This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Strict convexity and C1C^{1} regularity of solutions to generated Jacobian equations in dimension two

Cale Rankin Australian National University cale.rankin@anu.edu.au
Abstract.

We present a proof of strict gg-convexity in 2D for solutions of generated Jacobian equations with a gg-Monge–Ampère measure bounded away from 0. Subsequently this implies C1C^{1} differentiability in the case of a gg-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 gg-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 35J66
This research is supported by an Australian Government Research Training Program (RTP) Scholarship.

1. Introduction

Generated Jacobian equations are PDEs of the form

(1) detDY(,u,Du)=ψ(,u,Du),\det DY(\cdot,u,Du)=\psi(\cdot,u,Du),

where the vector field YY 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 YY are given in Section 2. For now we state our purpose is to show, for generalised notions of convexity, that solutions of

(2) detDY(,u,Du)c>0\displaystyle\det DY(\cdot,u,Du)\geq c>0

on a convex domain Ω𝐑2\Omega\subset\mathbf{R}^{2} are strictly convex. By considering an analogue of the Legendre transform we obtain that when (2) is instead a bound from above and uu satisfies the second boundary value problem

(3) Y(,u,Du)(Ω)=Ω,\displaystyle Y(\cdot,u,Du)(\Omega)=\Omega^{*},

for a domain Ω𝐑2\Omega\subset\mathbf{R}^{2} and a convex domain Ω𝐑2\Omega^{*}\subset\mathbf{R}^{2}, then uu is C1C^{1}.

The class of YY 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 Y=Y(,Du)Y=Y(\cdot,Du). They proved the C1C^{1} differentiability of solutions in two dimensions under a bound from above on the cc-Monge–Ampère measure. They assumed a uniformly cc-convex source Ω\Omega and a strictly cc^{*}-convex target Ω\Omega^{*}. Thanks largely to the maturity of the relevant convexity theory we are able to reduce these convexity conditions. Our C1C^{1} result requires no convexity condition on the source and only gg^{*}-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 C1C^{1} 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 YY 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 g(x,y,z)=xyzg(x,y,z)=x\cdot y-z and g(x,y,z)=c(x,y)zg(x,y,z)=c(x,y)-z where cc is a cost function from optimal transport. In these settings gg-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 gg, and require it satisfy the following assumptions.
A0. gC4(Γ¯)g\in C^{4}(\overline{\Gamma}) where Γ𝐑n×𝐑n×𝐑\Gamma\subset\mathbf{R}^{n}\times\mathbf{R}^{n}\times\mathbf{R} is a bounded domain satisfying that the projections

Ix,y:={z;(x,y,z)Γ}I_{x,y}:=\{z;(x,y,z)\in\Gamma\}

are open intervals. Moreover we assume there are domains Ω,Ω𝐑n\Omega,\Omega^{*}\subset\mathbf{R}^{n} such that for all xΩ¯,yΩ¯x\in\overline{\Omega},y\in\overline{\Omega^{*}} we have that Ix,yI_{x,y} is nonempty.
A1. For all (x,u,p)𝒰(x,u,p)\in\mathcal{U} defined as

𝒰={(x,g(x,y,z),gx(x,y,z));(x,y,z)Γ},\mathcal{U}=\{(x,g(x,y,z),g_{x}(x,y,z));(x,y,z)\in\Gamma\},

there is a unique (x,y,z)Γ(x,y,z)\in\Gamma such that

g(x,y,z)=u and gx(x,y,z)=p.\displaystyle g(x,y,z)=u\quad\text{ and }\quad g_{x}(x,y,z)=p.

A2. On Γ¯\overline{\Gamma} there holds gz<0g_{z}<0 and the matrix E with entries

Eij:=gxi,yj(gz)1gxi,zgyj,E_{ij}:=g_{x_{i},y_{j}}-(g_{z})^{-1}g_{x_{i},z}g_{y_{j}},

satisfies detE0\det E\neq 0.

2.2. Generated Jacobian equations

Assumption A1 allows us to define mappings Y:𝒰𝐑nY:\mathcal{U}\rightarrow\mathbf{R}^{n} and Z:𝒰𝐑Z:\mathcal{U}\rightarrow\mathbf{R} by requiring they solve

(4) g(x,Y(x,u,p),Z(x,u,p))\displaystyle g(x,Y(x,u,p),Z(x,u,p)) =u,\displaystyle=u,
(5) gx(x,Y(x,u,p),Z(x,u,p))\displaystyle g_{x}(x,Y(x,u,p),Z(x,u,p)) =p.\displaystyle=p.

We call a PDE of the form

(1) detDY(,u,Du)=ψ(,u,Du),\det DY(\cdot,u,Du)=\psi(\cdot,u,Du),

a generated Jacobian equation provided YY derives from solving (4) and (5) for some generating function.

Generated Jacobian equations may be rewritten as Monge–Ampère type equations. Suppose uC2(Ω)u\in C^{2}(\Omega) satisfies (1). Then differentiating (4) and (5) evaluated at (x,Y(x,u,Du),Z(x,u,Du))(x,Y(x,u,Du),Z(x,u,Du)) yields

Du\displaystyle Du =gx+gyDY+gzDZ,\displaystyle=g_{x}+g_{y}DY+g_{z}DZ,
and D2u\displaystyle\text{ and }\quad D^{2}u =gxx+gxyDY+gxzDZ.\displaystyle=g_{xx}+g_{xy}DY+g_{xz}DZ.

Since the first equation with (5) implies

(6) DjZ=1gzgykDjYk,D_{j}Z=-\frac{1}{g_{z}}g_{y_{k}}D_{j}Y^{k},

the second can be solved for DYDY by which we have

(7) DY(,u,Du)=E1[D2ugxx(Y(,u,Du),Z(,u,Du))],\displaystyle DY(\cdot,u,Du)=E^{-1}[D^{2}u-g_{xx}(Y(\cdot,u,Du),Z(\cdot,u,Du))],

with EE as in A2. Thus C2C^{2} solutions of (1) solve

(8) det[D2uA(,u,Du)]=B(,u,Du),\det[D^{2}u-A(\cdot,u,Du)]=B(\cdot,u,Du),

for

A(,u,Du)\displaystyle A(\cdot,u,Du) =gxx(Y(,u,Du),Z(,u,Du)),\displaystyle=g_{xx}(Y(\cdot,u,Du),Z(\cdot,u,Du)),
B(,u,Du)\displaystyle B(\cdot,u,Du) =detE(,u,Du)ψ(,u,Du).\displaystyle=\det E(\cdot,u,Du)\psi(\cdot,u,Du).

This PDE is elliptic when D2u>gxx(Y(,u,Du),Z(,u,Du))D^{2}u>g_{xx}(Y(\cdot,u,Du),Z(\cdot,u,Du)) as matrices. A necessary condition for ellipticity is B>0B>0.

2.3. gg-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 u:Ω𝐑u:\Omega\rightarrow\mathbf{R} is gg-convex provided for all x0Ωx_{0}\in\Omega there is y0,z0y_{0},z_{0} such that

(9) g(x0,y0,z0)\displaystyle g(x_{0},y_{0},z_{0}) =u(x0),\displaystyle=u(x_{0}),
(10) g(x,y0,z0)\displaystyle g(x,y_{0},z_{0}) u(x) for all xx0,xΩ.\displaystyle\leq u(x)\quad\text{ for all }x\neq x_{0},x\in\Omega.

In this case g(,y0,z0)g(\cdot,y_{0},z_{0}) is called a gg-support at x0x_{0}. We say uu is strictly gg-convex if the inequality in (10) is strict, that is, any given gg-support only touches uu at a single point. Note when uu is differentiable, since u()g(,y0,z0)u(\cdot)-g(\cdot,y_{0},z_{0}) has a minimum at x0x_{0}, we have Du(x0)=gx(x0,y0,z0)Du(x_{0})=g_{x}(x_{0},y_{0},z_{0}). This combined with (9) is equivalent to (5) and (4) so that y0=Y(x0,u(x0),Du(x0))y_{0}=Y(x_{0},u(x_{0}),Du(x_{0})) and z0=Z(x0,u(x0),Du(x0))z_{0}=Z(x_{0},u(x_{0}),Du(x_{0})). Moreover, if uu is C2C^{2}, again since u()g(,y0,z0)u(\cdot)-g(\cdot,y_{0},z_{0}) has a minimum at x0x_{0}, we have D2u(x0)gxx(x0,y0,z0)D^{2}u(x_{0})\geq g_{xx}(x_{0},y_{0},z_{0}) and (8) is degenerate elliptic. However when uu is not differentiable (9) and (10) may hold for more than one y0,z0y_{0},z_{0}. The set of y0y_{0} for which there is z0z_{0} such that (9) and (10) hold is denoted Yu(x0)Y_{u}(x_{0}). Similarly for Zu(x0)Z_{u}(x_{0}). When both are a singleton we identify these sets with their single element.

2.4. Alexandrov solutions

The mapping YuY_{u} allows us to define a notion of Alexandrov solution. A gg-convex function uu is called an Alexandrov solution of (1) on Ω\Omega provided for every Borel EΩE\subset\Omega

|Yu(E)|=Eψ(x,u,Du)𝑑x.|Y_{u}(E)|=\int_{E}\psi(x,u,Du)\ dx.

Since gg-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 uu is C2C^{2} and Y(,u,Du)Y(\cdot,u,Du) is a C1C^{1} diffeomorphism Alexandrov solutions are classical solutions. Moreover in the case of inequality (2) the notion of Alexandrov solution is that for every EΩE\subset\Omega

|Yu(E)|c|E|.|Y_{u}(E)|\geq c|E|.

Here we call the measure μ\mu defined on Borel EΩE\subset\Omega by μ(E)=|Yu(E)|\mu(E)=|Y_{u}(E)| the gg-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 {xθ}θ[0,1]\{x_{\theta}\}_{\theta\in[0,1]} is called a gg-segment with respect to yΩ,zθIxθ,yy\in\Omega^{*},z\in\cap_{\theta}I_{x_{\theta},y} joining x0x_{0} to x1x_{1} provided

gygz(xθ,y,z)=θgygz(x1,y,z)+(1θ)gygz(x0,y,z).\frac{g_{y}}{g_{z}}(x_{\theta},y,z)=\theta\frac{g_{y}}{g_{z}}(x_{1},y,z)+(1-\theta)\frac{g_{y}}{g_{z}}(x_{0},y,z).

Using this we say a set Ω\Omega is gg-convex with respect to (y,z)(y,z) provided for each x0,x1Ωx_{0},x_{1}\in\Omega the above gg-segment is contained in Ω\Omega. The gg-segment is unique via condition A1 in Section 2.7. If A𝐑nA\subset\mathbf{R}^{n} and B𝐑B\subset\mathbf{R} we say Ω\Omega is gg-convex with respect to A×BA\times B if Ω\Omega is gg-convex with respect to every (y,z)A×B.(y,z)\in A\times B.

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 gg is said to satisfy the condition A3w provided for all ξ,η𝐑n\xi,\eta\in\mathbf{R}^{n} with ξη=0\xi\cdot\eta=0 and (x,u,p)𝒰(x,u,p)\in\mathcal{U} there holds

(11) DpkplAij(x,u,p)ξiξjηkηl0.D_{p_{k}p_{l}}A_{ij}(x,u,p)\xi_{i}\xi_{j}\eta_{k}\eta_{l}\geq 0.

A4w. The matrix AA is non-decreasing in uu, in the sense that for any ξ𝐑n\xi\in\mathbf{R}^{n}

DuAij(x,u,p)ξiξj0.D_{u}A_{ij}(x,u,p)\xi_{i}\xi_{j}\geq 0.

In light of Lemma 13) we regard A3w and A4w as tools for controlling how gg-convex functions separate from their supporting hyperplanes.

2.7. The dual generating function

Strictly convex functions have C1C^{1} Legendre transforms. This provides a useful technique for proving solutions of Monge–Ampère inequalities are C1C^{1}. The same technique in the gg-convex case requires the gg^{*}-transform which is defined in terms of the dual generating function. We introduce the dual generating function here and the gg^{*} transform in Section 5. We set

Γ={(x,y,g(x,y,z));(x,y,z)Γ}.\Gamma^{*}=\{(x,y,g(x,y,z));(x,y,z)\in\Gamma\}.

The dual generating function, gg^{*}, is the unique function defined on Γ\Gamma^{*} by

(12) g(x,y,g(x,y,u))=u.\displaystyle g(x,y,g^{*}(x,y,u))=u.

It follows that if (x,y,z)Γ(x,y,z)\in\Gamma then

(13) g(x,y,g(x,y,z))=z.\displaystyle g^{*}(x,y,g(x,y,z))=z.

Further, if uu is gg-convex and yYu(x0)y\in Y_{u}(x_{0}) then the corresponding support is g(,y0,g(x0,y0,u(x0)))g(\cdot,y_{0},g^{*}(x_{0},y_{0},u(x_{0}))).

We introduce dual conditions on gg^{*}.
A1. For all (y,z,q)𝒱(y,z,q)\in\mathcal{V} defined as

𝒱={(y,g(x,y,u),gy(x,y,u));(x,y,u)Γ},\mathcal{V}=\{(y,g^{*}(x,y,u),g_{y}^{*}(x,y,u));(x,y,u)\in\Gamma^{*}\},

there is a unique (x,y,u)Γ(x,y,u)\in\Gamma^{*} such that

g(x,y,u)=z and gy(x,y,u)=q.\displaystyle g^{*}(x,y,u)=z\quad\text{ and }\quad g_{y}^{*}(x,y,u)=q.

Moreover we define mapping X(y,z,q),U(y,z,q)X(y,z,q),U(y,z,q) as the unique x,ux,u that satisfy these equations (cf. (4) and (5)). As remarked in Section 2.5 A1 implies uniqueness of the gg-segment between two points. For this note by differentiating (13) we obtain gygz(x,y,z)=gy(x,y,g(x,y,z))\frac{g_{y}}{g_{z}}(x,y,z)=-g^{*}_{y}(x,y,g(x,y,z)). The right hand side is injective in xx for fixed y,zy,z.

We define dual objects by swapping the roles of xx and yy as well as zz and uu. In particular gg^{*}-convex functions are those defined on Ω\Omega^{*} with gg^{*} supports, and for a gg^{*}-convex vv we have the mappings Xv(),Uv()X_{v}(\cdot),U_{v}(\cdot) which are analogues of Yu,ZuY_{u},Z_{u}. Similarly gg^{*}-segments are used to define gg^{*}-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 gg be a generating function satisfying A0, A1, A2, A1, A3w and A4w. Let Ω𝐑2\Omega\subset\mathbf{R}^{2} and u:Ω𝐑u:\Omega\rightarrow\mathbf{R} be a gg-convex function satisfying for c>0c>0 and all EΩE\subset\Omega

(14) |Yu(E)|c|E|.\displaystyle|Y_{u}(E)|\geq c|E|.

If Ω\Omega is gg-convex with respect to Yu(Ω)×Zu(Ω)Y_{u}(\Omega)\times Z_{u}(\Omega) then uu is strictly gg-convex.

Corollary 1.

Let gg be a generating function satisfying A0, A1, A2, A1, A3w, and A4w. Let Ω𝐑2\Omega\subset\mathbf{R}^{2} and u:Ω𝐑u:\Omega\rightarrow\mathbf{R} be a gg-convex function satisfying that for C>0C>0 and all EΩE\subset\Omega

(15) |Yu(E)|C|E|,\displaystyle|Y_{u}(E)|\leq C|E|,
(16) Yu(Ω)=Ω.\displaystyle Y_{u}(\Omega)=\Omega^{*}.

If Ω\Omega^{*} is gg^{*}-convex with respect to Ω×u(Ω)\Omega\times u(\Omega) then uC1(Ω)u\in C^{1}(\Omega).

Note conditions A4w and A4w are always satisfied in the optimal transport case. Indeed Corollary 1 implies the C1C^{1} 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 cc^{*}-convex, and the A3w condition is satisfied. These conditions are all known to be necessary. Necessity of the cc^{*}-convexity is due to Ma, Trudinger, and Wang [16] whilst necessity of A3w is due to Loeper [15].

The proof of Theorem 1 is given in Section 4. Finally the proof of Corollary 1 (which follows from Theorem 1) is given in Section 5.

3. Main lemma

We make frequent use of a differential inequality for the difference of a gg-convex function and a gg-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 gg be a generating function satisfying A0, A1 and A2. Let y0Ω,z0xΩIx,yy_{0}\in\Omega^{*},z_{0}\in\cap_{x\in\Omega}I_{x,y} be given, {xθ}\{x_{\theta}\} a gg-segment with respect to y0,z0y_{0},z_{0}, and uu a C2C^{2} gg-convex function. Then

h(θ):=u(xθ)g(xθ,y0,z0),h(\theta):=u(x_{\theta})-g(x_{\theta},y_{0},z_{0}),

satisfies

(17) d2dθ2h(θ)\displaystyle\frac{d^{2}}{d\theta^{2}}h(\theta) [Diju(xθ)gij(xθ,Yu(xθ),Zu(xθ))](xθ)˙i(xθ)˙j\displaystyle\geq[D_{ij}u(x_{\theta})-g_{ij}(x_{\theta},Y_{u}(x_{\theta}),Z_{u}(x_{\theta}))]\dot{(x_{\theta})}_{i}\dot{(x_{\theta})}_{j}
+DpkplAij(xθ)˙i(xθ)˙jDkh(θ)Dlh(θ)\displaystyle\quad+D_{p_{k}p_{l}}A_{ij}\dot{(x_{\theta})}_{i}\dot{(x_{\theta})}_{j}D_{k}h(\theta)D_{l}h(\theta)
Aij,u(xθ)˙i(xθ)˙jh(θ)K|h(θ)|,\displaystyle\quad\quad A_{ij,u}\dot{(x_{\theta})}_{i}\dot{(x_{\theta})}_{j}h(\theta)-K|h^{\prime}(\theta)|,

where KK depends only on the values of gg and its derivatives on (xθ,y0,z0)(x_{\theta},y_{0},z_{0}) and xθ˙=ddθxθ\dot{x_{\theta}}=\frac{d}{d\theta}x_{\theta}. We’ve used the shorthand Dkh(θ)=uk(xθ)gk(xθ,y0,z0)D_{k}h(\theta)=u_{k}(x_{\theta})-g_{k}(x_{\theta},y_{0},z_{0}). The arguments of Aij,uA_{ij,u} and DpkplAijD_{p_{k}p_{l}}A_{ij} 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 DpkplAijξiξjηkηlD_{p_{k}p_{l}}A_{ij}\xi_{i}\xi_{j}\eta_{k}\eta_{l} for arbitrary ξ,η\xi,\eta. We claim if ξ,η𝐑n\xi,\eta\in\mathbf{R}^{n} are arbitrary then A3w implies

(18) DpkplAijξiξjηkηlK|ξ||η||ξη|.D_{p_{k}p_{l}}A_{ij}\xi_{i}\xi_{j}\eta_{k}\eta_{l}\geq-K|\xi||\eta||\xi\cdot\eta|.

Here KK is a non-negative constant depending only on Dp2AC0\|D_{p^{2}}A\|_{C^{0}}. To obtain (18) from A3w first prove it for arbitrary unit vectors ξ,η\xi,\eta by using A3w with ξ\xi as given and η\eta replaced by ηξηξ\eta-\xi\cdot\eta\xi.

Thus when A3w,A4wA3w,A4w are satisfied and h(θ)0h(\theta)\geq 0 we have

h′′K|Dugx||xθ˙||h|C|h|,h^{\prime\prime}\geq-K|Du-g_{x}||\dot{x_{\theta}}||h^{\prime}|\geq-C|h^{\prime}|,

where CC depends on sup|Du|\sup|Du| and |xθ˙|.|\dot{x_{\theta}}|. This inequality appears in [8, Lemma 9.3],[22, pg. 310, pg. 315] and [18, 19]. It implies, amongst other things, estimates on hh^{\prime} in terms of sup|h|\sup|h| via the following lemma.

Lemma 2.

Let hC2([a,b])h\in C^{2}([a,b]) be a function satisfying h′′K|h|h^{\prime\prime}\geq-K|h^{\prime}|. For t(a,b)t\in(a,b) there holds

C0sup[a,t]|h|h(t)C1sup[t,b]|h|,-C_{0}\sup_{[a,t]}|h|\leq h^{\prime}(t)\leq C_{1}\sup_{[t,b]}|h|,

where C0,C1C_{0},C_{1} depend on ta,btt-a,b-t respectively and KK.

Proof.

First note if h(τ)=0h^{\prime}(\tau)=0 at any τ(a,b)\tau\in(a,b) then h(a)0h^{\prime}(a)\leq 0. To see this assume τ\tau is the infimum of points with h(τ)=0h^{\prime}(\tau)=0. By continuity if τ=a\tau=a we are done. Otherwise hh^{\prime} is single signed on (a,τ)(a,\tau) and if h<0h^{\prime}<0 on this interval then again by continuity we are done. Thus we assume h>0h^{\prime}>0 on (a,τ)(a,\tau). The inequality h′′K|h|h^{\prime\prime}\geq-K|h^{\prime}| implies ddtlog(h(t))K\frac{d}{dt}\log(h^{\prime}(t))\geq-K on (a,τ)(a,\tau). Subsequently for a<t1<t2<τa<t_{1}<t_{2}<\tau integration gives

(19) h(t1)eK(t2t1)h(t2),h^{\prime}(t_{1})\leq e^{K(t_{2}-t_{1})}h^{\prime}(t_{2}),

and sending t1a,t2τt_{1}\rightarrow a,t_{2}\rightarrow\tau gives h(a)0h^{\prime}(a)\leq 0.

Now to prove the inequality in the lemma let’s deal with the upper bound first. We assume h>0h^{\prime}>0 on [t,b][t,b] otherwise the argument just given implies h(t)0h^{\prime}(t)\leq 0. We obtain (19) for t1=tt_{1}=t and t2(t,b)t_{2}\in(t,b). Integrating with respect to t2t_{2} from tt to bb establishes the result. The other inequality follows by applying the same argument to the function kk defined by k(t):=h(t)k(t):=h(-t). ∎

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 gg-convexity estimate for C2C^{2} solutions of det[D2uA(,u,Du)]c\det[D^{2}u-A(\cdot,u,Du)]\geq c (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 C2C^{2} solutions
Initially we assume uu is C2C^{2}. Let g(,y0,z0)g(\cdot,y_{0},z_{0}) satisfy ug(,y0,z0)u\geq g(\cdot,y_{0},z_{0}) for y0Yu(Ω),z0Zu(Ω)y_{0}\in Y_{u}(\Omega),z_{0}\in Z_{u}(\Omega). Assume for some σ0\sigma\geq 0 there is distinct x1,x1Ωx_{-1},x_{1}\in\Omega with

(20) u(x1)g(x1,y0,z0)+σ,\displaystyle u(x_{-1})\leq g(x_{-1},y_{0},z_{0})+\sigma,
(21) u(x1)g(x1,y0,z0)+σ.\displaystyle u(x_{1})\leq g(x_{1},y_{0},z_{0})+\sigma.

Let {xθ}θ[1,1]\{x_{\theta}\}_{\theta\in[-1,1]} denote the gg-segment with respect to y0,z0y_{0},z_{0} that joins x1x_{-1} to x1x_{1} and set

hσ(x)=u(x)g(x,y0,z0)σ.h_{\sigma}(x)=u(x)-g(x,y_{0},z_{0})-\sigma.

We use the shorthand hσ(θ)=hσ(xθ)h_{\sigma}(\theta)=h_{\sigma}(x_{\theta}). Lemma 1 along with A3w and A4w yields h0′′(θ)K|h0(θ)|h^{\prime\prime}_{0}(\theta)\geq-K|h_{0}^{\prime}(\theta)| with the same inequality holding for hσh_{\sigma}. Hence, via the maximum principle, hσh_{\sigma} is less than or equal to its value at the end points. Thus

(22) 0hσ(θ)infθ[1,1]hσ(θ)=:H.\displaystyle 0\geq h_{\sigma}(\theta)\geq\inf_{\theta\in[-1,1]}h_{\sigma}(\theta)=:-H.

The convexity estimate we intend to derive is HC>0H\geq C>0 where CC depends only on |x0x1|,uC1,g,c|x_{0}-x_{1}|,\ \|u\|_{C^{1}},\ g,\ c. We use CC to indicate any positive constant depending only on these quantities.

Now, via semi-convexity, gg-convex functions are locally Lipschitz. Thus for θ[3/4,3/4]\theta\in[-3/4,3/4] and ξ𝐑n\xi\in\mathbf{R}^{n} sufficiently small, (22) implies

(23) hσ(xθ+ξ)\displaystyle h_{\sigma}(x_{\theta}+\xi) hσ(xθ)+C|ξ|C|ξ|,\displaystyle\leq h_{\sigma}(x_{\theta})+C|\xi|\leq C|\xi|,
(24) hσ(xθ+ξ)\displaystyle h_{\sigma}(x_{\theta}+\xi) hσ(xθ)C|ξ|HC|ξ|.\displaystyle\geq h_{\sigma}(x_{\theta})-C|\xi|\geq-H-C|\xi|.

We let ηθ\eta_{\theta} be a continuous unit normal vector field to the gg-segment {xθ}\{x_{\theta}\}. Fix δ>0\delta>0 so that x1/2+δη1/2x_{-1/2}+\delta\eta_{-1/2} and x1/2+δη1/2x_{1/2}+\delta\eta_{1/2} lie in Ω\Omega. For ε[0,δ]\varepsilon\in[0,\delta] let {xθε}θ[1/2,1/2]\{x^{\varepsilon}_{\theta}\}_{\theta\in[-1/2,1/2]} be the gg-segment with respect to y0,z0y_{0},z_{0} joining x1/2+εη1/2x_{-1/2}+\varepsilon\eta_{-1/2} to x1/2+εη1/2x_{1/2}+\varepsilon\eta_{1/2}. Using Lemma 2 for θ[1/4,1/4]\theta\in[-1/4,1/4] combined with (23) and (24) implies

(25) C(ε+H)ddθhσ(xθε)C(ε+H).\displaystyle-C(\varepsilon+H)\leq\frac{d}{d\theta}h_{\sigma}(x_{\theta}^{\varepsilon})\leq C(\varepsilon+H).

Here we have used that |xθxθε|<Cε|x_{\theta}-x_{\theta}^{\varepsilon}|<C\varepsilon for a Lipschitz constant independent of θ\theta.

This implies

(26) 1/41/4d2dθ2hσ(xθε)𝑑θC(ε+H),\displaystyle\int_{-1/4}^{1/4}\frac{d^{2}}{d\theta^{2}}h_{\sigma}(x_{\theta}^{\varepsilon})\ d\theta\leq C(\varepsilon+H),

and we come back to this in a moment. For now note that, since

det[D2uA(,u,Du)]cinfdetE>0,\det[D^{2}u-A(\cdot,u,Du)]\geq c\inf\det E>0,

for any two orthogonal unit vectors ξ,η\xi,\eta

[Dξξugξξ(x,Yu(x),Zu(x))][Dηηugηη(x,Yu(x),Zu(x))]C.[D_{\xi\xi}u-g_{\xi\xi}(x,Y_{u}(x),Z_{u}(x))][D_{\eta\eta}u-g_{\eta\eta}(x,Y_{u}(x),Z_{u}(x))]\geq C.

In particular for ηθε\eta_{\theta}^{\varepsilon} a choice of unit normal vector field orthogonal to x˙θε\dot{x}_{\theta}^{\varepsilon} (continuous in θ,ε\theta,\varepsilon) we have

C1\displaystyle C^{-1} [Dx˙θεx˙θεugx˙θεx˙θε(x,Yu(x),Zu(x))]\displaystyle[D_{\dot{x}_{\theta}^{\varepsilon}\dot{x}_{\theta}^{\varepsilon}}u-g_{\dot{x}_{\theta}^{\varepsilon}\dot{x}_{\theta}^{\varepsilon}}(x,Y_{u}(x),Z_{u}(x))]
|x˙θε|2[Dηθεηθεugηθεηθε(x,Yu(x),Zu(x))]1.\displaystyle\geq|\dot{x}_{\theta}^{\varepsilon}|^{2}[D_{\eta_{\theta}^{\varepsilon}\eta_{\theta}^{\varepsilon}}u-g_{\eta_{\theta}^{\varepsilon}\eta_{\theta}^{\varepsilon}}(x,Y_{u}(x),Z_{u}(x))]^{-1}.

Employing this and (25) in Lemma 1 gives

(27) d2dθ2hσ(xθε)\displaystyle\frac{d^{2}}{d\theta^{2}}h_{\sigma}(x_{\theta}^{\varepsilon}) C|x˙θε|2([Dηθεηθεu(xθε)gηθεηθε(xθε,Yu(xθε),Zu(xθε))])1\displaystyle\geq C|\dot{x}_{\theta}^{\varepsilon}|^{2}\big{(}[D_{\eta_{\theta}^{\varepsilon}\eta_{\theta}^{\varepsilon}}u(x_{\theta}^{\varepsilon})-g_{\eta_{\theta}^{\varepsilon}\eta_{\theta}^{\varepsilon}}(x_{\theta}^{\varepsilon},Y_{u}(x_{\theta}^{\varepsilon}),Z_{u}(x_{\theta}^{\varepsilon}))]\big{)}^{-1}
C(ε+H),\displaystyle\quad\quad-C(\varepsilon+H),

where initially this holds for h0h_{0}, and thus for hσh_{\sigma}. Note that by (18) the Dp2AD_{p}^{2}A term in Lemma 1 is bounded below by C|h|-C|h^{\prime}| and subsequently controlled by (25).

Substituting (27) into (26) we obtain

1/41/4|x˙θε|2([Diju(xθε)gij(xθε)](ηθε)i(ηθε)j)1𝑑θC(ε+H),\displaystyle\int_{-1/4}^{1/4}|\dot{x}_{\theta}^{\varepsilon}|^{2}\big{(}[D_{ij}u(x_{\theta}^{\varepsilon})-g_{ij}(x_{\theta}^{\varepsilon})](\eta_{\theta}^{\varepsilon})_{i}(\eta_{\theta}^{\varepsilon})_{j}\big{)}^{-1}\ d\theta\leq C(\varepsilon+H),

where we omit that gg is evaluated at (xθε,Yu(xθε),Zu(xθε))(x_{\theta}^{\varepsilon},Y_{u}(x_{\theta}^{\varepsilon}),Z_{u}(x_{\theta}^{\varepsilon})). An application of Jensen’s inequality implies

(28) 0δ1/41/4|x˙θε|2[Diju(xθε)gij(xθε)](ηθε)i(ηθε)j𝑑θ𝑑ε\displaystyle\int_{0}^{\delta}\int_{-1/4}^{1/4}|\dot{x}_{\theta}^{\varepsilon}|^{-2}[D_{ij}u(x_{\theta}^{\varepsilon})-g_{ij}(x_{\theta}^{\varepsilon})](\eta_{\theta}^{\varepsilon})_{i}(\eta_{\theta}^{\varepsilon})_{j}\ d\theta\ d\varepsilon
C0δ(1/41/4|x˙θε|2([Diju(xθε)gij(xθε)](ηθε)i(ηθε)j)1𝑑θ)1𝑑ε\displaystyle\geq C\int_{0}^{\delta}\Big{(}\int_{-1/4}^{1/4}|\dot{x}_{\theta}^{\varepsilon}|^{2}\big{(}[D_{ij}u(x_{\theta}^{\varepsilon})-g_{ij}(x_{\theta}^{\varepsilon})](\eta_{\theta}^{\varepsilon})_{i}(\eta_{\theta}^{\varepsilon})_{j}\big{)}^{-1}\ d\theta\Big{)}^{-1}\ d\varepsilon
(29) 0δCε+H𝑑ε.\displaystyle\quad\geq\int_{0}^{\delta}\frac{C}{\varepsilon+H}\ d\varepsilon.

This is the crux of the proof complete: the only way for the final integral to be bounded is if HH is bounded away from 0. We’re left to show the integral (28) is bounded in terms of the allowed quantities, and approximate when uu is not C2C^{2}.

To bound (28) use that detE0\det E\neq 0 implies |x˙θε||\dot{x}_{\theta}^{\varepsilon}| is bounded below by a positive constant depending on |x1x0||x_{1}-x_{0}| and gg. This gives the estimate

0δ1/41/4|x˙θε|2[Diju(xθε)gij(xθε)](ηθε)i(ηθε)j𝑑θ𝑑ε\displaystyle\int_{0}^{\delta}\int_{-1/4}^{1/4}|\dot{x}_{\theta}^{\varepsilon}|^{-2}[D_{ij}u(x_{\theta}^{\varepsilon})-g_{ij}(x_{\theta}^{\varepsilon})](\eta_{\theta}^{\varepsilon})_{i}(\eta_{\theta}^{\varepsilon})_{j}\ d\theta\ d\varepsilon
C0δ1/41/4[Diju(xθε)gij(xθε)](ηθε)i(ηθε)j𝑑θ𝑑ε\displaystyle\leq C\int_{0}^{\delta}\int_{-1/4}^{1/4}[D_{ij}u(x_{\theta}^{\varepsilon})-g_{ij}(x_{\theta}^{\varepsilon})](\eta_{\theta}^{\varepsilon})_{i}(\eta_{\theta}^{\varepsilon})_{j}\ d\theta\ d\varepsilon
C0δ1/41/4iDiiu(xθε)gii(xθε)dθdε.\displaystyle\leq C\int_{0}^{\delta}\int_{-1/4}^{1/4}\sum_{i}D_{ii}u(x_{\theta}^{\varepsilon})-g_{ii}(x_{\theta}^{\varepsilon})\ d\theta\ d\varepsilon.

The final line is obtained using positivity of DijugijD_{ij}u-g_{ij} and is bounded in terms of gC2\|g\|_{C^{2}} and sup|Du|\sup|Du| (compute the integral in the Cartesian coordinates and note the Jacobian for this transformation is bounded). Thus returning to (28) and (29) we obtain H>CH>C where CC 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 uu is an Alexandrov solution of (14) that is not strictly convex. There is a support g(,y0,z0)g(\cdot,y_{0},z_{0}) touching at two points x1,x1x_{1},x_{-1}. Using [18, Lemma 2.3] we also have ug(,y0,z0)u\equiv g(\cdot,y_{0},z_{0}) along the gg-segment joining these points (with respect to y0,z0y_{0},z_{0}). Balls with sufficiently small radius are gg-convex. This follows because, as noted in [14, §2.2], gg-convexity requires the boundary curvatures minus a function depending only on gC3\|g\|_{C^{3}} are positive. Thus we assume x1,x1x_{1},x_{-1} are sufficiently close to ensure that BB, the ball with radius |x1x1|/2|x_{1}-x_{-1}|/2 and centre (x1+x1)/2(x_{1}+x_{-1})/2 is gg-convex with respect to Yu(Ω)×Zu(Ω)Y_{u}(\Omega)\times Z_{u}(\Omega). Let ε>0\varepsilon>0 be given, and let uhu_{h} be the mollification of uu with hh taken small enough to ensure |uuh|<ε/2|u-u_{h}|<\varepsilon/2 on B.\partial B. The Dirichlet theory for GJE, [18, Lemma 4.6], yields a C3C^{3} solution of

detDY(,vh,Dvh)=c/2 in B,\displaystyle\det DY(\cdot,v_{h},Dv_{h})=c/2\text{ in }B,
vh=uh+ε on B,\displaystyle v_{h}=u_{h}+\varepsilon\text{ on }\partial B,

satisfying an estimate |Dvh|K|Dv_{h}|\leq K where KK depends on the local Lipschitz constant of uu. Since vhuv_{h}\geq u on B\partial B the comparison principle, ([18, Lemma 4.4]) implies vhuv_{h}\geq u in BB. Thus we can apply our previous argument and obtain strict gg-convexity of vhv_{h} provided we note (20) and (21) are satisfied for σ=2ε\sigma=2\varepsilon. Hence at xθεx_{\theta_{\varepsilon}} a point on the gg-segment where the infimum defining HH is obtained we have

u(xθε)g(xθε,y0,z0)2εvh(xθε)g(xθε,y0,z0)2εH<0.u(x_{\theta_{\varepsilon}})-g(x_{\theta_{\varepsilon}},y_{0},z_{0})-2\varepsilon\leq v_{h}(x_{\theta_{\varepsilon}})-g(x_{\theta_{\varepsilon}},y_{0},z_{0})-2\varepsilon\leq-H<0.

As ε0\varepsilon\rightarrow 0 we contradict that g(,y0,z0)g(\cdot,y_{0},z_{0}) supports uu. ∎

5. C1C^{1} regularity

For a gg-convex function defined on Ω\Omega with Yu(Ω)=ΩY_{u}(\Omega)=\Omega^{*} its gg^{*}-transform is the function defined on Ω\Omega^{*} by

v(y):=supxΩg(x,y,u(x)).v(y):=\sup_{x\in\Omega}g^{*}(x,y,u(x)).

We list a few essential properties. Let us suppose y0Yu(x0)y_{0}\in Y_{u}(x_{0}). This means

u(x)g(x,y0,g(x0,y0,u(x0))),\displaystyle u(x)\geq g(x,y_{0},g^{*}(x_{0},y_{0},u(x_{0}))),

taking g(x,y0,)g^{*}(x,y_{0},\cdot) of both sides and using gu<0g^{*}_{u}<0 yields

g(x,y0,u(x))g(x0,y0,u(x0)),\displaystyle g^{*}(x,y_{0},u(x))\leq g^{*}(x_{0},y_{0},u(x_{0})),

so that v(y0)=g(x0,y0,u(x0))v(y_{0})=g^{*}(x_{0},y_{0},u(x_{0})). The definition of vv implies for other yΩy\in\Omega^{*} we have v(y)g(x0,y,u(x0))v(y)\geq g^{*}(x_{0},y,u(x_{0})). Thus g(x0,,u(x0))g^{*}(x_{0},\cdot,u(x_{0})) is a gg^{*}-support at y0y_{0}. Which is to say x0Xv(y0)x_{0}\in X_{v}(y_{0}).

We use this as follows. Suppose in addition uu satisfies that for all EΩE\subset\Omega

|Yu(E)|c1|E|.\displaystyle|Y_{u}(E)|\leq c^{-1}|E|.

Take AΩA\subset\Omega^{*} and let EuE_{u} denote the measure 0 set of points where uu is not differentiable. Necessarily AYu(Eu)=Yu(E)A\setminus Y_{u}(E_{u})=Y_{u}(E) for some EΩE\subset\Omega. Our above reasoning implies EXv(Yu(E))E\subset X_{v}(Y_{u}(E)). Hence

(30) |Xv(A)||Xv(AYu(Eu))||E|c|A|.\displaystyle|X_{v}(A)|\geq|X_{v}(A\setminus Yu(E_{u}))|\geq|E|\geq c|A|.

Corollary 1 follows: Let uu be the function given in Corollary 1 and vv its gg^{*} transform defined on Ω\Omega^{*}. Theorem 1 holds in the dual form, that is, provided the relevant hypothesis are changed to their starred equivalents, Theorem 1 implies strict gg^{*}-convexity. Thus the hypothesis of Corollary 1 along with (30) allow us to conclude vv is strictly gg^{*}-convex.

Suppose for a contradiction uu is not C1C^{1}. Then for some x0x_{0} the set Yu(x0)Y_{u}(x_{0}) contains two distinct points, say y0,y1y_{0},y_{1}. Our above working implies g(x0,,u(x0))g^{*}(x_{0},\cdot,u(x_{0})) is a support touching at y0,y1y_{0},y_{1}. This contradicts strict gg^{*}-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 gg-segments. We suppose

(31) gygz(xθ,y0,z0)=θq1+(1θ)q0,\frac{g_{y}}{g_{z}}(x_{\theta},y_{0},z_{0})=\theta q_{1}+(1-\theta)q_{0},

and set q=q1q0q=q_{1}-q_{0}. We begin with a formula for first derivatives. Since

(32) ddθ=(xθ˙)iDxi\frac{d}{d\theta}=(\dot{x_{\theta}})_{i}D_{x_{i}}

we need to compute (xθ˙)i(\dot{x_{\theta}})_{i}. Differentiate (31) with respect to θ\theta and obtain111We use the convention that subscripts before the comma denote differentiation with respect to xx, and subscripts after the comma (which are not zz) denote differentiation with respect to yy.

[gi,mgzgi,zg,mgz2](xθ˙)i=qm,\left[\frac{g_{i,m}}{g_{z}}-\frac{g_{i,z}g_{,m}}{g_{z}^{2}}\right](\dot{x_{\theta}})_{i}=q_{m},

from which it follows that

(xθ˙)i=gzEm,iqm,(\dot{x_{\theta}})_{i}=g_{z}E^{m,i}q_{m},

where Em,iE^{m,i} denotes the m,ith{}^{\text{th}} entry of E1E^{-1}. Thus (32) becomes

(33) ddθ=gzEm,iqmDxi.\frac{d}{d\theta}=g_{z}E^{m,i}q_{m}D_{x_{i}}.

Using this expression to compute second derivatives we have

d2dθ2\displaystyle\frac{d^{2}}{d\theta^{2}} =gzEn,jDxj(gzEm,iDxi)qmqn\displaystyle=g_{z}E^{n,j}D_{x_{j}}(g_{z}E^{m,i}D_{x_{i}})q_{m}q_{n}
=gz2En,jEm,iqmqnDxixj+gz2qmqnEn,jDxj(Em,i)Dxi\displaystyle=g_{z}^{2}E^{n,j}E^{m,i}q_{m}q_{n}D_{x_{i}x_{j}}+g_{z}^{2}q_{m}q_{n}E^{n,j}D_{x_{j}}(E^{m,i})D_{x_{i}}
+gzgj,zEn,jEm,iqmqnDxi.\displaystyle\quad\quad+g_{z}g_{j,z}E^{n,j}E^{m,i}q_{m}q_{n}D_{x_{i}}.

The formula for differentiating an inverse yields

(34) d2dθ2=(xθ˙)i(xθ˙)j\displaystyle\frac{d^{2}}{d\theta^{2}}=(\dot{x_{\theta}})_{i}(\dot{x_{\theta}})_{j} Dxixjgz2qmqnEn,jEm,aDxj(Eab)Eb,iDxi\displaystyle D_{x_{i}x_{j}}-g_{z}^{2}q_{m}q_{n}E^{n,j}E^{m,a}D_{x_{j}}(E_{ab})E^{b,i}D_{x_{i}}
+gzgj,zEn,jEm,iqmqnDxi.\displaystyle+g_{z}g_{j,z}E^{n,j}E^{m,i}q_{m}q_{n}D_{x_{i}}.

Now compute

Dxj(Eab)\displaystyle D_{x_{j}}(E_{ab}) =Dxj[ga,bga,zg,bgz]\displaystyle=D_{x_{j}}\left[g_{a,b}-\frac{g_{a,z}g_{,b}}{g_{z}}\right]
=gaj,bgaj,zg,bgzga,zgj,bgz+gj,zga,zg,bgz2\displaystyle=g_{aj,b}-\frac{g_{aj,z}g_{,b}}{g_{z}}-\frac{g_{a,z}g_{j,b}}{g_{z}}+\frac{g_{j,z}g_{a,z}g_{,b}}{g_{z}^{2}}
(35) =ga,zgzEjb+El,bDplgaj.\displaystyle=-\frac{g_{a,z}}{g_{z}}E_{jb}+E_{l,b}D_{p_{l}}g_{aj}.

Here we have used that

El,bDplgaj(,Y(,u,p),Z(,u,p))=gaj,bgaj,zg,bgz,E_{l,b}D_{p_{l}}g_{aj}(\cdot,Y(\cdot,u,p),Z(\cdot,u,p))=g_{aj,b}-\frac{g_{aj,z}g_{,b}}{g_{z}},

which follows by computing DplgajD_{p_{l}}g_{aj}, differentiating (4) with respect to pp to express ZpZ_{p} in terms of YpY_{p}, and employing Ei,j=DpjYiE^{i,j}=D_{p_{j}}Y^{i} (which is obtained via calculations similar to those for (7)).

Substitute (35) into (34) to obtain

d2dθ2\displaystyle\frac{d^{2}}{d\theta^{2}} =(xθ˙)i(xθ˙)jDxixjgz2qmqnEn,jEm,aEl,bDplgajEb,iDxi\displaystyle=(\dot{x_{\theta}})_{i}(\dot{x_{\theta}})_{j}D_{x_{i}x_{j}}-g_{z}^{2}q_{m}q_{n}E^{n,j}E^{m,a}E_{l,b}D_{p_{l}}g_{aj}E^{b,i}D_{x_{i}}
+[gzga,zEn,jEm,aEj,bEb,iDxi+gzgj,zEn,jEm,iDxi]qmqn\displaystyle\quad\quad+[g_{z}g_{a,z}E^{n,j}E^{m,a}E_{j,b}E^{b,i}D_{x_{i}+}g_{z}g_{j,z}E^{n,j}E^{m,i}D_{x_{i}}]q_{m}q_{n}
=(xθ˙)i(xθ˙)jDxixjgz2qmqnEn,jEm,aDpigajDxi\displaystyle=(\dot{x_{\theta}})_{i}(\dot{x_{\theta}})_{j}D_{x_{i}x_{j}}-g_{z}^{2}q_{m}q_{n}E^{n,j}E^{m,a}D_{p_{i}}g_{aj}D_{x_{i}}
+[gzga,zEn,iEm,aDxi+gzgj,zEn,jEm,iDxi]qmqn\displaystyle\quad\quad+[g_{z}g_{a,z}E^{n,i}E^{m,a}D_{x_{i}+}g_{z}g_{j,z}E^{n,j}E^{m,i}D_{x_{i}}]q_{m}q_{n}
=(xθ˙)i(xθ˙)j(Dxi,xjDpkgijDxk)\displaystyle=(\dot{x_{\theta}})_{i}(\dot{x_{\theta}})_{j}(D_{x_{i},x_{j}}-D_{p_{k}}g_{ij}D_{x_{k}})
+gj,z(Em,jqmddθ+En,jqnddθ),\displaystyle\quad\quad+g_{j,z}\Big{(}E^{m,j}q_{m}\frac{d}{d\theta}+E^{n,j}q_{n}\frac{d}{d\theta}\Big{)},

where in the last equality we swapped the dummy indices ii and aa on the second term to allow us to collect like terms and also used (33).

Now let’s use this identity to compute h′′(θ)h^{\prime\prime}(\theta). We have

h′′(θ)\displaystyle h^{\prime\prime}(\theta) =[Diju(xθ)gij(xθ,y0,z0)\displaystyle=\big{[}D_{ij}u(x_{\theta})-g_{ij}(x_{\theta},y_{0},z_{0})
Dpkgij(xθ,y0,z0)(Dku(xθ)Dkg(xθ,y0,z0))](xθ˙)i(xθ˙)j\displaystyle\quad-D_{p_{k}}g_{ij}(x_{\theta},y_{0},z_{0})(D_{k}u(x_{\theta})-D_{k}g(x_{\theta},y_{0},z_{0}))\big{]}(\dot{x_{\theta}})_{i}(\dot{x_{\theta}})_{j}
+gj,z(Em,jqmh+En,jqnh).\displaystyle\quad\quad+g_{j,z}(E^{m,j}q_{m}h^{\prime}+E^{n,j}q_{n}h^{\prime}).

Terms on the final line are bounded below by K|h(θ)|-K|h^{\prime}(\theta)|. Now after adding and subtracting gij(xθ,y,z)g_{ij}(x_{\theta},y,z) for y=Yu(xθ),z=Zu(xθ)y=Y_{u}(x_{\theta}),z=Z_{u}(x_{\theta}) we have

h′′(θ)\displaystyle h^{\prime\prime}(\theta) [Diju(xθ)gij(xθ,y,z)](xθ˙)i(xθ˙)j+[gij(xθ,y,z)gij(xθ,y0,z0)\displaystyle\geq\big{[}D_{ij}u(x_{\theta})-g_{ij}(x_{\theta},y,z)\big{]}(\dot{x_{\theta}})_{i}(\dot{x_{\theta}})_{j}+[g_{ij}(x_{\theta},y,z)-g_{ij}(x_{\theta},y_{0},z_{0})
Dpkgij(xθ,y0,z0)(Dku(xθ)Dkg(xθ,y0,z0))](xθ˙)i(xθ˙)j\displaystyle\quad-D_{p_{k}}g_{ij}(x_{\theta},y_{0},z_{0})(D_{k}u(x_{\theta})-D_{k}g(x_{\theta},y_{0},z_{0}))\big{]}(\dot{x_{\theta}})_{i}(\dot{x_{\theta}})_{j}
K|h(θ)|.\displaystyle\quad\quad-K|h^{\prime}(\theta)|.

Set u0=g(xθ,y0,z0),u1=u(xθ)u_{0}=g(x_{\theta},y_{0},z_{0}),\ u_{1}=u(x_{\theta}) p0=gx(xθ,y0,z0),p_{0}=g_{x}(x_{\theta},y_{0},z_{0}), and p1=Du(xθ)p_{1}=Du(x_{\theta}). Then rewriting in terms of the matrix AA we have

h′′(θ)\displaystyle h^{\prime\prime}(\theta) [Diju(xθ)gij(xθ,y,z)](xθ˙)i(xθ˙)j+[Aij(xθ,u1,p1)Aij(xθ,u0,p0)\displaystyle\geq\big{[}D_{ij}u(x_{\theta})-g_{ij}(x_{\theta},y,z)\big{]}(\dot{x_{\theta}})_{i}(\dot{x_{\theta}})_{j}+\big{[}A_{ij}(x_{\theta},u_{1},p_{1})-A_{ij}(x_{\theta},u_{0},p_{0})
DpkAij(xθ,u0,p0)(p1p0)](xθ˙)i(xθ˙)jK|h(θ)|\displaystyle\quad-D_{p_{k}}A_{ij}(x_{\theta},u_{0},p_{0})(p_{1}-p_{0})\big{]}(\dot{x_{\theta}})_{i}(\dot{x_{\theta}})_{j}-K|h^{\prime}(\theta)|
=[Diju(xθ)gij(xθ,y,z)](xθ˙)i(xθ˙)j+Aij,u(xθ,uτ,p)(u1u0)(xθ˙)i(xθ˙)j\displaystyle=\big{[}D_{ij}u(x_{\theta})-g_{ij}(x_{\theta},y,z)\big{]}(\dot{x_{\theta}})_{i}(\dot{x_{\theta}})_{j}+A_{ij,u}(x_{\theta},u_{\tau},p)(u_{1}-u_{0})(\dot{x_{\theta}})_{i}(\dot{x_{\theta}})_{j}
+[Aij(xθ,u0,p1)Aij(xθ,u0,p0)\displaystyle\quad+\big{[}A_{ij}(x_{\theta},u_{0},p_{1})-A_{ij}(x_{\theta},u_{0},p_{0})
DpkAij(xθ,u0,p0)(p1p0)](xθ˙)i(xθ˙)jK|h(θ)|\displaystyle\quad\quad-D_{p_{k}}A_{ij}(x_{\theta},u_{0},p_{0})(p_{1}-p_{0})\big{]}(\dot{x_{\theta}})_{i}(\dot{x_{\theta}})_{j}-K|h^{\prime}(\theta)|

Here uτ=τu+(1τ)u0u_{\tau}=\tau u+(1-\tau)u_{0} results from a Taylor series. Another Taylor series for f(t):=Aij(xθ,u0,tp1+(1t)p0)f(t):=A_{ij}(x_{\theta},u_{0},tp_{1}+(1-t)p_{0}) and we obtain

h′′(θ)\displaystyle h^{\prime\prime}(\theta) [Diju(xθ)gij(xθ,y,z)](xθ˙)i(xθ˙)j+Aij,u(u1u0)(xθ˙)i(xθ˙)j\displaystyle\geq\big{[}D_{ij}u(x_{\theta})-g_{ij}(x_{\theta},y,z)\big{]}(\dot{x_{\theta}})_{i}(\dot{x_{\theta}})_{j}+A_{ij,u}(u_{1}-u_{0})(\dot{x_{\theta}})_{i}(\dot{x_{\theta}})_{j}
K|h(θ)|+DpkplAij(xθ,u0,pt)(xθ˙)i(xθ˙)j(p1p0)k(p1p0)l.\displaystyle-K|h^{\prime}(\theta)|+D_{p_{k}p_{l}}A_{ij}(x_{\theta},u_{0},p_{t})(\dot{x_{\theta}})_{i}(\dot{x_{\theta}})_{j}(p_{1}-p_{0})_{k}(p_{1}-p_{0})_{l}.

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 C1,αC^{1,\alpha} 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. C1C^{1} 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 cc-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 0<αrts2β<0<\alpha\leq rt-s^{2}\leq\beta<\infty. 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.