Skip to main content

Advertisement

Log in

Adaptive, anisotropic and hierarchical cones of discrete convex functions

  • Published:
Numerische Mathematik Aims and scope Submit manuscript

Abstract

We introduce a new class of adaptive methods for optimization problems posed on the cone of convex functions. Among the various mathematical problems which possess such a formulation, the Monopolist problem (Rochet and Choné, Econometrica 66:783–826, 1998; Ekeland and Moreno-Bromberg, Numer Math 115:45–69, 2010) arising in economics is our main motivation. Consider a two dimensional domain \(\Omega \), sampled on a grid \(X\) of N points. We show that the cone \({{\mathrm{Conv}}}(X)\) of restrictions to \(X\) of convex functions on \(\Omega \) is typically characterized by \(\approx N^2\) linear inequalities; a direct computational use of this description therefore has a prohibitive complexity. We thus introduce a hierarchy of sub-cones \({{\mathrm{Conv}}}(\mathcal{V})\) of \({{\mathrm{Conv}}}(X)\), associated to stencils \(\mathcal{V}\) which can be adaptively, locally, and anisotropically refined. We show, using the arithmetic structure of the grid, that the trace \(U_{|X}\) of any convex function U on \(\Omega \) is contained in a cone \({{\mathrm{Conv}}}(\mathcal{V})\) defined by only \(\mathcal{O}( N \ln ^2 N)\) linear constraints, in average over grid orientations. Numerical experiments for the Monopolist problem, based on adaptive stencil refinement strategies, show that the proposed method offers an unrivaled accuracy/complexity trade-off in comparison with existing methods. We also obtain, as a side product of our theory, a new average complexity result on edge flipping based mesh generation.

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
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9

Similar content being viewed by others

Notes

  1. Precisely, the constraints \(T_x^e\) were omitted in [7] for \(\Vert e\Vert >\sqrt{2}\).

  2. This number of constraints is empirically (and slightly erroneously) estimated to \(\mathcal{O}(N^{1.8})\) in [7].

  3. Strictly speaking, the optimal product Q(z) is an element of the subgradient \(\partial _z U\), which (Lebesgue)almost surely is a singleton \(\{\nabla U(z)\}\). Hence we may write (34) in terms of \(\nabla U(z)\), provided the density \(\mu \) of customers is absolutely continuous with respect to the Lebesgue measure.

  4. With this customer density [25] expected the bunching region to be triangular, and the image \(\nabla U\) to be the union of the segment [(0, 0), (1, 1)] and of the square \([1,2]^2\). After discussion with the author, and in view of the numerical experiments, we believe that these predictions are erroneous.

  5. Optimal solutions of (34) are not uniquely determined outside the customer density support \(T={{\mathrm{supp}}}(\mu )\).

  6. This property was not noticed in the original work [6].

  7. We use the description of \({{\mathrm{GradConv}}}(X)\) by \(\mathcal{O}(N^2)\) linear constraints given in [12], but (for simplicity) not their energy discretization, nor their method for globally extending elements \(u \in {{\mathrm{GradConv}}}(X)\).

  8. The methods (Ob\(_k\))\(_{k\ge 1}\) are closely related to our approach since they produce outputs with zero convexity defect (up to numerical precision), and the number of linear constraints only grows linearly with the domain cardinality: \(\mathcal{O}(k^2 N)\), with \(N:= \#(X)\). We suspect that better results could be obtained with these methods by selecting adaptively and locally the integer k.

  9. The (presumed) non-convergence of the methods OF\(_2 \) and OF\(_3\) is not visible in this graph.

References

  1. Aguilera, N.E., Morin, P.: Approximating optimization problems over convex functions. Numer. Math. 111, 1–34 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  2. Aguilera, N.E., Morin, P.: On convex functions and the finite element method. SIAM J. Numer. Anal. 47, 3139–3157 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  3. Benamou, J.-D. Collino, F., Mirebeau, J.-M.: Monotone and consistent discretization of the Monge–Ampere operator. Math. Comput. (2015, to appear). Preprint arXiv:1409.6694

  4. Bonnans, J.F., Ottenwaelter, E., Zidani, H.: A fast algorithm for the two dimensional HJB equation of stochastic control. Technical report (2004)

  5. Carlier, G.: A general existence result for the principal-agent problem with adverse selection. J. Math. Econ. 35, 129–150 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  6. Carlier, G., Ekeland, I., Touzi, N.: Optimal derivatives design for mean-variance agents under adverse selection. Math. Financ. Econ. 1, 57–80 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  7. Carlier, G., Lachand-Robert, T., Maury, B.: A numerical approach to variational problems subject to convexity constraint. Numer. Math. 88, 299–318 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  8. Chazelle, B.: An optimal convex hull algorithm in any fixed dimension. Discrete Comput. Geom. 5, 449–483 (1993)

    MathSciNet  MATH  Google Scholar 

  9. Choné, P., Le Meur, H.V.J.: Non-convergence result for conformal approximation of variational problems subject to a convexity constraint. Numer. Funct. Anal. Optim. 22, 529–547 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  10. Dobrzynski, C., Frey, P.: Anisotropic Delaunay mesh adaptation for unsteady simulations. In: Proceedings of the 17th International Meshing Roundtable (2008)

  11. Edelsbrunner, H., Seidel, R.: Voronoi diagrams and arrangements. Discrete Comput. Geom. 1, 25–44 (1986)

    Article  MathSciNet  MATH  Google Scholar 

  12. Ekeland, I., Moreno-Bromberg, S.: An algorithm for computing solutions of variational problems with global convexity constraints. Numer. Math. 115, 45–69 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  13. Fehrenbach, J., Mirebeau, J.-M.: Sparse non-negative stencils for anisotropic diffusion. J. Math. Imaging Vis. 49, 123–147 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  14. Figalli, A., Kim, Y.H., McCann, R.J.: When is multidimensional screening a convex program? J. Econ. Theory 146, 454–478 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  15. Hardy, G., Wright, E.M.: An Introduction to the Theory of Numbers. Oxford Science Publications, Oxford (1979)

    MATH  Google Scholar 

  16. Hurtado, F., Noy, M., Urrutia, J.: Flipping edges in triangulations. Discrete Comput. Geom. 22, 333–346 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  17. Lachand-Robert, T., Oudet, E.: Minimizing within convex bodies using a convex hull method. SIAM J. Optim. 16, 368–379 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  18. Manelli, A.M., Vincent, D.R.: Bundling as an optimal selling mechanism for a multiple-good monopolist. J. Econ. Theory 127, 1–35 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  19. Merigot, Q., Oudet, E.: Handling convexity-like constraints in variational problems. SIAM J. Numer. Anal. 52(5), 2466–2487 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  20. Mirebeau. J.-M.: Anisotropic fast-marching on cartesian grids using lattice basis reduction. SIAM J. Numer. Anal. 52(4), 1573–1599 (2014). http://dx.doi.org/10.1137/120861667

  21. Mirebeau, J.-M.: Efficient fast marching with Finsler metrics. Numer. Math. 126, 515–557 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  22. Oberman, A.M.: A numerical method for variational problems with convexity constraints. SIAM J. Sci. Comput. 35, A378–A396 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  23. Oberman, A.M., Friedlander, M.P.: A numerical method for variational problems over the cone of convex functions (2011, preprint). arXiv:1107.5290

  24. Qi, M., Cao, T.-T., Tan, T.-S.: Computing 2D constrained Delaunay triangulation using the GPU. In: The ACM SIGGRAPH Symposium (2012)

  25. Rochet, J.C., Choné, P.: Ironing, sweeping, and multidimensional screening. Econometrica 66, 783–826 (1998)

    Article  MATH  Google Scholar 

  26. Thanassoulis, J.: Haggling over substitutes. J. Econ. Theory 117, 217–245 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  27. Wachsmuth, G.: The numerical solution of Newton’s problem of least resistance. Math. Progr. 147(1–2), 331–350 (2014)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgments

The author thanks Pr. Ekeland and Pr. Rochet for introducing him to the monopolist problem, and the Mosek™ team for their free release policy for public research.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jean-Marie Mirebeau.

Additional information

This work was partly supported by ANR Grant NS-LBR, ANR-13-JS01-0003-01.

Appendices

Appendix A: Convergence of the discretization

For completeness, and at the request of a referee, we reproduce (with minor adaptations) arguments of [7] which establish the convergence of our discrete numerical approximations of the principal agent problem towards the continuous solution, as the grid scale tends to 0.

Let \(\Omega \subseteq {\mathbb R}^2\) be an open bounded convex polygonal domain, and let \((\mathcal{T}_h)_{h >0}\) be a family of regular triangulations of \(\Omega \): there exists a constant C such that, for all \(h >0\), \(T \in \mathcal{T}_h\), one has

$$\begin{aligned} \text {Correct scaling: } {{\mathrm{diam}}}(T)&\le h.&\text {Shape regularity: }{{\mathrm{diam}}}(T) \le&C \rho (T). \end{aligned}$$

We denoted by \({{\mathrm{diam}}}(T)\) the largest distance between two points of a triangle T, and by \(\rho (T)\) its inner circle radius. Let \(X_h\) be the collection of vertices of \(\mathcal{T}_h\), and let \({{\mathrm{\mathrm{I}}}}_h : \mathcal{F}(X_h) \rightarrow C^0(\Omega )\) be the piecewise linear interpolation operator associated to \(\mathcal{T}_h\).

The classical principal agent problem, and its finite element discretization, are as follows:

$$\begin{aligned} E_0&:= \inf \left\{ \mathcal{E}(u); \, u \!\in \! {{\mathrm{Conv}}}(\Omega ), \, u \!\ge \! 0\right\} , \text { with } \mathcal{E}(u)\!:=\! \int _\Omega \left( \frac{1}{2} \Vert \nabla u\Vert ^2 -\langle \nabla u, z\rangle \!+\! u\right) .\\ E_h&:= \inf \{ \mathcal{E}({{\mathrm{\mathrm{I}}}}_h v); \, v \in {{\mathrm{Conv}}}(\Omega ), \, v \ge 0\} = \inf \{ \mathcal{E}({{\mathrm{\mathrm{I}}}}_h v); \, v \in {{\mathrm{Conv}}}(X_h), \, v \ge 0\}. \end{aligned}$$

As proposed in [7], the triangulations \(\mathcal{T}_h\) are fixed, non-adaptive, and solely used for constructing the discrete energies \(u \mapsto \mathcal{E}({{\mathrm{\mathrm{I}}}}_h u)\). In particular \({{\mathrm{\mathrm{I}}}}_h u\) need not be convex; instead we use the rigidity result Proposition 8.1 to establish the convexity of the weak limit of the minimizers to the discretized problems. Our numerical experiments all involve rectangular domains \(\Omega \) and discretization points \(X_h\) distributed on a cartesian grid, for which \(\mathcal{T}_h\) is the standard structured mesh built of half squares.

Proposition 8.1

For all \(h>0\) let \(v_h \in {{\mathrm{Conv}}}(\Omega )\) and \(u_h := {{\mathrm{\mathrm{I}}}}_h v_h\). Assume that \(u_h \rightarrow u_0\) as \(h \rightarrow 0\) in the sense of distributions. Then \(u_0\) is convex, and both \(u_h\) and \(v_h\) converge uniformly to \(u_0\) on compact subsets of \(\Omega \).

Proof

We fix a radius \(r>0\), and limit our attention to grid scales \(0<h\le r\).

Non-uniform bound on \(u_h\). There exists M, independent of \(0<h \le r\), such that for any ball \(B(x,r) \subseteq \Omega \), one does not have \(u_h \ge M\) (resp. \(u_h \le -M\)) uniformly on B(xr). Indeed otherwise there exists a sequence \(h_n \rightarrow 0\), \(M_n\rightarrow \infty \), \(B(x_n,r) \subseteq \Omega \), such that \(u_n :=u_{h_n} \ge M_n\) (resp. \(\le -M_n\)) on \(B(x_n,r)\). By compactness we can assume that \(x_n \rightarrow x_* \in \Omega \). Considering a smooth bump function \(\varphi \) supported on \(B(x_*,r)\) we obtain \(\int u_n \varphi \rightarrow + \infty \) (resp. \(-\infty \)) which contradicts the distributional convergence of \(u_h\) as \(h\rightarrow 0\).

Non-uniform bound on \(v_h\). If for contradiction \(|v_h| \ge M\) uniformly on a ball \(B(z,2 r)\subseteq \Omega \), then by continuity \(v_h \ge M\) (resp. \(v_h \le -M\)) on B(z, 2r). Thus by interpolation \(u_h = {{\mathrm{\mathrm{I}}}}_h v_h \ge M\) (resp. \(u_h \le -M\)) on B(zr), which was excluded.

Upper bound on \(v_h\). Consider a ball \(B(z,4r) \subseteq \Omega \), a subgradient \(g \in \partial _z v_h\), and note that by convexity \(v_h(w) \ge v_h(z)+\langle w-z,g\rangle \ge v_h(z)\) for all \(w \in B(z+2r g / \Vert g\Vert , 2r) \subseteq B(z,4r) \subseteq \Omega \). From the non-uniform bound on \(v_h\), we conclude that \(v_h(z) \le M\).

Lower bound on \(v_h\). Consider a ball \(B(x,10 r) \subseteq \Omega \). By the non-uniform bound on \(v_h\), there exists \(y \in B(x,2r)\) such that \(v_h(y) \ge -M\). Denoting \(z := y+2(y-x)\) one has by convexity \(v_h(z) \ge v_h(y)+2 (v_h(y)-v_h(x))\). In addition \(B(z,4 r) \subseteq B(y,8 r) \subseteq B(x,10 r) \subseteq \Omega \), thus \(M \ge v_h(z) \ge v_h(y)+2 (v_h(y)-v_h(x)) \ge -3 M - 2 v_h(x)\), so that \(v_h(x) \ge -2 M\).

Lipschitz bound on \(v_h\). Consider a ball \(B(x,10r) \subseteq \Omega \), and a sub-gradient element \(g \in \partial _x v_h\). Let \(y := x+r g/\Vert g\Vert \). Then by convexity \(r \Vert g\Vert = \langle g, y-x\rangle \le v_h(y)-v_h(x) \le M-(-2M)=3M\), thus \(\Vert g\Vert \le C_L := 3M/r\). In particular, if \(B(x,11 r) \subseteq \Omega \) and \(x \in T \in \mathcal{T}_h\), then \(v_h\) is \(C_L\)-Lipschitz on T, so that by interpolation \(|u_h(x) - v_h(x)| \le C_L h\).

Conclusion. Consider the convex compact \(K := \{x \in \Omega ; \, B(x,11r)\}\). The maps \((v_h)_{0<h \le r}\) are uniformly bounded and Lipschitz on K, hence they form by Arzela–Ascoli’s theorem a pre-compact family: from any sub-sequence, one can extract a sub-sub-sequence which converges uniformly on K to a limit \(v_*\) as \(h \rightarrow 0\). Since \(\Vert u_h-v_h\Vert _{L^\infty (K)} \rightarrow 0\), the sub-sub-sequence limit can only be \(v_* = u_0\). This implies that \(u_0\) is convex and that \(v_h \rightarrow u_0\) uniformly on K as \(h \rightarrow 0\). Thus also \(u_h \rightarrow u_0\) uniformly on K, which concludes the proof. \(\square \)

The next steps involve the precise form of the optimized energy, which for non-negative \(u : \Omega \rightarrow {\mathbb R}\) reads

$$\begin{aligned} \mathcal{E}(u) = \Vert u\Vert _{L^1(\Omega )} + \frac{1}{2} \Vert \nabla u-z\Vert _{L^2(\Omega )} - \Vert z\Vert _{L^2(\Omega )} \end{aligned}$$
(38)

Lemma 8.2

(Upper estimate) One has \(\limsup _{h \rightarrow 0} E_h \le E_0\).

Proof

Consider \(u \in {{\mathrm{Conv}}}(\Omega )\), non-negative, and such that \(\mathcal{E}(u) < \infty \). In particular, \(u \in H^1(\Omega )\), by (38). Let \(x_*\) be a fixed point in the interior of \(\Omega \), let \(\alpha >0\), and let \(u_\alpha := u \circ D_\alpha \), where \(D_\alpha \) is the dilation of center \(x_*\) and factor \(1/(1+\alpha )\). By construction \(u_\alpha \) is convex, non-negative, and defined on a neighborhood of \(\Omega \). Furthermore \(\mathcal{E}(u_\alpha ) \rightarrow \mathcal{E}(u)\) as \(\alpha \rightarrow 0\), because \(u_\alpha \rightarrow u\) strongly in \(H^1(\Omega )\).

Fix \(\alpha \). Let \(\beta >0\), and let \(u_\alpha ^\beta := u_\alpha * K_\beta \), where \(K_\beta (x) := \frac{1}{\beta ^2} K( \frac{x}{\beta })\), and K is a mollifier: a smooth non-negative function of unit integral and supported on the unit ball. Since \(u_\alpha \) is defined on a neighborhood of \(\Omega \), the convolution \(u_\alpha * K_\beta \) is well defined on \(\Omega \) for sufficiently small \(\beta \). By construction \(u_\alpha ^\beta \) is convex, non-negative, and smooth. Furthermore \(\mathcal{E}(u_\alpha ^\beta ) \rightarrow \mathcal{E}(u_\alpha )\) as \(\beta \rightarrow 0\), because \(u_\alpha ^\beta \rightarrow u_\alpha \) strongly in \(H^1(\Omega )\).

Fix \(\alpha \) and \(\beta \). Since \(u_\alpha ^\beta \) is smooth one has \({{\mathrm{\mathrm{I}}}}_h u_\alpha ^\beta \rightarrow u_\alpha ^\beta \) strongly in \(H^1(\Omega )\) as \(h \rightarrow 0\), hence \(\mathcal{E}({{\mathrm{\mathrm{I}}}}_h u_\alpha ^\beta ) \rightarrow \mathcal{E}(u_\alpha ^\beta )\). Thus \(L := \limsup _{h \rightarrow 0} E_h \le \mathcal{E}(u_\alpha ^\beta )\). Letting \(\beta \rightarrow 0\) yields \(L\le \mathcal{E}(u_\alpha )\). Letting \(\alpha \rightarrow 0\) yields \(L \le \mathcal{E}(u)\). Optimizing over u yields the announced result. \(\square \)

Proposition 8.3

(Convergence of the discretization) One has \(\lim _{h\rightarrow 0} E_h = E_0\). For each \(h>0\) consider a minimizer \(v_h \in {{\mathrm{Conv}}}(X_h)\) for \(E_h\), and denote \(u_h = {{\mathrm{\mathrm{I}}}}_h v_h\). Then \((u_h)_{h > 0}\) is bounded in \(H^1(\Omega )\), and the limit of any weakly converging subsequence is a minimizer for \(E_0\).

Proof

In view of (38), the minimizers \(v_h\) exist by a finite dimensional compactness argument. Using again (38), and noting that \(\mathcal{E}(u_h) \le \mathcal{E}(0) < \infty \), we obtain as announced that \(u_h\) is bounded in \(H^1(\Omega )\). Since \(H^1(\Omega )\) is a Hilbert space, its bounded sets are pre-compact for the weak topology. Assuming, up to extraction, that \(u_h \rightarrow u_0\) weakly in \(H^1(\Omega )\) as \(h\rightarrow 0\) we obtain

$$\begin{aligned}&\lim _{h \rightarrow 0} \int _{\Omega } \langle \nabla u_h,z\rangle = \int _{\Omega } \langle \nabla u_0, z\rangle , \quad \lim _{h \rightarrow 0} \Vert u_h\Vert _{L^1(\Omega )} = \Vert u_0\Vert _{L^1(\Omega )},\\&\limsup _{h \rightarrow 0} \Vert \nabla u_h\Vert _{L^2(\Omega )} \ge \Vert \nabla u_0\Vert _{L^2(\Omega )}. \end{aligned}$$

We used weak \(H^1(\Omega )\) convergence for the first identity, the strong \(L^1(\Omega )\) convergence for the second identity (due to the compact inclusion \(H^1(\Omega ) \rightarrow L^1(\Omega )\)), and the weak lower semi-continuity of the \(H^1(\Omega )\) semi-norm for the third inequality. It follows that \(\mathcal{E}(u_0) \le \limsup _{h \rightarrow 0} \mathcal{E}(u_h) = \limsup _{h \rightarrow 0} E_h \le E_0\), using Lemma 8.2. The limit \(u_0\) non-negative, and convex by Proposition 8.1, hence it is a minimizer for \(E_0\) which concludes the proof. Note that by Proposition 8.1 the sub-sequence’s convergence is in fact locally uniform. \(\square \)

Minor modifications of the proof allow to address other optimization problems.

  • Projections in the \(H^1(\Omega )\) norm Sect. 7.3: the proof applies with almost no change.

  • Projections in the \(L^2(\Omega )\) norm Sect. 7.3: the Hilbert space \(H^1(\Omega )\) needs to be replaced with \(L^2(\Omega )\) in all steps of the proof, which does not introduce new difficulties.

  • Product bundles and lottery tickets Sect. 7.2. This model involves the Lipschitz regularity constraint \({\mathbb P}(u) : \forall x \in \Omega , \nabla u(x) \in [0,1]^2\). We make the additional assumption that each triangle \(T \in \mathcal{T}_h\), \(h>0\), has two edges aligned with the grid axes, so as to preserve this constraint: \({\mathbb P}(u) \Rightarrow {\mathbb P}({{\mathrm{\mathrm{I}}}}_h u)\). Lemma 8.2 is then easily adapted. In Proposition 8.3, replace Hilbert weak compactness with Arzela–Ascoli compactness, and rewrite the optimized functional without gradient terms:

    $$\begin{aligned} \int _\Omega (u - \langle \nabla u,z\rangle ) = \int _\Omega (u - {{\mathrm{div}}}(u z)+u{{\mathrm{div}}}z) = 3 \int _\Omega u - \int _{\partial \Omega } u \langle z,n\rangle . \end{aligned}$$

The convergence of higher order quantities, such as \(\nabla u\) or \(\det (\nabla ^2 u)\), is empirically observed but not claimed. The more complex functional (37) is not easily expressed using norms of classical functional spaces, and its analysis is cumbersome [6]; the convergence of the numerical method is thus left as an open question for this specific model.

Appendix B: Directional convexity

We briefly explore a weaker notion of discrete convexity [22, 23] based solely on the directional constraints \(S_z^e\), and omitting the triangular constraints \(T_z^e\).

Definition 9.1

We denote by \({{\mathrm{DConv}}}(X)\) the collection of elements in \(\mathcal{F}(X)\) on which all the linear forms \(S_x^e\), \(x \in X\), \(e\in {\mathbb Z}^2\) irreducible, supported on X, take non-negative values.

Some elements of \({{\mathrm{DConv}}}(X)\) cannot be extended into global convex maps on \({{\mathrm{Hull}}}(X)\). Their existence follows from the second point of Theorem 2.5 (minimality of the collection of constraints \(S_x^e\), \(T_x^e\)), but for completeness we give (without proof) a concrete example.

Proposition 9.2

Let \(u \in \mathcal{F}({\mathbb Z}^2)\) be defined by \(u(1,1) = 1\), \(u(-1,0)=u(0,-1) = -1\), and \(u(x) := 2\Vert x\Vert ^2\) for other \(x \in {\mathbb Z}^2\). Then for all \(x \in {\mathbb Z}^2\), and all irreducible \(e \in {\mathbb Z}^2\), one has \(S_x^e (u) \ge 1\), and \(T_x^e(u) \ge 2\) if \(\Vert e\Vert >1\), with the exception \(T_0^{(1,1)}(u) = -1\).

Elements of \({{\mathrm{DConv}}}(X)\) are nevertheless “almost” convex, in the sense that their restriction to a coarsened grid is convex. This might be a starting point for adapting the convergence analysis of Appendix A, which so far is limited to \({{\mathrm{Conv}}}(X)\).

Proposition 9.3

If \(u \in {{\mathrm{DConv}}}(X)\), then \(u_{|X'} \in {{\mathrm{Conv}}}(X')\), with \(X' : =X \cap 2 {\mathbb Z}^2\).

Proof

Let \(x \in X\), and let \(e \in {\mathbb Z}^2\), \(\Vert e\Vert >1\), be irreducible and of parents fg. Assuming that \(x+2 e, x-2 f, x-2 g \in X\), and observing that \(x \pm e \in X\) by convexity, we obtain

$$\begin{aligned} u(x+2 e)+u(x-2 f) +u(x- 2 g) - 3 u(x) = 2 S_x^e(u) + S_{x+e}^e(u) + S_{x-e}^{f-g}(u) \ge 0. \end{aligned}$$

Likewise \(u(x+2 e)- 2 u(x) + u(x-2 e) = S_{x-e}^e(u)+2 S_x^e(u)+S_{x+e}^e(u) \ge 0\). \(\square \)

The cone \({{\mathrm{DConv}}}(X)\) of directionally convex functions admits, just like \({{\mathrm{Conv}}}(X)\), a hierarchy of sub-cones \({{\mathrm{DConv}}}(\mathcal{V})\) associated to stencils. It is used in [3] to construct a monotone discretization of the Monge-Ampere operator.

Definition 9.4

Let \(\mathcal{V}\) be a family of stencils on X, and let \(u \in \mathcal{F}(X)\). The cone \({{\mathrm{DConv}}}(\mathcal{V})\) is defined by the non-negativity of the following linear forms: for all \(x \in X\)

  • For all \(e \in \mathcal{V}(x)\), the linear form \(S_x^e\), if supported on X.

  • For all \(e \in \hat{\mathcal{V}}(x)\), the linear form \( H_x^e := P_x^e + P_x^{-e} \), if supported on X.

Proposition 9.5

  • For any stencils \(\mathcal{V}, \mathcal{V}'\), one has \({{\mathrm{DConv}}}(\mathcal{V}) \subseteq {{\mathrm{DConv}}}(X)\), \({{\mathrm{DConv}}}(\mathcal{V}) \cap {{\mathrm{DConv}}}(\mathcal{V}') = {{\mathrm{DConv}}}(\mathcal{V}\cap \mathcal{V}')\), and \({{\mathrm{DConv}}}(\mathcal{V}) \cup {{\mathrm{DConv}}}(\mathcal{V}') \subseteq {{\mathrm{DConv}}}(\mathcal{V}\cup \mathcal{V}')\).

  • For any stencils \(\mathcal{V}\), one has

    $$\begin{aligned} {{\mathrm{DConv}}}(\mathcal{V}) \!=\! \{u \in {{\mathrm{DConv}}}(X); \, H_x^e(u) \!\ge \! 0 \text { for all } x \in X, \, e \in \mathcal{V}_{\max }(x) {\setminus } \mathcal{V}(x)\}.\nonumber \\ \end{aligned}$$
    (39)

Proof

As observed in Sect. 4.2, the second point of this proposition implies the first one. We denote by \({\mathbb P}(\mathcal{V})\) the identity (39), and prove it by decreasing induction over \(\#(\mathcal{V})\). Since \({\mathbb P}(\mathcal{V}_{\max })\) clearly holds, we consider stencils \(\mathcal{V}\subsetneq \mathcal{V}_{\max }\).

Let \(x \in X\), \(e \in \mathcal{V}_{\max }(x) {\setminus } \mathcal{V}(x)\), be such that \(\Vert e\Vert \) is minimal. Similarly to Proposition 4.9 we find that e belongs to the set \(\hat{\mathcal{V}}(x)\) of candidates for refinement at x, and define stencils \(\mathcal{V}'\) by \(\mathcal{V}(x) := \mathcal{V}(x) \cup \{e\}\), and \(\mathcal{V}'(y) := \mathcal{V}(y)\) for \(y \ne x\). The cones \({{\mathrm{DConv}}}(\mathcal{V})\) and \({{\mathrm{DConv}}}(\mathcal{V}')\) are defined by a common collection of constraints, with the addition respectively of \(H_x^e\) for \({{\mathrm{DConv}}}(\mathcal{V})\), and \(S_x^e\), \(H_x^{e+f}\), \(H_x^{e+g}\) for \({{\mathrm{DConv}}}(\mathcal{V}')\). Expressing the latter linear forms as combinations of those defining \({{\mathrm{DConv}}}(\mathcal{V})\)

$$\begin{aligned} S_x^e = H_x^e + S_x^f + S_x^g,\quad H_x^{e+f} = H_x^e + S_{x+e}^f + S_{x-e}^f, \quad H_x^{e+g} = H_x^e + S_{x+e}^g + S_{x-e}^g, \end{aligned}$$

and observing that \({\mathbb P}(\mathcal{V}')\) holds by induction, we conclude the proof of \({\mathbb P}(\mathcal{V})\). \(\square \)

Appendix C: Numerical results

See Figs. 10, 11, 12, 13, 14.

Fig. 10
figure 10

Results for the classical principal agent model, with a uniform density of customers on the domain obtained by rotating the square \([1,2]^2\) around its center (3 / 2, 3 / 2) by the indicated angle \(\theta \). From top to bottom level sets of U (vanishing set indicated in white), level sets of \(\det (\nabla ^2U)\) (allows to discriminate between the three categories of customers), density of products bought (which explodes on some curves), margin of the monopolist

Fig. 11
figure 11

Three dimensional plot of the optimal U for the product-bundles variant of the monopolist problem, with respect to various customer distributions. Left uniform distribution on \([0,1]^2\). Center, and right uniform distribution on the illustrated black polygon

Fig. 12
figure 12

Optimal pricing of risky assets, with the parameters \(\alpha =1\), \(\xi =1\), and a uniform customer density on \([1,2]^2\). Top left level sets of the estimated solution U of (37), with exclusion region \(U=0\) in white. Top right level sets of \(\det (\nabla ^2U)\), the darkest one is the estimated bunching region \(\det (\nabla ^2U)=0\). Bottom left and right (detail): optimal product line, colored with the monopolist margin (color figure online)

Fig. 13
figure 13

Top the projected function \(u_0\) is the sum of three gaussian bumps arranged at the vertices of a triangle. The \(L^2\) projection u inherits this triangular structure: curvature is concentrated at the triangle vertices, and the convexity is degenerate along the triangle edges as revealed by the u-Delaunay triangulation. Middle the projected function is a centered gaussian bump. The \(L^2\) projection is non-smooth along the four diagonals, in contrast with the \(H^1\) projection. Bottom \(L^2\) projection onto the convex functions is used as a parameterless denoising algorithm, so as to recover the orientation and axes-ratio of a convex quadratic function. Experiments on \(50\times 50\) grids, except for the \(20\times 20\) grid top right

Fig. 14
figure 14

Comparison of different numerical methods for the classical Monopolist problem

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mirebeau, JM. Adaptive, anisotropic and hierarchical cones of discrete convex functions. Numer. Math. 132, 807–853 (2016). https://doi.org/10.1007/s00211-015-0732-7

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00211-015-0732-7

Mathematics Subject Classification

Navigation