Skip to main content
Log in

Fast Continuous Dynamics Inside the Graph of Subdifferentials of Nonsmooth Convex Functions

  • Published:
Applied Mathematics & Optimization Submit manuscript

Abstract

In a Hilbert framework, we introduce a new class of second-order dynamical systems that combine viscous and geometric damping but also a time rescaling process for nonsmooth convex minimization. A main feature of these systems is to produce trajectories that lie in the graph of the Fenchel subdifferential of the objective. Moreover, they do not incorporate any regularization or smoothing processes. This new class originates from some combination of a continuous Nesterov-like dynamic and the Minty representation of subdifferentials. These models are investigated through first-order reformulations that amount to dynamics involving three variables: two solution trajectories (including an auxiliary one) and another one associated with subgradients. We prove the weak convergence towards equilibria for the solution trajectories, as well as properties of fast convergence to zero for their velocities. Remarkable convergence rates (possibly of exponential-type) are also established for the function values. We additionally state notable properties of fast convergence to zero for the subgradients trajectory and for its velocity. Some numerical experiments are performed so as to illustrate the efficiency of our approach. The proposed models offer a new and well-adapted framework for discrete counterparts, especially for structured minimization problems. Inertial algorithms with a correction term are then suggested relative to this latter context.

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
Fig. 10
Fig. 11

Similar content being viewed by others

Data Availibility

We do not analyse or generate any datasets, because our work proceeds within a theoretical and mathematical approach. One can obtain the relevant materials from the references below.

References

  1. Abbas, B., Attouch, H.: Dynamical systems and forward-backward algorithms associated with the sum of a convex subdifferential and a monotone cocoercive operator. Optimization 64, 2223–2252 (2015)

    Article  MathSciNet  Google Scholar 

  2. Abbas, B., Attouch, H., Svaiter, B.F.: Newton-like dynamics and forward-backward methods for structured monotone inclusions in Hilbert spaces. J. Optim. Theory Appl. 161(2), 331–360 (2014)

    Article  MathSciNet  Google Scholar 

  3. Alvarez, F., Attouch, H., Bolte, J., Redont, P.: A second-order gradient-like dissipative dynamical system with Hessian driven damping. Application to optimization and mechanics. J. Math. Pures Appl. 81(8), 747–779 (2002)

    Article  MathSciNet  Google Scholar 

  4. Apidopoulos, V., Aujol, J.-F., Dossal, Ch.: The differential inclusion modeling the FISTA algorithm and optimality of convergence rate in the case \(b \le 3\). SIAM J. Optim. 28(1), 551–574 (2018)

    Article  MathSciNet  Google Scholar 

  5. Attouch, H., Cabot, A.: Convergence of damped inertial dynamics governed by regularized maximally monotone operators. J. Differ. Equ. 264, 7138–7182 (2018)

    Article  MathSciNet  Google Scholar 

  6. Attouch, H., Cabot, A.: Convergence of a relaxed inertial forward-backward algorithm for structured monotone inclusions. Appl. Math. Optim. 80, 547–598 (2019)

    Article  MathSciNet  Google Scholar 

  7. Attouch, H., László, S.C.: Continuous Newton-like inertial dynamics for monotone inclusions. Set Valued Var. Anal. 29, 555–581 (2021)

    Article  MathSciNet  Google Scholar 

  8. Attouch, H., Peypouquet, J.: Convergence of inertial dynamics and proximal algorithms governed by maximally monotone operators. Math. Program. 174, 391–432 (2019). https://doi.org/10.1007/s10107-018-1252-x

    Article  MathSciNet  Google Scholar 

  9. Attouch, H., Svaiter, B.F.: A continuous dynamical Newton-like approach to solving monotone inclusions. SIAM J. Control Optim. 49(2), 574–598 (2011)

    Article  MathSciNet  Google Scholar 

  10. Attouch, H., Bolte, J., Redont, P.: Optimizing properties of an inertial dynamical system with geometric damping: Link with proximal methods. Control Cybern. 31, 643–657 (2002)

    MathSciNet  Google Scholar 

  11. Attouch, H., Peypouquet, J., Redont, P.: Fast convex minimization via inertial dynamics with Hessian driven damping. J. Differ. Equ. 261, 5734–5783 (2016)

    Article  Google Scholar 

  12. Attouch, H., Chbani, Z., Peypouquet, J., Redont, P.: Fast convergence of inertial dynamics and algorithms with asymptotic vanishing viscosity. Math. Program. 168(1–2), 123–175 (2018)

    Article  MathSciNet  Google Scholar 

  13. Attouch, H., Chbani, Z., Riahi, H.: Fast proximal methods via time scaling of damped inertial gradient dynamics. SIAM J. Optim. 29(3), 2227–2256 (2019)

    Article  MathSciNet  Google Scholar 

  14. Attouch, H., Chbani, Z., Riahi, H.: Fast convex optimization via time scaling of damped inertial gradients dynamics. Pure Appl. Funct. Anal. 6(6), 1081–1117 (2021)

    MathSciNet  Google Scholar 

  15. Attouch, H., Balhag, A., Chbani, Z., Riahi, H.: Fast convex optimization via inertial combining viscous and Hessian-driven damping with time rescaling dynamics. Evol. Equ. Control Theory 11(2), 487–514 (2022)

    Article  MathSciNet  Google Scholar 

  16. Attouch, H., Chbani, Z., Fadili, J., Riahi, H.: First-order optimization algorithms via inertial systems with Hessian driven damping. Math. Program. 193, 113–155 (2022). https://doi.org/10.1007/s10107-020-01591-1

    Article  MathSciNet  Google Scholar 

  17. Attouch, H., Chbani, Z., Fadili, J., Riahi, H.: Convergence of iterates for first-order optimization algorithms with inertia and Hessian driven damping. Optimization 72(5), 1199–1238 (2023). https://doi.org/10.1080/02331934.2021.2009828

    Article  MathSciNet  Google Scholar 

  18. Boţ, R.I., Hulett, D.A.: Second order splitting dynamics with vanishing damping for additively structured monotone inclusions. J. Dyn. Differ. Equ. (2022). https://doi.org/10.1007/s10884-022-10160-3

    Article  Google Scholar 

  19. Boţ, R.I., Karapetyants, M.A.: A fast continuous time approach with time scaling for nonsmooth convex optimization. Adv. Cont. Disc. Models (2022). https://doi.org/10.1186/s13662-022-03744-2

    Article  MathSciNet  Google Scholar 

  20. Boţ, R.I., Csetnek, E., László, S.C.: On the strong convergence of continuous Newton-like inertial dynamics with Tikhonov regularization for monotone inclusions. J. Math. Anal. Appl. (2023). https://doi.org/10.13140/RG.2.2.20539.18729

    Article  Google Scholar 

  21. Brezis, H.: Opérateurs maximaux monotones et semi-groupes de contractions dans les espaces de Hilbert. Math. Stud., vol. 5. North-Holland, Amsterdam (1973)

    Google Scholar 

  22. Brezis, H.: Function Analysis, Sobolev Spaces and Partial Differential Equations. Springer, New York (2010)

    Google Scholar 

  23. Cabot, A., Engler, H., Gadat, S.: On the long time behavior of second order differential equations with asymptotically small dissipation. Trans. Am. Math. Soc. 361, 5983–6017 (2009)

    Article  MathSciNet  Google Scholar 

  24. Cabot, A., Engler, H., Gadat, S.: Second order differential equations with asymptotically small dissipation and piecewise flat potentials. Electron. J. Differ. Equ. 17, 33–38 (2009)

    MathSciNet  Google Scholar 

  25. Haraux, A.: Systémes dynamiques dissipatifs et applications, RMA17. Masson, Paris (1991)

    Google Scholar 

  26. Kim, D.: Accelerated proximal point method for maximally monotone operators (2019). Math. Program. 190, 57–87 (2021)

    Article  MathSciNet  Google Scholar 

  27. Labarre, F., Maingé, P.E.: First-order frameworks for continuous Newton-like dynamics governed by maximally monotone operators. Set Valued Var. Anal. 20(2), 425–451 (2022). https://doi.org/10.1007/s11228-021-00593-1

    Article  MathSciNet  Google Scholar 

  28. Luo, H.: Accelerated differential inclusion for convex optimization. Optimization 72(5), 1139–1170 (2023). https://doi.org/10.1080/02331934.2021.2002327

    Article  MathSciNet  Google Scholar 

  29. Maingé, P.E.: Accelerated proximal algorithms with a correction term for monotone inclusions. Appl. Math. Optim. 84(Suppl 2), 2027–2061 (2021)

    Article  MathSciNet  Google Scholar 

  30. Maingé, P.E., Weng-Law, A.: Fast continuous dynamics inside the graph of maximally monotone operators. Set Valued Var. Anal. (2023). https://doi.org/10.1007/s11228-023-00663-6

    Article  MathSciNet  Google Scholar 

  31. May, R.: Asymptotic for a second order evolution equation with convex potential and vanishing damping term. Turk. J. Math. 41(3), 681–785 (2015). https://doi.org/10.3906/mat-1512-28

    Article  MathSciNet  Google Scholar 

  32. Minty, G.J.: Monotone (nonlinear) operators in Hilbert spaces. Duke Math. J. 29, 341–346 (1962)

    Article  MathSciNet  Google Scholar 

  33. Opial, Z.: Weak convergence of the sequence of successive approximations for nonexpansive mappings. Bull. Am. Math. Soc. 73, 591–597 (1967)

    Article  MathSciNet  Google Scholar 

  34. Polyak, B.T.: Some methods of speeding up the convergence of iterative methods. USSR Comput. Math. Math. Phys. 4(5), 1–17 (1964)

    Article  MathSciNet  Google Scholar 

  35. Qu, X., Bian, W.: Fast inertial dynamic algorithm with smoothing method for nonsmooth convex optimization. Comput. Optim. Appl. 83, 287–317 (2022)

    Article  MathSciNet  Google Scholar 

  36. Shi, B., Du, S.S., Jordan, M.I., Su, W.J.: Understanding the acceleration phenomenon via high-resolution differential equations. Math. Program. 195, 79–148 (2022). https://doi.org/10.1007/s10107-021-01681-8

    Article  MathSciNet  Google Scholar 

  37. Sontaq, E.D.: Mathematical Control Theory, 2nd edn. Springer, New York (1998)

    Book  Google Scholar 

  38. Su, W., Boyd, S., Candés, E.J.: A differential equation for modeling Nesterov’s accelerated gradient method: theory and insights. Neural Inf. Process. Syst. 27, 2510–2518 (2014)

    Google Scholar 

Download references

Funding

The authors declare that no funds, grants, or other support were received during the preparation of this manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Paul-Emile Maingé.

Ethics declarations

Conflict of interest

The authors have no competing interests to declare that are relevant to the content of this article.

Additional information

Publisher's Note

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

Appendix

Appendix

1.1 Proof of Proposition 2.1

Let us prove that (i1) \(\Rightarrow \) (i2). Consider a solution \((x,\xi ) \in {\mathcal {A}}_c\times {\mathcal {A}}_c\) to (2.2). Set \(\zeta ( \cdot )=\sigma ( \cdot ) \xi ( \cdot )\) and suppose that \(x( \cdot )+ \zeta ( \cdot )\) is of class \(C^1\) and that \(\left( x( \cdot )+ \zeta ( \cdot ) \right) ^{(1)} \in {\mathcal {A}}_c\). Clearly, for \(t \ge 0\), as \(\dot{x}\), \({\dot{\zeta }}\) and \(\zeta \) are integrable on [0, t] (since x and \(\zeta \) belong to \({\mathcal {A}}_c\)), we can set as a well-defined quantity

$$\begin{aligned} z(t)=\int _0^t \left( \alpha (r)\dot{x}(r)+\beta (r) {\dot{\zeta }}(r)+b(r) \zeta (r)\right) dr- q_0. \end{aligned}$$
(1.9)

Hence, \(z \in {\mathcal {A}}_c\), and by differentiating (1.9) we obtain

$$\begin{aligned} {\dot{z}(t)=\alpha (t)\dot{x}(t)+\beta (t) {\dot{\zeta }}(t)+b(t) \zeta (t), \text {for a.e. } t \ge 0}. \end{aligned}$$
(1.10)

Therefore, by (2.2b) together with the above equality, we get

$$\begin{aligned} {\big (x( \cdot ) + \zeta ( \cdot ) \big )^{(2)}(t) + \dot{z}(t)=0, \text {for a.e. } t\ge 0}. \end{aligned}$$
(1.11)

Moreover, recalling that \(\big (x( \cdot ) + \zeta ( \cdot ) \big )^{(1)} \in {\mathcal {A}}_c\) and \(z \in {\mathcal {A}}_c\), we get

\(\frac{d}{dt} \left( \big (x( \cdot ) + \zeta ( \cdot ) \big )^{(1)} + z( \cdot ) \right) (t) = \big (x( \cdot ) + \zeta ( \cdot ) \big )^{(2)}(t)+ \dot{z}(t)\), for a.e. \(t \ge 0\). Hence, we straightforwardly deduce

$$\begin{aligned} {\frac{d}{dt}\left( \big (x( \cdot ) + \zeta ( \cdot ) \big )^{(1)}(t) + z( \cdot ) \right) (t)=0, \text {for a.e. } t \ge 0}. \end{aligned}$$
(1.12)

It follows immediately that

$$\begin{aligned} {\big ( x( \cdot )+ \zeta ( \cdot ) \big )^{(1)}(t) + z(t) = \big ( x( \cdot )+ \zeta ( \cdot ) \big )^{(1)}(0) + z(0) \text { for } t \ge 0}, \end{aligned}$$
(1.13)

which, by the initial condition \(\big ( x( \cdot ) + \sigma ( \cdot )\xi ( \cdot )\big )^{(1)}(0) = q_0 \), yields

$$\begin{aligned} {\big ( x( \cdot )+ \zeta ( \cdot ) \big )^{(1)}(t) + z(t) = 0 \,\text {for}\, t \ge 0}, \end{aligned}$$
(1.14)

which readily implies that

$$\begin{aligned} {\dot{x}(t)+ {\dot{\zeta }}(t) + z(t) = 0\,\text {for a.e}\,t \ge 0}. \end{aligned}$$
(1.15)

Multiplying (1.15) by \(\beta (t)\) and adding the resulting equality to (1.10) give us

$$\begin{aligned} {\big (\beta (t)-\alpha (t)\big )\dot{x}(t) +\dot{z}(t)+\beta (t)z(t)-b(t)\zeta (t) = 0\text { for a.e }t \ge 0}. \end{aligned}$$
(1.16)

Now, since \(\theta ( \cdot )\) is positive, we introduce the function \(y( \cdot )\) defined for \(t \ge 0\) by

$$\begin{aligned} y(t):=\frac{1}{\theta (t)}z(t)+x(t)-\frac{\omega (t)}{\theta (t)}\zeta (t). \end{aligned}$$
(1.17)

For simplification we also set \(u(t)=y(t)-x(t)\). Observe from (1.17) that we equivalently have

$$\begin{aligned} {z(t)= \theta (t) u(t) +\omega (t) \zeta (t) }. \end{aligned}$$
(1.18)

Thus, for \(t \ge 0\), (1.14) in light of the above equality entails

$$\begin{aligned} {\big (x( \cdot )+ \zeta ( \cdot )\big )^{(1)}(t) + \theta (t) u(t) + \omega (t) \zeta (t) =0,} \end{aligned}$$
(1.19)

that is (2.4b). We now prove (2.4c). Differentiating (1.18), while noticing that \(\{x, \xi , u\} \subset {\mathcal {A}}_c\), readily implies, for a.e. \(t\ge 0\),

$$\begin{aligned} {\dot{z}(t)= \dot{\theta }(t) u(t) + \theta (t) \dot{u}(t) + {\dot{\omega }} (t) \zeta (t) + \omega (t) {\dot{\zeta }} (t)}. \end{aligned}$$
(1.20)

Moreover, using the definitions of \(\alpha ( \cdot )\), \(\beta ( \cdot )\) and \(b( \cdot )\) given by (1.3), namely \(\alpha (t) =-\frac{\dot{\theta }(t) }{\theta (t)} + \kappa -\theta (t)\), \(\beta (t) = -\frac{{\dot{\theta }}(t) }{\theta (t)} + \kappa +\omega (t) \) and \(b(t)= \omega (t) \left( \kappa +\frac{{\dot{\omega }} (t)}{\omega (t)}-\frac{\dot{\theta }(t)}{\theta (t)} \right) \), yields

$$\begin{aligned}{} & {} {\beta (t)-\alpha (t)-\theta (t)} = {\omega (t)}, \end{aligned}$$
(1.21a)
$$\begin{aligned}{} & {} {\beta (t) \theta (t) + \dot{\theta }(t)} = {\kappa \theta (t) + \omega (t) \theta (t)}, \end{aligned}$$
(1.21b)
$$\begin{aligned}{} & {} {\beta (t) + \frac{\dot{\omega }(t)}{\omega (t)}-\frac{b(t)}{\omega (t)}} = {\omega (t)}. \end{aligned}$$
(1.21c)

Hence, by (1.16) and using (1.20), (1.18), \(\dot{u}=\dot{y} - \dot{x}\) and (1.21a), successively, we get, for a.e. \(t\ge 0\),

$$\begin{aligned} \begin{array}{l} 0=\big (\beta (t)-\alpha (t)\big )\dot{x}(t) +\dot{z}(t)+\beta (t) z(t)-b(t) \zeta (t) \\ \quad =\big (\beta (t)-\alpha (t)\big )\dot{x}(t) + \dot{\theta }(t) u(t) + \theta (t) \dot{u}(t) + {\dot{\omega }} (t) \zeta (t) + \omega (t) {\dot{\zeta }} (t)\\ \qquad +\beta (t) \theta (t) u(t) + \beta (t) \omega (t) \zeta (t) -b(t) \zeta (t) \\ \quad =\big ( \beta (t)-\alpha (t)-\theta (t) \big ) \dot{x}(t) + \theta (t)\dot{y}(t) + \big (\beta (t)\theta (t)+\dot{\theta }(t) \big ) u(t) \\ \qquad + \omega (t) \big ( {\dot{\zeta }}(t) + \left( \frac{\dot{\omega }(t)}{\omega (t)} + \beta (t) - \frac{b(t) }{\omega (t)} \right) \zeta (t) \big ), \end{array} \end{aligned}$$
(1.22)

namely

$$\begin{aligned}\begin{array}{l} 0= \omega (t) \dot{x}(t) + \theta (t)\dot{y}(t) + \big ( \kappa \theta (t) + \omega (t) \theta (t) \big ) u(t) + \omega (t) \big ( {\dot{\zeta }}(t) + \omega (t) \zeta (t) \big ) \\ \quad = \omega (t) \big (\dot{x}(t) + {\dot{\zeta }} (t) +\theta (t) u(t) + \omega (t)\zeta (t) \big ) + \theta (t) \big ( \dot{y}(t) + \kappa u(t) \big ). \end{array} \end{aligned}$$

In addition, by (1.19), while recalling that \(\{x,\xi \} \subset {\mathcal {A}}_c\), we obtain

$$\begin{aligned} {\dot{x}(t)+ {\dot{\zeta }} (t) + \theta (t) u(t) + \omega (t) \zeta (t) =0, \text{ for } \text{ a.e. } t\ge 0}. \end{aligned}$$
(1.23)

Thus, combining (1.22) and (1.23), in light of \(\theta \ne 0\), yields

\(\dot{y}(t)+\kappa (u(t))=0\) for a.e. \(t\ge 0\),

that is (2.4c).

Finally, regarding the initial conditions, we have \((x(0),\xi (0))= (x_0,\xi _0)\) and \(\big (x( \cdot ) + \zeta ( \cdot )\big )(0)= q_0\) (according to (i1)) while (2.4b) at time \(t=0\) ensures that

\(\big ( x( \cdot ) + \zeta ( \cdot ) \big )^{(1)} (0) + \theta (0) \left( y(0)-x(0) \right) + \omega (0) \zeta (0) =0\).

Hence, we deduce that \(y(0) =x_0-\frac{1}{\theta (0)} \left( q_0 + \sigma (0) \omega (0) \xi _0 \right) \).

Let us prove that (i2) \(\Rightarrow \) (i1). Consider a solution \((x,\xi ,y) \in {\mathcal {A}}_c\times {\mathcal {A}}_c\times C^1\) to (2.4). For simplification, we set again \(u( \cdot )=y( \cdot )-x( \cdot )\) and \(\zeta ( \cdot )=\sigma ( \cdot ) \xi ( \cdot )\). Clearly, by (2.4b), we have, for \(t \ge 0\),

$$\begin{aligned} \big ( x( \cdot ) + \zeta ( \cdot ) \big )^{(1)} (t) + \theta (t) u(t) + \omega (t) \zeta (t)=0. \end{aligned}$$
(1.24)

This, by \((x,\xi ,y) \in {\mathcal {A}}_c\times {\mathcal {A}}_c\times C^1\) and by \(\{\omega ( \cdot )\), \(\theta ( \cdot )\), \(\sigma ( \cdot )\} \subset C^1([0, \infty ])\), entails that \(x( \cdot )+\zeta ( \cdot )\) is of class \(C^1\) and that \(\big ( x( \cdot ) + \sigma ( \cdot )\xi ( \cdot ) \big )^{(1)} \in {\mathcal {A}}_c\). Then, differentiating (1.24) gives us, for a.e \(t \ge 0\),

$$\begin{aligned} {(x( \cdot ) + \zeta ( \cdot ))^{(2)}(t) + {\dot{\theta }}(t) u(t) + \theta (t) \dot{u}(t) + \omega (t) {\dot{\zeta }}(t) + {\dot{\omega }}(t) \zeta (t) = 0}, \end{aligned}$$
(1.25)

while we know from (2.4c) that \(\dot{y}(t)= -\kappa u(t) \). Consequently, we readily obtain, for a.e \(t \ge 0\),

$$\begin{aligned} (x( \cdot ) + \zeta ( \cdot ))^{(2)}(t) + {\dot{\theta }}(t) u(t) + \theta (t)\big ( -\kappa u(t) - \dot{x}(t) \big ) + \omega (t) \dot{\zeta }(t) + {\dot{\omega }}(t) \zeta (t) = 0.\nonumber \\ \end{aligned}$$
(1.26)

Furthermore, for a.e. \(t \ge 0\), by (1.24) we readily have

\(u(t)=-\frac{1}{\theta (t)} \big ( ( x( \cdot )+ \zeta ( \cdot )) ^{(1)} + \omega (t) \zeta (t) \big )\), which, by (1.26), entails

$$\begin{aligned} \begin{array}{l} (x( \cdot ) + \zeta ( \cdot ))^{(2)}(t) \\ = \frac{{\dot{\theta }}(t)}{\theta (t)} \big (\dot{x}(t)+ {\dot{\zeta }}(t) + \omega (t) \zeta (t)\big ) - \theta (t) \left( \frac{\kappa }{\theta (t)}\big (\dot{x}(t)+ {\dot{\zeta }}(t) + \omega (t) \zeta (t) \big )-\dot{x}(t) \right) \\ \quad - \omega (t) {\dot{\zeta }}(t) - \dot{\omega }(t) \zeta (t) \\ =-\big (\kappa -\theta (t)-\frac{\dot{\theta }(t)}{\theta (t)} \big )\dot{x}(t) - \left( \kappa + \omega (t)-\frac{\dot{\theta }(t)}{\theta (t)} \right) \dot{\zeta }(t) - \omega (t) \left( \kappa + \frac{{\dot{\omega }}(t)}{\omega (t)} - \frac{{\dot{\theta }}(t)}{\theta (t)}\right) \zeta (t). \end{array} \end{aligned}$$

Hence the expressions of \(\alpha ( \cdot ), \beta ( \cdot )\) and \(b( \cdot )\) defined in (1.3) amounts to (2.2b). In addition, from the initial conditions in (i2), we have \(~x(0)=x_0\), \(~\xi (0)=\xi _0\) and \(y(0) =x_0-\frac{1}{\theta (0)} \left( q_0 + \sigma (0) \omega (0) \xi _0 \right) \). This implies that \(y(0) =x(0)-\frac{1}{\theta (0)} \left( q_0 + \sigma (0) \omega (0) \xi (0) \right) \), while (2.4b) at time \(t=0\) yields

\(\big ( x( \cdot ) + \zeta ( \cdot ) \big )^{(1)} (0) + \theta (0) \left( y(0)-x(0) \right) + \omega (0) \zeta (0) =0\).

Hence, regarding the last two equalities, substituting the former in the latter gives us \(\big ( x( \cdot ) + \zeta ( \cdot )\big )^{(1)}(0) = q_0 \).

1.2 Proof of Proposition 2.2

1.2.1 The Yosida Regularization

Some useful properties of the Yosida regularization are recalled through the lemma below established in [27] (see also [8, 21, 22]).

Lemma 1.1

Let \(A: {\mathcal {H}}\rightrightarrows {\mathcal {H}}\) be a maximally monotone operator such that \(S:=A^{-1}(0) \ne \emptyset \). Let \(\gamma ,\delta > 0\) and \(x, y \in {\mathcal {H}}\). Then for \(z \in A^{-1}(\{0\})\), we have

$$\begin{aligned} {\Vert \gamma A_{\gamma }x-\delta A_{\delta }y \Vert \le 2\Vert x-y\Vert + 2\frac{|\gamma - \delta |}{\gamma }\Vert x-z\Vert }, \end{aligned}$$
(1.27)
$$\begin{aligned} {\Vert A_{\gamma }x- A_{\delta }y \Vert \le \left( 3\frac{| \delta -\gamma |}{\delta \gamma } \times \Vert x-z\Vert + \frac{2}{\delta }\Vert x-y\Vert \right) }. \end{aligned}$$
(1.28)

Proof

The proof of (1.27) can be found in [8]. For proving (1.28), we simply observe that \(\begin{array}{l} A_{\gamma }x- A_{\delta }y= \frac{1}{ \delta } \left( \delta A_{\gamma }x- \delta A_{\delta }y\right) = \frac{1}{ \delta } \left( ( \delta -\gamma ) A_{\gamma }x + ( \gamma A_{\gamma }x -\delta A_{\delta }y\right) , \end{array}\) from which we get \(\begin{array}{l} \Vert A_{\gamma }x- A_{\delta }y \Vert \le \frac{1}{ \delta } \left( | \delta -\gamma | \times \Vert A_{\gamma }x \Vert + \Vert \gamma A_{\gamma }x -\delta A_{\delta }y\Vert \right) . \end{array}\)

Consequently, by \( \Vert A_{\gamma }x\Vert \le \frac{1}{\gamma } \Vert x-z\Vert \) and using (1.27), we obtain (1.28) \(\square \)

1.2.2 Main Proof of the Proposition

The proof follows the same lines as in [30] (see, also, [1, 2]), but it is developed through the following steps (s1)- (s3) with full details:

(s1) We begin by reformulating the (possibly) existing strong global solutions to (1.4)) (that are supposed to satisfy (2.3)) by means of the Minty representation of the maximal monotone operator \(\partial f\) (see [32]). Set \(J_{ \sigma ( \cdot )}^{ \partial f}=\big (I+\sigma ( \cdot ) \partial f\big )^{-1}\) and \((\partial f)_{\sigma ( \cdot ) }= \frac{1}{\sigma ( \cdot )} \big (I-J_{\sigma ( \cdot ) } ^{ \partial f}\big )\), namely the resolvent and the Yosida approximation of \(\partial f\) (with index \(\sigma ( \cdot )\)), respectively, which are well-known to be single-valued and everywhere defined. Associated with any strong global solution \((x( \cdot ),\xi ( \cdot ),y( \cdot ))\) to (1.4), we introduce the new unknown function

$$\begin{aligned} v( \cdot )=x( \cdot )+ \sigma ( \cdot ) \xi ( \cdot ). \end{aligned}$$
(1.29)

It is readily seen that \(v( \cdot )\) belongs to \({\mathcal {A}}_c\) (the set of absolutely continuous functions) and that

\(v(0)= x_0+ \sigma (0) \xi _0\).

Moreover, for \(t \ge 0\), by \(\xi (t) \in \partial f (x(t))\) we obtain \(v(t) \in x(t)+ \sigma (t) \partial f(x(t))\) and \( \xi (t)=\frac{1}{\sigma (t)}(v(t)-x(t))\), hence, by Minty’s representation we simply have

$$\begin{aligned} {x(t)=J_{\sigma (t)}^{ \partial f} v(t) \text{ and } \xi (t)= \frac{1}{\sigma (t)} \big (v(t)-J_{\sigma (t)}^{ \partial f}v(t)\big )=(\partial f)_{\sigma (t)} v(t)}. \end{aligned}$$
(1.30)

Differentiating (1.29), in light of (2.3b), gives us, for a.e. \(t \ge 0\),

\(\dot{v}(t)=\dot{x}(t)+ (\sigma ( \cdot ) \xi ( \cdot ))^{(1)}(t) = -\theta (t) (y(t)-x(t))-\omega (t) \sigma (t) \xi (t)\), hence, by (1.30), we obtain

$$\begin{aligned} \dot{v}(t)+\theta (t) \big (y(t)-J_{\sigma (t)}^{ \partial f} v(t)\big )+\omega (t) \sigma (t) (\partial f )_{\sigma (t)} v(t) =0. \end{aligned}$$

Hence, from (2.3), we deduce that \((v( \cdot ),y( \cdot ))\) are implicitly given, for a.e. \(t \in [0,\infty )\), by

$$\begin{aligned}{} & {} {\dot{v}(t) + \theta (t) \left( y(t)-J_{\sigma (t)}^{ \partial f}v(t) \right) + \omega (t) \sigma (t) (\partial f)_{\sigma (t)} v(t)=0}, \end{aligned}$$
(1.31a)
$$\begin{aligned}{} & {} \dot{y}(t)+ \kappa (y(t)- J_{\sigma (t)}^{\partial f }v(t))=0, \end{aligned}$$
(1.31b)

together with \(y(0)=y_0\) and \(v(0)=x_0+ \sigma (0) \xi _0\).

This shows us that any strong global solution \((x( \cdot ),\xi ( \cdot ),y( \cdot ))\) to (1.4) is entirely determined (thanks to the two formulas in (1.30)) by some (strong) solution \((v( \cdot ),y( \cdot ))\) to (1.31). So, for proving existence and uniqueness of a strong global solution to (1.4), we just state (as argued below) the existence and uniqueness of a (strong) global solution \((v( \cdot ),y( \cdot ))\) to (1.31), but also the existence of a strong global solution \((x( \cdot ),\xi ( \cdot ),y( \cdot ))\) to (1.4).

(s2) Existence, uniqueness and regularity of a (strong) global solution \((v( \cdot ),y( \cdot ))\) to (1.31). First, we show that (1.31) is relevant to the Cauchy–Lipschitz theorem. Indeed, (1.31) can be expressed as

$$\begin{aligned} \dot{U}(t)=F(t,U(t)), \end{aligned}$$
(1.32)

where \(U( \cdot )=(v( \cdot ),y( \cdot ))\) and \(F(t,.): {\mathcal {H}}^2 \rightarrow {\mathcal {H}}^2\) is defined for any \(t \ge 0\) and \( ({\bar{v}}, {\bar{y}}) \in {\mathcal {H}}^2\) by \(F(t,({\bar{v}},{\bar{y}}))= \big ( \phi _1(t,({\bar{v}},{\bar{y}})), \phi _2(t,({\bar{v}},{\bar{y}})) \big )\), together with

$$\begin{aligned}{} & {} { \phi _1(t,({\bar{v}},{\bar{y}})) = -\theta (t) \left( \bar{y}-J_{\sigma (t)}^{\partial f }{\bar{v}} \right) -\omega (t) \sigma (t) (\partial f)_{\sigma (t)}{\bar{v}}}, \end{aligned}$$
(1.33a)
$$\begin{aligned}{} & {} { \phi _2(t,({\bar{v}},{\bar{y}})) = -\kappa \left( {\bar{y}}- J_{\sigma (t)}^{\partial f }{\bar{v}} \right) .} \end{aligned}$$
(1.33b)

In view of applying the global Cauchy–Lipschitz theorem, we establish two main properties on F(., .) through the following items (a) and (b):

(a) Given \(({\bar{v}},{\bar{y}}) \in {\mathcal {H}}^2\), we prove that \(F(., (\bar{v},{\bar{y}}))\) is continuous on \([0,\infty )\). Indeed, let \(z \in (\partial f)^{-1}(0)\) and \((t_1,t_2) \in [0,\infty )^2\). By Lemma 1.1 with \(A= \partial f\) and \(\sigma ( \cdot ) \ge \sigma _0 >0\) (from (1.5b)), we obtain \( \Vert (\partial f)_{\sigma (t_1) }{\bar{y}}- (\partial f)_{\sigma (t_2) }{\bar{y}} \Vert \le 3 \frac{ | \sigma (t_1) - \sigma (t_2) |}{\sigma _0^2}\Vert {\bar{y}}-z\Vert \).

Then, the continuity of \(\sigma ( \cdot )\) on \([0,\infty )\) yields that the mappings \(t \rightarrow (\partial f)_{\sigma (t) }{\bar{y}}\) and \(t \rightarrow J_{\sigma (t)}^{\partial f}{\bar{v}} \) (given by \(J_{\sigma (t)}^{\partial f}{\bar{v}}:={\bar{v}}-\sigma (t) (\partial f)_{\sigma (t)}{\bar{v}}\)) are also continuous on \([0,\infty )\). So, in light of the definition of \(\phi _1(.,.)\) and \(\phi _2(.,.)\), together with the continuity of \(\{ \theta ( \cdot ), \omega ( \cdot ) \}\), we infer that \(F(.,(\bar{v},{\bar{y}}))\) is continuous on \([0,\infty )\) (as are \(\phi _1(.,.)\) and \(\phi _2(.,.)\)).

(b) Given \(t \ge 0\), we prove that F(t, .) is \(\iota (t)\)-Lipschitz continuous on \({\mathcal {H}}^2\), for some continuous mapping \(\iota : [0,\infty ) \rightarrow [0,\infty )\). Indeed, for \((v_i,y_i)\in {\mathcal {H}}^2\) (for \(i=1,2\)), while noticing that \(J_{\sigma (t)}^{\partial f}\) and \(\frac{1}{2}\sigma (t) (\partial f)_{\sigma (t)}\) are nonexpansive on  \({\mathcal {H}}\), by (1.33) we get

$$\begin{aligned} \begin{array}{l} \Vert \phi _1(t,(v_1,y_1)) - \phi _1(t,(v_2,y_2)) \Vert \\ \le \theta (t)\Vert y_1-y_2\Vert +\theta (t)\Vert J_{\sigma (t)}^{\partial f }v_1-J_{\sigma (t)}^{\partial f }v_2\Vert + \omega (t) \Vert {\sigma (t)} (\partial f)_{\sigma (t)}v_1-{\sigma (t)} (\partial f)_{\sigma (t)}v_2\Vert \\ \le (\theta (t)+2 \omega (t))(\Vert v_1-v_2\Vert +\Vert y_1-y_2\Vert ), \end{array} \end{aligned}$$

while an easy computation gives us

$$\begin{aligned} \begin{array}{l} \Vert \phi _2(t,v_1,y_1)-\phi _2(t,v_2,y_2)\Vert \le \kappa \left( \Vert v_1-v_2\Vert + \Vert y_1-y_2\Vert \right) .\end{array} \end{aligned}$$

It follows from the previous arguments that F(t, .) satisfies

$$\begin{aligned} \Vert F(t,(v_1,y_1))-F(t,(v_2,y_2))\Vert \le \big (\theta (t)+2 \omega (t)+ \kappa \big )\Vert (v_1,y_1)-(v_2,y_2)\Vert ,\nonumber \\ \end{aligned}$$
(1.34)

hence F(t., .) is \( \iota (t)\)-Lipschitz continuous on \({\mathcal {H}}^2\) along with \( \iota ( \cdot )=\theta ( \cdot )+2 \omega ( \cdot )+ \kappa \) which is continuous (by the continuity of \(\theta ( \cdot )\) and \(\omega ( \cdot )\)).

Thus, for any given \((x_0,y_0,\xi _0)\in {\mathcal {H}}^3\), applying the global Cauchy–Lipschitz theorem yields existence and uniqueness of a global classical solution \((v( \cdot ),y( \cdot ))\) to (1.31) (namely, \(y( \cdot )\) and \(v( \cdot )\) are of class \(C^1\)) such that \(y(0)=y_0\) and \(v(0)=x_0+ \sigma (0) \xi _0\). Furthermore, the previous arguments (a) and (b) guarantee existence and uniqueness of a strong global solution \((v( \cdot ),y( \cdot ))\) to the same problem (1.31), by invoking the version of the Cauchy–Lipschitz theorem involving absolutely continuous trajectories, see for example [25, Proposition 6.2.1.], [37, Theorem 54].

(s3) Existence of a strong global solution \((x( \cdot ),\xi ( \cdot ),y( \cdot ))\) to (1.4). Let \((x_0,\xi _0,y_0) \in {\mathcal {H}}^3\) be such that \( \xi _0 \in \partial f (x_0)\). Given a global classical solution \((v( \cdot ),y( \cdot ))\) to (1.31) such that \(y(0)=y_0\) and \(v(0)=x_0+ \sigma (0) \xi _0\), we consider the functions \(x( \cdot )\) and \(\xi ( \cdot )\) defined by (1.30), and we show through the following items (s3-a)(s3-b) that \((x( \cdot ),\xi ( \cdot ),y( \cdot ))\) is a strong global solution to (1.4):

(s3-a) Let us prove the absolute continuity of \(x( \cdot )\) and \(\xi ( \cdot )\) on any bounded subset of \([0, \infty )\). As \(v( \cdot )\) is of class \(C^1\) on \([0, \infty )\), we immediately see that \(v( \cdot )\) is absolutely continuous from the characterization (i1) of Definition 2.1. Hence, given \(\epsilon >0\) and finitely many intervals \(I_k=(a_k,b_k)\) such that \(I_k \cap I_j = \emptyset \) (for \(k \ne j\)), by using Definition 2.1-(i3) we know for some \(\eta >0\) that \(\sum _k |b_k-a_k| \le \eta \) implies that \(\sum _k \Vert v(b_k)-v(a_k)\Vert \le \min (\epsilon , \frac{2 \epsilon }{\sigma _0}).\) So, invoking the non-expansiveness of \(J_{\sigma ( \cdot )}^{ \partial f}\) and \( \frac{1}{2} \sigma ( \cdot ) (\partial f)_{\sigma ( \cdot )}\) while recalling that \(\sigma ( \cdot ) \ge \sigma _0 >0\) entails that

\(\sum _k \Vert J_{\sigma (t)}^{ \partial f} v(b_k)- J_{\sigma (t)}^{\partial f} v (a_k)\Vert \le \sum _k \Vert v(b_k)- v (a_k)\Vert \le \epsilon \)

and that

\(\sum _k \Vert (\partial f)_{\sigma (t)} v(b_k)-(\partial f)_{\sigma (t)} v(a_k)\Vert \le \sum _k \frac{2}{\sigma (t)}\Vert v(b_k)-v(a_k)\Vert \le \frac{2}{\sigma _0} \sum _k \Vert v(b_k)-v(a_k)\Vert \le \epsilon \).

Consequently, the mappings \(x( \cdot )= J_{\sigma ( \cdot )}^{\partial f }v( \cdot )\) and \(\xi ( \cdot ) = (\partial f)_{\sigma ( \cdot )} v( \cdot )\) also comply with characterization (i3) of Definition 2.1, which proves that \(x( \cdot )\) and \(\xi ( \cdot )\) are absolutely continuous on \([0, \infty )\).

(s3-b) Let us show that the triplet \((x( \cdot ),y( \cdot ),\xi ( \cdot ))\) satisfies system (2.3). Indeed, by \(x( \cdot )= J_{\sigma ( \cdot )}^{\partial f }v( \cdot )\) (from (1.30)), and \(\sigma ( \cdot ) \xi ( \cdot )=v( \cdot )-x( \cdot )\), we readily deduce that \(\sigma ( \cdot ) \xi ( \cdot ) \in \left( \sigma ( \cdot ) \partial f \right) (x( \cdot ))\) (because \(v( \cdot ) \in x( \cdot ) + {\sigma ( \cdot )}\partial f (x( \cdot ))\)), which by the positivity of \(\sigma ( \cdot )\) proves (2.3a). Moreover, in (1.31), substituting \(v( \cdot )\), \( J_{\sigma ( \cdot )}^{\partial f }v( \cdot )\) and \({\sigma ( \cdot )} (\partial f)_{\sigma ( \cdot )}v( \cdot )\), by \(x( \cdot ) + \sigma ( \cdot ) \xi ( \cdot )\), \(x( \cdot )\) and \(\sigma ( \cdot ) \xi ( \cdot )\), respectively, gives us immediately (2.3b) and (2.3c). In addition, regarding the initial conditions we obtain \(x(0)=J_{\sigma (0)}^{\partial f } v(0)=x_0\) (since \(v(0)=x_0+\sigma (0) \xi _0\) and \( \xi _0 \in \partial f (x_0)\)), \(y(0)=y_0\) and \(\sigma (0) \xi (0)=v(0)-x(0)=\sigma (0) \xi _0\).

Consequently, by items (s3-a) (s3-b), we get the existence of a strong global solution to (1.4) \(\square \)

1.3 Proof of Lemma 3.2

(See [30, Lemma 4.1]). Given \(t\in [ 0,\infty )\) and \(h\in (0,\infty )\), we have \(\xi (t)\in \partial fx(t)\) and \(\xi (t+h)\in \partial f x(t+h)\) hence, by monotonicity of \(\partial f\), we simply have

$$\begin{aligned} \begin{array}{l} \big <\frac{1}{h} \big (\xi (t+h)-\xi (t) \big ),\frac{1}{h} \big (x(t+h)-x(t) \big )\big >\ge 0. \end{array} \end{aligned}$$
(1.35)

Clearly, assuming that \(x( \cdot )\) and \(\xi ( \cdot )\) are absolutely continuous on \([0,\infty )\), yields that, for a.e. \(t \in [0,\infty )\), and as \(h \rightarrow 0^+\), we have

$$\begin{aligned} {\Vert \frac{1}{h}\big ( x(t+h)-x(t) \big ) \!-\! \dot{x}(t)\Vert \rightarrow 0 \text { and } \Vert \frac{ 1}{h}\big (\xi (t+h)-\xi (t)\big )\!-\! {\dot{\xi }}(t)\Vert \rightarrow 0}.\qquad \end{aligned}$$
(1.36)

Thus, letting h tend to \(0^+\) in (1.35), implies \(\langle \dot{\xi }(t),\dot{x}(t)\rangle \ge 0\), that is the desired result \(\square \)

1.4 Proof of Proposition 4.3

By Remark 3.1, we know under condition (1.15) that \(\theta ( \cdot )\) is well-defined and positive on \([0,\infty )\). Moreover, by \(\nu ( \cdot ) \in C^2\) and \(\theta ( \cdot )=\frac{\kappa \nu ( \cdot ) -\dot{\nu }( \cdot )}{\nu ( \cdot )+ e_*}\) (hence \( \theta ( \cdot )=\kappa -\frac{ \dot{\nu }( \cdot )+\kappa e_*}{\nu ( \cdot ) + e_*}\)), we can see that \(\theta ( \cdot ) \in C^1([0,\infty ))\) and (omitting the variable t) we readily get

$$\begin{aligned} \theta ( \kappa - \theta )= & {} \frac{(\kappa \nu - {\dot{\nu }})({\dot{\nu }} + \kappa e_*)}{(\nu + e_*)^2} \text { and } \dot{\theta }=\frac{\dot{\nu }( \dot{\nu } + \kappa e_*) - (\nu + e_*) \ddot{\nu }}{(\nu + e_*)^2} \nonumber \\{} & {} (\text {hence} \frac{\dot{\theta }}{\theta } = \frac{ \dot{\nu }( {\dot{\nu }} + \kappa e_*) - (\nu + e_*) \ddot{\nu } }{(\nu + e_*)(\kappa \nu -\dot{\nu })}). \end{aligned}$$
(1.37)

Let us recall (from (1.3)) that \(\alpha := \frac{\dot{\theta }}{\theta } + \kappa - \theta \). Consequently, by the previous arguments we obtain

\(\alpha := \frac{1}{\theta } (\theta ( \kappa - \theta ) - {\dot{\theta }}) = \frac{1}{\kappa \nu - {\dot{\nu }}} \left( (\dot{\nu } +\kappa e_*) \frac{\kappa \nu - 2 {\dot{\nu }}}{\nu + e_*} + \ddot{\nu }\right) \).

So, as \(t \rightarrow \infty \), by \(\nu (t) \rightarrow +\infty \), \(\dot{\nu }(t) \rightarrow l \in [0,\infty )\) and \(\ddot{\nu }(t) \rightarrow 0\) (from (4.40)), we immediately obtain that \(\alpha (t) \sim \frac{ l +\kappa e_*}{\nu (t) + e_*}\). We also recall (from (1.3)) that \(\beta :=-\frac{\dot{\theta }}{{\theta }}+ \kappa +\omega \), hence by the definition of \(\alpha \) we equivalently have \(\beta =\alpha + \theta + \omega \). Moreover, as \(t \rightarrow \infty \), by \(\omega :=\big (\kappa -\frac{\dot{\nu }}{\nu } \big )\vartheta -\frac{\delta }{\nu +e_*}\) (from (1.5c)), by \(\nu (t) \rightarrow \infty \), \({\dot{\nu }}(t) \rightarrow l\) and \(\vartheta (t) \rightarrow \vartheta _\infty \), we readily deduce that \(\omega (t) \rightarrow \kappa \vartheta _\infty \). Then, as \(t \rightarrow \infty \), by the latter formulation of \(\beta \), and remembering that (as \(t\rightarrow \infty \)) \(\theta (t) \rightarrow \kappa \), \(\alpha (t) \rightarrow 0\) and that \(\omega (t) \rightarrow \kappa \vartheta _\infty \), we get \(\beta (t) \rightarrow \kappa (1+\vartheta _\infty )\). Again from (1.3) we simply have

\(b(t):=\omega (t) \left( \kappa +\frac{{\dot{\omega }}(t)}{ \omega (t)}- \frac{\dot{\theta }(t)}{{\theta (t)}} \right) = \omega (t) \left( \kappa - \frac{\dot{\theta }(t)}{{\theta (t)}} \right) + {\dot{\omega }}(t)\).

In addition, we obviously see from its expression that \(\omega ( \cdot )\) is of class \(C^1\), and we have

$$\begin{aligned} \begin{array}{l} { {\dot{\omega }}(t)= \left( \kappa -\frac{\dot{\nu }(t)}{\nu (t)} \right) {\dot{\vartheta }}(t)- \frac{\ddot{\nu }(t) \nu (t)-{\dot{\nu }}(t) {\dot{\nu }}(t)}{\nu ^2(t)} \vartheta (t) + \frac{\delta {\dot{\nu }}(t)}{(\nu (t) + e_*)^2}}. \end{array} \end{aligned}$$
(1.38)

It is also classically deduced from the convergence of \(\vartheta \) and the Lipschitz continuity of \({\dot{\vartheta }}\) that \({\dot{\vartheta (t)}} \rightarrow 0\) (as \(t \rightarrow \infty \)). This, in light of condition (4.40) and \(\lim _{t \rightarrow \infty } \vartheta (t)= \vartheta _\infty \) entails that \({\dot{\omega }}(t) \rightarrow \kappa \vartheta _\infty \) (as \(t \rightarrow \infty \)). Consequently, as \(t \rightarrow \infty \), by the previous formulation of b together with \(\frac{\dot{\theta }(t)}{{\theta }(t)} \rightarrow 0\), \(\omega (t) \rightarrow \kappa \vartheta _\infty \) and \({\dot{\omega }}(t) \rightarrow 0\), we deduce that \( b(t) \rightarrow \kappa ^2 \vartheta _\infty \)

\(\square \)

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) 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

Maingé, PE., Weng-Law, A. Fast Continuous Dynamics Inside the Graph of Subdifferentials of Nonsmooth Convex Functions. Appl Math Optim 89, 1 (2024). https://doi.org/10.1007/s00245-023-10055-9

Download citation

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00245-023-10055-9

Keywords

Mathematics Subject Classification

Navigation