Skip to main content
Log in

Forward–backward quasi-Newton methods for nonsmooth optimization problems

  • Published:
Computational Optimization and Applications Aims and scope Submit manuscript

Abstract

The forward–backward splitting method (FBS) for minimizing a nonsmooth composite function can be interpreted as a (variable-metric) gradient method over a continuously differentiable function which we call forward–backward envelope (FBE). This allows to extend algorithms for smooth unconstrained optimization and apply them to nonsmooth (possibly constrained) problems. Since the FBE can be computed by simply evaluating forward–backward steps, the resulting methods rely on a similar black-box oracle as FBS. We propose an algorithmic scheme that enjoys the same global convergence properties of FBS when the problem is convex, or when the objective function possesses the Kurdyka–Łojasiewicz property at its critical points. Moreover, when using quasi-Newton directions the proposed method achieves superlinear convergence provided that usual second-order sufficiency conditions on the FBE hold at the limit point of the generated sequence. Such conditions translate into milder requirements on the original function involving generalized second-order differentiability. We show that BFGS fits our framework and that the limited-memory variant L-BFGS is well suited for large-scale problems, greatly outperforming FBS or its accelerated version in practice, as well as ADMM and other problem-specific solvers. The analysis of superlinear convergence is based on an extension of the Dennis and Moré theorem for the proposed algorithmic scheme.

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. http://github.com/kul-forbes/ForBES.

  2. http://wwwopt.mathematik.tu-darmstadt.de/spear/.

  3. http://www.csie.ntu.edu.tw/~cjlin/libsvmtools/datasets/.

  4. http://sun.stanford.edu/~rmunk/PROPACK/.

  5. http://grouplens.org/datasets/movielens/.

References

  1. Moreau, J.-J.: Proximité et dualité dans un espace Hilbertien. Bulletin de la Société mathématique de France 93, 273–299 (1965)

    Article  MathSciNet  MATH  Google Scholar 

  2. Lions, P.-L., Mercier, B.: Splitting algorithms for the sum of two nonlinear operators. SIAM J. Numer. Anal. 16(6), 964–979 (1979)

    Article  MathSciNet  MATH  Google Scholar 

  3. Combettes, P.L., Pesquet, J.-C.: Proximal splitting methods in signal processing. In: Fixed-Point Algorithms for Inverse Problems in Science and Engineering, pp. 185–212. Springer (2011)

  4. Łojasiewicz, S.: Une propriété topologique des sous-ensembles analytiques réels. Les équations aux dérivées partielles, pp. 87–89 (1963)

  5. Łojasiewicz, S.: Sur la géométrie semi- et sous-analytique. Annales de l’institut Fourier 43(5), 1575–1595 (1993)

    Article  MATH  Google Scholar 

  6. Kurdyka, K.: On gradients of functions definable in o-minimal structures. Annales de l’institut Fourier 48(3), 769–783 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  7. Attouch, H., Bolte, J., Svaiter, B.F.: Convergence of descent methods for semi-algebraic and tame problems: proximal algorithms, forward-backward splitting, and regularized Gauss–Seidel methods. Math. Program. 137(1–2), 91–129 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  8. Attouch, H., Bolte, J.: On the convergence of the proximal algorithm for nonsmooth functions involving analytic features. Math. Program. 116(1), 5–16 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  9. Attouch, H., Bolte, J., Redont, P., Soubeyran, A.: Proximal alternating minimization and projection methods for nonconvex problems: an approach based on the Kurdyka–Łojasiewicz inequality. Math. Oper. Res. 35(2), 438–457 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  10. Bolte, J., Sabach, S., Teboulle, M.: Proximal alternating linearized minimization for nonconvex and nonsmooth problems. Math. Program. 146(1), 459–494 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  11. Ochs, P., Chen, Y., Brox, T., Pock, T.: iPiano: inertial proximal algorithm for nonconvex optimization. SIAM J. Imaging Sci. 7(2), 1388–1419 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  12. Nesterov, Y.: A method for solving the convex programming problem with convergence rate \(O(1/k^2)\). Doklady Akademii Nauk SSSR 269(3), 543–547 (1983)

    MathSciNet  Google Scholar 

  13. Tseng, P.: On accelerated proximal gradient methods for convex-concave optimization. Department of Mathematics, University of Washington, Tech. Rep. (2008)

  14. Beck, A., Teboulle, M.: A fast iterative shrinkage-thresholding algorithm for linear inverse problems. SIAM J. Imaging Sci. 2(1), 183–202 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  15. Nesterov, Y.: Gradient methods for minimizing composite functions. Math. Program. 140(1), 125–161 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  16. Becker, S., Fadili, J.: A quasi-Newton proximal splitting method. In: Advances in Neural Information Processing Systems, pp. 2618–2626 (2012)

  17. Lee, J., Sun, Y., Saunders, M.: Proximal Newton-type methods for convex optimization. In: Advances in Neural Information Processing Systems, pp. 836–844 (2012)

  18. Scheinberg, K., Tang, X.: Practical inexact proximal quasi-Newton method with global complexity analysis. Math. Program. 160(1), 495–529 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  19. Patrinos, P., Bemporad, A.: Proximal Newton methods for convex composite optimization. In: IEEE Conference on Decision and Control, pp. 2358–2363 (2013)

  20. Fukushima, M.: Equivalent differentiable optimization problems and descent methods for asymmetric variational inequality problems. Math. Program. 53(1), 99–110 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  21. Yamashita, N., Taji, K., Fukushima, M.: Unconstrained optimization reformulations of variational inequality problems. J. Optim. Theory Appl. 92(3), 439–456 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  22. Facchinei, F., Pang, J.-S.: Finite-Dimensional Variational Inequalities and Complementarity Problems, vol. II. Springer, Berlin (2003)

    MATH  Google Scholar 

  23. Li, W., Peng, J.: Exact penalty functions for constrained minimization problems via regularized gap function for variational inequalities. J. Glob. Optim. 37, 85–94 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  24. Patrinos, P., Sopasakis, P., Sarimveis, H.: A global piecewise smooth Newton method for fast large-scale model predictive control. Automatica 47, 2016–2022 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  25. Liu, T., Pong, T.K.: Further properties of the forward–backward envelope with applications to difference-of-convex programming. Computational Optimization and Applications, pp. 1–32, 2017. doi:10.1007/s10589-017-9900-2

  26. Dennis, J.E., Moré, J.J.: A characterization of superlinear convergence and its application to quasi-Newton methods. Math. Comput. 28(126), 549–560 (1974)

    Article  MathSciNet  MATH  Google Scholar 

  27. Dai, Y.-H.: Convergence properties of the BFGS algorithm. SIAM J. Optim. 13(3), 693–701 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  28. Mascarenhas, W.F.: The BFGS method with exact line searches fails for non-convex objective functions. Math. Program. 99(1), 49–61 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  29. Mascarenhas, W.F.: On the divergence of line search methods. Comput. Appl. Math. 26, 129–169 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  30. Dai, Y.H.: A perfect example for the BFGS method. Math. Program. 138, 501–530 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  31. Rockafellar, R.T., Wets, R.J.-B.: Variational Analysis, vol. 317. Springer, Berlin (2011)

    MATH  Google Scholar 

  32. Bauschke, H.H., Combettes, P.L.: Convex Analysis and Monotone Operator Theory in Hilbert Spaces. Springer, Berlin (2011)

    Book  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  34. Dennis, J.E., Schnabel, R.B.: Numerical Methods for Unconstrained Optimization and Nonlinear Equations. Society for Industrial and Applied Mathematics (1996).

  35. Bertsekas, D.: Nonlinear Programming. Athena Scientific, Belmont (1999)

    MATH  Google Scholar 

  36. Lemaréchal, C., Sagastizábal, C.: Practical aspects of the Moreau–Yosida regularization: theoretical preliminaries. SIAM J. Optim. 7(2), 367–385 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  37. Bernstein, D.S.: Matrix Mathematics: Theory, Facts, and Formulas with Application to Linear Systems Theory. Princeton University Press, Woodstock (2009)

    Book  Google Scholar 

  38. Rockafellar, R.T.: First- and second-order epi-differentiability in nonlinear programming. Trans. Am. Math. Soc. 307, 75–108 (1988)

    Article  MathSciNet  MATH  Google Scholar 

  39. Rockafellar, R.: Second-order optimality conditions in nonlinear programming obtained by way of epi-derivatives. Math. Oper. Res. 14(3), 462–484 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  40. Poliquin, R.A., Rockafellar, R.T.: Amenable functions in optimization. In: Giannessi, F. (ed.) Nonsmooth Optimization: Methods and Applications, pp. 338–353. Gordon and Breach (1992).

  41. Poliquin, R.A., Rockafellar, R.T.: Second-order nonsmooth analysis in nonlinear programming. In: Du, D., Qi, L., Womersley, R. (eds.) Recent Advances in Nonsmooth Optimization, pp. 322–350. World Scientific Publishers (1995)

  42. Rockafellar, R.T.: Monotone operators and the proximal point algorithm. SIAM J. Control Optim. 14(5), 877–898 (1976)

    Article  MathSciNet  MATH  Google Scholar 

  43. Fukushima, M., Qi, L.: A globally and superlinearly convergent algorithm for nonsmooth convex minimization. SIAM J. Optim. 6(4), 1106–1120 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  44. Bonnans, J.F., Gilbert, J.C., Lemaréchal, C., Sagastizábal, C.A.: A family of variable metric proximal methods. Math. Program. 68(1), 15–47 (1995)

    MathSciNet  MATH  Google Scholar 

  45. Mifflin, R., Sun, D., Qi, L.: Quasi-Newton bundle-type methods for nondifferentiable convex optimization. SIAM J. Optim. 8(2), 583–603 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  46. Chen, X., Fukushima, M.: Proximal quasi-Newton methods for nondifferentiable convex optimization. Math. Program. 85(2), 313–334 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  47. Burke, J.V., Qian, M.: On the superlinear convergence of the variable metric proximal point algorithm using Broyden and BFGS matrix secant updating. Math. Program. 88(1), 157–181 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  48. Sagara, N., Fukushima, M.: A trust region method for nonsmooth convex optimization. J. Ind. Manage. Optim. 1(2), 171–180 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  49. Squire, W., Trapp, G.: Using complex variables to estimate derivatives of real functions. SIAM Rev. 40(1), 110–112 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  50. Nocedal, J., Wright, S.: Numerical Optimization. Springer, Berlin (2006)

    MATH  Google Scholar 

  51. Noll, D., Rondepierre, A.: Convergence of linesearch and trust-region methods using the Kurdyka–Łojasiewicz inequality. In: Bailey, D.H., Bauschke, H.H., Borwein, P., Garvan, F., Théra, M., Vanderwerff, J.D., Wolkowicz, H. (eds.) Computational and Analytical Mathematics: In Honor of Jonathan Borwein’s 60th birthday, pp. 593–611. Springer, New York (2013)

  52. Bolte, J., Daniilidis, A., Lewis, A.: The Łojasiewicz inequality for nonsmooth subanalytic functions with applications to subgradient dynamical systems. SIAM J. Optim. 17(4), 1205–1223 (2007)

    Article  MATH  Google Scholar 

  53. Dennis, J.E., Moré, J.J.: Quasi-Newton methods, motivation and theory. SIAM Rev. 19(1), 46–89 (1977)

    Article  MathSciNet  MATH  Google Scholar 

  54. Powell, M.J.D.: Some global convergence properties of a variable metric algorithm for minimization without exact line searches. In: Cottle, R.W., Lemke, C.E. (eds.) Nonlinear Programming. SIAM-AMS Proceedings 9, pp. 53–72. American Mathematical Society (1976)

  55. Liu, D.C., Nocedal, J.: On the limited memory BFGS method for large scale optimization. Math. Program. 45(1), 503–528 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  56. Ip, C.-M., Kyparisis, J.: Local convergence of quasi-Newton methods for B-differentiable equations. Math. Program. 56(1–3), 71–89 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  57. Nocedal, J.: Updating quasi-Newton matrices with limited storage. Math. Comput. 35(151), 773–782 (1980)

    Article  MathSciNet  MATH  Google Scholar 

  58. Li, D.-H., Fukushima, M.: On the global convergence of the BFGS method for nonconvex unconstrained optimization problems. SIAM J. Optim. 11(4), 1054–1064 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  59. Hiriart-Urruty, J.-B., Lemaréchal, C.: Convex Analysis and Minimization Algorithms I: Fundamentals, vol. 305. Springer, Berlin (1996)

    MATH  Google Scholar 

  60. Fletcher, R.: Practical Methods of Optimization. Wiley, Hoboken (1987)

    MATH  Google Scholar 

  61. Dai, Y.-H., Yuan, Y.: A nonlinear conjugate gradient method with a strong global convergence property. SIAM J. Optim. 10(1), 177–182 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  62. Wright, S.J., Nowak, R.D., Figueiredo, M.A.: Sparse reconstruction by separable approximation. IEEE Trans. Signal Process. 57(7), 2479–2493 (2009)

    Article  MathSciNet  Google Scholar 

  63. Wen, Z., Yin, W., Goldfarb, D., Zhang, Y.: A fast algorithm for sparse reconstruction based on shrinkage, subspace optimization, and continuation. SIAM J. Sci. Comput. 32(4), 1832–1857 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  64. Candès, E.J., Recht, B.: Exact matrix completion via convex optimization. Found. Comput. Math. 9(6), 717–772 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  65. Toh, K.-C., Yun, S.: An accelerated proximal gradient algorithm for nuclear norm regularized linear least squares problems. Pac. J. Optim. 6(3), 615–640 (2010)

    MathSciNet  MATH  Google Scholar 

  66. Boţ, R.I., Csetnek, E.R., László, S.C.: An inertial forward-backward algorithm for the minimization of the sum of two nonconvex functions. EURO J. Comput. Optim. 4(1), 3–25 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  67. Pang, J.-S.: Newton’s method for B-differentiable equations. Math. Oper. Res. 15(2), 311–341 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  68. Poliquin, R., Rockafellar, R.: Generalized Hessian properties of regularized nonsmooth functions. SIAM J. Optim. 6(4), 1121–1137 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  69. Horn, R.A., Johnson, C.R.: Matrix Analysis. Cambridge University Press, Cambridge (2012)

    Book  Google Scholar 

  70. Byrd, R.H., Nocedal, J.: A tool for the analysis of quasi-Newton methods with application to unconstrained minimization. SIAM J. Numer. Anal. 26(3), 727–739 (1989)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Lorenzo Stella.

Additional information

This work was supported by the KU Leuven Research Council under BOF/STG-15-043.

Appendices

Appendix 1: Definitions and known results

We say that \(G:\mathbb {R}^n\rightarrow \mathbb {R}^m\) is strictly continuous at \(\bar{x}\) if [31, Def. 9.1(b)]

$$\begin{aligned} \limsup _{ \begin{array}{c} (x,y)\rightarrow (\bar{x},\bar{x})\\ x\ne y \end{array} }{ \frac{\Vert G(y)-G(x)\Vert }{\Vert y-x\Vert } } {}<{} \infty . \end{aligned}$$

If G is differentiable, we let denote the Jacobian of G. When \(m=1\) we indicate with the gradient of G and with its Hessian, whenever it makes sense. We say that G is strictly differentiable at \(\bar{x}\) if it satisfies the stronger limit [31, Eq. 9(7)]

The next result states that strict differentiability is preserved by composition; its proof is a trivial computation and is therefore omitted.

Proposition 6.1

Let \(F:\mathbb {R}^n\rightarrow \mathbb {R}^m\), \(P:\mathbb {R}^m\rightarrow \mathbb {R}^k\). Suppose that F and P are (strictly) differentiable at \(\bar{x}\) and \(F(\bar{x})\), respectively. Then the composition \(T=P\circ F\) is (strictly) differentiable at \(\bar{x}\).

Similarly, the product of (strictly) differentiable functions is still (strictly) differentiable. However, if one of the two functions vanishes at one point, then we may relax some assumptions, as it is proved in the next result.

Proposition 6.2

Let \(Q:\mathbb {R}^n\rightarrow \mathbb {R}^{m\times k}\) and \(R:\mathbb {R}^n\rightarrow \mathbb {R}^k\), and suppose that \(R(\bar{x}) = 0\). If Q is (strictly) continuous at \(\bar{x}\) and R is (strictly) differentiable at \(\bar{x}\), then their product \(G:\mathbb {R}^n \rightarrow \mathbb {R}^m\) defined as \(G(x) = Q(x)R(x)\) is (strictly) differentiable at \(\bar{x}\) with .

Proof

Suppose first that Q is continuous at \(\bar{x}\) and R is differentiable at \(\bar{x}\). Then, expanding R(x) at \(\bar{x}\) and since \(G(\bar{x})=0\), we obtain

figure bn

The quantity is bounded, and continuity of Q at \(\bar{x}\) implies that taking the limit for \(\bar{x}\ne x\rightarrow \bar{x}\) yields 0. This proves that G is differentiable at \(\bar{x}\).

Suppose now that Q is strictly continuous at \(\bar{x}\), and that R is strictly differentiable at \(\bar{x}\). Then, expanding R(y) at x we obtain

figure bo

The quantity is bounded, and by strict continuity of Q at \(\bar{x}\) so is \( \frac{Q(x)-Q(y)}{\Vert x-y\Vert } \) for xy sufficiently close to \(\bar{x}\). Taking the limit for \((x,y)\rightarrow (\bar{x},\bar{x})\) with \(x\ne y\) in the above expression then yields 0, proving strict differentiability. Uniqueness of the Jacobian proves also the claimed form of . \(\square \)

Definition 6.3

A mapping \(G:\mathbb {R}^n\rightarrow \mathbb {R}^m\) is said to be semidifferentiable (or B -differentiable [56, 67]) at a point \(\bar{x}\in \mathbb {R}^n\) if there exists a positively homogeneous mapping such that

It is strictly semidifferentiable at \(\bar{x}\) if the stronger limit holds

is called semiderivative of G at \(\bar{x}\). If G is (strictly) semidifferentiable at every point of a set S, then it is said to be (strictly) semidifferentiable in S.

Proposition 6.4

([67, Thm. 2]) Suppose that \(G:\mathbb {R}^n\rightarrow \mathbb {R}^m\) is semidifferentiable in a neighborhood of \(\bar{x}\in \mathbb {R}^n\). Then, the following are equivalent:

  1. (a)

    is continuous in its first argument at \(\bar{x}\) for all \(d\in \mathbb {R}^n\);

  2. (b)

    G is strictly semidifferentiable at \(\bar{x}\);

  3. (c)

    G is strictly (Fréchet) differentiable at \(\bar{x}\).

Proposition 6.5

Suppose that \(G:\mathbb {R}^n\rightarrow \mathbb {R}^m\) is semidifferentiable in a neighborhood N of \(\bar{x}\) and that is calm at \(\bar{x}\), i.e., there exists \(L>0\) such that, for all \(x\in N\) and \(d\in \mathbb {R}^n\) with \(\Vert d\Vert =1\),

Then,

Proof

Follows from [56, Lem. 2.2] by observing that the assumption of Lipschitz-continuity may be relaxed to calmness. \(\square \)

Appendix 2: Proofs of Sect. 2

Proof of Lemma 2.9

We know from [68, Thm.s 3.8, 4.1] that \(\mathop {\mathrm{prox}}\nolimits _{\gamma g}\) is (strictly) differentiable at \(x-\gamma \nabla f(x)\) if and only if g satisfies Assumption 4 (Assumption 5) at x for \(-\nabla f(x)\). Since \(f\in C^2\) by assumption, then in particular \(\nabla f\) is strictly differentiable. The formula (2.7) follows from Proposition 6.1 with \(P = \mathop {\mathrm{prox}}\nolimits _{\gamma g}\) and \(F(x) = x - \gamma \nabla f(x)\).

Matrix \(Q_{\gamma }(x)\) is symmetric since \(f\in C^2\) and positive definite if \(\gamma < 1/L_f\). To obtain an expression for we can apply [31, Ex. 13.45] to the tilted function \(g+\langle \nabla f(x),{}\cdot {} \rangle \) so that, letting and the idempotent and symmetric projection matrix on S,

figure bp

where \(^\dagger \) indicates the pseudo-inverse, and last equality is due to [37, Facts 6.4.12(i)-(ii) and 6.1.6(xxxii)] and the properties of M as stated in Assumption 4. Apparently \(P_{\gamma }(x)\succeq 0\) is symmetric and \(\Vert P_{\gamma }(x)\Vert \le 1\). \(\square \)

Proof of Theorem 2.11

If follows from Theorem 2.10 that the Hessian \(\nabla ^2\varphi _{\gamma }(x)\) exists and is symmetric. Moreover, from [31, Ex. 13.18] we know that for all \(d\in \mathbb {R}^n\)

(6.1)

2.11(a) \(\Leftrightarrow \) 2.11(b): Follows directly from (6.1), using [31, Thm. 13.24(c)].

2.11(c) \(\Leftrightarrow \) 2.11(d): Letting \(Q = Q_{\gamma }(x)\), we see from (2.7) and (2.9) that is similar to the symmetric matrix \(Q^{-1/2}\nabla ^2\varphi _{\gamma }(x)Q^{-1/2}\), which is positive definite if and only if \(\nabla ^2\varphi _{\gamma }(x)\) is.

2.11(b) \(\Leftrightarrow \) 2.11(c): From the point above we know that has all real eigenvalues, and it can be easily seen to be similar to \(\gamma ^{-1}(I-QP)\), where \(P = P_{\gamma }(x)\). From [69, Thm. 7.7.3] it follows that \(\lambda _{\min }(I-QP) > 0\) if and only if \(Q^{-1} \succ P\). For all \(d\in S\), using (2.8) we have

$$\begin{aligned} \langle d,(Q^{-1}-P)d \rangle {}={}&\langle d,Q^{-1}d \rangle {}-{} \langle d, \Pi _S [I+\gamma M]^{-1} \Pi _Sd \rangle \\ {}={}&\langle d,Q^{-1}d \rangle {}-{} \langle \Pi _Sd, [I+\gamma M]^{-1} \Pi _Sd \rangle \\ {}={}&\langle d,Q^{-1}d \rangle {}-{} \langle d, [I+\gamma M]^{-1} d \rangle \end{aligned}$$

and last quantity is positive if and only if \(I+\gamma M\succ Q\) on S. By definition of Q, we then have that this holds if and only if \( \nabla ^2 f(x)+M\succ 0 \) on S, which is 2.11(b).

2.11(d) \(\Leftrightarrow \) 2.11(e): Trivial since \(\nabla ^2\varphi _{\gamma }(x)\) exists. \(\square \)

Appendix 3: Proofs of Sect. 3

The following results are instrumental in proving convergence of the iterates of

figure bq

.

Lemma 6.6

Under Assumption 1, consider the sequences and generated by

figure br

. If there exist \(\bar{\tau },c>0\) such that \(\tau _k\le \bar{\tau }\) and \(\Vert d^k\Vert \le c\Vert R_{\gamma _k}(x^k)\Vert \), then

$$\begin{aligned} \Vert x^{k+1}-x^k\Vert {}\le {}&\gamma _k\Vert R_{\gamma _k}(w^k)\Vert {}+{} \bar{\tau }c\Vert R_{\gamma _k}(x^k)\Vert \quad \forall k\in \mathbb {N}\end{aligned}$$
(6.2)

and, for k large enough,

$$\begin{aligned} \Vert x^{k+1}-x^k\Vert {}\le {}&\gamma _k\Vert R_{\gamma _k}(w^k)\Vert {}+{} \bar{\tau }c(1+\gamma _k L_f)\Vert R_{\gamma _{k-1}}(w^{k-1})\Vert . \end{aligned}$$
(6.3)

Proof

Equation (6.2) follows simply by

$$\begin{aligned} \Vert x^{k+1}-x^k\Vert = \Vert x^{k+1}-w^k+\tau _kd^k\Vert \le \gamma _k\Vert R_{\gamma _k}(w^k)\Vert +\bar{\tau }c\Vert R_{\gamma _k}(x^k)\Vert . \end{aligned}$$

Now, for k sufficiently large \(\gamma _k = \gamma _{k-1} = \gamma _\infty > 0\), see Lemma 3.1, and

$$\begin{aligned} \Vert R_{\gamma _{k}}(x^{k})\Vert&=\gamma _k^{-1}\Vert x^{k}-T_{\gamma _k}(x^{k})\Vert \\&= \gamma _k^{-1}\Vert T_{\gamma _k}(w^{k-1})-T_{\gamma _k}(x^{k})\Vert \\&\le \gamma _k^{-1}\Vert w^{k-1}-\gamma _k\nabla f(w^{k-1})-x^{k}+\gamma _k\nabla f(x^{k})\Vert \\&\le \gamma _k^{-1}\Vert w^{k-1}-x^{k}\Vert +\Vert \nabla f(w^{k-1})-\nabla f(x^{k})\Vert \\&\le (1+\gamma _k L_f)\Vert R_{\gamma _{k-1}}(w^{k-1})\Vert , \end{aligned}$$

where the first inequality follows from nonexpansiveness of \(\mathop {\mathrm{prox}}\nolimits _{\gamma g}\), and the last one from Lipschitz continuity of \(\nabla f\). Putting this together with (6.2) gives (6.3). \(\square \)

Lemma 6.7

Let and be real sequences satisfying \(\beta _k\ge 0\), \(\delta _k\ge 0\), \(\delta _{k+1}\le \delta _k\) and \(\beta _{k+1}^2\le (\delta _k-\delta _{k+1})\beta _k\) for all \(k\in \mathbb {N}\). Then \(\sum _{k=0}^\infty \beta _k<\infty \).

Proof

Taking the square root of both sides in \(\beta _{i+1}^2\le (\delta _i-\delta _{i+1})\beta _i\) and using

$$\begin{aligned} \sqrt{\zeta \eta }\le (\zeta +\eta )/2, \end{aligned}$$

for any nonnegative numbers \(\zeta \), \(\eta \), we arrive at \(2\beta _{i+1}\le (\delta _i-\delta _{i+1})+\beta _i\). Summing up the latter for \(i=0,\ldots ,k\), for any \(k\in \mathbb {N}\),

$$\begin{aligned} 2\textstyle {\sum _{i=0}^k}\beta _{i+1}&\le \textstyle {\sum _{i=0}^k}(\delta _i-\delta _{i+1})+\textstyle {\sum _{i=0}^k}\beta _{i}\\&=\delta _0-\delta _{k+1}+\beta _0-\beta _{k+1}+\textstyle {\sum _{i=0}^k}\beta _{i+1}\\&\le \delta _0+\beta _0+\textstyle {\sum _{i=0}^k}\beta _{i+1}. \end{aligned}$$

Hence

$$\begin{aligned} \sum _{i=0}^\infty \beta _{i+1}\le \delta _0+\beta _0<\infty , \end{aligned}$$
(6.4)

which concludes the proof. \(\square \)

Proposition 6.8

Suppose Assumption 1 is satisfied and that \(\varphi \) is lower bounded, and consider the sequences generated by

figure bs

. If \(\beta \in (0,1)\) and there exist \(\bar{\tau },c>0\) such that \(\tau _k\le \bar{\tau }\) and \(\Vert d^k\Vert \le c\Vert R_{\gamma _k}(x^k)\Vert \) then

$$\begin{aligned} \sum _{k=0}^\infty \Vert x^{k+1}-x^k\Vert ^2<\infty . \end{aligned}$$
(6.5)

If moreover is bounded, then

$$\begin{aligned} \lim _{k\rightarrow \infty }\mathop {\mathrm{dist}}\nolimits _{\omega (x^0)}(x^k)=0 \end{aligned}$$
(6.6)

and \(\omega (x^0)\) is a nonempty, compact and connected subset of \(\mathop {\mathrm{zer}}\nolimits \partial \varphi \) over which \(\varphi \) is constant.

Proof

(6.5) follows from (6.2), Proposition 3.4(ii) and 3.4(iv), and the fact that the sum of square-summable sequences is square summable.

If is bounded, that \(\omega (x^0)\) is nonempty, compact and connected and \(\lim _{k\rightarrow \infty }\mathop {\mathrm{dist}}\nolimits _{\omega (x^0)}(x^k)=0\) follow by [10, Lem.s 5(ii),(iii), Rem. 5]. That \(\varphi \) is constant on \(\omega (x^0)\) follows by a similar argument as in [10, Lem. 5(iv)]. \(\square \)

The following is [10, Lem. 6], therefore we state it with no proof.

Lemma 6.9

(Uniformized KL property) Let \(K\subset \mathbb {R}^n\) be a compact set and suppose that the proper lower semi-continuous function \(\varphi :\mathbb {R}^n\rightarrow \overline{{\mathbb {R}}}\) is constant on K and satisfies the KL property at every \({x}^\star \in K\). Then there exist \(\varepsilon >0\), \(\eta >0\), and a continuous concave function \(\psi :[0,\eta ]\rightarrow [0,+\infty )\) such that properties 3.9(i), 3.9(ii) and 3.9(iii) hold, and

  1. (iv’)

    for all \(x_\star \in K\) and x such that \(\mathop {\mathrm{dist}}\nolimits _K(x)<\varepsilon \) and \(\varphi (x_\star )< \varphi (x)<\varphi (x_\star )+\eta \),

    $$\begin{aligned} \psi '(\varphi (x)-\varphi (x_\star ))\mathop {\mathrm{dist}}\nolimits (0,\partial \varphi (x))\ge 1. \end{aligned}$$
    (6.7)

Proof of Lemma 3.1

Let be the sequence of stepsize parameters computed by

figure bt

. To arrive to a contradiction, suppose that \(k_0\) is the smallest element of \(\mathbb {N}\) such that

Clearly, \(k_0 \ge 1\). Moreover \(\sigma ^{-1}\gamma _{k_0}\) must satisfy the condition in step 4: for some \(w\in \mathbb {R}^n\) (corresponding to \(w^k=x^k+\tau _k d^k\) selected before going back to step 1 after the condition in step 4 is passed, which might differ from the final value of \(w^k\) after step 4 is passed)

$$\begin{aligned} \varphi (T_{\sigma ^{-1}\gamma _{k_0}}(w)) {}>{} \varphi _{\sigma ^{-1}\gamma _{k_0}}(w) {}-{} \frac{\beta \sigma ^{-1}\gamma _{k_0}}{2}\Vert R_{\sigma ^{-1}\gamma _{k_0}}(w)\Vert ^2. \end{aligned}$$

But from Proposition 2.2(ii) we also have

$$\begin{aligned} \varphi (T_{\sigma ^{-1}\gamma _{k_0}}(w)) {}\le {}&\varphi _{\sigma ^{-1}\gamma _{k_0}}(w) {}-{} \frac{\sigma ^{-1}\gamma _{k_0}}{2}(1-\sigma ^{-1}\gamma _{k_0}L_f)\Vert R_{\sigma ^{-1}\gamma _{k_0}}(w)\Vert ^2\\ {}\le {}&\varphi _{\sigma ^{-1}\gamma _{k_0}}(w) {}-{} \frac{\beta \sigma ^{-1}\gamma _{k_0}}{2}\Vert R_{\sigma ^{-1}\gamma _{k_0}}(w)\Vert ^2, \end{aligned}$$

where last inequality follows from \(\sigma ^{-1}\gamma _{k_0} < (1-\beta )/L_f\). This leads to a contradiction, therefore as claimed. That \(\gamma _k\) is asymptotically constant follows since the sequence is nonincreasing. \(\square \)

Proof of Proposition 3.4

We have

$$\begin{aligned} \varphi (x^{k+1})&\le \varphi _{\gamma _k}(w^k) - \tfrac{\beta \gamma _k}{2}\Vert R_{\gamma _k}(w^k)\Vert ^2 \nonumber \\&\le \varphi _{\gamma _k}(x^k) - \tfrac{\beta \gamma _k}{2}\Vert R_{\gamma _k}(w^k)\Vert ^2 \nonumber \\&\le \varphi (x^k)-\tfrac{\beta \gamma _k}{2}\Vert R_{\gamma _k}(w^k)\Vert ^2-\tfrac{{\gamma _k}}{2}\Vert R_{\gamma _k}(x^k)\Vert ^2, \end{aligned}$$
(6.8)

where the first inequality comes from step 4, the second from step 3 and the third from Proposition 2.2(i). This shows 3.4(i). Let \(\varphi _\star =\lim _{k\rightarrow \infty }\varphi (x^k)\), which exists since is monotone. If \(\varphi _\star =-\infty \), clearly \(\inf \varphi =-\infty \) and \(\omega (x^0)=\emptyset \) due to properness and lower semicontinuity of \(\varphi \) and to the monotonic behavior of . Otherwise, telescoping the inequality we get

$$\begin{aligned} \frac{1}{2} \sum _{i=0}^k{ \gamma _i \left( \beta \Vert R_{\gamma _i}(w^i)\Vert ^2 {}+{} \Vert R_{\gamma _i}(x^i)\Vert ^2 \right) } {}\le {} \varphi (x^0) {}-{} \varphi (x^{k+1}) {}\le {} \varphi (x^0) {}-{} \varphi _\star \end{aligned}$$
(6.9)

and since \(\gamma _k\) is uniformly lower bounded by a positive number (see Lemma 3.1) 3.4(ii) follows, hence 3.4(iii). If \(\beta >0\), observing that for k large enough such that \(\gamma _k\equiv \gamma _\infty \) we have

similar argumentations as those for proving 3.4(ii) show 3.4(iv). \(\square \)

Proof of Theorem 3.5

If \(\inf \varphi =-\infty \) there is nothing to prove. Otherwise, since the sequence is nonincreasing, from (6.9) we get

$$\begin{aligned} \frac{(k+1)\gamma _{k}}{2}\left( \min _{i=0\ldots k}\Vert R_{\gamma _{i}}(x^i)\Vert ^2 + \beta \min _{i=0\ldots k}\Vert R_{\gamma _{i}}(w^i)\Vert ^2\right) \le \varphi (x^0) - \inf \varphi . \end{aligned}$$

Rearranging the terms and invoking Lemma 3.1 gives the result. \(\square \)

Proof of Theorem 3.6

The proof is similar to that of [15, Thm. 4]. By Proposition 2.5(iii) we know that \(\varphi _{\gamma }\le \varphi ^\gamma \) for any \(\gamma >0\). Combining this with (6.8) we get

(6.10)

and in particular, for \(x_\star \in \mathop {\hbox {arg min}}\limits \varphi \),

figure bu

where the last inequality follows by convexity of \(\varphi \). If \(\varphi (x^0)-\inf \varphi \ge R^2/{\gamma _0}\), then the optimal solution of the latter problem for \(k=0\) is \(\alpha =1\) and we obtain (3.1). Otherwise, the optimal solution is

$$\begin{aligned} \alpha =\frac{{\gamma _{k}}(\varphi (x^k)-\inf \varphi )}{R^2}\le \frac{{\gamma _{k}}(\varphi (x^0)-\inf \varphi )}{R^2}\le 1, \end{aligned}$$

and we obtain

$$\begin{aligned} \varphi (x^{k+1})\le \varphi (x^k)-\frac{{\gamma _{k}}(\varphi (x^k)-\inf \varphi )^2}{2R^2}. \end{aligned}$$

Letting \(\lambda _k=\frac{1}{\varphi (x^k)-\inf \varphi }\) the latter inequality is expressed as

$$\begin{aligned} \frac{1}{\lambda _{k+1}}\le \frac{1}{\lambda _k}-\frac{{\gamma _{k}}}{2R^2\lambda _{k+1}^2}. \end{aligned}$$

Multiplying both sides by \(\lambda _{k}\lambda _{k+1}\) and rearranging

$$\begin{aligned} \lambda _{k+1}\ge \lambda _k+\frac{{\gamma _{k}}}{2R^2}\frac{\lambda _{k+1}}{\lambda _k}\ge \lambda _k+\frac{{\gamma _{k}}}{2R^2}, \end{aligned}$$

where the latter inequality follows from the fact that is nonincreasing, cf. Proposition 3.4(i). Telescoping the inequality and using Lemma 3.1, we obtain

Rearranging, we arrive at (3.2). \(\square \)

Proof of Theorem 3.7

If (3.3) holds, then \(\varphi \) has bounded level sets and . In particular, \(\omega (x^0)\ne \emptyset \) and Proposition 3.4(iii) then ensures \(x^k\rightarrow x_\star \). Therefore, there is \(k_0\in \mathbb {N}\) such that \(x^k \in N\) for all \(k\ge k_0\). Inequality (6.10) holds, and in particular for \(k\ge k_0\)

figure bv

where the second inequality follows by convexity of \(\varphi \) and (3.3). The minimum of last expression is achieved for . When \(\gamma _k < 2c^{-1}\) we have the bound

$$\begin{aligned} \varphi (x^{k+1}) - \inf \varphi \le (1-\tfrac{c}{4}\gamma _k)(\varphi (x^k) - \inf \varphi ). \end{aligned}$$

When instead \(\gamma _k\ge 2c^{-1}\) we have the bound

$$\begin{aligned} \varphi (x^{k+1}) - \inf \varphi \le (c\gamma _k)^{-1}(\varphi (x^k) - \inf \varphi ) \le \tfrac{1}{2}(\varphi (x^k) - \inf \varphi ). \end{aligned}$$

Therefore \(\varphi (x^{k+1}) - \inf \varphi \le \omega (\varphi (x^k) - \inf \varphi )\), where

$$\begin{aligned} \omega&\le \sup _k \max \bigl \{\tfrac{1}{2}, 1-\tfrac{c}{4}\gamma _k \bigr \}\\&\le \max \bigl \{\tfrac{1}{2}, 1-\tfrac{c}{4}\min \{\gamma _0,\sigma (1-\beta )/L_f\} \bigr \} \in \bigl [\tfrac{1}{2},1\bigr ), \end{aligned}$$

last inequality following from Lemma 3.1. This proves the claim on the sequence and using inequality (6.8) the same holds for . From the error bound (3.3) we obtain that \(x^k\rightarrow x_\star \) R-linearly. If the same error bound holds for \(\varphi _{\gamma _\infty }\), then also \(w^k\rightarrow x_\star \) R-linearly. \(\square \)

Proof of Theorem 3.10

The case where the sequence is finite does not deserve any further investigation, therefore we assume that is infinite. We then assume that \(R_{\gamma _k}(x^k)\ne 0\) which implies through Proposition 3.4 that \(\varphi (x^{k+1})<\varphi (x^k)\). Due to (6.6), the KL property for \(\varphi \), and Lemma 6.9, there exist \(\varepsilon ,\eta >0\) and a continuous concave function \(\psi :[0,\eta ]\rightarrow [0,+\infty )\) such that for all x with \(\mathop {\mathrm{dist}}\nolimits _{\omega (x^0)}(x)<\varepsilon \) and \(\varphi ({x}^\star )< \varphi (x)<\varphi (x_\star )+\eta \) one has

$$\begin{aligned} \psi '(\varphi (x)-\varphi (x_\star ))\mathop {\mathrm{dist}}\nolimits (0,\partial \varphi (x))\ge 1. \end{aligned}$$

According to Proposition 6.8 there exists a \(k_1\in \mathbb {N}\) such that \(\mathop {\mathrm{dist}}\nolimits _{\omega (x^0)}(x^k)<\varepsilon \) for all \(k\ge k_1\). Furthermore, since \(\varphi (x^k)\) converges to \(\varphi (x_\star )\) there exists a \(k_2\) such that \(\varphi (x^k)<\varphi (x_\star )+\eta \) for all \(k\ge k_2\). Take . Then for every \(k\ge \bar{k}\) we have

$$\begin{aligned} \psi '(\varphi (x^k)-\varphi (x_\star ))\mathop {\mathrm{dist}}\nolimits (0,\partial \varphi (x^k))\ge 1. \end{aligned}$$

From Proposition 3.4(i)

$$\begin{aligned} \varphi (x^{k+1})\le \varphi (x^k)-\tfrac{\beta \gamma _k}{2}\Vert R_{\gamma _k}(w^k)\Vert ^2. \end{aligned}$$

For every \(k>0\) let \( \tilde{\nabla }\varphi (x^k) {}={} \nabla f(x^{k})-\nabla f(w^{k-1})+R_{\gamma _{k-1}}(w^{k-1}) \). Since \( R_{\gamma _{k-1}}(w^{k-1}) {}\in {} \nabla f(w^{k-1}) + \partial g(x^k) \), then \( \tilde{\nabla }\varphi (x^k) {}\in {} \partial \varphi (x^k) \) and

$$\begin{aligned} \Vert \tilde{\nabla }\varphi (x^k)\Vert {}\le {}&\Vert \nabla f(x^{k})-\nabla f(w^{k-1})\Vert {}+{} \Vert R_{\gamma _{k-1}}(w^{k-1})\Vert \\ {}={}&(1+{\gamma _{k-1}} L_f) \Vert R_{\gamma _{k-1}}(w^{k-1})\Vert . \end{aligned}$$

From (6.7)

$$\begin{aligned} \psi '(\varphi (x^{k})-\varphi (x_\star )) {}\ge {} \frac{1}{\Vert \tilde{\nabla }\varphi (x^k)\Vert } {}\ge {} \frac{1}{(1+{\gamma _{k-1}} L_f) \Vert R_{\gamma _{k-1}}(w^{k-1})\Vert }. \end{aligned}$$

Let \(\Delta _k= \psi (\varphi (x^{k})-\varphi (x_\star ))\). By concavity of \(\psi \) and Proposition 3.4(i)

$$\begin{aligned} \Delta _k-\Delta _{k+1}&\ge \psi '(\varphi (x^{k})-\varphi (x_\star ))(\varphi (x^k)-\varphi (x^{k+1}))\\&\ge \frac{\beta \gamma _k}{2 (1+\gamma _{k-1} L_f)}\frac{\Vert R_{\gamma _{k}}(w^k)\Vert ^2}{\Vert R_{\gamma _{k-1}}(w^{k-1})\Vert }\\&\ge \frac{\beta \gamma _{\min }}{2 (1+\gamma _0 L_f)}\frac{\Vert R_{\gamma _{k}}(w^k)\Vert ^2}{\Vert R_{\gamma _{k-1}}(w^{k-1})\Vert } \end{aligned}$$

where , see Lemma 3.1, or

$$\begin{aligned} \Vert R_{\gamma _{k}}(w^{k})\Vert ^2\le \alpha (\Delta _k-\Delta _{k+1})\Vert R_{\gamma _{k-1}}(w^{k-1})\Vert \end{aligned}$$
(6.11)

where \(\alpha = 2 (1+\gamma _0 L_f)/(\beta \gamma _{\min })\). Applying Lemma 6.7 with

$$\begin{aligned} \delta _k=\alpha \Delta _k,\quad \beta _k=\Vert R_{\gamma _{k-1}}(w^{k-1})\Vert , \end{aligned}$$

we conclude that \(\sum _{k=0}^\infty \Vert R_{\gamma _{k}}(w^{k})\Vert <\infty \). From (6.3), using the fact that \(\gamma _k\le \gamma _0\) for all k, then it follows that

$$\begin{aligned} \sum _{k=0}^{\infty }\Vert x^{k+1}-x^k\Vert <\infty . \end{aligned}$$

Then is a Cauchy sequence, hence it converges to a point that, by Proposition 3.4, is a critical point \(x_\star \) of \(\varphi \). \(\square \)

Proof of Theorem 3.11

Theorem 3.10 ensures that converges to a critical point, be it \(x_\star \). We know from Lemma 3.1 that eventually \(\gamma _k = \gamma _\infty > 0\), therefore we assume k is large enough for this purpose and indicate \(\gamma \) in place of \(\gamma _k\) for simplicity. Denoting \(A_k=\sum _{i=k}^\infty \Vert x^{i+1}-x^i\Vert \) clearly \(A_k \ge \Vert x^k-x_\star \Vert \), so we will prove that \(A_k\) converges linearly to zero to obtain the result. Note that by (6.3) we know that

$$\begin{aligned} \Vert x^{i+1}-x^i\Vert \le \gamma \Vert R_{\gamma }(w^i)\Vert + \bar{\tau } c (1+\gamma L_f)\Vert R_{\gamma }(w^{i-1})\Vert . \end{aligned}$$

Therefore we can upper bound \(A_k\) as follows

$$\begin{aligned} A_k {}\le {}&\textstyle \bar{\tau } c (1+\gamma L_f)\Vert R_{\gamma }(w^{k-1})\Vert {}+{} \left( \gamma + \bar{\tau }c(1+\gamma L_f) \right) \sum _{i=k}^\infty { \Vert R_{\gamma }(w^i)\Vert } \nonumber \\ {}\le {}&\textstyle \left( \gamma + \bar{\tau }c(1+\gamma L_f) \right) \sum _{i=k-1}^\infty { \Vert R_{\gamma }(w^i)\Vert , } \end{aligned}$$
(6.12)

and reduce the problem to proving linear convergence of \(B_k = \sum _{i=k}^\infty \Vert R_{\gamma }(w^i)\Vert \). When \(\psi \) is as in (3.4), for sufficiently large k the KL inequality reads

$$\begin{aligned} \varphi (x^k)-\varphi (x_\star ) \le [\sigma (1-\theta )\Vert v^k\Vert ]^{\frac{1}{\theta }},\quad \forall v^k\in \partial \varphi (x^k). \end{aligned}$$

Taking \(v^k = \nabla f(x^k) - \nabla f(w^{k-1}) + R_{\gamma }(w^{k-1}) \in \partial \varphi (x^k)\), this in turn yields

$$\begin{aligned} \varphi (x^k)-\varphi (x_\star ) \le \left[ \sigma (1-\theta )(1+{\gamma } L_f)\Vert R_{\gamma }(w^{k-1})\Vert \right] ^{\frac{1}{\theta }}, \end{aligned}$$
(6.13)

(see the proof of Theorem 3.10). Inequality (6.11) holds, for sufficiently large k, with \(\Delta _k = \sigma (\varphi (x^k)-\varphi (x_\star ))^{1-\theta }\) in this case. Applying Lemma 6.7 with

$$\begin{aligned} \delta _k=\alpha \Delta _{k},\quad \beta _k=\Vert R_{\gamma }(w^{k-1})\Vert =B_{k-1}-B_k, \end{aligned}$$

we obtain

$$\begin{aligned} B_k&\le (B_{k-1} - B_k) + \sigma (\varphi (x^k)-\varphi (x_\star ))^{1-\theta } \\&\le (B_{k-1} - B_k) + \sigma \left[ \sigma (1-\theta )(1+{\gamma } L_f)(B_{k-1} - B_k)\right] ^{\frac{1-\theta }{\theta }}, \end{aligned}$$

where the second inequality is due to (6.13). Since \(B_{k-1}-B_k \rightarrow 0\), then for k large enough it holds that \(\sigma (1+{\gamma } L_f)(B_{k-1}-B_k) \le 1\), and the last term in the previous chain of inequalities is increasing in \(\theta \) when \(\theta \in (0,\tfrac{1}{2}]\). Therefore \(B_k\) eventually satisfies

$$\begin{aligned} B_k \le C(B_{k-1}-B_k), \end{aligned}$$

where \(C>0\), and so \(B_k \le [C/(1+C)] B_{k-1}\), i.e., \(B_k\) converges to zero Q-linearly. This in turn implies that \(\Vert x^k-x_\star \Vert \) converges to zero with R-linear rate. Furthermore,

$$\begin{aligned} \Vert w^k-x_\star \Vert&= \Vert x^k - x_\star + \tau _k d^k\Vert \\&\le \Vert x^k-x_\star \Vert + \bar{\tau }c\Vert R_{\gamma _k}(x^k)\Vert \\&= \Vert x^k-x_\star \Vert + \bar{\tau }c{\gamma _k}^{-1}\Vert T_{\gamma _k}(x^k)-x^k\Vert \\&\le (1+\bar{\tau }c\gamma _k^{-1})\Vert x^k-x_\star \Vert + \bar{\tau }c{\gamma _k}^{-1}\Vert T_{\gamma _k}(x^k)-T_{\gamma _k}(x_\star )\Vert \\&\le (1+\bar{\tau }c\gamma _k^{-1})\Vert x^k-x_\star \Vert + \bar{\tau }c{\gamma _k}^{-1}\Vert x^k - \gamma _k\nabla f(x^k)- x_\star + \gamma _k\nabla f(x_\star )\Vert \\&\le (1+\bar{\tau }c(2\gamma _k^{-1} + L_f))\Vert x^k-x_\star \Vert , \end{aligned}$$

where the last two inequalities follow by nonexpansiveness of \(\mathop {\mathrm{prox}}\nolimits _{\gamma g}\) and Lipschitz continuity of \(\nabla f\). Since \(\gamma _k\) is lower bounded by a positive quantity, then we deduce that also \(\Vert w^k-x_\star \Vert \) converges R-linearly to zero. \(\square \)

Appendix 4: Proofs of Sect. 4

Proof of Theorem 4.1

Since \(w^k = x^k - B_k^{-1}\nabla \varphi _{\gamma }(x^k)\), letting \(k\rightarrow \infty \) and using (4.1) we have that

$$\begin{aligned} 0 {}\leftarrow {} \frac{(B_k - \nabla ^2\varphi _{\gamma }(x_\star ))(w^k-x^k)}{\Vert w^k-x^k\Vert } {}={}&{}-\frac{\nabla \varphi _{\gamma }(x^k) + \nabla ^2\varphi _{\gamma }(x_\star )(w^k-x^k)}{\Vert w^k-x^k\Vert }\\ {}={}&{}-\frac{\nabla \varphi _{\gamma }(x^k) - \nabla \varphi _{\gamma }(w^k) + \nabla ^2\varphi _{\gamma }(x_\star )(w^k-x^k)}{\Vert w^k-x^k\Vert }\\&{}-\frac{\nabla \varphi _{\gamma }(w^k)}{\Vert w^k-x^k\Vert }. \end{aligned}$$

By strict differentiability of \(\nabla \varphi _{\gamma }\) at \(x_\star \) we obtain

$$\begin{aligned} \lim _{k\rightarrow \infty }{ \frac{ \Vert \nabla \varphi _{\gamma }(w^k)\Vert }{ \Vert w^k-x^k\Vert } } {}={} 0. \end{aligned}$$
(6.14)

By nonsingularity of \(\nabla ^2\varphi _{\gamma }(x_\star )\) and since \(w^k\rightarrow x^\star \), there exists \(\alpha >0\) such that \( \Vert \nabla \varphi _{\gamma }(x^k)\Vert \ge \alpha \Vert x^k-x_\star \Vert \) for k large enough. Therefore, for k sufficiently large,

$$\begin{aligned} \frac{ \Vert \nabla \varphi _{\gamma }(w^k)\Vert }{ \Vert w^k-x^k\Vert } {}\ge {} \frac{ \alpha \Vert w^k-x_\star \Vert }{ \Vert w^k-x^k\Vert } {}\ge {} \frac{ \alpha \Vert w^k-x_\star \Vert }{ \Vert w^k-x_\star \Vert +\Vert x^k-x_\star \Vert }. \end{aligned}$$

Using (6.14) we get

$$\begin{aligned} \lim _{k\rightarrow \infty }{ \frac{ \Vert w^k-x_\star \Vert }{ \Vert w^k-x_\star \Vert +\Vert x^k-x_\star \Vert } } {}={} \lim _{k\rightarrow \infty }{ \frac{ \Vert w^k-x_\star \Vert /\Vert x^k-x_\star \Vert }{ \Vert w^k-x_\star \Vert /\Vert x^k-x_\star \Vert +1 } } {}={} 0, \end{aligned}$$

from which we obtain

$$\begin{aligned} \lim _{k\rightarrow \infty }{ \frac{\Vert w^k-x_\star \Vert }{\Vert x^k-x_\star \Vert } } {}={} 0. \end{aligned}$$
(6.15)

Finally,

$$\begin{aligned} \Vert x^{k+1}-x_\star \Vert {}={}&\Vert T_{\gamma }(w^k)-T_{\gamma }(x_\star )\Vert \nonumber \\ {}={}&\left\| \mathop {\mathrm{prox}}\nolimits _{\gamma g}(w^k-\gamma \nabla f(w^k)) {}-{} \mathop {\mathrm{prox}}\nolimits _{\gamma g}(x_\star -\gamma \nabla f(x_\star )) \right\| \nonumber \\ {}\le {}&\left\| w^k - \gamma \nabla f(w^k) {}-{} x_\star + \gamma \nabla f(x_\star ) \right\| \nonumber \\ {}\le {}&(1+\gamma L_f)\Vert w^k-x_\star \Vert , \end{aligned}$$
(6.16)

where the first inequality follows from nonexpansiveness of \(\mathop {\mathrm{prox}}\nolimits _{\gamma g}\) and the second from Lipschitz continuity of \(\nabla f\). Using (6.16) in (6.15) we obtain that and converge Q-superlinearly to \(x_\star \). \(\square \)

Proof of Theorem 4.2

From Proposition 6.4(a) it follows that \(\nabla \varphi _{\gamma }\) is strictly differentiable and continuously semidifferentiable at \(x_\star \). Moreover, we know from Lemma 3.1 that eventually \(\gamma _k = \gamma _\infty > 0\). Therefore we assume that k is large enough for this purpose and indicate \(\gamma \) in place of \(\gamma _k\) for simplicity. We denote for short \(g^k = \nabla \varphi _{\gamma }(x^k)\). In

figure bw
$$\begin{aligned} w^k-x^k = \tau _k d^k = -\tau _k B_k^{-1}g^k, \end{aligned}$$

and by (4.1) and Cauchy-Schwarz inequality

$$\begin{aligned} \frac{\Vert (B_k-\nabla ^2\varphi _{\gamma }(x_\star ))(w^k-x^k)\Vert }{\Vert w^k-x^k\Vert }&= \frac{\Vert g^k+\nabla ^2\varphi _{\gamma }(x_\star )d^k\Vert }{\Vert d^k\Vert } \\&\ge \left| \frac{\langle d^k,g^k+\nabla ^2\varphi _{\gamma }(x_\star )d^k \rangle }{\Vert d^k\Vert ^2}\right| \rightarrow 0. \end{aligned}$$

Therefore

$$\begin{aligned} -\langle g^k,d^k \rangle = \langle d^k,\nabla ^2\varphi _{\gamma }(x_\star )d^k \rangle + o(\Vert d^k\Vert ^2). \end{aligned}$$
(6.17)

Since \(\nabla ^2\varphi _{\gamma }(x_\star )\) is positive definite, there is \(\eta >0\) such that for sufficiently large k

$$\begin{aligned} - \langle g^k,d^k \rangle \ge \eta \Vert d^k\Vert ^2. \end{aligned}$$
(6.18)

Since is continuous at \(x_\star \) and \(x^k\rightarrow x_\star \), we have

(6.19)

Next, since \(x^k\rightarrow x_\star \), for k large enough \(\nabla \varphi _{\gamma }\) is semidifferentiable at \(x^k\) and we can expand \(\varphi _{\gamma }\) around \(x^k\) using [31, Ex. 13.7(c)] to obtain

figure bx

where the second equality is due to (6.19), and the last equality is due to (6.17). Therefore, using (6.18), for sufficiently large k

$$\begin{aligned} \varphi _{\gamma }(x^k+d^k) - \varphi _{\gamma }(x^k) \le -\tfrac{\eta }{2}\Vert d^k\Vert ^2 < 0, \end{aligned}$$

i.e., \(\tau _k = 1\) satisfies the non-increase condition. As a consequence,

figure by

eventually reduces to the iterations of Theorem 4.1 and the proof follows. \(\square \)

Proof of Theorem 4.3

Suppose that Assumption 6(i) holds. Since \(x_\star \in \mathop {\mathrm{zer}}\nolimits \partial \varphi \) and \(\nabla ^2\varphi _{\gamma }(x_\star ) \succ 0\), it follows that \(x_\star \) is a strong local minimizer of \(\varphi _{\gamma }\), hence of \(\varphi \) in light of Proposition 2.2(i) and 2.3(i). Theorem 3.7 then ensures that and converge linearly to \(x_\star \). If instead Assumption 6(ii) holds, then we can invoke Theorem 3.11 (since \(\Vert \nabla \varphi _{\gamma _k}(x^k)\Vert \le (1+\gamma _0L_f)\Vert R_{\gamma _k}(x^k)\Vert \)) to infer that and converge linearly to a critical point, be it \(x_\star \). In both cases we can apply Proposition 6.5 and for k sufficiently large

(6.20)

Since the convergence is linear, then the right-hand side of (6.20) is summable. With similar arguments to those of [26, Lem. 3.2] we can see that eventually \(\langle s^k,y^k \rangle >0\). Therefore we can apply [70, Thm. 3.2], which ensures that condition (4.1) holds. The result follows then from Theorem 4.2. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Stella, L., Themelis, A. & Patrinos, P. Forward–backward quasi-Newton methods for nonsmooth optimization problems. Comput Optim Appl 67, 443–487 (2017). https://doi.org/10.1007/s10589-017-9912-y

Download citation

  • Received:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10589-017-9912-y

Keywords

Navigation