Skip to main content
Log in

The Minimum Principle for Convex Subequations

  • Published:
The Journal of Geometric Analysis Aims and scope Submit manuscript

Abstract

A subequation, in the sense of Harvey–Lawson, on an open subset \(X\subset \mathbb R^n\) is a subset F of the space of 2-jets on X with certain properties. A smooth function is said to be F-subharmonic if all of its 2-jets lie in F, and using the viscosity technique one can extend the notion of F-subharmonicity to any upper-semicontinuous function. Let \(\mathcal P\) denote the subequation consisting of those 2-jets whose Hessian part is semipositive. We introduce a notion of product subequation \(F\#\mathcal P\) on \(X\times \mathbb R^{m}\) and prove, under suitable hypotheses, that if F is convex and f(xy) is \(F\#\mathcal P\)-subharmonic then the marginal function

$$\begin{aligned} g(x):= \inf _y f(x,y) \end{aligned}$$

is F-subharmonic. This generalises the classical statement that the marginal function of a convex function is again convex. We also prove a complex version of this result that generalises the Kiselman minimum principle for the marginal function of a plurisubharmonic function.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1

Similar content being viewed by others

References

  1. Berndtsson, B.: Prekopa’s theorem and Kiselman’s minimum principle for plurisubharmonic functions. Math. Ann. 312(4), 785–792 (1998)

    Article  MathSciNet  Google Scholar 

  2. Caffarelli, L., Nirenberg, L., Spruck, J.: The Dirichlet problem for nonlinear second-order elliptic equations. III. Functions of the eigenvalues of the Hessian. Acta Math. 155(3–4), 261–301 (1985)

    Article  MathSciNet  Google Scholar 

  3. Cordero-Erausquin, D.: On Berndtsson’s generalization of Prékopa’s theorem. Math. Z. 249(2), 401–410 (2005)

    Article  MathSciNet  Google Scholar 

  4. Crandall, M.G., Ishii, H., Lions, P.-L.: User’s guide to viscosity solutions of second order partial differential equations. Bull. Am. Math. Soc. (NS) 27(1), 1–67 (1992)

    Article  MathSciNet  Google Scholar 

  5. Darvas, T., Rubinstein, Y.A.: Kiselman’s principle, the Dirichlet problem for the Monge-Ampère equation, and rooftop obstacle problems. J. Math. Soc. Jpn. 68(2), 773–796 (2016)

    Article  Google Scholar 

  6. Darvas, T., Rubinstein, Y.A.: A minimum principle for Lagrangian graphs. Commun. Anal. Geom. 27(4), 857–876 (2019)

    Article  MathSciNet  Google Scholar 

  7. Demailly, J.-P.: Complex Analytic and Differential Geometry. https://www-fourier.ujf-grenoble.fr/~demailly/manuscripts/agbook.pdf

  8. Deng, F.S., Zhang, H.P., Zhou, X.Y.: Minimum principle for plurisubharmonic functions and related topics. Acta Math. Sin. English Series (2018)

  9. Gallier, J.: The Schur complement and symmetric positive semidefinite (definite) matrices, p. jean/schur-comp.pdf (2010)

  10. Gallier, J.: Geometric Methods and Applications, volume 38 of Texts in Applied Mathematics, 2nd edn. Springer, New York (2011)

  11. Harvey, F.R., Lawson Jr., H.B.: The ae theorem and addition theorems for quasi-convex functions, (2013). arXiv:1309.1770

  12. Harvey, F.R., Lawson Jr., H.B.: Dirichlet duality and the nonlinear Dirichlet problem. Commun. Pure Appl. Math. 62(3), 396–443 (2009)

    Article  MathSciNet  Google Scholar 

  13. Harvey, F.R., Lawson Jr., H.B.: Dirichlet duality and the nonlinear Dirichlet problem on Riemannian manifolds. J. Differ. Geom. 88(3), 395–482 (2011)

    Article  MathSciNet  Google Scholar 

  14. Harvey, F.R., Lawson Jr., H.B.: The equivalence of viscosity and distributional subsolutions for convex subequations—a strong Bellman principle. Bull. Braz. Math. Soc. (NS) 44(4), 621–652 (2013)

    Article  MathSciNet  Google Scholar 

  15. Harvey, F.R., Lawson Jr., H.B.: The restriction theorem for fully nonlinear subequations. Ann. Inst. Fourier (Grenoble) 64(1), 217–265 (2014)

    Article  MathSciNet  Google Scholar 

  16. Hörmander, L.: An Introduction to Complex Analysis in Several Variables, vol. 7 of North-Holland Mathematical Library, 3rd edn. North-Holland Publishing Co., Amsterdam (1990)

  17. Hörmander, L.: Notions of Convexity. Modern Birkhäuser Classics. Birkhäuser Boston Inc, Boston, MA (2007)

    Google Scholar 

  18. Kiselman, C.O.: The partial Legendre transformation for plurisubharmonic functions. Invent. Math. 49(2), 137–148 (1978)

    Article  MathSciNet  Google Scholar 

  19. Kiselman, C.O.: Plurisubharmonic functions and their singularities. In: Complex Potential Theory (Montreal, PQ, 1993), vol. 439 of NATO Adv. Sci. Inst. Ser. C Math. Phys. Sci., pp. 273–323. Kluwer Acad. Publ., Dordrecht (1994)

  20. Kiselman, C.O.: Plurisubharmonic functions and potential theory in several complex variables. In: Development of Mathematics 1950–2000, pp. 655–714. Birkhäuser, Basel (2000)

    Chapter  Google Scholar 

  21. Krylov, N.V.: On the general notion of fully nonlinear second-order elliptic equations. Trans. Am. Math. Soc. 347(3), 857–895 (1995)

    Article  MathSciNet  Google Scholar 

  22. Poletsky, E.A.: The minimum principle. Indiana Univ. Math. J. 51(2), 269–303 (2002)

    Article  MathSciNet  Google Scholar 

  23. Prékopa, A.: On logarithmic concave measures and functions. Acta Sci. Math. (Szeged) 34, 335–343 (1973)

    MathSciNet  MATH  Google Scholar 

  24. Ross, J., Witt Nyström, D.: Differentiability of the argmin function and a minimum principle for semiconcave subsolutions. J. Convex Anal. 27(3), 811 (2020)

    MathSciNet  MATH  Google Scholar 

  25. Ross, J., Witt Nyström, D.: Harmonic discs of solutions to the complex homogeneous Monge-Ampère equation. Publ. Math. Inst. Hautes Études Sci. 122, 315–335 (2015)

    Article  MathSciNet  Google Scholar 

  26. Zeriahi, A.: A minimum principle for plurisubharmonic functions. Indiana Univ. Math. J. 56(6), 2671–2696 (2007)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

The authors wish to thank Tristan Collins and Yanir Rubinstein for conversations that stimulated this work.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Julius Ross.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: F-Subharmonic Functions

1.1 A.1 Types of Subequations

Definition A.1

Let \(F\subset J^2(X)\).

  1. (1)

    We say F is constant coefficient if \(F_x\) is independent of x, i.e.

    $$\begin{aligned} (x,r,p,A) \in F_x \Leftrightarrow (x',r,p,A)\in F_{x'} \text { for all }x,x',r,p,A. \end{aligned}$$
  2. (2)

    We say F is independent of the gradient part (or gradient-independent) if each \(F_x\) is independent of p, i.e.

    $$\begin{aligned} (r,p,A)\in F_x \Leftrightarrow (r,p',A) \in F_x \text { for all } x,r,p,p',A. \end{aligned}$$
  3. (3)

    We say F depends only on the Hessian part if each \(F_x\) is independent of (rp), i.e.

    $$\begin{aligned} (r,p,A) \in F_x \Leftrightarrow (r',p',A)\in F_{x} \text { for all }x,r,r',p,p',A. \end{aligned}$$

Definition A.2

(G-Invariance)The group \(GL_n(\mathbb R)\) acts on \(J^2(X)\) by

$$\begin{aligned} g^*(x,r,p,A):= (x,r,g^t p, g^t Ag) \text { for } g\in GL_n(\mathbb R). \end{aligned}$$

If G is a subgroup of \(GL_n(\mathbb R)\) we say \(F\subset J^2(X) \) is G-invariant if \(g^*\alpha \in F\) for all \(\alpha \in F\) and all \(g\in G\).

Remark A.3

Our action of \(GL_n\) comes from thinking of the jet space using the cotangent space to X, and is different in convention to that of [13].

1.2 A.2 Complex Subequations

Set

$$\begin{aligned} \mathbb J = \left( \begin{array}{cc} 0&{} -{\text {Id}}_{n} \\ {\text {Id}}_n &{} 0\end{array} \right) \in M_{2n\times 2n}(\mathbb R). \end{aligned}$$

If \(A\in M_{2n\times 2n}(\mathbb R)\) commutes with \(\mathbb J\) then making the standard identification \(\mathbb C\simeq \mathbb R^2\) we think of A as a complex matrix \(\hat{A} \in M_{n\times n}(\mathbb C)\). Explicitly if in block form

$$\begin{aligned} A = \left( \begin{array}{cc} a&{} c \\ b &{} d \end{array} \right) \end{aligned}$$

where \(a,b,c,d\in M_{n\times n}(\mathbb R)\) then A commutes with \(\mathbb J\) if and only if \(a=d\) and \(b=-c\), in which case

$$\begin{aligned} \hat{A}: = a+ib \in M_{n\times n}(\mathbb C). \end{aligned}$$

Observe \(\widehat{AB} = \hat{A} \hat{B}\) and \(\widehat{A^t} = \hat{A}^*\).

Let \({\text {Herm}}_n\) be the set of hermitian \(n\times n\) complex matrices, and

$$\begin{aligned} {\text {Pos}}_n^{\mathbb C} := \{ \hat{A} \in {\text {Herm}}_n : v^* \hat{A} v \ge 0 \text { for all } v\in \mathbb C^n\} \end{aligned}$$

the subset of semipositive hermitian matrices. From the above it easy to check that if \(A\mathbb J = \mathbb JA\) then

$$\begin{aligned} A\in {\text {Sym}}^2_{2n}&\Longleftrightarrow \hat{A}\in {\text {Herm}}(\mathbb C^n)\text { and }\\ A \in {\text {Pos}}_{2n}&\Longleftrightarrow \hat{A}\in {\text {Pos}}^{\mathbb C}_{n}. \end{aligned}$$

Now for any \(A\in M_{2n\times 2n}(\mathbb R)\) the matrix

$$\begin{aligned} A_{\mathbb C} : = \frac{1}{2} ( A - \mathbb J A \mathbb J) \end{aligned}$$

commutes with \(\mathbb J\) and thus we may think of \(A_{\mathbb C}\) as an element of \(M_{n\times n}(\mathbb C)\). Observe if A is symmetric then \(A_{\mathbb C}\) is hermitian.

Definition A.4

Let \(X\subset \mathbb R^{2n}\simeq \mathbb C^n\) be open. We say \(F\subset J^2(X)\) is a complex subequation if \((x,r,v,A)\in F\) if and only if \((x,r,v,A_{\mathbb C})\in F\).

So by abuse of notation if F is a complex subequation we may equivalently consider it as a subset

$$\begin{aligned} F\subset J^{2,\mathbb C}(X) := X\times \mathbb R \times \mathbb C^n \times {\text {Herm}}(\mathbb C^n) =: X\times J^{2,\mathbb C}_n \end{aligned}$$

without any loss of information. The group \(GL_n(\mathbb C)\) acts on \(J^{2,\mathbb C}(X)\) by

$$\begin{aligned} g^*(x,r,p,A) = (x,r,g^* p, g^* Ag). \end{aligned}$$

Observe also if F is complex, then having the Positivity property (1) is equivalent to

$$\begin{aligned} (x,r,p,A)\in F \Longrightarrow (x,r,p,A+P)\in F \text { for all }P\in {\text {Pos}}_n^{\mathbb C}. \end{aligned}$$

Example A.5

Let

$$\begin{aligned} \mathcal P^{\mathbb C}_X : = X\times \mathbb R\times \mathbb C^n \times {\text {Pos}}^{\mathbb C}_n \end{aligned}$$

which is a convex complex subequation. We will write \(\mathcal P^{\mathbb C}\) for \(\mathcal P^{\mathbb C}_X\) when X is clear from context.

Example A.6

(Convex and Plurisubharmonic) Recall \(\mathcal P_X = X\times \mathbb R\times \mathbb R^n\times {\text {Pos}}_n\). Then \(\mathcal P_X(X)\) consists of locally convex functions on X [13, Example 14.2]. Similarly if \(X\subset \mathbb C^n\) is open then \(\mathcal P^{\mathbb C}_X(X)\) consists of the plurisubharmonic functions on X [13, p. 63].

1.3 A.3 Basic Properties of F-Subharmonic Functions

The following lists some of the basic limit properties satisfied by F-subharmonic functions (under very mild assumptions on F).

Proposition A.7

Let \(F\subset J^2(X)\) be closed. Then

  1. (1)

    (Maximum Property) If \(f,g\in F(X)\) then \(\max \{f,g\}\in F(X)\).

  2. (2)

    (Decreasing Sequences) If \(f_j\) is decreasing sequence of functions in F(X) (so \(f_{j+1}\le f_j\) over X) then \(f:=\lim _j f_j\) is in F(X).

  3. (3)

    (Uniform limits) If \(f_j\) is a sequence of functions on F(X) that converge locally uniformly to f then \(f\in F(X)\).

  4. (4)

    (Families locally bounded above) Suppose \(\mathcal F\subset F(X)\) is a family of F-subharmonic functions locally uniformally bounded from above. Then the upper-semicontinuous regularisation of the supremum

    $$\begin{aligned} f:= {\sup }^*_{f\in \mathcal F} f \end{aligned}$$

    is in F(X).

  5. (5)

    If F is constant coefficient and f is F-subharmonic on X and \(x_0\in \mathbb R^{n}\) is fixed, then the function \(x\mapsto f(x-x_0)\) is F-subharmonic on \(X-x_0\).

Proof

See [13, Theorem 2.6] for (1-4). Item (5) is immediate from the definition. \(\square \)

Lemma A.8

(Limits under perturbations of subequations) Let X be open and \(F\subset J^2(X)\) be a primitive subequation. For \(\delta >0\) let \(F^\delta \subset J^2(X)\) be defined by

$$\begin{aligned} F^{\delta } = \{ (x,r,p,A) : \exists r',p' \text { with } (x,r',p',A)\in F \text { and } |r-r'|\le \delta \text { and } \Vert p-p'\Vert \le \delta \}. \end{aligned}$$

Then

  1. (1)

    \(F^{\delta }\) is a primitive subequation.

  2. (2)

    If F satisfies the Negativity property then so does \(F^\delta \).

  3. (3)

    \(\bigcap _{\delta >0} (F^\delta (X))=F(X)\).

Proof

That \(F^{\delta }\) has the Positivity property is immediate from the definition, and \(F^{\delta }\) is closed as F is closed giving (1). Statement (2) is also immediate from the definition. Finally using F is closed, \(\bigcap _{\delta >0} F^{\delta }_x = F_x\), and thus \(\bigcap _{\delta >0} (F^{\delta }(X)) = F(X).\) \(\square \)

1.4 F-Subharmonicity in Terms of Second Order Jets

It is useful to understand the property of being F-subharmonic in terms of second order jets. To do so we first discuss what it means to be twice differentiable at a point. Again let \(X\subset \mathbb R^n\) be open.

Definition A.9

(Twice differentiability at a point) We say that \(f:X\rightarrow \mathbb R\) is twice differentiable at \(x_0\in X\) if there exists a \(p\in \mathbb R^n\) and an \(L\in {\text {Sym}}_n^2\) such that for all \(\epsilon >0\) there is a \(\delta >0\) such that for \(\Vert x-x_0\Vert <\delta \) we have

$$\begin{aligned} |f(x) - f(x_0) - p.(x-x_0) - \frac{1}{2} (x-x_0)^tL (x-x_0) | \le \epsilon \Vert x-x_0\Vert ^2. \end{aligned}$$
(52)

When f is twice differentiable at \(x_0\) then the pL in (52) are unique, and moreover in this case f is differentiable at \(x_0\) and

$$\begin{aligned} p = \nabla f|_{x_0}= \left( \begin{array}{c} \frac{\partial f}{\partial x_1} \\ \frac{\partial f}{\partial x_2} \\ \vdots \\ \frac{\partial f}{\partial x_n} \end{array}\right) |_{x_0}\in \mathbb R^n. \end{aligned}$$

When f is twice differentiable at \(x_0\) we shall refer to L as the Hessian of f at \(x_0\) and denote it by \({\text {Hess}}(f)|_{x_0}\). Of course, by Taylor’s Theorem, when f is \(\mathcal C^{2}\) in a neighbourhood of \(x_0\) then \({\text {Hess}}_{x}(f)\) is the matrix with entries

$$\begin{aligned} ({\text {Hess}}(f)_{x_0} )_{ij}: = \frac{\partial ^2 f}{\partial x_i\partial x_j}|_{x_0}. \end{aligned}$$

Definition A.10

(Second order jet) Suppose that \(f: X\rightarrow \mathbb R\) is twice differentiable at \(x_0\). We denote the second order jet of f at \(x_0\) by

$$\begin{aligned} J^2_{x_0}(f):= (f(x_0), \nabla f|_{x_0}, {\text {Hess}}(f)|_{x_0}) \in J^2_{n} = \mathbb R\times \mathbb R^n\times {\text {Sym}}_n^2. \end{aligned}$$
(53)

The importance of the Positivity property is made apparent by the following that shows that F-subharmonicity behaves as expected for sufficiently smooth functions.

Lemma A.11

Let \(F\subset J^2(X)\) satisfy the Positivity assumption (1) and suppose \(f:X\rightarrow \mathbb R\) is \(\mathcal C^2\). Then \(f\in F(X)\) if and only if \(J^2_{x}(f)\in F_x\) for all \(x\in X\).

Proof

The reader may easily verify this, or consult [13, Equation 2.4 and Proposition 2.3]. \(\square \)

The definition F-subharmonicity given above says that at any upper-contact point x, with upper-second order jet (pA), the quadratic function

$$\begin{aligned} y\mapsto f(y) + p.(x-y) + \frac{1}{2} (y-x)^tA (y-x) \end{aligned}$$

has second-order jet lying in \(F_x\). The next statement says that this is equivalent to the more classical “viscosity definition". Given an upper-semicontinuous f we say that \(\phi \) is a \(\mathcal {C}^2\)-test function touching f from above at \(x_0\) if \(\phi \in \mathcal C^2\) in a neighbourhood of \(x_0\) with \(\phi \ge f\) on this neighbourhood and \(\phi (x_0) = f(x_0)\).

Lemma A.12

(Viscosity definition of F-subharmonicity) An upper-semicontinuous \(f:X\rightarrow \mathbb R\cup \{-\infty \}\) is in F(X) if and only if for all \(x_0\in X\) and test-functions \(\phi \) touching f from above at \(x_0\) it holds that \(J^2_{x_0}(\phi )\in F_{x_0}.\)

Proof

See [13, Lemma 2.4]. \(\square \)

It takes some work to understand how F-subharmonicity interacts with linearity in the space of functions. However when F is constant-coefficient and convex the following is true:

Proposition A.13

(Convex combinations of F-subharmonic functions) Let F be a constant coefficient convex primitive subequation. Then any convex combination of F-subharmonic functions is again F-subharmonic.

Proof

This is implied by [11, Theorem 5.1 ] (apply the cited theorem to \(F_x:=\lambda H_x\) and \(G_x:= (1-\lambda )H_x\) for a given \(\lambda \in [0,1]\)) \(\square \)

Appendix B: Associativity of Products

We prove Proposition 3.4 which states that if \(X_i\subset \mathbb R^{n_i}\) are open and \(F_i\subset J^2(X_i)\) for \(i=1,2,3\) then

$$\begin{aligned} (F_1\#F_2)\#F_3 = F_1\#(F_2\#F_3). \end{aligned}$$

Let xyz be coordinates on \(\mathbb R^{n_1},\mathbb R^{n_2},\mathbb R^{n_3}\) respectively. We will consider certain linear mappings

$$\begin{aligned} \Gamma&:\mathbb R^{n_1}\rightarrow \mathbb R^{n_2+n_3}\\ \Phi&:\mathbb R^{n_1+n_2}\rightarrow \mathbb R^{n_3}\\ \Psi&:\mathbb R^{n_1} \rightarrow \mathbb R^{n_2}\\ \Upsilon&:\mathbb R^{n_2}\rightarrow \mathbb R^{n_3} \end{aligned}$$

and write

$$\begin{aligned} \Phi (x,y) = \Phi _1(x) + \Phi _2(y) \end{aligned}$$

where \(\Phi _i:\mathbb R^{n_i} \rightarrow \mathbb R^{n_3}\) is linear. Recall that \(\iota _{\Gamma }:\mathbb R^{n_1} \rightarrow \mathbb R^{n_1+n_2+n_3}\) is \(\iota _{\Gamma }(x) = (x,\Gamma (x))\) and similarly for \(\iota _\Phi :\mathbb R^{n_1+n_2}\rightarrow \mathbb R^{n_1+n_2+n_3}\) and \(\iota _{\Psi }:\mathbb R^{n_1}\rightarrow \mathbb R^{n_1+n_2}\).

Lemma B.1

Suppose

$$\begin{aligned} \Gamma = (\Psi , \Phi _1 + \Phi _2\circ \Psi ). \end{aligned}$$
(54)

Then

$$\begin{aligned} \iota _{\Gamma } = \iota _{\Phi }\circ \iota _{\Psi } \end{aligned}$$

Proof

$$\begin{aligned} \iota _{\Phi }(\iota _{\Psi }(x))&= \iota _{\Phi }(x,\Psi (x)) = (x,\Psi (x),\Phi (x,\Psi (x))) \\ {}&= (x,\Psi (x),\Phi _1(x) + \Phi _2\circ \Psi (x)) = \iota _{\Gamma }(x).\end{aligned}$$

\(\square \)

Now set

$$\begin{aligned} \begin{array}{ll} j_2: \mathbb R^{n_2} \rightarrow \mathbb R^{n_1+n_2} &{}\text { } j_2(y) = (0,y)\\ j_3: \mathbb R^{n_3} \rightarrow \mathbb R^{n_1+n_2+n_3} &{}\text { } j_3(z) = (0,0,z)\\ j_{23}:\mathbb R^{n_2+ n_3} \rightarrow \mathbb R^{n_1+n_2+n_3} &{}\text { } j_{23}(y,z) = (0,y,z)\\ k:\mathbb R^{n_3} \rightarrow \mathbb R^{n_2+n_3} &{}\text { } k(z) = (0,z). \end{array} \end{aligned}$$

Fix \((x,y,z)\in X_1\times X_2\times X_3\). By definition of the product subequation we know \(\alpha \in (F_1\#(F_2\#F_3))_{(x,y,z)}\) if and only if

$$\begin{aligned} \forall \Gamma \text { we have }\iota _\Gamma ^*\alpha \in (F_1)_{x} \text { and } j_{23}^*\alpha \in (F_2\#F_3)_{(y,z)}. \end{aligned}$$
(55)

Observe for every \(\Gamma \) there is a pair \((\Phi ,\Psi )\) such that (54) holds. Thus by Lemma B.1, condition (55) is equivalent to

$$\begin{aligned} \forall \Psi ,\Phi \text { we have }\iota ^*_\Psi \iota ^*_\Phi \alpha \in (F_1)_{x} \text { and } j_{23}^*\alpha \in (F_2\#F_3)_{(y,z)}. \end{aligned}$$
(56)

Using the definition of \(F_2\#F_3\), condition (56) is in turn equivalent to

$$\begin{aligned} \forall \Psi ,\Phi ,\Upsilon \text { we have }\iota ^*_\Psi \iota ^*_\Phi \alpha \in (F_1)_{x} \text { and } \iota _{\Upsilon }^* j_{23}^*\alpha \in (F_2)_{y} \text { and }k^*j_{23}^*\alpha \in (F_3)_{z}. \end{aligned}$$
(57)

Now \(j_{23}\circ k=j_3\), and a simple check yields \(j_{23}\circ \iota _{\Phi _2} = \iota _{\Phi }\circ j_2\). Thus (57) is equivalent to

$$\begin{aligned} \forall \Psi ,\Phi \text { we have }\iota ^*_\Psi \iota ^*_\Phi \alpha \in (F_1)_{x} \text { and } j_{2}^*\iota _{\Phi }^*\alpha \in (F_2)_{y} \text { and }j_3^*\alpha \in (F_3)_{z}. \end{aligned}$$
(58)

So from the definition of \((F_1\#F_2)\), condition (58) is equivalent to

$$\begin{aligned} \forall \Phi \text { we have }\iota ^*_\Phi \alpha \in (F_1\#F_2)_{x} \text { and }j_3^*\alpha \in (F_3)_{z} \end{aligned}$$
(59)

which, by definition, is equivalent to \(\alpha \in ((F_1\#F_2)\#F_3)_{(x,y,z)}\).

Appendix C: Products of Gradient-Independent Subequations

Recall we say that a subequation \(F\subset J^2(X)\) has Property (P++) if the following holds. For all \(x\in X\) and all \(\epsilon >0\) there exists a \(\delta >0\) such that

$$\begin{aligned} (x,r,p,A)\in F_x \Rightarrow (x',r-\epsilon ,p, A+\epsilon {\text {Id}})\in F_{x'} \text { for all } \Vert x'-x\Vert <\delta \ \end{aligned}$$
($\hbox {P}^{++}$)

or said another way,

$$\begin{aligned} F_x + (x'-x,0,-\epsilon , \epsilon {\text {Id}}) \subset F_{x'} \text { for all } \Vert x'-x\Vert <\delta . \end{aligned}$$
($\hbox {P}^{++}$)

Lemma C.1

Assume that F and G have property (\(\text{ P}^{++}\)) and are independent of the gradient part. Then \(H: = F\#G\) is a subequation

Proof

We have already seen in Lemma 3.2 that H is closed, and satisfies the Positivity and Negativity properties (1) and (2). It remains to prove the Topological property (3) which we break up into a number of pieces. Since FG are independent of the gradient part, so is H, and thus the only non-trivial part of the topological property is to show [13, Section 4.8]

$$\begin{aligned} {\text {Int}}(H_{(x,y)}) =({\text {Int}}H)_{(x,y)} \end{aligned}$$

The fact that \({\text {Int}}(H_{(x,y)}) \subset ({\text {Int}}H)_{(x,y)}\) is obvious, so the task is to prove the other inclusion.

Let

$$\begin{aligned} \alpha \in {\text {Int}}( H_{(x,y)}), \end{aligned}$$

so there exists a \(\delta _1>0\) such that

$$\begin{aligned} \Vert \hat{\alpha } - \alpha \Vert <\delta _1 \text { and } \hat{\alpha } \in J^2(X\times Y)|_{(x,y)} \Rightarrow \hat{\alpha }\in H_{(x,y)}.\end{aligned}$$
(60)

By hypothesis there is a \(\delta _2>0\) such that

$$\begin{aligned}&F_x +\left( x'-x, -\frac{\delta _1}{4}, 0,\frac{\delta _1}{4} {\text {Id}}\right) \subset F_{x'} \text { for } \Vert x-x'\Vert <\delta _2 \end{aligned}$$
(61)
$$\begin{aligned}&G_y +\left( x'-x, -\frac{\delta _1}{4}, 0,\frac{\delta _1}{4} {\text {Id}}\right) \subset G_{y'} \text { for } \Vert y-y'\Vert <\delta _2. \end{aligned}$$
(62)

Set \(\delta = \min \{ \delta _1/2,\delta _2/2\}\) and pick any \(\alpha '\in J^2(X\times Y)\) with

$$\begin{aligned} \Vert \alpha '-\alpha \Vert <\delta . \end{aligned}$$

We will show that \(\alpha ' \in H\).

Denote the space coordinate of \(\alpha '\) by \((x',y')\), so \(\alpha '\in J^2(X\times Y)|_{(x',y')}\). Thus we certainly have \(\Vert x'-x\Vert<\delta <\delta _2\) and \(\Vert y-y'\Vert <\delta _2\). Define

$$\begin{aligned} \hat{\alpha } : = \alpha ' +\left( (x-x',y-y'),-\frac{\delta _1}{2},0, -\frac{\delta _1}{2} {\text {Id}}_{n+m}\right) . \end{aligned}$$

Then \(\hat{\alpha } \in J^2(X\times Y)_{(x,y)}\) and

$$\begin{aligned} \Vert \hat{\alpha } - \alpha \Vert \le \Vert \alpha ' - \alpha \Vert + \Vert \hat{\alpha } - \alpha '\Vert < \delta _1. \end{aligned}$$

Thus (60) applies, so \(\hat{\alpha }\in H_{(x,y)}\) which means

$$\begin{aligned} j^* \hat{\alpha } \in G_{y} \text { and } i_U^* \hat{\alpha }\in G_{x} \text { for all } U. \end{aligned}$$

Now using (62).

$$\begin{aligned} j^*\alpha ' = j^*\hat{\alpha } +\left( y'-y,\frac{\delta _1}{2},0, \frac{\delta _1}{2} {\text {Id}}_{m}\right) \in G_y +\left( y'-y,\frac{\delta _1}{2},0, \frac{\delta _1}{2} {\text {Id}}_{m}\right) \subset G_{y'}. \end{aligned}$$

Similarly using the Positivitiy property of F and (61)

$$\begin{aligned} i_U^*\alpha '&= i_U^*\hat{\alpha } + \left( x-x',\frac{\delta _1}{2},0,\frac{\delta _1}{2} {\text {Id}}_{n} + \frac{\delta _1}{2} U^tU\right) \\&\subset i_U^*\hat{\alpha } + \left( x-x',\frac{\delta _1}{2},0,\frac{\delta _1}{2} {\text {Id}}_{n}\right) \subset F_{x'}. \end{aligned}$$

Thus \(\alpha '\in H_{(x',y')}\). As this holds for all such \(\alpha '\) we conclude \(\alpha \in {\text {Int}}(H)\) completing the proof. \(\square \)

1.1 C.1 The Complex Case

Let \(X\subset \mathbb R^{2n}\simeq \mathbb C^n\) be open. If \(f:X\rightarrow \mathbb R\) is twice differentiable at a point \(z\in X \) its complex Hessian is

$$\begin{aligned} {\text {Hess}}^{\mathbb C}_z(f) = \frac{1}{2} ( {\text {Hess}}(f) - \mathbb J {\text {Hess}}_x(f) \mathbb J) \in {\text {Herm}}(\mathbb C^n). \end{aligned}$$

When f is sufficiently smooth we have

$$\begin{aligned} ( {\text {Hess}}^{\mathbb C}_z(f))_{jk} = 2 \frac{\partial ^2 f}{\partial z_j\partial \overline{z}_k}|_z \end{aligned}$$

where, as usual,

$$\begin{aligned} \frac{\partial }{\partial z_j} = \frac{1}{2} \left( \frac{\partial }{\partial x_j} - i \frac{\partial }{\partial y_j}\right) \text { for } z_j = x_j + i y_j. \end{aligned}$$

In terms of the gradient, under the identification \(\mathbb R^{2n}\simeq \mathbb C^n\) we have

$$\begin{aligned} \nabla f|_z = \left( \begin{array}{c} \frac{\partial f}{\partial x}|_z \\ \frac{\partial f}{\partial {y}}|_z \end{array}\right) = 2\frac{\partial f}{\partial \overline{z}}|_z. \end{aligned}$$

Definition C.2

(Complex 2-jet) The complex 2-jet of f at \(z\in X\) is

$$\begin{aligned} J^{2,\mathbb C}_{z} (f) := \left( f(z), 2\frac{\partial f}{\partial \overline{z}}|_z,{\text {Hess}}^{\mathbb C}_z(f)\right) \in J^{2,\mathbb C}_{z} = \mathbb R\times \mathbb C^n\times {\text {Herm}}_n. \end{aligned}$$

So if \(F\subset J^2(X)\) is complex then

$$\begin{aligned} J^2_{z} (f)\in F_z \Longleftrightarrow J^{2,\mathbb C}_{z} (f)\in F_z. \end{aligned}$$

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ross, J., Nyström, D.W. The Minimum Principle for Convex Subequations. J Geom Anal 32, 28 (2022). https://doi.org/10.1007/s12220-021-00782-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s12220-021-00782-2

Keywords

Mathematics Subject Classification

Navigation