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

On the exact self-similar finite-time blowup
of the Hou–Luo model with smooth profiles

De Huang1, Xiang Qin, Xiuyuan Wang, and Dongyi Wei2
Abstract.

We show that the 1D Hou–Luo model on the real line admits exact self-similar finite-time blowup solutions with smooth self-similar profiles. The existence of these profiles is established via a fixed-point method that is purely analytic. We also prove that the profiles satisfy some monotonicity and convexity properties that were unknown before, and we give rigorous estimates on the algebraic decay rates of the profiles in the far field. Our result supplements the previous computer-assisted proof of self-similar finite-time blowup for the Hou–Luo model with finer characterizations of the profiles.

School of Mathematical Sciences, Peking University

1. Introduction

We consider the 1D Hou–Luo (HL) model

ωt+uωx=θx,θt+uθx=0,ux=𝑯(ω),\begin{split}&\omega_{t}+u\omega_{x}=\theta_{x},\\ &\theta_{t}+u\theta_{x}=0,\\ &u_{x}=\bm{H}(\omega),\end{split} (1.1)

for xx\in\mathbb{R}, where 𝑯()\bm{H}(\cdot) denotes the Hilbert transform on the real line. This model was first proposed by Luo and Hou [10, 11] to acquire understanding of the numerically observed self-similar singularity formation of the 3D axisymmetric Euler equations on the solid boundary of an infinitely long cylinder. It also models the boundary induced singularity formation of the 2D Boussinesq equations [11, 3] in the half-space (x1,x2)×+(x_{1},x_{2})\in\mathbb{R}\times\mathbb{R}_{+},

ωt+u1ωx1+u2ωx2=θx1,θt+u1θx1+u2θx2=0,(u1,u2)=(Δ)1ω,\begin{split}&\omega_{t}+u_{1}\omega_{x_{1}}+u_{2}\omega_{x_{2}}=\theta_{x_{1}},\\ &\theta_{t}+u_{1}\theta_{x_{1}}+u_{2}\theta_{x_{2}}=0,\\ &(u_{1},u_{2})=\nabla^{\perp}(-\Delta)^{-1}\omega,\end{split} (1.2)

for the 2D Boussinesq equations behave similarly to the 3D axisymmetric Euler equations away from the symmetry axis; see e.g. [13]. In fact, the boundary finite-time blowup of 3D axisymmetric Euler equations can be approximated by the boundary finite-time blowup of the 2D Boussinesq equations up to an asymptotically small perturbation [11, 1].

Ever since the report of convincing numerical evidence of a self-similar finite-time blowup of the 3D axisymmetric Euler equations with boundary, namely the Hou–Luo scenario [10], vast amounts of effort have been made to try to rigorously prove the existence of such boundary singularity for the 3D Euler/2D Boussinesq equations as well as a number of simplified models including the HL model (1.1). We recommend the survey paper [7] for a comprehensive literature review. We only list some of the most relevant ones here. Shortly after the original work of Luo and Hou [11], Choi et al. [4] used a functional argument to prove the finite-time blowup of the HL model (1.1) and another 1D model known as the CKY model [5]. However, their approach was not able to capture the self-similar nature of the blowup. Years later, Chen, Hou, and Huang [3] developed a novel analysis framework based on rigorous computer-assisted proofs to establish asymptotically self-similar finite-time blowup of the HL model from smooth initial data. In particular, they constructed an approximate self-similar profile using numerical computation, and then they showed by an energy argument that any solution of the HL model that is initially close to the approximate self-similar profile (up to proper rescaling) will develop finite-time singularity in an asymptotically self-similar way. Recently, Chen and Hou [1] generalized this powerful computer-assisted approach to the higher dimension and used it to prove the asymptotically self-similar finite-time blowup of the 2D Boussinesq/3D Euler equations with boundary, thus finally settling the conjecture on the Hou–Luo scenario. Remarkably, their work showed for the first time that the 3D Euler equations can develop finite-time singularity from smooth initial data, though the presence of a solid boundary is critically necessary in this scenario. Whether this can happen in the free space 3\mathbb{R}^{3} still remains open.

As mentioned above, the asymptotically self-similar finite-time blowups of the 2D Boussinesq equations (1.2) and the 1D HL model (1.1) were both established via a computer-assisted approach. On the one hand, computer-assisted analysis can yield much sharper or tighter estimates far beyond the reach of pure analysis, which is critical in the proofs of these blowups. On the other hand, this novel proof framework relies heavily on computer assistance, which can be quite difficult to digest and to reproduce for most of the readers. Furthermore, though the existence of an exact self-similar solution (close to the approximate one constructed numerically) is also implied by the computer-assisted approach, characterizations of the solution profiles are very limited. For example, the computer-assisted proof did not tell whether the exact self-similar profiles for the 2D Boussinesq equations or the 1D HL model are smooth. Therefore, it would be helpful to develop pure analytic strategies to prove the existence of exact self-similar finite-time blowups and to provide finer characterizations of the corresponding self-similar profiles.

In this paper, we focus on the 1D HL model (1.1). More specifically, we look for exact self-similar solutions of (1.1) that take the form

ω(x,t)=(Tt)cωΩ(x(Tt)cl),θ(x,t)=(Tt)cθΘ(x(Tt)cl),\omega(x,t)=(T-t)^{c_{\omega}}\cdot\Omega\left(\frac{x}{(T-t)^{c_{l}}}\right),\quad\theta(x,t)=(T-t)^{c_{\theta}}\cdot\Theta\left(\frac{x}{(T-t)^{c_{l}}}\right), (1.3)

where Ω,Θ\Omega,\Theta are the self-similar profiles, cω,cθ,clc_{\omega},c_{\theta},c_{l} are the scaling factors, and TT is the finite blowup time. This particular self-similar ansatz is due to the natural scaling property of equations (1.1). Plugging this ansatz into (1.1) and balancing the equations as tTt\rightarrow T yields cω=1c_{\omega}=-1 and cθ=2cω+clc_{\theta}=2c_{\omega}+c_{l}. As we will see, the undetermined value clc_{l} is related to the far-field decay rates of Ω\Omega and Θ\Theta.

In contrast to an exact self-similar solution, an asymptotically self-similar finite-time blowup refers to a solution that exhibits clear self-similarity only as tt approaches the finite blowup time TT:

ω~(x,t)=(Tt)c~ω(t)Ω~(x(Tt)c~l(t),t),θ~(x,t)=(Tt)c~θΘ~(x(Tt)c~l,t),\tilde{\omega}(x,t)=(T-t)^{\tilde{c}_{\omega}(t)}\cdot\widetilde{\Omega}\left(\frac{x}{(T-t)^{\tilde{c}_{l}(t)}},t\right),\quad\tilde{\theta}(x,t)=(T-t)^{\tilde{c}_{\theta}}\cdot\widetilde{\Theta}\left(\frac{x}{(T-t)^{\tilde{c}_{l}}},t\right), (1.4)

where the profiles Ω~(,t),Θ~(,t)\widetilde{\Omega}(\cdot,t),\widetilde{\Theta}(\cdot,t) and the scaling factors c~ω(t),c~θ(t),c~l(t)\tilde{c}_{\omega}(t),\tilde{c}_{\theta}(t),\tilde{c}_{l}(t) all depend on time and will converge to some non-trivial steady state as tTt\rightarrow T. In particular, c~ω(t)\tilde{c}_{\omega}(t) must converge to 1-1. In their computer-assisted framework, Chen, Hou, and Huang [3] proved the existence of a family of asymptotically self-similar finite-time blowups of the form (1.4) by showing the nonlinear quasi-stability of the corresponding dynamic rescaling equations around an approximate steady state constructed by numerical methods. The term “a family of” means that the initial state of (1.4) can be taken arbitrarily as long as it falls in a small energy-norm ball centered at the approximate steady state (up to rescaling). Furthermore, they also used a limit argument to show that, near the approximate steady state, there lies a true stable steady state that corresponds to exact self-similar profiles, though they could not provide accurate descriptions of these profiles. Nevertheless, their result gives a very accurate estimate of the spacial scaling: |clc¯l|6×105|c_{l}-\bar{c}_{l}|\leq 6\times 10^{-5}, where c¯l=2.99870\bar{c}_{l}=2.99870 is a numerically computed constant corresponding to the approximate steady state, and the bound 6×1056\times 10^{-5} results from computer-assisted estimates. It is interesting that clc_{l} is extremely close to but strictly smaller than 33.

As supplementary to the existing results obtained via computer-assisted proofs, we prove the existence of an exact self-similar finite-time blowup of the form (1.3) for the HL model (1.1) in an alternative way that is pure analytic, and we provide more detailed characterizations of the self-similar profiles.

Theorem 1.1.

The 1D Hou–Luo model (1.1) on the real line admits an exact self-similar finite-time blowup solution of the form (1.3) with cω=1c_{\omega}=-1, cl(2,4.53)c_{l}\in(2\,,4.53), cθ=2cω+clc_{\theta}=2c_{\omega}+c_{l}, and a pair of profiles Ω,Θ\Omega,\Theta that satisfy the following:

  1. (1)

    Ω(x)\Omega(x) is odd in xx, and Θ(x)\Theta(x) is even in xx.

  2. (2)

    The functions f(x)=Ω(x)/xf(x)=\Omega(x)/x and m(x)=Θ(x)/xm(x)=\Theta^{\prime}(x)/x are both decreasing in xx and convex in x2x^{2} on [0,+)[0,+\infty); that is, f(x),m(x)<0f^{\prime}(x),m^{\prime}(x)<0 and (f(x)/x),(m(x)/x)0(f^{\prime}(x)/x)^{\prime},(m^{\prime}(x)/x)^{\prime}\geq 0 for all x>0x>0.

  3. (3)

    Ω\Omega and Θ\Theta are both infinitely smooth on \mathbb{R}. Moreover, ΩLLq()H˙p()\Omega\in L^{\infty}\cap L^{q}(\mathbb{R})\cap\dot{H}^{p}(\mathbb{R}) for any q>clq>c_{l} and any p1p\geq 1, and ΘLLq()H˙p()\Theta^{\prime}\in L^{\infty}\cap L^{q}(\mathbb{R})\cap\dot{H}^{p}(\mathbb{R}) for any q>cl/2q>c_{l}/2 and any p1p\geq 1.

  4. (4)

    Both the limits limx+x1/clΩ(x)\displaystyle\lim_{x\rightarrow+\infty}x^{1/c_{l}}\Omega(x) and limx+x2/clΘ(x)\displaystyle\lim_{x\rightarrow+\infty}x^{2/c_{l}}\Theta^{\prime}(x) exist and are positive and finite.

Let us remark on our result. The existence of the exact self-similar profiles is proved via a nonlinear fixed-point method. More precisely, we first construct two nonlinear nonlocal maps 𝑴\bm{M} and 𝑹\bm{R} over a suitable function DD such that, if f𝔻f\in\mathbb{D} is a fixed point of 𝑹\bm{R}, i.e. f=𝑹(f)f=\bm{R}(f), then Ω(x)=xf(x)\Omega(x)=xf(x) and Θ(x)=(cl/2)x𝑴(f)(x)\Theta^{\prime}(x)=(c_{l}/2)x\bm{M}(f)(x) are a pair of exact self-similar profiles of the HL model (1.1) with cω,cθ,clc_{\omega},c_{\theta},c_{l} given explicitly in terms of integrals of ff. We then prove the existence of a fixed point of 𝑹\bm{R} using the Schauder fixed-point theorem. A key observation in our proof is that the map 𝑹\bm{R} preserves the properties that f(x)f(x) is non-increasing in xx and convex in x2x^{2} for x0x\geq 0, which will be frequently used in our arguments. Furthermore, based on the fixed-point relation f=𝑹(f)f=\bm{R}(f), we are able to determine the regularity of f,𝑴(f)f,\bm{M}(f) and their far-field decay rates, which then transfer to desired properties of the corresponding self-similar profiles Ω\Omega and Θ\Theta.

This proof strategy is modified from the fixed-point method developed in a recent work by the same authors [9], where we proved the existence of exact self-similar finite-time blowups of the a-parameterized family of the generalized Constantin–Lax–Majda (gCLM) equation [8, 6, 14],

ωt+auωx=uxω,ux=𝑯(ω),\omega_{t}+au\omega_{x}=u_{x}\omega,\quad u_{x}=\bm{H}(\omega),

for all a1a\leq 1. This equation is also a 1D model for the vorticity formulation of the 3D incompressible Euler equations, with a parameter aa that controls the competition between advection and vortex stretching. The main difficulty of modifying our fixed-point method from the gCLM case to the HL case lies in that the gCLM model is an equation of one scalar ω\omega, while the HL model is a coupled system of two scalars ω,θ\omega,\theta. Surprisingly, though the nonlinear map 𝑹\bm{R} for the HL system is formally much more complicated than the one for the gCLM equation, it still enjoys the critical property that it preserves the previously mentioned monotonicity and convexity of f𝔻f\in\mathbb{D}. From a more essential perspective, this nice property of 𝑹\bm{R} owes to the kernel structure of the Hilbert transform 𝑯\bm{H}. However, we believe that this fixed-point framework can be further generalized to prove the existence of an exact self-similar finite-time blowup of the 2D Boussinesq equations. This shall be our next step in this line of research.

One unsatisfying thing about our result is the crude estimate of clc_{l}. Based on our fixed-point method, we can only show that 2<cl2(k+1)/(k1)4.52982<c_{l}\leq 2(k+1)/(k-1)\approx 4.5298, where k=1+10/2k=1+\sqrt{10}/2. This is of course much worse than the estimate 2.99870±6×1052.99870\pm 6\times 10^{-5} obtained in [3]. From this comparison, one sees clearly how rigorous computer-assisted estimates can outperform pure analytic estimates in providing sharp bounds.

Finally, we remark that our work does not prove the uniqueness (up to rescaling) of self-similar profiles for the HL model (1.1). Hence, we cannot conclude that the self-similar profiles Ω,Θ\Omega,\Theta obtained by our fixed-point method are identical (under proper rescaling) to those obtained via the computer-assisted proof. Nevertheless, it is likely that the exact self-similar solution is unique provided that cω=1c_{\omega}=-1 and Ω(0)=1\Omega^{\prime}(0)=1. In fact, we can numerically compute the fixed point f=𝑹(f)f=\bm{R}(f) by an iterative scheme and compare it to the one obtained numerically in [3] by solving the dynamic rescaling equation, and we see that they match perfectly well up to only scheme errors.

The remainder of this paper is organized as follows. In Section 2, we derive equations for the self-similar profiles and then transform them into an equivalent fixed-point formulation. Section 3 is devoted to proving the existence of exact self-similar profiles via a fixed-point method, and Section 4 is devoted to the establishment of the claimed properties. Finally, we perform some numerical simulations in Section 5 based on the fixed-point method to verify and visualize our theoretical results.

2. Equations for the self-similar profiles

Assuming that (1.3) is an exact self-similar solution of the HL model (1.1), we first derive a nonlocal ordinary differential system for the self-similar profiles Ω,Θ\Omega,\Theta and the scaling factors cl,cω,cθc_{l},c_{\omega},c_{\theta}. Under some natural regularity conditions on Ω\Omega and Θ\Theta, we perform a change of variables and then transform the self-similar equations into a fixed-point formulation in the new variables.

2.1. Self-similar profiles

Substituting the self-similar ansatz (1.3) into the equation (1.1) yields

(Tt)cω1cωΩ+(Tt)cω1clXΩX+(Tt)2cωUΩX=(Tt)cθclΘX,(Tt)cθ1cθΘ+(Tt)cθ1clXΘX+(Tt)cθ+cωUΘX=0,UX=𝑯(Ω),U(0)=0,\begin{split}&-(T-t)^{c_{\omega}-1}c_{\omega}\Omega+(T-t)^{c_{\omega}-1}c_{l}X\Omega_{X}+(T-t)^{2c_{\omega}}U\Omega_{X}=(T-t)^{c_{\theta}-c_{l}}\Theta_{X},\\ &-(T-t)^{c_{\theta}-1}c_{\theta}\Theta+(T-t)^{c_{\theta}-1}c_{l}X\Theta_{X}+(T-t)^{c_{\theta}+c_{\omega}}U\Theta_{X}=0,\\ &U_{X}=\bm{H}(\Omega),\quad U(0)=0,\end{split}

where X=x/(Tt)clX=x/(T-t)^{c_{l}}. Balancing the above equations as tTt\rightarrow T yields cω=1c_{\omega}=-1, cθ=cl+2cωc_{\theta}=c_{l}+2c_{\omega}, and a system for the self-similar profiles:

(clX+U)ΩX=cωΩ+ΘX,(clX+U)ΘX=cθΘ,UX=𝑯(Ω),\begin{split}&(c_{l}X+U)\Omega_{X}=c_{\omega}\Omega+\Theta_{X},\\ &(c_{l}X+U)\Theta_{X}=c_{\theta}\Theta,\\ &U_{X}=\bm{H}(\Omega),\end{split}

From now on, for notation simplicity, we will still use ω,θ,u,x\omega,\theta,u,x for Ω,Θ,U,X\Omega,\Theta,U,X, respectively. Moreover, it is more convenient to work with the variable v=θxv=\theta_{x} instead of θ\theta. We thus obtain our main equations:

(clx+u)ωx=cωω+v,(clx+u)vx=(2cωux)v,ux=𝑯(ω).\begin{split}&(c_{l}x+u)\omega_{x}=c_{\omega}\omega+v,\\ &(c_{l}x+u)v_{x}=(2c_{\omega}-u_{x})v,\\ &u_{x}=\bm{H}(\omega).\end{split} (2.1)

We have substituted cθ=cl+2cωc_{\theta}=c_{l}+2c_{\omega}, and we will always do so in what follows. The expressions of uu and uxu_{x} in terms of ω\omega are, respectively,

u(x)=(Δ)1/2ω(x)=1πω(y)ln|xy|dy,ux(x)=𝑯(ω)(x)=1πP.V.ω(y)xydy.\begin{split}&u(x)=-(-\Delta)^{-1/2}\omega(x)=\frac{1}{\pi}\int_{\mathbb{R}}\omega(y)\ln|x-y|\,\mathrm{d}{y},\\ &u_{x}(x)=\bm{H}(\omega)(x)=\frac{1}{\pi}P.V.\int_{\mathbb{R}}\frac{\omega(y)}{x-y}\,\mathrm{d}{y}.\end{split} (2.2)

Here, 𝑯\bm{H} is the Hilbert transform on the real line with P.V.P.V. denoting the Cauchy principal value.

It is worth mentioning that Chen, Hou, and Huang [3] studied the dynamic rescaling equations of the HL model (see [3, Equation (2.1)]),

ωt+(cl(t)x+u)ωx=cω(t)ω+v,vt+(cl(t)x+u)vx=(2cω(t)ux)v,ux=𝑯(ω),\begin{split}&\omega_{t}+(c_{l}(t)x+u)\omega_{x}=c_{\omega}(t)\omega+v,\\ &v_{t}+(c_{l}(t)x+u)v_{x}=(2c_{\omega}(t)-u_{x})v,\\ &u_{x}=\bm{H}(\omega),\end{split} (2.3)

which is in fact equivalent to the original HL model (1.1) under some time-dependent change of variables. They needed to choose how cl(t),cω(t)c_{l}(t),c_{\omega}(t) depend on the solution ω(x,t),v(x,t)\omega(x,t),v(x,t) (see (2.5) below) in order to capture the intrinsic blowup scaling. In this way, they reformulated the finite-time blowup problem into a stability problem. With the help of a numerically constructed approximate steady state, they showed that the solution of (2.3) converges to some true steady state near the approximate one, which then implies the self-similar finite-time blowup of the original HL model.

Different from their dynamic approach, our strategy is to directly find a nontrivial solution (ω,v,cl,cω)(\omega,v,c_{l},c_{\omega}) of equations (2.1)\eqref{eqt:main}, which is then an exact steady state of the dynamic rescaling equations (2.3). One important fact is that, if (ω,v,cl,cω)(\omega,v,c_{l},c_{\omega}) is a solution of (2.1), then

(ωα,β(x),vα,β(x),cl,α,cω,α)=(αω(βx),α2v(βx),αcl,αcω)(\omega_{\alpha,\beta}(x),v_{\alpha,\beta}(x),c_{l,\alpha},c_{\omega,\alpha})=(\alpha\omega(\beta x),\alpha^{2}v(\beta x),\alpha c_{l},\alpha c_{\omega}) (2.4)

is also a solution of (2.1) for any α,β>0\alpha\in\mathbb{R},\beta>0. Owing to this scaling property, we can release the restriction that cω=1c_{\omega}=-1. In fact, it is the ratio cl/cωc_{l}/c_{\omega} that matters. Moreover, in consistence with the previous computer-assisted work [3], we look for solutions to the profile equations (2.1) that satisfy the following assumptions:

  • Odd symmetry: ω\omega and vv are both odd functions of xx. As a consequence, uu is also odd in xx.

  • Non-degeneracy: ω(0)>0\omega^{\prime}(0)>0 and v(0)>0v^{\prime}(0)>0.

  • Out-pushing condition: cl+u/x>0c_{l}+u/x>0 for all xx\in\mathbb{R}.

These conditions are all satisfied by the approximate steady state constructed in [3], and they are actually critical to the stability argument there. The non-degeneracy condition not only ensures the solution is not trivial, but also determines how the scaling factors are related to the profiles:

cl=2v(0)ω(0),cω=12cl+u(0).c_{l}=2\frac{v^{\prime}(0)}{\omega^{\prime}(0)},\quad c_{\omega}=\frac{1}{2}c_{l}+u^{\prime}(0). (2.5)

To derive these relations, one simply evaluates the derivatives of the first two equations in (2.1) at x=0x=0 and uses the non-degeneracy condition and the odd symmetry. Conversely, imposing the relations (2.5) on the dynamic rescaling equations (2.3), as implemented in [3], ensures that ω(0,t)\omega^{\prime}(0,t) and v(0,t)v^{\prime}(0,t) are conserved quantities over time, i.e. ω(0,t)=ω(0,0)\omega^{\prime}(0,t)=\omega^{\prime}(0,0), v(0,t)=v(0,0)v^{\prime}(0,t)=v^{\prime}(0,0), which is essentially the source of stability. The odd symmetry, which is preserved by the dynamic rescaling equations, then ensures that the origin x=0x=0 is a stable stagnation point that “generates” stability. Finally, the out-pushing condition implies that the velocity clx+uc_{l}x+u has the same sign as xx, and thus it transports the stability from the origin towards the far-field.

Finally, we remark that it is possible to find odd nontrivial solutions of equations (2.1) that is more degenerate at x=0x=0 in the sense that ω(0)=v(0)=0\omega^{\prime}(0)=v^{\prime}(0)=0 while ω(n)(0)0,v(n)(0)0\omega^{(n)}(0)\neq 0,v^{(n)}(0)\neq 0 for some odd integer n>1n>1. Numerical evidence of the existence of these degenerate solutions has been reported in the thesis of Liu [12].

2.2. Reformulation of the problem

Now, we move on to transforming equations (2.1) into a fixed-point formulation. In view of (2.4), we may assume that ω(0)=1\omega^{\prime}(0)=1, in which case cl=2v(0)c_{l}=2v^{\prime}(0). This uses one degree of freedom in the scaling property (2.4). Since ω,v,u\omega,v,u are assumed to be odd functions of xx, we consider the change of variables:

f:=ωx,m:=2clvx,g:=cl+u/xcl+u(0).f:=\frac{\omega}{x},\quad m:=\frac{2}{c_{l}}\cdot\frac{v}{x},\quad g:=\frac{c_{l}+u/x}{c_{l}+u^{\prime}(0)}. (2.6)

Note that f(0)=m(0)=g(0)=1f(0)=m(0)=g(0)=1, and that the out-pushing condition implies g(x)>0g(x)>0 for all xx. Moreover, we define

b:=u(0)=𝑯(xf)(0),c:=cl+u(0)=clb,d:=cl2(cl+u(0))=b+c2c.b:=-u^{\prime}(0)=-\bm{H}(xf)(0),\quad c:=c_{l}+u^{\prime}(0)=c_{l}-b,\quad d:=\frac{c_{l}}{2(c_{l}+u^{\prime}(0))}=\frac{b+c}{2c}. (2.7)

Substituting these changes of variables into (2.1) yields

xgf=(1g)fdf+dm,xgm=2(1g)mxgm,g=1(Δ)1/2(xf)/xbc.\begin{split}&xgf^{\prime}=(1-g)f-df+dm,\\ &xgm^{\prime}=2(1-g)m-xg^{\prime}m,\\ &g=1-\frac{(-\Delta)^{-1/2}(xf)/x-b}{c}.\end{split} (2.8)

From the second equation of (2.8), we easily get

m(x)=1g(x)exp(20x1g(y)yg(y)dy),m(x)=\frac{1}{g(x)}\cdot\exp\left(2\int_{0}^{x}\frac{1-g(y)}{yg(y)}\,\mathrm{d}{y}\right),

meaning that mm can be explicitly determined by gg and hence by ff. Next, we rearrange the first equation of (2.8) to get

f(x)+df(x)x+(d1)1g(x)xg(x)f(x)=dm(x)xg(x).f^{\prime}(x)+d\,\frac{f(x)}{x}+\left(d-1\right)\frac{1-g(x)}{xg(x)}\cdot f(x)=d\,\frac{m(x)}{xg(x)}.

Multiplying both sides of this equation by xdx^{d} yields

(xdf(x))+(d1)1g(x)xg(x)xdf(x)=dxd1m(x)g(x),(x^{d}f(x))^{\prime}+\left(d-1\right)\frac{1-g(x)}{xg(x)}\cdot x^{d}f(x)=d\,x^{d-1}\frac{m(x)}{g(x)},

which leads to

f(x)=xdexp((d1)0xg(y)1yg(y)dy)0x𝑑yd1m(y)g(y)exp((d1)0y1g(z)zg(z)dz)dy.f(x)=x^{-d}\exp\left((d-1)\int_{0}^{x}\frac{g(y)-1}{yg(y)}\,\mathrm{d}{y}\right)\cdot\int_{0}^{x}d\,y^{d-1}\frac{m(y)}{g(y)}\exp\left((d-1)\int_{0}^{y}\frac{1-g(z)}{zg(z)}\,\mathrm{d}{z}\right)\,\mathrm{d}{y}.

Remember that, besides the normalization condition ω(0)=1\omega^{\prime}(0)=1, we still have one degree of freedom in the scaling property (2.4) at our disposal. Later, we will use it to determine the value of cc in (2.7) as a particular functional of ff:

cl+u(0)=c=c(f).c_{l}+u^{\prime}(0)=c=c(f).

In summary, we have obtained the following fixed-point formulation of (2.1):

fb=𝑯(xf)(0),c=c(f),d=b+c2c,g=1(Δ)1/2(xf)/xbc,m=1gexp(20x1gygdy),f=xdexp((d1)0xg1ygdy)0x𝑑yd1mgexp((d1)0y1gzgdz)dy.\begin{split}f\quad\longrightarrow\quad&b=-\bm{H}(xf)(0),\quad c=c(f),\quad d=\frac{b+c}{2c},\\ \longrightarrow\quad&g=1-\frac{(-\Delta)^{-1/2}(xf)/x-b}{c},\\ \longrightarrow\quad&m=\frac{1}{g}\cdot\exp\left(2\int_{0}^{x}\frac{1-g}{yg}\,\mathrm{d}{y}\right),\\ \longrightarrow\quad&f=x^{-d}\exp\left((d-1)\int_{0}^{x}\frac{g-1}{yg}\,\mathrm{d}{y}\right)\int_{0}^{x}d\,y^{d-1}\frac{m}{g}\exp\left((d-1)\int_{0}^{y}\frac{1-g}{zg}\,\mathrm{d}{z}\right)\,\mathrm{d}{y}.\end{split} (2.9)

From the derivation above, it is apparent that a solution of the integral equations (2.9) corresponds to a solution of our main equations (2.1) via the change of variables (2.6), (2.7).

3. Existence of solution by a fixed-point method

Our goal of this section is to show that the nonlinear integral equations (2.9) admit a non-trivial solution. As we can see, the system (2.9) naturally defines a fixed-point problem of a nonlinear nonlocal map. To prove the existence of a suitable fixed-point solution, we need to construct some appropriate function set in a Banach function space on which we can establish continuity and compactness of this nonlinear map, and then we use the Schauder fixed-point theorem.

3.1. A fixed-point problem

Consider a Banach space of continuous even functions,

𝕍={fC():f(x)=f(x),ρfL<+},\mathbb{V}=\{f\in C(\mathbb{R}):\ f(x)=f(-x),\ \|\rho f\|_{L^{\infty}}<+\infty\},

endowed with a weighted LL^{\infty}-norm ρfL\|\rho f\|_{L^{\infty}}, referred to as the LρL^{\infty}_{\rho}-norm, where ρ(x):=(1+|x|)1+δρ\rho(x):=(1+|x|)^{1+\delta_{\rho}} for some δρ>0\delta_{\rho}>0 to be defined below. In this topology, we introduce a nonempty, closed, and convex subset of 𝕍\mathbb{V}, in which we will prove the existence of a fixed point:

𝔻={f𝕍:\displaystyle\mathbb{D}=\Big{\{}f\in\mathbb{V}:\quad f(x)0,f(0)=1,m1(x)f(x)1 for all x,\displaystyle f(x)\geq 0,\ f(0)=1,\ \text{$m_{1}(x)\leq f(x)\leq 1$ for all $x$},
f(1)η,f(x) is non-increasing on [0,+) ,f(s) is convex in s,\displaystyle f^{\prime}_{-}(1)\leq-\eta,\ \text{$f(x)$ is non-increasing on $[0,+\infty)$ },\ \text{$f(\sqrt{s})$ is convex in $s$},
f(x)min{(1+3/δ1)(|x|/L1)1δ1, 5|x|δ0} }.\displaystyle\text{$f(x)\leq\min\big{\{}(1+3/\delta_{1})(|x|/L_{1})^{-1-\delta_{1}}\,,\,5|x|^{-\delta_{0}}\big{\}}$ }\Big{\}}.

Here and below, ff^{\prime}_{-} and f+f^{\prime}_{+} denote the left and the right derivatives of ff, respectively. For reasons that will become clear later, we set η:=1/(3112142)\eta:=1/(3^{11}\cdot 2^{14}\sqrt{2}), δ0:=54η/(8+27η)\delta_{0}:=54\eta/(8+27\eta), and

m1(x):=1g1(x)exp(20x1g1(y)yg1(y)dy),m_{1}(x):=\frac{1}{g_{1}(x)}\exp\left(2\int_{0}^{x}\frac{1-g_{1}(y)}{yg_{1}(y)}\,\mathrm{d}{y}\right), (3.1)

with

g1(x):=min{1+x22, 1+3L04η|x|}andL0:=0+|tln|1+t1t|2|dt<+.g_{1}(x):=\min\left\{1+\frac{x^{2}}{2}\,,\,1+\frac{3L_{0}}{4\eta}|x|\right\}\quad\text{and}\quad L_{0}:=\int_{0}^{+\infty}\left|t\ln\left|\frac{1+t}{1-t}\right|-2\right|\,\mathrm{d}{t}<+\infty.

Moreover, L1>0L_{1}>0 is an absolute constant that will be determined implicitly by δ0\delta_{0} and m1m_{1} in Lemma 3.15, and δ1>0\delta_{1}>0 is also an absolute constant that will be determined explicitly by the function m1m_{1} in Corollary 3.16. Once δ1\delta_{1} is determined, we choose δρ=δ1/2\delta_{\rho}=\delta_{1}/2 so that 𝔻\mathbb{D} is closed and bounded in the LρL^{\infty}_{\rho}-norm with ρ(x)=(1+|x|)1+δρ\rho(x)=(1+|x|)^{1+\delta_{\rho}}. It is also apparent that 𝔻\mathbb{D} is a convex set. Later, we will argue that 𝔻\mathbb{D} is nonempty. In fact, the function m1m_{1} defined above belongs to 𝔻\mathbb{D} by our choice of the constants.

One can check that f𝔻f\in\mathbb{D} implies

(1+x2/2)2f(x)max{1ηx2/2, 1η/2}<1,for x>0.(1+x^{2}/2)^{-2}\leq f(x)\leq\max\{1-\eta x^{2}/2\,,\,1-\eta/2\}<1,\quad\text{for $x>0$}. (3.2)

The lower bound above owes to the fact that m1(x)(1+x2/2)2m_{1}(x)\geq(1+x^{2}/2)^{-2}. Indeed, we have

m1(x)=1g1(x)exp(20x1g1(y)yg1(y)dy)11+x2/2exp(0x2y2+y2dy)=4(2+x2)2.m_{1}(x)=\frac{1}{g_{1}(x)}\exp\left(2\int_{0}^{x}\frac{1-g_{1}(y)}{yg_{1}(y)}\,\mathrm{d}{y}\right)\geq\frac{1}{1+x^{2}/2}\exp\left(-\int_{0}^{x}\frac{2y}{2+y^{2}}\,\mathrm{d}{y}\right)=\frac{4}{(2+x^{2})^{2}}. (3.3)

The upper bound of ff in (3.2) follows from the assumptions that f(s)f(\sqrt{s}) is convex in ss, f(1)ηf^{\prime}_{-}(1)\leq-\eta, and f(x)f(x) is non-increasing on [0,+)[0,+\infty).

We remark that, though a function f𝔻f\in\mathbb{D} is not required to be differentiable, the one-sided derivatives f(x)f^{\prime}_{-}(x) and f+(x)f^{\prime}_{+}(x) are both well defined at every point xx by the convexity of f(s)f(\sqrt{s}) in ss. In what follows, we will abuse notation and simply use f(x)f^{\prime}(x) for f(x)f^{\prime}_{-}(x) and f+(x)f^{\prime}_{+}(x) in both weak sense and strong sense. For example, when we write f(x)Cf^{\prime}(x)\leq C, we mean f(x)Cf^{\prime}_{-}(x)\leq C and f+(x)Cf^{\prime}_{+}(x)\leq C at the same time. In this context, the non-increasing property of ff on [0,+)[0,+\infty) can be represented as f(x)0f^{\prime}(x)\leq 0 for x0x\geq 0.

Now, we formally construct our nonlinear map in a few steps, guided by the fixed-point problem (2.9). We first define a linear map

𝑻(f)(x):=1π0+f(y)(yxln|x+yxy|2)dy.\bm{T}(f)(x):=\frac{1}{\pi}\int_{0}^{+\infty}f(y)\left(\frac{y}{x}\ln\left|\frac{x+y}{x-y}\right|-2\right)\,\mathrm{d}{y}.

Comparing this definition with (2.2), one finds that

𝑻(f)(x)=1x(Δ)1/2(xf)(x)+𝑯(xf)(0)=1x(Δ)1/2(xf)(x)b(f),\bm{T}(f)(x)=\frac{1}{x}(-\Delta)^{-1/2}(xf)(x)+\bm{H}(xf)(0)=\frac{1}{x}(-\Delta)^{-1/2}(xf)(x)-b(f), (3.4)

where

b(f):=2π0+f(y)dy.b(f):=\frac{2}{\pi}\int_{0}^{+\infty}f(y)\,\mathrm{d}{y}. (3.5)

Moreover, it is not hard to check that, for f𝔻f\in\mathbb{D},

limx+𝑻(f)(x)=2π0+(f(+)f(y))dy=2π0+f(y)dy=b(f).\lim_{x\rightarrow+\infty}\bm{T}(f)(x)=\frac{2}{\pi}\int_{0}^{+\infty}(f(+\infty)-f(y))\,\mathrm{d}{y}=-\frac{2}{\pi}\int_{0}^{+\infty}f(y)\,\mathrm{d}{y}=-b(f). (3.6)

Corresponding to the second line of (2.9), we define

𝑮(f):=1𝑻(f)c(f),\bm{G}(f):=1-\frac{\bm{T}(f)}{c(f)},

where we set

c(f):=43π0+f(y)ydy=43π0+f(0)f(y)y2dy.c(f):=-\frac{4}{3\pi}\int_{0}^{+\infty}\frac{f^{\prime}(y)}{y}\,\mathrm{d}{y}=\frac{4}{3\pi}\int_{0}^{+\infty}\frac{f(0)-f(y)}{y^{2}}\,\mathrm{d}{y}. (3.7)

As mentioned above, this is using the second degree of freedom in the scaling property (2.4) to determine the value of cc in (2.7) in terms of ff. It will be clear in the proof of Corollary 3.6 below that we choose to define c(f)c(f) in this way so that limx0𝑮(f)(x)/x=1\lim_{x\rightarrow 0}\bm{G}(f)^{\prime}(x)/x=1. Note that c(f)c(f) must be strictly positive and finite for any f𝔻f\in\mathbb{D}. Next, in view of the third line of (2.9), we define

𝑴(f)(x)=1𝑮(f)(x)exp(20x1𝑮(f)(y)y𝑮(f)(y)dy).\bm{M}(f)(x)=\frac{1}{\bm{G}(f)(x)}\exp\left(2\int_{0}^{x}\frac{1-\bm{G}(f)(y)}{y\bm{G}(f)(y)}\,\mathrm{d}{y}\right).

Finally, letting

d(f):=c(f)+b(f)2c(f)=12+b(f)2c(f),d(f):=\frac{c(f)+b(f)}{2c(f)}=\frac{1}{2}+\frac{b(f)}{2c(f)}, (3.8)

we define

𝑹(f)(x)=0xd(f)yd(f)1𝑴(f)(y)𝑮(f)(y)exp((d(f)1)0y1𝑮(f)(z)z𝑮(f)(z)dx)dyxd(f)exp((d(f)1)0x1𝑮(f)(y)y𝑮(f)(y)dy).\bm{R}(f)(x)=\frac{\displaystyle\int_{0}^{x}d(f)y^{d(f)-1}\frac{\bm{M}(f)(y)}{\bm{G}(f)(y)}\exp\left(\big{(}d(f)-1\big{)}\int_{0}^{y}\frac{1-\bm{G}(f)(z)}{z\bm{G}(f)(z)}\,\mathrm{d}{x}\right)\,\mathrm{d}{y}}{\displaystyle x^{d(f)}\exp\left(\big{(}d(f)-1\big{)}\int_{0}^{x}\frac{1-\bm{G}(f)(y)}{y\bm{G}(f)(y)}\,\mathrm{d}{y}\right)}.

The rest of this paper aims to study the fixed-point problem

𝑹(f)=f,f𝔻.\bm{R}(f)=f,\quad f\in\mathbb{D}.

The following proposition explains how a fixed-point of 𝑹\bm{R} relates to a solution of (2.1).

Proposition 3.1.

If f𝔻f\in\mathbb{D} is a fixed point of 𝐑\bm{R}, i.e. 𝐑(f)=f\bm{R}(f)=f, then ff is a solution to the integral equations (2.9) with g=𝐆(f)g=\bm{G}(f), m=𝐌(f)m=\bm{M}(f), and

b=b(f)=2π0+f(y)dy,c=c(f)=43π0+1f(y)y2dy.b=b(f)=\frac{2}{\pi}\int_{0}^{+\infty}f(y)\,\mathrm{d}{y},\quad c=c(f)=\frac{4}{3\pi}\int_{0}^{+\infty}\frac{1-f(y)}{y^{2}}\,\mathrm{d}{y}.

As a consequence, (ω,v,cl,cω)(\omega,v,c_{l},c_{\omega}) is a solution to equations (2.1) with ω=xf\omega=xf, v=clx𝐌(f)/2v=c_{l}x\bm{M}(f)/2, and

cl=b(f)+c(f),cω=c(f)b(f)2.c_{l}=b(f)+c(f),\quad c_{\omega}=\frac{c(f)-b(f)}{2}. (3.9)
Proof.

Denote r=𝑹(f)r=\bm{R}(f), m=𝑴(f)m=\bm{M}(f), g=𝑮(f)g=\bm{G}(f), b=b(f)b=b(f), c=c(f)c=c(f), and d=d(f)d=d(f). The first statement follows directly from the construction of these maps. Moreover, it is straightforward to compute that

xgr=(1g)rdr+dm,xgm=2(1g)mxgm,g=1(Δ)1/2(xf)/xbc.\begin{split}&xgr^{\prime}=(1-g)r-dr+dm,\\ &xgm^{\prime}=2(1-g)m-xg^{\prime}m,\\ &g=1-\frac{(-\Delta)^{-1/2}(xf)/x-b}{c}.\end{split} (3.10)

Hence, if f=rf=r, then (f,g,m,b,c,d)(f,g,m,b,c,d) is a solution of the equations (2.8). The second statement then follows from the change of variables (2.6), (2.7) and the relations in (2.5). ∎

We now explain the design of the set 𝔻\mathbb{D}. The guiding idea is to make 𝔻\mathbb{D} as specific as possible while keeping it compact in the LρL^{\infty}_{\rho}-norm and closed under the map 𝑹\bm{R}, so that we can apply the Schauder fixed-point theorem. The compactness of 𝔻\mathbb{D} is relatively easy, while its closedness under 𝑹\bm{R} takes most of the work. Firstly, the monotonicity in xx and the convexity in x2x^{2}, which provide powerful controls on the functions in 𝔻\mathbb{D}, are magically preserved by the map 𝑹\bm{R}. This is essentially due to the nice properties of the kernel of the linear map 𝑻\bm{T}. Secondly, the monotonicity, the convexity, and some calculations of 𝑻\bm{T} will lead to the control m1(x)𝑹(f)1m_{1}(x)\leq\bm{R}(f)\leq 1. Hence, we can impose this condition on 𝔻\mathbb{D} as well. Thirdly, we need 𝔻L1()\mathbb{D}\subset L^{1}(\mathbb{R}) in order for b(f)b(f) to be well defined on 𝔻\mathbb{D}, which explains the condition f(x)|x|1δ1f(x)\lesssim|x|^{-1-\delta_{1}}. However, to make sure this tail bound is not compromised by 𝑹\bm{R}, we need the weaker estimate f(x)|x|δ0f(x)\lesssim|x|^{-\delta_{0}}, and to further pass this weaker estimate to 𝑹(f)\bm{R}(f) we need to use the condition f(1)ηf^{\prime}_{-}(1)\leq-\eta. Also, f(1)ηf^{\prime}_{-}(1)\leq-\eta leads to the upper bound in (3.2), which further implies c(f)>0c(f)>0, so that 𝑮(f)\bm{G}(f) and d(f)d(f) are both well-defined. Finally, the estimate 𝑹(f)(1)η\bm{R}(f)^{\prime}(1)\leq-\eta can be derived from f(1)ηf^{\prime}_{-}(1)\leq-\eta and the convexity of f(x)f(x) in x2x^{2}, therefore closing the loop.

The remainder of this section is devoted to justifying the above arguments and to proving the existence of a fixed point of 𝑹\bm{R} in 𝔻\mathbb{D}.

3.2. Estimates of b(f)b(f) and c(f)c(f)

Recall the functionals b(f)b(f), c(f)c(f), d(f)d(f) defined in (3.5), (3.7), (3.8), respectively, which will be constantly used throughout this paper. We start with some estimates on b(f)b(f) and c(f)c(f) that will be used frequently in what follows.

Lemma 3.2.

For any f𝔻f\in\mathbb{D},

22<b(m1)b(f)1π(L1+4δ12),\frac{\sqrt{2}}{2}<b(m_{1})\leq b(f)\leq\frac{1}{\pi}\left(L_{1}+\frac{4}{\delta_{1}^{2}}\right),

and

4η3πc(f)c(m1)<22.\frac{4\eta}{3\pi}\leq c(f)\leq c(m_{1})<\frac{\sqrt{2}}{2}.

As a consequence,

1<d(m1)d(f)=12+b(f)2c(f)12+32η(L14+1δ12)<+.1<d(m_{1})\leq d(f)=\frac{1}{2}+\frac{b(f)}{2c(f)}\leq\frac{1}{2}+\frac{3}{2\eta}\left(\frac{L_{1}}{4}+\frac{1}{\delta_{1}^{2}}\right)<+\infty.
Proof.

Since f(x)min{1,(1+3/δ1)(x/L1)1δ1}f(x)\leq\min\{1\,,\,(1+3/\delta_{1})(x/L_{1})^{-1-\delta_{1}}\}, we can upper bound b(f)b(f) as

b(f)1π0L11dx+(1+3δ1)1πL1+(xL1)1δ1dx=1π(L1+1δ1+3δ12)1π(L1+4δ12).b(f)\leq\frac{1}{\pi}\int_{0}^{L_{1}}1\,\mathrm{d}{x}+\left(1+\frac{3}{\delta_{1}}\right)\frac{1}{\pi}\int_{L_{1}}^{+\infty}\left(\frac{x}{L_{1}}\right)^{-1-\delta_{1}}\,\mathrm{d}{x}=\frac{1}{\pi}\left(L_{1}+\frac{1}{\delta_{1}}+\frac{3}{\delta_{1}^{2}}\right)\leq\frac{1}{\pi}\left(L_{1}+\frac{4}{\delta_{1}^{2}}\right).

In view of (3.2), we can lower bound c(f)c(f) as

c(f)=43π0+1f(y)y2dy43π01η2dy+43π1+η2y2dy=4η3π.c(f)=\frac{4}{3\pi}\int_{0}^{+\infty}\frac{1-f(y)}{y^{2}}\,\mathrm{d}{y}\geq\frac{4}{3\pi}\int_{0}^{1}\frac{\eta}{2}\,\mathrm{d}{y}+\frac{4}{3\pi}\int_{1}^{+\infty}\frac{\eta}{2y^{2}}\,\mathrm{d}{y}=\frac{4\eta}{3\pi}.

By the definition of b(f)b(f) and c(f)c(f), it is easy to see that b(f)b(m1)b(f)\geq b(m_{1}) and c(f)c(m1)c(f)\leq c(m_{1}) for f𝔻f\in\mathbb{D}. Since g1(x)1+x2/2g_{1}(x)\leq 1+x^{2}/2, we have

m1(x)=1g1(x)exp(20x1g1(y)yg1(y)dy)11+x2/2exp(0x2y2+y2dy)=4(2+x2)2.m_{1}(x)=\frac{1}{g_{1}(x)}\exp\left(2\int_{0}^{x}\frac{1-g_{1}(y)}{yg_{1}(y)}\,\mathrm{d}{y}\right)\geq\frac{1}{1+x^{2}/2}\exp\left(-\int_{0}^{x}\frac{2y}{2+y^{2}}\,\mathrm{d}{y}\right)=\frac{4}{(2+x^{2})^{2}}.

In particular, by the definition of g1g_{1}, the inequality above is strict when xx is sufficiently large. It then follows that

b(m1)>2π0+4(2+x2)2dy=22,b(m_{1})>\frac{2}{\pi}\int_{0}^{+\infty}\frac{4}{(2+x^{2})^{2}}\,\mathrm{d}{y}=\frac{\sqrt{2}}{2},

and

c(m1)<43π0+4+y2(2+y2)2dy=22.c(m_{1})<\frac{4}{3\pi}\int_{0}^{+\infty}\frac{4+y^{2}}{(2+y^{2})^{2}}\,\mathrm{d}{y}=\frac{\sqrt{2}}{2}.

This proves the lemma. ∎

The next lemma provides an xx-dependent bound for c(f)c(f) that will later be used in the proof of the estimate 𝑹(f)(1)η\bm{R}(f)^{\prime}(1)\leq-\eta.

Lemma 3.3.

For any f𝔻f\in\mathbb{D} and any x>0x>0,

c(f)8(x+1)3πx(1f(x))1/2.c(f)\leq\frac{8(x+1)}{3\pi x}\big{(}1-f(x)\big{)}^{1/2}.
Proof.

Fix an x>0x>0. For 0yx0\leq y\leq x, f(y)max{(1+y2/2)2,f(x)}max{1y2,f(x)}f(y)\geq\max\{(1+y^{2}/2)^{-2},f(x)\}\geq\max\{1-y^{2},f(x)\}, so 1f(y)min{y2,1f(x)}1-f(y)\leq\min\{y^{2},1-f(x)\}. For y>xy>x, the convexity of f(s)f(\sqrt{s}) in ss implies that (1f(y))/y2(1f(x))/x2(1-f(y))/y^{2}\leq(1-f(x))/x^{2}, and so 1f(y)min{y2(1f(x))/x2,1}1-f(y)\leq\min\{y^{2}(1-f(x))/x^{2},1\}. Combining these estimates yields

1f(y)min{y2,1f(x)}+min{y2(1f(x))/x2,1},y0.1-f(y)\leq\min\{y^{2},1-f(x)\}+\min\{y^{2}(1-f(x))/x^{2},1\},\quad y\geq 0.

We thus obtain that

c(f)\displaystyle c(f) 43π0+min{y2,1f(x)}y2dy+43π0+min{y2(1f(x))/x2,1}y2dy\displaystyle\leq\frac{4}{3\pi}\int_{0}^{+\infty}\frac{\min\{y^{2},1-f(x)\}}{y^{2}}\,\mathrm{d}{y}+\frac{4}{3\pi}\int_{0}^{+\infty}\frac{\min\{y^{2}(1-f(x))/x^{2},1\}}{y^{2}}\,\mathrm{d}{y}
8(x+1)3πx(1f(x))1/2,\displaystyle\leq\frac{8(x+1)}{3\pi x}\big{(}1-f(x)\big{)}^{1/2},

which is the desired bound. ∎

We will need the continuity of b(f)b(f) and c(f)c(f) for proving the continuity of 𝑹\bm{R} in the LρL^{\infty}_{\rho}-norm.

Lemma 3.4.

b(f),c(f),d(f):𝔻b(f),c(f),d(f):\mathbb{D}\to\mathbb{R} are all continuous in the LρL^{\infty}_{\rho}-norm. In particular,

|b(f1)b(f2)|f1f2Lρ,|b(f_{1})-b(f_{2})|\lesssim\|f_{1}-f_{2}\|_{L^{\infty}_{\rho}},
|c(f1)c(f2)|f1f2Lρ1/2,|c(f_{1})-c(f_{2})|\lesssim\|f_{1}-f_{2}\|_{L^{\infty}_{\rho}}^{1/2},

and

|d(f1)d(f2)|f1f2Lρ1/2,|d(f_{1})-d(f_{2})|\lesssim\|f_{1}-f_{2}\|_{L^{\infty}_{\rho}}^{1/2},

for any f1,f2𝔻f_{1},f_{2}\in\mathbb{D}.

Proof.

Recall that ρ(x)=(1+|x|)1+δρ\rho(x)=(1+|x|)^{1+\delta_{\rho}}. Denote δ=f1f2Lρ=ρ(f1f2)L\delta=\|f_{1}-f_{2}\|_{L^{\infty}_{\rho}}=\|\rho(f_{1}-f_{2})\|_{L^{\infty}}. Since (1+x2/2)2fi1(1+x^{2}/2)^{-2}\leq f_{i}\leq 1, i=1,2i=1,2, we have

|f1(x)f2(x)|min{x2,δ(1+|x|)1δρ}|f_{1}(x)-f_{2}(x)|\leq\min\left\{x^{2}\,,\,\delta(1+|x|)^{-1-\delta_{\rho}}\right\}

Then,

|b(f1)b(f2)|1π0+|f1(x)f2(x)|dxδ0+(1+|x|)1δρdx=δδρ,|b(f_{1})-b(f_{2})|\lesssim\frac{1}{\pi}\int_{0}^{+\infty}|f_{1}(x)-f_{2}(x)|\,\mathrm{d}{x}\leq\delta\int_{0}^{+\infty}(1+|x|)^{-1-\delta_{\rho}}\,\mathrm{d}{x}=\frac{\delta}{\delta_{\rho}},

and

|c(f1)c(f2)|0+|f1(x)f2(x)|x2dx0δ1dx+δδ+(1+|x|)1δρx2dxδ.|c(f_{1})-c(f_{2})|\lesssim\int_{0}^{+\infty}\frac{|f_{1}(x)-f_{2}(x)|}{x^{2}}\,\mathrm{d}{x}\leq\int_{0}^{\sqrt{\delta}}1\,\mathrm{d}{x}+\delta\int_{\sqrt{\delta}}^{+\infty}\frac{(1+|x|)^{-1-\delta_{\rho}}}{x^{2}}\,\mathrm{d}{x}\lesssim\sqrt{\delta}.

The continuity of d(f)d(f) then follows from the continuity of b(f),c(f)b(f),c(f) and Lemma 3.2. ∎

3.3. Monotonicity and convexity

Next, we derive the key property of 𝑹\bm{R}, that it preserves the monotonicity in xx and convexity in x2x^{2} on [0,+)[0,+\infty) for functions in 𝔻\mathbb{D}. This nice property of 𝑹\bm{R} essentially arises from a kernel analysis of the linear map 𝑻\bm{T} as shown in the following lemma, which is a restatement of its counterpart in [9]. We still provide the proof here for the sake of completeness.

Lemma 3.5.

Given f𝔻f\in\mathbb{D}, 𝐓(f)(x)0\bm{T}(f)^{\prime}(x)\leq 0 on [0,+)[0,+\infty), and 𝐓(f)(s)\bm{T}(f)(\sqrt{s}) is convex in ss.

Proof.

We first show that 𝑻(f)(x)0\bm{T}(f)^{\prime}(x)\leq 0 on (0,+)(0,+\infty). We can use integration by parts to compute that, for x>0x>0,

𝑻(f)(x)=1π0+f(y)(yxln|x+yxy|2)dy=1π0+f(y)y(y2x22xln|x+yxy|y)dy=1π0+f(y)(x2y22xln|x+yxy|+y)dy=1π0+f(y)yF1(x/y)dy,\begin{split}\bm{T}(f)(x)&=\frac{1}{\pi}\int_{0}^{+\infty}f(y)\left(\frac{y}{x}\ln\left|\frac{x+y}{x-y}\right|-2\right)\,\mathrm{d}{y}\\ &=\frac{1}{\pi}\int_{0}^{+\infty}f(y)\cdot\partial_{y}\left(\frac{y^{2}-x^{2}}{2x}\ln\left|\frac{x+y}{x-y}\right|-y\right)\,\mathrm{d}{y}\\ &=\frac{1}{\pi}\int_{0}^{+\infty}f^{\prime}(y)\cdot\left(\frac{x^{2}-y^{2}}{2x}\ln\left|\frac{x+y}{x-y}\right|+y\right)\,\mathrm{d}{y}\\ &=\frac{1}{\pi}\int_{0}^{+\infty}f^{\prime}(y)\cdot yF_{1}(x/y)\,\mathrm{d}{y},\end{split} (3.11)

where the function F1F_{1} is defined in (A.1) in Appendix A.1, and the integration by parts can be justified by the properties of F1F_{1} proved in Lemma A.1. Therefore, we have

𝑻(f)(x)=1π0+f(y)yxF1(x/y)dy=1π0+f(y)F1(x/y)dy0,\bm{T}(f)^{\prime}(x)=\frac{1}{\pi}\int_{0}^{+\infty}f^{\prime}(y)\cdot y\partial_{x}F_{1}(x/y)\,\mathrm{d}{y}=\frac{1}{\pi}\int_{0}^{+\infty}f^{\prime}(y)\cdot F_{1}^{\prime}(x/y)\,\mathrm{d}{y}\leq 0, (3.12)

where the inequality follows from property (3) in Lemma A.1.

Next, we show that 𝑻(f)(s)\bm{T}(f)(\sqrt{s}) is convex in ss. By approximation theory, we may assume that f(s)f(\sqrt{s}) is twice differentiable in ss, so that the convexity of f(s)f(\sqrt{s}) in ss is equivalent to (f(x)/x)0(f^{\prime}(x)/x)^{\prime}\geq 0 for x>0x>0. Continuing the calculations above, we have

𝑻(f)(x)x\displaystyle\frac{\bm{T}(f)^{\prime}(x)}{x} =1π0+f(y)yyx(y2+x22x2ln|x+yxy|yx)dy\displaystyle=\frac{1}{\pi}\int_{0}^{+\infty}\frac{f^{\prime}(y)}{y}\cdot\frac{y}{x}\left(\frac{y^{2}+x^{2}}{2x^{2}}\ln\left|\frac{x+y}{x-y}\right|-\frac{y}{x}\right)\,\mathrm{d}{y}
=1π0+f(y)yy(y4+2x2y23x48x3ln|x+yxy|y34x27y12)dy+43π0+f(y)ydy\displaystyle=\frac{1}{\pi}\int_{0}^{+\infty}\frac{f^{\prime}(y)}{y}\cdot\partial_{y}\left(\frac{y^{4}+2x^{2}y^{2}-3x^{4}}{8x^{3}}\ln\left|\frac{x+y}{x-y}\right|-\frac{y^{3}}{4x^{2}}-\frac{7y}{12}\right)\,\mathrm{d}{y}+\frac{4}{3\pi}\int_{0}^{+\infty}\frac{f^{\prime}(y)}{y}\,\mathrm{d}{y}
=1π0+(f(y)y)(3x42x2y2y48x3ln|x+yxy|+y34x2+7y12)dy+43π0+f(y)ydy\displaystyle=\frac{1}{\pi}\int_{0}^{+\infty}\left(\frac{f^{\prime}(y)}{y}\right)^{\prime}\cdot\left(\frac{3x^{4}-2x^{2}y^{2}-y^{4}}{8x^{3}}\ln\left|\frac{x+y}{x-y}\right|+\frac{y^{3}}{4x^{2}}+\frac{7y}{12}\right)\,\mathrm{d}{y}+\frac{4}{3\pi}\int_{0}^{+\infty}\frac{f^{\prime}(y)}{y}\,\mathrm{d}{y}
=1π0+(f(y)y)yF2(x/y)dy+43π0+f(y)ydy.\displaystyle=\frac{1}{\pi}\int_{0}^{+\infty}\left(\frac{f^{\prime}(y)}{y}\right)^{\prime}\cdot yF_{2}(x/y)\,\mathrm{d}{y}+\frac{4}{3\pi}\int_{0}^{+\infty}\frac{f^{\prime}(y)}{y}\,\mathrm{d}{y}.

where the function F2F_{2} is defined in (A.2) in Appendix A.2, and the integration by parts can be justified by the properties of F2F_{2} proved in Lemma A.2. Therefore,

(𝑻(f)(x)x)=1π0+(f(y)y)yxF2(x/y)dy=1π0+(f(y)y)F2(x/y)dy0,\left(\frac{\bm{T}(f)^{\prime}(x)}{x}\right)^{\prime}=\frac{1}{\pi}\int_{0}^{+\infty}\left(\frac{f^{\prime}(y)}{y}\right)^{\prime}\cdot y\partial_{x}F_{2}(x/y)\,\mathrm{d}{y}=\frac{1}{\pi}\int_{0}^{+\infty}\left(\frac{f^{\prime}(y)}{y}\right)^{\prime}\cdot F_{2}^{\prime}(x/y)\,\mathrm{d}{y}\geq 0,

where the last inequality follows from property (3) in Lemma A.2. This implies the convexity of 𝑻(f)(s)\bm{T}(f)(\sqrt{s}) in ss. ∎

A similar property holds for the map 𝑮\bm{G} as it is just 11 plus a negative multiple of 𝑻\bm{T}. This then leads to a crude but useful universal estimate of 𝑮(f)\bm{G}(f) for f𝔻f\in\mathbb{D}.

Corollary 3.6.

For any f𝔻f\in\mathbb{D}, 𝐆(f)(x)0\bm{G}(f)^{\prime}(x)\geq 0 on [0,+)[0,+\infty), and 𝐆(f)(s)\bm{G}(f)(\sqrt{s}) is concave in ss. As a consequence, 1𝐆(f)(x)min{1+x2/2, 2d(f)}1\leq\bm{G}(f)(x)\leq\min\{1+x^{2}/2\,,\,2d(f)\} for all xx.

Proof.

Write g=𝑮(f)g=\bm{G}(f). The monotonicity of g(x)g(x) and the concavity of g(s)g(\sqrt{s}) follow directly from Lemma 3.5. More precisely, we have g(x)0g^{\prime}(x)\geq 0 and (g(x)/x)0(g(x)^{\prime}/x)^{\prime}\leq 0 for all x0x\geq 0. By monotonicity we immediately have that

1=g(0)g(x)g(+)=1+b(f)c(f)=2d(f).1=g(0)\leq g(x)\leq g(+\infty)=1+\frac{b(f)}{c(f)}=2d(f).

We have used (3.6) for the limit value g(+)g(+\infty).

Next, we show that g(x)1+x2/2g(x)\leq 1+x^{2}/2 for all xx. In fact, we can compute that

limx0g(x)x=1c(f)limx0𝑻(f)(x)x=1c(f)43π0+f(y)ydy=1.\lim_{x\rightarrow 0}\frac{g^{\prime}(x)}{x}=-\frac{1}{c(f)}\lim_{x\rightarrow 0}\frac{\bm{T}(f)^{\prime}(x)}{x}=-\frac{1}{c(f)}\cdot\frac{4}{3\pi}\int_{0}^{+\infty}\frac{f^{\prime}(y)}{y}\,\mathrm{d}{y}=1.

Then, by the concavity of g(s)g(\sqrt{s}) in ss, we have for x>0x>0,

g(x)1x2=g(x)g(0)x2limx+g(x)2x=12,\frac{g(x)-1}{x^{2}}=\frac{g(x)-g(0)}{x^{2}}\leq\lim_{x\rightarrow+\infty}\frac{g^{\prime}(x)}{2x}=\frac{1}{2},

which is the desired bound. ∎

The above properties of 𝑮\bm{G} imply similar properties of 𝑴\bm{M} through some straightforward derivative calculations.

Corollary 3.7.

For any f𝔻f\in\mathbb{D}, 𝐌(f)(x)0\bm{M}(f)^{\prime}(x)\leq 0 on [0,+)[0,+\infty), and 𝐌(f)(s)\bm{M}(f)(\sqrt{s}) is convex in ss.

Proof.

Write m=𝑴(f)m=\bm{M}(f) and g=𝑮(f)g=\bm{G}(f). From (3.10) we find that

m(x)=(g(x)g(x)+2g(x)1xg(x))m(x)0.\displaystyle m^{\prime}(x)=-\left(\frac{g^{\prime}(x)}{g(x)}+2\frac{g(x)-1}{xg(x)}\right)m(x)\leq 0.

We have used g(x)0g^{\prime}(x)\geq 0 and g(x)g(0)=1g(x)\geq g(0)=1.

To prove m(s)m(\sqrt{s}) is convex in ss, we only need to show that (m(x)/x)0(m^{\prime}(x)/x)^{\prime}\geq 0 for x>0x>0. Note that the concavity of g(s)g(\sqrt{s}) in ss implies that, for x>0x>0,

g(x)1x2g(x)2x,\frac{g(x)-1}{x^{2}}\geq\frac{g^{\prime}(x)}{2x},

which further implies

(g(x)1x2)=g(x)x22(g(x)1)x30.\left(\frac{g(x)-1}{x^{2}}\right)^{\prime}=\frac{g^{\prime}(x)}{x^{2}}-\frac{2(g(x)-1)}{x^{3}}\leq 0.

Therefore, for x>0x>0, we can compute

(m(x)x)\displaystyle\left(\frac{m^{\prime}(x)}{x}\right)^{\prime} =((g(x)x)+2(g(x)1x2))m(x)g(x)2(g(x)g(x)+g(x)1xg(x))m(x)x\displaystyle=-\left(\left(\frac{g^{\prime}(x)}{x}\right)^{\prime}+2\left(\frac{g(x)-1}{x^{2}}\right)^{\prime}\right)\cdot\frac{m(x)}{g(x)}-2\left(\frac{g^{\prime}(x)}{g(x)}+\frac{g(x)-1}{xg(x)}\right)\cdot\frac{m^{\prime}(x)}{x}
((g(x)x)+2(g(x)1x2))m(x)g(x)\displaystyle\geq-\left(\left(\frac{g^{\prime}(x)}{x}\right)^{\prime}+2\left(\frac{g(x)-1}{x^{2}}\right)^{\prime}\right)\cdot\frac{m(x)}{g(x)}
0,\displaystyle\geq 0,

as desired. ∎

Finally, we show that the map 𝑹\bm{R} also enjoys the same monotone and convex properties.

Corollary 3.8.

For any f𝔻f\in\mathbb{D}, 𝐑(f)(x)0\bm{R}(f)^{\prime}(x)\leq 0 on [0,+)[0,+\infty), and 𝐑(f)(s)\bm{R}(f)(\sqrt{s}) is convex in ss. Moreover, 𝐑(f)(x)𝐌(f)(x)\bm{R}(f)(x)\geq\bm{M}(f)(x) for all xx.

Proof.

Let r=𝑹(f),m=𝑴(f),g=𝑮(f),d=d(f)r=\bm{R}(f),m=\bm{M}(f),g=\bm{G}(f),d=d(f). For x0x\geq 0, define

A(x):=0x𝑑yd1m(y)g(y)exp((d1)0y1g(z)zg(z)dz)dy,B(x):=xdexp((d1)0x1g(y)yg(y)dy),\begin{split}A(x)&:=\int_{0}^{x}dy^{d-1}\cdot\frac{m(y)}{g(y)}\exp\left(\big{(}d-1\big{)}\int_{0}^{y}\frac{1-g(z)}{zg(z)}\,\mathrm{d}{z}\right)\,\mathrm{d}{y},\\ B(x)&:=x^{d}\exp\left(\big{(}d-1\big{)}\int_{0}^{x}\frac{1-g(y)}{yg(y)}\,\mathrm{d}{y}\right),\end{split} (3.13)

so that r(x)=A(x)/B(x)r(x)=A(x)/B(x). Note that A(x)0,B(x)0A(x)\geq 0,B(x)\geq 0, A(0)=B(0)=0A(0)=B(0)=0, and

r(0)=limx0A(x)B(x)=m(0)g(0)=1.r(0)=\lim_{x\rightarrow 0}\frac{A(x)}{B(x)}=\frac{m(0)}{g(0)}=1.

We can compute that

A(x)=dxd1m(x)g(x)exp((d1)0x1g(y)yg(y)dy)=dm(x)g(x)B(x)x0,A^{\prime}(x)=dx^{d-1}\frac{m(x)}{g(x)}\exp\left(\big{(}d-1\big{)}\int_{0}^{x}\frac{1-g(y)}{yg(y)}\,\mathrm{d}{y}\right)=d\frac{m(x)}{g(x)}\cdot\frac{B(x)}{x}\geq 0,
B(x)=(d+(d1)1g(x)g(x))B(x)x=(11g(x)+dg(x))B(x)x0,B^{\prime}(x)=\left(d+(d-1)\frac{1-g(x)}{g(x)}\right)\cdot\frac{B(x)}{x}=\left(1-\frac{1}{g(x)}+\frac{d}{g(x)}\right)\cdot\frac{B(x)}{x}\geq 0,

and thus

A(x)=dm(x)g(x)+d1B(x).A^{\prime}(x)=\frac{dm(x)}{g(x)+d-1}\cdot B^{\prime}(x).

Note that, by Corollaries 3.6 and 3.7, the function dm(x)/(g(x)+d1)dm(x)/(g(x)+d-1) is non-increasing in xx. Hence,

A(x)=0xdm(y)g(y)+d1B(y)dydm(x)g(x)+d10xB(y)dy=dm(x)g(x)+d1B(x),A(x)=\int_{0}^{x}\frac{dm(y)}{g(y)+d-1}\cdot B^{\prime}(y)\,\mathrm{d}{y}\geq\frac{dm(x)}{g(x)+d-1}\cdot\int_{0}^{x}B^{\prime}(y)\,\mathrm{d}{y}=\frac{dm(x)}{g(x)+d-1}B(x),

that is,

r(x)=A(x)B(x)dm(x)g(x)+d1=A(x)B(x).r(x)=\frac{A(x)}{B(x)}\geq\frac{dm(x)}{g(x)+d-1}=\frac{A^{\prime}(x)}{B^{\prime}(x)}.

We then find that, for x>0x>0,

r(x)=(A(x)B(x))=B(x)B(x)(A(x)B(x)A(x)B(x))0.r^{\prime}(x)=\left(\frac{A(x)}{B(x)}\right)^{\prime}=\frac{B^{\prime}(x)}{B(x)}\left(\frac{A^{\prime}(x)}{B^{\prime}(x)}-\frac{A(x)}{B(x)}\right)\leq 0.

For the record, we also compute that

r(x)=dm(x)(d+g(x)1)r(x)xg(x).r^{\prime}(x)=\frac{dm(x)-(d+g(x)-1)r(x)}{xg(x)}. (3.14)

From this we find that r(0)=(dm(0)g(0))/(1+d)=0r^{\prime}(0)=(dm^{\prime}(0)-g^{\prime}(0))/(1+d)=0.

Next, we show that 𝑹(f)(s)\bm{R}(f)(\sqrt{s}) is convex in ss. To this end, we first show that (r(x)m(x))/x2(r(x)-m(x))/x^{2} is non-increasing in xx in an analogous way. Define

A~(x)=A(x)B(x)m(x),B~(x)=x2B(x).\tilde{A}(x)=A(x)-B(x)m(x),\quad\tilde{B}(x)=x^{2}B(x). (3.15)

Consider the function

r(x)m(x)x2=B(x)r(x)B(x)m(x)x2B(x)=A~(x)B~(x).\frac{r(x)-m(x)}{x^{2}}=\frac{B(x)r(x)-B(x)m(x)}{x^{2}B(x)}=\frac{\tilde{A}(x)}{\tilde{B}(x)}.

We find that for x0x\geq 0,

A~(x)=A(x)B(x)m(x)B(x)m(x)=(g(x)+g(x)1x)B(x)m(x)g(x)0,\tilde{A}^{\prime}(x)=A^{\prime}(x)-B^{\prime}(x)m(x)-B(x)m^{\prime}(x)=\left(g^{\prime}(x)+\frac{g(x)-1}{x}\right)\cdot\frac{B(x)m(x)}{g(x)}\geq 0,

and

B~(x)=2xB(x)+x2B(x)=(3+d1g(x))xB(x)0.\tilde{B}^{\prime}(x)=2xB(x)+x^{2}B^{\prime}(x)=\left(3+\frac{d-1}{g(x)}\right)\cdot xB(x)\geq 0.

These calculations imply A~(x),B~(x)0\tilde{A}(x),\tilde{B}(x)\geq 0 since A~(0)=B~(0)=0\tilde{A}(0)=\tilde{B}(0)=0. It follows that

r(x)m(x)=x2A~(x)B~(x)0.r(x)-m(x)=x^{2}\cdot\frac{\tilde{A}(x)}{\tilde{B}(x)}\geq 0.

From the above we also find that

A~(x)B~(x)=(g(x)x+g(x)1x2)m(x)3g(x)+d1.\frac{\tilde{A}^{\prime}(x)}{\tilde{B}^{\prime}(x)}=\left(\frac{g^{\prime}(x)}{x}+\frac{g(x)-1}{x^{2}}\right)\cdot\frac{m(x)}{3g(x)+d-1}.

By the concavity of g(s)g(\sqrt{s}) in ss (Corollary 3.6), g(x)/xg^{\prime}(x)/x and (g(x)1)/x2(g(x)-1)/x^{2} are both non-increasing in xx. Moreover, the function m(x)/(3g(x)+d1)m(x)/(3g(x)+d-1) is also non-increasing in xx. Hence, A~(x)/B~(x)\tilde{A}^{\prime}(x)/\tilde{B}^{\prime}(x) is non-increasing in xx. This again implies that A~(x)/B~(x)A~(x)/B~(x)\tilde{A}(x)/\tilde{B}(x)\geq\tilde{A}^{\prime}(x)/\tilde{B}^{\prime}(x), and thus, for x>0x>0,

(r(x)m(x)x2)=(A~(x)B~(x))=B~(x)B~(x)(A~(x)B~(x)A~(x)B~(x))0.\left(\frac{r(x)-m(x)}{x^{2}}\right)^{\prime}=\left(\frac{\tilde{A}(x)}{\tilde{B}(x)}\right)^{\prime}=\frac{\tilde{B}^{\prime}(x)}{\tilde{B}(x)}\left(\frac{\tilde{A}^{\prime}(x)}{\tilde{B}^{\prime}(x)}-\frac{\tilde{A}(x)}{\tilde{B}(x)}\right)\leq 0.

Now, we can compute that

(r(x)x)\displaystyle\left(\frac{r^{\prime}(x)}{x}\right)^{\prime} =(d(m(x)r(x))(g(x)1)r(x)x2g(x))\displaystyle=\left(\frac{d(m(x)-r(x))-(g(x)-1)r(x)}{x^{2}g(x)}\right)^{\prime}
=g(x)g(x)2d(m(x)r(x))(g(x)1)r(x)x2dg(x)(r(x)m(x)x2)\displaystyle=-\frac{g^{\prime}(x)}{g(x)^{2}}\cdot\frac{d(m(x)-r(x))-(g(x)-1)r(x)}{x^{2}}-\frac{d}{g(x)}\cdot\left(\frac{r(x)-m(x)}{x^{2}}\right)^{\prime}
(g(x)1x2)r(x)g(x)g(x)1x2g(x)r(x)\displaystyle\qquad-\left(\frac{g(x)-1}{x^{2}}\right)^{\prime}\cdot\frac{r(x)}{g(x)}-\frac{g(x)-1}{x^{2}g(x)}\cdot r^{\prime}(x)
=(g(x)+g(x)1x)r(x)xg(x)dg(x)(r(x)m(x)x2)(g(x)1x2)r(x)g(x)\displaystyle=-\left(g^{\prime}(x)+\frac{g(x)-1}{x}\right)\cdot\frac{r^{\prime}(x)}{xg(x)}-\frac{d}{g(x)}\cdot\left(\frac{r(x)-m(x)}{x^{2}}\right)^{\prime}-\left(\frac{g(x)-1}{x^{2}}\right)^{\prime}\cdot\frac{r(x)}{g(x)}
0.\displaystyle\geq 0.

We have also used Corollary 3.6 for the last inequality. This completes the proof. ∎

We proceed to show that 𝑹\bm{R} preserves the uniform lower bound f(x)m1(x)f(x)\geq m_{1}(x) appearing in the definition of the set 𝔻\mathbb{D}. To this end, we first prove a finer uniform estimate for 𝑮(f)\bm{G}(f) for all f𝔻f\in\mathbb{D} that strictly improves the bound 𝑮(f)(x)1+x2/2\bm{G}(f)(x)\leq 1+x^{2}/2 (in Corollary 3.6) for sufficiently large xx.

Lemma 3.9.

For any f𝔻f\in\mathbb{D} and for all xx,

𝑮(f)(x)g1(x)=min{1+x22, 1+3L04η|x|},\bm{G}(f)(x)\leq g_{1}(x)=\min\left\{1+\frac{x^{2}}{2}\,,\,1+\frac{3L_{0}}{4\eta}|x|\right\},

where

L0:=0+|tln|1+t1t|2|dt<+.L_{0}:=\int_{0}^{+\infty}\left|t\ln\left|\frac{1+t}{1-t}\right|-2\right|\,\mathrm{d}{t}<+\infty.
Proof.

Using f(x)1f(x)\leq 1, we can estimate that

|𝑻(f)(x)|1π0+|yxln|x+yxy|2|dy=xπ0+|tln|1+t1t|2|dt=L0πx.|\bm{T}(f)(x)|\leq\frac{1}{\pi}\int_{0}^{+\infty}\left|\frac{y}{x}\ln\left|\frac{x+y}{x-y}\right|-2\right|\,\mathrm{d}{y}=\frac{x}{\pi}\int_{0}^{+\infty}\left|t\ln\left|\frac{1+t}{1-t}\right|-2\right|\,\mathrm{d}{t}=\frac{L_{0}}{\pi}x.

By Lemma 3.2 we have c(f)4η/3πc(f)\geq 4\eta/3\pi. It follows that, for x0x\geq 0,

g(x)=1𝑻(x)c(f)1+3L04ηx.g(x)=1-\frac{\bm{T}(x)}{c(f)}\leq 1+\frac{3L_{0}}{4\eta}x.

We have shown in Corollary 3.6 that g(x)1+x2/2g(x)\leq 1+x^{2}/2. Hence, g(x)g1(x)g(x)\leq g_{1}(x) for all xx. ∎

Comparing the definition of 𝑴(f)\bm{M}(f) and that of m1m_{1}, we immediately have the following.

Corollary 3.10.

For any f𝔻f\in\mathbb{D},

𝑹(f)(x)𝑴(f)(x)m1(x)=1g1(x)exp(20x1g1(y)yg1(y)dy).\bm{R}(f)(x)\geq\bm{M}(f)(x)\geq m_{1}(x)=\frac{1}{g_{1}(x)}\exp\left(2\int_{0}^{x}\frac{1-g_{1}(y)}{yg_{1}(y)}\,\mathrm{d}{y}\right).
Proof.

Let g=𝑮(f),m=𝑴(f)g=\bm{G}(f),m=\bm{M}(f), r=𝑹(f)r=\bm{R}(f). Since g(x)g1(x)g(x)\leq g_{1}(x) for all xx (Lemma 3.9), we have

m(x)\displaystyle m(x) =1g(x)exp(20x1g(y)yg(y)dy)1g1(x)exp(20x1g1(y)yg1(y)dy)=m1(x).\displaystyle=\frac{1}{g(x)}\exp\left(2\int_{0}^{x}\frac{1-g(y)}{yg(y)}\,\mathrm{d}{y}\right)\geq\frac{1}{g_{1}(x)}\exp\left(2\int_{0}^{x}\frac{1-g_{1}(y)}{yg_{1}(y)}\,\mathrm{d}{y}\right)=m_{1}(x).

Moreover, we have shown in the proof of Corollary 3.8 that r(x)m(x)r(x)\geq m(x) for all xx. This proves the corollary. ∎

Next, for the particular value of η\eta chosen in the definition of 𝔻\mathbb{D}, we will show that 𝑹(f)(1)η\bm{R}(f)^{\prime}(1)\leq-\eta for f𝔻f\in\mathbb{D}, which means the property f(1)ηf^{\prime}(1)\leq-\eta is preserved by 𝑹\bm{R}. To achieve this, we need the following lemma.

Lemma 3.11.

For any f𝔻f\in\mathbb{D}, 𝐆(f)(1)27η/4\bm{G}(f)^{\prime}(1)\geq 27\eta/4. As a consequence, 𝐆(f)(1)1+27η/8\bm{G}(f)(1)\geq 1+27\eta/8.

Proof.

Write g=𝑮(f)g=\bm{G}(f). We first upper bound 𝑻(f)(x)\bm{T}(f)^{\prime}(x) in two ways using some estimates in [9] (more precisely, in the proof of [9, Lemma 3.6]). We restate the detailed calculations below for the reader’s convenience. On the one hand, we can use the calculations in the proof of Lemma 3.5 to get that, for x>0x>0,

𝑻(f)(x)\displaystyle\bm{T}(f)^{\prime}(x) =1π0+f(y)F1(x/y)dy1π0xf(y)F1(x/y)dy\displaystyle=\frac{1}{\pi}\int_{0}^{+\infty}f^{\prime}(y)\cdot F_{1}^{\prime}(x/y)\,\mathrm{d}{y}\leq\frac{1}{\pi}\int_{0}^{x}f^{\prime}(y)\cdot F_{1}^{\prime}(x/y)\,\mathrm{d}{y}
1π0xf(x)xyF1(x/y)dy=xf(x)1π01tF1(1/t)dt=xf(x)2π.\displaystyle\leq\frac{1}{\pi}\int_{0}^{x}\frac{f^{\prime}(x)}{x}\cdot yF_{1}^{\prime}(x/y)\,\mathrm{d}{y}=xf^{\prime}(x)\cdot\frac{1}{\pi}\int_{0}^{1}tF_{1}^{\prime}(1/t)\,\mathrm{d}{t}=\frac{xf^{\prime}(x)}{2\pi}.

We have used the fact that 01tF1(1/t)dt=(4t/3tF2(1/t))|01=1/2\int_{0}^{1}tF_{1}^{\prime}(1/t)\,\mathrm{d}{t}=(4t/3-tF_{2}(1/t))\big{|}_{0}^{1}=1/2 (Lemma A.2). Recall that the special functions F1F_{1} and F2F_{2} are defined in Appendixes A.1 and A.2, respectively. On the other hand, for any 0<z<x0<z<x, we use F1(1/t)4t3/3F_{1}^{\prime}(1/t)\geq 4t^{3}/3 for t[0,1]t\in[0,1] to find that

𝑻(f)(x)\displaystyle\bm{T}(f)^{\prime}(x) 1π0xf(y)4y33x3dy1πzxf(y)4z33x3dy\displaystyle\leq\frac{1}{\pi}\int_{0}^{x}f^{\prime}(y)\cdot\frac{4y^{3}}{3x^{3}}\,\mathrm{d}{y}\leq\frac{1}{\pi}\int_{z}^{x}f^{\prime}(y)\cdot\frac{4z^{3}}{3x^{3}}\,\mathrm{d}{y}
=4z33πx3(f(x)f(z))4z33πx3(f(x)1+z2).\displaystyle=\frac{4z^{3}}{3\pi x^{3}}(f(x)-f(z))\leq\frac{4z^{3}}{3\pi x^{3}}(f(x)-1+z^{2}).

We then choose z=((1f(x))/2)1/2z=\big{(}(1-f(x))/2\big{)}^{1/2} to obtain

𝑻(f)(x)132πx3(1f(x))5/2.\bm{T}(f)^{\prime}(x)\leq-\frac{1}{3\sqrt{2}\pi x^{3}}\cdot(1-f(x))^{5/2}. (3.16)

Combining the two bounds above, we reach

𝑻(f)(x)(132πx3)1/5(1f(x))1/2(x|f(x)|2π)4/5=1π(x482)1/5|f(x)|4/5(1f(x))1/2.-\bm{T}(f)^{\prime}(x)\geq\left(\frac{1}{3\sqrt{2}\pi x^{3}}\right)^{1/5}(1-f(x))^{1/2}\cdot\left(\frac{x|f^{\prime}(x)|}{2\pi}\right)^{4/5}=\frac{1}{\pi}\left(\frac{x}{48\sqrt{2}}\right)^{1/5}|f^{\prime}(x)|^{4/5}(1-f(x))^{1/2}.

It follows that, for x0x\geq 0,

g(x)=𝑻(f)(x)c(f)3x8(1+x)(x482)1/5|f(x)|4/5.g^{\prime}(x)=\frac{-\bm{T}^{\prime}(f)(x)}{c(f)}\geq\frac{3x}{8(1+x)}\cdot\left(\frac{x}{48\sqrt{2}}\right)^{1/5}|f^{\prime}(x)|^{4/5}.

We have used the xx-dependent upper bound of c(f)c(f) in Lemma 3.3. Plugging in x=1x=1 and using f(1)ηf^{\prime}(1)\leq-\eta, we get

g(1)3η4/516(482)1/5=274η.g^{\prime}(1)\geq\frac{3\eta^{4/5}}{16\cdot(48\sqrt{2})^{1/5}}=\frac{27}{4}\eta.

We then use the concavity of g(s)g(\sqrt{s}) in ss to obtain

g(1)1g(1)2=278η.g(1)-1\geq\frac{g^{\prime}(1)}{2}=\frac{27}{8}\eta.

That is, g(1)1+27η/8g(1)\geq 1+27\eta/8. ∎

The above lemma will also be needed when we establish uniform decay bounds for 𝑹(f)\bm{R}(f) in the next subsection. For now, we use Lemma 3.11 to derive the following.

Corollary 3.12.

For any f𝔻f\in\mathbb{D}, 𝐑(f)(1)η\bm{R}(f)^{\prime}(1)\leq-\eta.

Proof.

Let g=𝑮(f),m=𝑴(f)g=\bm{G}(f),m=\bm{M}(f), and r=𝑹(f)r=\bm{R}(f). By Corollary 3.6 we have g(x)1+x2/2g(x)\leq 1+x^{2}/2, and by Corollary 3.10 we have r(x)m(x)m1(x)(1+x2/2)2r(x)\geq m(x)\geq m_{1}(x)\geq(1+x^{2}/2)^{-2}. We then find that, for x[0,1]x\in[0,1],

r(x)\displaystyle r^{\prime}(x) =g(x)1xr(x)g(x)dg(x)r(x)m(x)xg(x)1xr(x)g(x)\displaystyle=-\frac{g(x)-1}{x}\cdot\frac{r(x)}{g(x)}-\frac{d}{g(x)}\frac{r(x)-m(x)}{x}\leq-\frac{g(x)-1}{x}\cdot\frac{r(x)}{g(x)}
g(x)2r(x)g(x)4(2+x2)3g(x).\displaystyle\leq-\frac{g^{\prime}(x)}{2}\cdot\frac{r(x)}{g(x)}\leq-\frac{4}{(2+x^{2})^{3}}\cdot g^{\prime}(x).

It then follows from Lemma 3.11 that

r(1)427g(1)η,r^{\prime}(1)\leq-\frac{4}{27}g^{\prime}(1)\leq-\eta,

as desired. ∎

3.4. Decay estimates

In this subsection, we show that the map 𝑹\bm{R} preserves the uniform decay bounds for f𝔻f\in\mathbb{D}, i.e. 𝑹(f)(x)|x|δ0\bm{R}(f)(x)\lesssim|x|^{-\delta_{0}} and 𝑹(f)(x)|x|1δ1\bm{R}(f)(x)\lesssim|x|^{-1-\delta_{1}}. These estimates will be established in a general framework as presented in the following lemma.

Lemma 3.13.

Given f𝔻f\in\mathbb{D}, if 𝐆(f)(L)1+a\bm{G}(f)(L)\geq 1+a for some L>0L>0 and some a(0,(5d1)/(3d+1)]a\in(0,(5d-1)/(3d+1)], then for all x>0x>0,

exp(0x1𝑮(f)(y)y𝑮(f)(y)dy)(xL)a/(1+a),\exp\left(\int_{0}^{x}\frac{1-\bm{G}(f)(y)}{y\bm{G}(f)(y)}\,\mathrm{d}{y}\right)\leq\left(\frac{x}{L}\right)^{-a/(1+a)},
𝑴(f)(x)(xL)2a/(1+a),\bm{M}(f)(x)\leq\left(\frac{x}{L}\right)^{-2a/(1+a)},

and

𝑹(f)(x)(1+min{3(1+a)(1a)+,12dd1})(xL)2a/(1+a).\bm{R}(f)(x)\leq\left(1+\min\left\{\frac{3(1+a)}{(1-a)_{+}}\,,\,\frac{12d}{d-1}\right\}\right)\cdot\left(\frac{x}{L}\right)^{-2a/(1+a)}.
Proof.

Let r=𝑹(f)r=\bm{R}(f), m=𝑴(f)m=\bm{M}(f), g=𝑮(f)g=\bm{G}(f), and d=d(f)d=d(f). Bear in mind that g(x)g(x) is non-decreasing on [0,+)[0,+\infty), and 1=g(0)g(x)g(+)=2d1=g(0)\leq g(x)\leq g(+\infty)=2d. Define

ψ(x)=exp(0x1g(y)yg(y)dy).\psi(x)=\exp\left(\int_{0}^{x}\frac{1-g(y)}{yg(y)}\,\mathrm{d}{y}\right).

For xLx\geq L, g(x)g(L)1+ag(x)\geq g(L)\geq 1+a, and thus

ψ(x)=1g(x)xg(x)ψ(x)a1+aψ(x)x.\psi^{\prime}(x)=\frac{1-g(x)}{xg(x)}\psi(x)\leq-\frac{a}{1+a}\cdot\frac{\psi(x)}{x}.

This implies that, for xLx\geq L,

ψ(x)ψ(L)(xL)a/(1+a)(xL)a/(1+a).\psi(x)\leq\psi(L)\cdot\left(\frac{x}{L}\right)^{-a/(1+a)}\leq\left(\frac{x}{L}\right)^{-a/(1+a)}.

We have used that ψ(L)ψ(0)=1\psi(L)\leq\psi(0)=1. Note that for x(0,L)x\in(0,L) we also have

ψ(x)1(xL)a/(1+a).\psi(x)\leq 1\leq\left(\frac{x}{L}\right)^{-a/(1+a)}.

Moreover, this implies

m(x)=ψ(x)2g(x)ψ(x)2(xL)2a/(1+a).m(x)=\frac{\psi(x)^{2}}{g(x)}\leq\psi(x)^{2}\leq\left(\frac{x}{L}\right)^{-2a/(1+a)}.

Let A(x),B(x)A(x),B(x) be defined as in (3.13), and let A~(x)\tilde{A}(x) be defined as in (3.15). We can compute that,

B(x)=(1+d1g(x))B(x)x(1+d12d)B(x)x.B^{\prime}(x)=\left(1+\frac{d-1}{g(x)}\right)\cdot\frac{B(x)}{x}\geq\left(1+\frac{d-1}{2d}\right)\cdot\frac{B(x)}{x}.

This implies, for any xy>0x\geq y>0,

B(x)B(y)(xy)(3d1)/2d.\frac{B(x)}{B(y)}\geq\left(\frac{x}{y}\right)^{(3d-1)/2d}.

Moreover, we have

A~(x)=(g(x)+g(x)1x)B(x)m(x)g(x)3(g(x)1)xg(x)B(x)m(x)3B(x)m(x)x.\tilde{A}^{\prime}(x)=\left(g^{\prime}(x)+\frac{g(x)-1}{x}\right)\cdot\frac{B(x)m(x)}{g(x)}\leq\frac{3(g(x)-1)}{xg(x)}\cdot B(x)m(x)\leq\frac{3B(x)m(x)}{x}.

We have used the concavity of g(s)g(\sqrt{s}) in ss. It follows that

A~(x)\displaystyle\tilde{A}(x) 30xB(y)m(y)ydy3B(x)x(3d1)/2d0xm(y)yy(3d1)/2ddy\displaystyle\leq 3\int_{0}^{x}\frac{B(y)m(y)}{y}\,\mathrm{d}{y}\leq\frac{3B(x)}{x^{(3d-1)/2d}}\int_{0}^{x}\frac{m(y)}{y}\cdot y^{(3d-1)/2d}\,\mathrm{d}{y}
3B(x)x(3d1)/2d0x(yL)2a/(1+a)y(d1)/2ddy=3B(x)1+d12d2a1+a(xL)2a/(1+a).\displaystyle\leq\frac{3B(x)}{x^{(3d-1)/2d}}\int_{0}^{x}\left(\frac{y}{L}\right)^{-2a/(1+a)}\cdot y^{(d-1)/2d}\,\mathrm{d}{y}=\frac{3B(x)}{1+\frac{d-1}{2d}-\frac{2a}{1+a}}\cdot\left(\frac{x}{L}\right)^{-2a/(1+a)}.

Note that, since d>1d>1 (Lemma 3.2), for a(0,(5d1)/(3d+1)]a\in(0,(5d-1)/(3d+1)],

1+d12d2a1+amax{(1a)+1+a,d14d}>0,1+\frac{d-1}{2d}-\frac{2a}{1+a}\geq\max\left\{\frac{(1-a)_{+}}{1+a}\,,\,\frac{d-1}{4d}\right\}>0,

We then have

A~(x)min{1+a(1a)+,4dd1}3B(x)(xL)2a/(1+a).\tilde{A}(x)\leq\min\left\{\frac{1+a}{(1-a)_{+}}\,,\,\frac{4d}{d-1}\right\}\cdot 3B(x)\cdot\left(\frac{x}{L}\right)^{-2a/(1+a)}.

and thus

r(x)=A(x)B(x)=m(x)+A~(x)B(x)(1+min{3(1+a)(1a)+,12dd1})(xL)2a/(1+a).r(x)=\frac{A(x)}{B(x)}=m(x)+\frac{\tilde{A}(x)}{B(x)}\leq\left(1+\min\left\{\frac{3(1+a)}{(1-a)_{+}}\,,\,\frac{12d}{d-1}\right\}\right)\cdot\left(\frac{x}{L}\right)^{-2a/(1+a)}.

This completes the proof. ∎

To apply Lemma 3.13, we need to establish 𝑮(f)(L)1+a\bm{G}(f)(L)\geq 1+a for some L,a>0L,a>0. In fact, we have already obtained this type of condition in Lemma 3.11, which leads to the first uniform decay bound for 𝑹(f)\bm{R}(f).

Corollary 3.14.

For any f𝔻f\in\mathbb{D} and for all x0x\geq 0,

𝑴(f)(x)xδ0,\bm{M}(f)(x)\leq x^{-\delta_{0}},

and

𝑹(f)(x)5xδ0,\bm{R}(f)(x)\leq 5x^{-\delta_{0}},

where δ0=54η/(8+27η)\delta_{0}=54\eta/(8+27\eta).

Proof.

By Lemma 3.11 we have g(1)1+27η/8g(1)\geq 1+27\eta/8. We can thus apply Lemma 3.13 with L=1L=1 and a=27η/8a=27\eta/8 to obtain

𝑴(f)(x)(xL)2a/(1+a)=xδ0,\bm{M}(f)(x)\leq\left(\frac{x}{L}\right)^{-2a/(1+a)}=x^{-\delta_{0}},

and

𝑹(f)(x)(1+3(1+a)1a)(xL)2a/(1+a)5xδ0.\bm{R}(f)(x)\leq\left(1+\frac{3(1+a)}{1-a}\right)\cdot\left(\frac{x}{L}\right)^{-2a/(1+a)}\leq 5x^{-\delta_{0}}.

We have used the fact a=27η/8<1/7a=27\eta/8<1/7 so that 3(1+a)/(1a)<43(1+a)/(1-a)<4. ∎

Next, we use the condition f(x)5|x|δ0f(x)\leq 5|x|^{-\delta_{0}} to compute a new pair of (L,a)(L,a) for the assumption of Lemma 3.13, so that we can derive a stronger decay bound for 𝑹(f)\bm{R}(f).

Lemma 3.15.

There is some absolute constant L1>0L_{1}>0 (that is implicitly determined by δ0\delta_{0} and m1m_{1}) such that, for all f𝔻f\in\mathbb{D},

𝑮(f)(L1)1+d(m1).\bm{G}(f)(L_{1})\geq 1+d(m_{1}).
Proof.

For t0t\geq 0, define

φ(t)=2tln|1+t1t|.\varphi(t)=2-t\ln\left|\frac{1+t}{1-t}\right|. (3.17)

It is not hard to show that there is some t0(0,1)t_{0}\in(0,1) such that φ(t0)=0\varphi(t_{0})=0, 0<φ(t)20<\varphi(t)\leq 2 for t[0,t0)t\in[0,t_{0}), and φ(t)<0\varphi(t)<0 for t>t0t>t_{0}. Also note that

0+φ(t)dt=0+(tF1(1/t))dt=limt+tF1(1/t)limt0tF1(1/t)=0,\int_{0}^{+\infty}\varphi(t)\,\mathrm{d}{t}=\int_{0}^{+\infty}\left(tF_{1}(1/t)\right)^{\prime}\,\mathrm{d}{t}=\lim_{t\rightarrow+\infty}tF_{1}(1/t)-\lim_{t\rightarrow 0}tF_{1}(1/t)=0,

where F1(t)F_{1}(t) is defined in (A.1), and the limits above are given in Lemma A.1.

We then find that, for x0x\geq 0,

𝑻(f)(x)\displaystyle-\bm{T}(f)(x) =1π0+f(y)φ(y/x)dy=1π0+(f(y)f(t0x))φ(y/x)dy\displaystyle=\frac{1}{\pi}\int_{0}^{+\infty}f(y)\varphi(y/x)\,\mathrm{d}{y}=\frac{1}{\pi}\int_{0}^{+\infty}(f(y)-f(t_{0}x))\varphi(y/x)\,\mathrm{d}{y}
1π0t0x(f(y)f(t0x))φ(y/x)dy1π0t0x(m1(y)5(t0x)δ0)+φ(y/x)dy=:Φ(x).\displaystyle\geq\frac{1}{\pi}\int_{0}^{t_{0}x}(f(y)-f(t_{0}x))\varphi(y/x)\,\mathrm{d}{y}\geq\frac{1}{\pi}\int_{0}^{t_{0}x}(m_{1}(y)-5(t_{0}x)^{-\delta_{0}})_{+}\varphi(y/x)\,\mathrm{d}{y}=:\Phi(x).

The first inequality above owes to the non-increasing property of ff on [0,+)[0,+\infty), and the second inequality follows from m1(x)f(x)5xδ0m_{1}(x)\leq f(x)\leq 5x^{-\delta_{0}}. Note that for 0yt0x0\leq y\leq t_{0}x,

(m1(y)5(t0x)δ0)+φ(y/x)2m1(y).(m_{1}(y)-5(t_{0}x)^{-\delta_{0}})_{+}\varphi(y/x)\leq 2m_{1}(y).

Hence, we can use the dominated convergence theorem to obtain

limx+Φ(x)=2π0+m1(y)dy=b(m1).\lim_{x\rightarrow+\infty}\Phi(x)=\frac{2}{\pi}\int_{0}^{+\infty}m_{1}(y)\,\mathrm{d}{y}=b(m_{1}).

Recall d(m1)>1d(m_{1})>1 by Lemma 3.2. The above limit implies that, there exist some L1>0L_{1}>0 such that

𝑻(f)(L1)Φ(L1)d(m1)2d(m1)1b(m1),-\bm{T}(f)(L_{1})\geq\Phi(L_{1})\geq\frac{d(m_{1})}{2d(m_{1})-1}\cdot b(m_{1}),

and hence,

𝑮(f)(L1)=1𝑻(f)(L1)c(f)1+d(m1)2d(m1)1b(m1)c(m1)=1+d(m1).\bm{G}(f)(L_{1})=1-\frac{\bm{T}(f)(L_{1})}{c(f)}\geq 1+\frac{d(m_{1})}{2d(m_{1})-1}\cdot\frac{b(m_{1})}{c(m_{1})}=1+d(m_{1}).

The lemma is thus proved. ∎

We then derive a stronger uniform decay bound for 𝑹(f)\bm{R}(f) by again applying Lemma 3.13.

Corollary 3.16.

For any f𝔻f\in\mathbb{D} and for all x0x\geq 0,

exp(0x1𝑮(f)(y)y𝑮(f)(y)dy)(xL1)(1+δ1)/2,\exp\left(\int_{0}^{x}\frac{1-\bm{G}(f)(y)}{y\bm{G}(f)(y)}\,\mathrm{d}{y}\right)\leq\left(\frac{x}{L_{1}}\right)^{-(1+\delta_{1})/2},
𝑴(f)(x)(xL1)1δ1,\bm{M}(f)(x)\leq\left(\frac{x}{L_{1}}\right)^{-1-\delta_{1}},

and

𝑹(f)(x)(1+3δ1)(xL1)1δ1,\bm{R}(f)(x)\leq\left(1+\frac{3}{\delta_{1}}\right)\left(\frac{x}{L_{1}}\right)^{-1-\delta_{1}},

where

δ1:=d(m1)14d(m1)>0.\delta_{1}:=\frac{d(m_{1})-1}{4d(m_{1})}>0.
Proof.

Write d=d(f)d=d(f) and d1=d(m1)d_{1}=d(m_{1}). Since d1>1d_{1}>1 (Lemma 3.2), we have

d15d113d1+1=(3d11)(d11)3d1+1>0.d_{1}-\frac{5d_{1}-1}{3d_{1}+1}=\frac{(3d_{1}-1)(d_{1}-1)}{3d_{1}+1}>0.

Then, by Lemma 3.15, we have

𝑮(f)(L1)1+d11+5d113d1+1=:1+a1,\bm{G}(f)(L_{1})\geq 1+d_{1}\geq 1+\frac{5d_{1}-1}{3d_{1}+1}=:1+a_{1},

where

1<a1=5d113d1+15d13d+1.1<a_{1}=\frac{5d_{1}-1}{3d_{1}+1}\leq\frac{5d-1}{3d+1}.

Therefore, we can apply Lemma 3.13 with L=L1L=L_{1} and a=a1>1a=a_{1}>1 to obtain

exp(0x1𝑮(f)(y)y𝑮(f)(y)dy)(xL)a1/(1+a1)=(xL1)(1+δ1)/2,\exp\left(\int_{0}^{x}\frac{1-\bm{G}(f)(y)}{y\bm{G}(f)(y)}\,\mathrm{d}{y}\right)\leq\left(\frac{x}{L}\right)^{-a_{1}/(1+a_{1})}=\left(\frac{x}{L_{1}}\right)^{-(1+\delta_{1})/2},
𝑴(f)(x)exp(20x1𝑮(f)(y)y𝑮(f)(y)dy)(xL1)1δ1,\bm{M}(f)(x)\leq\exp\left(2\int_{0}^{x}\frac{1-\bm{G}(f)(y)}{y\bm{G}(f)(y)}\,\mathrm{d}{y}\right)\leq\left(\frac{x}{L_{1}}\right)^{-1-\delta_{1}},

and

𝑹(f)(x)(1+12dd1)(xL)2a1/(1+a1)(1+12d1d11)(xL)2a1/(1+a1)=(1+3δ1)(xL1)1δ1,\bm{R}(f)(x)\leq\left(1+\frac{12d}{d-1}\right)\cdot\left(\frac{x}{L}\right)^{-2a_{1}/(1+a_{1})}\leq\left(1+\frac{12d_{1}}{d_{1}-1}\right)\cdot\left(\frac{x}{L}\right)^{-2a_{1}/(1+a_{1})}=\left(1+\frac{3}{\delta_{1}}\right)\left(\frac{x}{L_{1}}\right)^{-1-\delta_{1}},

as desired. ∎

At this point, we are ready to conclude that the set 𝔻\mathbb{D} is nonempty and is closed under 𝑹\bm{R}.

Theorem 3.17.

𝔻\mathbb{D} is nonempty, and 𝐑\bm{R} maps 𝔻\mathbb{D} into itself.

Proof.

The claim that 𝑹\bm{R} maps 𝔻\mathbb{D} into itself follows from Corollaries 3.8, 3.14, 3.16, 3.10, and 3.12. In order to prove 𝔻\mathbb{D} is nonempty, it suffices to show that m1𝔻m_{1}\in\mathbb{D}. Recall the definition (3.1) of m1m_{1}. Firstly, it is nor hard to check that the function g1g_{1} satisfies g1(x)g_{1}(x) is non-decreasing on [0,)[0,\infty) and g1(s)g_{1}(\sqrt{s}) is concave in ss, which implies m1(x)m_{1}(x) is non-increasing on [0,)[0,\infty) and m1(s)m_{1}(\sqrt{s}) is convex in ss (by a similar argument as in the proof of Corollary 3.7). Secondly, a direct derivative calculation shows that m1(1)ηm_{1}^{\prime}(1)\leq-\eta (for that η\eta is very small). Thirdly, through the proofs of preceding lemmas and corollaries, it is straightforward to check that m1(x)min{(1+3/δ1)(|x|/L1)1δ1, 5|x|δ0}m_{1}(x)\leq\min\big{\{}(1+3/\delta_{1})(|x|/L_{1})^{-1-\delta_{1}}\,,\,5|x|^{-\delta_{0}}\big{\}}. The above together imply that m1𝔻m_{1}\in\mathbb{D}, and the theorem is thus proved. ∎

3.5. Continuity

In order to apply the Schauder fixed-point theorem, we also need the continuity of 𝑹\bm{R} on 𝔻\mathbb{D} in the LρL^{\infty}_{\rho} topology. This will rely on the continuity of 𝑻\bm{T} and Lemma 3.4.

Theorem 3.18.

𝑹:𝔻𝔻\bm{R}:\mathbb{D}\mapsto\mathbb{D} is continuous with respect to the LρL^{\infty}_{\rho}-norm.

Proof.

Recall that ρ(x)=(1+|x|)1+δρ\rho(x)=(1+|x|)^{1+\delta_{\rho}} with δρ=δ1/2\delta_{\rho}=\delta_{1}/2. Let f,f¯𝔻f,\bar{f}\in\mathbb{D} be arbitrary, and write r=𝑹(f)r=\bm{R}(f), m=𝑴(f)m=\bm{M}(f), g=𝑮(f)g=\bm{G}(f), d=d(f)d=d(f), r¯=𝑹(f¯)\bar{r}=\bm{R}(\bar{f}), m¯=𝑴(f¯)\bar{m}=\bm{M}(\bar{f}), g¯=𝑮(f¯)\bar{g}=\bm{G}(\bar{f}), d¯=d(f¯)\bar{d}=d(\bar{f}). Suppose that δ:=ff¯Lρ=ρ(ff¯)L\delta:=\|f-\bar{f}\|_{L^{\infty}_{\rho}}=\|\rho(f-\bar{f})\|_{L^{\infty}} is sufficiently small. We will show that ρ(rr¯)Lρ(ff¯)L1/2\|\rho(r-\bar{r})\|_{L^{\infty}}\lesssim\|\rho(f-\bar{f})\|_{L^{\infty}}^{1/2}, where the symbol “\lesssim” only hides a constant that does not depend on f,f¯f,\bar{f}.

For any x0x\geq 0, we have

|𝑻(f)(x)𝑻(f¯)(x)|\displaystyle|\bm{T}(f)(x)-\bm{T}(\bar{f})(x)| =1π|0+(f(y)f¯(y))(yxln|x+yxy|2)dy|\displaystyle=\frac{1}{\pi}\left|\int_{0}^{+\infty}(f(y)-\bar{f}(y))\left(\frac{y}{x}\ln\left|\frac{x+y}{x-y}\right|-2\right)\,\mathrm{d}{y}\right|
δπ0+(1+y)1δρ|yxln|x+yxy|2|dy\displaystyle\leq\frac{\delta}{\pi}\int_{0}^{+\infty}(1+y)^{-1-\delta_{\rho}}\left|\frac{y}{x}\ln\left|\frac{x+y}{x-y}\right|-2\right|\,\mathrm{d}{y}
=δπx0+(1+tx)1δρ|tln|t+1t1|2|dt\displaystyle=\frac{\delta}{\pi}\cdot x\int_{0}^{+\infty}(1+tx)^{-1-\delta_{\rho}}\left|t\ln\left|\frac{t+1}{t-1}\right|-2\right|\,\mathrm{d}{t}
=δπx0+(1+tx)1δρ|φ(t)|dt,\displaystyle=\frac{\delta}{\pi}\cdot x\int_{0}^{+\infty}(1+tx)^{-1-\delta_{\rho}}|\varphi(t)|\,\mathrm{d}{t},

where φ\varphi is defined as in (3.17). It is not hard to show that there is some t0(0,1)t_{0}\in(0,1) such that φ(t0)=0\varphi(t_{0})=0, 0<φ(t)20<\varphi(t)\leq 2 for t[0,t0)t\in[0,t_{0}), and φ(t)<0\varphi(t)<0 for t>t0t>t_{0}. Also note that φL1([0,+))\varphi\in L^{1}([0,+\infty)). We then decompose and estimate the last integral of tt above as

0+(1+tx)1δρ|φ(t)|dt\displaystyle\int_{0}^{+\infty}(1+tx)^{-1-\delta_{\rho}}|\varphi(t)|\,\mathrm{d}{t} =0t0(1+tx)1δρ|φ(t)|dt+t0+(1+tx)1δρ|φ(t)|dt\displaystyle=\int_{0}^{t_{0}}(1+tx)^{-1-\delta_{\rho}}|\varphi(t)|\,\mathrm{d}{t}+\int_{t_{0}}^{+\infty}(1+tx)^{-1-\delta_{\rho}}|\varphi(t)|\,\mathrm{d}{t}
20t0(1+tx)1δρdt+(1+t0x)1δρt0+|φ(t)|dt\displaystyle\leq 2\int_{0}^{t_{0}}(1+tx)^{-1-\delta_{\rho}}\,\mathrm{d}{t}+(1+t_{0}x)^{-1-\delta_{\rho}}\int_{t_{0}}^{+\infty}|\varphi(t)|\,\mathrm{d}{t}
=2xδρ(1(1+t0x)δρ)+(1+t0x)1δρ0+|φ(t)|dt\displaystyle=\frac{2}{x\delta_{\rho}}\left(1-(1+t_{0}x)^{-\delta_{\rho}}\right)+(1+t_{0}x)^{-1-\delta_{\rho}}\int_{0}^{+\infty}|\varphi(t)|\,\mathrm{d}{t}
min{1,x1}.\displaystyle\lesssim\min\{1\,,\,x^{-1}\}.

We then obtain

|𝑻(f)(x)𝑻(f¯)(x)|δmin{x, 1}.|\bm{T}(f)(x)-\bm{T}(\bar{f})(x)|\lesssim\delta\min\{x\,,\,1\}.

A similar argument shows that |𝑻(f)(x)|,|𝑻(f¯)(x)|min{x,1}|\bm{T}(f)(x)|,|\bm{T}(\bar{f})(x)|\lesssim\min\{x,1\}. Combining these estimates with Lemma 3.2 and Lemma 3.4 yields

|g(x)g¯(x)|δmin{x, 1}+δ1/2min{x, 1}δ1/2min{x, 1}.|g(x)-\bar{g}(x)|\lesssim\delta\min\{x\,,\,1\}+\delta^{1/2}\min\{x\,,\,1\}\lesssim\delta^{1/2}\min\{x\,,\,1\}. (3.18)

It then follows that, for any x0x\geq 0,

0x|1g(y)yg(y)1g¯(y)yg¯(y)|dy\displaystyle\int_{0}^{x}\left|\frac{1-g(y)}{yg(y)}-\frac{1-\bar{g}(y)}{y\bar{g}(y)}\right|\,\mathrm{d}{y} =0x|g(y)g¯(y)|yg(y)g¯(y)dyδ1/2min{x,|lnx|}.\displaystyle=\int_{0}^{x}\frac{|g(y)-\bar{g}(y)|}{yg(y)\bar{g}(y)}\,\mathrm{d}{y}\lesssim\delta^{1/2}\min\{x\,,\,|\ln x|\}.

Write

ψ(x)=exp(0x1g(y)yg(y)dy),ψ¯(x)=exp(0x1g¯(y)yg¯(y)dy).\psi(x)=\exp\left(\int_{0}^{x}\frac{1-g(y)}{yg(y)}\,\mathrm{d}{y}\right),\quad\bar{\psi}(x)=\exp\left(\int_{0}^{x}\frac{1-\bar{g}(y)}{y\bar{g}(y)}\,\mathrm{d}{y}\right).

Using Corollary 3.16, we find

|ψ(x)2ψ¯(x)2|\displaystyle|\psi(x)^{2}-\bar{\psi}(x)^{2}| (ψ(x)2+ψ¯(x)2)0x|1g(y)yg(y)1g¯(y)yg¯(y)|dy\displaystyle\lesssim(\psi(x)^{2}+\bar{\psi}(x)^{2})\int_{0}^{x}\left|\frac{1-g(y)}{yg(y)}-\frac{1-\bar{g}(y)}{y\bar{g}(y)}\right|\,\mathrm{d}{y}
min{1,x1δ1}δ1/2min{x,|lnx|}\displaystyle\lesssim\min\{1\,,\,x^{-1-\delta_{1}}\}\cdot\delta^{1/2}\min\{x\,,\,|\ln x|\}
δ1/2min{x,x1δρ}.\displaystyle\lesssim\delta^{1/2}\min\{x\,,\,x^{-1-\delta_{\rho}}\}.

We have used the fact |etes||ts|(et+es)|\mathrm{e}^{t}-\mathrm{e}^{s}|\leq|t-s|(\mathrm{e}^{t}+\mathrm{e}^{s}). This also implies

|m(x)m¯(x)|=|ψ(x)2g(x)ψ¯(x)2g¯(x)|δ1/2min{x,x1δρ}.|m(x)-\bar{m}(x)|=\left|\frac{\psi(x)^{2}}{g(x)}-\frac{\bar{\psi}(x)^{2}}{\bar{g}(x)}\right|\lesssim\delta^{1/2}\min\{x\,,\,x^{-1-\delta_{\rho}}\}. (3.19)

Next, we use (3.14) to obtain

rr¯\displaystyle r^{\prime}-\bar{r}^{\prime} =dm(d+g1)rxgd¯m¯(d¯+g¯1)r¯xg¯\displaystyle=\frac{dm-(d+g-1)r}{xg}-\frac{\bar{d}\bar{m}-(\bar{d}+\bar{g}-1)\bar{r}}{x\bar{g}}
=(1g1g¯)dm(d+g1)rx\displaystyle=\left(\frac{1}{g}-\frac{1}{\bar{g}}\right)\frac{dm-(d+g-1)r}{x}
+1xg¯((dd¯)(mr)+d¯(mm¯)d¯(rr¯)(gg¯)r(g¯1)(rr¯)).\displaystyle\qquad+\frac{1}{x\bar{g}}\big{(}(d-\bar{d})(m-r)+\bar{d}(m-\bar{m})-\bar{d}(r-\bar{r})-(g-\bar{g})r-(\bar{g}-1)(r-\bar{r})\big{)}.

Let h=rr¯h=r-\bar{r}. We can rearrange the display above to obtain

h+d¯+g¯1xg¯h\displaystyle h^{\prime}+\frac{\bar{d}+\bar{g}-1}{x\bar{g}}\cdot h =(1g1g¯)dm(d+g1)rx+1xg¯((dd¯)(mr)+d¯(mm¯)(gg¯)r)\displaystyle=\left(\frac{1}{g}-\frac{1}{\bar{g}}\right)\frac{dm-(d+g-1)r}{x}+\frac{1}{x\bar{g}}\big{(}(d-\bar{d})(m-r)+\bar{d}(m-\bar{m})-(g-\bar{g})r\big{)}
=:J(x).\displaystyle=:J(x).

We multiply both sides of the equation above by xd¯x^{\bar{d}} to obtain

(xd¯h)(d¯1)g¯1g¯xd¯1h=xd¯J.(x^{\bar{d}}h)^{\prime}-(\bar{d}-1)\frac{\bar{g}-1}{\bar{g}}x^{\bar{d}-1}h=x^{\bar{d}}J.

Since g¯1\bar{g}\geq 1 and d¯>1\bar{d}>1, we have (d¯1)(g¯1)/g¯0(\bar{d}-1)(\bar{g}-1)/\bar{g}\geq 0 for x0x\geq 0. It then follows that, for x0x\geq 0,

xd¯|J|\displaystyle x^{\bar{d}}|J| =|(xd¯h)(d¯1)g¯1g¯xd¯1h||(xd¯h)|(d¯1)g¯1g¯xd¯1|h|\displaystyle=\left|(x^{\bar{d}}h)^{\prime}-(\bar{d}-1)\frac{\bar{g}-1}{\bar{g}}x^{\bar{d}-1}h\right|\geq\left|(x^{\bar{d}}h)^{\prime}\right|-(\bar{d}-1)\frac{\bar{g}-1}{\bar{g}}x^{\bar{d}-1}|h|
(xd¯|h|)(d¯1)g¯1g¯xd¯1|h|(xd¯|h|)(d¯1)2d¯12d¯xd¯1|h|.\displaystyle\geq(x^{\bar{d}}|h|)^{\prime}-(\bar{d}-1)\frac{\bar{g}-1}{\bar{g}}x^{\bar{d}-1}|h|\geq(x^{\bar{d}}|h|)^{\prime}-(\bar{d}-1)\frac{2\bar{d}-1}{2\bar{d}}x^{\bar{d}-1}|h|.

We have used g¯(x)g¯(+)=2d¯\bar{g}(x)\leq\bar{g}(+\infty)=2\bar{d} for the last inequality above. Note that the derivative |h||h|^{\prime} is legit in weak sense. We have now reached

(xd¯|h|)+(p¯d¯)xd¯1|h|xd¯|J|,(x^{\bar{d}}|h|)^{\prime}+(\bar{p}-\bar{d})x^{\bar{d}-1}|h|\leq x^{\bar{d}}|J|,

where

p¯=d¯(d¯1)(2d¯1)2d¯=3d¯12d¯(1,d¯).\bar{p}=\bar{d}-\frac{(\bar{d}-1)(2\bar{d}-1)}{2\bar{d}}=\frac{3\bar{d}-1}{2\bar{d}}\in(1,\bar{d}).

It follows that

(xp¯|h(x)|)xp¯|J(x)|,(x^{\bar{p}}|h(x)|)^{\prime}\leq x^{\bar{p}}|J(x)|,

which leads to

|h(x)|xp¯0xyp¯|J(y)|dy.|h(x)|\leq x^{-\bar{p}}\int_{0}^{x}y^{\bar{p}}|J(y)|\,\mathrm{d}{y}.

We then need to estimate |J(x)||J(x)|. Note that, by Corollaries 3.16 and 3.10,

|rm|xmin{1,x2δ1}min{1,x2δρ}.\frac{|r-m|}{x}\lesssim\min\{1\,,\,x^{-2-\delta_{1}}\}\lesssim\min\{1\,,\,x^{-2-\delta_{\rho}}\}.

Using this, Lemma 3.2, Lemma 3.4, Corollary 3.16, and the preceding estimates (3.18) and (3.19), we can bound J(x)J(x) as

|J(x)|δ1/2min{1,x2δρ}.|J(x)|\lesssim\delta^{1/2}\min\{1\,,\,x^{-2-\delta_{\rho}}\}.

Moreover, since f¯𝔻\bar{f}\in\mathbb{D}, d¯d(m1)>1\bar{d}\geq d(m_{1})>1, and thus

p¯=3d¯12d¯5d¯14d¯5d114d1=1+d114d1=1+δ1>1+δρ.\bar{p}=\frac{3\bar{d}-1}{2\bar{d}}\geq\frac{5\bar{d}-1}{4\bar{d}}\geq\frac{5d_{1}-1}{4d_{1}}=1+\frac{d_{1}-1}{4d_{1}}=1+\delta_{1}>1+\delta_{\rho}.

Finally, we have

|r(x)r¯(x)|=|h(x)|\displaystyle|r(x)-\bar{r}(x)|=|h(x)| δ1/2xp¯0xyp¯min{1,y2δρ}dy\displaystyle\lesssim\delta^{1/2}x^{-\bar{p}}\int_{0}^{x}y^{\bar{p}}\min\{1\,,\,y^{-2-\delta_{\rho}}\}\,\mathrm{d}{y}
δ1/2p¯1δρmin{x,x1δρ}δ1/2(1+x)1δρ.\displaystyle\leq\frac{\delta^{1/2}}{\bar{p}-1-\delta_{\rho}}\min\{x\,,\,x^{-1-\delta_{\rho}}\}\lesssim\delta^{1/2}(1+x)^{-1-\delta_{\rho}}.

That is, for all x0x\geq 0,

ρ(x)|r(x)r¯(x)|=(1+x)1+δρ|r(x)r¯(x)|δ1/2.\rho(x)|r(x)-\bar{r}(x)|=(1+x)^{1+\delta_{\rho}}|r(x)-\bar{r}(x)|\lesssim\delta^{1/2}.

Therefore, rr¯Lρδ1/2=ff¯Lρ1/2.\|r-\bar{r}\|_{L^{\infty}_{\rho}}\lesssim\delta^{1/2}=\|f-\bar{f}\|_{L^{\infty}_{\rho}}^{1/2}. This proves the continuity of 𝑹:𝔻𝔻\bm{R}:\mathbb{D}\mapsto\mathbb{D} in the LρL^{\infty}_{\rho}-norm. ∎

3.6. Existence of a fixed point

One last ingredient for establishing existence of a fixed point of 𝑹\bm{R} is the compactness of 𝔻\mathbb{D}.

Lemma 3.19.

The set 𝔻\mathbb{D} is compact with respect to the LρL_{\rho}^{\infty}-norm.

Proof.

For any f𝔻f\in\mathbb{D}, we use convexity and monotonicity to obtain

f(x)2xf(0)f(x)x2min{1,1x2},x>0.-\frac{f^{\prime}(x)}{2x}\leq\frac{f(0)-f(x)}{x^{2}}\leq\min\{1\ ,\frac{1}{x^{2}}\},\quad x>0.

implying that |f(x)|min{2x,2x1}2|f^{\prime}(x)|\leq\min\{2x,2x^{-1}\}\leq 2. Based on this, we show that 𝔻\mathbb{D} is sequentially compact.

Recall ρ(x)=(1+|x|)1+δρ=(1+|x|)1+δ1/2\rho(x)=(1+|x|)^{1+\delta_{\rho}}=(1+|x|)^{1+\delta_{1}/2}. Let {fn}n=1+\{f_{n}\}_{n=1}^{+\infty} be an arbitrary sequence in 𝔻\mathbb{D}. Initialize n0,k=kn_{0,k}=k, k1k\geq 1. For each integer m1m\geq 1, let ϵm=2m\epsilon_{m}=2^{-m} and Xm=C1ϵm2/δ1X_{m}=C_{1}\epsilon_{m}^{-2/\delta_{1}} for some absolute constant C1>1C_{1}>1 that only depends on δ1\delta_{1} and L1L_{1}. We can choose C1C_{1} so that, for all n1n\geq 1 and for all xXmx\geq X_{m},

ρ(x)fn(x)(1+3δ11)(x/L1)1δ1(1+x)1+δ1/2ϵm\rho(x)f_{n}(x)\leq(1+3\delta_{1}^{-1})(x/L_{1})^{-1-\delta_{1}}\cdot(1+x)^{1+\delta_{1}/2}\leq\epsilon_{m}

Furthermore, since |fn(x)|2|f_{n}^{\prime}(x)|\leq 2 on [0,Xm][0,X_{m}], we can apply Ascoli’s theorem to select a sub-sequence {fnm,k}k=1+\{f_{n_{m,k}}\}_{k=1}^{+\infty} of {fnm1,k}k=1+\{f_{n_{m-1,k}}\}_{k=1}^{+\infty} such that ρ(fnm,ifnm,j)L2ϵm\|\rho(f_{n_{m,i}}-f_{n_{m,j}})\|_{L^{\infty}}\leq 2\epsilon_{m} for any i,j1i,j\geq 1. Then the diagonal sub-sequence {fnm,m}m=1+\{f_{n_{m,m}}\}_{m=1}^{+\infty} is a Cauchy sequence in the LρL_{\rho}^{\infty}-norm. This proves that 𝔻\mathbb{D} is sequentially compact. ∎

We are finally ready to prove the existence of a fixed point of 𝑹\bm{R} in 𝔻\mathbb{D}.

Theorem 3.20.

The map 𝐑\bm{R} has a fixed point f𝔻f\in\mathbb{D}, i.e. 𝐑(f)=f\bm{R}(f)=f.

Proof.

By Theorem 3.18 and Lemma 3.19, the nonempty set 𝔻\mathbb{D} is convex, closed and compact in the LρL_{\rho}^{\infty}-norm, and 𝑹\bm{R} continuously maps 𝔻\mathbb{D} into itself. The Schauder fixed-point theorem then guarantees that 𝑹\bm{R} has a fixed point in 𝔻\mathbb{D}. ∎

4. Properties of the fixed-point solution

Throughout this section, we always denote by ff a fixed point of 𝑹\bm{R} in 𝔻\mathbb{D}, i.e. f=𝑹(f)f=\bm{R}(f). We will use this fixed-point relation to derive finer characterizations of ff. As usual, we write g=𝑮(f),m=𝑴(f)g=\bm{G}(f),m=\bm{M}(f) and d=d(f)d=d(f). Provided that f=𝑹(f)f=\bm{R}(f), we have

xgf=(1g)fdf+dm,xgm=2(1g)mxgm.\begin{split}&xgf^{\prime}=(1-g)f-df+dm,\\ &xgm^{\prime}=2(1-g)m-xg^{\prime}m.\end{split} (4.1)

4.1. Regularity

We first show that a fixed point f=𝑹(f)f=\bm{R}(f) is actually infinitely smooth on \mathbb{R}, using the fact that the map 𝑻(f)\bm{T}(f) gains regularity by integration.

Lemma 4.1.

Given f𝔻f\in\mathbb{D}, suppose that b(f)<+b(f)<+\infty. If fHp()f\in H^{p}(\mathbb{R}) for some integer p0p\geq 0, then 𝐓(f)Hp()\bm{T}(f)^{\prime}\in H^{p}(\mathbb{R}).

Proof.

In view of (3.4), we have

(x𝑻(f))=𝑯(xf)b(f)=𝑯(xf)(0)𝑯(xf)=x𝑯(xf)(0)𝑯(xf)x=x𝑯(f).(x\bm{T}(f))^{\prime}=-\bm{H}(xf)-b(f)=\bm{H}(xf)(0)-\bm{H}(xf)=x\cdot\frac{\bm{H}(xf)(0)-\bm{H}(xf)}{x}=-x\bm{H}(f). (4.2)

We have used Lemma A.3 for the last identity above. It follows that

𝑻(f)(x)=1x0xy𝑯(f)(y)dy,\bm{T}(f)(x)=-\frac{1}{x}\int_{0}^{x}y\bm{H}(f)(y)\,\mathrm{d}{y},

and thus

𝑻(f)(x)=𝑯(f)(x)+1x20xy𝑯(f)(y)dy=𝑯(f)(x)+01t𝑯(f)(tx)dt.\bm{T}(f)^{\prime}(x)=-\bm{H}(f)(x)+\frac{1}{x^{2}}\int_{0}^{x}y\bm{H}(f)(y)\,\mathrm{d}{y}=-\bm{H}(f)(x)+\int_{0}^{1}t\bm{H}(f)(tx)\,\mathrm{d}{t}.

Then, for any integer p0p\geq 0, we have

𝑻(f)(p+1)(x)=𝑯(f)(p)(x)+01tp+1𝑯(f)(p)(tx)dt,\bm{T}(f)^{(p+1)}(x)=-\bm{H}(f)^{(p)}(x)+\int_{0}^{1}t^{p+1}\bm{H}(f)^{(p)}(tx)\,\mathrm{d}{t},

which easily implies that

𝑻(f)H˙p+1()Cp𝑯(f)H˙p()=CpfH˙p().\|\bm{T}(f)\|_{\dot{H}^{p+1}(\mathbb{R})}\leq C_{p}\|\bm{H}(f)\|_{\dot{H}^{p}(\mathbb{R})}=C_{p}\|f\|_{\dot{H}^{p}(\mathbb{R})}.

This proves the lemma. ∎

We can then prove the smoothness of f=𝑹(f)f=\bm{R}(f) by induction.

Theorem 4.2.

Let f𝔻f\in\mathbb{D} be a fixed point of 𝐑\bm{R}. Then, f,𝐌(f),(xf),(x𝐌(f))Hp()f,\bm{M}(f),(xf)^{\prime},(x\bm{M}(f))^{\prime}\in H^{p}(\mathbb{R}) for all p0p\geq 0.

Proof.

Write g=𝑮(f)g=\bm{G}(f), m=𝑴(f)m=\bm{M}(f), and d=d(f)d=d(f). Since f𝔻f\in\mathbb{D}, we know fL2()f\in L^{2}(\mathbb{R}). Moreover, by Corollary 3.16, we also have mL2()m\in L^{2}(\mathbb{R}). Now suppose that fHp()f\in H^{p}(\mathbb{R}) for some p0p\geq 0. Lemma 4.1 then implies that gHp()g^{\prime}\in H^{p}(\mathbb{R}). Recall that

m(x)=(g(x)+2g(x)1x)m(x)g(x).m^{\prime}(x)=-\left(g^{\prime}(x)+2\frac{g(x)-1}{x}\right)\frac{m(x)}{g(x)}. (4.3)

Also note that m(x)/g(x)m(0)/g(0)=1m(x)/g(x)\leq m(0)/g(0)=1. We immediately have mHp()m^{\prime}\in H^{p}(\mathbb{R}). Rearranging the first equation of (4.1), we obtain

f(x)+df(x)1x=(dm(x)1x+(d1)g(x)1xf(x)dg(x)1x)1g(x)=:h(x).f^{\prime}(x)+d\,\frac{f(x)-1}{x}=\left(d\,\frac{m(x)-1}{x}+(d-1)\,\frac{g(x)-1}{x}f(x)-d\,\frac{g(x)-1}{x}\right)\frac{1}{g(x)}=:h(x).

Since f,g,mHp()f,g^{\prime},m^{\prime}\in H^{p}(\mathbb{R}), we apparently have hHp(R)h\in H^{p}(R). Multiplying the equation above by xdx^{d} yields

(xd(f(x)1))=xdh(x),\big{(}x^{d}(f(x)-1)\big{)}^{\prime}=x^{d}h(x),

which implies

f(x)1x=xd10xydh(y)dy=01tdh(tx)dt.\frac{f(x)-1}{x}=x^{-d-1}\int_{0}^{x}y^{d}h(y)\,\mathrm{d}{y}=\int_{0}^{1}t^{d}h(tx)\,\mathrm{d}{t}.

This means (f(x)1)/xHp()(f(x)-1)/x\in H^{p}(\mathbb{R}). Then, since

f(x)=h(x)df(x)1x,f^{\prime}(x)=h(x)-d\,\frac{f(x)-1}{x},

we further have fHp()f^{\prime}\in H^{p}(\mathbb{R}). That is, f,mHp+1()f,m\in H^{p+1}(\mathbb{R}). Therefore, we can use induction to show that f,mHp()f,m\in H^{p}(\mathbb{R}) for all p0p\geq 0. Next, we can use (4.3), the above results, and the fact xmL()<+\|xm\|_{L^{\infty}(\mathbb{R})}<+\infty to inductively show that (xm)Hp()(xm)^{\prime}\in H^{p}(\mathbb{R}) for all p0p\geq 0. Moreover, we can compute that

(xf(x))=dm(x)g(x)(d1)f(x)g(x),(xf(x))^{\prime}=d\,\frac{m(x)}{g(x)}-(d-1)\,\frac{f(x)}{g(x)},

which implies (xf)Hp()(xf)^{\prime}\in H^{p}(\mathbb{R}) for all p0p\geq 0. This completes the proof. ∎

4.2. Asymptotic behavior

Next, we study the asymptotic behavior of f=𝑹(f)f=\bm{R}(f) as x+x\rightarrow+\infty. As we will see, both ff and 𝑴(f)\bm{M}(f) actually decay algebraically in the far field with decay rates given explicitly in terms of d(f)d(f). The next lemma explains how the asymptotic behavior of ff relates to that of 𝑻(f)\bm{T}(f).

Lemma 4.3.

Given f𝔻f\in\mathbb{D}, if Cδ:=supxx1+δf(x)<+C_{\delta}:=\sup_{x\in\mathbb{R}}\ x^{1+\delta}f(x)<+\infty for some δ(0,2)\delta\in(0,2), then

supxxδ(𝑻(f)(x)𝑻(f)(+))Cδ.\sup_{x\in\mathbb{R}}\ x^{\delta}\big{(}\bm{T}(f)(x)-\bm{T}(f)(+\infty)\big{)}\lesssim C_{\delta}.

Moreover, if the limit Dδ:=limx+x1+δf(x)D_{\delta}:=\lim_{x\rightarrow+\infty}x^{1+\delta}f(x) exists and is finite, then

limx+xδ(𝑻(f)(x)𝑻(f)(+))=Dδπ0+1tδln|t+1t1|dy.\lim_{x\rightarrow+\infty}x^{\delta}\big{(}\bm{T}(f)(x)-\bm{T}(f)(+\infty)\big{)}=\frac{D_{\delta}}{\pi}\int_{0}^{+\infty}\frac{1}{t^{\delta}}\ln\left|\frac{t+1}{t-1}\right|\,\mathrm{d}{y}.
Proof.

For any x>0x>0, we calculate that

𝑻(f)(x)𝑻(f)(+)\displaystyle\bm{T}(f)(x)-\bm{T}(f)(+\infty) =𝑻(f)(x)+b(f)=1π0+f(y)yxln|x+yxy|dy\displaystyle=\bm{T}(f)(x)+b(f)=\frac{1}{\pi}\int_{0}^{+\infty}f(y)\cdot\frac{y}{x}\ln\left|\frac{x+y}{x-y}\right|\,\mathrm{d}{y}
Cδπ1x0+1yδln|x+yxy|dy=Cδπ1xδ0+1tδln|t+1t1|dtCδxδ.\displaystyle\leq\frac{C_{\delta}}{\pi}\cdot\frac{1}{x}\int_{0}^{+\infty}\frac{1}{y^{\delta}}\ln\left|\frac{x+y}{x-y}\right|\,\mathrm{d}{y}=\frac{C_{\delta}}{\pi}\cdot\frac{1}{x^{\delta}}\int_{0}^{+\infty}\frac{1}{t^{\delta}}\ln\left|\frac{t+1}{t-1}\right|\,\mathrm{d}{t}\lesssim\frac{C_{\delta}}{x^{\delta}}.

We have used the fact that the non-negative function 1tδln|t+1t1|\frac{1}{t^{\delta}}\ln\left|\frac{t+1}{t-1}\right| is integrable on [0,+)[0,+\infty) for any δ(0,2)\delta\in(0,2). This proves the first result.

To prove the second result, we write the integral as

xδ(𝑻(f)(x)𝑻(f)(+))\displaystyle x^{\delta}\big{(}\bm{T}(f)(x)-\bm{T}(f)(+\infty)\big{)} =xδπ0+f(y)yxln|x+yxy|dy\displaystyle=\frac{x^{\delta}}{\pi}\int_{0}^{+\infty}f(y)\cdot\frac{y}{x}\ln\left|\frac{x+y}{x-y}\right|\,\mathrm{d}{y}
=x1+δπ0+f(tx)tln|t+1t1|dt\displaystyle=\frac{x^{1+\delta}}{\pi}\int_{0}^{+\infty}f(tx)\cdot t\ln\left|\frac{t+1}{t-1}\right|\,\mathrm{d}{t}
=1π0+(tx)1+δf(tx)1tδln|t+1t1|dt.\displaystyle=\frac{1}{\pi}\int_{0}^{+\infty}(tx)^{1+\delta}f(tx)\cdot\frac{1}{t^{\delta}}\ln\left|\frac{t+1}{t-1}\right|\,\mathrm{d}{t}.

Since the limit Dδ:=limx+x1+δf(x)D_{\delta}:=\lim_{x\rightarrow+\infty}x^{1+\delta}f(x) exists and is finite, we know Cδ:=supxx1+δf(x)<+C_{\delta}:=\sup_{x\in\mathbb{R}}\ x^{1+\delta}f(x)<+\infty and limx+(tx)1+δf(tx)=Dδ\lim_{x\rightarrow+\infty}(tx)^{1+\delta}f(tx)=D_{\delta} for any fixed t>0t>0. Also note that the function tδln|t+1t1|t^{-\delta}\ln\left|\frac{t+1}{t-1}\right| is absolutely integrable on [0,+)[0,+\infty) for any δ(0,2)\delta\in(0,2). We then use the dominated convergence theorem to obtain

limx+xδ(𝑻(f)(x)𝑻(f)(+))=Dδπ0+1tδln|t+1t1|dy,\lim_{x\rightarrow+\infty}x^{\delta}\big{(}\bm{T}(f)(x)-\bm{T}(f)(+\infty)\big{)}=\frac{D_{\delta}}{\pi}\int_{0}^{+\infty}\frac{1}{t^{\delta}}\ln\left|\frac{t+1}{t-1}\right|\,\mathrm{d}{y},

as claimed. ∎

We can now compute the decay rates of 𝑴(f)\bm{M}(f) and 𝑹(f)\bm{R}(f) based on the asymptotic behavior of 𝑮(f)\bm{G}(f).

Theorem 4.4.

For any f𝔻f\in\mathbb{D}, there are some finite constants Cm,Cr>0C_{m},C_{r}>0 such that

limx+x1+2δd𝑴(f)(x)=Cm,\lim_{x\rightarrow+\infty}x^{1+2\delta_{d}}\bm{M}(f)(x)=C_{m},

and

limx+x1+δd𝑹(f)(x)=Cr,\lim_{x\rightarrow+\infty}x^{1+\delta_{d}}\bm{R}(f)(x)=C_{r},

where

δd=d(f)12d(f).\delta_{d}=\frac{d(f)-1}{2d(f)}. (4.4)

As a consequence, if ff is a fixed point of 𝐑\bm{R} in 𝔻\mathbb{D}, then

limx+x1+δdf(x)=Cr.\lim_{x\rightarrow+\infty}x^{1+\delta_{d}}f(x)=C_{r}.
Proof.

Let r=𝑹(f)r=\bm{R}(f), m=𝑴(f)m=\bm{M}(f), g=𝑮(f)g=\bm{G}(f), and d=d(f)d=d(f). We first show that there is some finite constant C0>0C_{0}>0 such that

limx+x1/2+δdexp(0x1g(y)yg(y)dy)=C0.\lim_{x\rightarrow+\infty}x^{1/2+\delta_{d}}\exp\left(\int_{0}^{x}\frac{1-g(y)}{yg(y)}\,\mathrm{d}{y}\right)=C_{0}. (4.5)

Recall that g(+)=2dg(+\infty)=2d. We can compute

1x1g(y)yg(y)dy=1x1yg(+)g(y)g(+)g(y)dy+1g(+)g(+)1x1ydy=1x1yg(+)g(y)g(+)g(y)dy(12+δd)lnx.\begin{split}\int_{1}^{x}\frac{1-g(y)}{yg(y)}\,\mathrm{d}{y}&=\int_{1}^{x}\frac{1}{y}\cdot\frac{g(+\infty)-g(y)}{g(+\infty)g(y)}\,\mathrm{d}{y}+\frac{1-g(+\infty)}{g(+\infty)}\cdot\int_{1}^{x}\frac{1}{y}\,\mathrm{d}{y}\\ &=\int_{1}^{x}\frac{1}{y}\cdot\frac{g(+\infty)-g(y)}{g(+\infty)g(y)}\,\mathrm{d}{y}-\left(\frac{1}{2}+\delta_{d}\right)\ln x.\end{split} (4.6)

Since fmin{1,x1δ1}f\lesssim\min\{1\,,\,x^{-1-\delta_{1}}\}, we have by Lemma 4.3 that

g(+)g(x)=𝑻(f)(x)𝑻(f)(+)c(f)xδ1.g(+\infty)-g(x)=\frac{\bm{T}(f)(x)-\bm{T}(f)(+\infty)}{c(f)}\lesssim x^{-\delta_{1}}.

Hence, the first term in the last line of (4.6) is finite as x+x\rightarrow+\infty, that is,

1+1yg(+)g(y)g(+)g(y)dy<+.\int_{1}^{+\infty}\frac{1}{y}\cdot\frac{g(+\infty)-g(y)}{g(+\infty)g(y)}\,\mathrm{d}{y}<+\infty.

We then use (4.6) to obtain

limx+x1/2+δdexp(0x1g(y)yg(y)dy)\displaystyle\lim_{x\rightarrow+\infty}x^{1/2+\delta_{d}}\exp\left(\int_{0}^{x}\frac{1-g(y)}{yg(y)}\,\mathrm{d}{y}\right) =limx+exp(011g(y)yg(y)dy+1xg(+)g(y)yg(+)g(y)dy)\displaystyle=\lim_{x\rightarrow+\infty}\exp\left(\int_{0}^{1}\frac{1-g(y)}{yg(y)}\,\mathrm{d}{y}+\int_{1}^{x}\frac{g(+\infty)-g(y)}{yg(+\infty)g(y)}\,\mathrm{d}{y}\right)
=exp(011g(y)yg(y)dy+1+g(+)g(y)yg(+)g(y)dy)\displaystyle=\exp\left(\int_{0}^{1}\frac{1-g(y)}{yg(y)}\,\mathrm{d}{y}+\int_{1}^{+\infty}\frac{g(+\infty)-g(y)}{yg(+\infty)g(y)}\,\mathrm{d}{y}\right)
=:C0,\displaystyle=:C_{0},

as desired. It immediately follows that

limx+x1+2δdm(x)=limx+x1+2δd1g(x)exp(20x1g(y)yg(y)dy)=C022d=:Cm.\lim_{x\rightarrow+\infty}x^{1+2\delta_{d}}m(x)=\lim_{x\rightarrow+\infty}x^{1+2\delta_{d}}\frac{1}{g(x)}\exp\left(2\int_{0}^{x}\frac{1-g(y)}{yg(y)}\,\mathrm{d}{y}\right)=\frac{C_{0}^{2}}{2d}=:C_{m}.

Next, we prove the asymptotic behavior of rr. Let A(x),B(x)A(x),B(x) be defined as in (3.13). Recall that A(x)A(x) is non-decreasing on [0,+)[0,+\infty), and thus A(+)A(+\infty) is well-defined. We then use (4.5) to find that

A(+)\displaystyle A(+\infty) =0+𝑑yd11g(y)2exp((d+1)0y1g(z)zg(z)dz)dy\displaystyle=\int_{0}^{+\infty}dy^{d-1}\cdot\frac{1}{g(y)^{2}}\exp\left(\big{(}d+1\big{)}\int_{0}^{y}\frac{1-g(z)}{zg(z)}\,\mathrm{d}{z}\right)\,\mathrm{d}{y}
0+yd1min{1,y(1/2+δd)(d+1)}dy\displaystyle\lesssim\int_{0}^{+\infty}y^{d-1}\min\{1\,,\,y^{-(1/2+\delta_{d})(d+1)}\}\,\mathrm{d}{y}
01yd1dy+0+y1δddy<+.\displaystyle\lesssim\int_{0}^{1}y^{d-1}\,\mathrm{d}{y}+\int_{0}^{+\infty}y^{-1-\delta_{d}}\,\mathrm{d}{y}<+\infty.

Moreover, we can also use (4.5) to obtain

limx+x1+δdB(x)=limx+xd+1+δdexp((1d)0x1g(y)yg(y)dy)=1C0d1.\lim_{x\rightarrow+\infty}\frac{x^{1+\delta_{d}}}{B(x)}=\lim_{x\rightarrow+\infty}x^{-d+1+\delta_{d}}\exp\left((1-d)\int_{0}^{x}\frac{1-g(y)}{yg(y)}\,\mathrm{d}{y}\right)=\frac{1}{C_{0}^{d-1}}.

Therefore,

limx+x1+δdr(x)\displaystyle\lim_{x\rightarrow+\infty}x^{1+\delta_{d}}r(x) =limx+x1+δdA(x)B(x)=A(+)C0d1=:Cr.\displaystyle=\lim_{x\rightarrow+\infty}x^{1+\delta_{d}}\frac{A(x)}{B(x)}=\frac{A(+\infty)}{C_{0}^{d-1}}=:C_{r}.

This completes the proof. ∎

4.3. Estimates of d(f)d(f)

We have shown that the asymptotic decay rates of ff and 𝑴(f)\bm{M}(f) are given in terms of d(f)d(f), or equivalently, in terms of the ratio b(f)/c(f)b(f)/c(f). What is left undone is to estimate the value of d(f)d(f) for a fixed point f=𝑹(f)f=\bm{R}(f). We first prove a useful identity as follows.

Lemma 4.5.

Let f𝔻f\in\mathbb{D} be a fixed point of 𝐑\bm{R}. Then,

(b(f)c(f))b(f)(b(f)+c(f))b(m)=2Q(f),(b(f)-c(f))b(f)-(b(f)+c(f))b(m)=2Q(f),

where

Q(f):=2π0+xf(x)𝑻(f)(x)dx0.Q(f):=-\frac{2}{\pi}\int_{0}^{+\infty}xf(x)\bm{T}(f)^{\prime}(x)\,\mathrm{d}{x}\geq 0.
Proof.

Rearranging the first equation of (4.1) we get

df+xfdm=(1g)(xf).df+xf^{\prime}-dm=(1-g)(xf)^{\prime}.

Multiply both sides of the equation above by 2/π2/\pi and then integrating them over [0,+)[0,+\infty) yields

db(f)+2π0+xf(x)dxdb(m)=2π0+(1g(x))(xf(x))dx.db(f)+\frac{2}{\pi}\int_{0}^{+\infty}xf^{\prime}(x)\,\mathrm{d}{x}-db(m)=\frac{2}{\pi}\int_{0}^{+\infty}(1-g(x))(xf(x))^{\prime}\,\mathrm{d}{x}.

Since fx1δρf\lesssim x^{-1-\delta_{\rho}}, we can use integration by parts to obtain

2π0+xf(x)dx=2π0+f(x)dx=b(f),\frac{2}{\pi}\int_{0}^{+\infty}xf^{\prime}(x)\,\mathrm{d}{x}=-\frac{2}{\pi}\int_{0}^{+\infty}f(x)\,\mathrm{d}{x}=-b(f),

and

2π0+(1g(x))(xf(x))dx\displaystyle\frac{2}{\pi}\int_{0}^{+\infty}(1-g(x))(xf(x))^{\prime}\,\mathrm{d}{x} =2π0+g(x)xf(x)dx\displaystyle=\frac{2}{\pi}\int_{0}^{+\infty}g^{\prime}(x)\cdot xf(x)\,\mathrm{d}{x}
=1c(f)2π0+(𝑻(f)(x))xf(x)dx=Q(f)c(f).\displaystyle=\frac{1}{c(f)}\cdot\frac{2}{\pi}\int_{0}^{+\infty}(-\bm{T}(f)^{\prime}(x))\cdot xf(x)\,\mathrm{d}{x}=\frac{Q(f)}{c(f)}.

Hence, we obtain

(d1)b(f)db(m)=Q(f)c(f).(d-1)b(f)-db(m)=\frac{Q(f)}{c(f)}.

Substituting d=(b(f)+c(f))/2c(f)d=(b(f)+c(f))/2c(f) in this equation yields the desired identity.

Next, we show that Q(f)0Q(f)\geq 0. In view of (3.4), we can rewrite Q(f)Q(f) as

Q(f)\displaystyle Q(f) =2π0+(1x(Δ)1/2(xf)(x)b(f))xf(x)dx\displaystyle=-\frac{2}{\pi}\int_{0}^{+\infty}\left(\frac{1}{x}(-\Delta)^{-1/2}(xf)(x)-b(f)\right)^{\prime}xf(x)\,\mathrm{d}{x}
=2π0+𝑯(xf)(x)f(x)dx+2π0+(Δ)1/2(xf)(x)f(x)xdx\displaystyle=\frac{2}{\pi}\int_{0}^{+\infty}\bm{H}(xf)(x)\cdot f(x)\,\mathrm{d}{x}+\frac{2}{\pi}\int_{0}^{+\infty}\frac{(-\Delta)^{-1/2}(xf)(x)\cdot f(x)}{x}\,\mathrm{d}{x}
=b(f)22+2π0+(𝑻(f)(x)+b(f))f(x)dx\displaystyle=-\frac{b(f)^{2}}{2}+\frac{2}{\pi}\int_{0}^{+\infty}\big{(}\bm{T}(f)(x)+b(f)\big{)}f(x)\,\mathrm{d}{x}
=b(f)22+2π20+0+f(x)f(y)yxln|x+yxy|dxdy\displaystyle=-\frac{b(f)^{2}}{2}+\frac{2}{\pi^{2}}\int_{0}^{+\infty}\int_{0}^{+\infty}f(x)f(y)\frac{y}{x}\ln\left|\frac{x+y}{x-y}\right|\,\mathrm{d}{x}\,\mathrm{d}{y}
=b(f)22+1π20+0+f(x)f(y)(xy+yx)ln|x+yxy|dxdy\displaystyle=-\frac{b(f)^{2}}{2}+\frac{1}{\pi^{2}}\int_{0}^{+\infty}\int_{0}^{+\infty}f(x)f(y)\left(\frac{x}{y}+\frac{y}{x}\right)\ln\left|\frac{x+y}{x-y}\right|\,\mathrm{d}{x}\,\mathrm{d}{y}
=1π20+0+f(x)f(y)((xy+yx)ln|x+yxy|2)dxdy.\displaystyle=\frac{1}{\pi^{2}}\int_{0}^{+\infty}\int_{0}^{+\infty}f(x)f(y)\left(\left(\frac{x}{y}+\frac{y}{x}\right)\ln\left|\frac{x+y}{x-y}\right|-2\right)\,\mathrm{d}{x}\,\mathrm{d}{y}.

In the third line above, we have used Lemma A.3 to compute that

2π0+𝑯(xf)(x)f(x)dx=1π0+𝑯(xf)(x)xf(x)xdx=12(𝑯(fx)(0))2=b(f)22.\frac{2}{\pi}\int_{0}^{+\infty}\bm{H}(xf)(x)\cdot f(x)\,\mathrm{d}{x}=\frac{1}{\pi}\int_{0}^{+\infty}\frac{\bm{H}(xf)(x)\cdot xf(x)}{x}\,\mathrm{d}{x}=-\frac{1}{2}\left(\bm{H}(fx)(0)\right)^{2}=-\frac{b(f)^{2}}{2}.

Note that

(xy+yx)ln|x+yxy|20,for all x,y0,\left(\frac{x}{y}+\frac{y}{x}\right)\ln\left|\frac{x+y}{x-y}\right|-2\geq 0,\quad\text{for all $x,y\geq 0$},

Therefore, for all f𝔻f\in\mathbb{D},

Q(f)=1π20+0+f(x)f(y)((xy+yx)ln|x+yxy|2)dxdy0.Q(f)=\frac{1}{\pi^{2}}\int_{0}^{+\infty}\int_{0}^{+\infty}f(x)f(y)\left(\left(\frac{x}{y}+\frac{y}{x}\right)\ln\left|\frac{x+y}{x-y}\right|-2\right)\,\mathrm{d}{x}\,\mathrm{d}{y}\geq 0. (4.7)

This concludes the proof. ∎

Combining Lemma 4.5 and a few estimates in Lemma 3.2, we can bound d(f)d(f) from below as follows.

Corollary 4.6.

Let f𝔻f\in\mathbb{D} be a fixed point of 𝐑\bm{R}. Then, b(f)/c(f)1+10/2b(f)/c(f)\geq 1+\sqrt{10}/2 and d(f)1+10/4d(f)\geq 1+\sqrt{10}/4. As a consequence,

δd=d(f)12d(f)108+21014.5298.\delta_{d}=\frac{d(f)-1}{2d(f)}\geq\frac{\sqrt{10}}{8+2\sqrt{10}}\approx\frac{1}{4.5298}.
Proof.

Let m=𝑴(f)m=\bm{M}(f), d=d(f)d=d(f), and k=b(f)/c(f)k=b(f)/c(f). From Lemma 4.5 we find that

(k1)k(k+1)b(m)c(f)=2Q(f)c(f)2.(k-1)k-(k+1)\frac{b(m)}{c(f)}=\frac{2Q(f)}{c(f)^{2}}.

Since m0(x):=(1+x2/2)2m(x)𝑹(f)(x)=f(x)m_{0}(x):=(1+x^{2}/2)^{-2}\leq m(x)\leq\bm{R}(f)(x)=f(x) for all xx, we have

b(m)b(m0)=22=c(m0)c(f).b(m)\geq b(m_{0})=\frac{\sqrt{2}}{2}=c(m_{0})\geq c(f).

Moreover, we can use f(x)m0(x)f(x)\geq m_{0}(x) and the formula (4.7) to find that

Q(f)Q(m0)=2π0+xm0(x)𝑻(m0)(x)dx=162π0+x2(2+x2)4dx=18.Q(f)\geq Q(m_{0})=-\frac{2}{\pi}\int_{0}^{+\infty}xm_{0}(x)\bm{T}(m_{0})^{\prime}(x)\,\mathrm{d}{x}=\frac{16\sqrt{2}}{\pi}\int_{0}^{+\infty}\frac{x^{2}}{(2+x^{2})^{4}}\,\mathrm{d}{x}=\frac{1}{8}.

We have used the straightforward calculation that

𝑻(m0)(x)=22x22+x2.\bm{T}(m_{0})(x)=-\frac{\sqrt{2}}{2}\cdot\frac{x^{2}}{2+x^{2}}.

We thus obtain

(k1)k(k+1)12,(k-1)k-(k+1)\geq\frac{1}{2},

which implies k1+10/2k\geq 1+\sqrt{10}/2. Hence, d=(b(f)+c(f))/2c(f)=(1+k)/21+10/4d=(b(f)+c(f))/2c(f)=(1+k)/2\geq 1+\sqrt{10}/4. ∎

We conclude this section with a formal proof of our main theorem in the introduction.

Proof of Theorem 1.1.

From Proposition 3.1 and Theorem 3.20 we know the self-similar profiles equations (2.1) admit a solution (ω,v,cl,cω)(\omega,v,c_{l},c_{\omega}) with ω=xf,v=clx𝑴(f)/2\omega=xf,v=c_{l}x\bm{M}(f)/2, and

cl=b(f)+c(f),cω=c(f)b(f)2,c_{l}=b(f)+c(f),\quad c_{\omega}=\frac{c(f)-b(f)}{2},

where f𝔻f\in\mathbb{D} is a fixed point of 𝑹\bm{R}. By Lemma 3.2 we know b(f)>c(f)b(f)>c(f), and thus we have cω<0c_{\omega}<0. Moreover, we can compute that

cωcl=b(f)c(f)2(b(f)+c(f))=d(f)12d(f)=δd.-\frac{c_{\omega}}{c_{l}}=\frac{b(f)-c(f)}{2(b(f)+c(f))}=\frac{d(f)-1}{2d(f)}=\delta_{d}.

Since d(f)<+d(f)<+\infty, we have δd<1/2\delta_{d}<1/2; and by Corollary 4.6, we know δd>1/4.53\delta_{d}>1/4.53.

In view of the scaling property (2.4), we can renormalize the solution (by only tuning α\alpha) so that cω=1c_{\omega}=-1. Note that the ratio cω/clc_{\omega}/c_{l} is invariant under such rescaling. Hence, we have the estimate cl(2,4.53)c_{l}\in(2,4.53). Also note that the monotonicity and convexity properties of ω/x\omega/x and v/xv/x are invariant under renormalization. This proves the existence of an exact self-similar solution of the form (1.3) for the HL model (1.1), with the self-similar profiles satisfying properties (1) and (2) in Theorem 1.1. Moreover, (3) the regularity and (4) the asymptotic decay rates of the profiles follow form Theorems 4.2 and 4.4, respectively. This completes the proof. ∎

5. Numerical verification

In this final section, we perform a numerical study to verify and visualize our theoretical results. To this end, we need to obtain numerically accurate self-similar profiles of the HL model.

As a key step in their computer-assisted proof, Chen, Hou, and Huang [3] obtained accurate approximate self-similar profiles by numerically solving the dynamic rescaling equations (2.3) of the HL model with high-order numerical schemes. Owing to the nonlinear stability under the normalization conditions (2.5), the numerical solution (for a large class of initial data) can easily converge to an approximate steady state with extremely small point-wise residual errors. Their computer codes and stored output data can be found in [2].

Instead of using the codes and data off-the-shelf, we construct our own approximate self-similar profiles by numerically solving the fixed-point problem f=𝑹(f)f=\bm{R}(f) using a direct iterative method. That is, starting with some smooth initial function f(0)𝔻f^{(0)}\in\mathbb{D}, we numerically compute

f(n+1)=𝑹(f(n)),n=0,1,2,.f^{(n+1)}=\bm{R}(f^{(n)}),\quad n=0,1,2,.... (5.1)

More precisely, each iteration is computed in the order of (2.9). After the scheme converges numerically, the self-similar profiles ω,v\omega,v and the scaling factors cl,cωc_{l},c_{\omega} are then recovered as in Proposition 3.1 and renormalized as in (2.4). Note that a similar fixed-point method for computing the self-similar profiles of the 1D gCLM model was proposed in [9] by the same authors.

We perform the fixed-point computation for two purposes. The first one is to test the well-posedness of the fixed-point problem f=𝑹(f)f=\bm{R}(f). We have not been able to prove the uniqueness of a fixed-point of 𝑹\bm{R} nor the convergence of the scheme (5.1). Nevertheless, this iterative method converges quickly for a bunch of arbitrarily picked initial data in 𝔻\mathbb{D} with the maximum residual f(n)𝑹(f(n))L\|f^{(n)}-\bm{R}(f^{(n)})\|_{L^{\infty}} dropped below an extremely small tolerance (101010^{-10} in our computations, the same standard as in [3]) only within hundreds of iterations. For example, for the initial data f(0)=(1+x2)1𝔻f^{(0)}=(1+x^{2})^{-1}\in\mathbb{D}, the residual f(n)𝑹(f(n))L\|f^{(n)}-\bm{R}(f^{(n)})\|_{L^{\infty}} decreases below 101010^{-10} within 270 iterations; for the initial data f(0)=(1+x2/2)2f^{(0)}=(1+x^{2}/2)^{-2} that is not in 𝔻\mathbb{D}, the residual f(n)𝑹(f(n))L\|f^{(n)}-\bm{R}(f^{(n)})\|_{L^{\infty}} decreases below 101010^{-10} within 250 iterations. This makes us believe that the fixed-point of 𝑹\bm{R} in 𝔻\mathbb{D} is unique and the scheme (5.1) is convergent over 𝔻\mathbb{D} (and probably convergent for more general initial functions under weaker assumptions).

The second purpose is to check whether the self-similar profiles determined by the fixed-point method in this paper and those obtained by solving the dynamic rescaling equations as in [3] are identical under proper rescaling. It turns out that the numerical profiles obtained by the iterative method (5.1) and those obtained by solving the dynamic rescaling equations (data stored in [2]) under consistent normalization (f(0)=m(0)=1f(0)=m(0)=1) are almost identical up to only mesh-point-wise errors at the level of 10510^{-5}. The errors are likely due to the differences in the discretization methods. This convincingly supports our conjecture.

We remark that the most numerically expensive step in one iteration is to compute the function 𝑻(f(n))\bm{T}(f^{(n)}), which involves the evaluation of a linear transform with a dense kernel at all mesh points. The mesh points are distributed adaptively over a sufficiently large one-sided interval [0,1016][0,10^{16}] (solutions are truncated to 0 for x1016x\geq 10^{16}). An analogous computation (recovering uu from ω\omega) is also needed in every time step when solving the dynamic rescaling equations (2.3) numerically, and it takes more than tens of thousands of time steps for the solution to converge in time. Hence, our method is empirically more efficient than numerically solving the dynamic rescaling equations in obtaining approximate self-similar profiles.

Finally, we provide some plots of the numerically constructed profiles to verify and visualize some of their theoretically proved properties. Figure 5.1 plots the numerically obtained fixed point ff and the corresponding 𝑴(f)\bm{M}(f) in xx and in s=x2s=x^{2} respectively, verifying that they are both monotone decreasing in xx, convex in x2x^{2}, and lower bounded by (1+x2/2)2(1+x^{2}/2)^{-2} for x0x\geq 0. Figure 5.1 plots the corresponding 𝑮(f)\bm{G}(f) in a similar way, verifying that it is monotone increasing in xx, concave in x2x^{2}, and upper bounded by 1+x2/21+x^{2}/2 for x0x\geq 0. Figure 5.3 demonstrates asymptotic decay rates of ff and 𝑴(f)\bm{M}(f) for sufficiently large xx, verifying the statements in Theorem 4.4.

Refer to caption
(a) f(x)f(x) and 𝑴(f)(x)\bm{M}(f)(x)
Refer to caption
(b) f(s)f(\sqrt{s}) and 𝑴(f)(s)\bm{M}(f)(\sqrt{s})
Figure 5.1. The numerically obtained fixed point ff and the corresponding 𝑴(f)\bm{M}(f) plotted (a) in coordinate xx and (b) in coordinate s=x2s=x^{2}. The dashed line represents the lower bound (1+x2/2)2=(1+s/2)2(1+x^{2}/2)^{-2}=(1+s/2)^{-2} for functions in 𝔻\mathbb{D}.
Refer to caption
(a) 𝑮(f)(x)\bm{G}(f)(x)
Refer to caption
(b) 𝑮(f)(s)\bm{G}(f)(\sqrt{s})
Figure 5.2. 𝑮(f)\bm{G}(f) of the numerically obtained fixed point ff plotted (a) in coordinate xx and (b) in coordinate s=x2s=x^{2}. The dashed line represents the upper bound 1+x2/2=1+s/21+x^{2}/2=1+s/2.
Refer to caption
(a) x1+δdf(x)x^{1+\delta_{d}}f(x)
Refer to caption
(b) x1+2δd𝑴(f)(x)x^{1+2\delta_{d}}\bm{M}(f)(x)
Figure 5.3. Demonstrations of algebraic decay rates of (a) the numerically obtained fixed point ff and (b) the corresponding 𝑴(f)\bm{M}(f).

Appendix A Useful facts

A.1. Special function F1F_{1}

We define

F1(t):=t212tln|t+1t1|+1,t0.F_{1}(t):=\frac{t^{2}-1}{2t}\ln\left|\frac{t+1}{t-1}\right|+1,\quad t\geq 0. (A.1)

The derivative of FF reads

F1(t)=t2+12t2ln|t+1t1|1t.F_{1}^{\prime}(t)=\frac{t^{2}+1}{2t^{2}}\ln\left|\frac{t+1}{t-1}\right|-\frac{1}{t}.

For t[0,1)t\in[0,1), F1(t)F_{1}(t) and F1(t)F_{1}^{\prime}(t) have the Taylor expansions

F1(t)=n=12t2n4n21,F1(t)=n=14nt2n14n21.F_{1}(t)=\sum\limits_{n=1}^{\infty}\frac{2t^{2n}}{4n^{2}-1},\quad F_{1}^{\prime}(t)=\sum\limits_{n=1}^{\infty}\frac{4nt^{2n-1}}{4n^{2}-1}.

For t[0,1)t\in[0,1), F1(1/t)F_{1}(1/t) and F1(1/t)F_{1}^{\prime}(1/t) have the Taylor expansions

F1(1/t)=2n=12t2n4n21,F1(1/t)=n=14nt2n+14n21.F_{1}(1/t)=2-\sum\limits_{n=1}^{\infty}\frac{2t^{2n}}{4n^{2}-1},\quad F_{1}^{\prime}(1/t)=\sum\limits_{n=1}^{\infty}\frac{4nt^{2n+1}}{4n^{2}-1}.
Lemma A.1.

The function F1F_{1} defined in (A.1) satisfies

  1. (1)

    F1(1/t)=2F1(t)F_{1}(1/t)=2-F_{1}(t), F1(1/t)=t2F1(t)F_{1}^{\prime}(1/t)=t^{2}F_{1}^{\prime}(t);

  2. (2)

    F1C([0,+))F_{1}\in C([0,+\infty)), F1(0)=0F_{1}(0)=0, F1(1)=1F_{1}(1)=1, limt+F1(t)=2\lim_{t\rightarrow+\infty}F_{1}(t)=2, limt0F1(t)/t=0\lim_{t\rightarrow 0}F_{1}(t)/t=0;

  3. (3)

    F1(0)=0F_{1}^{\prime}(0)=0 and F1(t)>0F_{1}^{\prime}(t)>0 for t>0t>0.

Proof.

Property (11) is straightforward to check. (22) follows from the Taylor expansion of F1(t)F_{1}(t) and property (11). (33) follows from the Taylor expansion of F1(t)F_{1}^{\prime}(t) and property (11). ∎

A.2. Special function F2F_{2}

We define

F2(t):=3t42t218t3ln|t+1t1|+14t2+712,t0.F_{2}(t):=\frac{3t^{4}-2t^{2}-1}{8t^{3}}\ln\left|\frac{t+1}{t-1}\right|+\frac{1}{4t^{2}}+\frac{7}{12},\quad t\geq 0. (A.2)

The derivative of GG reads

F2(t)=3t4+2t2+38t4ln|t+1t1|3t2+34t3.F_{2}^{\prime}(t)=\frac{3t^{4}+2t^{2}+3}{8t^{4}}\ln\left|\frac{t+1}{t-1}\right|-\frac{3t^{2}+3}{4t^{3}}.

For t[0,1)t\in[0,1), F2(t)F_{2}(t) and F2(t)F_{2}^{\prime}(t) have the Taylor expansions

F2(t)=n=1+4(n+1)t2n(2n1)(2n+1)(2n+3),F2(t)=n=1+8n(n+1)t2n1(2n1)(2n+1)(2n+3).F_{2}(t)=\sum\limits_{n=1}^{+\infty}\frac{4(n+1)t^{2n}}{(2n-1)(2n+1)(2n+3)},\quad F_{2}^{\prime}(t)=\sum\limits_{n=1}^{+\infty}\frac{8n(n+1)t^{2n-1}}{(2n-1)(2n+1)(2n+3)}.

For t[0,1)t\in[0,1), F2(1/t)F_{2}(1/t) and F2(1/t)F_{2}^{\prime}(1/t) have the Taylor expansions

F2(1/t)=43n=1+4nt2n+2(2n1)(2n+1)(2n+3),F2(1/t)=n=1+8n(n+1)t2n+3(2n1)(2n+1)(2n+3).F_{2}(1/t)=\frac{4}{3}-\sum\limits_{n=1}^{+\infty}\frac{4nt^{2n+2}}{(2n-1)(2n+1)(2n+3)},\quad F_{2}^{\prime}(1/t)=\sum\limits_{n=1}^{+\infty}\frac{8n(n+1)t^{2n+3}}{(2n-1)(2n+1)(2n+3)}.
Lemma A.2.

The function F2F_{2} defined in (A.2) satisfies

  1. (1)

    F2(1/t)=t4F2(t)F_{2}^{\prime}(1/t)=t^{4}F_{2}^{\prime}(t);

  2. (2)

    F2C([0,+))F_{2}\in C([0,+\infty)), F2(0)=0F_{2}(0)=0, F2(1)=5/6F_{2}(1)=5/6, limt+F2(t)=4/3\lim_{t\rightarrow+\infty}F_{2}(t)=4/3, limt0F2(t)/t=0\lim_{t\rightarrow 0}F_{2}(t)/t=0;

  3. (3)

    F2(t)0F_{2}^{\prime}(t)\geq 0 for t0t\geq 0.

  4. (4)

    (4t/3tF2(1/t))=tF1(1/t)(4t/3-tF_{2}(1/t))^{\prime}=tF_{1}^{\prime}(1/t) for t0t\geq 0.

Proof.

Properties (1)(1) is straightforward to check. (2)(2) follows from the Taylor expansions of F2(t)F_{2}(t) and F2(1/t)F_{2}(1/t). (3)(3) follows from the Taylor expansion of F1(t)F_{1}^{\prime}(t) and property (1)(1). (4)(4) can be checked straightforwardly by the definitions of F2(t)F_{2}(t) and F1(t)F_{1}(t). ∎

A.3. The Hilbert transform

Lemma A.3.

For any suitable function ω\omega on \mathbb{R},

𝑯(ω)(x)𝑯(ω)(0)x=𝑯(ωω(0)x)(x).\frac{\bm{H}(\omega)(x)-\bm{H}(\omega)(0)}{x}=\bm{H}\left(\frac{\omega-\omega(0)}{x}\right)(x).

As a result,

1π𝑯(ω)(x)ω(x)xdx=12ω(0)212(𝑯(ω)(0))2.\frac{1}{\pi}\int_{\mathbb{R}}\frac{\bm{H}(\omega)(x)\cdot\omega(x)}{x}\,\mathrm{d}{x}=\frac{1}{2}\omega(0)^{2}-\frac{1}{2}\big{(}\bm{H}(\omega)(0)\big{)}^{2}.
Proof.

The first equation follows directly from the definition of the Hilbert transform on the real line. The second equation is derived from the first one as follows:

1π𝑯(ω)ωxdx\displaystyle\frac{1}{\pi}\int_{\mathbb{R}}\frac{\bm{H}(\omega)\cdot\omega}{x}\,\mathrm{d}{x} =1π(𝑯(ω)𝑯(ω)(0))xωdx+𝑯(ω)(0)1πωxdx\displaystyle=\frac{1}{\pi}\int_{\mathbb{R}}\frac{(\bm{H}(\omega)-\bm{H}(\omega)(0))}{x}\cdot\omega\,\mathrm{d}{x}+\bm{H}(\omega)(0)\cdot\frac{1}{\pi}\int_{\mathbb{R}}\frac{\omega}{x}\,\mathrm{d}{x}
=1π𝑯(ωω(0)x)ωdx(𝑯(ω)(0))2\displaystyle=\frac{1}{\pi}\int_{\mathbb{R}}\bm{H}\left(\frac{\omega-\omega(0)}{x}\right)\cdot\omega\,\mathrm{d}{x}-\big{(}\bm{H}(\omega)(0)\big{)}^{2}
=1πωω(0)x𝑯(ω)dx(𝑯(ω)(0))2\displaystyle=-\frac{1}{\pi}\int_{\mathbb{R}}\frac{\omega-\omega(0)}{x}\cdot\bm{H}(\omega)\,\mathrm{d}{x}-\big{(}\bm{H}(\omega)(0)\big{)}^{2}
=1πω𝑯(ω)xdx+ω(0)1π𝑯(ω)xdx(𝑯(ω)(0))2\displaystyle=-\frac{1}{\pi}\int_{\mathbb{R}}\frac{\omega\cdot\bm{H}(\omega)}{x}\,\mathrm{d}{x}+\omega(0)\cdot\frac{1}{\pi}\int_{\mathbb{R}}\frac{\bm{H}(\omega)}{x}\,\mathrm{d}{x}-\big{(}\bm{H}(\omega)(0)\big{)}^{2}
=1π𝑯(ω)ωxdx+ω(0)2(𝑯(ω)(0))2.\displaystyle=-\frac{1}{\pi}\int_{\mathbb{R}}\frac{\bm{H}(\omega)\cdot\omega}{x}\,\mathrm{d}{x}+\omega(0)^{2}-\big{(}\bm{H}(\omega)(0)\big{)}^{2}.

Rearranging the equation above yields the desired result. ∎

Acknowledgement

The authors are supported by the National Key R&D Program of China under the grant 2021YFA1001500.

References

  • CH [22] J. Chen and T. Y. Hou. Stable nearly self-similar blowup of the 2D Boussinesq and 3D Euler equations with smooth data. arXiv preprint arXiv:2210.07191, 2022.
  • [2] J. Chen, T. Y. Hou, and D. Huang. Matlab codes for computer-aided proofs in the paper “asymptotically self-similar blowup of the Hou–Luo model for the 3D Euler equations”. https://www.dropbox.com/sh/qjs6p6d9n3uiq8r/AABCDI-rZeVuTmBxGQuLJbUva?dl=0.
  • CHH [22] J. Chen, T. Y. Hou, and D. Huang. Asymptotically self-similar blowup of the Hou–Luo model for the 3D Euler equations. Annals of PDE, 8(2):24, 2022.
  • CHK+ [17] K. Choi, T. Y. Hou, A. Kiselev, G. Luo, V. Sverak, and Y. Yao. On the finite-time blowup of a one-dimensional model for the three-dimensional axisymmetric Euler equations. Communications on Pure and Applied Mathematics, 70(11):2218–2243, 2017.
  • CKY [15] K. Choi, A. Kiselev, and Y. Yao. Finite time blow up for a 1D model of 2D Boussinesq system. Communications in Mathematical Physics, 334:1667–1679, 2015.
  • CLM [85] P. Constantin, P. D. Lax, and A. Majda. A simple one-dimensional model for the three-dimensional vorticity equation. Communications on pure and applied mathematics, 38(6):715–724, 1985.
  • DE [23] T. D. Drivas and T. M. Elgindi. Singularity formation in the incompressible Euler equation in finite and infinite time. EMS Surveys in Mathematical Sciences, 10(1):1–100, 2023.
  • DG [90] S. De Gregorio. On a one-dimensional model for the three-dimensional vorticity equation. Journal of statistical physics, 59(5):1251–1263, 1990.
  • HQWW [24] D. Huang, X. Qin, X. Wang, and D. Wei. Self-similar finite-time blowups with smooth profiles of the generalized Constantin–Lax–Majda model. Archive for Rational Mechanics and Analysis, 248(2):22, 2024.
  • [10] G. Luo and T. Y. Hou. Potentially singular solutions of the 3D axisymmetric Euler equations. Proceedings of the National Academy of Sciences, 111(36):12968–12973, 2014.
  • [11] G. Luo and T. Y. Hou. Toward the finite-time blowup of the 3D axisymmetric Euler equations: a numerical investigation. Multiscale Modeling & Simulation, 12(4):1722–1776, 2014.
  • Liu [17] P. Liu. Spatial Profiles in the Singular Solutions of the 3D Euler Equations and Simplified Models. PhD thesis, California Institute of Technology, 2017.
  • MB [02] A. J. Majda and A. L. Bertozzi. Vorticity and Incompressible Flow, volume 27. Cambridge University Press, 2002.
  • OSW [08] H. Okamoto, T. Sakajo, and M. Wunsch. On a generalization of the Constantin–Lax–Majda equation. Nonlinearity, 21(10):2447, 2008.