Skip to main content
Log in

Faster first-order primal-dual methods for linear programming using restarts and sharpness

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

    We’re sorry, something doesn't seem to be working properly.

    Please try refreshing the page. If that doesn't work, please contact support so we can address the problem.

Abstract

First-order primal-dual methods are appealing for their low memory overhead, fast iterations, and effective parallelization. However, they are often slow at finding high accuracy solutions, which creates a barrier to their use in traditional linear programming (LP) applications. This paper exploits the sharpness of primal-dual formulations of LP instances to achieve linear convergence using restarts in a general setting that applies to alternating direction method of multipliers (ADMM), primal-dual hybrid gradient method (PDHG) and extragradient method (EGM). In the special case of PDHG, without restarts we show an iteration count lower bound of \(\Omega (\kappa ^2 \log (1/\epsilon ))\), while with restarts we show an iteration count upper bound of \(O(\kappa \log (1/\epsilon ))\), where \(\kappa \) is a condition number and \(\epsilon \) is the desired accuracy. Moreover, the upper bound is optimal for a wide class of primal-dual methods, and applies to the strictly more general class of sharp primal-dual problems. We develop an adaptive restart scheme and verify that restarts significantly improve the ability of PDHG, EGM, and ADMM to find high accuracy solutions to LP problems.

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

Similar content being viewed by others

Notes

  1. A robust algorithm for LP would need to detect violations of this assumption, i.e., when the problem is infeasible or unbounded. We refer readers to [6] to understand the behavior of primal-dual methods when applied to infeasible or unbounded LP problems.

  2. https://github.com/google-research/FirstOrderLp.jl.

  3. https://developers.google.com/optimization.

  4. The RHS of (4b) is well-defined because lim sup always exists [76, Section 5.3]. Later we show (Proposition 5) that \(\rho _r(z)\) is monotonically non-increasing in \(r \in (0,\infty )\) for fixed z which means that \(\rho _{0}(z) = \limsup _{r \rightarrow 0^{+}} \rho _r(z) = \lim _{r \rightarrow 0^{+}} \rho _{r}(z) < \infty \).

  5. PDHG is often presented in a form with different primal and dual step sizes [14, 15]. Here, we choose to use the same primal and dual step size for consistent notation with other primal-dual algorithms. Our results can easily extend to the case of different step sizes by rescaling. In particular, by setting \(\eta = \sqrt{\sigma \tau }\) and defining a rescaled space: \(( \hat{x}, \hat{y}) = ( x \sqrt{\eta / \tau }, y \sqrt{\eta / \sigma })\) where \(\tau \in (0,\infty )\) is the desired primal step and \(\sigma \in (0,\infty )\) is the desired dual step size. Applying (9) to this rescaled space, i.e., replacing f(x) with \(\hat{f}(\hat{x}) = f(\hat{x} / \sqrt{\eta / \tau })\), g(y) with \(\hat{g}(\hat{x}) = g(\hat{y} / \sqrt{\eta / \sigma })\), X with \(\hat{X} = \{ x \sqrt{\eta / \tau }: x \in X \}\), Y with \(\hat{Y} = \{ y \sqrt{\eta / \sigma }: y \in Y \}\), \(\Vert x - x^t \Vert \) with \(\Vert \hat{x}^t - x^t \Vert \), and \(\Vert y - y^t \Vert \) with \(\Vert \hat{y}^t - y^t \Vert \), then substituting back \((\hat{x},\hat{y}) = ( x \sqrt{\eta / \tau }, y \sqrt{\eta / \sigma })\) and \((\hat{x}^{t},\hat{y}^{t}) = ( x^t \sqrt{\eta / \tau }, y^t \sqrt{\eta / \sigma })\) yields the classic PDHG update:

    $$\begin{aligned} x^{t+1}&\in {\mathop {\textrm{argmin}}\limits _{x\in X}} f(x)+(y^t)^{\top } Ax+\frac{1}{2 \tau } \Vert x-x^t\Vert _2^2 \\ y^{t+1}&\in {\mathop {\textrm{argmin}}\limits _{y\in Y}} g(y)-y^{\top } A(2 x^{t+1}-x^t) +\frac{1}{2 \sigma } \Vert y-y^t\Vert _2^2 \ . \end{aligned}$$
  6. A function f grows \(\mu \)-quadratically if \({f(z) - f^{\star }} \ge {\mu }\mathbf{dist{}}(z, Z^{\star })^2\).

  7. The active variables are variables not at their bounds.

  8. In the LP case (19), non-degeneracy means that the algorithm converges to a primal-dual solution that satisfies strict complimentary.

  9. The restart lengths were ordered descending by iterations to find a normalized duality gap below \(10^{-7}\), and then by the normalized duality gap at the maximum number of iterations. The top three restart lengths were then selected for display.

References

  1. Alacaoglu, A., Fercoq, O., Cevher, V.: On the convergence of stochastic primal-dual hybrid gradient, arXiv preprint arXiv:1911.00799 (2019)

  2. Alamo, T., Limon, D., Krupa, P.: Restart FISTA with global linear convergence. In: 18th European Control Conference (ECC). IEEE, vol. 2019, pp. 1969–1974 (2019)

  3. Andersen, E.D., Andersen, K.D.: The MOSEK interior point optimizer for linear programming: an implementation of the homogeneous algorithm. In: High Performance Optimization, pp. 197–232. Springer (2000)

  4. Anderson, R.I., Fok, R., Scott, J.: Hotel industry efficiency: an advanced linear programming examination. Am. Bus. Rev. 18(1), 40 (2000)

    Google Scholar 

  5. Applegate, D., Díaz, M., Hinder, O., Lu, H., Lubin, M., O’Donoghue, B., Schudy, W.: Practical large-scale linear programming using primal-dual hybrid gradient. In: Advances in Neural Information Processing Systems, vol. 34 (2021)

  6. Applegate, D., Díaz, M., Lu, H., Lubin, M.: Infeasibility detection with primal-dual hybrid gradient for large-scale linear programming, arXiv preprint arXiv:2102.04592 (2021)

  7. Basu, K., Ghoting, A., Mazumder, R., Pan, Y.: ECLIPSE: an extreme-scale linear program solver for web-applications. In: Daumé III, H., Singh, A. (eds.) Proceedings of the 37th International Conference on Machine Learning (Virtual), Proceedings of Machine Learning Research, PMLR, vol. 119, pp. 704–714 (2020)

  8. Basu, K., Ghoting, A., Mazumder, R., Pan, Y.: Eclipse: an extreme-scale linear program solver for web-applications. In: International Conference on Machine Learning, PMLR, pp. 704–714 (2020)

  9. Blum, M., Floyd, R.W., Pratt, V., Rivest, R.L., Tarjan, R.E.: Time bounds for selection. J. Comput. Syst. Sci. 7(4), 448–461 (1973)

    MathSciNet  MATH  Google Scholar 

  10. Bowman, E.H.: Production scheduling by the transportation method of linear programming. Oper. Res. 4(1), 100–103 (1956)

    MATH  Google Scholar 

  11. Boyd, S., Parikh, N., Chu, E., Peleato, B., Eckstein, J., et al.: Distributed optimization and statistical learning via the alternating direction method of multipliers. Found. Trends® Mach. Learn. 3(1), 1–122 (2011)

    MATH  Google Scholar 

  12. Burke, J.V., Ferris, M.C.: Weak sharp minima in mathematical programming. SIAM J. Control Optim. 31(5), 1340–1359 (1993)

    MathSciNet  MATH  Google Scholar 

  13. Burke, V.J., Ferris, M.C.: A Gauss-Newton method for convex composite optimization. Math. Program. 71(2), 179–194 (1995)

    MathSciNet  MATH  Google Scholar 

  14. Chambolle, A., Pock, T.: A first-order primal-dual algorithm for convex problems with applications to imaging. J. Math. Imaging Vis. 40(1), 120–145 (2011)

    MathSciNet  MATH  Google Scholar 

  15. Chambolle, A., Pock, T.: On the ergodic convergence rates of a first-order primal-dual algorithm. Math. Program. 159(1–2), 253–287 (2016)

    MathSciNet  MATH  Google Scholar 

  16. Charnes, A., Cooper, W.W.: The stepping stone method of explaining linear programming calculations in transportation problems. Manag. Sci. 1(1), 49–69 (1954)

    MathSciNet  MATH  Google Scholar 

  17. Condat, L., Malinovsky, G., Richtárik, P.: Distributed proximal splitting algorithms with rates and acceleration. Front. Signal Process. 12 (2022)

  18. Dantzig, G.B.: Linear Programming and Extensions, vol. 48. Princeton University Press (1998)

  19. Daskalakis, C., Andrew, I., Syrgkanis, V., Zeng, H.: Training GANs with optimism. In: International Conference on Learning Representations (2018)

  20. Davis, D., Drusvyatskiy, D., MacPhee, K.J., Paquette, C.: Subgradient methods for sharp weakly convex functions. J. Optim. Theory Appl. 179(3), 962–982 (2018)

    MathSciNet  MATH  Google Scholar 

  21. Douglas, J., Rachford, H.H.: On the numerical solution of heat conduction problems in two and three space variables. Trans. Am. Math. Soc. 82(2), 421–439 (1956)

    MathSciNet  MATH  Google Scholar 

  22. Eckstein, J., Bertsekas, D.P.: On the Douglas–Rachford splitting method and the proximal point algorithm for maximal monotone operators. Math. Program. 55(1–3), 293–318 (1992)

    MathSciNet  MATH  Google Scholar 

  23. Eckstein, J., Bertsekas, D.P.: et al., An alternating direction method for linear programming (1990)

  24. Fercoq, O.: Quadratic error bound of the smoothed gap and the restarted averaged primal-dual hybrid gradient (2021)

  25. Fercoq, O., Zheng, Q.: Adaptive restart of accelerated gradient methods under local quadratic growth condition. IMA J. Numer. Anal. 39(4), 2069–2095 (2019)

    MathSciNet  MATH  Google Scholar 

  26. Ferris, M.C.: Finite termination of the proximal point algorithm. Math. Program. 50(1–3), 359–366 (1991)

    MathSciNet  MATH  Google Scholar 

  27. Freund, R.M., Haihao, L.: New computational guarantees for solving convex optimization problems with first order methods, via a function growth condition measure. Math. Program. 170(2), 445–477 (2018)

    MathSciNet  MATH  Google Scholar 

  28. Galabova, I.L., Hall, J.A.J.: The ‘idiot’ crash quadratic penalty algorithm for linear programming and its application to linearizations of quadratic assignment problems. Optim. Methods Softw. 35(3), 488–501 (2020)

    MathSciNet  MATH  Google Scholar 

  29. Gilpin, A., Pena, J., Sandholm, T.: First-order algorithm with \(\cal{O} (\ln (1/\epsilon ))\)-convergence for \(\epsilon \)-equilibrium in two-person zero-sum games. Math. Program. 133(1), 279–298 (2012)

    MathSciNet  MATH  Google Scholar 

  30. Giselsson, P., Boyd, S.: Monotonicity and restart in fast gradient methods. In: 53rd IEEE Conference on Decision and Control, pp. 5058–5063. IEEE (2014)

  31. Goldstein, T., Li, M., Yuan, X.: Adaptive primal-dual splitting methods for statistical learning and image processing. In: Advances in Neural Information Processing Systems, pp. 2089–2097 (2015)

  32. Gondzio, J.: Interior point methods 25 years later. Eur. J. Oper. Res. 218(3), 587–601 (2012)

    MathSciNet  MATH  Google Scholar 

  33. Güler, O., Hoffman, A.J., Rothblum, U.G.: Approximations to solutions to systems of linear inequalities. SIAM J. Matrix Anal. Appl. 16(2), 688–696 (1995)

    MathSciNet  MATH  Google Scholar 

  34. Gutman, D.H., Peña, J.F.: The condition number of a function relative to a set. Math. Program. (2020), to appear

  35. Hanssmann, F., Hess, S.W.: A linear programming approach to production and employment scheduling. Manag. Sci. 1, 46–51 (1960)

    MathSciNet  Google Scholar 

  36. Harker, P.T., Pang, J.-S.: Finite-dimensional variational inequality and nonlinear complementarity problems: a survey of theory, algorithms and applications. Math. Program. 48(1–3), 161–220 (1990)

    MathSciNet  MATH  Google Scholar 

  37. He, B., Yuan, X.: On the \({O}(1/n)\) convergence rate of the Douglas–Rachford alternating direction method. SIAM J. Numer. Anal. 50(2), 700–709 (2012)

    MathSciNet  MATH  Google Scholar 

  38. Hoffman, A.J.: On approximate solutions of systems of linear inequalities. J. Res. Natl. Bur. Stand. 49, 263–265 (1952)

    MathSciNet  Google Scholar 

  39. Hunter, J.K., Nachtergaele, B.: Applied Analysis. World Scientific Publishing Company (2001)

  40. Johnson, R., Zhang, T.: Accelerating stochastic gradient descent using predictive variance reduction. Adv. Neural. Inf. Process. Syst. 26, 315–323 (2013)

    Google Scholar 

  41. Karmarkar, N.: A new polynomial-time algorithm for linear programming. In: Proceedings of the Sixteenth Annual ACM Symposium on Theory of Computing, pp. 302–311 (1984)

  42. Klatte, D., Thiere, G.: Error bounds for solutions of linear equations and inequalities. Z. Oper. Res. 41(2), 191–214 (1995)

    MathSciNet  MATH  Google Scholar 

  43. Korpelevich, G.M.: The extragradient method for finding saddle points and other problems. Matecon 12, 747–756 (1976)

    MathSciNet  MATH  Google Scholar 

  44. Lewis, A.S, Liang, J.: Partial smoothness and constant rank, arXiv preprint arXiv:1807.03134 (2018)

  45. Li, X., Sun, D., Toh, K.-C.: An asymptotically superlinearly convergent semismooth newton augmented Lagrangian method for linear programming. SIAM J. Optim. 30(3), 2410–2440 (2020)

    MathSciNet  MATH  Google Scholar 

  46. Liang, J., Fadili, J., Peyré, G.: Local linear convergence analysis of primal-dual splitting methods. Optimization 67(6), 821–853 (2018)

    MathSciNet  MATH  Google Scholar 

  47. Lin, H., Mairal, J., Harchaoui, Z.: A universal catalyst for first-order optimization. In: Advances in Neural Information Processing Systems, pp. 3384–3392 (2015)

  48. Lin, Q., Xiao, L.: An adaptive accelerated proximal gradient method and its homotopy continuation for sparse optimization. In: International Conference on Machine Learning, pp. 73–81 (2014)

  49. Lin, T., Ma, S., Ye, Y., Zhang, S.: An ADMM-based interior-point method for large-scale linear programming. Optim. Methods Softw. 36(2–3), 389–424 (2021)

    MathSciNet  MATH  Google Scholar 

  50. Liu, Q., Van Ryzin, G.: On the choice-based linear programming model for network revenue management. Manuf. Serv. Oper. Manag. 10(2), 288–310 (2008)

    Google Scholar 

  51. Lu, H.: An \({O}(s^r)\)-resolution ODE framework for discrete-time optimization algorithms and applications to convex-concave saddle-point problems, arXiv preprint arXiv:2001.08826 (2020)

  52. Luo, Z.-Q., Tseng, P.: Error bounds and convergence analysis of feasible descent methods: a general approach. Ann. Oper. Res. 46(1), 157–178 (1993)

    MathSciNet  MATH  Google Scholar 

  53. Manne, A.S.: Linear programming and sequential decisions. Manag. Sci. 6(3), 259–267 (1960)

    MathSciNet  MATH  Google Scholar 

  54. Marcotte, P., Zhu, D.: Weak sharp solutions of variational inequalities. SIAM J. Optim. 9(1), 179–189 (1998)

    MathSciNet  MATH  Google Scholar 

  55. Mittelmann, H.D.: Benchmark of simplex LP solvers (2020). http://plato.asu.edu/ftp/lpsimp.html

  56. Mokhtari, A., Ozdaglar, A., Pattathil, S.: A unified analysis of extra-gradient and optimistic gradient methods for saddle point problems: proximal point approach. In: International Conference on Artificial Intelligence and Statistics (2020)

  57. Nemirovski, A.: Prox-method with rate of convergence O(1/t) for variational inequalities with Lipschitz continuous monotone operators and smooth convex–concave saddle point problems. SIAM J. Optim. 15(1), 229–251 (2004)

    MathSciNet  MATH  Google Scholar 

  58. Nesterov, Yu.: Subgradient methods for huge-scale optimization problems. Math. Program. 146(1), 275–297 (2014)

    MathSciNet  MATH  Google Scholar 

  59. Nesterov, Y.: Smooth minimization of non-smooth functions. Math. Program. 103(1), 127–152 (2005)

    MathSciNet  MATH  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

  61. Nesterov, Y.: Introductory Lectures on Convex Optimization: a basic course, vol. 87. Springer (2013)

  62. Nesterov, Y.E.: A method for solving the convex programming problem with convergence rate \({O} (1/k^2)\). Soviet Math. Doklady 27, 372–376 (1983)

    MATH  Google Scholar 

  63. Niao, H.: Mirror-prox algorithm, Fall (2016), http://niaohe.ise.illinois.edu/IE598_2016/pdf/IE598-lecture18-mirror%20prox%20algorithm%20for%20saddle%20point%20problems.pdf

  64. O’Donoghue, B., Candes, E.: Adaptive restart for accelerated gradient schemes. Found. Comput. Math. 15(3), 715–732 (2015)

    MathSciNet  MATH  Google Scholar 

  65. O’Donoghue, B., Chu, E., Parikh, N., Boyd, S.: Conic optimization via operator splitting and homogeneous self-dual embedding. J. Optim. Theory Appl. 169(3), 1042–1068 (2016)

    MathSciNet  MATH  Google Scholar 

  66. Peña, J., Vera, J.C., Zuluaga, L.F.: New characterizations of Hoffman constants for systems of linear constraints. Math. Program. (2020), to appear

  67. Pokutta, S.: Restarting algorithms: sometimes there is free lunch. In: International Conference on Integration of Constraint Programming, Artificial Intelligence, and Operations Research, pp. 22–38. Springer (2020)

  68. Polyak, B.: Sharp minima. In: Proceedings of the IIASA Workshop on Generalized Lagrangians and Their Applications, Laxenburg, Austria. Institute of Control Sciences Lecture Notes, Moscow (1979)

  69. Polyak, B.: Introduction to Optimization. Optimization Software Inc, New York (1987)

    MATH  Google Scholar 

  70. Ramakrishnan, K.G., Resende, M.G.C., Ramachandran, B., Pekny, J.F.: Tight QAP Bounds via Linear Programming, pp. 297–303. World Scientific Publishing Co. (2002)

  71. Renegar, J.: Incorporating condition measures into the complexity theory of linear programming. SIAM J. Optim. 5(3), 506–524 (1995)

    MathSciNet  MATH  Google Scholar 

  72. Renegar, J.: Linear programming, complexity theory and elementary functional analysis. Math. Program. 70(1–3), 279–351 (1995)

    MathSciNet  MATH  Google Scholar 

  73. Roulet, V., d’Aspremont, A.: Sharpness, restart, and acceleration. SIAM J. Optim. 30(1), 262–289 (2020)

    MathSciNet  MATH  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

  75. Tang, J., Golbabaee, M., Bach, F. et al.: Rest-katyusha: exploiting the solution’s structure via scheduled restart schemes. In: Advances in Neural Information Processing Systems, pp. 429–440 (2018)

  76. Thomson, B.S., Bruckner, J.B., Bruckner, A.M.: Elementary real analysis, vol. 1, ClassicalRealAnalysis.com (2008)

  77. Tseng, P.: On linear convergence of iterative methods for the variational inequality problem. J. Comput. Appl. Math. 60(1–2), 237–252 (1995)

    MathSciNet  MATH  Google Scholar 

  78. Yang, T., Lin, Q.: RSG: beating subgradient method without smoothness and strong convexity. J. Mach. Learn. Res. 19(1), 236–268 (2018)

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

We thank Brendan O’Donoghue, Vasilis Charisopoulos, and Warren Schudy for useful discussions and feedback on this work.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Oliver Hinder.

Additional information

Publisher's Note

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

Appendices

Proof of Lemma 2

Proof

Let \(U \Sigma V^{\top }\) be a singular value decomposition of H. Recall U and V are orthogonal matrices (\(U^{\top } U = U U^{\top } = {{\textbf {I}}}\) and \(\Vert U z \Vert = \Vert z \Vert \)), and \(\Sigma \) is diagonal. First, by assumption there exists some \(\bar{z}\) such that \(H \bar{z} = h\). This implies that for \(\bar{p} = V^{\top } \bar{z}\) we have \(U^{\top } h = U^{\top } (U \Sigma V^{\top }) \bar{z} = \Sigma V^{\top } \bar{z} = \Sigma \bar{p}\). Consider some \(z \in {\mathbb {R}}^{n}\). Define \(p = V^{\top } z\),

$$\begin{aligned} p^{*}_i = {\left\{ \begin{array}{ll} \Sigma _{ii}^{-1} (U^{\top } h)_i &{}\quad \Sigma _{ii} \ne 0 \\ z_i &{}\quad \text {otherwise} \end{array}\right. } \end{aligned}$$

and \(z^{\star }= V p^{*}\). If \(\Sigma _{ii} \ne 0\) then by definition of \(p^{*}\), \((\Sigma p^{*})_i = (U^{\top } h)_i\). If \(\Sigma _{ii} = 0\) then \((U^{\top } h)_i = (\Sigma \bar{p})_i = \Sigma _{ii} \bar{p}_i = 0 = \Sigma _{ii} p^{*}_i = (\Sigma p^{*})_i\). Hence \(\Sigma p^{*} = U^{\top } h\). Therefore,

$$\begin{aligned} \Vert z^{\star }- z \Vert= & {} \Vert V^{\top } z^{\star }- V^{\top } z \Vert \\= & {} \Vert p^{*} - p \Vert \\\le & {} \frac{1}{{\sigma _{\min }^+}{(H)}} \Vert \Sigma p^{*} - \Sigma p \Vert \\= & {} \frac{1}{{\sigma _{\min }^+}{(H)}} \Vert (U^{\top } h) - \Sigma p \Vert \\= & {} \frac{1}{{\sigma _{\min }^+}{(H)}} \Vert h - U \Sigma V^{\top } z \Vert \\= & {} \frac{1}{{\sigma _{\min }^+}{(H)}} \Vert h - H z \Vert . \end{aligned}$$

\(\square \)

Proof of Proposition 3 for EGM

Proof

Let \(z^{\star }= \min _{z \in Z^{\star }} \Vert z^{t} - z \Vert \). By definition of \(\hat{z}^t\) and that F is \(L\)-Lipschitz, we have

$$\begin{aligned}&F(z^{t-1})^{\top } (\hat{z}^t - z^{t-1}) + \frac{1}{2 \eta } \Vert \hat{z}^t - z^{t-1} \Vert ^2 \nonumber \\&\quad \le F(z^{t-1})^{\top } (z^{\star }- z^{t-1}) + \frac{1}{2 \eta } \Vert z^{\star }- z^{t-1} \Vert ^2 \nonumber \\&\quad = F(z^{\star })^{\top } (z^{\star }- z^{t-1}) + (F(z^{t-1}) - F(z^{\star }))^{\top } (z^{\star }- z^{t-1}) + \frac{1}{2 \eta } \Vert z^{\star }- z^{t-1} \Vert ^2 \ . \end{aligned}$$
(57)

Note that

$$\begin{aligned} \begin{aligned} F(z^{t-1})^{\top } (\hat{z}^t - z^{t-1}) =F(z^{\star })^{\top } (\hat{z}^t - z^{t-1}) + (F(z^{t-1}) - F(z^{\star }))^{\top } (\hat{z}^t - z^{\star }) + \\ (F(z^{t-1}) - F(z^{\star }))^{\top } (z^{\star }- z^{t-1}) \ . \end{aligned} \end{aligned}$$
(58)

Substituting (58) into the LHS of (57) and cancelling terms yields

$$\begin{aligned} \begin{aligned}&F(z^{\star })^{\top } (\hat{z}^t - z^{t-1}) + (F(z^{t-1}) - F(z^{\star }))^{\top } (\hat{z}^t - z^{\star }) + \frac{1}{2 \eta } \Vert \hat{z}^t - z^{t-1} \Vert ^2 \\&\le F(z^{\star })^{\top } (z^{\star }- z^{t-1}) + \frac{1}{2 \eta } \Vert z^{\star }- z^{t-1} \Vert ^2\ . \end{aligned} \end{aligned}$$
(59)

By \(F(z^{\star })^{\top } (z^{\star }- z^{t-1}) \le F(z^{\star })^{\top } (\hat{z}^t - z^{t-1})\), (59), F is \(L\)-Lipschitz, and the triangle inequality one gets

$$\begin{aligned} 0&\ge F(z^{\star })^{\top } (z^{\star }- z^{t-1}) - F(z^{\star })^{\top } (\hat{z}^t - z^{t-1}) \nonumber \\&\ge \frac{1}{2 \eta } \Vert \hat{z}^t - z^{t-1} \Vert ^2 + (F(z^{t-1}) - F(z^{\star }))^{\top } (\hat{z}^t - z^{\star }) - \frac{1}{2 \eta } \Vert z^{\star }- z^{t-1} \Vert ^2 \nonumber \\&\ge \frac{1}{2 \eta } \Vert \hat{z}^t - z^{t-1} \Vert ^2 - L\Vert z^{t-1} - z^{\star }\Vert \Vert \hat{z}^t - z^{\star }\Vert - \frac{1}{2 \eta } \Vert z^{\star }- z^{t-1} \Vert ^2 \nonumber \\&= \frac{1}{2 \eta } \Vert \hat{z}^t - z^{t-1} \Vert ^2 - L\Vert z^{t-1} - z^{\star }\Vert \Vert z^{t-1} - z^{\star }+ \hat{z}^t - z^{t-1} \Vert - \frac{1}{2 \eta } \Vert z^{\star }- z^{t-1} \Vert ^2 \nonumber \\&\ge \frac{1}{2 \eta } \Vert \hat{z}^t - z^{t-1} \Vert ^2 - L\Vert z^{t-1} - z^{\star }\Vert \Vert \hat{z}^t - z^{t-1} \Vert - \left( L+ \frac{1}{2 \eta } \right) \Vert z^{\star }- z^{t-1} \Vert ^2 \nonumber \\&= \left( \frac{1}{2\eta } \Vert \hat{z}^t - z^{t-1} \Vert - \left( \frac{1}{2\eta }+L\right) \Vert z^{\star }- z^{t-1} \Vert \right) \left( \Vert \hat{z}^t - z^{t-1} \Vert +\Vert z^{\star }- z^{t-1} \Vert \right) \ \end{aligned}$$
(60)

where the first inequality uses \(F(z^{\star })^{\top } (z^{\star }- z^{t-1}) \le F(z^{\star })^{\top } (\hat{z}^t - z^{t-1})\), the second inequality uses (59), the third inequality uses that F is \(L\)-Lipschitz, and the final inequality uses the triangle inequality.

Notice that \(\Vert \hat{z}^t - z^{t-1} \Vert +\Vert z^{\star }- z^{t-1} \Vert > 0\) otherwise the result trivially holds, thus by (60) we get \(\frac{1}{2\eta } \Vert \hat{z}^t - z^{t-1} \Vert - \left( \frac{1}{2\eta }+L\right) \Vert z^{\star }- z^{t-1} \Vert \le 0\), rearranging yields

$$\begin{aligned} \Vert \hat{z}^t - z^{t-1} \Vert \le (1+ 2\eta L) \Vert z^{\star }- z^{t-1} \Vert \le 3 \Vert z^{\star }- z^{t-1} \Vert \, \end{aligned}$$

where the last inequality uses that \(\eta \le 1 / L\) from the premise of Proposition 1. \(\square \)

Proof of Lemma 6

Proof

Recall for ADMM the norm is \(\Vert z \Vert = \sqrt{z^{\top } M z}\) with

$$\begin{aligned} M= \begin{pmatrix} 0 &{}\quad 0 &{}\quad 0 \\ 0 &{}\quad \eta V^{\top } V&{}\quad 0 \\ 0 &{}\quad 0 &{}\quad \frac{1}{\eta } {{\textbf {I}}}\end{pmatrix} = \begin{pmatrix} 0 &{}\quad 0 &{} \quad 0 \\ 0 &{}\quad \eta {{\textbf {I}}}&{}\quad 0 \\ 0 &{}\quad 0 &{}\quad \frac{1}{\eta } {{\textbf {I}}}\end{pmatrix}\, \end{aligned}$$

where we ustilize \(V=I\) in (28). Suppose that \(\eta = 1\). Consider \(z \in Z\), \(r \in (0, 2R]\). Let

$$\begin{aligned} v = \begin{pmatrix} 0 \\ (c + y)^{-} \\ x_{V} - x_{U} \end{pmatrix} \text { and note that } F(z) = \begin{pmatrix} -y \\ c + y \\ x_{U} - x_{V} \end{pmatrix}. \end{aligned}$$

Let \(z_1 = z + r \frac{v}{\Vert v \Vert }\), then \(z_1\in Z\) by the definition of Z. It follows from (21) that

$$\begin{aligned} \begin{aligned} \rho _r(z)&\ge \frac{F(z)^{\top } (z - z_1)}{r} = -\frac{1}{\Vert v \Vert } F(z)^{\top } v = \left\| v \right\| _{2} \ . \end{aligned} \end{aligned}$$
(61)

Consider any optimal solution \(z^{\star }= (x^{\star }_{V}, x^{\star }_{V}, y^{\star })\), define \(\hat{z} = (x^{\star }_{V}, x^{\star }_{V}, y) \in Z^{\star }\). Let \(z_2= z - \mu (z - \hat{z})\) for \(\mu = \min \left\{ \frac{r}{\Vert z - \hat{z} \Vert }, 1 \right\} \), then \(\Vert z_2-z\Vert \le \frac{r}{\Vert z - \hat{z} \Vert }\Vert z - \hat{z} \Vert \le r\). Meanwhile, we have \(z_2 = (1- \mu ) z + \mu \hat{z}\), thus \(z_2\in Z\) by convexity. We conclude that \(z_2 \in W_{r}(z)\). Therefore, it follows from (21) that

$$\begin{aligned} \begin{aligned} \rho _r(z)&\ge \frac{1}{r} \min \left\{ \frac{r}{\Vert z - \hat{z} \Vert }, 1 \right\} F(z)^{\top } (z - \hat{z}) \\&= \min \left\{ \frac{1}{\Vert z - \hat{z} \Vert }, \frac{1}{r} \right\} \begin{pmatrix} -y \\ c+y \\ x_{U} - x_{V} \end{pmatrix} \cdot \begin{pmatrix} x_U - x^{\star }_V \\ x_{V} - x^{\star }_V \\ y \end{pmatrix} \\&= \min \left\{ \frac{1}{\Vert z - \hat{z} \Vert }, \frac{1}{r} \right\} c^{\top } (x_{V} - x^{\star }_{V}) \ . \end{aligned} \end{aligned}$$
(62)

Substituting \(r \in (0, 2R]\) and \(\Vert z - \hat{z} \Vert \le \Vert z - z^{\star }\Vert \in [0,2R]\) into (62) and noticing \(\rho _r(z)\ge 0\) yields

$$\begin{aligned} \begin{aligned} \rho _r(z)&\ge \frac{1}{2R} (c^{\top } x_{V} - c^{\top } x^{\star }_{V})^+ \ . \end{aligned} \end{aligned}$$
(63)

Combining (61) and (63), we deduce there exists \(K'\) and \(h'\) such that

$$\begin{aligned} (4R^2 +1)\rho _r(z)^2&\ge ((c^{\top } x_{V} - c^{\top } x^{\star }_{V})^+)^2 + \Vert (c + y)^{-} \Vert _2^2 + \Vert x_V - x_U \Vert _2^2 \\&= ((c^{\top } x_{V} - c^{\top } x^{\star }_{V})^+)^2 + \Vert (c + y)^{-} \Vert _2^2\\&\qquad + \Vert x_V - x_U \Vert _2^2 + \Vert x_U^{-} \Vert _2^2 + \Vert A x_{V} - b \Vert _2^2\\&= \Vert (h'-K' z )^+\Vert ^2_2 \ge \frac{1}{H(K')^2}\mathbf{dist{}}(z,Z^{\star })^2 \ , \end{aligned}$$

where the last inequality is from duality, Hoffman condition (Equation (26)), and the fact that \(\Vert z \Vert _2 \ge \Vert z \Vert \). Taking the square root obtains the result for \(\eta = 1\).

Next, we consider the case \(\eta \ne 1\). Let us denote the corresponding normalized duality gap as \(\rho _r^{\eta }\) and distance as \(\mathbf{dist{}}^{\eta }\) then with \(\theta = \max \{ \eta , 1/\eta \}\) we get

$$\begin{aligned} \rho _{r}^{\eta }(z)\ge & {} \frac{\rho _{\theta r}^{1}(z)}{\theta } \ge \frac{1}{\theta H(K') \sqrt{4 R^2 + 1}} \textbf{dist}^{1}(z, Z^{\star })\\\ge & {} \frac{1}{\theta ^2 H(K') \sqrt{4 R^2 + 1}} \textbf{dist}^{\eta }(z, Z^{\star }) . \end{aligned}$$

\(\square \)

Proof of Proposition 10

The next lemma is used in the proof of Proposition 10.

Lemma 7

Consider the sequence \(\{ z^{n, 0} \}_{i=0}^{\infty }\), \(\{ \tau ^{n} \}_{i=1}^{\infty }\) generated by Algorithm 1 with the adaptive restart scheme and \(\beta <\frac{1}{q+3}\). Suppose (6) satisfies Property 3, and there exists \(a \in (0,\infty )\) s.t. \(\rho _r(z) \ge \frac{a}{1 + \Vert z \Vert } \mathbf{dist{}}(z, Z^{\star })\). Then \(z^{n,0}\) stays in a bounded region, in particular, \(z^{n,0} \in W_{R}(z^{0,0})\) for all \(n \in {\mathbb {N}}\) with \(R = \frac{2C (q + 3)^3}{\beta a (1 - \beta (q+3))} \mathbf{dist{}}( z^{0, 0}, Z^{\star }) (1 + \Vert z^{0, 0} \Vert + \mathbf{dist{}}(z^{0,0}, Z^{\star }))\).

Proof

First, notice that it holds for any \(n>0\) that \(\mathbf{dist{}}(z^{n, 0}, Z^{\star }) \le \mathbf{dist{}}(z^{n-1, 0}, Z^{\star }) + \Vert z^{n, 0} - z^{n-1,0} \Vert \le (q + 3) \mathbf{dist{}}(z^{n-1, 0}, Z^{\star })\) where the first inequality uses the triangle inequality, and the second inequality is from Property ii.. Recursively applying this inequality yields

$$\begin{aligned} \mathbf{dist{}}(z^{n, 0}, Z^{\star }) \le (q + 3)^n \mathbf{dist{}}(z^{0, 0}, Z^{\star }) \ . \end{aligned}$$
(64)

Thus, we have

$$\begin{aligned} \Vert z^{N, 0} - z^{0, 0} \Vert&\le \sum _{n=0}^{N-1} \Vert z^{n+1, 0} - z^{n,0} \Vert \le (q + 3) \sum _{n=0}^{N-1} \mathbf{dist{}}(z^{n, 0}, Z^{\star }) \nonumber \\&\le \mathbf{dist{}}(z^{0,0}, Z^{\star }) \sum _{n=0}^{N-1} (q+3)^{n+1}\nonumber \\&= \mathbf{dist{}}(z^{0,0}, Z^{\star }) \frac{(q + 3)^{N + 1} - (q+3)}{q+2} \nonumber \\&\le \frac{(q + 3)^{N + 1}}{2} \mathbf{dist{}}(z^{0,0}, Z^{\star }) \ , \end{aligned}$$
(65)

where the first inequality uses the triangle inequality, the second inequality is from Property ii., the third inequality utilizes (64) for \(n=0,\ldots ,N-1\), the equality uses the formula for the sum of a geometric series, and the last inequality uses \(q\ge 0\).

Furthermore, we have

$$\begin{aligned} \rho _{\Vert z^{n, 0} - z^{n-1, 0} \Vert }(z^{n,0})&\le \beta ^{n-1} \rho _{\Vert z^{1, 0} - z^{0, 0} \Vert }( z^{1,0} ) \le \beta ^{n-1} 2 C \Vert z^{1, 0} - z^{0, 0} \Vert \nonumber \\&\le \beta ^{n-1} 2 C (q + 2) \mathbf{dist{}}( z^{0, 0}, Z^{\star }) \ , \end{aligned}$$
(66)

where the first inequality recursively uses (30), the second inequality is from Property i. and \(\tau ^{0} \ge 1\), and the last inequality uses Property ii..

Therefore, it holds that

$$\begin{aligned} \begin{aligned} \Vert z^{n+1,0} - z^{n,0} \Vert&\le (q + 2) \mathbf{dist{}}(z^{n,0}, Z^{\star }) \\&\le (q + 2) \frac{\rho _{\Vert z^{n, 0} - z^{n-1, 0} \Vert }(z^{n,0})}{a} \left( 1 + \Vert z^{n,0} \Vert \right) \\&\le \beta ^{n-1} \frac{2 C (q + 2)^2 \mathbf{dist{}}( z^{0, 0}, Z^{\star })}{a} \left( 1 + \Vert z^{n,0} \Vert \right) \\&\le \beta ^{n-1} \frac{2 C (q + 2)^2 \mathbf{dist{}}( z^{0, 0}, Z^{\star })}{a} \left( 1 + \Vert z^{0, 0} \Vert + \Vert z^{n,0} - z^{0,0} \Vert \right) \\&\le \beta ^{n-1} \frac{2 C (q + 2)^2 \mathbf{dist{}}( z^{0, 0}, Z^{\star })}{a}\\&\qquad \left( 1 + \Vert z^{0, 0} \Vert + \frac{(q + 3)^{n+1}}{2} \mathbf{dist{}}(z^{0,0}, Z^{\star }) \right) \\&\le (\beta (q+3))^{n} \frac{C (q + 3)^3}{\beta a} \mathbf{dist{}}( z^{0, 0}, Z^{\star }) \\&\qquad \left( 1 + \Vert z^{0, 0} \Vert + \mathbf{dist{}}(z^{0,0}, Z^{\star }) \right) \ , \end{aligned} \end{aligned}$$
(67)

where the first inequality uses Property ii., the second inequality is from that \(\rho _r(z) \ge \frac{a}{1 + \Vert z \Vert } \mathbf{dist{}}(z, Z^{\star })\), the third inequality uses (66), the fourth inequality uses the triangle inequality, the fifth inequality uses (65), and the last inequality uses \(q+3 \ge 2\).

Therefore, we have

$$\begin{aligned} \Vert z^{N, 0} - z^{0, 0} \Vert&\le \sum _{n=0}^{N-1} \Vert z^{n+1,0} - z^{n,0} \Vert \\&\le \sum _{n=0}^{N-1} (\beta (q+3))^{n} \frac{C (q + 3)^3}{\beta a} \mathbf{dist{}}( z^{0, 0}, Z^{\star })\\&\qquad \left( 1 + \Vert z^{0, 0} \Vert + \mathbf{dist{}}(z^{0,0}, Z^{\star }) \right) \\&\le \frac{C (q + 3)^3}{\beta a (1 - \beta (q+3))} \mathbf{dist{}}( z^{0, 0}, Z^{\star }) \left( 1 + \Vert z^{0, 0} \Vert + \mathbf{dist{}}(z^{0,0}, Z^{\star }) \right) \ , \end{aligned}$$

where the first inequality is the triangle inequality, the second inequality uses (67), and the last inequality is from the bound on the sum of a geometric series by noticing \(\beta (q+3)<1\). This finishes the proof. \(\square \)

Proof of Proposition 10for EGM and ADMM appled to LP. Recall that the Lagrangian form of an LP instance is \(\alpha \)-sharp on \(S(R)=\{z \in Z: \Vert z \Vert \le R\}\) with \(\alpha =\frac{1}{H(K)\sqrt{1+4R^2}}\) (see Lemma 5) and the ADMM form of a LP instance is \(\alpha \)-sharp on \(S(R)=\{z \in Z: \Vert z \Vert \le R\}\) with \(\alpha =\frac{1}{ \max \{ \eta ^2, 1/\eta ^2 \} H(K')\sqrt{1+4R^2}}\) (see Lemma 6). Thus the sharpness constant \(\alpha \) satisfies the condition in Lemma 7 with \(a=\frac{1}{2\,H(K)}\) for standard form LP (Lemma 5) and and \(a=\frac{1}{2 \max \{ \eta ^2, 1/\eta ^2 \} H(K')}\) for ADMM form LP. We finish the proof by setting \(R = \frac{2C (q + 3)^3}{\beta a (1 - \beta (q+3))} \mathbf{dist{}}( z^{0, 0}, Z^{\star }) (1 + \Vert z^{0, 0} \Vert + \mathbf{dist{}}(z^{0,0}, Z^{\star }))\) as stated in Lemma 7. \(\square \)

Proof of Corollary 2

Proof

Let \(\Vert z \Vert _M = \sqrt{\Vert x \Vert ^2 - 2 \eta y^{\top } A x + \Vert y \Vert ^2}\). Recall that Property 3 holds for \(\Vert \cdot \Vert _M\) with \(q = 0\) and \(C = 2 / \eta \) (refer to Table 1). Define \(\hat{B}_{r}(z) = \{ z \in Z: \Vert z \Vert _{M} \le r \}\) then with \(\hat{r} = r \sqrt{\frac{1}{1 - \eta {\sigma _{\max }}{(A)}}}\) for any \(z \in Z\) we get

$$\begin{aligned} \rho _r(z)&= \frac{\max _{\hat{z}\in W_{r}(z)}\{ \mathcal {L}(x, \hat{y}) - \mathcal {L}(\hat{x}, y)\} }{r} \\&\le \frac{\max _{\hat{z}\in \hat{B}_{\hat{r}}(z)}\{ \mathcal {L}(x, \hat{y}) - \mathcal {L}(\hat{x}, y)\} }{r} \\&= \sqrt{\frac{1}{1 - \eta {\sigma _{\max }}{(A)}}} \frac{\max _{\hat{z}\in \hat{B}_{\hat{r}}(z)}\{ \mathcal {L}(x, \hat{y}) - \mathcal {L}(\hat{x}, y)\} }{\hat{r}} \\&\le \frac{1}{\eta } \sqrt{\frac{1}{1 - \eta {\sigma _{\max }}{(A)}}} \frac{2 \hat{r}}{t} = \frac{1}{\eta (1 - \eta {\sigma _{\max }}{(A)})} \frac{2 r}{t} \end{aligned}$$

where the first inequality uses that \(W_{r}(z) \subseteq \hat{B}_{\hat{r}}(z)\) by Proposition 7, the second inequality uses Property i.. This establishes \(C = \frac{2}{\eta (1 - \eta {\sigma _{\max }}{(A)})}\). Moreover, with \(z^{\star }\in {\mathop {\textrm{argmin}}\limits _{z \in Z^{\star }}} \Vert z^{0} - z \Vert _M\) we have

$$\begin{aligned} \Vert \bar{z}^{t} - z^{0} \Vert ^2 \le \frac{\Vert \bar{z}^{t} - z^{0} \Vert _M^2}{1 - \eta {\sigma _{\max }}{(A)}} \le \frac{4 \Vert z^{\star }- z^{0} \Vert _{M}^2}{1 - \eta {\sigma _{\max }}{(A)}} \le \frac{4 (1 + \eta {\sigma _{\max }}{(A)})}{1 - \eta {\sigma _{\max }}{(A)}} \Vert z^{\star }- z^{0} \Vert ^2 \end{aligned}$$

where the first inequality uses by Proposition 7, the second inequality uses Property ii., and the third inequality uses Proposition 7. This establishes \(q = 4 \frac{1 + \eta {\sigma _{\max }}{(A)}}{1 - \eta {\sigma _{\max }}{(A)}} - 2\). \(\square \)

Linear time algorithm for linear trust region problems

figure b

Theorem 4

Algorithm 2 exactly solves (49), the trust region problem with linear objective, Euclidean norm, and bounds.

Proof

If the algorithm returns from line 1, (51) is unbounded, and the algorithm returns \(\hat{z}(\infty )\). Otherwise, in each iteration of the while loop (line 5), the algorithm maintains the invariants

  • \(\mathcal {I} = \{i:\;\lambda _{\text {lo}}< \hat{\lambda }_i < \lambda _{\text {hi}}\}\)

  • For \(\lambda _{\text {lo}}< \lambda < \lambda _{\text {hi}}\), \(\Vert \hat{z}(\lambda ) - z \Vert ^2 = f_{\text {lo}}+ \lambda ^2 f_{\text {hi}}+ \sum _{i \in \mathcal {I}} (\hat{z}(\lambda )_i - z_i)^2\)

  • \(\Vert \hat{z}(\lambda _{\text {lo}}) - z\Vert ^2 \le r^2 \le \Vert \hat{z}(\lambda _{\text {hi}}) - z\Vert ^2\)

The while loop (line 5) is finite, since on each iteration \(|\mathcal {I}|\) is reduced by at least a factor of 2. When the while loop exits (with \(\mathcal {I} = \emptyset \)), these invariants mean that the final \(\lambda _{\text {mid}}\) computed on line 18 optimizes (51) and the returned \(\hat{z}(\lambda _{\text {mid}})\) solves (49).    \(\square \)

Theorem 5

Algorithm 2 runs in \(O(m+n)\) time.

Proof

The work outside the while loop (line 5) is clearly \(O(m+n)\). In each pass through the while loop, the median (line 6) can be found in \(O(|\mathcal {I}|)\) time ([9]), and the rest of the loop also takes \(O(|\mathcal {I}|)\) time. Since at least half of the elements of \(\mathcal {I}\) are removed in each iteration, and initially \(|\mathcal {I}| \le m+n\), the total time in the while loop is at most \(O(\sum _{i=0}^\infty \frac{m+n}{2^i}) = O(m+n)\) by the formula for the sum of an infinite geometric series. \(\square \)

Numerical results for restarted EGM and ADMM

In this section, we repeat the numerical experiments of Sect. 7 with EGM and ADMM instead of PDHG. We tune the primal weight for EGM and the step size \(\eta \) for ADMM using the same procedure as with PDHG (see Sect. 7).

Fig. 3
figure 3

Plots show normalized duality gap (in \(\log \) scale) versus number of iterations for (restarted) EGM. Results from left to right, top to bottom for qap10, qap15, nug08-3rd, and nug20. Each plot compares no restarts, adaptive restarts, and the three best fixed restart lengths. For the restart schemes we evaluate the normalized duality gap at the average, for no restarts we evaluate it at the last iterate as the average performs much worse. A star indicates the last iteration that the active set changed before the iteration limit was reached. Series markers indicate when restarts occur

Figure 3 (EGM) is almost identical to Fig. 2 (PDHG), which verifies again the improved performance of restarted algorithms. Furthermore, note that one iteration of EGM is about twice as expensive as a PDHG iteration, because EGM requires four matrix–vector multiplications per iteration, while PDHG requires two matrix–vector multiplications.

Indeed, to see the difference between EGM and PDHG requires a close examination of Table 4. Recall that termination criteria is only checked every 30 iterations, which causes the iteration counts to be identical in several instances.

Fig. 4
figure 4

Plots show normalized duality gap (in \(\log \) scale) versus number of iterations for (restarted) ADMM. Results from left to right, top to bottom for qap10, qap15, nug08-3rd, and nug20. Each plot compares no restarts, adaptive restarts, and the three best fixed restart lengths. For the restart schemes we evaluate the normalized duality gap at the average, for no restarts we evaluate it at the last iterate as the average performs much worse. A star indicates the last iteration that the active set changed before the iteration limit was reached. Series markers indicate when restarts occur

Figure 4 shows the performance of restarted ADMM for the same problems, which verifies again the improved performance of restarted algorithms. We observe that restarts do appear to make performance slower for nug08-3rd but this appears to be a very easy problem—the number of iterations for all methods is very low.

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Applegate, D., Hinder, O., Lu, H. et al. Faster first-order primal-dual methods for linear programming using restarts and sharpness. Math. Program. 201, 133–184 (2023). https://doi.org/10.1007/s10107-022-01901-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10107-022-01901-9

Mathematics Subject Classification

Navigation