Research article Special Issues

On the local theory of prescribed Jacobian equations revisited

  • In this paper we revisit our previous study of the local theory of prescribed Jacobian equations associated with generating functions, which are extensions of cost functions in the theory of optimal transportation. In particular, as foreshadowed in the earlier work, we provide details pertaining to the relaxation of a monotonicity condition in the underlying convexity theory and the consequent classical regularity. Taking advantage of recent work of Kitagawa and Guillen, we also extend our classical regularity theory to the weak form A3w of the critical matrix convexity conditions.

    Citation: Neil S. Trudinger. On the local theory of prescribed Jacobian equations revisited[J]. Mathematics in Engineering, 2021, 3(6): 1-17. doi: 10.3934/mine.2021048

    Related Papers:

    [1] Feida Jiang . Weak solutions of generated Jacobian equations. Mathematics in Engineering, 2023, 5(3): 1-20. doi: 10.3934/mine.2023064
    [2] Pengfei Guan . A weighted gradient estimate for solutions of Lp Christoffel-Minkowski problem. Mathematics in Engineering, 2023, 5(3): 1-14. doi: 10.3934/mine.2023067
    [3] Stefano Almi, Giuliano Lazzaroni, Ilaria Lucardesi . Crack growth by vanishing viscosity in planar elasticity. Mathematics in Engineering, 2020, 2(1): 141-173. doi: 10.3934/mine.2020008
    [4] Bin Deng, Xinan Ma . Gradient estimates for the solutions of higher order curvature equations with prescribed contact angle. Mathematics in Engineering, 2023, 5(6): 1-13. doi: 10.3934/mine.2023093
    [5] Jinghong Li, Hongyu Liu, Wing-Yan Tsui, Xianchao Wang . An inverse scattering approach for geometric body generation: a machine learning perspective. Mathematics in Engineering, 2019, 1(4): 800-823. doi: 10.3934/mine.2019.4.800
    [6] Edgard A. Pimentel, Miguel Walker . Potential estimates for fully nonlinear elliptic equations with bounded ingredients. Mathematics in Engineering, 2023, 5(3): 1-16. doi: 10.3934/mine.2023063
    [7] Giovanni Cupini, Paolo Marcellini, Elvira Mascolo . Local boundedness of weak solutions to elliptic equations with p,qgrowth. Mathematics in Engineering, 2023, 5(3): 1-28. doi: 10.3934/mine.2023065
    [8] Zaffar Mehdi Dar, M. Arrutselvi, Chandru Muthusamy, Sundararajan Natarajan, Gianmarco Manzini . Virtual element approximations of the time-fractional nonlinear convection-diffusion equation on polygonal meshes. Mathematics in Engineering, 2025, 7(2): 96-129. doi: 10.3934/mine.2025005
    [9] Marco Cirant, Kevin R. Payne . Comparison principles for viscosity solutions of elliptic branches of fully nonlinear equations independent of the gradient. Mathematics in Engineering, 2021, 3(4): 1-45. doi: 10.3934/mine.2021030
    [10] Claudia Lederman, Noemi Wolanski . Lipschitz continuity of minimizers in a problem with nonstandard growth. Mathematics in Engineering, 2021, 3(1): 1-39. doi: 10.3934/mine.2021009
  • In this paper we revisit our previous study of the local theory of prescribed Jacobian equations associated with generating functions, which are extensions of cost functions in the theory of optimal transportation. In particular, as foreshadowed in the earlier work, we provide details pertaining to the relaxation of a monotonicity condition in the underlying convexity theory and the consequent classical regularity. Taking advantage of recent work of Kitagawa and Guillen, we also extend our classical regularity theory to the weak form A3w of the critical matrix convexity conditions.



    In this paper we revisit our previous study [18] of the local theory of prescribed Jacobian equations associated with generating functions, which are extensions of cost functions in the theory of optimal transportation. In particular we elaborate further our remark there pertaining to relaxing the monotonicity condition on the matrix function A in our convexity theory, thereby enabling the use of duality properties in the ensuing convexity and local regularity theory.

    We begin by describing the class of equations under consideration, which we called generated prescribed Jacobian equations and is now typically abbreviated to just generated Jacobian equations (GJEs). Let Ω be a domain in Euclidean n-space, Rn, and Y a C1 mapping from Ω×R×Rn into Rn. The prescribed Jacobian equation (PJE), is a partial differential equation of the form,

    detDY(,u,Du)=ψ(,u,Du), (1.1)

    where ψ is a given scalar function on Ω×R×Rn and Du denotes the gradient of the function u:ΩR.

    Denoting points in Ω×R×Rn by (x,z,p), we always assume that the matrix Yp is invertible, that is detYp0, so that we may write (1.1) as a general Monge-Ampère type equation (MATE),

    det[D2uA(,u,Du)]=B(,u,Du), (1.2)

    where

    A=Y1p(Yx+Yzp),B=(detYp)1ψ. (1.3)

    A function uC2(Ω) is degenerate elliptic, (elliptic), for Eq (1.2), whenever

    D2uA(,u,Du)0,(>0), (1.4)

    in Ω. If u is an elliptic solution of (1.2), then the function B(,u,Du) is positive. Accordingly we assume throughout that B is at least non-negative in Ω×R×Rn, that is ψ and detYp have the same sign.

    The second boundary value problem for the prescribed Jacobian equation is to prescribe the image,

    Tu(Ω):=Y(,u,Du)(Ω)=Ω, (1.5)

    where Ω is another given domain in Rn. When ψ is separable, in the sense that

    |ψ(x,z,p)|=f(x)/fY(x,z,p), (1.6)

    for positive f,fL1(Ω), L1(Ω) respectively, then a necessary condition for the existence of an elliptic solution, for which the mapping T is a diffeomorphism, to the second boundary value problem (1.1), (1.5) is the mass balance condition,

    Ωf=Ωf. (1.7)

    We note that Y need only be defined on the one jet of u, J1=J1[u](Ω)=(,u,Du)(Ω) in order to formulate (1.1) and (1.5) and typically we will only have Y defined on an open set URn×R×Rn with the resultant Monge-Ampère type Eq (1.2), accompanied by a constraint, J1[u]U.

    We will also change our notation slightly from [18] and let gC2(Γ) denote a generating function, where Γ is a domain in Rn×Rn×R whose projections,

    I(x,y)={zR| (x,y,z)Γ},

    are open intervals. For convenience, we also denote the following projections,

    Γx={(y,z)Rn×R| (x,y,z)Γ},Γy,z={xRn| (x,y,z)Γ}.

    Note that these projections may be empty for some values of x,y and z.

    Denoting points in Rn×Rn×R, by (x,y,z), we assume that gz0 in Γ, together with the following two conditions which extend the corresponding conditions in the optimal transportation case [14]:

    A1: The mapping (gx,g)(x,,) is one-to-one in Γx, for each xRn.

    A2: detE0 in Γ, where E is the n×n matrix given by

    E=[Ei,j]=gx,y(gz)1gx,zgy. (1.8)

    From A1 and A2, the vector field Y, together with the dual function Z, are generated by g through the equations,

    gx(x,Y,Z)=p,g(x,Y,Z)=u. (1.9)

    The Jacobian determinant of the mapping (y,z)(gx,g)(x,y,z) is gzdetE,0 by A2, so that Yand Z are accordingly C1 smooth. Also by differentiating (1.9), with respect to p, we obtain Yp=E1. Using (1.3) or differentiating (1.9) for p=Du, with respect to x, we obtain that the corresponding prescribed Jacobian Eq (1.1) is a Monge-Ampère equation of the form (1.2) with

    A(x,u,p)=gxx[x,Y(x,u,p),Z(x,u,p)],B(x,u,p)=detE(x,Y,Z)ψ(x,u,p) (1.10)

    and is well defined in domains Ω for J1=J1[u](Ω)U, where

    U={(x,u,p)Rn×R×Rnu=g(x,y,z),p=gx(x,y,z),(x,y,z)Γ}. (1.11)

    Following [18] we also have the dual condition to A1:

    A1: The mapping Q:=gy/gz(,y,z) is one-to-one in Γy,z, for all (y,z)Rn×R.

    Condition A1* arises through the notion of duality introduced in [18], where the dual generating function g is defined by

    g[x,y,g(x,y,u)]=u. (1.12)

    Clearly g is well defined on the dual set,

    Γ={(x,y,u)Rn×Rn×RuJ(x,y)},

    where J(x,y)=g(x,y,)I(x,y), and gy(x,y,u)=Q(x,y,z) for u=g(x,y,z) so that condition A1* may be equivalently expressed as the mapping gy is one-to-one in x,u for all (x,y,u)Γ. Furthermore the Jacobian matrix of the mapping xQ(x,y,z) is Et/gz, where Et denotes the transpose of E, so its determinant is automatically non-zero when condition A2 holds. From condition A1*, we then infer the existence of a C1 dual mapping X, defined uniquely by

    Q(X(y,z,q),y,z)=q (1.13)

    for all qQ(,y,z)(Γy,z). Note also that by setting

    P(x,y,u)=gx(x,y,g(x,y,u)),

    we may express condition A1 in the same form as A1*, namely the mapping P is one-to-one in y, for all (x,u) such that (x,y,u)Γ.

    In the special case of optimal transportation, we have

    g(x,y,z)=c(x,y)z,Γ=D×Rgz=1,I=I(x,y)=J(x,y)=R,E=cx,y,g(x,y,u)=c(x,y)u, (1.14)

    where D is a domain in Rn×Rn and cC2(D) is a cost function, satisfying conditions A1 and A2 in [14]. The essential difference here is that Y and A are independent of u so that our arguments here and in [18] are primarily concerned with handling such a dependence.

    As in [18] we will assume throughout that g has been normalised so that gz<0 in accordance with (1.14).

    Our next conditions extend the conditions A3 and A3w introduced for optimal transportation in [14,16,20] and are expressed in terms of the matrix function A in (1.2), which for the purpose of classical regularity is assumed twice differentiable.

    A3 (A3w)

    Aklijξiξjηkηl:=(DpkplAij)ξiξjηkηl>() 0,

    for all (x,u,p)U,ξ,ηRn such that ξ.η=0.

    Conditions A3w (A3) express a co-dimension one convexity (strict codimension one convexity) of the matrix function A with respect to the gradient variable p in the set U, which we can generally assume is convex in p for fixed x and u. As in [16], we may write equivalently that A is regular, (strictly regular), in U. It is proved in [18] that conditions A3 and A3w are invariant under duality, through explicit formulae for D2pA in terms of the generating function g and its derivatives up to order four. This result is extended to non smooth A in [13], where A co-dimension one convex (strictly co-dimension one convex) means that the form Aξ.ξ=Aijξiξj is convex, (locally uniformly convex), along line segments in p, orthogonal to ξ for all ξRn.

    In [18] we also introduced conditions expressing the monotonicity of A with respect to u, namely:

    A4 (A4w)

    DuAijξiξj>(0),

    for all (x,u,p)U,ξRn.

    Only the weak monotonicity A4w was used in [18].

    In the next section, we revisit the corresponding section in [18] and show that conditions A1, A2, A1* and A3w suffice for the convexity theory developed there, without using the monotonicity condition A4w. This will entail some upgrading of our conditions on domains, relative to Γ, but will facilitate better the use of duality in ensuing regularity arguments. In fact we had already worked out versions of these out at the time of writing [18] but omitted them in order to avoid the messier statements which pertained to ensuring the sets where condition A3w is used are contained in Γ.

    In Section 3, as foreshadowed in [18] we revisit our existence and interior regularity theory. We work with a modified version of our gradient control assumption in [18]. For this and throughout this paper, it will be convenient to fix domains U and V in Rn such that U×V×g{U×V×J(U,V)}Γ, where J(U,V)=U×VJ. For domains Ω and Ω with ˉΩU and ˉΩV, we will then assume for our existence results:

    A5: There exists an open interval J0=(m0,M0)J(U,V), m0<M0 and positive constant K0<(M0m0)/2d, d=diam(Ω), such that

    |gx(x,y,z)|K0

    for all xˉΩ,yˉΩ,g(x,y,z)J0.

    Following [18], we then have the following classical existence theorem, which improves the corresponding result in Corollary 4.7 there. For the notions of domain convexity used here, the reader is referred to [18] or Section 2 of this paper.

    Theorem 1.1. Let Ω and Ω be bounded domains in Rn, and let g be a generating function satisfying A1, A2, A1*, A3 and A5. Suppose that f>0,C1,1(Ω),f>0,C1,1(Ω), with f,1/fL(Ω),f,1/fL(Ω) and that f and f satisfy the mass balance condition (1.7). Then for any x0Ω and u0 satisfying m0+K1<u0<M0K1, K1=K0diam(Ω), there exists a g-convex, elliptic solution uC3(Ω) of the second boundary value problem (1.2), (1.5), satisfying u(x0)=u0, provided Ω is g-convex with respect to Ω×J1, where J1=(u0K1,u0+K1), and Ω is g-convex with respect to all yΩ and zg(,y,J1)(Ω).

    Theorem 1.1 is an immediate consequence of the local regularity result Theorem 3.2, which extends Theorem 4.6 in [18]. Taking account of recent developments, notably the strict convexity result in [2], we can now extend Theorem 1.1 to A3w, by extending the regularity argument in [9] for the optimal transportation case and moreover by [15] the solution is unique. We will also treat this extension in Section 3, (see Theorem 3.4 and Corollaries 3.5, 3.6), together with the necessary local Pogorelov estimate for its proof, Lemma 3.3. These results have also been presented by us in recent lectures, at Peking University in 2019 and Okinawa Institute of Science and Technology in early 2020.

    Finally in Section 4, we revisit again our convexity theory, providing an extension of Theorem 3.2 to non-smooth densities and a variant of our key convexity property Lemma 2.2, which does not need duality for its proof.

    We conclude this introduction by noting that our introduction of the concept of generating function in [17,18] was to provide a framework for extending the theory of optimal transportation to embrace near field geometric optics, where the associated ray mappings depended also on the position of a reflecting or refracting surface as well as its gradient. Particular motivation came from the point source reflection regularity theory in [7] which for graph targets is modelled by the generating function in equation (4.15) in [18]. Note that it is A in [18] (4.17) which satisfies A4 for τ<0 so the local regularity theory in this case is covered here. The reader is also referred to the papers [2,5] for further examples of generating functions in optics which fit our theory here.

    We begin by repeating the definitions in [18]. We consider bounded domains Ω and ΩRn and a generating function g, satisfying conditions A1 and A2 on Γ. For x0,y0Rn, we also denote

    I(Ω,y0)=ΩI(,y0),J(x0,Ω)=ΩJ(x0,).

    A function uC0(Ω) is called g-convex in Ω, if for each x0Ω, there exists y0Rn and z0I(Ω,y0) such that

    u(x0)=g(x0,y0,z0),u(x)g(x,y0,z0) (2.1)

    for all xΩ. If u(x)>g(x,y0,z0) for all xx0, then u is called strictly g-convex. If u is differentiable at x0, then y0=Tu(x0):=Y(x0,u(x0),Du(x0)), while if u is twice differentiable at x0, then

    D2u(x0)gxx(x0,y0,z0)=A(,u,Du)(x0) (2.2)

    that is, u is degenerate elliptic for Eq (1.4) at x0. If uC2(Ω), we call u locally g-convex in Ω if J1[u](Ω)U and (2.2) holds for all x0Ω. We will also refer to functions of the form g(,y0,z0) as g-affine and as a g-support at x0 if (2.1) is satisfied. Note also that the g-convexity of a function u in Ω implies its local semi-convexity. When the inequality in (2.2) is strict, that is u is elliptic for equation (1.4), then we call u locally uniformly g-convex. It follows readily that a locally g-convex C2 function u will also be locally g-convex in the sense that ug(,y0,z0) in some neighbourhood of x0, while a locally uniformly g-convex function is strictly g-convex in some neighbourhood of x0.

    The domain Ω is g-convex with respect to y0Rn, z0I(Ω,y0) if the image Q0(Ω):=gy/gz(,y0,z0)(Ω) is convex in Rn.

    The domain Ω is g-convex with respect to x0Rn, u0J(x0,Ω), if the image P0(Ω):=P(x0,,u0)(Ω)=gx[x0,,g(x0,,u0)](Ω) is convex in Rn.

    We may also consider a corresponding notion of domain convexity when u is fixed which agrees with that associated with the vector field Y in [16]. Namely, the domain Ω is Y-convex with respect to y0Rn, u0J(Ω,y0) if the image Q0(Ω):=Q(Ω,y0,u0)=gy/gz[,y0,g(,y0,u0)](Ω) is convex in Rn. It follows then that Ω is g-convex with respect to y0Rn, z0I(Ω,y0) if Ω is Y-convex with respect to y0 and u0=g(x,y0,z0) for every xΩ. We remind the reader that our definition of g*-convexity is already a special case of the notion of Y-convexity in [16] since P0(Ω)={pRnY(x0,u0,p)Ω}.

    It will also be convenient to introduce a more general "sub convexity" notion as follows. The domain Ω is sub g-convex with respect to y0Rn, z0I(Ω,y0) if the convex hull of Q0(Ω)Q(Γ) and the domain Ω is sub g-convex with respect to x0Rn, u0J(x0,Ω) if the convex hull of P0(Ω)P(Γ). Analogously the domain Ω is sub Y-convex with respect to y0Rn, u0J(Ω,y0) if the image Q0(Ω)Q(Γ).

    We also define the domain Ω to be g-convex (sub g-convex) with respect to a function uC0(Ω) if Ω is g-convex (sub g-convex) with respect to each point on the graph of u.

    Note that the above definitions also can be applied to general sets, in place of the domains Ω and Ω.

    Next we define the relevant notions of normal mapping and section.

    Let uC0(Ω) be g-convex in Ω. We define the g-normal mapping of u at x0Ω to be the set:

    Tu(x0)={y0RnΩΓy0,z0 and u(x)g(x,y0,z0) for all xΩ},

    where z0=g(x0,y0,u0),u0=u(x0). Clearly Tu agrees with our previous terminology when u is differentiable. In the non differentiable case we at least have the inclusion,

    Tu(x0)Σ0:=Y(x0,u(x0),u(x0)), (2.3)

    where u denotes the subdifferential of u, provided the extended one jet, J1[u](x0)=[x0,u(x0),u(x0)]U. Moreover from the semi-convexity of u, it follows that u(x0) is the convex hull of P0(Tu(x0)) and dist{Tu(x),Tu(x0)}0 as xx0.

    Next if g0=g(,y0,z0) is a g-affine function on Ω, we define the g-section S of a g-convex function u with respect to g0 by

    S=S(u,g0)=S(u,y0,z0)={xΩu(x)<g(x,y0,z0)}

    If g0 is also a g-affine support to u at x0, we define the contact set S0 by

    S0=S0(u,g0)=S0(u,y0,z0)={xΩu(x)=g(x,y0,z0)}.

    Note that we have defined sections here differently to [18].

    We now have the following variant of Lemma 2.1 in [18].

    Lemma 2.1. Assume that g satisfies A1, A2, A1* and A3w and that uC2(Ω)C0(ˉΩ) is locally g-convex in Ω and u(Ω)⊂⊂J(Ω,Tu(Ω)) with Tu(Ω) sub g-convex with respect to u. Then if Ω is g-convex with respect to (y,z) for all yTu(Ω),zg(,y,u)(Ω), it follows that u is g-convex in Ω.

    Remark 2.1. More specifically, we have for any x0Ω,y0=Tu(x0),z0=g(x0,y0,u(x0)), the g-affine function g(,y0,z0) is a g-support, provided g(,y0,z1)u in Ω for some z1<z0,I(Ω,y0) and Ω is g-convex with respect to (y0,z), for all z(z1,z0). More generally we can weaken the assumption that u(Ω)⊂⊂J(Ω,Tu(Ω)) to just requiring (,Tu(Ω),u)(ˉΩ)Γ.

    In order to prove Lemma 2.1 and the ensuing results concerning g-normal mappings and sections, we first recall a fundamental inequality from [18], for which we assume the generating function g satisfies conditions A1, A2, A1* and A3w. Let uC2(Ω) be locally g-convex in Ω and g0=g(,y0,z0) be a g-affine function defined on Ω, where the domain Ω is assumed to be g-convex with respect to (y0,z0). Defining the height function h=ug0 and making the coordinate transformation xq=Q(x,y0,z0), we then have, following the computation in (2.9) and (2.10) in [18], (without using condition A4w), the differential inequality,

    DqξqξhK|Dqξh|K0|h|, (2.4)

    for any unit vector ξ, at any point ˆxΩ⊂⊂Ω, for which (,[g0,u][u,g0],Dg0)(ˆx)U and the set Tu(Ω){y0}, is sub g-convex with respect to (,g0)(ˆx), where K and K0 are constants depending on g,g0,Ω,Ω and J1[u]. If additionally condition A4w holds, the differential inequality (2.4) holds with K0=0, for h(ˆx)0, as in inequality (2.10) in [18], while if A3w is replaced by the strict condition A3 or u is locally uniformly g-convex at ˆx, we have strict inequality in (2.4), and our proofs here can be simplified by only using the simpler inequalities

    Dqξqξh>0 (2.5)

    whenever h(ˆx)=Dqξh(ˆx)=0.

    Lemma 2.1 now follows by a modification of the proof of the corresponding lemma in [18], which we indicate after formulating the extensions to normal mappings and sections. The crucial convexity property for our regularity considerations is the characterisation of the g-normal mapping in terms of the sub differential through equality in the inclusion (2.3). In our earlier versions of this paper going back to 2014, we had various hypotheses for this result depending on whether we raise or lower g-affine functions or use duality in the proofs. The following version, depending on duality, will be convenient for our purposes here.

    Lemma 2.2. Assume g satisfies A1, A2, A1* and A3w and suppose uC0(Ω) is g-convex in Ω, with Ω sub g-convex with respect to all yΣ0, z=zy for some x0Ω. Then we have Tu(x0)=Σ0.

    We will deduce Lemma 2.2 from a convexity result for g-sections which extends Lemma 2.3 in [18].

    Lemma 2.3. Assume g satisfies A1, A2, A1* and A3w and suppose uC0(Ω) is g-convex in Ω. Assume also:

    (i) Ω is g-convex with respect to some y0Rn, z0I(Ω,y0);

    (ii) Tu(Ω){y0} is sub g-convex with respect to g0:=g(,y0,z0) in Ω.

    Then the section S=S(u,g0) is also g-convex with respect with respect to (y0,z0), while if g0 is a g-support to u, the contact set S0=S0(u,g0) is g-convex with respect with respect to (y0,z0).

    We note that Lemmas 2.2 and 2.3 are direct extensions of the corresponding lemmas in [18] under condition A4w but Lemma 2.1 needs stronger domain convexity conditions in its hypotheses. We will consider further variants of our convexity results in Section 4, including a version of Lemma 2.2 which does not need duality in its proof. We also remark here that Lemmas 2.2 and 2.3 are also extended to C2 generating functions in [13].

    Proofs. We indicate here the necessary modifications of the corresponding proofs of Lemmas 2.1 and 2.3 in [18] which follow by using the more general differential inequality (2.4) in conjunction with decreasing or increasing the appropriate g-affine functions. Letting u be locally g-convex in Ω, for a fixed point x0Ω, y0=Tu(x0), z0=g(x0,y0,u0), we modify the height function h in Eq (2.8) in [18] by setting for δ0,

    h(x)=hδ(x)=u(x)g(x,y0,z0δ). (2.6)

    As in [18], the function g0=g(,y0,z0) is a local support near x0, that is h00 near x0. If h0(x)<0 at some point xΩ, from the hypotheses of Lemma 2.1, there exists δ0 such that hδ attains a zero maximum along the closed g-segment joining x0 and x, with respect to y0,z0δ, (which will lie in some subdomain Ω⊂⊂Ω) and hδ(x)<0. Setting q0=Q(x0,y0,z0δ), qt=tq+(1t)q0, xt=X(qt,y0,z0δ), 0t1 and defining f(t)=hδ(xt), it follows that f attains a zero maximum at some point ˆt(0,1) for δ>0 and ˆt=0 for δ=0, so that also f(ˆt)=0 in both cases. From the differential inequality (2.4) and our sub g-convexity hypothesis, we have the corresponding differential inequality for f,

    fK|f|K0|f| (2.7)

    holding, at least in some neighbourhood of the set where f vanishes, which is a contradiction. The latter assertion is easily seen as (2.7) clearly implies a uniform bound from above for v where v=log(ϵf), ϵ>0, and is a one dimensional version of the strong maximum principle. Accordingly g0 must be a g-support and Lemma 2.1 is proved.

    Next we show that Lemma 2.3 also can be proved by a corresponding modification of the proof of Lemma 2.3 in [18]. By modifying Ω we may assume if necessary that u and g0 extend to a neighbourhood of ˉΩ. If S is not g-convex, with respect to y0,z0, there must be a g-segment in ˉΩ joining two points in S containing a point x1S. Setting u1=u(x1),y1Tu(x1),z1=g(x1,y1,u1), the inequality

    g(x,y1,z1)<g(x,y0,z0), (2.8)

    holds for all xS. Now replacing h in (2.6) by

    h(x)=hδ(x)=g(x,y1,z1+δ)g(x,y0,z0), (2.9)

    where g is extended so that gδ(x):=g(x,y1,z1+δ)= where (x,y1,z1+δ)Γ, we again obtain a contradiction with the differential inequality (2.4) for some δ0. Note that in this case we are lowering the support g(,y1,z1) by increasing z1 and also using condition (ii) in the sense that (x,y1,z1+δ)Γ whenever gδ=g0.

    Finally, the g-convexity of the contact set S0 follows by replacing S by S0 in the above proof, (or alternatively using S0=σ>0Sσ, where Sσ=S(u,y0,z0σ)), and hence Lemma 2.3 is proved.

    Remark 2.2. In the case when u is also a g-affine function g1=g(,y1,z1), we can improve Lemma 2.3 so that Ω can be replaced by a closed set, with possibly empty interior, and the section S replaced by the closed section ˜S=˜S(g1,g0)={xΩg1g0}, thereby inferring that ˜S is g-convex with respect to g0. In this case we can take x1Ω˜S in the proof since Tg1=y1.

    Now we can prove Lemma 2.2. Note that we can infer a version of Lemma 2.2 from Lemma 2.3 through the g-transform v, given by

    v(y)=ug(y)=supΩ g(,y,u) (2.10)

    as the g-convexity of Tu(x0) is equivalent to that of the contact set S0[v,x0,u0] of v. Here we will proceed somewhat differently with the technicalities. We begin by taking two g affine supports to u at x0, g0=g(,y0,z0),g1=g(,y1,z1) and fixing a point x=x1 in Ω, where g1g0. Letting y=yθ=Y(x0,u0,pθ), for pθ=θDg0(x0)+(1θ)Dg1(x0), θ[0,1] denote a point on the closed g-segment I0, with respect to (x0,u0), joining y0 and y1, we now define uy=g(x1,y0,zy). We then obtain, from Remark 2.2 and the g-convexity of I0, that g(x1,yθ,uθ)g(x0,y,u0)=zy and hence g(x1,y,zy)g(x1,y0,z0), for all yI0.

    Consequently we have the following extension of the Loeper maximum principle in optimal transportation [11]

    g(,y,zy)max{g0,g1} (2.11)

    in Ω for all yI0 which, by virtue of the semi-convexity of u implies that Tu(x0) is g- convex with respect to x0,u0 and hence completes the proof of Lemma 2.2.

    Remark 2.3. From the proof of Lemma 2.2 we see that condition (i) can be weakened to just requiring the pair of points {x0,x} is sub g-convex with respect to all yΣ0, z=zy, for all xΩ, that is the g-segment joining x0 to any point in Ω is well defined with respect to all yΣ0, z=zy.

    We also remark that the technicalities in using the differential inequalities (2.4) in the general A3w case can also be simplified if we use an approximation of g0 by a uniformly g-convex function, as in [6], so that we only need the simpler strict inequality f>0 from (2.5) whenever f=f=0, which requires less smoothness of g for its validity when, additionally to gC2 and A regular, either u is uniformly g-convex or A is strictly regular. In fact here the differentiability of Aξ.ξ with respect to p in directions orthogonal to ξ would suffice. However in using our differential inequality approach, (which for the optimal transportation case goes back to [8], with simplified versions in [19,22]), we would still need some smoothness of the generating function g beyond C2 smoothness. A substantially different geometric approach to our convexity theory, which has its optimal transportation roots in [21], is presented in [12,13], where we do not need any derivatives beyond second order. A different analytic approach, based on a weak form of condition A3w corresponding to a sharpening of the quasi-convexity property (2.11), is developed in [2].

    We will return to the convexity theory in Section 4, in conjunction with consideration of the strict convexity and continuous differentiability results foreshadowed in [18], which are applications of Lemma 2.3.

    Since the arguments in Section 3 of [18] are largely independent of condition A4w, they extend readily to the more general case as a consequence of Lemma 2.2. First we recall from [18] the definition of generalized solution. For convenience we let Ω and Ω be bounded domains in Rn, whose closures lie in the domains U and V respectively, as introduced in Section 1, and uC0(¯Ω) be g-convex in Ω, with Tu(Ω)⊂⊂V and conditions A1, A2, A1* satisfied. Then there is a measure μ=μ[u]=μ(u,f) on Ω, for f0L1(V), such that for any Borel set EΩ,

    μ(E)=Tu(E)f (3.1)

    which is also weakly continuous with respect to local uniform convergence. A g-convex function u on Ω is now defined to be a generalized solution of the second boundary value problem (1.5) for Eqs (1.1) and (1.6), under the mass balance condition (1.7), if

    μ[u]=νf (3.2)

    where νf=fdx and f is extended to vanish outside Ω. We then have the following extension of Theorem 3.1 in [18].

    Theorem 3.1. Let Ω and Ω be domains satisfying ˉΩU and ˉΩV and let g be a generating function satisfying A1, A2, A1* and A5. Suppose that f and f are positive densities in L1(Ω) and L1(Ω) satisfying the mass balance condition (1.7). Then for any x0Ω and u0>m0+K1, where K1=K0diam(Ω), there exists a generalized solution of (1.2), (1.5) satisfying u(x0)=u0. Furthermore if Ω is g-convex with respect to u, then any generalized solution of (1.5) satisfies Tu(Ω)ˉΩ.

    Using Lemma 2.2 in place of the corresponding Lemma 2.2 in [18], we then have the following extension of the local regularity result in Theorem 4.6 in [18].

    Theorem 3.2. Let uC0(ˉΩ) be a generalized solution of (1.5) with positive densities fC1,1(Ω),fC1,1(ˉΩ) with f,1/fL(Ω),f,1/fL(Ω) and with generating function g satisfying conditions A1, A2, A1* and A3. Suppose that u(Ω)⊂⊂J0, Ω is g-convex with respect to u and Ω is sub g-convex with respect to the dual function v=ug. Then uC3(Ω) and is an elliptic solution of (1.2), (1.10). Furthermore if Ω is g-convex with respect to v, then Tu is also a diffeomorphism from Ω to Ω, with v an elliptic solution of the dual boundary value problem.

    For the explicit form of the dual boundary value problem, the reader is referred to Eqs (3.5) and (3.7) in [18].

    Theorem 1.1 now follows as a consequence of Theorems 3.1 and 3.2, (and approximating f near Ω if only locally C1,1). From Rankin [15], the solution in Theorem 1.1 is unique. As foreshadowed in [18], we will also consider the extension of Theorem 3.2 to fC1,1(Ω) in Section 4, by adapting the strict convexity argument in [21], as this also relates to our extension to A3w.

    In the rest of this section we will treat the extension to A3w which follows for strictly convex generalized solutions from modification of the Pogorelov estimates in [10,20]. For this we consider classical elliptic solutions u of the Monge-Ampère type Eq (1.2) in sections Ω=S(u,g0), with respect to a g-affine function g0 on ˉΩ, so that we have a Dirichlet boundary condition u=g0 on Ω. Accordingly we assume A and B are C2 smooth on an open set URn×R×Rn, with A regular and B positive. We may also assume U is convex in p, for fixed x,u. We then have the following estimate:

    Lemma 3.3. Let uC4(Ω)C0,1(ˉΩ) be an elliptic solution of (1.2) in Ω and u0C2(Ω)C0,1(ˉΩ) a degenerate elliptic solution of the homogeneous equation with B=0, such that J1[u],J1[u0]U and u=u0 on Ω, u<u0 in Ω. Then there exist positive constants β, δ and C depending on n,U,A,B,u0 and |u|1=supΩ(|u|+|Du|), such that if d:=diam(Ω)<δ,

    supΩ(u0u)β|D2u|C. (3.3)

    For generated Jacobian equations, satisfying A1, A2, A1*, A3w and A4w, Lemma 3.3 corresponds to Theorem 1.2 in [5] and does not need the smallness condition on Ω. We can prove Lemma 3.3 by modification of the global estimate Theorem 3.1 in [20], incorporating an appropriate cut-off function, or by modification of the interior estimate Theorem 2.1 in [10], extending to u dependence in A. Dealing with the general u dependence is the critical issue in the treatment of the cut-off function so we will largely focus on this in the proof. For this it is more convenient for us to begin with the calculations in [20]. Accordingly with u and u0 satisfying the hypotheses of Lemma 2.3, we consider, as in [10] an auxiliary function

    v=v(,ξ)=log(wijξiξj)+τ|Du|2+κφ+βlog(u0u) (3.4)

    where |ξ|=1, τ, κ and β are positive constants to be chosen, w=D2uA(,u,Du) and φC2(ˉΩ) satisfies the global barrier condition,

    [DijφDpkAij(,u,Du)Dkφ]ξiξj|ξ|2 (3.5)

    Using the smallness condition on Ω we can fix such a barrier by taking φ=|xx1|2 for some point x1Ω.

    For the linearized operator, L given by

    L=L[u]=wij[DijDpkAij(,u,Du)Dk]DpklogB(,u,Du)Dk, (3.6)

    with [wij] denoting the inverse of [wij], we can now follow the calculations in [20] to obtain, from inequalities (3.11) and (3.12) there, at a maximum point x0 and vector ξ=e1 of v in Ω,

    Lvτwii+κwiiC(τ+κ)+12w211i>1wii(Diw11)2+βLlog(u0u), (3.7)

    provided τC and κCτ. Here, as is customary, we use C to denote a constant depending on the same quantities as in the estimate being proved, (3.3). We also note here that a term τwii is missing in the bracketed terms in (3.11), (3.12) in [20], (which is controlled, as in (3.7), by taking κCτ). To handle the last term in (3.7), we now need to estimate Lη, where η=u0u by extending analogous estimates in the special cases of optimal transportation in [10] and generated Jacobian equations satisfying A4w in [5]. We then have in our general case, at x=x0Ω, with similar computation to that underlying our differential inequality(2.4), using condition A3w,

    Lη=wij[wij+Aij(x,u0,Du0)Aij(x,u,Du)DpkAij(x,u,Du)Dkη]Dpk(logB)(x,u,Du)DkηC(1+wii|η|+wii|Diη|), (3.8)

    Substituting in (3.7), we then obtain, with [wij] assumed diagonal at x0,

    Lvτwii+κwiiC(τ+κ)+12w211i>1wii(Diw11)2CβηCβη2wii|Diη|2, (3.9)

    provided also κCβ. Now we follow [10] to control the last term in (3.9). Using Dv(x0)=0, together with |Dφ|d, we then estimate, for each i=1,n,

    β2(Diηη)24(Diw11w11)2+Cτ2(w2ii+1)+4κ2d2. (3.10)

    Assuming ηw11(x0)β, we can now estimate the last term in (3.9),

    βη2wii|Diη|2Cη+4βw211i>1wii(Diw11)2+Cτ2β(wii+wii)+4βκ2d2wii. (3.11)

    We now conclude the proof of Lemma 3.3 by choosing τC, κCβ, βCτ2 and d1/κ.

    From Lemma 3.3 we now have the following extension of Theorem 3.2 to A3w.

    Theorem 3.4. Let u be a strictly g-convex generalized solution of (1.5) with positive densities fC1,1(Ω),fC1,1(Ω) with f,1/fL(Ω),f,1/fL(Ω) and with generating function g satisfying conditions, A1, A2, A1* and A3w. Suppose that u(Ω)⊂⊂J0, Ω is g-convex with respect to u and Ω is sub g-convex with respect to v=ug. Then uC3(Ω) and is an elliptic solution of (1.2), (1.10). Furthermore if Ω is g-convex with respect to v=ug, then Tu is also a diffeomorphism from Ω to Ω, with v an elliptic solution of the dual boundary value problem.

    To prove Theorem 3.4, we need to adjust the regularity argument in [18] by using the interior estimate (3.3) for the solutions w of the approximating Dirichlet problems in Lemma 4.6, in place of the estimate (4.12) in [18], when the strong condition A3 is satisfied. To be more specific using Lemma 4.6 in [18] we construct a sequence {ˉum} of smooth elliptic solutions of the Dirichlet problem for Eq (1.2) in a small ball of radius r, BrΩ⊂⊂Ω, with boundary condition ˉum=um on Br, where {um} is an appropriately chosen sequence of smooth functions converging uniformly to u in Ω, uniformly bounded in C1 and uniformly semi-convex. Here the strict g-convexity of u ensures Tum(Ω)˜Ω for some ˜Ω⊂⊂Ω. Otherwise we need to assume fC1,1(ˉΩ), as done for Theorem 3.2. The sequence {ˉum} will also be uniformly bounded in C1 and moreover with r sufficiently small, {ˉum} is g-convex in Br. Then using the strict g-convexity of u and Lemma 3.3, we can follow the proof of Theorem 3.1 in [9] to obtain uniform interior second derivative bounds for ˉum in Br. Subsequently, returning to the proof of Theorem 4.6 in [18] and using Lemmas 4.3 and 4.4 in [18], we conclude that ˉum converges to u in Br. From standard elliptic regularity theory [1], we then infer uC3(Ω), is an elliptic solution of (1.2), while from the mass balance condition (1.7), Tu(Ω)=ΩE, for some closed null set E. Now letting Tv denote the g-normal mapping of v on Ω, we then have, (as in the A3 case), from the g-convexity of Ω and smoothness of u, that Tv(Ω)=Ω. Hence Tu is a diffeomorphism from Ω to Ω with (Tu)1=Tv and we conclude from [18], Section 3 that v is a C3 elliptic solution of the dual boundary value problem.

    In order to apply the strict convexity result of Kitagawa and Guillen [2], it will be convenient to strengthen our domain conditions by assuming that the domain U is g-convex with respect to all yV and zg(,y,J0)(U) and V is g-convex with respect to U×J0. Then we have the following corollaries of Theorem 3.4.

    Corollary 3.5. Let u be a g-convex generalized solution of (1.5) with positive densities fC1,1(Ω),fC1,1(Ω) with f,1/fL(Ω),f,1/fL(Ω), ˉΩU, ˉΩV and with generating function g satisfying conditions, A1, A2, A1* and A3w. Suppose that u(Ω)⊂⊂J0, together with its g-affine supports, and Ω is g-convex with respect to u on Ω. Then uC3(Ω) and is an elliptic solution of (1.2) and (1.10). Furthermore if Ω is g-convex with respect to v=ug, then Tu is also a diffeomorphism from Ω to Ω, with v an elliptic solution of the dual boundary value problem.

    Note that from Theorems 2.3 and 2.4 in [2] we have, in Corollary 3.5, that uC1(Ω) is strictly g-convex in Ω, using just the boundedness of the densities f,f and their reciprocals.

    From Theorem 3.1 and Corollary 3.5, and taking account of the uniqueness result of Rankin [15], we now have the following extension of Theorem 1.1 to the situation when A3 is weakened to A3w.

    Corollary 3.6. Let Ω and Ω be bounded domains in Rn, with closures ˉΩU and ˉΩV and let g be a generating function satisfying A1, A2, A1*, A3 and A5. Suppose that f>0,C1,1(Ω),f>0,C1,1(Ω), with f,1/fL(Ω),f,1/fL(Ω) and that f and f satisfy the mass balance condition (1.7). Then for any x0Ω and u0 satisfying m0+2K1<u0<M0K1, there exists a unique g-convex, elliptic solution uC3(Ω) of the second boundary value problem (1.2), (1.5), satisfying u(x0)=u0, provided Ω is g-convex with respect to Ω×J1, where J1=(u0K1,u0+K1), and Ω is g-convex with respect to all yΩ and zg(,y,J1)(Ω).

    To conclude this section we remark that we can also use Lemma 3.3 to extend the second derivative estimates in [5,9] and consequently the classical existence theory in [6] to generated Jacobian equations under just conditions A1, A1*, A2 and A3w.

    In [18] we also mentioned that the optimal transportation strict convexity and C1 regularity results in [21] extended readily to the generated Jacobian case. The key convexity lemmas for this result are Lemmas 2.3 and its predecessor, Lemma 2.3 in [18], on the g-convexity of contact sets.

    Lemma 4.1. Let u and g satisfy the hypotheses of Lemma 2.3 with g0=g(,y0,z0) a g-support at x0Ω and condition A3w strengthened to A3. Suppose for some ball B=BR(x0)⊂⊂Ω, there exists a positive constant λ0 such that

    |Tu(ω)|λ0|ω|, (4.1)

    for all open subsets ωB. Then u is strictly g-convex at x0. Alternatively suppose that Tu(Ω) is g-convex with respect to u on Ω, where Ω is sub g-convex with respect to g(x,,u(x)) on Tu(Ω) for all xΩ, and

    |Tu(ω)|Λ0|ω| (4.2)

    for all open subsets ωΩ and some positive constant Λ0. Then uC1(Ω).

    To prove Lemma 4.1, we suppose that u is not strictly g-convex at x0 so that by Lemma 2.3 there must exist a g-segment γ, with respect to y0,z0 joining x0 to another point x1B and lying in B. For any sufficiently small radius ρ, there then exists a ball Bρ⊂⊂B of radius ρ intersecting γ, with x0,x1Bρ. Now let ˜uC3(Bρ)C0(ˉBρ) be the unique elliptic solution of the Dirichlet problem,

    det[D2˜uA(,˜u,D˜u)]=λ02 in Bρ,˜u=u on Bρ, (4.3)

    which, as in the proofs of regularity, is well defined for sufficiently small ρ. By our previous comparison arguments, we also have ˜u>ug0 in Bρ and ˜u=u=g0 on γBρ. This can be seen readily from the finer inequality ˜u>˜u0u, where ˜u0 solves the Dirichlet problem (4.3) with λ0/2 replaced by λ0. Again Lemma 4.6 in [18] is also crucial here as in the regularity theory in Section 3. Applying now Lemma 2.3 again in Bρ, with z0 slightly reduced, or more simply the proof of Lemma 2.1, we thus reach a contradiction. The second assertion in Lemma 4.1 on C1 regularity now follows by duality.

    Note that in the above argument we have taken advantage of the g-convexity of the contact set S0 to simplify the technicalities in [21], where just the connectedness of S0 is used, although the basic approach though reduction to a Dirichlet problem in small balls is the same. Also, from either the ellipticity of (4.3) or condition A3, we only need the strict differential inequality (2.5) in replicating the differential inequality argument in Lemma 2.1.

    For application to regularity we need to enhance our domain sub convexity conditions. Namely we will call Ω sub g-convex in U, with respect to (y0,z0), if z0I(U,y0) and the convex hull of Q(Ω,y0,z0)Q(U,y0,z0) and Ω sub g-convex in V, with respect to (x0,u0), if u0J(x0,V) and the convex hull of P(x0,Ω,u0)P(x0,V,u0). Then we can weaken the g-convexity of Ω in Lemma 2.3, in the case of arbitrary contact sets S0, by Ω being sub g-convex in a larger domain U, with respect to (y,z), for all g-supports, g(,y,z) to u, or equivalently with respect to the dual function v=ug. In this case we need to strengthen condition (ii) so that Tu(Ω) is sub g-convex in a larger domain V with respect to g(,y,zy) on ˆΩy, for all yTu(Ω),zy=v(y), where ˆΩy denotes the Q(,y,zy) convex hull of Ω. To be more explicit, we define ˆΩy as the image under the inverse mapping Q1(,y,zy) of the convex hull of Q(Ω,y,zy).

    Then using duality we have the following extension of Theorem 3.2 to non smooth densities.

    Theorem 4.2. Let u be a generalized solution of (1.5) with positive densities f and f, with f,1/fL(Ω),f,1/fL(Ω) and with generating function g satisfying conditions A1, A2, A1* and A3. Suppose that Ω⊂⊂U is sub g-convex in U with respect to g(x,,u(x) on Tu(Ω), for all xΩ and Ω⊂⊂V is sub g-convex in V with respect to g(,y,zy) on ˆΩy for all yΩ,zy=v(y). Then uC1(ˉΩ) is strictly g-convex in Ω if Ω is g-convex with respect to u and Tu is a homeomorphism from Ω to Ω if also Ω is g-convex with respect to v. Furthermore if fC1,1(Ω),fC1,1(Ω), then uC3(Ω).

    We can also simplify the statement of Theorem 4.2, (as well as earlier results), by replacing Γ by U×V×J0 for sufficiently large domains U and V and interval J0. We note here also that the approach of Loeper [11] to local C1,α regularity in optimal transportation under A3 was used in near field geometric optics in [3] and has been recently extended to general generated Jacobian equations in [4]. As a byproduct of our local convexity theory in [13], these results also extend to C2 cost and generating functions.

    We prove here a version of Lemma 2.2 which does not use duality and corresponds in some sense to Lemma 2.3. We use the same notation as in Lemma 2.2.

    Lemma 4.3. Assume g satisfies A1, A2, A1* and A3w and suppose uC0(Ω) is g-convex in Ω. Assume also:

    (i) Ω is g-convex with respect to some ˜yΣ0, ˜z=g(x0,˜y,u0) for some x0 in Ω;

    (ii) Tu(x0) is sub g-convex with respect to ˜g=g(,˜y,˜z) in Ω.

    Then ˜yTu(x0).

    To prove Lemma 4.3 we first choose two extreme points p0 and p1 in u(x0) such that pθ=θp0+(1θ)p1=gx(x0,˜y,˜z) for some θ(0,1). Setting yθ,zθ=Y,Z(x0,u0,pθ), and gθ=g(,yθ,zθ) for any θ[0,1], it follows that for either i=0 or 1, ˜h=gi˜g>0 at some point on the g-segment Ig, with respect to ˜y,˜z, joining x0 and a point xΩ, provided Dηg0(x0)Dηg1(x0), where ηj=Ei,j(x0,˜y,˜z)(qiq0i), q=Q(x,˜y,˜z),q0=Q(x0,˜y,˜z). By decreasing ˜z we can then apply the argument of Lemma 2.3 in the domain Ω to obtain a contradiction if max{g0,g1}(x)<˜g(x). On the other hand, if Dηg0(x0)=Dηg1(x0), then the function f, given by f(t)=˜h(xt), satisfies f(0)=f(0)=0 for both i=0 and 1 so that if also f0 on [0,1], f(1)<0, we also obtain a contradiction with the differential inequality (2.7). Alternatively, we can approximate x to reduce to the case Dηg0(x0)Dηg1(x0), as in [18]. Consequently ˜gmax{g0,g1} on Ω whence ˜yTu(x0) as asserted.

    Note that the domain Ω in Lemma 4.3 can be made arbitrary by replacing it by its Q(,˜y,˜z) convex hull in condition (ii). By extending to all ˜yΣ0, we can then obtain another version of Lemma 2.2, though under stricter sub convexity hypotheses.

    We also have another version of Lemma 2.3 from our earlier drafts, which essentially follows from the proof of Lemma 2.1.

    Lemma 4.4. Assume g satisfies A1, A2, A1* and A3w and suppose uC0(Ω) is g-convex in Ω and g0=g(,y0,z0) is a g-affine function on Ω. Assume also:

    (i) Ω is g-convex with respect to y0 and all zg(,y0,sup{u,g0})(Ω);

    (ii) Tu(Ω){y0} is sub g-convex with respect to sup{u,g0} in Ω.

    Then the section S=S(u,g0) is g-convex with respect to (y0,z0), while if g0 is a g-support to u, the contact set S0=S0(u,g0) is g-convex with respect to (y0,z0).

    Note that when we extend Lemma 4.4 to general domains, when g0 is an arbitrary g-support to u, we end up with the same sub convexity hypotheses as the corresponding extension above of Lemma 2.3.

    Finally we also indicate that by modification of the proof of Lemma 4.3, we may obtain a refinement of the local version of the Loeper maximum principle (2.11) under condition A3, corresponding to those used to show C1,α regularity in [3,4,11]. This is also extended to gC2, through the geometric approach in [13].

    Lemma 4.5. Assume g satisfies A1, A2, A1* and A3 and (x0,u0,[p0,p1])U⊂⊂U, where [p0,p1] denotes the straight line joining p0 and p1 in Rn. Then there exist small positive constants, ϵ0 and γ0, depending on g and U, such that

    g(x,yθ,zθ)max{g0(x),g1(x)}γ0[θ(1θ)|p1p0||xx0|]2 (4.4)

    for all |xx0|ϵ0 and θ[0,1].

    To prove (4.4), we replace ˜g=gθ in the proof of Lemma 4.3 by the function ˜gγ given by

    ˜gγ=gθ+γ[θ(1θ)|p1p0||xx0|]2 (4.5)

    for sufficiently small γ>0 and distance |xx0|, using more explicitly the differential inequalities (2.9) and (2.10) in [18], in conjunction with condition A3 in U, to extend (2.7) to ˜gγ.

    Research supported by Australian Research Council Grants DP170100929, DP180100431.

    The author declares no conflict of interest.



    [1] D. Gilbarg, N. S. Trudinger, Elliptic partial differential equations of second order, Springer, 1983.
    [2] N. Guillen, J. Kitagawa, Pointwise inequalities in geometric optics and other generated Jacobian equations, Commun. Pure Appl. Math., 70 (2017), 1146–1220. doi: 10.1002/cpa.21691
    [3] C. E. Gutiérrez, F. Tournier, Regularity for the near field parallel refractor and reflector problems, Calc. Var., 45 (2015), 917–949.
    [4] S. Jeong, Local Hölder regularity of solutions to generated Jacobian equations, arXiv: 2004.12004.
    [5] F. Jiang, N. S. Trudinger, On Pogorelev estimates in optimal transportation and geometric optics, Bull. Math. Sci., 4 (2014), 407–431. doi: 10.1007/s13373-014-0055-5
    [6] F. Jiang, N. S. Trudinger, On the second boundary value problem for Monge Ampère type equations and geometric optics, Arch. Ration. Mech. Anal., 229 (2018), 547–567. doi: 10.1007/s00205-018-1222-8
    [7] A. Karakhanyan, X. J. Wang, On the reflector shape design, J. Differ. Geom., 84 (2010), 561–610.
    [8] Y. H. Kim, R. J. McCann, Continuity, curvature and the general covariance of optimal transportation, J. Eur. Math. Soc., 12 (2010), 1009–1040.
    [9] J. K. Liu, N. S. Trudinger, On Pogorelov estimates for Monge-Ampère type equations, Discrete Contin. Dyn. A, 28 (2010), 1121–1135. doi: 10.3934/dcds.2010.28.1121
    [10] J. K. Liu, N. S. Trudinger, X. J. Wang, Interior C2,α regularity for potential functions in optimal transportation, Commun. Part. Diff. Eq., 35 (2010), 165–184.
    [11] G. Loeper, On the regularity of solutions of optimal transportation problems, Acta Math., 202 (2009), 241–283. doi: 10.1007/s11511-009-0037-8
    [12] G. Loeper, N. S. Trudinger, Weak formulation of the MTW condition and convexity properties of potentials, arXiv: 2007.02665.
    [13] G. Loeper, N. S. Trudinger, On the convexity theory of generating functions, 2020, preprint.
    [14] X. N. Ma, N. S. Trudinger, X. J. Wang, Regularity of potential functions of the optimal transportation problem, Arch. Ration. Mech. Anal., 177 (2005), 151–183. doi: 10.1007/s00205-005-0362-9
    [15] C. Rankin, Distinct solutions to GPJE cannot intersect, Bull. Aust. Math. Soc., 102 (2020), 462–470. doi: 10.1017/S0004972720000052
    [16] N. S. Trudinger, Recent developments in elliptic partial differential equations of Monge-Ampère type, In: Proceedings of the International Congress of Mathematicians, 2006,291–302.
    [17] N. S. Trudinger, On generated prescribed Jacobian equations, Oberwolfach Reports, 38 (2011), 32–36.
    [18] N. S. Trudinger, The local theory of prescribed Jacobian equations, Discrete Contin. Dyn. A, 34 (2014), 1663–1681. doi: 10.3934/dcds.2014.34.1663
    [19] N. S. Trudinger, X. J. Wang, On convexity notions in optimal transportation, 2008, preprint.
    [20] N. S. Trudinger, X. J. Wang, On the second boundary value problem for Monge-Ampère type equations and optimal transportation, Ann. Scuola Norm. Sup. Pisa Cl. Sci., (2009), 143–174.
    [21] N. S. Trudinger, X. J. Wang, On strict convexity and continuous differentiability of potential functions in optimal transportation, Arch. Ration. Mech. Anal., 192 (2009), 403–418. doi: 10.1007/s00205-008-0147-z
    [22] C. Villani, Optimal transportation: old and new, Springer, 2008.
  • This article has been cited by:

    1. Neil S. Trudinger, A note on second derivative estimates for Monge-Ampère-type equations, 2023, 23, 2169-0375, 10.1515/ans-2022-0036
    2. Feida Jiang, Weak solutions of generated Jacobian equations, 2023, 5, 2640-3501, 1, 10.3934/mine.2023064
    3. Cale Rankin, Strict convexity and C1
    regularity of solutions to generated Jacobian equations in dimension two, 2021, 60, 0944-2669, 10.1007/s00526-021-02093-4
    4. Cale Rankin, First and second derivative Hölder estimates for generated Jacobian equations, 2023, 62, 0944-2669, 10.1007/s00526-022-02406-1
    5. Cale Rankin, Strict g-Convexity for Generated Jacobian Equations with Applications to Global Regularity, 2023, 55, 0036-1410, 5685, 10.1137/22M1518852
    6. Grégoire Loeper, Neil S. Trudinger, On the convexity theory of generating functions, 2024, 1536-1365, 10.1515/ans-2023-0160
  • Reader Comments
  • © 2021 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2685) PDF downloads(681) Cited by(6)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog