First and Second derivative Hölder estimates for generated Jacobian equations
Abstract.
We prove two Hölder regularity results for solutions of generated Jacobian equations. First, that under the A3 condition and the assumption of nonnegative valued data solutions are for an that is sharp. Then, under the additional assumption of positive Dini continuous data, we prove a estimate. Thus the equation is uniformly elliptic and when the data is Hölder continuous solutions are in .
1. Introduction
Generated Jacobian equations are a class of PDE which model problems in geometric optics and have recently seen applications to monopolist problems in economics [19, 15]. These equations are of the form
(1) |
where , is a domain, and satisfies a convexity condition that ensures the PDE is elliptic.
The precise form of this condition on and the structure of requires a number of definitions which we introduce in Section 2. However, the framework we work in includes, in addition to the just listed applications, two well-known special cases. First, the Monge–Ampère equation (take ). Second, the Monge–Ampère type equations from optimal transport (take as the optimal transport map depending on the corresponding potential function for the optimal transport problem). Indeed, generated Jacobian equations (GJEs) were introduced to treat the aforementioned new applications using the techniques from optimal transport. That’s what we do here: Two fundamental results for the regularity theory in optimal transport are the result of Liu [10] and the results of Liu, Trudinger, Wang [11]. These are based on the corresponding results of Caffarelli in the Monge–Ampère case, respectively, [1] and [2]. In this paper we extend the results of [10, 11] to generated Jacobian equations. Our main results are stated precisely at the conclusion of Section 2, once the necessary definitions have been introduced. The importance of our results is that, first, the (or at least some ) result is a necessary assumptions in the derivation of models in geometric optics and, second, the result puts us in the regime of classical elliptic PDE and lets us bootstrap higher regularity.
We note that the corresponding result of Loeper in the optimal transport setting [13] has been extended to generated Jacobian equations by Jeong [8]. Thus our key contribution for the result is improving the value of so that it is sharp. As we explain at the conclusion of Section 2 this yields a corresponding improvement on the result. An outline of the paper is provided by the table of contents.
2. Generating functions, -convexity, and GJEs
In this section we state the essential definitions. Further introductory material can be found in the expository article of Guillen [6], and more detailed outlines of the whole theory in [19, 20, 7, 16].
The framework for GJEs was introduced by Trudinger [19] and is built around a generalized notion of convexity. The generating function, which we now define, plays a central role, essentially that of affine hyperplanes in classical convexity.
Definition 1.
A generating function is a function, which we denote by , satisfying the conditions A0,A1,A1∗, and A2.
A0. where is a bounded domain of the form
for domains and an open interval for each . Moreover we assume there is an open interval such that for each .
A1. For each defined by
there is a unique , whose components we denote by , , satisfying
(2) |
A1∗. For each fixed the mapping is injective on its domain of definition.
A2. On there holds and the matrix
satisfies Here subscripts before and after a comma denote, respectively, differentiation in and .
Two examples are which generates (in accordance with Definition 3) the Monge–Ampère equation and standard convexity, and , where is a cost function from optimal transport, which generates the Monge–Ampère type equation from optimal transport, and the cost convexity theory [21, Chapter 5]. By a duality structure, which we do not need and thus don’t introduce here, the condition A1∗ is dual to , thereby justifying the name. The A0 condition is weakened by some authors who treat generating functions [7, 9]. However our interior Pogorelov estimate, an essential tool for the result, is obtained by differentiating the PDE twice which relies on a generating function.
Definition 2.
The set of all such that there is for which is a -support at is denoted . When is differentiable and is a -support at then has a minimum at implying by (2) that . Similarly, if is a -support at and is then is a nonnegative definite matrix. When is not differentiable at , is not a singleton.
Definition 3.
A generated Jacobian equation is an equation of the form (1) where the mapping derives from a generating function as in the A1 condition.
For a GJE to make sense we must have . By calculations which are now standard [19], a solution of (1) satisfies the Monge–Ampère type equation
(6) |
where are defined on by
(7) | ||||
(8) |
This equation is degenerate elliptic provided is -convex.
The following definition extends the definition of Aleksandrov solution for the Monge–Ampère equation to generated Jacobian equations.
Definition 4.
A -convex function is called an Aleksandrov solution of
(9) |
for a Borel measure on , provided for every Borel
Whilst (9) would, classically, require a function, we’ve defined Aleksandrov solutions for merely -convex functions. However it is a consequence of the change of variables formula and the Lebesgue differentiation theorem that Aleksandrov solutions are classical solutions.
We introduce one final condition on the generating function. It was introduced by Ma, Trudinger, and Wang [14] and was extended to GJEs in [19]. The necessity of the weakened form, for even regularity, was proved by Loeper [13].
A3. There is such that
for all unit vectors satisfying .
The A3 weak (A3w) condition, is the same but with .
2.1. Statement of main theorems
Theorem 1.
Let be a generating function satisfying A3. Let be a -convex Aleksandrov solution of (9) with for . Then with
Remark 1.
Liu proved this value, when , is sharp. That is, there exists a function which solves a GJE satisfying the hypothesis of the theorem, and which is in for the stated , but not in for any . We note that Loeper [13] and Jeong [8] have proved the Hölder regularity for the right-hand side a measure satisfying . Our proof is easily adapted to this condition (which is more general than above but comes at the expense of a smaller ). We indicate the necessary changes in a remark after the proof of Theorem 1.
Our estimate is for where is Dini continuous.
Definition 5 (Dini continuity).
Let . The oscillation of is
Then is called Dini continuous if
Theorem 2.
Let be a generating function satisfying A3. Let be an Aleksandrov solution of (9) with . If and is Dini continuous then . If then .
Remark 2.
Our result can be stated more precisely as follows. Under the above hypothesis for each there is depending on such that
-
(1)
If is Dini continuous we have
(10) where and .
-
(2)
If for some then with
(11)
We do not prove (10) and (11) directly — we just prove a estimate. This ensures the equation is uniformly elliptic then (10) and (11) follow from [22, Theorem 3.1] (details in Section 6).
Remark 3.
A common form of arising in applications is
Theorem 1 applies whenever and for some positive constant . Then the assumption of Hölder continuous right-hand side in Theorem 2 is satisfied when and are Hölder continuous. More precisely if then Theorem 1 implies the sharp exponent with and we subsequently obtain for .
As is standard we prove apriori estimates for smooth solutions. The results then hold by approximation and uniqueness of the Dirichlet problem in the small [19, Lemma 4.6]. The results above all hold without boundary conditions, that is they are interior local results. This is possible because we are considering Aleksandrov solutions under A3. With A3 weakened to A3w such results are not possible. Moreover for applications to optimal transport to conclude that a potential function is an Aleksandrov solution we require also a boundary condition - the second boundary value problem with a target satisfying a convexity condition [14].
3. Background results and normalization lemma
We will use a number of background results. These originally appeared in [7] though we’ll use the formulation in [18].
We assume is a generating function satisfying A3w and is a strictly -convex function with in the Aleksandrov sense. Let be given and a support at put where and define the section
which by the strict -convexity is compactly contained in for sufficiently small .
A lemma [18, Theorems 3,5], which we employ repeatedly is the following.
Lemma 1.
There exists which depend only on , such that if and is a doubling measure, then
Note the requirement that is a doubling measure is only necessary for the lower bound. Also, for depending only on . In the special case where and we have
(12) |
We introduce new coordinates
(13) |
When A3w is satisfied is convex in the coordinates [20, Lemma 2.3]. We often use this result in conjunction with the minimum ellipsoid. The minimum ellipsoid of an arbitrary open convex set is the unique ellipsoid of minimal volume, denoted , containing . It satisfies where this is dilation with respect to the centre of the ellipsoid [12, Lemma 2.1]. We assume, after a rotation and translation that the minimum ellipsoid of is
(14) |
Then elementary convex geometry implies and (12) becomes
(15) |
We also define here the good shape constant. Let be any convex set, and assume its minimum ellipsoid is given by (14). Then a good shape constant is any satisfying and the good shape constant is just , explicitly, the infimum of all good shape constants. For solutions of generated Jacobian equations the good shape constant of carries information about (see Lemmas 5 and 6).
When A3w is strengthened to A3 we have a particularly strong estimate concerning the geometry of sections and their height. It is used repeatedly throughout this paper. In optimal transport this estimate is due to Liu [10, Lemma 4] and we largely follow his proof.
Lemma 2.
Assume is a generating function satisfying A3 and is a -convex function. Assume that and that the minimum ellipsoid of (in the coordinates) is (14). Then there is depending only on such that
Proof.
We work in the coordinates, though keep the notation . Define by
(16) |
Note
(17) |
Let be the boundary point of on the negative axis. In a neighbourhood of , denoted , we represent as a graph of some function , that is
Using (17) we may assume is defined for . Similarly by (17) and the convexity of we conclude when . Let be the curve
Because (17) implies at some the proof will be complete provided we can show that
(18) |
for every .
Let be defined on by
Differentiating once, then twice, with respect to gives
(19) | ||||
(20) |
To estimate we compute at
The inequality is by -convexity of . Put . Then use the definition of (equation (7)), on , and a Taylor series to obtain
(21) | ||||
(22) |
where for some . A direct, but involved, calculation which we relegate to Appendix A.1 implies
Thus, using also , (21) becomes
Now returning to we have
(23) |
Since by (19) is orthogonal to we also have orthogonality of and . Thus employing the A3 condition in (23) yields
Now we substitute this into (20) to obtain
If we can show then we’ve obtained (18).
For this final estimate fix and set
where supports at so . A standard argument [16, Eq. A.14] using the A3w condition implies that for depending only on ,
Then we follow [17, Eq. 19] (there are similar arguments in [21, 7, 20, 19]) to obtain
for . Choosing and integrating from to we have
(24) |
Now and
where we’ve used that by the minimum ellipsoid . Thus (24) becomes . To conclude we control by . Indeed, by (19) , similar reasoning implies for . Thus
Recalling we complete the proof with the observation . ∎
4. regularity
The regularity is essentially an immediate consequence of Lemmas 1 and 2. The proof of Theorem 1 is as follows.
Proof.
Step 1. [Proof for strictly -convex functions] Fix , without loss of generality equal to , and then sufficiently small to ensure . By Lemma 1
(25) |
Then assuming the minimum ellipsoid of is given by (14), (25) becomes
Using Hölder’s inequality (with the second function equal to 1) we have
(26) |
Now we conclude as in [10] (which is where Lemma 2 is used). More precisely (26) is the 5th inequality on [10, pg. 446], so the rest of this step is exactly as given there.
Step 2. [Proof for -convex functions] When is not strictly -convex, we may consider on a small enough neighbourhood of the function . Indeed by the proof of [16, Theorem 2.22] this function is strictly -convex on a neighbourhood of depending only on (in particular, independent of and ). Moreover it is an Aleksandrov solution of a generated Jacobian equation with right-hand side in the original space. This is an consequence of the identity for the subgradient. Thus by the previous proof
as required. ∎
5. Interior Pogorelov estimate for constant right-hand side
Now we start work on the estimate. Recall we will prove this by establishing a (i.e. uniform ellipticity) estimate when the right-hand side is Dini continuous. Then we obtain the estimate from the elliptic theory [22]. The estimate for Dini continuous right hand side is a perturbation of the same result for the special case of constant right-hand side, the proof of which is the goal of this section. First we introduce a strengthening of the result and a strict -convexity estimate that holds by duality.
Lemma 3.
Let be a generating function satisfying A3 and be a -convex solution of . For each there is depending on for which the following holds. Whenever , is the -support at and there holds
(27) |
The right-hand inequality is the estimate of Guillen and Kitagawa [7] which follows from the strict convexity in [20, Lemma 4.1]. The left-hand side inequality follows from the right-hand side by duality. We give the proof in Appendix A.2.
Lemma 4.
Assume that is a -convex solution of
(28) |
where is a support at some and is a real number. We assume satisfies A3w, is sufficiently small (determined in the proof), and is the good shape constant of . For each there is depending on such that in
we have
(29) |
Proof.
This is essentially the estimate
(30) |
which was given in [20]. Since we need to ensure the constant only depends on we will provide full details. The proof is via a Pogorelov type estimate: We consider a certain test function which attains a maximum in . Our choice of test function ensures it both controls the second derivatives and is controlled at its maximum point.
We let and introduce both the function
and the differential operator111Here differentiation is with respect to , never .
(31) |
where and is evaluated at .
We use the notation and . Because the nonnegative function is on , attains an interior maximum at and some assumed without loss of generality to be . We also assume, again without loss of generality, that at is diagonal. At an interior maximum and so that
(32) |
We will compute each term in (32) and from this obtain (30).
Term 1: . This one’s immediate – provided the domain (and subsequently ) is chosen sufficiently small depending on we have
(33) |
Term 2: We compute
(34) |
We note (using ) that
(35) |
and by differentiating the PDE in the direction at
(36) |
Hence (34) becomes
(37) |
Term 3: . To begin, we differentiate the PDE twice in the direction and obtain, with the notation , that
(38) |
We use A3w to deal with the second term. Use to write
Then by applying A3w with for we see so that
Thus (38) becomes
We perform similar calculations for to obtain
(39) |
We use this to compute as follows. First,
Hence by (39)
(40) |
When in the first term and in the second these terms cancel. At the expense of an inequality we discard terms with neither nor . Thus
(41) | ||||
Rewriting the second sum in terms of the matrix yields
Now, Cauchy’s inequality implies
Thus
and on returning to (40)
(42) |
Now, using (33),(37), and (42) in (32) implies
(43) | ||||
Term 3: . First, write
(44) |
We compute (using )
(45) |
For each write
Then by A3w the first term is nonnegative, so that
Returning to (45) we see
Which into (44) implies
(46) |
Here we’ve used that we can assume , and also used Cauchy’s to note
Now we deal with the final term in (46). We can assume that , for if not we have (29) with . Since we are at a maximum , that is
This implies
Choosing and returning to (46) we obtain
Substituting into (43) completes the proof:
Take , and subsequently , small enough to ensure (our choice of will only depend on allowed quantities). A further choice of , large depending only on , and finally large depending on implies
This implies at the maximum point, and the proof is complete. ∎
6. regularity
In this section we will prove the estimate via a estimate. We adapt the method of proof used by Liu, Trudinger, and Wang in the optimal transport case [11]. We also use some details from Figalli’s exposition in the Monge–Ampère case [3, Section 4.10]. Here’s how we obtain the estimate. When the right hand side is constant the estimate is true by the interior Pogorelov’s estimate of the previous section. Then the argument for Dini-continuous is to perturb the argument for constant . That is we zoom in, treating a series of normalized approximating problems with constant right hand side.
6.1. Normalization of sections
Here we explain the procedure for normalizing a solution on a section. We assume we are given a strictly -convex function which is an Aleksandrov solution of
as well as a point and corresponding -support . As usual, we consider the section . The definition of Aleksandrov solution is coordinate independent so we may assume we are in the coordinates given by (13) and is convex. Assume the minimum ellipsoid is given by (14) and is given by (16). We want to consider the PDE solved by
(47) |
on . Importantly , , and for depending only on . Thus this is a natural generalization of the normalization procedure for Monge–Ampère equations. We show that solves a MATE
(48) |
for satisfying and for depending only on . When is defined by (47) we use the notation
and similarly for . We compute directly solves the equation (48) for
In this form we can argue exactly as in [10, pg. 440] to obtain
Then is bounded by Lemma 2 when A3 is satisfied. For the estimates for note is bounded by the strict convexity and differentiability estimate (27). Finally the pinching estimate on follows from , , and (15).
6.2. Lemmas for the estimate
Here we show we can estimate the good shape constant of sections by the second derivatives of the function and vice versa. A subtlety is that sections of different height are convex in different coordinates. We assume at the outset some initial coordinates are fixed. We will say a section has good shape constant if after performing the change of variables (13) the section has good shape constant . Note because the Jacobian of the transformation (13) is (from A2) if has good shape constant then it is still true, in the initial coordinates with respect to which may not be constant, that contains a ball of radius and is contained in a ball of radius for where are respectively the minimum and maximum eigenvalues of over .
Lemma 5.
Let be a strictly -convex solution of
(49) |
Fix and a support at . Assume , and for some . Then has a good shape constant which depends only on and the constant in the Pogorelov Lemma 4.
Proof.
We normalize the section and solution as in Section 6.1 and let the normalized solution be denoted by . By the Pogorelov interior estimates we have
for . Similarly by the PDE and we have for . Using in addition
we obtain (see [3, Eq. 4.26]) a bi-Lipschitz estimate . Since we can assume is of the form (16) we obtain , the desired good shape estimate on . ∎
Lemma 6.
Assume is a strictly -convex solution of
Fix and assume is a -support at . If there is a sequence such that each
has a good shape constant less than some then for depending on and .
Proof.
Without loss of generality and the minimum ellipsoid of has axis . Our assumption is . Then, by (15)
that is . Moreover, because is the largest axis of the minimum ellipsoid so that
(50) |
Let and denote the minimum eigenvalue of by and corresponding normalized eigenvector by . Using a Taylor series we have
Thus provided is taken sufficiently large for . That is, there is with
(51) |
Combining (50) and (51) we obtain a lower bound on which implies an upper bound on the largest eigenvalue of by the PDE. ∎
Before proving the estimate we state one final lemma.
Lemma 7.
Assume for is a solution of
where is a Hölder continuous function and For any there is depending on such that
6.3. Proof of the estimate
As stated at the start of this section Theorem 2 follows from an interior estimate. This is what we now prove.
Theorem 3.
Let be a generating function satisfying A3 and be a -convex solution of
If is Dini-continuous with then for each we have an estimate .
Proof.
Step 1. [Setup of approximating problems] At the outset we fix , where without loss of generality . Consider
where is the support at and is chosen small enough to ensure Lemma 4 applies. We normalize so that with the minimum ellipsoid. Moreover, we assume is chosen small enough to ensure that after this normalization
(52) |
for an to be chosen in the proof (recall is from Definition 5). Note such a choice of is controlled by (27) and thus up to rescaling we assume .
W introduce a sequence of approximating problems. Define the domains
and let .
Let be the solution (whose existence is guaranteed by the Perron method) of
(53) | |||
(54) |
In addition put
Using (15) the section is contained in a ball of of radius for depending on the good shape constant of . Thus when the good shape constant of is controlled which we’ll use at the conclusion of the proof. Lemmas 5 and 6 suggest the structure of our proof: It suffices to show a good shape constant of each is controlled by a fixed constant independent of and this, in turn, follows from a uniform estimate on each inside a subsection. More precisely, we’ll prove by induction that for each we have
(55) |
for any which is defined by
By the uniform estimates in [18]222Use the upper bound [18, Theorem 4] and corresponding lower bound obtained as in [4, Theorem 6.2]. there is a choice of sufficiently close to such that
(56) | ||||
(57) |
Step 2. [Induction base case: ] It is clear that (55) holds for . However we note here that is controlled by the interior Pogorelov estimate, that is in terms of and our initial normalizing transformation which is, in turn, controlled by (15).
Step 3. [Inductive step] Now we assume (55) up to some fixed . We rescale our solution and domain by introducing
where . The function solves
for and and similarly for . Note this transformation does not change the magnitude of the second derivatives. Thus the inductive hypothesis (55) and Lemma 5 implies has a good shape constant depending only on and the constant in the Pogorelov lemma. We claim that has a good shape constant depending on the same parameters and the constants in (15). To see this assume the minimum ellipsoids of and have axis and respectively. By (15) applied to the section we have
Using this to compute an upper bound on , we obtain
(58) |
where the final inequality is because is the larger section. Now, let be the good shape constant of , that is . Using (15) again, this time applied to the section , we have
(59) |
Combining (58) and (59) implies the claimed fact that when has good shape so does 333To be explicit, because we are proving (55) by induction (not a good shape estimate by induction) it does not matter that the good shape constant of is worse than ..
Noting that we obtain by Lemma 4 a estimate depending on allowed quantities and in . Then the Evans–Krylov interior estimates imply estimates in and subsequently higher estimates by the elliptic theory. Similarly for in the corresponding sections .
Linearising and using the maximum principle on small domains(see [16, Lemma A.3]) we obtain
Thus in . Then, by Lemma 7, in we have
(Note we first obtain this for then use .) We’ve used that by (57) the estimates for in hold on the entire smaller section . Since all we’ve used is the induction hypothesis we can conclude the same inequality for replaced by . Moreover because these sections have a controlled good shape constant we have
(60) |
for .
Appendix A Omitted calculations
A.1. Proof that
Let some initial coordinates, denoted , be given and define . For notation put . We compute that for in
(61) | ||||
Thus if our initial coordinates are the coordinates and
(62) |
We recall from [19, Eq. 2.3] that
(63) |
From the definition of we compute
(64) |
Applying (64) twice and using the identity for differentiating an inverse matrix gives
(65) |
Now, by direct calculation
which when substituted into (65) implies
Combined with (63) this implies (61). We note if our initial coordinates are the coordinates then
This implies (62).
A.2. Quantitative convexity via duality
Here we prove the first inequality in (27) assuming the second inequality. We follow the duality argument in [11] simplified by the transformation in [18]. Indeed by [18, Lemma 3] it suffices to prove the result at the origin for the generating function
(66) | ||||
where are functions. Moreover we assume is a strictly -convex function satisfying and . We need to prove . Throughout the proof we use the notation to denote any function satisfying an estimate on a neighbourhood of the origin for depending only on . Similarly for the notation .
The -transform of is defined by
for satisfying and the dual generating function (see [18]). The function is -convex with -support at and the result, which holds by duality, implies . Thus, by duality,
(67) |
We also note by the estimate for , , so we can assume throughout that the neighbourhood over which the supremum is taken, and subsequently , is sufficiently small. The supremum is obtained for satisfying
(68) |
Now using (66)
(69) |
Now we use (68) to estimate from below. First note
(70) |
Thus substituting into (68) and taking an inner product with we obtain (near the origin in )
and subsequently returing to (69) implies
Since is as close to as desired on a sufficiently small neighbourhood of the origin we have
However also from (68) and (70)
This completes the proof.
References
- [1] 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.
- [2] Luis A. Caffarelli. Interior estimates for solutions of the Monge-Ampère equation. Ann. of Math. (2), 131(1):135–150, 1990.
- [3] Alessio Figalli. The Monge-Ampère equation and its applications. Zurich Lectures in Advanced Mathematics. European Mathematical Society (EMS), Zürich, 2017.
- [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] David Gilbarg and Neil S. Trudinger. Elliptic partial differential equations of second order. Classics in Mathematics. Springer-Verlag, Berlin, 2001.
- [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. Pointwise estimates and regularity in geometric optics and other generated Jacobian equations. Comm. Pure Appl. Math., 70(6):1146–1220, 2017.
- [8] Seonghyeon Jeong. Local Hölder regularity of solutions to generated Jacobian equations. Pure Appl. Anal., 3(1):163–188, 2021.
- [9] Yash Jhaveri. Partial regularity of solutions to the second boundary value problem for generated Jacobian equations. Methods Appl. Anal., 24(4):445–475, 2017.
- [10] Jiakun Liu. Hölder regularity of optimal mappings in optimal transportation. Calc. Var. Partial Differential Equations, 34(4):435–451, 2009.
- [11] Jiakun Liu, Neil S. Trudinger, and Xu-Jia Wang. Interior regularity for potential functions in optimal transportation. Comm. Partial Differential Equations, 35(1):165–184, 2010.
- [12] 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.
- [13] Grégoire Loeper. On the regularity of solutions of optimal transportation problems. Acta Math., 202(2):241–283, 2009.
- [14] 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.
- [15] 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.
- [16] Cale Rankin. Regularity and uniqueness results for generated Jacobian equations. PhD thesis, Australian National University, 2021.
- [17] Cale Rankin. Strict convexity and regularity of solutions to generated Jacobian equations in dimension two. Calc. Var. Partial Differential Equations, 60(6):Paper No. 221, 14, 2021.
- [18] Cale Rankin. Strict -convexity for generated Jacobian equations with applications to global regularity. ArXiV:2111.00448, 2021.
- [19] Neil S. Trudinger. On the local theory of prescribed Jacobian equations. Discrete Contin. Dyn. Syst., 34(4):1663–1681, 2014.
- [20] Neil S. Trudinger. On the local theory of prescribed Jacobian equations revisited. Math. Eng., 3(6):Paper No. 048, 17, 2021.
- [21] Cédric Villani. Optimal transport, volume 338 of Grundlehren der mathematischen Wissenschaften. Springer-Verlag, Berlin, 2009. Old and new.
- [22] Xu-Jia Wang. Schauder estimates for elliptic and parabolic equations. Chinese Ann. Math. Ser. B, 27(6):637–642, 2006.