𝒟(u;𝐬¯)=4(θ12Υ(|θ|)|θ|+θ22Υ′′(|θ|)sin2|θ|)>0.\displaystyle\mathbf{H}^{\prime\prime}(1)=\partial^{2}_{s_{1}}\mathcal{D}(u;\mathbf{\overline{s}})=4\left(\frac{\theta_{1}^{2}}{-\Upsilon^{\prime}(|\theta|)|\theta|}+\frac{\theta_{2}^{2}}{-\Upsilon^{\prime\prime}(|\theta|)\sin^{2}|\theta|}\right)>0. (7.32)

It is a simple but vital fact that both 𝒟p\mathcal{D}_{\mathrm{p}} and 𝐇\mathbf{H} enjoy some monotonicity properties, which will be used extensively in Section 8. Actually, one has:

Lemma 7.5.

The following conclusions hold:

  1. (i)

    For fixed w\mathrm{w}, the even function γ𝒟p(w,γ)\gamma\mapsto\mathcal{D}_{\mathrm{p}}(\mathrm{w},\gamma) is increasing w.r.t. |γ||\gamma|;

  2. (ii)

    The function 𝐇()\mathbf{H}(\cdot) is decreasing on the interval (0,1)(0,1) and increasing on (1,+)(1,+\infty).

Proof.

The proof of (i) is easy. Since w\mathrm{w} is fixed, then by (7.16) the positive function

𝐀~p(w,γ)\displaystyle\widetilde{\mathbf{A}}_{\mathrm{p}}(\mathrm{w},\gamma) =u12+(u22𝐰¯wcosγ)2+4𝐰¯2w2sin2γ\displaystyle=\sqrt{u_{1}^{2}+(u_{2}-2\mathbf{\overline{w}}\,\mathrm{w}\cos\gamma)^{2}+4\,\mathbf{\overline{w}}^{2}\,\mathrm{w}^{2}\sin^{2}\gamma} (7.33)
=|u|2+4𝐰¯2w24u2𝐰¯wcosγ\displaystyle=\sqrt{|u|^{2}+4\mathbf{\overline{w}}^{2}\,\mathrm{w}^{2}-4u_{2}\mathbf{\overline{w}}\,\mathrm{w}\cos{\gamma}}

is increasing w.r.t. |γ||\gamma|, so is 𝐔~p(w,γ)=𝐰¯2w2𝐀~p(w,γ)\widetilde{\mathbf{U}}_{\mathrm{p}}(\mathrm{w},\gamma)=\mathbf{\overline{w}}^{-2}\mathrm{w}^{-2}\widetilde{\mathbf{A}}_{\mathrm{p}}(\mathrm{w},\gamma). Thus (i) follows from the fact that 𝒟p(w,γ)=𝐰¯2w2Φ(𝐔~p(w,γ))\mathcal{D}_{\mathrm{p}}(\mathrm{w},\gamma)=\mathbf{\overline{w}}^{2}\,\mathrm{w}^{2}\,\Phi(\widetilde{\mathbf{U}}_{\mathrm{p}}(\mathrm{w},\gamma)) and the first claim in Lemma 7.2.

To prove (ii), it suffices to show that 𝐇(w)<0\mathbf{H}^{\prime}(\mathrm{w})<0 on (0,1)(0,1) and 𝐇(w)>0\mathbf{H}^{\prime}(\mathrm{w})>0 on (1,+)(1,+\infty). Since 𝐇(1)=0\mathbf{H}^{\prime}(1)=0 and 𝐇′′(1)>0\mathbf{H}^{\prime\prime}(1)>0, then 𝐇(w)<0\mathbf{H}^{\prime}(\mathrm{w})<0 whenever 1w>01-\mathrm{w}>0 is small enough, and 𝐇(w)>0\mathbf{H}^{\prime}(\mathrm{w})>0 for w1>0\mathrm{w}-1>0 small enough. As a result, by the smoothness of 𝐇\mathbf{H}, it remains to show that there is no w(0,1)(1,+)\mathrm{w}_{*}\in(0,1)\cup(1,+\infty) such that 𝐇(w)=0\mathbf{H}^{\prime}(\mathrm{w}_{*})=0. If this is not the case, from (i) it will deduce that

w𝒟p(u;w,0)=0,γ𝒟p(u;w,0)=0,γ2𝒟p(u;w,0)0,\displaystyle\partial_{\mathrm{w}}\mathcal{D}_{\mathrm{p}}(u;\mathrm{w}_{*},0)=0,\quad\partial_{\gamma}\mathcal{D}_{\mathrm{p}}(u;\mathrm{w}_{*},0)=0,\quad\partial^{2}_{\gamma}\mathcal{D}_{\mathrm{p}}(u;\mathrm{w}_{*},0)\geq 0,

which is equivalent to

s1𝒟(u;w,0)=0,s2𝒟(u;w,0)=0,s22𝒟(u;w,0)0.\displaystyle\partial_{s_{1}}\mathcal{D}(u;-\mathrm{w}_{*},0)=0,\quad\partial_{s_{2}}\mathcal{D}(u;-\mathrm{w}_{*},0)=0,\quad\partial^{2}_{s_{2}}\mathcal{D}(u;-\mathrm{w}_{*},0)\geq 0.

Then following the argument in the proof of Theorem 2.10, we have (w,0)=(1,0)(-\mathrm{w}_{*},0)=(-1,0) and this leads to a contradiction, which completes the proof. ∎

We are now in the position to introduce

7.3 Some key parameters related to 41|x|2𝒟(u;)4^{-1}|x|^{2}\,\mathcal{D}(u;\cdot)

Under Assumption (A) (cf. (2.18)), recalling (7.32) and (4.20), we set in the sequel,

𝔪=𝔪(g):=|x|24𝐇(1)=|x|24𝒟(u;𝐬¯),\displaystyle\mathfrak{m}=\mathfrak{m}(g):=\frac{|x|^{2}}{4}\mathbf{H}(1)=\frac{|x|^{2}}{4}\mathcal{D}(u;\mathbf{\overline{s}}), (7.34)
𝔏1=𝔏1(g):=|x|24s12𝒟(u;𝐬¯)=|x|24𝐇′′(1)=(θ12Υ(|θ|)|θ|+θ22Υ′′(|θ|)sin2|θ|)|x|2,\displaystyle\mathfrak{L}_{1}=\mathfrak{L}_{1}(g):=\frac{|x|^{2}}{4}\partial^{2}_{s_{1}}\mathcal{D}(u;\mathbf{\overline{s}})=\frac{|x|^{2}}{4}\mathbf{H}^{\prime\prime}(1)=\left(\frac{\theta_{1}^{2}}{-\Upsilon^{\prime}(|\theta|)|\theta|}+\frac{\theta_{2}^{2}}{-\Upsilon^{\prime\prime}(|\theta|)\sin^{2}|\theta|}\right)|x|^{2}, (7.35)
𝔏2=𝔏2(g):=|x|24s22𝒟(u;𝐬¯)=ψ(|θ|)|θ|2ψ(|θ|)K3(θ1,θ2)|x|2=u2|x|2𝐰¯|θ|2𝐀(1),\displaystyle\mathfrak{L}_{2}=\mathfrak{L}_{2}(g):=\frac{|x|^{2}}{4}\,\partial^{2}_{s_{2}}\mathcal{D}(u;\mathbf{\overline{s}})=\frac{\psi(|\theta|)\,|\theta|}{2\psi^{\prime}(|\theta|)}\,\mathrm{K}_{3}(\theta_{1},\theta_{2})\,|x|^{2}=\frac{u_{2}\,|x|^{2}\,\mathbf{\overline{w}}\,|\theta|}{2\mathbf{A}(1)}, (7.36)

where we have used in the last “==” of (7.36) the fact that 𝐀(1)=θ22ψ(|θ|)\mathbf{A}(1)=\theta_{2}^{2}\,\psi^{\prime}(|\theta|). Note that 𝔪\mathfrak{m} is exactly the minimum of |x|24𝒟(u;)\frac{|x|^{2}}{4}\mathcal{D}(u;\cdot), and in fact 𝔏1,𝔏2>0\mathfrak{L}_{1},\,\mathfrak{L}_{2}>0 are two eigenvalues of the Hessian matrix of |x|24𝒟(u;)\frac{|x|^{2}}{4}\mathcal{D}(u;\cdot) at its unique minimum point (1,0)(-1,0).

These parameters are important for the asymptotics of the heat kernel in Sections 8 and 9 below. We summarize some useful properties of them in the following lemma, which is a direct application of Theorems 2.1 and 2.10.

Lemma 7.6.

Supposing that Assumption (A) (cf. (2.18)) holds, then we have:

  1. (i)

    𝐇(1)u1θ11θ22|π|θ||2uθϵ2𝐰¯2|θ|1𝐀(1)\mathbf{H}(1)\sim u_{1}\,\theta_{1}^{-1}\sim\theta_{2}^{2}\,|\pi-|\theta|\,|^{-2}\sim u\cdot\theta\sim\epsilon^{-2}\,\mathbf{\overline{w}}^{2}\sim|\theta|^{-1}\,\mathbf{A}(1).

  2. (ii)

    𝔏2ϵu2|x|3𝔪12\mathfrak{L}_{2}\sim\epsilon\,u_{2}\,|x|^{3}\,\mathfrak{m}^{-\frac{1}{2}}, 𝔏1ϵ2d(g)2\mathfrak{L}_{1}\sim\epsilon^{2}\,d(g)^{2}, and 𝔏1𝔏2+ϵ𝔪\mathfrak{L}_{1}\sim\mathfrak{L}_{2}+\epsilon\,\mathfrak{m}.

  3. (iii)

    If |u|1|u|\gtrsim 1, then

    |θ|1,ϵ1,|u|θ22(π|θ|)2,𝐇(1)|u|,𝐰¯|u|12.\displaystyle|\theta|\sim 1,\quad\epsilon\sim 1,\quad|u|\sim\frac{\theta_{2}^{2}}{(\pi-|\theta|)^{2}},\quad\mathbf{H}(1)\sim|u|,\quad\mathbf{\overline{w}}\sim|u|^{\frac{1}{2}}. (7.37)
  4. (iv)

    If |u|1|u|\ll 1 and |θ|1|\theta|\geq 1, then

    ϵu1+u2u112,𝐇(1)u1,𝐰¯ϵu112,θ11.\displaystyle\epsilon\sim u_{1}+u_{2}\,u_{1}^{-\frac{1}{2}},\quad\mathbf{H}(1)\sim u_{1},\quad\mathbf{\overline{w}}\sim\epsilon\,u_{1}^{\frac{1}{2}},\quad\theta_{1}\sim 1. (7.38)

    In particular, u2/u1u_{2}/\sqrt{u_{1}} is bounded.

  5. (v)

    If |θ|3|\theta|\leq 3, then

    u1θ1θ22,u2θ2,𝐇(1)θ22,𝔏2|x|2.\displaystyle u_{1}\sim\theta_{1}\theta_{2}^{2},\quad u_{2}\sim\theta_{2},\quad\mathbf{H}(1)\sim\theta_{2}^{2},\quad\mathfrak{L}_{2}\sim|x|^{2}. (7.39)
  6. (vi)

    It holds that

    |u22𝐰¯|=Υ(|θ|)|θ||θ2|𝐰¯2|θ2|ϵ2𝐰¯2|π|θ||𝐇(1)32𝐇(1)32.|u_{2}-2\mathbf{\overline{w}}|=-\frac{\Upsilon^{\prime}(|\theta|)}{|\theta|}\,|\theta_{2}|\,\mathbf{\overline{w}}^{2}\sim|\theta_{2}|\,\epsilon^{-2}\,\mathbf{\overline{w}}^{2}\sim|\pi-|\theta|\,|\,\mathbf{H}(1)^{\frac{3}{2}}\lesssim\mathbf{H}(1)^{\frac{3}{2}}. (7.40)
Proof.

Recall that 𝐇(1)=𝒟(u;𝐬¯)\mathbf{H}(1)=\mathcal{D}(u;\mathbf{\overline{s}}), ϵ=ϑ1|θ|\epsilon=\vartheta_{1}-|\theta|, 𝐰¯=θ2ψ(|θ|)\mathbf{\overline{w}}=\theta_{2}\,\psi(|\theta|) and d(g)2|x|2(1+|u|)d(g)^{2}\sim|x|^{2}\,(1+|u|) (cf. (2.7)).

For item (i), let us begin with the proof of 𝒟(u;𝐬¯)u1/θ1\mathcal{D}(u;\mathbf{\overline{s}})\sim u_{1}/\theta_{1}. By the penultimate equality in (2.36), it remains to prove that the continuous function φ2(r)r=r2(r2sin2r)/h(r)1\varphi_{2}(r)\,r=r^{2}\,(r^{2}-\sin^{2}{r})/h(r)\sim 1 for all 0<rϑ10<r\leq\vartheta_{1}. In fact, it follows from Corollary 3.6 that φ2(r)r>0\varphi_{2}(r)\,r>0. Next, a simple calculation, via Taylor’s expansion at 0 for r2(r2sin2r)r^{2}\,(r^{2}-\sin^{2}{r}) and h(r)h(r), shows that limr0+φ2(r)r>0\lim\limits_{r\to 0^{+}}\varphi_{2}(r)\,r>0, which implies the desired result. Similarly, using the third and the second “==” in (2.36) successively, we get that

𝒟(u;𝐬¯)uθ=|uθ|θ22(π|θ|)2.\mathcal{D}(u;\mathbf{\overline{s}})\sim u\cdot\theta=|u\cdot\theta|\sim\theta_{2}^{2}\,(\pi-|\theta|)^{-2}.

On the other hand, applying (3.12) with r=|θ|r=|\theta|, we have that

𝐰¯2=θ22ψ2(|θ|)θ22(π|θ|)2ϵ2,and𝐀(1)=θ22ψ(|θ|)θ22(π|θ|)2|θ|.\mathbf{\overline{w}}^{2}=\theta_{2}^{2}\,\psi^{2}(|\theta|)\sim\frac{\theta_{2}^{2}}{(\pi-|\theta|)^{2}}\,\epsilon^{2},\quad\mbox{and}\quad\mathbf{A}(1)=\theta_{2}^{2}\,\psi^{\prime}(|\theta|)\sim\frac{\theta_{2}^{2}}{(\pi-|\theta|)^{2}}\,|\theta|.

Hence 𝐇(1)ϵ2𝐰¯2|θ|1𝐀(1)\mathbf{H}(1)\sim\epsilon^{-2}\,\mathbf{\overline{w}}^{2}\sim|\theta|^{-1}\mathbf{A}(1).

For item (ii), to show the first claim, it suffices to use the last “=” in (7.36), since it holds that 𝐰¯ϵ|x|1𝔪12\mathbf{\overline{w}}\sim\epsilon\,|x|^{-1}\mathfrak{m}^{\frac{1}{2}} and 𝐀(1)|θ||x|2𝔪\mathbf{A}(1)\sim|\theta|\,|x|^{-2}\mathfrak{m} by item (i) and the fact that 𝐇(1)=4𝔪/|x|2\mathbf{H}(1)=4\mathfrak{m}/|x|^{2}. Next we prove items (iii)-(v) before proving the remaining claims.

For item (iii), it follows from Assumption (A) that

1|u|u1+u2=ψ(|θ|)|θ|θ22θ1+θ2(ψ(|θ|)|θ|θ22+2ψ(|θ|))θ22(π|θ|)2+|θ2||π|θ||,\displaystyle 1\lesssim|u|\sim u_{1}+u_{2}=\frac{\psi^{\prime}(|\theta|)}{|\theta|}\theta_{2}^{2}\,\theta_{1}+\theta_{2}\left(\frac{\psi^{\prime}(|\theta|)}{|\theta|}\theta_{2}^{2}+2\psi(|\theta|)\right)\lesssim\frac{\theta_{2}^{2}}{(\pi-|\theta|)^{2}}+\frac{|\theta_{2}|}{|\pi-|\theta|\,|}, (7.41)

where we have used (3.12) in the last inequality. Hence |θ|1|\theta|\sim 1 and |θ2|/|π|θ||1|\theta_{2}|/|\pi-|\theta|\,|\gtrsim 1.

In the case |θ|<π|\theta|<\pi (so ϵ1\epsilon\sim 1), by (3.26) and (3.27), the estimate (7.41) can be improved as

|u|θ22(π|θ|)2+θ2π|θ|θ22(π|θ|)2.\displaystyle|u|\sim\frac{\theta_{2}^{2}}{(\pi-|\theta|)^{2}}+\frac{\theta_{2}}{\pi-|\theta|}\sim\frac{\theta_{2}^{2}}{(\pi-|\theta|)^{2}}.

In the opposite case π<|θ|<ϑ1\pi<|\theta|<\vartheta_{1}, it deduces from (3.28) that ϵ/(|θ|π)θ22/(|θ|π)21\epsilon/(|\theta|-\pi)\gtrsim\theta_{2}^{2}/(|\theta|-\pi)^{2}\gtrsim 1, and thus ϵ1\epsilon\sim 1. Using (3.28) again we obtain u2|θ2|/(|θ|π)u1u_{2}\lesssim|\theta_{2}|/(|\theta|-\pi)\sim\sqrt{u_{1}} by (3.26) and the fact that θ11\theta_{1}\sim 1. This also leads to |u|θ22/(|θ|π)2|u|\sim\theta_{2}^{2}/(|\theta|-\pi)^{2}.

In conclusion, we establish the first three estimates in (7.37). The others are clear by (i).

We are in a position to prove item (iv). Indeed by item (i) we have 𝐇(1)|u|\mathbf{H}(1)\lesssim|u| and |θ2||π|θ||𝐇(1)12|\theta_{2}|\sim|\pi-|\theta|\,|\,\mathbf{H}(1)^{\frac{1}{2}}, whence |θ2|1|\theta_{2}|\ll 1 and θ11\theta_{1}\sim 1. By (i) again, it turns out that

𝐇(1)u1θ22(π|θ|)21,𝐰¯ϵu1.\displaystyle\mathbf{H}(1)\sim u_{1}\sim\theta_{2}^{2}\,(\pi-|\theta|)^{-2}\ll 1,\qquad\mathbf{\overline{w}}\sim\epsilon\,\sqrt{u_{1}}. (7.42)

To show the first estimate, consider the following three cases: (1) ϵ1\epsilon\ll 1, (2) ϵ1\epsilon\sim 1 with |θ|<π|\theta|<\pi, and (3) ϵ1\epsilon\sim 1 with |θ|>π|\theta|>\pi. In the first case, it follows from (3.28) that ϵu1+u2|θ2|1\epsilon\sim u_{1}+u_{2}\,|\theta_{2}|^{-1}. And in such case, we remark that |θ2|u112|\theta_{2}|\sim u_{1}^{\frac{1}{2}}, thereby showing that ϵu1+u2u112\epsilon\sim u_{1}+u_{2}\,u_{1}^{-\frac{1}{2}}. In the second case, by (3.27) and the first equation in (7.42), it holds that

u2|θ2||π|θ||(1+|θ2||θ2||π|θ||)|θ2||π|θ||u112,u_{2}\sim\frac{|\theta_{2}|}{|\pi-|\theta|\,|}\left(1+|\theta_{2}|\frac{|\theta_{2}|}{|\pi-|\theta|\,|}\right)\sim\frac{|\theta_{2}|}{|\pi-|\theta|\,|}\sim u_{1}^{\frac{1}{2}},

and therefore, u1+u2u1121ϵu_{1}+u_{2}u_{1}^{-\frac{1}{2}}\sim 1\sim\epsilon. In the third case, the argument is similar except using (3.28) instead.

For item (v), since |θ|3|\theta|\leq 3 (so ϵ1\epsilon\sim 1), then by (3.26) and (3.27) we have u1θ1θ22,u2θ2u_{1}\sim\theta_{1}\theta_{2}^{2},\,u_{2}\sim\theta_{2}. Thus, by item (i) we obtain 𝐇(1)θ22u22\mathbf{H}(1)\sim\theta_{2}^{2}\sim u_{2}^{2}. It follows from the first claim in item (ii) that 𝔏2u2|x|3𝔪12|x|2\mathfrak{L}_{2}\sim u_{2}|x|^{3}\mathfrak{m}^{-\frac{1}{2}}\sim|x|^{2}.

Now we return to the rest claims in item (ii). Recall (7.35). By (7.5) and the trivial fact that sin2rr2(πr)2\sin^{-2}{r}\sim r^{-2}\,(\pi-r)^{-2} for 0<rπ<ϑ10<r\neq\pi<\vartheta_{1}, we obtain

𝔏1|x|2ϵ2|θ|2(θ12+ϵ|π|θ||2θ22).\mathfrak{L}_{1}\sim|x|^{2}\epsilon^{2}|\theta|^{-2}\left(\,\theta_{1}^{2}+\epsilon\,|\pi-|\theta|\,|^{-2}\,\theta_{2}^{2}\,\right). (7.43)

There are two possible cases:

(ii-1) |u|1|u|\gtrsim 1. We have d(g)2|x|2|u|d(g)^{2}\sim|x|^{2}\,|u|. Then combining item (i) with item (iii) yields that ϵ1\epsilon\sim 1, 𝔏1|x|2(θ12+|u|)|x|2|u|ϵ2d(g)2ϵ𝔪\mathfrak{L}_{1}\sim|x|^{2}\,(\theta_{1}^{2}+|u|)\sim|x|^{2}\,|u|\sim\epsilon^{2}\,d(g)^{2}\sim\epsilon\,\mathfrak{m} and 𝔏2|x|2u2|u|12|x|2|u|𝔏1\mathfrak{L}_{2}\sim|x|^{2}\,u_{2}\,|u|^{-\frac{1}{2}}\lesssim|x|^{2}\,|u|\sim\mathfrak{L}_{1}, whence 𝔏2+ϵ𝔪𝔏1\mathfrak{L}_{2}+\epsilon\,\mathfrak{m}\sim\mathfrak{L}_{1}.

(ii-2) |u|1|u|\ll 1. We have d(g)2|x|2d(g)^{2}\sim|x|^{2}. (ii-2a) If |θ|1|\theta|\geq 1, recalling that θ11\theta_{1}\sim 1 (cf. (7.38)), then by (7.43) and the first equation in (7.42), we get 𝔏1|x|2ϵ2(1+ϵu1)ϵ2d(g)2\mathfrak{L}_{1}\sim|x|^{2}\,\epsilon^{2}\,(1+\epsilon\,u_{1})\sim\epsilon^{2}\,d(g)^{2}. By using (7.38) and again the first claim in item (ii), we yield that ϵ𝔪+𝔏2ϵ(|x|2u1+|x|2u2u112)ϵ2|x|2𝔏1\epsilon\,\mathfrak{m}+\mathfrak{L}_{2}\sim\epsilon\,(|x|^{2}\,u_{1}+|x|^{2}\,u_{2}\,u_{1}^{-\frac{1}{2}})\sim\epsilon^{2}\,|x|^{2}\sim\mathfrak{L}_{1}. (ii-2b) If |θ|3|\theta|\leq 3, by (7.43) and item (v) we obtain 𝔏1|x|2|θ|2(θ12+θ22)=|x|2ϵ2d(g)2𝔏2\mathfrak{L}_{1}\sim|x|^{2}\,|\theta|^{-2}\,(\theta_{1}^{2}+\theta_{2}^{2})=|x|^{2}\sim\epsilon^{2}\,d(g)^{2}\sim\mathfrak{L}_{2}, and therefore ϵ𝔪|x|2|u|𝔏1\epsilon\,\mathfrak{m}\lesssim|x|^{2}\,|u|\ll\mathfrak{L}_{1}. Hence item (ii) is fully verified.

Finally, item (vi) follows easily from (4.15), (7.5) and item (i). ∎

Remark 7.7.

For Lemma 7.6 (iv), if we replace the condition “|θ|1|\theta|\geq 1” by “|θ|α0|\theta|\geq\alpha_{0} with α0(0,1]\alpha_{0}\in(0,1]”, then (7.38) still holds with the implicit constants depending additionally on α0\alpha_{0}.

Similarly for Lemma 7.6 (v), if we replace the condition “|θ|3|\theta|\leq 3” by “|θ|β0|\theta|\leq\beta_{0} with β0[3,π)\beta_{0}\in[3,\pi)”, then (7.39) still holds with the implicit constants depending additionally on β0\beta_{0}.

8 Uniform asymptotics for the subtlest case: |θ|1|\theta|\geq 1, 𝔏11\mathfrak{L}_{1}\gg 1 and 𝔪+\mathfrak{m}\to+\infty

Recall that d(g)2=|x|2+4𝔪d(g)^{2}=|x|^{2}+4\mathfrak{m}, 𝔪=|x|24𝐇(1)\mathfrak{m}=\frac{|x|^{2}}{4}\mathbf{H}(1) is the minimum of |x|24𝒟(u;)\frac{|x|^{2}}{4}\mathcal{D}(u;\cdot), and 𝔏1,𝔏2>0\mathfrak{L}_{1},\,\mathfrak{L}_{2}>0 are two eigenvalues of the Hessian matrix of |x|24𝒟(u;)\frac{|x|^{2}}{4}\mathcal{D}(u;\cdot) at its unique minimum point (1,0)(-1,0). Our target in this section is the following:

Theorem 8.1.

Let |θ|1|\theta|\geq 1 and 𝔏1,𝔪ζ0\mathfrak{L}_{1},\mathfrak{m}\geq\zeta_{0} with ζ01\zeta_{0}\gg 1. Then

p(g)=16π2πϑ1𝔮(|θ|)e𝔏2I0(𝔏2)ϵ𝔏1eϑ1|x|2𝐰¯22ϵI0(ϑ1|x|2𝐰¯22ϵ)ed(g)24(1+oζ0(1)).p(g)=16\pi^{2}\sqrt{\pi\vartheta_{1}}\,\mathfrak{q}(|\theta|)\,\frac{e^{-\mathfrak{L}_{2}}\,I_{0}(\mathfrak{L}_{2})}{\sqrt{\epsilon\,\mathfrak{L}_{1}}}\,e^{-\frac{\vartheta_{1}|x|^{2}\,\mathbf{\overline{w}}^{2}}{2\epsilon}}\,I_{0}\left(\frac{\vartheta_{1}|x|^{2}\,\mathbf{\overline{w}}^{2}}{2\epsilon}\right)\,e^{-\frac{d(g)^{2}}{4}}\,(1+o_{\zeta_{0}}(1)).

Its proof is based on Propositions 2.8 and 2.9. As mentioned in Subsection 2.5, Laplace’s method is not enough for our purpose, since 𝔏2\mathfrak{L}_{2} could be bounded. To prove it in a uniform way, we will use Proposition 8.2 below, via the modified polar coordinates provided by (7.19). In short, we introduce a suitable coordinate system, give a nice expression for the integrand (especially, (8.1) for the phase), and determine the main contributing region 𝒲1{\mathcal{W}}_{1} of the integral.

Structure of 𝒲1\boldsymbol{{\mathcal{W}}}_{1}. This set depends on two parameters δ\delta and η\eta (cf. (8.2) below), which in fact involves three different cases. To avoid redundant computation, we list some preparatory estimates for the aforementioned parameters in the following Table 1. These results, which can be checked without difficulties by Lemma 7.6 (i)-(iv) and the fact that d(g)2|x|2(1+|u|)d(g)^{2}\sim|x|^{2}\,(1+|u|), will be used iteratively throughout the proof of Theorem 8.1.

More explanations for Table 1:

  • The condition for all estimates in Table 1 to hold is the same as that in Theorem 8.1, i.e., |θ|1|\theta|\geq 1, 𝔏11\mathfrak{L}_{1}\gg 1 and 𝔪1\mathfrak{m}\gg 1, while the last two columns need an additional hypothesis that 𝔏21\mathfrak{L}_{2}\gg 1.

  • These estimates should be used in the sense of “\sim”. For instance, “1” in the ϵ\epsilon-Column means ϵ1\epsilon\sim 1.

Table 1: Preparatory estimates for Theorem 8.1
Case Condition ϵ\epsilon 𝐇(1)\mathbf{H}(1) δ\delta 𝔏1δ2\mathfrak{L}_{1}\,\delta^{2} η\eta 𝔏2η2\mathfrak{L}_{2}\,\eta^{2}
(1) |u|1|u|\gtrsim 1 1 |u||u| 𝔏138\mathfrak{L}_{1}^{-\frac{3}{8}} 𝔪14\mathfrak{m}^{\frac{1}{4}} 𝔏238\mathfrak{L}_{2}^{-\frac{3}{8}} 𝔏214\mathfrak{L}_{2}^{\frac{1}{4}}
(2) |u|1|u|\ll 1, u2u1u_{2}\gtrsim u_{1} u2u112u_{2}\,u_{1}^{-\frac{1}{2}} u1u_{1} |x|34u21u158|x|^{-\frac{3}{4}}\,u_{2}^{-1}\,u_{1}^{\frac{5}{8}} 𝔪14\mathfrak{m}^{\frac{1}{4}} |x|34u21u158|x|^{-\frac{3}{4}}\,u_{2}^{-1}\,u_{1}^{\frac{5}{8}} 𝔪14\mathfrak{m}^{\frac{1}{4}}
(3) |u|1|u|\ll 1, u2u1u_{2}\ll u_{1} u1+u2u1u_{1}+\frac{u_{2}}{\sqrt{u_{1}}} u1u_{1} 𝔏138\mathfrak{L}_{1}^{-\frac{3}{8}} 𝔏114\mathfrak{L}_{1}^{\frac{1}{4}} 𝔏238\mathfrak{L}_{2}^{-\frac{3}{8}} 𝔏214\mathfrak{L}_{2}^{\frac{1}{4}}
Proposition 8.2.

Let |θ|1|\theta|\geq 1 and 𝔏1,𝔪ζ0\mathfrak{L}_{1},\mathfrak{m}\geq\zeta_{0} with ζ01\zeta_{0}\gg 1. Then it holds that

𝒟p(w,γ)=𝐇(1)+12𝐇′′(1)(w1)2+4𝔏2|x|2(1cosγ)+oζ0(|x|2),(w,γ)𝒲1,\mathcal{D}_{\mathrm{p}}(\mathrm{w},\gamma)=\mathbf{H}(1)+\frac{1}{2}\mathbf{H}^{\prime\prime}(1)(\mathrm{w}-1)^{2}+\frac{4\mathfrak{L}_{2}}{|x|^{2}}(1-\cos{\gamma})+o_{\zeta_{0}}(|x|^{-2}),\quad\forall(\mathrm{w},\gamma)\in{\mathcal{W}}_{1}, (8.1)

where 𝒲1:={(w,γ);|w1|δ,|γ|η}{\mathcal{W}}_{1}:=\{(\mathrm{w},\gamma);|\mathrm{w}-1|\leq\delta,\,|\gamma|\leq\eta\} with

δ=δ(g):=𝔏138(𝐇(1)|u|)141,η=η(g):={𝔏238(𝐇(1)|u|)141,𝔏21,π,𝔏21.\delta=\delta(g):=\mathfrak{L}_{1}^{-\frac{3}{8}}\left(\frac{\mathbf{H}(1)}{|u|}\right)^{\frac{1}{4}}\ll 1,\quad\eta=\eta(g):=\begin{cases}\mathfrak{L}_{2}^{-\frac{3}{8}}\left(\frac{\mathbf{H}(1)}{|u|}\right)^{\frac{1}{4}}\ll 1,&\mathfrak{L}_{2}\gg 1,\\ \pi,&\mathfrak{L}_{2}\lesssim 1.\end{cases} (8.2)

Moreover, we have uniformly on 𝒲1{\mathcal{W}}_{1} (cf. (2.31), (2.33) and the convention (2.18) for pertinent definitions)

w𝒫p(w,γ)\displaystyle\mathrm{w}\,\mathcal{P}_{\mathrm{p}}(\mathrm{w},\gamma) =2ϑ116π2𝔮(|θ|)ϵ12|x|2𝐰¯2exp(𝔏2ϑ1|x|2𝐰¯22ϵ)I0(ϑ1|x|2𝐰¯22ϵ)e𝔪\displaystyle=\sqrt{2\vartheta_{1}}16\pi^{2}\,\mathfrak{q}(|\theta|)\,\frac{\epsilon^{-\frac{1}{2}}}{|x|^{2}\,\mathbf{\overline{w}}^{2}}\,\exp\left(-\mathfrak{L}_{2}-\frac{\vartheta_{1}|x|^{2}\,\mathbf{\overline{w}}^{2}}{2\epsilon}\right)\,I_{0}\left(\frac{\vartheta_{1}|x|^{2}\,\mathbf{\overline{w}}^{2}}{2\epsilon}\right)e^{-\mathfrak{m}} (8.3)
exp(12𝔏1(w1)2+𝔏2cosγ)(1+oζ0(1)).\displaystyle\qquad\cdot\exp\left(-\frac{1}{2}\mathfrak{L}_{1}(\mathrm{w}-1)^{2}+\mathfrak{L}_{2}\cos{\gamma}\right)\,(1+o_{\zeta_{0}}(1)).

The proof of Proposition 8.2 is postponed to Subsection 8.3 below. At this point we see how to apply it to obtain Theorem 8.1. For the sake of brevity, from now on we will omit the dependence of undetermined constants in the proof.

Let us decompose (0,+)×(π,π)=i=13𝒲i(0,+\infty)\times(-\pi,\pi)=\bigcup\limits_{i=1}^{3}{\mathcal{W}}_{i} with 𝒲1{\mathcal{W}}_{1} defined as in Proposition 8.2 and

𝒲2:={(w,γ);w<C|u|d(gu)1𝐰¯1}𝒲1,𝒲3:={(w,γ);wC|u|d(gu)1𝐰¯1},{\mathcal{W}}_{2}:=\{(\mathrm{w},\gamma);\mathrm{w}<C|u|\,d(g_{u})^{-1}\,\mathbf{\overline{w}}^{-1}\}\setminus{\mathcal{W}}_{1},\quad{\mathcal{W}}_{3}:=\{(\mathrm{w},\gamma);\mathrm{w}\geq C|u|\,d(g_{u})^{-1}\,\mathbf{\overline{w}}^{-1}\}, (8.4)

where the universal constant CC will be chosen later. Then by (2.30) and (7.19) we can write

p(g)\displaystyle p(g) =14πe|x|24|x|2𝐰¯20+ππw𝒫p(w,γ)𝑑w𝑑γ\displaystyle=\frac{1}{4\pi}\,e^{-\frac{|x|^{2}}{4}}|x|^{2}\,\mathbf{\overline{w}}^{2}\int_{0}^{+\infty}\int_{-\pi}^{\pi}\mathrm{w}\,\mathcal{P}_{\mathrm{p}}(\mathrm{w},\gamma)\,d\mathrm{w}d\gamma
=14πe|x|24i=13𝒲i|x|2𝐰¯2w𝒫p(w,γ)dwdγ=:14πe|x|24i=13𝒬i.\displaystyle=\frac{1}{4\pi}\,e^{-\frac{|x|^{2}}{4}}\sum_{i=1}^{3}\int_{{\mathcal{W}}_{i}}|x|^{2}\,\mathbf{\overline{w}}^{2}\,\mathrm{w}\,\mathcal{P}_{\mathrm{p}}(\mathrm{w},\gamma)\,d\mathrm{w}d\gamma=:\frac{1}{4\pi}\,e^{-\frac{|x|^{2}}{4}}\sum_{i=1}^{3}\mathcal{Q}_{i}. (8.5)

We will see that the main contribution comes from 𝒬1\mathcal{Q}_{1}.

Estimate of 𝒬1\mathcal{Q}_{1}. We use (8.3) to treat 𝒬1\mathcal{Q}_{1}. Under our assumptions, observe that 𝔏1δ2min{𝔪,𝔏1}1/4ζ01/41\mathfrak{L}_{1}\delta^{2}\gtrsim\min\{\mathfrak{m},\mathfrak{L}_{1}\}^{1/4}\geq\zeta_{0}^{1/4}\gg 1. Then the standard Laplace’s method implies that:

1δ1+δe12𝔏1(w1)2𝑑w=2π𝔏1𝔏1δ22π𝔏1δ22πeπr2𝑑r=2π𝔏1(1+O(ec𝔏1δ2)).\displaystyle\int_{1-\delta}^{1+\delta}e^{-\frac{1}{2}\mathfrak{L}_{1}(\mathrm{w}-1)^{2}}d\mathrm{w}=\sqrt{\frac{2\pi}{\mathfrak{L}_{1}}}\int_{-\sqrt{\frac{\mathfrak{L}_{1}\delta^{2}}{2\pi}}}^{\sqrt{\frac{\mathfrak{L}_{1}\delta^{2}}{2\pi}}}e^{-\pi r^{2}}dr=\sqrt{\frac{2\pi}{\mathfrak{L}_{1}}}\,\left(1+O(e^{-c\,\mathfrak{L}_{1}\,\delta^{2}})\right).

On the other hand, if 𝔏21\mathfrak{L}_{2}\lesssim 1, then η=π\eta=\pi and therefore,

|γ|<ηe𝔏2cosγ𝑑γ=2πI0(𝔏2).\int_{|\gamma|<\eta}e^{\mathfrak{L}_{2}\cos\gamma}d\gamma=2\pi I_{0}(\mathfrak{L}_{2}).

If 𝔏21\mathfrak{L}_{2}\gg 1, notice that 𝔏2η2min{𝔏2,𝔪}1/41\mathfrak{L}_{2}\,\eta^{2}\gtrsim\min\{\mathfrak{L}_{2},\mathfrak{m}\}^{1/4}\gg 1 (cf. also Table 1). Then Lemma 7.1 gives:

|γ|<ηe𝔏2cosγ𝑑γ=2πI0(η;𝔏2)=2πI0(𝔏2)(1+o(1)).\int_{|\gamma|<\eta}e^{\mathfrak{L}_{2}\cos\gamma}d\gamma=2\pi I_{0}(\eta;\mathfrak{L}_{2})=2\pi I_{0}(\mathfrak{L}_{2})\,(1+o(1)).

Combining all the estimates above yields that

𝒬1\displaystyle\mathcal{Q}_{1} =64π3πϑ1𝔮(|θ|)e𝔏2I0(𝔏2)ϵ𝔏1eϑ1|x|2𝐰¯22ϵI0(ϑ1|x|2𝐰¯22ϵ)e𝔪(1+o(1))\displaystyle=64\,\pi^{3}\,\sqrt{\pi\vartheta_{1}}\,\mathfrak{q}(|\theta|)\,\frac{e^{-\mathfrak{L}_{2}}I_{0}(\mathfrak{L}_{2})}{\sqrt{\epsilon\,\mathfrak{L}_{1}}}e^{-\frac{\vartheta_{1}|x|^{2}\,\mathbf{\overline{w}}^{2}}{2\epsilon}}\,I_{0}\left(\frac{\vartheta_{1}|x|^{2}\,\mathbf{\overline{w}}^{2}}{2\epsilon}\right)e^{-\mathfrak{m}}(1+o(1))
ϵ2𝔏112(1+ϵ𝔪)12(1+𝔏2)12e𝔪,\displaystyle\sim\epsilon^{2}\,\mathfrak{L}_{1}^{-\frac{1}{2}}\,(1+\epsilon\,\mathfrak{m})^{-\frac{1}{2}}(1+\mathfrak{L}_{2})^{-\frac{1}{2}}\,e^{-\mathfrak{m}}, (8.6)

where we have used in “\sim” (7.15), (7.3) and the fact that |x|2𝐰¯2/ϵϵ𝔪|x|^{2}\,\mathbf{\overline{w}}^{2}/\epsilon\sim\epsilon\,\mathfrak{m} (cf. Lemma 7.6 (i) with the definition of 𝔪\mathfrak{m} (7.34)).

As a consequence, to finish the proof of Theorem 8.1, it is enough to show that

𝒬2+𝒬3=o(𝒬1).\mathcal{Q}_{2}+\mathcal{Q}_{3}=o(\mathcal{Q}_{1}). (8.7)

This is done in the following two cases, |u|1|u|\gtrsim 1 and |u|1|u|\ll 1, which will be treated in Subsections 8.1 and 8.2, respectively.

The following estimates will be used repeatedly hereafter:

𝒟p𝐰¯2w2+𝐀~p,𝒫p𝔪1e|x|24𝒟p,(w,γ).\mathcal{D}_{\mathrm{p}}\sim\mathbf{\overline{w}}^{2}\,\mathrm{w}^{2}+\widetilde{\mathbf{A}}_{\mathrm{p}},\qquad\mathcal{P}_{\mathrm{p}}\lesssim\mathfrak{m}^{-1}\,e^{-\frac{|x|^{2}}{4}\mathcal{D}_{\mathrm{p}}},\quad\forall\,(\mathrm{w},\gamma). (8.8)

Indeed, by recalling the definition of 𝒟\mathcal{D} (cf. (2.35)), 𝐀~\widetilde{\mathbf{A}} (cf. (7.16)) and 𝒫\mathcal{P} (cf. (2.31)) (with the coordinates (7.19)), the first estimate comes from (4.14), and the second one from Proposition 2.9 (ii) and the fact that the minimum of |x|2𝒟/4|x|^{2}\,\mathcal{D}/4 equals to |x|2𝐇(1)/4=𝔪1|x|^{2}\,\mathbf{H}(1)/4=\mathfrak{m}\gg 1 (cf. (7.34) again and Theorem 2.10).

8.1 Proof of (8.7) in the case where |u|1|u|\gtrsim 1

In this case, notice that d(gu)21+|u||u|d(g_{u})^{2}\sim 1+|u|\sim|u| (cf. (2.7)), so d(g)2|x|2|u|d(g)^{2}\sim|x|^{2}\,|u|. First from Table 1 with Lemma 7.6 (i)-(iii), we remark that

ϵ1,d(g)2𝔏1𝔪|x|2𝐰¯2|x|2|u|1,𝔏2𝔪,𝔏1δ2𝔪141.\displaystyle\epsilon\sim 1,\ d(g)^{2}\sim\mathfrak{L}_{1}\sim\mathfrak{m}\sim|x|^{2}\,\mathbf{\overline{w}}^{2}\sim|x|^{2}\,|u|\gg 1,\ \mathfrak{L}_{2}\lesssim\mathfrak{m},\ \mathfrak{L}_{1}\delta^{2}\sim\mathfrak{m}^{\frac{1}{4}}\gg 1. (8.9)

Then (8.6) can be simplified as

𝒬1(1+𝔏2)12𝔪1e𝔪𝔪32e𝔪.\mathcal{Q}_{1}\sim(1+\mathfrak{L}_{2})^{-\frac{1}{2}}\,\mathfrak{m}^{-1}\,e^{-\mathfrak{m}}\gtrsim\mathfrak{m}^{-\frac{3}{2}}\,e^{-\mathfrak{m}}. (8.10)

Estimate of 𝒬3\mathcal{Q}_{3}. On 𝒲3{\mathcal{W}}_{3}, by (8.8) we obtain |x|2𝒟p|x|2𝐰¯2w2C2𝔪|x|^{2}\,\mathcal{D}_{\mathrm{p}}\gtrsim|x|^{2}\,\mathbf{\overline{w}}^{2}\,\mathrm{w}^{2}\gtrsim C^{2}\,\mathfrak{m}, whence there is a positive constant cc such that |x|24𝒟p𝔪c𝔪+c|x|2𝐰¯2w2\frac{|x|^{2}}{4}\mathcal{D}_{\mathrm{p}}-\mathfrak{m}\geq c\,\mathfrak{m}+c\,|x|^{2}\,\mathbf{\overline{w}}^{2}\,\mathrm{w}^{2} by selecting CC large enough. Combining this with the second claim in (8.8), we get that

𝒬3|x|2𝐰¯2𝔪1e𝔪ec𝔪0ec|x|2𝐰¯2w2w𝑑w𝔪1e𝔪ec𝔪=𝒬1o(ec𝔪).\mathcal{Q}_{3}\lesssim|x|^{2}\,\mathbf{\overline{w}}^{2}\,\mathfrak{m}^{-1}\,e^{-\mathfrak{m}}\,e^{-c\,\mathfrak{m}}\int_{0}^{\infty}e^{-c\,|x|^{2}\,\mathbf{\overline{w}}^{2}\,\mathrm{w}^{2}}\,\mathrm{w}\,d\mathrm{w}\sim\mathfrak{m}^{-1}\,e^{-\mathfrak{m}}\,e^{-c\,\mathfrak{m}}=\mathcal{Q}_{1}\,o(e^{-c^{\prime}\,\mathfrak{m}}).

Estimate of 𝒬2\mathcal{Q}_{2}. We have to deal with two kinds of subcases: 𝔏21\mathfrak{L}_{2}\lesssim 1 and 𝔏21\mathfrak{L}_{2}\gg 1. For 𝔏21\mathfrak{L}_{2}\lesssim 1, recalling η=π\eta=\pi, then by applying Lemma 7.5 and (8.1) (with γ=0\gamma=0) successively, we have

|x|24𝒟p|x|24min{𝐇(1δ),𝐇(1+δ)}=𝔪+12𝔏1δ2+o(1)𝔪+c𝔪14\frac{|x|^{2}}{4}\mathcal{D}_{\mathrm{p}}\geq\frac{|x|^{2}}{4}\min\{\mathbf{H}(1-\delta),\mathbf{H}(1+\delta)\}=\mathfrak{m}+\frac{1}{2}\mathfrak{L}_{1}\delta^{2}+o(1)\geq\mathfrak{m}+c\,\mathfrak{m}^{\frac{1}{4}} (8.11)

on 𝒲2{\mathcal{W}}_{2} for some positive constant cc. Hence by the second estimation in (8.8) again, the fact that d(gu)2|u|𝐰¯2d(g_{u})^{2}\sim|u|\sim\mathbf{\overline{w}}^{2} and |x|2𝐰¯2𝔪11|x|^{2}\,\mathbf{\overline{w}}^{2}\mathfrak{m}^{-1}\sim 1 (cf. the second equation in (8.9)), we obtain

𝒬2|x|2𝐰¯2𝔪1e𝔪ec𝔪14𝒲2w𝑑w𝑑γe𝔪ec𝔪14=𝒬1o(exp{c𝔪14}).\mathcal{Q}_{2}\lesssim|x|^{2}\,\mathbf{\overline{w}}^{2}\,\mathfrak{m}^{-1}\,e^{-\mathfrak{m}}\,e^{-c\mathfrak{m}^{\frac{1}{4}}}\int_{{\mathcal{W}}_{2}}\mathrm{w}\,d\mathrm{w}d\gamma\lesssim e^{-\mathfrak{m}}e^{-c\mathfrak{m}^{\frac{1}{4}}}=\mathcal{Q}_{1}\,o(\exp\{-c^{\prime}\,\mathfrak{m}^{\frac{1}{4}}\}).

In the opposite case 𝔏21\mathfrak{L}_{2}\gg 1, we decompose 𝒲2=𝒲2𝒲2′′{\mathcal{W}}_{2}={\mathcal{W}}_{2}^{\prime}\bigcup{\mathcal{W}}_{2}^{\prime\prime} with

𝒲2:={(w,γ);|w1|δ,|γ|>η},𝒲2′′:={(w,γ);|w1|>δ,w<C|u|d(gu)1𝐰¯1},\begin{gathered}{\mathcal{W}}_{2}^{\prime}:=\{(\mathrm{w},\gamma);|\mathrm{w}-1|\leq\delta,\,|\gamma|>\eta\},\\[2.84526pt] {\mathcal{W}}_{2}^{\prime\prime}:=\{(\mathrm{w},\gamma);|\mathrm{w}-1|>\delta,\,\mathrm{w}<C\,|u|\,d(g_{u})^{-1}\,\mathbf{\overline{w}}^{-1}\},\end{gathered} (8.12)

on which the integrals of |x|2𝐰¯2w𝒫p|x|^{2}\,\mathbf{\overline{w}}^{2}\,\mathrm{w}\,\mathcal{P}_{\mathrm{p}} are denoted by 𝒬2\mathcal{Q}_{2}^{\prime} and 𝒬2′′\mathcal{Q}_{2}^{\prime\prime} respectively. Notation for such decomposition will be used repeatedly in the rest of this section without particular statements.

The same argument as in the case 𝔏21\mathfrak{L}_{2}\lesssim 1 yields 𝒬2′′=𝒬1o(ec𝔪14)\mathcal{Q}_{2}^{\prime\prime}=\mathcal{Q}_{1}\,o(e^{-c^{\prime}\,\mathfrak{m}^{\frac{1}{4}}}).

To estimate 𝒬2\mathcal{Q}_{2}^{\prime}, first notice that 𝔏2η2𝔏2141\mathfrak{L}_{2}\,\eta^{2}\sim\mathfrak{L}_{2}^{\frac{1}{4}}\gg 1. Similar to the proof of (8.11), there exists some positive constant cc such that

|x|24𝒟p(w,γ)|x|24𝒟p(w,η)𝔪+𝔏12(w1)2+c𝔏214,(w,γ)𝒲2.\frac{|x|^{2}}{4}\mathcal{D}_{\mathrm{p}}(\mathrm{w},\gamma)\geq\frac{|x|^{2}}{4}\mathcal{D}_{\mathrm{p}}(\mathrm{w},\eta)\geq\mathfrak{m}+\frac{\mathfrak{L}_{1}}{2}(\mathrm{w}-1)^{2}+c\,\mathfrak{L}_{2}^{\frac{1}{4}},\qquad\forall(\mathrm{w},\gamma)\in{\mathcal{W}}_{2}^{\prime}. (8.13)

Combining this with |x|2𝐰¯2𝔪|x|^{2}\,\mathbf{\overline{w}}^{2}\sim\mathfrak{m} and Proposition 2.9 (ii), we conclude

w𝒫p(w,γ)𝔪3/2exp{𝔪𝔏12(w1)2c𝔏214},\mathrm{w}\,\mathcal{P}_{\mathrm{p}}(\mathrm{w},\gamma)\lesssim\mathfrak{m}^{-3/2}\,\exp\left\{-\mathfrak{m}-\frac{\mathfrak{L}_{1}}{2}(\mathrm{w}-1)^{2}-c\,\mathfrak{L}_{2}^{\frac{1}{4}}\right\}, (8.14)

which implies immediately that 𝒬2=𝒬1o(exp{c𝔏214})\mathcal{Q}_{2}^{\prime}=\mathcal{Q}_{1}\,o(\exp\{-c^{\prime}\,\mathfrak{L}_{2}^{\frac{1}{4}}\}) because of the estimate 𝔏1𝔪\mathfrak{L}_{1}\sim\mathfrak{m} and (8.10).

This completes the proof of (8.7) under our assumptions.

8.2 Proof of (8.7) in the case where |u|1|u|\ll 1

In such case, d(gu)1d(g_{u})\sim 1, d(g)|x|1d(g)\sim|x|\gg 1, 𝐇(1)u1\mathbf{H}(1)\sim u_{1} and 𝔪|x|2u11\mathfrak{m}\sim|x|^{2}u_{1}\gg 1 (cf. (7.38)). Moreover, Lemma 7.6 (ii) says that 𝔏2𝔏1\mathfrak{L}_{2}\lesssim\mathfrak{L}_{1} and 𝔏1ϵ2|x|21\mathfrak{L}_{1}\sim\epsilon^{2}\,|x|^{2}\gg 1. Thus it follows from (8.6) that

𝒬1ϵ2𝔏11𝔪12e𝔪|x|2𝔪12e𝔪.\mathcal{Q}_{1}\gtrsim\epsilon^{2}\,\mathfrak{L}_{1}^{-1}\,\mathfrak{m}^{-\frac{1}{2}}\,e^{-\mathfrak{m}}\sim|x|^{-2}\,\mathfrak{m}^{-\frac{1}{2}}\,e^{-\mathfrak{m}}. (8.15)

Estimate of 𝒬3\mathcal{Q}_{3}. It’s very similar to the case |u|1|u|\gtrsim 1. Since u2|u||u|d(gu)1u_{2}\leq|u|\sim|u|\,d(g_{u})^{-1}, then on 𝒲3{\mathcal{W}}_{3}, we can choose CC large enough so that u2<C|u|d(gu)1/2𝐰¯w/2u_{2}<C|u|d(g_{u})^{-1}/2\leq\mathbf{\overline{w}}\mathrm{w}/2. Hence it deduces from (7.33) that 𝐀~p2|u|2+4𝐰¯w(𝐰¯wu2)|u|2+2𝐰¯2w2\widetilde{\mathbf{A}}_{\mathrm{p}}^{2}\geq|u|^{2}+4\mathbf{\overline{w}}\mathrm{w}(\mathbf{\overline{w}}\mathrm{w}-u_{2})\geq|u|^{2}+2\mathbf{\overline{w}}^{2}\mathrm{w}^{2}, and therefore, 𝔪|x|2|u||x|2𝐰¯w/C|x|2𝐀~p/C\mathfrak{m}\lesssim|x|^{2}|u|\lesssim|x|^{2}\mathbf{\overline{w}}\mathrm{w}/C\lesssim|x|^{2}\widetilde{\mathbf{A}}_{\mathrm{p}}/C. As a result, with CC large enough, by the first estimate in (8.8) we have

|x|24𝒟p(w,γ)>2𝔪+c|x|2𝐰¯w,(w,γ)𝒲3.\frac{|x|^{2}}{4}\mathcal{D}_{\mathrm{p}}(\mathrm{w},\gamma)>2\mathfrak{m}+c\,|x|^{2}\,\mathbf{\overline{w}}\,\mathrm{w},\qquad\forall(\mathrm{w},\gamma)\in{\mathcal{W}}_{3}.

Consequently, combining this with the second estimate in (8.8) and (8.15), we yield

𝒬3|x|2𝐰¯2𝔪1e2𝔪𝒲3ec|x|2𝐰¯ww𝑑w𝑑γ|x|2𝔪1e2𝔪=𝒬1o(e𝔪).\mathcal{Q}_{3}\lesssim|x|^{2}\,\mathbf{\overline{w}}^{2}\,\mathfrak{m}^{-1}e^{-2\,\mathfrak{m}}\,\int_{{\mathcal{W}}_{3}}e^{-c\,|x|^{2}\,\mathbf{\overline{w}}\,\mathrm{w}}\,\mathrm{w}\,d\mathrm{w}d\gamma\lesssim|x|^{-2}\mathfrak{m}^{-1}e^{-2\,\mathfrak{m}}=\mathcal{Q}_{1}\,o(e^{-\mathfrak{m}}). (8.16)

Estimate of 𝒬2\mathcal{Q}_{2}. We divide this remaining estimate into two cases.

Case I. u2u1.u_{2}\gtrsim u_{1}. By Table 1 with Lemma 7.6 (i)-(ii), we have ϵu2u112\epsilon\sim u_{2}\,u_{1}^{-\frac{1}{2}}, 𝔏1|x|2u22u11𝔏21\mathfrak{L}_{1}\sim|x|^{2}\,u_{2}^{2}\,u_{1}^{-1}\sim\mathfrak{L}_{2}\gg 1 and 𝔏1δ2𝔏2η2𝔪141\mathfrak{L}_{1}\delta^{2}\sim\mathfrak{L}_{2}\eta^{2}\sim\mathfrak{m}^{\frac{1}{4}}\gg 1. Now we write 𝒲2=𝒲2𝒲2{\mathcal{W}}_{2}={\mathcal{W}}_{2}^{*}\bigcup{\mathcal{W}}_{2}^{**} in the rectangular coordinates with

𝒲2={s𝒲2;|s1+1|<Cu1𝐰¯1,|s2|<Cu1𝐰¯1},\displaystyle{\mathcal{W}}_{2}^{*}=\{s\in{\mathcal{W}}_{2};|s_{1}+1|<C_{*}u_{1}\mathbf{\overline{w}}^{-1},|s_{2}|<C_{*}u_{1}\mathbf{\overline{w}}^{-1}\},
𝒲2={s𝒲2;|s1+1|Cu1𝐰¯1or|s2|Cu1𝐰¯1},\displaystyle{\mathcal{W}}_{2}^{**}=\{s\in{\mathcal{W}}_{2};|s_{1}+1|\geq C_{*}u_{1}\mathbf{\overline{w}}^{-1}\,\,{\rm or}\,\,|s_{2}|\geq C_{*}u_{1}\mathbf{\overline{w}}^{-1}\},

where the constant C1C_{*}\geq 1 will be determined later.

As in the proof of (8.11) and (8.13), there exists a constant c>0c>0 such that for all (w,γ)𝒲2(\mathrm{w},\gamma)\in{\mathcal{W}}_{2}^{*},

|x|24𝒟p(w,γ)|x|24inf(w,γ)𝒲1𝒟p(w,γ)𝔪+c𝔪14.\frac{|x|^{2}}{4}\mathcal{D}_{\mathrm{p}}(\mathrm{w},\gamma)\geq\frac{|x|^{2}}{4}\inf_{(\mathrm{w},\gamma)\in\partial{\mathcal{W}}_{1}}\mathcal{D}_{\mathrm{p}}(\mathrm{w},\gamma)\geq\mathfrak{m}+c\,\mathfrak{m}^{\frac{1}{4}}.

As a consequence,

𝒬2|x|2𝐰¯2𝔪1e𝔪ec𝔪14𝒲2𝑑s|x|2𝔪e𝔪ec𝔪14=𝒬1o(ec𝔪14)\mathcal{Q}_{2}^{*}\lesssim|x|^{2}\,\mathbf{\overline{w}}^{2}\,\mathfrak{m}^{-1}\,e^{-\mathfrak{m}}\,e^{-c\,\mathfrak{m}^{\frac{1}{4}}}\int_{{\mathcal{W}}_{2}^{*}}\,ds\lesssim|x|^{-2}\,\mathfrak{m}\,e^{-\mathfrak{m}}\,e^{-c\,\mathfrak{m}^{\frac{1}{4}}}=\mathcal{Q}_{1}\,o(e^{-c^{\prime}\,\mathfrak{m}^{\frac{1}{4}}})

where we have used 𝔪|x|2u1\mathfrak{m}\sim|x|^{2}\,u_{1} in the second “\lesssim” and (8.15) in “==”.

To deal with 𝒬2\mathcal{Q}_{2}^{**}, we use an argument similar to that in the proof of (8.16). In brief, on 𝒲2{\mathcal{W}}_{2}^{**}, recalling that |u22𝐰¯|u1|u_{2}-2\,\mathbf{\overline{w}}|\ll u_{1} (from (7.40) and 𝐇(1)u11\mathbf{H}(1)\sim u_{1}\ll 1), we obtain:

𝐀~p2\displaystyle\widetilde{\mathbf{A}}_{\mathrm{p}}^{2} =u12+(u2+2𝐰¯s1)2+4𝐰¯2s22=u12+(u22𝐰¯+2𝐰¯(s1+1))2+4𝐰¯2s22\displaystyle=u_{1}^{2}+(u_{2}+2\,\mathbf{\overline{w}}s_{1})^{2}+4\mathbf{\overline{w}}^{2}\,s_{2}^{2}=u_{1}^{2}+(u_{2}-2\,\mathbf{\overline{w}}+2\,\mathbf{\overline{w}}\,(s_{1}+1))^{2}+4\mathbf{\overline{w}}^{2}\,s_{2}^{2}
(2𝐰¯|s1+1||u22𝐰¯|)2+4𝐰¯2s22(𝐰¯|s1+1|)2+4𝐰¯2s22C2u12,\displaystyle\geq\left(2\,\mathbf{\overline{w}}\,|s_{1}+1|-|u_{2}-2\,\mathbf{\overline{w}}|\right)^{2}+4\mathbf{\overline{w}}^{2}\,s_{2}^{2}\geq(\mathbf{\overline{w}}\,|s_{1}+1|)^{2}+4\mathbf{\overline{w}}^{2}\,s_{2}^{2}\geq C_{*}^{2}u_{1}^{2},

for CC_{*} large enough. Then by the first estimate in (8.8), there is a constant c>0c>0 such that

|x|24𝒟p2𝔪+c|x|2(𝐰¯|s1+1|+2𝐰¯|s2|),\frac{|x|^{2}}{4}\,\mathcal{D}_{\mathrm{p}}\geq 2\,\mathfrak{m}+c\,|x|^{2}\,(\mathbf{\overline{w}}\,|s_{1}+1|+2\,\mathbf{\overline{w}}|s_{2}|),

which further implies that

𝒬2|x|2𝐰¯2𝔪1e2𝔪𝒲2ec|x|2(𝐰¯|s1+1|+2𝐰¯|s2|)𝑑s|x|2𝔪1e2𝔪=𝒬1o(e𝔪).\mathcal{Q}_{2}^{**}\lesssim|x|^{2}\,\mathbf{\overline{w}}^{2}\,\mathfrak{m}^{-1}\,e^{-2\,\mathfrak{m}}\int_{{\mathcal{W}}_{2}^{**}}e^{-c\,|x|^{2}\,(\mathbf{\overline{w}}\,|s_{1}+1|+2\,\mathbf{\overline{w}}|s_{2}|)}\,ds\lesssim|x|^{-2}\,\mathfrak{m}^{-1}\,e^{-2\,\mathfrak{m}}=\mathcal{Q}_{1}\,o(e^{-\mathfrak{m}}).

Thus we have shown that 𝒬2=𝒬2+𝒬2=o(𝒬1)\mathcal{Q}_{2}=\mathcal{Q}_{2}^{*}+\mathcal{Q}_{2}^{**}=o(\mathcal{Q}_{1}).

Case II. u2u1.u_{2}\ll u_{1}. From Table 1 we can read that ϵ1\epsilon\ll 1 and 𝔏1δ2𝔏1141\mathfrak{L}_{1}\delta^{2}\sim\mathfrak{L}_{1}^{\frac{1}{4}}\gg 1. There are two subcases as well: 𝔏21\mathfrak{L}_{2}\lesssim 1 and 𝔏21\mathfrak{L}_{2}\gg 1.

Subcase II-1. 𝔏21.\mathfrak{L}_{2}\lesssim 1. In such subcase, by (8.6) and the fact that ϵ𝔪+𝔏2𝔏1ϵ2|x|2\epsilon\,\mathfrak{m}+\mathfrak{L}_{2}\sim\mathfrak{L}_{1}\sim\epsilon^{2}\,|x|^{2} (cf. Lemma 7.6 (ii)), we have

𝒬1ϵ2𝔏132e𝔪|x|2𝔏112e𝔪.\mathcal{Q}_{1}\gtrsim\epsilon^{2}\mathfrak{L}_{1}^{-\frac{3}{2}}e^{-\mathfrak{m}}\sim|x|^{-2}\mathfrak{L}_{1}^{-\frac{1}{2}}e^{-\mathfrak{m}}. (8.17)

Recall that η=π\eta=\pi (cf. (8.2)). Fix selected constant CC. Without loss of generality, we may assume that Cu1C|u|1Cu_{1}\leq C|u|\ll 1. And we decompose 𝒲2=𝒲2𝒲2{\mathcal{W}}_{2}={\mathcal{W}}_{2}^{\sharp}\bigcup{\mathcal{W}}_{2}^{\sharp\sharp} with

𝒲2={(w,γ);wC,|w1|>δ},𝒲2={(w,γ);C<w<C|u|d(gu)1𝐰¯1},\displaystyle{\mathcal{W}}_{2}^{\sharp}=\{(\mathrm{w},\gamma);\mathrm{w}\leq C_{\sharp},\,|\mathrm{w}-1|>\delta\},\quad{\mathcal{W}}_{2}^{\sharp\sharp}=\{(\mathrm{w},\gamma);C_{\sharp}<\mathrm{w}<C\,|u|\,d(g_{u})^{-1}\mathbf{\overline{w}}^{-1}\},

where the constant C2C_{\sharp}\geq 2 will be chosen later. The argument used in the estimation (8.11) shows that we have uniformly on 𝒲2{\mathcal{W}}_{2}^{\sharp}, |x|24𝒟p(w,γ)𝔪+12𝔏1δ2+o(1)\frac{|x|^{2}}{4}\mathcal{D}_{\mathrm{p}}(\mathrm{w},\gamma)\geq\mathfrak{m}+\frac{1}{2}\mathfrak{L}_{1}\delta^{2}+o(1). Then, using (8.17), a simple calculation yields 𝒬2=𝒬1OC(ec𝔏114)\mathcal{Q}_{2}^{\sharp}=\mathcal{Q}_{1}\,O_{C_{\sharp}}(e^{-c\,\mathfrak{L}_{1}^{\frac{1}{4}}}).

The estimate for 𝒬2\mathcal{Q}_{2}^{\sharp\sharp} is somewhat more complicate. Recall that 𝐀(1)𝐇(1)u1\mathbf{A}(1)\sim\mathbf{H}(1)\sim u_{1} (cf. Lemma 7.6 (i) and (7.38)). Since now it holds that u1|u|u2u_{1}\sim|u|\gg u_{2} and 𝐰¯ϵu112u132+u2u1\mathbf{\overline{w}}\sim\epsilon u_{1}^{\frac{1}{2}}\sim u_{1}^{\frac{3}{2}}+u_{2}\ll u_{1} by (7.38), then from the first equality in (7.33) and (7.23), a direct calculation shows that, for any w<C|u|d(gu)1𝐰¯1\mathrm{w}<C\,|u|\,d(g_{u})^{-1}\mathbf{\overline{w}}^{-1} and γ[π,π]\gamma\in[-\pi,\pi],

u1+w𝐰¯𝐀~pC𝐀(1)u11,𝐔(1)1ϵ21,𝐔~p=𝐀~p𝐰¯2w21w𝐰¯1Cu11.u_{1}+\mathrm{w}\mathbf{\overline{w}}\sim\widetilde{\mathbf{A}}_{\mathrm{p}}\sim_{C}\mathbf{A}(1)\sim u_{1}\ll 1,\quad\mathbf{U}(1)\sim\frac{1}{\epsilon^{2}}\gg 1,\quad\widetilde{\mathbf{U}}_{\mathrm{p}}=\frac{\widetilde{\mathbf{A}}_{\mathrm{p}}}{\mathbf{\overline{w}}^{2}\,\mathrm{w}^{2}}\gtrsim\frac{1}{\mathrm{w}\mathbf{\overline{w}}}\gtrsim\frac{1}{Cu_{1}}\gg 1. (8.18)

Consequently, we can invoke (7.18) and (7.7) to obtain

|x|24𝒟p(w,γ)=|x|24𝐰¯2w2Φ(𝐔~p)|x|24(ϑ1𝐀~p2ϑ1𝐀~p𝐰¯w+32𝐰¯2w2),\displaystyle\frac{|x|^{2}}{4}\mathcal{D}_{\mathrm{p}}(\mathrm{w},\gamma)=\frac{|x|^{2}}{4}\mathbf{\overline{w}}^{2}\,\mathrm{w}^{2}\Phi(\widetilde{\mathbf{U}}_{\mathrm{p}})\geq\frac{|x|^{2}}{4}\left(\vartheta_{1}\widetilde{\mathbf{A}}_{\mathrm{p}}-2\sqrt{\vartheta_{1}\widetilde{\mathbf{A}}_{\mathrm{p}}}\,\mathbf{\overline{w}}\,\mathrm{w}+\frac{3}{2}\mathbf{\overline{w}}^{2}\mathrm{w}^{2}\right),
𝔪=|x|24𝐰¯2Φ(𝐔(1))|x|24(ϑ1𝐀(1)2ϑ1𝐀(1)𝐰¯+52𝐰¯2),\displaystyle\mathfrak{m}=\frac{|x|^{2}}{4}\mathbf{\overline{w}}^{2}\,\Phi(\mathbf{U}(1))\leq\frac{|x|^{2}}{4}\left(\vartheta_{1}\mathbf{A}(1)-2\sqrt{\vartheta_{1}\mathbf{A}(1)}\,\mathbf{\overline{w}}+\frac{5}{2}\mathbf{\overline{w}}^{2}\right),

which follows that

|x|24𝒟p𝔪|x|24[(ϑ1𝐀~p𝐰¯w)2(ϑ1𝐀(1)𝐰¯)2+12𝐰¯2(w23)].\frac{|x|^{2}}{4}\mathcal{D}_{\mathrm{p}}-\mathfrak{m}\geq\frac{|x|^{2}}{4}\left[(\sqrt{\vartheta_{1}\widetilde{\mathbf{A}}_{\mathrm{p}}}-\mathbf{\overline{w}}\mathrm{w})^{2}-(\sqrt{\vartheta_{1}\mathbf{A}(1)}-\mathbf{\overline{w}})^{2}+\frac{1}{2}\mathbf{\overline{w}}^{2}(\mathrm{w}^{2}-3)\right]. (8.19)

Note that w24\mathrm{w}^{2}\geq 4 on 𝒲2{\mathcal{W}}_{2}^{\sharp\sharp} and u2𝐰¯u_{2}\lesssim\mathbf{\overline{w}}. Then by (8.18) we can choose CC_{\sharp} properly such that, for any (w,γ)𝒲2(\mathrm{w},\gamma)\in{\mathcal{W}}_{2}^{\sharp\sharp},

𝐀~p𝐀(1)\displaystyle\sqrt{\widetilde{\mathbf{A}}_{\mathrm{p}}}-\sqrt{\mathbf{A}(1)} =4u2𝐰¯(1wcosγ)+4𝐰¯2(w21)(𝐀~p+𝐀(1))(𝐀~p+𝐀(1))𝐰¯w(𝐰¯w2u2)(u1+𝐰¯w)32\displaystyle=\frac{4u_{2}\mathbf{\overline{w}}(1-\mathrm{w}\cos\gamma)+4\mathbf{\overline{w}}^{2}(\mathrm{w}^{2}-1)}{(\sqrt{\widetilde{\mathbf{A}}_{\mathrm{p}}}+\sqrt{\mathbf{A}(1)})(\widetilde{\mathbf{A}}_{\mathrm{p}}+\mathbf{A}(1))}\,\gtrsim\,\frac{\mathbf{\overline{w}}\mathrm{w}\,(\mathbf{\overline{w}}\mathrm{w}-2\,u_{2})}{(u_{1}+\mathbf{\overline{w}}\mathrm{w})^{\frac{3}{2}}}
min{u132𝐰¯w,(𝐰¯w)12}𝐰¯wmin{C,(Cu1)12}𝐰¯w,\displaystyle\gtrsim\min\left\{u_{1}^{-\frac{3}{2}}\mathbf{\overline{w}}\mathrm{w},(\mathbf{\overline{w}}\mathrm{w})^{-\frac{1}{2}}\right\}\mathbf{\overline{w}}\mathrm{w}\gtrsim\min\left\{C_{\sharp},(Cu_{1})^{-\frac{1}{2}}\right\}\mathbf{\overline{w}}\mathrm{w},

where we have used (7.33) in the equality and 𝐰¯u132\mathbf{\overline{w}}\gtrsim u_{1}^{\frac{3}{2}} in the last inequality. Hence

ϑ1𝐀~pϑ1𝐀(1)𝐰¯w+𝐰¯u132𝐰¯2w2/C2,\sqrt{\vartheta_{1}\widetilde{\mathbf{A}}_{\mathrm{p}}}-\sqrt{\vartheta_{1}\mathbf{A}(1)}-\mathbf{\overline{w}}\mathrm{w}+\mathbf{\overline{w}}\,\gtrsim\,u_{1}^{-\frac{3}{2}}\mathbf{\overline{w}}^{2}\mathrm{w}^{2}/C^{2}, (8.20)

as long as CC_{\sharp} is large enough. Besides, observing on 𝒲2{\mathcal{W}}_{2}^{\sharp\sharp} it holds that 𝐰¯<𝐰¯wCu1\mathbf{\overline{w}}<\mathbf{\overline{w}}\mathrm{w}\,\lesssim\,C\,u_{1}, thus by (8.18) we get

ϑ1𝐀~p+ϑ1𝐀(1)𝐰¯w𝐰¯u112.\sqrt{\vartheta_{1}\widetilde{\mathbf{A}}_{\mathrm{p}}}+\sqrt{\vartheta_{1}\mathbf{A}(1)}-\mathbf{\overline{w}}\mathrm{w}-\mathbf{\overline{w}}\,\gtrsim\,u_{1}^{\frac{1}{2}}. (8.21)

Consequently, it follows from (8.19)-(8.21) that

|x|24𝒟p(w,γ)𝔪u11|x|2𝐰¯2w2/C2.\frac{|x|^{2}}{4}\mathcal{D}_{\mathrm{p}}(\mathrm{w},\gamma)-\mathfrak{m}\,\gtrsim\,u_{1}^{-1}|x|^{2}\,\mathbf{\overline{w}}^{2}\,\mathrm{w}^{2}/C^{2}. (8.22)

On the other hand, as in the proof of (8.11), we also have |x|24𝒟p𝔪𝔏1δ2𝔏1141\frac{|x|^{2}}{4}\mathcal{D}_{\mathrm{p}}-\mathfrak{m}\gtrsim\mathfrak{L}_{1}\delta^{2}\sim\mathfrak{L}_{1}^{\frac{1}{4}}\gg 1. Hence there is a positive constant cc such that

|x|24𝒟p(w,γ)𝔪cu11|x|2𝐰¯2w2+c𝔏114,(w,γ)𝒲2.\frac{|x|^{2}}{4}\mathcal{D}_{\mathrm{p}}(\mathrm{w},\gamma)-\mathfrak{m}\geq c\,u_{1}^{-1}\,|x|^{2}\,\mathbf{\overline{w}}^{2}\,\mathrm{w}^{2}+c\,\mathfrak{L}_{1}^{\frac{1}{4}},\quad\forall\,(\mathrm{w},\gamma)\in{\mathcal{W}}_{2}^{\sharp\sharp}. (8.23)

By the second estimate in (8.8) again, we yield that

𝒬2|x|2𝐰¯2𝔪1ec𝔏114e𝔪0ecu11|x|2𝐰¯2w2w𝑑w|x|2ec𝔏114e𝔪=o(𝒬1)\mathcal{Q}_{2}^{\sharp\sharp}\lesssim|x|^{2}\,\mathbf{\overline{w}}^{2}\,\mathfrak{m}^{-1}\,e^{-c\,\mathfrak{L}_{1}^{\frac{1}{4}}}\,e^{-\mathfrak{m}}\int_{0}^{\infty}e^{-c\,u_{1}^{-1}\,|x|^{2}\,\mathbf{\overline{w}}^{2}\,\mathrm{w}^{2}}\,\mathrm{w}d\mathrm{w}\lesssim|x|^{-2}\,e^{-c\,\mathfrak{L}_{1}^{\frac{1}{4}}}\,e^{-\mathfrak{m}}=o(\mathcal{Q}_{1})

where we have used that 𝔪|x|2u1\mathfrak{m}\sim|x|^{2}\,u_{1} in the second “\lesssim”, and (8.17) in “==”. In conclusion, 𝒬2=𝒬2+𝒬2=o(𝒬1)\mathcal{Q}_{2}=\mathcal{Q}_{2}^{\sharp}+\mathcal{Q}_{2}^{\sharp\sharp}=o(\mathcal{Q}_{1}) and the proof for 𝔏21\mathfrak{L}_{2}\lesssim 1 is finished.

Subcase II-2. 𝔏21.\mathfrak{L}_{2}\gg 1. We decompose 𝒲2{\mathcal{W}}_{2} as in (8.12) and write 𝒬2=𝒬2+𝒬2′′\mathcal{Q}_{2}=\mathcal{Q}_{2}^{\prime}+\mathcal{Q}_{2}^{\prime\prime}. To show 𝒬2′′=o(𝒬1)\mathcal{Q}_{2}^{\prime\prime}=o(\mathcal{Q}_{1}) we can argue as in Subcase II-1, in fact the two key estimates (8.17) and (8.23) (on 𝒲2′′{\mathcal{W}}_{2}^{\prime\prime} instead of 𝒲2{\mathcal{W}}_{2}^{\sharp\sharp}) are still valid.

To show 𝒬2=o(𝒬1)\mathcal{Q}_{2}^{\prime}=o(\mathcal{Q}_{1}) we can argue as in the case where 𝔏21\mathfrak{L}_{2}\gg 1 in Subsection 8.1, whereas the key estimates at this point we yield are (8.14) with the non-exponential factor replaced by 𝔪1(1+ϵ𝔪)12\mathfrak{m}^{-1}\,(1+\epsilon\,\mathfrak{m})^{-\frac{1}{2}} and (8.6).

In summary, the estimate (8.7) is established. To complete the proof of Theorem 8.1, we are left to provide the

8.3 Proof of Proposition 8.2

Proof.

The proof of (8.1) is elementary, mainly by means of Taylor’s formula and Chain Rule. However, it is very cumbersome and we have to make use of Lemma 7.6 and Table 1 by distinguishing three cases therein repeatedly, as explained earlier.

Let us begin with the “radial” part. Under our assumptions, it holds that for any |w1|δ|\mathrm{w}-1|\leq\delta,

𝐀𝐇(1),|𝐀|ϵ2d(gu)2𝐇(1)|u|1,|𝐀′′|ϵ2d(gu)2,|𝐀′′′|ϵ4d(gu)4|u|1,𝐔ϵ2,|𝐔|ϵ2𝐇(1)1|u|,|𝐔′′|ϵ2𝐇(1)2|u|2,|𝐔′′′|ϵ2𝐇(1)3|u|3.\begin{gathered}\mathbf{A}\sim\mathbf{H}(1),\quad|\mathbf{A}^{\prime}|\lesssim\epsilon^{2}d(g_{u})^{2}\,\mathbf{H}(1)|u|^{-1},\quad|\mathbf{A}^{\prime\prime}|\lesssim\epsilon^{2}\,d(g_{u})^{2},\quad|\mathbf{A}^{\prime\prime\prime}|\lesssim\epsilon^{4}\,d(g_{u})^{4}|u|^{-1},\\[5.69054pt] \mathbf{U}\sim\epsilon^{-2},\quad|\mathbf{U}^{\prime}|\lesssim\epsilon^{-2}\,\mathbf{H}(1)^{-1}\,|u|,\quad|\mathbf{U}^{\prime\prime}|\lesssim\epsilon^{-2}\,\mathbf{H}(1)^{-2}\,|u|^{2},\quad|\mathbf{U}^{\prime\prime\prime}|\lesssim\epsilon^{-2}\,\mathbf{H}(1)^{-3}\,|u|^{3}.\end{gathered} (8.24)

Here we only prove the first two estimates, which together with (7.24)-(7.26), the fact that d(gu)21+|u|d(g_{u})^{2}\sim 1+|u| (cf. (2.7)), and Table 1 with Lemma 7.6 will allow us to conclude the remaining claims.

To this end, recall the definition of 𝐀()\mathbf{A}(\cdot) (cf. (7.21)), (7.40) and the fact that 𝐰¯ϵ𝐇(1)12\mathbf{\overline{w}}\sim\epsilon\,\mathbf{H}(1)^{\frac{1}{2}}. Then

|𝐀(w)2𝐀(1)2|\displaystyle|\mathbf{A}(\mathrm{w})^{2}-\mathbf{A}(1)^{2}| =4𝐰¯|w1||u2𝐰¯(w+1)|4𝐰¯|w1|(|u22𝐰¯|+𝐰¯|w1|)\displaystyle=4\mathbf{\overline{w}}\,|\mathrm{w}-1|\,|u_{2}-\mathbf{\overline{w}}(\mathrm{w}+1)|\leq 4\mathbf{\overline{w}}\,|\mathrm{w}-1|(|u_{2}-2\mathbf{\overline{w}}|+\mathbf{\overline{w}}\,|\mathrm{w}-1|)
δϵ𝐇(1)12[𝐇(1)32+ϵ𝐇(1)12δ]𝐇(1)2𝐀(1)2,\displaystyle\lesssim\delta\epsilon\,\mathbf{H}(1)^{\frac{1}{2}}\,\left[\mathbf{H}(1)^{\frac{3}{2}}+\epsilon\,\mathbf{H}(1)^{\frac{1}{2}}\,\delta\right]\ll\mathbf{H}(1)^{2}\sim\mathbf{A}(1)^{2},

where “\sim” trivially holds since at this point |θ|1|\theta|\sim 1 and 𝐀(1)|θ|𝐇(1)\mathbf{A}(1)\sim|\theta|\mathbf{H}(1) (cf. Lemma 7.6 (i)). We explain how to obtain “\ll”, or equivalently, δ2ϵ2𝐇(1)\delta^{2}\,\epsilon^{2}\ll\mathbf{H}(1). Indeed, its proof is based on Table 1 and Lemma 7.6. For instance, we consider the subtlest situation: |u|1|u|\ll 1 and u2u1u_{2}\gtrsim u_{1}. In such case, Table 1 says that ϵu2/u1\epsilon\sim u_{2}/\sqrt{u_{1}}, 𝐇(1)u1\mathbf{H}(1)\sim u_{1} and δ|x|3/4u21u15/8\delta\sim|x|^{-3/4}\,u_{2}^{-1}\,u_{1}^{5/8}. Hence 𝐇(1)/(δ2ϵ2)(|x|2u1)3/4𝔪3/41\mathbf{H}(1)/(\delta^{2}\,\epsilon^{2})\sim(|x|^{2}\,u_{1})^{3/4}\sim\mathfrak{m}^{3/4}\gg 1, which implies the desired estimate. As a consequence we obtain 𝐀𝐀(1)𝐇(1)\mathbf{A}\sim\mathbf{A}(1)\sim\mathbf{H}(1).

Similarly, |2𝐰¯wu2||u22𝐰¯|+2𝐰¯|w1|𝐇(1)32+ϵ𝐇(1)12δ|2\mathbf{\overline{w}}\mathrm{w}-u_{2}|\leq|u_{2}-2\mathbf{\overline{w}}|+2\mathbf{\overline{w}}|\mathrm{w}-1|\lesssim\mathbf{H}(1)^{\frac{3}{2}}+\epsilon\,\mathbf{H}(1)^{\frac{1}{2}}\delta, thereby from (7.24)

|𝐀|ϵ𝐇(1)12𝐇(1)32+ϵ𝐇(1)12δ𝐇(1)ϵ𝐇(1)+δϵ2ϵ2d(gu)2𝐇(1)|u|1,\displaystyle|\mathbf{A}^{\prime}|\lesssim\epsilon\,\mathbf{H}(1)^{\frac{1}{2}}\,\frac{\mathbf{H}(1)^{\frac{3}{2}}+\epsilon\,\mathbf{H}(1)^{\frac{1}{2}}\,\delta}{\mathbf{H}(1)}\sim\epsilon\,\mathbf{H}(1)+\delta\,\epsilon^{2}\lesssim\epsilon^{2}\,d(g_{u})^{2}\,\mathbf{H}(1)\,|u|^{-1}, (8.25)

where the last “\lesssim” can also be checked using Table 1 and Lemma 7.6 again.

Note that we have shown in (8.24) that 𝐔ϵ21\mathbf{U}\sim\epsilon^{-2}\gtrsim 1. It follows from the first estimate in (7.5) and the definition of 𝒵()\mathcal{Z}(\cdot) (cf. (4.10)) that

𝒵(𝐔)(ϑ1𝒵(𝐔))2Υ(𝒵(𝐔))=𝐔ϵ21,\frac{\mathcal{Z}(\mathbf{U})}{(\vartheta_{1}-\mathcal{Z}(\mathbf{U}))^{2}}\sim-\Upsilon^{\prime}(\mathcal{Z}(\mathbf{U}))=\mathbf{U}\sim\epsilon^{-2}\gtrsim 1,

which implies that 𝒵(𝐔)1\mathcal{Z}(\mathbf{U})\sim 1 and ϑ1𝒵(𝐔)ϵ\vartheta_{1}-\mathcal{Z}(\mathbf{U})\sim\epsilon.

Thus using (7.29), (8.24), (7.6) and Table 1, we can show that:

|𝐇′′′|ϵ2d(gu)2𝐇(1)1|u|.|\mathbf{H}^{\prime\prime\prime}|\lesssim\epsilon^{2}\,d(g_{u})^{2}\,\mathbf{H}(1)^{-1}\,|u|.

Hence by the mean value theorem, we get

𝐇(w)=𝐇(1)+12𝐇′′(1)(w1)2+O(δ3ϵ2d(gu)2𝐇(1)1|u|),w[1δ,1+δ].\displaystyle\mathbf{H}(\mathrm{w})=\mathbf{H}(1)+\frac{1}{2}\mathbf{H}^{\prime\prime}(1)(\mathrm{w}-1)^{2}+O(\delta^{3}\,\epsilon^{2}\,d(g_{u})^{2}\,\mathbf{H}(1)^{-1}\,|u|),\quad\forall\,\mathrm{w}\in[1-\delta,1+\delta]. (8.26)

Now we handle the “angular” part. Recall that 𝔪=|x|2𝐇(1)/4\mathfrak{m}=|x|^{2}\,\mathbf{H}(1)/4, 𝐰¯ϵ𝐇(1)12\mathbf{\overline{w}}\sim\epsilon\,\mathbf{H}(1)^{\frac{1}{2}}, and 𝔏2ϵu2|x|3𝔪12\mathfrak{L}_{2}\sim\epsilon u_{2}|x|^{3}\mathfrak{m}^{-\frac{1}{2}} (cf. Lemma 7.6 (i)-(ii)), it follows from the first estimate in (8.24) that:

2u2𝐰¯w(1cosγ)𝐀(w)2ϵ𝐇(1)32u2γ2𝔏2η2𝔪11,(w,γ)𝒲1.\frac{2u_{2}\,\mathbf{\overline{w}}\,\mathrm{w}(1-\cos{\gamma})}{\mathbf{A}(\mathrm{w})^{2}}\sim\epsilon\,\mathbf{H}(1)^{-\frac{3}{2}}u_{2}\,\gamma^{2}\lesssim\mathfrak{L}_{2}\,\eta^{2}\,\mathfrak{m}^{-1}\ll 1,\quad\forall(\mathrm{w},\gamma)\in{\mathcal{W}}_{1}. (8.27)

Here to show “\ll”, we have distinguished the following cases: (I) 𝔏21\mathfrak{L}_{2}\lesssim 1; (II) 𝔏21\mathfrak{L}_{2}\gg 1 with: (II-1) |u|1|u|\gtrsim 1; (II-2a) |u|1|u|\ll 1 and u2u1u_{2}\gtrsim u_{1}; and (II-2b) |u|1|u|\ll 1 and u2u1u_{2}\ll u_{1}. We only show the subtlest case (II-2b), and the others are trivial. In such case, from Lemma 7.6 and Table 1, we yield

𝔏2η2𝔪𝔏214𝔪𝔏114𝔪(|x|2ϵ2(|x|2u1)4)14(|x|2u1(|x|2u1)4)14𝔪3/41.\frac{\mathfrak{L}_{2}\,\eta^{2}}{\mathfrak{m}}\lesssim\frac{\mathfrak{L}_{2}^{\frac{1}{4}}}{\mathfrak{m}}\lesssim\frac{\mathfrak{L}_{1}^{\frac{1}{4}}}{\mathfrak{m}}\sim\left(\frac{|x|^{2}\,\epsilon^{2}}{(|x|^{2}\,u_{1})^{4}}\right)^{\frac{1}{4}}\lesssim\left(\frac{|x|^{2}\,u_{1}}{(|x|^{2}\,u_{1})^{4}}\right)^{\frac{1}{4}}\lesssim\mathfrak{m}^{-3/4}\ll 1.

Therefore, using (7.21) and the Taylor’s expansion of 1+ρ\sqrt{1+\rho} at 0, we can write 𝐔~p(w,γ)\widetilde{\mathbf{U}}_{\mathrm{p}}(\mathrm{w},\gamma) as

𝐰¯2w2𝐀~p(w,γ)\displaystyle\mathbf{\overline{w}}^{-2}\mathrm{w}^{-2}\widetilde{\mathbf{A}}_{\mathrm{p}}(\mathrm{w},\gamma) =𝐰¯2w2|u|2+4𝐰¯2w24u2𝐰¯wcosγ\displaystyle=\mathbf{\overline{w}}^{-2}\mathrm{w}^{-2}\sqrt{|u|^{2}+4\mathbf{\overline{w}}^{2}\mathrm{w}^{2}-4u_{2}\mathbf{\overline{w}}\,\mathrm{w}\cos{\gamma}}
=𝐰¯2w2𝐀(w)(1+4u2𝐰¯w(1cosγ)𝐀(w)2)12\displaystyle=\,\mathbf{\overline{w}}^{-2}\mathrm{w}^{-2}\mathbf{A}(\mathrm{w})\left(1+\frac{4u_{2}\mathbf{\overline{w}}\,\mathrm{w}(1-\cos{\gamma})}{\mathbf{A}(\mathrm{w})^{2}}\right)^{\frac{1}{2}}
=𝐔(w)[1+2u2𝐰¯w(1cosγ)𝐀(w)2+O(ϵ2𝐇(1)3u22γ4)]\displaystyle=\,\mathbf{U}(\mathrm{w})\left[1+\frac{2u_{2}\mathbf{\overline{w}}\,\mathrm{w}(1-\cos{\gamma})}{\mathbf{A}(\mathrm{w})^{2}}+O\left(\epsilon^{2}\,\mathbf{H}(1)^{-3}u_{2}^{2}\gamma^{4}\right)\right]
=𝐔(w)+2u2(1cosγ)𝐰¯w𝐀(w)+O(𝐇(1)3u22γ4),\displaystyle=\,\mathbf{U}(\mathrm{w})+\frac{2u_{2}(1-\cos{\gamma})}{\mathbf{\overline{w}}\,\mathrm{w}\mathbf{A}(\mathrm{w})}+O\left(\mathbf{H}(1)^{-3}u_{2}^{2}\gamma^{4}\right), (8.28)

where we have used 𝐔(w)ϵ2\mathbf{U}(\mathrm{w})\sim\epsilon^{-2} (cf. (8.24)) in the last equality. In particular, from (8.27) again, we find that

|𝐔~p𝐔(w)|ϵ1𝐇(1)32u2γ2ϵ2,(w,γ)𝒲1.|\widetilde{\mathbf{U}}_{\mathrm{p}}-\mathbf{U}(\mathrm{w})|\lesssim\epsilon^{-1}\mathbf{H}(1)^{-\frac{3}{2}}u_{2}\gamma^{2}\ll\epsilon^{-2},\qquad(\mathrm{w},\gamma)\in{\mathcal{W}}_{1}. (8.29)

Now applying the Taylor’s expansion to the function 𝒵\mathcal{Z} at 𝐔(1)\mathbf{U}(1) (also recalling that 𝒵(𝐔(1))=|θ|\mathcal{Z}(\mathbf{U}(1))=|\theta| by (7.23)), we obtain that

|(ϑ1𝒵(𝐔~p))ϵ|=|(ϑ1𝒵(𝐔~p))(ϑ1|θ|)|=|𝒵(𝐔~p)𝒵(𝐔(1))|\displaystyle|(\vartheta_{1}-\mathcal{Z}(\widetilde{\mathbf{U}}_{\mathrm{p}}))-\epsilon|=|(\vartheta_{1}-\mathcal{Z}(\widetilde{\mathbf{U}}_{\mathrm{p}}))-(\vartheta_{1}-|\theta|)|=|\mathcal{Z}(\widetilde{\mathbf{U}}_{\mathrm{p}})-\mathcal{Z}(\mathbf{U}(1))|
\displaystyle\leq supι[0,1]|𝒵(ι𝐔~p+(1ι)𝐔(1))||𝐔~p𝐔(1)|.\displaystyle\sup_{\iota\in[0,1]}|\mathcal{Z}^{\prime}(\iota\widetilde{\mathbf{U}}_{\mathrm{p}}+(1-\iota)\mathbf{U}(1))|\,|\widetilde{\mathbf{U}}_{\mathrm{p}}-\mathbf{U}(1)|.

It follows from (8.29) and (8.24) that for all ι[0,1]\iota\in[0,1] we have ι𝐔(1)+(1ι)𝐔~pϵ2\iota\mathbf{U}(1)+(1-\iota)\widetilde{\mathbf{U}}_{\mathrm{p}}\sim\epsilon^{-2} and thus by (7.8)

supι[0,1]|𝒵(ι𝐔(1)+(1ι)𝐔~p)|=O(ϵ3),\sup_{\iota\in[0,1]}|\mathcal{Z}^{\prime}(\iota\mathbf{U}(1)+(1-\iota)\widetilde{\mathbf{U}}_{\mathrm{p}})|=O(\epsilon^{3}),

which implies

|(ϑ1𝒵(𝐔~p))ϵ|ϵ,(w,γ)𝒲1\displaystyle|(\vartheta_{1}-\mathcal{Z}(\widetilde{\mathbf{U}}_{\mathrm{p}}))-\epsilon|\ll\epsilon,\qquad(\mathrm{w},\gamma)\in{\mathcal{W}}_{1} (8.30)

and also

|𝒵(𝐔(w))|θ||=|(ϑ1𝒵(𝐔(w))ϵ|=O(ϵ3|𝐔(w)𝐔(1)|)=O(δϵ𝐇(1)1|u|)\displaystyle|\mathcal{Z}(\mathbf{U}(\mathrm{w}))-|\theta||=|(\vartheta_{1}-\mathcal{Z}(\mathbf{U}(\mathrm{w}))-\epsilon|=O(\epsilon^{3}\,|\mathbf{U}(\mathrm{w})-\mathbf{U}(1)|)=O(\delta\,\epsilon\,\mathbf{H}(1)^{-1}\,|u|) (8.31)

by another Taylor’s expansion and (8.24).

Similarly, using the Taylor’s expansion to the function Φ\Phi at 𝐔(w)\mathbf{U}(\mathrm{w}) gives

Φ(𝐔~p)=Φ(𝐔(w))+𝒵(𝐔(w))(𝐔~p𝐔(w))+O(ϵ3|𝐔~p𝐔(w)|2)\displaystyle\Phi(\widetilde{\mathbf{U}}_{\mathrm{p}})=\Phi(\mathbf{U}(\mathrm{w}))+\mathcal{Z}(\mathbf{U}(\mathrm{w}))(\widetilde{\mathbf{U}}_{\mathrm{p}}-\mathbf{U}(\mathrm{w}))+O(\epsilon^{3}\,|\widetilde{\mathbf{U}}_{\mathrm{p}}-\mathbf{U}(\mathrm{w})|^{2})
=\displaystyle= Φ(𝐔(w))+𝒵(𝐔(w))2u2(1cosγ)𝐰¯w𝐀(w)+O(𝐇(1)3u22γ4),\displaystyle\,\Phi(\mathbf{U}(\mathrm{w}))+\mathcal{Z}(\mathbf{U}(\mathrm{w}))\frac{2u_{2}\,(1-\cos{\gamma})}{\mathbf{\overline{w}}\,\mathrm{w}\mathbf{A}(\mathrm{w})}+O\left(\mathbf{H}(1)^{-3}\,u_{2}^{2}\,\gamma^{4}\right), (8.32)

where in the first “==” we have used (7.10) and (7.8), and in the second “==” (8.30) and (8.29). Writing

𝒵(𝐔(w))w𝐀(w)|θ|𝐀(1)=𝒵(𝐔(w))(w1)𝐀(w)+𝒵(𝐔(w))|θ|𝐀(w)+|θ|(1𝐀(w)1𝐀(1)),\frac{\mathcal{Z}(\mathbf{U}(\mathrm{w}))\,\mathrm{w}}{\mathbf{A}(\mathrm{w})}-\frac{|\theta|}{\mathbf{A}(1)}=\frac{\mathcal{Z}(\mathbf{U}(\mathrm{w}))\,(\mathrm{w}-1)}{\mathbf{A}(\mathrm{w})}+\frac{\mathcal{Z}(\mathbf{U}(\mathrm{w}))-|\theta|}{\mathbf{A}(\mathrm{w})}+|\theta|\left(\frac{1}{\mathbf{A}(\mathrm{w})}-\frac{1}{\mathbf{A}(1)}\right),

by (8.24), the first term is bounded by δ𝐇(1)1\delta\,\mathbf{H}(1)^{-1}. Using (8.31) in addition, the second term is bounded by

δϵ𝐇(1)2|u|δϵ2d(gu)2𝐇(1)1|u|1.\delta\,\epsilon\,\mathbf{H}(1)^{-2}\,|u|\lesssim\delta\,\epsilon^{2}\,d(g_{u})^{2}\,\mathbf{H}(1)^{-1}\,|u|^{-1}.

As before, the “\lesssim” here can be checked using Table 1 and Lemma 7.6. For the last term, we just use Taylor’s expansion, together with (8.24) again to obtain the same upper bound. In conclusion, we obtain that

|𝒵(𝐔(w))w𝐀(w)|θ|𝐀(1)|δ[𝐇(1)1+ϵ2d(gu)2𝐇(1)1|u|1].\displaystyle\left|\frac{\mathcal{Z}(\mathbf{U}(\mathrm{w}))\,\mathrm{w}}{\mathbf{A}(\mathrm{w})}-\frac{|\theta|}{\mathbf{A}(1)}\right|\lesssim\delta\left[\mathbf{H}(1)^{-1}+\epsilon^{2}\,d(g_{u})^{2}\,\mathbf{H}(1)^{-1}\,|u|^{-1}\right]. (8.33)

Furthermore, substituting (8.32), (8.26) and (8.33) into 𝒟p=𝐰¯2w2Φ(𝐔~p)\mathcal{D}_{\mathrm{p}}=\mathbf{\overline{w}}^{2}\mathrm{w}^{2}\,\Phi(\widetilde{\mathbf{U}}_{\mathrm{p}}), we can yield:

𝒟p(w,γ)\displaystyle\mathcal{D}_{\mathrm{p}}(\mathrm{w},\gamma) =𝐇(w)+𝒵(𝐔(w))2u2𝐰¯w(1cosγ)𝐀(w)+O(ϵ2𝐇(1)2u22γ4)\displaystyle=\mathbf{H}(\mathrm{w})+\mathcal{Z}(\mathbf{U}(\mathrm{w}))\frac{2u_{2}\,\mathbf{\overline{w}}\,\mathrm{w}(1-\cos{\gamma})}{\mathbf{A}(\mathrm{w})}+O\left(\epsilon^{2}\,\mathbf{H}(1)^{-2}\,u_{2}^{2}\,\gamma^{4}\,\right)
=𝐇(1)+12𝐇′′(1)(w1)2+4𝔏2|x|2(1cosγ)+o(|x|2),\displaystyle=\mathbf{H}(1)+\frac{1}{2}\mathbf{H}^{\prime\prime}(1)(\mathrm{w}-1)^{2}+\frac{4\mathfrak{L}_{2}}{|x|^{2}}(1-\cos{\gamma})+o(|x|^{-2}),

where in the last equality, we have used |γ|η|\gamma|\leq\eta and repeated the trick of combining Table 1 with Lemma 7.6. This establishes (8.1).

It remains to show (8.3). By Lemma 7.3, (8.30) and the mean value theorem we can check

𝔮(𝒵(𝐔~p))=𝔮(|θ|)(1+o(1)),(ϑ1𝒵(𝐔~p))12|x|2𝐰¯2w2=ϵ12|x|2𝐰¯2(1+o(1)),\displaystyle\mathfrak{q}(\mathcal{Z}(\widetilde{\mathbf{U}}_{\mathrm{p}}))=\mathfrak{q}(|\theta|)\,(1+o(1)),\qquad\frac{(\vartheta_{1}-\mathcal{Z}(\widetilde{\mathbf{U}}_{\mathrm{p}}))^{-\frac{1}{2}}}{|x|^{2}\,\mathbf{\overline{w}}^{2}\,\mathrm{w}^{2}}=\frac{\epsilon^{-\frac{1}{2}}}{|x|^{2}\,\mathbf{\overline{w}}^{2}}(1+o(1)),

and temporarily take it for granted that

eϑ1|x|2𝐰¯2w22(ϑ1𝒵(𝐔~p))I0(ϑ1|x|2𝐰¯2w22(ϑ1𝒵(𝐔~p)))=eϑ1|x|2𝐰¯22ϵI0(ϑ1|x|2𝐰¯22ϵ)(1+o(1)),e^{-\frac{\vartheta_{1}|x|^{2}\,\mathbf{\overline{w}}^{2}\mathrm{w}^{2}}{2(\vartheta_{1}-\mathcal{Z}(\widetilde{\mathbf{U}}_{\mathrm{p}}))}}I_{0}\left(\frac{\vartheta_{1}|x|^{2}\,\mathbf{\overline{w}}^{2}\mathrm{w}^{2}}{2(\vartheta_{1}-\mathcal{Z}(\widetilde{\mathbf{U}}_{\mathrm{p}}))}\right)=e^{-\frac{\vartheta_{1}|x|^{2}\,\mathbf{\overline{w}}^{2}}{2\epsilon}}I_{0}\left(\frac{\vartheta_{1}|x|^{2}\,\mathbf{\overline{w}}^{2}}{2\epsilon}\right)(1+o(1)), (8.34)

then by (2.31), Proposition 2.9 (iii) (with the fact that |x|24𝒟p(w,γ)𝔪1\frac{|x|^{2}}{4}\mathcal{D}_{\mathrm{p}}(\mathrm{w},\gamma)\geq\mathfrak{m}\gg 1) and (8.1), we get (8.3).

Finally, to prove (8.34), the above procedure allows us to obtain:

ϑ1|x|2𝐰¯2w22(ϑ1𝒵(𝐔~p))=ϑ1|x|2𝐰¯22ϵ(1+o(1))ϵ𝔪.\displaystyle\frac{\vartheta_{1}|x|^{2}\,\mathbf{\overline{w}}^{2}\,\mathrm{w}^{2}}{2(\vartheta_{1}-\mathcal{Z}(\widetilde{\mathbf{U}}_{\mathrm{p}}))}=\frac{\vartheta_{1}|x|^{2}\,\mathbf{\overline{w}}^{2}}{2\epsilon}(1+o(1))\sim\epsilon\,\mathfrak{m}. (8.35)

Thus, if ϵ𝔪1\epsilon\,\mathfrak{m}\gg 1 then (8.34) follows from (7.2); if ϵ𝔪1\epsilon\,\mathfrak{m}\lesssim 1 then from (8.35) we see that

ϑ1|x|2𝐰¯2w22(ϑ1𝒵(𝐔~p))=ϑ1|x|2𝐰¯22ϵ+o(1),\frac{\vartheta_{1}|x|^{2}\,\mathbf{\overline{w}}^{2}\mathrm{w}^{2}}{2(\vartheta_{1}-\mathcal{Z}(\widetilde{\mathbf{U}}_{\mathrm{p}}))}=\frac{\vartheta_{1}|x|^{2}\,\mathbf{\overline{w}}^{2}}{2\epsilon}+o(1),

which also leads to (8.34), since erI0(r)e^{-r}I_{0}(r) is smooth and positive on \mathbb{R}. ∎

Remark 8.3.

From the above proof and Remark 7.7, we observe that Theorem 8.1 still holds if the condition “|θ|1|\theta|\geq 1” is replaced by “|θ|α0|\theta|\geq\alpha_{0} with α0(0,1]\alpha_{0}\in(0,1]”, the cost of which is that the remainder “oζ0(1)o_{\zeta_{0}}(1)” will depend on one more parameter α0\alpha_{0} than before. As before, the choice of ζ0\zeta_{0} depends on α0\alpha_{0}.

9 Uniform asymptotics for 𝔏11\mathfrak{L}_{1}\lesssim 1 and 𝔪\mathfrak{m}\to\infty

We continue to use the notation as before, and for the sake of brevity we still omit the dependence of undetermined constants in the proof of our theorems. The aim of this section is to establish the following:

Theorem 9.1.

Let ζ0>0\zeta_{0}>0. Then there is a constant C(ζ0)1C(\zeta_{0})\gg 1 such that, for all gg satisfying 𝔏1ζ0\mathfrak{L}_{1}\leq\zeta_{0} and 𝔪C(ζ0)\mathfrak{m}\geq C(\zeta_{0}),

p(g)=4πϑ12sinϑ1ed(g)240w0ππ𝒫~p(w,γ)𝑑w𝑑γ(1+oζ0(1)),p(g)=4\pi\,\frac{\vartheta_{1}^{2}}{-\sin{\vartheta_{1}}}\,e^{-\frac{d(g)^{2}}{4}}\int_{0}^{\mathrm{w}_{0}}\int_{-\pi}^{\pi}\widetilde{\mathcal{P}}_{\mathrm{p}}(\mathrm{w},\gamma)\,d\mathrm{w}\,d\gamma\,(1+o_{\zeta_{0}}(1)),

where w0:=|x|12u134𝐰¯1\mathrm{w}_{0}:=|x|^{-\frac{1}{2}}\,u_{1}^{\frac{3}{4}}\,\mathbf{\overline{w}}^{-1} and

𝒫~p(w,γ)\displaystyle\widetilde{\mathcal{P}}_{\mathrm{p}}(\mathrm{w},\gamma) :=exp(|x|24[(ϑ1𝐀~p(w,γ)𝐰¯2w)2(ϑ1𝐀(1)𝐰¯)2])\displaystyle:=\exp\left(-\frac{|x|^{2}}{4}\left[\left(\sqrt{\vartheta_{1}\widetilde{\mathbf{A}}_{\mathrm{p}}(\mathrm{w},\gamma)}-\mathbf{\overline{w}}^{2}\mathrm{w}\right)^{2}-\left(\sqrt{\vartheta_{1}\mathbf{A}(1)}-\mathbf{\overline{w}}\right)^{2}\right]\right)
exp(ϑ1|x|2𝐰¯2w22(ϑ1𝒵(𝐔~p)))I0(ϑ1|x|2𝐰¯2w22(ϑ1𝒵(𝐔~p)))(ϑ1𝒵(𝐔~p))2w.\displaystyle\qquad\cdot\exp\left(-\frac{\vartheta_{1}|x|^{2}\mathbf{\overline{w}}^{2}\mathrm{w}^{2}}{2(\vartheta_{1}-\mathcal{Z}(\widetilde{\mathbf{U}}_{\mathrm{p}}))}\right)I_{0}\left(\frac{\vartheta_{1}|x|^{2}\mathbf{\overline{w}}^{2}\mathrm{w}^{2}}{2(\vartheta_{1}-\mathcal{Z}(\widetilde{\mathbf{U}}_{\mathrm{p}}))}\right)\,\frac{(\vartheta_{1}-\mathcal{Z}(\widetilde{\mathbf{U}}_{\mathrm{p}}))^{2}}{\mathrm{w}}. (9.1)

Recall that even in the special 55-dimensional non-isotropic Heisenberg group, it indeed happens that the leading term of the asymptotic expansion for the heat kernel cannot be represented in an explicit expression but an integral form. See [36, Theorem 4]. Hence we believe that in our group N3,2N_{3,2}, which is 66-dimensional, it is unpractical to find all the explicit expressions as in the setting of isotropic Heisenberg groups (cf. [30]). Nevertheless, in such case precise estimates for the heat kernel can still be provided in a very concise form:

Theorem 9.2.

Let ζ0>0\zeta_{0}>0. Then there is a constant C(ζ0)1C(\zeta_{0})\gg 1 such that, for all gg satisfying 𝔏1ζ0\mathfrak{L}_{1}\leq\zeta_{0} and 𝔪C(ζ0)\mathfrak{m}\geq C(\zeta_{0}),

p(g)ζ0|x|2ed(g)24.p(g)\sim_{\zeta_{0}}|x|^{-2}\,e^{-\frac{d(g)^{2}}{4}}. (9.2)

Let us begin with the

Proof of Theorem 9.2.

Under our assumptions, it deduces from Lemma 7.6 that 1𝔪|x|2|u|d(g)21\ll\mathfrak{m}\lesssim|x|^{2}\,|u|\lesssim d(g)^{2} and ϵ2d(g)2𝔏11\epsilon^{2}d(g)^{2}\sim\mathfrak{L}_{1}\lesssim 1, which yield d(g)21d(g)^{2}\gg 1 and ϵ1\epsilon\ll 1. This occurs only in the case of Lemma 7.6 (iv), whence |u|1|u|\ll 1 and (7.38) holds. In particular, 𝔪=𝐇(1)|x|2/4|x|2u1\mathfrak{m}=\mathbf{H}(1)\,|x|^{2}/4\sim|x|^{2}\,u_{1}. Now from (7.38), d(g)2|x|2d(g)^{2}\sim|x|^{2} and 𝔏1ϵ2|x|21\mathfrak{L}_{1}\sim\epsilon^{2}|x|^{2}\lesssim 1 force that

u1|x|11,u2|x|1u112u1𝔪u1,𝐰¯ϵu112|x|1u112.\displaystyle u_{1}\lesssim|x|^{-1}\ll 1,\qquad u_{2}\lesssim|x|^{-1}\,u_{1}^{\frac{1}{2}}\sim\frac{u_{1}}{\sqrt{\mathfrak{m}}}\ll u_{1},\qquad\mathbf{\overline{w}}\sim\epsilon\,u_{1}^{\frac{1}{2}}\lesssim|x|^{-1}\,u_{1}^{\frac{1}{2}}. (9.3)

We split the integral (2.30), by means of the modified polar coordinates (7.19), into the following three parts:

p(g)\displaystyle p(g) =14πe|x|24|x|2𝐰¯20+ππw𝒫p(w,γ)𝑑w𝑑γ\displaystyle=\frac{1}{4\pi}\,e^{-\frac{|x|^{2}}{4}}|x|^{2}\,\mathbf{\overline{w}}^{2}\int_{0}^{+\infty}\int_{-\pi}^{\pi}\mathrm{w}\,\mathcal{P}_{\mathrm{p}}(\mathrm{w},\gamma)\,d\mathrm{w}d\gamma
=14πe|x|24|x|2𝐰¯2(Ξ1+Ξ2+Ξ3)=:14πe|x|24i=13Li=:14πe|x|24L,\displaystyle=\frac{1}{4\pi}e^{-\frac{|x|^{2}}{4}}\,|x|^{2}\,\mathbf{\overline{w}}^{2}\left(\int_{\Xi_{1}}+\int_{\Xi_{2}}+\int_{\Xi_{3}}\right)=:\frac{1}{4\pi}e^{-\frac{|x|^{2}}{4}}\sum_{i=1}^{3}L_{i}=:\frac{1}{4\pi}e^{-\frac{|x|^{2}}{4}}L,

where

Ξ1:={(w,γ);wC|x|1u112𝐰¯1},\displaystyle\Xi_{1}:=\{(\mathrm{w},\gamma);\mathrm{w}\leq C^{\prime}\,|x|^{-1}\,u_{1}^{\frac{1}{2}}\,\mathbf{\overline{w}}^{-1}\}, (9.4)
Ξ2:={(w,γ);C|x|1u112𝐰¯1<wCu1𝐰¯1},\displaystyle\Xi_{2}:=\{(\mathrm{w},\gamma);C^{\prime}\,|x|^{-1}\,u_{1}^{\frac{1}{2}}\,\mathbf{\overline{w}}^{-1}<\mathrm{w}\leq C\,u_{1}\,\mathbf{\overline{w}}^{-1}\}, (9.5)
Ξ3:={(w,γ);w>Cu1𝐰¯1}.\displaystyle\Xi_{3}:=\{(\mathrm{w},\gamma);\mathrm{w}>Cu_{1}\,\mathbf{\overline{w}}^{-1}\}. (9.6)

Here CC and CC^{\prime} are two constants to be chosen later.

The choice of CC and the estimate of L3L_{3} are similar to those for 𝒬3\mathcal{Q}_{3} in Subsection 8.2, which yield that

L3|x|2𝔪1e2𝔪.L_{3}\lesssim|x|^{-2}\,\mathfrak{m}^{-1}e^{-2\,\mathfrak{m}}. (9.7)

The estimate of L2L_{2} follows in a similar argument but with some simplification for 𝒬2\mathcal{Q}_{2}^{\sharp\sharp} as in Subsection 8.2. Indeed, using (9.3) one can check (8.22) is still valid on Ξ2\Xi_{2} provided that CC^{\prime} is large enough. Consequently, by the second claim in (8.8) we have

L2|x|2𝐰¯2𝔪1e𝔪Ξ2ecu11|x|2𝐰¯2w2w𝑑w𝑑γ|x|2e𝔪.L_{2}\lesssim|x|^{2}\,\mathbf{\overline{w}}^{2}\,\mathfrak{m}^{-1}\,e^{-\mathfrak{m}}\int_{\Xi_{2}}e^{-c\,u_{1}^{-1}\,|x|^{2}\,\mathbf{\overline{w}}^{2}\,\mathrm{w}^{2}}\mathrm{w}\,d\mathrm{w}d\gamma\lesssim|x|^{-2}\,e^{-\mathfrak{m}}. (9.8)

Aiming now at L1L_{1}, as in the proof of (8.18), we can show that 𝐔(1)ϵ21\mathbf{U}(1)\sim\epsilon^{-2}\gg 1, 𝐀~p𝐀(1)u1\widetilde{\mathbf{A}}_{\mathrm{p}}\sim\mathbf{A}(1)\sim u_{1} and 𝐔~p=𝐀~p/(𝐰¯2w2)|x|21\widetilde{\mathbf{U}}_{\mathrm{p}}=\widetilde{\mathbf{A}}_{\mathrm{p}}/(\mathbf{\overline{w}}^{2}\,\mathrm{w}^{2})\gtrsim|x|^{2}\gg 1 on Ξ1\Xi_{1}. Then an application of (7.7) with the fact that |x|u11|x|u_{1}\lesssim 1 gives

|x|24𝒟p=|x|24𝐰¯2w2Φ(𝐔~p)=|x|24ϑ1𝐀~p+O(1),𝔪=|x|24𝐰¯2Φ(𝐔(1))=|x|24ϑ1𝐀(1)+O(1).\frac{|x|^{2}}{4}\mathcal{D}_{\mathrm{p}}=\frac{|x|^{2}}{4}\mathbf{\overline{w}}^{2}\mathrm{w}^{2}\Phi(\widetilde{\mathbf{U}}_{\mathrm{p}})=\frac{|x|^{2}}{4}\vartheta_{1}\widetilde{\mathbf{A}}_{\mathrm{p}}+O(1),\qquad\mathfrak{m}=\frac{|x|^{2}}{4}\mathbf{\overline{w}}^{2}\Phi(\mathbf{U}(1))=\frac{|x|^{2}}{4}\vartheta_{1}\mathbf{A}(1)+O(1).

Moreover, it follows from (7.33) and (9.3) that

|𝐀~p𝐀(1)|=|𝐀~p2𝐀(1)2|𝐀~p+𝐀(1)4u2𝐰¯(w+1)+4𝐰¯2(w2+1)𝐀~p+𝐀(1)|x|2.\displaystyle|\widetilde{\mathbf{A}}_{\mathrm{p}}-\mathbf{A}(1)|=\frac{|\widetilde{\mathbf{A}}_{\mathrm{p}}^{2}-\mathbf{A}(1)^{2}|}{\widetilde{\mathbf{A}}_{\mathrm{p}}+\mathbf{A}(1)}\leq\frac{4u_{2}\,\mathbf{\overline{w}}\,(\mathrm{w}+1)+4\,\mathbf{\overline{w}}^{2}\,(\mathrm{w}^{2}+1)}{\widetilde{\mathbf{A}}_{\mathrm{p}}+\mathbf{A}(1)}\lesssim|x|^{-2}.

Hence |x|24𝒟p=𝔪+O(1)(|x|2u1)1\frac{|x|^{2}}{4}\mathcal{D}_{\mathrm{p}}=\mathfrak{m}+O(1)(\sim|x|^{2}u_{1})\gg 1 and therefore, |x|𝐰¯w|x|2𝒟pu1|x|1|x|\,\mathbf{\overline{w}}\,\mathrm{w}\cdot\sqrt{|x|^{2}\,\mathcal{D}_{\mathrm{p}}}\lesssim u_{1}|x|\lesssim 1 on Ξ1\Xi_{1}. As a result, from (2.30) and Proposition 2.9 (ii), we get

L1|x|2𝐰¯2𝔪1(|x|1u112𝐰¯1)2e𝔪|x|2e𝔪.L_{1}\sim|x|^{2}\,\mathbf{\overline{w}}^{2}\,\mathfrak{m}^{-1}\,(|x|^{-1}\,u_{1}^{\frac{1}{2}}\,\mathbf{\overline{w}}^{-1})^{2}\,e^{-\mathfrak{m}}\sim|x|^{-2}\,e^{-\mathfrak{m}}. (9.9)

To summarize, by (9.7)-(9.9) and the assumption that 𝔪1\mathfrak{m}\gg 1, we obtain

L|x|2e𝔪,p(g)|x|2e𝔪|x|24.L\sim|x|^{-2}e^{-\mathfrak{m}},\qquad p(g)\sim|x|^{-2}e^{-\mathfrak{m}-\frac{|x|^{2}}{4}}. (9.10)

This together with the fact that d(g)2=|x|2+4𝔪d(g)^{2}=|x|^{2}+4\mathfrak{m} concludes the proof of Theorem 9.2. ∎

Now, we return to the

Proof of Theorem 9.1.

We split the integral (2.30) again and write this time

p(g)=14πe|x|24|x|2𝐰¯2(Ξ~1+Ξ~2+Ξ~3)=:14πe|x|24(L~1+L~2+L~3)=:14πe|x|24L~,\displaystyle p(g)=\frac{1}{4\pi}\,e^{-\frac{|x|^{2}}{4}}|x|^{2}\,\mathbf{\overline{w}}^{2}\left(\int_{\widetilde{\Xi}_{1}}+\int_{\widetilde{\Xi}_{2}}+\int_{\widetilde{\Xi}_{3}}\right)=:\frac{1}{4\pi}e^{-\frac{|x|^{2}}{4}}\,(\widetilde{L}_{1}+\widetilde{L}_{2}+\widetilde{L}_{3})=:\frac{1}{4\pi}e^{-\frac{|x|^{2}}{4}}\widetilde{L},

where Ξ~3\widetilde{\Xi}_{3} is equal to Ξ3\Xi_{3} and

Ξ~1:={(w,γ);w<|x|12u134𝐰¯1},Ξ~2:={(w,γ);|x|12u134𝐰¯1wCu1𝐰¯1}.\displaystyle\widetilde{\Xi}_{1}:=\{(\mathrm{w},\gamma);\mathrm{w}<|x|^{-\frac{1}{2}}\,u_{1}^{\frac{3}{4}}\,\mathbf{\overline{w}}^{-1}\},\quad\widetilde{\Xi}_{2}:=\{(\mathrm{w},\gamma);|x|^{-\frac{1}{2}}\,u_{1}^{\frac{3}{4}}\,\mathbf{\overline{w}}^{-1}\leq\mathrm{w}\leq Cu_{1}\,\mathbf{\overline{w}}^{-1}\}.

Recall that we have already shown that L~3=L3=L1O(e𝔪)\widetilde{L}_{3}=L_{3}=L_{1}\,O(e^{-\mathfrak{m}}). To treat L~2\widetilde{L}_{2}, remark that

|x|12u134𝐰¯1|x|1u112𝐰¯1=(|x|2u1)14𝔪141.\frac{|x|^{-\frac{1}{2}}\,u_{1}^{\frac{3}{4}}\,\mathbf{\overline{w}}^{-1}}{|x|^{-1}\,u_{1}^{\frac{1}{2}}\,\mathbf{\overline{w}}^{-1}}=\left(|x|^{2}\,u_{1}\right)^{\frac{1}{4}}\sim\mathfrak{m}^{\frac{1}{4}}\gg 1.

Then one follows the argument just given for L2L_{2}. The main difference is that we can write now |x|24𝒟p𝔪cu11|x|2𝐰¯2w2+c𝔪12\frac{|x|^{2}}{4}\mathcal{D}_{\mathrm{p}}-\mathfrak{m}\geq c\,u_{1}^{-1}\,|x|^{2}\,\mathbf{\overline{w}}^{2}\,\mathrm{w}^{2}+c\,\mathfrak{m}^{\frac{1}{2}} with the additional term c𝔪12c\,\mathfrak{m}^{\frac{1}{2}}, due to the fact that u11|x|2𝐰¯2w2𝔪12u_{1}^{-1}\,|x|^{2}\,\mathbf{\overline{w}}^{2}\,\mathrm{w}^{2}\gtrsim\mathfrak{m}^{\frac{1}{2}} on Ξ~2\widetilde{\Xi}_{2}. Thus,

L~2|x|2𝐰¯2𝔪1e𝔪ec𝔪12Ξ~2ecu11|x|2𝐰¯2w2w𝑑w𝑑γ|x|2e𝔪ec𝔪12=L1O(ec𝔪12).\displaystyle\widetilde{L}_{2}\lesssim|x|^{2}\,\mathbf{\overline{w}}^{2}\,\mathfrak{m}^{-1}\,e^{-\mathfrak{m}}\,e^{-c\,\mathfrak{m}^{\frac{1}{2}}}\int_{\widetilde{\Xi}_{2}}e^{-c\,u_{1}^{-1}\,|x|^{2}\,\mathbf{\overline{w}}^{2}\,\mathrm{w}^{2}}\mathrm{w}\,d\mathrm{w}d\gamma\lesssim|x|^{-2}\,e^{-\mathfrak{m}}\,e^{-c\,\mathfrak{m}^{\frac{1}{2}}}=L_{1}\,O(e^{-c\,\mathfrak{m}^{\frac{1}{2}}}).

Notice that Ξ~1Ξ1\widetilde{\Xi}_{1}\supseteq\Xi_{1}, then the above estimates imply that L~=L~1(1+o(1))\widetilde{L}=\widetilde{L}_{1}(1+o(1)).

On the other hand, restricting ourselves to the region Ξ~1\widetilde{\Xi}_{1}, some simplifications can be made to the expression of the heat kernel. In fact, using (9.3) it can be checked without difficulties that 𝐀~p𝐀(1)u1\widetilde{\mathbf{A}}_{\mathrm{p}}\sim\mathbf{A}(1)\sim u_{1}, 𝐔(1)ϵ21\mathbf{U}(1)\sim\epsilon^{-2}\gg 1, and 𝐔~p𝐰¯2w2u1|x|u1121\widetilde{\mathbf{U}}_{\mathrm{p}}\sim\mathbf{\overline{w}}^{-2}\mathrm{w}^{-2}u_{1}\gtrsim|x|u_{1}^{-\frac{1}{2}}\gg 1. Then by the first estimate in (7.8) we have ϑ1𝒵(𝐔~p)𝐔~p12|x|12u1141\vartheta_{1}-\mathcal{Z}(\widetilde{\mathbf{U}}_{\mathrm{p}})\sim\widetilde{\mathbf{U}}_{\mathrm{p}}^{-\frac{1}{2}}\lesssim|x|^{-\frac{1}{2}}u_{1}^{\frac{1}{4}}\ll 1, which, together with (7.14), implies that

𝔮(𝒵(𝐔~p))=2ϑ1ϑ12sinϑ1(ϑ1𝒵(𝐔~p))52(1+o(1)).\mathfrak{q}(\mathcal{Z}(\widetilde{\mathbf{U}}_{\mathrm{p}}))=\frac{\sqrt{2\vartheta_{1}}\vartheta_{1}}{-2\sin\vartheta_{1}}(\vartheta_{1}-\mathcal{Z}(\widetilde{\mathbf{U}}_{\mathrm{p}}))^{\frac{5}{2}}(1+o(1)).

And moreover, we can apply (7.7) with the fact that |x|u132u11|x|u_{1}^{\frac{3}{2}}\lesssim\sqrt{u_{1}}\ll 1 and obtain that, for all (w,γ)Ξ~1(\mathrm{w},\gamma)\in\widetilde{\Xi}_{1},

|x|24𝒟p(w,γ)=|x|24(ϑ1𝐀~p(w,γ)𝐰¯w)2+o(1),𝔪=|x|24(ϑ1𝐀(1)𝐰¯)2+o(1).\displaystyle\frac{|x|^{2}}{4}\mathcal{D}_{\mathrm{p}}(\mathrm{w},\gamma)=\frac{|x|^{2}}{4}(\sqrt{\vartheta_{1}\widetilde{\mathbf{A}}_{\mathrm{p}}(\mathrm{w},\gamma)}-\mathbf{\overline{w}}\mathrm{w})^{2}+o(1),\quad\mathfrak{m}=\frac{|x|^{2}}{4}(\sqrt{\vartheta_{1}\mathbf{A}(1)}-\mathbf{\overline{w}})^{2}+o(1).

Thus, combining these estimates with Proposition 2.9 (iii) yields that

|x|2𝐰¯2w𝒫p(w,γ)=16π2ϑ12sinϑ1e𝔪𝒫~p(w,γ)(1+o(1)),(w,γ)Ξ~1,\displaystyle|x|^{2}\,\mathbf{\overline{w}}^{2}\,\mathrm{w}\,\mathcal{P}_{\mathrm{p}}(\mathrm{w},\gamma)=16\pi^{2}\,\frac{\vartheta_{1}^{2}}{-\sin{\vartheta_{1}}}\,e^{-\mathfrak{m}}\,\widetilde{\mathcal{P}}_{\mathrm{p}}(\mathrm{w},\gamma)\,(1+o(1)),\quad\forall\,(\mathrm{w},\gamma)\in\widetilde{\Xi}_{1},

where 𝒫~p(w,γ)\widetilde{\mathcal{P}}_{\mathrm{p}}(\mathrm{w},\gamma) is defined by (9.1).

Now set

L~:=16π2ϑ12sinϑ1e𝔪Ξ~1𝒫~p𝑑w𝑑γ.\displaystyle\widetilde{L}^{\prime}:=16\pi^{2}\,\frac{\vartheta_{1}^{2}}{-\sin{\vartheta_{1}}}\,e^{-\mathfrak{m}}\int_{\widetilde{\Xi}_{1}}\widetilde{\mathcal{P}}_{\mathrm{p}}\,d\mathrm{w}d\gamma.

Then our estimates above imply L~=L~1(1+o(1))=L~(1+o(1))\widetilde{L}^{\prime}=\widetilde{L}_{1}(1+o(1))=\widetilde{L}\,(1+o(1)), which completes the proof of Theorem 9.1. ∎

10 Uniform asymptotics for the remaining cases: points near the abnormal set

Recall that 𝔪=|x|2𝐇(1)/4=(d(g)2|x|2)/4\mathfrak{m}=|x|^{2}\,\mathbf{H}(1)/4=(d(g)^{2}-|x|^{2})/4. In this section, our goal is to deal with the following two remaining cases:

(C1). d(g)2+d(g)^{2}\to+\infty, |θ|3|\theta|\leq 3, and θ2|x|1\theta_{2}|x|\lesssim 1.

(C2). d(g)2+d(g)^{2}\to+\infty, |θ|1|\theta|\geq 1 and 𝔪1\mathfrak{m}\lesssim 1.

However, we shall give asymptotics under more general conditions as follows:

(C3). d(g)2+d(g)^{2}\to+\infty and 𝔪1\mathfrak{m}\lesssim 1.

(C4). d(g)2+d(g)^{2}\to+\infty, u1|x|2u_{1}\lesssim|x|^{-2} and u2|x|1u_{2}\lesssim|x|^{-1}.

More precisely, the relations among them are the following:

(C1) or (C2) \Rightarrow (C3) \Leftrightarrow (C4). (10.1)

Indeed, (C2) \Rightarrow (C3) trivially holds and (C1) \Rightarrow (C3) can be explained by the fact that 𝔪θ22|x|2\mathfrak{m}\sim\theta_{2}^{2}|x|^{2} (cf. Lemma 7.6 (i)). To show (C3) \Rightarrow (C4), noting now that d(g)2=|x|2+4𝔪|x|21d(g)^{2}=|x|^{2}+4\,\mathfrak{m}\sim|x|^{2}\gg 1. Then Lemma 7.6 (i)-(ii) imply that u1θ1𝔪|x|2|x|2u_{1}\sim\theta_{1}\,\mathfrak{m}\,|x|^{-2}\lesssim|x|^{-2} and u2ϵ𝔏2𝔏11𝔪12|x|1|x|1u_{2}\sim\epsilon\,\mathfrak{L}_{2}\,\mathfrak{L}_{1}^{-1}\,\mathfrak{m}^{\frac{1}{2}}\,|x|^{-1}\lesssim|x|^{-1}, which yield the desired results. Finally, suppose (C4) holds. First we claim that |u|1|u|\ll 1. Suppose the contrary. Then 1|u||x|2+|x|11\lesssim|u|\lesssim|x|^{-2}+|x|^{-1}, whence |x|1|x|\lesssim 1 and |u||x|2|u|\lesssim|x|^{-2}, which, however, implies that d(g)2|x|2|u|1d(g)^{2}\sim|x|^{2}|u|\lesssim 1, a contradiction! Now 𝔪1\mathfrak{m}\lesssim 1 follows at once from (7.38) as |θ|1|\theta|\geq 1 and from (7.39) as |θ|3|\theta|\leq 3, respectively. Hence (C4) \Rightarrow (C3).

To conclude, we reduce the remaining cases to (C4), or equivalently, (C3). Roughly speaking, these points are near the “(shortest) abnormal set”, i.e., the points gg such that the Hessian matrices of the reference functions Hessθϕ(g;θ)\mathrm{Hess}_{\theta}\,\phi(g;\theta) are degenerate at the critical points θ\theta. Hence in such case, the method of stationary phase no longer works even for small-time heat kernel asymptotics. The interested readers may consult [33]. More precisely, we shall prove the following theorem.

Theorem 10.1.

Let ζ0>0\zeta_{0}>0. Then there is a constant C(ζ0)1C(\zeta_{0})\gg 1 such that, for all gg satisfying 𝔪ζ0\mathfrak{m}\leq\zeta_{0} and d(g)2C(ζ0)d(g)^{2}\geq C(\zeta_{0}),

p(g)=e|x|244π|x|2𝐅(12|x|u2,14|x|2u1)(1+oζ0(1)),\displaystyle p(g)=e^{-\frac{|x|^{2}}{4}}\,\frac{4\pi}{|x|^{2}}\,\mathbf{F}\left(\frac{1}{2}|x|u_{2},\,\frac{1}{4}|x|^{2}u_{1}\right)(1+o_{\zeta_{0}}(1)), (10.2)

where

𝐅(v1,v2):=r3rcoshrsinhre14r2rcothr1v12+iv2r𝑑r,v1,v2.\displaystyle\mathbf{F}(v_{1},v_{2}):=\int_{\mathbb{R}}\frac{r^{3}}{r\cosh{r}-\sinh{r}}\,e^{-\frac{1}{4}\frac{r^{2}}{r\coth{r}-1}v_{1}^{2}+iv_{2}r}\,dr,\qquad v_{1},v_{2}\in{\mathbb{R}}. (10.3)

Before showing Theorem 10.1, we state the positivity of the smooth function 𝐅\mathbf{F} in the following lemma:

Lemma 10.2.

It holds that 𝐅(v1,v2)>0\mathbf{F}(v_{1},v_{2})>0, v1,v2\forall\,v_{1},v_{2}\in{\mathbb{R}}.

Its proof is essentially the same as in Appendix A.1 below, but simpler in the 1D case. We now proceed with the proof of Theorem 10.1.

Proof.

Recall we have shown that |x|d(g)1|x|\sim d(g)\gg 1 in the proof of (10.1), and the heat kernel is given by (cf. (2.19))

p(g)\displaystyle p(g) =e|x|243|λ|sinh|λ|exp{|x|24[|λ|coth|λ|1|λ|2(λ22+λ32)iuλ]}𝑑λ\displaystyle=e^{-\frac{|x|^{2}}{4}}\,\int_{\mathbb{R}^{3}}\frac{|\lambda|}{\sinh{|\lambda|}}\,\exp\left\{-\frac{|x|^{2}}{4}\left[\frac{|\lambda|\coth{|\lambda|}-1}{|\lambda|^{2}}\,(\lambda_{2}^{2}+\lambda_{3}^{2})-iu\cdot\lambda\right]\right\}\,d\lambda (10.4)
=e|x|24(Λ1+Λ2+Λ3)=:e|x|24(1+2+3)=:e|x|24,\displaystyle=e^{-\frac{|x|^{2}}{4}}\left(\int_{\Lambda_{1}}+\int_{\Lambda_{2}}+\int_{\Lambda_{3}}\right)=:e^{-\frac{|x|^{2}}{4}}\,(\mathcal{I}_{1}+\mathcal{I}_{2}+\mathcal{I}_{3})=:e^{-\frac{|x|^{2}}{4}}\,\mathcal{I},

where

Λ1:={(λ1,λ)×2;|λ||x|34,|λ1||x|18},\displaystyle\Lambda_{1}:=\,\{(\lambda_{1},\lambda^{\prime})\in\mathbb{R}\times\mathbb{R}^{2};\,|\lambda^{\prime}|\leq|x|^{-\frac{3}{4}},\,|\lambda_{1}|\leq|x|^{\frac{1}{8}}\},
Λ2:={(λ1,λ)×2;|λ|>|x|34,|λ1||x|18},\displaystyle\Lambda_{2}:=\,\{(\lambda_{1},\lambda^{\prime})\in\mathbb{R}\times\mathbb{R}^{2};\,|\lambda^{\prime}|>|x|^{-\frac{3}{4}},\,|\lambda_{1}|\leq|x|^{\frac{1}{8}}\},
Λ3:={(λ1,λ)×2;|λ1|>|x|18}.\displaystyle\Lambda_{3}:=\,\{(\lambda_{1},\lambda^{\prime})\in\mathbb{R}\times\mathbb{R}^{2};\,|\lambda_{1}|>|x|^{\frac{1}{8}}\}.

Notice that the functions

(ln(rsinhr))=1rcothr=sinhrrcoshrrsinhr,\displaystyle\left(\ln\left(\frac{r}{\sinh{r}}\right)\right)^{\prime}=\frac{1}{r}-\coth{r}=\frac{\sinh{r}-r\,\cosh{r}}{r\,\sinh{r}},
(rcothr1r2)=1r2(cothrr(sinhr)2)2(rcothr1r3)\displaystyle\left(\frac{r\coth{r}-1}{r^{2}}\right)^{\prime}=\frac{1}{r^{2}}\left(\coth{r}-\frac{r}{(\sinh{r})^{2}}\right)-2\left(\frac{r\coth{r}-1}{r^{3}}\right)

are all bounded on \mathbb{R}, and ||λ||λ1|||λ|.|\,|\lambda|-|\lambda_{1}|\,|\leq|\lambda^{\prime}|. Then on Λ1\Lambda_{1}, using the finite-increment theorem, we can write

|λ|sinh|λ|e|x|24|λ|coth|λ|1|λ|2|λ|2\displaystyle\frac{|\lambda|}{\sinh{|\lambda|}}e^{-\frac{|x|^{2}}{4}\frac{|\lambda|\coth{|\lambda|}-1}{|\lambda|^{2}}\,|\lambda^{\prime}|^{2}} =λ1sinhλ1e|x|24λ1cothλ11λ12|λ|2eO(|λ|)+O(|λ|3|x|2)\displaystyle=\,\frac{\lambda_{1}}{\sinh{\lambda_{1}}}e^{-\frac{|x|^{2}}{4}\frac{\lambda_{1}\coth{\lambda_{1}}-1}{\lambda_{1}^{2}}\,|\lambda^{\prime}|^{2}}e^{O(|\lambda^{\prime}|)+O(|\lambda^{\prime}|^{3}\,|x|^{2})}
=λ1sinhλ1e|x|24λ1cothλ11λ12|λ|2(1+O(|x|14))\displaystyle=\,\frac{\lambda_{1}}{\sinh{\lambda_{1}}}e^{-\frac{|x|^{2}}{4}\frac{\lambda_{1}\coth{\lambda_{1}}-1}{\lambda_{1}^{2}}\,|\lambda^{\prime}|^{2}}(1+O(|x|^{-\frac{1}{4}})) (10.5)
=:𝐀1+𝐀2 (with the obvious meaning),\displaystyle=:\mathbf{A}_{1}+\mathbf{A}_{2}\mbox{\,\,(with the obvious meaning)},

where we have used the fact that d(g)2|x|2+d(g)^{2}\sim|x|^{2}\to+\infty.

Put 1,k=Λ1𝐀kexp{i|x|2uλ/4}𝑑λ\mathcal{I}_{1,k}=\int_{\Lambda_{1}}\mathbf{A}_{k}\exp\{i|x|^{2}u\cdot\lambda/4\}\,d\lambda, k=1,2k=1,2. We then write

1,1\displaystyle\mathcal{I}_{1,1} =Λ1λ1sinhλ1e|x|24λ1cothλ11λ12|λ|2e|x|24iuλ𝑑λ\displaystyle=\int_{\Lambda_{1}}\frac{\lambda_{1}}{\sinh{\lambda_{1}}}e^{-\frac{|x|^{2}}{4}\frac{\lambda_{1}\coth{\lambda_{1}}-1}{\lambda_{1}^{2}}\,|\lambda^{\prime}|^{2}}e^{\frac{|x|^{2}}{4}iu\cdot\lambda}\,d\lambda
=3Λ2Λ3=:1,1,11,1,21,1,3.\displaystyle=\int_{\mathbb{R}^{3}}\cdots-\int_{\Lambda_{2}}\cdots-\int_{\Lambda_{3}}\cdots=:\mathcal{I}_{1,1,1}-\mathcal{I}_{1,1,2}-\mathcal{I}_{1,1,3}.

Next, we shall illustrate that the term 1,1,1\mathcal{I}_{1,1,1} is principal and the other terms 1,2,1,1,2,1,1,3,2,3\mathcal{I}_{1,2},\,\mathcal{I}_{1,1,2},\,\mathcal{I}_{1,1,3},\,\mathcal{I}_{2},\,\mathcal{I}_{3} are negligible.

Estimate of 1,1,1\mathcal{I}_{1,1,1}. Applying (4.3) with q=2q=2, A=|x|22λ1cothλ11λ12𝕀2A=\frac{|x|^{2}}{2}\frac{\lambda_{1}\coth\lambda_{1}-1}{\lambda_{1}^{2}}\,{\mathbb{I}}_{2} and Y=|x|24(u2,0)Y=\frac{|x|^{2}}{4}(u_{2},0) respectively w.r.t. the variable λ\lambda^{\prime}, we see that

1,1,1=4π|x|2𝐅(12|x|u2,14|x|2u1)\displaystyle\mathcal{I}_{1,1,1}=\frac{4\pi}{|x|^{2}}\,\mathbf{F}\left(\frac{1}{2}|x|u_{2},\,\frac{1}{4}|x|^{2}u_{1}\right)

from the definition of 𝐅\mathbf{F} in (10.3).

By Lemma 10.2, 𝐅(v1,v2)1\mathbf{F}(v_{1},v_{2})\sim 1 whenever v1v_{1} and v2v_{2} are both bounded. Then 1,1,1ζ0|x|2\mathcal{I}_{1,1,1}\sim_{\zeta_{0}}|x|^{-2} under our assumptions. It is therefore sufficient to show

|1,2|+|1,1,2|+|1,1,3|+|2|+|3|=o(|x|2).|\mathcal{I}_{1,2}|+|\mathcal{I}_{1,1,2}|+|\mathcal{I}_{1,1,3}|+|\mathcal{I}_{2}|+|\mathcal{I}_{3}|=o(|x|^{-2}). (10.6)

Bound of 1,2\mathcal{I}_{1,2}. Applying (10.5) and the exponential decay of r/sinhrr/\sinh r, we obtain

|1,2||x|14Λ1λ1sinhλ1e|x|24λ1cothλ11λ12|λ|2𝑑λ𝑑λ1|x|94.|\mathcal{I}_{1,2}|\lesssim|x|^{-\frac{1}{4}}\,\int_{\Lambda_{1}}\frac{\lambda_{1}}{\sinh{\lambda_{1}}}e^{-\frac{|x|^{2}}{4}\frac{\lambda_{1}\coth{\lambda_{1}}-1}{\lambda_{1}^{2}}\,|\lambda^{\prime}|^{2}}d\lambda^{\prime}\,d\lambda_{1}\lesssim|x|^{-\frac{9}{4}}.

Evaluation of 2\mathcal{I}_{2} and 1,1,2\mathcal{I}_{1,1,2}. On Λ2\Lambda_{2}, first observe that |λ|2|x|2|x|32|x|2=|x|12|\lambda^{\prime}|^{2}\,|x|^{2}\geq|x|^{-\frac{3}{2}}\,|x|^{2}=|x|^{\frac{1}{2}}. Moreover, the simple estimate

rcothr1r211+|r|,r\frac{r\coth{r}-1}{r^{2}}\sim\frac{1}{1+|r|},\quad\forall\,r\in\mathbb{R} (10.7)

implies that

λ1cothλ11λ1211+|λ1|11+|x|181|x|18.\displaystyle\frac{\lambda_{1}\coth{\lambda_{1}}-1}{\lambda_{1}^{2}}\sim\frac{1}{1+|\lambda_{1}|}\geq\frac{1}{1+|x|^{\frac{1}{8}}}\sim\frac{1}{|x|^{\frac{1}{8}}}.

Then

|x|24|λ|2λ1cothλ11λ12|x|38+|x|158|λ|2.\frac{|x|^{2}}{4}|\lambda^{\prime}|^{2}\,\frac{\lambda_{1}\coth{\lambda_{1}}-1}{\lambda_{1}^{2}}\gtrsim|x|^{\frac{3}{8}}+|x|^{\frac{15}{8}}|\lambda^{\prime}|^{2}.

From this and the trivial estimate r/sinhr1r/\sinh{r}\lesssim 1, we obtain

|1,1,2|ec|x|38Λ2exp{|x|158|λ|2}𝑑λ1𝑑λec|x|38.|\mathcal{I}_{1,1,2}|\leq e^{-c\,|x|^{\frac{3}{8}}}\int_{\Lambda_{2}}\exp\{-|x|^{\frac{15}{8}}|\lambda^{\prime}|^{2}\}\,d\lambda_{1}d\lambda^{\prime}\lesssim e^{-c\,|x|^{\frac{3}{8}}}.

To estimate 2\mathcal{I}_{2}, using (10.7) again for r=|λ|r=|\lambda|, we see

|λ|coth|λ|1|λ|2|λ|2|x|2|x|38,λΛ2,\displaystyle\frac{|\lambda|\coth{|\lambda|}-1}{|\lambda|^{2}}\,|\lambda^{\prime}|^{2}\,|x|^{2}\gtrsim|x|^{\frac{3}{8}},\qquad\forall\,\lambda\in\Lambda_{2}, (10.8)

which implies that

|2|ec|x|38Λ2|λ|sinh|λ|𝑑λec|x|383|λ|sinh|λ|𝑑λec|x|38.|\mathcal{I}_{2}|\lesssim e^{-c\,|x|^{\frac{3}{8}}}\int_{\Lambda_{2}}\frac{|\lambda|}{\sinh|\lambda|}\,d\lambda\leq e^{-c\,|x|^{\frac{3}{8}}}\int_{\mathbb{R}^{3}}\frac{|\lambda|}{\sinh|\lambda|}\,d\lambda\lesssim e^{-c\,|x|^{\frac{3}{8}}}.

Estimations of 3\mathcal{I}_{3} and 1,1,3\mathcal{I}_{1,1,3}. It follows from the monotonicity and the exponential decay of ssinhs\frac{s}{\sinh{s}} on [0,+)[0,+\infty) that

λ1sinhλ1=(λ1sinhλ1)2|x|18sinh|x|18λ1sinhλ1λ1sinhλ1e|x|184,(λ1,λ)Λ3.\displaystyle\frac{\lambda_{1}}{\sinh{\lambda_{1}}}=\left(\sqrt{\frac{\lambda_{1}}{\sinh{\lambda_{1}}}}\right)^{2}\leq\sqrt{\frac{|x|^{\frac{1}{8}}}{\sinh{|x|^{\frac{1}{8}}}}}\sqrt{\frac{\lambda_{1}}{\sinh{\lambda_{1}}}}\lesssim\sqrt{\frac{\lambda_{1}}{\sinh{\lambda_{1}}}}\,e^{-\frac{|x|^{\frac{1}{8}}}{4}},\quad\forall(\lambda_{1},\lambda^{\prime})\in\Lambda_{3}.

Inserting this estimate into 1,1,3\mathcal{I}_{1,1,3}, we have

|1,1,3|e|x|184Λ3λ1sinhλ1exp{14λ1cothλ11λ12|λ|2|x|2}𝑑λ1𝑑λe|x|184|x|2.|\mathcal{I}_{1,1,3}|\lesssim e^{-\frac{|x|^{\frac{1}{8}}}{4}}\int_{\Lambda_{3}}\sqrt{\frac{\lambda_{1}}{\sinh{\lambda_{1}}}}\,\exp\left\{-\frac{1}{4}\frac{\lambda_{1}\coth{\lambda_{1}}-1}{\lambda_{1}^{2}}\,|\lambda^{\prime}|^{2}\,|x|^{2}\right\}\,d\lambda_{1}\,d\lambda^{\prime}\\ \lesssim e^{-\frac{|x|^{\frac{1}{8}}}{4}}|x|^{-2}.

Similarly,

|3|e|x|184Λ3|λ|sinh|λ|𝑑λe|x|184.|\mathcal{I}_{3}|\lesssim e^{-\frac{|x|^{\frac{1}{8}}}{4}}\int_{\Lambda_{3}}\sqrt{\frac{|\lambda|}{\sinh{|\lambda|}}}d\lambda\lesssim e^{-\frac{|x|^{\frac{1}{8}}}{4}}.

From these estimates we conclude the proof of (10.6) and hence the theorem. ∎

Remark 10.3.

The argument above also works for u1=u2=0u_{1}=u_{2}=0.

Notice that the above proof and the fact that d(g)2=|x|2+4𝔪d(g)^{2}=|x|^{2}+4\,\mathfrak{m} also give the following

Corollary 10.4.

Let ζ0>0\zeta_{0}>0. Then there is a constant C(ζ0)1C(\zeta_{0})\gg 1 such that

p(g)ζ0|x|2ed(g)24,for d(g)2C(ζ0) with 𝔪ζ0.p(g)\,\sim_{\zeta_{0}}\,|x|^{-2}\,e^{-\frac{d(g)^{2}}{4}},\quad\mbox{for $d(g)^{2}\geq C(\zeta_{0})$ with $\mathfrak{m}\leq\zeta_{0}$.} (10.9)

11 Summaries of main results

Recall that in previous sections we have established the uniform asymptotic expansions of the heat kernel at infinity in four possible cases, namely Theorems 6.1, 8.1, 9.1 and 10.1 with Remarks 6.4 and 8.3. Now we summarize these results and deduce the precise estimates as well as the small-time asymptotics of the heat kernel.

11.1 Uniform asymptotics

Recall (7.34)-(7.36), (2.33), (9.1) and (10.3) for pertinent definitions. We have:

Theorem 11.1.

Under the Assumption (A) (cf. (2.18)), it holds that:

  1. (i)

    Let α0[3,π)\alpha_{0}\in[3,\pi). Then there exists a constant C(α0)1C(\alpha_{0})\gg 1 such that for all gg satisfying |θ|α0|\theta|\leq\alpha_{0} with θ2|x|C(α0)\theta_{2}|x|\geq C(\alpha_{0}),

    p(g)=(8π)32ed(g)24|θ|sin|θ|1det(Hessθϕ(g;θ))(1+oα0(1)).p(g)=(8\pi)^{\frac{3}{2}}\,e^{-\frac{d(g)^{2}}{4}}\,\frac{|\theta|}{\sin{|\theta|}}\,\frac{1}{\sqrt{\det(-\mathrm{Hess}_{\theta}\phi(g;\theta))}}\,(1+o_{\alpha_{0}}(1)).
  2. (ii)

    Let β0(0,1]\beta_{0}\in(0,1]. Then there exists a constant C(β0)1C(\beta_{0})\gg 1 such that for all gg satisfying |θ|β0|\theta|\geq\beta_{0}, 𝔏1,𝔪C(β0)\mathfrak{L}_{1},\mathfrak{m}\geq C(\beta_{0}),

    p(g)=16π2πϑ1𝔮(|θ|)e𝔏2I0(𝔏2)ϵ𝔏1eϑ1|x|2𝐰¯22ϵI0(ϑ1|x|2𝐰¯22ϵ)ed(g)24(1+oβ0(1)).p(g)=16\pi^{2}\sqrt{\pi\vartheta_{1}}\,\mathfrak{q}(|\theta|)\,\frac{e^{-\mathfrak{L}_{2}}\,I_{0}(\mathfrak{L}_{2})}{\sqrt{\epsilon\,\mathfrak{L}_{1}}}\,e^{-\frac{\vartheta_{1}|x|^{2}\,\mathbf{\overline{w}}^{2}}{2\epsilon}}\,I_{0}\left(\frac{\vartheta_{1}|x|^{2}\,\mathbf{\overline{w}}^{2}}{2\epsilon}\right)\,e^{-\frac{d(g)^{2}}{4}}\,(1+o_{\beta_{0}}(1)).
  3. (iii)

    Let ζ0>0\zeta_{0}>0. Then there is a constant C(ζ0)1C(\zeta_{0})\gg 1 such that, for all gg satisfying 𝔏1ζ0\mathfrak{L}_{1}\leq\zeta_{0} and 𝔪C(ζ0)\mathfrak{m}\geq C(\zeta_{0}),

    p(g)=4πϑ12sinϑ1ed(g)240w0ππ𝒫~p(w,γ)𝑑w𝑑γ(1+oζ0(1)).p(g)=4\pi\,\frac{\vartheta_{1}^{2}}{-\sin{\vartheta_{1}}}\,e^{-\frac{d(g)^{2}}{4}}\int_{0}^{\mathrm{w}_{0}}\int_{-\pi}^{\pi}\widetilde{\mathcal{P}}_{\mathrm{p}}(\mathrm{w},\gamma)\,d\mathrm{w}\,d\gamma\,(1+o_{\zeta_{0}}(1)).
  4. (iv)

    Let ζ0>0\zeta_{0}>0. Then there is a constant C(ζ0)1C(\zeta_{0})\gg 1 such that, for all gg satisfying 𝔪ζ0\mathfrak{m}\leq\zeta_{0} and d(g)2C(ζ0)d(g)^{2}\geq C(\zeta_{0}),

    p(g)=ed(g)24e𝔪4π|x|2𝐅(12|x|u2,14|x|2u1)(1+oζ0(1)).p(g)=e^{-\frac{d(g)^{2}}{4}}\,e^{\mathfrak{m}}\,\frac{4\pi}{|x|^{2}}\,\mathbf{F}\left(\frac{1}{2}|x|u_{2},\,\frac{1}{4}|x|^{2}u_{1}\right)(1+o_{\zeta_{0}}(1)).
Remark 11.2.

Supposing the requirements of Theorem 11.1 (i) and (ii) are fulfilled simultaneously, then the leading terms of the asymptotic formulas for the heat kernel given by them indeed coincide. This fact can be verified via an elementary but somewhat tedious calculation, by means of (7.2) (noting now that 𝔏21\mathfrak{L}_{2}\gg 1 and ϑ1|x|2𝐰¯22ϵ1\frac{\vartheta_{1}|x|^{2}\,\mathbf{\overline{w}}^{2}}{2\epsilon}\gg 1 by Lemma 7.6 and Remark 7.7), with equalities (7.35), (7.36), (2.33), (3.30), (3.25) and 𝐰¯=θ2ψ(|θ|)\mathbf{\overline{w}}=\theta_{2}\,\psi(|\theta|) (cf. (2.18)).

As a by-product, we get the following important equality under Assumption (A):

(8π)32|θ|θ2sin|θ||x|3K(θ1,θ2)K3(θ1,θ2)=8ππ𝔮(|θ|)(|x|2𝐰¯22𝔏1𝔏2)12.\displaystyle(8\pi)^{\frac{3}{2}}\,\frac{|\theta|}{\theta_{2}\sin{|\theta|}}\,\frac{|x|^{-3}}{\sqrt{\mathrm{K}(\theta_{1},\theta_{2})\,\mathrm{K}_{3}(\theta_{1},\theta_{2})}}=8\pi\sqrt{\pi}\,\mathfrak{q}(|\theta|)\,\left(\frac{|x|^{2}\,\mathbf{\overline{w}}^{2}}{2}\mathfrak{L}_{1}\,\mathfrak{L}_{2}\right)^{-\frac{1}{2}}. (11.1)

An application of Theorem 11.1 leads to the following consequence:

11.2 Sharp upper and lower bounds

The statement of this result has already appeared in Section 2.4. We restate it here for the reader’s convenience.

Theorem 11.3.

Under Assumption (A) (cf. (2.18)), we have

p(x,t)(1+d(g))21+ϵd(g)1+ϵd(g)+ϵt212|x|12(d(g)2|x|2)14ed(g)24.\displaystyle p(x,t)\sim(1+d(g))^{-2}\frac{1+\epsilon\,d(g)}{1+\epsilon\,d(g)+\epsilon\,t_{2}^{\frac{1}{2}}\,|x|^{\frac{1}{2}}\,(d(g)^{2}-|x|^{2})^{\frac{1}{4}}}\,e^{-\frac{d(g)^{2}}{4}}. (11.2)
Proof.

Recall that d(g)2=|x|2+4𝔪d(g)^{2}=|x|^{2}+4\,\mathfrak{m} and 𝔪=𝐇(1)|x|2/4\mathfrak{m}=\mathbf{H}(1)\,|x|^{2}/4. Then from Lemma 7.6 (ii), the right hand side of (11.2) is bounded below and above by the quantity

BND:=11+d(g)2(1+𝔏1)12(1+𝔏1+ϵ𝔪𝔏2)12ed(g)24.\displaystyle{\rm BND}:=\frac{1}{1+d(g)^{2}}\,(1+\mathfrak{L}_{1})^{\frac{1}{2}}\,(1+\mathfrak{L}_{1}+\epsilon\,\mathfrak{m}\,\mathfrak{L}_{2})^{-\frac{1}{2}}\,e^{-\frac{d(g)^{2}}{4}}. (11.3)

Thus the estimate (11.2) just amounts to pBNDp\sim{\rm BND}.

For the case that d(g)21d(g)^{2}\lesssim 1, using Lemma 7.6 (ii) again we obtain d(g)+ϵd(g)+ϵ𝔪+𝔏21d(g)+\epsilon\,d(g)+\epsilon\,\mathfrak{m}+\mathfrak{L}_{2}\lesssim 1, whence BND1{\rm BND}\sim 1, while by the positivity of the heat kernel we have p1p\sim 1. Thus, pBNDp\sim{\rm BND}.

The opposite case d(g)21d(g)^{2}\gg 1 can be deduced directly from Theorem 11.1, according to which we split it into four subcases:

If |θ|3|\theta|\leq 3 (so ϵ1\epsilon\sim 1) and θ2|x|\theta_{2}|x|\to\infty, then |x|1|x|\gg 1, and 𝔪|θ2|2|x|21\mathfrak{m}\sim|\theta_{2}|^{2}|x|^{2}\gg 1 by Lemma 7.6 (i), which, together with Theorem 11.1 (i) and (6.2) show that

p|x|2(θ2|x|)1ed(g)24|x|2𝔪12ed(g)24.p\sim|x|^{-2}(\,\theta_{2}|x|\,)^{-1}e^{-\frac{d(g)^{2}}{4}}\sim|x|^{-2}\mathfrak{m}^{-\frac{1}{2}}\,e^{-\frac{d(g)^{2}}{4}}.

Since d(g)2|x|2(1+θ22)|x|2d(g)^{2}\sim|x|^{2}(1+\theta_{2}^{2})\sim|x|^{2}, then 𝔏1|x|2\mathfrak{L}_{1}\sim|x|^{2} by Lemma 7.6 (ii). From this and Lemma 7.6 (v) we conclude that BNDd(g)1|x|1𝔪12ed(g)24p{\rm BND}\sim d(g)^{-1}|x|^{-1}\mathfrak{m}^{-\frac{1}{2}}e^{-\frac{d(g)^{2}}{4}}\sim p.

If |θ|1|\theta|\geq 1, 𝔪\mathfrak{m}\to\infty and 𝔏1\mathfrak{L}_{1}\to\infty, then the desired result follows at once from (8.6) and the fact that 𝔏1ϵ2d(g)2ϵ𝔪+𝔏2\mathfrak{L}_{1}\sim\epsilon^{2}d(g)^{2}\sim\epsilon\,\mathfrak{m}+\mathfrak{L}_{2} (cf. Lemma 7.6 (ii)).

If 𝔪\mathfrak{m}\to\infty and 𝔏11\mathfrak{L}_{1}\lesssim 1, then from the proof of Theorem 9.2 we see that d(g)|x|d(g)\sim|x|. Thus, by Lemma 7.6 (ii) again we have ϵd(g)+ϵ𝔪+𝔏21\epsilon\,d(g)+\epsilon\,\mathfrak{m}+\mathfrak{L}_{2}\lesssim 1, whence BND|x|2ed(g)24{\rm BND}\sim|x|^{-2}e^{-\frac{d(g)^{2}}{4}}. Consequently, the estimate (9.2) yields that pBNDp\sim{\rm BND}.

If 𝔪1\mathfrak{m}\lesssim 1, then by (10.9) we have p|x|2ed(g)24p\sim|x|^{-2}e^{-\frac{d(g)^{2}}{4}}. And from the relations (10.1) we see that d(g)|x|1d(g)\sim|x|\gg 1, u1|x|2u_{1}\lesssim|x|^{-2} and u2|x|1u_{2}\lesssim|x|^{-1}. Hence BND|x|2ed(g)24p{\rm BND}\sim|x|^{-2}e^{-\frac{d(g)^{2}}{4}}\sim p, since ϵ𝔪𝔏2ϵ𝔏2𝔏1\epsilon\,\mathfrak{m}\,\mathfrak{L}_{2}\,\lesssim\epsilon\,\mathfrak{L}_{2}\lesssim\mathfrak{L}_{1}. This completes the proof of the theorem. ∎

Thanks to this theorem, we can derive the precise estimates for the heat kernel for all points g=(x,t)N3,2g=(x,t)\in N_{3,2} at infinity. Notice that the properties (2.8) allow us to consider only specific points as in the following:

Corollary 11.4.

Assuming that t=(t1,t2,0)t=(t_{1},t_{2},0) with t1,t20t_{1},t_{2}\geq 0 and x=|x|e1x=|x|e_{1}, we have:

  1. (i)

    If |x|2|t||x|^{2}\lesssim|t|, then

    p(x,t)(1+t2|x||t|12)12|t|1ed(g)24,as |t|+.\displaystyle p(x,t)\sim(1+t_{2}\,|x|\,|t|^{-\frac{1}{2}})^{-\frac{1}{2}}\,|t|^{-1}\,e^{-\frac{d(g)^{2}}{4}},\qquad\mbox{as\,\,}|t|\to+\infty. (11.4)
  2. (ii)

    If |x|2|t||x|^{2}\gg|t|, then

    p(x,t)(1+|x|1t2+|x|12t114t212)1|x|2ed(g)24,as |x|+.p(x,t)\sim(1+|x|^{-1}t_{2}+|x|^{-\frac{1}{2}}\,t_{1}^{\frac{1}{4}}\,t_{2}^{\frac{1}{2}})^{-1}|x|^{-2}\,e^{-\frac{d(g)^{2}}{4}},\qquad\mbox{as\,\,}|x|\to+\infty. (11.5)
Proof.

Via a limit argument, it is enough to prove these estimates under Assumption (A) (cf. (2.18)) only. Recall that d(g)2|x|2=4𝔪=|x|2𝐇(1)d(g)^{2}-|x|^{2}=4\mathfrak{m}=|x|^{2}\,\mathbf{H}(1), and d(g)2|x|2+|t|d(g)^{2}\sim|x|^{2}+|t|.

For item (i), it suffices to use (11.2) and the facts that

ϵ1,𝔪|t|d(g)21,\epsilon\sim 1,\quad\mathfrak{m}\sim|t|\sim d(g)^{2}\gg 1,

deduced from Lemma 7.6 (iii).

For item (ii), using Lemma 7.6 (iv)-(v) we have 𝔪t1(1+u22u11)\mathfrak{m}\sim t_{1}(1+u_{2}^{2}\,u_{1}^{-1}). In fact, when |θ|1|\theta|\geq 1, this follows from the facts that 𝔪t1\mathfrak{m}\sim t_{1} and u22u11ϵ21u_{2}^{2}\,u_{1}^{-1}\lesssim\epsilon^{2}\lesssim 1 (by (7.38)); while when |θ|3|\theta|\leq 3, by (7.39) it holds that 𝔪|x|2u22\mathfrak{m}\sim|x|^{2}u_{2}^{2} and u22u11θ111u_{2}^{2}\,u_{1}^{-1}\sim\theta_{1}^{-1}\gtrsim 1, yielding the desired estimate as well. Consequently, appealing to (11.2) again gives

p(x,t)1+ϵ|x|1+ϵ|x|+ϵt2+ϵ|x|12t114t212|x|2ed(g)24.p(x,t)\sim\frac{1+\epsilon|x|}{1+\epsilon\,|x|+\epsilon\,t_{2}+\epsilon\,|x|^{\frac{1}{2}}\,t_{1}^{\frac{1}{4}}\,t_{2}^{\frac{1}{2}}}\,|x|^{-2}\,e^{-\frac{d(g)^{2}}{4}}. (11.6)

If ϵ|x|1\epsilon\,|x|\leq 1, then ϵ1\epsilon\ll 1. So together with the first estimate in (7.38), we get that: t1=|x|u1|x|/4|x|(ϵ|x|)|x|t_{1}=|x|u_{1}|x|/4\lesssim|x|(\epsilon|x|)\lesssim|x| and t2t1t_{2}\lesssim\sqrt{t_{1}}. As a result, we obtain p(x,t)|x|2exp{d(g)2/4}RHSof(11.5)p(x,t)\sim|x|^{-2}\exp\{-d(g)^{2}/4\}\sim{\rm RHS\,\,of\,\,\eqref{sim22}}. Conversely, if ϵ|x|1\epsilon\,|x|\geq 1, then 1+ϵ|x|ϵ|x|1+\epsilon\,|x|\sim\epsilon\,|x|. From this and (11.6) the estimate (11.5) follows immediately. ∎

Remark 11.5.

One can replace |t||t| by d(g)2d(g)^{2} in (11.4), and |x||x| by d(g)d(g) in (11.5) respectively.

Another consequence of Theorem 11.1 is the

11.3 Small-time asymptotics

To obtain the desired asymptotics, the uniform asymptotics of Theorem 11.1 (ii) and Remark 10.3 are sufficient. Since g=og=o is trivial, we may assume that gog\neq o. Recall that one can reduce the matters to the case where x=|x|e1x=|x|e_{1} and t=(t1,t2,0)t=(t_{1},t_{2},0) with t1,t20t_{1},t_{2}\geq 0.

The first corollary is due to Remark 10.3.

Corollary 11.6.

Let x0x\neq 0 and t=0t=0. Then

ph(x,t)=𝐂 4π𝐅(0,0)h72|x|2ed(g)24h(1+og(1)),ash0+.p_{h}(x,t)=\mathbf{C}\,4\pi\,\mathbf{F}(0,0)\,h^{-\frac{7}{2}}\,|x|^{-2}\,e^{-\frac{d(g)^{2}}{4h}}\,(1+o_{g}(1)),\quad{\rm as}\,\,h\to 0^{+}.

Applying Theorem 11.1 (ii) and (11.1), one obtains immediately the following Corollary 11.7, which shows in particular that (2.29) is still valid for gu<,+2g_{u}\in\mathbb{R}^{2}_{<,+}.

Corollary 11.7.

Let x0x\neq 0, t1,t2>0t_{1},t_{2}>0 and πt22|x|2t1\pi t_{2}^{2}\neq|x|^{2}t_{1}. Then as h0+h\to 0^{+},

ph(x,t)=𝐂(8π)32h3|θ|Missing Operator\displaystyle p_{h}(x,t)=\mathbf{C}\,(8\pi)^{\frac{3}{2}}\,h^{-3}\,\frac{|\theta|}{\sin|\theta|}\,\theta_{2}^{-1}[\mathrm{K}(\theta_{1},\theta_{2})\,\mathrm{K}_{3}(\theta_{1},\theta_{2})]^{-\frac{1}{2}}\,|x|^{-3}\,e^{-\frac{d(g)^{2}}{4h}}\,(1+o_{g}(1)),

where θ=(θ1,θ2,0):=(Λ1(4|x|2t1,4|x|2t2),0)\theta=(\theta_{1},\theta_{2},0):=(\Lambda^{-1}\left(4\,|x|^{-2}\,t_{1},4\,|x|^{-2}\,t_{2}\right),0).

To deduce the following four corollaries, we can appeal to Theorem 11.1 (ii) with an argument of limit. Indeed, by (7.35), Corollary 2.3 and the fact that 4𝔪=d(g)2|x|24\mathfrak{m}=d(g)^{2}-|x|^{2}, one can check that the approximating sequence (whose choice is similar as in the proof of Corollary 2.3) of the given point (xh,th)(\frac{x}{\sqrt{h}},\frac{t}{h}) satisfies the asymptotic condition uniformly provided hh is small enough, then a passage to limit in both sides of the asymptotic formula will give us the asserted results.

Corollary 11.8.

Let x0x\neq 0, t1=0t_{1}=0 and t2>0t_{2}>0. Then

ph(x,t)\displaystyle p_{h}(x,t) =𝐂 8π32h3Υ(r)32[Υ(r)]12t212|x|2ed(g)24h(1+og(1)),ash0+,\displaystyle=\mathbf{C}\,8\pi^{\frac{3}{2}}\,h^{-3}\,\Upsilon(r)^{\frac{3}{2}}\,[-\Upsilon^{\prime}(r)]^{-\frac{1}{2}}\,t_{2}^{-\frac{1}{2}}|x|^{-2}e^{-\frac{d(g)^{2}}{4h}}\,(1+o_{g}(1)),\quad{\rm as}\,\,h\to 0^{+},

where rr is the unique solution of μ(r)=4|x|2t2\mu(r)=4\,|x|^{-2}\,t_{2}.

Corollary 11.9.

Let x0x\neq 0, t1,t2>0t_{1},t_{2}>0 and πt22=|x|2t1\pi t_{2}^{2}=|x|^{2}t_{1}. Then

ph(x,t)=𝐂 22ππ3h3t21|x|(|x|4+π2t22)12ed(g)24h(1+og(1)),ash0+.p_{h}(x,t)=\mathbf{C}\,2\,\sqrt{2\pi}\,\pi^{3}\,h^{-3}\,t_{2}^{-1}\,|x|\,(|x|^{4}+\pi^{2}t_{2}^{2})^{-\frac{1}{2}}\,e^{-\frac{d(g)^{2}}{4h}}\,(1+o_{g}(1)),\quad{\rm as}\,\,h\to 0^{+}.
Corollary 11.10.

Let x0x\neq 0, t1>0t_{1}>0 and t2=0t_{2}=0. Then as h0+h\to 0^{+} it holds that

ph(x,t)=𝐂 16π2h72[Υ(r)r4Υ(r)ψ(r)[2rsin(2r)]+4r2ψ(r)]12|x|2ed(g)24h(1+og(1)),\displaystyle p_{h}(x,t)=\mathbf{C}\,16\pi^{2}\,h^{-\frac{7}{2}}\left[\frac{-\Upsilon^{\prime}(r)\,r^{4}}{\Upsilon^{\prime\prime}(r)\psi(r)\,[2r-\sin(2r)]+4r^{2}\,\psi^{\prime}(r)}\right]^{\frac{1}{2}}|x|^{-2}\,e^{-\frac{d(g)^{2}}{4h}}\,(1+o_{g}(1)),

where rr is the unique solution of the equation (2.26) with β=4|x|2t1\beta=4\,|x|^{-2}\,t_{1}.

Corollary 11.11.

Let x=0x=0 and t0t\neq 0. Then

ph(x,t)=𝐂 4π4h72|t|1ed(g)24h(1+og(1)),ash0+.p_{h}(x,t)=\mathbf{C}\,4\pi^{4}\,h^{-\frac{7}{2}}\,|t|^{-1}e^{-\frac{d(g)^{2}}{4h}}\,(1+o_{g}(1)),\quad{\rm as}\,\,h\to 0^{+}.

12 Sharp bounds for derivatives of the heat kernel

Recall that we use g=(x,t)g=(x,t) to denote an element in N3,2N_{3,2} and =(X1,X2,X3)\nabla=({\mathrm{X}}_{1},\,{\mathrm{X}}_{2},\,{\mathrm{X}}_{3}) to denote the horizontal gradient on this group, with Xi{\mathrm{X}}_{i} given by (2.1). Let X^i\widehat{{\mathrm{X}}}_{i} (1i31\leq i\leq 3) represent the corresponding right-invariant vector field, that is,

X^1=x1+12x3t212x2t3,X^2=x212x3t1+12x1t3,X^3=x3+12x2t112x1t2.\begin{gathered}\widehat{{\mathrm{X}}}_{1}=\frac{\partial}{\partial x_{1}}+\frac{1}{2}x_{3}\frac{\partial}{\partial t_{2}}-\frac{1}{2}x_{2}\frac{\partial}{\partial t_{3}},\quad\widehat{{\mathrm{X}}}_{2}=\frac{\partial}{\partial x_{2}}-\frac{1}{2}x_{3}\frac{\partial}{\partial t_{1}}+\frac{1}{2}x_{1}\frac{\partial}{\partial t_{3}},\\ \widehat{{\mathrm{X}}}_{3}=\frac{\partial}{\partial x_{3}}+\frac{1}{2}x_{2}\frac{\partial}{\partial t_{1}}-\frac{1}{2}x_{1}\frac{\partial}{\partial t_{2}}.\end{gathered}

Now set ^=(X^1,X^2,X^3)\widehat{\nabla}=(\widehat{{\mathrm{X}}}_{1},\widehat{{\mathrm{X}}}_{2},\widehat{{\mathrm{X}}}_{3}) and c=(,^)\nabla^{c}=(\nabla,\widehat{\nabla}). Then we have the following theorem. We mention that B. Qian has established |lnp(g)|d(g)|\nabla\ln{p(g)}|\lesssim d(g) for any free step-two Carnot group with kk generators (cf. [43, Proposition 5.5]), and its proof is based on the Harnack inequality and Bakry–Émery criterion (cf. [6] for more details on this approach).

Theorem 12.1.

It holds for any gN3,2g\in N_{3,2} that

|cp(g)|d(g)p(g),|(c)lp(g)|l(1+d(g))lp(g)for l=2,3,.|\nabla^{c}\,p(g)|\lesssim d(g)\,p(g),\qquad|(\nabla^{c})^{l}\,p(g)|\lesssim_{l}(1+d(g))^{l}\,p(g)\quad\mbox{for $l=2,3,\ldots$.}
Proof.

Recalling also that d(g)2|x|2+|t|d(g)^{2}\sim|x|^{2}+|t|, we start with the reduction of the problem. When d(g)d(g) is bounded, since pp is smooth and positive, we see the second estimate is trivial, and the first one follows at once from the the exponential decay of 𝐕(λ)\mathbf{V}(\lambda) as well as the facts that |x|d(g)|x|\lesssim d(g) and 0<r2(rcothr1)10<r^{-2}\,(r\coth r-1)\lesssim 1 for all r>0r>0. So only the case where d(g)d(g) is large needs handling with greater care, then it is enough to show the following

|αxβtp(g)|α,βd(g)|α|p(g),gN3,2,α,β3.|\partial^{\alpha}_{x}\partial^{\beta}_{t}p(g)|\lesssim_{\alpha,\beta}d(g)^{|\alpha|}\,p(g),\qquad\forall\,g\in N_{3,2},\,\alpha,\beta\in\mathbb{N}^{3}. (12.1)

By an argument of limit and the second equation of (2.8), we only need to prove (12.1) under the following assumption:

x0,xt>0,ttx|x|x|x|0,π|ttx|x|x|x||2(xt)|x|.\displaystyle x\neq 0,\qquad x\cdot t>0,\qquad t-\frac{t\cdot x}{|x|}\frac{x}{|x|}\neq 0,\qquad\pi\left|t-\frac{t\cdot x}{|x|}\frac{x}{|x|}\right|^{2}\neq(x\cdot t)|x|. (12.2)

Note that this assumption is equivalent to that we can find a suitable orthogonal matrix OgO_{g} such that g~=(Ogx,Ogt)\widetilde{\mathrm{g}}=(O_{g}\,x,O_{g}\,t) satisfies Assumption (A) (cf. (2.18)). Notice that there exists only one such g~\widetilde{\mathrm{g}}, we have also Ogx=|x|e1O_{g}\,x=|x|\,e_{1}, d(g)=d(g~)d(g)=d(\widetilde{\mathrm{g}}), and 𝔪(g~)=(d(g)2|x|2)/4=𝔪(g)\mathfrak{m}(\widetilde{\mathrm{g}})=(d(g)^{2}-|x|^{2})/4=\mathfrak{m}(g). In principle, the methods in establishing the asymptotics of the heat kernel should be enough to estimate the sharp bounds of its derivatives. As before, we want to reduce the point to Assumption (A) naturally. However, it turns out that the orthogonal matrix OgO_{g} may also depend on gg and we should be very careful. In fact, we prove (12.1) in the following four cases (conditions are stated for the point g~\widetilde{\mathrm{g}}):

Case (1): |θ(g~)|3|\theta(\widetilde{\mathrm{g}})|\leq 3 and θ2(g~)|x|+\theta_{2}(\widetilde{\mathrm{g}})|x|\to+\infty.

Case (2): 𝔪(g~)1\mathfrak{m}(\widetilde{\mathrm{g}})\lesssim 1 and d(g)+d(g)\to+\infty.

Case (3): |x|d(g)1|x|\,d(g)\leq 1 and d(g)+d(g)\to+\infty.

Case (4): |x|d(g)1|x|\,d(g)\geq 1, |θ(g~)|1|\theta(\widetilde{\mathrm{g}})|\geq 1 and 𝔪(g~)+\mathfrak{m}(\widetilde{\mathrm{g}})\to+\infty.

For Cases (1)-(3), it transpires that one can first directly taking derivatives in the original expression of the heat kernel (cf. (2.3)), and can reduce the proof of (12.1) to the one of (12.3) or (12.5) below. Then performing the orthogonal transform as before solves the problem. However, we should pay attention to Case (3) since there emerges singularity of higher order in the amplitude owing to derivation. Unlike the known cases (Heisenberg groups, non-isotropic Heisenberg groups, and GM-Métivier groups), in Case (3) we neither have the estimates for higher order singularities in the amplitude on N3,2N_{3,2}, nor can apply the technique in Case (4) to deal with it. So we have to establish the upper bound estimates for higher order singularities. Fortunately, the method to obtain the uniform heat kernel can be modified to overcome this difficulty in an elegant manner, though the integral in this case is converted to a higher dimensional one.

For the remaining Case (4), we first choose an orthogonal matrix O=OgO=O_{g} in (2.8) to enable us to use the new formula (2.30). However, in this way we should be careful since we need also take derivatives on such orthogonal matrix and thus we require a lower bound of |x||x| (namely d(g)1d(g)^{-1}, which indeed can be observed from (12.1)) to treat it.

12.1 Proof of (12.1) in Cases (1)-(3)

In both Cases (1)-(2), by the fact that |x|d(g)|x|\lesssim d(g), it suffices to establish the following estimate

|pn0,α0(x,t)|n0,α0p(x,t),n0,α03\displaystyle|p^{n_{0},\alpha_{0}}(x,t)|\lesssim_{n_{0},\alpha_{0}}p(x,t),\qquad\forall\,n_{0}\in\mathbb{N},\,\alpha_{0}\in\mathbb{N}^{3} (12.3)

where g=(x,t)g=(x,t) as in (12.2), and

pn0,α0(x,t):=3𝐕(λ)(|λ|coth|λ|1|λ|2)n0λα0e14ϕ~((x,t);λ)dλ.\displaystyle p^{n_{0},\alpha_{0}}(x,t):=\int_{\mathbb{R}^{3}}\mathbf{V}(\lambda)\left(\frac{|\lambda|\coth{|\lambda|}-1}{|\lambda|^{2}}\right)^{n_{0}}\lambda^{\alpha_{0}}\,e^{-\frac{1}{4}\widetilde{\phi}((x,t);\lambda)}\,d\lambda. (12.4)

Notice that from the definition above we have |pn0,α0(g)|α0|α|=|α0||pn0,α(g~)||p^{n_{0},\alpha_{0}}(g)|\lesssim_{\alpha_{0}}\sum_{|\alpha|=|\alpha_{0}|}|p^{n_{0},\alpha}(\widetilde{\mathrm{g}})| by means of an orthogonal transform. Combining this with the fact that p(g)=p(g~)p(g)=p(\widetilde{\mathrm{g}}). It remains to prove (12.3) for gg as in Assumption (A) (cf. (2.18)). Note that we are only concerned with the upper bound, the desired estimate can be deduced easily by using an argument similar to (and simpler than) that of Theorem 6.1 for Case (1), and that of Theorem 10.1 for Case (2).

For Case (3), some modification is now required. Firstly, by the assumption that |x|d(g)1|x|\leq d(g)^{-1}, it remains to prove

|pn0,α0(x,t)|n0,α0d(x,t)2n0p(x,t),n0,α03.\displaystyle|p^{n_{0},\alpha_{0}}(x,t)|\lesssim_{n_{0},\alpha_{0}}d(x,t)^{2n_{0}}\,p(x,t),\qquad\forall\,n_{0}\in\mathbb{N},\,\alpha_{0}\in\mathbb{N}^{3}. (12.5)

As before we can reduce the proof to the points satisfying (2.18) or equivalently Assumption (A). Secondly, it follows from |x|d(g)1|x|\leq d(g)^{-1} that d(g)|t|1d(g)\sim\sqrt{|t|}\gg 1 and |u|1|u|\gg 1, which occurs only in the case of Lemma 7.6 (iii). As a result, we obtain

𝐰¯|t|/|x||u|,|x|2𝐰¯2𝔪d(g)2,|π|θ|||u|121.\mathbf{\overline{w}}\sim\sqrt{|t|}/|x|\sim\sqrt{|u|},\quad|x|^{2}\,\mathbf{\overline{w}}^{2}\sim\mathfrak{m}\sim d(g)^{2},\quad|\pi-|\theta|\,|\lesssim|u|^{-\frac{1}{2}}\ll 1. (12.6)

Moreover, from Corollary 11.4 (i) it follows that

p(x,t)|t|1ed(g)24d(g)2ed(g)24.p(x,t)\sim|t|^{-1}e^{-\frac{d(g)^{2}}{4}}\sim d(g)^{-2}\,e^{-\frac{d(g)^{2}}{4}}.

Thirdly, we need a more suitable integral expression for pn0,α0p^{n_{0},\alpha_{0}}, which will play an important role. To be more precise, applying (4.3) with q=2+2n0q=2+2n_{0}, Y=|x|2𝐰¯2(λ2,λ3,0,,0)××2n0Y=\frac{|x|^{2}\mathbf{\overline{w}}}{2}(\lambda_{2},\lambda_{3},0,\ldots,0)\in\mathbb{R}\times\mathbb{R}\times\mathbb{R}^{2n_{0}} and A=12|λ|2|λ|coth|λ|1|x|2𝐰¯2𝕀2+2n0A=\frac{1}{2}\frac{|\lambda|^{2}}{|\lambda|\coth|\lambda|-1}\,|x|^{2}\,\mathbf{\overline{w}}^{2}\,{\mathbb{I}}_{2+2n_{0}}, respectively, we find that

pn0,α0(x,t)=(|x|2𝐰¯2)1+n0(4π)1+n0e|x|242+2n0𝐏n0,α0(s|x|𝐰¯,14|x|2(u+2𝐰¯s1e2+2𝐰¯s2e3))dsp^{n_{0},\alpha_{0}}(x,t)=\frac{(|x|^{2}\,\mathbf{\overline{w}}^{2})^{1+n_{0}}}{(4\pi)^{1+n_{0}}}e^{-\frac{|x|^{2}}{4}}\,\int_{\mathbb{R}^{2+2n_{0}}}\mathbf{P}^{n_{0},\alpha_{0}}\left(s|x|\,\mathbf{\overline{w}},\frac{1}{4}|x|^{2}(u+2\mathbf{\overline{w}}s_{1}e_{2}+2\mathbf{\overline{w}}s_{2}e_{3})\right)\,ds

where

𝐏n0,α0(X,T):=3𝒱(λ)λα0e14Γ~((X,T);λ)dλ,(X,T)2+2n0×3.\mathbf{P}^{n_{0},\alpha_{0}}(X,T):=\int_{\mathbb{R}^{3}}\mathcal{V}(\lambda)\,\lambda^{\alpha_{0}}\,e^{-\frac{1}{4}\widetilde{\Gamma}((X,T);\lambda)}\,d\lambda,\qquad(X,T)\in\mathbb{R}^{2+2n_{0}}\times\mathbb{R}^{3}.

A similar and simpler argument used in Appendix A yields the following upper bound for 𝐏n0,α0\mathbf{P}^{n_{0},\alpha_{0}}:

|𝐏n0,α0(X,T)|α0exp{𝐃(X,T)24}(1+𝐃(X,T))2(1+|X|𝐃(X,T))12,|\mathbf{P}^{n_{0},\alpha_{0}}(X,T)|\lesssim_{\alpha_{0}}\frac{\exp\left\{-\frac{\mathbf{D}(X,T)^{2}}{4}\right\}}{(1+\mathbf{D}(X,T))^{2}(1+|X|\mathbf{D}(X,T))^{\frac{1}{2}}}, (12.7)

where 𝐃\mathbf{D} is now is defined by (4.12) with X2+2n0X\in\mathbb{R}^{2+2n_{0}}.

Now adopting similar notation as in Subsection 7.2, we write

𝒟n0(s):=𝐃(s𝐰¯,14(u+2𝐰¯s1e2+2𝐰¯s2e3))2,s=(s1,s2,s)××2n0,\displaystyle\mathcal{D}_{n_{0}}(s):=\mathbf{D}\left(s\,\mathbf{\overline{w}},\frac{1}{4}(u+2\mathbf{\overline{w}}s_{1}e_{2}+2\mathbf{\overline{w}}s_{2}e_{3})\right)^{2},\quad s=(s_{1},s_{2},s^{\prime})\in\mathbb{R}\times\mathbb{R}\times{\mathbb{R}}^{2n_{0}},
𝐇0(w):=𝐰¯2|w|2Φ(𝐔0(w))with𝐔0(w):=𝐀(w1)𝐰¯2|w|2,w=(w1,w2)(0,)×(0,).\displaystyle\mathbf{H}_{0}(\mathrm{w}):=\mathbf{\overline{w}}^{2}|\mathrm{w}|^{2}\,\Phi(\mathbf{U}_{0}(\mathrm{w}))\,\,{\rm with}\,\,\mathbf{U}_{0}(\mathrm{w}):=\frac{\mathbf{A}(\mathrm{w}_{1})}{\mathbf{\overline{w}}^{2}|\mathrm{w}|^{2}},\quad\mathrm{w}=(\mathrm{w}_{1},\mathrm{w}_{2})\in(0,\infty)\times(0,\infty).

Then 𝒟n0(s)𝐇0(|(s1,s2)|,|s|)\mathcal{D}_{n_{0}}(s)\geq\mathbf{H}_{0}(|(s_{1},s_{2})|,|s^{\prime}|). From the above estimates and using polar coordinates (7.19) for (s1,s2)(s_{1},s_{2}) and 2n02n_{0}-dimensional ones for ss^{\prime} respectively, together with the fact that d(g)2=|x|2+4𝔪d(g)^{2}=|x|^{2}+4\mathfrak{m}, we deduce that (12.5) will hold provided that

00exp{|x|24𝐇0(w)}w1w22n01(1+|x|2𝐇0(w))1+|x|2𝐰¯|w|𝐇0(w)dw1dw2𝔪2e𝔪.\int_{0}^{\infty}\int_{0}^{\infty}\frac{\exp\left\{-\frac{|x|^{2}}{4}\mathbf{H}_{0}(\mathrm{w})\right\}\,\mathrm{w}_{1}\mathrm{w}_{2}^{2n_{0}-1}}{\left(1+|x|^{2}\mathbf{H}_{0}(\mathrm{w})\right)\sqrt{1+|x|^{2}\,\mathbf{\overline{w}}|\mathrm{w}|\sqrt{\mathbf{H}_{0}(\mathrm{w})}}}\,d\mathrm{w}_{1}d\mathrm{w}_{2}\lesssim\mathfrak{m}^{-2}\,e^{-\mathfrak{m}}. (12.8)

To prove this we write LHS of (12.8) as 𝒪1+𝒪2+𝒪3\int_{\mathcal{O}_{1}}+\int_{\mathcal{O}_{2}}+\int_{\mathcal{O}_{3}} with

𝒪1:={w;|w|δ0},𝒪2:={w;δ0<|w|<C0},𝒪3:={w;|w|C0},\mathcal{O}_{1}:=\left\{\mathrm{w};|\mathrm{w}|\leq\delta_{0}\right\},\quad\mathcal{O}_{2}:=\left\{\mathrm{w};\delta_{0}<|\mathrm{w}|<C_{0}\right\},\quad\mathcal{O}_{3}:=\left\{\mathrm{w};|\mathrm{w}|\geq C_{0}\right\},

where δ01\delta_{0}\ll 1 and C01C_{0}\gg 1 to be determined later. The estimate 𝒪3𝔪2e𝔪\int_{\mathcal{O}_{3}}\lesssim\mathfrak{m}^{-2}\,e^{-\mathfrak{m}} follows from a similar argument as in the estimate of 𝒬3{\mathcal{Q}}_{3} in Subsection 8.1 with C0C_{0} large enough. The estimate for 𝒪1\int_{\mathcal{O}_{1}} is also similar to that by choosing δ0\delta_{0} small enough and using the following assertion: 0.841|x|2𝐇0(w)𝔪0.8\cdot 4^{-1}|x|^{2}\,\mathbf{H}_{0}(\mathrm{w})\geq\mathfrak{m} for w𝒪1\mathrm{w}\in\mathcal{O}_{1}. In fact, from (7.21) and (12.6) we can check that

𝐀(w1)=|u|(1+OC0(|u|12))𝐀(|w|)𝐰¯2|u|,|w|<C0.\mathbf{A}(\mathrm{w}_{1})=|u|(1+O_{C_{0}}(|u|^{-\frac{1}{2}}))\sim\mathbf{A}(|\mathrm{w}|)\sim\mathbf{\overline{w}}^{2}\sim|u|,\quad\forall\,|\mathrm{w}|<C_{0}. (12.9)

Therefore, 𝐔0(w)|w|2δ02\mathbf{U}_{0}(\mathrm{w})\sim|\mathrm{w}|^{-2}\gtrsim\delta_{0}^{-2} on 𝒪1\mathcal{O}_{1}. By this and (7.7), we get that:

𝐇0(w)\displaystyle\mathbf{H}_{0}(\mathrm{w}) =ϑ1𝐰¯2|w|2𝐔0(w)(1+O(𝐔0(w)12))\displaystyle=\vartheta_{1}\mathbf{\overline{w}}^{2}\,|\mathrm{w}|^{2}\,\mathbf{U}_{0}(\mathrm{w})\,\Big{(}1+O(\mathbf{U}_{0}(\mathrm{w})^{-\frac{1}{2}})\Big{)}
=ϑ1|u|[1+OC0(|u|12)+O(δ0)]=ϑ1|u|(1+o(1)).\displaystyle=\vartheta_{1}|u|\,\left[1+O_{C_{0}}(|u|^{-\frac{1}{2}})+O(\delta_{0})\right]=\vartheta_{1}|u|\,(1+o(1)).

However, recalling φ1(π)=π\varphi_{1}(\pi)=\pi (cf. (2.23)), the third equality in (2.36) implies that 𝐇(1)φ1(|θ|)|u|=π|u|(1+o(1))\mathbf{H}(1)\leq\varphi_{1}(|\theta|)|u|=\pi|u|(1+o(1)) since |θ|π|\theta|\to\pi by the third estimate in (12.6), which readily yields the assertion.

To estimate 𝒪2\int_{\mathcal{O}_{2}}, first recall that |u|1|u|\gg 1 and |x|d(g)1|x|\,d(g)\leq 1. Fixing selected δ0\delta_{0} and C0C_{0}, then one can conclude similarly as in the proof of (8.32) (a rougher argument is enough) that

||x|2𝐇0(w)|x|2𝐇(|w|)|1,w𝒪2.\left|\,|x|^{2}\mathbf{H}_{0}(\mathrm{w})-|x|^{2}\mathbf{H}(|\mathrm{w}|)\,\right|\lesssim 1,\quad\forall\,\mathrm{w}\in\mathcal{O}_{2}. (12.10)

In particular, |x|2𝐇0(w)|x|2𝐇(|w|)4𝔪1|x|^{2}\mathbf{H}_{0}(\mathrm{w})\sim|x|^{2}\mathbf{H}(|\mathrm{w}|)\geq 4\mathfrak{m}\gg 1. Applying these estimates with the 22-dimensional polar coordinates w=ρ(cosγ,sinγ)\mathrm{w}=\rho\,(\cos\gamma,\sin\gamma) to 𝒪2\int_{\mathcal{O}_{2}}, we infer it is bounded by

𝔪32δ0C0ρ2n0exp{|x|24𝐇(ρ)}dρ.\mathfrak{m}^{-\frac{3}{2}}\int_{\delta_{0}}^{C_{0}}\rho^{2n_{0}}\exp\left\{-\frac{|x|^{2}}{4}\mathbf{H}(\rho)\right\}\,d\rho.

Recalling that 11 is the only minimal point of 𝐇\mathbf{H} with 𝔏1=|x|2𝐇(1)/4d(g)2𝔪\mathfrak{L}_{1}=|x|^{2}\mathbf{H}^{\prime\prime}(1)/4\sim d(g)^{2}\sim\mathfrak{m}, so it is exactly majorized by 𝔪2e𝔪\mathfrak{m}^{-2}e^{-\mathfrak{m}} via the standard Laplace’s method. This finishes the proof of Case (3).

12.2 Proof of (12.1) in Case (4)

To prove (12.1) in this case, the following estimate is crucial:

2|sι𝐏α(s)|dsι,αd(g)|ι|p(g),ι2,α3,\int_{{\mathbb{R}}^{2}}\left|s^{\iota}\,\mathbf{P}_{\alpha}(s)\right|\,ds\lesssim_{\iota,\alpha}d(g)^{|\iota|}\,p(g),\quad\forall\,\iota\in{\mathbb{N}}^{2},\,\alpha\in{\mathbb{N}}^{3}, (12.11)

where |θ(g~)|1|\theta(\widetilde{\mathrm{g}})|\geq 1, 𝔪(g~)\mathfrak{m}(\widetilde{\mathrm{g}})\to\infty, and

𝐏α(s):=1πe|x|24𝐏α(2s,t+|x|s1e2+|x|s2e3),s2,\mathbf{P}_{\alpha}(s):=\frac{1}{\pi}\,e^{-\frac{|x|^{2}}{4}}\,\mathbf{P}_{\alpha}\left(2s,t+|x|s_{1}\,e_{2}+|x|s_{2}\,e_{3}\right),\quad s\in{\mathbb{R}}^{2},

with

𝐏α(X,T)=𝐏0,α(X,T)=3𝒱(λ)λαe14Γ~((X,T);λ)dλ,(X,T)2×3.\mathbf{P}_{\alpha}(X,T)=\mathbf{P}^{0,\alpha}(X,T)=\int_{\mathbb{R}^{3}}\,\mathcal{V}(\lambda)\,\lambda^{\alpha}\,e^{-\frac{1}{4}\widetilde{\Gamma}((X,T);\lambda)}\,d\lambda,\quad(X,T)\in{\mathbb{R}}^{2}\times{\mathbb{R}}^{3}.

Remark that the upper bound (12.7) is valid for |𝐏α(X,T)||\mathbf{P}_{\alpha}(X,T)|. Then following a similar line of reasoning as in Sections 8 and 9, we can obtain (12.11) under the condition that |θ(g~)|1|\theta(\widetilde{\mathrm{g}})|\geq 1 and 𝔪(g~)\mathfrak{m}(\widetilde{\mathrm{g}})\to\infty.

The rest proof of (12.1) is then simple based on (12.11) and we shall go on in the following three steps. Firstly, without loss of generality we may assume that |x|2x22+x32|x|2/3>0|x|^{2}\geq x_{2}^{2}+x_{3}^{2}\geq|x|^{2}/3>0, then taking the orthogonal transform

O0(x):=(|x|1x1|x|1x2|x|1x30x3x22+x32x2x22+x32x22+x32|x|x1x2|x|x22+x32x1x3|x|x22+x32)O_{0}(x):=\begin{pmatrix}|x|^{-1}x_{1}&|x|^{-1}x_{2}&|x|^{-1}x_{3}\\ 0&\frac{x_{3}}{\sqrt{x_{2}^{2}+x_{3}^{2}}}&\frac{-x_{2}}{\sqrt{x_{2}^{2}+x_{3}^{2}}}\\ -\frac{\sqrt{x_{2}^{2}+x_{3}^{2}}}{|x|}&\frac{x_{1}x_{2}}{|x|\sqrt{x_{2}^{2}+x_{3}^{2}}}&\frac{x_{1}x_{3}}{|x|\sqrt{x_{2}^{2}+x_{3}^{2}}}\end{pmatrix} (12.12)

in (2.8) and applying (4.3) to (2.3) we arrive at that

p(g)=1πe|x|242𝐏(2s,t~+|x|s1e2+|x|s2e3)ds,p(g)=\frac{1}{\pi}\,e^{-\frac{|x|^{2}}{4}}\,\int_{\mathbb{R}^{2}}\mathbf{P}\left(2s,\tilde{t}+|x|s_{1}\,e_{2}+|x|s_{2}\,e_{3}\right)ds, (12.13)

where

t~:=O0(x)t=(xt|x|,t2x3t3x2x22+x32,t1(x22+x32)+t2x1x2+t3x1x3|x|x22+x32).\tilde{t}:=O_{0}(x)\,t=\left(\frac{x\cdot t}{|x|},\,\frac{t_{2}x_{3}-t_{3}x_{2}}{\sqrt{x_{2}^{2}+x_{3}^{2}}},\,\frac{-t_{1}(x_{2}^{2}+x_{3}^{2})+t_{2}\,x_{1}x_{2}+t_{3}\,x_{1}x_{3}}{|x|\sqrt{x_{2}^{2}+x_{3}^{2}}}\right).

Note that to reduce such points to ones obeying (2.9), one only needs to use some orthogonal transform (in 𝐏\mathbf{P} or 𝐏α\mathbf{P}_{\alpha}) again. This will be used implicitly in what follows. Moreover, the parameter t~\tilde{t}, viewed as a function of (x,t)(x,t), is homogeneous on xx of degree 0 and on tt of degree 11 respectively.

Secondly, suppose for the moment that |x|2|t||x|^{2}\geq|t|. Take derivatives in (12.13). By Leibniz’s rule, the chain rule, (12.11) and the homogeneous property of t~\tilde{t}, we see that to show (12.1) it is enough to observe that, for instance, for any 1i,j31\leq i,j\leq 3,

|xi(t~)||x|1|t|d(g),|ti(t~)|1;\displaystyle|\partial_{x_{i}}(\tilde{t})|\lesssim|x|^{-1}|t|\lesssim d(g),\quad|\partial_{t_{i}}(\tilde{t})|\lesssim 1;
|xixj(t~)||x|2|t|d(g)2,|xitj(t~)||x|1d(g),|titj(t~)|=0.\displaystyle|\partial_{x_{i}}\partial_{x_{j}}(\tilde{t})|\lesssim|x|^{-2}|t|\lesssim d(g)^{2},\quad|\partial_{x_{i}}\partial_{t_{j}}(\tilde{t})|\lesssim|x|^{-1}\lesssim d(g),\quad|\partial_{t_{i}}\partial_{t_{j}}(\tilde{t})|=0.

Other possible situations can also be verified easily.

Thirdly, let us improve the above result to the desired case |x|d(g)1|x|\,d(g)\geq 1, for which we only need to cope with the remaining case |x|2|t||x|^{2}\leq|t|. In particular one has |t|d(g)21|t|\sim d(g)^{2}\gg 1. We may assume t~22+t~32|t|2/3\tilde{t}_{2}^{2}+\tilde{t}_{3}^{2}\geq|t|^{2}/3 as well, since only homogeneous property plays a role when estimating the derivatives of t~\tilde{t}. Notice also that a change of variables λOλ\lambda\mapsto O\lambda in the definition of 𝐏\mathbf{P} (cf. (4.4)) gives the following

𝐏(X,T)=𝐏(X,OT),OO3.\mathbf{P}(X,T)=\mathbf{P}(X,O\,T),\quad\forall\,O\in\mathrm{O}_{3}.

Then applying the previous formula to (12.13) with O=O0(t~)O=O_{0}(\tilde{t}) (cf. (12.12)), we get:

p(g)=1πe|x|242𝐏(2s,|t|+|x|t~2s1+t~3s2|t|,|x|t~3s1t~2s2t~22+t~32,|x|t~1t~2s1+t~1t~3s2|t|t~22+t~32)ds.p(g)=\frac{1}{\pi}\,e^{-\frac{|x|^{2}}{4}}\,\int_{\mathbb{R}^{2}}\mathbf{P}\left(2s,\,|t|+|x|\,\frac{\tilde{t}_{2}\,s_{1}+\tilde{t}_{3}\,s_{2}}{|t|},\,|x|\,\frac{\tilde{t}_{3}\,s_{1}-\tilde{t}_{2}\,s_{2}}{\sqrt{\tilde{t}_{2}^{2}+\tilde{t}_{3}^{2}}},\,|x|\,\frac{\tilde{t}_{1}\tilde{t}_{2}\,s_{1}+\tilde{t}_{1}\tilde{t}_{3}\,s_{2}}{|t|\sqrt{\tilde{t}_{2}^{2}+\tilde{t}_{3}^{2}}}\right)ds. (12.14)

Observe that all coefficients of ss in the TT-variable of (12.14) is homogeneous on xx of degree 11 and on tt of degree 0. Then using a similar argument as in the case |x|2|t||x|^{2}\geq|t|, one can easily check that the condition |x|d(g)1|x|\,d(g)\geq 1 and the estimate (12.11) are sufficient for (12.1). This proves Case (4), and hence Theorem 12.1. ∎

Appendix A Proof of Proposition 2.9

Recall the functions 𝐏(X,T)\mathbf{P}(X,T), 𝐃(X,T)2\mathbf{D}(X,T)^{2} and τ(X,T)\tau^{*}(X,T) introduced in Subsections 2.5 and 4.1-4.2, and set ϵ:=ϑ1|τ|\epsilon_{*}:=\vartheta_{1}-|\tau^{*}|. Then (7.9) implies that

𝐃(X,T)2|X|2(ϑ1|τ|)2=ϵ2|X|2,if|X|0.\displaystyle\mathbf{D}(X,T)^{2}\sim\frac{|X|^{2}}{(\vartheta_{1}-|\tau^{*}|)^{2}}=\epsilon_{*}^{-2}\,|X|^{2},\quad\mbox{if}\ |X|\neq 0. (A.1)

Let us now begin with the positivity of 𝐏\mathbf{P}. Indeed, the argument below also allows us to establish [37, Proposition 2.2]. The fact is implicitly indicated therein.

A.1 Proof of Proposition 2.9 (i)

Proof.

It follows from (4.4)-(4.6) that

𝐏(X,T)=3e34|X|2liml0+3eiTλl=1l0𝐏l(X;λ)dλ,\displaystyle\mathbf{P}(X,T)=3\,e^{-\frac{3}{4}|X|^{2}}\,\lim_{l_{0}\to+\infty}\int_{\mathbb{R}^{3}}e^{iT\cdot\lambda}\,\prod_{l=1}^{l_{0}}\mathbf{P}_{l}(X;\lambda)\,d\lambda, (A.2)

where

𝐏l(X;λ):=(1+|λ|2ϑl2)1exp{12(1+|λ|2ϑl2)1|X|212|X|2}.\displaystyle\mathbf{P}_{l}(X;\lambda):=\left(1+\frac{|\lambda|^{2}}{\vartheta_{l}^{2}}\right)^{-1}\,\exp\left\{\frac{1}{2}\left(1+\frac{|\lambda|^{2}}{\vartheta_{l}^{2}}\right)^{-1}|X|^{2}-\frac{1}{2}\,|X|^{2}\right\}.

We apply (4.3) with q=2q=2, A=(1+|λ|2ϑl2)𝕀2A=\left(1+\frac{|\lambda|^{2}}{\vartheta_{l}^{2}}\right){\mathbb{I}}_{2} and Y=iXY=-iX, respectively, to get that

𝐏l(X;λ)=12π2e12|slX|2e|λ|22ϑl2|sl|2dsl.\displaystyle\mathbf{P}_{l}(X;\lambda)=\frac{1}{2\pi}\,\int_{\mathbb{R}^{2}}e^{-\frac{1}{2}|s_{l}-X|^{2}}e^{-\frac{|\lambda|^{2}}{2\vartheta_{l}^{2}}|s_{l}|^{2}}ds_{l}.

As a result, for every j0,l0j_{0},l_{0}\in\mathbb{N}^{*} with l0j01l_{0}-j_{0}\geq 1, X2X\in\mathbb{R}^{2}, and T3T\in\mathbb{R}^{3}, we obtain

Q(j0,l0;X,T):=\displaystyle Q(j_{0},l_{0};X,T):= 3eiTλl=j0l0𝐏l(X;λ)dλ\displaystyle\int_{\mathbb{R}^{3}}e^{iT\cdot\lambda}\,\prod_{l=j_{0}}^{l_{0}}\mathbf{P}_{l}(X;\lambda)\,d\lambda
=\displaystyle= 1(2π)l0j0+1(2)l0j0+1l=j0l0e12|slX|2ds3exp{|λ|22l=j0l0|sl|2ϑl2+iTλ}dλ\displaystyle\frac{1}{(2\pi)^{l_{0}-j_{0}+1}}\,\int_{(\mathbb{R}^{2})^{l_{0}-j_{0}+1}}\prod_{l=j_{0}}^{l_{0}}e^{-\frac{1}{2}|s_{l}-X|^{2}}ds\int_{\mathbb{R}^{3}}\exp\left\{-\frac{|\lambda|^{2}}{2}\sum_{l=j_{0}}^{l_{0}}\frac{|s_{l}|^{2}}{\vartheta_{l}^{2}}+iT\cdot\lambda\right\}d\lambda
=\displaystyle= 1(2π)l0j012(2)l0j0+1l=j0l0e12|slX|2(l=j0l0|sl|2ϑl2)32exp{|T|22l=j0l0|sl|2ϑl2}ds>0,\displaystyle\frac{1}{(2\pi)^{l_{0}-j_{0}-\frac{1}{2}}}\,\int_{(\mathbb{R}^{2})^{l_{0}-j_{0}+1}}\frac{\prod\limits_{l=j_{0}}^{l_{0}}e^{-\frac{1}{2}|s_{l}-X|^{2}}\,}{\left(\sum\limits_{l=j_{0}}^{l_{0}}\frac{|s_{l}|^{2}}{\vartheta_{l}^{2}}\right)^{\frac{3}{2}}}\,\exp\left\{-\frac{|T|^{2}}{2\sum\limits_{l=j_{0}}^{l_{0}}\frac{|s_{l}|^{2}}{\vartheta_{l}^{2}}}\right\}\,ds>0,

where s=(sj0,,sl0)(2)l0j0+1s=(s_{j_{0}},\ldots,s_{l_{0}})\in(\mathbb{R}^{2})^{l_{0}-j_{0}+1} and we have used (4.3) again in the last “==”. Hence we get that Q(1,2;X,T)>0Q(1,2;X,T)>0, and by induction Q(3,+;X,T)0Q(3,+\infty;X,T)\geq 0.

It remains to show that Q(1,+;X,T)>0Q(1,+\infty;X,T)>0. In fact, the basic properties of the Fourier transform and convolution give that

Q(1,+;X,T)=(2π)33Q(1,2;X,ξ)Q(3,+;X,Tξ)dξ.\displaystyle Q(1,+\infty;X,T)=(2\pi)^{-3}\,\int_{\mathbb{R}^{3}}Q(1,2;X,\xi)\,Q(3,+\infty;X,T-\xi)\,d\xi. (A.3)

And it suffices to recall the continuous function Q(1,2;X,)>0Q(1,2;X,\cdot)>0 and notice that the nonnegative continuous function Q(3,+;X,)Q(3,+\infty;X,\cdot) satisfies Q(3,+;X,0)>0Q(3,+\infty;X,0)>0. ∎

A.2 Proof of Proposition 2.9 (ii) and (iii)

It is enough to prove (iii) in detail, since (ii) will follow easily from it and the fact that 𝐏\mathbf{P} is positive and continuous we have just obtained, with estimates (7.3), (7.15) and (A.1).

To show (iii), we consider two cases as in the proposition below:

Proposition A.1.

The following uniform asymptotic estimates hold:

  1. (I)

    If X0X\neq 0, 𝐃(X,T)+\mathbf{D}(X,T)\to+\infty, and ϑ1|τ|1\vartheta_{1}-|\tau^{*}|\gtrsim 1, then we have

    𝐏(X,T)=(8π)32e𝐃(X,T)24det(HessτΓ((X,T);τ))12𝒱(iτ)(1+o(1)).\displaystyle\mathbf{P}(X,T)=\frac{(8\pi)^{\frac{3}{2}}\,e^{-\frac{\mathbf{D}(X,T)^{2}}{4}}}{\det(-\mathrm{Hess}_{\tau^{*}}\,\Gamma((X,T);\tau^{*}))^{\frac{1}{2}}}\,\mathcal{V}(i\tau^{*})\,(1+o(1)). (A.4)
  2. (II)

    If X0X\neq 0, 𝐃(X,T)+\mathbf{D}(X,T)\to+\infty and |τ|ϑ1|\tau^{*}|\to\vartheta_{1}^{-}, then we have

    𝐏(X,T)\displaystyle\mathbf{P}(X,T) =(2π)24ϑ12sinϑ1e𝐃(X,T)24eϑ1|X|22(ϑ1|τ|)I0(ϑ1|X|22(ϑ1|τ|))(ϑ1|τ|)2|X|2(1+o(1)).\displaystyle=(2\pi)^{2}\,\frac{4\vartheta_{1}^{2}}{-\sin{\vartheta_{1}}}\,e^{-\frac{\mathbf{D}(X,T)^{2}}{4}}\,e^{-\frac{\vartheta_{1}|X|^{2}}{2(\vartheta_{1}-|\tau^{*}|)}}I_{0}\left(\frac{\vartheta_{1}|X|^{2}}{2(\vartheta_{1}-|\tau^{*}|)}\right)\frac{(\vartheta_{1}-|\tau^{*}|)^{2}}{|X|^{2}}\,(1+o(1)). (A.5)

Now we show how (iii) of Proposition 2.9 follows from Proposition A.1. First we consider the case where 𝐃(X,T)+\mathbf{D}(X,T)\to+\infty with X0X\neq 0 and ϑ1|τ|1\vartheta_{1}-|\tau^{*}|\gtrsim 1. In such case, a direct computation shows that

HessτΓ((X,T);τ)=|X|2[Υ(|τ|)|τ|𝕀3+(Υ(|τ|)+Υ(|τ|)|τ|)τ|τ|(τ|τ|)T],-\mathrm{Hess}_{\tau^{*}}\,\Gamma((X,T);\tau^{*})=|X|^{2}\left[-\frac{\Upsilon^{\prime}(|\tau^{*}|)}{|\tau^{*}|}{\mathbb{I}}_{3}+\left(-\Upsilon^{\prime\prime}(|\tau^{*}|)+\frac{\Upsilon^{\prime}(|\tau^{*}|)}{|\tau^{*}|}\right)\frac{\tau^{*}}{|\tau^{*}|}\left(\frac{\tau^{*}}{|\tau^{*}|}\right)^{\mathrm{T}}\right], (A.6)

and thus, by Schur’s Lemma,

det(HessτΓ((X,T);τ))=Υ(|τ|)(Υ(|τ|)|τ|)2|X|6.\det(-\mathrm{Hess}_{\tau^{*}}\,\Gamma((X,T);\tau^{*}))=-\Upsilon^{\prime\prime}(|\tau^{*}|)\left(-\frac{\Upsilon^{\prime}(|\tau^{*}|)}{|\tau^{*}|}\right)^{2}\,|X|^{6}.

Combining this, (A.4) with (7.2) and the following observation:

𝒱(iτ)=Υ(|τ|)|τ|sin|τ|,\displaystyle\mathcal{V}(i\tau^{*})=\Upsilon(|\tau^{*}|)\frac{|\tau^{*}|}{\sin{|\tau^{*}|}}, (A.7)

we obtain the asserted asymptotic in Proposition 2.9 without difficulties.

For the case where 𝐃(X,T)+\mathbf{D}(X,T)\to+\infty with X0X\neq 0 and 0<ϑ1|τ|10<\vartheta_{1}-|\tau^{*}|\ll 1, one can appeal to (A.5) and (7.14).

A.3 Proof of (A.4)

It suffices to use the argument introduced in [33], namely, the conjunction of the method of stationary phase and the operator convexity. To this end, using (4.7) and (4.8), the argument in the proof of Lemma 6.2 implies its counterpart:

Lemma A.2.

There exists a constant c>0c>0 such that for any (X,T)2×3(X,T)\in\mathbb{R}^{2}\times\mathbb{R}^{3}, we have

[Γ((X,T);τiλ)Γ((X,T);τ)]c|λ|21+|λ|2|X|2,λ3.\Re[\Gamma((X,T);\tau^{*}-i\lambda)-\Gamma((X,T);\tau^{*})]\geq c\,\frac{|\lambda|^{2}}{1+|\lambda|^{2}}|X|^{2},\quad\forall\,\lambda\in\mathbb{R}^{3}.

For 0<ϑ1|r|ζ010<\vartheta_{1}-|r|\leq\zeta_{0}\leq 1, it follows from (4.9) that

Υ(r)rζ01,0Υ(r)+Υ(r)r=16k=1+ϑk2r2(ϑk2r2)3ζ01.-\frac{\Upsilon^{\prime}(r)}{r}\sim_{\zeta_{0}}1,\qquad 0\leq-\Upsilon^{\prime\prime}(r)+\frac{\Upsilon^{\prime}(r)}{r}=16\sum_{k=1}^{+\infty}\frac{\vartheta_{k}^{2}\,r^{2}}{(\vartheta_{k}^{2}-r^{2})^{3}}\lesssim_{\zeta_{0}}1.

By this and (A.6), we get that

HessτΓ((X,T);τ)ζ0|X|2𝕀3.\displaystyle-\mathrm{Hess}_{\tau^{*}}\,\Gamma((X,T);\tau^{*})\sim_{\zeta_{0}}|X|^{2}\,{\mathbb{I}}_{3}. (A.8)

Then we split 3\mathbb{R}^{3} into disjoint sets {λ;|λ||X|34}\{\lambda;\,|\lambda|\leq|X|^{-\frac{3}{4}}\} and {λ;|λ|>|X|34}\{\lambda;\,|\lambda|>|X|^{-\frac{3}{4}}\}. Using a similar argument as in the proof of Theorem 6.1, one obtains immediately the desired result.

A.4 Proof of (A.5)

Our assumptions are now X0X\neq 0, 𝐃(X,T)+\mathbf{D}(X,T)\to+\infty and |τ|ϑ1|\tau^{*}|\to\vartheta_{1}. In principle we use the same argument as in [35, §5 – §6]. The main difference is that in this case q=2<m=3q=2<m=3. Fortunately, the functions 𝒱\mathcal{V} and Γ~\widetilde{\Gamma} in the definition of 𝐏\mathbf{P} (cf. (4.4)) are explicit enough for us to get a better leading term than the ones in [35]. In fact, the function 𝐏\mathbf{P} is more or less a heat kernel at time 11 on HH-type groups, which has a more concrete leading term. See [31] for more details.

Let us start with the

A.4.1 Preliminaries

Recalling (A.1), now we have 𝐃(X,T)2|X|2\mathbf{D}(X,T)^{2}\gg|X|^{2}. As in Subsection 2.5, we need a new formula of 𝐏\mathbf{P} to proceed our estimates. In fact, set in the sequel for rr\in\mathbb{R} and λ3\lambda\in\mathbb{R}^{3}

𝒱2(λ):=3k=2+(1+λλϑk2)1,Υ~2(r):=3+2k=2+r2ϑk2+r2,\displaystyle\mathcal{V}_{2}(\lambda):=3\,\prod_{k=2}^{+\infty}\left(1+\frac{\lambda\cdot\lambda}{\vartheta_{k}^{2}}\right)^{-1},\quad\widetilde{\Upsilon}_{2}(r):=3+2\sum_{k=2}^{+\infty}\frac{r^{2}}{\vartheta_{k}^{2}+r^{2}}, (A.9)
Γ~2((X,T);λ):=Υ~2(|λ|)|X|24iTλ.\displaystyle\widetilde{\Gamma}_{2}((X,T);\lambda):=\widetilde{\Upsilon}_{2}(|\lambda|)\,|X|^{2}-4i\,T\cdot\lambda. (A.10)

Then from the definition we find that

𝒱(λ)eΓ~((X,T);λ)4\displaystyle\mathcal{V}(\lambda)\,e^{-\frac{\widetilde{\Gamma}((X,T);\lambda)}{4}} =(1+|λ|2ϑ12)1e12ϑ12ϑ12+|λ|2|X|2e|X|22𝒱2(λ)eΓ~2((X,T);λ)4\displaystyle=\left(1+\frac{|\lambda|^{2}}{\vartheta_{1}^{2}}\right)^{-1}e^{\frac{1}{2}\frac{\vartheta_{1}^{2}}{\vartheta_{1}^{2}+|\lambda|^{2}}|X|^{2}}\,e^{-\frac{|X|^{2}}{2}}\,\mathcal{V}_{2}(\lambda)\,e^{-\frac{\widetilde{\Gamma}_{2}((X,T);\lambda)}{4}}
=e|X|22𝒱2(λ)eΓ~2((X,T);λ)412π2e12(1+|λ|2ϑ12)|η|2+ηXdη,\displaystyle=\,e^{-\frac{|X|^{2}}{2}}\,\mathcal{V}_{2}(\lambda)\,e^{-\frac{\widetilde{\Gamma}_{2}((X,T);\lambda)}{4}}\frac{1}{2\pi}\int_{\mathbb{R}^{2}}e^{-\frac{1}{2}\left(1+\frac{|\lambda|^{2}}{\vartheta_{1}^{2}}\right)|\eta|^{2}+\eta\cdot X}d\eta,

where we have used in the last equality (4.3) with q=2q=2, A=(1+|λ|2ϑ12)𝕀2A=\left(1+\frac{|\lambda|^{2}}{\vartheta_{1}^{2}}\right){\mathbb{I}}_{2}, and Y=iXY=-iX, respectively. Hence

𝐏(X,T)=12π2e12|ηX|2dη3𝒱2(λ)e14Γ~2((X,T);λ)|η|22ϑ12|λ|2dλ.\displaystyle\mathbf{P}(X,T)=\frac{1}{2\pi}\int_{\mathbb{R}^{2}}e^{-\frac{1}{2}|\eta-X|^{2}}d\eta\int_{\mathbb{R}^{3}}\,\mathcal{V}_{2}(\lambda)e^{-\frac{1}{4}\widetilde{\Gamma}_{2}((X,T);\lambda)-\frac{|\eta|^{2}}{2\vartheta_{1}^{2}}|\lambda|^{2}}\,d\lambda. (A.11)

As in (4.7), we define

Γ2((X,T);τ):=Γ~2((X,T);iτ).\displaystyle\Gamma_{2}((X,T);\tau):=\widetilde{\Gamma}_{2}((X,T);i\tau). (A.12)

Here we state a counterpart of Lemma A.2, which follows from the same argument.

Lemma A.3.

There exists a constant c>0c>0 such that

[Γ2((X,T);τiλ)Γ2((X,T);τ)]c|λ|21+|λ|2|X|2,λ3.\Re[\Gamma_{2}((X,T);\tau^{*}-i\lambda)-\Gamma_{2}((X,T);\tau^{*})]\geq c\,\frac{|\lambda|^{2}}{1+|\lambda|^{2}}|X|^{2},\quad\forall\,\lambda\in\mathbb{R}^{3}.

It follows from the definition that

𝐃(X,T)2=Γ((X,T);τ)=Γ2((X,T);τ)2|τ|2ϑ12|τ|2|X|2,\displaystyle\mathbf{D}(X,T)^{2}=\Gamma((X,T);\tau^{*})=\Gamma_{2}((X,T);\tau^{*})-\frac{2|\tau^{*}|^{2}}{\vartheta_{1}^{2}-|\tau^{*}|^{2}}|X|^{2},
0=τΓ((X,T);τ)=τΓ2((X,T);τ)4τϑ12(ϑ12|τ|2)2|X|2.\displaystyle 0=\nabla_{\tau^{*}}\,\Gamma((X,T);\tau^{*})=\nabla_{\tau^{*}}\,\Gamma_{2}((X,T);\tau^{*})-\frac{4\tau^{*}\vartheta_{1}^{2}}{(\vartheta_{1}^{2}-|\tau^{*}|^{2})^{2}}|X|^{2}.

Then deforming the contour from 3\mathbb{R}^{3} to 3+iτ\mathbb{R}^{3}+i\tau^{*} in the inner integral of (A.11), and applying the change of variables η=(1|τ|2ϑ12)12ξ\eta=\left(1-\frac{|\tau^{*}|^{2}}{\vartheta_{1}^{2}}\right)^{-\frac{1}{2}}\xi for the outer integral, we yield that

𝐏(X,T)=12π(1|τ|2ϑ12)1e𝐃(X,T)242e12|ξ𝐘|2𝐅2(X,T;ξ)dξ\displaystyle\mathbf{P}(X,T)=\frac{1}{2\pi}\left(1-\frac{|\tau^{*}|^{2}}{\vartheta_{1}^{2}}\right)^{-1}e^{-\frac{\mathbf{D}(X,T)^{2}}{4}}\,\int_{\mathbb{R}^{2}}e^{-\frac{1}{2}|\xi-\mathbf{Y}|^{2}}\mathbf{F}_{2}(X,T;\xi)\,d\xi (A.13)

where

𝐅2(X,T;ξ)\displaystyle\mathbf{F}_{2}(X,T;\xi) :=3𝒱2(λ+iτ)e14𝐒(X,T;λ)+iλ𝐖(X,T;ξ)12|ξ|2|λ|2ϑ12|τ|2dλ,\displaystyle:=\int_{\mathbb{R}^{3}}\,\mathcal{V}_{2}(\lambda+i\tau^{*})e^{-\frac{1}{4}\mathbf{S}(X,T;\lambda)+i\lambda\cdot\mathbf{W}(X,T;\xi)-\frac{1}{2}\frac{|\xi|^{2}|\lambda|^{2}}{\vartheta_{1}^{2}-|\tau^{*}|^{2}}}\,d\lambda, (A.14)

with

𝐒(X,T;λ):=Γ2((X,T);τiλ)Γ2((X,T);τ)+iτΓ2((X,T);τ)λ,\displaystyle\mathbf{S}(X,T;\lambda):=\Gamma_{2}((X,T);\tau^{*}-i\lambda)-\Gamma_{2}((X,T);\tau^{*})+i\nabla_{\tau^{*}}\,\Gamma_{2}((X,T);\tau^{*})\cdot\lambda,
𝐖(X,T;ξ):=|𝐘|2|ξ|2ϑ12|τ|2τ,with 𝐘:=(1|τ|2ϑ12)12X.\displaystyle\mathbf{W}(X,T;\xi):=\frac{|\mathbf{Y}|^{2}-|\xi|^{2}}{\vartheta_{1}^{2}-|\tau^{*}|^{2}}\tau^{*},\quad\mbox{with }\quad\mathbf{Y}:=\left(1-\frac{|\tau^{*}|^{2}}{\vartheta_{1}^{2}}\right)^{-\frac{1}{2}}X. (A.15)

Finally, we split up the integral in (A.13) as

2:=i=15i:=i=15𝐊i,\int_{\mathbb{R}^{2}}:=\sum_{i=1}^{5}\int_{\blacklozenge_{i}}:=\sum_{i=1}^{5}\mathbf{K}_{i},

where

1:=B2(0,12|𝐘|),2:=B2(0,|𝐘|ϵ𝐃(X,T)14)B2(0,12|𝐘|),\displaystyle\blacklozenge_{1}:=B_{\mathbb{R}^{2}}(0,\frac{1}{2}\,|\mathbf{Y}|),\quad\blacklozenge_{2}:=B_{\mathbb{R}^{2}}\left(0,|\mathbf{Y}|-\sqrt{\epsilon_{*}}\,\mathbf{D}(X,T)^{\frac{1}{4}}\right)\setminus B_{\mathbb{R}^{2}}(0,\frac{1}{2}\,|\mathbf{Y}|),
3:=B2(0,|𝐘|+ϵ𝐃(X,T)14)B2(0,|𝐘|ϵ𝐃(X,T)14),\displaystyle\blacklozenge_{3}:=B_{\mathbb{R}^{2}}\left(0,|\mathbf{Y}|+\sqrt{\epsilon_{*}}\,\mathbf{D}(X,T)^{\frac{1}{4}}\right)\setminus B_{\mathbb{R}^{2}}\left(0,|\mathbf{Y}|-\sqrt{\epsilon_{*}}\,\mathbf{D}(X,T)^{\frac{1}{4}}\right),
4:=B2(0,2|𝐘|)B2(0,|𝐘|+ϵ𝐃(X,T)14),5:=B2(0,2|𝐘|)c.\displaystyle\blacklozenge_{4}:=B_{\mathbb{R}^{2}}(0,2\,|\mathbf{Y}|)\setminus B_{\mathbb{R}^{2}}\left(0,|\mathbf{Y}|+\sqrt{\epsilon_{*}}\,\mathbf{D}(X,T)^{\frac{1}{4}}\right),\quad\blacklozenge_{5}:=B_{\mathbb{R}^{2}}(0,2\,|\mathbf{Y}|)^{c}.

Notice that from the definitions of 𝐘\mathbf{Y} and ϵ\epsilon_{*}, (A.1) implies that

|𝐘||X|ϵϵ𝐃(X,T)ϵ𝐃(X,T)14.\displaystyle|\mathbf{Y}|\sim\frac{|X|}{\sqrt{\epsilon_{*}}}\sim\sqrt{\epsilon_{*}}\,\mathbf{D}(X,T)\gg\sqrt{\epsilon_{*}}\,\mathbf{D}(X,T)^{\frac{1}{4}}. (A.16)

A.4.2 Properties of 𝐅2\mathbf{F}_{2}

We will see that the main contribution of the integral in (A.13) comes from the part of the integral taken only over the domain 3\blacklozenge_{3}. To this end, we would like to study the asymptotic behavior of 𝐅2\mathbf{F}_{2} on 3\blacklozenge_{3} and the upper bound of 𝐅2\mathbf{F}_{2} for other regions.

For this purpose, let us first introduce

𝔸(X,T;ξ):=14HessτΓ2((X,T);τ)+|ξ|2ϑ12|τ|2𝕀3,\displaystyle\mathbb{A}(X,T;\xi):=-\frac{1}{4}\mathrm{Hess}_{\tau^{*}}\,\Gamma_{2}((X,T);\tau^{*})+\frac{|\xi|^{2}}{\vartheta_{1}^{2}-|\tau^{*}|^{2}}{\mathbb{I}}_{3}, (A.17)
𝐒(X,T;λ):=𝐒(X,T;λ)+12λTHessτΓ2((X,T);τ)λ.\displaystyle\mathbf{S}_{*}(X,T;\lambda):=\mathbf{S}(X,T;\lambda)+\frac{1}{2}\lambda^{\mathrm{T}}\,\mathrm{Hess}_{\tau^{*}}\,\Gamma_{2}((X,T);\tau^{*})\,\lambda.

Then 𝐒(X,T;)\mathbf{S}_{*}(X,T;\cdot) vanishes to order 2 at the origin, and the phase function of the integral in (A.14) becomes

14𝐒(X,T;λ)12λT𝔸(X,T;ξ)λ+iλ𝐖(X,T;ξ).-\frac{1}{4}\mathbf{S}_{*}(X,T;\lambda)-\frac{1}{2}\lambda^{\mathrm{T}}\,\mathbb{A}(X,T;\xi)\,\lambda+i\lambda\cdot\mathbf{W}(X,T;\xi).

Now we can establish:

Proposition A.4.

We have uniformly for all ξ3\xi\in\blacklozenge_{3} that:

𝐅2(X,T;ξ)=(2π)32𝒱2(iτ)exp{12𝐖(X,T;ξ)T𝔸(X,T;ξ)1𝐖(X,T;ξ)}det𝔸(X,T;ξ)(1+o(1)).\mathbf{F}_{2}(X,T;\xi)=(2\pi)^{\frac{3}{2}}\,\mathcal{V}_{2}(i\tau^{*})\,\frac{\exp\left\{-\frac{1}{2}\mathbf{W}(X,T;\xi)^{\mathrm{T}}\,\mathbb{A}(X,T;\xi)^{-1}\,\mathbf{W}(X,T;\xi)\right\}}{\sqrt{\det{\mathbb{A}(X,T;\xi)}}}\,(1+o(1)).
Proof.

Recall ϵ:=ϑ1|τ|\epsilon_{*}:=\vartheta_{1}-|\tau^{*}|. By (A.16) and the first equality of (A.15), we have for all ξ3\xi\in\blacklozenge_{3} that:

|ξ|2ϑ12|τ|2|𝐘|2ϵ𝐃(X,T)2,\displaystyle\frac{|\xi|^{2}}{\vartheta_{1}^{2}-|\tau^{*}|^{2}}\sim\frac{|\mathbf{Y}|^{2}}{\epsilon_{*}}\sim\mathbf{D}(X,T)^{2}, (A.18)
|𝐖(X,T;ξ)|(|ξ|+|𝐘|)||ξ||𝐘||ϵ𝐃(X,T)54.\displaystyle|\mathbf{W}(X,T;\xi)|\lesssim\frac{(|\xi|+|\mathbf{Y}|)\,|\,|\xi|-|\mathbf{Y}|\,|}{\epsilon_{*}}\lesssim\mathbf{D}(X,T)^{\frac{5}{4}}. (A.19)

Then from the definition of Γ2((X,T);)\Gamma_{2}((X,T);\cdot) (cf. (A.12)), a simple computation (compared with (A.8)) shows that

HessτΓ2((X,T);τ)|X|2𝕀3𝐃(X,T)2𝕀3|ξ|2ϑ12|τ|2𝕀3,ξ3,\displaystyle-\mathrm{Hess}_{\tau^{*}}\,\Gamma_{2}((X,T);\tau^{*})\sim|X|^{2}\,{\mathbb{I}}_{3}\ll\mathbf{D}(X,T)^{2}\,{\mathbb{I}}_{3}\sim\frac{|\xi|^{2}}{\vartheta_{1}^{2}-|\tau^{*}|^{2}}\,{\mathbb{I}}_{3},\quad\forall\,\xi\in\blacklozenge_{3},

which implies that, for all ξ3\xi\in\blacklozenge_{3},

𝔸(X,T;ξ)=|ξ|2ϑ12|τ|2𝕀3(1+o(1))=|𝐘|2ϑ12|τ|2𝕀3(1+o(1))𝐃(X,T)2𝕀3.\displaystyle\mathbb{A}(X,T;\xi)=\frac{|\xi|^{2}}{\vartheta_{1}^{2}-|\tau^{*}|^{2}}\,{\mathbb{I}}_{3}\,(1+o(1))=\frac{|\mathbf{Y}|^{2}}{\vartheta_{1}^{2}-|\tau^{*}|^{2}}\,{\mathbb{I}}_{3}\,(1+o(1))\sim\mathbf{D}(X,T)^{2}\,{\mathbb{I}}_{3}. (A.20)

Now setting

ρ(X,T;ξ):=𝔸(X,T;ξ)1𝐖(X,T;ξ),\rho_{*}(X,T;\xi):=\mathbb{A}(X,T;\xi)^{-1}\,\mathbf{W}(X,T;\xi),

we have |ρ(X,T;ξ)|=O(𝐃(X,T)34)|\rho_{*}(X,T;\xi)|=O(\mathbf{D}(X,T)^{-\frac{3}{4}}) by (A.20) and (A.19). Then lifting the contour again by iρ(X,T;ξ)i\rho_{*}(X,T;\xi) in (A.14), and splitting the integral into the ones over {λ;|λ|𝐃(X,T)34}\{\lambda;\,|\lambda|\leq\mathbf{D}(X,T)^{-\frac{3}{4}}\} and {λ;|λ|>𝐃(X,T)34}\{\lambda;\,|\lambda|>\mathbf{D}(X,T)^{-\frac{3}{4}}\}. It follows from the estimate of ρ(X,T;ξ)\rho_{*}(X,T;\xi) above that

𝐒(X,T;λ+iρ(X,T;ξ))=O(𝐃(X,T)14)\mathbf{S}_{*}(X,T;\lambda+i\rho_{*}(X,T;\xi))=O(\mathbf{D}(X,T)^{-\frac{1}{4}})

in the first region and thus using a similar argument in the proof of Theorem 6.1, we obtain the desired proposition. ∎

Next we give an upper bound of 𝐅2\mathbf{F}_{2} for all ξ2\xi\in\mathbb{R}^{2}.

Proposition A.5.

There exists a constant c>0c>0 such that

|𝐅2(X,T;ξ)|exp{c|𝐖(X,T;ξ)||X|2+|ξ|2/ϵ+1}.|\mathbf{F}_{2}(X,T;\xi)|\lesssim\exp\left\{-c\,\frac{|\mathbf{W}(X,T;\xi)|}{\sqrt{|X|^{2}+|\xi|^{2}/\epsilon_{*}+1}}\right\}.
Proof.

We deform the contour in (A.14) from 3\mathbb{R}^{3} to 3+iρ(X,T;ξ)\mathbb{R}^{3}+i\rho^{*}(X,T;\xi), where we choose

ρ(X,T;ξ):=c|X|2+|ξ|2/ϵ+1𝐖(X,T;ξ)|𝐖(X,T;ξ)|\rho^{*}(X,T;\xi):=\frac{c}{\sqrt{|X|^{2}+|\xi|^{2}/\epsilon_{*}+1}}\frac{\mathbf{W}(X,T;\xi)}{|\mathbf{W}(X,T;\xi)|}

for some small c>0c>0. From our choice and the analyticity of Υ~2\widetilde{\Upsilon}_{2} on {z;ϑ2<(z)<ϑ2}\{z\in\mathbb{C};\ -\vartheta_{2}<\Im(z)<\vartheta_{2}\}, we have

𝐒~(X,T;ξ):=14𝐒(X,T;iρ(X,T;ξ))+12ρ(X,T;ξ)T𝔸(X,T;ξ)ρ(X,T;ξ)=O(1)\widetilde{\mathbf{S}}(X,T;\xi):=-\frac{1}{4}\mathbf{S}_{*}(X,T;i\rho^{*}(X,T;\xi))+\frac{1}{2}\rho^{*}(X,T;\xi)^{\mathrm{T}}\mathbb{A}(X,T;\xi)\rho^{*}(X,T;\xi)=O(1)

since 𝔸(X,T;ξ)(|X|2+|ξ|2/ϵ)𝕀3\mathbb{A}(X,T;\xi)\sim\left(|X|^{2}+|\xi|^{2}/\epsilon_{*}\right){\mathbb{I}}_{3}. Finally it follows from Lemma A.3 and the inequality |𝒱2(iτ+ξ)|𝒱2(iτ)𝒱2(ξ)|\mathcal{V}_{2}(i\tau+\xi)|\leq\mathcal{V}_{2}(i\tau)\mathcal{V}_{2}(\xi) for any |τ|<ϑ2|\tau|<\vartheta_{2} and any ξ3\xi\in\mathbb{R}^{3} (compared with (6.3)) that

|𝐅2(X,T;ξ)|\displaystyle|\mathbf{F}_{2}(X,T;\xi)| exp{𝐒~(X,T;ξ)𝐖(X,T;ξ)ρ(X,T;ξ)}\displaystyle\lesssim\exp\left\{\widetilde{\mathbf{S}}(X,T;\xi)-\mathbf{W}(X,T;\xi)\cdot\rho^{*}(X,T;\xi)\right\}
exp{𝐖(X,T;ξ)ρ(X,T;ξ)}=exp{c|𝐖(X,T;ξ)||X|2+|ξ|2/ϵ+1}.\displaystyle\lesssim\,\exp\left\{-\mathbf{W}(X,T;\xi)\cdot\rho^{*}(X,T;\xi)\right\}=\exp\left\{-c\,\frac{|\mathbf{W}(X,T;\xi)|}{\sqrt{|X|^{2}+|\xi|^{2}/\epsilon_{*}+1}}\right\}.

This proves the proposition as required. ∎

A.4.3 Estimation of 𝐊3\mathbf{K}_{3}

To treat 𝐊3\mathbf{K}_{3}, we need the following lemma, which is an analogue of [32, Proposition 4.2].

Lemma A.6.

Let Y2{0}Y\in\mathbb{R}^{2}\setminus\{0\}, r>0r>0, δ(0,|Y|)\delta\in(0,\ |Y|) and set

~δ(Y,r):={w2;|Y|δ|w|<|Y|+δ}e12|wY|2e(|w|2|Y|2)2rdw.\displaystyle\widetilde{\mathcal{I}}_{\delta}(Y,r):=\int_{\{w\in\mathbb{R}^{2};\,|Y|-\delta\leq|w|<|Y|+\delta\}}e^{-\frac{1}{2}|w-Y|^{2}}e^{-\frac{(|w|^{2}-|Y|^{2})^{2}}{r}}\,dw.

Then we have uniformly

~δ(Y,r)=π32re|Y|2I0(|Y|2)(1+o(1)),\displaystyle\widetilde{\mathcal{I}}_{\delta}(Y,r)=\pi^{\frac{3}{2}}\,\sqrt{r}\,e^{-|Y|^{2}}I_{0}(|Y|^{2})\,(1+o(1)),

provided

δ|Y|12,δ2|Y|2r+,and|Y|2r+.\displaystyle\frac{\delta}{|Y|}\leq\frac{1}{2},\quad\frac{\delta^{2}|Y|^{2}}{r}\to+\infty,\quad\mbox{and}\quad\frac{|Y|^{2}}{r}\to+\infty. (A.21)
Proof.

First notice that the change of variables wOww\mapsto Ow gives ~δ(Y,r)=~δ(OY,r)\widetilde{\mathcal{I}}_{\delta}(Y,r)=\widetilde{\mathcal{I}}_{\delta}(OY,r) for any OO2O\in{\rm O}_{2}. Then without loss of generality we can assume Y=|Y|(1,0)Y=|Y|(1,0). Using polar coordinates w=ρ(cosγ,sinγ)w=\rho\,(\cos{\gamma},\sin{\gamma}) and the integral representation for I0I_{0} (cf. (2.32)), we get

~δ(Y,r)\displaystyle\widetilde{\mathcal{I}}_{\delta}(Y,r) =e12|Y|2{w2;|Y|δ|w|<|Y|+δ}e12|w|2+|Y|w1e(|w|2|Y|2)2rdw\displaystyle=e^{-\frac{1}{2}|Y|^{2}}\int_{\{w\in\mathbb{R}^{2};\,|Y|-\delta\leq|w|<|Y|+\delta\}}e^{-\frac{1}{2}|w|^{2}+|Y|w_{1}}e^{-\frac{(|w|^{2}-|Y|^{2})^{2}}{r}}\,dw
=e12|Y|2|Y|δ|Y|+δρe12ρ2(ρ2|Y|2)2rdρππeρ|Y|cosγdγ\displaystyle=e^{-\frac{1}{2}|Y|^{2}}\int_{|Y|-\delta}^{|Y|+\delta}\rho\,e^{-\frac{1}{2}\rho^{2}-\frac{(\rho^{2}-|Y|^{2})^{2}}{r}}\,d\rho\int_{-\pi}^{\pi}e^{\rho\,|Y|\cos{\gamma}}\,d\gamma
=2πe12|Y|2|Y|δ|Y|+δρI0(|Y|ρ)e12ρ2(ρ2|Y|2)2rdρ\displaystyle=2\pi\,e^{-\frac{1}{2}|Y|^{2}}\int_{|Y|-\delta}^{|Y|+\delta}\rho I_{0}(|Y|\rho)\,e^{-\frac{1}{2}\rho^{2}-\frac{(\rho^{2}-|Y|^{2})^{2}}{r}}\,d\rho
=2πe12|Y|2|Y|21δ|Y|1+δ|Y|uI0(|Y|2u)e|Y|22u2e|Y|4r(u+1)2(u1)2du,\displaystyle=2\pi\,e^{-\frac{1}{2}|Y|^{2}}|Y|^{2}\int_{1-\frac{\delta}{|Y|}}^{1+\frac{\delta}{|Y|}}uI_{0}(|Y|^{2}u)e^{-\frac{|Y|^{2}}{2}u^{2}}e^{-\frac{|Y|^{4}}{r}(u+1)^{2}(u-1)^{2}}\,du,

where in the last “==” we have used the change of variables ρ=|Y|u\rho=|Y|u additionally. Then we split the proof into two cases:

Case 1: |Y|1|Y|\lesssim 1.

Noting |Y|4/r4δ2|Y|2/r+|Y|^{4}/r\geq 4\delta^{2}|Y|^{2}/r\to+\infty, the standard Laplace’s method gives

~δ(Y,r)=π32re|Y|2I0(|Y|2)(1+o(1)),\widetilde{\mathcal{I}}_{\delta}(Y,r)=\pi^{\frac{3}{2}}\sqrt{r}\,e^{-|Y|^{2}}I_{0}(|Y|^{2})\,(1+o(1)),

which ends the proof.

Case 2: |Y|+|Y|\to+\infty.

In such case, from (7.2) we get

~δ(Y,r)=2π|Y|1δ|Y|1+δ|Y|ue|Y|22(u1)2[1+2|Y|2r(u+1)2]du(1+o(1)).\widetilde{\mathcal{I}}_{\delta}(Y,r)=\sqrt{2\pi}\,|Y|\int_{1-\frac{\delta}{|Y|}}^{1+\frac{\delta}{|Y|}}\sqrt{u}\,e^{-\frac{|Y|^{2}}{2}(u-1)^{2}\left[1+\frac{2|Y|^{2}}{r}(u+1)^{2}\right]}du\,(1+o(1)).

Then the usage of the Laplace’s method yields

~δ(Y,r)=π2r|Y|(1+o(1))=π32re|Y|2I0(|Y|2)(1+o(1)),\widetilde{\mathcal{I}}_{\delta}(Y,r)=\frac{\pi}{\sqrt{2}}\,\frac{\sqrt{r}}{|Y|}\,(1+o(1))=\pi^{\frac{3}{2}}\sqrt{r}e^{-|Y|^{2}}I_{0}(|Y|^{2})\,(1+o(1)),

where in the last “==” we have used (7.2) again. ∎

Recall the estimate (A.20). The result of Proposition A.4 can be further simplified. To be more precise, it follows from (A.20) and (A.15) that:

det𝔸(X,T;ξ)=|X|38ϵ3(1+o(1)),\displaystyle\sqrt{\det\mathbb{A}(X,T;\xi)}=\frac{|X|^{3}}{8\,\epsilon_{*}^{3}}(1+o(1)),
𝐖(X,T;ξ)T𝔸(X,T;ξ)1𝐖(X,T;ξ)=(|𝐘|2|ξ|2)2|X|2(1+o(1)).\displaystyle\mathbf{W}(X,T;\xi)^{\mathrm{T}}\mathbb{A}(X,T;\xi)^{-1}\mathbf{W}(X,T;\xi)=\frac{(|\mathbf{Y}|^{2}-|\xi|^{2})^{2}}{|X|^{2}}(1+o(1)).

Consequently, if we set

𝔯(X,T):=infξ3[𝐖(X,T;ξ)T𝔸(X,T;ξ)1𝐖(X,T;ξ)(|𝐘|2|ξ|2)2|X|2(|𝐘|2|ξ|2)2|X|2],\displaystyle\mathfrak{r}_{*}(X,T):=\inf_{\xi\in\blacklozenge_{3}}\left[\frac{\mathbf{W}(X,T;\xi)^{\mathrm{T}}\mathbb{A}(X,T;\xi)^{-1}\mathbf{W}(X,T;\xi)-\frac{(|\mathbf{Y}|^{2}-|\xi|^{2})^{2}}{|X|^{2}}}{\frac{(|\mathbf{Y}|^{2}-|\xi|^{2})^{2}}{|X|^{2}}}\right], (A.22)
𝔯(X,T):=supξ3[𝐖(X,T;ξ)T𝔸(X,T;ξ)1𝐖(X,T;ξ)(|𝐘|2|ξ|2)2|X|2(|𝐘|2|ξ|2)2|X|2],\displaystyle\mathfrak{r}^{*}(X,T):=\sup_{\xi\in\blacklozenge_{3}}\left[\frac{\mathbf{W}(X,T;\xi)^{\mathrm{T}}\mathbb{A}(X,T;\xi)^{-1}\mathbf{W}(X,T;\xi)-\frac{(|\mathbf{Y}|^{2}-|\xi|^{2})^{2}}{|X|^{2}}}{\frac{(|\mathbf{Y}|^{2}-|\xi|^{2})^{2}}{|X|^{2}}}\right], (A.23)

then 𝔯(X,T)=o(1)\mathfrak{r}_{*}(X,T)=o(1) and 𝔯(X,T)=o(1)\mathfrak{r}^{*}(X,T)=o(1). Hence it follows from Proposition A.4 that

𝐅2(X,T;ξ)(2π)328ϵ3|X|3𝒱2(iτ)e(|ξ|2|𝐘|2)22|X|2(1+𝔯(X,T))(1+o(1)),ξ3.\mathbf{F}_{2}(X,T;\xi)\leq(2\pi)^{\frac{3}{2}}\frac{8\,\epsilon_{*}^{3}}{|X|^{3}}\,\mathcal{V}_{2}(i\tau^{*})\,e^{-\frac{(|\xi|^{2}-|\mathbf{Y}|^{2})^{2}}{2|X|^{2}}\,(1+\mathfrak{r}_{*}(X,T))}\,(1+o(1)),\quad\forall\,\xi\in\blacklozenge_{3}.

which implies that

𝐊3(2π)328ϵ3|X|3𝒱2(iτ)𝐐(X,T)(1+o(1)),\displaystyle\mathbf{K}_{3}\leq(2\pi)^{\frac{3}{2}}\,\frac{8\,\epsilon_{*}^{3}}{|X|^{3}}\,\mathcal{V}_{2}(i\tau^{*})\,\mathbf{Q}_{*}(X,T)\,(1+o(1)),

where

𝐐(X,T):=3e12|ξ𝐘|2e(|ξ|2|𝐘|2)22|X|2(1+𝔯(X,T))dξ.\displaystyle\mathbf{Q}_{*}(X,T):=\int_{\blacklozenge_{3}}e^{-\frac{1}{2}|\xi-\mathbf{Y}|^{2}}e^{-\frac{(|\xi|^{2}-|\mathbf{Y}|^{2})^{2}}{2|X|^{2}}(1+\mathfrak{r}_{*}(X,T))}d\xi.

Now taking Y=𝐘Y=\mathbf{Y}, r=2|X|2/(1+𝔯(X,T))=2|X|2(1+o(1))r=2\,|X|^{2}/(1+\mathfrak{r}_{*}(X,T))=2\,|X|^{2}\,(1+o(1)), and δ=ϵ𝐃(X,T)14\delta=\sqrt{\epsilon_{*}}\,\mathbf{D}(X,T)^{\frac{1}{4}}, in the above lemma, by (A.16) we can check that the condition (A.21) is fulfilled, and hence

𝐊332π3ϵ3|X|2𝒱2(iτ)e|𝐘|2I0(|𝐘|2)(1+o(1)).\displaystyle\mathbf{K}_{3}\leq 32\,\pi^{3}\,\frac{\epsilon_{*}^{3}}{|X|^{2}}\,\mathcal{V}_{2}(i\tau^{*})\,e^{-|\mathbf{Y}|^{2}}I_{0}(|\mathbf{Y}|^{2})\,(1+o(1)).

On the other hand, using the same argument with 𝔯\mathfrak{r}_{*} replaced by 𝔯\mathfrak{r}^{*}, we can obtain the same lower bound for 𝐊3\mathbf{K}_{3}. So it is actually an equality. In conclusion, under the assumption in Proposition A.1 (II), it holds that

𝐊3\displaystyle\mathbf{K}_{3} =32π3ϵ3|X|2𝒱2(iτ)e|𝐘|2I0(|𝐘|2)(1+o(1))\displaystyle=32\pi^{3}\,\frac{\epsilon_{*}^{3}}{|X|^{2}}\,\mathcal{V}_{2}(i\tau^{*})\,e^{-|\mathbf{Y}|^{2}}I_{0}(|\mathbf{Y}|^{2})\,(1+o(1))
ϵ𝐃(X,T)211+ϵ𝐃(X,T)2ϵ𝐃(X,T)3,\displaystyle\sim\epsilon_{*}\mathbf{D}(X,T)^{-2}\frac{1}{\sqrt{1+\epsilon_{*}\mathbf{D}(X,T)^{2}}}\gtrsim\epsilon_{*}\,\mathbf{D}(X,T)^{-3}, (A.24)

where we have used in “\sim” (7.3), (A.1) and (A.16).

A.4.4 Bounds for the remaining terms

We begin with the estimate of 𝐊2\mathbf{K}_{2}. In fact, on 2\blacklozenge_{2}, we have that |ξ||𝐘||\xi|\sim|\mathbf{Y}|, |𝐖(X,T;ξ)|𝐃(X,T)54|\mathbf{W}(X,T;\xi)|\gtrsim\mathbf{D}(X,T)^{\frac{5}{4}} and |X|2+|ξ|2ϵ+1𝐃(X,T)2|X|^{2}+\frac{|\xi|^{2}}{\epsilon_{*}}+1\sim\mathbf{D}(X,T)^{2} by the first equality in (A.15) and (A.16). Then it follows from Proposition A.5 that

|𝐊2|2ec𝐃(X,T)14dξϵ𝐃(X,T)2ec𝐃(X,T)14=o(𝐊3),|\mathbf{K}_{2}|\lesssim\int_{\blacklozenge_{2}}e^{-c\,\mathbf{D}(X,T)^{\frac{1}{4}}}d\xi\sim\epsilon_{*}\,\mathbf{D}(X,T)^{2}\,e^{-c\,\mathbf{D}(X,T)^{\frac{1}{4}}}=o(\mathbf{K}_{3}),

where we have used in “\sim” (A.16) again.

Similarly, we can prove |𝐊4|=o(𝐊3)|\mathbf{K}_{4}|=o(\mathbf{K}_{3}) as well.

Consider now 𝐊1\mathbf{K}_{1}. On 1\blacklozenge_{1}, notice that ||ξ||𝐘|||𝐘||\,|\xi|-|\mathbf{Y}|\,|\sim|\mathbf{Y}|, |𝐖(X,T;ξ)|𝐃(X,T)2|\mathbf{W}(X,T;\xi)|\sim\mathbf{D}(X,T)^{2}, and |X|2+|ξ|2ϵ+1𝐃(X,T)2|X|^{2}+\frac{|\xi|^{2}}{\epsilon_{*}}+1\lesssim\mathbf{D}(X,T)^{2}. As a result, Proposition A.5 yields

|𝐊1|1ec𝐃(X,T)dξ=o(𝐊3).|\mathbf{K}_{1}|\lesssim\int_{\blacklozenge_{1}}e^{-c\,\mathbf{D}(X,T)}d\xi=o(\mathbf{K}_{3}).

We are now left with the estimation of 𝐊5\mathbf{K}_{5}. Remark that we have on 5\blacklozenge_{5} that |𝐖(X,T,ξ)||ξ|2ϵ𝐃(X,T)2|\mathbf{W}(X,T,\xi)|\sim\frac{|\xi|^{2}}{\epsilon_{*}}\gtrsim\mathbf{D}(X,T)^{2} and |X|2+|ξ|2ϵ+1|ξ|2ϵ|X|^{2}+\frac{|\xi|^{2}}{\epsilon_{*}}+1\sim\frac{|\xi|^{2}}{\epsilon_{*}}. Then Proposition A.5 gives

|𝐊5|5ec|ξ|ϵdξϵ{η;|η|c𝐃(X,T)}ec|η|dηϵec2𝐃(X,T)=o(𝐊3),|\mathbf{K}_{5}|\lesssim\int_{\blacklozenge_{5}}e^{-c\,\frac{|\xi|}{\sqrt{\epsilon_{*}}}}d\xi\lesssim\epsilon_{*}\int_{\{\eta;\,|\eta|\geq c\mathbf{D}(X,T)\}}e^{-c\,|\eta|}d\eta\lesssim\epsilon_{*}e^{-\frac{c}{2}\,\mathbf{D}(X,T)}=o(\mathbf{K}_{3}),

where we have used in the second “\lesssim” the change of variables ξ=ϵη\xi=\sqrt{\epsilon_{*}}\,\eta.

A.4.5 Summary

From all the estimates obtained above, we conclude that

𝐏(X,T)\displaystyle\mathbf{P}(X,T) =12π(1|τ|2ϑ12)1e𝐃(X,T)24i=15𝐊i\displaystyle=\frac{1}{2\pi}\left(1-\frac{|\tau^{*}|^{2}}{\vartheta_{1}^{2}}\right)^{-1}e^{-\frac{\mathbf{D}(X,T)^{2}}{4}}\,\sum_{i=1}^{5}\mathbf{K}_{i}
=12π(1|τ|2ϑ12)1e𝐃(X,T)24𝐊3(1+o(1))\displaystyle=\frac{1}{2\pi}\left(1-\frac{|\tau^{*}|^{2}}{\vartheta_{1}^{2}}\right)^{-1}e^{-\frac{\mathbf{D}(X,T)^{2}}{4}}\,\mathbf{K}_{3}\,(1+o(1))
=16π2ϵ3|X|2𝒱(iτ)e𝐃(X,T)24e|𝐘|2I0(|𝐘|2)(1+o(1)).\displaystyle=16\pi^{2}\,\frac{\epsilon_{*}^{3}}{|X|^{2}}\,\mathcal{V}(i\tau^{*})\,e^{-\frac{\mathbf{D}(X,T)^{2}}{4}}\,e^{-|\mathbf{Y}|^{2}}I_{0}(|\mathbf{Y}|^{2})\,(1+o(1)).

Then to show (A.5), it suffices to use (A.7), (7.11) and a similar argument as in the proof of (8.34).

Acknowledgement

This work is partially supported by NSF of China (Grants No. 12271102 and No. 11625102). The first author would like to thank D. Bakry for bringing the free step-two Carnot group with 33 generators to our attention in 2008.

References

  • [1] J.-P. Anker and L. Ji. Heat kernel and Green function estimates on noncompact symmetric spaces. Geom. Funct. Anal., 9(6):1035–1091, 1999.
  • [2] T. Ariyoshi and M. Hino. Small-time asymptotic estimates in local Dirichlet spaces. Electron. J. Probab., 10:no. 37, 1236–1259, 2005.
  • [3] P. Auscher, T. Coulhon, X. T. Duong, and S. Hofmann. Riesz transform on manifolds and heat kernel regularity. Ann. Sci. École Norm. Sup. (4), 37(6):911–957, 2004.
  • [4] D. Bakry, F. Baudoin, M. Bonnefont, and D. Chafaï. On gradient bounds for the heat kernel on the Heisenberg group. J. Funct. Anal., 255(8):1905–1938, 2008.
  • [5] D. Bakry, I. Gentil, and M. Ledoux. Analysis and geometry of Markov diffusion operators, volume 348 of Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer, Cham, 2014.
  • [6] F. Baudoin and M. Bonnefont. The subelliptic heat kernel on SU(2){\rm SU}(2): representations, asymptotics and gradient bounds. Math. Z., 263(3):647–672, 2009.
  • [7] R. Beals, B. Gaveau, and P. Greiner. The Green function of model step two hypoelliptic operators and the analysis of certain tangential Cauchy Riemann complexes. Adv. Math., 121(2):288–345, 1996.
  • [8] R. Beals, B. Gaveau, and P. C. Greiner. Hamilton-Jacobi theory and the heat kernel on Heisenberg groups. J. Math. Pures Appl. (9), 79(7):633–689, 2000.
  • [9] M. Berger. A panoramic view of Riemannian geometry. Springer-Verlag, Berlin, 2003.
  • [10] N. Berline, E. Getzler, and M. Vergne. Heat kernels and Dirac operators. Grundlehren Text Editions. Springer-Verlag, Berlin, 2004. Corrected reprint of the 1992 original.
  • [11] R. Bhatia. Matrix analysis, volume 169 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1997.
  • [12] A. Bonfiglioli, E. Lanconelli, and F. Uguzzoni. Stratified Lie groups and potential theory for their sub-Laplacians. Springer Monographs in Mathematics. Springer, Berlin, 2007.
  • [13] M. Bonnefont, D. Chafaï, and R. Herry. On logarithmic Sobolev inequalities for the heat kernel on the Heisenberg group. Annales de la Faculté des sciences de Toulouse : Mathématiques, 29(2):335–355, 2020. identifier: 10.5802/afst.1633.
  • [14] S. Boutayeb, T. Coulhon, and A. Sikora. A new approach to pointwise heat kernel upper bounds on doubling metric measure spaces. Adv. Math., 270:302–374, 2015.
  • [15] R. W. Brockett. Nonlinear control theory and differential geometry. In Proceedings of the International Congress of Mathematicians, Vol. 1, 2 (Warsaw, 1983), pages 1357–1368. PWN, Warsaw, 1984.
  • [16] T. Coulhon. Itération de Moser et estimation gaussienne du noyau de la chaleur. J. Operator Theory, 29(1):157–165, 1993.
  • [17] T. Coulhon and X. T. Duong. Riesz transforms for 1p21\leq p\leq 2. Trans. Amer. Math. Soc., 351(3):1151–1169, 1999.
  • [18] T. Coulhon and A. Sikora. Gaussian heat kernel upper bounds via the Phragmén-Lindelöf theorem. Proc. Lond. Math. Soc. (3), 96(2):507–544, 2008.
  • [19] J. Cygan. Heat kernels for class 22 nilpotent groups. Studia Math., 64(3):227–238, 1979.
  • [20] E. B. Davies and M. M. H. Pang. Sharp heat kernel bounds for some Laplace operators. Quart. J. Math. Oxford Ser. (2), 40(159):281–290, 1989.
  • [21] B. Gaveau. Principe de moindre action, propagation de la chaleur et estimées sous elliptiques sur certains groupes nilpotents. Acta Math., 139(1-2):95–153, 1977.
  • [22] I. S. Gradshteyn and I. M. Ryzhik. Table of integrals, series, and products. Elsevier/Academic Press, Amsterdam, eighth edition, 2015. Translated from the Russian, Translation edited and with a preface by Daniel Zwillinger and Victor Moll, Revised from the seventh edition [MR2360010].
  • [23] A. Grigor’yan. Heat kernel and analysis on manifolds, volume 47 of AMS/IP Studies in Advanced Mathematics. American Mathematical Society, Providence, RI; International Press, Boston, MA, 2009.
  • [24] W. Hebisch and B. Zegarliński. Coercive inequalities on metric measure spaces. J. Funct. Anal., 258(3):814–851, 2010.
  • [25] L. Hörmander. The analysis of linear partial differential operators. I. Springer Study Edition. Springer-Verlag, Berlin, second edition, 1990. Distribution theory and Fourier analysis.
  • [26] H. Hueber and D. Müller. Asymptotics for some Green kernels on the Heisenberg group and the Martin boundary. Math. Ann., 283(1):97–119, 1989.
  • [27] S. G. Krantz and H. R. Parks. The implicit function theorem. Modern Birkhäuser Classics. Birkhäuser/Springer, New York, 2013. History, theory, and applications, Reprint of the 2003 edition.
  • [28] R. Léandre. Majoration en temps petit de la densité d’une diffusion dégénérée. Probab. Theory Related Fields, 74(2):289–294, 1987.
  • [29] H.-Q. Li. Estimation optimale du gradient du semi-groupe de la chaleur sur le groupe de Heisenberg. J. Funct. Anal., 236(2):369–394, 2006.
  • [30] H.-Q. Li. Estimations asymptotiques du noyau de la chaleur sur les groupes de Heisenberg. C. R. Math. Acad. Sci. Paris, 344(8):497–502, 2007.
  • [31] H.-Q. Li. Estimations optimales du noyau de la chaleur sur les groupes de type Heisenberg. J. Reine Angew. Math., 646:195–233, 2010.
  • [32] H.-Q. Li. Estimations asymptotiques du noyau de la chaleur pour l’opérateur de Grushin. Comm. Partial Differential Equations, 37(5):794–832, 2012.
  • [33] H.-Q. Li. The Carnot-Carathéodory distance on 22-step groups. arXiv e-prints, page arXiv:2112.07822, Dec. 2021.
  • [34] H.-Q. Li and P. Sjögren. Estimates for operators related to the sub-Laplacian with drift in Heisenberg groups. J. Fourier Anal. Appl., 28(1):Paper No. 10, 29, 2021.
  • [35] H.-Q. Li and Y. Zhang. Heat kernel asymptotics for a class of Métivier groups. Preprint 2022.
  • [36] H.-Q. Li and Y. Zhang. Revisiting the heat kernel on isotropic and nonisotropic Heisenberg groups. Comm. Partial Differential Equations, 44(6):467–503, 2019.
  • [37] H.-Q. Li and Y. Zhang. A complete answer to the Gaveau–Brockett problem. arXiv e-prints, page arXiv:2112.07927, Dec. 2021.
  • [38] H.-Q. Li and Y. Zhang. Sub-Riemannian geometry on some step-two Carnot groups. arXiv e-prints, page arXiv:2102.09860, Feb. 2021.
  • [39] P. Li and S.-T. Yau. On the parabolic kernel of the Schrödinger operator. Acta Math., 156(3-4):153–201, 1986.
  • [40] F. Lust-Piquard. A simple-minded computation of heat kernels on Heisenberg groups. Colloq. Math., 97(2):233–249, 2003.
  • [41] A. Montanari and D. Morbidelli. On the subRiemannian cut locus in a model of free two-step Carnot group. Calc. Var. Partial Differential Equations, 56(2):Paper No. 36, 26, 2017.
  • [42] C. P. Niculescu and L.-E. Persson. Convex functions and their applications. CMS Books in Mathematics/Ouvrages de Mathématiques de la SMC. Springer, Cham, 2018. A contemporary approach, Second edition of [ MR2178902].
  • [43] B. Qian. Positive curvature property for sub-Laplacian on nilpotent Lie group of rank two. Potential Anal., 39(4):325–340, 2013.
  • [44] S. Rigot. Mass transportation in groups of type HH. Commun. Contemp. Math., 7(4):509–537, 2005.
  • [45] L. Saloff-Coste. Analyse sur les groupes de Lie à croissance polynômiale. Ark. Mat., 28(2):315–331, 1990.
  • [46] L. Saloff-Coste. A note on Poincaré, Sobolev, and Harnack inequalities. Internat. Math. Res. Notices, (2):27–38, 1992.
  • [47] L. Saloff-Coste. Aspects of Sobolev-type inequalities, volume 289 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 2002.
  • [48] C. Séguin and A. Mansouri. Short-time asymptotics of heat kernels of hypoelliptic Laplacians on unimodular Lie groups. J. Funct. Anal., 262(9):3891–3928, 2012.
  • [49] A. Sikora. Sharp pointwise estimates on heat kernels. Quart. J. Math. Oxford Ser. (2), 47(187):371–382, 1996.
  • [50] A. Sikora. Riesz transform, Gaussian bounds and the method of wave equation. Math. Z., 247(3):643–662, 2004.
  • [51] K.-H. Tan and X.-P. Yang. Characterisation of the sub-Riemannian isometry groups of HH-type groups. Bull. Austral. Math. Soc., 70(1):87–100, 2004.
  • [52] N. T. Varopoulos. Small time Gaussian estimates of heat diffusion kernels. II. The theory of large deviations. J. Funct. Anal., 93(1):1–33, 1990.
  • [53] N. T. Varopoulos, L. Saloff-Coste, and T. Coulhon. Analysis and geometry on groups, volume 100 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge, 1992.


Hong-Quan Li, Sheng-Chen Mao
School of Mathematical Sciences/Shanghai Center for Mathematical Sciences
Fudan University
220 Handan Road
Shanghai 200433
People’s Republic of China
E-Mail: [email protected]
[email protected]  or  [email protected]


Ye Zhang
Analysis on Metric Spaces Unit
Okinawa Institute of Science and Technology Graduate University
1919-1 Tancha, Onna-son, Kunigami-gun
Okinawa, 904-0495, Japan
E-Mail: [email protected]  or  [email protected]