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

Extremal of Log-Sobolev Functionals and Li-Yau Estimate on RCD(K,N)\mathrm{RCD}^{*}(K,N) Spaces

Samuel Drapeau Shanghai Jiao Tong University, School of Mathematical Sciences & Shanghai Advanced Institute for Finance, Shanghai, China [email protected] http://www.samuel-drapeau.info  and  Liming Yin Shanghai Jiao Tong University, School of Mathematical Sciences, Shanghai, China [email protected]
(Date: April 7, 2025)
Abstract.

In this work, we study the extremal functions of the log-Sobolev functional on compact metric measure spaces satisfying the RCD(K,N)\mathrm{RCD}^{*}(K,N) condition for KK in \mathbb{R} and NN in (2,)(2,\infty). We show the existence, regularity and positivity of non-negative extremal functions. Based on these results, we prove a Li-Yau type estimate for the logarithmic transform of any non-negative extremal functions of the log-Sobolev functional. As applications, we show a Harnack type inequality as well as lower and upper bounds for the non-negative extremal functions.
Keywords: Log-Sobolev functional; metric measure space; Li-Yau inequality; Curvature-dimension condition; extremal function.

National Science Foundation of China, Grants Numbers: 11971310 and 11671257 are gratefully acknowledged.

1. Introduction

In this work, we study the extremal functions of the following variational problem of the log-Sobolev functional

λ(α1,α2):=inffX(|f|2+α1f2α22f2logf2)𝑑m\lambda(\alpha_{1},\alpha_{2}):=\inf_{f\in\mathcal{F}}\int_{X}\left(|\nabla f|^{2}+\alpha_{1}f^{2}-\frac{\alpha_{2}}{2}f^{2}\log f^{2}\right)dm (1.1)

where α1\alpha_{1} and α2\alpha_{2} are in \mathbb{R} and :={fW1,2:fL2=1}\mathcal{F}:=\{f\in W^{1,2}:\|f\|_{L^{2}}=1\}. Of interest are the existence, regularity, positivity of non-negative extremal functions, and analytic results such as Li-Yau or Harnack type estimates. While those results are well-known in the smooth Riemannian setting, it seems natural to ask whether they can be extended and in which form to more general non-smooth metric spaces.

While the log-Sobolev inequality has vast applications in different branches of mathematics—see Gross [27], Otto and Villani [41], Bakry, Gentil, and Ledoux [10]—studying its extremal functions is important on its own right. For example, Zhang [52] shows that in the noncompact smooth manifold, the geometry of the manifold at infinity will affect the existence of extremal of the log-Sobolev functional and Perelman’s WW-entropy. Using these points, the author further shows that under the mild assumptions, noncompact shrinking breathers of Ricci flow are gradient shrinking solitons. Very recently, the extremal functions of the log-Sobolev inequality are used together with the needle decomposition technique by Ohta and Takatsu [40] to show some rigidity result of the underlying weighted Riemannian manifold.

From the viewpoint of the underlying space, starting with the works of Sturm [47, 48] and Lott and Villani [37], the synthetic notion of Ricci curvature—referred to as CD(K,N)\mathrm{CD}(K,N) condition—bounded from below by KK in \mathbb{R} and the dimension bounded from above by 1N1\leq N\leq\infty on a general metric measure space without having a smooth structure was introduced and developed greatly in the last decade. The key property of this notion is that it is compatible with the smooth Riemannian setting and stable with respect to the measured Gromov-Hausdorff convergence so that it includes Ricci limit spaces and Alexandrov spaces. Later, to rule out the Finsler geometry, the finer RCD(K,)\mathrm{RCD}(K,\infty) condition was introduced by Ambrosio, Gigli, and Savaré [4] and the finite dimensional counterpart RCD(K,N)\mathrm{RCD}^{*}(K,N) condition was introduced and studied in [20, 8, 7]. Recently, building upon the abstract module theory, the first and second order differential structure on RCD(K,)\mathrm{RCD}(K,\infty) spaces was developed by Gigli [23], and finer geometric results such as the rectifiability of RCD(K,N)\mathrm{RCD}(K,N) spaces were studied by Mondino and Naber [38].

With these analytic tools, the geometric analysis on metric measure spaces satisfying the synthetic Ricci curvature condition also developed quickly. For instance, Li-Yau-Hamilton type inequalities for the heat flow [21, 31, 32] and the localized gradient and a local Li-Yau estimate for the heat equation [30, 51]. In particular, Zhang and Zhu [51] develop an Omori-Yau type maximum principle on the RCD(K,N)\mathrm{RCD}^{*}(K,N) space and use it to show a pointwise Li-Yau type estimate for locally weak solutions of the heat equation which may not have the semigroup property.

Motivated by these works, we study the extremal functions of the log-Sobolev functional (1.1) on more general metric measure spaces. In particular, we are interested in whether analytic results such as Li-Yau type estimates for the non-negative extremal functions of the log-Sobolev functional holding on smooth Riemannian manifolds can be extended to non-smooth metric measure spaces, in particular, those satisfying the synthetic Ricci curvature condition.

To do so, one of the key points is to show the existence, boundedness, regularity and positivity of the non-negative extremal functions of the log-Sobolev functional. Our first main result, Theorem 3.2, states that the log-Sobolev functional (1.1) with α2>0\alpha_{2}>0 on a compact metric measure space satisfying the RCD(K,N)\mathrm{RCD}^{*}(K,N) condition with KK in \mathbb{R} and NN in (2,)(2,\infty) admits non-negative extremal functions which satisfy certain Euler-Lagrange equation. Moreover, we show that all the non-negative extremal functions are bounded, Lipschitz continuous and bounded away from 0.

We remark that while the existence and Euler-Lagrange equation problems are quite standard and similar to the smooth compact cases solved by Rothaus [45], several problems arise on metric measure spaces. For instance, the positivity of the non-negative extremal functions [45, page 114] is shown by relying heavily on the underlying smooth differential structure of the Riemannian manifold, the polar coordinates and the exact asymptotic volume ratio near the pole so that the problem can be reduced to a one-dimensional ODE problem. However, these smooth structures are lost on general metric measure spaces. While for RCD(K,N)\mathrm{RCD}^{*}(K,N) spaces, the polar decomposition still works by using the “needle decomposition” generalized to essentially non-branching CD(K,N)\mathrm{CD}(K,N) spaces by Cavalletti and Mondino [15] (see also [17]), the similar asymptotic volume ratio analysis seems to fail without further assumptions on the underlying metric measure space. To overcome this difficulty, we make use of a maximum principle type argument for the De Giorgi class on some local domain proved by Kinnunen and Shanmugalingam [34], to show that non-negative extremal functions are either bounded away from 0 or vanish on the whole space. This method works in very general metric measure spaces supporting the local doubling property and weak Poincaré inequality, which is in particular the case for RCD(K,N)\mathrm{RCD}^{*}(K,N) spaces.

Our second main result is Theorem 4.1. Based on the regularity and positivity results obtained above, we recover a Li-Yau type estimate for the logarithmic transform of all non-negative extremal functions of (1.1). More precisely, for any non-negative extremal functions uu, it holds that

|v|2+(α2βK)vNα2(1β)4β(1β((2β)Kα2)2α2(1β))2m-a.e.,|\nabla v|^{2}+(\alpha_{2}-\beta K)v\leq\frac{N\alpha_{2}(1-\beta)}{4\beta}\left(1-\frac{\beta((2-\beta)K-\alpha_{2})}{2\alpha_{2}(1-\beta)}\right)^{2}\quad\text{$m$-a.e.,} (1.2)

for any β\beta in (0,1)(0,1) and v=logu+(λα1)/α2v=\log u+(\lambda-\alpha_{1})/\alpha_{2}. The same estimate for the smooth Riemannian case was shown by Wang [50], where the argumentation relies on the pointwise Bochner formula and a pointwise characterization of local maximum points of the smooth function in the left-hand side of (1.2). However, in the RCD\mathrm{RCD} setting, neither the function in the left-hand side of (1.2) is smooth and pointwise defined, nor the pointwise Bochner formula is available. To overcome the difficulty, we follow a similar argument as in [51] by using an Omori-Yau type maximum principle. Note that, to avoid the sign problem of |v|2+(α2βK)v|\nabla v|^{2}+(\alpha_{2}-\beta K)v, we use a different auxiliary function ϕ\phi from those in [51, Theorem 1.4], which are constructed from the distance function so that they have the measure-valued Laplacian. Our construction is based on the “good” cut-off functions from Mondino and Naber [38, Lemma 3.1], which are smoother than those in [51] and has the L2L^{2}-valued Laplacian. Furthermore, for our purpose, we slightly extend the Omori-Yau maximum principle in [51], which holds on RCD(K,N)\mathrm{RCD}^{*}(K,N) spaces with KK\in\mathbb{R} and N1N\geq 1, to proper RCD(K,)\mathrm{RCD}(K,\infty) spaces. While most arguments are similar, our proof follows the so-called “Sobolev-to-Lip” property, a property shared by all RCD\mathrm{RCD} spaces, rather than the “weak maximum principle” from [51].

Finally, we provide some applications of the regularity and positivity results and the Li-Yau type estimate. We show a Harnack type estimate as well as lower and upper bounds of the non-negative extremal functions of (1.1) depending only on the geometry of the space. These generalize the results in [50] proved in the Riemannian setting. Using the weak Bochner inequality, we also show that all non-negative extremal functions are constant when 0<α2KN/(N1)0<\alpha_{2}\leq KN/(N-1), which is well-known in the smooth setting.

The paper is organized as follows: in Section 2, we introduce the notations and definitions about metric measure spaces and RCD\mathrm{RCD} conditions as well as the analytic results needed later. In Section 3, we study the variational problem and show the existence, regularity and positivity of non-negative extremal functions of (1.1). Section 4 is dedicated to the Li-Yau type estimate for the non-negative extremal functions. In Section 5, we present some applications of the previous results. Finally, in an Appendix, we prove the Omori-Yau type maximum principle for proper RCD(K,)\mathrm{RCD}(K,\infty) spaces.

2. Preliminary and notations

We briefly introduce the terminologies and notations related to calculus. For more details, we refer the readers to [25, 23].

Throughout this work, we denote by (X,d,m)(X,d,m) a metric measure space where (X,d)(X,d) is a complete, separable and proper metric space and mm is a non-negative Radon measure with full support which is finite on every bounded and measurable set. We denote by B(x,r)B(x,r) and B¯(x,r)\bar{B}(x,r) the open and closed metric balls centered at xx in XX with radius r>0r>0, respectively. By Lp:=Lp(X,m)L^{p}:=L^{p}(X,m) for 1p1\leq p\leq\infty we denote the standard LpL^{p} spaces with LpL^{p}-norm p\|\cdot\|_{p}. By LlocpL^{p}_{loc} we denote those measurable functions f:Xf:X\to\mathbb{R} such that fχBf\chi_{B} is in LpL^{p} for any bounded and measurable subset BB of XX, where χB\chi_{B} denote the indicator function of the set BB. Let further Lip:=Lip(X)\mathrm{Lip}:=\mathrm{Lip}(X), Liploc:=Liploc(X)\mathrm{Lip}_{loc}:=\mathrm{Lip}_{loc}(X), and Lipbs:=Lipbs(X)\mathrm{Lip}_{bs}:=\mathrm{Lip}_{bs}(X) be the spaces of real-valued functions f:Xf:X\to\mathbb{R} which are Lipschitz, locally Lipschitz, and Lipschitz with bounded support, respectively. For ff in Liploc\mathrm{Lip}_{loc}, we denote by lip(f)\mathrm{lip}(f) the local Lipschitz constant, or slope, defined for any xx in XX as

lipf(x):=lim supyx|f(y)f(x)|d(y,x),\mathrm{lip}f(x):=\limsup_{y\rightarrow x}\frac{|f(y)-f(x)|}{d(y,x)},

if xx is not isolated and lipf(x)=0\mathrm{lip}f(x)=0 if xx is isolated.

2.1. Cheeger Energy, Laplacian and Calculus Tools

The Cheeger energy is a L2L^{2}-lower-semicontinuous and convex functional Ch:L2[0,]\mathrm{Ch}:L^{2}\rightarrow[0,\infty] defined as

Ch(f):=inf{lim inf12X(lip(fn))2𝑑m:(fn)LipL2,fnf20}.\mathrm{Ch}(f):=\inf\left\{\liminf\frac{1}{2}\int_{X}\left(\mathrm{lip}(f_{n})\right)^{2}dm\colon(f_{n})\subseteq\mathrm{Lip}\cap L^{2},\|f_{n}-f\|_{2}\rightarrow 0\right\}.

The domain of Ch\mathrm{Ch} is a linear space denoted by W1,2:=W1.2(X)W^{1,2}:=W^{1.2}(X) and is called the Sobolev space. For ff in W1,2W^{1,2}, we identify the canonical element |f||\nabla f| called the minimal relaxed gradient as the unique element with minimal L2L^{2}-norm, also minimal in the mm-a.e. sense, in the set

{GL2:G=limlipfn in L2 for some (fn)Lip such that fnf in L2},\left\{G\in L^{2}:G=\lim\mathrm{lip}f_{n}\text{ in }L^{2}\text{ for some }(f_{n})\subseteq\mathrm{Lip}\text{ such that }f_{n}\rightarrow f\text{ in $L^{2}$}\right\},

which provides an integral representation Ch(f)=12X|f|2𝑑m\mathrm{Ch}(f)=\frac{1}{2}\int_{X}|\nabla f|^{2}dm. The Sobolev space equipped with the norm fW1,22:=f22+2Ch(f)\|f\|^{2}_{W^{1,2}}:=\|f\|^{2}_{2}+2\mathrm{Ch}(f) is a Banach space and is dense in L2L^{2}, see [3, Proposition 4.1]. We further denote by Wloc1,2:={fLloc2:ηfW1,2 for any ηLipbs}W^{1,2}_{loc}:=\{f\in L^{2}_{loc}:\eta f\in W^{1,2}\text{ for any }\eta\in\mathrm{Lip}_{bs}\} the space of local Sobolev functions, and define the minimal relaxed gradient as |f|:=|(ηf)||\nabla f|:=|\nabla(\eta f)| mm-a.e. on {η=1}\{\eta=1\} for ff in Wloc1,2W^{1,2}_{loc} where η\eta is in Lipbs\mathrm{Lip}_{bs}.

We say that (X,d,m)(X,d,m) is infinitesimally Hilbertian if the Cheeger energy is a quadratic form, or equivalently, W1,2W^{1,2} is a Hilbert space. Under these assumptions, it can be proved that for any ff and gg in W1,2W^{1,2}, the limit

f,g:=limε0|(f+εg)|2|f|22ε\langle\nabla f,\nabla g\rangle:=\lim_{\varepsilon\rightarrow 0}\frac{|\nabla(f+\varepsilon g)|^{2}-|\nabla f|^{2}}{2\varepsilon}

exists in L1L^{1} and it is a bilinear form from W1,2×W1,2W^{1,2}\times W^{1,2} to L1L^{1}, see [4].

Definition 2.1.

Let (X,d,m)(X,d,m) be infinitesimally Hilbertian.

  • Laplacian: We say that ff in W1,2W^{1,2}, is in the domain of the Laplacian, denoted by D(Δ)D(\Delta), provided that there exists hh in L2L^{2} such that

    Xf,g𝑑m=Xhg𝑑m,for any gW1,2.-\int_{X}\langle\nabla f,\nabla g\rangle dm=\int_{X}hgdm,\quad\text{for any }g\in W^{1,2}. (2.1)

    In this case, we denote Δf=h\Delta f=h.

  • Measure-Valued Laplacian: We say that ff in Wloc1,2W^{1,2}_{loc} is in the domain of the measure-valued Laplacian, denoted by D(𝚫)D(\bm{\Delta}), provided that there exists a signed Radon measure μ\mu on XX such that,111Recall that XX is assumed to be proper, and therefore any bounded and closed set is compact on which Radon measures are finite.

    Xf,g𝑑m=Xg𝑑μ,for any gLipbs.-\int_{X}\langle\nabla f,\nabla g\rangle dm=\int_{X}gd\mu,\quad\text{for any }g\in\mathrm{Lip}_{bs}. (2.2)

    In this case, we denote 𝚫f=μ\bm{\Delta}f=\mu.

By the separating property of Lipbs\mathrm{Lip}_{bs} for Radon measures and the infinitesimal Hilbertian property, it’s clear that both Δ\Delta and 𝚫\bm{\Delta} are well-defined, unique and linear operators. Moreover, the two definitions are compatible in the following sense: on the one hand, if ff is in W1,2W^{1,2} with 𝚫f=ρm\bm{\Delta}f=\rho m for some ρ\rho in L2L^{2}, then ff is in D(Δ)D(\Delta) and Δf=ρ\Delta f=\rho. On the other hand, if ff is in W1,2W^{1,2} such that ΔfL1\Delta f\in L^{1}, then ff is in D(𝚫)D(\bm{\Delta}) and 𝚫f=(Δf)m\bm{\Delta}f=(\Delta f)m, see [25, Proposition 6.2.13]. For ff in the domain of 𝚫\bm{\Delta}, we denote the Lebesgue decomposition with respect to mm

𝚫f=(𝚫acf)m+𝚫sf,\bm{\Delta}f=(\bm{\Delta}^{ac}f)\cdot m+\bm{\Delta}^{s}f, (2.3)

where 𝚫acf\bm{\Delta}^{ac}f is the Radon-Nikodym density and 𝚫sf\bm{\Delta}^{s}f is the singular part of 𝚫f\bm{\Delta}f.

For ww in W1,2LW^{1,2}\cap L^{\infty}, we define the weighted Laplacian Δw\Delta_{w} similarly but with respect to the reference measure mw:=ewmm_{w}:=e^{w}\cdot m and test functions in W1,2(X,mw)W^{1,2}(X,m_{w}). For such ww, it can be shown that W1,2W^{1,2} coincides with W1,2(X,mw)W^{1,2}(X,m_{w}), and the minimal relaxed gradient induced by mwm_{w} coincides with the one induced by mm, see [3, Lemma 4.11]. Moreover, it holds that Δwf=Δf+w,f\Delta_{w}f=\Delta f+\langle\nabla w,\nabla f\rangle, see [26, Lemma 3.4]

The following Lemma recaps the calculus rules, whose proofs can be found in [25, 51, 24].

Lemma 2.2.

Let (X,d,m)(X,d,m) be infinitesimally Hilbertian. Then:

  1. (i)

    Locality: |f|=|g||\nabla f|=|\nabla g| on {fg=c}\{f-g=c\} for any ff, gg in W1,2W^{1,2} and constant cc.

  2. (ii)

    Chain rule: for any ff in W1,2W^{1,2} and Lipschitz function ϕ:\phi:\mathbb{R}\to\mathbb{R}, it follows that

    |(ϕf)|=|ϕ(f)||f||\nabla(\phi\circ f)|=|\phi^{\prime}(f)||\nabla f|

    In particular, if ϕ\phi is a contraction, then |(ϕf)||f||\nabla(\phi\circ f)|\leq|\nabla f|.

  3. (iii)

    Leibniz rule: for any ff, gg and hh in W1,2LW^{1,2}\cap L^{\infty}, it follows that fgfg is in W1,2W^{1,2} and

    (fg),h=fg,h+gf,h.\langle\nabla(fg),\nabla h\rangle=f\langle\nabla g,\nabla h\rangle+g\langle\nabla f,\nabla h\rangle.
  4. (iv)

    Chain rule: for any ff in D(Δ)LipbD(\Delta)\cap\mathrm{Lip}_{b} and C2C^{2}-function ϕ:\phi:\mathbb{R}\rightarrow\mathbb{R}, it follows that ϕ(f)\phi(f) is in D(Δ)D(\Delta) and

    Δϕ(f)=ϕ(f)Δf+ϕ′′(f)|f|2.\Delta\phi(f)=\phi^{\prime}(f)\Delta f+\phi^{\prime\prime}(f)|\nabla f|^{2}. (2.4)
  5. (v)

    Leibniz rule: for any ff and gg in D(𝚫)LD(\bm{\Delta})\cap L^{\infty} such that gg is continuous and 𝚫g\bm{\Delta}g is absolutely continuous with respect to mm, then fgfg is in D(𝚫)D(\bm{\Delta}) and

    𝚫(fg)=f𝚫g+g𝚫f+2f,gm.\bm{\Delta}(fg)=f\bm{\Delta}g+g\bm{\Delta}f+2\langle\nabla f,\nabla g\rangle\cdot m. (2.5)

By the L2L^{2}-lower semicontinuity and convexity of the Cheeger energy, the heat semigroup PtfP_{t}f is defined as the gradient flow in L2L^{2} of the Cheeger energy starting from fL2f\in L^{2} based on the classical Brezis-Komura theory, which provides the existence and uniqueness results. Moreover, for any fL2f\in L^{2}, it holds that tPtft\mapsto P_{t}f is locally absolutely continuous on (0,)(0,\infty) and PtfP_{t}f is in D(Δ)D(\Delta) for all t>0t>0 and

ddtPtf=ΔPtf,for almost all t(0,).\frac{d}{dt}P_{t}f=\Delta P_{t}f,\quad\text{for almost all }t\in(0,\infty). (2.6)

Under the further assumption that (X,d,m)(X,d,m) is infinitesimally Hilbertian, the heat semigroup PtP_{t} is linear, strongly continuous, contractive and order-preserving in L2L^{2}. Moreover, PtP_{t} can be extended into a linear, mass preserving and strongly continuous operator in LpL^{p} for any 1p<1\leq p<\infty, see [3] and further results therein.

Finally, we recall the definition (see for example [1, Definition 2.14]) of the local Sobolev space W1,2(Ω)W^{1,2}(\Omega) on an open subset ΩX\Omega\subset X. We denote by Lipc(Ω)\mathrm{Lip}_{c}(\Omega) the space of Lipschitz functions on Ω\Omega with compact support in Ω\Omega and by Liploc(Ω)\mathrm{Lip}_{loc}(\Omega) the space of locally Lipschitz functions on Ω\Omega.

Definition 2.3 (Local Sobolev space).

Let ΩX\Omega\subset X be an open subset. We say that fL2(Ω)f\in L^{2}(\Omega) is in W1,2(Ω)W^{1,2}(\Omega) if

  1. (i)

    ηf\eta f is in W1,2W^{1,2} for all ηLipc(Ω)\eta\in\mathrm{Lip}_{c}(\Omega);

  2. (ii)

    |f|L2(Ω)|\nabla f|\in L^{2}(\Omega) where |f|:=|(ηf)||\nabla f|:=|\nabla(\eta f)| mm-a.e. on {η=1}\{\eta=1\} for ηLipc(Ω)\eta\in\mathrm{Lip}_{c}(\Omega).

We remark that by locality of the minimal relaxed gradient and Condition (i), Condition (ii) makes sense. As for the Laplacian on Ω\Omega, it is modified as follows: A function ff in W1,2(Ω)W^{1,2}(\Omega) belongs to D(Δ,Ω)D(\Delta,\Omega) provided that there exists gg in L2(Ω)L^{2}(\Omega) such that

Ωf,ϕ𝑑m=Ωgϕ𝑑m,for all ϕLipc(Ω),-\int_{\Omega}\langle\nabla f,\nabla\phi\rangle dm=\int_{\Omega}g\phi dm,\quad\text{for all }\phi\in\mathrm{Lip}_{c}(\Omega),

and we denote ΔΩf:=g\Delta_{\Omega}f:=g. Clearly, ΔΩ\Delta_{\Omega} is linear by infinitesimal Hilbertianty and one can easily check that for ff in D(Δ)D(\Delta), its restriction to Ω\Omega belongs to D(Δ,Ω)D(\Delta,\Omega).

2.2. RCD metric measure spaces

Erbar, Kuwada, and Sturm [20], Ambrosio, Mondino, and Savaré [7, 8] introduced the notion of the Riemannian curvature-dimension condition RCD(K,N)\mathrm{RCD}^{*}(K,N) as the finite dimensional counterpart to RCD(K,)\mathrm{RCD}(K,\infty), itself introduced by Ambrosio, Gigli, and Savaré [4] based on the curvature-dimension condition proposed by [37, 47, 48] to rule out the Finsler geometry. It is shown in [20, 7] that the RCD(K,N)\mathrm{RCD}^{*}(K,N) spaces satisfy the so-called “local-to-global” property (see also [9]). Very recently, Cavalletti and Milman [12] show that RCD(K,N)\mathrm{RCD}(K,N) and RCD(K,N)\mathrm{RCD}^{*}(K,N) conditions are equivalent if the reference measure is finite.

The notion of the RCD(K,N)\mathrm{RCD}^{*}(K,N) condition can be defined in several equivalent ways, see [20, 7, 8] for RCD(K,N)\mathrm{RCD}^{*}(K,N) cases and [4, 5, 6] for RCD(K,)\mathrm{RCD}(K,\infty) cases. In this work we give a definition including the case where N=N=\infty from an Eulerian point of view based on the abstract Γ\Gamma-calculus.

Definition 2.4.

We say that a metric measure space (X,d,m)(X,d,m) satisfies the RCD(K,N)\mathrm{RCD}^{*}(K,N) condition for KK in \mathbb{R} and NN in [1,][1,\infty] provided that

  1. (i)

    m(B(x,r))Cexp(Cr2)m(B(x,r))\leq C\exp(Cr^{2}) for some xx in XX and C>0C>0;

  2. (ii)

    Sobolev-to-Lip property: any ff in W1,2W^{1,2} with |f||\nabla f| in LL^{\infty} admits a Lipschitz representation f~\tilde{f} in Lip\mathrm{Lip} such that f=f~f=\tilde{f} mm-a.e and Lip(f)=|f|\mathrm{Lip}(f)=\||\nabla f|\|_{\infty}.

  3. (iii)

    (X,d,m)(X,d,m) is infinitesimally Hilbertian.

  4. (iv)

    Weak Bochner inequality: for any ff and gg in D(Δ)D(\Delta) with Δf\Delta f in W1,2W^{1,2}, Δg\Delta g in LL^{\infty} and g0g\geq 0, it holds

    XΔg|f|22𝑑mXg(1N(Δf)2+f,Δf+K|f|2)𝑑m.\int_{X}\Delta g\frac{|\nabla f|^{2}}{2}dm\geq\int_{X}g\left(\frac{1}{N}(\Delta f)^{2}+\langle\nabla f,\nabla\Delta f\rangle+K|\nabla f|^{2}\right)dm. (2.7)

In the following, we assume that (X,d,m)(X,d,m) is a compact RCD(K,N)\mathrm{RCD}^{\ast}(K,N) space with NN in (1,)(1,\infty), KK\in\mathbb{R} and diam(X)=D<\mathrm{diam}(X)=D<\infty, which is the framework of the results in this work

First, XX satisfies the generalized Bishop-Gromov inequality, that is, for any 0<r<R0<r<R and xXx\in X,

m(B(x,R))vK,N(R)m(B(x,r))vK,N(r),\frac{m(B(x,R))}{v_{K,N}(R)}\leq\frac{m(B(x,r))}{v_{K,N}(r)}, (2.8)

where vK,N(r)v_{K,N}(r) is the volume of ball with radius r>0r>0 in the model space (see [9, Theorem 6.2]). In particular, XX is globally doubling with the constant 2N2^{N} when K0K\geq 0, that is,

m(B(x,2r))2Nm(B(x,r)),for any xX and r>0,m(B(x,2r))\leq 2^{N}m(B(x,r)),\quad\text{for any }x\in X\text{ and }r>0, (2.9)

and with the constant C(K,N,D)C(K,N,D) depending only on K,NK,N and DD when K<0K<0. It also holds that m(B(x,r))>0m(B(x,r))>0 for any r>0r>0 and xx in XX and that m(X)<m(X)<\infty. Thus by [12], XX also satisfies the RCD(K,N)\mathrm{RCD}(K,N) condition.

Second, XX supports the weak (1,1)(1,1)-Poincaré inequality, see [43, Theorem 1.1], that is, for any xx in XX, r>0r>0 and any continuous function f:Xf:X\rightarrow\mathbb{R} and any upper gradient gg of ff, we have

B(x,r)|f(f)x,r|𝑑mCrB(x,2r)g𝑑m,\fint_{B(x,r)}\left|f-(f)_{x,r}\right|dm\leq Cr\fint_{B(x,2r)}gdm, (2.10)

where the constant CC only depends on KK, NN and rr, and Ωf𝑑m:=1m(Ω)Ωf𝑑m\fint_{\Omega}fdm:=\frac{1}{m(\Omega)}\int_{\Omega}fdm and (f)x,r:=B(x,r)f𝑑m(f)_{x,r}:=\int_{B(x,r)}fdm. Clearly, by Hölder inequality, XX also supports the weak (1,q)(1,q)-Poincaré inequality for any q(1,)q\in(1,\infty). This implies that XX is connected, see [11, Proposition 4.2]. In fact, one can show that XX is also a geodesic space.

Third, when K>0K>0, the Bonnet-Myers theorem implies that diam(X)π(N1)/K\mathrm{diam}(X)\leq\pi\sqrt{(N-1)/K}, see [48, Corollary 2.6]. Since the normalization of the reference measure does not affect RCD(K,N)\mathrm{RCD}^{*}(K,N) conditions, it is not restrictive to assume that m(X)=1m(X)=1 when XX is compact.

Fourth, in the RCD\mathrm{RCD} setting, for Sobolev functions fW1,2f\in W^{1,2}, it is possible to identify the gradient f\nabla f, rather than the modulus of the gradient |f||\nabla f|, as the unique element in the tangent module L2(TX)L^{2}(TX), which is a L2(m)L^{2}(m)-normed L(m)L^{\infty}(m)-module, see [25, 23]. Therein, a second-order calculus on RCD(K,)\mathrm{RCD}(K,\infty) spaces is also developed such that the notions of Hessian Hessf\mathrm{Hess}f and its pointwise norm |Hessf|HSL2|\mathrm{Hess}f|_{HS}\in L^{2} are well-defined. For the complete theory, we refer readers to [25, 23]. Here we only mention the inclusion D(Δ)H2,2D(\Delta)\subseteq H^{2,2}. Consequently, for any ff in D(Δ)LipD(\Delta)\cap\mathrm{Lip}, we have |f|2|\nabla f|^{2} in W1,2W^{1,2} and

||f|2|2|Hessf|HS|f|,|\nabla|\nabla f|^{2}|\leq 2|\mathrm{Hess}f|_{HS}|\nabla f|, (2.11)

see [23, Theorem 3.3.18] or [19, Lemma 3.5].

Remark 2.5.

Although there exist different notions of the Sobolev space such as the Newtonian space and the Cheeger-Sobolev space, see [11, 46, 18], and different notions of weak upper gradients on metric measure spaces, all these notions are equivalent to each other in the setting of RCD(K,N)\mathrm{RCD}^{*}(K,N) spaces. In particular, the minimal relaxed gradient coincides with the minimal weak upper gradient, which is defined via test plans and geodesics, see [25, Definition 2.1.8], and W1,2W^{1,2} is reflexive. For more details, we refer to [2].

Fifth, we recall a regularity result of the Poisson equation on the ball, see [51, Lemma 3.4] and [35, Theorem 3.1] proved in the setting of the heat semigroup curvature condition.

Lemma 2.6.

Let (X,d,m)(X,d,m) be a RCD(K,N)\mathrm{RCD}^{*}(K,N) space with KK in \mathbb{R} and NN in [1,)[1,\infty). Let further gg be in L(BR)L^{\infty}(B_{R}) where BR:=B(x0,R)B_{R}:=B(x_{0},R) is a geodesic ball centered at some x0x_{0} in XX with radius R>0R>0. Assume that ff is in W1,2(BR)W^{1,2}(B_{R}) and ΔBRf=g\Delta_{B_{R}}f=g on BRB_{R}. Then it holds that |f||\nabla f| is in Lloc(BR)L^{\infty}_{loc}(B_{R}) and

|f|L(BR/2)C(N,K,R)(1m(BR)fL1(BR)+gL(BR)).\||\nabla f|\|_{L^{\infty}(B_{R/2})}\leq C(N,K,R)\left(\frac{1}{m(B_{R})}\|f\|_{L^{1}(B_{R})}+\|g\|_{L^{\infty}(B_{R})}\right).
Remark 2.7.

Note that our definition of the local Sobolev space W1,2(Ω)W^{1,2}(\Omega) on some open domain Ω\Omega is a priori different from the one H1,2(Ω)H^{1,2}(\Omega) used in [35], which is the completion of Liploc(Ω)\mathrm{Lip}_{loc}(\Omega) with respect to the norm fH1,2(Ω):=fL2(Ω)+lip(f)L2(Ω)\|f\|_{H^{1,2}(\Omega)}:=\|f\|_{L^{2}(\Omega)}+\|\mathrm{lip}(f)\|_{L^{2}(\Omega)}. However, on the one hand, as shown in [1, Remark 2.15], this space coincides with the Cheeger-Sobolev space Ch1,2(Ω)Ch^{1,2}(\Omega) defined by Cheeger [18] on RCD(K,N)\mathrm{RCD}^{*}(K,N) spaces. On the other hand, since Liploc(Ω)\mathrm{Lip}_{loc}(\Omega) is dense in Ch1,2(Ω)Ch^{1,2}(\Omega) (see [11, Theorem 5.47]) and the norms of Ch1,2(Ω)Ch^{1,2}(\Omega) and H1,2(Ω)H^{1,2}(\Omega) are equal for Liploc(Ω)\mathrm{Lip}_{loc}(\Omega) on RCD(K,N)\mathrm{RCD}^{*}(K,N) spaces222Since the local Lipschitz constant of Lipschitz functions coincides with their minimal weak upper gradient in the Cheeger sense, see [11, Theorem A.7], these two spaces coincide with each other. Consequently, H1,2(Ω)H^{1,2}(\Omega) and W1,2(Ω)W^{1,2}(\Omega) coincide with each other.

We also recall the following result of the Sobolev inequality for the compact RCD(K,N)\mathrm{RCD}(K,N) space with KK\in\mathbb{R} and NN in (2,)(2,\infty), proved in [39]. As a result, we have a metric version of the Rellich-Kondrachov type theorem for compact CD(K,N)\mathrm{CD}(K,N) spaces with KK\in\mathbb{R} and N(2,)N\in(2,\infty). Similar results have been previously proved by Profeta [42] for RCD(K,N)\mathrm{RCD}^{*}(K,N) spaces with K>0K>0 and N(2,)N\in(2,\infty).

Lemma 2.8.

Let (X,d,m)(X,d,m) be a compact RCD(K,N)\mathrm{RCD}^{*}(K,N) space with KK\in\mathbb{R} and NN in (2,)(2,\infty) and diam(X)D\mathrm{diam}(X)\leq D for some D>0D>0 and m(X)=1m(X)=1.

  1. (i)

    Sobolev inequality ([39, Proposition 5.1]): there exist a constant A>0A>0 depending only on K,N,DK,N,D, such that for every ff in W1,2W^{1,2}, it holds

    f22f22+ACh(f),\|f\|^{2}_{2^{*}}\leq\|f\|^{2}_{2}+A\cdot\mathrm{Ch}(f), (2.12)

    where 2=2N/(N2)2^{*}=2N/(N-2) is the Sobolev conjugate of 22.

  2. (ii)

    Rellich-Kondrachov: let (fn)(f_{n}) be a sequence in W1,2W^{1,2} with supnfnW1,2<\sup_{n}\|f_{n}\|_{W^{1,2}}<\infty. Then there exists ff in W1,2W^{1,2} and a subsequence (fnk)k(f_{n_{k}})_{k} such that for every 1q<21\leq q<2^{*}, it holds that

    fnkfin Lq(X,m).f_{n_{k}}\rightarrow f\quad\text{in }L^{q}(X,m).

The proof of the Rellich-Kondrachov type theorem above follows the argument in [28, Theorem 8.1] and the equivalent characterizations of weak Poincaré inequalities from Keith [33], which is actually the same as the one in [42, Proposition 4.2] for RCD(K,N)\mathrm{RCD}^{*}(K,N) spaces with K>0K>0 and N(2,)N\in(2,\infty). For the sake of completeness, we give the proof in the Appendix B.

Finally, we mention a key result about the heat semigroup and the resolvent of the Laplacian. Recall that for the heat semigroup PtP_{t}, we say that PtP_{t} is ultracontractive if for 1p<q1\leq p<q\leq\infty, there exists a constant C(t)>0C(t)>0 such that for any ff in LpL^{p}, it holds that

PtfqC(t)fp,t>0,\|P_{t}f\|_{q}\leq C(t)\|f\|_{p},\quad t>0,

and we denote Pt(p,q):=supfp1Ptfq\|P_{t}\|_{(p,q)}:=\sup_{\|f\|_{p}\leq 1}\|P_{t}f\|_{q}. The ultracontractive property of the heat semigroup is equivalent to the Sobolev inequality for the Markov triple associated with the heat semigroup, see [10, Theorem 6.3.1]. For compact RCD(K,N)\mathrm{RCD}^{*}(K,N) spaces with KK\in\mathbb{R} and NN in (2,)(2,\infty), since the Sobolev inequality holds, it follows that PtP_{t} has the ultracontractive property. More precisely, for 1p<q1\leq p<q\leq\infty and 0<t10<t\leq 1, it holds that

Pt(p,q)CtN2(1p1q),\|P_{t}\|_{(p,q)}\leq Ct^{-\frac{N}{2}(\frac{1}{p}-\frac{1}{q})},

where C>0C>0 is a constant depending on KK and NN. As a consequence, the ultracontractive property of the heat semigroup provides the following boundedness result for the resolvent operator of the Laplacian Rλ:=(λIΔ)1R_{\lambda}:=(\lambda I-\Delta)^{-1} for λ>0\lambda>0, the proof of which can be found in [42, Lemma 4.1] and [10, Corollary 6.3.3].

Lemma 2.9.

On a compact RCD(K,N)\mathrm{RCD}^{*}(K,N) space with KK\in\mathbb{R} and NN in (2,)(2,\infty), let λ>0\lambda>0. If 1pN/21\leq p\leq N/2, the resolvent Rλ:LpLqR_{\lambda}:L^{p}\rightarrow L^{q} is bounded for each 1q<pN/(N2p)1\leq q<pN/(N-2p). If p>N/2p>N/2, the resolvent Rλ:LpLR_{\lambda}:L^{p}\rightarrow L^{\infty} is bounded.

3. Existence of positive extremal functions

From now on, we assume that (X,d,m)(X,d,m) is a compact RCD(K,N)\mathrm{RCD}^{*}(K,N) space with KK\in\mathbb{R} and NN in (2,)(2,\infty), and mm is a Borel probability measure with full support. We consider the following variational problem

inf{X(|f|2+α1f2α22f2logf2)dm:fW1,2,f2=1},\inf\left\{\int_{X}\left(|\nabla f|^{2}+\alpha_{1}f^{2}-\frac{\alpha_{2}}{2}f^{2}\log f^{2}\right)dm:f\in W^{1,2},\|f\|_{2}=1\right\}, (3.1)

where α1\alpha_{1} and α2\alpha_{2} are in \mathbb{R}. The infimum quantity of variational problem (3.1) is called the log Sobolev constant λ:=λ(α1,α2)\lambda:=\lambda(\alpha_{1},\alpha_{2}) on XX with parameters α1\alpha_{1} and α2\alpha_{2}, and the functional in (3.1) is called the log-Sobolev functional.

Definition 3.1.

Provided that λ=λ(α1,α2)\lambda=\lambda(\alpha_{1},\alpha_{2}) is a finite number, we call a function uu in W1,2W^{1,2} the extremal function of the variational problem (3.1) if u2=1\|u\|_{2}=1 and

λ=X(|u|2+α1u2α22u2logu2)𝑑m.\lambda=\int_{X}\left(|\nabla u|^{2}+\alpha_{1}u^{2}-\frac{\alpha_{2}}{2}u^{2}\log u^{2}\right)dm. (3.2)

In the following, we provide our main result of this section. It states the existence of non-negative extremal functions of the variational problem (3.1). Moreover, we show that all non-negative extremal functions are actually Lipschitz continuous and bounded away from zero on XX. As a corollary, we show that the logarithmic transform of any non-negative extremal functions is Lipschitz and in the domain of the Laplacian and satisfies some Euler-Lagrange equation.

Theorem 3.2.

Let (X,d,m)(X,d,m) be a compact RCD(K,N)\mathrm{RCD}^{*}(K,N) spaces with KK\in\mathbb{R} and N(2,)N\in(2,\infty), and let α1\alpha_{1} and α2\alpha_{2} be given constants with α2>0\alpha_{2}>0. Then the log Sobolev constant λ=λ(α1,α2)\lambda=\lambda(\alpha_{1},\alpha_{2}) has finite value and the variational problem (3.1) admits non-negative extremal functions. Moreover, any non-negative extremal function uu satisfies the following properties:

  1. (i)

    uu is in D(Δ)D(\Delta) and satisfies

    Δu=α2ulogu+(λα1)u.-\Delta u=\alpha_{2}u\log u+(\lambda-\alpha_{1})u. (3.3)

    Furthermore, if λα1\lambda\neq\alpha_{1}, then uu is non-constant.

  2. (ii)

    uu is Lipschitz continuous;

  3. (iii)

    uu is positive.333Since XX is compact, u>δ>0u>\delta>0 for some δ\delta.

Corollary 3.3.

Let α1\alpha_{1} and α2\alpha_{2} be given constants with α2>0\alpha_{2}>0 and uu be an arbitrary non-negative extremal function of (3.1). Then v:=loguv:=\log u is Lipschitz and in D(Δ)D(\Delta) and satisfies

Δv=|v|2+α2v+λα1.-\Delta v=|\nabla v|^{2}+\alpha_{2}v+\lambda-\alpha_{1}. (3.4)

In particular, the equation Δf=|f|2+α2f-\Delta f=|\nabla f|^{2}+\alpha_{2}f admits Lipschitz weak solutions which are non-constant whenever λα1\lambda\neq\alpha_{1}.

Remark 3.4.

A classical example of the above variational problem is the weak log-Sobolev inequalities, that is, the variational problems

λε=inf{2αεX|f|2dm+X(εf2f2logf2)dm:fW1,2,f2=1},\lambda_{\varepsilon}=\inf\left\{\frac{2}{\alpha_{\varepsilon}}\int_{X}|\nabla f|^{2}dm+\int_{X}\left(\varepsilon f^{2}-f^{2}\log f^{2}\right)dm:f\in W^{1,2},\|f\|_{2}=1\right\}, (3.5)

where ε0\varepsilon\geq 0 and αε\alpha_{\varepsilon} is defined as the supremum of those C>0C>0 such that

Xf2logf2dmX(2C|f|2+εf2)𝑑m\int_{X}f^{2}\log f^{2}dm\leq\int_{X}\left(\frac{2}{C}|\nabla f|^{2}+\varepsilon f^{2}\right)dm

for all ff in W1,2W^{1,2} and f2=1\|f\|_{2}=1.

The log-Sobolev inequality on the RCD(K,N)\mathrm{RCD}^{*}(K,N) space with K>0K>0 and N(1,)N\in(1,\infty), see [14], implies that αεKN/(N1)\alpha_{\varepsilon}\geq KN/(N-1) for all ε0\varepsilon\geq 0, and αε<\alpha_{\varepsilon}<\infty when ε\varepsilon is small enough. A straightforward inspection shows that if αε\alpha_{\varepsilon} has finite value, then λε=0\lambda_{\varepsilon}=0. In this case, by using Theorem 3.2 and Corollary 3.3, we obtain that the weak Sobolev inequality admits Lipschitz and positive extremal functions which are further non-constant if ε>0\varepsilon>0. Furthermore, we derive that the equation Δf=|f|2+αεf-\Delta f=|\nabla f|^{2}+\alpha_{\varepsilon}f admits non-constant, Lipschitz and positive weak solutions if ε>0\varepsilon>0.

We first prove all assertions of Theorem 3.2 but positivity. While the existence result follows classical methods in [45] using variational techniques together with the compact Sobolev embedding, the proofs of the boundedness and Lipschitz regularity are different. Our methods are based on the ultracontractive property of the resolvent and the local regularity result of the Poisson equation on a ball in Lemma 2.6. In the following proofs, we denote by C>0C>0 a universal constant which may vary from one step to the next.

First part of the proof of Theorem 3.2.
  1. Step 1:

    We show the existence of non-negative extremal functions of variational problem (3.1). Let :={fW1,2:f2=1}\mathcal{F}:=\{f\in W^{1,2}:\|f\|_{2}=1\} and F:F:\mathcal{F}\rightarrow\mathbb{R} be the log Sobolev functional in (3.1), that is,

    F(f):=Fα1,α2(f)=X(|f|2+α1f2α22f2logf2)𝑑m,for f.F(f):=F_{\alpha_{1},\alpha_{2}}(f)=\int_{X}\left(|\nabla f|^{2}+\alpha_{1}f^{2}-\frac{\alpha_{2}}{2}f^{2}\log f^{2}\right)dm,\quad\text{for }f\in\mathcal{F}. (3.6)

    We claim that FF is coercive on \mathcal{F}. Indeed, choosing 0<δ<N/(N2)0<\delta<N/(N-2), since f2=1\|f\|_{2}=1, by Jensen’s inequality, it follows that

    f2logf2dm=1δlog|f|2δ(f2dm)1+δδlogf2+2δ2.\int f^{2}\log f^{2}dm=\frac{1}{\delta}\int\log|f|^{2\delta}(f^{2}dm)\leq\frac{1+\delta}{\delta}\log\|f\|^{2}_{2+2\delta}. (3.7)

    Note that the Sobolev inequality, see Lemma 2.8, implies that

    f2+2δ2f22f22+ACh(f)C(1+Ch(f)),\|f\|^{2}_{2+2\delta}\leq\|f\|^{2}_{2^{*}}\leq\|f\|^{2}_{2}+A\cdot\mathrm{Ch}(f)\leq C(1+\mathrm{Ch}(f)), (3.8)

    and therefore

    F(f)2Ch(f)+α1(1+δ)α22δlogC(1+Ch(f)),F(f)\geq 2\mathrm{Ch}(f)+\alpha_{1}-\frac{(1+\delta)\alpha_{2}}{2\delta}\log C\left(1+\mathrm{Ch}(f)\right), (3.9)

    which implies that FF\rightarrow\infty for ff in \mathcal{F} with fW1,2\|f\|_{W^{1,2}}\rightarrow\infty. Let (fn)(f_{n})\subseteq\mathcal{F} be a minimizing sequence which can be assumed non-negative since Ch(|f|)Ch(f)\mathrm{Ch}(|f|)\leq\mathrm{Ch}(f). By the coercivity of FF, it follows that (fn)(f_{n}) is bounded in W1,2W^{1,2}. The compact Sobolev embedding implies the existence of a non-negative uu in W1,2W^{1,2}\cap\mathcal{F} and a subsequence of (fn)(f_{n}), relabelled as (fn)(f_{n}), such that fnuf_{n}\rightarrow u strongly in LqL^{q} for all qq in [1,2)[1,2^{*}). By the L2L^{2}-lower semicontinuity of the Cheeger energy as well as the L2+δL^{2+\delta}-continuity of f2logf2\int f^{2}\log f^{2} for small δ>0\delta>0,444Which can be shown from mean value theorem and equality f2logf2g2logg2=2(|f||g|)(1+logθ2)θf^{2}\log f^{2}-g^{2}\log g^{2}=2(|f|-|g|)(1+\log\theta^{2})\theta where θ\theta is between |f||f| and |g||g|, see [45, page 112] it follows that FF reaches its minimum at uu. Furthermore, if α1λ\alpha_{1}\neq\lambda, a straightforward inspection shows that u1u\equiv 1 cannot be an extremal function.

  2. Step 2:

    Given any non-negative extremal function uu, we show that uD(Δ)u\in D(\Delta) and satisfies (3.3) by using the classical Euler-Lagrangian method. Let ϕLipbs\phi\in\mathrm{Lip}_{bs} be an arbitrary Lipschitz function and define the functional GG as follows

    G(t,β)=F(u+tϕ)+β(u+tϕ221).G(t,\beta)=F(u+t\phi)+\beta\left(\|u+t\phi\|_{2}^{2}-1\right). (3.10)

    The mean value theorem yields

    |u2logu2(u+tϕ)2log(u+tϕ)2|2t|ϕ||(1+logθ2)θ|\left|u^{2}\log u^{2}-(u+t\phi)^{2}\log(u+t\phi)^{2}\right|\leq 2t|\phi|\left|(1+\log\theta^{2})\theta\right| (3.11)

    with a function θ\theta taking values between |u||u| and |u+tϕ||u+t\phi|. From the inequality |θlogθ|max(1/e,C(δ)θ1+δ)|\theta\log\theta|\leq\max(1/e,C(\delta)\theta^{1+\delta}) for some small δ>0\delta>0 together with the dominated convergence of Lebesgue, it follows that

    0=12dG(t,β)dt|t=0=(u,ϕ+α1uϕα22uϕlogu2α2uϕ+βuϕ)𝑑m.0=\frac{1}{2}\frac{dG(t,\beta)}{dt}\Big{|}_{t=0}=\int\left(\langle\nabla u,\nabla\phi\rangle+\alpha_{1}u\phi-\frac{\alpha_{2}}{2}u\phi\log u^{2}-\alpha_{2}u\phi+\beta u\phi\right)dm. (3.12)

    Since Lipbs\mathrm{Lip}_{bs} is dense in W1,2W^{1,2} and uloguu\log u is in L2L^{2} by the Sobolev inequality, it follows that

    u,ϕ𝑑m=(α22ϕulogu2+(α2α1β)ϕu)𝑑m,for all ϕW1,2.\int\langle\nabla u,\nabla\phi\rangle dm=\int\left(\frac{\alpha_{2}}{2}\phi u\log u^{2}+(\alpha_{2}-\alpha_{1}-\beta)\phi u\right)dm,\quad\text{for all }\phi\in W^{1,2}. (3.13)

    Plugging ϕ=u\phi=u into (3.13) and using the real value of the log Sobolev constant λ\lambda, it follows that α2β=λ\alpha_{2}-\beta=\lambda. By the definition of the Laplacian we deduce that uu is in D(Δ)D(\Delta) and that Δu=α2ulogu+(λα1)u-\Delta u=\alpha_{2}u\log u+(\lambda-\alpha_{1})u.

  3. Step 3:

    We show that uu is in Lip\mathrm{Lip}. We start by showing that uu is in LL^{\infty}. Let β>0\beta>0. Note that using the resolvent of the Laplacian, (3.3) can be rewritten as

    u=(βIΔ)1(α2ulogu+(β+λα1)u)=Rβ(α2ulogu+α3u),u=\left(\beta I-\Delta\right)^{-1}\left(\alpha_{2}u\log u+(\beta+\lambda-\alpha_{1})u\right)=R_{\beta}\left(\alpha_{2}u\log u+\alpha_{3}u\right), (3.14)

    where II is the identity map and RβR_{\beta} is the resolvent of the Laplacian and α3:=β+λα1\alpha_{3}:=\beta+\lambda-\alpha_{1}. Note that in order to prove that uu is in LL^{\infty}, it is sufficient to show that g:=α2ulogu+α3ug:=\alpha_{2}u\log u+\alpha_{3}u is in LrL^{r} for some r>N/2r>N/2, according to Lemma 2.9. By the Sobolev inequality, we know that uu is in L2L^{2^{*}}, which implies that gg is in LrL^{r} for all r(2,2)r\in(2,2^{*}), (recall that 2=2N/(N2)2^{\ast}=2N/(N-2)). Fix δ2N\delta\geq 2N. It implies that 2<r1<22<r_{1}<2^{*} for 1/r1=1/2+1/δ1/r_{1}=1/2^{*}+1/\delta. If r1>N/2r_{1}>N/2, then Lemma 2.9 and the identity (3.14) implies that uLu\in L^{\infty}. If r1N/2r_{1}\leq N/2, then Lemma 2.9 implies that uLru\in L^{r} for all r<r1N/(N2r1)r<r_{1}N/(N-2r_{1}). Repeating the previous step, we can obtain that gLr2g\in L^{r_{2}} for some r2<r1N/(N2r1)r_{2}<r_{1}N/(N-2r_{1}) such that

    1r2=1r12N+1δ=122N+2δ.\frac{1}{r_{2}}=\frac{1}{r_{1}}-\frac{2}{N}+\frac{1}{\delta}=\frac{1}{2^{*}}-\frac{2}{N}+\frac{2}{\delta}. (3.15)

    Define by induction 1/rk:=1/22(k1)/N+k/δ1/r_{k}:=1/2^{*}-2(k-1)/N+k/\delta if rk1N/2r_{k-1}\leq N/2 which by definition of δ\delta implies that after finitely many iterations gg is in LrkL^{r_{k}} for some rk>N/2r_{k}>N/2 and deduce that uu is in LL^{\infty}.

    We now show that uu is actually in Lip\mathrm{Lip}. Let x0x_{0} be any point in XX and BR=B(x0,R)B_{R}=B(x_{0},R) be an open ball with 0<R<diam(X)/30<R<\mathrm{diam}(X)/3 and g:=α2ulogu+(λα1)ug:=\alpha_{2}u\log u+(\lambda-\alpha_{1})u. Since uu is in D(Δ)D(\Delta) and the identity (3.3), it follows by the definition that the restriction of uu on BRB_{R}, which we still denote by uu, belongs to W1,2(BR)W^{1,2}(B_{R}) as well as D(Δ,BR)D(\Delta,B_{R}) and that ΔBRu=g\Delta_{B_{R}}u=-g holds on BRB_{R} in the distributional sense. Hence, by Lemma 2.6, it follows that |u|Lloc(BR)|\nabla u|\in L^{\infty}_{loc}(B_{R}) and

    |u|L(BR/2)C(N,K,R)(1m(BR)uL2+gL).\||\nabla u|\|_{L^{\infty}(B_{R/2})}\leq C(N,K,R)\left(\frac{1}{m(B_{R})}\|u\|_{L^{2}}+\|g\|_{L^{\infty}}\right). (3.16)

    However, XX is compact and m(BR)>0m(B_{R})>0 by the doubling property, implying that |u||\nabla u| belongs to LL^{\infty}. The Sobolev-to-Lipschitz property thus implies that uu has a Lipschitz representation ending the proof of Theorem 3.2 but positivity.

As for the last assertion of Theorem 3.2, the positivity of non-negative extremal functions, we address the following auxiliary lemma stating that any non-negative extremal function vanishing at one point must also vanish in a neighborhood of that point. Our approach is based on a maximum principle type result for the De Giorgi class proved in [34].

Lemma 3.5.

Suppose that the hypothesis of Theorem 3.2 holds. Let uu be any non-negative extremal function of variational problem (3.1). Assume that u(x0)=0u(x_{0})=0 for some x0x_{0} in XX, then u0u\equiv 0 on a neighborhood of x0x_{0}.

Proof.

From the first part of the proof of Theorem 3.2, any non-negative extremal function of (3.2) is Lipschitz continuous. Furthermore, since α2>0\alpha_{2}>0 and λ\lambda is finite, the function g(t):=α2tlogt+(λα1)t0g(t):=\alpha_{2}t\log t+(\lambda-\alpha_{1})t\leq 0 for all small enough t>0t>0. Hence we can find r0>0r_{0}>0 such that g(u)0g(u)\leq 0 on B(x0,r0)B(x_{0},r_{0}). We first claim that u-u is of De Giorgi class DG2(B(x0,r0))\mathrm{DG}_{2}(B(x_{0},r_{0})). In other terms, there exists C>0C>0 such that for all kk in \mathbb{R} and zz in B(x0,r0)B(x_{0},r_{0}) and all 0<ρ<Rdiam(X)/30<\rho<R\leq\mathrm{diam}(X)/3 with B(z,R)B(x0,r0)B(z,R)\subseteq B(x_{0},r_{0}), it holds that

Az(k,ρ)|u|2𝑑mC(Rρ)2Az(k,R)(uk)2𝑑m,\int_{A_{z}(k,\rho)}|\nabla u|^{2}dm\leq\frac{C}{(R-\rho)^{2}}\int_{A_{z}(k,R)}(-u-k)^{2}dm, (3.17)

where Az(k,r):={xB(z,r):u(x)>k}A_{z}(k,r):=\{x\in B(z,r):-u(x)>k\}. In the following, CC denotes a positive constant, which is independent of the choice of z,ρ,r,R,kz,\rho,r,R,k, and may vary from line to line. Let η\eta be a Lipschitz cut-off function such that η=1\eta=1 on B(z,ρ)B(z,\rho) and supp(η)B(z,R)\mathrm{supp}(\eta)\subseteq B(z,R) with |η|C/(Rρ)|\nabla\eta|\leq C/(R-\rho) for some C>0C>0. Taking ϕ=η(uk)+\phi=\eta(-u-k)_{+} as test function for (3.3), it follows that

(η(uk)+),u𝑑m=η(uk)+(α2ulogu+(λα1)u)𝑑m.\int\langle\nabla(\eta(-u-k)_{+}),\nabla u\rangle dm=\int\eta(-u-k)_{+}\left(\alpha_{2}u\log u+(\lambda-\alpha_{1})u\right)dm. (3.18)

As for the left-hand side of (3.18), the Leibniz rule yields

(η(uk)+),u𝑑m{u>k}η|u|2𝑑m{u>k}(uk)+|η||u|𝑑mAz(k,ρ)|u|2𝑑mC(Rρ)2Az(k,R)(uk)+2𝑑mAz(k,R)Az(k,ρ)C|u|2𝑑m,\int\langle\nabla(\eta(-u-k)_{+}),\nabla u\rangle dm\geq\int_{\{-u>k\}}\eta|\nabla u|^{2}dm-\int_{\{-u>k\}}(-u-k)_{+}|\nabla\eta||\nabla u|dm\\ \geq\int_{A_{z}(k,\rho)}|\nabla u|^{2}dm-\frac{C}{(R-\rho)^{2}}\int_{A_{z}(k,R)}(-u-k)^{2}_{+}dm-\int_{A_{z}(k,R)\setminus A_{z}(k,\rho)}C|\nabla u|^{2}dm, (3.19)

where we use the locality of the minimal weak upper gradient |(uk)|=|u||\nabla(-u-k)|=|\nabla u| for the first inequality, and Young’s inequality together with |η|C/(Rρ)|\nabla\eta|\leq C/(R-\rho) and |η|=0|\nabla\eta|=0 on B(z,ρ)B(z,\rho) for the second inequality. As for the right hand side of (3.18), by the very choice of B(x0,r0)B(x_{0},r_{0}) and the non-negativity of uu, it follows that α2ulogu+(λα1)u0\alpha_{2}u\log u+(\lambda-\alpha_{1})u\leq 0. Hence

η(uk)+(α2ulogu+(λα1)u)𝑑m0.\int\eta(-u-k)_{+}\left(\alpha_{2}u\log u+(\lambda-\alpha_{1})u\right)dm\leq 0. (3.20)

With (3.19) and (3.20), we obtain that

Az(k,ρ)|u|2𝑑mC(Rρ)2Az(k,R)(uk)+2𝑑m+CAz(k,R)Az(k,ρ)|u|2𝑑m.\int_{A_{z}(k,\rho)}|\nabla u|^{2}dm\leq\frac{C}{(R-\rho)^{2}}\int_{A_{z}(k,R)}(-u-k)^{2}_{+}dm+C\int_{A_{z}(k,R)\setminus A_{z}(k,\rho)}|\nabla u|^{2}dm. (3.21)

Adding CAz(k,ρ)|u|2𝑑mC\int_{A_{z}(k,\rho)}|\nabla u|^{2}dm and then dividing by (1+C)(1+C) on both sides, it follows that

Az(k,ρ)|u|2𝑑mC(Rρ)2Az(k,R)(uk)+2𝑑m+θAz(k,R)|u|2𝑑m,\int_{A_{z}(k,\rho)}|\nabla u|^{2}dm\leq\frac{C}{(R-\rho)^{2}}\int_{A_{z}(k,R)}(-u-k)^{2}_{+}dm+\theta\int_{A_{z}(k,R)}|\nabla u|^{2}dm, (3.22)

where θ=C/(1+C)(0,1)\theta=C/(1+C)\in(0,1). Applying the same argument to 0<ρ<rR0<\rho<r\leq R and enlarging the domain of integral, we derive that for all 0<ρ<rR0<\rho<r\leq R,

Az(k,ρ)|u|2𝑑mC(rρ)2Az(k,R)(uk)+2𝑑m+θAz(k,r)|u|2𝑑m.\int_{A_{z}(k,\rho)}|\nabla u|^{2}dm\leq\frac{C}{(r-\rho)^{2}}\int_{A_{z}(k,R)}(-u-k)^{2}_{+}dm+\theta\int_{A_{z}(k,r)}|\nabla u|^{2}dm. (3.23)

By [34, Lemma 3.2] (see also [22, Lemma 3.1 in p. 161]) with f(r):=Az(k,r)|u|2𝑑mf(r):=\int_{A_{z}(k,r)}|\nabla u|^{2}dm, we obtain that

Az(k,ρ)|u|2𝑑mC(Rρ)2Az(k,R)(uk)+2𝑑m,\int_{A_{z}(k,\rho)}|\nabla u|^{2}dm\leq\frac{C}{(R-\rho)^{2}}\int_{A_{z}(k,R)}(-u-k)^{2}_{+}dm, (3.24)

which implies that u-u is of De Giorgi class DG2(B(x0,r0))\mathrm{DG}_{2}(B(x_{0},r_{0})).

Note that by the assumption that XX is a compact RCD(K,N)\mathrm{RCD}^{*}(K,N) space with KK\in\mathbb{R} and N(2,)N\in(2,\infty), it follows that XX is global doubling and supports the global weak (1,1)(1,1)-Poincaré inequality, see [43]. Together with Hölder’s inequality, XX supports the global weak (1,q)(1,q)-Poincare inequality for any qq in (1,2)(1,2). Since the minimal weak upper gradient in the RCD\mathrm{RCD} setting coincides with the minimal weak upper gradient in the Newtonian setting, see [2], it follows that the RCD(K,N)\mathrm{RCD}^{*}(K,N) space together with that u0u\geq 0 and uDG2(B(x0,r0))-u\in\mathrm{DG}_{2}(B(x_{0},r_{0})) satisfies the assumptions of [34, Lemma 6.1 and Lemma 6.2].

We now prove the assertion by contradiction. Suppose that the assertion of our Lemma does not hold. Let then 0<R<r00<R<r_{0} and xx in B(x0,R)B(x_{0},R) and τ:=u(x)>0\tau:=u(x)>0 be fixed. By the Lipschitz continuity of uu, we can find 0<rR0<r\leq R with B(x,r)B(x0,R)B(x,r)\subseteq B(x_{0},R) such that uτ/2u\geq\tau/2 on B(x,r)B(x,r). The generalized Bishop-Gromov inequality yields555For this fact, see [39, Theorem 2.14] or [49, Corollary 30.12]

m(B(x,r))m(B(x0,R))C(K,N,R)(rR)N,\frac{m(B(x,r))}{m(B(x_{0},R))}\geq C(K,N,R)\left(\frac{r}{R}\right)^{N}, (3.25)

for some constant C(K,N,R)>0C(K,N,R)>0 depending only on K,N,RK,N,R. Taking rr even smaller if necessary, we can assume that 0<C(K,N,R)(r/R)N<10<C(K,N,R)(r/R)^{N}<1. Hence it follows that

m({zB(x0,R):u(z)τ/2})m(B(x,r))C(K,N,R)(rR)Nm(B(x0,R)).m\left(\left\{z\in B(x_{0},R):u(z)\geq\tau/2\right\}\right)\geq m\left(B(x,r)\right)\geq C(K,N,R)\left(\frac{r}{R}\right)^{N}m(B(x_{0},R)). (3.26)

Taking γ:=1C(K,N,R)(r/R)N\gamma:=1-C(K,N,R)(r/R)^{N}, the inequality (3.26) implies that

m({xB(x0,R):u<τ/2})γm(B(x0,R)).m\left(\left\{x\in B(x_{0},R):u<\tau/2\right\}\right)\leq\gamma\cdot m(B(x_{0},R)). (3.27)

Since 0<γ<10<\gamma<1, [34, Lemma 6.2] yields the existence of λ¯=λ¯(γ)>0\bar{\lambda}=\bar{\lambda}(\gamma)>0 such that

u(x0)=infB(x0,R/2)uλ¯τ2>0,u(x_{0})=\inf_{B(x_{0},R/2)}u\geq\frac{\bar{\lambda}\tau}{2}>0, (3.28)

which is a contradiction. ∎

We can now address the positivity assertion of Theorem 3.2.

Final part of the proof of Theorem 3.2.
  1. Step 4:

    We show the positivity of non-negative extremal functions. From the continuity of uu, the set A={x:u(x)=0}A=\{x\colon u(x)=0\} is closed. By contradiction, suppose that AA is non-empty. By Lemma 3.5, it follows that AA is also open. Since XX is connected, it follows that A=XA=X and therefore u0u\equiv 0. This however contradicts the fact that u2=1\|u\|_{2}=1.

Finally we address the proof of Corollary 3.3.

Proof of Corollary 3.3.

Let ϕ(t)=logt,t>0\phi(t)=\log t,t>0 and uu be an arbitrary non-negative extremal function of the variational problem (3.1). By Theorem 3.2, we know that 0<cuC0<c\leq u\leq C for some some positive constants cc and CC and that uu is Lipschitz. Since ϕ\phi is a C2C^{2}-function with bounded first and second derivatives on [c,C][c,C], it follows by the chain rule that v:=ϕ(u)v:=\phi(u) is Lipschitz, vv is in D(Δ)D(\Delta) and

Δ(ϕ(u))=ϕ(u)Δu+ϕ′′(u)|u|2=1uΔu1u2|u|2.\Delta(\phi(u))=\phi^{\prime}(u)\Delta u+\phi^{\prime\prime}(u)|\nabla u|^{2}=\frac{1}{u}\Delta u-\frac{1}{u^{2}}|\nabla u|^{2}. (3.29)

By (3.3) we get

Δv=|v|2+α2v+(λα1).-\Delta v=|\nabla v|^{2}+\alpha_{2}v+(\lambda-\alpha_{1}). (3.30)

Taking v~=v+(λα1)/α2\tilde{v}=v+(\lambda-\alpha_{1})/\alpha_{2}. By locality of the minimal weak upper gradient and the Laplacian, we obtain that v~D(Δ)\tilde{v}\in D(\Delta) satisfies Δv~=|v~|2+α2v~-\Delta\tilde{v}=|\nabla\tilde{v}|^{2}+\alpha_{2}\tilde{v}. If λα1\lambda\neq\alpha_{1}, then Theorem 3.2 implies that uu is non-constant, which also implies that v~\tilde{v} is non-constant. Since v~\tilde{v} satisfies (3.4), we obtain the result. ∎

4. Li-Yau type inequality for logarithmic extremal functions

In this section, we derive a Li-Yau type estimate for the Lipschitz solutions of the equation Δv=|v|2+αv-\Delta v=|\nabla v|^{2}+\alpha v, whose existence is guaranteed by Corollary 3.3. In particular, based on the regularity and positivity results obtained in the previous section, this Li-Yau estimate holds for any logarithmic transform of non-negative extremal functions of (3.1).

Theorem 4.1.

Let (X,d,m)(X,d,m) be a compact RCD(K,N)\mathrm{RCD}^{*}(K,N) space with KK\in\mathbb{R} and NN in (2,)(2,\infty). Let vLipD(Δ)v\in\mathrm{Lip}\cap D(\Delta) such that Δv=|v|2+αv-\Delta v=|\nabla v|^{2}+\alpha v for some α>0\alpha>0. Then, for all 0<β<10<\beta<1 it holds

|v|2+(αβK)vNα(1β)4β(1β((2β)Kα)2α(1β))2,m-a.e.|\nabla v|^{2}+(\alpha-\beta K)v\leq\frac{N\alpha(1-\beta)}{4\beta}\left(1-\frac{\beta((2-\beta)K-\alpha)}{2\alpha(1-\beta)}\right)^{2},\quad\text{$m$-a.e.} (4.1)
Corollary 4.2.

Let uu be any non-negative extremal function of (3.1) with the log-Sobolev constant λ(α1,α2)\lambda(\alpha_{1},\alpha_{2}) and α2>0\alpha_{2}>0. Then v:=logu+(λα1)/α2v:=\log u+(\lambda-\alpha_{1})/\alpha_{2} satisfies the Li-Yau type estimate (4.1) with α=α2\alpha=\alpha_{2}. Moreover, if K>0K>0 and 0<α2K0<\alpha_{2}\leq K, then any non-negative extremal function is constant.

The proof of Theorem 4.1 is divided into three parts: In the first step, we show the regularity for |v|2+(αβKv)|\nabla v|^{2}+(\alpha-\beta Kv) using the weak Bochner inequality (2.7). In the second step, following the similar computation arguments as in [50] we derive a lower bound of the absolutely continuous part of 𝚫(|v|2+(αβKv))\bm{\Delta}(|\nabla v|^{2}+(\alpha-\beta Kv)), where all inequalities are understood in the mm-a.e. sense. In the last step, we make use of a slightly generalized Omori-Yau type maximum principle proved in Appendix A together with a “good” cut-off function inspired by [38] to derive the desired Li-Yau estimate.

Proof of Theorem 4.1.

Recall that Test:={fLipD(Δ)L:ΔfW1,2L}\mathrm{Test}^{\infty}:=\{f\in\mathrm{Lip}\cap D(\Delta)\cap L^{\infty}:\Delta f\in W^{1,2}\cap L^{\infty}\} and Test+:={fTest:f0 m-a.e. on X}\mathrm{Test}^{\infty}_{+}:=\{f\in\mathrm{Test}^{\infty}:f\geq 0\text{ $m$-a.e. on $X$}\}.

  1. Step 1:

    We claim that |v|2|\nabla v|^{2} is in W1,2LW^{1,2}\cap L^{\infty} and |v|2|\nabla v|^{2} is in D(𝚫)D(\bm{\Delta}) with 𝚫s(|v|2)0\bm{\Delta}^{s}(|\nabla v|^{2})\geq 0. Indeed, by the assumption that vv is in D(Δ)D(\Delta), it follows that vv belongs to H2,2H^{2,2} and

    ||v|2|2|Hessv|HS|v|,|\nabla|\nabla v|^{2}|\leq 2|\mathrm{Hess}v|_{HS}|\nabla v|, (4.2)

    with |Hessv|HSL2|\mathrm{Hess}v|_{HS}\in L^{2}. By the fact that |v||\nabla v| is in LL^{\infty}, we obtain that |v|2|\nabla v|^{2} is also in W1,2W^{1,2}. We now show that |v|2D(𝚫)|\nabla v|^{2}\in D(\bm{\Delta}). For any ϕTest+\phi\in\mathrm{Test}^{\infty}_{+}, by the weak Bochner inequality (2.7), it follows that

    X|v|2Δϕ𝑑m2Xϕ((Δv)2N+v,Δv+K|v|2)𝑑m=Xϕ𝑑μ,\int_{X}|\nabla v|^{2}\Delta\phi dm\geq 2\int_{X}\phi\left(\frac{(\Delta v)^{2}}{N}+\langle\nabla v,\nabla\Delta v\rangle+K|\nabla v|^{2}\right)dm=\int_{X}\phi d\mu, (4.3)

    where μ=2((Δv)2/N+v,Δv+K|v|2)m\mu=2((\Delta v)^{2}/N+\langle\nabla v,\nabla\Delta v\rangle+K|\nabla v|^{2})m. By the standard regularization via the mollified heat flow, see [25, Corollary 6.2.17], the inequality (4.3) holds for all ϕLipbs+\phi\in\mathrm{Lip}_{bs}^{+}. Then by [25, Proposition 6.2.16], it follows that |v|2D(𝚫)|\nabla v|^{2}\in D(\bm{\Delta}) and that

    𝚫(|v|2)2((Δv)2N+v,Δv+K|v|2)m.\bm{\Delta}\left(|\nabla v|^{2}\right)\geq 2\left(\frac{(\Delta v)^{2}}{N}+\langle\nabla v,\nabla\Delta v\rangle+K|\nabla v|^{2}\right)\cdot m. (4.4)

    In particular we obtain that 𝚫s(|v|2)0\bm{\Delta}^{s}(|\nabla v|^{2})\geq 0.

  2. Step 2:

    We provide a lower bound for (𝚫acg)(\bm{\Delta}^{ac}g) based on the inequality (4.4) where g:=|v|2+(αβK)vg:=|\nabla v|^{2}+(\alpha-\beta K)v for 0<β<10<\beta<1. First note that since Δv=|v|2+αv-\Delta v=|\nabla v|^{2}+\alpha v, it follows that Δv=g+βKv-\Delta v=g+\beta Kv. Hence, from the first step, we get that gg is in D(𝚫)D(\bm{\Delta}) and 𝚫sg=𝚫s(|v|2)0\bm{\Delta}^{s}g=\bm{\Delta}^{s}(|\nabla v|^{2})\geq 0. Then inequality (4.4) implies that

    𝚫acg=𝚫ac(|v|2+(αβK)v)2(Δv)2N+2v,Δv+2K|v|2+(αβK)Δv.\bm{\Delta}^{ac}g=\bm{\Delta}^{ac}(|\nabla v|^{2}+(\alpha-\beta K)v)\\ \geq 2\frac{(\Delta v)^{2}}{N}+2\langle\nabla v,\nabla\Delta v\rangle+2K|\nabla v|^{2}+(\alpha-\beta K)\Delta v. (4.5)

    Plugging Δv=gβKv\Delta v=-g-\beta Kv into (4.5), it follows that

    𝚫acg2(g+βKv)2N2v,(g+βKv)+2K|v|2(αβK)(g+βKv)=2N(g2+2βKgv+β2K2v2)2βK|v|22v,g+2K|v|2(αβK)gβK(αβK)v.\bm{\Delta}^{ac}g\geq 2\frac{(g+\beta Kv)^{2}}{N}-2\langle\nabla v,\nabla(g+\beta Kv)\rangle+2K|\nabla v|^{2}-(\alpha-\beta K)(g+\beta Kv)\\ =\frac{2}{N}\left(g^{2}+2\beta Kgv+\beta^{2}K^{2}v^{2}\right)-2\beta K|\nabla v|^{2}-2\langle\nabla v,\nabla g\rangle+2K|\nabla v|^{2}\\ -(\alpha-\beta K)g-\beta K(\alpha-\beta K)v. (4.6)

    Plugging identity |v|2=g(αβK)v|\nabla v|^{2}=g-(\alpha-\beta K)v into the right hand of (4.6), it follows that

    𝚫acg2Ng2+(4βKNv+(2β)Kα)g+2β2K2Nv2K(2β)(αβK)v2v,g=2N[g+(βKv+N[(2β)Kα]4)]22N(βKv+N[(2β)Kα]4)2+2β2K2Nv2K(2β)(αβK)v2v,g.\bm{\Delta}^{ac}g\geq\frac{2}{N}g^{2}+\left(\frac{4\beta K}{N}v+(2-\beta)K-\alpha\right)g+\frac{2\beta^{2}K^{2}}{N}v^{2}\\ -K(2-\beta)(\alpha-\beta K)v-2\langle\nabla v,\nabla g\rangle\\ =\frac{2}{N}\left[g+\left(\beta Kv+\frac{N[(2-\beta)K-\alpha]}{4}\right)\right]^{2}-\frac{2}{N}\left(\beta Kv+\frac{N[(2-\beta)K-\alpha]}{4}\right)^{2}\\ +\frac{2\beta^{2}K^{2}}{N}v^{2}-K(2-\beta)(\alpha-\beta K)v-2\langle\nabla v,\nabla g\rangle. (4.7)

    For a=N((2β)Kα)a=N((2-\beta)K-\alpha) and b=Nα(1β)/βb=N\alpha(1-\beta)/\beta, inequality (4.7) simplifies to

    𝚫acg2N(g+βKv+a4)22N(Kβbv)2Na2162v,g.\bm{\Delta}^{ac}g\geq\frac{2}{N}(g+\beta Kv+\frac{a}{4})^{2}-\frac{2}{N}(K\beta bv)-\frac{2}{N}\frac{a^{2}}{16}-2\langle\nabla v,\nabla g\rangle. (4.8)
  3. Step 3:

    We show that the assumptions of Lemma A.2 are valid for gg based on the estimate (4.8). Let D:=diam(X)D:=\mathrm{diam}(X) and fix 0<R<D/40<R<D/4. Since XX is compact and gLg\in L^{\infty}, we can find x0x_{0} in XX such that:

    M1:=esssupB(x0,R)g=esssupXgM_{1}:=\operatorname*{ess\,sup}_{B(x_{0},R)}g=\operatorname*{ess\,sup}_{X}g (4.9)

    Define

    M2:=esssupXB(x0,R)gM1M_{2}:=\operatorname*{ess\,sup}_{X\setminus B(x_{0},R)}g\leq M_{1}

    By our choice of RR and the doubling property of XX, we have m(B(x0,R))>0m(B(x_{0},R))>0 and m(XB(x0,R))>0m(X\setminus B(x_{0},R))>0. Without loss of generality, we assume that M1>0M_{1}>0, otherwise nothing needs to be shown. We consider different cases for M1M_{1} and M2M_{2}.

    1. Case 1:

      M1>M2M_{1}>M_{2}. By the regularity result in Step 1, we can apply Lemma A.2 to gg. Hence taking w=2vw=2v which is in W1,2LipbW^{1,2}\cap\mathrm{Lip}_{b}, it follows from Lemma A.2 that we can find a sequence (xj)X(x_{j})\subseteq X such that g(xj)>M11/jg(x_{j})>M_{1}-1/j and

      𝚫acg(xj)+g,w(xj)1/j.\bm{\Delta}^{ac}g(x_{j})+\langle\nabla g,\nabla w\rangle(x_{j})\leq 1/j. (4.10)

      Plugging (4.8) into (4.10) and letting hv,j=(Kβbv+a2/16+N/(2j))1/2(xj)h_{v,j}=(K\beta bv+a^{2}/16+N/(2j))^{1/2}(x_{j}), it follows that

      g(xj)βKv(xj)a4+hv,j=1bhv,j2+hv,j+a216ba4+N2bj=1b(hv,jb2)2+b4(1a2b)2+N2bj.g(x_{j})\leq-\beta Kv(x_{j})-\frac{a}{4}+h_{v,j}\\ =-\frac{1}{b}h_{v,j}^{2}+h_{v,j}+\frac{a^{2}}{16b}-\frac{a}{4}+\frac{N}{2bj}\\ =-\frac{1}{b}\left(h_{v,j}-\frac{b}{2}\right)^{2}+\frac{b}{4}\left(1-\frac{a}{2b}\right)^{2}+\frac{N}{2bj}. (4.11)

      For jj\rightarrow\infty together with g(xj)>M11/jg(x_{j})>M_{1}-1/j and b>0b>0, we obtain that

      esssupXgb4(1a2b)2.\operatorname*{ess\,sup}_{X}g\leq\frac{b}{4}\left(1-\frac{a}{2b}\right)^{2}. (4.12)
    2. Case 2:

      M1=M2M_{1}=M_{2}. Let ε(0,1/2)\varepsilon\in(0,1/2). By [38, Lemma 3.1], we can find a Lipschitz cut-off function ψ:X\psi:X\rightarrow\mathbb{R} with ψD(Δ)\psi\in D(\Delta) such that 0ψ10\leq\psi\leq 1 and ψ1\psi\equiv 1 on B(x0,R)B(x_{0},R) and supp(ψ)B(x0,2R)\mathrm{supp}(\psi)\subseteq B(x_{0},2R), and

      R2|Δψ|+R|ψ|C,R^{2}|\Delta\psi|+R|\nabla\psi|\leq C, (4.13)

      where C>0C>0 is a constant depending only on K,NK,N and RR. Let ϕε:X\phi_{\varepsilon}:X\rightarrow\mathbb{R} be defined as ϕε:=1ε+εψ\phi_{\varepsilon}:=1-\varepsilon+\varepsilon\psi and let Gε:=ϕεgG_{\varepsilon}:=\phi_{\varepsilon}\cdot g. Note that since ϕεD(Δ)\phi_{\varepsilon}\in D(\Delta), it holds that 𝚫ϕε=Δϕεm\bm{\Delta}\phi_{\varepsilon}=\Delta\phi_{\varepsilon}\cdot m. Together with the fact that ϕε\phi_{\varepsilon} is continuous, by the Leibniz rule for the measure-valued Laplacian, it follows that

      𝚫(Gε)=ϕε𝚫g+gΔϕεm+2ϕε,gm.\bm{\Delta}(G_{\varepsilon})=\phi_{\varepsilon}\bm{\Delta}g+g\Delta\phi_{\varepsilon}\cdot m+2\langle\nabla\phi_{\varepsilon},\nabla g\rangle\cdot m. (4.14)

      Together with the result in the first step of this proof that 𝚫sg0\bm{\Delta}^{s}g\geq 0, we deduce that

      𝚫s(Gε)=ϕε𝚫sg0.\bm{\Delta}^{s}(G_{\varepsilon})=\phi_{\varepsilon}\bm{\Delta}^{s}g\geq 0. (4.15)

      Furthermore, from (4.13), it follows that

      |ϕε|2ϕεε2CR2and|Δϕε|εCR2.\frac{|\nabla\phi_{\varepsilon}|^{2}}{\phi_{\varepsilon}}\leq\varepsilon^{2}\frac{C}{R^{2}}\quad\text{and}\quad|\Delta\phi_{\varepsilon}|\leq\varepsilon\frac{C}{R^{2}}. (4.16)

      Denote by HvH_{v} the right side of (4.8) without the last term 2v,g-2\langle\nabla v,\nabla g\rangle. Then, by using g=Gε/ϕεg=G_{\varepsilon}/\phi_{\varepsilon} and the inequality (4.8), it follows that

      𝚫ac(Gε)=ϕε𝚫acg+GεϕεΔϕε+2ϕε,(Gεϕε)ϕε(Hv2v,g)+2ϕε,Gε/ϕε+Gεϕε(Δϕε2|ϕε|2ϕε).\bm{\Delta}^{ac}(G_{\varepsilon})=\phi_{\varepsilon}\bm{\Delta}^{ac}g+\frac{G_{\varepsilon}}{\phi_{\varepsilon}}\Delta\phi_{\varepsilon}+2\left\langle\nabla\phi_{\varepsilon},\nabla\left(\frac{G_{\varepsilon}}{\phi_{\varepsilon}}\right)\right\rangle\\ \geq\phi_{\varepsilon}\left(H_{v}-2\left\langle\nabla v,\nabla g\right\rangle\right)+2\left\langle\nabla\phi_{\varepsilon},\nabla G_{\varepsilon}\right\rangle/\phi_{\varepsilon}+\frac{G_{\varepsilon}}{\phi_{\varepsilon}}\left(\Delta\phi_{\varepsilon}-2\frac{|\nabla\phi_{\varepsilon}|^{2}}{\phi_{\varepsilon}}\right). (4.17)

      Using the estimate (4.16) with ε2<ε\varepsilon^{2}<\varepsilon, it follows that

      𝚫ac(Gε)ϕεHv2ϕε(v,Gεϕεv,Gεϕεϕε2)+2ϕε,Gε/ϕεεgCR2ϕεHv2(vlogϕε),Gε+2Gεϕεv,ϕεεgCR2ϕεHv2(vlogϕε),Gεεg|v|CRεgCR2.\bm{\Delta}^{ac}(G_{\varepsilon})\geq\phi_{\varepsilon}H_{v}-2\phi_{\varepsilon}\left(\left\langle\nabla v,\frac{\nabla G_{\varepsilon}}{\phi_{\varepsilon}}\right\rangle-\left\langle\nabla v,\frac{G_{\varepsilon}\nabla\phi_{\varepsilon}}{\phi_{\varepsilon}^{2}}\right\rangle\right)\\ +2\left\langle\nabla\phi_{\varepsilon},\nabla G_{\varepsilon}\right\rangle/\phi_{\varepsilon}-\varepsilon\|g\|_{\infty}\frac{C}{R^{2}}\\ \geq\phi_{\varepsilon}H_{v}-2\left\langle\nabla(v-\log\phi_{\varepsilon}),\nabla G_{\varepsilon}\right\rangle+2\frac{G_{\varepsilon}}{\phi_{\varepsilon}}\langle\nabla v,\nabla\phi_{\varepsilon}\rangle-\varepsilon\|g\|_{\infty}\frac{C}{R^{2}}\\ \geq\phi_{\varepsilon}H_{v}-2\langle\nabla(v-\log\phi_{\varepsilon}),\nabla G_{\varepsilon}\rangle-\varepsilon\|g\|_{\infty}\||\nabla v|\|_{\infty}\frac{C}{R}-\varepsilon\|g\|_{\infty}\frac{C}{R^{2}}. (4.18)

      Since esssupXB(x0,2R)gesssupXB(x0,R)g<M1\operatorname*{ess\,sup}_{X\setminus B(x_{0},2R)}g\leq\operatorname*{ess\,sup}_{X\setminus B(x_{0},R)}g<M_{1}, by the definition of ϕε\phi_{\varepsilon}, it follows that

      esssupXB(x0,2R)Gε(1ε)M1<M1=esssupB(x0,R)Gε=esssupB(x0,2R)Gε.\operatorname*{ess\,sup}_{X\setminus B(x_{0},2R)}G_{\varepsilon}\leq(1-\varepsilon)M_{1}<M_{1}=\operatorname*{ess\,sup}_{B(x_{0},R)}G_{\varepsilon}=\operatorname*{ess\,sup}_{B(x_{0},2R)}G_{\varepsilon}. (4.19)

      The doubling property and R<D/4R<D/4 implies that m(B(x0,2R))>0m(B(x_{0},2R))>0 as well as m(XB(x0,2R))>0m(X\setminus B(x_{0},2R))>0, which together with the fact that 𝚫s(Gε)0\bm{\Delta}^{s}(G_{\varepsilon})\geq 0 and GεW1,2LipbG_{\varepsilon}\in W^{1,2}\cap\mathrm{Lip}_{b}, allows us to apply Lemma A.2 to GεG_{\varepsilon}. Taking wε=2v2logϕεW1,2Lipbw_{\varepsilon}=2v-2\log\phi_{\varepsilon}\in W^{1,2}\cap\mathrm{Lip}_{b}, by Lemma A.2, it follows that there exists a sequence (xj)X(x_{j})\subseteq X such that Gε(xj)>esssupXGε1/jG_{\varepsilon}(x_{j})>\operatorname*{ess\,sup}_{X}G_{\varepsilon}-1/j and

      𝚫acGε(xj)+Gε,wε(xj)1/j.\bm{\Delta}^{ac}G_{\varepsilon}(x_{j})+\langle\nabla G_{\varepsilon},\nabla w_{\varepsilon}\rangle(x_{j})\leq 1/j. (4.20)

      Plugging (4.18), we obtain that

      ϕε(xj)Hv(xj)1j+εC1,\phi_{\varepsilon}(x_{j})H_{v}(x_{j})\leq\frac{1}{j}+\varepsilon C_{1}, (4.21)

      where C1=C(K,N,R)g(|f|/R+1/R2)C_{1}=C(K,N,R)\|g\|_{\infty}\left(\||\nabla f|\|_{\infty}/R+1/R^{2}\right). Following the similar argument as in Case 1 and noting that b>0b>0, we obtain that

      g(xj)b4(1a2b)2+N(j1+εC1)2bϕε(xj).g(x_{j})\leq\frac{b}{4}\left(1-\frac{a}{2b}\right)^{2}+\frac{N(j^{-1}+\varepsilon C_{1})}{2b\phi_{\varepsilon}(x_{j})}. (4.22)

      Since esssupXGε=esssupXg=M1\operatorname*{ess\,sup}_{X}G_{\varepsilon}=\operatorname*{ess\,sup}_{X}g=M_{1}, so multiplying ϕε(xj)\phi_{\varepsilon}(x_{j}) on both side of (4.22) and letting jj\rightarrow\infty, it follows that

      esssupXgb4(1a2b)2+εNC12b.\operatorname*{ess\,sup}_{X}g\leq\frac{b}{4}\left(1-\frac{a}{2b}\right)^{2}+\frac{\varepsilon NC_{1}}{2b}. (4.23)

      The inequality (4.23) holding for any 0<ε<1/20<\varepsilon<1/2, we send ε\varepsilon to 0 to obtain

      esssupXgb4(1a2b)2.\operatorname*{ess\,sup}_{X}g\leq\frac{b}{4}\left(1-\frac{a}{2b}\right)^{2}. (4.24)

    With Case 1 and Case 2, together with the definitions of aa and bb, we obtain the inequality (4.1).

Proof of Corollary 4.2.

By Theorem 3.2 and Corollary 3.3, we know that w:=loguw:=\log u is in LipD(Δ)\mathrm{Lip}\cap D(\Delta) and satisfies Δw=|w|2+α2w+λα1-\Delta w=|\nabla w|^{2}+\alpha_{2}w+\lambda-\alpha_{1}. Then by the definition of the Laplacian and locality of the minimal weak upper gradient, it follows that v=w+(λα1)/α2v=w+(\lambda-\alpha_{1})/\alpha_{2} is in LipD(Δ)\mathrm{Lip}\cap D(\Delta) and satisfies the equation Δv=|v|2+α2v-\Delta v=|\nabla v|^{2}+\alpha_{2}v. Hence, by Theorem 4.1, the Li-Yau estimate (4.1) holds for vv. For the second claim, let C(β)C(\beta) denote the term in the right-hand side of (4.1). If α2(0,K]\alpha_{2}\in(0,K], then taking β(0,α2/K)\beta\in(0,\alpha_{2}/K). One can check that C(β)C(\beta) goes to 0 as βα2/K\beta\nearrow\alpha_{2}/K. Hence, taking βα2/K\beta\nearrow\alpha_{2}/K on the both sides of (4.1), it follows that

|v|2=0,m-a.e.\left|\nabla v\right|^{2}=0,\quad\text{$m$-a.e.}

Then the Sobolev-to-Lip property implies that uu is constant. ∎

5. Applications

In this section, we present applications of the regularity and positivity results in Theorem 3.2 and the Li-Yau type estimate in Theorem 4.1, which generalize the results in [50] on the smooth Riemannian manifold to the RCD(K,N)\mathrm{RCD}^{*}(K,N) spaces.

For the notational simplicity, we define the non-negative constants C1C_{1} and C2C_{2} as follows:

C1(β):=αβKandC2(β):=Nα(1β)4β(1β((2β)Kα)2α(1β))2,C_{1}(\beta):=\alpha-\beta K\quad\text{and}\quad C_{2}(\beta):=\frac{N\alpha(1-\beta)}{4\beta}\left(1-\frac{\beta((2-\beta)K-\alpha)}{2\alpha(1-\beta)}\right)^{2}, (5.1)

where 0<β<10<\beta<1. Note that C1C_{1} is positive whenever α>max{K,0}\alpha>\max\{K,0\}.

The first direct consequence is a Harnack type inequality for the non-negative extremal functions.

Corollary 5.1.

Let (X,d,m)(X,d,m) be a compact RCD(K,N)\mathrm{RCD}^{*}(K,N) space with KK\in\mathbb{R} and NN in (2,)(2,\infty). Suppose that vLipv\in\mathrm{Lip} satisfies Δv=|v|2+αv-\Delta v=|\nabla v|^{2}+\alpha v for some α>max{K,0}\alpha>\max\{K,0\}. Then for any x,yXx,y\in X, it holds that

ev(x)e(1ε)v(y)exp(C1(β)d2(x,y)4ε+εC2(β)C1(β)),e^{v(x)}\leq e^{(1-\varepsilon)v(y)}\exp\left(\frac{C_{1}(\beta)d^{2}(x,y)}{4\varepsilon}+\frac{\varepsilon C_{2}(\beta)}{C_{1}(\beta)}\right), (5.2)

for any 0<ε<10<\varepsilon<1 and β(0,1)\beta\in(0,1).

Proof.

First note that by Theorem 4.1, we have

|v|C2(β)C1(β)v,m-a.e.|\nabla v|\leq\sqrt{C_{2}(\beta)-C_{1}(\beta)v},\quad\text{$m$-a.e.} (5.3)

Let x0,y0Xx_{0},y_{0}\in X be arbitrary points in XX and let μ0:=1m(B(x0,r))m|B(x0,r)\mu_{0}:=\frac{1}{m(B(x_{0},r))}m|_{B(x_{0},r)} and μ1:=1m(B(y0,r))m|B(y0,r)\mu_{1}:=\frac{1}{m(B(y_{0},r))}m|_{B(y_{0},r)} for 0<r<diam(X)/30<r<\mathrm{diam}(X)/3. By [44, Corollary 1.2], RCD(K,N)\mathrm{RCD}^{*}(K,N) spaces are essentially non-branching. Together with [13, Corollary 5.3], there exists a unique W2W_{2}-geodesic (μt){t[0,1]}(\mu_{t})_{\{t\in[0,1]\}} joining μ0\mu_{0} and μ1\mu_{1} with μtCm\mu_{t}\leq Cm for any tt in [0,1][0,1] for some C>0C>0, and a test plan π\pi in 𝒫(C([0,1];X))\mathcal{P}(C([0,1];X)) such that (et)π=μt(e_{t})_{\sharp}\pi=\mu_{t} for any tt in [0,1][0,1] and π\pi is concentrated on the set Geo(X)\mathrm{Geo}(X) of geodesics on XX. By the Fubini’s theorem together with the fact that π\pi has bounded compression and the inequality (5.3), it follows that for π\pi-a.s γGeo(X)\gamma\in\mathrm{Geo}(X),

|v(γt)|C2(β)C1(β)v(γt),for a.e. t[0,1].\left|\nabla v(\gamma_{t})\right|\leq\sqrt{C_{2}(\beta)-C_{1}(\beta)v(\gamma_{t})},\quad\text{for a.e. $t\in[0,1]$}.

Let γGeo(X)\gamma\in\mathrm{Geo}(X) be any such a geodesic. By the continuity of vv, let s1s_{1} and s2s_{2} in [0,1][0,1] be the maximum and minimum points respectively, that is

v(γs1)=maxs[0,1]v(γs)andv(γs2)=mins[0,1]v(γs).v(\gamma_{s_{1}})=\max_{s\in[0,1]}v(\gamma_{s})\quad\text{and}\quad v(\gamma_{s_{2}})=\min_{s\in[0,1]}v(\gamma_{s}). (5.4)

Then by the definition of weak upper gradients (see [25, Proposition 1.20]), it follows that

|v(γs1)v(γs2)|min(s1,s2)max(s1,s2)|v|(γt)|γ˙t|𝑑td(γ0,γ1)C2(β)C1(β)v(γs2).\left|v(\gamma_{s_{1}})-v(\gamma_{s_{2}})\right|\leq\int_{\min(s_{1},s_{2})}^{\max(s_{1},s_{2})}\left|\nabla v\right|(\gamma_{t})\left|\dot{\gamma}_{t}\right|dt\leq d(\gamma_{0},\gamma_{1})\sqrt{C_{2}(\beta)-C_{1}(\beta)v(\gamma_{s_{2}})}. (5.5)

For 0<ε<10<\varepsilon<1, together with v(γ0)v(γs1)v(\gamma_{0})\leq v(\gamma_{s_{1}}), it follows that

v(γ0)(1ε)v(γs2)+εv(γs2)+d(γ0,γ1)C2(β)C1(β)v(γs2).v(\gamma_{0})\leq(1-\varepsilon)v(\gamma_{s_{2}})+\varepsilon v(\gamma_{s_{2}})+d(\gamma_{0},\gamma_{1})\sqrt{C_{2}(\beta)-C_{1}(\beta)v(\gamma_{s_{2}})}. (5.6)

Let a=C2(β)C1(β)v(γs2)a=\sqrt{C_{2}(\beta)-C_{1}(\beta)v(\gamma_{s_{2}})}. Since C1(β)>0C_{1}(\beta)>0, it follows that

v(γ0)(1ε)v(γs2)εC1(β)a2+d(γ0,γ1)a+εC2(β)C1(β)(1ε)v(γ1)+C1(β)d2(γ0,γ1)4ε+εC2(β)C1(β).v(\gamma_{0})\leq(1-\varepsilon)v(\gamma_{s_{2}})-\frac{\varepsilon}{C_{1}(\beta)}a^{2}+d(\gamma_{0},\gamma_{1})a+\frac{\varepsilon C_{2}(\beta)}{C_{1}(\beta)}\\ \leq(1-\varepsilon)v(\gamma_{1})+\frac{C_{1}(\beta)d^{2}(\gamma_{0},\gamma_{1})}{4\varepsilon}+\frac{\varepsilon C_{2}(\beta)}{C_{1}(\beta)}. (5.7)

Integrating (5.7) with respect to π\pi on the both sides and noting that (et)π=μt(e_{t})_{\sharp}\pi=\mu_{t}, it follows that

v(x)𝑑μ0((1ε)v(y)+C1(β)d2(x,y)4ε+εC2(β)C1(β))𝑑π~(x,y),\int v(x)d\mu_{0}\leq\int\left((1-\varepsilon)v(y)+\frac{C_{1}(\beta)d^{2}(x,y)}{4\varepsilon}+\frac{\varepsilon C_{2}(\beta)}{C_{1}(\beta)}\right)d\tilde{\pi}(x,y), (5.8)

where π~=(e0,e1)π\tilde{\pi}=(e_{0},e_{1})_{\sharp}\pi is the optimal transport plan between μ0\mu_{0} and μ1\mu_{1}. Since μ0\mu_{0} and μ1\mu_{1} weakly converges to δx0\delta_{x_{0}} and δy0\delta_{y_{0}} in the duality of CbC_{b} as r0r\rightarrow 0, it follows that, up to a subsequence, π~\tilde{\pi} weakly converges to δx0×δy0\delta_{x_{0}}\times\delta_{y_{0}}. Using the identity u=exp(v)u=\exp(v), we obtain that

u(x0)u(y0)1εexp(C1(β)d2(x0,y0)4ε+εC2(β)C1(β)).u(x_{0})\leq u(y_{0})^{1-\varepsilon}\exp\left(\frac{C_{1}(\beta)d^{2}(x_{0},y_{0})}{4\varepsilon}+\frac{\varepsilon C_{2}(\beta)}{C_{1}(\beta)}\right). (5.9)

Remark 5.2.

Note that using the result proved in [18] that the minimal weak upper gradient of Lipschitz function coincides with its local Lipschitz slope in the complete doubling metric space supporting the weak (1,p)(1,p)-Poincaré inequality for p>1p>1, Corollary 5.1 can be shown directly following the lines of [50, Corollary 2.2] instead of using the optimal transport method.

The next corollary addresses the estimates on the upper and lower bounds of vv based on the dimension-free Harnack inequality on the RCD(K,)\mathrm{RCD}(K,\infty) space in [36]. The proof of which is essentially the same one as in the Riemannian case from [50, Corollary 2.4]. For the sake of completeness, we provide the proof.

Corollary 5.3.

Let (X,d,m)(X,d,m) be a compact RCD(K,N)\mathrm{RCD}^{*}(K,N) space with KK\in\mathbb{R} and NN in (2,)(2,\infty). Then for any non-negative extremal function uu of (3.1) with the log Sobolev constant λ(α1,α2)\lambda(\alpha_{1},\alpha_{2}) with α2>max{K,0}\alpha_{2}>\max\{K,0\}, it holds that

supXloguλα1α2+C3(α2,K)diam(X)2,\displaystyle\sup_{X}\log u\leq\frac{\lambda-\alpha_{1}}{\alpha_{2}}+C_{3}(\alpha_{2},K)\mathrm{diam}(X)^{2}, (5.10)
infXloguλα1α22C3(α2,K)diam(X)2\displaystyle\inf_{X}\log u\geq-\frac{\lambda-\alpha_{1}}{\alpha_{2}}-2C_{3}(\alpha_{2},K)\mathrm{diam}(X)^{2} (5.11)

where C3(α2,K)=α2C_{3}(\alpha_{2},K)=\alpha_{2} when K0K\geq 0 and C3(α2,K)=K/(exp(K/α2)1)C_{3}(\alpha_{2},K)=K/(\exp(K/\alpha_{2})-1) when K<0K<0. In particular, it holds that |logu|2C2(β)+2C1(β)C3(α2,K)diam(X)2|\nabla\log u|^{2}\leq C_{2}(\beta)+2C_{1}(\beta)C_{3}(\alpha_{2},K)\mathrm{diam}(X)^{2} mm-a.e. for any 0<β<10<\beta<1.

Proof.

Let D:=diam(X)D:=\mathrm{diam}(X) and xx and yy be the maximum and minimum point of uu on XX, respectively, the existence of which is guaranteed by the regularity of non-negative extremal functions proved in Theorem 3.2.

As for the upper bound, firstly, by the dimension-free Harnack inequality on RCD(K,)\mathrm{RCD}(K,\infty) spaces shown in [36, Theorem 3.1] for p>1p>1, it follows that

(Ptu(x))pPt(up)(z)exp(pKd2(x,z)2(p1)(e2Kt1)),for any zX.\left(P_{t}u(x)\right)^{p}\leq P_{t}(u^{p})(z)\exp\left(\frac{pKd^{2}(x,z)}{2(p-1)(e^{2Kt}-1)}\right),\quad\text{for any }z\in X. (5.12)

Since Ptu21=u21=1\|P_{t}u^{2}\|_{1}=\|u^{2}\|_{1}=1 by the mass-preserving property of the heat semigroup, we deduce from (5.12) by taking p=2p=2 that

(Ptu(x))2exp(KD2e2Kt1).\left(P_{t}u(x)\right)^{2}\leq\exp\left(\frac{KD^{2}}{e^{2Kt}-1}\right). (5.13)

Secondly, by the equation (3.3) and the commutation between PtP_{t} and Δ\Delta, it follows that

Ptu(z)Psu(z)=stPτ(Δu)(z)𝑑τ=stPτ(α2ulogu+(λα1)u)(z)𝑑τ,for any 0<s<t,P_{t}u(z)-P_{s}u(z)=\int_{s}^{t}P_{\tau}\left(\Delta u\right)(z)d\tau\\ =-\int_{s}^{t}P_{\tau}\left(\alpha_{2}u\log u+(\lambda-\alpha_{1})u\right)(z)d\tau,\quad\text{for any $0<s<t$}, (5.14)

for mm-a.e zXz\in X. By the regularization of the heat semigroup, the equality (5.14) holds for any zXz\in X. Since uu is positive and xx is the maximum point, it follows that (ulogu)(z)u(z)logu(x)(u\log u)(z)\leq u(z)\log u(x) for any zz in XX which by the comparison principle of the heat semigroup, yields

Ptu(x)Psu(x)α2logu(x)stPτu(x)𝑑τ(λα1)stPτu(x)𝑑τ.P_{t}u(x)-P_{s}u(x)\geq-\alpha_{2}\log u(x)\int_{s}^{t}P_{\tau}u(x)d\tau-(\lambda-\alpha_{1})\int_{s}^{t}P_{\tau}u(x)d\tau. (5.15)

Grönwall’s inequality further implies that

Ptu(x)u(x)exp(α2tlogu(x)(λα1)t)=e(λα1)tu(x)1α2t.P_{t}u(x)\geq u(x)\exp\left(-\alpha_{2}t\log u(x)-(\lambda-\alpha_{1})t\right)=e^{-(\lambda-\alpha_{1})t}u(x)^{1-\alpha_{2}t}. (5.16)

So, together with (5.13), it follows that for any 0<t<1/α20<t<1/\alpha_{2}:

logu(x)(λα1)t1α2t+KD22(1α2t)(e2Kt1).\log u(x)\leq\frac{(\lambda-\alpha_{1})t}{1-\alpha_{2}t}+\frac{KD^{2}}{2(1-\alpha_{2}t)(e^{2Kt}-1)}. (5.17)

By using K/(e2Kt1)1/(2t)K/(e^{2Kt}-1)\leq 1/(2t) when K>0K>0 and letting t=1/(2α2)t=1/(2\alpha_{2}), it follows that

logu(x)λα1α2+C3(α2,K)D2,\log u(x)\leq\frac{\lambda-\alpha_{1}}{\alpha_{2}}+C_{3}(\alpha_{2},K)D^{2}, (5.18)

where C3(α2,K)=α2C_{3}(\alpha_{2},K)=\alpha_{2} when K0K\geq 0 and C3(α2,K)=K/(exp(K/α2)1)C_{3}(\alpha_{2},K)=K/(\exp(K/\alpha_{2})-1) when K<0K<0.

As for the lower bound, from (ulogu)(z)u(z)logu(y)(u\log u)(z)\geq u(z)\log u(y) and similar arguments as above, it follows that

Ptu(y)e(λα1)tu(y)1α2t.P_{t}u(y)\leq e^{-(\lambda-\alpha_{1})t}u(y)^{1-\alpha_{2}t}. (5.19)

Taking z=yz=y in the dimension-free inequality (5.12) and plugging (5.16) and (5.19), it follows that

ep(λα1)tu(x)ppα2t(Ptu(x))pPt(up)(y)exp(pKD22(p1)(e2Kt1))u(x)p1(Ptu)(y)exp(pKD22(p1)(e2Kt1))u(x)p1e(λα1)tu(y)1α2texp(pKD22(p1)(e2Kt1)),e^{-p(\lambda-\alpha_{1})t}u(x)^{p-p\alpha_{2}t}\leq\left(P_{t}u(x)\right)^{p}\leq P_{t}(u^{p})(y)\exp\left(\frac{pKD^{2}}{2(p-1)(e^{2Kt}-1)}\right)\\ \leq u(x)^{p-1}(P_{t}u)(y)\exp\left(\frac{pKD^{2}}{2(p-1)(e^{2Kt}-1)}\right)\\ \leq u(x)^{p-1}e^{-(\lambda-\alpha_{1})t}u(y)^{1-\alpha_{2}t}\exp\left(\frac{pKD^{2}}{2(p-1)(e^{2Kt}-1)}\right), (5.20)

where in the third inequality we use the comparison principle of the heat semigroup that Ptusupu=u(x)P_{t}u\leq\sup u=u(x), and in the last inequality we use the inequality (5.19). After reorganizing the inequality, it follows that

u(x)1pα2te(p1)(λα1)tu(y)1α2texp(pKD22(p1)(e2Kt1)).u(x)^{1-p\alpha_{2}t}\leq e^{(p-1)(\lambda-\alpha_{1})t}u(y)^{1-\alpha_{2}t}\exp\left(\frac{pKD^{2}}{2(p-1)(e^{2Kt}-1)}\right). (5.21)

Taking p=1/(α2t)p=1/(\alpha_{2}t) and t=1/(2α2)t=1/(2\alpha_{2}), we obtain that

logu(y)λα1α22C3(α2,K)D2,\log u(y)\geq-\frac{\lambda-\alpha_{1}}{\alpha_{2}}-2C_{3}(\alpha_{2},K)D^{2}, (5.22)

where we use the inequality K/(e2Kt1)1/(2t)K/(e^{2Kt}-1)\leq 1/(2t) when K>0K>0 again. ∎

As a final application, following the similar methods as in [42, Lemma 5.3] together with Theorem 3.2, we recover the classical result from [10, Theorem 5.7.4] that any non-negative extremal function with the log-Sobolev constant λ(α1,α2)\lambda(\alpha_{1},\alpha_{2}) is constant when K>0K>0 and 0<α2KN/(N1)0<\alpha_{2}\leq KN/(N-1).

Corollary 5.4.

Let (X,d,m)(X,d,m) be a RCD(K,N)\mathrm{RCD}^{*}(K,N) space with K>0K>0 and N(2,)N\in(2,\infty). Then any non-negative extremal function of (3.1) with the log-Sobolev constant λ(α1,α2)\lambda(\alpha_{1},\alpha_{2}) is constant whenever 0<α2KN/(N1)0<\alpha_{2}\leq KN/(N-1).

Proof.

Let uu be an arbitrary non-negative extremal function of (3.1) with a log-Sobolev constant λ(α1,α2)\lambda(\alpha_{1},\alpha_{2}) with 0<α2KN/(N1)0<\alpha_{2}\leq KN/(N-1), and define v=loguv=\log u. Let further aa, bb, and dd be real numbers to be determined later. We first estimate ebv(Δv)2𝑑m\int e^{bv}(\Delta v)^{2}dm from both the PDE equation (3.3) and the weak Bochner inequality (2.7), and then derive the desired result.

  1. Step 1:

    We first derive the estimate from the equation (3.3). Let α3:=λα1\alpha_{3}:=\lambda-\alpha_{1} in (3.3). By the Lipschitz regularity of vv and the regularity of |v|2|\nabla v|^{2} proved in Step 1 in Theorem 4.1, it follows that ϕ:=e(b1)vΔvW1,2\phi:=e^{(b-1)v}\Delta v\in W^{1,2} and ψ:=e(b1)v|v|2W1,2\psi:=e^{(b-1)v}|\nabla v|^{2}\in W^{1,2}. On the one hand, applying ϕ\phi and then ψ\psi to the right-hand side of (3.3) it follows that

    I:=(α2v+α3)evϕ𝑑m=(α2v+α3)eve(b1)vΔv𝑑m=α2|v|2ebv𝑑mb(α2v+α3)|v|2ebv𝑑m=α2|v|2ebv𝑑mb(α2v+α3)evψ𝑑m=α2|v|2ebv𝑑m+bΔ(ev)e(b1)v|v|2𝑑m=α2|v|2ebv𝑑m+b|v|4ebv𝑑m+b(Δv)|v|2ebv𝑑m.I:=\int(\alpha_{2}v+\alpha_{3})e^{v}\phi dm=\int(\alpha_{2}v+\alpha_{3})e^{v}e^{(b-1)v}\Delta vdm\\ =-\alpha_{2}\int|\nabla v|^{2}e^{bv}dm-b\int(\alpha_{2}v+\alpha_{3})|\nabla v|^{2}e^{bv}dm\\ =-\alpha_{2}\int|\nabla v|^{2}e^{bv}dm-b\int(\alpha_{2}v+\alpha_{3})e^{v}\psi dm\\ =-\alpha_{2}\int|\nabla v|^{2}e^{bv}dm+b\int\Delta(e^{v})e^{(b-1)v}|\nabla v|^{2}dm\\ =-\alpha_{2}\int|\nabla v|^{2}e^{bv}dm+b\int|\nabla v|^{4}e^{bv}dm+b\int(\Delta v)|\nabla v|^{2}e^{bv}dm. (5.23)

    On the other hand, applying ϕ\phi to the left-hand side of (3.3), we get

    I=Δ(ev)e(b1)vΔvdm=(Δv)|v|2ebv𝑑m(Δv)2ebv𝑑m.I=\int-\Delta(e^{v})e^{(b-1)v}\Delta vdm=-\int(\Delta v)|\nabla v|^{2}e^{bv}dm-\int(\Delta v)^{2}e^{bv}dm. (5.24)

    Showing that

    (Δv)2ebv𝑑m=α2|v|2ebv𝑑mb|v|4ebv𝑑m(b+1)(Δv)|v|2ebv𝑑m.\int(\Delta v)^{2}e^{bv}dm=\alpha_{2}\int|\nabla v|^{2}e^{bv}dm-b\int|\nabla v|^{4}e^{bv}dm-(b+1)\int(\Delta v)|\nabla v|^{2}e^{bv}dm. (5.25)
  2. Step 2:

    We derive the estimate from the weak Bochner inequality (2.7). Note that f:=eavf:=e^{av} and g:=edvg:=e^{dv} satisfies the regularity requirement in (2.7). So plugging ff and gg into (2.7), it follows that the left-hand side of (2.7) can be expressed as

    12Δ(g)|f|2𝑑m=a2d2e(2a+d)v(Δv)|v|2𝑑m+a2d22e(2a+d)v|v|4𝑑m,\frac{1}{2}\int\Delta(g)|\nabla f|^{2}dm=\frac{a^{2}d}{2}\int e^{(2a+d)v}(\Delta v)|\nabla v|^{2}dm+\frac{a^{2}d^{2}}{2}\int e^{(2a+d)v}|\nabla v|^{4}dm, (5.26)

    and the right-hand side of (2.7) can be expressed as

    N1Ng(Δf)2𝑑mΔfg,f𝑑m+Kg|f|2𝑑m=a2N1Ne(2a+d)(Δv)2𝑑ma2(2aN1N+d)e(2a+d)v(Δv)|v|2𝑑ma2(a2N1N+ad)e(2a+d)v|v|4+a2Ke(2a+d)v|v|2𝑑m.-\frac{N-1}{N}\int g(\Delta f)^{2}dm-\int\Delta f\langle\nabla g,\nabla f\rangle dm+K\int g|\nabla f|^{2}dm\\ =-a^{2}\frac{N-1}{N}\int e^{(2a+d)}(\Delta v)^{2}dm-a^{2}\left(2a\frac{N-1}{N}+d\right)\int e^{(2a+d)v}(\Delta v)|\nabla v|^{2}dm\\ -a^{2}\left(a^{2}\frac{N-1}{N}+ad\right)\int e^{(2a+d)v}|\nabla v|^{4}+a^{2}K\int e^{(2a+d)v}|\nabla v|^{2}dm. (5.27)

    From (5.26) and (5.27), it follows that (2.7) reads as follows:

    e(2a+d)v(Δv)2𝑑mKNN1e(2a+d)v|v|2(a2+NN1ad+N2(N1)d2)e(2a+d)v|v|4𝑑m(2a+3N2(N1)d)e(2a+d)v(Δv)|v|2𝑑m.\int e^{(2a+d)v}(\Delta v)^{2}dm\geq\frac{KN}{N-1}\int e^{(2a+d)v}|\nabla v|^{2}\\ -\left(a^{2}+\frac{N}{N-1}ad+\frac{N}{2(N-1)}d^{2}\right)\int e^{(2a+d)v}|\nabla v|^{4}dm\\ -\left(2a+\frac{3N}{2(N-1)}d\right)\int e^{(2a+d)v}(\Delta v)|\nabla v|^{2}dm. (5.28)
  3. Step 3:

    We conclude by comparing the coefficients of (5.25) and (5.28) and choosing particular values for b,a,db,a,d. Let ε>0\varepsilon>0 be determined later and let b,a,db,a,d\in\mathbb{R} satisfy the following system of equations:

    {b=2a+d,bε=a2+NN1ad+N2(N1)d2,b+1=2a+3N2(N1)d.\left\{\begin{aligned} &b=2a+d,\\ &b-\varepsilon=a^{2}+\frac{N}{N-1}ad+\frac{N}{2(N-1)}d^{2},\\ &b+1=2a+\frac{3N}{2(N-1)}d.\end{aligned}\right. (5.29)

    This system (5.29) admits real-valued solutions if and only if 0<ε4N/(N+2)20<\varepsilon\leq 4N/(N+2)^{2}. Hence, choosing arbitrary ε(0,4N/(N+2)2]\varepsilon\in(0,4N/(N+2)^{2}] and

    d\displaystyle d =2(N1)N+2,\displaystyle=\frac{2(N-1)}{N+2},
    a\displaystyle a =2N+2+4N(N+2)2ε,\displaystyle=\frac{2}{N+2}+\sqrt{\frac{4N}{(N+2)^{2}}-\varepsilon},
    b\displaystyle b =2a+d.\displaystyle=2a+d.

    By comparing (5.25) and (5.28), it follows that

    (α2KNN1)|v|2ebv𝑑mε|v|4ebv𝑑m.\left(\alpha_{2}-\frac{KN}{N-1}\right)\int|\nabla v|^{2}e^{bv}dm\geq\varepsilon\int|\nabla v|^{4}e^{bv}dm. (5.30)

    Since 0<α2KN/(N1)0<\alpha_{2}\leq KN/(N-1), then |v||\nabla v| has to be 0, implying that uu is constant.

Appendix A Omori-Yau Maximum Principle

In the appendix, we provide a slightly generalized version of the Omori-Yau type maximum principle for the whole metric space of proper RCD(K,)\mathrm{RCD}(K,\infty) spaces with KK in \mathbb{R} which may not support the doubling property. To show it, we first show the Kato’s inequality in the proper RCD(K,)\mathrm{RCD}(K,\infty) setting whose proof follows a similar argument as in [51]. For the sake of completeness, we provide the complete argumentation.

Beforehand, recall the definition of the weak Laplacian. We say that an operator LL on Wloc1,2W^{1,2}_{loc} is the weak Laplacian provided that for each fWloc1,2f\in W^{1,2}_{loc}, LfLf is a linear functional acting on W1,2LW^{1,2}\cap L^{\infty} with bounded support defined as:

Lf(g):=f,g𝑑m,for all gW1,2L with bounded support.Lf(g):=-\int\langle\nabla f,\nabla g\rangle dm,\quad\text{for all }g\in W^{1,2}\cap L^{\infty}\text{ with bounded support}. (A.1)

For each hh in Wloc1,2LW^{1,2}_{loc}\cap L^{\infty}, hLfh\cdot Lf is the linear functional given by hLf(g):=Lf(hg)h\cdot Lf(g):=Lf(hg) for each gg in W1,2LW^{1,2}\cap L^{\infty} with bounded support. We say that LfLf is a signed Radon measure provided that there exists a signed Radon measure μ\mu such that Lf(g)=g𝑑μLf(g)=\int gd\mu for all gW1,2Lg\in W^{1,2}\cap L^{\infty} with bounded support. It is clear that in this case, we have fD(𝚫)f\in D(\bm{\Delta}) and Lf=𝚫fLf=\bm{\Delta}f. For wW1,2Lw\in W^{1,2}\cap L^{\infty} and mw:=ewmm_{w}:=e^{w}\cdot m, LwL_{w} denotes the weighted weak Laplacian on Wloc1,2W^{1,2}_{loc} defined as

Lwf(g):=f,g𝑑mw,for all gW1,2L with bounded support.L_{w}f(g):=-\int\langle\nabla f,\nabla g\rangle dm_{w},\quad\text{for all }g\in W^{1,2}\cap L^{\infty}\text{ with bounded support}. (A.2)

It is also easy to check that Lwf=ew(Lf+w,fm)L_{w}f=e^{w}\cdot(Lf+\langle\nabla w,\nabla f\rangle m). When LwfL_{w}f is a signed Radon measure, we denote by Lwf=(Lwacf)mw+LwsfL_{w}f=(L^{ac}_{w}f)\cdot m_{w}+L_{w}^{s}f its Lebesgue decomposition with respect to mwm_{w}. Finally, we remark that using the similar arguments as in [51, Lemma 3.2], we have the following chain rule: for fWloc1,2Lf\in W^{1,2}_{loc}\cap L^{\infty} and ϕC2()\phi\in C^{2}(\mathbb{R}), it holds that ϕ(f)Wloc1,2L\phi(f)\in W^{1,2}_{loc}\cap L^{\infty} and that

L(ϕ(f))=ϕ(f)Lf+ϕ′′(f)|f|2m.L\left(\phi(f)\right)=\phi^{\prime}(f)\cdot Lf+\phi^{\prime\prime}(f)|\nabla f|^{2}\cdot m. (A.3)
Lemma A.1.

(Kato’s inequality) Let (X,d,m)(X,d,m) be a proper RCD(K,)\mathrm{RCD}(K,\infty) space and ww be in W1,2LW^{1,2}\cap L^{\infty}. Suppose ff in Wloc1,2LW^{1,2}_{loc}\cap L^{\infty} is such that LwfL_{w}f is a signed Radon measure and that Lwsf0L^{s}_{w}f\geq 0. Then Lw(f+)L_{w}(f_{+}) is a signed Radon measure such that

Lw(f+)χ{f>0}Lwacfmw,L_{w}(f_{+})\geq\chi_{\{f>0\}}L_{w}^{ac}f\cdot m_{w}, (A.4)

where f+:=max{f,0}f_{+}:=\max\{f,0\} and f:=max{f,0}f_{-}:=\max\{-f,0\}.

Proof.

It suffices to prove that Lw(|f|)L_{w}(|f|) is a signed Radon measure and that

Lw(|f|)sgn(f)Lwf,L_{w}(|f|)\geq\mathrm{sgn}(f)\cdot L_{w}f, (A.5)

where sgn(t)=1\mathrm{sgn}(t)=1 if t>0t>0 and sgn(t)=1\mathrm{sgn}(t)=-1 if t<0t<0 and sgn(t)=0\mathrm{sgn}(t)=0 if t=0t=0. Indeed, if the inequality (A.5) holds, then it follows that both Lw(f+)L_{w}(f_{+}) and Lw(f)L_{w}(f_{-}) are signed Radon measures and (A.5) implies that

Lw(f+)+Lw(f)χ{f>0}Lwfχ{f<0}Lwf.L_{w}(f_{+})+L_{w}(f_{-})\geq\chi_{\{f>0\}}\cdot L_{w}f-\chi_{\{f<0\}}\cdot L_{w}f. (A.6)

By locality of the minimal weak upper gradient and the inner regularity of Radon measures, it is immediate to check that Lw(f+)L_{w}(f_{+}) is concentrated on the set {f0}\{f\geq 0\}. Then the inequality (A.6) with the assumption that Lwsf0L^{s}_{w}f\geq 0 implies that

Lw(f+)χ{f>0}Lwf=χ{f>0}(Lwacfmw+Lwsf)χ{f>0}Lwacfmw.L_{w}(f_{+})\geq\chi_{\{f>0\}}L_{w}f=\chi_{\{f>0\}}\left(L^{ac}_{w}f\cdot m_{w}+L^{s}_{w}f\right)\geq\chi_{\{f>0\}}L^{ac}_{w}f\cdot m_{w}. (A.7)

We are left to show (A.5). Let ε>0\varepsilon>0 and ϕε(t):=t2+ε2ε\phi_{\varepsilon}(t):=\sqrt{t^{2}+\varepsilon^{2}}-\varepsilon and fε:=ϕε(f)f_{\varepsilon}:=\phi_{\varepsilon}(f). Since ϕε\phi_{\varepsilon} is in C2()C^{2}(\mathbb{R}), it follows that fε|f|f_{\varepsilon}\leq|f| and that

|fε|=|ϕε(f)||f|=|f|f2+ε2|f||f|.|\nabla f_{\varepsilon}|=|\phi^{\prime}_{\varepsilon}(f)||\nabla f|=\frac{|f|}{\sqrt{f^{2}+\varepsilon^{2}}}|\nabla f|\leq|\nabla f|. (A.8)

Further, the chain rule of LwL_{w} for fWloc1,2Lf\in W^{1,2}_{loc}\cap L^{\infty} and ψ(t)=t2\psi(t)=t^{2} yields

2fLwf+2|f|2mw=Lwf2=Lw((fε+ε)2ε2)=2(fε+ε)Lwfε+2|fε|2mw.2f\cdot L_{w}f+2|\nabla f|^{2}\cdot m_{w}=L_{w}f^{2}=L_{w}\left((f_{\varepsilon}+\varepsilon)^{2}-\varepsilon^{2}\right)\\ =2(f_{\varepsilon}+\varepsilon)L_{w}f_{\varepsilon}+2|\nabla f_{\varepsilon}|^{2}\cdot m_{w}. (A.9)

So by (A.8), it follows that

Lwfεffε+εLwf.L_{w}f_{\varepsilon}\geq\frac{f}{f_{\varepsilon}+\varepsilon}\cdot L_{w}f. (A.10)

Now, let ΩX\Omega\subseteq X be any arbitrary bounded open subset. Let ηLipc\eta\in\mathrm{Lip}_{c} be a positive cut-off function such that η1\eta\equiv 1 on Ω\Omega. Since |fε||f||\nabla f_{\varepsilon}|\leq|\nabla f| and 0fε|f|0\leq f_{\varepsilon}\leq|f|, it follows that ηfεη|f|\eta f_{\varepsilon}\leq\eta|f| and that

|(ηfε)|fε|η|+η|fε||f||η|+η|f|.|\nabla(\eta f_{\varepsilon})|\leq f_{\varepsilon}|\nabla\eta|+\eta|\nabla f_{\varepsilon}|\leq|f||\nabla\eta|+\eta|\nabla f|.

Since ff is in Wloc1,2LW^{1,2}_{loc}\cap L^{\infty}, it follows that (ηfε)ε(\eta f_{\varepsilon})_{\varepsilon} is uniformly bounded in W1,2W^{1,2}. Note that by the assumption on ww, both the Sobolev spaces W1,2W^{1,2} and W1,2(X,mw)W^{1,2}(X,m_{w}) and the minimal weak upper gradients induced by mm and mwm_{w} coincide (see [3, Lemma 4.11]) and hence (ηfε)ε(\eta f_{\varepsilon})_{\varepsilon} is also uniformly bounded in W1,2(X,mw)W^{1,2}(X,m_{w}). Since W1,2W^{1,2} is a Hilbert space, W1,2(X,mw)W^{1,2}(X,m_{w}) is also Hilbert and hence reflexive. Therefore, there exists a subsequence (ηfεj)(\eta f_{\varepsilon_{j}}) with εj0\varepsilon_{j}\searrow 0 converging weakly in W1,2(X,mw)W^{1,2}(X,m_{w}) to some gW1,2(X,mw)g\in W^{1,2}(X,m_{w}). As fεj|f|f_{\varepsilon_{j}}\rightarrow|f| pointwise, we obtain that g=η|f|g=\eta|f| mwm_{w}-a.e. and in particular g=|f|g=|f| mwm_{w}-a.e. on Ω\Omega since η1\eta\equiv 1 on Ω\Omega. Since fεj(x)+εj|f|(x)f_{\varepsilon_{j}}(x)+\varepsilon_{j}\rightarrow|f|(x) for all xx in XX and |f|/(fεj+εj)1|f|/(f_{\varepsilon_{j}}+\varepsilon_{j})\leq 1, for any non-negative Lipschitz ϕLipc(Ω)\phi\in\mathrm{Lip}_{c}(\Omega), we obtain that

LΩ,f(ϕ):=Ωϕ,|f|𝑑mw=Ωϕ,(η|f|)𝑑mw=limjΩϕ,(ηfεj)dmw=limjΩϕ,fεjdmw=limjLwfεj(ϕ)limjΩϕffεj+εj𝑑Lwf=Ωϕsgn(f)𝑑Lwf.L_{\Omega,f}(\phi):=-\int_{\Omega}\langle\nabla\phi,\nabla|f|\rangle dm_{w}=-\int_{\Omega}\langle\nabla\phi,\nabla(\eta|f|)\rangle dm_{w}\\ =-\lim_{j\rightarrow\infty}\int_{\Omega}\langle\nabla\phi,\nabla(\eta f_{\varepsilon_{j})}\rangle dm_{w}=-\lim_{j\rightarrow\infty}\int_{\Omega}\langle\nabla\phi,\nabla f_{\varepsilon_{j}}\rangle dm_{w}\\ =\lim_{j\rightarrow\infty}L_{w}f_{\varepsilon_{j}}(\phi)\geq\lim_{j\rightarrow\infty}\int_{\Omega}\frac{\phi f}{f_{\varepsilon_{j}}+\varepsilon_{j}}dL_{w}f=\int_{\Omega}\phi\mathrm{sgn}(f)dL_{w}f. (A.11)

Following a similar argument as in [16, (4-28),(4-29) in Theorem 4.14], there exists a constant CΩ,f>0C_{\Omega,f}>0 such that |LΩ,f(ϕ)|CΩ,fmax|ϕ||L_{\Omega,f}(\phi)|\leq C_{\Omega,f}\max|\phi| for any Lipschitz function ϕ\phi with supp(ϕ)Ω\mathrm{supp}(\phi)\subset\Omega. Since XX is proper, by the Riesz representation theorem, there exists a signed Radon measure μΩ,f\mu_{\Omega,f} on Ω\Omega such that LΩ,f(ϕ)=Ωϕ𝑑μΩ,fL_{\Omega,f}(\phi)=\int_{\Omega}\phi d\mu_{\Omega,f} for all ϕ\phi in Lipc(Ω)\mathrm{Lip}_{c}(\Omega) and that μΩ,fsgn(f)Lwf\mu_{\Omega,f}\geq\mathrm{sgn}(f)L_{w}f on Ω\Omega (see the remark before Theorem 1.2 in [51] or [25, Proposition 6.2.16] for finite signed Radon measures). Clearly μΩ1,f\mu_{\Omega_{1},f} and μΩ2,f\mu_{\Omega_{2},f} coincide on Ω1Ω2\Omega_{1}\cap\Omega_{2} for any bounded open subsets Ω1\Omega_{1} and Ω2\Omega_{2}. Henceforth, there exists a unique signed Radon measure ν\nu on XX such that ν|Ω=μΩ,f\nu|_{\Omega}=\mu_{\Omega,f} for all bounded open domain Ω\Omega, and thus, we obtain that Lw|f|L_{w}|f| is a signed Radon measure with Lw|f|sgn(f)LwfL_{w}|f|\geq\mathrm{sgn}(f)\cdot L_{w}f. ∎

We now show the Omori-Yau type maximum principle in proper RCD(K,)\mathrm{RCD}(K,\infty) spaces based on the previous Kato’s inequality. While the proof in [51] relies on the doubling property and the weak Poincaré inequality of the underlying metric measure space as well as the weak maximum principle, our proof is based on the "Sobolev-to-Lip" property of RCD(K,)\mathrm{RCD}(K,\infty) spaces which in general does not support the doubling property.

Lemma A.2.

(Omori-Yau type maximum principle) Let (X,d,m)(X,d,m) be a proper RCD(K,)\mathrm{RCD}(K,\infty) space with KK in \mathbb{R}. Let further ff be in W1,2LD(𝚫)W^{1,2}\cap L^{\infty}\cap D(\bm{\Delta}) such that 𝚫sf0\bm{\Delta}^{s}f\geq 0. Suppose that ff achieves one of its strict maximum in XX in the sense that there exists a bounded and measurable subset UXU\subset X satisfying m(U)>0m(U)>0 and m(XU)>0m(X\setminus U)>0 with

esssupUf>esssupXUf.\operatorname*{ess\,sup}_{U}f>\operatorname*{ess\,sup}_{X\setminus U}f. (A.12)

Then, given any ww in W1,2LipbW^{1,2}\cap\mathrm{Lip}_{b}, for any ε>0\varepsilon>0, we have

m({xX:f(x)esssupXfε and (𝚫acf)(x)+f,w(x)ε})>0.m\left(\left\{x\in X\colon f(x)\geq\operatorname*{ess\,sup}_{X}f-\varepsilon\text{ and }(\bm{\Delta}^{ac}f)(x)+\langle\nabla f,\nabla w\rangle(x)\leq\varepsilon\right\}\right)>0. (A.13)

In particular, there exists a sequence (xj)(x_{j}) in XX such that f(xj)esssupf(xj)1/jf(x_{j})\geq\operatorname*{ess\,sup}f(x_{j})-1/j and (𝚫acf)(xj)+f,w(xj)1/j(\bm{\Delta}^{ac}f)(x_{j})+\langle\nabla f,\nabla w\rangle(x_{j})\leq 1/j.

Proof.

We adapt the proof in [51]. Let M:=esssupXfM:=\operatorname*{ess\,sup}_{X}f. Suppose by contradiction that there exists ε0>0\varepsilon_{0}>0 and ww in W1,2LipbW^{1,2}\cap\mathrm{Lip}_{b} such that (A.13) fails. Then for possibly smaller ε0>0\varepsilon_{0}>0 such that Mε0>esssupXUfM-\varepsilon_{0}>\operatorname*{ess\,sup}_{X\setminus U}f, it follows that g:=(f(Mε0))+g:=(f-(M-\varepsilon_{0}))_{+} is in W1,2W^{1,2} with g=0g=0 mm-a.e. in XUX\setminus U, and that

m({xX:f(x)>Mε0 and 𝚫acf(x)+f,w(x)ε0})=0.m\left(\left\{x\in X:f(x)>M-\varepsilon_{0}\text{ and }\bm{\Delta}^{ac}f(x)+\langle\nabla f,\nabla w\rangle(x)\leq\varepsilon_{0}\right\}\right)=0. (A.14)

It follows that for mm-a.e. xx in {yX:f(y)>Mε0}\{y\in X:f(y)>M-\varepsilon_{0}\}, we have

𝚫acf(x)+f,w(x)>ε0.\bm{\Delta}^{ac}f(x)+\langle\nabla f,\nabla w\rangle(x)>\varepsilon_{0}. (A.15)

Note that since wW1,2Lipbw\in W^{1,2}\cap\mathrm{Lip}_{b} and 𝚫f\bm{\Delta}f is a signed Radon measure, it follows that ew(𝚫f+w,fm)e^{w}(\bm{\Delta}f+\langle\nabla w,\nabla f\rangle m) is a well-defined signed Radon measure. By the identity of the weighted weak Laplacian Lwf=ew(Lf+w,fm)L_{w}f=e^{w}(Lf+\langle\nabla w,\nabla f\rangle m) and Lf=𝚫fLf=\bm{\Delta}f, it follows that

Lwacfmw=ew(Lacf+w,f)mewLε0m>0L^{ac}_{w}f\cdot m_{w}=e^{w}\left(L^{ac}f+\langle\nabla w,\nabla f\rangle\right)\cdot m\geq e^{-\|w\|_{L^{\infty}}}\varepsilon_{0}\cdot m>0 (A.16)

on {yX:f(y)>Mε0}\{y\in X:f(y)>M-\varepsilon_{0}\}. Moreover, the identity of the weighted weak Laplacian together with 𝚫sf0\bm{\Delta}^{s}f\geq 0 implies that Lwsf0L^{s}_{w}f\geq 0 and f(Mε0)f-(M-\varepsilon_{0}) is in Wloc1,2LW^{1,2}_{loc}\cap L^{\infty}, applying Lemma A.1 to f(Mε0)f-(M-\varepsilon_{0}), yields

Lwg=Lw(f(Mε0))+χ{f>Mε0}Lwacfmw0.L_{w}g=L_{w}\left(f-(M-\varepsilon_{0})\right)_{+}\geq\chi_{\{f>M-\varepsilon_{0}\}}L^{ac}_{w}f\cdot m_{w}\geq 0. (A.17)

By the definition of LwL_{w}, it follows from (A.17) that

Xg,g𝑑mw=Lwg(g)0,-\int_{X}\langle\nabla g,\nabla g\rangle dm_{w}=L_{w}g(g)\geq 0, (A.18)

which implies that |g|=0|\nabla g|=0 mwm_{w}-a.e. From mm being equivalent to mwm_{w} since wLw\in L^{\infty}, we get that |g|=0|\nabla g|=0 mm-a.e. Together with gg being in W1,2W^{1,2} and |g||\nabla g| in LL^{\infty}, by the Sobolev-to-Lip property of RCD(K,)\mathrm{RCD}(K,\infty) spaces, it follows that gg admits a Lipschitz representation g~\tilde{g} with Lip(g~)|g|\mathrm{Lip}(\tilde{g})\leq\||\nabla g|\|_{\infty}. Then, from the fact that |g|=0|\nabla g|=0 mm-a.e. as well as g=0g=0 mm-a.e on XUX\setminus U, it follows that g~\tilde{g} is constant and g~0\tilde{g}\equiv 0. Hence, fMε0f\leq M-\varepsilon_{0} mm-a.e., which is a contradiction. ∎

Appendix B Proof of Rellich-Kondrachov type theorem

In this subsection, we show the second claim of Lemma 2.8. The Rellich-Kondrachov type theorem has been previously proved for RCD(K,N)\mathrm{RCD}^{*}(K,N) spaces with K>0K>0 and N(2,)N\in(2,\infty) in [42] by following the argument in [28, Theorem 8.1]. A careful inspection of their proof shows that the result also holds for compact CD(K,N)\mathrm{CD}(K,N) spaces with KK\in\mathbb{R} and N(2,)N\in(2,\infty) without the infinitesimal Hilbertianty assumption. For the sake of completeness, we provide the proof.

Proof of (ii) of Lemma 2.8.

We follow the similar arguments from [28]. Let D:=diam(X)D:=\mathrm{diam}(X). By the assumptions and the generalized Bishop-Gromov inequality, without loss of generality, we can assume that m(X)=1m(X)=1. By the Sobolev inequality (2.12) and the assumption that supnfnW1,2<\sup_{n}\|f_{n}\|_{W^{1,2}}<\infty, it follows that (fn)n(f_{n})_{n} is bounded in L2L^{2^{*}}. Since L2L^{2^{*}} is reflexive, there exists fL2f\in L^{2^{*}} such that there exists a subsequence (fnk)k(f_{n_{k}})_{k} of (fn)n(f_{n})_{n} such that fnkf_{n_{k}} converges to ff weakly in L2L^{2^{*}}. It suffices to show that fnkf_{n_{k}} converges to ff strongly in LqL^{q} for all q[1,2)q\in[1,2^{*}). For the notation’s convenience, we relabel (fnk)k(f_{n_{k}})_{k} as (fk)k(f_{k})_{k}.

Let ε>0\varepsilon>0 and 0<rD/20<r\leq D/2 be arbitrary and q[1,2)q\in[1,2^{*}). Let f¯r(x):=B(x,r)f𝑑m\bar{f}_{r}(x):=\fint_{B(x,r)}fdm and (fk)¯r(x):=B(x,r)fk𝑑m\bar{(f_{k})}_{r}(x):=\fint_{B(x,r)}f_{k}dm for any xXx\in X. Then it follows that

m(|fk(x)f(x)|>ε)m(|fk(x)(fk)¯r(x)|>ε/3)+m(|(fk)¯r(x)f¯r(x)|>ε/3)+m(|f¯r(x)f(x)|>ε/3).m\left(\left|f_{k}(x)-f(x)\right|>\varepsilon\right)\leq m\left(\left|f_{k}(x)-\bar{(f_{k})}_{r}(x)\right|>\varepsilon/3\right)\\ +m\left(\left|\bar{(f_{k})}_{r}(x)-\bar{f}_{r}(x)\right|>\varepsilon/3\right)+m\left(\left|\bar{f}_{r}(x)-f(x)\right|>\varepsilon/3\right). (B.1)

Note that the compact CD(K,N)\mathrm{CD}(K,N) space with bounded diameter DD support the global doubling property and the weak (1,1)(1,1)-Poincaré inequality (2.10) with the global constant C(K,N,D)C(K,N,D). By the equivalent characterizations of Poincaré inequalities [33, Theorem 2] and [11, Proposition 4.13], it follows that the weak (1,2)(1,2)-Poincaré inequality (2.10) holds for all pairs (fk,|fk|)k(f_{k},|\nabla f_{k}|)_{k}. Moreover, by the doubling property of CD(K,N)\mathrm{CD}(K,N) spaces, compact CD(K,N)\mathrm{CD}(K,N) spaces support the Lebesgue differentiation theorem and the maximal theorem (see [29]). Let xXx\in X be any Lebesgue point of XX and ri:=r/2ir_{i}:=r/2^{i} for ii\in\mathbb{N}. Then for the first term of the right-hand side of (B.1), it follows that

|fk(x)(fk)¯r(x)|i=0|(fk)¯ri(x)(fk)¯ri+1(x)|i=0B(x,ri)|fk(y)(fk)¯ri(x)|𝑑m(y)C(D)i=0ri(B(x,2ri)|fk|2𝑑m)1/22C(D)rM(|fk|2)1/2(x),\left|f_{k}(x)-\bar{(f_{k})}_{r}(x)\right|\leq\sum_{i=0}^{\infty}\left|\bar{(f_{k})}_{r_{i}}(x)-\bar{(f_{k})}_{r_{i+1}}(x)\right|\\ \leq\sum_{i=0}^{\infty}\fint_{B(x,r_{i})}\left|f_{k}(y)-\bar{(f_{k})}_{r_{i}}(x)\right|dm(y)\leq C(D)\sum_{i=0}^{\infty}r_{i}\left(\fint_{B(x,2r_{i})}|\nabla f_{k}|^{2}dm\right)^{1/2}\\ \leq 2C(D)r\cdot M\left(|\nabla f_{k}|^{2}\right)^{1/2}(x), (B.2)

where Mf(x):=supr>0B(x,r)|f(y)|𝑑mMf(x):=\sup_{r>0}\fint_{B(x,r)}|f(y)|dm be the centered maximal operator. So by applying the maximal theorem [29, Theorem 3.5.6], it follows that

m(|fk(x)(fk)¯r(x)|>ε/3)m(M(|fk|2)1/2(x)>ε6C(D)r)36C1C(D)2r2ε2supk|fk|22.m\left(\left|f_{k}(x)-\bar{(f_{k})}_{r}(x)\right|>\varepsilon/3\right)\leq m\left(M\left(\left|\nabla f_{k}\right|^{2}\right)^{1/2}(x)>\frac{\varepsilon}{6C(D)r}\right)\\ \leq\frac{36C_{1}\cdot C(D)^{2}r^{2}}{\varepsilon^{2}}\sup_{k}\left\|\left|\nabla f_{k}\right|\right\|^{2}_{2}. (B.3)

For the second term of the right-hand side of (B.1), note that since fkf_{k} converges to ff weakly in L2(X)L^{2^{*}}(X), so taking the test function η:=χB(x,r)/m(B(x,r))\eta:=\chi_{B(x,r)}/m(B(x,r)), it follows that

1m(B(x,r))B(x,r)fk𝑑m1m(B(x,r))B(x,r)f𝑑m,as k,\frac{1}{m(B(x,r))}\int_{B(x,r)}f_{k}dm\rightarrow\frac{1}{m(B(x,r))}\int_{B(x,r)}fdm,\quad\text{as $k\rightarrow\infty$}, (B.4)

which implies that (fk)¯r\bar{(f_{k})}_{r} converges to f¯r\bar{f}_{r} pointwise and thus (fk)¯r\bar{(f_{k})}_{r} converges to f¯\bar{f} in probability. For the third term of right-hand side of (B.1), applying the Lebesgue differentiation theorem, it follows that f¯r\bar{f}_{r} converges to ff mm-a.e. on XX as r0r\searrow 0. Together with all arguments above, taking kk\rightarrow\infty first and then taking r0r\searrow 0 on both side of (B.1), it follows that

limkm(|fk(x)f(x)|>ε)=0.\lim_{k\rightarrow\infty}m\left(\left|f_{k}(x)-f(x)\right|>\varepsilon\right)=0. (B.5)

Then by [28, Lemma 8.2], we obtain that fkff_{k}\rightarrow f strongly in LqL^{q} for all q[1,2)q\in[1,2^{*}).

References

  • Ambrosio and Honda [2018] Luigi Ambrosio and Shouhei Honda. Local spectral convergence in RCD(K,N)\mathrm{RCD}^{*}({K},{N}) spaces. Nonlinear Analysis, 177:1–23, 2018.
  • Ambrosio et al. [2013] Luigi Ambrosio, Nicola Gigli, and Giuseppe Savaré. Density of Lipschitz functions and equivalence of weak gradients in metric measure spaces. Revista Matemática Iberoamericana, 29(3):969–996, 2013.
  • Ambrosio et al. [2014a] Luigi Ambrosio, Nicola Gigli, and Giuseppe Savaré. Calculus and heat flow in metric measure spaces and applications to spaces with Ricci bounds from below. Inventiones Mathematicae, 195(2):289–391, 2014a.
  • Ambrosio et al. [2014b] Luigi Ambrosio, Nicola Gigli, and Giuseppe Savaré. Metric measure spaces with Riemannian Ricci curvature bounded from below. Duke Mathematical Journal, 163(7):1405–1490, 2014b.
  • Ambrosio et al. [2015a] Luigi Ambrosio, Nicola Gigli, Andrea Mondino, and Tapio Rajala. Riemannian Ricci curvature lower bounds in metric measure spaces with σ\sigma-finite measure. Transactions of the American Mathematical Society, 367(7):4661–4701, 2015a.
  • Ambrosio et al. [2015b] Luigi Ambrosio, Nicola Gigli, and Giuseppe Savaré. Bakry–Émery curvature-dimension condition and Riemannian Ricci curvature bounds. The Annals of Probability, 43(1):339–404, 2015b.
  • Ambrosio et al. [2016] Luigi Ambrosio, Andrea Mondino, and Giuseppe Savaré. On the Bakry–Émery condition, the gradient estimates and the local-to-global property of RCD(K,N)\mathrm{RCD}^{*}({K},{N}) metric measure spaces. The Journal of Geometric Analysis, 26:24–56, 2016.
  • Ambrosio et al. [2019] Luigi Ambrosio, Andrea Mondino, and Giuseppe Savaré. Nonlinear Diffusion Equations and Curvature Conditions in Metric Measure Spaces, volume 262. American mathematical society, 2019.
  • Bacher and Sturm [2010] Kathrin Bacher and Karl-Theodor Sturm. Localization and tensorization properties of the curvature-dimension condition for metric measure spaces. Journal of Functional Analysis, 259(1):28–56, 2010.
  • Bakry et al. [2014] Dominique Bakry, Ivan Gentil, and Michel Ledoux. Analysis and Geometry of Markov Diffusion Operators, volume 103. Springer, 2014.
  • Björn and Björn [2011] Anders Björn and Jana Björn. Nonlinear Potential Theory on Metric Spaces, volume 17. European Mathematical Society, 2011.
  • Cavalletti and Milman [2021] Fabio Cavalletti and Emanuel Milman. The globalization theorem for the curvature-dimension condition. Inventiones Mathematicae, 226:1–137, 2021.
  • Cavalletti and Mondino [2017a] Fabio Cavalletti and Andrea Mondino. Optimal maps in essentially non-branching spaces. Communications in Contemporary Mathematics, 19(06):1750007, 2017a.
  • Cavalletti and Mondino [2017b] Fabio Cavalletti and Andrea Mondino. Sharp geometric and functional inequalities in metric measure spaces with lower Ricci curvature bounds. Geometry & Topology, 21(1):603–645, 2017b.
  • Cavalletti and Mondino [2017c] Fabio Cavalletti and Andrea Mondino. Sharp and rigid isoperimetric inequalities in metric-measure spaces with lower Ricci curvature bounds. Inventiones Mathematicae, 208:803–849, 2017c.
  • Cavalletti and Mondino [2020] Fabio Cavalletti and Andrea Mondino. New formulas for the Laplacian of distance functions and applications. Analysis & PDE, 13(7):2091–2147, 2020.
  • Cavalletti and Sturm [2012] Fabio Cavalletti and Karl-Theodor Sturm. Local curvature-dimension condition implies measure-contraction property. Journal of Functional Analysis, 262(12):5110–5127, 2012.
  • Cheeger [1999] Jeff Cheeger. Differentiability of Lipschitz functions on metric measure spaces. Geometric & Functional Analysis GAFA, 9:428–517, 1999.
  • Debin et al. [2021] Clément Debin, Nicola Gigli, and Enrico Pasqualetto. Quasi-continuous vector fields on RCD spaces. Potential Analysis, 54:183–211, 2021.
  • Erbar et al. [2015] Matthias Erbar, Kazumasa Kuwada, and Karl-Theodor Sturm. On the equivalence of the entropic curvature-dimension condition and Bochner’s inequality on metric measure spaces. Inventiones Mathematicae, 201(3):993–1071, 2015.
  • Garofalo and Mondino [2014] Nicola Garofalo and Andrea Mondino. Li–Yau and Harnack type inequalities in RCD(K,N)\mathrm{RCD}^{*}({K},{N}) metric measure spaces. Nonlinear Analysis: Theory, Methods & Applications, 95:721–734, 2014.
  • Giaquinta [1983] Mariano Giaquinta. Multiple Integrals in the Calculus of Variations and Nonlinear Elliptic Systems. Number 105 in Annals of Mathematics Studies. Princeton University Press, 1983.
  • Gigli [2018] Nicola Gigli. Nonsmooth Differential Geometry–An Approach Tailored for Spaces with Ricci Curvature Bounded from Below, volume 251. American Mathematical Society, 2018.
  • Gigli and Mondino [2013] Nicola Gigli and Andrea Mondino. A PDE approach to nonlinear potential theory in metric measure spaces. Journal de Mathématiques Pures et Appliquées, 100(4):505–534, 2013.
  • Gigli and Pasqualetto [2020] Nicola Gigli and Enrico Pasqualetto. Lectures on Nonsmooth Differential Geometry, volume 2. Springer, 2020.
  • Gigli et al. [2020] Nicola Gigli, Christian Ketterer, Kazumasa Kuwada, and Shin-ichi Ohta. Rigidity for the spectral gap on RCD(K,)\mathrm{RCD}({K},\infty)-spaces. American Journal of Mathematics, 142(5):1559–1594, 2020.
  • Gross [1975] Leonard Gross. Logarithmic Sobolev inequalities. American Journal of Mathematics, 97(4):1061–1083, 1975.
  • Hajłasz and Koskela [2000] Piotr Hajłasz and Pekka Koskela. Sobolev Met Poincaré, volume 688. American Mathematical Society, 2000.
  • Heinonen et al. [2015] Juha Heinonen, Pekka Koskela, Nageswari Shanmugalingam, and Jeremy T. Tyson. Sobolev Spaces on Metric Measure Spaces, volume 27. Cambridge University Press, 2015.
  • Huang and Zhang [2020] Jia-Cheng Huang and Hui-Chun Zhang. Localized elliptic gradient estimate for solutions of the heat equation on RCD(K,N)\mathrm{RCD}^{*}({K},{N}) metric measure spaces. Manuscripta Mathematica, 161:303–324, 2020.
  • Jiang [2015] Renjin Jiang. The Li–Yau inequality and heat kernels on metric measure spaces. Journal de Mathématiques Pures et Appliquées, 104(1):29–57, 2015.
  • Jiang and Zhang [2016] Renjin Jiang and Huichun Zhang. Hamilton’s gradient estimates and a monotonicity formula for heat flows on metric measure spaces. Nonlinear Analysis, 131:32–47, 2016.
  • Keith [2003] Stephen Keith. Modulus and the Poincaré inequality on metric measure spaces. Mathematische Zeitschrift, 245:255–292, 2003.
  • Kinnunen and Shanmugalingam [2001] Juha Kinnunen and Nageswari Shanmugalingam. Regularity of quasi-minimizers on metric spaces. Manuscripta Mathematica, 105(3):401–423, 2001.
  • Koskela et al. [2003] Pekka Koskela, Kai Rajala, and Nageswari Shanmugalingam. Lipschitz continuity of Cheeger-harmonic functions in metric measure spaces. Journal of Functional Analysis, 202(1):147–173, 2003.
  • Li [2016] Huaiqian Li. Dimension-free Harnack inequalities on RCD(K,)\mathrm{RCD}({K},\infty) spaces. Journal of Theoretical Probability, 29:1280–1297, 2016.
  • Lott and Villani [2009] John Lott and Cédric Villani. Ricci curvature for metric-measure spaces via optimal transport. Annals of Mathematics, 169:903–991, 2009.
  • Mondino and Naber [2019] Andrea Mondino and Aaron Naber. Structure theory of metric measure spaces with lower Ricci curvature bounds. Journal of the European Mathematical Society, 21(6):1809–1854, 2019.
  • Nobili and Violo [2022] Francesco Nobili and Ivan Yuri Violo. Rigidity and almost rigidity of Sobolev inequalities on compact spaces with lower Ricci curvature bounds. Calculus of Variations and Partial Differential Equations, 61(5):1–65, 2022.
  • Ohta and Takatsu [2020] Shin-ichi Ohta and Asuka Takatsu. Equality in the logarithmic Sobolev inequality. Manuscripta Mathematica, 162(1):271–282, 2020.
  • Otto and Villani [2000] Felix Otto and Cédric Villani. Generalization of an inequality by Talagrand and links with the logarithmic Sobolev inequality. Journal of Functional Analysis, 173(2):361–400, 2000.
  • Profeta [2015] Angelo Profeta. The sharp Sobolev inequality on metric measure spaces with lower Ricci curvature bounds. Potential Analysis, 43:513–529, 2015.
  • Rajala [2012] Tapio Rajala. Interpolated measures with bounded density in metric spaces satisfying the curvature-dimension conditions of Sturm. Journal of Functional Analysis, 263(4):896–924, 2012.
  • Rajala and Sturm [2014] Tapio Rajala and Karl-Theodor Sturm. Non-branching geodesics and optimal maps in strong CD(K,)\mathrm{CD}({K},\infty)-spaces. Calculus of Variations and Partial Differential Equations, 50(3):831–846, 2014.
  • Rothaus [1981] O. S. Rothaus. Logarithmic Sobolev inequalities and the spectrum of Schrödinger operators. Journal of functional analysis, 42(1):110–120, 1981.
  • Shanmugalingam [2000] Nageswari Shanmugalingam. Newtonian spaces: an extension of Sobolev spaces to metric measure spaces. Revista Matemática Iberoamericana, 16(2):243–279, 2000.
  • Sturm [2006a] Karl-Theodor Sturm. On the geometry of metric measure spaces. Acta Mathematica, 196(1):65–131, 2006a.
  • Sturm [2006b] Karl-Theodor Sturm. On the geometry of metric measure spaces. II. Acta Mathematica, 196(1):133–177, 2006b.
  • Villani [2008] Cédric Villani. Optimal Transport, Old and New. Grundlehren der mathematischen Wissenschaften. Springer, 1 edition, 2008.
  • Wang [1999] Feng-Yu Wang. Harnack inequalities for log-Sobolev functions and estimates of log-Sobolev constants. The Annals of Probability, 27(2):653–663, 1999.
  • Zhang and Zhu [2016] Hui-Chun Zhang and Xi-Ping Zhu. Local Li–Yau’s estimates on RCD(K,N)\mathrm{RCD}^{*}({K},{N}) metric measure spaces. Calculus of Variations and Partial Differential Equations, 55(4):1–30, 2016.
  • Zhang [2012] Qi S. Zhang. Extremal of Log Sobolev inequality and W entropy on noncompact manifolds. Journal of Functional Analysis, 263(7):2051–2101, 2012.