Collective Variables from Local Fluctuations - The Journal of Physical

We introduce a method to obtain one-dimensional collective variables for studying rarely occurring transitions between two metastable states separated...
2 downloads 0 Views 2MB Size
Letter Cite This: J. Phys. Chem. Lett. 2018, 9, 2776−2781

pubs.acs.org/JPCL

Collective Variables from Local Fluctuations Dan Mendels,†,‡,¶ GiovanniMaria Piccini,†,‡,¶ and Michele Parrinello*,†,‡ †

Department of Chemistry and Applied Biosciences, ETH Zurich, c/o USI Campus, Via Giuseppe Buffi 13, CH-6900 Lugano, Ticino, Switzerland ‡ Facoltá di Informatica, Istituto di Scienze Computazionali, Universitá della Svizzera italiana (USI), Via Giuseppe Buffi 13, CH-6900 Lugano, Ticino, Switzerland S Supporting Information *

ABSTRACT: We introduce a method to obtain one-dimensional collective variables for studying rarely occurring transitions between two metastable states separated by a high free energy barrier. No previous information, not even approximated, on the path followed during the transition is needed. The only requirement is to know the fluctuations of the system while in the two metastable states. With this information in hand, we build the collective variable using a modified version of Fisher’s linear discriminant analysis. The usefulness of this approach is tested on the metadynamics simulation of two representative systems. The first is the freezing of silver iodide into the superionic α-phase, and the second is the study of a classical Diels−Alder reaction. The collective variable works very well in these two diverse cases.

A

Over the years, a vast amount of experience has been gathered and a large number of CVs has been programmed and made publically available for use in connection with several enhanced sampling methods.17 However, whenever a new class of problems needs to be tackled, the construction of new CVs requires some effort. While this construction can be very instructive per se,18 it would be useful to have a simple and effortless method to construct efficient and low-dimensional CVs. One could distinguish between two types of CVs. The first type does not assume that the final state is known a priori and is generally used in an exploratory mode. Typical examples are the potential energy19 or CVs that are used as a surrogate for the entropy.20,21 The second type instead assumes that the initial (A) and final state (B) are known and is used to measure free energy profiles. Typical examples are the so-called path CVs.22 The rational behind our theory is best understood if we recall the scenario that underpins the success of the finite temperature string method23,24 or path-CV metadynamics. These theories are based on the observation that in a rare event scenario the ensemble of all the reactive trajectories that go from A to B form a tube whose center lies along a minimum free energy path. The finite temperature string method is based on finding the minimum free energy path, while in path CV metadynamics this condition can be somewhat relaxed, as considerable empirical evidence has shown.

central problem in modern day atomistic simulations is the study of systems exhibiting complex free energy landscapes in which long-lived metastable states separated by kinetic bottlenecks are present. The kinetic bottlenecks impede transition between metastable states and restrict the time scale that can be explored.1 Special purpose machines2 have been constructed to push this limit. However, in spite of remarkable progress, problems like drug unbinding or first-order phase transitions such as nucleation occur on a time scale that puts them out of the reach of direct simulations, not to mention the fact that such machines can be accessed only by a restricted number of researchers. This has motivated a vast community of modelers to develop enhanced sampling methods that allow investigating the properties of complex systems in an affordable computational time, overcoming kinetic bottlenecks, and exploring different metastable states. It is not practical to review here the vast literature on the subject, and we refer the reader to recent publications for an updated review.3,4 Following the pioneering work of Torrie and Valleau,5 a large class of enhanced sampling methods rely on the identification of an appropriate set s(R) of collective variables (CVs). The s(R) are appropriately chosen functions of the atomic coordinates R. They should be able to distinguish between different metastable states and describe those modes that are difficult to sample. Their choice is crucial for a successful simulation. Many enhanced sampling methods can be described as ways of enhancing the fluctuations of the CVs so as to favor transitions from one metastable state to another. Two such methods are metadynamics6−8 or variationally enhanced sampling,9 but other methods can be similarly described.10−16 © XXXX American Chemical Society

Received: March 9, 2018 Accepted: May 7, 2018 Published: May 7, 2018 2776

DOI: 10.1021/acs.jpclett.8b00733 J. Phys. Chem. Lett. 2018, 9, 2776−2781

Letter

The Journal of Physical Chemistry Letters

Now the stage is set to describe our approach. If we look at the problem from the more forgiving foliation point of view, it is then worthwhile to consider the simplest possible foliation, namely, the one in which the space is partitioned into planar surfaces of dimension Nd − 1. Any such foliation is defined, once the direction orthogonal to the planes is given. Let this direction be determined by the vector W. As argued above, we have some leeway in the choice of W. Of course the transition region, namely, the region separating the two basins, is the one requiring the most attention. The ideal choice would be to choose W to be perpendicular to the FES discriminant surface that distinguishes between basin A and basin B. Because calculating the FES is the very object of our method, this is not possible, and even if we had the FES in all its full glory, calculating the discriminant would be somewhat challenging. Here we take an approach that has many times been followed in quantum chemistry, namely we try to infer properties of transition states from the fluctuations of the systems while in the two metastable states. The most famous example of such an approach is the celebrated Marcus theory of electron transfer. A saving grace is that we do not need to know the precise position of the discriminant but only its direction. To help us in this endeavor, we call in the help of artificial intelligence methods and use linear discriminant analysis (LDA) to identify the optimal direction in which to foliate the space for the purpose of driving the reaction. This type of analysis was introduced to classify a set of data into two classes and recently used to distinguish between biomolecules’ conformations.27,28 Not surprisingly, because here the purpose is different, we had to modify for optimal performance the original LDA applied to distinguish between different protein formulations and introduce a variant that we call harmonic linear discriminant analysis (HLDA). We exemplify the usefulness of HLDA on a crystal phase transformation of the ionic compound AgI and to a prototypical Diels−Alder reaction before finishing the Letter with some discussion. Linear Discriminant Analysis for Metastable States. As discussed above, we assume that the system is well-described once the free energy surface F(d) as a function of the Nd descriptors is given. On this surface lie the metastable states A and B. While trapped in these two states, the descriptors will have an expectation value μA,B and a multivariate variance ΣA,B. These quantities can be evaluated in short unbiased runs, and we assume here this to be case. The data generated in such runs will be separated in the Nd dimensional space of the descriptors. Our first goal is to find a one-dimensional projection of these two sets of data along which they still do not overlap. If we take a generic projection x = WT d, where W is a vector in the Nd space, there is no guarantee that they will not overlap. The linear discriminant analysis aims at finding the direction W along which the projected data are best separated. To do this one needs a measure of their degree of separation. Following Fisher, this is given by the ratio between the so-called between class Sb and the within class Sw scatter matrices. The former is measured by the square of the distance between the projected means

In order to understand why, we recall some of the features of path CV metadynamics. Let the free energy surface (FES) be spanned by Nd descriptors di(R) that are a function of the atomic coordinates R and are arranged to form a vector d(R). In this free energy space, one defines a reference path d(t) parametrized such that one is in A for t = 0 and reaches the B state when t = 1. In ref 25, two CVs where defined: one that measures the progression along the path (s) and another that measures the distance from the path (z). Here, only the first one is of interest and reads 1

s = lim

λ →∞

2

∫0 t e−λ(d(R) − d(t )) dt 1

2

∫0 e−λ(d(R) − d(t )) dt

(1)

In the first applications it was suggested to optimize d(t), but very soon it was realized that even reference paths that depart within limits from the optimal one could be successfully employed. To understand this empirical fact, we note that the surfaces of constant s partition the space into hyper surfaces of dimension Nd − 1. This is often referred to as a foliation of the space. Close to d(t), the surfaces become flat and locally orthogonal to d(t) (see Figure 1). In the following we shall

Figure 1. Free energy surface showing two basins of attraction connected by the MFEP (dashed black line). The solid white lines are the contour isolines relative to the path variable (s) calculated using the MFEP as reference. The data are taken from ref 26 for the SN2 reaction CH3F + Cl− → CH3Cl + F−.

suppose that d(t) threads only once the foliating surfaces. If this were not the case, it would signal a poor choice of descriptors, a problem that could possibly be solved by increasing the number of descriptors. To understand the observed relative insensitivity to the choice of d(t), it is convenient to change our perspective and focus on the dual point view of the surfaces that foliate the space rather than on d(t) itself. We then imagine to start a simulation from a position not too distant from the minimum free energy path and then let the system evolve under the condition that s is kept constant. The spontaneous evolution of the system will bring it close to the minimum of the intersection between the fixed s surface and the underlining FES. Thus, a bias based on an imperfect d(t) still drags the system along the minimum free energy path. Of course, the more orthogonal the foliating surfaces are to the path direction, the more efficient the simulation. It is this self-focusing effect that we will count on in the following.

μà − μ̃B = WT(μA − μB )

(2)

and can be written as WTS bW

(3)

where 2777

DOI: 10.1021/acs.jpclett.8b00733 J. Phys. Chem. Lett. 2018, 9, 2776−2781

Letter

The Journal of Physical Chemistry Letters

Figure 2. Upper panel: free energy surface along entropy, sS, and enthalpy, sH, collective variables. The green and orange dots represent the points sampled in the unbiased basins, and the red dots are the points sampled using metadynamics with a standard (arithmetic average) LDA CV (coefficients: sS, − 0.7686 and sH, 0.6397), left panel, and HLDA (coefficients: sS, −0.9957 and sH, 0.09244), right panel. The dashed lines are the distribution discriminants to which the CV is normal. Lower panel: time evolution of the arithmetic LDA CV, left panel, and of the HLDA CV, right panel. The green and orange lines represent the projections of the unbiased distributions on the discriminant plane. The FES along sS and sH reported here has been calculated from the data of ref 29.

S b = (μA − μB )(μA − μB )T

the state with the smaller fluctuations, which is more difficult to get into and escape from. For these reasons, we propose a different measure of the scatter, and rather than using of the arithmetic average, we base the measure of the within scatter matrix on the harmonic average as follows:

(4)

The within spread is estimated from the sum of the two spreads, leading to the expression WTSw W

(5)

with

Sw =

Sw = ΣA + ΣB

(6)

The Fisher’s object function then reads like a Rayleigh ratio W S bW WTSw W

(10)

(11)

From a machine learning point of view, the more compact states are better defined and thus have a larger weight in determining the discriminants. In our experience, this second choice has proven to be far superior. Liquid−Superionic Phase Transition in AgI. For completeness sake and following a consolidated tradition, we have studied alanine dipeptide. The results are shown in the Supporting Information. Here we have instead chosen to present a more challenging case, namely, the transition from liquid to the αphase of AgI. In this interesting transition, the iodine anions order to form a BCC lattice while the silver cations remain highly mobile. This partially ordered phase is an example of a superionic phase.

(8)

Following up on the introductory discussion, this would suggest that a useful one-dimensional CV is s LDA (R) = (μA − μB )T (ΣA + ΣB)−1d(R)

1 ΣB

⎛ 1 1 ⎞ s HLDA (R) = (μA − μB )T ⎜ + ⎟d(R) ΣB ⎠ ⎝ ΣA

(7)

that is maximized by W * = S−w1(μA − μB )

+

This leads to a different expression for the collective variable that reads

T

1(W) =

1 1 ΣA

(9)

As we shall see below, the performance of such a CV proved to be mediocre. One possible explanation could be that when taking the arithmetic averages of the covariances, it is the larger one that carries more weight. From a data analysis point of view, this seems a bit counterintuitive because the data with smaller variance are better defined. Also, from the rare event point, it would seem more appropriate to give more weight to 2778

DOI: 10.1021/acs.jpclett.8b00733 J. Phys. Chem. Lett. 2018, 9, 2776−2781

Letter

The Journal of Physical Chemistry Letters

Thus, we used these six parameters as a basis for our discriminant analysis, performing two unbiased runs of 40 ps for the reactant and product states, respectively. Because, in principle, in the gas phase the reactants molecules can be uniformly distributed all over the volume, we applied harmonic restraints on the d4 and d6 distances for values greater than 6 Å. This allows the reactants to explore a significant part of the conformational space. Having considered these two distributions for the six distances, we applied HLDA to them. In our hands, the result is qualitatively intuitive. The coefficients of the linear combination reported in Table 1 show that the largest

Recently, we have studied this compound using metadynamics29 and found that a free energy surface can be drawn using enthalpy sH and a surrogate for the entropy sS as CVs. The ability to represent this complex transition with only two CVs offers the possibility of demonstrating how our prescription for constructing a one-dimensional CV works. The fluctuations around these two minima are, to a good approximation, Gaussian in form and can be described by their fluctuation matrices ΣA and ΣB. In Figure 2, the points sampled during these two unbiased runs are shown and are clearly separated. We can now construct two different one-dimensional CVs, one built using LDA sLDA (eq 9) and another using HLDA sHLDA (eq 11), making sure that the covariance matrices are well-converged. In Figure 2 we contrast the orientations of the line orthogonal to the direction defined by W. They appear to be rather different (the computational details of the calculations can be found in the Supporting Information). In the transition region between A and B this different orientation has significant influence on the quality of the CV. In the case of sHLDA, the points sampled bunch nicely to form a tube whose center closely follows the minimum free energy path. In the sLDA case, instead the points in the transition path branch out to follow different pathways. This is a reflection of a somewhat hysteretic behavior, as can be seen in the lower panels of Figure 2, where it is also shown that the points generated in the first two unbiased runs are better separated in the one-dimensional projection along sHLDA. Diels−Alder Reaction: [4 + 2] Cycloaddition of 1,3-Butadiene and Ethene. As an example of application to chemical reactions, we study a classical Diels−Alder process, namely the [4 + 2] cycloaddition of 1,3-butadiene and ethene. The essence of a chemical reaction can be summarized as a combination of simultaneous bonds breaking and formation. This is accompanied by a time-crucial conformational reorganization30 and a local change of distances. Such a bond reorganization can be a dramatic event involving breaking and/or formation of σ-type bonds but can also be accompanied by local electronic rearrangements, resulting in a nonlocal strengthening or weakening of other bonds, like the π−σ conversion. The chemical scheme for a classical Diels−Alder reaction is reported in Figure 3. The peculiarity of a Diels−Alder reaction

Table 1. LDA Coefficients for the Reaction Collective Variable W*

d1

d2

d3

d4

d5

d6

0.05

−0.21

0.08

−0.69

0.00

−0.68

contributions are related to the d4 and d6 distances that have the same sign and approximately the same amplitude, indicating that in the concerted mechanism the two partners have to come close simultaneously. Moreover, they show that while for instance d4 and d6 are shrinking the d2 distance is shrinking as well as going from a σ-type to a π-type bonding. At the same time, all the bonds that go from a shorter π-type to a longer σtype, namely, distances d1 and d3, have an opposite sign reflecting the fact that they must elongate while the other three are shrinking and vice versa. In Table 1 in the Supporting Information, the coefficients of arithmetic LDA are reported. It is clear from the magnitude of the coefficients that the chemically meaningful arguments described above do not hold anymore in this case; thus, any enhanced sampling along such a variable would not lead to any desired transition. To demonstrate the power of such a method, we report in Figure 4 the free energy profile along sHLDA (the computational

Figure 3. Reaction scheme of the [4 + 2] cycloaddition of 1,3butadiene and ethene. In the stick model in the lower part of the figure, the carbon−carbon distances used as descriptors of the process are reported.

Figure 4. Free energy profile for the [4 + 2] cycloaddition of 1,3butadiene and ethene along the HLDA CV.

lays in the fact that the formation of the two σ bonds in the product state implies that three π bonds of the reactant state become two σ bonds while a σ bond becomes a π bond. Therefore, although it is clear from Figure 3 that distances d4 and d6 will play a major role in the reaction progression as they govern the approach of the two molecules to each other, the concerted elongation−contraction of distances d1, d2,d3, and d5 will also play a significant role.

details of the calculations can be found in the Supporting Information). The HLDA CV, besides being rather efficient, clearly brings out the chemistry of the problem with a wide entropic basin and a narrow and deep enthalpic state. Furthermore, the configurations extracted from the apparent transition region do correspond to those that are described in standard text books as transition states (see Figure 5). In the final state, two possible configurations are possible, namely endo 2779

DOI: 10.1021/acs.jpclett.8b00733 J. Phys. Chem. Lett. 2018, 9, 2776−2781

Letter

The Journal of Physical Chemistry Letters

data is the sought-after collective variable. Arguments are given why this procedure should be useful. The validity of our approach is borne out by two very successful calculations to two typical areas of application of rare events methods, i.e. nucleation and chemical reactions. The CVs obtained by using our HLDA approach perform very satisfactorily and lead to highly meaningful results. The reasons for this success will need to be investigated further because they provide information regarding the nature of rare transitions, in particular for those concerning chemical reactions. Much work needs to be done also on the practical side. LDA, although modified to the present HLDA version, is probably the simplest approach to statistical classification of data. More sophisticated methods are possible and might lead to further improvement; in addition, the extension of LDA to multiple class cases might prove to be useful in all those cases in which several transitions are possible.

Figure 5. Overlap of an ensemble of configurations extracted from the barrier top region (s(R) ≈ − 2.8) resembling the typical textbook example of a Diels−Alder transition structure. Here, the potentially chiral stereospecific configurations endo and exo are depicted in magenta and cyan, respectively.

and exo. In our case, they are symmetry-related by reflection. However, if some specific asymmetry of the reactants would be present this may lead to the well-known endo−exo chirality. Such a procedure can be applied to more complex cases where the bonding transformation cannot be clearly attributed by means of simple chemical intuition. Moreover, as the discrimination of the states is so well-defined in cases like this the method can be applied to understand more deeply the nature of the transition region that can be a key in the understanding of many chemical properties. We underscore the fact that we do not claim these are optimal CVs. What sets them apart from those derived in other approaches is that we do not need to guess the reaction mechanism beforehand. The mechanism emerges in all its clarity and in remarkable detail at the very beginning of the calculation as soon as the CV is computed (see the Supporting Information). In the Diels−Alder case, the apparent free energy surface transition state coincides with the expected real one. This is not necessarily the case in all circumstances. However, if the CV is able to go reversibly from one minimum to the other, it is to be expected that in the course of the simulation the trajectory passes close to the transition-state ensemble. This was the case also in the more complex case of AgI, and some members of this ensemble could be identified. However, we stress that this is not our main concern here, and to us a good CV is one for which the free energy profile can be converged in a practical time. Even better CVs can be obtained by analyzing the biased trajectories generated with HLDA and applying methods like SGOOB31 or TICA.32 Typically, the first step, namely, obtaining reversible transitions, is the most difficult one. In the cases in which our theory applies, this has been made almost painless. It would be foolhardy to expect that our approach could work well in all circumstances. We have implicitly described the scenario in which HLDA is expected to work: well-defined local minima with short-lived fluctuations and a narrow tube of lowlying conformations that link one minimum with another. These conditions are obeyed by very many rare events. It remains to be seen how far HLDA can be pushed. In this contribution we present a simple way of obtaining efficient collective variables, to be used in collective variablesbased enhanced sampling methods. The only information needed is a set of descriptors that identify the metastable states of the system and the fluctuations of these descriptors while in the basins. We use a modified version of a method often used in statistics to classify data into different classes, namely, Fisher’s classical LDA. In the space of descriptors, the direction orthogonal to the hyper plane that best separates the sets of



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jpclett.8b00733. Computational details (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Dan Mendels: 0000-0002-2408-3133 GiovanniMaria Piccini: 0000-0002-3511-4281 Author Contributions ¶

D.M. and G.P. contributed equally to this work

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This research was supported by the European Union Grant No. ERC-2014-AdG-670227/VARMET. Calculations were carried out on the Mönch cluster at the Swiss National Supercomputing Center (CSCS).



REFERENCES

(1) Peters, B. React. Rate Theory Rare Events Simulations 2017, 227− 271. (2) Lindorff-Larsen, K.; Piana, S.; Dror, R. O.; Shaw, D. E. How fastfolding proteins fold. Science 2011, 334, 517−520. (3) Peters, B. Reaction Coordinates and Mechanistic Hypothesis Tests. Annu. Rev. Phys. Chem. 2016, 67, 669−690. (4) Tiwary, P.; van de Walle, A. In Multiscale Materials Modeling for Nanomechanics; Weinberger, C. R., Tucker, G. J., Eds.; Springer International Publishing: Cham, 2016; pp 195−221. (5) Torrie, G. M.; Valleau, J. P. Nonphysical sampling distributions in Monte Carlo free-energy estimation: Umbrella sampling. J. Comput. Phys. 1977, 23, 187−199. (6) Laio, A.; Parrinello, M. Escaping free-energy minima. Proc. Natl. Acad. Sci. U. S. A. 2002, 99, 12562−12566. (7) Dama, J. F.; Parrinello, M.; Voth, G. A. Well-tempered metadynamics converges asymptotically. Phys. Rev. Lett. 2014, 112, 240602. (8) Barducci, A.; Bussi, G.; Parrinello, M. Well-tempered metadynamics: A smoothly converging and tunable free-energy method. Phys. Rev. Lett. 2008, 100, 020603.

2780

DOI: 10.1021/acs.jpclett.8b00733 J. Phys. Chem. Lett. 2018, 9, 2776−2781

Letter

The Journal of Physical Chemistry Letters (9) Valsson, O.; Parrinello, M. Variational Approach to Enhanced Sampling and Free Energy Calculations. Phys. Rev. Lett. 2014, 113, 090601. (10) Huber, T.; Torda, A. E.; van Gunsteren, W. F. Local elevation: a method for improving the searching properties of molecular dynamics simulation. J. Comput.-Aided Mol. Des. 1994, 8, 695−708. (11) Darve, E.; Pohorille, A. Calculating free energies using average force. J. Chem. Phys. 2001, 115, 9169−9183. (12) Wang, F.; Landau, D. Efficient, multiple-range random walk algorithm to calculate the density of states. Phys. Rev. Lett. 2001, 86, 2050. (13) Hansmann, U. H.; Wille, L. T. Global optimization by energy landscape paving. Phys. Rev. Lett. 2002, 88, 068105. (14) Maragakis, P.; van der Vaart, A.; Karplus, M. Gaussian-mixture umbrella sampling. J. Phys. Chem. B 2009, 113, 4664−4673. (15) Piana, S.; Laio, A. A bias-exchange approach to protein folding. J. Phys. Chem. B 2007, 111, 4553−4559. (16) Valsson, O.; Tiwary, P.; Parrinello, M. Enhancing important fluctuations: Rare events and metadynamics from a conceptual viewpoint. Annu. Rev. Phys. Chem. 2016, 67, 159−184. (17) Tribello, G. A.; Bonomi, M.; Branduardi, D.; Camilloni, C.; Bussi, G. PLUMED 2: New feathers for an old bird. Comput. Phys. Commun. 2014, 185, 604−613. (18) Gervasio, F. L.; Laio, A.; Parrinello, M. Flexible docking in solution using metadynamics. J. Am. Chem. Soc. 2005, 127, 2600− 2607. (19) Bonomi, M.; Parrinello, M. Enhanced sampling in the welltempered ensemble. Phys. Rev. Lett. 2010, 104, 190601. (20) Piaggi, P. M.; Valsson, O.; Parrinello, M. Enhancing entropy and enthalpy fluctuations to drive crystallization in atomistic simulations. Phys. Rev. Lett. 2017, 119, 015701. (21) Palazzesi, F.; Valsson, O.; Parrinello, M. Conformational entropy as a collective variable for proteins. J. Phys. Chem. Lett. 2017, 8, 4752−4756. (22) Branduardi, D.; Gervasio, F. L.; Parrinello, M. From A to B in free energy space. J. Chem. Phys. 2007, 126, 054103. (23) E, W.; Ren, W.; Vanden-Eijnden, E. Finite temperature string method for the study of rare events. J. Phys. Chem. B 2005, 109, 6688− 6693. (24) Vanden-Eijnden, E.; Venturoli, M. Revisiting the finite temperature string method for the calculation of reaction tubes and free energies. J. Chem. Phys. 2009, 130, 194103. (25) Branduardi, D.; Gervasio, F. L.; Parrinello, M. From A to B in free energy space. J. Chem. Phys. 2007, 126, 054103. (26) Piccini, G.; McCarty, J. J.; Valsson, O.; Parrinello, M. Variational Flooding Study of a SN2 Reaction. J. Phys. Chem. Lett. 2017, 8, 580− 583. (27) Sakuraba, S.; Kono, H. Spotting the difference in molecular dynamics simulations of biomolecules. J. Chem. Phys. 2016, 145, 074116. (28) Uyar, A.; Karamyan, V.; Dickson, A. Long-Range Changes in Neurolysin Dynamics Upon Inhibitor Binding. J. Chem. Theory Comput. 2018, 14, 444−452. (29) Mendels, D.; McCarty, J.; Piaggi, P. M.; Parrinello, M. Searching for Entropically Stabilized Phases: The Case of Silver Iodide. J. Phys. Chem. C 2018, 122, 1786−1790. (30) Piccini, G.; Polino, D.; Parrinello, M. Identifying Slow Molecular Motions in Complex Chemical Reactions. J. Phys. Chem. Lett. 2017, 8, 4197−4200. (31) Tiwary, P.; Berne, B. Spectral gap optimization of order parameters for sampling complex molecular systems. Proc. Natl. Acad. Sci. U. S. A. 2016, 113, 2839−2844. (32) McCarty, J.; Parrinello, M. A variational conformational dynamics approach to the selection of collective variables in metadynamics. J. Chem. Phys. 2017, 147, 204109.

2781

DOI: 10.1021/acs.jpclett.8b00733 J. Phys. Chem. Lett. 2018, 9, 2776−2781