correction of estimation bias feliks nüske, hao wu, jan-hendrik … · 2019. 5. 16. · feliks...

17
Markov state models from short non-equilibrium simulations—Analysis and correction of estimation bias Feliks Nüske, Hao Wu, Jan-Hendrik Prinz, Christoph Wehmeyer, Cecilia Clementi, and Frank Noé Citation: The Journal of Chemical Physics 146, 094104 (2017); doi: 10.1063/1.4976518 View online: http://dx.doi.org/10.1063/1.4976518 View Table of Contents: http://aip.scitation.org/toc/jcp/146/9 Published by the American Institute of Physics Articles you may be interested in Reconciling transition path time and rate measurements in reactions with large entropic barriers The Journal of Chemical Physics 146, 071101071101 (2017); 10.1063/1.4977177 Identification of simple reaction coordinates from complex dynamics The Journal of Chemical Physics 146, 044109044109 (2017); 10.1063/1.4974306 Perspective: Dissipative particle dynamics The Journal of Chemical Physics 146, 150901150901 (2017); 10.1063/1.4979514 Exploring high dimensional free energy landscapes: Temperature accelerated sliced sampling The Journal of Chemical Physics 146, 094108094108 (2017); 10.1063/1.4977704 Rate constants from instanton theory via a microcanonical approach The Journal of Chemical Physics 146, 074105074105 (2017); 10.1063/1.4976129 Beyond histograms: Efficiently estimating radial distribution functions via spectral Monte Carlo The Journal of Chemical Physics 146, 094107094107 (2017); 10.1063/1.4977516

Upload: others

Post on 26-Feb-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: correction of estimation bias Feliks Nüske, Hao Wu, Jan-Hendrik … · 2019. 5. 16. · Feliks Nuske,¨ 1,a) Hao Wu,1 Jan-Hendrik Prinz,1 Christoph Wehmeyer,1 Cecilia Clementi,2

Markov state models from short non-equilibrium simulations—Analysis andcorrection of estimation biasFeliks Nüske, Hao Wu, Jan-Hendrik Prinz, Christoph Wehmeyer, Cecilia Clementi, and Frank Noé

Citation: The Journal of Chemical Physics 146, 094104 (2017); doi: 10.1063/1.4976518View online: http://dx.doi.org/10.1063/1.4976518View Table of Contents: http://aip.scitation.org/toc/jcp/146/9Published by the American Institute of Physics

Articles you may be interested inReconciling transition path time and rate measurements in reactions with large entropic barriersThe Journal of Chemical Physics 146, 071101071101 (2017); 10.1063/1.4977177

Identification of simple reaction coordinates from complex dynamicsThe Journal of Chemical Physics 146, 044109044109 (2017); 10.1063/1.4974306

Perspective: Dissipative particle dynamicsThe Journal of Chemical Physics 146, 150901150901 (2017); 10.1063/1.4979514

Exploring high dimensional free energy landscapes: Temperature accelerated sliced samplingThe Journal of Chemical Physics 146, 094108094108 (2017); 10.1063/1.4977704

Rate constants from instanton theory via a microcanonical approachThe Journal of Chemical Physics 146, 074105074105 (2017); 10.1063/1.4976129

Beyond histograms: Efficiently estimating radial distribution functions via spectral Monte CarloThe Journal of Chemical Physics 146, 094107094107 (2017); 10.1063/1.4977516

Page 2: correction of estimation bias Feliks Nüske, Hao Wu, Jan-Hendrik … · 2019. 5. 16. · Feliks Nuske,¨ 1,a) Hao Wu,1 Jan-Hendrik Prinz,1 Christoph Wehmeyer,1 Cecilia Clementi,2

THE JOURNAL OF CHEMICAL PHYSICS 146, 094104 (2017)

Markov state models from short non-equilibrium simulations—Analysisand correction of estimation bias

Feliks Nuske,1,a) Hao Wu,1 Jan-Hendrik Prinz,1 Christoph Wehmeyer,1 Cecilia Clementi,2

and Frank Noe1,a)1Department of Mathematics and Computer Science, Freie Universitat Berlin, Arnimallee 6,14195 Berlin, Germany2Center for Theoretical Biological Physics and Department of Chemistry, Rice University, Houston,Texas 77005, USA

(Received 11 October 2016; accepted 30 January 2017; published online 2 March 2017)

Many state-of-the-art methods for the thermodynamic and kinetic characterization of large and com-plex biomolecular systems by simulation rely on ensemble approaches, where data from large numbersof relatively short trajectories are integrated. In this context, Markov state models (MSMs) areextremely popular because they can be used to compute stationary quantities and long-time kineticsfrom ensembles of short simulations, provided that these short simulations are in “local equilibrium”within the MSM states. However, over the last 15 years since the inception of MSMs, it has beencontroversially discussed and not yet been answered how deviations from local equilibrium can bedetected, whether these deviations induce a practical bias in MSM estimation, and how to correctfor them. In this paper, we address these issues: We systematically analyze the estimation of MSMsfrom short non-equilibrium simulations, and we provide an expression for the error between unbiasedtransition probabilities and the expected estimate from many short simulations. We show that theunbiased MSM estimate can be obtained even from relatively short non-equilibrium simulations inthe limit of long lag times and good discretization. Further, we exploit observable operator model(OOM) theory to derive an unbiased estimator for the MSM transition matrix that corrects for theeffect of starting out of equilibrium, even when short lag times are used. Finally, we show how theOOM framework can be used to estimate the exact eigenvalues or relaxation time scales of the systemwithout estimating an MSM transition matrix, which allows us to practically assess the discretiza-tion quality of the MSM. Applications to model systems and molecular dynamics simulation dataof alanine dipeptide are included for illustration. The improved MSM estimator is implemented inPyEMMA of version 2.3. Published by AIP Publishing. [http://dx.doi.org/10.1063/1.4976518]

I. INTRODUCTION

Ensemble approaches, where many fairly short simu-lations are produced in parallel or on distributed computerarchitectures, are widely used in order to characterize the ther-modynamics and kinetics of large biological macromolecules.Markov state models (MSMs)1–3 have become standard toolsfor the analysis of such data sets generated by moleculardynamics (MD) simulations.4–13 An MSM provides a sim-plified model of the underlying Markov process, which iscontinuous in both time and space, by a discrete time Markovchain on finitely many states. These states are defined by par-titioning the continuous state space into finitely many disjointsets. Time is discretized by choosing a discrete time step, calledthe lag time, and the full process is replaced by a snapshotprocess that only keeps track of the discrete state visited atthe discrete time steps, discarding any time information inbetween and any spatial information within the discrete sets.The quality of this approximation critically depends on thechoice of both discretization and lag time.14 One of the

a)Electronic addresses: [email protected] and [email protected]

strengths of Markov models is that the simulations used toconstruct them do not necessarily need to sample from theglobal equilibrium distribution, as only conditional transitionprobabilities between the states are required.4 In particular, atleast in principle, these transition probabilities can be obtainedwithout bias from simulations started out of local equilibriumin each state which only run for the length of a single lag timestep. However, it is much more practical to produce simula-tions that are longer than one lag time and estimate MSMs bycounting transitions along these trajectories. Even if the sim-ulations are started out of local equilibrium, the distributiondeviates from local equilibrium over time until global equilib-rium is restored. The estimation of transition probabilities istherefore subjected to a bias.1 In order to keep the bias small,it must be assumed that local equilibrium is approximatelyrestored after every time step.

The effect of the initial distribution onto the MSM qualityor even the justification of using an MSM for data analy-sis has been controversially discussed, and this issue has notbeen resolved yet. At least three ideas have been discussedas follows:15 (1) This effect exists,1 but may be small andcan be ignored in practice. (2) We can reduce the effect ofnon-equilibrium starting points by discarding the first bit of

0021-9606/2017/146(9)/094104/16/$30.00 146, 094104-1 Published by AIP Publishing.

Page 3: correction of estimation bias Feliks Nüske, Hao Wu, Jan-Hendrik … · 2019. 5. 16. · Feliks Nuske,¨ 1,a) Hao Wu,1 Jan-Hendrik Prinz,1 Christoph Wehmeyer,1 Cecilia Clementi,2

094104-2 Nuske et al. J. Chem. Phys. 146, 094104 (2017)

simulation trajectories, enough to reach local equilibrium.11

(3) We can avoid this problem by preparing local equilibriumdistributions in the starting states using biased simulations andthen shooting trajectories out of them.16–19

Here we qualify and quantify these ideas by systemati-cally analyzing the effect of non-equilibrium starting condi-tions onto MSM quality, and we suggest effective correctionmechanisms. Throughout the manuscript, we use the term“non-equilibrium” to describe the problem that simulationsare started from a distribution which is not in global equi-librium, and their simulation time is too short to reach thatglobal equilibrium. Note, however, that the dynamics itself isassumed to possess a unique equilibrium distribution, and iflong enough simulations would be run, they would samplefrom the equilibrium distribution. Briefly, our main results arethe following:

1. We provide an expression for the error between unbi-ased transition probabilities and the expected estimatefrom many simulations running for multiple discrete timesteps, see Sec. II. We find that there is no fundamentaladvantage of starting simulations in local equilibrium.Rather, the estimation error depends on the discretiza-tion, the simulation length, and the lag time. In the limitof long lag times and fine discretization, MSMs are esti-mated without bias even when non-equilibrium startingpoints are used. However, for a given discretization thelag time required to practically achieve a small estimationbias might be large.

2. We derive an unbiased MSM estimator that correctsthe error due to non-equilibrium starting conditions atshort lag times, by exploiting the framework of observ-able operator models (OOMs)—see Sec. III. OOMsare powerful finite-dimensional models that provide

unbiased estimates of stationary and kinetic properties ofstochastic processes under fairly mild assumptions, seeRefs. 20–22. Most importantly, OOMs can be estimatedfrom non-equilibrium simulations22 and are not limitedto a local equilibrium assumption.

3. We utilize the fact that exact relaxation time scales thatare not contaminated by the MSM projection error (i.e.,quality of the coordinates and the clustering used) canbe estimated using the OOM framework. The differencebetween the unbiased estimate and the uncorrected orcorrected MSM estimate is very insightful as it providesan indicator of the quality of the MSM discretization.If this difference is too large, it is suggested to ratherimprove the coordinate selection or discretization usedfor MSM construction and re-analyze. Note that whileOOMs offer the more general theory, they are not as easyto interpret and their estimation from finite data is not asstable and mature as MSM estimation.

As a technical advance, we provide a meaningful strategyto select the model rank of an OOM which is required in orderto obtain practically useful estimates, by using a statisticalanalysis of singular values of the count matrix (Sec. III D).

Sec. IV demonstrates the usefulness of the OOM frame-work for two model systems and MD simulation data of alaninedipeptide. We show that accurate estimates of spectral and sta-tionary properties can be obtained from short non-equilibriumsimulations, even for short lag times or poor discretizations.We explain how the discretization quality is revealed by thedifference between spectral estimates of the MSM and OOM.We also show that the rank selection strategy helps to choose asuitable model rank even for small lag times, when no apparenttime scale separation can be utilized.

As an illustration, consider the one-dimensional modelsystem governed by the potential shown in Fig. 1(a), see

FIG. 1. (a) One-dimensional potential function and dis-cretization into two states. (b) The same potential witha five-state discretization. ((c) and (d)) Estimates for theequilibrium probability of state 1 from the direct MSM(green) and the unbiased MSM (red), reference in black.((e) and (f)) Estimates for the slowest relaxation time scalet2 from a direct MSM (green), c.f. Eq. (19), the unbiasedMSM (red), c.f. Eqs. (38) and (39), and the spectral OOMestimation (blue), Eqs. (62) and (63). The black dashedline corresponds to the reference value.

Page 4: correction of estimation bias Feliks Nüske, Hao Wu, Jan-Hendrik … · 2019. 5. 16. · Feliks Nuske,¨ 1,a) Hao Wu,1 Jan-Hendrik Prinz,1 Christoph Wehmeyer,1 Cecilia Clementi,2

094104-3 Nuske et al. J. Chem. Phys. 146, 094104 (2017)

Sec. IV A for details. We study the estimation of a Markovmodel using the two-state discretization indicated in panel (a)of Fig. 1. For various lag times, we investigate the expectedtransition matrix if 90% of the simulations are started fromlocal equilibrium within state 1, while the other 10% are startedfrom local equilibrium within state 2. Note that we do not useany simulation data here, we only compute expected valuesover an ensemble of trajectories, with the trajectory length setto 2000 steps, which is shorter than the slowest relaxation timescale.

For short lag times, the standard MSM provides a stronglybiased estimate of the equilibrium population of the two wells(Fig. 1(c), green curve). For longer lag times, the MSM con-verges towards the correct equilibrium population, but the biasonly disappears when the lag time approaches the longestrelaxation time scale of the system, so if the initial distri-bution is far from equilibrium this can entail a significanterror at practically feasible lag times. In contrast, the cor-rected MSM estimate proposed in this paper achieves thecorrect estimate of equilibrium populations even at short lagtimes (Fig. 1(c), red curve). The standard MSM relaxation timescales are underestimated at short lag times, consistent withprevious variational results,23–25 but they can be improved byusing the unbiased MSM estimator proposed here (Fig. 1(e)).The OOM can provide a model-free estimate of the relax-ation time scale that is unbiased at a relatively short lag time(Fig. 1(e), blue line). The difference between the OOM and thecorrected MSM estimate (blue versus red lines in Fig. 1(e)) isan indicator of the MSM model error due to the state spacediscretization. Please note that all MSM results in this fig-ure can be dramatically improved if a finer clustering is used.For example, if the five-state partitioning from Fig. 1(b) isused instead, the estimation of stationary properties convergesmuch faster (Fig. 1(d)), and there is hardly a difference betweenthe time scales estimated by a direct and an unbiased MSM(Fig. 1(f)).

II. ANALYSIS OF THE STATE OF THE ART: MSMESTIMATION FROM SIMULATIONS WITH ARBITRARYSTARTING POINTSA. Molecular dynamics, count matrix,and transition matrix

In this work, we consider the setting described in detail inRef. 1, that is, an ergodic and reversible Markov process X t oncontinuous state spaceΩ, which possesses a unique stationarydistribution π. We denote by τ > 0 the lag time and by

p(x, y; τ) = P(Xτ ∈ dy|X0 = x) (1)

the conditional transition density function, that is the probabil-ity that the process, when located at configuration x at time t,will be found at configuration y at time t + τ. The correspond-ing transfer operator is denoted by T(τ) and is defined by itsaction on a function of state space u,

T(τ)u(y) =∫Ω

p(x, y; τ)π(x)π(y)

u(x) dx. (2)

Its eigenvalues are called

λm(τ) = exp(−τ/tm), (3)

where tm are the implied relaxation time scales. We denotethe transfer operator eigenfunctions by ψm, m = 1, . . . . Inparticular, we have that ψ1 ≡ 1. If the transfer operator is ofrank M at lag time τ, the transition density can be writtenas

p(x, y; τ) =M∑

m=1

λm(τ)ψm(x) π(y)ψm(y). (4)

Note that exact equality in Eq. (4) is an assumption, but oftenit is satisfied approximately for a large range of lag timesτ. Throughout the paper, we will consider decompositionsof state space into disjoint sets S1, . . . , SN , where Ω=

⋃iSi.

The indicator function of set Si is called χi. For a simulationof the continuous dynamics which samples positions at dis-crete time steps, we will denote the position at the kth timestep by Xk , k = 1, . . . , K , such that K is the total number oftime steps in the simulation. We use the symbol Y as a short-hand notation for an entire simulation. If multiple differentsimulations need to be distinguished, we will denote themby Yq, q= 1, . . . , Q, i.e., Q is the total number of availablesimulations.

Most of this work is based on correlations between thediscrete sets. For a trajectory as above, we define the empiricalhistograms and correlations (also called state-to-state time-correlations) as follows:

s(i) :=1

K − 2τ

K−2τ∑k=1

χi(Xk), (5)

Sτ(i, j) :=1

K − 2τ

K−2τ∑k=1

χi(Xk)χj(Xk+τ), (6)

S2τr (i, j) :=

1K − 2τ

K−2τ∑k=1

χi(Xk)χr(Xk+τ)χj(Xk+2τ). (7)

Up to the normalization, the matrix Sτ∈ RN×N is a count matrixbecause it simply counts the number of transitions from stateSi to Sj over a time window τ that have occurred in the simu-lation, while the vector s ∈RN counts the total visits to state Si

and corresponds to the ith row sum of Sτ . For each set Sr , thematrix S2τ

r ∈ RN ×N is proportional to a two-step count matrix

counting subsequent transitions from state Si to Sr and on tostate Sj. At first sight, it may seem confusing that Sτ and s onlycount transitions and visits up to time K−2τ, but further below,we will use all three matrices in conjunction which requiresestimating all of them over the same part of the data. We willcontinue to refer to these matrices as a count matrix, count vec-tor, and two-step count matrix in what follows. Also note thatin the literature, the count matrix and vector are often denotedby Cτ , c, but we will use these symbols differently in whatfollows. Let us note at this point that s, Sτ , S2τ

r can be seen asrandom variables that map a (stochastic) trajectory Y of dis-crete time steps to the values given in Eqs. (5)–(7). To empha-size this dependence, we will also write s(Y), Sτ(Y), S2τ

r (Y) ifappropriate.

We are concerned with the estimation of a transition prob-ability matrix between the sets Si of a given discretization of

Page 5: correction of estimation bias Feliks Nüske, Hao Wu, Jan-Hendrik … · 2019. 5. 16. · Feliks Nuske,¨ 1,a) Hao Wu,1 Jan-Hendrik Prinz,1 Christoph Wehmeyer,1 Cecilia Clementi,2

094104-4 Nuske et al. J. Chem. Phys. 146, 094104 (2017)

state space. If the process is in equilibrium, the conditionaltransition probabilities can be expressed as

TτEq(i, j) =P

(Xt ∈ Si, Xt+τ ∈ Sj

)P(Xt ∈ Si)

(8)

=∫Si

dx ∫Sjdy π(x) p(x, y; τ)

∫Sidx π(x)

(9)

=Cτ

Eq(i, j)

πi. (10)

Here, we have defined the equilibrium correlation between setsSi and Sj by the nominator of Eq. (9) and denoted it by Cτ

Eq(i, j).Also, we have adopted the usual notation πi = ∫Si

dx π(x) forthe equilibrium probabilities of the discrete states. Such amatrix of conditional transition probabilities is called a Markovstate model (MSM) or Markov model. It can be used as a sim-plified model for the dynamics allowing extensive analysis,see Ref. 1.

From a long simulation Xk , k = 1, . . . , K that samplespoints from the stationary density π, the matrix TτEq can beestimated by the formula

TτEq(i, j) ≈Sτ(i, j)

s(i). (11)

B. Starting from local equilibrium

In practice, producing simulation data that sample fromthe global equilibrium density π is often not tractable. One ofthe strengths of Markov models is the fact that the transitionmatrix can also be expressed in terms of local equilibriumdensities

πSi (x) =1πiχi(x)π(x). (12)

The density πSi is the normalized restriction of π to state Si.A Markov model transition matrix can also be estimated bypreparing an ensemble of trajectories in such a way that, withineach state, the distribution of starting points equals the localdensity Eq. (12). These trajectories are simulated for a veryshort time, and the fraction of trajectories starting in Si andending up in Sj provides an estimate for the transition matrixentry TτEq(i, j).16,26 To see this, note that in the setting justdescribed, the initial distribution is a convex combination ρL

of the local densities πSi,

ρL =

N∑i=1

aiπSi , (13)

N∑i=1

ai = 1. (14)

Here, ai is the probability to start in state Si. Upon replacingπ by ρL in Eq. (9), it follows that

TτEq(i, j) =∫Si

dx ∫Sjdy ρL(x)p(x, y; τ)

∫Sidx ρL(x)

. (15)

Only very short trajectories and knowledge of the local densi-ties are needed for the application of this method. However,this

method suffers from three major disadvantages: first, the inter-mediate data points of the simulations cannot be used. Second,estimation of the local densities requires the use of biasedsampling methods, which is a significant extra effort andentails additional difficulties. Third, changing the discretiza-tion requires redoing the simulations, which is not acceptableif a suitable discretization is not easy to find.

C. Multiple-step estimator

A common way to construct MSMs in practice is by con-ducting a large set of distributed simulations Yq, q= 1, . . . , Qof lengths that are shorter than the largest relaxation time scalesof the system but are longer than the lag time τ. For our theoret-ical investigation, we will assume that each of these trajectorieshas the same length of K stored simulation steps, but for theestimators we will be deriving later that uniform length is nota requirement, see Appendix C.

The simulations are started from some arbitrary initialdistribution at time k = 1. The transition probability matrix isestimated by replacing S(i, j) and s(i) by their empirical meanvalues over all simulations Yq. These are defined by the follow-ing equations, where we include the corresponding definitionfor S2τ

r for later use,

s =1Q

Q∑q=1

s(Yq), (16)

Sτ=

1Q

Q∑q=1

Sτ(Yq), (17)

S2τr =

1Q

Q∑q=1

S2τr (Yq). (18)

In analogy to Eq. (11), the transition matrix is then estimatedby

Tτ(i, j) =

Sτ(i, j)

s(i). (19)

Additional constraints can be incorporated in order to obtainmore specific estimators than Eq. (19), such as estimatorsobeying detailed balance.1,10,27

The argument from Sec. II B cannot be transferred directlyto a multiple-step estimator like Eqs. (16) and (17): Even if thesimulations are started from local equilibrium, this property islost after the first simulation step, and the resulting estimatesare no longer unbiased. A detailed illustration of this phe-nomenon has been provided by Ref. 1 [Fig. 4], and we repeatit here in Fig. 2. It can be argued that if the discretization ischosen well enough such that the dynamics equilibrates to anapproximate local equilibrium within all states over a singletime step, the bias can be expected to be very small. Thisassumption is difficult to check or quantify in practice. InSec. II D, we analyze the bias introduced by the multiple-step estimator, as well as its dependence on the lag time andsimulation length.

Page 6: correction of estimation bias Feliks Nüske, Hao Wu, Jan-Hendrik … · 2019. 5. 16. · Feliks Nuske,¨ 1,a) Hao Wu,1 Jan-Hendrik Prinz,1 Christoph Wehmeyer,1 Cecilia Clementi,2

094104-5 Nuske et al. J. Chem. Phys. 146, 094104 (2017)

FIG. 2. Loss of local equilibrium property illustrated bycomparing the dynamics of the diffusion in a double-wellpotential (a) and (e) at time steps 0 (b), 250 (c), 500 (d)with the predictions of a Markov model parameterized atlag time τ = 250 at the same times 0 (f), 250 (g), 500 (h).Please refer to the supplementary material of Ref.1 fordetails of the system. (b)–(d) show the true distribution ofthe system (solid black line) and the probabilities asso-ciated with the two discrete states left and right of thedashed line. The numbers in (f)–(h) are the discrete stateprobabilities pi(kτ), i= 1, 2, k = 0, 1, 2, predicted by theMarkov model. The solid black lines show the hypothet-ical density pi(kτ)πSi that is inherently assumed whenestimating a Markov model by counting transitions overmultiple steps. This figure has been re-used with permis-sion from Prinz et al., J. Chem. Phys. 134, 174105 (2011)[Fig. 4]. Copyright 2011 American Institute of Physics.

D. Estimation error from non-equilibrium simulations

Now we study the effect of using an initial distributionof simulation data that is not in local equilibrium when thetransitions are counted. This deviation from local equilibriumcould come either from the fact that we start trajectories inan arbitrary initial condition, or that our trajectories exceedthe lag time τ such that an initially prepared local equilibriumis lost for all transition counts harvested after the first one(Sec. II C).

Let ρ denote the empirical distribution sampled by thesimulations. We need to study the error between the equilib-rium transition matrix TτEq and the asymptotic limit of Eq. (19).

To this end, we study the asymptotic limits of Sτ(i, j) and s(i)

in the limit of infinitely many simulations, Q → ∞, but eachhaving finite lengths,

Cτρ(i, j) := E (Sτ(i, j)) , (20)

cρ(i) := E (s(i)) , (21)

Tτρ(i, j) :=Cτρ(i, j)

cρ(i). (22)

Thus, we use the symbols Cτρ, cρ for the expected count matrix

and vector of total counts associated with the empirical dis-tribution ρ. Using the spectral decomposition, Eq. (4), theexpected count matrix can be expressed in terms of the spectralcomponents of the dynamics,

Cτρ(i, j) =

M∑m=1

λm(τ)〈χi,ψm〉ρ〈χj,ψm〉π , (23)

〈χi,ψm〉ρ =

∫Ω

dx χi(x)ψm(x)ρ(x), (24)

〈χi,ψm〉π =

∫Ω

dx χi(x)ψm(x)π(x). (25)

In matrix form, Eq. (23) can be written as

Cτρ = QρΛ(τ)QT

π , (26)

Qρ(i, m) = 〈χi,ψm〉ρ, (27)

Qπ(j, m) = 〈χj,ψm〉π . (28)

These matrices contain the MSM projections of the true eigen-functions, i.e., their approximations by step functions, whichare extensively discussed in Refs. 1 and 14. Let us emphasizethat Eq. (23) also holds for arbitrary basis functions, i.e., χi

is not required to be a basis of indicator functions. Thus, itis the most general expression for a correlation matrix fromMarkovian dynamics.

Summation over j shows that

cρ(i) = 〈χi〉ρ. (29)

It follows from Eq. (23) that the spectral expansion of CτEq is

given by

CτEq(i, j) =

M∑m=1

λm(τ)〈χi,ψm〉π〈χj,ψm〉π , (30)

using the fact that for trajectories started from global equilib-rium we have ρ= π. Combining Eqs. (23), (29), and (30), weobtain an expression for the estimation error Eτ := Tτρ − TτEq,

Page 7: correction of estimation bias Feliks Nüske, Hao Wu, Jan-Hendrik … · 2019. 5. 16. · Feliks Nuske,¨ 1,a) Hao Wu,1 Jan-Hendrik Prinz,1 Christoph Wehmeyer,1 Cecilia Clementi,2

094104-6 Nuske et al. J. Chem. Phys. 146, 094104 (2017)

Eτ(i, j) =Cτρ(i, j)

cρ(i)−

CτEq(i, j)

πi(31)

=

M∑m=2

λm(τ)〈χj,ψm〉π

[〈χi,ψm〉ρ

〈χi〉ρ−〈χi,ψm〉π

〈χi〉π

](32)

=

M∑m=2

λm(τ)〈χj,ψm〉π

[〈χi,ψm − qimψ1〉ρ

〈χi〉ρ

], (33)

where qim =〈χi ,ψm〉π〈χi〉π

, and we were able to drop the m = 1 termson both sides as they are equal. Inspecting this expressionleads to a number of insights that are practically important foranalyzing simulation data with MSMs:

1. MSM estimation from long trajectories: In the limit thatour trajectories are longer than the time scale of theslowest process, the empirical distribution ρ convergesto the equilibrium distribution π, and the bias becomeszero. This offers an explanation why MSMs built fromultra-long simulations28,29 are quite well-behaved andhave been extensively used for benchmarking and methodvalidation.

2. MSM estimation from short trajectories: Even if the tra-jectories are not long enough to reach global equilibrium,because of Eq. (3), the bias decays multi-exponentiallywith the lag time τ. This is an important insight becauseMSMs are in practice constructed in the limit of longenough lag times in which the time scale estimates con-verge,1,5 and the above equation shows that this limit ismeaningful as it approaches an unbiased estimate.

3. Dependence of bias on the discretization error: The aboveformula reflects the well-known insight that Markovmodels are free of bias if the discretization perfectlyapproximates the dominant eigenfunctions, meaning thatthe eigenfunctions are constant on the states Si.1,5

4. Consequences for adaptive sampling: Previous adaptivesampling approaches have suggested preparing an initiallocal equilibrium distribution in order to shoot trajecto-ries out of selected states.16 The above analysis showsthat this strategy is effective if we only count a singletransition out of the state but is ineffective when longertrajectories are shot. In the latter case, it is simpler to

ignore the initial distribution and to reduce the effect ofbias by extending the lag time τ, see again Fig. 1 and alsothe next example.

E. Example

Before proceeding, we illustrate these findings by re-visiting the one-dimensional model system presented in theIntroduction. We study the same two different discretizations,the two-state model from panel (a) of Fig. 3 and the five-statediscretization shown in Fig. 3(b). Again, simulations are ini-tiated from local equilibrium in states 1 and 2 of the coarsediscretization, with a1 = 0.9, a2 = 0.1. We study the expectedestimate of the equilibrium probability of state 1, which equalsthe equilibrium probability of states I and II for the finer statedefinition. Panels (c) and (d) of Fig. 3 show the respectiveestimates for the coarse and fine discretization as a functionof the lag time, for simulation lengths K = 1000, 2000, 5000,10 000, 50 000. Indeed, the estimates improve if the lag timeis increased, if the simulation length is increased, or if thediscretization is improved. From the coarse partitioning exam-ple, we conclude that relaxation to global equilibrium can berequired in order to obtain unbiased estimates from simulationsinitiated out of local equilibrium.

III. CORRECTION OF ESTIMATION BIAS USINGOBSERVABLE OPERATOR MODELS

In this section, we show how to go beyond just using alonger lag time τ and suggest correction mechanisms to obtainthe correct equilibrium transition matrix TτEq (Eqs. (8)–(10))from an ensemble of short simulations. This can be accom-plished regardless of the starting distribution being in globalequilibrium, in local equilibrium, or far from any equilibrium.

As discussed above, limitations of MSMs include theassumption of Markovianity, sensitivity to projection error,and sensitivity to the distribution of trajectory starting points.All of these limitations can be overcome by realizing thatmolecular dynamics that is observed in a chosen set of vari-ables, reaction coordinates, or order parameters at a certain lagtime τ can be exactly described by projected Markov models

FIG. 3. ((a) and (b)) One-dimensional potential functionwith two different discretizations into two states and fivestates, respectively. (c) Expected estimate of the equi-librium probability of state 1 as a function of the lagtime, for simulation lengths K = 1000, 2000, 5000, 10 000,50 000, and using the discretization from panel (a). Thesimulations are initiated in local equilibrium in both states1 and 2 but predominantly in state 1 (a1 = 0.9, a2 = 0.1).(d) The same for the five-state discretization frompanel (b).

Page 8: correction of estimation bias Feliks Nüske, Hao Wu, Jan-Hendrik … · 2019. 5. 16. · Feliks Nuske,¨ 1,a) Hao Wu,1 Jan-Hendrik Prinz,1 Christoph Wehmeyer,1 Cecilia Clementi,2

094104-7 Nuske et al. J. Chem. Phys. 146, 094104 (2017)

(PMMs).30 This insight allows us to employ estimators thatare not affected by the MSM limitations, such as hiddenMarkov models (HMMs)30 or observable operator models(OOMs),20–22 that operate on the discretized state space.

Here, we employ OOMs in order to get improved MSMestimators that are not subject to the bias caused by a non-equilibrium distribution of the trajectories used. In a nutshell,OOMs are spectral estimators able to provide unbiased esti-mates of stationary and dynamical quantities for dynamicalsystems that can be well described by a finite number ofdynamical components. Here we only summarize a few aspectsof OOMs that are relevant to the present paper and presentan algorithm that can be used to estimate MSMs withoutbias from the initial trajectory distribution. To fully under-stand the theoretical background and derivation, please refer toRefs. 20–22.

A. Observable operator models

Observable operator models (OOMs) provide a frame-work that completely captures the dynamics of a stochasticdynamical system by a finite-dimensional algebraic system ifonly a finite number M of relaxation processes contribute inEq. (4), see Refs. 20 and 21. For molecular dynamics, thisproperty is achieved if we observe and model the dynamics ata finite lag time τ. The full-state observable operator ΞΩ is anM ×M matrix which contains the scalar products between theeigenfunctions,

ΞΩ(m, m′) = λm(τ)∫Ω

dxψm(x)ψm′(x)π(x). (34)

In statistical terms, ΞΩ is the expectation value of the covari-ance matrix between eigenfunctions. As eigenfunctions areorthogonal with respect to the equilibrium distribution π, orin other words, statistically uncorrelated, ΞΩ is just a diagonalmatrix of the eigenvalues,

ΞΩ = Λ. (35)

If we do not integrate over the full state spaceΩ in Eq. (34), butonly over a subset A ⊂ Ω, we can define a matrix ΞA of sizeM ×M, called the set-observable operator for set A. All set-observable operators and two vectors ω, σ ∈ RM are the keyingredients of OOM theory. The vectors ω, σ equal the firstcanonical unit vector e1, i.e.,ω =σ = e1 = (1, 0, . . . , 0)T , andthey are called information state and evaluator, respectively.If the finite-rank assumption Eq. (4) holds, these componentsform an algebraic system that allows to compute equilibriumprobabilities of finite observation sequences. Let A1, . . . , Al

be arbitrary subsets of Ω that do not need to form a partitionof the state space. If Eq. (4) is satisfied, we can compute theprobability that a trajectory in equilibrium visits set A1 at timeτ, set A2 at time 2τ, . . ., and set Al at time lτ by the followingmatrix-vector product:

P(Xτ ∈ A1, X2τ ∈ A2, . . . , Xlτ ∈ Al) = ωTΞA1 . . .ΞAlσ. (36)

The proof can be found in Ref. 21, we also repeat it inAppendix B. Note that, in case that A1, . . . , Al form a par-tition of state space, the probability of such an observationsequence cannot be computed from a Markov model transi-tion matrix between the sets A1, . . . , Al, unless the dynamics is

Markovian on these sets. This clearly distinguishes an OOMfrom a Markov model: An OOM can correctly describearbitrary projected dynamics as long as Eq. (4) holds.

As a Markov process is determined entirely by finiteobservation probabilities like Eq. (36), it follows that we cancompute several key equilibrium, kinetic, and mechanisticquantities in an unbiased fashion if we can somehow esti-mate the OOM components. For a fixed decomposition of statespace into sets Sr , r = 1, . . . , N as before, let us denote theset-observable operators of sets Sr by Ξr , which implies that

ΞΩ =

N∑r=1

Ξr . (37)

It follows from Eq. (36) that we can compute the unbiasedequilibrium correlation matrix and the stationary probabilitiesby the formulas

CτEq(i, j) = ωTΞiΞjσ, (38)

πi = ωTΞiσ. (39)

In practice we cannot directly estimate Ξr but only a similaroperator Ξr . However, it follows directly from Eqs. (38) and(39) that if an unknown similarity transform R ∈ RM×M affectsall OOM quantities via

Ξr = RΞrR−1, (40)

ωT = ωT R−1, (41)

σ = Rσ, (42)

then Eqs. (38) and (39) remain exactly valid using ω, Ξr , σ.In other words, all OOMs that can be constructed by choos-ing some transformation matrix R form a family of equivalentOOMs. A specific member of this family can be estimateddirectly from simulation data, and thus we can use it in orderto obtain unbiased estimates of Eqs. 38 and 39 even froma large ensemble of trajectories that do not need to samplefrom global equilibrium. It has been shown in Ref. 22 thatEqs. (47)–(50) in Sec. III B indeed provide the components ofan equivalent OOM, i.e., there is an invertible matrix R suchthat Eqs. (40)–(42) are satisfied in the absence of statisticalnoise.

B. Unbiased estimation of Markov state models

To construct an exact unbiased estimator, we need threeingredients: (i) the expectation values of the empirical countmatrix Cτ

ρ, (ii) the vector of total counts cρ from Eqs. (20) and(21), and additionally (iii) the two-step count matrices,

C2τρ,r := E

(S2τ

r

). (43)

As a reminder, expectation values here denote the expectationover a trajectory ensemble sampling from the empirical (non-equilibrium) distribution ρ. In practice, only finitely manysimulations are available, and we thus replace cρ, Cτ

ρ and C2τρ,r

by count vectors and matrices s, Sτ

and S2τr (Eqs. (16)–(18)),

which are asymptotically unbiased estimators. The unbiasedestimation algorithm can be summarized as follows:

1. Obtain the empirical mean s, count matrix Sτ, and

two-step count matrices S2τr from simulation data using

Eqs (16)–(18).

Page 9: correction of estimation bias Feliks Nüske, Hao Wu, Jan-Hendrik … · 2019. 5. 16. · Feliks Nuske,¨ 1,a) Hao Wu,1 Jan-Hendrik Prinz,1 Christoph Wehmeyer,1 Cecilia Clementi,2

094104-8 Nuske et al. J. Chem. Phys. 146, 094104 (2017)

2. Decompose the count matrix Sτ

by singular value decom-position (SVD),

Sτ= VΣWT , (44)

and compute weighted projections onto the leading M leftand right singular vectors by

F1 = VMΣ−1/2M , (45)

F2 =WMΣ−1/2M . (46)

We have used the symbols VM , WM , ΣM to denote therestriction of these matrices to their first M columns.

3. Use F1, F2 to obtain the set-observable operators Ξr andthe evaluation state vector σ of an equivalent OOM via

Ξr = FT1 S

2τr F2, (47)

σ = FT1 s. (48)

Compute the full-state observable operator ΞΩ =∑Nr=1 Ξr and obtain the information state vector ω as the

solution to the eigenvalue problem,

ωT ΞΩ = ωT , (49)

ωT σ = 1. (50)

The normalization Eq. (50) can be achieved by dividingthe arbitrarily scaled solution ωT by ωT σ.

4. Compute the unbiased equilibrium correlation matrix andunbiased equilibrium distribution by

CτEq(i, j) = ωT ΞiΞjσ, (51)

πi = ωT Ξiσ (52)

=

N∑j=1

CτEq(i, j) (53)

and then obtain the unbiased MSM transition matrix TτEqeither using the nonreversible estimator

TτEq(i, j) =Cτ

Eq(i, j)

πi, (54)

or the reversible estimator

TτEq(i, j) =Cτ

Eq(i, j) + CτEq(j, i)∑N

j=1 CτEq(i, j) +

∑Nj=1 Cτ

Eq(j, i). (55)

Let us briefly comment on the central idea behind this algo-rithm, which is the estimation of an equivalent OOM in thethird step, particularly in Eq. (47). Using the path probabilityformula Eq. (36), it can be shown that the expected two-stepcount matrix is given by

C2τρ,r = QρΞrΛ(τ)QT

π , (56)

where the matrices Qρ, Qπ are the same as in Eqs. (27) and(28). Thus, by the intermediate step, the set-observable opera-tor is introduced into the decomposition of the two-step countmatrix. Now, the idea is to find two matrices F1, F2∈ R

N×M ,such that R1 :=FT

1 Qρ and R2 :=Λ(τ)QTπF2 are inverse to each

other, because this implies that

FT1 C2τ

ρ,rF2 = R1ΞrR2 (57)

= RΞrR−1 (58)

is the rth component of an equivalent OOM. The properties ofSVD and the decomposition Eq. (26) guarantee that the choiceof F1, F2 in the second step above achieves this goal,

Id = FT1 Cτ

ρF2 (59)

=(FT

1 Qρ

) (Λ(τ)QT

πF2

)(60)

= R1R2. (61)

Similar arguments can be used to justify the equations forω, σ. We also note that different choices of F1, F2 in step 2are possible. For detailed explanations and proofs, please referto the previous publications.20–22

C. Recovery of exact relaxation time scales

A remarkable by-product of the procedure describedabove is that the transformed full-state two-step count matrixΞΩ is similar to a diagonal matrix of the system eigenval-ues λm(τ) without any MSM projection error. This has beenshown for equilibrium data in Ref. 31 and also applies tonon-equilibrium data,21

ΞΩ = RΞΩR−1 (62)

= RΛ(τ)R−1. (63)

Thus, diagonalization of ΞΩ provides an estimate of theleading system eigenvalues, and consequently also of therelaxation rates or time scales, that is not distorted by thefact that we coarse-grain the dynamics to a Markov chainbetween coarse sets in state space. These eigenvalue and timescale estimates are only subject to statistical error but notto any MSM model error. It is impossible to directly buildan MSM that produces these time scales—when an MSM isdesired, the time scales can only be approximated, and theywill only be correct in the limit of long lag times and gooddiscretization.

However, the fact that we can get a model-free estimate ofthe eigenvalues and relaxation time scales can be used to assessthe discretization quality: According to the variational princi-ple of conformation dynamics,24 the exact system eigenvaluesprovide an upper bound to the eigenvalues of the equilibriumtransition matrix TτEq. By comparing the eigenvalues of TτEq tothose from Eqs. (62) and (63), the MSM discretization errortheoretically studied in Refs. 1, 14, and 23 can be practicallyquantified.

D. Selection of model rank

The above method is theoretically guaranteed to workwhenever the number of MSM states N is at least equal tothe number M of relaxation processes in Eq. (4), and the countmatrix Cτ

ρ is of rank M. In the absence of statistical noise,the model rank M can then be determined by the number ofnon-zero singular values of Cτ

ρ. For finite data, the numerical

rank of Sτ

is not necessarily equal to M, as the singular valuescan be perturbed by noise. Classical matrix perturbation theorypredicts that small singular values will be particularly affectedby noise, see, e.g., Ref. 32, and also Fig. 4(a). Including noisyand small singular values can severely affect the accuracy ofthe method, most likely due to the presence of the matrix ofinverse singular values in Eqs. (45) and (46). Also, we expect

Page 10: correction of estimation bias Feliks Nüske, Hao Wu, Jan-Hendrik … · 2019. 5. 16. · Feliks Nuske,¨ 1,a) Hao Wu,1 Jan-Hendrik Prinz,1 Christoph Wehmeyer,1 Cecilia Clementi,2

094104-9 Nuske et al. J. Chem. Phys. 146, 094104 (2017)

FIG. 4. Analysis of statistical uncertainties for sin-gular values of the count matrix. We use the one-dimensional model system and seven-state discretizationas in Sec. IV A, the sample consists of Q = 5000 tra-jectories of length K = 2000. (a) For each of the sevensingular values (distinguished in descending order bythe colors black, blue, cyan, green, magenta, red, andyellow), we show the ratio of the true singular valueσr (Cτ

ρ), r = 1, . . . , 7 of the expected count matrix Cτρ to

the corresponding singular value σr (Sτ

) of the empiricalcount matrix S

τ, as a function of the lag time. As the small

singular values decay quickly with the lag time, they aredominated by the noise even for small lag times. Includ-ing these noisy singular values would ruin the results.(b) Ratio between mean value and uncertainty (signal-to-noise ratio) from the bootstrapping for the seven singularvalues as a function of the lag time. The thin black dashedline indicates the cutoff we have used in applications. Onlysingular values above this line are included in the estima-tion, the number of points above this line corresponds tothe OOM model rank, see Fig. 5(h).

small singular values to have little impact on the dominantspectral and stationary properties of the final OOM, but thiswill be backed up by further theoretical investigation.

Consequently, it seems appropriate to cut off small and sta-tistically unreliable singular values and select a smaller modelrank M < M in Eqs. (45) and (46). In order to determine theuncertainties of the singular values, we use the bootstrappingprocedure, and we discard all singular values with a signal-to-noise ratio of less than 10. This has proven to be a usefulchoice in all applications presented further below. Fig. 4(b)illustrates this procedure for a simple model system.

E. Software, algorithmic details, and analysisof computational effort

We close the Section III of this paper by pointing out afew more details of practical importance. First, while it wasconvenient for the theoretical analysis to assume that all tra-jectories sample the same number of simulation steps K, thisis not required (see Appendix C). Moreover, we also arguein Appendix C that all normalizations in Eqs. (5)–(7) and(16)–(18) can be dropped in practice. All of the matrices

s, Sτ, S

2τr used in the estimation algorithm can be replaced

by integer valued matrices that simply count the number ofvisits, transitions, and two-step transitions.

Second, we have suggested using the bootstrapping pro-cedure in order to estimate uncertainties for the singular values

of the count matrix. One way to realize this is to re-drawtrajectories with replacement from the set of all availablesimulations and to re-estimate the count matrix from thismodified set of simulations. As individual simulations arestatistically independent, this procedure is theoretically jus-tified and can also be used to estimate uncertainties of furtherderived quantities, like time scales and stationary probabilities.We used the trajectory-based bootstrapping in all examplesshown below. However, if only a small number of ratherlong simulations is available, it may be more practical tore-draw individual transitions from the set of all available tran-sitions in the data set. Let T denote the total number of datapoints, which equals T = KQ for the uniform trajectory length,and Eq. (C1) otherwise. If the transitions were statisticallyindependent, one could simply re-sample T transition pairsfrom the set of all N2 possible pairs, where the probabilityof drawing the pair (i, j) is given by S

τ(i, j). In fact, transi-

tions are not statistically independent. Therefore, we suggestto replace the count matrix S

τby the effective count matrix

described in Ref. 33, but it should be noted that this procedurerelies on several approximations and must be improved in thefuture.

Third, we present an overview of the computational costof each step in the estimation algorithm in Table I, assumingthat dense matrix algebra is used in every step. It is expressedin terms of the total number of data points T, the number of

Page 11: correction of estimation bias Feliks Nüske, Hao Wu, Jan-Hendrik … · 2019. 5. 16. · Feliks Nuske,¨ 1,a) Hao Wu,1 Jan-Hendrik Prinz,1 Christoph Wehmeyer,1 Cecilia Clementi,2

094104-10 Nuske et al. J. Chem. Phys. 146, 094104 (2017)

TABLE I. Analysis of computational effort required by the OOM-basedestimation algorithm, if all operations are performed in dense matrix algebra.

Operation Cost

Count matrix estimation ∝TBootstrapping ∝nbTN3

SVD of Sτ

∝N3

Computation of OOM components σ : MN + N2

Ξ : N(N2M + NM2

)ω : ∝M3 + NM2

Transition matrix TτEq N

(2M2 +M

)

MSM states N, the OOM model rank M, and the number ofbootstrapping samples nb.

The first step requires an effort which is linear in the datasize and can be performed efficiently. In most cases, we can

also assume the count matrices Sτ, S

2τr to be sparse, and the

model rank M to be small. In this case, the cubic term appearingfor the calculation of Ξ becomes quadratic, while the contribu-tions of the model rank are small. The only real bottleneck isthe singular value decomposition of S

τ, accounting for the fac-

tor N3 in the second and third steps. As we generally require allsingular values of the count matrix, this step must be performedusing dense matrix algebra, which can be time-consuming.Future research may provide a method that only requires thecomputation of the leading singular values, thus allowing forsparse algebra to be employed.

Lastly, we note that the MSM correction method describedin Sec. III B is available as part of the PyEMMA package,34

version 2.3 or later, see http://pyemma.org.

IV. EXAMPLES

For each of the following examples, we use the trajectory-based bootstrapping strategy to determine the OOM modelrank. Mean values and standard errors for the singular valuesare estimated from nb = 10 000 re-samplings, singular valueswith a signal-to-noise ratio of at least 10.0 are accepted. Wealso generate error estimates for all quantities derived fromthe OOM-based Markov model by trajectory bootstrapping,using 1000 re-samplings. In addition, we compute a con-ventional Markov model without OOM-based correction asa comparison.

A. One-dimensional toy potential

As the first example, we study in more detail the one-dimensional system used in the Introduction. The systemis defined by the double-well potential function shown inFig. 5(a). The dynamics here is a finite state space Markovchain with 100 microstates distributed along the x-axis, wheretransitions can occur between neighboring states based on aMetropolis criterion. The system is kinetically two-state, asthe slowest relaxation time scale of the system, correspondingto the transition process between the two wells, is t2 = 3708steps and clearly dominates all others (Fig. 5(b)).

We investigate the estimation of a seven-state Markovmodel (N = 7) using the discretization indicated by dashed

lines in Fig. 5(a). Using seven states instead of two acceler-ates the convergence of OOM estimates. Still, the seven-statediscretization is a poor one—note that state 4 contains largeparts of the transition region as well as parts of the right min-imum. This choice was made deliberately in order to test therobustness of our method with respect to poor MSM cluster-ings. We produced two different data sets, each comprisingQ = 5000 simulations. The first set contains short simula-tions of length K = 250, while the simulations of the secondset are K = 2000 steps long. For the analysis of the smallerdata set, we can use lag times up to τ = 30, while we can goup to τ = 200 for the larger data set. Panels (c), (e), and (g)of Fig. 5 display the results for the short simulations, whilethe corresponding results for the larger data set are shown inpanels (d), (f), and (h). All simulations were initiated froma non-equilibrium starting distribution, where the probabil-ities to start in each of the seven states are given by thevector

ρ1 =[

0.3 0.3 0.3 0 0.05 0.05 0]

, (64)

that is, 90% of the simulations were started in the left threestates, while only 10% were initialized in the deeper mini-mum on the right. Within each state, the actual microstate wasselected from a uniform distribution.

Figs. 5(c) and 5(d) compare estimates of stationary prob-abilities from direct MSMs based on Eq. (19) and correctedMSMs with the transition matrix given by Eq. (55). Due tothe non-equilibrium initial distribution, the simulations visitthe left minimum much more frequently than a simulation inequilibrium would do. While the MSM estimates of the station-ary distribution converge to the true equilibrium distribution atlong lag times, they are surprisingly inaccurate at short times,where the effect of the non-equilibrium starting distributionstill has a strong effect. Even at the largest lag time τ = 200,the bias is still visible. In contrast, the corrected MSM pro-vides an excellent and stable estimate at lag times of 15 stepsor longer.

In Figs. 5(e) and 5(f), we compare estimates of the slowestimplied relaxation time scale t2 from three different estima-tors: a direct Markov model based on Eq. (19), the correctedMarkov model based on Eq. (55), and the OOM-based spectralestimation, Eqs. (62) and (63). First, we notice that the directand corrected MSMs provide different estimates because ofthe combination of non-equilibrium starting points and thepoor discretization quality. The corrected MSM time scalesconverge faster to the true time scales than the uncorrectedones. Second, the OOM-based direct estimation of relaxationtime scales by Eq. (63) provides accurate results already atlag time τ = 15, which is a regime where the number of rel-evant relaxation processes cannot be easily determined bya time scale separation, see again panel (b) of Fig. 5. TheOOM time scale estimates become very accurate for largerlag times if more data can be used. Third, the large devia-tion between the corrected MSM and the OOM time scalesare indicative of the poor discretization quality employedhere.

Finally, in Figs. 5(g) and 5(h) we show the model rankselected by the bootstrapping procedure as a function of thelag time. We can observe how our criterion based on statistical

Page 12: correction of estimation bias Feliks Nüske, Hao Wu, Jan-Hendrik … · 2019. 5. 16. · Feliks Nuske,¨ 1,a) Hao Wu,1 Jan-Hendrik Prinz,1 Christoph Wehmeyer,1 Cecilia Clementi,2

094104-11 Nuske et al. J. Chem. Phys. 146, 094104 (2017)

FIG. 5. (a) One-dimensional potential function and dis-cretization of the landscape into seven states. (b) Decadiclogarithm of the first nine implied time scales of themodel system. ((c) and (d)) Estimates of the stationaryprobability of states 1-3 from the direct MSM (green)and the corrected MSM (red), compared to the reference(black dashed line). ((e) and (f)) Estimates of the slowestrelaxation time scale t2 from a direct MSM (green), thecorrected MSM (red), and the OOM-based spectral esti-mation (blue), compared to the reference (black dashedline). ((g) and (h)) Model rank selected by the bootstrap-ping procedure. For all quantities derived from the OOM,the dashed lines indicate the estimated values using thecomplete data set, whereas the bullets and errorbars corre-spond to mean and standard error from the bootstrappingprocedure. Note that the errorbars are hardly visible inpanels (d) and (f).

uncertainties helps to select an appropriate model rank for eachlag time, even when it is not obvious from the time scale plot.As expected, the system becomes effectively of rank 2 for lagtimes τ ≥ 80.

B. Molecular dynamics simulationsof alanine dipeptide

Our second example is molecular dynamics simulationdata of alanine dipeptide (Ac-A-NHMe) in explicit water.Alanine dipeptide has been used as a model system in numer-ous previous studies, see Refs. 25 and 35 and many others. Itis well known that its dynamics can be described by the two-dimensional space of backbone dihedral angles φ, ψ. Fig. 6(a)shows the equilibrium probability distribution in this spacewith its three metastable minima in the upper left, central left,and central right parts of the plane. The slow dynamics consistsof exchanges between the left and right parts (t2 ≈ 1400 ps)and between the two minima on the left (t3 ≈ 70 ps). Westudy the estimation of a Markov model using the discretiza-tion also indicated in panel (a) of Fig. 6. It was generatedby k-means clustering of the data set described below usingN = 40 cluster centers. We produced an ensemble of roughly11 000 very short simulations of length 20 ps each. Simulationswere initiated from eight different starting structures labelledby the numbers 1-8 in Fig. 6(b), see Appendix A for details.

It can be seen that the resulting empirical distribution doesnot even reach local equilibrium within the three metastableregions.

Like in the previous example, we find that it is possible toobtain precise estimates of stationary probabilities as soon asthe convergence of the OOM-based time scales is achieved. Inpanel (c) of Fig. 6, we compare results for the equilibrium prob-ability of all states in the right part of the plane, from a directMSM and the corrected MSM. For lag times τ ≥ 500 fs, we areable to correct the bias introduced by strong non-equilibriumsampling.

In panels (d) and (f) of Fig. 6, we present estimates of thetwo slowest time scales t2, t3 produced by the same estimatorsas before (OOM in blue, direct MSM in green, and correctedMSM in red). Additionally, the cyan lines correspond to thetime scale estimates of an MSM using equilibrium simula-tions and the same discretization (see Appendix A). We findthat the OOM-based spectral estimation provides accurate timescale estimates for short lag times starting at τ = 500 fs. More-over, we notice that for lag times as small as these, MSM timescales are clearly lower than the true time scales, althougha decent discretization is employed. The difference betweenOOM and MSM estimates indicates that an even finer dis-cretization would be required to match the references at theselag times. The direct estimates, the reference equilibrium time

Page 13: correction of estimation bias Feliks Nüske, Hao Wu, Jan-Hendrik … · 2019. 5. 16. · Feliks Nuske,¨ 1,a) Hao Wu,1 Jan-Hendrik Prinz,1 Christoph Wehmeyer,1 Cecilia Clementi,2

094104-12 Nuske et al. J. Chem. Phys. 146, 094104 (2017)

FIG. 6. Results for alanine dipeptide. (a) Equilibriumdistribution (logarithmic scale) in the space of backbonedihedral angles φ, ψ and cluster centers of a forty-state k-means discretization used to analyze the data. (b) Empir-ical distribution (logarithmic scale) sampled by the datainitiated from eight starting structures indicated by thenumbers 1-8. (c) Equilibrium probability of all states inthe right part of the plane estimated from the direct MSM(green) and the corrected MSM (red). Reference in black.(d) Estimates for the slowest relaxation time scale t2 froma direct MSM (green), the corrected MSM (red), andthe OOM-based estimation (blue). Reference values fromequilibrium simulations are displayed in black. We alsoshow the expected time scale estimate using the sameforty-state discretization if equilibrium data were used(cyan line). (e) Model rank used for the OOM estima-tion as determined by the bootstrapping. (f) The same as(d) for the second slowest time scales t3. For all quan-tities derived from the OOM, the dashed lines indicatethe estimated values using the complete data set, whereasthe bullets and errorbars correspond to mean and stan-dard error from the bootstrapping procedure. Note thatthe errorbars are hardly visible in panels (c) and (f).

scales, and our OOM-based estimates of equilibrium timescales are nearly identical. Only the mean values extractedfrom bootstrapping for t2 seem to be a bit low. This will beinvestigated further.

Finally, the selected model ranks shown in Fig. 6(e) con-firm that our framework can work in situations where low-rankdescriptions of the dynamics using only a few processes arenot adequate.

C. Two-dimensional model system with poordiscretization

Our final example is another finite state space Markovchain in the two-dimensional energy landscape shown inFig. 7(a), defined by 40 × 40 microstates. Here we show thebehavior of different estimators in an extreme case, where thediscretization is so poor that MSM estimates fail completely.Transitions between neighboring states are now possible inboth x- and y-directions, again based on a Metropolis cri-terion. We study the estimation of a Markov model using adiscretization into 16 MSM states, also shown in Fig. 7(a). Ascan be seen in Fig. 7(b), there are two dominant time scales:t2 ≈ 144 000 steps and t3 ≈ 17 000 steps. The next time scale isclearly separated from the first two, after that, there is no moreapparent time scale separation. This time, we fix the simulationlength at K = 5000 steps, i.e., the trajectories are approximately30 times shorter than the slowest time scale. The simulationsare started from a uniform distribution over all microstates.In panels (c)-(h) of Fig. 7, we display the results if the num-ber of simulations is set to Q = 2000 ((c), (e), and (g)) andQ = 10 000 ((d), (f), and (h)).

In Figs. 7(c) and 7(d), we show the estimation resultsfor the equilibrium probability of the states labeled 13, 14,

and 15. We expect it to be difficult to estimate this proba-bility, as the states are blending different metastable regionsand transition regions. It can be observed that the estimationof stationary probabilities is more sensitive to noise, see theresults for Q = 2000. This observation is not surprising, asthe stationary probabilities require accurate estimation of thetwo-step count matrices Eq. (7) from the data, which can bemore difficult for rarely visited states. Still, for Q = 10 000,a reliable estimate is achieved and the biased estimate ofthe direct MSM can be corrected. Another comparison wemake is between the estimates from the corrected MSM andthose from long equilibrium simulations that use the samenumber of total data points, i.e., K = 2000 · 5000= 107 forQ = 2000 and K = 10 000 · 5000= 5 · 107 for Q = 10 000. Weshow mean values and standard errors from roughly 400 longsimulations for Q = 2000, and roughly 900 simulations forQ = 10 000. In both cases, the estimates from long equilib-rium trajectories provide more accurate estimates. In practice,however, one needs to strike a balance between long tra-jectories that are more beneficial for the analysis and shorttrajectories that can be more efficient for sampling and stateexploration.36–38

Again, we also compare the estimates for the slowesttime scales t2 ((e) and (f)) and t3 ((g) and (h)) from a directMSM, the corrected MSM, and the OOM-based spectral esti-mation. In both cases, correct estimates of both time scalescan be obtained from the OOM, while both the direct andcorrected MSMs estimate time scales one order of magnitudetoo small. This suggests that for a bad enough discretization,correcting for the effect of the non-equilibrium starting distri-bution will not be sufficient to achieve convergence in the timescales. However, the poor discretization quality is revealedby a large error between the OOM-based estimate and the

Page 14: correction of estimation bias Feliks Nüske, Hao Wu, Jan-Hendrik … · 2019. 5. 16. · Feliks Nuske,¨ 1,a) Hao Wu,1 Jan-Hendrik Prinz,1 Christoph Wehmeyer,1 Cecilia Clementi,2

094104-13 Nuske et al. J. Chem. Phys. 146, 094104 (2017)

FIG. 7. (a) Two-dimensional potential function with dis-cretization into 16 MSM states indicated by dashed lines.(b) Leading nine implied time scales tm of the system. ((c)and (d)) Estimates of equilibrium probability of states 13,14, and 15 from direct MSM (green) and the correctedMSM (red), compared to the reference (black line) andestimates from 400 (c) / 900 (d) different equilibriumsimulations, shown by the cyan lines. ((e) and (f)) Esti-mates of slowest relaxation time scale t2 from a directMSM (green), the corrected MSM (red), and the OOM-based spectral estimation (blue), compared to the refer-ence (black dashed line). ((g) and (h)) The same for t3.For all quantities derived from the OOM, the dashed linesindicate the estimated values using the complete data set,whereas the bullets and errorbars correspond to mean andstandard error from the bootstrapping procedure. Notethat the errorbars are hardly visible in panels (f) and (h).

corrected MSM, and this observation can be exploited in orderto improve the discretization and repeat the analysis.

V. CONCLUSIONS

We have investigated the quality of Markov state mod-els when estimated from many simulations of short length,initiated from non-equilibrium starting conditions. We havederived an expression for the error between unbiased MSMtransition probabilities and the expected estimate from manyshort simulations. This error is shown to depend on the sim-ulation length, the lag time, and the state discretization. Ifultra-long trajectories are employed, i.e., trajectories that arelong compared to the slowest relaxation time scales, then theeffect of the initial distribution is negligible and no furthercorrection is needed. For ensembles of short trajectories, thesituation is more complex. Preparing simulation trajectories insuch a way that they emerge from a local equilibrium distri-bution does not appear to be of much practical use: this wouldonly correct the first transition count of every trajectory whilethe subsequent trajectory segments are still biased. The localequilibrium will be lost for intermediate times along the tra-jectory as the trajectory ensemble is not in global equilibrium.In a similar sense discarding initial simulation fragments canreduce the bias but cannot systematically remove it. In parti-cular, since the effect of the bias disappears with the slowest

relaxation times of the system, discarding pieces of simula-tion trajectories appears more harmful in terms of reducingthe statistics than it is useful to reduce the bias. With the stan-dard MSM estimator, the most effective and simplest methodto reduce the bias from the initial trajectory distribution in factseems to be using a longer lag time or a better state spacediscretization. These are already the usual objectives of MSMconstruction. However, if the discretization is poor, the estima-tion bias due to a non-equilibrium distribution can be dramaticat practically usable lag times.

The main result of this paper is that we propose animproved estimator of the MSM transition matrix which isnot biased by the initial distribution. This new estimator isbased on the theory of observable operator models. In contrastto the standard MSM estimator, the corrected MSM estimatordoes not only use the number of transitions observed betweenpairs of states at lag time τ but also the number of transitionsat lag time 2τ. These statistics are combined to get a transitionmatrix estimate at lag time τ that is unbiased by the initial tra-jectory distribution. While it may seem that having to estimatestatistics at 2τ is a deficiency compared to standard MSM esti-mation when only short simulation trajectories are available,please note that the corrected MSM estimator can get signifi-cantly better estimates at short lag times, so in practice the lagtimes needed for a converged MSM will be smaller than forthe standard estimator.

Page 15: correction of estimation bias Feliks Nüske, Hao Wu, Jan-Hendrik … · 2019. 5. 16. · Feliks Nuske,¨ 1,a) Hao Wu,1 Jan-Hendrik Prinz,1 Christoph Wehmeyer,1 Cecilia Clementi,2

094104-14 Nuske et al. J. Chem. Phys. 146, 094104 (2017)

Finally, we report a result from the OOM frameworkthat shows how the model-free relaxation time scales canbe computed from the same statistics used for the correctedMSM estimator (i.e., transition matrices at lag times τ and2τ). These estimates are only impaired by statistical errorbut are not affected by systematic MSM error as no MSMis used in the process of obtaining them. The differencebetween the corrected MSM time scales and the OOM timescales can be used in order to assess the discretization qual-ity, as this difference goes to zero in the limit of gooddiscretization.

This paper addresses the long-standing controversy aboutthe correct use of simulation data from short non-equilibriumsimulations for MSM estimation and their effect on theestimation of equilibrium expectations and kinetics.

ACKNOWLEDGMENTS

This work was funded by Deutsche Forschungsgemein-schaft through SFB 958 and SFB 1114 and by the Euro-pean Commission through ERC starting grant “pcCell.” C.C.is supported by the National Science Foundation (GrantNo. CHE-1265929) and the Welch Foundation (Grant No.C-1570).

APPENDIX A: SIMULATION SETUPOF ALANINE DIPEPTIDE

Molecular dynamics simulations of alanine dipeptidein explicit water at temperature 300 K were generated withACEMD39 software using the AMBER ff-99SB-ILDN forcefield40 and an integration time step of 2 fs. The peptidewas simulated inside a cubic box of volume (2.3222 nm)3

containing 651 TIP3P water molecules. The Langevin ther-mostat was used. The electrostatics were computed every twotime steps by the particle-mesh Ewald (PME) method,41 usingreal-space cutoff 0.9 nm and grid spacing 0.1 nm. All bondsbetween hydrogens and heavy atoms were constrained.

We have produced 11 388 ultra short simulations of length20 ps each, with 50 fs saving interval. The simulations wereinitiated from eight different structures; their projections intoφ−ψ-space are indicated by the numbers 1-8 in Fig. 6(b). Theprobabilities to start in each of these structures are given bythe vector

ρ1 =[

0.05 0.05 0.2 0.2 0.2 0.1 0.1 0.1]

. (A1)

These simulations were used to perform the analyses describedin Sec. IV B. Using the same setup, we produced 2363 longruns of 1 ns simulation time each, with 1 ps saving interval. Weestimated a Markov model on the 40-state k-means discretiza-tion at lag time τ = 100 ps using this data set, and extracted thereference time scales and equilibrium probabilities shown asblack lines in Fig. 6. Also, we used the stationary probabilitiesestimated from this model to initialize 203 short equilibriumruns of 500 ps simulation time each, with 100 fs saving inter-val. This data set was used to compute the equilibrium timescales of the k-means discretization shown as cyan lines inFigs. 6(d) and 6(f).

APPENDIX B: OOM PROBABILITYOF OBSERVATION SEQUENCE

Here, we show the derivation of the path probability for-mula Eq. (36), that can also be found in Ref. 21. In general,the left-hand side of Eq. (36) can be expressed by repeatedintegrals over the transition kernel,

P(Xτ ∈ A1, . . . , Xlτ ∈ Al) =∫Ω

∫A1

. . .

∫Al

dx0 . . . dxl π(x0)p(x0, x1; τ) . . . p(xl−1, xl; τ). (B1)

Note that π appears in the first integral as we assumed that the dynamics is in equilibrium, i.e., the initial distribution equals π.Next, we replace all transition kernels by the expansion in Eq. (4),

P(Xτ ∈ A1, . . . , Xlτ ∈ Al) =M∑

m0=1

M∑m1=1

· · ·

M∑ml−1=1

[∫Ω

dx0 π(x0)ψm0 (x0)

]λm0 (τ)

×

[∫A1

dx1 ψm0 (x1)π(x1)ψm1 (x1)

]. . . λml−1 (τ)

[∫Al

dxl ψml−1 (xl)π(xl)

](B2)

=

M∑m0=1

M∑m1=1

· · ·

M∑ml−1=1

δ1,m0ΞA1 (m0, m1) . . .ΞAl (ml−1, 1). (B3)

In the second equation, we have used the π-orthogonalityof the eigenfunctions ψm0 and the fact that ψ1 ≡ 1 in orderto replace the x0-integral by δ1,m0 . For the last integral, wehave also used that ψ1 ≡ 1. This is a sequence of matrix-vector products. It remains to use δ1,m0 =ω(m0) and thatΞAl (ml−1, 1)=

[ΞAlσ

](ml−1). In matrix notation, Eq. (36)

follows:

P(Xτ ∈ A1, . . . , Xlτ ∈ Al) = ωTΞA1 . . .ΞAlσ (B4)

Finally, note that this derivation also works if the dynam-ics is not in equilibrium. In this case, the vector ω isgiven by ω(m0) = ∫Ω dx0 ρ0(x0)ψm0 (x0), where ρ0 is thenon-equilibrium initial condition.

APPENDIX C: VARIABLE SIMULATION LENGTH

Here, we verify that the estimation algorithm fromSec. III B can be applied to data sets comprising simulations of

Page 16: correction of estimation bias Feliks Nüske, Hao Wu, Jan-Hendrik … · 2019. 5. 16. · Feliks Nuske,¨ 1,a) Hao Wu,1 Jan-Hendrik Prinz,1 Christoph Wehmeyer,1 Cecilia Clementi,2

094104-15 Nuske et al. J. Chem. Phys. 146, 094104 (2017)

non-uniform length. We assume that for j = 1, . . . , J , there isan ensemble of Qj simulations of length Kj + 2τ, i.e., K j tran-sition pairs/triples will be used from each of these trajectories.We assume that Qj → ∞ for all j, such that every sub-ensemblesamples from an empirical distribution ρj. Define the numberof data points generated by the jth ensemble as T j = QjK j andthe total number of data points by

T :=J∑

j=1

QjKj. (C1)

Moreover, we assume thatTj

T → αj, i.e., the fraction of datapoints generated by the jth ensemble approaches a constant forall j. Let us define the distribution

ρ =

J∑j=1

αj ρj. (C2)

Trajectories of length Kj + 2τ are enumerated by qj andlabelled Yqj . Further, let sKj (Yqj ) be any of the estimators fromEqs. (5)–(7), where the subscript K j indicates that K − 2τ inEqs. (5)–(7) must be replaced by K j. In addition, denote bys(Yqj ) the same estimator but without the normalization. Also,let cρj denote the corresponding correlation from Eqs. (20),(21), and (43) with respect to the density ρj. It follows that

sT :=1T

Q1∑q1=1

s(Yq1 ) + · · · +QJ∑

qJ=1

s(YqJ )

(C3)

=T1

T

1T1

Q1∑q1=1

s(Yq1 )

+ · · · +

TJ

T

1TJ

QJ∑qJ=1

s(YqJ )

(C4)

=T1

T

1Q1

Q1∑q1=1

sK1 (Yq1 )

+ · · · +

TJ

T

1QJ

QJ∑qJ=1

sKJ (YqJ )

(C5)

→ α1E(sK1

)+ · · · + αJE

(sKJ

)(C6)

= α1cρ1 + · · · + αJcρJ (C7)

= cρ. (C8)

The convergence in Eq. (C6) is convergence in probability.Thus, if we sum up all visits/transitions/two-step transitions,and divide by the total number of data points in the end, wearrive at an asymptotically correct estimator of the correlationswith respect to the density ρ. As the OOM estimation algo-rithm only relies on consistent estimators for correlations withrespect to some empirical density ρ, it can still be applied inthis setting. Finally, the normalization by 1

T can be omitted inpractice because it cancels out in Eqs. (47)-(52).

1J.-H. Prinz, H. Wu, M. Sarich, B. Keller, M. Senne, M. Held, J. D. Chodera,C. Schutte, and F. Noe,“Markov models of molecular kinetics: Generationand validation,” J. Chem. Phys. 134, 174105 (2011).

2C. Schutte and M. Sarich, Metastability and Markov State Models in Molec-ular Dynamics, Modeling, Analysis, Algorithmic Approaches, Courant Lec-ture Notes Vol. 24 (American Mathematical Society and Courant Instituteof Mathematical Sciences, 2013).

3An Introduction to Markov State Models and Their Application to LongTimescale Molecular Simulation, edited by G. R. Bowman, V. S. Pande,and F. Noe (Springer, Netherlands, 2014).

4C. Schutte, A. Fischer, W. Huisinga, and P. Deuflhard, “A direct approachto conformational dynamics based on hybrid Monte Carlo,” Comput. Phys.151, 146–168 (1999).

5W. C. Swope, J. W. Pitera, and F. Suits, “Describing protein folding kineticsby molecular dynamics simulations. 1. Theory,” J. Phys. Chem. B 108(21),6571–6581 (2004).

6F. Noe, I. Horenko, C. Schutte, and J. C. Smith, “Hierarchical analy-sis of conformational dynamics in biomolecules: Transition networks ofmetastable states,” J. Chem. Phys. 126, 155102 (2007).

7J. D. Chodera, K. A. Dill, N. Singhal, V. S. Pande, W. C. Swope, andJ. W. Pitera, “Automatic discovery of metastable states for the constructionof Markov models of macromolecular conformational dynamics,” J. Chem.Phys. 126, 155101 (2007).

8F. Noe, “Probability distributions of molecular observables computed fromMarkov models,” J. Chem. Phys. 128(24), 244103 (2008).

9N. V. Buchete and G. Hummer, “Coarse master equations for peptide foldingdynamics,” J. Phys. Chem. B 112, 6057–6069 (2008).

10G. R. Bowman, K. A. Beauchamp, G. Boxer, and V. S. Pande, “Progressand challenges in the automated construction of Markov state models forfull protein systems,” J. Chem. Phys. 131, 124101 (2009).

11F. Noe, C. Schutte, E. Vanden-Eijnden, L. Reich, and T. R. Weikl, “Con-structing the full ensemble of folding pathways from short off-equilibriumsimulations,” Proc. Natl. Acad. Sci. 106, 19011–19016 (2009).

12V. A. Voelz, G. R. Bowman, K. Beauchamp, and V. S. Pande, “Molecularsimulation of ab initio protein folding for a millisecond folder NTL9,” J. Am.Chem. Soc. 132(5), 1526–1528 (2010).

13W. Zhuang, R. Z. Cui, D.-A. Silva, and X. Huang, “Simulating the T-jump-triggered unfolding dynamics of trpzip2 peptide and its time-resolved IRand two-dimensional IR signals using the Markov state model approach,”J. Phys. Chem. B 115, 5415–5424 (2011).

14M. Sarich, F. Noe, and C. Schutte, “On the approximation quality of Markovstate models,” Multiscale Model. Simul. 8, 1154–1177 (2010).

15J. Chodera, V. S. Pande, M. Weber, F. Noe, and C. Schutte, personalcommunication.

16S. Roblitz, “Statistical error estimation and grid-free hierarchical refine-ment in conformation dynamics,” Ph.D. thesis, Freie Universitat Berlin,2009.

17M. Weber, “A subspace approach to molecular Markov state models via aninfinitesimal generator,” Technical Report 09–27, ZIB, 2009.

18X. Huang, G. R. Bowman, S. Bacallado, and V. S. Pande, “Rapid equilib-rium sampling initiated from nonequilibrium data,” Proc. Natl. Acad. Sci.106(47), 19765–19769 (2009).

19G.R. Bowman, X. Huang, and V. S. Pande, “Using generalized ensemblesimulations and Markov state models to identify conformational states,”Methods 49(2), 197–201 (2009).

20H. Jaeger,“Observable operator models for discrete stochastic time series,”Neural Comput. 12(6), 1371–1398 (2000).

21H. Wu, J.-H. Prinz, and F. Noe, “Projected metastable Markov processes andtheir estimation with observable operator models,” J. Chem. Phys. 143(14),144101 (2015).

22H. Wu and F. Noe, “Spectral learning of dynamic systems from nonequi-librium data,” in Advances in Neural Information Processing Systems 29,edited by D. D. Lee, M. Sugiyama, U. V. Luxburg, I. Guyon, and R. Garnett(Curran Associates, Inc., 2016), pp. 4179–4187.

23N. Djurdjevac, M. Sarich, and C. Schutte, “Estimating the eigenvalueerror of Markov state models,” Multiscale Model. Simul. 10(1), 61–81(2012).

24F. Noe and F. Nuske, “A variational approach to modeling slow processesin stochastic dynamical systems,” Multiscale Model. Simul. 11, 635–655(2013).

25F. Nuske, B. G. Keller, G. Perez-Hernandez, A. S. J. S. Mey, and F. Noe,“Variational approach to molecular kinetics,” J. Chem. Theory Comput. 10,1739–1752 (2014).

26M. Weber, “Meshless methods in conformation dynamics,” Ph.D. thesis,Freie Universitat Berlin, 2006.

27B. Trendelkamp-Schroer, H. Wu, F. Paul, and F. Noe, “Estimation anduncertainty of reversible Markov models,” J. Chem. Phys. 143(17), 174101(2015).

28D. E. Shaw, P. Maragakis, K. Lindorff-Larsen, S. Piana, R. O. Dror,M. P. Eastwood, J. A. Bank, J. M. Jumper, J. K. Salmon, Y. Shan, and W.Wriggers, “Atomic-level characterization of the structural dynamics ofproteins,” Science 330, 341–346 (2010).

29K. Lindorff-Larsen, S. Piana, R. O. Dror, and D. E. Shaw, “How fast-foldingproteins fold,” Science 334(6055), 517–520 (2011).

30F. Noe, H. Wu, J.-H. Prinz, and N. Plattner, “Projected and hidden Markovmodels for calculating kinetics and metastable states of complex molecules,”J. Chem. Phys. 139, 184114 (2013).

Page 17: correction of estimation bias Feliks Nüske, Hao Wu, Jan-Hendrik … · 2019. 5. 16. · Feliks Nuske,¨ 1,a) Hao Wu,1 Jan-Hendrik Prinz,1 Christoph Wehmeyer,1 Cecilia Clementi,2

094104-16 Nuske et al. J. Chem. Phys. 146, 094104 (2017)

31J.-H. Prinz, “Advanced estimation methods for Markov models of dynami-cal systems,” Ph.D. thesis, Freie Universitat Berlin, 2012.

32G. W. Stewart, “Perturbation theory for the singular value decomposition,”in SVD and Signal Processing, II: Algorithms, Analysis and Applications,edited by R. J. Vaccaro (Elsevier, 1991).

33F. Noe, “Statistical inefficiency of Markov model count matrices”(unpublished).

34M. K. Scherer, B. Trendelkamp-Schroer, F. Paul, G. Perez-Hernandez,M. Hoffmann, N. Plattner, C. Wehmeyer, J.-H. Prinz, and F. Noe, “PyEMMA2: A software package for estimation, validation, and analysis of Markovmodels,” J. Chem. Theory Comput. 11, 5525 (2015).

35F. Vitalini, A. S. J. S. Mey, F. Noe, and B. G. Keller, “Dynamic propertiesof force fields,” J. Chem. Phys. 142(8), 084101 (2015).

36J. Preto and C. Clementi, “Fast recovery of free energy landscapes viadiffusion-map-directed molecular dynamics,” Phys. Chem. Chem. Phys.16, 19181–19191 (2014).

37S. Doerr and G. De Fabritiis, “On-the-fly learning and sampling of lig-and binding by high-throughput molecular simulations,” J. Chem. TheoryComput. 10(5), 2064–2069 (2014).

38S. Doerr, M. J. Harvey, F. Noe, and G. De Fabritiis, “HTMD: High-throughput molecular dynamics for molecular discovery,” J. Chem. TheoryComput. 12(4), 1845–1852 (2016).

39M. J. Harvey, G. Giupponi, and G. De Fabritiis, “ACEMD: Acceleratingbiomolecular dynamics in the microsecond time scale,” J. Chem. TheoryComput. 5(6), 1632–1639 (2009).

40K. Lindorff-Larsen, S. Piana, K. Palmo, P. Maragakis, J. L. Klepeis,R. O. Dror, and D. E. Shaw, “Improved side-chain torsion potentials forthe Amber ff99SB protein force field,” Proteins: Struct., Funct., Bioinf.78(8), 1950–1958 (2010).

41T. Darden, D. York, and L. Pedersen, “Particle mesh Ewald: An N -log(N)method for Ewald sums in large systems,” J. Chem. Phys. 98, 10089–10092(1993).