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

An infinite dimensional balanced embedding problem II: uniqueness

Jingzhou Sun Department of Mathematics, Shantou University, Shantou City, Guangdong Province 515063, China jzsun@stu.edu.cn
Abstract.

This is the sequel to our first paper concerning the balanced embedding of a non-compact complex manifold into an infinite-dimensional projective space. We prove the uniqueness of such an embedding. The proof relies on fine estimates of the asymptotics of a balanced embedding.

The author is partially supported by NNSF of China no.11701353.

1. Introduction

In this article, we continue our study of a model problem concerning the balanced embedding of a non-compact complex manifold into a projective Hilbert space started in [8].

For the general set-up and motivations of this problem, the readers are referred to [8] and the references there. The readers are also referred to [6], [7]and [3] etc. for the applications of balanced metric in algebraic geometry. For the statement of our results, we repeat the necessary notations.

Let \mathcal{H} be a separable complex Hilbert space, and let ()\mathbb{P}(\mathcal{H}) be the associated projective Hilbert space, viewed as a set. We denote by 𝔥()\mathfrak{h}(\mathcal{H}) the space of bounded self-adjoint operators on \mathcal{H}. There exists a natural injective map

ι:()𝔥();[z]zz|z|2,\iota:\mathbb{P}(\mathcal{H})\rightarrow\mathfrak{h}(\mathcal{H});[z]\mapsto\frac{z\otimes z^{*}}{|z|^{2}},

where zzz\otimes z^{*} is the operator defined by yy,zz,yy\mapsto\langle y,z\rangle z,y\in\mathcal{H}. Let XX be a finite-dimensional (possibly non-compact) complex manifold with a volume form dμd\mu. We say a map F:X()F:X\rightarrow\mathbb{P}(\mathcal{H}) is a holomorphic embedding if it is injective and locally induced by a holomorphic embedding into \mathcal{H}.

We say a holomorphic embedding F:XF:X\rightarrow\mathcal{H} is balanced if it satisfies the equation

(1.1) c(F)Xι(F(z))𝑑μ=CId,c(F)\equiv\int_{X}\iota(F(z))d\mu=C\cdot\text{Id},

where CC is a constant. More generally, we consider a pair (X,D)(X,D), where DD is a divisor in XX. Let dμd\mu be a volume form on XX and dνd\nu be a volume form on DD. We say that a holomorphic embedding F:X()F:X\rightarrow\mathbb{P}(\mathcal{H}) is (β,C)(\beta,C)-balanced for some β,C\beta,C\in\mathbb{R} if

(1.2) Xι(F(z))𝑑μ+βDι(F(z))𝑑ν=CId,\int_{X}\iota(F(z))d\mu+\beta\int_{D}\iota(F(z))d\nu=C\cdot\text{Id},

where ι(F(z))\iota(F(z)) denotes the Kähler form on ()\mathbb{P}(\mathcal{H}) induced by the Fubini-Study metric.

We consider the setting where XX is the complex plane \mathbb{C} equipped with the standard volume form dμ=1πω0d\mu=\frac{1}{\pi}\omega_{0}, and DD is the origin 0{0} equipped with the dirac measure δ0\delta_{0}. Let \mathcal{H} be the dual of the Hilbert space of entire holomorphic functions on \mathbb{C} endowed with the L2L^{2} inner product associated with dμd\mu. Then a basis of \mathcal{H}^{*} induces a map F:X()F:X\to\mathbb{P}(\mathcal{H}). And we are interested to know if we can find a basis so that the induced map FF is (β,1)(\beta,1)-balanced, β[0,1)\beta\in[0,1). Then our main result is the following.

Theorem 1.1.

For β[0,1)\beta\in[0,1) there is a (β,1)(\beta,1)-balanced embedding F:X()F:X\rightarrow\mathbb{P}(\mathcal{H}) which is unique up to the action of U()U(\mathcal{H}).

Since the existence has been proved in [8], we only need to show the uniqueness part. As explained in [8], this is equivalent to proving the uniqueness of the solution of the following equation

(1.3) 0f(sx)f(x)𝑑x=11sβ,\int_{0}^{\infty}\frac{f(sx)}{f(x)}dx=\frac{1}{1-s}-\beta,

for all s[0,1)s\in[0,1), where

f(x)=i=0eλixi,f(x)=\sum_{i=0}^{\infty}e^{-\lambda_{i}}x^{i},

is an analytic function on [0,)[0,\infty). As equation 1.3 does not change when we multiply a constant on f(x)f(x), we need to prove the uniqueness up to a scalar multiplication.

In the case of β=0\beta=0, the uniqueness is already non-trivial, and it follows from a result in analysis by Miles-Williamson [5]. In [1], Cuccu-Loi proved the uniqueness of balanced metric(β=0\beta=0) on n\mathbb{C}^{n} under some symmetry condition, by using the result of Miles-Williamson [5]. It is also worth mentioning that Wang [9] and Mossa-Loi [4] proved the uniqueness of balanced metrics on vector bundles on compact Kähler manifolds, which is related to algebraic stability of vector bundles. Of course, our method is very different from theirs due to the infinite-dimensional nature of our problem.

A corollary of our result is the continuous dependence of the solutions on β\beta.

Corollary 1.2.

Let 𝛌β=(βλi){}^{\beta}\bm{\lambda}=(^{\beta}\lambda_{i}) be the unique solution to (1.3) with λ0β=1{}^{\beta}\lambda_{0}=1. Then for each ii, λiβ{}^{\beta}\lambda_{i} depends on continuously on β\beta. Equivalently, the corresponding function fβ(x){}^{\beta}f(x) depends continuously on β\beta for xx in any finite interval.

As our setting on (,0)(\mathbb{C},0) is expected to play the role of a model, it is then important to understand the asymptotics of fβ(x){}^{\beta}f(x) as xx\to\infty. We make the following conjecture in this direction.

Conjecture 1.3.

Let fβ(x){}^{\beta}f(x) be the unique solution to (1.3) with (0)β=1{}^{\beta}(0)=1, then fβ{}^{\beta}f should asymptotically be CβxβexC_{\beta}x^{\beta}e^{x} as x+x\to+\infty.

We will now talk about the proof. Let u(x)=xf(x)/f(x)u(x)=xf^{\prime}(x)/f(x). In [5], the two main ingredients are a result of [2] and the explicit formula for the coefficients of the Taylor expansion of exe^{x}. More precisely, [2] showed that when β=0\beta=0, if f(x)f(x) satisfies the equation 1.3, then ux=O(x)u-x=O(\sqrt{x}) as xx\to\infty. Although this estimate likely holds for β>0\beta>0, we no longer have explicit formulas for the coefficients of f(x)f(x). So we have to take a different route, in which we do not rely on the result from [2].

Since the coefficients eλie^{-\lambda_{i}} and the function ff are mutually determined, our proof involves improving our knowledge of both eλie^{-\lambda_{i}} and ff simultaneously. To obtain good estimates of the λi\lambda_{i}’s, we need to control dudx=xd2dt2logf(x)\frac{du}{dx}=x\frac{d^{2}}{dt^{2}}\log f(x), where t=logxt=\log x. Noticing that d2dt2logf(x)>0\frac{d^{2}}{dt^{2}}\log f(x)>0 and λ′′(i)>0\lambda^{\prime\prime}(i)>0, our main strategy is to make use of the convexity. In particular, in section 3, we will show that there is a specific type of convex function(called dominating function), which is computable and can be used to control the needed integrals of general convex functions. And that is the main observation and the most technical part of this article. Then, using this technique, we can show the asymptotic behavior of λi\lambda_{i} as ii approaches infinity and the asymptotic behavior of f(x)f(x) as xx approaches infinity, which then implies the uniqueness of f(x)f(x).

This paper is organized as follows: In Section 2, we build up the rough estimates; In Section 3, we develop the tool of dominating function to control the integrals of general convex functions; Then in Section 4, we use the dominating function to prove more accurate estimates, which can then be used in section 5 to prove the uniqueness of the balanced metric.

Acknowledgements. The author would like to thank Professor Song Sun for many very helpful discussions.

2. rough estimates

We fix β(0,1)\beta\in(0,1), and let f(x)f(x) be an entire function satisfying the equation 1.3. We make the normalization so that f(0)=1f(0)=1, i.e. λ0=0\lambda_{0}=0.

2.1. degree 1 estimate

Recall that we have defined u=xf(x)f(x)u=x\frac{f^{\prime}(x)}{f(x)}.

Proposition 2.1.
(2.1) limxux=1.\lim_{x\to\infty}\frac{u}{x}=1.

The proof basically repeats the arguments in section 3.0.1 in [8], but for the convenience of the readers, we include it here.

Proof.

We use the arguments in section 3 in [8]. First, we have

0f((1t)x)f(x)𝑑x=1tβ,\int_{0}^{\infty}\frac{f((1-t)x)}{f(x)}dx=\frac{1}{t}-\beta,

and

0f(x/(1+t))f(x)𝑑x=1+ttβ.\int_{0}^{\infty}\frac{f(x/(1+t))}{f(x)}dx=\frac{1+t}{t}-\beta.

Then we get for all t>0t>0 that

|0etuφ(u)𝑑u1t|1+β.|\int_{0}^{\infty}e^{-tu}\varphi(u)du-\frac{1}{t}|\leq 1+\beta.

So for any nn\in\mathbb{N}, we have

|0etuentuφ(u)𝑑u1(n+1)t|1+β.|\int_{0}^{\infty}e^{-tu}e^{-ntu}\varphi(u)du-\frac{1}{(n+1)t}|\leq 1+\beta.

Let gg be any measurable function on [0,1][0,1] with bounded variation. Fix any ε>0\varepsilon>0 small, by the Stone-Weierstrass approximation theorem there are polynomials pp and PP such that

p(x)<g(x)<P(x),p(x)<g(x)<P(x),

and

0eu(P(eu)p(eu))𝑑t<ε.\int_{0}^{\infty}e^{-u}(P(e^{-u})-p(e^{-u}))dt<\varepsilon.

For any nn\in\mathbb{N}, we have

|0etuentuφ(u)𝑑u1(n+1)t|C01+(n+1)1/2t1/2(n+1)1/2t1/2.|\int_{0}^{\infty}e^{-tu}e^{-ntu}\varphi(u)du-\frac{1}{(n+1)t}|\leq C_{0}\frac{1+(n+1)^{1/2}t^{1/2}}{(n+1)^{1/2}t^{1/2}}.

Let NN be the maximum of degp\deg p and degP\deg P. So

|0etuP(etu)φ(u)𝑑u1t0euP(eu)𝑑u|(1+β)(1+N).|\int_{0}^{\infty}e^{-tu}P(e^{-tu})\varphi(u)du-\frac{1}{t}\int_{0}^{\infty}e^{-u}P(e^{-u})du|\leq(1+\beta)(1+N).

So

0etug(etu)φ(u)𝑑u1t(0eug(eu)𝑑u+ε)+(1+β)(1+N).\int_{0}^{\infty}e^{-tu}g(e^{-tu})\varphi(u)du\leq\frac{1}{t}(\int_{0}^{\infty}e^{-u}g(e^{-u})du+\varepsilon)+(1+\beta)(1+N).

Similarly, we have

0etug(etu)φ(t)𝑑t1t(0eug(eu)𝑑uε)(1+β)(1+N).\int_{0}^{\infty}e^{-tu}g(e^{-tu})\varphi(t)dt\geq\frac{1}{t}(\int_{0}^{\infty}e^{-u}g(e^{-u})du-\varepsilon)-(1+\beta)(1+N).

Now define g(x)g(x) to be zero for x<e1x<e^{-1} and 1/x1/x for x[e1,1]x\in[e^{-1},1]. Then

0eug(eu)𝑑u=1,\int_{0}^{\infty}e^{-u}g(e^{-u})du=1,

and

0etug(etu)φ(u)𝑑u=01/tφ(u)𝑑u.\int_{0}^{\infty}e^{-tu}g(e^{-tu})\varphi(u)du=\int_{0}^{1/t}\varphi(u)du.

Let T=1/tT=1/t and fix ε>0\varepsilon>0 sufficiently small we obtain

0Tφ(u)𝑑u(1+ε)T+(1+β)(1+N).\int_{0}^{T}\varphi(u)du\leq(1+\varepsilon)T+(1+\beta)(1+N).

Similarly, we have

(2.2) 0Tφ(u)𝑑u(1ε)T(1+β)(1+N).\int_{0}^{T}\varphi(u)du\geq(1-\varepsilon)T-(1+\beta)(1+N).

So we get

1εlim infxuxlim supxux1+ε.1-\varepsilon\leq\liminf_{x\to\infty}\frac{u}{x}\leq\limsup_{x\to\infty}\frac{u}{x}\leq 1+\varepsilon.

Since ε\varepsilon is arbitrary, the conclusion follows.

2.2. rough degree 2 estimates

Denote by t=logxt=\log x. Then we have

0xndxf(x)=e(n+1)tlogf(x)𝑑t,\int_{0}^{\infty}\frac{x^{n}dx}{f(x)}=\int_{-\infty}^{\infty}e^{(n+1)t-\log f(x)}dt,

and

f(x)=i=0eλi+ilogx.f(x)=\sum_{i=0}^{\infty}e^{-\lambda_{i}+i\log x}.

In both the integral and the summation, the exponents have concavity of which we will make heavy use. For any aa\in\mathbb{R}, we denote

ga(t)=atlogf(x).g_{a}(t)=at-\log f(x).

It is easy to check that gag_{a} is a strictly concave function of tt. Since ga(t)=au(x)g_{a}^{\prime}(t)=a-u(x), and by proposition 2.1 limxf(x)f(x)=1\lim_{x\to\infty}\frac{f^{\prime}(x)}{f(x)}=1, we see that ga(t)g_{a}(t) has a unique maximum at ta=logxat_{a}=\log x_{a} with u(xa)=au(x_{a})=a. Moreover, we have dxada>0\frac{dx_{a}}{da}>0 and limaa1xa=1\lim_{a\rightarrow\infty}a^{-1}x_{a}=1. In particular, for xx large there is a unique aa such that x=xax=x_{a}.

We also define the function

(2.3) c(a)=(0xaf(x)𝑑x)1c(a)=(\int_{0}^{\infty}\frac{x^{a}}{f(x)}dx)^{-1}

and λ(a)=logc(a)\lambda(a)=-\log c(a) for a[0,)a\in[0,\infty). Notice that when a>0a>0 is an integer we have c(a)=cac(a)=c_{a} and λ(a)=λa\lambda(a)=\lambda_{a}. One also notices that c(0)c0c(0)\neq c_{0}, which can be ignored, since it will not affect our arguments.

One can compute

(2.4) λ(a)=c(a)0(logx)xadxf(x),\lambda^{\prime}(a)=c(a)\int_{0}^{\infty}\frac{(\log x)x^{a}dx}{f(x)},

and

(2.5) d2da2λ(a)=c2(a)[0xadxf(x)0(logx)2xadxf(x)(0(logx)xadxf(x))2]>0.\displaystyle\frac{d^{2}}{da^{2}}\lambda(a)=c^{2}(a)[\int_{0}^{\infty}\frac{x^{a}dx}{f(x)}\int_{0}^{\infty}\frac{(\log x)^{2}x^{a}dx}{f(x)}-(\int_{0}^{\infty}\frac{(\log x)x^{a}dx}{f(x)})^{2}]>0.

For x>1x>1 we denote

Fx(n)=nlogxλ(n),F_{x}(n)=n\log x-\lambda(n),

then Fx(n)F_{x}(n) is a strictly concave function of nn. It follows that Fx(n)F_{x}(n) has at most one maximum. Estimate (3.9) in [8] implies that λ(n)+\lambda^{\prime}(n)\to+\infty as n+n\to+\infty. Therefore when xx is large Fx(n)F_{x}(n) attains its maximum at nx>1n_{x}>1. Moreover, as xx\rightarrow\infty, we have nxn_{x}\rightarrow\infty.

Now for all a1a\geq 1 we denote

ha(x)=ega(x)=f(x)c(a)xa.h_{a}(x)=e^{-g_{a}(x)}=\frac{f(x)}{c(a)x^{a}}.
Lemma 2.2.

For a1a\gg 1, we have ha(xa)a13h_{a}(x_{a})\geq\frac{\sqrt{a}}{13}.

Proof.

For simplicity, we assume aa is an integer. The case when aa is not an integer is similar, just with more complicated notations. We write

ha(x)=i=0δi(x),h_{a}(x)=\sum_{i=0}^{\infty}\delta_{i}(x),

where δi(x)=ca1cixia\delta_{i}(x)=c_{a}^{-1}c_{i}x^{i-a}. Then for a1a\geq 1, we have 0ha(x)1𝑑x=1\int_{0}^{\infty}h_{a}(x)^{-1}dx=1.

Assume that ha(xa)<a13h_{a}(x_{a})<\frac{\sqrt{a}}{13} for some a1a\gg 1. Let λx(i)=Fx(i)=logx+c(i)c(i)\lambda_{x}(i)=F_{x}^{\prime}(i)=\log x+\frac{c^{\prime}(i)}{c(i)}. By definition, λx(nx)=0\lambda_{x}(n_{x})=0 and δa(xa)=1\delta_{a}(x_{a})=1 . Let γ=δa+[a](xa)\gamma=\delta_{a+[\sqrt{a}]}(x_{a}). Then we have, by concavity,

i=aa+[a]δi(xa)[a]min(γ,1).\sum_{i=a}^{a+[\sqrt{a}]}\delta_{i}(x_{a})\geq[\sqrt{a}]\min(\gamma,1).

It follows that γ<1\gamma<1 and

(2.6) nxa<a+a.n_{x_{a}}<a+\sqrt{a}.

Then we have γ<a13[a]\gamma<\frac{\sqrt{a}}{13[\sqrt{a}]}. Furthermore we must have λxa(a+[a])<0\lambda_{x_{a}}(a+[\sqrt{a}])<0.

By the mean value theorem, λxa(a+[a])[a]>log1γ-\lambda_{x_{a}}(a+[\sqrt{a}])[\sqrt{a}]>\log\frac{1}{\gamma}. Therefore

λxa(a+[a])<1[a]log(13[a]a).\lambda_{x_{a}}(a+[\sqrt{a}])<\frac{-1}{[\sqrt{a}]}\log(\frac{13[\sqrt{a}]}{\sqrt{a}}).

Then

λxa+a(a+[a])\displaystyle\lambda_{x_{a}+\sqrt{a}}(a+[\sqrt{a}]) =\displaystyle= λxa(a+[a])+log(1+axa)\displaystyle\lambda_{x_{a}}(a+[\sqrt{a}])+\log(1+\frac{\sqrt{a}}{x_{a}})
<\displaystyle< axa1[a]log(13[a]a)\displaystyle\frac{\sqrt{a}}{x_{a}}-\frac{1}{[\sqrt{a}]}\log(\frac{13[\sqrt{a}]}{\sqrt{a}})
<\displaystyle< 0,\displaystyle 0,

if aa is sufficiently large. Therefore, nxa+a<a+[a].n_{x_{a}+\sqrt{a}}<a+[\sqrt{a}]. So for 0<y<a0<y<\sqrt{a}, we have

(2.7) i>a+aδi(xa+y)<γ(1+yxa)[a](axa1[a]log(13[a]a))<e2a26,\sum_{i>a+\sqrt{a}}\delta_{i}(x_{a}+y)<\frac{\gamma(1+\frac{y}{x_{a}})^{[\sqrt{a}]}}{-(\frac{\sqrt{a}}{x_{a}}-\frac{1}{[\sqrt{a}]}\log(\frac{13[\sqrt{a}]}{\sqrt{a}}))}<\frac{e^{2}\sqrt{a}}{26},

where the second inequality is because the denominator is larger than 2a\frac{2}{\sqrt{a}} and (1+yxa)[a]<e2(1+\frac{y}{x_{a}})^{[\sqrt{a}]}<e^{2} for aa large.

Since δi(xa+y)δi(xa)=(1+yxa)ia\frac{\delta_{i}(x_{a}+y)}{\delta_{i}(x_{a})}=(1+\frac{y}{x_{a}})^{i-a}, we also have

(2.8) aia+aδi(xa+y)<(1+yxa)[a]aia+aδi(xa)<e2a13.\sum_{a\leq i\leq a+\sqrt{a}}\delta_{i}(x_{a}+y)<(1+\frac{y}{x_{a}})^{[\sqrt{a}]}\sum_{a\leq i\leq a+\sqrt{a}}\delta_{i}(x_{a})<\frac{e^{2}\sqrt{a}}{13}.

So we have ha(xa+y)<(1+e2+e22)a13<ah_{a}(x_{a}+y)<(1+e^{2}+\frac{e^{2}}{2})\frac{\sqrt{a}}{13}<\sqrt{a}. Therefore

01ha(x)𝑑x>1,\int_{0}^{\infty}\frac{1}{h_{a}(x)}dx>1,

which is a contradiction. ∎

We can write

ha(x)=1cai=0e(ia)t+logc(i).h_{a}(x)=\frac{1}{c_{a}}\sum_{i=0}^{\infty}e^{(i-a)t+\log c(i)}.

Then

ddtlogha(x)=i=0(ia)e(ia)t+logc(i)i=0e(ia)t+logc(i),\frac{d}{dt}\log h_{a}(x)=\frac{\sum_{i=0}^{\infty}(i-a)e^{(i-a)t+\log c(i)}}{\sum_{i=0}^{\infty}e^{(i-a)t+\log c(i)}},

and

d2dt2logha(x)\displaystyle\frac{d^{2}}{dt^{2}}\log h_{a}(x) =\displaystyle= had2dt2ha(ddtha)2ha2\displaystyle\frac{h_{a}\frac{d^{2}}{dt^{2}}h_{a}-(\frac{d}{dt}h_{a})^{2}}{h^{2}_{a}}
=\displaystyle= i=0(iaσa(x))2τx(i),\displaystyle\sum_{i=0}^{\infty}(i-a-\sigma_{a}(x))^{2}\tau_{x}(i),

where τx(i)=hi(x)1\tau_{x}(i)=h_{i}(x)^{-1}, and σa(x)=ddtha(x)ha(x)\sigma_{a}(x)=\frac{\frac{d}{dt}h_{a}(x)}{h_{a}(x)}.

We denote t~a=logx~a=ddaλ(a)\tilde{t}_{a}=\log\tilde{x}_{a}=\frac{d}{da}\lambda(a). Then nx~a=an_{\tilde{x}_{a}}=a and ha(x~a)ha(xa)a13.h_{a}(\tilde{x}_{a})\geq h_{a}(x_{a})\geq\frac{\sqrt{a}}{13}. The convexity of λ(n)\lambda(n) implies that x~a\tilde{x}_{a} is a strictly increasing function of aa. In particular, for xx large, there is a unique aa such that x=x~ax=\tilde{x}_{a}.

Lemma 2.3.

Suppose there exists A>0A>0 such that ha(xa)>aAh_{a}(x_{a})>\frac{\sqrt{a}}{A} for all aa large enough. Then for any δ>0\delta>0, we have

d2dt2logf(x)>x27A2+δ,\frac{d^{2}}{dt^{2}}\log f(x)>\frac{x}{27A^{2}+\delta},

for xx large enough.

Proof.

Fix δ>0\delta>0. For xx large we may write x=x~ax=\tilde{x}_{a} for a unique large aa. By concavity of FxF_{x} we have τx(i)1ha(x)\tau_{x}(i)\leq\frac{1}{h_{a}(x)} for all ii. So if we take the interval I=(σa(x)+aha(x)3,σa(x)+aha(x)3)I=(\sigma_{a}(x)+a-\frac{h_{a}(x)}{3},\sigma_{a}(x)+a-\frac{h_{a}(x)}{3}), then iIτx(i)<23\sum_{i\in I}\tau_{x}(i)<\frac{2}{3}. So iIτx(i)>13\sum_{i\notin I}\tau_{x}(i)>\frac{1}{3}. Therefore

i=0(iaσa(x))2τx(i)>13(ha(x)3)2>a27A2\sum_{i=0}^{\infty}(i-a-\sigma_{a}(x))^{2}\tau_{x}(i)>\frac{1}{3}(\frac{h_{a}(x)}{3})^{2}>\frac{a}{27A^{2}}

If x~a<xa\tilde{x}_{a}<x_{a}, then since limaaxa=1\lim_{a\to\infty}\frac{a}{x_{a}}=1, we get that for aa large that

d2dt2logha(x~a)>x~a27A2+δ.\frac{d^{2}}{dt^{2}}\log h_{a}(\tilde{x}_{a})>\frac{\tilde{x}_{a}}{27A^{2}+\delta}.

If x~a>xa\tilde{x}_{a}>x_{a}, then nxa<an_{x_{a}}<a. So λxa(a)=logxax~a<0\lambda_{x_{a}}(a)=\log\frac{x_{a}}{\tilde{x}_{a}}<0. Hence

ca1ia+1(ia)cixaia<j1jejlogxax~a=xax~a(1xa/x~a)2.c_{a}^{-1}\sum_{i\geq a+1}(i-a)c_{i}x_{a}^{i-a}<\sum_{j\geq 1}je^{j\log\frac{x_{a}}{\tilde{x}_{a}}}=\frac{x_{a}}{\tilde{x}_{a}(1-x_{a}/\tilde{x}_{a})^{2}}.

Since i(ia)cixaia=0\sum_{i}(i-a)c_{i}x_{a}^{i-a}=0, we have

ca1i<a(ai)cixaia<xax~a(1xa/x~a)2+1c_{a}^{-1}\sum_{i<a}(a-i)c_{i}x_{a}^{i-a}<\frac{x_{a}}{\tilde{x}_{a}(1-x_{a}/\tilde{x}_{a})^{2}}+1

On the other hand,

ca1nxai<a(ai)cixaiaj=1[a]nxaj12([a]nxa)2.c_{a}^{-1}\sum_{n_{x_{a}}\leq i<a}(a-i)c_{i}x_{a}^{i-a}\geq\sum_{j=1}^{[a]-\ulcorner n_{x_{a}}\urcorner}j\geq\frac{1}{2}([a]-\ulcorner n_{x_{a}}\urcorner)^{2}.

We denote ra=[a]nxar_{a}=[a]-\ulcorner n_{x_{a}}\urcorner and βa=xax~a\beta_{a}=\frac{x_{a}}{\tilde{x}_{a}}. So

12ra2<βa(1βa)2+1.\frac{1}{2}r_{a}^{2}<\frac{\beta_{a}}{(1-\beta_{a})^{2}}+1.

Let ξ(nxa,a)\xi\in(n_{x_{a}},a) such that c(ξ)c(ξ)c(a)c(a)=12logx~axa\frac{c^{\prime}(\xi)}{c(\xi)}-\frac{c^{\prime}(a)}{c(a)}=\frac{1}{2}\log\frac{\tilde{x}_{a}}{x_{a}}. Then we can estimate

hξ(x~ξ)<ra+4+2βa1βa.h_{\xi}(\tilde{x}_{\xi})<r_{a}+4+\frac{2\sqrt{\beta_{a}}}{1-\sqrt{\beta_{a}}}.

We have hξ(x~ξ)ξ13h_{\xi}(\tilde{x}_{\xi})\geq\frac{\sqrt{\xi}}{13}. When aa\to\infty, we have xax_{a}\to\infty, so nxan_{x_{a}}\to\infty. In particular, when aa is large, ξ\xi is also large. We have

βa27A2xa(27A2+δ)a.\beta_{a}\geq\frac{27A^{2}x_{a}}{(27A^{2}+\delta)a}.

So we always have

d2dt2logf(x)=d2dt2logha(x~a)>x~a27A2+δ,\frac{d^{2}}{dt^{2}}\log f(x)=\frac{d^{2}}{dt^{2}}\log h_{a}(\tilde{x}_{a})>\frac{\tilde{x}_{a}}{27A^{2}+\delta},

Similarly, we get

Lemma 2.4.

Suppose there exists A>0A>0 such that ha(xa)>aAh_{a}(x_{a})>\frac{\sqrt{a}}{A} for all aa large enough. Then for any δ>0\delta>0, we have

d2da2λ(a)>1(27A2+δ)a,\frac{d^{2}}{da^{2}}\lambda(a)>\frac{1}{(27A^{2}+\delta)a},

for aa large enough.

Proof.

From (2.4) we have

t~a=0logxdxha(x).\tilde{t}_{a}=\int_{0}^{\infty}\frac{\log xdx}{h_{a}(x)}.

So by (2.5) we get

d2da2λ(a)=0(logxt~a)2dxha(x).\frac{d^{2}}{da^{2}}\lambda(a)=\int_{0}^{\infty}\frac{(\log x-\tilde{t}_{a})^{2}dx}{h_{a}(x)}.

We take I=(x~aa3A,x~a+a3A)I=(\tilde{x}_{a}-\frac{\sqrt{a}}{3A},\tilde{x}_{a}+\frac{\sqrt{a}}{3A}), then since 1ha(x)1ha(xa)<Aa\frac{1}{h_{a}(x)}\leq\frac{1}{h_{a}(x_{a})}<\frac{A}{\sqrt{a}}, we have xIdxha(x)<23\int_{x\in I}\frac{dx}{h_{a}(x)}<\frac{2}{3}. So xIdxha(x)>13\int_{x\notin I}\frac{dx}{h_{a}(x)}>\frac{1}{3}. Now

d2da2λ(a)>xI(logxt~a)2dxha(x)>13log2(1+a3Ax~a).\frac{d^{2}}{da^{2}}\lambda(a)>\int_{x\notin I}\frac{(\log x-\tilde{t}_{a})^{2}dx}{h_{a}(x)}>\frac{1}{3}\log^{2}(1+\frac{\sqrt{a}}{3A\tilde{x}_{a}}).

As we have shown in the last lemma for any ε>0\varepsilon>0, we have x~aa>1ε\frac{\tilde{x}_{a}}{a}>1-\varepsilon for aa large enough. Therefore, we have

d2da2λ(a)>1(27A2+δ)a,\frac{d^{2}}{da^{2}}\lambda(a)>\frac{1}{(27A^{2}+\delta)a},

for aa large enough. ∎

Now let A=13A=13, the above results imply the following.

Lemma 2.5.

For x1x\gg 1 we have

d2dt2logf(x)>x5000,\frac{d^{2}}{dt^{2}}\log f(x)>\frac{x}{5000},

and for a1a\gg 1 we have

d2da2λ(a)>15000a.\frac{d^{2}}{da^{2}}\lambda(a)>\frac{1}{5000a}.

As corollaries, we have the following.

Proposition 2.6.

For aa large we have t~a=ta+1+O(logaa)\tilde{t}_{a}=t_{a+1}+O(\frac{\log a}{\sqrt{a}})

Proof.

Notice tlogha(x)=ga+1(t)+logc(a)t-\log h_{a}(x)=g_{a+1}(t)+\log c(a), and ga+1(t)g_{a+1}(t) has unique maximum at ta+1t_{a+1}. By Lemma 2.5 and the standard Laplace’s method we see that there exists a constant C>0C>0 so that

|tta+1|>Clogaatha(x)𝑑x=ε(a),\int_{|t-t_{a+1}|>C\frac{\log a}{\sqrt{a}}}\frac{t}{h_{a}(x)}dx=\varepsilon(a),

where ε(a)\varepsilon(a) denotes a quantity which is O(ak)O(a^{-k}) as aa\rightarrow\infty for all k>0k>0. So by (2.4) we have

t~a=ta+1+|tta+1|Clogaatta+1ha(x)𝑑x+ε(a)=ta+1+O(logaa).\tilde{t}_{a}=t_{a+1}+\int_{|t-t_{a+1}|\leq C\frac{\log a}{\sqrt{a}}}\frac{t-t_{a+1}}{h_{a}(x)}dx+\varepsilon(a)=t_{a+1}+O(\frac{\log a}{\sqrt{a}}).

Proposition 2.7.

We have

d2da2λ(a)<21071a\frac{d^{2}}{da^{2}}\lambda(a)<2*10^{7}\frac{1}{a}

for aa large enough, and

d2dt2logf(x)<2107x\frac{d^{2}}{dt^{2}}\log f(x)<2*10^{7}x

for xx large enough.

Proof.

Let x=x~ax=\tilde{x}_{a}, then τx(i)\tau_{x}(i) achieves its maximum at i=ai=a. So

d2dt2logf(x)=d2dt2logha(x)\displaystyle\frac{d^{2}}{dt^{2}}\log f(x)=\frac{d^{2}}{dt^{2}}\log h_{a}(x) =\displaystyle= (iaddtha(x))2τx(i)\displaystyle\sum(i-a-\frac{d}{dt}h_{a}(x))^{2}\tau_{x}(i)
\displaystyle\leq (ia)2τx(i)\displaystyle\sum(i-a)^{2}\tau_{x}(i)
\displaystyle\leq 1ha(x~a)(O(1)+y2ey210000a𝑑y)\displaystyle\frac{1}{h_{a}(\tilde{x}_{a})}(O(1)+\int_{-\infty}^{\infty}y^{2}e^{-\frac{y^{2}}{10000a}}dy)
\displaystyle\leq 13a(O(1)+(10000a)3/2π2)\displaystyle\frac{13}{\sqrt{a}}(O(1)+\frac{(10000a)^{3/2}\sqrt{\pi}}{2})
\displaystyle\leq 1.2107a\displaystyle 1.2*10^{7}a
\displaystyle\leq 2107x~a,\displaystyle 2*10^{7}\tilde{x}_{a},

Therefore dudx<2107\frac{du}{dx}<2*10^{7}. So xa+1=xa+O(1)x_{a+1}=x_{a}+O(1). Similarly, since

d2da2λ(a)=(logxt~a)2dxha(x)(logxta+1)2dxha(x),\frac{d^{2}}{da^{2}}\lambda(a)=\int\frac{(\log x-\tilde{t}_{a})^{2}dx}{h_{a}(x)}\leq\int\frac{(\log x-t_{a+1})^{2}dx}{h_{a}(x)},

we can substitute y=logxta+1y=\log x-t_{a+1} to get

d2da2λ(a)\displaystyle\frac{d^{2}}{da^{2}}\lambda(a) \displaystyle\leq xa+1ha(xa+1)ea11000y2𝑑y\displaystyle\frac{x_{a+1}}{h_{a}(x_{a+1})}\int_{-\infty}^{\infty}e^{-\frac{a}{11000}}y^{2}dy
\displaystyle\leq 14aπ(11000)3/22a3/2\displaystyle 14\sqrt{a}\frac{\sqrt{\pi}(11000)^{3/2}}{2a^{3/2}}
\displaystyle\leq 21071a.\displaystyle 2*10^{7}\frac{1}{a}.

Now we get a more quantitative estimate on xax_{a}.

Proposition 2.8.

xa=a+O(aloga)x_{a}=a+O(\sqrt{a}\log a) for aa large enough.

Proof.

First, we have

0xa+alogacixif(x)𝑑x=1βδ0i+ε(a)\int_{0}^{x_{a}+\sqrt{a}\log a}\frac{c_{i}x^{i}}{f(x)}dx=1-\beta\delta_{0i}+\varepsilon(a)

for 0ia0\leq i\leq a. So

xa+aloga>0xa+alogai=0acixif(x)𝑑x=a+1β+ε(a)x_{a}+\sqrt{a}\log a>\int_{0}^{x_{a}+\sqrt{a}\log a}\frac{\sum_{i=0}^{a}c_{i}x^{i}}{f(x)}dx=a+1-\beta+\varepsilon(a)

For the other direction, we have

i>acixif(x)=ε(a)\frac{\sum_{i>a}c_{i}x^{i}}{f(x)}=\varepsilon(a)

for xxaalogax\leq x_{a}-\sqrt{a}\log a. Therefore,

a+1>0xaalogai=0acixif(x)𝑑x=xaaloga+ε(a).a+1>\int_{0}^{x_{a}-\sqrt{a}\log a}\frac{\sum_{i=0}^{a}c_{i}x^{i}}{f(x)}dx=x_{a}-\sqrt{a}\log a+\varepsilon(a).

Remark: With these estimates, one can improve the estimates in Lemma 2.2 to get A=2.5A=2.5, then we have that d2dt2logf(x)>x150\frac{d^{2}}{dt^{2}}\log f(x)>\frac{x}{150} for xx large and d2da2λ(a)>1150a\frac{d^{2}}{da^{2}}\lambda(a)>\frac{1}{150a} for aa large. Then one can generate better upper bounds by repeating the proof of Proposition 2.7. But this refinement will not be essential to us. To further narrow down the gap between the upper and lower bounds, we need the techniques in the following subsection.

3. Dominating function

In this section, we will develop the tool of dominating functions to narrow down the estimates for d2da2λ(a)\frac{d^{2}}{da^{2}}\lambda(a) and d2dt2logf(x)\frac{d^{2}}{dt^{2}}\log f(x).

3.1. half real line

Theorem 3.1.

Let g(t)g(t) be a C1C^{1} function on +\mathbb{R}_{+} satisfying:

  • (1)

    g(0)=0g(0)=0;

  • (2)

    g(0)=0g^{\prime}(0)=0;

  • (3)

    The left derivative and right derivative of g(t)g^{\prime}(t) exist for every tt;

  • (4)

    g(t)C2g(t)\in C^{2} except at finite points;

  • (5)

    there exist 0<M1<M2<0<M_{1}<M_{2}<\infty such that M1<g′′(t)<M2M_{1}<g^{\prime\prime}(t)<M_{2} for all tt\in\mathbb{R}.

Denote

a(g)=0eg(t)𝑑t,a(g)=\int_{0}^{\infty}e^{-g(t)}dt,
b(g)=0t2eg(t)𝑑t,b(g)=\int_{0}^{\infty}t^{2}e^{-g(t)}dt,

and

d(g)=b(g)4(a(g))3.d(g)=\frac{b(g)}{4(a(g))^{3}}.

Then there exist 112<q<p<12\frac{1}{12}<q<p<\frac{1}{2} depending only on M1M_{1} and M2M_{2} such that

112<q<d(g)<p<12.\frac{1}{12}<q<d(g)<p<\frac{1}{2}.

Remark: Notice that if g~(y)=g(ρy)\tilde{g}(y)=g(\rho y) for some ρ>0\rho>0, then d(g~)=d(g)d(\tilde{g})=d(g). So the bounds pp and qq depend only on the ratio M2M1\frac{M_{2}}{M_{1}}.

Proof.

For each pair 0<s<l<0<s<l<\infty, we can define a function gs,l(t)g_{s,l}(t) satisfying gs,l(0)=gs,l(0)=0g_{s,l}(0)=g_{s,l}^{\prime}(0)=0 and

gs,l′′(t)={M2,t<sg′′(t),l>tsM1,tlg_{s,l}^{\prime\prime}(t)=\begin{cases}M_{2},&t<s\\ g^{\prime\prime}(t),&l>t\geq s\\ M_{1},&t\geq l\end{cases}

We call a pair 0<s<l<0<s<l<\infty an admissible pair if a(gs,l)=1a(g_{s,l})=1. It is easy to see that when ss is small enough, there exists l>sl>s such that (s,l)(s,l) is admissible. Notice that for an admissible pair (s,l)(s,l) the following holds:

l1>lsatisfying{g(t)gs,l(t),tl1g(t)gs,l(t),t>l1\exists\ l_{1}>l\ \ \textit{satisfying}\begin{cases}g(t)\leq g_{s,l}(t),&t\leq l_{1}\\ g(t)\geq g_{s,l}(t),&t>l_{1}\end{cases}

Therefore

b(g)b(gs,l)=0(t2l12)(eg(t)egs,l(t))𝑑t0.b(g)-b(g_{s,l})=\int_{0}^{\infty}(t^{2}-l_{1}^{2})(e^{-g(t)}-e^{-g_{s,l}(t)})dt\leq 0.

Hence, we have d(g)d(gs,l).d(g)\leq d(g_{s,l}).

For later references, we will call this operation that replaces gg with gs,lg_{s,l} operation A.

Let l(s)l(s) be the smallest ll such that (s,l)(s,l) is admissible. Then by a straightforward continuity argument, one sees that s0>0\exists s_{0}>0 such that s0=l(s0)s_{0}=l(s_{0}). Clearly, gs0,l(s0)g_{s_{0},l(s_{0})} is determined by the pair (M1,M2)(M_{1},M_{2}), so we can denote by dM1,M2=d(gs0,l(s0))d_{M_{1},M_{2}}=d(g_{s_{0},l(s_{0})}). Then one easily sees that for 0<M1M1M2M20<M_{1}^{\prime}\leq M_{1}\leq M_{2}\leq M_{2}^{\prime}, we have d(g)dM1,M2d(g)\leq d_{M_{1}^{\prime},M_{2}^{\prime}}. So for δ>0\delta>0 small enough, we have

d(g)dδ,1/δ.d(g)\leq d_{\delta,1/\delta}.

We denote by sδs_{\delta} the s0s_{0} corresponding to the pair (δ,1/δ)(\delta,1/\delta), and by gδg_{\delta} the corresponding function gs0,l(s0)g_{s_{0},l(s_{0})}. Then we have

gδ(t)={t22δ,tsδsδ22δ+sδδ|tsδ|+δ2(tsδ)2,t>sδg_{\delta}(t)=\begin{cases}\frac{t^{2}}{2\delta},&t\leq s_{\delta}\\ \frac{s_{\delta}^{2}}{2\delta}+\frac{s_{\delta}}{\delta}|t-s_{\delta}|+\frac{\delta}{2}(t-s_{\delta})^{2},&t>s_{\delta}\end{cases}

Then in order for a(gδ)=1a(g_{\delta})=1, sδs_{\delta} has to converge to 0 when δ0\delta\to 0 and limδ0sδδ=1\lim_{\delta\to 0}\frac{s_{\delta}}{\delta}=1. So

limδ0gδ(t)=g0(t)=t.\lim_{\delta\to 0}g_{\delta}(t)=g_{0}(t)=t.

Then by the dominated convergence limδ0d(gδ)=d(g0)=12\lim_{\delta\to 0}d(g_{\delta})=d(g_{0})=\frac{1}{2}. Therefore we have proved the upper bound.

For the lower bound, the argument is similar. We also replace gg with g~s,l\tilde{g}_{s,l} satisfying

g~s,l′′(t)={M1,t<sg′′(t),l>tsM2,tl\tilde{g}_{s,l}^{\prime\prime}(t)=\begin{cases}M_{1},&t<s\\ g^{\prime\prime}(t),&l>t\geq s\\ M_{2},&t\geq l\end{cases}

For later references, we will call this operation that replaces gg with g~s,l\tilde{g}_{s,l} operation B.

And we also get a function

gδ(t)={δ2t2,tsδδ2sδ2+δsδ(tsδ)+12δ(tsδ)2,t>sδg_{\delta}(t)=\begin{cases}\frac{\delta}{2}t^{2},&t\leq s_{\delta}\\ \frac{\delta}{2}s_{\delta}^{2}+\delta s_{\delta}(t-s_{\delta})+\frac{1}{2\delta}(t-s_{\delta})^{2},&t>s_{\delta}\end{cases}

satisfying d(g)>d(gδ)d(g)>d(g_{\delta}). Again in order for a(gδ)=1a(g_{\delta})=1, s(δ)s(\delta) has to converge to 1/21/2 when δ0\delta\to 0 and

limδ0gδ(t)=g0(t)=χI(t),\lim_{\delta\to 0}g_{\delta}(t)=g_{0}(t)=\chi_{I}(t),

where χI(t)\chi_{I}(t) is the characteristic function of the interval I=[0,1]I=[0,1]. Moreover, a(g0)=1,b(g0)=13a(g_{0})=1,b(g_{0})=\frac{1}{3}, so d(g0)=112d(g_{0})=\frac{1}{12}. Again by dominated convergence limδ0d(gδ)=d(g0)=112\lim_{\delta\to 0}d(g_{\delta})=d(g_{0})=\frac{1}{12}. This proves the lower bound. ∎

By dilation, in the proof of Theorem 3.1 we have actually proved the following two lemmas:

Lemma 3.2.

Let g(y)g(y) be a function satisfying the conditions in Theorem 3.1. Let m=M1M2m=\frac{M_{1}}{M_{2}}. Then there exist c>0c>0 and a function f(y)f(y) satisfying:

  • f(y)=y22f(y)=\frac{y^{2}}{2} for ycy\leq c;

  • f(y)=my22+(1m)cy+mc22c22f(y)=\frac{my^{2}}{2}+(1-m)cy+\frac{mc^{2}}{2}-\frac{c^{2}}{2} for y>cy>c;

  • d(f)d(g)d(f)\geq d(g).

Lemma 3.3.

Let g(y)g(y) be a convex function satisfying the conditions in Theorem 3.1. Let m=M2M1m=\frac{M_{2}}{M_{1}}. Then there exist c>0c>0 and a function f(y)f(y) satisfying:

  • f(y)=y22f(y)=\frac{y^{2}}{2} for ycy\leq c;

  • f(y)=my22+(1m)cy+mc22c22f(y)=\frac{my^{2}}{2}+(1-m)cy+\frac{mc^{2}}{2}-\frac{c^{2}}{2} for y>cy>c.

  • d(f)d(g)d(f)\leq d(g).

Now for fixed m>0m>0, we consider the function

dm(c)=0y2egc(y)𝑑y(0egc(y)𝑑y)3,d_{m}(c)=\frac{\int_{0}^{\infty}y^{2}e^{-g_{c}(y)}dy}{(\int_{0}^{\infty}e^{-g_{c}(y)}dy)^{3}},

where gc(y)g_{c}(y) satisfies:

  • gc(y)=y22g_{c}(y)=\frac{y^{2}}{2} for ycy\leq c;

  • gc(y)=my22+(1m)cy+mc22c22g_{c}(y)=\frac{my^{2}}{2}+(1-m)cy+\frac{mc^{2}}{2}-\frac{c^{2}}{2} for y>cy>c.

And clearly dm(0)=2πd_{m}(0)=\frac{2}{\pi} and limcdm(c)=2π\lim_{c\to\infty}d_{m}(c)=\frac{2}{\pi}. So both the maximum and minimum of dm(c)d_{m}(c) in [0,][0,\infty] exist. Remark: Later on, when we need to emphasize the fact that gc(y)g_{c}(y) also depends on mm, we will denote it by gm,c(y)g_{m,c}(y)

Let

a(c)=0egc(y)𝑑ya(c)=\int_{0}^{\infty}e^{-g_{c}(y)}dy

and

b(c)=0y2egc(y)𝑑y.b(c)=\int_{0}^{\infty}y^{2}e^{-g_{c}(y)}dy.

Then since

cgc(y)=(1m)(yc),\frac{\partial}{\partial c}g_{c}(y)=(1-m)(y-c),

we have

a(c)=e12c212mc2c(m1)(xc)e12mx2+(m1)cx𝑑xa^{\prime}(c)=e^{\frac{1}{2}c^{2}-\frac{1}{2}mc^{2}}\int_{c}^{\infty}(m-1)(x-c)e^{-\frac{1}{2}mx^{2}+(m-1)cx}dx

and

b(c)=e12c212mc2c(m1)(xc)x2e12mx2+(m1)cx𝑑x.b^{\prime}(c)=e^{\frac{1}{2}c^{2}-\frac{1}{2}mc^{2}}\int_{c}^{\infty}(m-1)(x-c)x^{2}e^{-\frac{1}{2}mx^{2}+(m-1)cx}dx.

Denote

γ(c)=a(c)b(c)3b(c)a(c),\gamma(c)=a(c)b^{\prime}(c)-3b(c)a^{\prime}(c),

then d(c)=γ(c)a(c)4d^{\prime}(c)=\frac{\gamma(c)}{a(c)^{4}}, and

γ(0)\displaystyle\gamma(0) =\displaystyle= (m1)[0e12mx2dx0x3e12mx2dx\displaystyle(m-1)[\int_{0}^{\infty}e^{-\frac{1}{2}mx^{2}}dx\int_{0}^{\infty}x^{3}e^{-\frac{1}{2}mx^{2}}dx
30x2e12mx2dx0xe12mx2dx]\displaystyle-3\int_{0}^{\infty}x^{2}e^{-\frac{1}{2}mx^{2}}dx\int_{0}^{\infty}xe^{-\frac{1}{2}mx^{2}}dx]
=\displaystyle= (m1)[π21m21m23π21m3/21m]\displaystyle(m-1)[\sqrt{\frac{\pi}{2}}\frac{1}{\sqrt{m}}\cdot 2\frac{1}{m^{2}}-3\sqrt{\frac{\pi}{2}}\frac{1}{m^{3/2}}\cdot\frac{1}{m}]
=\displaystyle= (1m)π2m5/2\displaystyle(1-m)\sqrt{\frac{\pi}{2}}m^{-5/2}

So d(0)>1d^{\prime}(0)>1 when m<1m<1, and d(0)<1d^{\prime}(0)<1 when m>1m>1.

By Theorem 3.1, we have b(c)a3(c)<124=2\frac{b(c)}{a^{3}(c)}<\frac{1}{2}\cdot 4=2. Since

a(c)π2max{1,1m},a(c)\leq\sqrt{\frac{\pi}{2}}\max\{1,\sqrt{\frac{1}{m}}\},

we get that

b(c)a(c)<πmax{1,1m}.\frac{b(c)}{a(c)}<\pi\max\{1,\frac{1}{m}\}.

On the other hand clearly we have |b(c)a(c)|c2|\frac{b^{\prime}(c)}{a^{\prime}(c)}|\geq c^{2}. Therefore, we have

  • when m<1m<1, d(c)<0d^{\prime}(c)<0 for cc large enough;

  • when m>1m>1, d(c)>0d^{\prime}(c)>0 for cc large enough.

So when m<1m<1, dm(c)d_{m}(c) attains its maximum for some c>0c>0, and when m>1m>1, dm(c)d_{m}(c) attains its minimum for some c>0c>0. For m1m\geq 1, we denote

p(m)=14maxc0d1/m(c),p(m)=\frac{1}{4}\max_{c\geq 0}d_{1/m}(c),

and

q(m)=14minc0dm(c),q(m)=\frac{1}{4}\min_{c\geq 0}d_{m}(c),
F(m)=(p(m)q(m))2.F(m)=(\frac{p(m)}{q(m)})^{2}.

Clearly, limm1F(m)=1\lim_{m\to 1}F(m)=1. The rest of this subsection is devoted to the proof of the following quantitative estimate

Proposition 3.4.

For m[1,1.01)m\in[1,1.01) we have

0<F(m)<0.84.0<F^{\prime}(m)<0.84.

Let

G(m,c)=a(c)c(xc)x2egc(x)𝑑x3b(c)c(xc)egc(x)𝑑x,G(m,c)=a(c)\int_{c}^{\infty}(x-c)x^{2}e^{-g_{c}(x)}dx-3b(c)\int_{c}^{\infty}(x-c)e^{-g_{c}(x)}dx,

then G(1,c)=π2c(xc)(x23)ex2/2𝑑xG(1,c)=\sqrt{\frac{\pi}{2}}\int_{c}^{\infty}(x-c)(x^{2}-3)e^{-x^{2}/2}dx. It is not hard to see that G(1,c)=0G(1,c)=0 has exactly one solution c=c0c=c_{0} for c0c\geq 0, and we can use Mathematica to estimate c00.612003c_{0}\approx 0.612003. Also, we have:

cG\displaystyle\frac{\partial}{\partial c}G =\displaystyle= 3b(c)cegc𝑑xa(c)cx2egc𝑑x\displaystyle 3b(c)\int_{c}^{\infty}e^{-g_{c}}dx-a(c)\int_{c}^{\infty}x^{2}e^{-g_{c}}dx
+\displaystyle+ (1m)[c(xc)egcdx+a(c)c(xc)2x2egcdx\displaystyle(1-m)[\int_{c}^{\infty}(x-c)e^{-g_{c}}dx+a(c)\int_{c}^{\infty}(x-c)^{2}x^{2}e^{-g_{c}}dx
\displaystyle- 3c(xc)x2egcdxc(xc)egcdx3b(c)c(xc)2egcdx)]\displaystyle 3\int_{c}^{\infty}(x-c)x^{2}e^{-g_{c}}dx\int_{c}^{\infty}(x-c)e^{-g_{c}}dx-3b(c)\int_{c}^{\infty}(x-c)^{2}e^{-g_{c}}dx)]
=\displaystyle= A1(c)+(1m)A2(c).\displaystyle A_{1}(c)+(1-m)A_{2}(c).

Of course, A1(c)A_{1}(c) and A2(c)A_{2}(c) also depend on mm. Then

cG(1,c0)1.06\frac{\partial}{\partial c}G(1,c_{0})\approx 1.06

So by the implicit function theorem, the equation G(m,c)=0G(m,c)=0 has exactly one solution c(m)c(m) close to c0c_{0} for mm in a neighborhood of 11. Also, since

mgc(y)=12(yc)2,\frac{\partial}{\partial m}g_{c}(y)=\frac{1}{2}(y-c)^{2},

we have

ma(c)=12c(xc)2egc𝑑x,\frac{\partial}{\partial m}a(c)=-\frac{1}{2}\int_{c}^{\infty}(x-c)^{2}e^{-g_{c}}dx,

and

mb(c)=12c(xc)2x2egc𝑑x.\frac{\partial}{\partial m}b(c)=-\frac{1}{2}\int_{c}^{\infty}(x-c)^{2}x^{2}e^{-g_{c}}dx.

Therefore,

mG\displaystyle\frac{\partial}{\partial m}G =\displaystyle= 12c(xc)2egc𝑑xc(xc)x2egc𝑑x12a(c)c(xc)3egc𝑑x\displaystyle-\frac{1}{2}\int_{c}^{\infty}(x-c)^{2}e^{-g_{c}}dx\int_{c}^{\infty}(x-c)x^{2}e^{-g_{c}}dx-\frac{1}{2}a(c)\int_{c}^{\infty}(x-c)^{3}e^{-g_{c}}dx
+\displaystyle+ 32c(xc)2x2egc𝑑xc(xc)egc𝑑x+32b(c)c(xc)3egc𝑑x\displaystyle\frac{3}{2}\int_{c}^{\infty}(x-c)^{2}x^{2}e^{-g_{c}}dx\int_{c}^{\infty}(x-c)e^{-g_{c}}dx+\frac{3}{2}b(c)\int_{c}^{\infty}(x-c)^{3}e^{-g_{c}}dx

We can use Mathematica again to estimate mG(1,c0)1.557\frac{\partial}{\partial m}G(1,c_{0})\approx 1.557. So ddmc(1)1.47\frac{d}{dm}c(1)\approx 1.47.

When 1m<1.011\leq m<1.01 and c00.03cc0+0.03c_{0}-0.03\leq c\leq c_{0}+0.03, we have

mcegc𝑑x<0,\frac{\partial}{\partial m}\int_{c}^{\infty}e^{-g_{c}}dx<0,
mcx2egc𝑑x<0,\frac{\partial}{\partial m}\int_{c}^{\infty}x^{2}e^{-g_{c}}dx<0,
ccegc𝑑x<0,\frac{\partial}{\partial c}\int_{c}^{\infty}e^{-g_{c}}dx<0,

and

ccx2egc𝑑x<0.\frac{\partial}{\partial c}\int_{c}^{\infty}x^{2}e^{-g_{c}}dx<0.

Therefore, we can estimate

0.932<A1(c)<1.182.0.932<A_{1}(c)<1.182.

Similarly, we can estimate |A2(c)|<3.36|A_{2}(c)|<3.36. Hence we have

0.89<cG(m,c)<1.22,0.89<\frac{\partial}{\partial c}G(m,c)<1.22,

for 1m<1.011\leq m<1.01 and c00.03cc0+0.03c_{0}-0.03\leq c\leq c_{0}+0.03. In the same way, we can estimate

1.29<mG(m,c)<1.7921.29<\frac{\partial}{\partial m}G(m,c)<1.792

So for 1m<1.011\leq m<1.01 the equation G(m,c)=0G(m,c)=0 has exactly one solution c(m)c(m) satisfying

1.05<c(m)<2.02.1.05<c^{\prime}(m)<2.02.

When 0.99<m10.99<m\leq 1, we have similar conclusions. More precisely, we have 0.943<A1(c)<1.1970.943<A_{1}(c)<1.197 and |A2(c)|<4.21|A_{2}(c)|<4.21. So we have

0.9<cG(m,c)<1.24.0.9<\frac{\partial}{\partial c}G(m,c)<1.24.

Also,

1.32<mG(m,c)<1.84.1.32<\frac{\partial}{\partial m}G(m,c)<1.84.

Therefore

1.06<c(m)<2.051.06<c^{\prime}(m)<2.05

Now the argument above, we have for 1m<1.011\leq m<1.01, q(m)=14dm(c(m))q(m)=\frac{1}{4}d_{m}(c(m)) and p(m)=14d1/m(c(1m))p(m)=\frac{1}{4}d_{1/m}(c(\frac{1}{m})). So

q(m)=14(a(c(m)))4[ddmb(c(m))a(c(m))3ddma(c(m))b(c(m))].q^{\prime}(m)=\frac{1}{4(a(c(m)))^{4}}[\frac{d}{dm}b(c(m))a(c(m))-3\frac{d}{dm}a(c(m))b(c(m))].

Let η=ddmb(c(m))a(c(m))3ddma(c(m))b(c(m))\eta=\frac{d}{dm}b(c(m))a(c(m))-3\frac{d}{dm}a(c(m))b(c(m)). We have

η\displaystyle-\eta =\displaystyle= c12(xc)2x2egc𝑑xa(c(m))c(m)b(c)a(c)\displaystyle\int_{c}^{\infty}\frac{1}{2}(x-c)^{2}x^{2}e^{-g_{c}}dxa(c(m))-c^{\prime}(m)b^{\prime}(c)a(c)
3b(c)c12(xc)2egc𝑑x+3b(c)c(m)a(c)\displaystyle-3b(c)\int_{c}^{\infty}\frac{1}{2}(x-c)^{2}e^{-g_{c}}dx+3b(c)c^{\prime}(m)a^{\prime}(c)
=\displaystyle= 120egc𝑑xc(xc)2x2egc𝑑x\displaystyle\frac{1}{2}\int_{0}^{\infty}e^{-g_{c}}dx\int_{c}^{\infty}(x-c)^{2}x^{2}e^{-g_{c}}dx
320x2egc𝑑xc(xc)2egc𝑑x.\displaystyle-\frac{3}{2}\int_{0}^{\infty}x^{2}e^{-g_{c}}dx\int_{c}^{\infty}(x-c)^{2}e^{-g_{c}}dx.

Then η(1)0.318018\eta(1)\approx-0.318018. Since limm1a(c(m))=π2\lim_{m\to 1}a(c(m))=\sqrt{\frac{\pi}{2}} and

limm1q(m)=limm1p(m)=12π,\lim_{m\to 1}q(m)=\lim_{m\to 1}p(m)=\frac{1}{2\pi},

we get

F(1)0.81F^{\prime}(1)\approx 0.81

If 1m<1.011\leq m<1.01 and |cc0|<0.021|c-c_{0}|<0.021, we can estimate |ddmη(c(m))|<4|\frac{d}{dm}\eta(c(m))|<4, so we have 0.314>η>0.323-0.314>\eta>-0.323. Since ddma(c(m))<0\frac{d}{dm}a(c(m))<0, we have 1.251<a(c)<π21.251<a(c)<\sqrt{\frac{\pi}{2}}, so

0.132<4q(m)<0.127.-0.132<4q^{\prime}(m)<-0.127.

We can also estimate 4q(m)>0.6354q(m)>0.635 and 4p(m)<0.6384p(m)<0.638, so

p(m)q(m)<1.005\frac{p(m)}{q(m)}<1.005

Similarly if 0.99<m<10.99<m<1 and |cc0|<0.021|c-c_{0}|<0.021, since π2<a(c)<1.2554\sqrt{\frac{\pi}{2}}<a(c)<1.2554, we have

0.131<4p(1/m)<0.126.-0.131<4p^{\prime}(1/m)<-0.126.

Therefore, we have when m[1,1.01]m\in[1,1.01],

0<F(m)<0.84.0<F^{\prime}(m)<0.84.

This finishes the proof of Proposition 3.4.

3.2. full real line

Definition 3.5.

We say that a function g(t)g(t) is a (M1,M2)(M_{1},M_{2})-admissible function if g(t)g(t) is a convex function on \mathbb{R} such that g(t)g(t) and g(t)g(-t), t0t\geq 0, satisfy the conditions in theorem 3.1 with constants M1,M2M_{1},M_{2}.

Let g(t)g(t) be a (M1,M2)(M_{1},M_{2})-admissible function, in order to control d2da2λ(a)\frac{d^{2}}{da^{2}}\lambda(a) and d2dt2logf(x)\frac{d^{2}}{dt^{2}}\log f(x), we actually need to consider the following functions. Let

t¯(g)=teg(t)𝑑teg(t)𝑑t,\bar{t}(g)=\frac{\int_{-\infty}^{\infty}te^{-g(t)}dt}{\int_{-\infty}^{\infty}e^{-g(t)}dt},

be the center of mass. We define the following functions of gg:

a~(g)=eg(t)𝑑t,\tilde{a}(g)=\int_{-\infty}^{\infty}e^{-g(t)}dt,
b~(g)=(tt¯)2eg(t)𝑑t,\tilde{b}(g)=\int_{-\infty}^{\infty}(t-\bar{t})^{2}e^{-g(t)}dt,

and

d~(g)=b~(g)(a~(g))3.\tilde{d}(g)=\frac{\tilde{b}(g)}{(\tilde{a}(g))^{3}}.

Remark: Notice that d~(g)\tilde{d}(g) does not change if we dilate tt, namely d~(g(t))=d~(g(λt))\tilde{d}(g(t))=\tilde{d}(g(\lambda t)).

When a~(g)=1\tilde{a}(g)=1, we have a simpler formula

d~(g)=b~(g)=t2eg(t)𝑑tt¯2(g).\tilde{d}(g)=\tilde{b}(g)=\int_{-\infty}^{\infty}t^{2}e^{-g(t)}dt-\bar{t}^{2}(g).

Recall we have defined the function gm,c(y)g_{m,c}(y) right after lemma 3.3.

Definition 3.6.

For fixed c10,c20c_{1}\geq 0,c_{2}\geq 0, we denote by gm,c1,c2(y)g_{m,c_{1},c_{2}}(y) the function satisfying the following:

  • gm,c1,c2(y)=gm,c1(y)g_{m,c_{1},c_{2}}(y)=g_{m,c_{1}}(y) for y0y\geq 0;

  • gm,c1,c2(y)=gm,c2(y)g_{m,c_{1},c_{2}}(y)=g_{m,c_{2}}(-y) for y<0y<0.

We will need to apply operations that replace gm,c1,c2g_{m,c_{1},c_{2}} with gm,c1ε,c2+δg_{m,c_{1}-\varepsilon,c_{2}+\delta}. We will make sure that the total mass a~(g)\tilde{a}(g) does not change. Then in order to show that we change t2eg𝑑t(eg𝑑t)3\frac{\int_{-\infty}^{\infty}t^{2}e^{-g}dt}{(\int_{-\infty}^{\infty}e^{-g}dt)^{3}} in a direction we need, we only need to consider the ratio of the differences of the quantities a(g)a(g) and b(g)b(g) when we replace gm,cg_{m,c} with gm,c+εg_{m,c+\varepsilon}. And we can consider it from the infinitesimal point of view. It is not hard to see that

ddca(gm,c)\displaystyle\frac{d}{dc}a(g_{m,c}) =\displaystyle= (m1)e12c212mc2c(xc)e12mx2+(m1)cx𝑑x\displaystyle(m-1)e^{\frac{1}{2}c^{2}-\frac{1}{2}mc^{2}}\int_{c}^{\infty}(x-c)e^{-\frac{1}{2}mx^{2}+(m-1)cx}dx
=\displaystyle= (m1)e12c20xe12mx2cx𝑑x\displaystyle(m-1)e^{-\frac{1}{2}c^{2}}\int_{0}^{\infty}xe^{-\frac{1}{2}mx^{2}-cx}dx
=\displaystyle= (m1)m1e12c20xe12x2mcx𝑑x\displaystyle(m-1)m^{-1}e^{-\frac{1}{2}c^{2}}\int_{0}^{\infty}xe^{-\frac{1}{2}x^{2}-\sqrt{m}cx}dx

Similarly

ddcb(gm,c)\displaystyle\frac{d}{dc}b(g_{m,c}) =\displaystyle= (m1)e12c212mc2cx2(xc)e12mx2+(m1)cx𝑑x\displaystyle(m-1)e^{\frac{1}{2}c^{2}-\frac{1}{2}mc^{2}}\int_{c}^{\infty}x^{2}(x-c)e^{-\frac{1}{2}mx^{2}+(m-1)cx}dx
=\displaystyle= (m1)e12c20(x+c)2xe12mx2cx𝑑x\displaystyle(m-1)e^{-\frac{1}{2}c^{2}}\int_{0}^{\infty}(x+c)^{2}xe^{-\frac{1}{2}mx^{2}-cx}dx
=\displaystyle= (m1)m2e12c20(x+mc)2xe12x2mcx𝑑x\displaystyle(m-1)m^{-2}e^{-\frac{1}{2}c^{2}}\int_{0}^{\infty}(x+\sqrt{m}c)^{2}xe^{-\frac{1}{2}x^{2}-\sqrt{m}cx}dx

So we need to consider the ratio function

f1(c)=0x(x+c)2ex2/2cx𝑑x0xex2/2cx𝑑x,c0.f_{1}(c)=\frac{\int_{0}^{\infty}x(x+c)^{2}e^{-x^{2}/2-cx}dx}{\int_{0}^{\infty}xe^{-x^{2}/2-cx}dx},\quad c\geq 0.

Recall the complementary error function erfc(x)=2πxet2𝑑t{\operatorname{erfc}}(x)=\frac{2}{\sqrt{\pi}}\int_{x}^{\infty}e^{-t^{2}}dt Then we can calculate the derivative:

f1(c)=22π(c2+1)ec22erfc(c2)4c(2πcec22erfc(c2)2)2f_{1}^{\prime}(c)=\frac{2\sqrt{2\pi}\left(c^{2}+1\right)e^{\frac{c^{2}}{2}}\text{erfc}\left(\frac{c}{\sqrt{2}}\right)-4c}{\left(\sqrt{2\pi}ce^{\frac{c^{2}}{2}}\text{erfc}\left(\frac{c}{\sqrt{2}}\right)-2\right)^{2}}

Since erfc{\operatorname{erfc}} satisfies the inequality

(3.1) πex2erfc(x)>2x+x2+2.\sqrt{\pi}e^{x^{2}}{\operatorname{erfc}}(x)>\frac{2}{x+\sqrt{x^{2}+2}}.

It follows that

(3.2) f1(c)>0,f_{1}^{\prime}(c)>0,

for c0c\geq 0.

Similarly we can consider

f2(c)=0x(x+c)ex2/2cx𝑑x0xex2/2cx𝑑x.f_{2}(c)=\frac{\int_{0}^{\infty}x(x+c)e^{-x^{2}/2-cx}dx}{\int_{0}^{\infty}xe^{-x^{2}/2-cx}dx}.

We then have

f2(2c)=2[πe2c2erfc2(c)+2πcec2erfc(c)2](22πcec2erfc(c))2.f_{2}^{\prime}(\sqrt{2}c)=\frac{2[\pi e^{2c^{2}}{\operatorname{erfc}}^{2}(c)+2\sqrt{\pi}ce^{c^{2}}{\operatorname{erfc}}(c)-2]}{(2-2\sqrt{\pi}ce^{c^{2}}{\operatorname{erfc}}(c))^{2}}.

The numerator can be considered as a quadratic function x2+2cx2x^{2}+2cx-2 of x=πec2erfc(c)x=\sqrt{\pi}e^{c^{2}}{\operatorname{erfc}}(c). And the larger root of the quadratic function is c+c2+2-c+\sqrt{c^{2}+2}. Then by formula 3.1 again, we have

(3.3) f2(c)>0,f_{2}^{\prime}(c)>0,

for c0c\geq 0, which implies that the function f3(c)=b1(c)a(c)f_{3}(c)=\frac{b_{1}(c)}{a(c)}, where b1(c)=0xegm,c(x)𝑑xb_{1}(c)=\int_{0}^{\infty}xe^{-g_{m,c}(x)}dx satisfies the following.

Lemma 3.7.

When m<1m<1, we have

(3.4) f3(c)<0,f_{3}^{\prime}(c)<0,

for c0c\geq 0.

Proof.

Then

f3(c)=a(c)b1(c)b1(c)a(c)a2(c).f_{3}^{\prime}(c)=\frac{a(c)b_{1}^{\prime}(c)-b_{1}(c)a^{\prime}(c)}{a^{2}(c)}.

We have

b1(c)=(m1)m3/2e12c20(x+mc)xe12x2mcx𝑑x,\displaystyle b_{1}^{\prime}(c)=(m-1)m^{-3/2}e^{-\frac{1}{2}c^{2}}\int_{0}^{\infty}(x+\sqrt{m}c)xe^{-\frac{1}{2}x^{2}-\sqrt{m}cx}dx,

and

a(c)=(m1)m1e12c20xe12x2mcx𝑑x.a^{\prime}(c)=(m-1)m^{-1}e^{-\frac{1}{2}c^{2}}\int_{0}^{\infty}xe^{-\frac{1}{2}x^{2}-\sqrt{m}cx}dx.

So at 0, we have a(0)=π2ma(0)=\sqrt{\frac{\pi}{2m}}, b1(0)=π21mb_{1}(0)=\sqrt{\frac{\pi}{2}}\frac{1}{m}, a(0)=(m1)/ma^{\prime}(0)=(m-1)/m and b1(0)=π2m1m3/2b_{1}^{\prime}(0)=\sqrt{\frac{\pi}{2}}\frac{m-1}{m^{3/2}}. Therefore

f3(0){<0,m<1>0,m>1.f_{3}^{\prime}(0)\begin{cases}<0,&m<1\\ >0,&m>1\end{cases}.

Write

f3(c)\displaystyle f_{3}^{\prime}(c) =\displaystyle= a(c)a(c)[b1(c)a(c)b1(c)a(c)]\displaystyle\frac{a^{\prime}(c)}{a(c)}[\frac{b_{1}^{\prime}(c)}{a^{\prime}(c)}-\frac{b_{1}(c)}{a(c)}]
=\displaystyle= a(c)a(c)[1mf2(mc)f3(c)]\displaystyle\frac{a^{\prime}(c)}{a(c)}[\frac{1}{\sqrt{m}}f_{2}(\sqrt{m}c)-f_{3}(c)]

Then formula 3.3 implies that we always have b1(c)a(c)>b1(c)a(c)\frac{b_{1}^{\prime}(c)}{a^{\prime}(c)}>\frac{b_{1}(c)}{a(c)}, so f3(c)<0f_{3}^{\prime}(c)<0. ∎

Then we can control t¯\bar{t} with the following proposition:

Proposition 3.8.

Let g(t)g(t) be a (M1,M2)(M_{1},M_{2})-admissible function. We have

|t¯(g)|2πM1(11m),|\bar{t}(g)|\leq\sqrt{\frac{2}{\pi M_{1}}}(1-\frac{1}{\sqrt{m}}),

where m=M2M1m=\frac{M_{2}}{M_{1}}.

Proof.

By a dilation of tt, we can assume M1=1M_{1}=1 and M2=mM_{2}=m. Then we show that the extremal case is that g′′(t)=1g^{\prime\prime}(t)=1 for t0t\geq 0 and g′′(t)=mg^{\prime\prime}(t)=m for t<0t<0. To see this, we assume without loss of generality that t¯(g)>0\bar{t}(g)>0. Then we apply operation B on the part where t<0t<0 and apply operation A on the part where t>0t>0, both of these operations will increase t¯(g)\bar{t}(g). So we arrive at a function g1g_{1} satisfying

  • eg1(t)𝑑t=eg(t)𝑑t\int_{-\infty}^{\infty}e^{-g_{1}(t)}dt=\int_{-\infty}^{\infty}e^{-g(t)}dt and t¯(g)<t¯(g1)\bar{t}(g)<\bar{t}(g_{1}).

  • c1>0>c2\exists c_{1}>0>c_{2} such that

    g1′′(t)={m,0<t<c11,tc11,c2<t0m,tc2.g_{1}^{\prime\prime}(t)=\begin{cases}m,&0<t<c_{1}\\ 1,&t\geq c_{1}\\ 1,&c_{2}<t\leq 0\\ m,&t\leq c_{2}\end{cases}.

We will denote by fm,c1,c2f_{m,c_{1},c_{2}} a function satisfying the condition in the second item. Then we apply the following operation: we replace g1=fm,c1,c2g_{1}=f_{m,c_{1},c_{2}} with g2=fm,c1ε,c2+δg_{2}=f_{m,c_{1}-\varepsilon,c_{2}+\delta} for some positive ε,δ\varepsilon,\delta satisfying

eg1(t)𝑑t=eg2(t)𝑑t.\int_{-\infty}^{\infty}e^{-g_{1}(t)}dt=\int_{-\infty}^{\infty}e^{-g_{2}(t)}dt.

Clearly, this operation increases t¯(g)\bar{t}(g).

And we can repeat applying this operation until c1=0c_{1}=0 or c2=0c_{2}=0. In the second case where c2=0c_{2}=0, by formula 3.4, we can increase t¯(g)\bar{t}(g) by replacing the obtained function with the extremal case where g′′(t)=1g^{\prime\prime}(t)=1 for t0t\geq 0 and g′′(t)=mg^{\prime\prime}(t)=m for t<0t<0. In the first case where c1=0c_{1}=0, recalling that t¯=teg(t)𝑑teg(t)𝑑t\bar{t}=\frac{\int_{-\infty}^{\infty}te^{-g(t)}dt}{\int_{-\infty}^{\infty}e^{-g(t)}dt}, if we increase c2c_{2} we simultaneously decrease the denominator and increase the numerator so we also increase t¯(g)\bar{t}(g) by replacing the function with the extremal case. And t¯\bar{t} in this case is

11mπ2(1+1m)=2π(11m).\frac{1-\frac{1}{m}}{\sqrt{\frac{\pi}{2}}(1+\sqrt{\frac{1}{m}})}=\sqrt{\frac{2}{\pi}}(1-\frac{1}{\sqrt{m}}).

By symmetry, we will also have a lower bound. Then notice that a dilation of tt makes the same dilation to t¯\bar{t}, and we get the conclusion. ∎

Theorem 3.9.

Let g(t)g(t) be a (M1,M2)(M_{1},M_{2})-admissible function. Let m=M2M1m=\frac{M_{2}}{M_{1}}. Then there exist c1c20c_{1}\geq c_{2}\geq 0 such that

d~(g)d~(gm,c1,c2),\tilde{d}(g)\geq\tilde{d}(g_{m,c_{1},c_{2}}),

and

d~(g)<p(m).\tilde{d}(g)<p(m).
Proof.

We can assume without loss of generality that t¯0\bar{t}\geq 0. Recall that in the proof of theorem 3.1 we have defined operations A and B to change 0t2eg(t)𝑑t\int_{0}^{\infty}t^{2}e^{-g(t)}dt while keeping 0eg(t)𝑑t\int_{0}^{\infty}e^{-g(t)}dt. The idea here is similar. For the lower bound, we first notice that if we get a function g1g_{1} satisfying 0eg1(t)𝑑t=0eg(t)𝑑t=1\int_{0}^{\infty}e^{-g_{1}(t)}dt=\int_{0}^{\infty}e^{-g(t)}dt=1 and (tt¯(g))2eg1(t)𝑑t(tt¯(g))2eg(t)𝑑t\int_{-\infty}^{\infty}(t-\bar{t}(g))^{2}e^{-g_{1}(t)}dt\leq\int_{-\infty}^{\infty}(t-\bar{t}(g))^{2}e^{-g(t)}dt, then

(tt¯(g1))2eg1(t)𝑑t(tt¯(g))2eg(t)𝑑t,\int_{-\infty}^{\infty}(t-\bar{t}(g_{1}))^{2}e^{-g_{1}(t)}dt\leq\int_{-\infty}^{\infty}(t-\bar{t}(g))^{2}e^{-g(t)}dt,

since

(tt¯(g1))2eg1(t)𝑑t(ta)2eg1(t)𝑑t,\int_{-\infty}^{\infty}(t-\bar{t}(g_{1}))^{2}e^{-g_{1}(t)}dt\leq\int_{-\infty}^{\infty}(t-a)^{2}e^{-g_{1}(t)}dt,

for any aa. Then we first apply operation B to the part t0t\leq 0. By abuse of notations, we will still denote by gg the new function after this operation. Clearly, this will decrease b~(g)\tilde{b}(g). And since such an operation B will increase t¯(g)\bar{t}(g), we can continue this operation until g(t)g(-t) for t0t\geq 0 is a dilation of gm,cg_{m,c} for some cc. Then we apply operation B to the part t0t\geq 0. To show that we can also continue applying operation B, we denote by

σ(g)=inf{x>0|g′′(t)=M2fortx}.\sigma(g)=\inf\{x>0|g^{\prime\prime}(t)=M_{2}\quad\textit{for}\quad t\geq x\}.

Then since t¯(g)\bar{t}(g) is the center of mass, we must have 2t¯(g)<σ(g)2\bar{t}(g)<\sigma(g). Therefore small operation B on the part t0t\geq 0 will always decrease b~(g)\tilde{b}(g). Let g2g_{2} be the limit function, then g2g_{2} must be a dilation of gm,c1,c2g_{m,c_{1},c_{2}} for some c1,c20c_{1},c_{2}\geq 0.

For the upper bound, we first make a dilation to make M2=1M_{2}=1. Assume again that t¯0\bar{t}\geq 0. We first apply operation A on g(t)g(t) for t0t\leq 0. Notice that this will increase t2eg(t)𝑑t\int_{-\infty}^{\infty}t^{2}e^{-g(t)}dt and decrease t¯\bar{t}. And as long as the new t¯0\bar{t}\geq 0, t¯2\bar{t}^{2} decreases. Therefore such an operation will increase d~(g)\tilde{d}(g). Then we keep doing this operation A until it cannot be continued. There are two possibilities:

  • (1)

    c0\exists c\leq 0 such that g(t)g(-t), t0t\geq 0, is g1/m,cg_{1/m,c};

  • (2)

    t¯=0\bar{t}=0.

In the second case, we can apply operation A for the two sides t0t\geq 0 and t0t\leq 0 simultaneously while keeping t¯=0\bar{t}=0. As long as neither side is some g1/m,cg_{1/m,c}, we will continue this operation. This operation will end until one side, which we can assume is the part t0t\leq 0, is some g1/m,cg_{1/m,c}.

Then we keep applying operation A on the part t0t\geq 0 until the limit function g3g_{3} is some g1/m,cg_{1/m,c^{\prime}}. We denote the new function by g4g_{4}. Notice that during this process, both t2eg(t)𝑑t\int t^{2}e^{-g(t)}dt and t¯(g)\bar{t}(g) increases. It is not guaranteed that g~\tilde{g} increases.

For tt\in\mathbb{R}, g4=g1/m,c,cg_{4}=g_{1/m,c^{\prime},c}. Then t¯>0\bar{t}>0 imples that c>cc>c^{\prime}. Then we apply a new operation D by replacing g1/m,c,cg_{1/m,c^{\prime},c} with g1/m,c+ε,cδg_{1/m,c^{\prime}+\varepsilon,c-\delta} for some small positve ε,δ\varepsilon,\delta satisfying

a~(g1/m,c+ε,cδ)=a~(g1/m,c,c).\tilde{a}(g_{1/m,c^{\prime}+\varepsilon,c-\delta})=\tilde{a}(g_{1/m,c^{\prime},c}).

Then, formula 3.2 implies that when ε\varepsilon is small enough, we have

t2eg1/m,c,c(t)𝑑t<t2eg1/m,c+ε,cδ𝑑t.\int_{-\infty}^{\infty}t^{2}e^{-g_{1/m,c^{\prime},c}(t)}dt<\int_{-\infty}^{\infty}t^{2}e^{-g_{1/m,c^{\prime}+\varepsilon,c-\delta}}dt.

So we can continue applying operation D until we get c=c=c3c=c^{\prime}=c_{3}. Notice now that t¯(g1/m,c3,c3)=0\bar{t}(g_{1/m,c_{3},c_{3}})=0 and we have t2eg(t)𝑑t<t2eg1/m,c3,c3\int t^{2}e^{-g(t)}dt<\int_{-\infty}^{\infty}t^{2}e^{-g_{1/m,c_{3},c_{3}}} and eg(t)𝑑t=eg1/m,c3,c3\int e^{-g(t)}dt=\int_{-\infty}^{\infty}e^{-g_{1/m,c_{3},c_{3}}}, therefore, g~<p(m)\tilde{g}<p(m).

Remark: We expect that d~(g)>q(m)\tilde{d}(g)>q(m) also holds. However, we are not able to show this due to the complexity of the calculations.

Corollary 3.10.

Let g(t)g(t) be a (M1,M2)(M_{1},M_{2})-admissible function. Let m=M2M1m=\frac{M_{2}}{M_{1}}. Then we have

d~(g)q(m)4π2(11m)2(1+1m)2.\tilde{d}(g)\geq q(m)-\frac{4}{\pi^{2}}\frac{(1-\sqrt{\frac{1}{m}})^{2}}{(1+\sqrt{\frac{1}{m}})^{2}}.
Proof.

Write A=0eg(y)𝑑yA=\int_{0}^{\infty}e^{-g(y)}dy, B=0eg(y)𝑑yB=\int_{0}^{\infty}e^{-g(y)}dy. We have

d~(g)=t2eg(t)𝑑t(eg(t)𝑑t)3t¯2(g)(eg(t)𝑑t)2\tilde{d}(g)=\frac{\int_{-\infty}^{\infty}t^{2}e^{-g(t)}dt}{(\int_{-\infty}^{\infty}e^{-g(t)}dt)^{3}}-\frac{\bar{t}^{2}(g)}{(\int_{-\infty}^{\infty}e^{-g(t)}dt)^{2}}

Then since

A3+B3(A+B)314,\frac{A^{3}+B^{3}}{(A+B)^{3}}\geq\frac{1}{4},

and

|t¯|2πM1(11m),|\bar{t}|\leq\sqrt{\frac{2}{\pi M_{1}}}(1-\sqrt{\frac{1}{m}}),

we have the conclusion. ∎

This corollary does not give us a priori lower bound for d~(g)\tilde{d}(g). So we need to prove the following:

Corollary 3.11.

Let g(t)g(t) be a (M1,M2)(M_{1},M_{2})-admissible function. Then we have

d~(g)112.\tilde{d}(g)\geq\frac{1}{12}.
Proof.

One can repeat the proof of theorem 3.1, and then relax M10M_{1}\to 0 and M2M_{2}\to\infty. In order for the total mass to be 1, we arrive at the limit function

g(t)={1,c1<t<c20,t>c2ort<c1,g_{\infty}(t)=\begin{cases}1,&c_{1}<t<c_{2}\\ 0,&t>c_{2}\quad\textit{or}\quad t<c_{1}\end{cases},

where c1<0<c2c_{1}<0<c_{2} and c2c1=1c_{2}-c_{1}=1. Then clearly, d~(g)=112\tilde{d}(g_{\infty})=\frac{1}{12}. ∎

We end this subsection with the following proposition:

Proposition 3.12.

For fixed m>1m>1, d~(gm,c1,c2)\tilde{d}(g_{m,c_{1},c_{2}}) considered as a function of c10,c20c_{1}\geq 0,c_{2}\geq 0 attains its minimum.

Proof.

We denote by dm=infc10,c20d~(gm,c1,c2)d_{m}=\inf_{c_{1}\geq 0,c_{2}\geq 0}\tilde{d}(g_{m,c_{1},c_{2}}). We first notice that for cc large enough, the three integrals cegm,c(t)𝑑t\int_{c}^{\infty}e^{-g_{m,c}(t)}dt, ctegm,c(t)𝑑t\int_{c}^{\infty}te^{-g_{m,c}(t)}dt and ct2egm,c(t)𝑑t\int_{c}^{\infty}t^{2}e^{-g_{m,c}(t)}dt are all very small. Therefore, for c1c_{1} and c2c_{2} large engough, d~(gm,c1,c2)\tilde{d}(g_{m,c_{1},c_{2}}) is close to 12π\frac{1}{2\pi}. Since it is clear that

dmq(m)<12π,d_{m}\leq q(m)<\frac{1}{2\pi},

for m>1m>1, we have that for any ε1>0\varepsilon_{1}>0, N1>0\exists N_{1}>0 such that when c1N1c_{1}\geq N_{1} and c2N1c_{2}\geq N_{1}, we have

d~(gm,c1,c2)>infd~(gm,c1,c2)+ε1.\tilde{d}(g_{m,c_{1},c_{2}})>\inf\tilde{d}(g_{m,c_{1},c_{2}})+\varepsilon_{1}.

We claim that N2>0\exists N_{2}>0 such that for N1c20N_{1}\geq c_{2}\geq 0, for c1>N2c_{1}>N_{2} we have

d~(gm,c1,c2)>d~(gm,c1ε,c2+δ),\tilde{d}(g_{m,c_{1},c_{2}})>\tilde{d}(g_{m,c_{1}-\varepsilon,c_{2}+\delta}),

for some ε>0,δ>0\varepsilon>0,\delta>0. By symmetry, the same holds if we switch c1c_{1} and c2c_{2}. It will then follow that d~(gm,c1,c2)\tilde{d}(g_{m,c_{1},c_{2}}) attains its minimum. To see that the claim holds one only need to notice that if g1(t)>g(t)g_{1}(t)>g(t) for t>l>0t>l>0, then

tlt2(eg(t)eg1(t))𝑑ttl(eg(t)eg1(t))𝑑t>l2.\frac{\int_{t\geq l}t^{2}(e^{-g(t)}-e^{-g_{1}(t)})dt}{\int_{t\geq l}(e^{-g(t)}-e^{-g_{1}(t)})dt}>l^{2}.

Now we apply an operation by replacing gm,c1,c2g_{m,c_{1},c_{2}} with gm,c1ε,c2+δg_{m,c_{1}-\varepsilon,c_{2}+\delta}, where δ>0\delta>0 is chosen according to ε>0\varepsilon>0 so that the total mass

egm,c1ε,c2+δ𝑑t=egm,c1,c2𝑑t.\int_{-\infty}^{\infty}e^{-g_{m,c_{1}-\varepsilon,c_{2}+\delta}}dt=\int_{-\infty}^{\infty}e^{-g_{m,c_{1},c_{2}}}dt.

Since t¯(g)\bar{t}(g) is bounded in terms of mm, and

0t2(egm,c2+δegm,c2)𝑑t0(egm,c2+δegm,c2)𝑑t\frac{\int_{0}^{\infty}t^{2}(e^{-g_{m,c_{2}+\delta}}-e^{-g_{m,c_{2}}})dt}{\int_{0}^{\infty}(e^{-g_{m,c_{2}+\delta}}-e^{-g_{m,c_{2}}})dt}

is bounded in terms of mm and N1N_{1}, it is then clear that when N2N_{2} is large enough, we have

d~(gm,c1,c2)>d~(gm,c1ε,c2+δ).\tilde{d}(g_{m,c_{1},c_{2}})>\tilde{d}(g_{m,c_{1}-\varepsilon,c_{2}+\delta}).

4. Refined estimates

In order to apply our estimates on d~(g)\tilde{d}(g), we first notice that

d2da2λ(a)\displaystyle\frac{d^{2}}{da^{2}}\lambda(a) =\displaystyle= 0(logxt~a)2dxha(x)\displaystyle\int_{0}^{\infty}\frac{(\log x-\tilde{t}_{a})^{2}dx}{h_{a}(x)}
=\displaystyle= xa+1ha(xa+1)(tt~a)2eg(t)𝑑t.\displaystyle\frac{x_{a+1}}{h_{a}(x_{a+1})}\int_{-\infty}^{\infty}(t-\tilde{t}_{a})^{2}e^{-g(t)}dt.

We have g(t)g(t) a convex function satisfying g(ta+1)=g(ta+1)=0g(t_{a+1})=g^{\prime}(t_{a+1})=0 and g′′(t)=d2dt2logf(x)g^{\prime\prime}(t)=\frac{d^{2}}{dt^{2}}\log f(x). Also

t~a=t¯(g),\tilde{t}_{a}=\bar{t}(g),

in the notations of the previous subsection. Also since xa+1ha(xa+1)eg(t)𝑑t=1\frac{x_{a+1}}{h_{a}(x_{a+1})}\int_{-\infty}^{\infty}e^{-g(t)}dt=1, we have the total mass

eg(t)𝑑t=ha(xa+1)xa+1.\int_{-\infty}^{\infty}e^{-g(t)}dt=\frac{h_{a}(x_{a+1})}{x_{a+1}}.

Since

dadta=d2dt2logf(xa),\frac{da}{dt_{a}}=\frac{d^{2}}{dt^{2}}\log f(x_{a}),

we have ta+1=ta+O(1a)t_{a+1}=t_{a}+O(\frac{1}{a}), therefore xa+1=xa+O(1)x_{a+1}=x_{a}+O(1). We have d2dt2logha(x)=O(x)\frac{d^{2}}{dt^{2}}\log h_{a}(x)=O(x) and ddtlogha(xa)=0\frac{d}{dt}\log h_{a}(x_{a})=0. So

logha(xa+1)=logha(xa)+O(1a).\log h_{a}(x_{a+1})=\log h_{a}(x_{a})+O(\frac{1}{a}).

Therefore

(4.1) eg(t)𝑑t=ha(xa+1)xa+1=(1+O(1a))ha(xa)xa.\int_{-\infty}^{\infty}e^{-g(t)}dt=\frac{h_{a}(x_{a+1})}{x_{a+1}}=(1+O(\frac{1}{a}))\frac{h_{a}(x_{a})}{x_{a}}.

For d2dt2logf(x)\frac{d^{2}}{dt^{2}}\log f(x), we have

d2dt2logf(xa)=d2dt2logha(xa)=(ia)2τxa(i),\frac{d^{2}}{dt^{2}}\log f(x_{a})=\frac{d^{2}}{dt^{2}}\log h_{a}(x_{a})=\sum(i-a)^{2}\tau_{x_{a}}(i),

where

τxa(i)\displaystyle\tau_{x_{a}}(i) =\displaystyle= 1ha(xa)cicaxaia\displaystyle\frac{1}{h_{a}(x_{a})}\frac{c_{i}}{c_{a}}x_{a}^{i-a}
=\displaystyle= 1ha(xa)cnxacaxanxaaeG(i),\displaystyle\frac{1}{h_{a}(x_{a})}\frac{c_{n_{x_{a}}}}{c_{a}}x_{a}^{n_{x_{a}}-a}e^{-G(i)},

where G(i)G(i) is a convex function satisfying G(nxa)=G(nxa)=0G(n_{x_{a}})=G^{\prime}(n_{x_{a}})=0 and G′′(i)=d2di2λ(i)G^{\prime\prime}(i)=\frac{d^{2}}{di^{2}}\lambda(i). We will denote by Δa=cnxacaxanxaa\Delta_{a}=\frac{c_{n_{x_{a}}}}{c_{a}}x_{a}^{n_{x_{a}}-a}. Since d2da2λ(a)=O(1a)\frac{d^{2}}{da^{2}}\lambda(a)=O(\frac{1}{a}), we can use integral to estimate the summations τxa(i)\sum\tau_{x_{a}}(i) and (ia)2τxa(i)\sum(i-a)^{2}\tau_{x_{a}}(i) with relative error of size O(1a)O(\frac{1}{\sqrt{a}}). Since τxa(i)=1\sum\tau_{x_{a}}(i)=1, we have

(4.2) i0eG(i)𝑑i=(1+O(1a))ha(xa)Δa.\int_{i\geq 0}e^{-G(i)}di=(1+O(\frac{1}{\sqrt{a}}))\frac{h_{a}(x_{a})}{\Delta_{a}}.

Now we are ready to prove the following theorem.

Theorem 4.1.

We have the following:

(4.3) limaha(xa)a=2π,\lim_{a\to\infty}\frac{h_{a}(x_{a})}{\sqrt{a}}=\sqrt{2\pi},
(4.4) limaad2da2λ(a)=1,\lim_{a\to\infty}a\frac{d^{2}}{da^{2}}\lambda(a)=1,

and

(4.5) limxd2dt2logf(x)x=1\lim_{x\to\infty}\frac{\frac{d^{2}}{dt^{2}}\log f(x)}{x}=1
Proof.

Let A=lim supaha(xa)aA=\limsup_{a\to\infty}\frac{h_{a}(x_{a})}{\sqrt{a}}, B=lim infaha(xa)aB=\liminf_{a\to\infty}\frac{h_{a}(x_{a})}{\sqrt{a}}. We first assume that A>BA>B and draw a contradiction. For all ε>0\varepsilon>0, there exists an α\alpha such that for aαa\geq\alpha, we have

Bε<ha(xa)a<A+ε.B-\varepsilon<\frac{h_{a}(x_{a})}{\sqrt{a}}<A+\varepsilon.

Denote ρ(a)=logha(xa)a\rho(a)=\log\frac{h_{a}(x_{a})}{\sqrt{a}}. Let bb be a local minimum point of ρ\rho such that ρ(b)<log(B+ε)\rho(b)<\log(B+\varepsilon). Then we have ρ(b)=0\rho^{\prime}(b)=0 and ρ′′(b)0\rho^{\prime\prime}(b)\geq 0. We can calculate

ρ(a)=f(xa)f(xa)dxada+dλ(a)dataaxadxada12a=dλ(a)data12a,\rho^{\prime}(a)=\frac{f^{\prime}(x_{a})}{f(x_{a})}\frac{dx_{a}}{da}+\frac{d\lambda(a)}{da}-t_{a}-\frac{a}{x_{a}}\frac{dx_{a}}{da}-\frac{1}{2a}=\frac{d\lambda(a)}{da}-t_{a}-\frac{1}{2a},

since axa=f(xa)f(xa)\frac{a}{x_{a}}=\frac{f^{\prime}(x_{a})}{f(x_{a})}. Therefore,

ρ′′(a)=d2da2λ(a)1d2dt2logf(xa)+12a2\rho^{\prime\prime}(a)=\frac{d^{2}}{da^{2}}\lambda(a)-\frac{1}{\frac{d^{2}}{dt^{2}}\log f(x_{a})}+\frac{1}{2a^{2}}

So we have t~btb=12b\tilde{t}_{b}-t_{b}=\frac{1}{2b} and at a=ba=b,

d2da2λ(a)1d2dt2logf(xa)12a2.\frac{d^{2}}{da^{2}}\lambda(a)-\frac{1}{\frac{d^{2}}{dt^{2}}\log f(x_{a})}\geq-\frac{1}{2a^{2}}.

Since nx~a=an_{\tilde{x}_{a}}=a, we have nxa=a+O(1)n_{x_{a}}=a+O(1) at a=ba=b. So

Δb=eO(1/b).\Delta_{b}=e^{O(1/b)}.

Suppose we have positve numbers M1,M2,M1,M2M_{1},M_{2},M_{1}^{\prime},M_{2}^{\prime} such that

M1\displaystyle M_{1} <\displaystyle< ad2da2λ(a)<M2\displaystyle a\frac{d^{2}}{da^{2}}\lambda(a)<M_{2}
M1\displaystyle M_{1}^{\prime} <\displaystyle< d2dt2logf(x)x<M2\displaystyle\frac{\frac{d^{2}}{dt^{2}}\log f(x)}{x}<M_{2}^{\prime}

Let m=M2M1m=\frac{M_{2}}{M_{1}} and m=M2M1m^{\prime}=\frac{M^{\prime}_{2}}{M^{\prime}_{1}} Let

q~(m)=maxc10,c20d~(gm,c1,c2),\tilde{q}(m)=\max_{c_{1}\geq 0,c_{2}\geq 0}\tilde{d}(g_{m,c_{1},c_{2}}),

whose existence is guanrenteed by proposition 3.12. For simplify, we write p=p(m)p=p(m), p=p(m)p^{\prime}=p(m^{\prime}), q=q~(m)q=\tilde{q}(m) and q=q~(m)q^{\prime}=\tilde{q}(m^{\prime}). Then

d2da2λ(a)<p(B+ε)2a(1+ε),\displaystyle\frac{d^{2}}{da^{2}}\lambda(a)<\frac{p(B+\varepsilon)^{2}}{a}(1+\varepsilon),

and

d2dt2logf(xa)<pa(B+ε)2(1+ε),\frac{d^{2}}{dt^{2}}\log f(x_{a})<p^{\prime}a(B+\varepsilon)^{2}(1+\varepsilon),

for a=ba=b large enough. Therefore at a=ba=b, we have pp(B+ε)4(1+ε)2>1O(1b)pp^{\prime}(B+\varepsilon)^{4}(1+\varepsilon)^{2}>1-O(\frac{1}{b}). Since ε\varepsilon is arbitrary, we get that

B(pp)1/4.B\geq(pp^{\prime})^{-1/4}.

We can repeat this argument for a local maximum to get that

A(qq)1/4.A\leq(qq^{\prime})^{-1/4}.

So we get that for aa large enough, we have

(pp)1/4ε<ha(xa)a<(qq)1/4+ε.(pp^{\prime})^{-1/4}-\varepsilon<\frac{h_{a}(x_{a})}{\sqrt{a}}<(qq^{\prime})^{-1/4}+\varepsilon.

Then by theorem 3.9 and formula 4.1, we have

(4.6) q((pp)1/4ε)2<ad2da2λ(a)<p((qq)1/4+ε)2,q^{\prime}((pp^{\prime})^{-1/4}-\varepsilon)^{2}<a\frac{d^{2}}{da^{2}}\lambda(a)<p^{\prime}((qq^{\prime})^{-1/4}+\varepsilon)^{2},

for aa large enough. And by theorem 3.9 and formula 4.2, we have

(4.7) 1Δa2q((pp)1/4ε)2<d2dt2logf(xa)xa<1Δa2p((qq)1/4+ε)2,\frac{1}{\Delta_{a}^{2}}q((pp^{\prime})^{-1/4}-\varepsilon)^{2}<\frac{\frac{d^{2}}{dt^{2}}\log f(x_{a})}{x_{a}}<\frac{1}{\Delta_{a}^{2}}p((qq^{\prime})^{-1/4}+\varepsilon)^{2},

for aa large enough. Since Δa1\Delta_{a}\geq 1, we have

d2dt2logf(xa)xa<p((qq)1/4+ε)2.\frac{\frac{d^{2}}{dt^{2}}\log f(x_{a})}{x_{a}}<p((qq^{\prime})^{-1/4}+\varepsilon)^{2}.

To better control the lower bound, we notice that by formula 4.6, we have

i0eG(i)𝑑i>(1+O(1a))2πγ,\int_{i\geq 0}e^{-G(i)}di>(1+O(\frac{1}{a}))\sqrt{\frac{2\pi}{\gamma}},

where γ=p((qq)1/4+ε)2\gamma=p^{\prime}((qq^{\prime})^{-1/4}+\varepsilon)^{2}. Therefore we get

(4.8) 2πqγ<d2dt2logf(xa)xa<p((qq)1/4+ε)2.\frac{2\pi q}{\gamma}<\frac{\frac{d^{2}}{dt^{2}}\log f(x_{a})}{x_{a}}<p((qq^{\prime})^{-1/4}+\varepsilon)^{2}.

Let M2=M2=2106,M1=M1=1/5000M_{2}=M_{2}^{\prime}=2*10^{6},M_{1}=M_{1}^{\prime}=1/5000, namely m=m=1010m=m^{\prime}=10^{10}, then by theorem 3.1 we can let p=p=2p=p^{\prime}=2 and q=q=112q=q^{\prime}=\frac{1}{12} and start a process of iteration as follows. We let

m¯=pqppqq+ε,\bar{m}=\frac{p^{\prime}}{q^{\prime}}\sqrt{\frac{pp^{\prime}}{qq^{\prime}}}+\varepsilon,
m¯=pp2πq2q+ε\bar{m}^{\prime}=\frac{pp^{\prime}}{2\pi q^{2}q^{\prime}}+\varepsilon

Then let

p¯\displaystyle\bar{p} =\displaystyle= p(m¯)\displaystyle p(\bar{m})
p¯\displaystyle\bar{p}^{\prime} =\displaystyle= p(m¯)\displaystyle p(\bar{m}^{\prime})
q¯\displaystyle\bar{q} =\displaystyle= q~(m¯)\displaystyle\tilde{q}(\bar{m})
q¯\displaystyle\bar{q}^{\prime} =\displaystyle= q~(m¯)\displaystyle\tilde{q}(\bar{m})

Then we redefine p=p¯p=\bar{p}, p=p¯p^{\prime}=\bar{p}^{\prime}, q=q¯q=\bar{q} and q=q¯q^{\prime}=\bar{q}^{\prime} and repeat the preceding arguments to get new bounds for ha(xa)a\frac{h_{a}(x_{a})}{\sqrt{a}}, ad2da2λ(a)a\frac{d^{2}}{da^{2}}\lambda(a) and d2dt2logf(x)x\frac{\frac{d^{2}}{dt^{2}}\log f(x)}{x} and so on. We use the software Mathematica to compute p(m)p(m) and q~(m)\tilde{q}(m) and iterate the process 67 times to reach the point where 1<m<m<1.011<m<m^{\prime}<1.01, 0.99<M1<M1<M2<M2<1.010.99<M_{1}^{\prime}<M1<M_{2}<M_{2}^{\prime}<1.01 and 0.158818<q<q<p<p<0.1594930.158818<q^{\prime}<q<p<p^{\prime}<0.159493. In order to use proposition 3.4 we now use the lower bound in formula 4.7 for d2dt2logf(xa)xa\frac{\frac{d^{2}}{dt^{2}}\log f(x_{a})}{x_{a}}. Since

anxa=t¯(G)+O(1aa),a-n_{x_{a}}=\bar{t}(G)+O(\frac{1}{\sqrt{a}}\sqrt{a}),

by proposition 3.8 we have

|anxa|2<2.1aπ(11m)2.|a-n_{x_{a}}|^{2}<\frac{2.1a}{\pi}(1-\frac{1}{\sqrt{m}})^{2}.

Therefore, since λ(nxa)=logxa\lambda^{\prime}(n_{x_{a}})=\log x_{a},

1Δa<e122.2π(11m)2.1\leq\Delta_{a}<e^{\frac{1}{2}\frac{2.2}{\pi}(1-\frac{1}{\sqrt{m}})^{2}}.

Now we let M1=M1M_{1}=M_{1}^{\prime} and M2=M2M_{2}=M_{2}^{\prime}. We let

p=p=p(m)p=p^{\prime}=p(m)

, and let

q=q=q(m)4π2(11m)2(1+1m)2,q=q^{\prime}=q(m)-\frac{4}{\pi^{2}}\frac{(1-\sqrt{\frac{1}{m}})^{2}}{(1+\sqrt{\frac{1}{m}})^{2}},

and repeat arguments on local extremal points of ha(xa)a\frac{h_{a}(x_{a})}{\sqrt{a}} to get

B1p,A1q,B\geq\frac{1}{\sqrt{p}},\quad A\leq\frac{1}{\sqrt{q}},

and

(4.9) q(p1/2ε)2<ad2da2λ(a)<p(q1/2+ε)2.q(p^{-1/2}-\varepsilon)^{2}<a\frac{d^{2}}{da^{2}}\lambda(a)<p(q^{-1/2}+\varepsilon)^{2}.

Then we have

(4.10) e1.1π(11m)2q(p1/2ε)2<d2dt2logf(xa)xa<p(q1/2+ε)2.e^{-\frac{1.1}{\pi}(1-\frac{1}{\sqrt{m}})^{2}}q(p^{-1/2}-\varepsilon)^{2}<\frac{\frac{d^{2}}{dt^{2}}\log f(x_{a})}{x_{a}}<p(q^{-1/2}+\varepsilon)^{2}.

Then we let

m¯=m¯=p2q2e1.1π(11m)2+ε.\bar{m}=\bar{m}^{\prime}=\frac{p^{2}}{q^{2}}e^{\frac{1.1}{\pi}(1-\frac{1}{\sqrt{m}})^{2}}+\varepsilon.

Plugging in the formulas for pp and qq, we get

m¯=F(m)H2(m)Q(m)+ε,\bar{m}=F(m)H^{-2}(m)Q(m)+\varepsilon,

where H(m)=14q(m)π2(1α)2(1+α)2H(m)=1-\frac{4}{q(m)\pi^{2}}\frac{(1-\alpha)^{2}}{(1+\alpha)^{2}} and Q(m)=e1.1π(1α)2Q(m)=e^{\frac{1.1}{\pi}(1-\alpha)^{2}}. Denote by I(m)=F(m)H2(m)Q(m)I(m)=F(m)H^{-2}(m)Q(m). Clearly I(1)=1I(1)=1, so in order to show that the iteration will force m1m\to 1, we only need to show that I(m)<1I^{\prime}(m)<1 for 1m<1.011\leq m<1.01. We have shown that 0<F(m)<0.840<F^{\prime}(m)<0.84. When 1m<1.011\leq m<1.01, 0.99998H(m)10.99998\leq H(m)\leq 1. We have

H(m)=4(1α)2[(1+α)2qq(1+α)α3]π2q2(1+α)44(1α)α39π2(1+α)2.H^{\prime}(m)=\frac{4(1-\alpha)^{2}[(1+\alpha)^{2}q^{\prime}-q(1+\alpha)\alpha^{3}]}{\pi^{2}q^{2}(1+\alpha)^{4}}-\frac{4(1-\alpha)\alpha^{3}}{9\pi^{2}(1+\alpha)^{2}}.

So

0.000067<H(m)0.-0.000067<H^{\prime}(m)\leq 0.

Therefore

0(H1(m))<0.00007.0\leq(H^{-1}(m))^{\prime}<0.00007.

When 1m<1.011\leq m<1.01, Q(m)<1.00001Q(m)<1.00001 and

Q(m)=1.1π(1α)α3e1.1π(1α)2<0.0018.Q^{\prime}(m)=\frac{1.1}{\pi}(1-\alpha)\alpha^{3}e^{\frac{1.1}{\pi}(1-\alpha)^{2}}<0.0018.

So, when 1m<1.011\leq m<1.01,

0<I(m)<0.86.0<I^{\prime}(m)^{\prime}<0.86.

So this iteration will indeed end up with the limit situation, namely p=q=12πp=q=\frac{1}{2\pi} and M1=M2=1M_{1}=M_{2}=1. So this gives a contradiction. This proves that we must have A=BA=B.

For all ε>0\varepsilon>0, there exists NN such that for aNa\geq N, we have

Aε<ha(xa)a<A+ε.A-\varepsilon<\frac{h_{a}(x_{a})}{\sqrt{a}}<A+\varepsilon.

Then if ρ(a)\rho(a) is not monotone for aa large enough, then the same argument as above actually proves (4.3), (4.4) and (4.5). If ρ(a)\rho(a) is monotone, we assume, for example, that ρ(a)=t~ata12a0\rho^{\prime}(a)=\tilde{t}_{a}-t_{a}-\frac{1}{2a}\geq 0 for aa large enough. Then a0ρ(a)𝑑a<\int_{a_{0}}^{\infty}\rho^{\prime}(a)da<\infty, so the asymptotic density of the set ET={aT|ρ(a)1a}E_{T}=\{a\leq T|\rho^{\prime}(a)\geq\frac{1}{a}\}

limTλ(ET)T=0,\lim_{T\to\infty}\frac{\lambda(E_{T})}{T}=0,

where λ\lambda is the Lebesgue measure. Since t~a=ta+1+O(logaa)\tilde{t}_{a}=t_{a+1}+O(\frac{\log a}{\sqrt{a}}), we have limaρ(a)=0\lim_{a\to\infty}\rho^{\prime}(a)=0. There are two possibilities:

  • ρ(a)\rho^{\prime}(a) is monotone, then we have ρ(a)=o(1a)\rho^{\prime}(a)=o(\frac{1}{a}) and there is a sequence aja_{j}\to\infty so that ρ′′(aj)=O(1ajlogaj)\rho^{\prime\prime}(a_{j})=O(\frac{1}{a_{j}\log a_{j}});

  • ρ(a)\rho^{\prime}(a) is not monotone, then around a local minimum where |ρ(a)|<1a|\rho^{\prime}(a)|<\frac{1}{a}, we have ρ′′(a)=0\rho^{\prime\prime}(a)=0.

In both cases, we can apply the above argument to get the claimed estimates in the theorem. The case ρ(a)=t~ata12a<0\rho^{\prime}(a)=\tilde{t}_{a}-t_{a}-\frac{1}{2a}<0 is similar.

As a direct application, we get that limxdudx=1\lim_{x\to\infty}\frac{du}{dx}=1. Then just as Proposition 2.6, we have the following:

Corollary 4.2.

t~ata+1=o(1a)\tilde{t}_{a}-t_{a+1}=o(\frac{1}{\sqrt{a}})

Proof.

With the notation used in the proof of the preceding theorem, we have

ta+1(tta+1)eg(t)𝑑t=(1+o(1))a1/20yey2/2𝑑y\int_{t_{a+1}}^{\infty}(t-t_{a+1})e^{-g(t)}dt=(1+o(1))a^{-1/2}\int_{0}^{\infty}ye^{-y^{2}/2}dy

and similarly

ta+1(tta+1)eg(t)𝑑t=(1+o(1))a1/20yey2/2𝑑y.\int_{-\infty}^{t_{a+1}}(t-t_{a+1})e^{-g(t)}dt=-(1+o(1))a^{-1/2}\int_{0}^{\infty}ye^{-y^{2}/2}dy.

Summing up, we get the corollary. ∎

Corollary 4.3.

For xx large enough, let a(x)a(x) be the number satisfying x=x~ax=\tilde{x}_{a}. Then

f(x)=(1+o(1))c[a]x[a]2πx,f(x)=(1+o(1))c_{[a]}x^{[a]}\sqrt{2\pi x},

where [a][a] is the round down of aa.

Proof.

Since ha(x)=2πa(1+o(1))h_{a}(x)=\sqrt{2\pi a}(1+o(1)), we have f(x)=(1+o(1))caxa2πxf(x)=(1+o(1))c_{a}x^{a}\sqrt{2\pi x}. But since d2da2λ(a)=(1+o(1))1a\frac{d^{2}}{da^{2}}\lambda(a)=(1+o(1))\frac{1}{a}, we have c[a]x[a]=(1+O(1a))caxac_{[a]}x^{[a]}=(1+O(\frac{1}{a}))c_{a}x^{a}. ∎

Corollary 4.4.

We have

dx~ada=1+o(1).\frac{d\tilde{x}_{a}}{da}=1+o(1).
Proof.

Since logx~a=λ(a)\log\tilde{x}_{a}=\lambda^{\prime}(a),

1x~adx~ada=dlogx~ada=λ′′(a).\frac{1}{\tilde{x}_{a}}\frac{d\tilde{x}_{a}}{da}=\frac{d\log\tilde{x}_{a}}{da}=\lambda^{\prime\prime}(a).

By corollary 4.2, x~aa=1\frac{\tilde{x}_{a}}{a}=1, so we get the conclusion. ∎

5. Proof of the uniqueness

Now we prove the uniqueness part in Theorem 1.1. Let f0=i=0cixif_{0}=\sum_{i=0}^{\infty}c_{i}x^{i}, f1=i=0c¯ixif_{1}=\sum_{i=0}^{\infty}\bar{c}_{i}x^{i} be two functions both satisfying

0cixif0𝑑x=0c¯ixif1𝑑x={1β,i=01,i>0\int_{0}^{\infty}\frac{c_{i}x^{i}}{f_{0}}dx=\int_{0}^{\infty}\frac{\bar{c}_{i}x^{i}}{f_{1}}dx=\left\{\begin{array}[]{rr}1-\beta,&\quad i=0\\ 1,&i>0\end{array}\right.

The goal is to prove that f1f_{1} is a scalar multiple of f0f_{0}. Suppose not, we will draw a contradiction below. As before we may define the functions for i[1,)i\in[1,\infty)

c(i)=(0xif0𝑑x)1,c(i)=(\int_{0}^{\infty}\frac{x^{i}}{f_{0}}dx)^{-1},
c¯(i)=(0xif1𝑑x)1.\bar{c}(i)=(\int_{0}^{\infty}\frac{x^{i}}{f_{1}}dx)^{-1}.

Then c(i)=cic(i)=c_{i} and c¯(i)=c¯i\bar{c}(i)=\bar{c}_{i} when ii is a positive integer. We also define c(0)=c0c(0)=c_{0} and c¯(0)=c¯0\bar{c}(0)=\bar{c}_{0}.

For t[0,1]t\in[0,1], let ft(x)=tf1(x)+(1t)f0(x)f_{t}(x)=tf_{1}(x)+(1-t)f_{0}(x). For i{0}[1,)i\in\{0\}\cup[1,\infty) we define ai(t)=tc¯(i)+(1t)c(i)a_{i}(t)=t\bar{c}(i)+(1-t)c(i), and

bi(t)=0ai(t)xift(x)𝑑x.b_{i}(t)=\int_{0}^{\infty}\frac{a_{i}(t)x^{i}}{f_{t}(x)}dx.

Then

bi(t)=0ai(t)ftai(t)ftft2(x)xi𝑑xb_{i}^{\prime}(t)=\int_{0}^{\infty}\frac{a^{\prime}_{i}(t)f_{t}-a_{i}(t)f^{\prime}_{t}}{f^{2}_{t}(x)}x^{i}dx

and

bi′′(t)=02xift3(x)(ai(t)ftai(t)f(t))ft𝑑x.b^{\prime\prime}_{i}(t)=\int_{0}^{\infty}\frac{2x^{i}}{f^{3}_{t}(x)}(a_{i}(t)f^{\prime}_{t}-a_{i}^{\prime}(t)f(t))f^{\prime}_{t}dx.
Lemma 5.1.

We have for all i{0}[1,)i\in\{0\}\cup[1,\infty) that bi(0)<0<bi(1)b_{i}^{\prime}(0)<0<b_{i}^{\prime}(1).

Proof.

Notice that if we replace f1f_{1} by λf1\lambda f_{1} for some λ>0\lambda>0 and let gt=tλf1+(1t)f0g_{t}=t\lambda f_{1}+(1-t)f_{0}, then gtg_{t^{\prime}} is a scalar multiple of ftf_{t} for

t=tλ+(1λ)t.t^{\prime}=\frac{t}{\lambda+(1-\lambda)t}.

The map ttt\mapsto t^{\prime} is a bijection from [0,1][0,1] to itself with a strictly positive derivative. This means that it suffices to prove bi(0)<0b_{i}^{\prime}(0)<0 for a particular choice of λ\lambda. For a given ii we can choose λ\lambda so that c(i)=c¯(i)c(i)=\bar{c}(i). Then

bi′′(t)=02xiai(t)(ft)2ft3(x)𝑑x>0.b^{\prime\prime}_{i}(t)=\int_{0}^{\infty}\frac{2x^{i}a_{i}(t)(f^{\prime}_{t})^{2}}{f^{3}_{t}(x)}dx>0.

Since bi(0)=bi(1)b_{i}(0)=b_{i}(1), we have bi(0)<0<bi(1)b_{i}^{\prime}(0)<0<b_{i}^{\prime}(1).

In particular, we have

(5.1) c¯(i)c(i)(1βδ0i)<0c(i)xif0f1f0𝑑x,\frac{\bar{c}(i)}{c(i)}(1-\beta\delta_{0i})<\int_{0}^{\infty}\frac{c(i)x^{i}}{f_{0}}\frac{f_{1}}{f_{0}}dx,

and

(5.2) c(i)c¯(i)(1βδ0i)<0c¯(i)xif1f0f1𝑑x.\frac{c(i)}{\bar{c}(i)}(1-\beta\delta_{0i})<\int_{0}^{\infty}\frac{\bar{c}(i)x^{i}}{f_{1}}\frac{f_{0}}{f_{1}}dx.

From now on, we fix the normalization that c0=c¯0=1c_{0}=\bar{c}_{0}=1. Then it is not hard to see the following:

Lemma 5.2.

Consider the set Q={c(i)c¯(i)|i,i1}Q=\{\frac{c(i)}{\bar{c}(i)}|i\in\mathbb{R},i\geq 1\}. Then we have supQQ\sup Q\notin Q and infQQ\inf Q\notin Q.

Proof.

First of all, we have supQ>1\sup Q>1 and infQ<1\inf Q<1. Assume that the maximum MM of QQ is attained at i=ni=n, then we have f0f1<M\frac{f_{0}}{f_{1}}<M for all xx. Then we have

M(1βδ0n)=c(n)c¯(n)(1βδ0n)0c¯(n)xnf1f0f1𝑑x<M(1βδ0n),M(1-\beta\delta_{0n})=\frac{c(n)}{\bar{c}(n)}(1-\beta\delta_{0n})\leq\int_{0}^{\infty}\frac{\bar{c}(n)x^{n}}{f_{1}}\frac{f_{0}}{f_{1}}dx<M(1-\beta\delta_{0n}),

a contradiction. The proof for the minimum is the same. ∎

In the following, we will put a bar over a quantity or function to denote the corresponding quantity or function for f1f_{1}. Denote ki=c(i)c¯(i)k_{i}=\frac{c(i)}{\bar{c}(i)}. From Lemma 5.2, one sees that

infQ=lim infi+ki,\inf Q=\liminf_{i\to+\infty}k_{i},

and

supQ=lim supi+ki.\sup Q=\limsup_{i\to+\infty}k_{i}.

So lim infi+kilim supi+ki\liminf_{i\to+\infty}k_{i}\neq\limsup_{i\to+\infty}k_{i}. Let a2a_{2} (not necessarily an integer) be a local minimum point of kik_{i} as a function of ii, satisfying that ki>ka2k_{i}>k_{a_{2}} for all i<a2i<a_{2}. Let a1a_{1} be the integer satisfying the following:

  • a1<a2a_{1}<a_{2};

  • ki<ka1k_{i}<k_{a_{1}} for a1<ia2a_{1}<i\leq a_{2};

  • kika1k_{i}\leq k_{a_{1}} for i<a1i<a_{1}.

By Lemma 5.2, it is clear that we can choose a1a_{1}, hence a2a_{2}, arbitrarily large. So our notations o(1a1)o(\frac{1}{a_{1}}) and o(1)o(1) in the following make sense.

With the notation from the preceding section, let λ(i)=logci\lambda(i)=-\log c_{i} and λ¯=logc¯i\bar{\lambda}=-\log\bar{c}_{i}. Let

ε=maxa1aa2|λ′′(a)λ¯′′(a)|a=o(1),\varepsilon=\max_{a_{1}\leq a\leq a_{2}}|\lambda^{\prime\prime}(a)-\bar{\lambda}^{\prime\prime}(a)|\cdot a=o(1),

then since λ(a2)λ¯(a2)=0\lambda^{\prime}(a_{2})-\bar{\lambda}^{\prime}(a_{2})=0,

logka1ka2ε2a1(a2a1)2.\log\frac{k_{a_{1}}}{k_{a_{2}}}\leq\frac{\varepsilon}{2a_{1}}(a_{2}-a_{1})^{2}.

So a2a12logka1ka2εa1a_{2}-a_{1}\geq\sqrt{\frac{2\log\frac{k_{a_{1}}}{k_{a_{2}}}}{\varepsilon}}\sqrt{a_{1}}. In other words, a2a_{2} is not close to a1a_{1}.

When f=f0f=f_{0}, let x1x_{1} satisfy x1=xa1x_{1}=x^{\prime}_{a_{1}}, namely logx1=λ(a1)\log x_{1}=\lambda^{\prime}(a_{1}). Similarly, we define logx¯1=λ¯(a1)\log\bar{x}_{1}=\bar{\lambda}^{\prime}(a_{1}), logx¯2=logx2=λ(a2)\log\bar{x}_{2}=\log x_{2}=\lambda^{\prime}(a_{2}). So logx1logx¯1=o(1a1)\log x_{1}-\log\bar{x}_{1}=o(\frac{1}{a_{1}}). Since f0(x1)2πa1ca1x1a1f_{0}(x_{1})\approx\sqrt{2\pi a_{1}}c_{a_{1}}x_{1}^{a_{1}} and f1(x1)2πa1c¯a1x1a1f_{1}(x_{1})\approx\sqrt{2\pi a_{1}}\bar{c}_{a_{1}}x_{1}^{a_{1}}, we have

f0(x1)f1(x1)=(1+o(1))ca1c¯a1.\frac{f_{0}(x_{1})}{f_{1}(x_{1})}=(1+o(1))\frac{c_{a_{1}}}{\bar{c}_{a_{1}}}.

And similarly

f0(x2)f1(x2)=(1+o(1))ca2c¯a2.\frac{f_{0}(x_{2})}{f_{1}(x_{2})}=(1+o(1))\frac{c_{a_{2}}}{\bar{c}_{a_{2}}}.

Consider the integral

I1=0c¯a1xa1f1f0f1𝑑x.I_{1}=\int_{0}^{\infty}\frac{\bar{c}_{a_{1}}x^{a_{1}}}{f_{1}}\frac{f_{0}}{f_{1}}dx.

First notice that G(t)=log(c¯a1xa1f1f0f1)G(t)=-\log(\frac{\bar{c}_{a_{1}}x^{a_{1}}}{f_{1}}\frac{f_{0}}{f_{1}}) is again a convex function of t=logxt=\log x with G′′(t)=(1+o(1))xG^{\prime\prime}(t)=\frac{(1+o(1))}{x}. We have 0c¯a1xa1f1𝑑x=1\int_{0}^{\infty}\frac{\bar{c}_{a_{1}}x^{a_{1}}}{f_{1}}dx=1, we now estimate f0f1\frac{f_{0}}{f_{1}}.

Lemma 5.3.

For xx2x\leq x_{2}, we have

f0(x)f1(x)<ca1c¯a1.\frac{f_{0}(x)}{f_{1}(x)}<\frac{c_{a_{1}}}{\bar{c}_{a_{1}}}.
Proof.

Recall that the summation i=0cixi\sum_{i=0}^{\infty}c_{i}x^{i}, when xx is large, is concentrated on the terms cixic_{i}x^{i} for ii close to a(x)a(x). By our choice, we have λ(a1)λ¯(a1)=o(1a1)\lambda^{\prime}(a_{1})-\bar{\lambda}^{\prime}(a_{1})=o(\frac{1}{a_{1}}). So x1x¯1=o(1)x_{1}-\bar{x}_{1}=o(1). Since d2di2logki=o(1i)\frac{d^{2}}{di^{2}}\log k_{i}=o(\frac{1}{i}), we have that within a neighborhood of the form |ia2|a2o(1)|i-a_{2}|\leq\frac{\sqrt{a_{2}}}{o(1)},

ki1.k_{i}\leq 1.

Then we can use integral to estimate the summations cix1i\sum c_{i}x_{1}^{i} and c¯ix1i\sum\bar{c}_{i}x_{1}^{i}. For any δ>0\delta>0, we have

Aex2𝑑x<εA2Aex2𝑑x,\int_{A}^{\infty}e^{-x^{2}}dx<\varepsilon\int_{A-2}^{A}e^{-x^{2}}dx,

for AA large enough. Then it is easy to see that we can find a3>a2a_{3}>a_{2}, so that

  • ki1k_{i}\leq 1 for a2ia3a_{2}\leq i\leq a_{3};

  • a3a2>logka2εa2a_{3}-a_{2}>\sqrt{\frac{-\log k_{a_{2}}}{\varepsilon}}\sqrt{a_{2}}

Then for any δ>0\delta>0, we have

ia2cix1i\displaystyle\sum_{i\geq a_{2}}c_{i}x_{1}^{i} <\displaystyle< (1+δ)a3i>a2cix1i\displaystyle(1+\delta)\sum_{a_{3}\geq i>a_{2}}c_{i}x_{1}^{i}
<\displaystyle< (1+δ)2a3i>a2c¯ix1i,\displaystyle(1+\delta)^{2}\sum_{a_{3}\geq i>a_{2}}\bar{c}_{i}x_{1}^{i},

for a2a_{2} large enough. So we have

f0(x1)f1(x1)<ca1c¯a1.\frac{f_{0}(x_{1})}{f_{1}(x_{1})}<\frac{c_{a_{1}}}{\bar{c}_{a_{1}}}.

Then since x2=x2x_{2}=x_{2}^{\prime}, we can repeat the argument to show that the conclusion holds for x=x2x=x_{2}. Then for x<x2x<x_{2}, we have that nx<a2n_{x}<a_{2} and n¯x<a2\bar{n}_{x}<a_{2}. So the argument above still works. ∎

Now we are ready to draw a contradiction.

We consider the function ν(x)=c¯a1xa1f1f0f1\nu(x)=\frac{\bar{c}_{a_{1}}x^{a_{1}}}{f_{1}}\frac{f_{0}}{f_{1}}. Firstly, we have

f0(x1)f1(x1)=(1+o(1))ca1c¯a1.\frac{f_{0}(x_{1})}{f_{1}(x_{1})}=(1+o(1))\frac{c_{a_{1}}}{\bar{c}_{a_{1}}}.

By corollary 4.2, we have x¯a1=x1+o(x1)\bar{x}_{a_{1}}^{\prime}=x_{1}+o(\sqrt{x_{1}}), namely the maximum point of the function c¯a1xa1f1\frac{\bar{c}_{a_{1}}x^{a_{1}}}{f_{1}} is close to x1x_{1}. So we have h¯a1(x1)=(1+o(1))h¯a1(x¯a1)\bar{h}_{a_{1}}(x_{1})=(1+o(1))\bar{h}_{a_{1}}(\bar{x}_{a_{1}}^{\prime}). Let xνx_{\nu} be the maximum point of λ(x)\lambda(x), then we have

ν(xν)<ka11h¯a1(x¯a1).\nu(x_{\nu})<k_{a_{1}}\frac{1}{\bar{h}_{a_{1}}(\bar{x}_{a_{1}}^{\prime})}.

Therefore

xλx1=o(x1).x_{\lambda}-x_{1}=o(\sqrt{x_{1}}).

We adapt the notations in the proof of Lemma 5.3 and define x3=x¯a3x_{3}=\bar{x}^{\prime}_{a_{3}}, namely logx3=λ¯(a3)\log x_{3}=\bar{\lambda}^{\prime}(a_{3}). Then by Corollary 4.4, we have x3x2=(1+o(1))(a3a2)x_{3}-x_{2}=(1+o(1))(a_{3}-a_{2}) and

x2x1>(1+o(1))2logka1ka2εa1.x_{2}-x_{1}>(1+o(1))\sqrt{\frac{2\log\frac{k_{a_{1}}}{k_{a_{2}}}}{\varepsilon}}\sqrt{a_{1}}.

Then as in the proof of Lemma 5.3, for any δ>0\delta>0, we have

x2ν(x)𝑑x(1+δ)x2x3ν(x)𝑑x,\int_{x_{2}}^{\infty}\nu(x)dx\leq(1+\delta)\int_{x_{2}}^{x_{3}}\nu(x)dx,

for a2a_{2} large enough.

For x(x2,x3)x\in(x_{2},x_{3}), we can use Corollary 4.3 to estimate ν(x)<(1+δ)c¯a1xa1f1\nu(x)<(1+\delta)\frac{\bar{c}_{a_{1}}x^{a_{1}}}{f_{1}}. Therefore, we get

x2ν(x)𝑑x(1+δ)2x2x3c¯a1xa1f1𝑑x,\int_{x_{2}}^{\infty}\nu(x)dx\leq(1+\delta)^{2}\int_{x_{2}}^{x_{3}}\frac{\bar{c}_{a_{1}}x^{a_{1}}}{f_{1}}dx,

for a2a_{2} large enough.

So we just need to take δ\delta so that (1+δ)2<ka1(1+\delta)^{2}<k_{a_{1}}, then by Lemma 5.3,

0c¯a1xa1f1f0f1𝑑x<cac¯a.\int_{0}^{\infty}\frac{\bar{c}_{a_{1}}x^{a_{1}}}{f_{1}}\frac{f_{0}}{f_{1}}dx<\frac{c_{a}}{\bar{c}_{a}}.

This contradicts (5.2). Therefore we have completed the proof of the uniqueness part in Theorem 1.1.

Proof of Corollary 1.2.

We write eβ(x)=cn(β)xne_{\beta}(x)=\sum c_{n}(\beta)x^{n}. If we let f0=eβ,β>0f_{0}=e_{\beta},\beta>0 and f1(x)=exf_{1}(x)=e^{x}, then assume that lim infnknlim supnkn\liminf_{n\to\infty}k_{n}\neq\limsup_{n\to\infty}k_{n}, we can run the argument above again to get a contradiction. So we must have lim infnkn=lim supnkn\liminf_{n\to\infty}k_{n}=\limsup_{n\to\infty}k_{n}, which may be infinity. Since 01f0(x)𝑑x<1\int_{0}^{\infty}\frac{1}{f_{0}(x)}dx<1, we get that supQ>1\sup Q>1. Then if we repeat the proof of Lemma 5.2, we get that supQQ\sup Q\notin Q. So we have lim supnkn=supQ\limsup_{n\to\infty}k_{n}=\sup Q. Then if infQ1\inf Q\leq 1, then infQ\inf Q is attained for some knQk_{n}\in Q, and then we can repeat the proof of Lemma 5.2 again to get a contradiction. Therefore

cn(β)>1n!,c_{n}(\beta)>\frac{1}{n!},

for all n1n\geq 1.

We can also get cn(β)>cn(γ)c_{n}(\beta)>c_{n}(\gamma) for γ<β\gamma<\beta, by letting f1=eγf_{1}=e_{\gamma}. Therefore, if we let

c¯n(γ)=infβ>γcn(β),\bar{c}_{n}(\gamma)=\inf_{\beta>\gamma}c_{n}(\beta),

then the entire function f(x)=i=0c¯n(γ)xnf(x)=\sum_{i=0}^{\infty}\bar{c}_{n}(\gamma)x^{n} satisfies (1.3) with β\beta replaced by γ\gamma. To see this, one only needs to use again the property that for fixed nn, for all ε>0\varepsilon>0, There exists N>0N>0 such that Nxnex𝑑x<ε\int_{N}^{\infty}\frac{x^{n}}{e^{x}}dx<\varepsilon. So by the uniqueness theorem, f(x)=eγ(x)f(x)=e_{\gamma}(x). So we have proved the upper semi-continuity of cn(β)c_{n}(\beta). The lower semi-continuity is similar. So by Dini’s Theorem, we have proved corollary 1.2. ∎

References

  • [1] Fabrizio Cuccu and Andrea Loi. Balanced metrics on n\mathbb{C}^{n}. J. Geom. Phys., 57(4):1115–1123, 2007.
  • [2] WK Hayman and I Vincze. A problem on entire functions. Complex Analysis and its Applications (dedicated to IN Vekua on his 70th Birthday). Izd. Nauka, Moscow, pages 591–594, 1978.
  • [3] Julien Keller and Julius Ross. A note on chow stability of the projectivization of gieseker stable bundles. The Journal of Geometric Analysis, 24(3):1526–1546, Jul 2014.
  • [4] Andrea Loi and Roberto Mossa. Uniqueness of balanced metrics on holomorphic vector bundles. J. Geom. Phys., 61(1):312–316, 2011.
  • [5] Joseph Miles and Jack Williamson. A characterization of the exponential function. Journal of the London Mathematical Society, s2-33(1):110–116, 1986.
  • [6] Reza Seyyedali. Balanced metrics and Chow stability of projective bundles over Kähler manifolds. Duke Mathematical Journal, 153(3):573 – 605, 2010.
  • [7] Reza Seyyedali. Balanced metrics and chow stability of projective bundles over kähler manifolds ii. Journal of Geometric Analysis, 23(4):1944–1975, Oct 2013.
  • [8] Jingzhou Sun and Song Sun. An infinite dimensional balanced embedding problem I:existence. Preprint, arxiv 2309.01097, Available at http://arxiv.org/abs/2309.01097.
  • [9] Xiaowei Wang. Canonical metrics on stable vector bundles. Communications in Analysis and Geometry, 13(2):253–286, 2005.