Fixing the flux: a Dual Approach to Computing Transport Coefficients

CECAM Summer School on Sampling High-Dimensional Probability Measures with Applications in (Non)Equilibrium Molecular Dynamics and Statistics, University of Birmingham

Noé Blassel (joint work with Gabriel Stoltz)

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

2025-07-06

Transport coefficients: a computational challenge

  • Sensitivities of fluxes in response to nonequilibrium forcings.
  • Dynamical properties of molecular systems: thermal transport, diffusion, shear viscosity…
  • Crucial to parametrize macroscopic models (e.g. evolution PDEs like Navier–Stokes).
  • Standard methods require long simulations to get useful confidence intervals \(\rightarrow\) need for innovative variance reduction techniques.

Fourier’s law: J \propto \Delta T Fourier’s law: \(J = \kappa (T_+-T_-)\), \(\kappa=\) thermal conductivity.

Formal setting: linear responses of stochastic dynamics

  • Perturbed stochastic differential equation: \[ \mathrm{d}X^\eta_t = b(X^\eta_t)\,\mathrm{d}t+\sigma(X^\eta_t)\,\mathrm{d}W_t +\eta F(X^\eta_t)\,\mathrm{d}t,\qquad X^\eta_t\in\mathcal X. \] Parameters: vector fields \(\mathcal X\to T\mathcal X\), \(b\) (equilibrium drift), \(F\) (nonequilibrium forcing), \(\sigma\) (diffusion matrix), \(W\) Wiener process.

  • Strength of the perturbation: \(\eta\). Equilibrium when \(\eta = 0\).

  • Nonequilibrium ensemble: \(\psi_\eta\in\mathcal M_1(\mathcal X)=\) invariant measure for \(X^\eta\). Generally non-explicit for \(\eta\neq 0\).

  • Define a response observable of flux \(R:\mathcal X \to\mathbb R\), with \(\mathbb E_{\psi_0} R = 0\)

  • Linear response is the derivative of the expected flux at equilibrium: \[ \begin{equation}\alpha(F,R) = \underset{\eta \to 0}{\lim}\, \frac1\eta\mathbb E_{\psi_\eta}[R] \end{equation} \]

Classical estimators of the linear response

  • Nonequilibrium static averages (NEMD): \[ \begin{equation}\widehat\alpha_{T,\eta}^{\mathrm{NEMD}}(F,R) = \frac1{T\eta}\int_0^T R(X_t^\eta)\,\mathrm{d}t,\quad \eta>0. \end{equation} \]
  • Variance scales like \(T^{-1}\eta^{-2}\): simulation times \(\mathcal{O}(\eta^{-2}\varepsilon^{-2})\) for confidence intervals of width \(\mathcal{O}(\varepsilon)\).
  • Equilibrium dynamical averages (Green-Kubo): \[ \begin{equation} \widehat\alpha_T^{\mathrm{GK}}(F,R) = \int_0^{T} \mathbb E_{\psi_0}[R(X_t^0)S(X_0^0)]\,\mathrm{d}t, \end{equation} \] where \(S\) is the so-called conjugate response, which is explicit in terms of \(F\).

Dual approach: fixing the flux

  • Macroscopically: does the forcing cause a flux in the nonequilibrium steady-state, or does the flux cause a resisting forcing ?

  • Microscopically: dynamics with fixed forcing / variable flux vs dynamics with a fixed flux / variable forcing.

  • Constant-flux dynamics \[ \mathrm{d}Y^r_t = b(Y^r_t)\,\mathrm{d}t+\sigma(Y^r_t)\,\mathrm{d}W_t + F(Y^r_t)\mathrm{d}\Lambda^r_t,\quad \mathrm{d}\Lambda^r_t = \lambda(Y_t^r)\,\mathrm{d}t + \widetilde{\lambda}(Y_t^r)\,\mathrm{d}W_t. \]

  • The forcing process/stochastic Lagrange multiplier \(\Lambda^r\) is constructed to enforce the constant flux condition: \(R(Y_t^{r}) = r \in {\mathbb{{R}}}\) for all \(t>0\).

  • \(\Lambda^r\) satisfies a SDE with explicit coefficients \(\lambda,\widetilde{\lambda}\), which can be obtained under the controllability condition \(F^\top\nabla R\neq 0\).

  • Generalization of the “Norton” method of Evans & Morriss (1980s) to stochastic dynamics.

  • Also works for vector-valued fluxes/time-dependent constraints.

Dual approach: fixing the flux

Linear response

  • Denote by \(\psi^r\) the constant-flux steady-state (or “Norton ensemble”).
  • Dual definitions of the linear response \[ \alpha^*(F,R) = \underset{r\to 0}{\lim}\, \frac{r}{\mathbb E_{\psi^r}[\lambda]},\qquad \alpha(F,R) = \underset{\eta\to 0}{\lim}\,\frac{\mathbb E_{\psi_\eta}[R]}{\eta}. \]
  • Set \(r(\eta):=\mathbb{E}_{\psi_\eta}[R],\,\eta(r):=\mathbb E_{\psi^r}[\lambda]\). If \(\eta(r(\eta)) = \eta + o(\eta)\) and \(r(\eta(r))=r + o(r)\), then \(\alpha^*(F,R)=\alpha(F,R)\).
  • Reciprocal estimator of the linear response: \[ \begin{equation}\widehat\alpha_{T,r}^{*}(F,R) = \left(\frac1{T r}\int_0^T \lambda(Y_t^r)\,\mathrm{d}t\right)^{-1},\quad r>0. \end{equation} \]
  • Explicit expression for \(\lambda,\widetilde{\lambda}\): \[ \lambda = -\frac1{F^\top \nabla R}\left(b^\top \nabla R +\frac12 \mathrm{Tr}\left[\sigma^\top \overline{P}_{F,\nabla R}^{\top}\nabla^2 R \overline{P}_{ F,\nabla R}\sigma\right]\right),\qquad \overline{P}_{F,\nabla R} = \mathrm{Id} - \frac{F\nabla R^\top}{F^\top\nabla R} \]
  • \(\rightarrow\) link with variance reduction: uses \(-\frac1T\int_0^T \widetilde{\lambda}(Y_t^r)\mathrm{d}W_t\) as a control variate.

Example: mobility of a Langevin particle

  • Specialize to underdamped Langevin dynamics. Coefficients \(b(q,p)=\begin{pmatrix} p \\-\nabla V(q) - \gamma p\end{pmatrix}\), \(\sigma(q,p) = \sqrt{2\gamma}\begin{pmatrix} 0 & 0 \\ 0 & \mathrm{Id} \end{pmatrix}\) in phase space \((q,p)\in ({\mathbb{{R}}}/\mathbb Z)^d\times {\mathbb{{R}}}^d\).
  • Forcing/Flux pair: fixed direction \(\mathcal F = \begin{pmatrix}0\\ F\end{pmatrix} \in {\mathbb{{R}}}^{2d}\). Response is the system’s velocity in the \(F\)-direction: \(R(q,p) = F^\top p\). Controllability condition: \(\mathcal F^\top \nabla R = |F|^2>0\).
  • Apply Itô’s formula to the constant-flux condition \(R(q^r_t,p^r_t) = r\): \[ \nabla R(q^r_t,p^r_t)^\top\left(\left[b(q_t^r,p_t^r) + \lambda(q_t^r,p_t^r)\mathcal F\right]\,\mathrm{d}t + \left[\sigma+\mathcal F\widetilde\lambda(q_t^r,p_t^r)\right]\mathrm{d}W_t\right) + \frac12\sigma\sigma^\top \nabla^2 R(q_t^r,p_t^r)\mathrm{d}t = 0 \]
  • \[ \implies F^\top\left[-\nabla V(q^r_t) -\gamma p_t^r + \lambda(q_t^r,p_t^r)F\right] = 0,\qquad F^\top\left[\sqrt{2\gamma}\begin{pmatrix} 0 & 0 \\ 0 & \mathrm{Id} \end{pmatrix}+\mathcal F\widetilde\lambda(q_t^r,p_t^r)\right] = 0 \]
  • \[ \implies \lambda(q,p) = -\frac{F^\top (-\nabla V(q)-\gamma p)}{|F|^2},\qquad \widetilde\lambda(q,p) = -\frac{\sqrt{2\gamma}}{|F|^2}\begin{pmatrix}0_q^\top & F^\top \end{pmatrix}. \]

Norton-Langevin dynamics for mobility

  • Plugging these in, we get a dynamics \[ \left\{\begin{aligned} \mathrm{d}q_t^r &= p_t^r\,\mathrm{d}t\\ \mathrm{d}p_t^r &= \overline{P}_{F}\left(-\nabla V(q^r_t)\mathrm{d}t - \gamma p_t^r \mathrm{d}t + \sqrt{\frac{2\gamma}{\beta}}\,\mathrm{d}W_t\right) \end{aligned}\right.,\qquad \overline{P}_F = \mathrm{Id}_p - P_F,\qquad P_F = \frac{F F^\top}{F^\top F} \]
  • Projection of the equilibrium dynamics on the constant-flux manifold.

General case: oblique projection

General case: oblique projection

General case: oblique projection

General case: oblique projection

General case: oblique projection

General case: oblique projection

Does it work ?

Does \(\alpha(F,R) = \alpha^*(F,R)\) ?

Sometimes. Take \(F =N^{-1/2} \begin{pmatrix}1 & 0 & 0 & -1 & 0 & 0 \cdots \end{pmatrix}^\top\): “color-drift” constant forcing.

Linear responses coincide (\(N=1000\)).

Non-linear responses coincide ! \(\rightarrow\) question of equivalence of nonequilibrium ensembles.

Does it always work ?

  • No.
  • Take \(F = \begin{pmatrix} 1 & 0 & 0 & 0 & \cdots\end{pmatrix}^\top\): “single-drift” forcing.
  • \(\rightarrow\) question of sufficient conditions for the equivalence to hold.

Exercise: Take a Langevin particle on \({\mathbb{{R}}}/\mathbb Z \times {\mathbb{{R}}}\). Show that, for any \(V:{\mathbb{{R}}}/\mathbb Z\to {\mathbb{{R}}}\) smooth, for \(F(q,p)=(0,1)\) and \(R(q,p)=p\), \(\alpha^*(F,R)=\gamma^{-1}\). Conclude \(\alpha(F,R)\neq \alpha^*(F,R)\) in general.

Is it efficient ?

  • Assuming equivalence of the constant-forcing and constant-flux ensembles, we have a choice to measure a given response property.
  • Which ensemble to choose: the one with lower asymptotic variance for linear response estimators.

Asymptotic variance as a function of system size, for equivalent nonequilibrium state points (shear-forcing of a Lennard-Jones fluid).

Anomalous scaling of the stationary variance: \(\mathbb E_{\pi^r_{N}}[(\lambda - \langle \lambda_{\pi_N^r} \rangle)^2]\propto N^{-5/3}\)

Some perspectives

  • Many theoretical questions are open:
    • Ergodicity of the Norton dynamics
    • Equivalence of linear responses
    • Equivalence of nonequilibrium ensembles
    • Forcing fluctuations in the Norton ensemble, anomalous scaling
  • Some ongoing work by Shiva Darshan & Gabriel Stoltz.
  • Application to DPD: Xiaocheng Shang & Xinyi Wu (next talk!).

Some bibliography

Blassel, N., and G. Stoltz. 2024. “Fixing the Flux: A Dual Approach to Computing Transport Coefficients.” Journal of Statistical Physics 191.
Evans, D. J. 1993. “The Equivalence of Norton and Thévenin Ensembles.” Molecular Physics 80.
Evans, D. J., W. G. Hoover, B. H. Failor, B. Moran, and A. J. C. Ladd. 1983. “Nonequilibrium Molecular Dynamics via Gauss’s Principle of Least Constraint.” Physical Review A 28 (2).
Evans, D. J., and G. P. Morriss. 2007. Statistical Mechanics of Nonequilbrium Liquids. ANU Press.
Wu, X., and X. Shang. 2025. “Stochastic Norton Dynamics: An Alternative Approach for the Computation of Transport Coefficients in Dissipative Particle Dynamics.” arXiv:2504.14479.