The monopolist’s free boundary problem in the plane
Abstract.
We study the Monopolist’s problem with a focus on the free boundary separating bunched from unbunched consumers, especially in the plane, and give a full description of its solution for the family of square domains . The Monopolist’s problem is fundamental in economics, yet widely considered analytically intractable when both consumers and products have more than one degree of heterogeneity. Mathematically, the problem is to minimize a smooth, uniformly convex Lagrangian over the space of nonnegative convex functions. What results is a free boundary problem between the regions of strict and nonstrict convexity. Our work is divided into three parts: a study of the structure of the free boundary problem on convex domains in showing that the product allocation map remains Lipschitz up to most of the fixed boundary and that each bunch extends to this boundary; a proof in that the interior free boundary can only fail to be smooth in one of four specific ways (cusp, high frequency oscillations, stray bunch, nontranversal bunch); and, finally, the first complete solution to Rochet and Choné’s example on the family of squares , where we discover bifurcations first to targeted and then to blunt bunching as the distance to the origin is increased. We use techniques from the study of the Monge–Ampére equation, the obstacle problem, and localization for measures in convex-order.
∗Robert McCann’s work was supported in part by the Canada Research Chairs program CRC-2020-00289, Natural Sciences and Engineering Research Council of Canada Discovery Grants RGPIN–2020–04162, and a grant from the Simons Foundation (923125, McCann). Cale Rankin’s was supported in part by postdoctoral fellowships from the Fields Institute for Research in Mathematical Sciences and the University of Toronto. Kelvin Shuangjian Zhang’s was supported by a Start-Up Grant from Fudan University. The authors are grateful to Krzysztof Ciosmak and Ting-Kam Leonard Wong for fruitful conversations. ©
1. Introduction
The Monopolist’s problem is a principal-agent model for making decisions facing asymmetric information; it has fundamental importance in microeconomic theory. A simple form from [50] capturing multiple dimensions of heterogeneity that we rederive below is to
(1) |
We always take to be open and convex with compact closure . Our goal in this paper is to elucidate the properties of solutions to this minimization problem, with an eventual focus on the two-dimensional setting. Because the minimization takes place over the set of convex functions the problem has a free boundary structure. The free boundary separates the region where the function is convex, but not strictly convex, from the region where the function is strictly convex. In this paper we refine our understanding of the free boundary structure in all dimensions by showing regions where is not strictly convex always extend to the fixed boundary . We show the regularity known for often extends from the interior to the fixed boundary. In two-dimensions, we show in a neighbourhood of a certain class of free boundary points (that we call tame), the minimizer solves the classical obstacle problem. From this we obtain the tame free boundary is locally piecewise Lipschitz except at accumulation points of its local maxima; it has Hausdorff dimension strictly less than two and is the graph of a function continuous -almost everywhere. We also establish a boot strapping procedure: if the free boundary is suitably Lipschitz, it is . As an application of our techniques we completely describe the solution on the square domains with . Despite significant numerical [7, 17, 18, 19, 25, 42, 44, 48], and analytic attempts [50, 35, 41] the description of the solution on the square has previously remained incomplete, and has come to be regarded as analytically intractable [7, 19, 34]. At least on the plane, we rebut this view by confirming for the solution recently hypothesized by McCann and Zhang [41]. We show how their solution can also be modified to accommodate smaller values of and other convex, planar domains. We show that the nature of the bunching undergoes unanticipated changes — from absent to targeted to blunt — as is increased. We rigorously prove the support of the unknown distribution of products consumed has a lower boundary which is concave nondecreasing — as the above-mentioned numerics and stingray description suggest — and that all products selected by more than one type of consumer lie on this boundary or its reflection through the diagonal.
The problem (1) arises from the question of how a monopolist should price goods for optimal profit in the face of information asymmetry. Here is a simple derivation of (1). We assume a closed set of products where each coordinate represents some attribute of the product, and an open set of consumers where each coordinate represents some attribute of the consumer. Consumers are distributed according to a Borel probability measure . The monopolist’s goal is to determine a price at which to sell product in a way which maximizes their profit. If consumer attains benefit from product then the consumer will choose the product which maximizes their utility
(2) |
Provided it exists, we denote the which realizes this supremum by . Assuming the monopolist pays cost for product , then the monopolist’s goal is to maximize their profit, the integral of price they sell for minus cost they pay,
The problem has been considered in this generality in e.g. [3] [12] [30] [38], and for even more general (non quasilinear) utility functions in [47], [39]. For this paper, to highlight the mathematical properties of most interest, we adopt several standard simplifying assumptions proposed by Rochet and Choné [50]: that products lie in the nonnegative orthant and the monopolist’s direct cost to produce them is quadratic , furthermore, that consumers are uniformly distributed on their domain and their product preferences are bilinear . In this case (2) implies is the Legendre transform of (and thus a convex function), , and when the Monopolist’s goal becomes to maximize
over nonnegative convex functions which, up to an irrelevant constant, is the problem (1). Since convex functions are differentiable almost everywhere the integrand is well-defined. The nonnegativity constraint on represents the additional requirement that , meaning consumers need not consume if the monopolist raises prices too high, or equivalently, are always free to pay nothing by choosing the zero product as an outside option.
This problem was first considered by Mussa and Rosen in the one-dimensional setting [46], (after related models of taxation [45], matching [4], and signaling [53] were introduced and analyzed by Mirrlees, Becker and Spence). The multidimensional problem was considered by Wilson [55] and Armstrong [2], while our formulation above is essentially that of Rochet and Choné [50]. Although this model is of significant importance to economists it presents serious mathematical difficulties. Indeed, were there only the nonnegativity constraint in (1) we would have a variant of the obstacle problem [8, 29]; with the nonnegativity and convexity constraint we have a free boundary problem for three different regions. In the region where the function is positive but not strictly convex, the fundamental tool of two-sided perturbation by an arbitrary test function is no longer applicable. As a result, until recent work of the first and third authors [41] it has not even been possible to write down the Euler–Lagrange equation in the region of nonstrict convexity. Despite this, other aspects of the problem have been studied, notably by Rochet and Choné [50], who derived a necessary and sufficient condition for optimality in terms of convex-ordering between the positive and negative parts of the variational derivative of the objective functional, Basov [3] who advanced a control theoretic approach to such problems, Carlier [12, 13] who considered existence and first-order conditions for the minimizer, and Carlier and Lachand-Robert [14, 20, 16] who studied regularity and gave a description of the polar cone.
In this paper we prove results of mathematical and economic interest. We invoke tools from diverse areas of mathematics: the theory of sweeping and convex orders of measures, Monge–Ampére equations, regularity theory for the obstacle problem, and the theory of optimal transport (which has deep links to the Monopolist’s problem). We also indicate a striking connection to the classical obstacle problem: Locally the minimizer solves an obstacle problem where the obstacle is the minimal convex extension of from its region of nonstrict convexity. We now outline our results.
If solves (1) and is a convex open subset of it is known from the work of Rochet–Choné and Carlier–Lachand-Robert that and from the work of Caffarelli and Lions [9] (see [38]) that . Any convex function partitions according to its sets of contact with supporting hyperplanes; these sets are convex. Namely for each set
(3) | ||||
Here is the equivalence class of under the equivalence relation if and only if . We call an equivalence class trivial if , in which case we say is strictly convex at . We call equivalence classes leaves, since they foliate the interior of . They are also called isochoice sets [23] or bunches if nontrivial [50]. We also call one-dimensional leaves rays. We set
(4) |
Thus, for example, consists of all points at which is strictly convex and consists of all points lying in the closure of some open set on which is affine. These disjoint sets partition and our first result describes the qualitative behavior in each set.
Theorem 1.1 (Partition into foliations by leaves that extend to the boundary).
Let solve (1) where is bounded, open and convex. Then
-
(1)
If then hence is closed and convex.
-
(2)
are a union of equivalence classes on which is affine and each such equivalence class intersects the boundary .
-
(3)
is an open set on which solves .
It is immediate from the definition that for are a union of nontrivial equivalence classes; the key conclusion is these extend to the boundary (i.e. if for then ). The PDE in point (3) has been considered in more generality by Rochet and Choné [50]. Note the economic interpretation of (1): no bunches of positive measure are sold apart from the null product; as we shall see during the proof of (2), any product sold to more than one consumer lies on the boundary of the set of products sold. Thus the entire interior of the set of products sold consists of individually customized products. Our proof of Theorem 1.1 and some subsequent results requires a new proposition asserting that a.e. on the boundary of a convex domain , the minimizer of (1) satisfies the boundary condition where is the outer unit normal to . Established in Proposition 2.3, it can be interpreted to mean that the normal component of any boundary distortion in product selected can never be inward. Moreover, in convex polyhedral domains and certain other situations, we are able to extend the interior regularity of Caffarelli and Lions [9] [38] to the smooth parts of the fixed boundary, Theorem 4.1.
The remainder of our results are restricted to the planar case . Theorem 1.1 provides a complete description of the solution in and : it remains to better understand the behavior of the solution in as well as the properties of the domains and , (noting is, by Theorem 1.1, a closed convex set).
By Theorem 1.1, the free boundary between and consists only of points in rays which foliate and extend to . In §5 we prove that the Neumann condition
(5) |
where is the outer unit normal to the fixed boundary , can be used to characterize the presence of these rays. Namely if (5) is not satisfied at then . Conversely, if lies in a boundary neighbourhood on which the Neumann condition (5) is satisfied then is a point of strict convexity for . The remaining case — rays which satisfy (5) — is subtle: we call such rays stray and conjecture the union of stray rays has zero area in general, but are able to show it is empty only for the squares .
Let be a point in the free boundary which, necessarily, lies on the ray . Let be the boundary endpoint of . Note that provided is in a neighbourhood of and , the same inequality holds for all and thus such are also the boundary endpoints of nontrivial rays. In this case we call a tame free boundary point (and a tame ray); we denote the set of tame free boundary points by .
Theorem 1.2 (Regularity results for the free boundary).
Let solve (1) where is a bounded, open, and
convex with smooth boundary except possibly at finitely many points. For every
and there is such that
(1) has Hausdorff dimension less than in ;
(2) the function is continuous on , where denotes the accumulation points of ’s local maxima;
(3) is Lipschitz, except at those
local maxima of where the Lebesgue density of happens to vanish;
(4) if is Lipschitz on near , then a bootstrapping procedure
yields is a curve and .
Since we don’t know whether local maxima of can be dense, we augment (2) by proving the function is continuous -almost everywhere on the fixed boundary in Theorem 7.2. Establishing the Lipschitz regularity of , which permits the above-mentioned bootstrapping to a free boundary, remains an interesting open problem.
Remark 1.3 (Lipschitzianity, convexity, and smoothness).
Note the Lipschitz requirement on from (4) is not necessarily satisfied even when the corresponding portion of lies in a Lipschitz submanifold given by (3). If happens to be convex this distinction disappears for every smaller value of ; simulations [44] suggest this occurs in the square examples from Theorem 1.4. Thus if the region depicted in Figure 1 is convex, as Mirebeau’s simulations lead us to conjecture, Theorem 1.2 guarantees the curved portion of its boundary is smooth (away from ).
Theorem 1.2 is proved using new coordinates for the problem, new Euler–Lagrange equations, and a new observation: That in a neighbourhood of a tame free boundary point the difference between the minimizer and the minimal convex extension of solves the classical obstacle problem. A priori, the obstacle is , i.e. has a merely Laplacian, and thus, without first improving the regularity of , the above results are the best one can obtain from the theory of the obstacle problem.
For general convex domains it is difficult to study the structure of the stray set which may include points in the relative interior of rays. We also have not ruled out, in general, that the relative boundary of
in might have positive -measure and the corresponding free boundary be nonsmooth. However, in specific cases, a more complete description is possible. For example, on the squares , we show is a single connected component of . In fact, we are able to provide the first explicit and complete description of the solution on , including an unexpected trichotomy for , sufficiently small, and sufficiently large. To describe the solution we label the edges of by their compass direction and set
The minimizer is described by the following bifurcation theorem (visualized in Figure 1).
Theorem 1.4 (Blunt bunching is a symptom of a seller’s market).
Let solve (1) with where . Then
-
(1)
is a convex set which includes a neighbourhood of in .
-
(2)
The portion of consisting of rays having both endpoints on the boundary is connected and denoted by . It is nonempty when . All rays in are orthogonal to the diagonal with one endpoint on and the other on . On the other hand there is such that is empty when .
- (3)
-
(4)
The set of strict convexity of contains and the Neumann condition (5) holds at each apart from the 3 vertices.
The following corollary may be of purely mathematical interest from the point of view of calculus of variations and partial differential equations; the smoothness asserted follows from the results of [10, 8].
Corollary 1.5 (Convexity of solution to, and contact set for, an obstacle problem).
For , the minimizer of over non-negative functions is convex. Its zero set is smooth, convex, has positive area, and is compactly contained in the centered square .
Remark 1.6 (Concave nondecreasing profile of stingray’s tail).
Numerical simulations of the square example show the region of products consumed to be shaped like a stingray, e.g. Figure 1 of [19]. Theorem 1.4 combines with Lemma 8.4 below to provide a rigorous proof that the lower edge of stingray is concave non-decreasing — as the simulations suggest — while Theorem 1.1 shows that every product selected by more than one type of consumer lies on this boundary or its mirror image across the diagonal.
Remark 1.7 (Absence and ordering of blunt vs targeted bunching).
The potential absence of blunt bunching from the square example — established on a nonempty interval by the preceding theorem — has been overlooked in all previous investigations that we are aware of. It can be understood as a symptom of a buyer’s market, in which a lack of enthusiasm on the part of qualified buyers incentivizes the monopolist to sell to fewer buyers but cater more to the tastes of those who do buy. The persistence of targeted bunching for all reflects the need to transition continuously from vanishing Neumann condition (5) — satisfied on the customization region where is strictly convex — to the uniformly positive Neumann condition on the exclusion region where vanishes, in light of the known regularity [15][50]. Such bunching is neither needed nor present when : in this case on shows the Neumann conditions in and coincide. When the blunt bunching region is present, our proof of Theorem 1.4 shows it separates from , which in turn separate all but one point of from . In particular, blunt bunching implies is a triangle, which is exceedingly rare in its absence.
Assuming is Lipschitz (or at least has finite perimeter), we arrive at a characterization of the solution to (1) on for every value of . Namely minimizes (1) if and only if (A) bunching is absent (, as for ), in which case solves , i.e the classical obstacle problem [49, 29, 27] and on (Figure 1(a)), or (B) bunching is present but blunt bunching is absent, (, as for ) in which case we derive below necessary conditions, whose sufficiency can be confirmed as in [41] (Figure 1(b)), or (C) blunt bunching is present, (, as for ) (Figure 1(c)), in which case the sufficient conditions for a minimum established by two of us [41] are also shown to be necessary below. (The only gap separating the necessary from the sufficient condition is the question of whether or not must have finite perimeter.)
If instead (B) blunt bunching is absent but bunching is present, , Theorem 1.4 asserts that splits into two connected components
(6) |
placed symmetrically below and above the diagonal. The region and its reflection below the diagonal are foliated by isochoice segments making continuously varying angles with the horizontal. The limit of these segments is a segment of length lying on the boundary of the convex set , having endpoint and making angle with the horizontal.
Fix any closed convex neighbourhood of in which is reflection symmetric around the diagonal and contains such a segment in its boundary. We describe the solution in using an Euler-Lagrange equation (9) from [41], rederived more simply in Section 8 below. Index each isochoice segment in by its angle ; (angles which are less than or non-negative are ruled out in the proof of Theorem 1.4). Let denote its left-hand endpoint and parameterize the segment by distance to this boundary point . Along the hypothesized length of this segment assume increases linearly with slope and offset :
(7) |
Given the initial (angle, height) pair , and piecewise Lipschitz with , solve
(8) |
(9) |
then set
(10) | |||||
(11) |
Given and as above, the triple satisfying (9)–(11) exists and is unique on the interval where . Thus the shape of and the value of on it will be uniquely determined by and . We henceforth restrict our attention to choices of and for which the resulting set lies above the diagonal. In this case and the value of on are determined by reflection symmetry across the diagonal. This defines on and provides the boundary data on needed for the mixed Dirichlet / Neumann boundary value problem for Poisson’s equation,
(12) |
which determines on . The duality discovered in [41], implies that for at most one choice of and Lipschitz can convex (pieced together from as above with on ) satisfy the supplemental Neumann conditions
(13) | ||||
(14) |
required on the free boundaries (since [15][50]); here denotes the outer unit normal to at . In the course of proving Theorem 1.4 in Section 8 below we complete this circle of ideas — apart from the piecewise Lipschitz hypothesis which Theorem 1.2 falls just short of proving — by showing at least one such choice exists; this choice uniquely solves (1) on the square in case (B). In case (C), Theorem 1.4 shows at least (and therefore exactly [41]) one choice exists satisfying the free boundary problem from [40] in the analogous sense.
We conclude this introduction by outlining the structure of the paper. Section 2 contains preliminaries: the variational inequality associated to (1), some background on Alexandrov second derivatives, and localization results of Rochet–Choné. In Section 3 we prove Theorem 1.1 using perturbation techniques previously used to study the Monge–Ampère equation. In Sections 4 and 5 and we prove some technical results which facilitate our later work in Sections 6 and 7. First, in Section 4, a boundary result which is new for this problem and extends the interior regularity result of Caffarelli and Lions [9] [38]. Next, in Section 5, propositions quantifying how at points of nonstrict convexity the Neumann boundary condition fails to be satisfied. Section 6 and Section 7 establish Theorem 1.2 using techniques from the study of the obstacle problem. Here we indicate a new connection to the classical obstacle problem. Namely, that the Monopolist’s problem gives rise to an obstacle problem where the obstacle is the minimal convex extension of the function defined on . The proof of Theorem 1.4 is completed in Section 8 using a case by case analysis based on a careful choice of coordinates. It confirms the economic intuition that the degree to which product selection (hence bunching) is influenced by the market presence of competing consumers decreases as we move away from the exclusion region, i.e. from the lower left toward the upper right region of the square, while on the other hand, increasing as we move the entire square of consumer types away from the outside option by increasing . We conclude with appendices containing some relevant background results. Table 1 contains a list of notation.
Notation | Meaning |
---|---|
A bounded open convex subset of . | |
The closure of . | |
The interior of . | |
Set complement of . | |
Outer unit normal at a point where is differentiable. | |
Bunch . | |
The relative interior of the convex set . | |
Subset (4) of foliated by ()-dimensional bunches. | |
Compact containment. | |
The positive part of a function, . | |
The positive part of a measure . | |
The support of , i.e. . | |
The set of Borel probability measures on . | |
Measure restriction: . | |
-dimensional Hausdorff measure. | |
-dimensional Lebesgue measure, i.e. volume measure. | |
Surface area measure (or arclength in special case ). |
2. Variational inequalities and Alexandrov second derivatives
2.1. Variational inequalities
Our basic tools for studying the unique minimizer of the functional
(15) |
over
are the variational inequalities stated in the following lemma.
Lemma 2.1 (Variational inequalities).
Proof.
We begin with (16). Let be the minimizer and observe is convex. Thus for any , and we have
so, in particular,
(19) |
Note with so we may apply the divergence theorem and obtain
where is the outer unit normal to which exists -a.e. for the convex domain .
Remark 2.2.
(1) It is straightforward to see, again by arguing using a perturbation, that inequality (17) holds not just for but for any convex with (the support of the negative part of ) disjoint from the set . The key observation is that for sufficiently small , .
(2) In any neighbourhood where is and uniformly convex, that is
satisfies an estimate , one may perturb — as is standard in the calculus of variations — by smooth compactly supported functions and obtain in the interior and on the fixed boundary of .
Without a local uniform convexity estimate, even for smooth functions , it may be that there is no small enough to ensure is convex.
Inequality (18) is useful when one chooses as paraboloid with prescribed Laplacian. We give an example now — the result we prove is required in subsequent sections. It is interesting to contrast the following result with the one-dimensional case, in which minimizers on domains satisfy .
Proposition 2.3 (Normal distortion is not inward).
Let solve (1) where is bounded, open, and convex. Then for any where the outer normal is defined,
Proof.
By approximation it suffices to prove the result for smooth strictly convex domains. Indeed, [30, Corollary 4.7] and its proof imply if is a sequence of smooth strictly convex approximating domains and is the solution of (1) on then at every where and each is differentiable. The result of [50, 15] implies this is every .
Thus we take to be smooth and strictly convex. Up to a choice of coordinates we assume and (see Figure 2). Recall is the affine support at . For sufficiently small and we consider the family of admissible perturbations (see Figure 2)
Note has positive measure since at , . Moreover if then and thus . Thus, strict convexity of implies in Hausdorff distance [52, §1.8] as . Clearly on . To derive a contradiction assume . Then for sufficiently small we also have
(20) |
which holds by the continuity of and because . But (18) with implies
which contradicts (20) to conclude the proof.
∎
An essential tool is that the variational inequality (17) holds not only on but restricted to the contact set — at least for almost every . This novel and powerful technique was pioneered in this context by Rochet and Chonè [50] who exploited the sweeping theory of measures in convex order. In this section we recall the statement of their localization result; for completeness we include a proof in Appendix A.
We introduce the following notation for the variational derivative of our objective :
(21) |
which turns out to be a measure with finite total variation. The equivalence relation induced by , namely if and only if yields the partitioning of each into leaves. We let denote the equivalence class of and can disintegrate by conditioning on this equivalence relation. Let the conditional measures be defined by disintegrating separately the positive and negative parts of with respect to the given equivalence relation, we recall how
for all convex functions in Corollary A.9 of Appendix A below.
2.2. Legendre transforms and Alexandrov second derivatives
Recall if is a convex function, then its Legendre transform is defined by
(22) |
A function is called Alexandrov second differentiable at , with Alexandrov Hessian (an matrix), provided as that
Alexandrov proved convex functions are twice differentiable in this sense almost everywhere.
It’s well known that if a differentiable convex function is Alexandrov differentiable at and its Legendre transform is Alexandrov second differentiable at then
We have an analogous result even when Alexandrov second differentiability is not assumed.
Lemma 2.4 (Legendre transform of Hessian bounds).
Assume is a convex function with Legendre transform , that and is an invertible symmetric positive definite matrix. Assume . Then
if and only if
Proof.
We prove the “only if” statement; the “if” statement is proved similarly. Up to a choice of coordinates and subtracting an affine support we may assume and . Whereby we’re assuming
(23) |
It is straightforward to see that (23) holds if and only if for every there is a neighbourhood of on which
(24) |
Now, for we have
Provided lies in the possibly smaller neighbourhood the supremum is obtained at and
The aforementioned equivalence between (24) and (23) gives the desired result. ∎
3. Partition into foliations by leaves that intersect the boundary
In this section we present the proof of Theorem 1.1. We use localization (Corollary A.9) to obtain the vanishing Neumann condition throughout , but apart from that the only technique we use is energy comparison, and the variational inequalities (16) – (17) coupled with careful choice of the comparison functions, many of which are inspired by those used to study the Monge–Ampère equation [28, 36]. We denote a subsection to each point of Theorem 1.1.
3.1. Point 1: if nonempty.
We shall only prove ; equality follows easily if is nonempty, which is known to be true on strictly convex domains [2].
Proof of Theorem 1.1 (1).
For a contradiction we assume there is with . Applying Rochet–Chonè’s localization with we have
(25) |
We plan to show equality holds. By assumption on . Since is convex and we may apply Corollary A.9 with and obtain
(26) |
Using on , (25) and (26) imply
(27) |
We know and on (Proposition 2.3) and . Therefore (27) implies on and this contradiction completes the proof. ∎
3.2. Point 3 of Theorem 1.1
Now we present point 3: that in and that is open. One can immediately obtain a.e. in via Rochet–Choné’s localization (Corollary A.9). However for point 3 of Theorem 1.1 we require in addition, an inequality for at all points where . Thus we prove the following lemma directly using perturbations.
Lemma 3.1 (Sub- and super-Poisson for interior vs. customized consumption).
Assume solves (1) and is a point of Alexandrov second differentiability satisfying . Then
-
(1)
There holds .
-
(2)
If, in addition is strictly convex at , then .
Proof.
(1) We take with assumed to be a point of Alexandrov second differentiability. For convenience translate and subtract the affine support at so that and all vanish.
For a contradiction assume and take satisfying . There is a neighborhood of on which .
Let denote the Legendre transform (22) of and set . Lemma 2.4 implies in a punctured neighborhood of the origin. Thus as the connected component of containing the origin, which we denote , converges to in the Hausdorff distance. Set
(28) |
Let be the Legendre transform of . Note and this inequality is strict at . Moreover because at each Lemma 2.4 implies on the set . This contradicts inequality (18) to establish (1), where for small enough has been used.
(2) Now suppose, in addition, is a point of strict convexity for and, for a contradiction, that . Set with chosen so small that . Note that in a punctured neighborhood of ; (this relies on the strict convexity of at in the case has a zero eigenvalue). Thus for sufficiently small the connected component of containing , which we call , converges to in the Hausdorff distance. Set
Then is an admissible interior perturbation of with on . Once again we contradict inequality (18). ∎
At any interior point of strict convexity, , we have since is closed. It’s now immediate that at each point of Alexandrov second differentiability in . It remains to show is open. We prove in the next subsection that , that is if then is in the boundary of the set of gradients. Combined with the regularity of we obtain if the same is true for all sufficiently close (this is because is also in ) and this is a sufficient condition for . We conclude is open.
Since we now know is a (equivalently, ) solution of almost everywhere on the open set the elliptic regularity [31, Theorem 9.19] implies .
3.3. Point 2 of Theorem 1.1
To conclude the proof of Theorem 1.1 we show if for then extends to the boundary.
For a contradiction assume otherwise. Because is convex, then there exists with and since is closed. Because is and the sections
converge to in the Hausdorff distance as , we obtain for sufficiently small. In particular by Lemma 3.1 we have in . Using as a perturbation function in inequality (16) we see almost everywhere in (if on a subset of with positive measure, inequality (16) is violated). As in the previous subsection the elliptic regularity implies is a classical solution of in . Differentiating this PDE twice implies the second derivatives of are harmonic (and nonnegative by convexity). The strong maximum principle for harmonic functions says in fact in for all , so cannot be affine on . This contradiction completes the proof. Note our use of the strong maximum principle requires at some point in . This, however, follows by considering . If throughout then , and is independent of , hence , which would contradict as .
In the course of the above proof we’ve proved the following lemma which we record here since we require it again and again.
Lemma 3.2 (Interior regularity and strong maximum principle).
Assume is a convex function. Let . Assume, in the sense of Alexandrov second derivatives, almost everywhere in . Then and satisfies that for each unit vector either throughout or throughout .
4. Product selection remains Lipschitz up to the boundary
One of our key techniques for studying the planar Monopolist’s problem is the introduction of a new coordinate system defined in terms of the rays which foliate . These new coordinates are a powerful tool for studying the behaviour of the minimizer but their justification requires two significant technicalities. The first is a boundary regularity result in arbitrary dimensions, namely that in convex polyhedral domains is up to the boundary (away from the nondifferentiabilities of the boundary); furthermore, in smooth convex domains is on the set of rays having only one end on the boundary (Theorem 4.1). The second required technicality, proved in Section 5, is an equivalence between the Neumann condition and strict convexity stated more precisely in Propositions 5.3 and 5.4. Readers who are interested primarily in the consequences of these technicalities rather than their proof may proceed directly to Section 6.
Boundary regularity beyond [15][50] for the Monopolist’s problem is new. Previously only an interior result was known [9, 38] and regularity is known to be sharp.
Theorem 4.1 (Boundary regularity on convex polyhedral domains).
Let minimize (1) where is open, bounded, and convex.
-
(1)
There is depending only on such that if is a point of Alexandrov second differentiability and a singleton or empty then
-
(2)
Assume, in addition, is a convex polyhedron (i.e. an intersection of finitely many half spaces). Let be excluding an -ball about each point where is not smooth. Then there is depending only on and such that
Proof.
The key energy comparison ideas are inspired by Caffarelli and Lions’s proof of interior regularity [9] and its generalization [38]. However new ideas are required for perturbation near the boundary. We prove there exists depending only on and such that for all there holds
(29) |
Equivalently, there is some such that for all there holds . We emphasize that will be chosen small depending on quantities which the constant in (29) is not permitted to depend on, however this does not affect the estimate111This is analogous to an estimate for all sufficiently small implying regardless of what dictates our small choice of . It is interesting to note such an approach would not work for boundary Hölder or estimates with . .
We begin by explaining the proof for part (2), that is, when is a polyhedron and then explain the changes required for part (1) namely, points on rays having an end in the interior of . Note the result for is a straightforward consequence of the convexity of and in .
Step 1. (Construction of section and comparison function on polyhedrons) We fix and translate and subtract a support plane after which we may assume and all vanish and thus, 222It is worth noting that is not translation invariant, so after this transformation we should work with . Inspection of the proof reveals such a change is inconsequential.. Now, after a rotation we may assume the face closest to is
We assume . (If then we already have a estimate; Caffarelli and Lions’s estimate is .) Note that may be close to a single face of the polyhedron but, because we work in , satisfies
(30) |
for a positive constant depending only on and .
For to be chosen sufficiently small, but initially , set
where the latter equality defines the unit vector as the direction in which the supremum is obtained. The section
satisfies the slab containment condition
(31) |
The lower estimate is because when and . For the upper estimate note
is the outer unit normal to at . However, because attains its maximum over the boundary of the ball at , has zero tangential component and so, by convexity, for some meaning the outer normal is
Step 2. (Tilting and shifting at the boundary on polyhedrons) The possibility that intersects complicates the boundary estimate. The existing interior estimates use a bound which does not suffice near the boundary. Thus we must tilt the affine support to ensure points where intersects lie sufficiently far (distance greater than ) from .
We consider the modified plane and section (see Figure 3)
(32) |
where is the outer unit normal to along and is a small positive or negative shift to be specified. The key idea is that provided , so is a small perturbation of (in both direction and magnitude). Observe that slab containment, (31), implies and on , we have , so provided (which we enforce below), is disjoint from . We claim if is initially chosen sufficiently small, then
(33) |
where the fact that is bounded has been used.
First we prove the lower bound . This follows because a choice of sufficiently small ensures the plane makes an arbitrarily small angle with the original plane . Moreover we can ensure for some . Indeed, the plane , which we used as our original lower bound for the slab , is necessarily orthogonal to . Similarly, the plane , which we use as our lower bound for , is orthogonal to
provided the vectors and make arbitrarily small angle. Thus the planes and make arbitrarily small angle. Since and , provided and there is a point on where .
Next we prove the upper bound . This is where we choose our vertical shift. Recall
where . Note that
Because the choice gives and we have equality of and at , i.e. . Then, as before, is a normal to a support of the convex set . However
Recalling we see again for sufficiently small this vector makes arbitrarily small angle with . This yields the upper containment in (33).
Our choice of tilted support implies is disjoint from . Moreover we have
(34) |
where the second inequality is because each point in the containment slab is of distance less than from the plane and . These properties are enough for us to employ the dilation argument used by the authors in [38]. For completeness we include full details, but first explain how to obtain the section in the other setting of the theorem: in arbitrary convex domains at rays with one endpoint on the boundary.
Step 3. (Rays with one endpoint on the boundary in convex domains) Now we explain the choice of a suitable perturbation for case (1) of Theorem 4.1, namely when is merely open, bounded, and convex and satisfies that has only one point on the boundary. In this case a similar tilting procedure yields a section which is in fact strictly contained in . Assume, where . Note, if there is such that for all sufficiently small, (i.e. the angle between and is bounded away from ) then the slab containment (31) yields . On the other hand, if (i.e. approaches the orthogonal direction to ) we consider the new perturbation
Provided , which we can enforce as above, we have for sufficiently small . Since, in addition, as , we have that for sufficiently small (obtained by an initial choice of sufficiently small) will be strictly contained in and satisfy both the slab containment condition (33) and the height estimates (34). Note in the steps that follow the argument is simpler for this case. Indeed, implies we do not need to consider the boundary terms in what follows or the dilation argument in Step 5.
Step 4. Initial estimates. Now we use the minimality of the function for the functional (defined in (1)) to derive the desired inequality . Set where is defined in (32). Minimality implies
(35) |
The divergence theorem implies
(36) |
Next, using that is linear in (in particular, ), we compute
(37) |
Substituting (36) and (37) into (35) we have for depending on
(38) |
Step 5. (Final estimates) To complete the proof we prove for , which in the case of polyhedra depends in particular on , there holds
(39) | ||||
(40) |
For the first, in the case of polyhedral domains, recall is empty so is of distance from . Thus the estimate (39), for depending on , is standard in convex geometry and may be proved either as in the work of Chen [22], or the authors [38]. In case (1) of the theorem the estimate is trivial because
Now we obtain (40). We let denote the dilation of by a factor of with respect to . What is again crucial is that so for defined as in (30), consists of interior points of on which . It is helpful now to choose coordinates such that . For , let be the projection onto . For each the set is two disjoint line segments. We let be the line segment with greater component and write where .
Choose in case (2) which is bounded below by a positive constant depending on and . Case (1) is simpler as this dilation is not required. Note that each line segment for has at the upper endpoint. This is because from the slab containment condition the upper endpoint lies distance less than from whereas lies distance at least from . Clearly on we have .
5. Strict convexity implies the Neumann condition
In this section we continue establishing technical conditions required for the coordinates introduced in Section 6. For planar domains we prove the equivalence between the Neumann condition and strict convexity stated precisely in Propositions 5.3 and 5.4. We begin with two lemmas concerning convex functions in . The first states the upper semicontinuity of the function and the second yields the convexity of when restricted to a contact set .
Lemma 5.1 (Upper semicontinuity of leaf diameter).
Let be a bounded open convex subset of and a convex function. Then the function is upper semicontinuous.
Proof.
We fix a sequence converging to some and note it suffices to prove that
To this end, let and take , in realizing
The convergence properties of the subdifferential of a convex function imply and we may assume that, up to a subsequence, for . Thus we may send in the identity
(46) |
to obtain that for we have and thus has diameter greater than or equal to
∎
Let denote the relative interior of the convex set .
Lemma 5.2 (Existence a.e. of on a leaf implies convexity of ).
Let be a differentiable convex function defined on an open convex subset . Fix any . If , where denotes the set of second differentiability of , then and is a convex function for each .
Proof.
Fix satisfying . After subtracting the support at , we may assume and that is not a singleton (since otherwise the result holds trivially). We will show is convex along any of those line segments contained in for which Alexandrov second differentiability holds a.e. . To this end fix along such a segment. Choose orthonormal coordinates such that for some . Then since is affine on and we have a.e. on .
Next, for and any , convexity of implies for sufficiently small
(47) |
Here, by sufficiently small we mean small enough to ensure the above arguments of are contained in . The definition of Alexandrov second differentiability along with on implies
for a.e. and similarly at and . Thus (47) becomes
Dividing by and sending yields that is the restriction of a convex function to the segment — hence continuous on . The polarization identity implies the continuity of mixed second order partial derivatives. It follows that . ∎
Proposition 5.3 (No normal distortion nearby implies strict convexity).
Let minimize (1) where is open and convex. Let be a point where and . Assume there is with
Then , that is is strictly convex at .
Proof.
Because and Lemma 5.1 implies the upper semicontinuity of , we may find a possibly smaller such that and for each . Rochet and Choné’s localization (Corollary A.9) with boundary term implies almost everywhere on
To see this, note, by Lemma 5.2, restricted to any for which is a convex function. Thus is a permissible test function in the localization Corollary A.9 from which we obtain
(48) |
where denotes the disintegration of the Lebesgue measure with respect to the contact sets (an explanation of disintegration is provided in Appendix A). Inequality (48) implies that almost everywhere on there holds and thus the same equality holds on .
Let be the interior endpoint of (if then we are already done). In a sufficiently small ball , we have just shown a.e. in . Moreover, Theorem 1.1 implies in , which is nonempty. Our usual maximum principle argument, Lemma 3.2, implies is strictly convex inside , contradicting that is the endpoint of a ray. ∎
The next proof requires Lemma A.6 which gives the pushforward of the variational derivative — and is proved in Appendix A by combining the neutrality implied by localization away from the excluded region with the fact that our objective responds proportionately to a uniform increase in indirect utility. We use these to estimate the following:
Proposition 5.4 (One-ended ray lengths bound normal distortion).
Let solve (1) where is open and convex. Let with smooth in a neighbourhood of and . Set . Then
(49) |
where depends only on a bound for in a neighbourhood of .
Proof.
The lower bound (49) was established in Proposition 2.3. We first prove (49) assuming nontrivial and at the conclusion of the proof explain why it holds for all in the theorem. Let and let be locally represented by a smooth curve with an arc length parametrization traversing in the anticlockwise direction with and without loss of generality .
The upper semicontinuity of from Lemma 5.1 implies
On the other hand we know from Lemma 3.2 that no subinterval of can be exposed to by which we mean there is no and with . Note in two-dimensions, if satisfies and then in the Hausdorff distance (and such sequences can be found).
As a result there exists sufficiently small such that
-
(1)
The leaves and have length at least .
-
(2)
The leaves and can be chosen to make fixed but arbitrarily small angle with , by e.g. [11, Lemma 16].
-
(3)
All leaves intersecting the boundary in have length less than (this holds by the upper semicontinuity of ).
Moreover can be taken as close to as desired. With the smoothness of , our two-dimensional setting, and the fact that leaves cannot intersect, this significantly constrains the geometry of
The set is strictly contained in a set with left edge and a vertical right edge of lengths bounded by , and top and bottom side lengths bounded by (see Figure 4). Finally we note we can choose sequences satisfying the above requirements and .
Now, by Lemma A.6, . That is,
(50) |
Using the boundary estimate from Theorem 4.1 (proved in Section 4) near 333Because doesn’t intersect at both endpoints and is upper semicontinuous the same is true for all sufficiently close leaves., the constrained geometry of , and nonnegativity of already established, we see (50) implies
which is precisely the desired estimate. Indeed, employing this estimate with in place of , dividing by , and sending we obtain (49) (after dividing we have an average and is continuous).
Now, we explain how to obtain the estimate when is trivial. If is such that is trivial and there is a sequence of nontrivial leaves then the estimate follows by the upper semicontinuity of . If there is no such sequence then lies in a relatively open subset of the boundary on which is strictly convex. Using the nonnegativity of and Corollary A.9 with applied to yields on . ∎
6. Leafwise coordinates parameterizing bunches in the plane
In this section and the next we study the behavior of the minimizer on and the free boundary in two-dimensions. We introduce one of our main tools: a coordinate system to study the problem on which is flexible enough to include the coordinates proposed earlier by the first and third authors [41], and for which we are finally able to provide a rigorous foundation by proving biLipschitz equivalence to Cartesian coordinates. Moreover, by combining these coordinates with Rochet and Choné’s localization technique (Corollary A.9) we are able to provide a radically simpler derivation of the Euler-Lagrange equations (9)–(10) first expressed in [41]; c.f. (67)–(68) and (80)–(81) below,
Let be a curve parameterizing in the clockwise direction with and write .
First we give conditions to ensure a neighbourhood of a ray is foliated by rays.
Lemma 6.1 (Local foliation around each tame ray).
Let be smooth in a neighbourhood of satisfying , and . Then there exist such that and for all .
Proof.
We assume without loss of generality that and thus there is such that is a smooth curve. Lemma 5.4 implies is nontrivial. For a possibly smaller and all , the regularity of implies is nontrivial, with length bounded below by some determined by (49). Furthermore, the upper semicontinuity of (Lemma 5.1) implies for a possibly smaller , for each . ∎
Recall is a tame point of the free boundary provided for an satisfying the hypothesis of Lemma 6.1.
We define as the unit direction vector of the leaf pointing into . This means, with ,
and subsequently we can write a subset of the connected component of containing as
(51) | ||||
(52) |
and we take as new coordinates for . Because each ray is a contact set along which is affine there exists functions such that
(53) |
and is independent of .
Our goal is to derive the Euler–Lagrange equations of Lemma 6.3 below which describe the equations the minimizer satisfies in terms of , , and . First, we record the key structural equalities for the new coordinates in the following lemma, which holds under the biLipschitz hypothesis we eventually establish in Corollary 6.6. The quantities (55)–(57) from this lemma also yield a formula for the Laplacian of :
(54) |
Lemma 6.2 (Gradient and Hessian of in coordinates along tame rays).
Suppose solves (1) where is bounded, open and convex. Let on be as above (52). If the transformation is biLipschitz on , then its Jacobian determinant is positive and given by
(55) |
where and similarly are evaluated at . In addition the following formulas for the gradient and entries of the Hessian of (53) hold -a.e.:
(56) | ||||
(57) |
Proof.
Where the transformation and its inverse are Lipschitz, their Jacobian derivatives are easily compute to be:
(58) | ||||
(59) |
with from (55). Next, to obtain the gradient expressions we differentiate equation (53) with respect to to obtain
(60) |
Similarly, differentiating (53) with respect to and equating coefficients of yields
(61) | ||||
(62) |
where and that is independent from has been used. We solve (60) and (62) for , and obtain (56). Note the functions and in (56) are Lipschitz because . Thus, differentiating the expressions for and given by (56) with respect to, respectively, and and using the Jacobian (59) gives the formula (57).
We note two facts about the functions and which determine the Jacobian determinant. First
(63) |
where the nonnegativity is by our chosen orientation: traverses in a clockwise direction and points into the convex domain . Inequality (63) is strict because is nontrivial and has only one endpoint on for each . Next, because is a unit vector, whence is orthogonal to , we have
(64) |
where the value of is determined by the sign of . From our biLipschitz hypothesis (or nonnegativity of the Laplacian (54)) the sign of is independent of . Combined with (63) we obtain for and . ∎
Now we combine our raywise coordinates with Rochet–Choné’s localization (Corollary A.9). We obtain the following Euler–Lagrange equations which are central to the remainder of our work.
Lemma 6.3 (Poisson data along tame rays).
Proof.
We compute the disintegration (110) and obtain for -a.e. that
The terms outside the integral are evaluated at . We consider four choices of test functions, and . Using these in the localization formula we obtain the equalities
(67) | |||
(68) |
These combine to imply
(69) |
We recall (64) and note (69) determines the sign . Thus the Euler–Lagrange equation (65) for the free boundary holds. Equation (65) implies is bounded away from zero and from above depending only on estimates for the continuous function and our previously given estimate (Lemma 6.1).
Remark 6.4 (Tame rays must spread as they leave the boundary).
Recalling that in , we see (66) quantifies how Poisson’s equation fails to be satisfied along leaves (with the equation only satisfied at the point ). It also implies that when we move from into the Laplacian jumps discontinuously across their common boundary.
In Section 7 we show this discontinuity yields quadratic separation of from its contact sets and exploit this to obtain estimates on the Hausdorff dimension of .
Let us conclude by justifying the aforementioned Lipschitz continuity of . Note both (66) and (65) yield Lipschitz estimates. However, since their derivation assumed was Lipschitz we need to redo these calculations with a perturbed, Lipschitz, and obtain uniform estimates as the perturbation parameter approaches . Notice the Lipschitz constant from the following lemma does not depend on the Neumann values .
Lemma 6.5 (Tame ray directions are Lipschitz on the fixed boundary).
With defined as above, the function is Lipschitz on the interval provided by Lemma 6.1, with Lipschitz constant depending only on an upper bound for and a lower bound on .
Proof.
Assume a collection of non-intersecting rays foliate an open set in and pierce a smooth curve . Provided the intersection of each ray with the curve occurs some fixed distance from either endpoint of the ray, the assertion of Caffarelli, Feldman, and McCann [11, Lemma 16] says the directions (of the ray passing through ) is a locally Lipschitz function with Lipschitz constant depending on and .
Thus, in our setting, if we could extend each ray by length outside the domain, would be locally Lipschitz with a constant depending on . Then, once we’ve obtained (66) the Lipschitz constant of is independent of . Apriori such an extension outside the domain may not be possible. Thus our strategy below will be to translate the boundary distance and use the translated boundary to redefine the axis of the coordinates. In this setting the corresponding direction vector is locally Lipschitz, the above calculations are justified, and sending gives the Lipschitz estimate on .
Thus, with as above let be the outer unit normal at and set
Because the length of rays intersecting is bounded below by , up to a smaller choice of we may assume for each that and lies distance at least from each endpoint of . Let denote the unit vector parallel to , where by [11, Lemma 16] is a locally Lipschitz function of . We may redefine the coordinates and write a connected component of containing as
where is defined so that is the point where the ray intersects . The function is locally Lipschitz because the curves are smooth and is Lipschitz (though we don’t assert that the Lipschitz constant of is independent of ). Thus all our earlier computations may be repeated in these new coordinates. The computations leading to (66) now yield the equation
(70) |
which we note satisfies when and thus agrees with our earlier coordinate system (in our modified coordinates is the point of distance from the endpoint of the ray).
Corollary 6.6 (Raywise coordinates (52) are biLipschitz).
Proof.
It suffices to estimate each of the entries in the Jacobians and computed earlier in (58) and (59). From (58) it’s clear that the entries of this Jacobian permit an estimate from above in terms of and , where the latter may, in turn, be estimated in terms of and thanks to (65). The only additional term we must estimate for the second Jacobian, that is , is . Since is decreasing in by Remark 6.4, it suffices to estimate and this is an immediate consequence of (55) which we recall says . ∎
7. On the regularity of the free boundary
In this section we study local properties of the free boundary by transformation to an obstacle problem and prove Theorem 1.2. If denotes the minimal convex extension of we show that solves an obstacle problem with the same free boundary as our original problem. Standard results for the obstacle problem then imply the free boundary has Lebesgue measure and, the stronger result, has Hausdorff dimension strictly less than .
We use these estimates to establish that the function from Section 6 is continuous almost everywhere on , Theorem 7.2. In Proposition 7.3 we describe a bootstrapping procedure which shows that if the free boundary — or rather the function — is Lipschitz, then it is In Theorem 7.5 and its corollary, we show has a Lipschitz graph away from its local maxima.
7.1. Transformation to the classical obstacle problem
Let denote the endpoint of a tame ray and as in Lemma 6.1. Using the coordinates from the previous section we consider a subset of ,
In Remark 6.4 we observed that rays spread out as they leave the boundary. Thus the coordinates remain well-defined on an extension of . We denote this extension by :
On we define the minimal convex extension of by the formula (53)
Note there is some such that . Moreover on we have, by (66), for depending only on a lower bound for and . Let and let denote the characteristic function of . Then in implies that
(71) | ||||
(72) | ||||
(73) |
where . Thus solves the classical obstacle problem in with the same free boundary as our original problem. The regularity theory for the obstacle problem yields estimates for the measure of the free boundary. What prevents us from using higher regularity theory for the obstacle problem is that on ,
may not be continuous — the minimum regularity required to apply typical regularity results for the obstacle problem is Hölder continuous. If, — i.e. the free boundary — were Lipschitz, would also be Lipschitz in which case one can bootstrap to regularity of ; see Proposition 7.3. As a partial result in this direction we prove that is continuous almost everywhere. We begin with Hausdorff dimension estimates for the free boundary.
Lemma 7.1 (Hausdorff dimension estimate for the free boundary).
Let be as given above, so that is the endpoint of a tame ray; c.f. Lemma 6.1. Then the Hausdorff dimension of equals for some depending only on , and .
Proof.
This is a standard result for the obstacle problem once one notes that in (71) satisfies on for depending only on and . We follow the clear exposition of Petrosyan, Shahgholian, and Uraltseva [49, §3.1, 3.2] to establish first quadratic detachment, then porosity.
Step 1. (Quadratic detachment at free boundary points) We claim if then
(74) |
Fix such an and ; we will eventually take . Set On the set we have . Thus, the maximum principle implies
Because on the supremum is attained at some on . Because we obtain
for some . We send to establish (74).
Step 2. (Nondegeneracy implies porosity) We recall a measurable set is called porous with porosity constant if for all and there is with
We prove that nondegeneracy, i.e. (74), and Caffarelli–Lions’s implies is porous. Take . Note (74) implies Indeed, with such that we have
Now, redefine as a point in where . Using that (where , gives the obvious estimate ), we have if for then
Since along this proves lies in . Thus we’ve established the porosity condition for balls centered on . To establish the porosity condition for any ball in we argue as follows. Let . We take , noting if no such exists we’re done. Our porosity result applied on gives porosity of .
Theorem 7.2 (Continuity of the free boundary a.e.).
Taking as in Lemma 6.1, the function defined by is continuous for almost every .
Proof.
Assume is not continuous at some . Because is upper semicontinuous this implies there exists a sequence with . We show each for lies in the free boundary Then, because Lemma 7.1 implies , Fubini’s theorem and Corollary 6.6 imply the union of all such has -measure .
To show for each , we suppose otherwise. Then since no such point can be interior to without violating Lemma 3.2, there is such that is an interior point of .
Upper semicontinuity of implies rays sufficiently close to have intersection with the boundary close to . More precisely for every there is such that
Thus, our planar foliation implies that because is an interior point of then is also an interior point of for each which contradicts that and completes the proof. ∎
Next we show prove the fourth point of Theorem 1.2: that if this continuity could be strengthened to Lipschitz continuity then one can bootstrap to a smooth free boundary and minimizer on .
Proposition 7.3 (Tame part of free boundary is smooth where Lipschitz).
Proof.
We prove by induction on that there is some neighbourhood on which the curve and the function are both for some , while the coordinate transformations and are in . For our assumption is that is Lipschitz, and from Corollary 6.6 the coordinate transformations are biLipschitz. From the formula (65), reproduced here for the readers convenience
(75) |
and Caffarelli and Lions (or Theorem 4.1) it follows that is also Lipschitz, hence the coordinate transformations improve to by the Jacobian expressions (58)–(59). From (66), i.e.
(76) |
we see , hence the regularity theory for Poisson’s equation implies when (one derivative more than needed).
Now assume the inductive hypothesis for some fixed . From (75)–(76) we again deduce has Laplacian in . Thus the regularity theory for Poisson’s equation implies . The regularity theory for the obstacle problem (where the obstacle has a Laplacian; (due to Caffarelli [10, 8], and Kinderlehrer [33] with Nirenberg [32], though the clearest statement we’ve found is by Blank [6, 5]) implies the free boundary is . Note to apply the classical regularity theory for the obstacle problem we are using that the Lipschitz regularity of implies the set has positive density at each boundary point; here as in (71) . Now that is the same is again true for by equation (75) since the smoothness established in propagates down the rays from the free to the fixed boundary using the coordinate transformations; these transformations then improve to by equations (58)–(59) so the induction is established and the proof is complete. ∎
7.2. Criteria for the tame part of the free boundary to be Lipschitz
Having deduced regularity of the free boundary when it is Lipschitz we now turn our attention to the question of characterising the set on which the free boundary is Lipschitz. We will rely on the well known Caffarelli dichotomy for the blow-up of solutions to the obstacle problem. We recall that blowing-up at the edge of the contact region in the classical obstacle problem (without convexity constraints) led Caffarelli to formulate his celebrated alternative [10, 8]: If satisfies
(77) |
then is convex and either a quadratic polynomial or a rotated translate of the half-parabola solution
At each point in the free boundary, the density of the contact region is therefore either (called singular) or (called regular); it cannot equal because of quadratic detachment (as in e.g. the proof of Lemma 7.1). Furthermore, the dichotomy holds for blowups of solution to equations of the form in a domain where is continuous in the following sense: Take and a sequence . Note that up to taking a sequence the limit
exists and is a globally defined solution of so that Caffarelli’s dichotomy applies to the function . Unfortunately, in our setting we only know and not the Hölder continuity required for higher regularity [33].
A real-valued function on an interval is called unimodal if it is monotone, or else if it attains its maximum on a (possibly degenerate interval) , with being non-decreasing throughout the connected component of to the left of , and non-increasing through the connected component to the right of . The following lemma shows lower semicontinuous functions are unimodal away from their local minima.
Lemma 7.4 (Lower semicontinuous unimodality away from local minima).
Let be lower semicontinuous on an interval . Let denote the subset of consisting of local minima for , and its closure. Then is unimodal on each connected component of .
Proof.
Fix any open interval . We claim is unimodal on . Since is lower semicontinuous but has no local minima on , for each it follows that is a countable union of open intervals on which with on . If there were more than one open interval in this union, say and with , then would attain a local minimum on the compact set , contradicting the fact . Thus the set consists of at most one open interval, which is monotone nonincreasing with . Let denote the infimum of for which is empty, and set . Then is non-decreasing to the left of , non-increasing to the right of , and — if is nonempty — attains its maximum value on . ∎
We apply this lemma to the diameter of the rays along the tame part of the free boundary to deduce the free boundary is Lipschitz away from its local maxima.
Theorem 7.5 (Tame free boundary is Lipschitz away from local maxima).
Let with for smoothly parameterize a fixed boundary interval throughout which the Neumann condition (5) is violated. Let denote the subset of consisting of local maxima for , and any connected component of , where is the closure of . Then extends continuously to and its graph is a Lipschitz submanifold of . Similarly, the graph of over is a Lipschitz submanifold of (except perhaps at ), and is continuous on .
Proof.
Corollary 6.6 shows the coordinates are locally biLipschitz on , so the final sentence follows from showing has a continuous extension to whose graph is a Lipschitz submanifold.
Proposition 5.4 asserts is upper semicontinuous on . Unless is monotone on , Lemma 7.4 shows decomposes into two subintervals on which is monotone and they overlap at least at one point . Although a monotone function need not be Lipschitz — or even continuous — its graph has Lipschitz constant at most . A discontinuity in on the closure of either of these subintervals would correspond to a line segment of length in the free boundary along which is affine. This contradicts the strong maximum principle (Lemma 3.2) after constructing the appropriate reflection of across this segment. Thus is continuous on . The graph of on is obviously Lipschitz, except perhaps when the minimum value of is uniquely attained at some . Since is a local minimum, is continuous at hence Caffarelli’s alternative holds for the blow-up at : the Lebesgue density of at cannot be zero since is a local minimum, so it must be exactly [8, 10]. The blow-up limits of at all coincide with the same half-parabola, and is differentiable at . The Lipschitz graph of to the left of shares the same tangent at as the Lipschitz graph of to the right of , which completes the proof. ∎
Our next corollary shows that the tame part of the free boundary can only fail to be locally Lipschitz when oscillations with unbounded frequency cause local maxima of to accumulate, or when an isolated local maximum forms a cusp. In the latter case, the tame free boundary is locally piecewise Lipschitz and the perimeter of is locally finite in this region.
Corollary 7.6 (Is the tame free boundary piecewise Lipschitz?).
Assume has only finitely many connected components in Theorem 7.5 and is constant on each of them — as when has only finitely many local maxima on . Then the graph of is a piecewise Lipschitz submanifold of for each . Moreover, if the graph of fails to be Lipschitz at for some , then has an isolated local maximum at and has Lebesgue density zero at .
Proof.
Under the hypotheses of Theorem 7.5, assume has only finitely many connected components and is constant on each of them. Then these components must be intervals which are relatively closed in : otherwise the upper semicontinuous function has a jump increase at the end of one of them, which leads to a segment in the graph of — producing the same contradiction to Lemma 3.2 as in the proof of Theorem 7.5. Thus . For each of the open intervals comprising , Theorem 7.5 already asserts that is continuous and has Lipschitz graph on ; the only question is whether the graph extends past each endpoint of in in a Lipschitz fashion. If the endpoint of belongs to a nondegenerate interval in this is obvious. When the endpoint of is an isolated point in , then Lemma 7.4 shows nearby is monotone on either side hence must be continuous at to avoid an affine segment in the graph of as before. Now Caffarelli’s alternative applies, so the density of at must be either or . In the latter case has a Lipschitz graph in a neighbourhood of , as in the proof of Theorem 7.5, hence the corollary is established. ∎
Remark 7.7 (A partial converse).
It remains to see whether accumulation points of local maxima of and/or cusps might be ruled out by combining quadratic detachment shown in Lemma 7.1 with estimates in the spirit of the following lemma.
Lemma 7.8 (A variant on Clarke’s implicit function theorem).
Let be a Lipschitz manifold whose topology is metrized by . Fix a Lipschitz function such that for each , has Lipschitz constant on . Assume there exists a set , and nonempty interval such that for each there exists satisfying
(78) | ||||
(79) |
Then the restriction of to has Lipschitz constant .
8. Bifurcations to bunching in the family of square examples
In this section we apply our results and techniques to a concrete example and completely describe the solution on the domain . We prove Theorem 1.4. For the solution is as hypothesized in the earlier work of McCann and Zhang [41]. Recall denotes the set of leaves with one endpoint on and the other on and here the solution is given explicitly in [50, 41]. We let denote the set of leaves with one endpoint in and the other on , and, finally, let denote the set of leaves with one endpoint in and the other on . Because, in the course of our proof, we prove does not intersect or we have .
Our main tool to study the minimizer is the coordinates introduced in Section 6. Let us consider a component of consisting of leaves with one endpoint on the boundary and the other interior to . The argument is similar on each side. We may take the angle made by leaves with the horizontal (that is, with the vector ), as the parametrization coordinate of our boundary (i.e. and ). Then and satisfy
where is the height at which the leaf that makes angle with the horizontal intersects . We work in a connected subset of
In this setting (67) and (68) become
(80) | ||||
(81) |
where we use the prime notation for derivatives of roman characters as opposed to the dot notation for derivatives of greek characters, and equation (65) becomes
(82) |
Note when we parameterize with respect to , so is Lipschitz by (61). We’ve used that as is easily seen by first working with the parametrization and the angles , for which the identity derived in Section 6 implies . Equations (80) – (82) yield a new, expedited, proof of the Euler–Lagrange equations in originally derived by the first and third author [41] via a complicated perturbation argument. Solving (60) and (61) gives
(83) | ||||
(84) |
Thus, along
Substituting into (82) we obtain
After multiplying by and solving (81) for we obtain (9), which coincides precisely with equation (4.22) of [41].
We obtain Theorem 1.4 as a combination of Lemmas. As required by the theorem, we henceforth make the tacit assumption .
Lemma 8.1 (Exclusion includes right-angled triangle in lower left corner).
Let minimize (1) with . Then and for some satisfying .
Proof.
Whenever is a subset of the first quadrant, symmetry shows the minimizer satisfies . Since the inclusion of Theorem 1.1 becomes an equality if is nonempty, monotonicity of convex gradients implies for some such that = in this case; here symmetry across the diagonal and the fact that is closed have been used. Armstrong has proved that has positive measure whenever is strictly convex [2] and this result has been extended to general benefit functions by Figalli, Kim, and McCann [30]. It is straightforward to adapt their proof to our setting. Indeed, convexity of and symmetry across the diagonal means implies and this implication is the only place strict convexity is used in [30, Theorem 4.8]. ∎
Proposition 8.2 (No normal distortion along top right boundaries).
Let solve (1) with . Then throughout and . Consequently, .
Proof.
For a contradiction we assume (without loss of generality, by Proposition 2.3) there is , at which ; when is a vertex of we interpret . With this interpretation the continuity of implies we may in fact assume, without loss of generality, that lies in the relative interior of . Thus has positive diameter and the same is true in a relatively open portion of the boundary, by Proposition 5.4. Working on , it is convenient to let denote the clockwise angle a ray makes with the inward normal . Thus corresponds to a ray with nonpositive slope. Parametrizing the boundary as using this , the derivation of the equation (82) is unchanged along . In particular , which is most easily seen by beginning with the clockwise oriented parametrization and in the coordinate arguments of Section 6 and recalling .
Clearly or ; we will derive a contradiction in either case.
Case 1. (Nonpositively sloped leaf). First let’s assume the leaf has nonpositive slope (i.e. ). The inequality implies leaves above, but in the same connected component of , as with one endpoint on are also nonpositively sloped.
At the endpoint of each leaf the Neumann condition is not satisfied, that is , equivalently (by the sign condition on the Neumann value). On the boundary portion where leaves have nonnegative slope, is a nondecreasing function (see Figure 5(A)). Thus for all and, by Proposition 5.4, each is the endpoint of a nontrivial leaf of nonpositive slope. We consider the following dichotomy and derive a contradiction in either case: (1) there is a sequence of leaves approaching with one endpoint on and the other in or else there is not, in which case all sequences of leaves approaching have one end on and the other on .
(B) Since is constant along , monotonicity of the gradient implies and for all .
Case 1a. (All leaves approaching the vertex have one end in the interior). In the first case take a sequence with satisfying and that is increasing in . We can take to contain points of Alexandrov second differentiability of since the leaves occupy positive area by Corollary 6.6 and Fubini’s theorem. Let the corresponding angles be . Because the leaves don’t intersect other sides of the square, symmetry and the sign of the angle yield . Theorem 4.1(1) provides a estimate along leaves with one endpoint on the boundary. Thus from (66),
evaluated at we obtain an estimate
Combined with (82), i.e. , we contradict that but is positive and increasing.
Case 1b. (There exists a leaf crossing the domain). In the second case we pick any leaf with one endpoint (call it ) on and the other on . Then by symmetry. Note (by the Neumann inequality on and ). Also this leaf bounds a right triangle with sides , and segments of (Figure 5(B)). Define
(85) |
Because is defined by extension of an affine support for , for all , . Moreover for we have
this is because monotonicity of the gradient and the Neumann condition implies for , , whereas . Thus is admissible for (1) and strictly decreases , a contradiction, given that minimizes .
Case 2. (Positively sloped leaf). If our originally chosen leaf has positive slope (i.e. ) the proof is similar, with slight modifications in the lower right corner. Indeed, implies all leaves below our chosen leaf also have positive slope and on such leaves is a decreasing function (by monotonicity of , as in Case 1). Thus the Neumann value increases as we move towards the lower right corner. Proposition 5.4 then implies each is the endpoint of a nontrivial leaf with positive slope. Consider the same two alternatives as above: there is a sequence of leaves whose endpoints on converge to and whose other endpoint is interior to , or there is not.
In the first case the contradiction is the same as in Case 1a above. In the second case choose a leaf with endpoint on and other endpoint on . By the Neumann inequality, Proposition 2.3, while . For in the interior of the right triangle formed by and segments of , monotonicity of the gradient implies
(86) | |||
(87) |
Thus the affine extension as in (85) once again satisfies (because ) and in . Thus — the same contradiction as in Case 1b above.
Conclusion: . It remains to be shown that the corners and are not the endpoints of rays. This follows by the maximum principle, Lemma 3.2, combined with a reflection argument. For example, suppose a ray is the endpoint of a nontrivial ray which has negative slope and thus enters . Fix any point in the relative interior of and let be the normal to that has positive components. Then for sufficiently small and the half ball
is contained in (because no rays intersect the relative interior of ). After subtracting from its support at and extending the resulting function to
via reflection from we obtain a function which violates Lemma 3.2. We conclude no rays intersect and, via an identical argument, no rays intersect and . ∎
Remark 8.3 (No ray has positive slope).
A similar argument to the above implies no leaf intersecting or has positive slope. Indeed, if a leaf on has positive slope its other endpoint is interior to (the leaf cannot intersect or ). The same argument as Case 1 above implies as one moves vertically up each boundary point remains the endpoint of a nontrivial leaf of nonpositive slope. Leaves must have length shrinking to as they approach and thus we obtain the same contradiction as in Case 1a above. As a result along the function is nondecreasing (equivalently is nonincreasing; see again Figure 5(A)).
Another key point that it will be helpful to have at our disposal is the following.
Lemma 8.4 (Concave nondecreasing profile of stingray’s tail).
Proof.
Connectivity of combines with Theorem 1.1(2) and the definition of to imply the set of fixed boundary endpoints of rays intersecting is also connected — hence forms an interval on the left boundary of the square; it cannot intersect the relative interior of according to Proposition 8.2. On the set of rays whose endpoints lie in the relative interior of this interval , the Lipschitz regularity of from Theorem 4.1 and Corollary 6.6 justifies the following computations.
We wish to consider the convexity of the curve where by (83) and (84)
Using Lipschitz regularity of we have for almost every
(88) |
Here the sign condition comes from Remark 8.3. Note this implies the curve is such that is an increasing function of . We see is an increasing function of , namely
(89) |
and subsequently, by Remark 8.3 which implies is an increasing function of , is an increasing function .Thus we have the required monotonicity of the derivative to conclude is a convex function of . ∎
Symmetry implies connected components of are curves with a concave function of .
Now we can combine all the Lemmas we’ve just proved and complete the proof of Theorem 1.4.
Proof.
(Theorem 1.4). By symmetry about the diagonal and we can prove each point of the theorem by an analysis of the function . Lemma 8.1 asserts is in as is for some . On we have . whereas on we have . Thus, for , [15, 50] implies some portion of must abut as one moves up .
Now we consider the configuration of . Since leaves in reach the diagonal, by symmetry they are orthogonal to it and on such leaves, i.e. the product selected lies on the diagonal.
Step 1. (Configuration of domains) We claim as one moves vertically up there is, in order, (i) a closed interval of with positive length, (ii) a half-open interval of which is empty for sufficiently small and nonempty for sufficiently large, (iii) a nonempty open interval of , and finally an interval of . All we must show is there is at most a single component of , and it is followed by . This is because, if and exist their ordering follows from Lemma 8.4. Indeed if a portion of is preceded by or then followed by we have the contradiction of a strictly convex curve lying above the diagonal with a start and endpoint on the diagonal in the stingray’s profile.
Step 2. (Blunt bunching (i.e. ) for ). Recall cannot be affine on any segment in the closure of , by a reflection argument combined with the strong maximum principle of Lemma 3.2. It follows that is nonempty whenever is nonempty; this was previously established by a different approach in [40]. Next we assume is empty and show . Let be the upper endpoint of and let be the lower endpoint of . The segment consists of endpoints of leaves in . By the Neumann condition we have and . Thus,
As in (4.17) of [41], using the coordinates we compute
From (81), which reads we have
We’ve used the inverse function theorem to rewrite . Using from Remark 8.3 and and for convexity of we conclude
Step 3. (No blunt bunching (i.e. ) for sufficiently small) Suppose for a contradiction that there is a sequence such that the minimizer on has nonempty. Let be the minimal convex extension to of the corresponding minimizer to (1). Let, for example denote for the problem on , and domains with no superscript denote the corresponding domain for . The convergence result [30, Corollary 4.7] implies locally uniformly for any where is the minimizer for .
It is clear that is empty when . Indeed, for the solution on is the restriction of the solution on . The solution on satisfies : Theorem 1.1(2) asserts the rays all extend to the boundary but Proposition 8.2, which is valid also on asserts there can be no ray intersecting the boundaries and . By symmetry there are no rays intersecting anywhere on and hence no rays whatsoever.
Recall (4.18) of [41] asserts is the triangle defined by . The limit from Theorem 1.1 must therefore contain the triangle of area in . Nor can be larger than this triangle, since Proposition 2.3 implies both integrands are non-negative in the identity
asserted by Lemma A.6. But the previous paragraph implies , so outside the triangle the minimizing is a strictly convex solution to Poisson’s equation . Reflecting this solution across the line where it vanishes contradicts the strong maximum principle (Lemma 3.2). This is the desired contradiction which establishes Step 3.
Remark 8.5 (Estimating the bifurcation point).
It is clear that is not sharp for the existence of . However the existence of a bifurcation reflects the radically different behavior we have shown the model to display for small and large . We expect there is a single bifurcation value such that is nonempty for while is empty for . It would be interesting to confirm this expectation, and to find or estimate more precisely.
Appendix A Rochet and Choné’s sweeping and localization
We recall as in (17) the minimizer of (1) satisfies the variational inequality
(90) |
for all convex with disjoint from . Since [15], we know
(91) |
is a measure of finite total-variation and we can rewrite (90) as
(92) |
The disintegration theorem, which we state in Section A.3 (see also [1, Theorem 5.3.1], [24, 78-III]) implies we may disintegrate the measure (by separately disintegrating its positive and negative parts and ) with respect to the map (equivalently with respect to the contact sets ). Our goal in this section is to prove Corollary A.9, namely to show that for almost every (90) holds for the disintegration on . In fact, provided we will prove the result for general convex , and for we will prove the result for satisfying . More precisely, the disintegration theorem implies there exists families of measures and such that for all Borel
We show for almost every
(93) |
for all convex with disjoint from . As we will prove the same result (with replacing ) then holds for almost every and almost every .
We emphasize that this appendix, whilst included for completeness, merely provides some more details on Rochet and Chonè’s proof of this localization property [50].
A.1. Measures in convex order and sweeping operators.
We begin with the following definition which is used in the theory of martingales and clearly related to (92).
Definition A.1 (Convex order).
Let be a Borel set. Let be Borel probability measures. We write (read as precedes in convex order) provided for every continuous convex there holds
The “Sweeping Theorem” characterizes measure in convex order and requires some more definitions. We take this Theorem from the work of Strassen [54, Theorem 2] where it’s attributed to “Hardy–Littlewood–Pólya–Blackwell–Stein–Sherman–Cartier–Fell–Meyer” (see also [21, Théorème 1], [43, T51, T53]).
Definition A.2 (Sweeping operators and Markov kernels).
(1) By a Markov
kernel on we mean a function where for each ,
is a probability measure. As a technicality we require for
each Borel that is Borel measurable.
(2) A Markov kernel
is called a sweeping operator444The term sweeping is also sometimes
known as a balayage or dilation. provided it satisfies that for each affine
,
(94) |
(3) If and is a Markov kernel on , then we define by
(95) |
We note two points: First, (94) is equivalent to the requirement
Second, for each integrable , from (95) we have
(96) |
The sweeping theorem gives a necessary and sufficient condition for to precede in convex order.
Theorem A.3 (Sweeping characterization of convex order; see [54]).
The measures satisfy if and only if there exists a sweeping operator such that .
A.2. First characterization; Lagrange multiplier and sweeping
We aim to apply Theorem A.3 to the positive and negative parts of . However Theorem A.3 does not apply directly because we do not have the condition but only the weaker condition obtained by testing against nonnegative convex functions. Nevertheless, we obtain the following lemma.
Lemma A.4 (Restoring neutrality).
Let be as in (91). There exists a nonnegative measure supported on and a sweeping operator on such that
(97) | ||||
(98) |
for
The representation (97) is, of course, trivial. The essential conclusion is that after subtracting the Lagrange multiplier (our reason for designating it so will be clear in the proof) .
Proof of Lemma A.4.
First note for all , is admissible for the minimization problem. Thus by minimality
(99) |
Combined with the minimality condition (92) we have
with realizing the infimum. A classical theorem in the calculus of variations says the constraint may be realized by a Lagrange multiplier (see, for example, [37, Theorem 1,pg. 217]). Thus there is a nonnegative Radon measure such that
(100) |
with still realizing the infimum. Using that attains the infimum along with (99) yields
Since is nonnegative we conclude . Since (100) implies we see (98) follows by Theorem A.3. Note that and may not be probability measures but are of finite and equal mass (finiteness follows from [15]), equality from testing against ) and this justifies our use of Theorem A.3. ∎
The following property of the sweeping operator is essential to our arguments.
Lemma A.5 (Localization to leaves).
Let be the sweeping operator given by Lemma A.4. Then for almost every the measure is supported on .
Proof.
At this point we have almost everything needed to obtain the localization property. Now we establish the following lemma which was was used earlier in the proof of Proposition 5.4.
Lemma A.6 (Objective responds proportionately to uniform utility increase).
Let denote the variational derivative (recall equation (21)). Then that is is a unit Dirac mass at the origin.
Proof.
That the variational derivative assigns measure to was observed by Rochet and Chonè [50]: this follows from the leafwise neutrality outside implied by Lemmas A.4–A.5, and from for all . Thus it suffices to prove for any not containing or, equivalently,
(102) |
Because is supported on for a.e. we have
where we’ve used for . Since for we have (up to normalization) this proves (102). ∎
A.3. Disintegration and localization
In this section we complete the proof that the variational inequality (18) holds on almost every contact set. We use the Disintegration Theorem in the following form, which can be viewed as a continuum generalization of Bayes’ theorem (see [1, Theorem 5.3.1] and [24, 78-III]).
Theorem A.7 (Disintegration of measure).
Let be Radon separable metric spaces, and let be a Borel map used to define the push forward . Then there exists a -a.e. uniquely determined Borel family of probability measures such that vanishes outside for -a.e. , and
for every Borel .
Theorem A.8 (Leafwise Euler-Lagrange condition).
We work away from and use without further reference to this restriction. The idea of the proof is that because is supported on the conditioning of , denoted , is obtained by sweeping the conditioning of , denoted . More succinctly
(104) |
Proof.
We apply the Disintegration Theorem a number of times in this proof, each time with . Applying the Disintegration Theorem to the measure from Lemma A.4, we obtain a family of measures such that for any Borel
(105) |
We consider two ways of expressing the result of disintegrating . First, by a direct application of the disintegration theorem
(106) |
On the other hand using the sweeping operator and the disintegration of , (105), we obtain
(107) |
Using that is supported on the inner two integrals in (107) become integration against . Namely,
(108) |
To conclude we recall from Lemma A.6 that . Thus comparing (108) and (106) and using the uniqueness a.e of the conditional measures we obtain
which is (104) so by the sweeping characterization of convex order (Theorem A.3) we obtain (103) for almost every .
To finish the proof we must translate back to stating the result in terms of a.e. . Let be the set of for which the leafwise localized inequality (103) does not hold. Then and by Lemma A.6 . Hence
(109) |
It follows that
Now on each with because the leafwise inequality does not hold we have on a positive measure subset of . Indeed, since the leafwise inequality does not hold either or and in this latter case mass balance, Lemma A.6, implies . Thus , which clearly contains , has measure . Thus .
Finally to obtain the result for almost every , note at boundary points of strict convexity the leafwise (now, pointwise) inequality is satisfied by Lemma A.6. Moreover if the set of which are contained in nontrivial contact sets and has then, by Fubini’s theorem . We conclude ∎
Corollary A.9 (Rochet and Chonè’s leafwise localization).
Every convex satisfies
(110) |
for almost every and for almost every . (The same conclusions extend to if also .)
References
- [1] Luigi Ambrosio, Nicola Gigli, and Giuseppe Savaré. Gradient flows in metric spaces and in the space of probability measures. Lectures in Mathematics ETH Zürich. Birkhäuser Verlag, Basel, 2005.
- [2] M. Armstrong. Multiproduct nonlinear pricing. Econometrica, 64:51–75, 1996.
- [3] S. Basov. Multidimensional Screening. Springer-Verlag, Berlin, 2005.
- [4] G.S. Becker. A theory of marriage. Part I. J. Political Econom., 81:813–846, 1973.
- [5] Ivan Blank. Sharp results for the regularity and stability of the free boundary in the obstacle problem. Indiana Univ. Math. J., 50(3):1077–1112, 2001.
- [6] Ivan Alexander Blank. Sharp results for the regularity and stability of the free boundary in the obstacle problem. ProQuest LLC, Ann Arbor, MI, 2000. Thesis (Ph.D.)–New York University.
- [7] Job Boerma, Aleh Tsyvinski, and Alexander P. Zimin. Bunching and taxing multidimensional skills. Preprint at arXiv:2204.13481, 2022+.
- [8] L. A. Caffarelli. The obstacle problem revisited. J. Fourier Anal. Appl., 4(4-5):383–402, 1998.
- [9] L.A. Caffarelli and P.-L. Lions. Untitled notes. Unpublished, 2006+.
- [10] Luis A. Caffarelli. The regularity of free boundaries in higher dimensions. Acta Math., 139(3-4):155–184, 1977.
- [11] L.A. Caffarelli, M. Feldman and R.J. McCann. Constructing optimal maps for Monge’s transport problem as a limit of strictly convex costs. J. Amer. Math. Soc., 15:1–26, 2002.
- [12] G. Carlier. A general existence result for the principal-agent problem with adverse selection. J. Math. Econom., 35:129–150, 2001.
- [13] G. Carlier. Calculus of variations with convexity constraints. J. Nonlinear Convex Anal., 3:125–143, 2002.
- [14] G. Carlier and T. Lachand-Robert. Regularité des solutions d’un problème variationnel sous contrainte de convexité. C.R. Acad. Sci. Paris Sér. I Math., 332:79–83, 2001.
- [15] G. Carlier and T. Lachand-Robert. Regularity of solutions for some variational problems subject to convexity constraint. Comm. Pure Appl. Math., 54:583–594, 2001.
- [16] G. Carlier and T. Lachand-Robert. Representation of the polar cone of convex functions and applications. J. Convex Anal., 15(3):535–546, 2008.
- [17] G. Carlier, T. Lachand-Robert, and B. Maury. A numerical approach to variational problems subject to convexity constraint. Numer. Math., 88(2):299–318, 2001.
- [18] Guillaume Carlier and Xavier Dupuis. An iterated projection approach to variational problems under generalized convexity constraints. Applied Mathematics & Optimization, 76(3):565–592, 2017.
- [19] Guillaume Carlier, Xavier Dupuis, Jean-Charles Rochet, and John Thanassoulis. A general solution to the quasi linear screening problem. J. Math. Econom., 114:Paper No. 103025, 14, 2024.
- [20] Guillaume Carlier and Thomas Lachand-Robert. Regularity of solutions for some variational problems subject to a convexity constraint. Communications on Pure and Applied Mathematics: A Journal Issued by the Courant Institute of Mathematical Sciences, 54(5):583–594, 2001.
- [21] Pierre Cartier, J. M. G. Fell, and Paul-André Meyer. Comparaison des mesures portées par un ensemble convexe compact. Bull. Soc. Math. France, 92:435–445, 1964.
- [22] Shibing Chen. Regularity of the solution to the principal-agent problem. In Geometric and functional inequalities and recent topics in nonlinear PDEs, volume 781 of Contemp. Math., pages 41–47. Amer. Math. Soc., [Providence], RI, 2023.
- [23] Pierre-Andre Chiappori, Robert J. McCann, and Brendan Pass. Multi- to one-dimensional optimal transport. Comm. Pure Appl. Math., 70:2405–2444, 2017.
- [24] Claude Dellacherie and Paul-André Meyer. Probabilities and potential, volume 29 of North-Holland Mathematics Studies. North-Holland Publishing Co., Amsterdam-New York, 1978.
- [25] Ivar Ekeland and Santiago Moreno-Bromberg. An algorithm for computing solutions of variational problems with global convexity constraints. Numer. Math., 115(1):45–69, 2010.
- [26] Lawrence C. Evans and Ronald F. Gariepy. Measure theory and fine properties of functions. Studies in Advanced Mathematics. CRC Press, Boca Raton, FL, 1992.
- [27] Xavier Fernández-Real and Xavier Ros-Oton. Regularity theory for elliptic PDE, volume 28 of Zurich Lectures in Advanced Mathematics. EMS Press, Berlin, [2022] ©2022.
- [28] Alessio Figalli. The Monge-Ampère equation and its applications. Zurich Lectures in Advanced Mathematics. European Mathematical Society (EMS), Zürich, 2017.
- [29] Alessio Figalli. Free boundary regularity in obstacle problems. Journées équations aux dérivées partielles, 2(2):1–24, 2018.
- [30] Alessio Figalli, Young-Heon Kim, and Robert J. McCann. When is multidimensional screening a convex program? J. Econom Theory, 146:454–478, 2011.
- [31] D. Gilbarg and N.S. Trudinger. Elliptic Partial Differential Equations of Second Order. Springer-Verlag, New York, second edition, 1983.
- [32] D. Kinderlehrer and L. Nirenberg. Regularity in free boundary problems. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4), 4(2):373–391, 1977.
- [33] David Kinderlehrer. The free boundary determined by the solution to a differential equation. Indiana Univ. Math. J., 25(2):195–208, 1976.
- [34] Alexander V. Kolesnikov. Auctions and mass transportation. Preprint at arXiv:2312.08077.
- [35] Alexander V Kolesnikov, Fedor Sandomirskiy, Aleh Tsyvinski, and Alexander P Zimin. Beckmann’s approach to multi-item multi-bidder auctions. Preprint at arXiv:2203.06837v2, 2022+.
- [36] Jiakun Liu and Xu-Jia Wang. Interior a priori estimates for the Monge-Ampère equation. In Surveys in differential geometry 2014. Regularity and evolution of nonlinear equations, volume 19 of Surv. Differ. Geom., pages 151–177. Int. Press, Somerville, MA, 2015.
- [37] David G. Luenberger. Optimization by vector space methods. John Wiley & Sons, Inc., New York-London-Sydney, 1969.
- [38] Robert J McCann, Cale Rankin, and Kelvin Shuangjian Zhang. regularity for principal-agent problems. Preprint at arXiv:2303.04937, 2023.
- [39] Robert J. McCann and Kelvin Shuangjian Zhang. On concavity of the monopolist’s problem facing consumers with nonlinear price preferences. Comm. Pure Appl. Math., 72(7):1386–1423, 2019.
- [40] Robert J. McCann and Kelvin Shuangjian Zhang. Remark on Rochet and Choné’s square screening example. Preprint at arXiv:2311.13012, to appear in J. Convex Anal., 2023+.
- [41] Robert J. McCann and Kelvin Shuangjian Zhang. A duality and free boundary approach to adverse selection. Math. Models Methods Appl. Sci., 34(12):2351–2394, 2024. arXiv:2301.07660.
- [42] Quentin Mérigot and Édouard Oudet. Handling convexity-like constraints in variational problems. SIAM J. Numer. Anal., 52(5):2466–2487, 2014.
- [43] Paul-A. Meyer. Probability and potentials. Blaisdell Publishing Co. [Ginn and Co.], Waltham, Mass.-Toronto, Ont.-London, 1966.
- [44] Jean-Marie Mirebeau. Adaptive, anisotropic and hierarchical cones of discrete convex functions. Numer. Math., 132(4):807–853, 2016.
- [45] J.A. Mirrlees. An exploration in the theory of optimum income taxation. Rev. Econom. Stud., 38:175–208, 1971.
- [46] M. Mussa and S. Rosen. Monopoly and product quality. J. Econom. Theory, 18:301–317, 1978.
- [47] Georg Nöldeke and Larry Samuelson. The implementation duality. Econometrica, 86(4):1283–1324, 2018.
- [48] Adam M. Oberman. A numerical method for variational problems with convexity constraints. SIAM J. Sci. Comput., 35(1):A378–A396, 2013.
- [49] Arshak Petrosyan, Henrik Shahgholian, and Nina Uraltseva. Regularity of free boundaries in obstacle-type problems, volume 136 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2012.
- [50] J.-C. Rochet and P. Choné. Ironing, sweeping and multidimensional screening. Econometrica, 66:783–826, 1998.
- [51] Jukka Sarvas. The Hausdorff dimension of the branch set of a quasiregular mapping. Ann. Acad. Sci. Fenn. Ser. A I Math., 1(2):297–307, 1975.
- [52] R. Schneider. Convex Bodies: The Brunn-Minkowski Theory. Cambridge University Press, Cambridge, 1993.
- [53] M. Spence. Job market signaling. Quarterly J. Econom., 87:355–374, 1973.
- [54] V. Strassen. The existence of probability measures with given marginals. Ann. Math. Statist., 36:423–439, 1965.
- [55] R. Wilson. Nonlinear Pricing. Oxford University Press, Oxford, 1993.