Extremal of Log-Sobolev Functionals and Li-Yau Estimate on Spaces
Abstract.
In this work, we study the extremal functions of the log-Sobolev functional on compact metric measure spaces satisfying the condition for in and in .
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.
1. Introduction
In this work, we study the extremal functions of the following variational problem of the log-Sobolev functional
(1.1) |
where and are in and . 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 -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 condition—bounded from below by in and the dimension bounded from above by 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 condition was introduced by Ambrosio, Gigli, and Savaré [4] and the finite dimensional counterpart condition was introduced and studied in [20, 8, 7]. Recently, building upon the abstract module theory, the first and second order differential structure on spaces was developed by Gigli [23], and finer geometric results such as the rectifiability of 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 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 on a compact metric measure space satisfying the condition with in and in 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 spaces, the polar decomposition still works by using the “needle decomposition” generalized to essentially non-branching 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 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 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 , it holds that
(1.2) |
for any in and . 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 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 , we use a different auxiliary function 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 -valued Laplacian. Furthermore, for our purpose, we slightly extend the Omori-Yau maximum principle in [51], which holds on spaces with and , to proper spaces. While most arguments are similar, our proof follows the so-called “Sobolev-to-Lip” property, a property shared by all 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 , 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 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 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 a metric measure space where is a complete, separable and proper metric space and is a non-negative Radon measure with full support which is finite on every bounded and measurable set. We denote by and the open and closed metric balls centered at in with radius , respectively. By for we denote the standard spaces with -norm . By we denote those measurable functions such that is in for any bounded and measurable subset of , where denote the indicator function of the set . Let further , , and be the spaces of real-valued functions which are Lipschitz, locally Lipschitz, and Lipschitz with bounded support, respectively. For in , we denote by the local Lipschitz constant, or slope, defined for any in as
if is not isolated and if is isolated.
2.1. Cheeger Energy, Laplacian and Calculus Tools
The Cheeger energy is a -lower-semicontinuous and convex functional defined as
The domain of is a linear space denoted by and is called the Sobolev space. For in , we identify the canonical element called the minimal relaxed gradient as the unique element with minimal -norm, also minimal in the -a.e. sense, in the set
which provides an integral representation . The Sobolev space equipped with the norm is a Banach space and is dense in , see [3, Proposition 4.1]. We further denote by the space of local Sobolev functions, and define the minimal relaxed gradient as -a.e. on for in where is in .
We say that is infinitesimally Hilbertian if the Cheeger energy is a quadratic form, or equivalently, is a Hilbert space. Under these assumptions, it can be proved that for any and in , the limit
exists in and it is a bilinear form from to , see [4].
Definition 2.1.
Let be infinitesimally Hilbertian.
-
•
Laplacian: We say that in , is in the domain of the Laplacian, denoted by , provided that there exists in such that
(2.1) In this case, we denote .
-
•
Measure-Valued Laplacian: We say that in is in the domain of the measure-valued Laplacian, denoted by , provided that there exists a signed Radon measure on such that,111Recall that is assumed to be proper, and therefore any bounded and closed set is compact on which Radon measures are finite.
(2.2) In this case, we denote .
By the separating property of for Radon measures and the infinitesimal Hilbertian property, it’s clear that both and are well-defined, unique and linear operators. Moreover, the two definitions are compatible in the following sense: on the one hand, if is in with for some in , then is in and . On the other hand, if is in such that , then is in and , see [25, Proposition 6.2.13]. For in the domain of , we denote the Lebesgue decomposition with respect to
(2.3) |
where is the Radon-Nikodym density and is the singular part of .
For in , we define the weighted Laplacian similarly but with respect to the reference measure and test functions in . For such , it can be shown that coincides with , and the minimal relaxed gradient induced by coincides with the one induced by , see [3, Lemma 4.11]. Moreover, it holds that , see [26, Lemma 3.4]
Lemma 2.2.
Let be infinitesimally Hilbertian. Then:
-
(i)
Locality: on for any , in and constant .
-
(ii)
Chain rule: for any in and Lipschitz function , it follows that
In particular, if is a contraction, then .
-
(iii)
Leibniz rule: for any , and in , it follows that is in and
-
(iv)
Chain rule: for any in and -function , it follows that is in and
(2.4) -
(v)
Leibniz rule: for any and in such that is continuous and is absolutely continuous with respect to , then is in and
(2.5)
By the -lower semicontinuity and convexity of the Cheeger energy, the heat semigroup is defined as the gradient flow in of the Cheeger energy starting from based on the classical Brezis-Komura theory, which provides the existence and uniqueness results. Moreover, for any , it holds that is locally absolutely continuous on and is in for all and
(2.6) |
Under the further assumption that is infinitesimally Hilbertian, the heat semigroup is linear, strongly continuous, contractive and order-preserving in . Moreover, can be extended into a linear, mass preserving and strongly continuous operator in for any , see [3] and further results therein.
Finally, we recall the definition (see for example [1, Definition 2.14]) of the local Sobolev space on an open subset . We denote by the space of Lipschitz functions on with compact support in and by the space of locally Lipschitz functions on .
Definition 2.3 (Local Sobolev space).
Let be an open subset. We say that is in if
-
(i)
is in for all ;
-
(ii)
where -a.e. on for .
We remark that by locality of the minimal relaxed gradient and Condition (i), Condition (ii) makes sense. As for the Laplacian on , it is modified as follows: A function in belongs to provided that there exists in such that
and we denote . Clearly, is linear by infinitesimal Hilbertianty and one can easily check that for in , its restriction to belongs to .
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 as the finite dimensional counterpart to , 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 spaces satisfy the so-called “local-to-global” property (see also [9]). Very recently, Cavalletti and Milman [12] show that and conditions are equivalent if the reference measure is finite.
The notion of the condition can be defined in several equivalent ways, see [20, 7, 8] for cases and [4, 5, 6] for cases. In this work we give a definition including the case where from an Eulerian point of view based on the abstract -calculus.
Definition 2.4.
We say that a metric measure space satisfies the condition for in and in provided that
-
(i)
for some in and ;
-
(ii)
Sobolev-to-Lip property: any in with in admits a Lipschitz representation in such that -a.e and .
-
(iii)
is infinitesimally Hilbertian.
-
(iv)
Weak Bochner inequality: for any and in with in , in and , it holds
(2.7)
In the following, we assume that is a compact space with in , and , which is the framework of the results in this work
First, satisfies the generalized Bishop-Gromov inequality, that is, for any and ,
(2.8) |
where is the volume of ball with radius in the model space (see [9, Theorem 6.2]). In particular, is globally doubling with the constant when , that is,
(2.9) |
and with the constant depending only on and when . It also holds that for any and in and that . Thus by [12], also satisfies the condition.
Second, supports the weak -Poincaré inequality, see [43, Theorem 1.1], that is, for any in , and any continuous function and any upper gradient of , we have
(2.10) |
where the constant only depends on , and , and and . Clearly, by Hölder inequality, also supports the weak -Poincaré inequality for any . This implies that is connected, see [11, Proposition 4.2]. In fact, one can show that is also a geodesic space.
Third, when , the Bonnet-Myers theorem implies that , see [48, Corollary 2.6]. Since the normalization of the reference measure does not affect conditions, it is not restrictive to assume that when is compact.
Fourth, in the setting, for Sobolev functions , it is possible to identify the gradient , rather than the modulus of the gradient , as the unique element in the tangent module , which is a -normed -module, see [25, 23]. Therein, a second-order calculus on spaces is also developed such that the notions of Hessian and its pointwise norm are well-defined. For the complete theory, we refer readers to [25, 23]. Here we only mention the inclusion . Consequently, for any in , we have in and
(2.11) |
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 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 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 be a space with in and in . Let further be in where is a geodesic ball centered at some in with radius . Assume that is in and on . Then it holds that is in and
Remark 2.7.
Note that our definition of the local Sobolev space on some open domain is a priori different from the one used in [35], which is the completion of with respect to the norm . However, on the one hand, as shown in [1, Remark 2.15], this space coincides with the Cheeger-Sobolev space defined by Cheeger [18] on spaces. On the other hand, since is dense in (see [11, Theorem 5.47]) and the norms of and are equal for on 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, and coincide with each other.
We also recall the following result of the Sobolev inequality for the compact space with and in , proved in [39]. As a result, we have a metric version of the Rellich-Kondrachov type theorem for compact spaces with and . Similar results have been previously proved by Profeta [42] for spaces with and .
Lemma 2.8.
Let be a compact space with and in and for some and .
-
(i)
Sobolev inequality ([39, Proposition 5.1]): there exist a constant depending only on , such that for every in , it holds
(2.12) where is the Sobolev conjugate of .
-
(ii)
Rellich-Kondrachov: let be a sequence in with . Then there exists in and a subsequence such that for every , it holds that
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 spaces with and . 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 , we say that is ultracontractive if for , there exists a constant such that for any in , it holds that
and we denote . 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 spaces with and in , since the Sobolev inequality holds, it follows that has the ultracontractive property. More precisely, for and , it holds that
where is a constant depending on and . As a consequence, the ultracontractive property of the heat semigroup provides the following boundedness result for the resolvent operator of the Laplacian for , the proof of which can be found in [42, Lemma 4.1] and [10, Corollary 6.3.3].
Lemma 2.9.
On a compact space with and in , let . If , the resolvent is bounded for each . If , the resolvent is bounded.
3. Existence of positive extremal functions
From now on, we assume that is a compact space with and in , and is a Borel probability measure with full support. We consider the following variational problem
(3.1) |
where and are in . The infimum quantity of variational problem (3.1) is called the log Sobolev constant on with parameters and , and the functional in (3.1) is called the log-Sobolev functional.
Definition 3.1.
Provided that is a finite number, we call a function in the extremal function of the variational problem (3.1) if and
(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 . 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 be a compact spaces with and , and let and be given constants with . Then the log Sobolev constant has finite value and the variational problem (3.1) admits non-negative extremal functions. Moreover, any non-negative extremal function satisfies the following properties:
-
(i)
is in and satisfies
(3.3) Furthermore, if , then is non-constant.
-
(ii)
is Lipschitz continuous;
-
(iii)
is positive.333Since is compact, for some .
Corollary 3.3.
Let and be given constants with and be an arbitrary non-negative extremal function of (3.1). Then is Lipschitz and in and satisfies
(3.4) |
In particular, the equation admits Lipschitz weak solutions which are non-constant whenever .
Remark 3.4.
A classical example of the above variational problem is the weak log-Sobolev inequalities, that is, the variational problems
(3.5) |
where and is defined as the supremum of those such that
for all in and .
The log-Sobolev inequality on the space with and , see [14], implies that for all , and when is small enough. A straightforward inspection shows that if has finite value, then . 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 . Furthermore, we derive that the equation admits non-constant, Lipschitz and positive weak solutions if .
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 a universal constant which may vary from one step to the next.
First part of the proof of Theorem 3.2.
-
Step 1:
We show the existence of non-negative extremal functions of variational problem (3.1). Let and be the log Sobolev functional in (3.1), that is,
(3.6) We claim that is coercive on . Indeed, choosing , since , by Jensen’s inequality, it follows that
(3.7) Note that the Sobolev inequality, see Lemma 2.8, implies that
(3.8) and therefore
(3.9) which implies that for in with . Let be a minimizing sequence which can be assumed non-negative since . By the coercivity of , it follows that is bounded in . The compact Sobolev embedding implies the existence of a non-negative in and a subsequence of , relabelled as , such that strongly in for all in . By the -lower semicontinuity of the Cheeger energy as well as the -continuity of for small ,444Which can be shown from mean value theorem and equality where is between and , see [45, page 112] it follows that reaches its minimum at . Furthermore, if , a straightforward inspection shows that cannot be an extremal function.
-
Step 2:
Given any non-negative extremal function , we show that and satisfies (3.3) by using the classical Euler-Lagrangian method. Let be an arbitrary Lipschitz function and define the functional as follows
(3.10) The mean value theorem yields
(3.11) with a function taking values between and . From the inequality for some small together with the dominated convergence of Lebesgue, it follows that
(3.12) Since is dense in and is in by the Sobolev inequality, it follows that
(3.13) Plugging into (3.13) and using the real value of the log Sobolev constant , it follows that . By the definition of the Laplacian we deduce that is in and that .
-
Step 3:
We show that is in . We start by showing that is in . Let . Note that using the resolvent of the Laplacian, (3.3) can be rewritten as
(3.14) where is the identity map and is the resolvent of the Laplacian and . Note that in order to prove that is in , it is sufficient to show that is in for some , according to Lemma 2.9. By the Sobolev inequality, we know that is in , which implies that is in for all , (recall that ). Fix . It implies that for . If , then Lemma 2.9 and the identity (3.14) implies that . If , then Lemma 2.9 implies that for all . Repeating the previous step, we can obtain that for some such that
(3.15) Define by induction if which by definition of implies that after finitely many iterations is in for some and deduce that is in .
We now show that is actually in . Let be any point in and be an open ball with and . Since is in and the identity (3.3), it follows by the definition that the restriction of on , which we still denote by , belongs to as well as and that holds on in the distributional sense. Hence, by Lemma 2.6, it follows that and
(3.16) However, is compact and by the doubling property, implying that belongs to . The Sobolev-to-Lipschitz property thus implies that 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.
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 and is finite, the function for all small enough . Hence we can find such that on . We first claim that is of De Giorgi class . In other terms, there exists such that for all in and in and all with , it holds that
(3.17) |
where . In the following, denotes a positive constant, which is independent of the choice of , and may vary from line to line. Let be a Lipschitz cut-off function such that on and with for some . Taking as test function for (3.3), it follows that
(3.18) |
As for the left-hand side of (3.18), the Leibniz rule yields
(3.19) |
where we use the locality of the minimal weak upper gradient for the first inequality, and Young’s inequality together with and on for the second inequality. As for the right hand side of (3.18), by the very choice of and the non-negativity of , it follows that . Hence
(3.20) |
With (3.19) and (3.20), we obtain that
(3.21) |
Adding and then dividing by on both sides, it follows that
(3.22) |
where . Applying the same argument to and enlarging the domain of integral, we derive that for all ,
(3.23) |
By [34, Lemma 3.2] (see also [22, Lemma 3.1 in p. 161]) with , we obtain that
(3.24) |
which implies that is of De Giorgi class .
Note that by the assumption that is a compact space with and , it follows that is global doubling and supports the global weak -Poincaré inequality, see [43]. Together with Hölder’s inequality, supports the global weak -Poincare inequality for any in . Since the minimal weak upper gradient in the setting coincides with the minimal weak upper gradient in the Newtonian setting, see [2], it follows that the space together with that and 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 and in and be fixed. By the Lipschitz continuity of , we can find with such that on . The generalized Bishop-Gromov inequality yields555For this fact, see [39, Theorem 2.14] or [49, Corollary 30.12]
(3.25) |
for some constant depending only on . Taking even smaller if necessary, we can assume that . Hence it follows that
(3.26) |
Taking , the inequality (3.26) implies that
(3.27) |
Since , [34, Lemma 6.2] yields the existence of such that
(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.
-
Step 4:
We show the positivity of non-negative extremal functions. From the continuity of , the set is closed. By contradiction, suppose that is non-empty. By Lemma 3.5, it follows that is also open. Since is connected, it follows that and therefore . This however contradicts the fact that .
∎
Finally we address the proof of Corollary 3.3.
Proof of Corollary 3.3.
Let and be an arbitrary non-negative extremal function of the variational problem (3.1). By Theorem 3.2, we know that for some some positive constants and and that is Lipschitz. Since is a -function with bounded first and second derivatives on , it follows by the chain rule that is Lipschitz, is in and
(3.29) |
By (3.3) we get
(3.30) |
Taking . By locality of the minimal weak upper gradient and the Laplacian, we obtain that satisfies . If , then Theorem 3.2 implies that is non-constant, which also implies that is non-constant. Since 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 , 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 be a compact space with and in . Let such that for some . Then, for all it holds
(4.1) |
Corollary 4.2.
The proof of Theorem 4.1 is divided into three parts: In the first step, we show the regularity for 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 , where all inequalities are understood in the -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 and .
-
Step 1:
We claim that is in and is in with . Indeed, by the assumption that is in , it follows that belongs to and
(4.2) with . By the fact that is in , we obtain that is also in . We now show that . For any , by the weak Bochner inequality (2.7), it follows that
(4.3) where . By the standard regularization via the mollified heat flow, see [25, Corollary 6.2.17], the inequality (4.3) holds for all . Then by [25, Proposition 6.2.16], it follows that and that
(4.4) In particular we obtain that .
-
Step 2:
We provide a lower bound for based on the inequality (4.4) where for . First note that since , it follows that . Hence, from the first step, we get that is in and . Then inequality (4.4) implies that
(4.5) Plugging into (4.5), it follows that
(4.6) Plugging identity into the right hand of (4.6), it follows that
(4.7) For and , inequality (4.7) simplifies to
(4.8) -
Step 3:
We show that the assumptions of Lemma A.2 are valid for based on the estimate (4.8). Let and fix . Since is compact and , we can find in such that:
(4.9) Define
By our choice of and the doubling property of , we have and . Without loss of generality, we assume that , otherwise nothing needs to be shown. We consider different cases for and .
- Case 1:
-
Case 2:
. Let . By [38, Lemma 3.1], we can find a Lipschitz cut-off function with such that and on and , and
(4.13) where is a constant depending only on and . Let be defined as and let . Note that since , it holds that . Together with the fact that is continuous, by the Leibniz rule for the measure-valued Laplacian, it follows that
(4.14) Together with the result in the first step of this proof that , we deduce that
(4.15) Furthermore, from (4.13), it follows that
(4.16) Denote by the right side of (4.8) without the last term . Then, by using and the inequality (4.8), it follows that
(4.17) Using the estimate (4.16) with , it follows that
(4.18) Since , by the definition of , it follows that
(4.19) The doubling property and implies that as well as , which together with the fact that and , allows us to apply Lemma A.2 to . Taking , by Lemma A.2, it follows that there exists a sequence such that and
(4.20) Plugging (4.18), we obtain that
(4.21) where . Following the similar argument as in Case 1 and noting that , we obtain that
(4.22) Since , so multiplying on both side of (4.22) and letting , it follows that
(4.23) The inequality (4.23) holding for any , we send to to obtain
(4.24)
With Case 1 and Case 2, together with the definitions of and , we obtain the inequality (4.1).
∎
Proof of Corollary 4.2.
By Theorem 3.2 and Corollary 3.3, we know that is in and satisfies . Then by the definition of the Laplacian and locality of the minimal weak upper gradient, it follows that is in and satisfies the equation . Hence, by Theorem 4.1, the Li-Yau estimate (4.1) holds for . For the second claim, let denote the term in the right-hand side of (4.1). If , then taking . One can check that goes to as . Hence, taking on the both sides of (4.1), it follows that
Then the Sobolev-to-Lip property implies that 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 spaces.
For the notational simplicity, we define the non-negative constants and as follows:
(5.1) |
where . Note that is positive whenever .
The first direct consequence is a Harnack type inequality for the non-negative extremal functions.
Corollary 5.1.
Let be a compact space with and in . Suppose that satisfies for some . Then for any , it holds that
(5.2) |
for any and .
Proof.
First note that by Theorem 4.1, we have
(5.3) |
Let be arbitrary points in and let and for . By [44, Corollary 1.2], spaces are essentially non-branching. Together with [13, Corollary 5.3], there exists a unique -geodesic joining and with for any in for some , and a test plan in such that for any in and is concentrated on the set of geodesics on . By the Fubini’s theorem together with the fact that has bounded compression and the inequality (5.3), it follows that for -a.s ,
Let be any such a geodesic. By the continuity of , let and in be the maximum and minimum points respectively, that is
(5.4) |
Then by the definition of weak upper gradients (see [25, Proposition 1.20]), it follows that
(5.5) |
For , together with , it follows that
(5.6) |
Let . Since , it follows that
(5.7) |
Integrating (5.7) with respect to on the both sides and noting that , it follows that
(5.8) |
where is the optimal transport plan between and . Since and weakly converges to and in the duality of as , it follows that, up to a subsequence, weakly converges to . Using the identity , we obtain that
(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 -Poincaré inequality for , 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 based on the dimension-free Harnack inequality on the 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 be a compact space with and in . Then for any non-negative extremal function of (3.1) with the log Sobolev constant with , it holds that
(5.10) | |||
(5.11) |
where when and when . In particular, it holds that -a.e. for any .
Proof.
Let and and be the maximum and minimum point of on , 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 spaces shown in [36, Theorem 3.1] for , it follows that
(5.12) |
Since by the mass-preserving property of the heat semigroup, we deduce from (5.12) by taking that
(5.13) |
Secondly, by the equation (3.3) and the commutation between and , it follows that
(5.14) |
for -a.e . By the regularization of the heat semigroup, the equality (5.14) holds for any . Since is positive and is the maximum point, it follows that for any in which by the comparison principle of the heat semigroup, yields
(5.15) |
Grönwall’s inequality further implies that
(5.16) |
So, together with (5.13), it follows that for any :
(5.17) |
By using when and letting , it follows that
(5.18) |
where when and when .
As for the lower bound, from and similar arguments as above, it follows that
(5.19) |
Taking in the dimension-free inequality (5.12) and plugging (5.16) and (5.19), it follows that
(5.20) |
where in the third inequality we use the comparison principle of the heat semigroup that , and in the last inequality we use the inequality (5.19). After reorganizing the inequality, it follows that
(5.21) |
Taking and , we obtain that
(5.22) |
where we use the inequality when 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 is constant when and .
Corollary 5.4.
Let be a space with and . Then any non-negative extremal function of (3.1) with the log-Sobolev constant is constant whenever .
Proof.
Let be an arbitrary non-negative extremal function of (3.1) with a log-Sobolev constant with , and define . Let further , , and be real numbers to be determined later. We first estimate from both the PDE equation (3.3) and the weak Bochner inequality (2.7), and then derive the desired result.
-
Step 1:
We first derive the estimate from the equation (3.3). Let in (3.3). By the Lipschitz regularity of and the regularity of proved in Step 1 in Theorem 4.1, it follows that and . On the one hand, applying and then to the right-hand side of (3.3) it follows that
(5.23) On the other hand, applying to the left-hand side of (3.3), we get
(5.24) Showing that
(5.25) -
Step 2:
We derive the estimate from the weak Bochner inequality (2.7). Note that and satisfies the regularity requirement in (2.7). So plugging and into (2.7), it follows that the left-hand side of (2.7) can be expressed as
(5.26) and the right-hand side of (2.7) can be expressed as
(5.27) From (5.26) and (5.27), it follows that (2.7) reads as follows:
(5.28) -
Step 3:
We conclude by comparing the coefficients of (5.25) and (5.28) and choosing particular values for . Let be determined later and let satisfy the following system of equations:
(5.29) This system (5.29) admits real-valued solutions if and only if . Hence, choosing arbitrary and
By comparing (5.25) and (5.28), it follows that
(5.30) Since , then has to be , implying that 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 spaces with in which may not support the doubling property. To show it, we first show the Kato’s inequality in the proper 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 on is the weak Laplacian provided that for each , is a linear functional acting on with bounded support defined as:
(A.1) |
For each in , is the linear functional given by for each in with bounded support. We say that is a signed Radon measure provided that there exists a signed Radon measure such that for all with bounded support. It is clear that in this case, we have and . For and , denotes the weighted weak Laplacian on defined as
(A.2) |
It is also easy to check that . When is a signed Radon measure, we denote by its Lebesgue decomposition with respect to . Finally, we remark that using the similar arguments as in [51, Lemma 3.2], we have the following chain rule: for and , it holds that and that
(A.3) |
Lemma A.1.
(Kato’s inequality) Let be a proper space and be in . Suppose in is such that is a signed Radon measure and that . Then is a signed Radon measure such that
(A.4) |
where and .
Proof.
It suffices to prove that is a signed Radon measure and that
(A.5) |
where if and if and if . Indeed, if the inequality (A.5) holds, then it follows that both and are signed Radon measures and (A.5) implies that
(A.6) |
By locality of the minimal weak upper gradient and the inner regularity of Radon measures, it is immediate to check that is concentrated on the set . Then the inequality (A.6) with the assumption that implies that
(A.7) |
We are left to show (A.5). Let and and . Since is in , it follows that and that
(A.8) |
Further, the chain rule of for and yields
(A.9) |
So by (A.8), it follows that
(A.10) |
Now, let be any arbitrary bounded open subset. Let be a positive cut-off function such that on . Since and , it follows that and that
Since is in , it follows that is uniformly bounded in . Note that by the assumption on , both the Sobolev spaces and and the minimal weak upper gradients induced by and coincide (see [3, Lemma 4.11]) and hence is also uniformly bounded in . Since is a Hilbert space, is also Hilbert and hence reflexive. Therefore, there exists a subsequence with converging weakly in to some . As pointwise, we obtain that -a.e. and in particular -a.e. on since on . Since for all in and , for any non-negative Lipschitz , we obtain that
(A.11) |
Following a similar argument as in [16, (4-28),(4-29) in Theorem 4.14], there exists a constant such that for any Lipschitz function with . Since is proper, by the Riesz representation theorem, there exists a signed Radon measure on such that for all in and that on (see the remark before Theorem 1.2 in [51] or [25, Proposition 6.2.16] for finite signed Radon measures). Clearly and coincide on for any bounded open subsets and . Henceforth, there exists a unique signed Radon measure on such that for all bounded open domain , and thus, we obtain that is a signed Radon measure with . ∎
We now show the Omori-Yau type maximum principle in proper 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 spaces which in general does not support the doubling property.
Lemma A.2.
(Omori-Yau type maximum principle) Let be a proper space with in . Let further be in such that . Suppose that achieves one of its strict maximum in in the sense that there exists a bounded and measurable subset satisfying and with
(A.12) |
Then, given any in , for any , we have
(A.13) |
In particular, there exists a sequence in such that and .
Proof.
We adapt the proof in [51]. Let . Suppose by contradiction that there exists and in such that (A.13) fails. Then for possibly smaller such that , it follows that is in with -a.e. in , and that
(A.14) |
It follows that for -a.e. in , we have
(A.15) |
Note that since and is a signed Radon measure, it follows that is a well-defined signed Radon measure. By the identity of the weighted weak Laplacian and , it follows that
(A.16) |
on . Moreover, the identity of the weighted weak Laplacian together with implies that and is in , applying Lemma A.1 to , yields
(A.17) |
By the definition of , it follows from (A.17) that
(A.18) |
which implies that -a.e. From being equivalent to since , we get that -a.e. Together with being in and in , by the Sobolev-to-Lip property of spaces, it follows that admits a Lipschitz representation with . Then, from the fact that -a.e. as well as -a.e on , it follows that is constant and . Hence, -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 spaces with and 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 spaces with and 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 . By the assumptions and the generalized Bishop-Gromov inequality, without loss of generality, we can assume that . By the Sobolev inequality (2.12) and the assumption that , it follows that is bounded in . Since is reflexive, there exists such that there exists a subsequence of such that converges to weakly in . It suffices to show that converges to strongly in for all . For the notation’s convenience, we relabel as .
Let and be arbitrary and . Let and for any . Then it follows that
(B.1) |
Note that the compact space with bounded diameter support the global doubling property and the weak -Poincaré inequality (2.10) with the global constant . By the equivalent characterizations of Poincaré inequalities [33, Theorem 2] and [11, Proposition 4.13], it follows that the weak -Poincaré inequality (2.10) holds for all pairs . Moreover, by the doubling property of spaces, compact spaces support the Lebesgue differentiation theorem and the maximal theorem (see [29]). Let be any Lebesgue point of and for . Then for the first term of the right-hand side of (B.1), it follows that
(B.2) |
where be the centered maximal operator. So by applying the maximal theorem [29, Theorem 3.5.6], it follows that
(B.3) |
For the second term of the right-hand side of (B.1), note that since converges to weakly in , so taking the test function , it follows that
(B.4) |
which implies that converges to pointwise and thus converges to in probability. For the third term of right-hand side of (B.1), applying the Lebesgue differentiation theorem, it follows that converges to -a.e. on as . Together with all arguments above, taking first and then taking on both side of (B.1), it follows that
(B.5) |
Then by [28, Lemma 8.2], we obtain that strongly in for all .
∎
References
- Ambrosio and Honda [2018] Luigi Ambrosio and Shouhei Honda. Local spectral convergence in 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 -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 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 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 -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 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 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 -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 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.