This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Gentlest ascent dynamics on manifolds defined by adaptively sampled point-clouds

Juan M. Bello-Rivas [email protected] [    Anastasia Georgiou [    Hannes Vandecasteele [    Ioannis G. Kevrekidis [email protected] [
Abstract

Finding saddle points of dynamical systems is an important problem in practical applications such as the study of rare events of molecular systems. Gentlest ascent dynamics (GAD) 1 is one of a number of algorithms in existence that attempt to find saddle points in dynamical systems. It works by deriving a new dynamical system in which saddle points of the original system become stable equilibria. GAD has been recently generalized to the study of dynamical systems on manifolds (differential algebraic equations) described by equality constraints 2 and given in an extrinsic formulation. In this paper, we present an extension of GAD to manifolds defined by point-clouds, formulated using the intrinsic viewpoint. These point-clouds are adaptively sampled during an iterative process that drives the system from the initial conformation (typically in the neighborhood of a stable equilibrium) to a saddle point. Our method requires the reactant (initial conformation), does not require the explicit constraint equations to be specified, and is purely data-driven.

JHU]Department of Chemical and Biomolecular Engineering, Whiting School of Engineering, Johns Hopkins University, 3400 North Charles Street, Baltimore, 21218, MD, USA JHU]Department of Chemical and Biomolecular Engineering, Whiting School of Engineering, Johns Hopkins University, 3400 North Charles Street, Baltimore, 21218, MD, USA KULeuven]Department of Computer Science, KU Leuven, Celestijnenlaan 200A, 3001 Leuven JHU]Department of Chemical and Biomolecular Engineering, Whiting School of Engineering, Johns Hopkins University, 3400 North Charles Street, Baltimore, 21218, MD, USA \alsoaffiliationDepartments of Applied Mathematics and Statistics, Johns Hopkins University, 3400 North Charles Street, Baltimore, 21218, MD, USA

{tocentry}[Uncaptioned image]

1 Introduction

The problem of finding saddle points of dynamical systems has one of its most notable applications in the search for transition states of chemical systems described at the atomistic level, since saddle points coincide with transition states 3 at the zero temperature limit. While finding local stable equilibria (sinks) is a relatively straightforward matter, finding saddle points is a more complicated endeavor for which a number of algorithms have been presented in the literature 4.

Saddle point search methods can be classified according to whether they require one single input state (usually the reactant, located at the minimum of a free energy well) or two states (reactant and product). The Gentlest Ascent Dynamics (GAD) 1 belongs to the class of methods requiring a single reactant state as input and its applicability has been demonstrated with atomistic chemical systems 5, 6. GAD can be regarded as a variant of the dimer method that is formulated as a continuous dynamical system whose integral curves with initial condition at the reactant state can lead to saddle points. Variants of GAD such as high-index saddle dynamics (HiSD) have been the subject of recent research efforts and applications 7, 8, 9, 2, 10, 11, 12, 13.

While many search schemes attempt to find an optimal path between reactant and product (or between reactant state and transition state), it is interesting that there exist continuous curves joining the desired states in a variety of ways: following isoclines 14, 15, gradient extremals 16, 17, and the GAD studied in this paper. In most cases the study of these curves has been carried out in the Euclidean setting with some exceptions on the manifold of internal coordinates 18, 6 and on manifolds defined by the zeros of smooth maps 2 (in these cases, the algorithms are formulated extrinsically on the ambient space). Our contribution is formulated intrinsically and is valid on arbitrary manifolds, not necessarily explicitly defined by an atlas or by the zeros of maps. Importantly, algorithms like the one presented here or our previous work 19 do not rely on a priori knowledge of good collective coordinates, but rather use manifold learning to find them on the fly. In our method, there is a feedback loop of data collection that drives progress towards a saddle point.

In this paper, we study an application of the GAD to manifolds defined by point-clouds. The manifold does not need to be characterized in advance either by the zeros of a smooth function or by an atlas, and it is only assumed that the user is capable of sampling the vicinity of arbitrary points on the manifold (e.g., umbrella sampling based on reduced local coordinates). The algorithm uses dimensionality reduction (namely, diffusion map coordinates 20) to define a dynamical system intrinsically on the reduced coordinates that can lead to a saddle point. Since the saddle point in general is not expected to lie in the vicinity of the reactant, our algorithm works by iteratively sampling the manifold on the fly, resolving the path on the local chart, and repeatedly switching charts until convergence. Our approach shares algorithmic elements with our previous work 19 which, however, instead of GAD dynamics on manifolds, was following isoclines on manifolds.

2 Gentlest Ascent Dynamics and Idealized Saddle Dynamics

Gentlest Ascent Dynamics (GAD) 1 is an algorithm for finding saddle points of dynamical systems. We propose an extension of GAD to manifolds defined by point-clouds that finds saddle points combining nonlinear dimensionality reduction and adaptive sampling.

Let U:nU\colon\mathbb{R}^{n}\to\mathbb{R} be a smooth potential energy function and consider the associated gradient vector field XX in n\mathbb{R}^{n} given by x˙=X(x)\dot{x}=X(x), where X=DUX=-DU with DD denoting the gradient (or, equivalently, the Jacobian matrix). We restrict ourselves here, for the sake of simplicity, to the case of gradient systems and index-1 saddle points. The GAD algorithm consists of integrating the equations of motion of the related dynamical system X^\hat{X} on an extended phase space 2n\mathbb{R}^{2n} given by

{x˙=H(v)DU(x),v˙=D2U(x)v+r(x,v)v,\left\{\begin{aligned} &\dot{x}=-H(v)DU(x),\\ &\dot{v}=-D^{2}U(x)v+r(x,v)v,\end{aligned}\right. (1)

where x,vnx,v\in\mathbb{R}^{n}, H(v)w=w(2vw)vH(v)w=w-(2\,v\!\cdot\!w)v is the Householder reflection 21 of wnw\in\mathbb{R}^{n} across the hyperplane v={znvz=0}\langle v\rangle^{\perp}=\{z\in\mathbb{R}^{n}\mid v\cdot z=0\} and r(x,v)=v2vD2U(x)vr(x,v)=\|v\|^{-2}\,v\cdot D^{2}U(x)v is the Rayleigh quotient of the Hessian matrix D2U(x)D^{2}U(x) corresponding to the vector vnv\in\mathbb{R}^{n}.

Remark 1.

The right hand side of the ordinary differential equation for v˙\dot{v} is the term

D2U(x)v+r(x,v)v.-D^{2}U(x)v+r(x,v)v.

If vv is an eigenvector of D2U(x)D^{2}U(x) with eigenvalue λ\lambda, then r(x,v)=λr(x,v)=\lambda. Now note that the constrained optimization problem consisting of finding the extrema of Z(x)=12DU(x)2Z(x)=\frac{1}{2}\|-DU(x)\|^{2} along the level sets C={xnU(x)=c}C=\{x\in\mathbb{R}^{n}\mid U(x)=c\} is precisely given by the Lagrange equation

D2U(x)DU(x)λDU(x)=0,D^{2}U(x)DU(x)-\lambda DU(x)=0,

which says that the extrema of the magnitude of the gradient, ZZ, along CC are attained wherever the gradient field X=DUX=-DU happens to be an eigenvector of the Hessian D2UD^{2}U.

The intuition behind GAD is that the force DU-DU can be written in a basis determined by the eigenvectors of the Hessian D2UD^{2}U, which gives us the stable and unstable directions in the vicinity of an index-1 saddle point. The second differential equation in (1) acts as a (continuous) eigensolver yielding the unstable direction. Then, we can flip the sign of the component of the force corresponding to the unstable direction to make the resulting vector point towards the saddle point (see Figure 1).

Refer to caption
Figure 1: Vector field corresponding to the first equation of GAD on the potential U(x1,x2)=(x1)2(x2)2U(x^{1},x^{2})=(x^{1})^{2}-(x^{2})^{2} with v=(0,1)2v=(0,1)\in\mathbb{R}^{2} being the unstable direction. Note how the steepest descent direction is reflected across the orthogonal complement v\langle v\rangle^{\perp} to obtain x˙\dot{x}.

2.1 Idealized saddle-point dynamics

We consider a variant of the GAD algorithm named Idealized Saddle Dynamics (ISD) 22, given by

x˙=H(v)DU(x),\dot{x}=-H(v)DU(x),

where vnv\in\mathbb{R}^{n} is an eigenvector of D2U(x)D^{2}U(x) corresponding to the smallest eigenvalue λ\lambda.

We now leave the Euclidean setting behind and study the Riemannian setting (see the appendix for a summary of the topic and the notation). Let MM be a dd-dimensional smooth manifold with Riemannian metric gg and let UC(M)U\in C^{\infty}(M) be a potential energy function. We consider the gradient field X=gradU𝔛(M)X=-\operatorname{grad}U\in\mathfrak{X}(M) and we write the following equation for the ISD vector field on MM,

X^=X2g(V,X)V𝔛(M),\hat{X}=X-2g(V,X)V\in\mathfrak{X}(M), (2)

where VV is the vector field on MM defined by choosing the eigenvector (normalized so that g(V,V)=1g(V,V)=1) corresponding to the smallest eigenvalue of the Hessian matrix. However, in this case the Hessian must be defined in terms of the covariant derivative on (M,g)(M,g) with respect to the Levi-Civita connection 23, 24. To be precise, given two vector fields S,T𝔛(M)S,T\in\mathfrak{X}(M), we have

2U(S,T)=T(SU)TSU,\nabla^{2}U(S,T)=\nabla_{T}(\nabla_{S}U)-\nabla_{\nabla_{T}S}U,

which is a tensor of type (0,2)(0,2), where \nabla denotes the covariant derivative induced by the Levi-Civita connection (the notation TS\nabla_{T}S represents the covariant derivative of SS in the direction of TT). We apply the sharp (\sharp) isomorphism (raising or lowering indices) to turn 2U\nabla^{2}U into a (1,1)(1,1)-tensor. Therefore,

(HessU)T=TgradU.(\operatorname{Hess}U)T=\nabla_{T}\operatorname{grad}U.

The eigenvector corresponding to the lowest eigenvalue of HessU\operatorname{Hess}U at each point pMp\in M induces a vector field VV whose integral curves are curves that may join an initial point with a saddle point.

The ISD formulation of GAD is particularly amenable to be coupled with dimensionality reduction approaches because the resulting eigenproblem is often of much lower dimensionality than the ambient space in which the original dynamical system is defined.

Example 1 (Exact solution of a model system).

Let us compute a concrete case of ISD, first with known exact formulas and later on with approximations using diffusion maps and Gaussian processes. Consider the sphere

𝕊2={(x1,x2,x3)3(x1)2+(x2)2+(x3)2=1}.\mathbb{S}^{2}=\{(x^{1},x^{2},x^{3})\in\mathbb{R}^{3}\mid(x^{1})^{2}+(x^{2})^{2}+(x^{3})^{2}=1\}.

and the stereographic projection from the North pole onto the tangent plane at the South pole. The system of coordinates is given by

ϕ(x1,x2,x3)=(x11x3,x21x3)2,\phi(x^{1},x^{2},x^{3})=(\frac{x^{1}}{1-x^{3}},\frac{x^{2}}{1-x^{3}})\in\mathbb{R}^{2},

Let (u1,u2)2(u^{1},u^{2})\in\mathbb{R}^{2}. The corresponding parameterization, ψ=ϕ1\psi=\phi^{-1}, is given by

ψ(u1,u2)=11+(u1)2+(u2)2(2u1,2u2,(u1)2+(u2)21).\psi(u^{1},u^{2})=\frac{1}{1+(u^{1})^{2}+(u^{2})^{2}}(2u^{1},2u^{2},(u^{1})^{2}+(u^{2})^{2}-1).

The pullback of the Euclidean metric h=dx1dx1+dx2dx2+dx3dx3h=\mathrm{d}x^{1}\otimes\mathrm{d}x^{1}+\mathrm{d}x^{2}\otimes\mathrm{d}x^{2}+\mathrm{d}x^{3}\otimes\mathrm{d}x^{3} by ψ\psi gives us the metric gg

g=ψh=4du1du1+4du2du2(1+(u1)2+(u2)2)2g=\psi^{\star}h=\frac{4\mathrm{d}u^{1}\otimes\mathrm{d}u^{1}+4\mathrm{d}u^{2}\otimes\mathrm{d}u^{2}}{\left(1+(u^{1})^{2}+(u^{2})^{2}\right)^{2}}

The non-redundant Christoffel symbols Γijk\Gamma_{ij}^{k} that characterize the Levi-Civita connection \nabla are

Γ1,11=2u11+(u1)2+(u2)2,Γ1,21=2u21+(u1)2+(u2)2,\Gamma_{1,1}^{1}=\frac{-2u^{1}}{1+(u^{1})^{2}+(u^{2})^{2}},\quad\Gamma_{1,2}^{1}=\frac{-2u^{2}}{1+(u^{1})^{2}+(u^{2})^{2}},
Γ2,21=Γ1,11,Γ1,12=Γ1,21,Γ1,22=Γ1,11,andΓ2,22=Γ1,21.\Gamma_{2,2}^{1}=-\Gamma_{1,1}^{1},\quad\Gamma_{1,1}^{2}=-\Gamma_{1,2}^{1},\quad\Gamma_{1,2}^{2}=\Gamma_{1,1}^{1},\quad\text{and}\quad\Gamma_{2,2}^{2}=\Gamma_{1,2}^{1}.

The energy E(x1,x2,x3)=x1x2x3E(x^{1},x^{2},x^{3})=x^{1}x^{2}x^{3}, constrained on 𝕊23\mathbb{S}^{2}\subset\mathbb{R}^{3} is transformed to U=ψEU=\psi^{\star}E on 𝕊2\mathbb{S}^{2},

U(u1,u2)=4u1u2((u1)2+(u2)21)((u1)2+(u2)2+1)3,U(u^{1},u^{2})=\frac{4u^{1}u^{2}\left((u^{1})^{2}+(u^{2})^{2}-1\right)}{\left((u^{1})^{2}+(u^{2})^{2}+1\right)^{3}},

The force on 𝕊2\mathbb{S}^{2} is the negative of the gradient

gradU=((u2)52(u1)2(u2)3+(3(u1)4+8(u1)21)u2(u2)4+(2(u1)2+2)(u2)2+(u1)4+2(u1)2+1)u1+(3u1(u2)4+(2(u1)38u1)(u2)2(u1)5+u1(u2)4+(2(u1)2+2)(u2)2+(u1)4+2(u1)2+1)u2\operatorname{grad}U=\left({{(u^{2})^{5}-2\,(u^{1})^{2}\,(u^{2})^{3}+\left(-3\,(u^{1})^{4}+8\,(u^{1})^{2}-1\right)\,{u^{2}}}\over{(u^{2})^{4}+\left(2\,(u^{1})^{2}+2\right)\,(u^{2})^{2}+(u^{1})^{4}+2\,(u^{1})^{2}+1}}\right)\,\frac{\partial}{\partial{u^{1}}}\\ +\left(-{{3\,{u^{1}}\,(u^{2})^{4}+\left(2\,(u^{1})^{3}-8\,{u^{1}}\right)\,(u^{2})^{2}-(u^{1})^{5}+{u^{1}}}\over{(u^{2})^{4}+\left(2\,(u^{1})^{2}+2\right)\,(u^{2})^{2}+(u^{1})^{4}+2\,(u^{1})^{2}+1}}\right)\,\frac{\partial}{\partial{u^{2}}}

The potential energy and the gradient field are shown in Figure 2(a).

Refer to caption
(a) Vector field corresponding to the potential energy function considered in the text.
Refer to caption
(b) Idealized saddle dynamics vector field corresponding to the original vector field. Notice how saddle points in the original vector field become sinks (stable equilibria) of the ISD vector field.
Figure 2: Vector fields on the sphere, seen in the stereographic projection from the North pole onto the tangent plane to the South pole. The equilibria of the original vector field appear represented as \bullet for sinks, \blacksquare for saddle points, and \blacktriangle for sources.

The Hessian is then given by the (1,1)(1,1) tensor

gradU=Au1du1+Bu2du1+Bu1du2+Du2du2,\nabla\operatorname{grad}U=A\frac{\partial}{\partial{u^{1}}}\otimes\mathrm{d}u^{1}+B\frac{\partial}{\partial{u^{2}}}\otimes\mathrm{d}u^{1}+B\frac{\partial}{\partial{u^{1}}}\otimes\mathrm{d}u^{2}+D\frac{\partial}{\partial{u^{2}}}\otimes\mathrm{d}u^{2},

where

A\displaystyle A =4u1u2((u2)4+(u2)2(u1)4+11(u1)26)((u2)2+(u1)2+1)3\displaystyle=-{{4\,{u^{1}}\,{u^{2}}\,\left((u^{2})^{4}+(u^{2})^{2}-(u^{1})^{4}+11\,(u^{1})^{2}-6\right)}\over{\left((u^{2})^{2}+(u^{1})^{2}+1\right)^{3}}}
B\displaystyle B =(u2)65(u1)2(u2)45(u2)45(u1)4(u2)2+30(u1)2(u2)25(u2)2+(u1)65(u1)45(u1)2+1((u2)2+(u1)2+1)3\displaystyle=-{{(u^{2})^{6}-5\,(u^{1})^{2}\,(u^{2})^{4}-5\,(u^{2})^{4}-5\,(u^{1})^{4}\,(u^{2})^{2}+30\,(u^{1})^{2}\,(u^{2})^{2}-5\,(u^{2})^{2}+(u^{1})^{6}-5\,(u^{1})^{4}-5\,(u^{1})^{2}+1}\over{\left((u^{2})^{2}+(u^{1})^{2}+1\right)^{3}}}
D\displaystyle D =4u1u2((u2)411(u2)2(u1)4(u1)2+6)((u2)2+(u1)2+1)3.\displaystyle={{4\,{u^{1}}\,{u^{2}}\,\left((u^{2})^{4}-11\,(u^{2})^{2}-(u^{1})^{4}-(u^{1})^{2}+6\right)}\over{\left((u^{2})^{2}+(u^{1})^{2}+1\right)^{3}}}.

The smallest eigenvector of gradU\nabla\operatorname{grad}U determines the vector field V𝔛(𝕊2)V\in\mathfrak{X}(\mathbb{S}^{2}) that is used in the formulation of the ISD vector field (2) and is shown in Figure 2(b).

Example 1 required the knowledge of a particular system of coordinates (namely, the stereographic projection from the North pole to the tangent space at the South pole) mapping three-dimensional points on the sphere to two-dimensional coordinates; here this allows us to work with a single chart. In some settings the system of coordinates of the underlying manifold is either unknown or difficult to obtain. It is possible under those circumstances to replicate the steps in Example 1 in the absence of a system of coordinates given as a closed-form expression by resorting to manifold learning / dimensionality reduction techniques. In our case, as we shall discuss next, we use diffusion maps on points sampled from a local neighborhood of the manifold to extract a suitable system of coordinates. Fitting a Gaussian process to the diffusion map coordinates of the point-cloud yields a local system of coordinates ϕ\phi that can be evaluated at arbitrary points (not necessarily those in the sampled point-cloud). Once we have a system of coordinates, we can again proceed to estimate the Riemannian metric as well as the Levi-Civita connection, and compute the flow of the ISD vector field X^\hat{X} given in (2) to find saddle points.

Revisiting Example 1 and approaching it with the aforementioned procedure allows us to obtain trajectories that lead to saddle points, as shown in Figures 3 to 5.

The choice of diffusion map coordinates for dimensionality reduction and Gaussian processes for non-linear regression is inessential and both methods could be replaced by, for instance, neural networks (e.g., a variational autoencoder and a graph neural network, respectively). An important aspect of our approach is the fact that it operates on local neighborhoods of points, as opposed to using data from longtime simulations, and only then applying dimensionality reduction techniques to extract global collective variables. From a geometric viewpoint, this is motivated by the fact that a sufficiently small neighborhood of a manifold can always be transformed onto a subset of its tangent space by a smooth invertible map. From a physical viewpoint, different sets of collective variables govern different stages of a reaction (e.g., the distance between a ligand and a receptor is important when both molecules are far away whereas their relative orientation might be more important to ligand-binding, when both molecules become close).

One interesting aspect of diffusion maps is its relation to the infinitesimal generator of a diffusion process 25, which in turn has connections to the committor function, which is an optimal reaction coordinate 26, 27, 28. We conclude this discussion on collective variables by pointing out that a priori knowledge of good reaction coordinates (some recent examples can be found in 29, 30) can often be put in one-to-one correspondence with diffusion map coordinates 31, 32.

2.2 Mean force

In computational statistical mechanics we are often interested in dynamics on Riemannian manifolds endowed with a probability distribution. For instance, for simulations at constant temperature, we use the Boltzmann distribution with probability density proportional to exp{βU(x)}\exp\{-\beta U(x)\}, where β>0\beta>0 is the inverse temperature. The relevant vector field at uu in the local chart is the mean force 33, 34, 35, given by

Xu=grad(Uψ12β1logdet(DψDψ))u,X_{u}=-\langle\operatorname{grad}\left(U\circ\psi-\tfrac{1}{2}\beta^{-1}\log\det(D\psi^{\top}D\psi)\right)\rangle_{u},

where u\langle\cdot\rangle_{u} denotes the ensemble average with respect to the Boltzmann distribution, conditioned on {xnϕ(x)=u}\{x\in\mathbb{R}^{n}\mid\phi(x)=u\}. There are a number of numerical methods that estimate the mean force as a step in their calculations. Adaptive biasing force 36 and the string method 37 are examples of those.

3 Algorithm

Consider the manifold MnM\subset\mathbb{R}^{n} and the gradient dynamical system X𝔛(M)X\in\mathfrak{X}(M). We begin by drawing a total of NN samples from the manifold MM in the neighborhood of an initial point pMp\in M. This can be done in a variety of ways depending on the application. If MM is the inertial manifold of a dynamical system XX, then a reasonable way to approach this problem is to generate NN distinct perturbations {p(i)ni=1,,N}\{p_{(i)}\in\mathbb{R}^{n}\mid i=1,\dotsc,N\} of pp and propagate them according to the flow of XX during a (short) time horizon τ>0\tau>0. Doing so, we obtain a data set 𝒟={q(i)=expτp(i)Mi=1,,N}\mathscr{D}=\{q_{(i)}=\exp_{\tau}p_{(i)}\in M\mid i=1,\dotsc,N\} approximately on the manifold. Alternatively, one may numerically solve a stochastic differential equation such as the Brownian dynamics equation,

dqt=X(qt)dt+σdBt,\mathrm{d}q_{t}=X(q_{t})\,\mathrm{d}t+\sigma\,\mathrm{d}B_{t}, (3)

where the drift is the vector field XX, σ>0\sigma>0 is a constant, and BtB_{t} is a standard nn-dimensional Brownian motion. Solving (3) (possibly with an added RMSD-based restraint around the initial conformation) up to a certain time τ>0\tau>0 and extracting an uncorrelated subset of the states at different time steps yields a data set 𝒟={q(i)=qtii=1,,N,0t1tNτ}\mathscr{D}=\{q_{(i)}=q_{t_{i}}\mid i=1,\dotsc,N,0\leq t_{1}\leq\cdots\leq t_{N}\leq\tau\}.

Next, we apply a dimensionality reduction algorithm on the data set 𝒟\mathscr{D} to obtain a set of reduced coordinates ϕ\phi. In our case, we use diffusion maps 20, 25 to obtain a set of vectors ϕ(i)d\phi_{(i)}\in\mathbb{R}^{d} with dnd\leq n but other methods, such as local tangent space alignment 38, may be used as well. It is important to note that our dimensionality reduction method is applied to a local neighborhood of an initial point and, therefore, it is expected to yield a reasonable approximation to a chart on that neighborhood.

We fit a Gaussian process regressor ϕ\phi to the pairs of points (q(i),ϕ(i))n×d(q_{(i)},\phi_{(i)})\in\mathbb{R}^{n}\times\mathbb{R}^{d} to obtain a smooth map ϕ:Mnd\phi\colon M\subset\mathbb{R}^{n}\to\mathbb{R}^{d} that will act as a system of coordinates (in particular, ϕ(q(i))ϕ(i)\phi(q_{(i)})\approx\phi_{(i)}). Proceeding in an analogous fashion, we compute the inverse mapping ψ=ϕ1\psi=\phi^{-1}.

Note that one possible way of estimating the dimensionality is by computing the average of the approximate rank of the Jacobian matrix of ϕ\phi at (a subset of) the data points and retaining the components that yield a local chart.

Remark 2.

We can reduce the computational expense of the Gaussian process regression by reusing the kernel matrix with entries eq(i)q(j)2/2ε\mathrm{e}^{-\|q_{(i)}-q_{(j)}\|^{2}/2\varepsilon} (for some ε>0\varepsilon>0) calculated during the computation of the diffusion map coordinates as the covariance matrix for the Gaussian process (assuming that it is formulated using the squared exponential kernel).

Remark 3.

It is not always possible to obtain a suitable Gaussian process regressor for ψ=ϕ1\psi=\phi^{-1}. An alternative is to add an Ornstein-Uhlenbeck process to the stochastic differential equation (3) in order to obtain

dqt=(X(qt)κ(ϕ(qt)ϕ0))dt+σdBt,\mathrm{d}q_{t}=\left(X(q_{t})-\kappa(\phi(q_{t})-\phi_{0})\right)\,\mathrm{d}t+\sigma\,\mathrm{d}B_{t},

where κ>0\kappa>0 is a hyperparameter and ϕ0\phi_{0} is a prescribed point not necessarily in ϕ(𝒟)\phi(\mathscr{D}). Computing the ensemble average qt\langle q_{t}\rangle of the solution to the above equation yields a point q(0)q_{(0)} such that ϕ(q(0))ϕ0\phi(q_{(0)})\approx\phi_{0} or, equivalently, q(0)=ϕ1(ϕ(0))q_{(0)}=\phi^{-1}(\phi_{(0)}). This works because the new term added to the drift nudges the system towards a point in ambient space such that its image by ϕ\phi is the prescribed point ϕ0\phi_{0}.

We consider the values Xq(i)TMX_{q_{(i)}}\in TM of the vector field XX at the points in the data set 𝒟\mathscr{D} and map them via the system of coordinates in order to obtain the vector field ϕXq(i)=Dϕ(q(i))Xq(i)\phi_{\star}X_{q_{(i)}}=D\phi(q_{(i)})X_{q_{(i)}} in the new coordinates, where DϕD\phi is the Jacobian matrix of ϕ\phi (note that the Jacobian-vector product can be computed either as a closed-form formula or efficiently using automatic differentiation —e.g., using jvp in JAX 39). We fit another Gaussian process to the pushforward vector field ϕX\phi_{\star}X in order to be able to evaluate it at arbitrary points in the new coordinates and we abuse notation in what follows by also referring to ϕX\phi_{\star}X as XX.

At this stage, we can readily compute the Riemannian metric gg as

g=i,j=1dgijdϕidϕj,g=\sum_{i,j=1}^{d}g_{ij}\,\mathrm{d}\phi^{i}\otimes\mathrm{d}\phi^{j}, (4)

where gij=DψDψg_{ij}=D\psi^{\top}D\psi for i,j=1,,di,j=1,\dotsc,d. The metric induces an inner product, denoted by the bracket ,\langle\cdot,\cdot\rangle, between tangent vectors such that if S=i=1dSiϕiS=\sum_{i=1}^{d}S^{i}\frac{\partial}{\partial\phi^{i}} and T=i=1dTiϕiT=\sum_{i=1}^{d}T^{i}\frac{\partial}{\partial\phi^{i}} are the expressions in local coordinates of two tangent vectors S,TTuMS,T\in T_{u}M at a point uMu\in M, then S,T=i,j=1dgijSiTi\langle S,T\rangle=\sum_{i,j=1}^{d}g_{ij}S^{i}T^{i}.

Using (4), we obtain the coefficients of the Levi-Civita connection 23,

Γjk=i=1dgi(gijϕk+gikϕjgjkϕi)\Gamma^{\ell}_{jk}=\sum_{i=1}^{d}g^{\ell i}\left(\frac{\partial g_{ij}}{\partial\phi^{k}}+\frac{\partial g_{ik}}{\partial\phi^{j}}-\frac{\partial g_{jk}}{\partial\phi^{i}}\right)

for j,k,{1,,d}j,k,\ell\in\{1,\dotsc,d\}, where gijg^{ij} denotes the entries of the inverse of the matrix with components gijg_{ij}. This, in turn, allows us to take the covariant derivative and the Hessian, defined by

HessU(S,T)=TgradU,S\operatorname{Hess}U(S,T)=\langle\nabla_{T}\operatorname{grad}U,S\rangle

for arbitrary tangent vectors S,TS,T. Observe that the Hessian is defined on the local chart, not on the ambient space. The eigenvector VV, normalized with respect to gg, with smallest eigenvalue of the Hessian then yields a vector field

X^=X2Vg(V,X)\hat{X}=X-2\,Vg(V,X)

such that the index-1 saddle points of XX become stable equilibria 1 of X^\hat{X}. Consequently, an integral curve of X^\hat{X} in the vicinity of a saddle point leads to the saddle point.

In general, in order to carry out the computation until convergence, we must frequently switch charts. This is due to the fact that at each step, we sample a point-cloud 𝒟\mathscr{D} in a small neighborhood of a given point and the Gaussian process regressor ϕ\phi yields an approximation to XX on the chart that cannot extrapolate far away from the sampled points. Therefore, when we reach the confines of 𝒟\mathscr{D} (which can be determined by the density of points), we ought to map the latest point of our integral curve of X^\hat{X} from the chart back to the ambient space using ψ=ϕ1\psi=\phi^{-1}, as discussed earlier. After that, we start the whole procedure again: sample a new data set 𝒟\mathscr{D}, compute ϕ\phi, trace an integral curve of X^\hat{X}, etc.

Remark 4.

An alternative approach to the one presented here could consist of exploiting the fact that GAD trajectories are geodesics 40 of a Finsler metric 41, 42 and to numerically compute said geodesics in each learned local chart in a similar manner to what we propose.

The factor that impacts the algorithm the most is the deviation of the sampled points in the data set from the manifold MM. In other words, the farther away from MM our data points are, the more noisy is the estimate of X^\hat{X}, and, consequently, the harder it is to find the saddle points.

3.1 Numerical examples

The computations that follow were carried out using the JAX 39 and Diffrax 43 libraries, and the code to reproduce our results is available at https://github.com/jmbr/gentlest_ascent_dynamics_on_manifolds.

In our two examples below, we automatically set the bandwidth parameter (usually denoted by ε\varepsilon) in diffusion maps to be equal to the squared median of the pairwise distances between points. The regularization parameters in the Gaussian processes ϕ\phi and ψ\psi are chosen so that their scores are sufficiently close to 1 on a test set.

3.1.1 Sphere

The preceding algorithm applied to the vector field on the sphere 𝕊2\mathbb{S}^{2} introduced in Example 1 produces the iterations shown in Figures 3, 4, and 5. These figures depict the integral curves (highlighted) in the local neighborhoods obtained by sampling and integrating X^\hat{X}. The full integral curve joining the initial point to a saddle point at the equator of the sphere in Figure 6.

We sampled 10310^{3} points per iteration and integrated the ISD vector field using an explicit Euler integrator for a total 10310^{3} steps with a time-step length of 10410^{-4}. The algorithm converges to a saddle point in ten iterations.

Refer to caption
Figure 3: Iteration #1 of ISD on the sphere by sampling point-clouds and using diffusion maps. The points and the vector field are drawn from the neighborhood of a stable equilibrium. The solid curve represents the GAD/ISD path.
Refer to caption
Figure 4: Iteration #5 of ISD on the sphere by sampling point-clouds and using diffusion maps.
Refer to caption
Figure 5: Iteration #10 of ISD on the sphere by sampling point-clouds and using diffusion maps. The algorithm has reached a saddle point of the original vector field (notice how it becomes a sink for the ISD vector field).
Refer to caption
Figure 6: Resulting trajectory (in orange) of GAD/ISD on the sphere (color indicates value of the potential energy function) using diffusion maps to navigate from the sink (bottom right) to the saddle (center). The insets depict the corresponding vector fields in local coordinates at different iterations and also the GAD curve (orange). See also Figures 3, 4, and 5.

3.1.2 Regular surface

Our second example is the Müller-Brown (MB) potential 44 mapped onto a regular surface. Namely, the manifold MM is the graph of the function (see Figure 7),

f(x1,x2)=k1=0Kk2=0Kak1,k2cos(k1x1+k2x2+bk1,k2),f(x^{1},x^{2})=\sum_{k^{1}=0}^{K}\sum_{k^{2}=0}^{K}a_{k^{1},k^{2}}\cos(k^{1}x^{1}+k^{2}x^{2}+b_{k^{1},k^{2}}),

where K=3K=3 and the coefficients are given in Table 1.

k1k^{1} k2k^{2} ak1,k2a_{k^{1},k^{2}} bk1,k2b_{k^{1},k^{2}}
0 1 0.9490 0.8838
0 2 0.4575 0.6564
1 0 0.4152 0.7449
1 2 0.2911 0.3619
2 0 0.4121 0.5469
3 2 0.2817 0.4719
Table 1: Non-zero coefficients characterizing the regular surface.

After eight iterations, sampling 5×1035\times 10^{3} points and integrating for 10310^{3} steps per iteration with a time step length of 10410^{-4}, our algorithm successfully constructs a path joining the initial point located at a minimum (sink) of the MB potential to the nearest saddle point. The relative errors between the points in the constructed path and the known coordinates of the saddle point are shown in Figure 8.

Refer to caption
Figure 7: Path (in pink) from sink (right endpoint) to saddle point (upper endpoint) obtained by eight iterations of GAD on the Müller-Brown potential (blue heat map) defined on a regular surface.
Refer to caption
Figure 8: Relative error between each point γn\gamma_{n} in the GAD path and the saddle point xx_{\star}. Each iteration comprises 1000 points.

4 Conclusions

We have presented a formulation of GAD on manifolds defined by point-clouds that are meant to be sampled on-demand. Our formulation is intrinsic and does not require the specification of the manifolds either by a given atlas or by the zeros of a smooth map. The required charts are discovered through a data-driven, iterative process that only requires knowledge of an initial conformation of the reactant. We illustrated our approach with two simple examples and we expect the results to transfer to the high-dimensional dynamical systems of interest in computational statistical mechanics.

{acknowledgement}

This work was supported by the US Air Force Office of Scientific Research (AFOSR) and the US Department of Energy DOE with IIT: SA22-0052-S001 and AFOSR-MURI: FA9550-21-1-0317.

Appendix A Appendix

In this section we review basic notions of Riemannian geometry relevant to this paper. Our aim is to emphasize intuition over formalism as much as possible. We refer the reader to general treatises on the topic such as 23, 45 for a deeper presentation and we especially recommend the relevant material in 46, 47 for the working physical chemist.

A.1 Tensor spaces

Let VV and WW be two vector spaces over the real numbers and denote by 𝒱(V×W)\mathscr{V}(V\times W) the vector space generated by all finite linear combinations of elements of the Cartesian product V×WV\times W. The tensor product VWV\otimes W is a vector subspace of 𝒱(V×W)\mathscr{V}(V\times W) with elements of the form vwv\otimes w, where vVv\in V, wWw\in W, such that:

  1. 1.

    (v1+v2)w=v1w+v2w(v_{1}+v_{2})\otimes w=v_{1}\otimes w+v_{2}\otimes w, where v1,v2Vv_{1},v_{2}\in V.

  2. 2.

    v(w1+w2)=vw1+vw2v\otimes(w_{1}+w_{2})=v\otimes w_{1}+v\otimes w_{2}, where w1,w2Ww_{1},w_{2}\in W.

  3. 3.

    (αv)w=αvw(\alpha v)\otimes w=\alpha\,v\otimes w, where α\alpha\in\mathbb{R}.

  4. 4.

    v(αw)=αvwv\otimes(\alpha w)=\alpha\,v\otimes w.

The above construction can be recursively extended to tensor spaces of arbitrary numbers of factors (i.e., given vector spaces V1,,VnV_{1},\dotsc,V_{n}, with nn\in\mathbb{N}, we construct V1V2VnV_{1}\otimes V_{2}\otimes\dotsb\otimes V_{n} as V1(V2Vn)V_{1}\otimes(V_{2}\otimes\dotsb\otimes V_{n}) and so on and so forth).

Given a vector space VV, its dual space, denoted by VV^{\star}, is the vector space formed by all linear maps f:Vf\colon V\to\mathbb{R}. A tensor space of type (p,q)(p,q) is a tensor product

Tqp(V)=(i=1pV)(i=1qV).T_{q}^{p}(V)=\left(\bigotimes_{i=1}^{p}V\right)\otimes\left(\bigotimes_{i=1}^{q}V^{\star}\right). (5)

Elements of Tqp(V)T_{q}^{p}(V) are called tensors of type (p,q)(p,q).

Similarly to how all vector spaces of fixed dimension over the real numbers are linearly isomorphic to each other, all tensor spaces of type (p,q)(p,q) over the reals are linearly isomorphic to each other and, therefore, we can assume that the factors are arranged as in (5).

Example 2.

The set of linear maps from V=nV=\mathbb{R}^{n} to W=mW=\mathbb{R}^{m} is a vector space. If {e1,,en}\{e_{1},\dotsc,e_{n}\} is an orthonormal basis of VV and {f1,,fm}\{f_{1},\dotsc,f_{m}\} is an orthonormal basis of WW, then {ei:Vi=1,,n}\{e^{i}\colon V\to\mathbb{R}\mid i=1,\dotsc,n\} is the basis of the dual space VV^{\star} determined by

ei(ej)={1,if i=j,0,otherwise.e^{i}(e_{j})=\begin{cases}1,&\text{if $i=j$,}\\ 0,&\text{otherwise}.\end{cases}

Consequently, the linear map A:VWA\colon V\to W can be written as

A=i=1nj=1maijfjei,A=\sum_{i=1}^{n}\sum_{j=1}^{m}a_{i}^{j}\,f_{j}\otimes e^{i},

where aija_{i}^{j}\in\mathbb{R}. Additionally, if {e1,en}\{e_{1},\dotsc e_{n}\} and {f1,,fm}\{f_{1},\dotsc,f_{m}\} are, respectively, the canonical bases of V=nV=\mathbb{R}^{n} and W=mW=\mathbb{R}^{m}, then fjeiWVf_{j}\otimes e^{i}\in W\otimes V^{\star} coincides with the outer product fjeim×nf_{j}e_{i}^{\top}\in\mathbb{R}^{m\times n} and AA coincides with the m×nm\times n real matrix

A=[a11a1nam1amn].A=\begin{bmatrix}a_{1}^{1}&\cdots&a_{1}^{n}\\ \vdots&\ddots&\vdots\\ a_{m}^{1}&\cdots&a_{m}^{n}\end{bmatrix}.

The wedge product is defined as

v1vk=1k!σSksgn(σ)vσ1(1)vσ1(k),v_{1}\wedge\dotsb\wedge v_{k}=\frac{1}{k!}\sum_{\sigma\in S_{k}}\operatorname{sgn}(\sigma)\,v_{\sigma^{-1}(1)}\otimes\dotsb\otimes v_{\sigma^{-1}(k)},

where SkS_{k} is the group of permutations of {1,,k}\{1,\dotsc,k\} and sgn(σ)\operatorname{sgn}(\sigma) is equal to 1-1 if the permutation σ\sigma is odd and +1+1 if it is even. This is equivalent to the projection of the tensor v1vkv_{1}\otimes\dotsb\otimes v_{k} onto the subspace ΛkV\Lambda^{k}V of anti-symmetric tensors of i=1kV\bigotimes_{i=1}^{k}V.

Remark 5.

The signature sgn(σ)\operatorname{sgn}(\sigma) coincides with the Levi-Civita symbol. The latter should not be confused with the Levi-Civita connection, to be introduced later.

A.2 Smooth manifolds and smooth curves

In what follows, we consider a smooth map to be a map that has sufficiently many derivatives and whose derivatives are continuous functions.

A chart is a tuple (U,ϕ)(U,\phi) such that UU is an open set in a topological space and ϕ:Um\phi\colon U\to\mathbb{R}^{m} is a diffeomorphism (i.e., a smooth mapping with a smooth inverse ψ1\psi^{-1}). A smooth manifold MM is a topological space together with a family of charts (i.e., an atlas),

{(Uα,ϕα)|ϕα:UαMm},\left\{(U_{\alpha},\phi_{\alpha})\,|\,\phi_{\alpha}\colon U_{\alpha}\subset M\to\mathbb{R}^{m}\right\},

such that for every pMp\in M there exists a chart (U,ϕ)(U,\phi) with pUp\in U. Moreover, for any other chart (V,χ)(V,\chi) such that pVp\in V, the function ϕχ1\phi\circ\chi^{-1} is smooth (see Figure 9).

\begin{overpic}[width=303.53267pt]{figures/charts.png} \put(45.0,45.0){\vector(-1,-1){15.0}} \put(55.0,45.0){\vector(1,-1){15.0}} \put(40.0,20.0){\vector(1,0){20.0}} \put(32.5,40.0){$\chi$} \put(65.0,40.0){$\phi$} \put(42.5,12.5){$\phi\circ\chi^{-1}$} \put(5.0,30.0){$\mathbb{R}^{m}$} \put(85.0,30.0){$\mathbb{R}^{m}$} \end{overpic}
Figure 9: Smooth 22-dimensional manifold (a torus) with two overlapping charts UU and VV and their corresponding systems of coordinates ϕ(U)\phi(U) and χ(V)\chi(V).
Example 3.

The sets

N={(x1,x2,x3)3|x3=f(x1,x2)=1(x1)2(x2)2}N=\left\{(x^{1},x^{2},x^{3})\in\mathbb{R}^{3}\,|\,x^{3}=f(x^{1},x^{2})=\sqrt{1-(x^{1})^{2}-(x^{2})^{2}}\right\}

and

𝕊2={(x1,x2,x3)3|F(x1,x2,x3)=0},\mathbb{S}^{2}=\left\{(x^{1},x^{2},x^{3})\in\mathbb{R}^{3}\,|\,F(x^{1},x^{2},x^{3})=0\right\},

where F(x1,x2,x3)=(x1)2+(x2)2+(x3)21F(x^{1},x^{2},x^{3})=(x^{1})^{2}+(x^{2})^{2}+(x^{3})^{2}-1, are both smooth manifolds. The former is the graph of the function ff and the latter is implicitly defined by the zeros of the equation of a 2-sphere (i.e., the points in 3\mathbb{R}^{3} at which the function FF vanishes).

Differential Geometry is often explained following the intrinsic point of view in which nothing outside of the manifold is considered. Nevertheless, for didactic purposes it is sometimes useful to think of a manifold as being embedded in a sufficiently high-dimensional Euclidean space. This can always be done due to the celebrated embedding theorem of Whitney 48, which guarantees that we can embed any smooth mm-dimensional manifold MM in 2m\mathbb{R}^{2m}.

A smooth curve on a manifold is a smooth mapping γ:IM\gamma\colon I\subset\mathbb{R}\to M where II is an interval of \mathbb{R}.

A.3 Tangent spaces and tangent bundles

Again, let MM be a smooth mm-dimensional manifold and let pMp\in M be some arbitrary point on it. A tangent vector vv at pp is a map that assigns a real number to each smooth, real-valued function of MM such that

  1. 1.

    v(f+g)=v(f)+v(g)v(f+g)=v(f)+v(g),

  2. 2.

    v(λf)=λv(f)v(\lambda f)=\lambda v(f),

  3. 3.

    v(fg)=v(f)g+fv(g)v(fg)=v(f)\,g+f\,v(g),

For any f,g:Mf,g\colon M\to\mathbb{R} smooth and λ\lambda\in\mathbb{R}.

The set of tangent vectors at pMp\in M is a vector space called the tangent space, denoted by TpMT_{p}M. Indeed, TpMT_{p}M is a vector space because if v,wv,w are tangent vectors at pp and λ\lambda\in\mathbb{R}, then the axioms

  1. 1.

    (v+w)(f)=v(f)+w(f)(v+w)(f)=v(f)+w(f),

  2. 2.

    (λv)(f)=λv(f)(\lambda\,v)(f)=\lambda\,v(f),

are satisfied.

In a local chart, the vectors

x1|p,,xm|p\frac{\partial}{\partial x^{1}}\bigg{|}_{p},\dotsc,\frac{\partial}{\partial x^{m}}\bigg{|}_{p}

constitute a basis of TpMT_{p}M.

Example 4.

Let M=mM=\mathbb{R}^{m}. In this case, the directional derivative of f:Mf\colon M\to\mathbb{R} at pMp\in M in the direction v=(v1,,vm)mv=(v^{1},\dotsc,v^{m})\in\mathbb{R}^{m} is

f(p)v=ivifxi(p).\nabla f(p)\cdot v=\sum_{i}v^{i}\frac{\partial f}{\partial x^{i}}(p).

In our notation, we would instead say that v=ivixi|pv=\sum_{i}v^{i}\,\frac{\partial}{\partial x^{i}}\big{|}_{p} is a tangent vector of MM at pp.

We can view each tangent vector at pp as the velocity vector of a smooth curve γ:[0,1]M\gamma\colon[0,1]\subset\mathbb{R}\to M such that γ(0)=p\gamma(0)=p. In that case,

v(f)=ddtf(γ(t))=iγ˙i(t)fxi(γ(t))=(iγ˙i(t)=vixi|γ(t))(f).v(f)=\frac{\mathrm{d}}{\mathrm{d}t}f(\gamma(t))=\sum_{i}\dot{\gamma}^{i}(t)\,\frac{\partial f}{\partial x^{i}}(\gamma(t))=\left(\sum_{i}\underbrace{\dot{\gamma}^{i}(t)}_{=v^{i}}\,\frac{\partial}{\partial x^{i}}\bigg{|}_{\gamma(t)}\right)(f).

The geometric interpretation of tangent vectors and tangent spaces is shown in Figure 10.

\begin{overpic}[width=281.85034pt,trim=42.67912pt 99.58464pt 42.67912pt 184.9429pt,clip]{figures/tangentspace.png} \put(49.0,52.0){$p$} \put(83.0,30.0){$T_{p}M$} \put(48.0,10.0){$M$} \end{overpic}
Figure 10: A tangent vector can be regarded intuitively as an “arrow” based at a point pMp\in M. We use tangent vectors as a tool to evaluate directional derivatives of real-valued functions defined on MM. The tangent space of a manifold at pp can be viewed as the span of all the tangent vectors at pp.

The collection

TM=pMTpMTM=\bigcup_{p\in M}T_{p}M

of the tangent spaces at each point of MM is called the tangent bundle and it is a smooth manifold in its own right.

Example 5 (The tangent bundle in classical mechanics).

If we denote the position of a mechanical system by qMq\in M, where MM is the configurational manifold of the system, and we consider an arbitrary trajectory t[0,1]γ(t)Mt\in[0,1]\mapsto\gamma(t)\in M starting at the point γ(0)=q\gamma(0)=q with initial velocity γ˙(0)=v\dot{\gamma}(0)=v, then we see that the tangent bundle is the collection of all the positions and velocities of the mechanical system under consideration. Moreover, the Lagrangian is a real-valued function of the tangent bundle,

L\displaystyle L :TM\displaystyle\colon TM\to\mathbb{R}
(q,v)L(q,v)=12vmvU(q),\displaystyle(q,v)\mapsto L(q,v)=\tfrac{1}{2}v\cdot mv-U(q),

where mm is the mass matrix of the system.

The momenta pp are obtained from the velocities vv by duality (via the Legendre transform 49). Because of that, the phase space of a mechanical system is a manifold that is closely related to the tangent bundle of the configurational manifold MM. Indeed, the phase space is the co-tangent bundle TMTM^{\star} of the configurational manifold MM.

The tangent bundle of a manifold is the prototype of a more general construction called a vector bundle that we will introduce next (albeit not fully rigorously). Informally, a vector bundle is a collection of vector spaces that are in correspondence to each point of MM. Slightly more formally, we say that a vector bundle over MM is a tuple (E,π,M)(E,\pi,M) where:

  1. 1.

    The total space EE and the base space MM are smooth manifolds.

  2. 2.

    The projection π:EM\pi\colon{E}\to{M} is a smooth map whose preimages (known as fibers) Ep=π1(p)E_{p}=\pi^{-1}(p) for pMp\in M are vector spaces (there are additional conditions that must be satisfied but they are beyond the scope of this appendix).

A trivial example of a vector bundle is shown in Figure 11.

\begin{overpic}[width=216.81pt,trim=42.67912pt 0.0pt 56.9055pt 85.35826pt,clip]{figures/trivialbundle.png} \put(50.0,60.0){\vector(0,-1){30.0}} \put(55.0,42.5){$\pi$} \put(80.0,5.0){$M$} \put(85.0,55.0){$E$} \end{overpic}
Figure 11: An example of a vector bundle. The base space in this case is the circle MM and the total space is the cylinder EE shown on top of MM. The vertical lines on the cylinder represent the fibers (i.e., the one-dimensional vector spaces that are the preimages by π\pi of each point of MM).

In the case of the tangent bundle, the fiber EpE_{p} is the tangent space TpMT_{p}M of MM at pp. More generally, we can take the fibers to be tensor spaces.

A.4 Vector fields and sections

A vector field is a correspondence of a tangent vector vpv_{p} to each point pMp\in M. The set of all vector fields over a manifold MM is denoted by 𝔛(M)\mathfrak{X}(M). An example of a vector field is shown in Figure 12.

Refer to caption
Figure 12: A vector field on the 2-sphere 𝕊2\mathbb{S}^{2}. A vector field is a (smooth) correspondence between each point pp of the manifold MM (the base of a black arrow) and a vector of TpMT_{p}M (the black arrow itself).

In local coordinates,

vp=ivi(p)xi|pv_{p}=\sum_{i}v^{i}(p)\,\frac{\partial}{\partial x^{i}}\bigg{|}_{p}

From now on, we adopt the convention

i=xi.\partial_{i}=\frac{\partial}{\partial x^{i}}.

To any smooth vector field v𝔛(M)v\in\mathfrak{X}(M) and any tt\in\mathbb{R}, we can associate the flow map, Φt:MM\Phi_{t}\colon M\to M, that satisfies

ddtΦt(p)=vΦt(p).\frac{\mathrm{d}}{\mathrm{d}t}\Phi_{t}(p)=v_{\Phi_{t}(p)}.

The smooth curve tΦt(p)t\mapsto\Phi_{t}(p) is also called an integral curve of vv.

In local coordinates, if v=iviiv=\sum_{i}v^{i}\,\partial_{i}, then tΦt(p)=(γ1(t),,γm(t))t\mapsto\Phi_{t}(p)=(\gamma^{1}(t),\dotsc,\gamma^{m}(t)) solves the initial value problem

{ddtγi(t)=vi(γ(t)),for each i=1,,m,\left\{\begin{aligned} &\tfrac{\mathrm{d}}{\mathrm{d}t}{\gamma}^{i}(t)=v^{i}(\gamma(t)),\\ &\text{for each $i=1,\dotsc,m$},\end{aligned}\right.

with the initial condition Φ0(p)=γ(0)=p\Phi_{0}(p)=\gamma(0)=p. We often denote dΦdt\frac{\mathrm{d}\Phi}{\mathrm{d}t} by Φ˙\dot{\Phi}.

Let π:EM{\pi}\colon{E}\to{M} be a vector bundle. A section is a smooth map

s:M\displaystyle{s}\colon{M} E\displaystyle\to{E}
p\displaystyle{p} s(p)Ep\displaystyle\mapsto{s(p)\in E_{p}}

such that the composition of ss followed by the projection π\pi is the identity in MM (or, equivalently, πs=1M\pi\circ s=1_{M}). Indeed, we have the following diagram:

E\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces E}π\scriptstyle{\pi}M\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}s\scriptstyle{s}

The set of sections of a vector bundle EE, denoted by Γ(E)\Gamma(E), is a vector space with the operations of addition,

(s+s)(p)=s(p)+s(p)(s+s^{\prime})(p)=s(p)+s^{\prime}(p)

and scalar product,

(λs)(p)=λs(p)(\lambda\,s)(p)=\lambda\,s(p)

In local coordinates (x1,,xm)(x^{1},\dotsc,x^{m}) around a point pp with a basis e1,,emEpe_{1},\dotsc,e_{m}\in E_{p}, the expression of a section ss can be written as

s=isiei.s=\sum_{i}s^{i}e_{i}.

A.5 Pullbacks and pushforwards

Consider a smooth map ϕ:MN\phi\colon M\to N between manifolds MM and NN with respective local coordinates (x1,,xm)(x^{1},\dotsc,x^{m}) and (y1,,yn)(y^{1},\dotsc,y^{n}). Let j1,,jk{1,,n}j_{1},\dotsc,j_{k}\in\{1,\dotsc,n\} and let f:Nf\colon N\to\mathbb{R} be a smooth function. The pullback of the expression f(y1,,yn)dyj1dyjkf(y^{1},\dotsc,y^{n})\,\mathrm{d}y^{j_{1}}\otimes\dotsb\otimes\mathrm{d}y^{j_{k}} is defined by

ϕ(f(y1,,yn)dyj1dyjk)=f(ϕ1(x1,,xm),,ϕn(x1,,xm))i1,,ik=1mϕj1xi1ϕjkxikdxi1dxik.\phi^{\star}\left(f(y^{1},\dotsc,y^{n})\,\mathrm{d}y^{j_{1}}\otimes\dotsb\otimes\mathrm{d}y^{j_{k}}\right)\\ =f(\phi^{1}(x^{1},\dotsc,x^{m}),\dotsc,\phi^{n}(x^{1},\dotsc,x^{m}))\sum_{i_{1},\dotsc,i_{k}=1}^{m}\frac{\partial\phi^{j_{1}}}{\partial x^{i_{1}}}\dotsb\frac{\partial\phi^{j_{k}}}{\partial x^{i_{k}}}\,\mathrm{d}x^{i_{1}}\otimes\dotsb\otimes\mathrm{d}x^{i_{k}}.

Consider a vector field X𝔛(M)X\in\mathfrak{X}(M), the pushforward of XX by ϕ\phi is a vector field in 𝔛(N)\mathfrak{X}(N) defined by

(ϕXp)(h)=Xp(hϕ)(\phi_{\star}X_{p})(h)=X_{p}(h\circ\phi)

for all smooth functions hC(N)h\in C^{\infty}(N) and all points pMp\in M.

Remark 6.

The pullback generalizes the transformation of the integrand in the change of variables formula known from integral calculus. The pushforward of a vector field is an expression of the chain rule from differential calculus.

A.6 Riemannian metrics

Let MM be an mm-dimensional smooth manifold. A Riemannian metric is a smooth map gg such that each point pMp\in M is mapped to a symmetric and positive definite matrix g(p):TpM×TpMg(p)\colon T_{p}M\times T_{p}M\to\mathbb{R}. If i,jTpM\partial_{i},\partial_{j}\in T_{p}M are vectors from an orthonormal frame in the local chart UU, the components of the Riemannian metric can be written as

gij(p)=gp(i,j)=ig(p)j=i,j,g_{ij}(p)=g_{p}(\partial_{i},\partial_{j})=\partial_{i}\cdot g(p)\,\partial_{j}=\langle\partial_{i},\partial_{j}\rangle,

for all pUp\in U and i,j=1,,mi,j=1,\dotsc,m. The inner product gp(v,w)g_{p}(v,w) between v=iviiTpMv=\sum_{i}v^{i}\partial_{i}\in T_{p}M and w=iwiiTpMw=\sum_{i}w^{i}\partial_{i}\in T_{p}M is defined by linearity. The norm of an arbitrary vector v=iviiTpMv=\sum_{i}v^{i}\,\partial_{i}\in T_{p}M is v=gp(v,v)\|v\|=\sqrt{g_{p}(v,v)}.

Remark 7.

If we take V=TpMV=T_{p}M, then its dual is V=TpMV^{\star}=T_{p}M^{\star} with orthonormal basis given by {ei:V}\{e^{i}\colon V\to\mathbb{R}\}. When V=TpMV=T_{p}M, as in this case, it is customary to write i=ei\partial_{i}=e_{i} and dxi=ei\mathrm{d}x^{i}=e^{i}. From the previous discussion, we see that the Riemannian metric gg is a section in Γ(TMTM)\Gamma(TM^{\star}\otimes TM^{\star}) because g(p):TpM×TpMg(p)\colon T_{p}M\times T_{p}M\to\mathbb{R} is linear and we can write it as

g(p)=ijgij(p)dxpidxpj.g(p)=\sum_{ij}g_{ij}(p)\,\mathrm{d}x^{i}_{p}\otimes\mathrm{d}x^{j}_{p}.
Example 6 (Euclidean metric in n\mathbb{R}^{n}).

The Euclidean metric in n\mathbb{R}^{n} is

g=i=1ndxidxi.g=\sum_{i=1}^{n}\mathrm{d}x^{i}\otimes\mathrm{d}x^{i}. (6)

Consider the polar coordinates in 2\mathbb{R}^{2} given by

{x1(r,θ)=rcosθ,x2(r,θ)=rsinθ,\left\{\begin{aligned} x^{1}(r,\theta)&=r\cos\theta,\\ x^{2}(r,\theta)&=r\sin\theta,\end{aligned}\right.

where r0r\geq 0 and 0<θ<2π0<\theta<2\pi. We have

dx1=cosθdrrsinθdθanddx2=sinθdr+rcosθdθ.\mathrm{d}x^{1}=\cos\theta\,\mathrm{d}r-r\sin\theta\,\mathrm{d}\theta\quad\text{and}\quad\mathrm{d}x^{2}=\sin\theta\,\mathrm{d}r+r\cos\theta\,\mathrm{d}\theta.

Therefore, the Euclidean metric (6) is pulled back as:

g=dx1dx1+dx2dx2=drdr+r2dθdθ.g=\mathrm{d}x^{1}\otimes\mathrm{d}x^{1}+\mathrm{d}x^{2}\otimes\mathrm{d}x^{2}=\mathrm{d}r\otimes\mathrm{d}r+r^{2}\mathrm{d}\theta\otimes\mathrm{d}\theta.

While in Euclidean coordinates the components of the metric tensor are gij=δijg_{ij}=\delta_{ij}, in polar coordinates they are g11=1g_{11}=1, g22=r2g_{22}=r^{2}, g12=g21=0g_{12}=g_{21}=0.

Let I=[0,1]I=[0,1]\subset\mathbb{R} and let γ:IM\gamma\colon I\to M be a smooth curve such that tγ(t)=(γ1(t),,γm(t))t\mapsto\gamma(t)=(\gamma^{1}(t),\dotsc,\gamma^{m}(t)). The arc-length of a curve γ\gamma is

(γ)=01γ˙(t)g(γ(t))γ˙(t)dt=01γ˙(t)dt.\ell(\gamma)=\int_{0}^{1}\sqrt{\dot{\gamma}(t)\cdot g(\gamma(t))\,\dot{\gamma}(t)}\,\mathrm{d}t=\int_{0}^{1}\|\dot{\gamma}(t)\|\,\mathrm{d}t.

Given an arbitrary smooth manifold MM, we can regard it as the configurational space of a (constrained) mechanical system. Let γ:[0,1]M\gamma\colon[0,1]\to M be a curve on MM. If we set the potential energy to be constant (U(q)=0U(q)=0, for simplicity) and we think of the Riemannian metric gg as a (position-dependent) mass matrix, then it turns out that the action integral is

01L(γ(t),γ˙(t))dt,\int_{0}^{1}L(\gamma(t),\dot{\gamma}(t))\,\mathrm{d}t, (7)

where the Lagrangian LL contains only a kinetic energy term,

L(γ(t),γ˙(t))=12γ˙(t)g(γ(t))γ˙(t),L(\gamma(t),\dot{\gamma}(t))=\tfrac{1}{2}\dot{\gamma}(t)\cdot g(\gamma(t))\,\dot{\gamma}(t),

A curve γ\gamma that locally minimizes the action 01L(γ,γ˙)dt\int_{0}^{1}L(\gamma,\dot{\gamma})\mathrm{d}t is called a geodesic curve of MM. By Jensen’s inequality 50, curves that minimize (7) also minimize the arc-length.

A.7 Connections

Consider a vector bundle EE on the manifold MM and let Γ(E)\Gamma(E) denote the sections of EE. We are interested in studying how a section sΓ(E)s\in\Gamma(E) changes along a given direction vTpMv\in T_{p}M. In essence, we want to define what it means to take the limit

limh0τh1s(γ(h))s(γ(0))h,\lim_{h\to 0}\frac{\tau_{h}^{-1}s(\gamma(h))-s(\gamma(0))}{h}, (8)

where τh:EpEγ(h)\tau_{h}\colon E_{p}\to E_{\gamma(h)} is a linear isomorphism between fibers and γ:[0,1]M\gamma\colon[0,1]\to M is a smooth curve on MM such that γ(0)=p\gamma(0)=p and γ˙(0)=v\dot{\gamma}(0)=v. If M=nM=\mathbb{R}^{n}, then we can take τh\tau_{h} to be the identity map but in general we need a non-trivial τh\tau_{h} to account for the fact that the fibers Eγ(h)E_{\gamma(h)} and Eγ(0)E_{\gamma(0)} are different vector spaces. To formalize this notion, we define the concept of a connection.

Let v,w𝔛(M)v,w\in\mathfrak{X}(M), f:Mf\colon M\to\mathbb{R} smooth, and λ\lambda\in\mathbb{R}. A connection :𝔛(M)×Γ(E)Γ(E)\nabla\colon\mathfrak{X}(M)\times\Gamma(E)\to\Gamma(E) acts on sections s,tΓ(E)s,t\in\Gamma(E) as follows:

  1. 1.

    v(λs)=λvs\nabla_{v}(\lambda s)=\lambda\nabla_{v}s.

  2. 2.

    v(s+t)=vs+vt\nabla_{v}(s+t)=\nabla_{v}s+\nabla_{v}t.

  3. 3.

    v(fs)=v(f)s+fvs\nabla_{v}(f\,s)=v(f)\,s+f\,\nabla_{v}s.

  4. 4.

    v+ws=vs+ws\nabla_{v+w}s=\nabla_{v}s+\nabla_{w}s.

  5. 5.

    fvs=fvs\nabla_{fv}s=f\nabla_{v}s.

In short, the connection \nabla is linear on its first argument (a vector field) and acts as a derivation (i.e., as Leibniz’s rule) on its second argument (a section). The covariant derivative of the section ss in the direction of vv is defined by vs\nabla_{v}s.

We can characterize the covariant derivative by writing it as:

vs=¯vs+Avs,\nabla_{v}s=\bar{\nabla}_{v}s+A_{v}s,

where ¯v\bar{\nabla}_{v} is known as the flat connection. If we introduce local coordinates (x1,,xm)(x^{1},\dotsc,x^{m}) around a point in the open set UMU\subset M, and consider the coordinate vector fields k\partial_{k} and a basis eie_{i} of Γ(E)\Gamma(E), then ¯kei=0\bar{\nabla}_{k}e_{i}=0 where k=k\nabla_{k}=\nabla_{\partial_{k}}. Therefore,

kei=jAkijej,\nabla_{k}e_{i}=\sum_{j}A^{j}_{ki}e_{j},

where AkijA^{j}_{ki} are the components of the vector potential. In full generality, we have

vs\displaystyle\nabla_{v}s =kvkk(isiei)=kvkk(isiei)\displaystyle=\nabla_{\sum_{k}v^{k}\partial_{k}}\left(\sum_{i}s^{i}e_{i}\right)=\sum_{k}v^{k}\nabla_{k}\left(\sum_{i}s^{i}e_{i}\right)
=kivk((ksi)ei+sikei)\displaystyle=\sum_{k}\sum_{i}v^{k}\left(\left(\nabla_{k}s^{i}\right)e_{i}+s^{i}\nabla_{k}e_{i}\right)
=kivk((ksi)ei+siAkijej)\displaystyle=\sum_{k}\sum_{i}v^{k}\left(\left(\partial_{k}s^{i}\right)e_{i}+s^{i}A^{j}_{ki}e_{j}\right)
=ikvk(ksi+sjAkji)ei.\displaystyle=\sum_{i}\sum_{k}v^{k}\left(\partial_{k}s^{i}+s^{j}A^{i}_{kj}\right)e_{i}.

Consequently, each component of the covariant derivative of the section ss along the vector vv is of the form

(vs)i=kvk(ksi+sjAkji).(\nabla_{v}s)^{i}=\sum_{k}v^{k}\left(\partial_{k}s^{i}+s^{j}A^{i}_{kj}\right).
Remark 8.

The definition of a covariant derivative vs\nabla_{v}s is equivalent to (8) (see 51 Chapter 6). In other words, the flat connection and the vector potential fully characterize all the possible ways in which we could define a derivative as a limit of the form (8). We will see next that there is a special type of connection that is determined by the Riemannian metric.

A.8 The Levi-Civita connection

A Riemannian metric induces a unique connection called the Levi-Civita connection. The components of the vector potential of the Levi-Civita connection are called the Christoffel symbols and we shall derive them next using classical mechanics.

Consider a free particle with unit mass moving on the manifold MM with position qq and velocity q˙\dot{q}. Its Lagrangian is

L(q,q˙)=12i=1mj=1mgij(q)q˙iq˙j.L(q,\dot{q})=\frac{1}{2}\sum_{i=1}^{m}\sum_{j=1}^{m}g_{ij}(q)\,\dot{q}^{i}\dot{q}^{j}.

The Euler-Lagrange equations for this mechanical system are

ddtLq˙iLqi=0,fori=1,,m.\frac{\mathrm{d}}{\mathrm{d}t}\frac{\partial L}{\partial\dot{q}^{i}}-\frac{\partial L}{\partial q^{i}}=0,\quad\text{for}\quad i=1,\dotsc,m.

This implies that

0\displaystyle 0 =j=1mddt(gij(q)q˙j)12j=1mk=1mgjkqi(q)q˙jq˙k\displaystyle=\sum_{j=1}^{m}\frac{\mathrm{d}}{\mathrm{d}t}\left(g_{ij}(q)\,\dot{q}^{j}\right)-\frac{1}{2}\sum_{j=1}^{m}\sum_{k=1}^{m}\frac{\partial g_{jk}}{\partial q^{i}}(q)\,\dot{q}^{j}\dot{q}^{k}
=j=1mgij(q)q¨j+j=1mk=1m(gijqk(q)12gjkqi(q))q˙jq˙k,\displaystyle=\sum_{j=1}^{m}g_{ij}(q)\,\ddot{q}^{j}+\sum_{j=1}^{m}\sum_{k=1}^{m}\left(\frac{\partial g_{ij}}{\partial q^{k}}(q)-\frac{1}{2}\frac{\partial g_{jk}}{\partial q^{i}}(q)\right)\dot{q}^{j}\dot{q}^{k},

for i=1,,mi=1,\dotsc,m. Using the identity j=1mk=1mgijqk=j=1mk=1mgikqj,\sum_{j=1}^{m}\sum_{k=1}^{m}\frac{\partial g_{ij}}{\partial q^{k}}=\sum_{j=1}^{m}\sum_{k=1}^{m}\frac{\partial g_{ik}}{\partial q^{j}}, we find

0=j=1mgij(q)q¨j+12j=1mk=1m(gijqk(q)+gikqj(q)gjkqi(q))q˙jq˙k,0=\sum_{j=1}^{m}g_{ij}(q)\,\ddot{q}^{j}+\frac{1}{2}\sum_{j=1}^{m}\sum_{k=1}^{m}\left(\frac{\partial g_{ij}}{\partial q^{k}}(q)+\frac{\partial g_{ik}}{\partial q^{j}}(q)-\frac{\partial g_{jk}}{\partial q^{i}}(q)\right)\dot{q}^{j}\dot{q}^{k},

for i=1,,mi=1,\dotsc,m. Finally, multiplying both sides above by i=1gi\sum_{i=1}g^{\ell i}, where gig^{\ell i} are the components of the inverse metric tensor (i.e., i=1gigij=δ,j\sum_{i=1}g^{\ell i}g_{ij}=\delta_{\ell,j} where δj\delta_{\ell j} is the Kronecker delta), we arrive at the geodesic equation

0=q¨+j=1mk=1mΓjk(q)q˙jq˙k,0=\ddot{q}^{\ell}+\sum_{j=1}^{m}\sum_{k=1}^{m}\Gamma^{\ell}_{jk}(q)\,\dot{q}^{j}\dot{q}^{k}, (9)

where

Γjk(q)=12i=1mgi(q)(gijqk(q)+gikqj(q)gjkqi(q))\Gamma^{\ell}_{jk}(q)=\frac{1}{2}\sum_{i=1}^{m}g^{\ell i}(q)\left(\frac{\partial g_{ij}}{\partial q^{k}}(q)+\frac{\partial g_{ik}}{\partial q^{j}}(q)-\frac{\partial g_{jk}}{\partial q^{i}}(q)\right)

is the expression of the Christoffel symbol for i,j,{1,,m}i,j,\ell\in\{1,\dotsc,m\}.

Remark 9.

Equation (9) is known as the geodesic equation and can be written more succinctly as q˙q˙=0\nabla_{\dot{q}}\dot{q}=0. The phase flow of this equation when the initial velocity has unit length is known as the exponential map expt\exp_{t} for t0t\geq 0. Therefore, exp0p=p\exp_{0}p=p, ddtexptp=q˙(t)\frac{\mathrm{d}}{\mathrm{d}t}\exp_{t}p=\dot{q}(t) such that q˙(0)=1\|\dot{q}(0)\|=1, and q(t)=exptpq(t)=\exp_{t}p solves the second order ODE (9).

Remark 10.

In the Euclidean case, M=mM=\mathbb{R}^{m} and gij=δijg_{ij}=\delta_{ij}, leading to the usual expression of the kinetic energy and to the free particle moving in a straight line. When MM is an arbitrary Riemannian manifold and the free particle moves along a geodesic, we can interpret the Christoffel symbols as the terms giving rise to the corrections to an otherwise straight trajectory so that it remains on the manifold MM.

A.9 Raising and lowering indices

The inner product g:TM×TMg\colon TM\times TM\to\mathbb{R} between two vectors X,YTpMX,Y\in T_{p}M determined by the Riemannian metric gives rise to a linear isomorphism (called flat) between TpMT_{p}M and TpMT_{p}^{\star}M by mapping

XTpMX=g(X,)TpM.X\in T_{p}M\mapsto X^{\flat}=g(X,\cdot)\in T_{p}^{\star}M.

The inverse linear map (called sharp) is

ωTpMωTpM\omega\in T_{p}^{\star}M\mapsto\omega^{\sharp}\in T_{p}M

such that ω=g(ω,)\omega=g(\omega^{\sharp},\cdot). The isomorphisms introduced above are called the musical isomorphisms.

Denoting by (gij)(g^{ij}) the inverse of the metric tensor gg, we write the components of ω\omega^{\sharp} and XX^{\flat} as

(ω)i=jgijωjand(X)i=jgijXi.{(\omega^{\sharp})^{i}=\sum_{j}g^{ij}\omega_{j}}\quad\text{and}\quad{(X^{\flat})_{i}=\sum_{j}g_{ij}X^{i}}.
Example 7.

The gradient of a function fC(M)f\in C^{\infty}(M) is defined as gradf=df\operatorname{grad}f=\mathrm{d}f^{\sharp}, where df=i=1mfxidxi\mathrm{d}f=\sum_{i=1}^{m}\frac{\partial f}{\partial x^{i}}\mathrm{d}x^{i}. We have seen that in the case of the plane 2\mathbb{R}^{2} with polar coordinates, g=drdr+r2dθdθg=\mathrm{d}r\otimes\mathrm{d}r+r^{2}\mathrm{d}\theta\otimes\mathrm{d}\theta. Consequently, gradf=frr+1r2fθθ\operatorname{grad}f=\frac{\partial f}{\partial r}\frac{\partial}{\partial r}+\frac{1}{r^{2}}\frac{\partial f}{\partial\theta}\frac{\partial}{\partial\theta}.

The musical isomorphisms can be used with individual factors of a tensor product to map between the primal and dual vector spaces.

References

  • E and Zhou 2011 E, W.; Zhou, X. The Gentlest Ascent Dynamics. Nonlinearity 2011, 24, 1831–1842
  • Yin et al. 2022 Yin, J.; Huang, Z.; Zhang, L. Constrained High-Index Saddle Dynamics for the Solution Landscape with Equality Constraints. Journal of Scientific Computing 2022, 91, 62
  • Hänggi et al. 1990 Hänggi, P.; Talkner, P.; Borkovec, M. Reaction-Rate Theory: Fifty Years after Kramers. Reviews of Modern Physics 1990, 62, 251–341
  • Olsen et al. 2004 Olsen, R. A.; Kroes, G. J.; Henkelman, G.; Arnaldsson, A.; Jónsson, H. Comparison of Methods for Finding Saddle Points without Knowledge of the Final States. The Journal of Chemical Physics 2004, 121, 9776–9792, Publisher: American Institute of Physics
  • Samanta et al. 2014 Samanta, A.; Chen, M.; Yu, T.-Q.; Tuckerman, M.; E, W. Sampling saddle points on a free energy surface. The Journal of Chemical Physics 2014, 140, 164109–164109
  • Quapp and Bofill 2014 Quapp, W.; Bofill, J. M. Locating saddle points of any index on potential energy surfaces by the generalized gentlest ascent dynamics. Theoretical Chemistry Accounts 2014, 133, 1510
  • Gu and Zhou 2017 Gu, S.; Zhou, X. Multiscale gentlest ascent dynamics for saddle point in effective dynamics of slow-fast system. Communications in Mathematical Sciences 2017, 15, 2279–2302, Publisher: International Press of Boston
  • Gu and Zhou 2018 Gu, S.; Zhou, X. Simplified gentlest ascent dynamics for saddle points in non-gradient systems. Chaos: An Interdisciplinary Journal of Nonlinear Science 2018, 28, 123106, Publisher: American Institute of Physics
  • Yin et al. 2019 Yin, J.; Zhang, L.; Zhang, P. High-Index Optimization-Based Shrinking Dimer Method for Finding High-Index Saddle Points. SIAM Journal on Scientific Computing 2019, 41, A3576–A3595, Publisher: Society for Industrial and Applied Mathematics
  • Yin et al. 2022 Yin, J.; Zhang, L.; Zhang, P. Solution Landscape of the Onsager Model Identifies Non-Axisymmetric Critical Points. Physica D: Nonlinear Phenomena 2022, 430, 133081
  • Gu et al. 2022 Gu, S.; Wang, H.; Zhou, X. Active Learning for Saddle Point Calculation. Journal of Scientific Computing 2022, 93, 78
  • Luo et al. 2022 Luo, Y.; Zheng, X.; Cheng, X.; Zhang, L. Convergence Analysis of Discrete High-Index Saddle Dynamics. 2022; arXiv:2204.00171 [cs, math]
  • Zhang et al. 2022 Zhang, L.; Zhang, P.; Zheng, X. Error Estimates for Euler Discretization of High-Index Saddle Dynamics. SIAM Journal on Numerical Analysis 2022, 60, 2925–2944, Publisher: Society for Industrial and Applied Mathematics
  • Quapp et al. 1998 Quapp, W.; Hirsch, M.; Imig, O.; Heidrich, D. Searching for saddle points of potential energy surfaces by following a reduced gradient. Journal of Computational Chemistry 1998, 19, 1087–1100
  • Quapp et al. 1998 Quapp, W.; Hirsch, M.; Heidrich, D. Bifurcation of reaction pathways: the set of valley ridge inflection points of a simple three-dimensional potential energy surface. Theoretical Chemistry Accounts 1998, 100, 285–299
  • Basilevsky 1982 Basilevsky, M. V. The topography of potential energy surfaces. Chemical Physics 1982, 67, 337–346
  • Hoffman et al. 1986 Hoffman, D. K.; Nord, R. S.; Ruedenberg, K. Gradient extremals. Theoretica chimica acta 1986, 69, 265–279
  • Quapp 2004 Quapp, W. Newton Trajectories in the Curvilinear Metric of Internal Coordinates. Journal of Mathematical Chemistry 2004, 36, 365–379
  • Bello-Rivas et al. 2022 Bello-Rivas, J. M.; Georgiou, A.; Guckenheimer, J.; Kevrekidis, I. G. Staying the Course: iteratively locating equilibria of dynamical systems on Riemannian manifolds defined by point-clouds. Journal of Mathematical Chemistry 2022,
  • Coifman and Lafon 2006 Coifman, R. R.; Lafon, S. Diffusion Maps. Applied and Computational Harmonic Analysis 2006, 21, 5–30
  • Golub and Van Loan 2013 Golub, G. H.; Van Loan, C. F. Matrix computations, 4th ed.; Johns Hopkins Studies in the Mathematical Sciences; Johns Hopkins University Press, 2013
  • Levitt and Ortner 2017 Levitt, A.; Ortner, C. Convergence and Cycling in Walker-type Saddle Search Algorithms. SIAM Journal on Numerical Analysis 2017, 55, 2204–2227
  • do Carmo 1992 do Carmo, M. P. Riemannian geometry; Mathematics: Theory & Applications; Birkhäuser Boston, Inc., Boston, MA, 1992
  • Absil et al. 2008 Absil, P.-A.; Mahony, R.; Sepulchre, R. Optimization Algorithms on Matrix Manifolds; Princeton University Press, 2008
  • Nadler et al. 2006 Nadler, B.; Lafon, S.; Coifman, R. R.; Kevrekidis, I. G. Diffusion Maps, Spectral Clustering and Reaction Coordinates of Dynamical Systems. Applied and Computational Harmonic Analysis 2006, 21, 113–127
  • Peters 2016 Peters, B. Reaction Coordinates and Mechanistic Hypothesis Tests. Annual Review of Physical Chemistry 2016, 67, 669–690
  • Berezhkovskii and Szabo 2019 Berezhkovskii, A. M.; Szabo, A. Committors, First-Passage Times, Fluxes, Markov states, Milestones, and All That. The Journal of Chemical Physics 2019, MMMK, 054106
  • Roux 2022 Roux, B. Transition Rate Theory, Spectral Analysis, and Reactive Paths. The Journal of Chemical Physics 2022, 156, 134111, Publisher: American Institute of Physics
  • Wu et al. 2022 Wu, S.; Li, H.; Ma, A. Exact Reaction Coordinates for Flap Opening in HIV-1 Protease. Proceedings of the National Academy of Sciences 2022, 119, e2214906119
  • Manuchehrfar et al. 2022 Manuchehrfar, F.; Li, H.; Ma, A.; Liang, J. Reactive Vortexes in a Naturally Activated Process: Non-Diffusive Rotational Fluxes at Transition State Uncovered by Persistent Homology. The Journal of Physical Chemistry B 2022, 126, 9297–9308
  • Chiavazzo et al. 2017 Chiavazzo, E.; Covino, R.; Coifman, R. R.; Gear, C. W.; Georgiou, A. S.; Hummer, G.; Kevrekidis, I. G. Intrinsic Map Dynamics Exploration for Uncharted Effective Free-Energy Landscapes. Proceedings of the National Academy of Sciences 2017, 201621481–201621481
  • Fujisaki et al. 2018 Fujisaki, H.; Moritsugu, K.; Mitsutake, A.; Suetani, H. Conformational Change of a Biomolecule Studied by the Weighted Ensemble Method: Use of the Diffusion Map Method to Extract Reaction Coordinates. The Journal of Chemical Physics 2018, 149, 134112
  • Fixman 1974 Fixman, M. Classical Statistical Mechanics of Constraints: A Theorem and Application to Polymers. Proceedings of the National Academy of Sciences 1974, 71, 3050–3053
  • Carter et al. 1989 Carter, E. A.; Ciccotti, G.; Hynes, J. T.; Kapral, R. Constrained Reaction Coordinate Dynamics for the Simulation of Rare Events. Chemical Physics Letters 1989, 156, 472–477
  • den Otter 2000 den Otter, W. K. Thermodynamic Integration of the Free Energy along a Reaction Coordinate in Cartesian Coordinates. The Journal of Chemical Physics 2000, 112, 7283–7292, Publisher: American Institute of Physics
  • Darve et al. 2008 Darve, E.; Rodríguez-Gómez, D.; Pohorille, A. Adaptive Biasing Force Method for Scalar and Vector Free Energy Calculations. The Journal of Chemical Physics 2008, 128, 144120–144120
  • Maragliano et al. 2006 Maragliano, L.; Fischer, A.; Vanden-Eijnden, E.; Ciccotti, G. String Method in Collective Variables: Minimum Free Energy Paths and Isocommittor Surfaces. The Journal of Chemical Physics 2006, 125, 024106
  • Zhang and Zha 2004 Zhang, Z.; Zha, H. Principal Manifolds and Nonlinear Dimensionality Reduction via Tangent Space Alignment. SIAM Journal on Scientific Computing 2004, 26, 313–338, Publisher: Society for Industrial and Applied Mathematics
  • Bradbury et al. 2018 Bradbury, J.; Frostig, R.; Hawkins, P.; Johnson, M. J.; Leary, C.; Maclaurin, D.; Necula, G.; Paszke, A.; VanderPlas, J.; Wanderman-Milne, S.; Zhang, Q. JAX: composable transformations of Python+NumPy programs. http://github.com/google/jax (Last accessed 2023-04-20), 2018
  • Bofill and Quapp 2016 Bofill, J. M.; Quapp, W. The variational nature of the gentlest ascent dynamics and the relation of a variational minimum of a curve and the minimum energy path. Theoretical Chemistry Accounts 2016, 135, 1–14
  • Randers 1941 Randers, G. On an Asymmetrical Metric in the Four-Space of General Relativity. Physical Review 1941, 59, 195–199
  • Chern and Shen 2005 Chern, S.-S.; Shen, Z. Riemann-Finsler Geometry; Nankai Tracts in Mathematics; WORLD SCIENTIFIC, 2005; Vol. 6
  • Kidger 2021 Kidger, P. On Neural Differential Equations. Ph.D. thesis, University of Oxford, 2021
  • Müller and Brown 1979 Müller, K.; Brown, L. D. Location of Saddle Points and Minimum Energy Paths by a Constrained Simplex Optimization procedure. Theoretica chimica acta 1979, 53, 75–93
  • Frankel 2011 Frankel, T. The Geometry of Physics: An Introduction, 3rd ed.; Cambridge University Press, 2011
  • Mezey 1987 Mezey, P. G. Potential Energy Hypersurfaces; Studies in physical and theoretical chemistry; Elsevier Scientific Publishing Co., Amsterdam, 1987; Vol. 53
  • Wales 2001 Wales, D. Energy Landscapes: Applications to Clusters, Biomolecules and Glasses, 1st ed.; Cambridge University Press, 2001
  • Whitney 1944 Whitney, H. The Self-Intersections of a Smooth nn-manifold in 2n2n-space. The Annals of Mathematics 1944, 45, 220–220
  • Arnold 1989 Arnold, V. I. Mathematical Methods of Classical Mechanics; Graduate Texts in Mathematics; Springer New York, 1989; Vol. 60
  • Rudin 1987 Rudin, W. Real and Complex Analysis, 3rd ed.; McGraw-Hill Book Co., 1987
  • Spivak 1999 Spivak, M. A Comprehensive Introduction to Differential Geometry. Vol. II, 3rd ed.; Publish or Perish, Inc., 1999
  • 52 Ref. 51, Chapter 6