The Cheeger problem in abstract measure spaces
Abstract
We consider nonnegative -finite measure spaces coupled with a proper functional that plays the role of a perimeter. We introduce the Cheeger problem in this framework and extend many classical results on the Cheeger constant and on Cheeger sets to this setting, requiring minimal assumptions on the pair measure space perimeter. Throughout the paper, the measure space will never be asked to be metric, at most topological, and this requires the introduction of a suitable notion of Sobolev spaces, induced by the coarea formula with the given perimeter.
1 INTRODUCTION
In the past decades, the Cheeger constant has been extensively studied in view of its many applications: the lower bounds on the first eigenvalue of the Dirichlet -Laplacian operator [88] and the equivalence of such an inequality with Poincaré's one [104] (up to some convexity assumptions); the relation with the torsion problem [32, 33]; the existence of sets with prescribed mean curvature [5, 95]; the existence of graphs with prescribed mean curvature [77, 94]; the reconstruction of noisy images [46, 60, 73, 113]; the minimum flow-maximum cut problem [79, 117]; and its medical applications [16]. In addition, the Cheeger constant has been employed in elastic-plastic models of plate failure [90] and (its Euclidean-weighted variant) has found applications to Bingham fluids [83] and landslide models [84]. Moreover, the Cheeger constant of a square has been recently used to provide an elementary proof of the Prime Number Theorem [17]. For more literature and a general overview of the problem, we refer the interested reader to the two surveys [91, 112].
Because of its numerous applications, several authors have been drawn to the subject and started to investigate the constants and the above-mentioned links with other problems in several frameworks: weighted Euclidean spaces [7, 26, 96, 114]; finite-dimensional Gaussian spaces [51, 86]; anisotropic Euclidean and Riemannian spaces [8, 21, 49, 89]; the fractional perimeter [28] or nonsingular nonlocal perimeter functionals [100]; Carnot groups [108]; -spaces [70, 71]; and lately smooth metric-measure spaces [1].
In the settings mentioned above, the proofs mostly follow those available in the usual Euclidean space. In the present paper, we are interested in pinpointing the minimal assumptions needed on the space and on the perimeter functional in order to establish the fundamental properties of Cheeger sets, as well as the links to other problems. In the following, we shall be interested in nonnegative -finite measure spaces endowed with a perimeter functional satisfying some suitable assumptions.
Our approach fits into a broader current of research that has gained popularity in the past decade, aiming to study variational problems, well known in the Euclidean setting, in general spaces under the weakest possible assumptions. Quite often, the ambient space is a (metric) measure spaces. For example, such a general point of view has been adopted for the variational mean curvature of a set [19], for shape optimization problems [37], for Anzellotti–Gauss–Green formulas [78], for the total variation flow [34, 35], and, very recently, for the existence of isoperimetric clusters [111].
1.1 Structure of the paper and results
In Section 2, we introduce the basic setting of perimeter -finite measure spaces, that is, nonnegative -finite measure spaces endowed with a proper functional possibly satisfying suitable properties (P.1)–(P.7) that we shall require from time to time.
A considerable effort goes toward defining functions in measure spaces, where a metric is not available. Indeed, usually, the perimeter functional is induced by the metric. In our setting, instead, only a perimeter functional is at disposal, so we use it to define functions by defining the total variation via the coarea formula with the given perimeter.
To properly define Sobolev functions, we need a slightly richer structure, requiring the measure space to be endowed with a topology, and the perimeter functional to arise from a relative (with respect to open sets ) perimeter functional . By using the relative perimeter, we refine the notion of function by requiring that the variation is a finite measure. This, in turn, allows us to define functions as functions whose variation is absolutely continuous with respect to the reference measure. Afterward, via relaxation, we can define functions for any . For a more detailed overview, we refer the reader to Section 1.1.3 and Section 1.1.4.
Once the general framework is set, we then shall start to tackle the problem of our interest.
1.1.1 Definition and existence
In Theorem 3.6, we provide a general existence result. Unsurprisingly, the key assumptions on the perimeter are the lower semicontinuity and the compactness of its sublevels with respect to the norm, besides an isoperimetric-type property that prevents minimizing sequences to converge toward sets with null -measure.
Further, we provide inequalities between the -Cheeger and -Cheeger constants and prove some basic properties of -Cheeger sets, with a particular attention to the case .
1.1.2 Link to sets with prescribed mean curvature
In Section 4, we introduce the notion of -mean curvature in the spirit of [19]. With this notion at our disposal, we show that any 1-Cheeger set has as one of its -mean curvatures, see Corollary 4.3. An analogous result holds for the chambers of an -cluster minimizing , see Corollary 4.4.
1.1.3 Link to the first eigenvalue of the Dirichlet 1-Laplacian
On a general set in the Euclidean space, the infimum in (1.2) is less than or equal to that in (1.1), since one only has the inclusion .
This approach allows us to consider problem (1.2) without any underlying metric structure. In addition, no regularity of the set is required, since there is no need for the problem (1.2) to be equivalent to its regular counterpart (1.1) that, in the present abstract framework, cannot be even formally stated.
With this notion of total variation at hand, we prove that the constant coincides with the 1-Cheeger constant under minimal assumptions on the perimeter, that is, we require that the perimeter of negligible sets and of the whole space is zero, the perimeter is lower semicontinuous with respect to the norm, and that the perimeter of a set coincides with that of its complement set, see Theorem 5.4. Moreover, we prove some inequalities relating the -Cheeger constant with a cluster counterpart of (1.2). As observed in Remark 5.9, if one slightly modifies the definition of by considering nonnegative functions as the only competitors, then one can obtain the relation with the Cheeger constant even for perimeter functionals that are not symmetric with respect to the complement-set operation.
1.1.4 Link to the Dirichlet -Laplacian and the -torsion
In the Euclidean space, the 1-Cheeger constant comes into play in estimating some quantities related to the Laplace equation and to the torsional creep equation. More precisely, it provides lower bounds on the first eigenvalue of the Dirichlet -Laplacian for and to the norm of the -torsional creep function. In Section 6, we extend these results to our more general framework.
Both these problems require an extensive preliminary work to define Sobolev spaces in our general (nonmetric) context. In order to do so, we need a little more structure on the perimeter-measure space: we require it to be endowed with a topology, we require the class of measurable sets to be that of Borel sets, and we require the perimeter to stem from a relative perimeter when evaluated relatively to the whole space .
We here quickly sketch how we construct these Sobolev spaces, and we refer the interested reader to Section 2.3. A relative perimeter functional allows, again via the relative coarea formula in a similar fashion to (1.3), to define the relative variation of an function with respect to a measurable set. When this happens to define a measure, we shall say that the function is in , and this extends the notion briefly discussed in Section 1.1.3 and formally introduced in Section 2.2. When this measure happens to be absolutely continuous with respect to , we shall say that the function is in and that the density of the measure with respect to is the 1-slope of . Via approximation arguments, one can then define the -slope of a function and the associated spaces. In turn, the approximation properties allow to define the Sobolev space , refer to Definition 6.1.
1.1.5 Examples
In the last section of the paper, we collect several examples of spaces that meet our hypotheses. In particular, our very general approach basically covers all results known so far about the existence of Cheeger sets in finite-dimensional spaces, and the relation of the constant with the first eigenvalue of the Dirichlet -Laplacian in numerous contexts. In some of the frameworks presented in Section 7, the results are new, up to our knowledge.
Unfortunately, our approach does not cover the case of the infinite-dimensional Wiener space. In this case, one can suitably define the Cheeger constant and prove the existence of Cheeger sets. Nonetheless, this requires ad hoc notions of function and of perimeter that are quite different from the ones adopted in the present paper. We refer the interested reader to [51, Sect. 6] for a more detailed exposition about this specific framework.
2 PERIMETER-MEASURE SPACES
2.1 Perimeter functional
In the same spirit of [19, Sect. 3], we introduce the following definition.
Definition 2.1.A perimeter functional is any map
- (P.1) ;
- (P.2) ;
- (P.3) for all ;
- (P.4) is lower semicontinuous with respect to the convergence;
- (P.5) for any with , the family
is compact in for all ;
- (P.6) there exists a function such that
with the following property: if and with , then ;
- (P.7) for all .
Assuming property (P.7) true, properties (P.1) and (P.2) become equivalent. Throughout the paper, we often refer to (P.6) as an isoperimetric property. Notice that, in case an isoperimetric inequality holds true for suitable and , and for all with , then (P.6) clearly follows. Depending on the situation, it could be more convenient to prove (P.6) directly or to rely on a finer isoperimetric-type inequality, see Section 7. We remark that all the properties listed above will appear every now and then throughout the paper, but they are not enforced throughout — every statement will precisely contain the bare minimum for its validity.
Remark 2.2. ( is invariant under -negligible modifications)Let property (P.4) be in force. If are such that , then . To see this, consider any measurable set and any -negligible set , look at the constant sequence converging to in , and at the constant one converging to and exploit (P.4).
Remark 2.3.Let property (P.6) be in force. If , then the set is -negligible, that is, . Conversely, if , then . Thus, property (P.6) says that the only sets with finite measure that could possibly have zero perimeter are -negligible sets. Moreover, if properties (P.1) and (P.4) are in force as well, then -negligible sets have zero perimeter, thanks to Remark 2.2.
2.2 Variation and functions
We begin with the following result, proving that assuming the validity of properties (P.1) and (P.2), the variation coincides with the perimeter functional on characteristic functions.
Lemma 2.4. (Total variation of sets)Let properties (P.1) and (P.2) be in force. If , then .
Proof.By definition, (P.1), and (P.2), the function
Remark 2.5.As an immediate consequence of Lemma 2.4, if (P.1) and (P.2) are in force, then is a proper functional and whenever is such that and . In particular, with .
The following result rephrases [59, Prop. 3.2] in the present context.
Lemma 2.6. (Basic properties of total variation)The following hold:
- (i) for all and ;
- (ii) for all and ;
- (iii) if (P.1) and (P.2) are in force, then for all ;
- (iv) if (P.4) is in force, then is lower semicontinuous with respect to the (strong) convergence in .
Proof.The proofs of the first three points are natural consequences of the definition.
Proof of (iv). Let be such that in as . Without loss of generality, we can assume that
The following result, which can be proved as in [59] up to minor modifications, states that the variation functional is convex as soon as the perimeter functional is sufficiently well behaved.
Proposition 2.7. (Convexity of variation)Let properties (P.1), (P.2), (P.3), and (P.4) be in force. Then, is convex. As a consequence, is a convex cone in .
2.3 Relative perimeter and relative variation
In this subsection, we assume that the set is endowed with a topology such that , the Borel -algebra generated by .
Definition 2.8.A relative perimeter functional is any map
- (RP.1) for all ;
- (RP.2) for all ;
- (RP.3) for all and ;
- (RP.4) for each , is lower semicontinuous with respect to the (strong) convergence in .
Below, we rephrase Lemma 2.4, Lemma 2.6, and Proposition 2.7 in the present setting. Their proofs are omitted, because they are similar to those already given or referred to.
Lemma 2.9. (Relative variation of sets)Let properties (RP.1) and (RP.2) be in force. If , then for all .
Lemma 2.10. (Basic properties of relative variation)The following hold:
- (i) for all , , and ;
- (ii) for all , , and ;
- (iii) if (RP.1) and (RP.2) are in force, then for all and ;
- (iv) if (RP.4) is in force, then, for each , the relative variation is lower semicontinuous with respect to the (strong) convergence in .
Proposition 2.11. (Convexity of relative variation)Let properties (RP.1), (RP.2), (RP.3), and (RP.4) be in force. Then, for each , the functional is convex.
2.3.1 Variation measure
We now define the perimeter and variation measures by rephrasing the validity of the relative coarea formula (2.5) in a measure-theoretic sense.
Definition 2.12. (Perimeter and variation measures)We say that a set has finite perimeter measure if its relative perimeter
Adopting the usual notation, if has finite perimeter measure, then we write for all . Similarly, if has finite variation measure, then we write for all .
It is worth noticing that Definition 2.12 is well posed in the following sense. As soon as properties (RP.1) and (RP.2) are in force, if has finite perimeter measure, then has finite variation measure with , since they are outer regular Borel measures on agreeing on open sets. This is a simple consequence of Lemma 2.9.
Corollary 2.13. (Generalized coarea formula)If has finite variation measure, then
- (RP.+) .
We now outline some consequences of Lemma 2.10 and Proposition 2.11, and leave the simple proofs of these statements to the interested reader, see also the proof of Lemma 2.6.
Corollary 2.14. (Basic properties of variation measure)Let properties (RP.1), (RP.2), (RP.3), and (RP.4) be in force. The following hold:
- (i) if has finite variation measure, then has finite variation measure, with , for all ;
- (ii) if has finite variation measure, then has finite variation measure, with , for all ;
- (iii) constant functions have finite variation measure and for all (in particular, );
- (iv) if and in as for some , then
for all ;
- (v) if also property (RP.+) is in force and , then
as outer regular Borel measures on .
2.3.2 Chain rule
- (RP.L) for all with .
Theorem 2.15. (Chain rule)Let properties (RP.1), (RP.2), and (RP.L) be in force and let be a strictly increasing function. If has finite variation measure, then also has finite variation measure, with
Proof.Since is strictly increasing, its inverse function is well defined, continuous and strictly increasing, and we can write
Thus, letting , we have that for all ; hence, the following equalities hold:
2.3.3 -Slope and Sobolev functions
Corollary 2.16. (Basic properties of 1-slope)Let properties (RP.1), (RP.2), (RP.3), (RP.4), and (RP.+) be in force. The following hold:
- (i) if , then , with , for all ;
- (ii) , with ;
- (iii) if , then with
Having the notion of 1-slope at our disposal, following the standard approach about slopes (see [13], e.g.), we can introduce the notion of -relaxed 1-slope, for .
Definition 2.17. (-Relaxed 1-slope)Let . We shall say that a function is a -relaxed 1-slope of if there exist a function and a sequence such that:
- (i) in ;
- (ii) for all and weakly in ;
- (iii) -a.e. in .
Lemma 2.18. (Basic properties of -relaxed 1-slope)Let properties (RP.1), (RP.2), (RP.3), (RP.4), and (RP.+) be in force and let . The following hold:
- (i) is a convex subset (possibly empty) for all ;
- (ii) if and , then there exist a sequence , a sequence , and a function , such that and both in , with for all and ;
- (iii) if and are sequences in , with for all , such that and weakly in , then .
Under the assumptions of the above Lemma 2.18, for each , the set is a (possibly empty) closed convex subset of , and thus, the following definition is well posed.
Definition 2.19. (Weak -slope)Let and let properties (RP.1), (RP.2), (RP.3), (RP.4), and (RP.+) be in force. If is such that , we let be the element of of minimal -norm and we call it the weak -slope of . Finally, we let
Following the same line of [13], one can show that the weak -slope can be actually approximated in in the strong sense.
Corollary 2.20. (Strong approximation of weak -slope)Let and let properties (RP.1), (RP.2), (RP.3), (RP.4), and (RP.+) be in force. If , then there exists a sequence such that for all and
3 CHEEGER SETS IN PERIMETER-MEASURE SPACES
In this section, we work in a measure space endowed with a perimeter functional as in Section 2.1.
3.1 -Cheeger constant and -Cheeger sets
We begin by introducing the central notions of the present paper.
Definition 3.1.Let . An -cluster is a collection of measurable sets satisfying:
- for all ;
- for all with ;
- for all .
Definition 3.2. (-admissible set)Let . We say that is N-admissible if there exists an -cluster .
Remark 3.3.Let . Trivially, if is -admissible, then it is -admissible for all integers .
Definition 3.4. (-Cheeger constant and -Cheeger sets)Let and let be an -admissible set. The -Cheeger constant of is
3.1 Existence of -Cheeger sets
We prove that the existence of -Cheeger clusters of is ensured whenever the perimeter functional possesses properties (P.4), (P.5), and (P.6), and the set is -admissible with finite -measure. These requests are not necessary though, as some examples at the end of this section show.
Theorem 3.6.Let properties (P.4), (P.5), and (P.6) be in force. Let , and let be an -admissible set with . Then there exists an -Cheeger set of .
Proof.On the one hand, since is -admissible, there exists an -cluster , which immediately implies that . On the other hand, for any -cluster of , property (P.6) gives
By (P.4) and (P.5) (recall also Remark 2.2), possibly passing to a subsequence, for each , there exists such that , with , , and as . Now, using (P.6), for all sufficiently large and any , we get
It remains to be proved that is an -cluster contained in , that is, that the chambers are pairwise disjoint, and the reader can easily check it on its own.
Consequently, thanks to (P.4), we find that
Let us point out that properties (P.4), (P.5), and (P.6) are all crucial in the above proof. Among them (P.6) looks as the “most artificial”; nevertheless, it is essential in the reasoning: an example where existence fails when (P.6) is missing is given in Example 3.7 below. It is also relevant to point out that these properties provide a sufficient but in no way a necessary condition, as Example 3.8 and Example 3.9 show.
Example 3.7.Consider the measure space , where denotes the Borel -algebra, is defined by
Within this framework, one has , for any set containing an open neighborhood of the origin. Indeed, it is enough to consider the sequence of balls centered at the origin (for sufficiently small), for which we have
For the sake of completeness, we shall note that, in the situation depicted in this remark, -Cheeger sets exist in any open set not containing the origin, since the weight would be , refer to [19, Prop. 3.3] or to [114, Prop. 3.2].
We now present two simple examples in which the existence of Cheeger sets is ensured even if properties (P.5) and (P.6) do not hold.
Example 3.8.Consider any nonnegative (-finite) measure space , and consider , for all , as perimeter functional. For this choice, while (P.4) holds, neither property (P.5) nor (P.6) hold, the latter because any isoperimetric function is bounded from above by 1. Nevertheless, fixed any , we have , for any integer , and any -cluster is an -Cheeger set.
Example 3.9.Consider any nonnegative (-finite) measure space , and consider , for all , as perimeter functional. While (P.4) holds, neither property (P.5) nor (P.6) hold. Nevertheless, fixed any , we have , for any integer , and any -cluster is an -Cheeger set.
3.2 Inequalities between the - and -Cheeger constants
Proposition 3.10.Let be an -admissible set. Then, for all with , one has
Proof.Let and be fixed integers, with . Let be any fixed -cluster. For any subset of of cardinality , the -cluster provides an upper bound to , whereras the -cluster to .
Hence, no matter how we choose , we have
Corollary 3.11.Let be an -admissible set. Then, for all such that for some integer one has , one has
Remark 3.12.The inequalities (3.2) and (3.3) hold as equalities in some cases, as, for instance, it happens anytime a set has multiple disjoint 1-Cheeger sets. A trivial example of this behavior is given by disjoint and equal balls in the usual Euclidean space.
One can also build connected sets that have this feature. For , it is enough to consider a standard dumbbell in the usual two-dimensional Euclidean space, that is, the set given by two disjoint equal balls, spaced sufficiently far apart, and connected via a thin tube. Such a set has two connected 1-Cheeger sets and given by small perturbations of the two balls, and the 2-cluster is necessarily a 2-Cheeger set, refer, for instance, to [93, Ex. 4.5].
An easy connected example for is instead given by an -dumbbell in the usual two-dimensional Euclidean space, that is, a set formed by disjoint equal balls and linked by a thin tube, say
3.3 -subclusters of -Cheeger sets
Given an -Cheeger set of , consider any of its -subcluster. It is natural to imagine that such an -cluster is an -Cheeger set in the ambient space given by minus the chambers not belonging to the subcluster. In this short section, we prove that this is true.
For the sake of clarity of notation, we let be the cardinality of a set .
Proposition 3.13.Let be an -admissible set, and assume that it has an -Cheeger set . For any proper subset , let
Proof.It is enough to prove the claim for a subset of cardinality , and then to reason by induction. In particular, up to relabeling, we can assume to be the proper subset .
As both and are measurable, so it is the set . Moreover, this latter is -admissible because there exists at least the -cluster .
By contradiction, assume that is not an -Cheeger set of . Then, for small enough, we find a different -cluster with
3.4 Properties of -Cheeger sets
Proposition 3.14. (Basic properties of -Cheeger sets)Let be a collection of -admissible sets. The following hold for all integers :
- (i) if , then ;
- (ii) if (P.6) is in force, and , then ;
- (iii) if (P.4), (P.5), and (P.6) are in force, and in , with , then
Moreover, if also (P.3) is in force, is finite, and , then
Proof.Recall that an -admissible set is also -admissible for all integers , see Remark 3.3.
Proof of (i). For any two fixed -admissible sets with , any -cluster of is also an -cluster of . The inequality immediately follows by definition of -Cheeger constant.
Proof of (ii). In virtue of (3.3) and the positivity of , it is enough to prove the claim for . Fix , and for all , let be such that
Proof of (iii). Without loss of generality, we can assume that there exists a constant independent of such that
To show the second part, let us pick an -Cheeger cluster of , which exists since we are under the assumptions of Theorem 3.6. Let us consider the collections
Remark 3.15.Notice that to prove point (iii), all requests come into play. Indeed, a priori, one can work with an “almost-infimizing” -cluster for and find an analogous of (3.5) up to an additive factor . Then, the compactness granted by (P.5) is needed, but in order to ensure that the limiting collection