Skip to main content
Log in

A proximal method for composite minimization

  • Full Length Paper
  • Series A
  • Published:
Mathematical Programming Submit manuscript

Abstract

We consider minimization of functions that are compositions of convex or prox-regular functions (possibly extended-valued) with smooth vector functions. A wide variety of important optimization problems fall into this framework. We describe an algorithmic framework based on a subproblem constructed from a linearized approximation to the objective and a regularization term. Properties of local solutions of this subproblem underlie both a global convergence result and an identification property of the active manifold containing the solution of the original problem. Preliminary computational results on both convex and nonconvex examples are promising.

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

Similar content being viewed by others

References

  1. Bolte, J., Daniilidis, A., Lewis, A.S.: Generic optimality conditions for semialgebraic convex problems. Math. Oper. Res. 36, 55–70 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  2. Bonnans, J.F., Shapiro, A.: Perturbation Analysis of Optimization Problems. Springer Series in Operations Research. Springer, Berlin (2000)

    Book  MATH  Google Scholar 

  3. Burke, J.V.: Descent methods for composite nondifferentiable optimization problems. Math. Program. Ser. A 33, 260–279 (1985)

    Article  MathSciNet  MATH  Google Scholar 

  4. Burke, J.V.: On the identification of active constraints II: the nonconvex case. SIAM J. Numer. Anal. 27, 1081–1102 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  5. Burke, J.V., Moré, J.J.: On the identification of active constraints. SIAM J. Numer. Anal. 25, 1197–1211 (1988)

    Article  MathSciNet  MATH  Google Scholar 

  6. Byrd, R., Gould, N.I.M., Nocedal, J., Waltz, R.A.: On the convergence of successive linear-quadratic programming algorithms. SIAM J. Optim. 16, 471–489 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  7. Cai, J.-F., Candès, E., Shen, Z.: A singular value thresholding algorithm for matrix completion. SIAM J. Optim. 20, 1956–1982 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  8. Candès, E., Recht, B.: Exact matrix completion via convex optimization. Found. Comput. Math. 9, 717–772 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  9. Candès, E.J.: Compressive sampling. In: Proceedings of the International Congress of Mathematicians, Madrid (2006)

  10. Chen, S.S., Donoho, D.L., Saunders, M.A.: Atomic decomposition by basis pursuit. SIAM J. Sci. Comput. 20, 33–61 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  11. Combettes, P., Pennanen, T.: Proximal methods for cohypomonotone operators. SIAM J. Control Optim. 43, 731–742 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  12. Combettes, P.L., Wajs, V.R.: Signal recovery by proximal forward–backward splitting. Multiscale Model. Simul. 4, 1168–1200 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  13. Daniilidis, A., Hare, W., Malick, J.: Geometrical interpretation of the predictor-corrector type algorithms in structured optimization problems. Optimization 55, 481–503 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  14. Dmitruk, A.V., Kruger, A.Y.: Metric regularity and systems of generalized equations. J. Math. Anal. Appl. 342, 864–873 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  15. Dontchev, A.L., Lewis, A.S., Rockafellar, R.T.: The radius of metric regularity. Trans. Am. Math. Soc. 355, 493–517 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  16. Efron, B., Hastie, T., Johnstone, I., Tibshirani, R.: Least angle regression. Ann. Stat. 32, 407–499 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  17. Fan, J., Li, R.: Variable selection via nonconvex penalized likelihood and its oracle properties. J. Am. Stat. Assoc. 96, 1348–1361 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  18. Fletcher, R., Sainz de la Maza, E.: Nonlinear programming and nonsmooth optimization by successive linear programming. Math. Program. 43, 235–256 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  19. Friedlander, M.P., Gould, N.I.M., Leyffer, S., Munson, T.S.: A filter active-set trust-region method, Preprint ANL/MCS-P1456-0907, Mathematics and Computer Science Division, Argonne National Laboratory, 9700 S. Cass Avenue, Argonne IL 60439, September 2007

  20. Fukushima, M., Mine, H.: A generalized proximal point algorithm for certain nonconvex minimization problems. Int. J. Syst. Sci. 12, 989–1000 (1981)

    Article  MathSciNet  MATH  Google Scholar 

  21. Hale, E.T., Yin, W., Zhang, Y.: A fixed-point continuation method for \(\ell _1\)-minimization: methodology and convergence. SIAM J. Optim. 19, 1107–1130 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  22. Hare, W., Lewis, A.: Identifying active constraints via partial smoothness and prox-regularity. J. Convex Anal. 11, 251–266 (2004)

    MathSciNet  MATH  Google Scholar 

  23. Iusem, A., Pennanen, T., Svaiter, B.: Inexact variants of the proximal point algorithm without monotonicity. SIAM J. Optim. 13, 1080–1097 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  24. Jokar, S., Pfetsch, M.E.: Exact and approximate sparse solutions of underdetermined linear equations. SIAM J. Sci. Comput. 31, 23–44 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  25. Kaplan, A., Tichatschke, R.: Proximal point methods and nonconvex optimization. J. Glob. Optim. 13, 389–406 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  26. Kim, T., Wright, S.J.: An \(\text{ S }\ell _1\text{ LP }\)-active set approach for feasibility restoration in power systems, tech. rep., Computer Science Department, University of Wisconsin-Madison, May 2014. arXiv:1405.0322

  27. Lan, G.: Bundle-level type methods uniformly optimal for smooth and nonsmooth convex optimization. Math. Program. Ser. A 149, 1–45 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  28. Lemaréchal, C., Oustry, F., Sagastizábal, C.: The \({\cal {U}}\)-Lagrangian of a convex function. Trans. Am. Math. Soc. 352, 711–729 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  29. Levy, A.: Lipschitzian multifunctions and a Lipschitzian inverse mapping theorem. Math. Oper. Res. 26, 105–118 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  30. Lewis, A.: Active sets, nonsmoothness, and sensitivity. SIAM J. Optim. 13, 702–725 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  31. Mangasarian, O.L.: Minimum-support solutions of polyhedral concave programs. Optimization 45, 149–162 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  32. Martinet, B.: Régularisation d’inéquations variationnelles par approximations successives. Rev. Française Informat. Recherche Opérationnelle 4, 154–158 (1970)

    MathSciNet  MATH  Google Scholar 

  33. Mifflin, R., Sagastizábal, C.: A VU-algorithm for convex minimization. Math. Program. Ser. B 104, 583–608 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  34. Miller, S.A., Malick, J.: Newton methods for nonsmooth convex minimization: connections among \({\cal {U}}\)-Lagrangian, Reimannian Newton, and SQP methods. Math. Program. Ser. B 104, 609–633 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  35. Mordukhovich, B.: Variational Analysis and Generalized Differentiation, I: Basic Theory; II: Applications. Springer, New York (2006)

    Google Scholar 

  36. Pennanen, T.: Local convergence of the proximal point algorithm and multiplier methods without monotonicity. Math. Oper. Res. 27, 170–191 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  37. Recht, B., Fazel, M., Parrilo, P.: Guaranteed minimum-rank solutions of matrix equations via nuclear norm minimization. SIAM Rev. 52, 471–501 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  38. Rockafellar, R.: Monotone operators and the proximal point algorithm. SIAM J. Control Optim. 14, 877–898 (1976)

    Article  MathSciNet  MATH  Google Scholar 

  39. Rockafellar, R.T.: Convex Analysis. Princeton University Press, Princeton (1970)

    Book  MATH  Google Scholar 

  40. Rockafellar, R.T., Wets, R.J.: Variational Analysis. Springer, Berlin (1998)

    Book  MATH  Google Scholar 

  41. Rudin, L., Osher, S., Fatemi, E.: Nonlinear total variation based noise removal algorithms. Phys. D 60, 259–268 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  42. Sagastizábal, C.: Composite proximal bundle method. Math. Program. Ser. B 140, 189–233 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  43. Sagastizábal, C., Mifflin, R.: Proximal points are on the fast track. J. Convex Anal. 9, 563–579 (2002)

    MathSciNet  MATH  Google Scholar 

  44. Shapiro, A.: On a class of nonsmooth composite functions. Math. Oper. Res. 28, 677–692 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  45. Shi, W., Wahba, G., Wright, S.J., Lee, K., Klein, R., Klein, B.: LASSO-Patternsearch algorithm with application to opthalmology data. Stat. Interface 1, 137–153 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  46. Spingarn, J.: Submonotone mappings and the proximal point algorithm. Numer. Funct. Anal. Optim. 4, 123–150 (1981/82)

  47. Tibshirani, R.: Regression shrinkage and selection via the LASSO. J. R. Stat. Soc. B 58, 267–288 (1996)

    MathSciNet  MATH  Google Scholar 

  48. Wen, Z., Yin, W., Zhang, H., Goldfarb, D.: On the convergence of an active set method for \(\ell _1\) minimization. SIAM J. Sci. Comput. 32, 1832–1857 (2010)

    Article  MathSciNet  Google Scholar 

  49. Wright, S.J.: Convergence of an inexact algorithm for composite nonsmooth optimization. IMA J. Numer. Anal. 9, 299–321 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  50. Wright, S.J.: Identifiable surfaces in constrained optimization. SIAM J. Control Optim. 31, 1063–1079 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  51. Wright, S.J., Nowak, R.D., Figueiredo, M.A.T.: Sparse reconstruction by separable approximation. IEEE Trans. Signal Process. 57, 2479–2493 (2009)

    Article  MathSciNet  Google Scholar 

  52. Yuan, Y.: Conditions for convergence of a trust-region method for nonsmooth optimization. Math. Program. 31, 220–228 (1985)

    Article  MATH  Google Scholar 

  53. Yuan, Y.: On the superlinear convergence of a trust region algorithm for nonsmooth optimization. Math. Program. 31, 269–285 (1985)

    Article  MathSciNet  MATH  Google Scholar 

  54. Zhang, C.H.: Nearly unbiased variable selection under minimax concave penalty. Ann. Stat. 38, 894–942 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  55. Zimmerman, R.D., Murillo-Sánchez, C.E., Thomas, R.J.: MATPOWER: steady-state operations, planning, and analysis tools for power systems research and education. IEEE Trans. Power Syst. 26, 12–19 (2011)

    Article  Google Scholar 

Download references

Acknowledgments

We acknowledge the support of NSF Grants 0430504 and DMS-0806057. We are grateful for the comments of two referees, which were most helpful in revising earlier versions. We thank Mr. Taedong Kim for obtaining computational results for the formulation (6.4).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to A. S. Lewis.

Additional information

A.S. Lewis’s research supported in part by NSF Award DMS-1208338.

S.J. Wright’s research supported in part by NSF Awards DMS-1216318 and IIS-1447449, ONR Award N00014-13-1-0129, AFOSR Award FA9550-13-1-0138, and Subcontract 3F-30222 from Argonne National Laboratory.

Appendix

Appendix

The basic building block for variational analysis (see Rockafellar and Wets [40] or Mordukhovich [35]) is the normal cone to a (locally) closed set S at a point \(s \in S\), denoted by \(N_S(s)\). It consists of all normal vectors: limits of sequences of vectors of the form \(\lambda (u-v)\) for points \(u,v \in \mathfrak {R}^m\) approaching s such that v is a closest point to u in S, and scalars \(\lambda > 0\). On the other hand, tangent vectors are limits of sequences of vectors of the form \(\lambda (u-s)\) for points \(u \in S\) approaching s and scalars \(\lambda > 0\). The set S is Clarke regular at s when the inner product of any normal vector with any tangent vector is always nonpositive. Closed convex sets and smooth manifolds are everywhere Clarke regular.

The epigraph of a function \(h:\mathfrak {R}^m \rightarrow {\bar{\mathfrak {R}}}\) is the set

$$\begin{aligned} \text{ epi }\,h=\{(c,r)\in \mathfrak {R}^m\times \mathfrak {R}:r\ge h(c)\}. \end{aligned}$$

If the value of h is finite at some point \(\bar{c} \in \mathfrak {R}^m\), then h is lower semicontinuous nearby if and only if its epigraph is locally closed around the point \(\big (\bar{c}, h(\bar{c})\big )\). Henceforth we focus on that case.

The subdifferential of h at \(\bar{c}\) is the set

$$\begin{aligned} \partial h(\bar{c})=\big \{v\in \mathfrak {R}^m\,:\,(v,-1) \in N_{\mathrm{epi}\,h}(\bar{c},h\big (\bar{c})\big ) \big \} \end{aligned}$$

and the horizon subdifferential is

$$\begin{aligned} \partial ^{\infty } h(\bar{c})=\big \{v\in \mathfrak {R}^m:(v,0) \in N_{\mathrm{epi}\,h}\big (\bar{c},h(\bar{c})\big ) \big \} \end{aligned}$$
(6.5)

(see [40, Theorem 8.9]). The function h is subdifferentially regular at \(\bar{c}\) if its epigraph is Clarke regular at \(\big (\bar{c}, h(\bar{c})\big )\) (as holds in particular if h is convex lower semicontinuous, or smooth). Subdifferential regularity implies that \(\partial h(\bar{c})\) is a closed and convex set in \(\mathfrak {R}^m\), and its recession cone is exactly \(\partial ^{\infty } h(\bar{c})\) (see [40, Corollary 8.11]). In the case when h is locally Lipschitz, it is almost everywhere differentiable: h is then subdifferentially regular at \(\bar{c}\) if and only if its directional derivative for every direction \(d \in \mathfrak {R}^m\) equals

$$\begin{aligned} \limsup _{c \rightarrow \bar{c}} \langle \nabla h(c),d \rangle , \end{aligned}$$

where the \(\limsup \) is taken over points c where h is differentiable.

Consider a subgradient \(\bar{v} \in \partial h(\bar{c})\), and a localization of the subdifferential mapping \(\partial h\) around the point \((\bar{c},\bar{v})\), by which we mean a set-valued mapping \(T:\mathfrak {R}^m \rightrightarrows \mathfrak {R}^m\) defined by

$$\begin{aligned} T(y)=\left\{ \begin{array}{ll} \partial h(y) \cap B_{\epsilon }(\bar{v}) &{} (|y - \bar{c}| \le \epsilon ,\,|h(y) - h(\bar{c})| \le \epsilon ) \\ \emptyset &{}(\text{ otherwise }) \end{array}\right. \end{aligned}$$

for some constant \(\epsilon >0\). The function h is prox-regular at \(\bar{c}\) for \(\bar{v}\) if some such localization is hypomonotone: that is, for some constant \(\rho > 0\), we have

$$\begin{aligned} z \in T(y)\quad \text{ and }\quad z'\in T(y')\Rightarrow \langle z'-z,y'-y \rangle \ge -\rho |y'-y|^2. \end{aligned}$$

This definition is equivalent to Definition 1.1 (with the same constant \(\rho \)) [40, Example 12.28 and Theorem 13.36]. Prox-regularity at \(\bar{c}\) (for all subgradients v) implies subdifferential regularity.

A general class of prox-regular functions common in engineering applications is “lower \({\mathcal {C}}^2\)” functions [40, Definition 10.29]. A function \(h:\mathfrak {R}^m \rightarrow \mathfrak {R}\) is lower \({\mathcal {C}}^2\) around a point \(\bar{c} \in \mathfrak {R}^m\) if h has the local representation

$$\begin{aligned} h(c)=\max _{t\in T}f(c,t)\quad \text{ for }\,c\in \mathfrak {R}^m\,\text{ near }\,{\bar{c}}, \end{aligned}$$

for some function \(f:\mathfrak {R}^m \times T \rightarrow \mathfrak {R}\), where the space T is compact and the quantities f(ct), \(\nabla _c f(c,t)\), and \(\nabla ^2_{cc} f(c,t)\) all depend continuously on (ct). All lower \({\mathcal {C}}^2\) functions are prox-regular [40, Proposition 13.3]. A simple equivalent property, useful in theory though harder to check in practice, is that h has the form \(g-\kappa |\cdot |^2\) around the point \(\bar{c}\) for some continuous convex function g and some constant \(\kappa \).

The normal cone is crucial to the definition of another central variational-analytic tool. Given a set-valued mapping \(F : \mathfrak {R}^p \rightrightarrows \mathfrak {R}^q\) with closed graph,

$$\begin{aligned} \text{ gph }\,F=\{(u,v):v\in F(v)\}, \end{aligned}$$

at any point \((\bar{u},\bar{v}) \in \text{ gph }\,F\), the coderivative \(D^*F(\bar{u}|\bar{v}):\mathfrak {R}^q \rightrightarrows \mathfrak {R}^p\) is defined by

$$\begin{aligned} w \in D^* F(\bar{u} | \bar{v})(y)\Leftrightarrow (w,-y) \in N_{\mathrm{gph}\,F}(\bar{u},\bar{v}). \end{aligned}$$

The coderivative generalizes the adjoint of the derivative of smooth vector function: for smooth \(c : \mathfrak {R}^n \rightarrow \mathfrak {R}^m\), the set-valued mapping \(x \mapsto F(x) := \{c(x)\}\) has coderivative given by \(D^*F(x|c(x))(y) = \{\nabla c(x)^* y\}\) for all \(x \in \mathfrak {R}^n\) and \(y\in \mathfrak {R}^m\). As we see next, coderivative calculations drive two of the arguments in Sect. 4.1.

Proof of Corollary 4.3

Corresponding to any linear map \(A :\mathfrak {R}^p \rightarrow \mathfrak {R}^q\), define a set-valued mapping \(F_A :\mathfrak {R}^p \rightrightarrows \mathfrak {R}^q\) by \(F_A(u) = Au-S\). A coderivative calculation shows, for vectors \(v \in \mathfrak {R}^p\),

$$\begin{aligned} D^* F_A(0|0)(v) =\left\{ \begin{array}{ll} \{A^*v\} &{} \quad \big (v \in N_S(0)\big ) \\ \emptyset &{} \quad (\text{ otherwise }). \end{array}\right. \end{aligned}$$

Hence, by assumption, the only vector \(v \in \mathfrak {R}^p\) satisfying \(0 \in D^* F_{\bar{A}}(0|0)(v)\) is zero, so by [40, Thm 9.43], the mapping \(F_{\bar{A}}\) is metrically regular at zero for zero. Applying Theorem 4.2 shows that there exist constants \(\delta ,\gamma > 0\) such that, if \(\Vert A-\bar{A}\Vert < \delta \) and \(|v| < \delta \), then we have

$$\begin{aligned} \mathrm{dist}\,\!\big (0,F_A^{-1}(-v)\big ) \le \gamma \, \mathrm{dist}\,\!\big (-v,F_A(0)\big ), \end{aligned}$$

or equivalently,

$$\begin{aligned} \mathrm{dist}\,\!\big (0,A^{-1}(S-v)\big ) \le \gamma \, \mathrm{dist}\,(v,S). \end{aligned}$$

Since \(0 \in S\), the right-hand side is bounded above by \(\gamma |v|\), so the result follows. \(\square \)

Proof of Theorem 4.4

We simply need to check that the set-valued mapping \(G :\mathfrak {R}^p \!\rightrightarrows \mathfrak {R}^q\) defined by \(G(z) = F(z) - S\) is metrically regular at \(\bar{z}\) for zero. Much the same coderivative calculation as in the proof of Corollary 4.3 shows, for vectors \(v \in \mathfrak {R}^p\), the formula

$$\begin{aligned} D^* G(\bar{z}|0)(v) =\left\{ \begin{array}{ll} \{\nabla F(\bar{z})^*v\} &{} \big (v \in N_S(\bar{z})\big ) \\ \emptyset &{} (\text{ otherwise }). \end{array}\right. \end{aligned}$$

Hence, by assumption, the only vector \(v \in \mathfrak {R}^p\) satisfying \(0 \in D^* G(\bar{z}|0)(v)\) is zero, so metric regularity follows by [40, Thm 9.43]. \(\square \)

Alternative proof of Theorem 4.2

In the text we gave a short ad hoc proof of Theorem 4.2. Here we present a more formal approach. Denote the space of linear maps from \(\mathfrak {R}^p\) to \(\mathfrak {R}^q\) by \(L(\mathfrak {R}^p,\mathfrak {R}^q)\), and define a mapping \(g :L(\mathfrak {R}^p,\mathfrak {R}^q) \times \mathfrak {R}^p \rightarrow \mathfrak {R}^q\) and a parametric mapping \(g_H :\mathfrak {R}^p \rightarrow \mathfrak {R}^q\) by \(g(H,u)= g_H(u) = Hu\) for maps \(H \in L(\mathfrak {R}^p,\mathfrak {R}^q)\) and points \(u \in \mathfrak {R}^p\). Using the notation of [14, Section 3], the Lipschitz constant \(l[g](0;\bar{u},0)\), is by definition the infimum of the constants \(\rho \) for which the inequality

$$\begin{aligned} d\big (w,g_H(u)\big ) \le \rho d\big (u,g_H^{-1}(w)\big ) \end{aligned}$$
(6.6)

holds for all triples (uwH) sufficiently near the triple \((\bar{u}, 0, 0)\). Inequality (6.6) says simply

$$\begin{aligned} |w-Hu| \le \rho |u-z|\quad \text{ for } \text{ all }\,z \in \mathfrak {R}^p\,\hbox {satisfying}\, Hz=w, \end{aligned}$$

a property that holds providing \(\rho \ge \Vert H\Vert \). We deduce

$$\begin{aligned} l[g](0;\bar{u},0) = 0. \end{aligned}$$
(6.7)

We can also consider \(F+g\) as a set-valued mapping from \(L(\mathfrak {R}^p,\mathfrak {R}^q) \times \mathfrak {R}^p\) to \(\mathfrak {R}^q\), defined by \((F+g)(H,u) = F(u) + Hu\), and then the parametric mapping \((F+g)_H :\mathfrak {R}^p \rightrightarrows \mathfrak {R}^q\) is defined in the obvious way: in other words, \((F+g)_H(u) = F(u) + Hu\). According to [14, Theorem 2], Equation (6.7) implies the following relationship between the “covering rates” for F and \(F+g\):

$$\begin{aligned} r[F+g](0;\bar{u},\bar{v}) = r[F](\bar{u}, \bar{v}). \end{aligned}$$

The reciprocal of the right-hand side is, by definition, the infimum of the constants \(\kappa > 0\) such that inequality (4.1) holds for all pairs (uv) sufficiently near the pair \((\bar{u}, \bar{v})\). By metric regularity, this number is strictly positive. On the other hand, the reciprocal of the left-hand side is, by definition, the infimum of the constants \(\gamma > 0\) such that inequality (4.2) holds for all triples (uvH) sufficiently near the pair \((\bar{u}, \bar{v},0)\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lewis, A.S., Wright, S.J. A proximal method for composite minimization. Math. Program. 158, 501–546 (2016). https://doi.org/10.1007/s10107-015-0943-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10107-015-0943-9

Keywords

Mathematics Subject Classfication

Navigation