Optimization of metastable timescales for accelerated sampling of trajectories

SMAI 2025

Noé Blassel (avec Tony Lelièvre et Gabriel Stoltz)

CERMICS lab, École des Ponts ParisTech, MATHERIALS team, Inria Paris. Funding from ERC EMC2 and ANR SINEQ.

2025-04-22

Molecular dynamics

Molecular dynamics (MD) probes the behavior of atomistic system at the level of classical motion of nuclei.

5 picoseconds of the protein Foldit1 (PDB ID 6MRR) simulated using Molly.jl, (courtesy of Joe Greener).

  • Aims to replace physical experiments in extreme or expensive conditions with numerical experiments.
  • Long trajectories are necessary to access dynamical properties, and parametrize models on larger scales.
  • Many algorithmic tricks are needed to simulate long trajectories reliably and efficiently, and sample the whole configuration space.
  • Length of sampled trajectories is hindered by metastability/stiffness/separation of timescales, the manifestation of free-energetic barriers larger than typical thermal fluctuations.
  • Configurational ensemble has many modes, corresponding to metastable states.

The timescale problem in MD

Accessible systems for direct MD Scaling in simulations: from petascale to exascale. Courtesy of Danny Perez (LANL)

  • Parallelization in space is “easy”, parallelization in time is much harder.
  • Accelerated dynamics methods: (Hyperdynamics (Voter 1997), Temperature Accelerated Dynamics (Sørensen and Voter 2000), Parallel Replica (Voter 1998)) can help.
  • Metastable dynamics jump rarely and suddenly between metastable states.
  • Accelerated MD aims at predicting the sequence of states and transition times correctly by sacrificing resolution of the dynamics inside each state.
  • Spatial coarse-graining strategy.
  • Key idea: the system will typically converge in a statistical sense to a local equilibrium within a metastable state \(\Omega\) and remain there for a long time before exiting.
  • For this idea to be valid, and for efficient algorithms, we require: \[ \frac{\text{Typical time to converge to local equilibrium in }\Omega}{\text{Typical time to exit from }\Omega} \ll 1. \]

Dynamical setting

Molecular trajectories are solutions to a stochastic differential equation: \[ \mathrm{d}X_t = -a(X_t)\nabla V(X_t)\,\mathrm{d}t + \frac1\beta\mathrm{div}\,a(X_t)\,\mathrm{d}t + \sqrt{\frac{2}{\beta}}a(X_t)^{1/2}\,\mathrm{d}W_t, \]

  • \(V:{\mathbb{{R}}}^d\to {\mathbb{{R}}}\) potential energy function, \(a:{\mathbb{{R}}}^{d\times d}\) s.p.d. diffusion coefficient, \(W\) standard Brownian motion in \({\mathbb{{R}}}^d\).

  • Ergodic and reversible with respect to the standard Gibbs measure at temperature \((k\beta)^{-1}\) \[ \pi(\mathrm{d}x) = \mathrm{e}^{-\beta V(x)}\,\mathrm{d}x/Z_\beta. \]

  • Continuous-time preconditioned Langevin Monte-Carlo / diffusion on the Riemannian manifold \(\mathcal M=({\mathbb{{R}}}^d,a^{-1})\)

  • Evolution semigroup generator is symmetric on \(L^2({\mathbb{{R}}}^d,\pi)\): \[ \mathcal{L}_\beta \varphi = \frac1\beta\mathrm{e}^{\beta V}\mathrm{div}\,\left(\mathrm{e}^{-\beta V}a\nabla\varphi\right). \]

  • We assume everywhere that \(V,a\in\mathcal W^{2,\infty}_{\mathrm{loc}}\), and \(a\) is locally elliptic: \(a^{-1}\in L^\infty_{\mathrm{loc}}\).

Local approach to metastability

  • For \(\Omega\subset{\mathbb{{R}}}^d\), local equilibrium is given by a quasi-stationary distribution (QSD) in \(\Omega\) i.e. a probability \(\nu_\Omega\in \mathcal M_1(\Omega)\) for which \[ \forall\,A\subset \Omega,\qquad \mathbb P_{\nu_\Omega}\left(X_t \in A\,\middle|\,t<T_{\mathrm{exit}}(\Omega)\right) = \nu_\Omega(A). \]
  • Since \(V\), \(\Omega\) are bounded and \(a\) is uniformly elliptic on \(\Omega\), the QSD is unique, and related to the principal Dirichlet eigenvector: \[ \nu_\Omega(\mathrm{d}x) = \frac{u_{1,\Omega}(x)\mathrm{e}^{-\beta V(x)}\,\mathrm{d}x}{\int_\Omega u_{1,\Omega}\mathrm{e}^{-\beta V}},\qquad -\mathcal{L}_\beta u_{1,\Omega}(x) = \lambda_{1,\Omega}u_{1,\Omega}(x), \]
  • Here, we consider the generator \(\mathcal{L}_\beta\) with absorbing Dirichlet boundary conditions on \(\partial \Omega\): \[ \mathcal D(\mathcal{L}_\beta) = H_0^1(\Omega) \cap \{u\in L^2(\Omega):\,\mathcal{L}_\beta u\in L^2(\Omega)\subset\mathcal D'({\mathbb{{R}}}^d)\}. \] It is self-adjoint with compact resolvent on \(L^2_\beta(\Omega):=L^2(\Omega,\mathrm{e}^{-\beta V(x)}\,\mathrm{d}x)\) for bounded \(\Omega\).

Spectral characterization of timescales

Denote \(\mu_t = \mathrm{Law}\left(X_t \middle| t<T_{\mathrm{exit}}(\Omega)\right)\) the law of the process conditioned on remaining trapped.

  • Theorem (Le Bris et al. (2012))

    \[ \mu_0=\nu_\Omega \implies T_{\mathrm{exit}}(\Omega)\sim \mathcal E(\color{blue}{\lambda_{1,\Omega}}),\qquad T_{\mathrm{exit}}(\Omega) \perp\!\!\!\!\perp X_{T_{\mathrm{exit}(\Omega)}}. \]

  • This result justifies the use of the Parallel Replica algorithm.

  • The exit rate is given by the principal Dirichlet eigenvalue \(\lambda_{1,\Omega} = 1/\mathbb E_{\nu_\Omega}[T_{\mathrm{exit}}(\Omega)]\).

  • Theorem (Le Bris et al. (2012))

    Convergence to the QSD and bias on the exit event are exponentially decaying. There exist \(C_1,C_2>0\) depending on \(\mu_0\) such that \[ \left\|\mu_t-\nu_\Omega\right\|_{\mathrm{TV}} \leqslant C_1 \mathrm{e}^{-(\color{red}{\lambda_{2,\Omega}-\lambda_{1,\Omega}})t}, \] \[ \left\|\mathrm{Law}_{\mu_t}(T_{\mathrm{exit}(\Omega)}-t,X_{T_{\mathrm{exit}}(\Omega)}) - \mathrm{Law}_{\nu_\Omega}(T_{\mathrm{exit}(\Omega)},X_{T_{\mathrm{exit}}(\Omega)})\right\|_{\mathrm{TV}}, \leqslant C_2 \mathrm{e}^{-(\color{red}{\lambda_{2,\Omega}-\lambda_{1,\Omega}})t}. \]

  • The decorrelation rate is given by the spectral gap of the Dirichlet generator \(\lambda_{2,\Omega}-\lambda_{1,\Omega}\).

  • Set \(T_{\mathrm{corr}}(\Omega) = -\log\,\varepsilon_{\mathrm{corr}}/(\lambda_{2,\Omega}-\lambda_{1,\Omega})\) in ParRep, for some fixed tolerance parameter \(0<\varepsilon_{\mathrm{corr}}<1\).

  • In practice, raises the need for quantitative or statistical estimates of \(\lambda_{2,\Omega}-\lambda_{1,\Omega}\) (and also \(C_1,C_2\)).

A spectral shape optimization problem

  • The goal becomes to maximize the efficiency of ParRep by maximizing the separation of timescales: \[ \underset{\Omega \subset {\mathbb{{R}}}^d\text{ open, bounded, connected}}{\max}\, \left(\color{red}{\lambda_{2,\Omega}-\lambda_{1,\Omega}}\right)/\color{blue}{\lambda_{1,\Omega}}. \]
  • Make the exit rate as small as possible compared to the decorrelation rate. Quantitative measure of local metastability / timescale separation inside \(\Omega\).
  • Standard choice takes \(\Omega=\mathcal A(z_0)\), the basin of attraction of some local minimum \(z_0\) for \(V\) under the steepest-descent dynamics \(\dot X(t) = -a(X(t))\nabla V(X(t))\).
  • The case of the Dirichlet Laplacian \(a=\mathrm{Id}\), \(V=0\) is a famous question in spectral geometry (Payne-Polya-Weinberger conjecture).
  • Question has been identified by MD practitioners, e.g. in Perez, Uberuaga, and Voter (2015), but not studied in depth.
  • Problem is ill-posed as a global optimization problem, e.g., for \(V(x) = |x|^2/2 + \delta V(x)\), \(\delta V\in \mathcal C^\infty_c({\mathbb{{R}}}^d)\): \[ \lambda_{2,B(0,R)}/\lambda_{1,B(0,R)}\xrightarrow{R\to +\infty} +\infty. \]
  • Typical situation in spectral optimization: introduce constraints, or focus on local optimization.

Local optimization with steepest ascent method

  • Expressions for shape-perturbations of the Dirichlet eigenvalues, using Hadamard calculus.
  • Given a space of perturbations: \(\theta\in\mathcal W^{1,\infty}({\mathbb{{R}}}^d,{\mathbb{{R}}}^d)\), consider the transported domain \[ \Omega_\theta = \{x+\theta(x),\,x\in\Omega\}. \]
  • Study properties of the eigenvalue perturbation maps \[ \lambda_k:\theta\mapsto \lambda_{k,\Omega_\theta}. \]
  • The spectrum (of \(-\mathcal{L}_\beta\) with Dirichlet b.c. on \(\Omega\)) is enumerated with multiplicity in order of increasing magnitude \[ 0<\lambda_{1,\Omega}<\lambda_{2,\Omega}\leqslant\lambda_{3,\Omega}\leqslant\cdots \] with an associated basis of eigenvectors \[ \int_{\Omega}u_{i,\Omega}u_{j,\Omega}\mathrm{e}^{-\beta V} = \delta_{ij},\qquad \forall\, i,j\geqslant 1. \]

A domain \(\Omega\) and its perturbation \(\Omega_\theta\).

Explicit formulas for shape perturbations

Assume \(\Omega\subset{\mathbb{{R}}}^d\) is a bounded \(\mathcal C^{1,1}\) (or convex Lipschitz) domain and \(\lambda_{k,\Omega}\) has multiplicity \(m_k\geqslant 1\): \[ \lambda_{k-1,\Omega}<\lambda_{k,\Omega}=\cdots=\lambda_{k+m_k-1,\Omega}<\lambda_{k+m_k,\Omega}. \]

Theorem (Blassel, Lelièvre, and Stoltz (2025b), in preparation)

The map \(\theta\mapsto (\lambda_{k+\ell,\Omega_\theta})_{0\leqslant\ell < m_k}\) is locally Lipschitz in a \(\mathcal W^{1,\infty}({\mathbb{{R}}}^d;{\mathbb{{R}}}^d)\)-neighborhood of \(\theta=0\).

If \(m_k=1\), \(\lambda_k\) is \(\mathcal C^1\) in a \(\mathcal W^{1,\infty}({\mathbb{{R}}}^d;{\mathbb{{R}}}^d)\)-neighborhood of \(\theta=0\), with Fréchet derivative \[ D\lambda_k \theta = -\frac1\beta \int_{\partial\Omega}\left(\frac{\partial u_{k,\Omega}}{\partial\mathrm{n}}\right)^2\mathrm{n}^\top a\mathrm{n}\theta^\top\mathrm{n}\mathrm{e}^{-\beta V}, \] the shape derivative of \(\lambda_k\).

If \(m_k>1\), \(\lambda_k\) is only Gateaux semi-differentiable: \(t\mapsto\lambda_{k,\Omega_{t\theta}}\) is right-differentiable at \(t=0\). Its derivative is the bottom eigenvalue of the matrix \[ M^{\Omega,k}_{ij}(\theta) = -\frac1\beta\int_{\partial\Omega}\frac{\partial u_{k+i-1,\Omega}}{\partial\mathrm{n}}\frac{\partial u_{k+j-1,\Omega}}{\partial\mathrm{n}}\mathrm{n}^\top a\mathrm{n}\theta^\top\mathrm{n}\mathrm{e}^{-\beta V},\qquad\forall\, 1\leqslant i,j\leqslant m_k. \]

Idea of proof

Transport variational formulation of eigenvalue PDE on \(\Omega_{\theta} = \Phi_\theta\Omega\) back to \(\Omega\), transforming it into a \(\theta\)-dependent PDE on a fixed domain.

Regularity results follow from perturbation theory (Kato (2013)), using an approach inspired by Haug and Rousselet (1980), Rousselet (1983) (for problems in structural mechanics).

Perturbation of the eigenvalue \(\lambda_{k,\Omega}\) in the direction \(\theta\) (here \(m_k=3\))

Slopes of the right-tangents are the eigenvalues of \(M^{\Omega,k}(\theta)\) (with multiplicity).

Efficient numerical optimization requires taking eigenvalue degeneracies into account, and regularization of the ascent direction.

Galerkin objective for high-dimensional systems

For systems of interest, \(d\gg 1\). In MD, common to use a collective variable (CV) \(\xi:{\mathbb{{R}}}^d\to{\mathbb{{R}}}^m\) with \(m\ll d\) for interpretation and biasing.

It is natural to consider states defined in CV space: \(\Omega = \xi^{-1}(\Omega^\xi)\) for some \(\Omega^\xi\subset{\mathbb{{R}}}^m\)

Coarse-grained spectrum

Consider the test space \[ \mathcal V_\xi(\Omega^\xi) = \left\{\varphi\circ\xi,\,\varphi\in H_0^1(\Omega_\xi)\right\} \subset H_0^1(\xi^{-1}(\Omega_\xi)), \] introduce the Rayleigh-Ritz (or “coarse-grained”) eigenpairs: \[ Q_\xi(\varphi,\psi;\Omega_\xi)=Q(\varphi\circ\xi,\psi\circ\xi;\xi^{-1}(\Omega_\xi)),\qquad Q_\xi(u_{k,\Omega_\xi}^\xi,u_{\ell,\Omega_\xi}^\xi) = \lambda_{k,\Omega_\xi}^\xi\delta_{k\ell}, \] where \(Q\) is the Rayleigh quotient for \(-\mathcal{L}_\beta\) \[ Q(u,v;\Omega)=\frac1\beta\frac{\int_\Omega \nabla u^\top a \nabla v \mathrm{e}^{-\beta V}}{\int_\Omega uv\mathrm{e}^{-\beta V}}. \]

Co-area formula gives (Blassel, Lelièvre, and Stoltz (2025b))

\[ \forall\,\varphi,\psi\in\mathcal V_\xi(\Omega^\xi),\qquad Q_\xi(\varphi,\psi;\Omega_\xi) = \frac1\beta\frac{\int_{\Omega_\xi}\nabla\varphi^\top a_\xi \nabla\psi \mathrm{e}^{-\beta F_\xi}}{\int_{\Omega_\xi}\varphi\psi\mathrm{e}^{-\beta F_\xi}}, \] where \(F_\xi,a_\xi\) are the free-energy and effective diffusion through \(\xi\): \[ F_\xi(z) = -\frac1\beta\log\,\int_{\xi^{-1}(z)}\mathrm{e}^{-\beta V}(\det [\nabla\xi^\top\nabla\xi])^{-1/2}\mathrm{d}\sigma,\quad a_\xi(z) = \int_{\xi^{-1}(z)}\nabla\xi^\top a\nabla\xi \mathrm{d}\mu_z, \] where \(\mu_z\) is the conditional measure of \(\mu\) on the level-set \(\xi^{-1}(z)\).

Error estimate for all \(k\geqslant 1\): \[ 0 \leqslant\lambda^\xi_k-\lambda_k \leqslant\frac1\beta\int_{\xi^{-1}(\Omega_\xi)}\nabla\left[u_j-u_j^\xi\circ\xi\right]^\top a\nabla\left[u_j-u_j^\xi\circ\xi\right]\mathrm{e}^{-\beta V}. \]

Rates \(\lambda_{k,\Omega_\xi}^\xi\) correspond to rates for an effective dynamics in \({\mathbb{{R}}}^m\): \[ \mathrm{d}Z_t^\xi = \left(-a_\xi\nabla F_\xi + \frac1\beta\mathrm{div}\,a_\xi\right)(Z_t^\xi)\,\mathrm{d}t + \sqrt{\frac{2}{\beta}}a_\xi(Z_t^\xi)^{1/2}\,\mathrm{d}B_t. \] Shape perturbation results apply verbatim if \(\xi\in\mathcal W^{3,\infty}({\mathbb{{R}}}^d,{\mathbb{{R}}}^m)\) and \(\mathrm{rank}\,\nabla\xi \equiv m\).

Example: solvated Alanine dipeptide

Consider alanine-dipeptide in solution \((d=1857)\).

We use standard dihedral angles \(\xi=(\phi,\psi)\) as a CV.

Computed free-energy (\(F_\xi\)) landscape at \(T=300K\).

Computed effective diffusion tensor (\(a_\xi\)).

As \(F_\xi,a_\xi\) are thermodynamic quantities, can be computed efficiently with enhanced sampling methods. Here, umbrella sampling using Tinker-HP and Colvars.

Shape-optimization results

Shape-gradient ascent, implemented with FreeFem++.

Convergence of the Galerkin objective.

Behavior of eigenvalues throughout the ascent iterations.

Shape-optimization results

Degeneracy-aware steepest ascent method implemented in FreeFem++.

Initial and optimized domains, associated effective QSDs.

Quantifying gains in the original objective

Measure gain in timescale separation with respect to reference state definition, in a practical dynamical setting.

Simulate Fleming-Viot process (in a modded Tinker-HP with 50 replicas), sampling initial configurations in transition regions.

\(\xi\)-marginal of the QSD for \(\mathcal A(z_5)\).

\(\xi\)-marginal for \(\Omega_\xi^*\) around \(z_5\).

Inference of decorrelation timescale

For the various ensembles of initial conditions, infer decorrelation rate at the level of histogram convergence.

Free-energy basin (\(\gamma=2\,\mathrm{ps}^{-1}\)).

Optimized state (\(\gamma=2\,\mathrm{ps}^{-1}\)).

Comparison of free-energy basin and optimized state

State Exit rate \(\pm 1.96\sigma\) (\(\mathrm{ps}^{-1}\)) Decorrelation rate (\(\mathrm{ps}^{-1}\)) Ratio \(\pm 1.96\sigma\) Mixing time (\(\mathrm{ps}\), \(\mathrm{tol}=0.05\))
\(\mathcal{A}(z_5)\ (\gamma = 1\,\mathrm{ps}^{-1})\) \(5.6 \times 10^{-3} \pm 1.6 \times 10^{-4}\) \(0.49\) \(88.1 \pm 2.5\) \(5.8\)
\(\Omega_\xi^*\ (\gamma = 1\,\mathrm{ps}^{-1})\) \(1.8 \times 10^{-3} \pm 8 \times 10^{-5}\) \(0.33\) \(177.5 \pm 7.7\) \(7.8\)
\(\mathcal{A}(z_5)\ (\gamma = 2\,\mathrm{ps}^{-1})\) \(5.4 \times 10^{-3} \pm 1.4 \times 10^{-4}\) \(0.46\) \(85.3 \pm 2.3\) \(7.1\)
\(\Omega_\xi^*\ (\gamma = 2\,\mathrm{ps}^{-1})\) \(1.8 \times 10^{-3} \pm 8.5 \times 10^{-5}\) \(0.33\) \(187.0 \pm 9.0\) \(8.4\)
\(\mathcal{A}(z_5)\ (\gamma = 5\,\mathrm{ps}^{-1})\) \(5.1 \times 10^{-3} \pm 1.4 \times 10^{-4}\) \(0.39\) \(76.9 \pm 2.2\) \(9.0\)
\(\Omega_\xi^*\ (\gamma = 5\,\mathrm{ps}^{-1})\) \(1.5 \times 10^{-3} \pm 8 \times 10^{-5}\) \(0.34\) \(233.0 \pm 13.0\) \(10.8\)
\(\mathcal{A}(z_5)\ (\gamma = 10\,\mathrm{ps}^{-1})\) \(4.2 \times 10^{-3} \pm 1.4 \times 10^{-4}\) \(0.3\) \(70.4 \pm 2.3\) \(11.1\)
\(\Omega_\xi^*\ (\gamma = 10\,\mathrm{ps}^{-1})\) \(1.2 \times 10^{-3} \pm 7.8 \times 10^{-5}\) \(0.27\) \(230.0 \pm 15.0\) \(12.1\)
\(\mathcal{A}(z_5)\ (\gamma = 50\,\mathrm{ps}^{-1})\) \(2.0 \times 10^{-3} \pm 1.2 \times 10^{-4}\) \(0.088\) \(44.0 \pm 2.6\) \(41.2\)
\(\Omega_\xi^*\ (\gamma = 50\,\mathrm{ps}^{-1})\) \(3.3 \times 10^{-4} \pm 4.7 \times 10^{-5}\) \(0.077\) \(233.0 \pm 33.0\) \(55.0\)

Throughout range of of conditions, optimized state definition leads to \(\approx 2-5\times\) increase in timescale separation with respect to reference state.

Asymptotic optimization methodology in the low-temperature limit

  • Introduce a parametric family of domains \((\Omega_{\beta,\alpha})_{\alpha\in\mathcal X,\beta>0}\), jointly parametrized by the asymptotic parameter \(\beta\) and the shape parameter \(\alpha\).
  • Compute quantitative asymptotics of the objective: \[ \frac{\lambda_2(\Omega_{\beta,\alpha})}{\lambda_1(\Omega_{\beta,\alpha})} = \mathcal F(\beta,\alpha)(1+\mathrm{o}(1)) \] for some explicit \(\mathcal F\) in the limit \(\beta\to+\infty\).
  • Optimize \(\mathcal F\) with respect to \(\alpha\) (the optimal parameter may itself depend on \(\beta\)).
  • Main difficulty is to compute spectral asymptotics for domains which change with the temperature.
  • Equivalent to the semiclassical analysis of the Witten Laplacian \[ -\mathrm{e}^{-\beta V/2}\mathcal{L}_\beta\mathrm{e}^{\beta V/2} = -\frac1\beta \Delta + \frac{\beta}{4}|\nabla V|^2 - \frac12\Delta V \] with \(\hbar:=1/\beta\)-dependent Dirichlet boundary conditions.

Main geometric assumption

  • We restrict to \(a=\mathrm{Id}\), assume \(V\) is a Morse function (non-degenerate Hessians at critical points), and \(\Omega_\beta \subset \mathcal K\) with \(\mathrm{Crit} V|_{\mathcal K}=\{z_0,\dots,z_{N-1}\}\).
  • Assume the following limit is well-defined in \((-\infty, +\infty]\) for each \(0\leqslant i < N\). \[ \alpha_i = \underset{\beta\to\infty}{\lim}\, \sqrt\beta\sigma_{\Omega_\beta}(z_i),\qquad \sigma_{\Omega_\beta}(x) = \begin{cases} d(x,\partial\Omega_\beta),& x\in \Omega_\beta,\\ -d(x,\partial\Omega_\beta)&x\in\mathcal X\setminus \Omega_\beta \end{cases}. \]
  • If \(\alpha_i = +\infty\), \(z_i\) is far from the boundary, and if not, it is close.
  • Assume there exists a function \(\delta :{\mathbb{{R}}}_+\to (0,\varepsilon(V))\) such that \[ \sqrt\beta\delta(\beta)\xrightarrow{\beta\to\infty}\,+\infty,\qquad\mathcal{O}_{i}^{-}(\beta) \subseteq B(z_i,\delta(\beta))\cap\Omega_\beta \subseteq \mathcal{O}_{i}^{+}(\beta), \] where the sets \(\mathcal{O}_{i}^{\pm}\) constrain the geometry of the boundary near the \(z_i\) in a specific way.

Main geometric assumption

  • In the drawing, \(\mathcal{O}_{i}^{\pm}\) are the two capped balls, \(\sqrt\beta\gamma(\beta)\to 0\), and \(v_1^{(i)}\) is an eigenvector of \(\nabla^2 V(z_i)\).
  • If \(z_i\) is an index 1 saddle points, we assume \(v_1^{(i)}\) is the unique eigenvector for the negative eigenvalue.
  • Boundary near a saddle \(z_i \approx\) level set of the position along the minimum energy path through the saddle.
  • Standard settings (\(\Omega_\beta = \Omega\) satisfying a transversality condition) are included.

First-order behavior for the full spectrum

Theorem: Harmonic approximation (Blassel, Lelièvre, and Stoltz (2025a)):

For any \(k\), it holds \[ \underset{\beta\to\infty}{\lim}\, \lambda_{k,\Omega_\beta} = \lambda_k^{\mathrm H}, \] where \(\lambda_k^{\mathrm H}\) is the \(k\)-th (temperature independent) eigenvalue of a direct sum of local harmonic models \[ H^\alpha_\beta = \bigoplus_{i=0}^{N-1} H_\beta^{(i),\alpha_i},\qquad H_\beta^{(i),\alpha_i} = \beta(x-z_i)^\top\nabla V^2(z_i)(x-z_i)/4-\Delta V(z_i)/2 - \frac1\beta \Delta, \] acting on the half-space \(\left\{ (x-z_i)^\top v_1^{(i)} < \alpha^{(i)}/\beta\right\}\) with Dirichlet boundary conditions.

Example

Assume only one minimum \(z_0\) and index 1 saddle points \(z_1,\dots,z_{N-1}\). \[ \lambda_{1,\Omega_\beta}\xrightarrow{\beta\to\infty}0,\quad \lambda_{2,\Omega_\beta}\xrightarrow{\beta\to\infty} \min\left[\nu^{(0)}_{1},\underset{0<i<N}{\min}|\nu^{(i)}_{1}|\left(\color{blue}{\mu_{0,\alpha^{(i)}\sqrt{|\nu^{(i)}_{1}|/2}}}+\frac12\right)\right], \] where \(\color{blue}{\mu_{0,\theta}}=\)ground-state energy of \(\frac12(x^2-\partial_x^2)\) acting on \((-\infty,\theta)\) with Dirichlet boundary conditions and \(\nu^{(i)}_{k}=\) \(k\)-th eigenvalue of \(\nabla^2 V(z_i)\).

  • In particular, quantitative asymptotics for the decorrelation rate \(\lambda_{2,\Omega_\beta}-\lambda_{1,\Omega_\beta}\) whenever \(\lambda_2^{\mathrm{H}}>0\).
  • Theorem says that \(-\mathcal{L}_\beta\) is well-approximated by a block-diagonal operator, with each block corresponds to a local harmonic model of \(-\mathcal{L}_\beta\) around a critical point of \(V\).
  • Proof relies on the construction of approximate eigenvectors, harmonic quasimodes, constructed using Dirichlet quantum oscillators localized around each critical point.
  • Adaptation of a method of Simon (1983) from semiclassical analysis to the (new) setting of moving Dirichlet boundaries.
  • Dependence on shape parameter \(\alpha\): eigenvalues of 1D Dirichlet harmonic oscillators.

Further geometric assumptions

  • We now restrict the setting to the case of only one minimum far from the boundary \(z_0\) (plus arbitrary other critical points).
  • Harmonic approximation then implies \(\lambda_{1,\Omega_\beta}\to 0\), \(\lambda_{2,\Omega_\beta}\to \lambda_{2,\alpha}^{\mathrm{H}}>0\).
  • Define \(V^* = \min\left\{V(z),\,z\in\partial \left[{\mathbb{{R}}}^d\setminus \mathcal A(z_0)\right]\right\}\), where \(\mathcal A(z_0)=\) basin of attraction of \(z_0\) for \(X'(t)=-\nabla V(X(t))\).
  • The minimizers are index one saddle points \(\left\{z_i,i\in I_{\min}\right\}\).
  • Assume also the energetic condition \(\left(\mathcal A(z_0)\cap \{V<V^*+c\delta(\beta)^2\} \setminus \cup_{i\in I_{\min}} B(z_i,\delta(\beta))\right) \subset \Omega_\beta\) for all \(\beta\) and some \(c>0\).
  • This condition avoids the presence of generalized saddle points with energy \(\leqslant V^*\), ensuring that the likeliest exit pathway is intrinsic to \(V\), and not an artefact of the domain geometry.

Modified Eyring-Kramers formula

Theorem (Blassel, Lelièvre, and Stoltz (2025a)):

In the limit \(\beta\to\infty\), it holds \[ \lambda_{1,\Omega_\beta} = \mathrm{e}^{-\beta(V^*-V(z_0))} \left[\sum_{i\in I_{\min}} \frac{|\nu^{(i)}_{1}|}{2\pi\color{blue}{\Phi\left(|\nu^{(i)}_{1}|^{\frac12}\alpha^{(i)}\right)}}\sqrt{\frac{\det \nabla^2 V(z_0)}{\left|\det \nabla ^2 V(z_i)\right|}} \left(1 +\mathcal{O}(1/\sqrt\beta)\right)\right], \] where \[ \Phi(x) = \frac1{\sqrt{2\pi}}\int_{-\infty}^x \mathrm{e}^{-\frac{t^2}2}\,\mathrm{d}t. \]

  • Result provides sharp quantitative estimate for the metastable exit rate \(\lambda_{1,\Omega_\beta}\) Dependence on \(\alpha\) is only in the prefactor \(\color{blue}{P}\).
  • Adds geometric corrections to standard Eyring-Kramers (\(\alpha=+\infty\), \(P=1\)) and (\(\alpha=0\), \(P=1/2\)) formulas.
  • Proof relies on the construction of very precise approximations for \(u_{1,\Omega_\beta}\), inspired by works in potential theory (Bovier, Gayrard, and Klein (2005)) and semiclassical analysis of Witten Laplacians (Le Peutrec and Nectoux (2021)).

Illustration on a 1D energy well

1D potential with several domains. \(\Omega_{\beta,\alpha} = (z_1 - \alpha_1/\sqrt\beta,z_2+\alpha_2/\sqrt\beta)\). Parameters \(\alpha=\color{green}{(0.5,0.3)},\,\color{blue}{(1.0,-0.3)},\color{red}{(0.0,0.0)}\).

We aim to maximize \[ \frac{\lambda_{2,\beta}(\Omega_{\beta,\alpha})-\lambda_{1,\beta}(\Omega_{\beta,\alpha})}{\lambda_{1,\beta}(\Omega_{\beta,\alpha})} \] with respect to \(\alpha\). Equivalently, maximize \[ J_\beta(\alpha) = \frac{\lambda_{2,\beta}(\Omega_{\beta,\alpha})\lambda_{1,\beta}(\Omega_{\beta,0})}{\lambda_{1,\beta}(\Omega_{\beta,\alpha})\lambda_{2,\beta}(\Omega_{\beta,0})}, \] where \(\Omega_{\beta,0}\) is the basin of attraction for the local minimum \(z_0\). According to the results of Blassel, Lelièvre, and Stoltz (2025a), \(J_\beta~\xrightarrow{\beta\to+\infty}J_\infty\) pointwise, where \(J_\infty\) is a complicated but fully explicit function of \(\alpha\).

Semiclassical approximation of shape-optimization landscape

Shape-optimization landscape for \(J_\beta(\alpha)\) for \(\beta=10\). Basin of attraction: \(\color{blue}{+}\). Optimal domain: \(\times\).

Semiclassical approximation \(J_\infty(\alpha)\). Semiclassical optimizer: \(\color{red}{\times}\).

Bibliography

Aristoff, D., M. Johnson, and D. Perez. 2023. “Arbitrarily Accurate, Nonparametric Coarse Graining with Markov Renewal Processes and the Mori–Zwanzig Formulation.” AIP Advances 13 (9): 095131-1 - 11.
Blassel, N., T. Lelièvre, and G. Stoltz. 2025a. “Quantitative Low-Temperature Spectral Asymptotics for Reversible Diffusions in Temperature-Dependent Domains.” ArXiv Preprint: 2501.16082.
———. 2025b. “Shape Optimization of Metastable States.”
Bovier, A., V. Gayrard, and M. Klein. 2005. “Metastability in Reversible Diffusion Processes. II: Precise Asymptotics for Small Eigenvalues.” Journal of the European Mathematical Society 7 (1): 69–99.
Cox, S. J. 1995. “The Generalized Gradient at a Multiple Eigenvalue.” Journal of Functional Analysis 133 (1): 30–40.
Haug, E. J., and B. Rousselet. 1980. “Design Sensitivity Analysis in Structural Mechanics. II. Eigenvalue Variations.” Journal of Structural Mechanics 8 (2): 161–86.
Kato, T. 2013. Perturbation Theory for Linear Operators. Vol. 132. Springer Science & Business Media.
Kirsch, W., H. L. Cycon, R. G. Froese, and B. Simon. 1987. Schrödinger Operators: With Applications to Quantum Mechanics and Global Geometry. Springer, Berlin.
Le Bris, C., T. Lelievre, M. Luskin, and D. Perez. 2012. “A Mathematical Formalization of the Parallel Replica Dynamics.” Monte–Carlo Methods and Applications 18: 119–46.
Le Peutrec, D., and B. Nectoux. 2021. “Small Eigenvalues of the Witten Laplacian with Dirichlet Boundary Conditions: The Case with Critical Points on the Boundary.” Analysis & PDE 14 (8): 2595–2651.
Perez, D., B. P. Uberuaga, and A. F. Voter. 2015. “The Parallel Replica Dynamics Method – Coming of Age.” Computational Materials Science, Special Issue on Advanced Simulation Methods, 100: 90–103.
Rousselet, B. 1983. “Shape Design Sensitivity of a Membrane.” Journal of Optimization Theory and Applications 40: 595–623.
Simon, B. 1983. “Semiclassical Analysis of Low Lying Eigenvalues. I. Non-Degenerate Minima: Asymptotic Expansions.” Annales de l’IHP: Physique Théorique 38 (3): 295–308.
Sørensen, M.., and A. F. Voter. 2000. “Temperature-Accelerated Dynamics for Simulation of Infrequent Events.” The Journal of Chemical Physics 112 (21): 9599–9606.
Voter, A. F. 1997. “Hyperdynamics: Accelerated Molecular Dynamics of Infrequent Events.” Physical Review Letters 78 (20): 3908–11.
———. 1998. “Parallel Replica Method for Dynamics of Infrequent Events.” Physical Review B 57 (22): 9599–9606.