Strict convexity and regularity of solutions to generated Jacobian equations in dimension two
Abstract.
We present a proof of strict -convexity in 2D for solutions of generated Jacobian equations with a -Monge–Ampère measure bounded away from 0. Subsequently this implies differentiability in the case of a -Monge–Ampère measure bounded from above. Our proof follows one given by Trudinger and Wang in the Monge–Ampère case. Thus, like theirs, our argument is local and yields a quantitative estimate on the -convexity. As a result our differentiability result is new even in the optimal transport case: we weaken previously required domain convexity conditions. Moreover in the optimal transport case and the Monge–Ampère case our key assumptions, namely A3w and domain convexity, are necessary.
2010 Mathematics Subject Classification:
35J96 and 35J661. Introduction
Generated Jacobian equations are PDEs of the form
| (1) |
where the vector field has a particular structure. This class of equations includes the Monge–Ampère equation and the Jacobian equation from optimal transport as special cases. Precise statements concerning the structure of are given in Section 2. For now we state our purpose is to show, for generalised notions of convexity, that solutions of
| (2) |
on a convex domain are strictly convex. By considering an analogue of the Legendre transform we obtain that when (2) is instead a bound from above and satisfies the second boundary value problem
| (3) |
for a domain and a convex domain , then is .
The class of we work with, those obtained from a generating function, embraces applications in optimal transport and geometric optics, whilst allowing the use of a large family of techniques developed for the study of the Monge–Ampère equation. Thus these equations, first studied by Trudinger [18], provide a good combination of applicability and tractability. For example, our proof follows the corresponding proof for the Monge–Ampère equation as given by Trudinger and Wang [20, Remark 3.2]. Though there the result is originally due to Alexandrov [1] and Heinz [9].
The work most relevant to ours is that of Figalli and Loeper [5] who dealt with optimal transport, that is when . They proved the differentiability of solutions in two dimensions under a bound from above on the -Monge–Ampère measure. They assumed a uniformly -convex source and a strictly -convex target . Thanks largely to the maturity of the relevant convexity theory we are able to reduce these convexity conditions. Our result requires no convexity condition on the source and only -convexity on the target. This condition is necessary even for the Monge–Ampère equation.
Our results are inherently two dimensional — in higher dimensions strict convexity does not hold under only a bound from below. However one of the key applications of generated Jacobian equations (GJE) is geometric optics and this takes place in two dimensions. In higher dimensions a two sided bound on the Monge–Ampère measure is required for strict convexity and differentiability. Here the relevant work is that of Caffarelli [2] for the Monge–Ampere equation, Chen, Wang [3], Figalli, Kim, McCann [4], Guillen, Kitagawa [7], and Vétois [21] for optimal transport, and Guillen, Kitagawa [8] for GJE.
Our plan is to define generating functions, GJE, and the related convexity notions in Section 2. Here we also state precisely our main results: Theorem 1 and Corollary 1. In Section 3 we state a versatile differential inequality (essentially taken from [18, 19]) whose proof we relegate to the appendix. We prove strict convexity in Section 4 and conclude with differentiability in Section 5.
Acknowledgements
Thanks to Neil Trudinger who suggested extending the proof in [20] to generated Jacobian equations.
2. Generating functions and GJE
Generated Jacobian equations are equations of the form (1) where derives from a generating function (defined below). This requirement allows us to develop a framework extending convexity theory. The material below is largely due to Trudinger [18] with other presentations in [8, 10, 11, 12, 14, 17]. Guillen’s survey article [6] is a good introduction and lists the 2D theory we develop here as an open problem. Throughout it is helpful to keep in mind the cases and where is a cost function from optimal transport. In these settings -convexity simplifies to standard convexity and the cost convexity of optimal transport respectively.
2.1. Generating functions
The structure of a particular generated Jacobian equation derives from a generating function. We denote this function by , and require it satisfy the following assumptions.
A0. where is a bounded domain satisfying that the projections
are open intervals. Moreover we assume there are domains such that for all we have that is nonempty.
A1. For all defined as
there is a unique such that
A2. On there holds and the matrix E with entries
satisfies .
2.2. Generated Jacobian equations
Assumption A1 allows us to define mappings and by requiring they solve
| (4) | ||||
| (5) |
We call a PDE of the form
| (1) |
a generated Jacobian equation provided derives from solving (4) and (5) for some generating function.
Generated Jacobian equations may be rewritten as Monge–Ampère type equations. Suppose satisfies (1). Then differentiating (4) and (5) evaluated at yields
Since the first equation with (5) implies
| (6) |
the second can be solved for by which we have
| (7) |
with as in A2. Thus solutions of (1) solve
| (8) |
for
This PDE is elliptic when as matrices. A necessary condition for ellipticity is .
2.3. -convex functions
We introduce an analogue of convexity theory in which the generating function plays the role of a supporting hyperplane. We say a function is -convex provided for all there is such that
| (9) | ||||
| (10) |
In this case is called a -support at . We say is strictly -convex if the inequality in (10) is strict, that is, any given -support only touches at a single point. Note when is differentiable, since has a minimum at , we have . This combined with (9) is equivalent to (5) and (4) so that and . Moreover, if is , again since has a minimum at , we have and (8) is degenerate elliptic. However when is not differentiable (9) and (10) may hold for more than one . The set of for which there is such that (9) and (10) hold is denoted . Similarly for . When both are a singleton we identify these sets with their single element.
2.4. Alexandrov solutions
The mapping allows us to define a notion of Alexandrov solution. A -convex function is called an Alexandrov solution of (1) on provided for every Borel
Since -convex functions are locally semi-convex and thus differentiable almost everywhere the integrand on the right-hand side is well defined almost everywhere. We see, via the change of variables formula, that when is and is a diffeomorphism Alexandrov solutions are classical solutions. Moreover in the case of inequality (2) the notion of Alexandrov solution is that for every
Here we call the measure defined on Borel by the -Monge–Ampère measure.
2.5. Domain Convexity
With the goal of introducing a notion of domain convexity we introduce a generalisation of line segments. A collection of points is called a -segment with respect to joining to provided
Using this we say a set is -convex with respect to provided for each the above -segment is contained in . The -segment is unique via condition A1∗ in Section 2.7. If and we say is -convex with respect to if is -convex with respect to every
2.6. Conditions for regularity
Domain convexity and a Monge–Ampère measure bounded below is necessary and sufficient for strict convexity in 2D in the Monge–Ampère case. In the optimal transport case we need the now well known A3w condition. This condition was introduced for optimal transport by Ma, Trudinger, and Wang [16] and generalised to GJE by Trudinger. For GJE we use an additional condition A4w (due to Trudinger [18]).
A3w. A generating function is said to satisfy the condition A3w provided for all with and there holds
| (11) |
A4w. The matrix is non-decreasing in , in the sense that for any
2.7. The dual generating function
Strictly convex functions have Legendre transforms. This provides a useful technique for proving solutions of Monge–Ampère inequalities are . The same technique in the -convex case requires the -transform which is defined in terms of the dual generating function. We introduce the dual generating function here and the transform in Section 5. We set
The dual generating function, , is the unique function defined on by
| (12) |
It follows that if then
| (13) |
Further, if is -convex and then the corresponding support is .
We introduce dual conditions on .
A1∗. For all defined as
there is a unique such that
Moreover we define mapping as the unique that satisfy these equations (cf. (4) and (5)). As remarked in Section 2.5 A1∗ implies uniqueness of the -segment between two points. For this note by differentiating (13) we obtain . The right hand side is injective in for fixed .
We define dual objects by swapping the roles of and as well as and . In particular -convex functions are those defined on with supports, and for a -convex we have the mappings which are analogues of . Similarly -segments are used to define -convex sets. We also define dual conditions A2∗, A3∗, A4w∗. However Trudinger [18] showed the conditions A2∗ and A3w∗ are satisfied provided A2 and A3w are.
2.8. Main results
Theorem 1.
Let be a generating function satisfying A0, A1, A2, A1∗, A3w and A4w. Let and be a -convex function satisfying for and all
| (14) |
If is -convex with respect to then is strictly -convex.
Corollary 1.
Let be a generating function satisfying A0, A1, A2, A1∗, A3w, and A4w∗. Let and be a convex function satisfying that for and all
| (15) | |||
| (16) |
If is -convex with respect to then .
Note conditions A4w and A4w∗ are always satisfied in the optimal transport case. Indeed Corollary 1 implies the differentiability of potentials for the optimal transport problem (and subsequently continuity of the optimal transport map) whenever the right hand side of the associated Monge–Ampère type equation is bounded from above, the target is -convex, and the A3w condition is satisfied. These conditions are all known to be necessary. Necessity of the -convexity is due to Ma, Trudinger, and Wang [16] whilst necessity of A3w is due to Loeper [15].
3. Main lemma
We make frequent use of a differential inequality for the difference of a -convex function and a -affine function. Similar inequalities are used in [18, 19] and the works of Kim and McCann [13, Proposition 4.6] and Guillen and Kitagawa [8, Lemma 9.3].
Lemma 1.
Let be a generating function satisfying A0, A1 and A2. Let be given, a -segment with respect to , and a -convex function. Then
satisfies
| (17) | ||||
where depends only on the values of and its derivatives on and . We’ve used the shorthand . The arguments of and are given in the proof.
The proof can be found in [18]. For completeness we include a proof in Appendix A. To use A3w in (17) we need to control for arbitrary . We claim if are arbitrary then A3w implies
| (18) |
Here is a non-negative constant depending only on . To obtain (18) from A3w first prove it for arbitrary unit vectors by using A3w with as given and replaced by .
Thus when are satisfied and we have
where depends on and This inequality appears in [8, Lemma 9.3],[22, pg. 310, pg. 315] and [18, 19]. It implies, amongst other things, estimates on in terms of via the following lemma.
Lemma 2.
Let be a function satisfying . For there holds
where depend on respectively and .
Proof.
First note if at any then . To see this assume is the infimum of points with . By continuity if we are done. Otherwise is single signed on and if on this interval then again by continuity we are done. Thus we assume on . The inequality implies on . Subsequently for integration gives
| (19) |
and sending gives .
Now to prove the inequality in the lemma let’s deal with the upper bound first. We assume on otherwise the argument just given implies . We obtain (19) for and . Integrating with respect to from to establishes the result. The other inequality follows by applying the same argument to the function defined by . ∎
4. Strict convexity in 2D
In this section we present the proof of Theorem 1. The proof follows closely the proof of Trudinger and Wang [20, Remark 3.2] who obtained the same result in the Monge–Ampère case. The key ideas of our proof will be more transparent if the reader is familiar with their proof. There are two key steps: First we obtain a quantitative -convexity estimate for solutions of (importantly the estimate is independent of bounds on second derivatives). Then we obtain a convexity estimate for Alexandrov solutions via a barrier argument.
Theorem 1..
Step 1. Quantitative convexity for solutions
Initially we assume is . Let satisfy for . Assume for some there is distinct with
| (20) | |||
| (21) |
Let denote the -segment with respect to that joins to and set
We use the shorthand . Lemma 1 along with A3w and A4w yields with the same inequality holding for . Hence, via the maximum principle, is less than or equal to its value at the end points. Thus
| (22) |
The convexity estimate we intend to derive is where depends only on . We use to indicate any positive constant depending only on these quantities.
Now, via semi-convexity, -convex functions are locally Lipschitz. Thus for and sufficiently small, (22) implies
| (23) | ||||
| (24) |
We let be a continuous unit normal vector field to the -segment . Fix so that and lie in . For let be the -segment with respect to joining to . Using Lemma 2 for combined with (23) and (24) implies
| (25) |
Here we have used that for a Lipschitz constant independent of .
This implies
| (26) |
and we come back to this in a moment. For now note that, since
for any two orthogonal unit vectors
In particular for a choice of unit normal vector field orthogonal to (continuous in ) we have
Employing this and (25) in Lemma 1 gives
| (27) | ||||
where initially this holds for , and thus for . Note that by (18) the term in Lemma 1 is bounded below by and subsequently controlled by (25).
Substituting (27) into (26) we obtain
where we omit that is evaluated at . An application of Jensen’s inequality implies
| (28) | ||||
| (29) |
This is the crux of the proof complete: the only way for the final integral to be bounded is if is bounded away from 0. We’re left to show the integral (28) is bounded in terms of the allowed quantities, and approximate when is not .
To bound (28) use that implies is bounded below by a positive constant depending on and . This gives the estimate
The final line is obtained using positivity of and is bounded in terms of and (compute the integral in the Cartesian coordinates and note the Jacobian for this transformation is bounded). Thus returning to (28) and (29) we obtain where depends on the stated quantities.
Step 2: Convexity estimates for Alexandrov solutions via a barrier argument
We extend to Alexandrov solutions via a barrier argument. Suppose is an Alexandrov solution of (14) that is not strictly convex. There is a support touching at two points . Using [18, Lemma 2.3] we also have along the -segment joining these points (with respect to ). Balls with sufficiently small radius are -convex. This follows because, as noted in [14, §2.2], -convexity requires the boundary curvatures minus a function depending only on are positive. Thus we assume are sufficiently close to ensure that , the ball with radius and centre is -convex with respect to . Let be given, and let be the mollification of with taken small enough to ensure on The Dirichlet theory for GJE, [18, Lemma 4.6], yields a solution of
satisfying an estimate where depends on the local Lipschitz constant of . Since on the comparison principle, ([18, Lemma 4.4]) implies in . Thus we can apply our previous argument and obtain strict -convexity of provided we note (20) and (21) are satisfied for . Hence at a point on the -segment where the infimum defining is obtained we have
As we contradict that supports . ∎
5. regularity
For a -convex function defined on with its -transform is the function defined on by
We list a few essential properties. Let us suppose . This means
taking of both sides and using yields
so that . The definition of implies for other we have . Thus is a -support at . Which is to say .
We use this as follows. Suppose in addition satisfies that for all
Take and let denote the measure set of points where is not differentiable. Necessarily for some . Our above reasoning implies . Hence
| (30) |
Corollary 1 follows: Let be the function given in Corollary 1 and its transform defined on . Theorem 1 holds in the dual form, that is, provided the relevant hypothesis are changed to their starred equivalents, Theorem 1 implies strict -convexity. Thus the hypothesis of Corollary 1 along with (30) allow us to conclude is strictly -convex.
Suppose for a contradiction is not . Then for some the set contains two distinct points, say . Our above working implies is a support touching at . This contradicts strict -convexity and proves the corollary.
Appendix A Proof of main lemma
In this appendix we provide the proof of Lemma 1.
Proof.
We first compute a differentiation formula for second derivatives along -segments. We suppose
| (31) |
and set . We begin with a formula for first derivatives. Since
| (32) |
we need to compute . Differentiate (31) with respect to and obtain111We use the convention that subscripts before the comma denote differentiation with respect to , and subscripts after the comma (which are not ) denote differentiation with respect to .
from which it follows that
where denotes the m,i entry of . Thus (32) becomes
| (33) |
Using this expression to compute second derivatives we have
The formula for differentiating an inverse yields
| (34) | ||||
Now compute
| (35) |
Here we have used that
which follows by computing , differentiating (4) with respect to to express in terms of , and employing (which is obtained via calculations similar to those for (7)).
Substitute (35) into (34) to obtain
where in the last equality we swapped the dummy indices and on the second term to allow us to collect like terms and also used (33).
Now let’s use this identity to compute . We have
Terms on the final line are bounded below by . Now after adding and subtracting for we have
Set and . Then rewriting in terms of the matrix we have
Here results from a Taylor series. Another Taylor series for and we obtain
This is the desired formula. ∎
References
- [1] A. Alexandroff. Smoothness of the convex surface of bounded Gaussian curvature. C. R. (Doklady) Acad. Sci. URSS (N.S.), 36:195–199, 1942.
- [2] L. A. Caffarelli. A localization property of viscosity solutions to the Monge-Ampère equation and their strict convexity. Ann. of Math. (2), 131(1):129–134, 1990.
- [3] Shibing Chen and Xu-Jia Wang. Strict convexity and regularity of potential functions in optimal transportation under condition A3w. J. Differential Equations, 260(2):1954–1974, 2016.
- [4] Alessio Figalli, Young-Heon Kim, and Robert J. McCann. Hölder continuity and injectivity of optimal maps. Arch. Ration. Mech. Anal., 209(3):747–795, 2013.
- [5] Alessio Figalli and Grégoire Loeper. regularity of solutions of the Monge-Ampère equation for optimal transport in dimension two. Calc. Var. Partial Differential Equations, 35(4):537–550, 2009.
- [6] Nestor Guillen. A primer on generated Jacobian equations: geometry, optics, economics. Notices Amer. Math. Soc., 66(9):1401–1411, 2019.
- [7] Nestor Guillen and Jun Kitagawa. On the local geometry of maps with -convex potentials. Calc. Var. Partial Differential Equations, 52(1-2):345–387, 2015.
- [8] Nestor Guillen and Jun Kitagawa. Pointwise estimates and regularity in geometric optics and other generated Jacobian equations. Comm. Pure Appl. Math., 70(6):1146–1220, 2017.
- [9] Erhard Heinz. Über die Differentialungleichung . Math. Z, 72:107–126, 1959/1960.
- [10] Seonghyeon Jeong. Local Hölder regularity of solutions to generated Jacobian equations. Pure Appl. Anal., 3(1):163–188, 2021.
- [11] Yash Jhaveri. Partial regularity of solutions to the second boundary value problem for generated Jacobian equations. Methods Appl. Anal., 24(4):445–475, 2017.
- [12] Feida Jiang and Neil S. Trudinger. On the second boundary value problem for Monge-Ampère type equations and geometric optics. Arch. Ration. Mech. Anal., 229(2):547–567, 2018.
- [13] Young-Heon Kim and Robert J. McCann. Continuity, curvature, and the general covariance of optimal transportation. J. Eur. Math. Soc. (JEMS), 12(4):1009–1040, 2010.
- [14] Jiakun Liu and Neil S. Trudinger. On the classical solvability of near field reflector problems. Discrete Contin. Dyn. Syst., 36(2):895–916, 2016.
- [15] Grégoire Loeper. On the regularity of solutions of optimal transportation problems. Acta Math., 202(2):241–283, 2009.
- [16] Xi-Nan Ma, Neil S. Trudinger, and Xu-Jia Wang. Regularity of potential functions of the optimal transportation problem. Arch. Ration. Mech. Anal., 177(2):151–183, 2005.
- [17] Cale Rankin. Distinct solutions to generated jacobian equations cannot intersect. Bull. Aust. Math. Soc., 102(3):462–470, 2020.
- [18] Neil S. Trudinger. On the local theory of prescribed Jacobian equations. Discrete Contin. Dyn. Syst., 34(4):1663–1681, 2014.
- [19] Neil S. Trudinger. On the local theory of prescribed Jacobian equations revisited. Math. Eng., 3(6):Paper No. 048, 17, 2021.
- [20] Neil S. Trudinger and Xu-Jia Wang. The Monge-Ampère equation and its geometric applications. In Handbook of geometric analysis. No. 1, volume 7 of Adv. Lect. Math. (ALM), pages 467–524. Int. Press, Somerville, MA, 2008.
- [21] Jérôme Vétois. Continuity and injectivity of optimal maps. Calc. Var. Partial Differential Equations, 52(3-4):587–607, 2015.
- [22] Cédric Villani. Optimal transport, volume 338 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 2009. Old and new.