Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/233564925

Blue Moon Approach to Rare Events

Article  in  Molecular Simulation · September 2004


DOI: 10.1080/0892702042000270214

CITATIONS READS

24 213

2 authors:

Giovanni Ciccotti Mauro Ferrario


Sapienza University of Rome Università degli Studi di Modena e Reggio Emilia
247 PUBLICATIONS   22,335 CITATIONS    130 PUBLICATIONS   3,594 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Infrared spectroscopy of small protonated water clusters at room temperature View project

Molecular Dynamics in general View project

All content following this page was uploaded by Giovanni Ciccotti on 12 April 2017.

The user has requested enhancement of the downloaded file.


Molecular Simulation

ISSN: 0892-7022 (Print) 1029-0435 (Online) Journal homepage: http://www.tandfonline.com/loi/gmos20

Blue Moon Approach to Rare Events

Giovanni Ciccotti & Mauro Ferrario

To cite this article: Giovanni Ciccotti & Mauro Ferrario (2004) Blue Moon Approach to Rare
Events, Molecular Simulation, 30:11-12, 787-793, DOI: 10.1080/0892702042000270214

To link to this article: http://dx.doi.org/10.1080/0892702042000270214

Published online: 26 Oct 2010.

Submit your article to this journal

Article views: 120

View related articles

Citing articles: 15 View citing articles

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=gmos20

Download by: [Universita Studi la Sapienza] Date: 12 April 2017, At: 08:04
Molecular Simulation, Vol. 30 (11–12), 15 September–15 October 2004, pp. 787–793

Blue Moon Approach to Rare Events


GIOVANNI CICCOTTIa,* and MAURO FERRARIOb,†
a
INFM and Dipartimento di Fisica dell’Università di Roma “La Sapienza”, P.le A. Moro 2, 00185 Roma, Italy; bS3-INFM and Dipartimento di Fisica
dell’Università di Modena e Reggio Emilia, Via G. Campi 213/A, 41100 Modena, Italy

(Received September 2003; In final form January 2004)

The “Blue Moon” ensemble is a computationally when a property j ¼ j ðr N Þ exists which takes the
efficient molecular dynamics method to estimate the corresponding value j ‡ and otherwise fluctuates
rate constants of rare activated events when the process between the values jA and jB corresponding to the
can be described by a reaction coordinate j (r), a well-
defined function in configuration space. By means of reactant and product states, respectively. Escape from
holonomic constraints a number of values of j (r) can be the region around j ¼ j ‡ is fast. See Fig. 1 for a
prescribed along the relevant path to identify the pictorial interpretation.
“bottleneck” region first and to sample an ensemble of In order to give a Statistical Mechanics description
starting conditions to generate activated trajectories. of rare events we start from the Boltzmann’s
These MD trajectories sample phase space according to a
biased configurational distribution. With a suitable definition of probability for a state identified by a
re-weighting of averages from such ensemble of particular value j ¼ j 0 ; introducing the reversible
trajectories one can characterize completely rare events. work Wj (j0 ) (free energy) associated to the “progress
variable” j (r N):
Keywords: Blue moon approach; Rare events; Reactive flux
correlation; MD trajectories P j ðj 0 Þ ¼ kd ðj ðr N Þ 2 j 0 Þl ; C exp{ 2 bW j ðj 0 Þ} ð1Þ
where k. . .l is the canonical ensemble average with
b ¼ ðkB TÞ21 : The constant C is determined by the
normalization condition on P(j ). A typical behavior
INTRODUCTION of Wj (j 0 ) is sketched in Fig. 2.
The quantities of physical interest in rare events
Rare events are activated processes in which a (chemical reactions) are the rate constants, for
transition takes place between two stable states, i.e. example, the one related to the population of the
states where the system spends the overwhelming reactant which goes to the product state. In terms of
majority of its time. Stable states are separated by a the progress variable (reaction coordinate), the
region of very low probability so that the transition population of the reactant is given by the space
can be viewed as the infrequent crossing of a free function u ðj ðr N Þ 2 j ‡ Þ with its associated flux u_ ¼
_ ðj ðr N Þ 2 j ‡ Þ: Using the standard linear response
jd
energy barrier much higher than kBT. As such, the
transition is of very short duration compared with theory approach for “transport” properties [1 – 3], for
the time one has to wait for it to occur and cannot be a review see also Refs. [4,5], one finds the analogous
simulated using conventional techniques. of a Green-Kubo formula for the rate constant kf of
For a classical system, described by an Hamiltonian the activated processes
Hðr N ; p N Þ ¼ K þ V whose time evolution can be _ ðj 2 j ‡ Þu ðj ðtÞ 2 j ‡ Þl
kjd
viewed as a curve in the 6N-dimensional phase space kf ðtÞ ¼ ; ð2Þ
ku ðj ‡ 2 jÞl
identified by the initial conditions and generated
according to the equations of motion r_ ¼ ð›H=›pÞ; Given that j ¼ j ‡ is a rare event and happens very
p_ ¼ 2ð›H=›rÞ; the transition state can be identified infrequently it is difficult, if not impossible, to collect

*Corresponding author. E-mail: giovanni.ciccotti@roma1.infn.it



E-mail: ferrario.mauro@unimo.it

ISSN 0892-7022 print/ISSN 1029-0435 online q 2004 Taylor & Francis Ltd
DOI: 10.1080/0892702042000270214
788 G. CICCOTTI AND M. FERRARIO

Molecular Dynamics (MD) simulations [12], in


place of stiff restoring forces, to model strong
chemical bonds which are not to be broken during
the simulation, only because of the availability of a
suitable algorithm. Thus it became possible to
integrate efficiently the dynamics and simulate
molecular systems composed of large molecules.
However, the use of holonomic constraints is not
restricted to the modeling of molecular systems, but
naturally arises in other situations such as, for
example, the Car-Parrinello ab-initio [13 – 15], where
one has to satisfy orthonormality conditions for the
FIGURE 1 A typical dynamical evolution of the reaction electronic orbitals contributing to the electron density
coordinate j (t), characterized by transition of very short duration of the system, or, for the case of interest here, when
between values close to jA or j B, that identify the two stable states
around which the systems spend the majority of its time as holonomic constraints can be shown to be a useful
expressed by the two maxima of P(j), the probability distribution tool to frequently produce rare events in molecular
along j, pictured on the left. dynamics simulations. In particular, we refer to the
approach known as the Blue Moon ensemble [16,17],
good statistics for the rate constants by “brute force” which is not restricted to the computation of static
molecular dynamics simulation. Special methods equilibrium properties (free energies) related to
have to be used to determine the probability of a activated processes but has been generalized to
highly improbable value of j. They include special include the full calculation of dynamic properties
sampling schemes, for example umbrella sampling such as the rate constants.
[6], to compute free energy differences Wðj 0 Þ 2 Wðj Þ: In order to solve the statistical problem we use the
The rate constant in turn can be computed, knowing following steps: (i) rewriting the expression for the
this probability, as the product of the P(j ‡) and the rate constants in terms of conditional averages and
average fraction of trajectories which, starts at the (ii) establishing a theorem that states that
transition point j ‡ and end in the product region. (infrequent) conditional averages can be estimated
The classical treatment of chemical reactions we have by corresponding “constrained” averages with good
just described was first introduced by Keck [7,8] and statistics at relatively low computational prices.
Anderson [9] to treat gas phase chemical reactions
and later applied by Bennett [10,11] and Chandler [2]
for the treatment of condensed-phase rate processes. REACTIVE FLUX CORRELATION FORMULAS
In this paper, we discuss a particular sampling
scheme which enhances the statistical occurrence of We begin this section by recalling the autocorrelation
highly improbable values of the reaction coordinate function expressions for the rate constant of a
(i.e. a function of the whole set of particle coordinates) reaction. These expressions were first derived by
by fixing it to any desired value j 0 through the Yamamoto [1] in 1960 and have continued to
holonomic constraints j ðr N Þ ¼ j 0 : Historically, holo- fascinate researchers in the area as evidenced by
nomic constraints could be first introduced in the number of times they have been rederived and
reinterpreted in the literature [2,3,18 – 21].
In order to give a statistical mechanical definition
of the rate of the reaction
A O B; ð3Þ
we must first introduce a microscopic definition of
the chemical species A and B. This is achieved by
using the progress variable j and defining the states
corresponding to species A such that j , j ‡ while the
complementary range defines species B. The divi-
ding surface between the two species, j ¼ j ‡ ;
identifies the rare value of j which must be crossed
when transforming A to B (or vice versa). The
microscopic variable
(
1 j , j‡
‡ N
n^ A ðj Þ ¼ u ðj 2 j ðr ÞÞ ¼ ð4Þ
FIGURE 2 The potential of mean force W(j ) corresponding 0 j . j‡
system whose behaviour was sketched in Fig. 1.
BLUE MOON APPROACH 789

characterizes microscopically the species A with


(nA ¼ kn^ A l; as usual) while ð
1
n_^ A ðj Þ ¼ 2jd
_ ðj 2 jÞ

ð5Þ P j ðj 0 Þ ¼ dr N dp N exp{ 2 bH}d ðj ðr N Þ 2 j 0 Þ: ð10Þ
Q
is the microscopic expression for its flux.
Indeed, if we try to obtain Pj (j 0 ) via thermodynamic
If a phenomenological law is valid, one can write
integration, immediately we see that the reversible
down a rate equation for the reaction and identify the
work is given by
forward and reverse rate coefficient, kf and kr ,
respectively. ðj
dWðj 0 Þ
By applying the fluctuation –dissipation theorem Wðj Þ ¼ dj 0
dj 0
one finds that the rate coefficient, say kf , can be
ðj
computed in microscopic terms from the plateau k 2 ››Hj d ðj ðr N Þ 2 j 0 Þl
value of the quantity kf (t) defined by Refs. [1,2] ¼ dj 0 ð11Þ
kd ðj ðr N Þ 2 j 0 Þl
ð
1 t _ i.e. W(j ) is the potential of mean force associated with
kf ðtÞ ¼ kn^ A ðj Þn_^ A ðj ðt0 ÞÞl dt0
nA 0 the conditional average of the generalized force Fj ¼
2ð›HÞ=ð›j Þ: Note that to derive this simple result we
1 _
;2 kjd ðj 2 j ‡ Þu ðj ‡ 2 j ðtÞÞl have to go through a change of variables in the
nA statistical mechanical derivation: from {r N ; p N } to
_ ðj 2 j ‡ Þu ðj ðtÞ 2 j ‡ Þl
kjd {u; p u } ¼ {j; q; p j ; p q }: Explicitly, in terms of the
¼ ; ð6Þ Jacobian matrix of the transformation J ia ¼ ð›ri =›ua Þ
ku ðj ‡ 2 jÞl
and the metric matrix M
where the last equality follows by observing that
_ ðj 2 j ‡ Þ
u ðj ‡ 2 j ðtÞÞ ¼ 1 2 u ðj ðtÞ 2 j ‡ Þ and that , jd X ›r i ›r i
M ab ¼ mi ¼ ðJ T mJÞab ; mij ¼ dij mi ð12Þ
.¼ 0: The limit t ! 0þ of Eq. (6) plays a special i ›ua ›ub
role in the theory since it gives the transition state
theory value for the rate constant [22] for a system of N particles interacting through the
1 _ potential V N ðr 1 ; . . .; r N Þ the effective mean force can be
kTST
f ¼2 kju ð2j_Þd ðj 2 j ‡ Þl computed in terms of the matrix determinants as
nA
ð
_ ðj_Þd ðj 2 j ‡ Þl dW 1
kju ¼ dr N d ðj 2 j 0 Þexpð2bV N Þ
¼ ; ð7Þ dj 0 kd ðj 2 j 0 Þl
ku ðj ‡ 2 jÞl
 
› 1
where the last equality follows by time reversal £ V N 2 kB T lnjMj
invariance. kTST is the reference value for most ›j 2
f
elementary treatments of the rate constant. We can ð
1
separate the static and dynamic contributions to the ¼ dr N d ðj 2 j 0 Þ
kd ðj 2 j 0 Þl
rate constant by multiplying and dividing the right
X ›r i ›
hand side of Eq. (6) by kd ðj 2 j ‡ Þl: We then obtain £ expð2bV N Þ · ½V N þ kB T lnjJ 21 j ð13Þ
 _  ›j ›r
i i
1 ‡ kjd ðj 2 j ‡ Þu ðj ðtÞ 2 j ‡ Þl
kf ðtÞ ¼ eq kd ðj 2 j Þl
nA kd ðj 2 j ‡ Þl where one needs to compute explicitly both
1 the Jacobian of the transformation and the vectors
¼ £ C exp{ 2 bW j ðj ‡ Þ} ð›r i =›j Þ in order to perform the differentiation with
ku ðj ‡ 2 jÞl
respect to j.
_ ðj 2 j ‡ Þu ðj ðtÞ 2 j ‡ Þl
kjd 1
£ ¼Ð 0 kd ðj 2 j 0 Þl
kd ðj 2 j ‡ Þl j, j ‡ d j
( ð ‡ ) CONSTRAINTS AND BLUE MOON ENSEMBLE
j 0 0
0 k›H=›j Þd ðj 2 j Þl
£ exp 2b dj
reference kd ðj 2 j 0 Þl The question now is: how can we obtain good
_ ðj 2 j Þu ðj ðtÞ 2 j Þl
‡ ‡ statistics from a MD simulation for a conditional
kjd
£ ð8Þ average when the conditioning value we are
kd ðj 2 j ‡ Þl interested is very infrequent? A very natural answer
In this way we have explicitly introduced the is to use the infrequent value of the reaction
reversible work needed to bring the system from coordinate as a holonomic constraint, j ðr N Þ ¼ j † ; in
some reference state to j 0 : the simulation and to unbias the sampling along the
constrained trajectory by using the relations between
Wðj 0 Þ ¼ 2kB T lnP j ðj 0 Þ; ð9Þ constrained and unconstrained ensembles [16,23].
790 G. CICCOTTI AND M. FERRARIO

In Molecular Dynamics, a configurational conditional straightforward calculation [24]) after some Gaus-
average sian integrations, the effective force can be rewritten
as
^ N Þd ðj ðr N Þ 2 j 0 Þl
kOðr
Ocond ¼ ð14Þ D 1 E
kd ðj ðr N Þ 2 j 0 Þl Z 22 2l 2 p_ j þ kB2T › ln Z
dW ›j j0
can be computed by using an alternative trajectory, ¼ ð18Þ
dj 0 1
kZ 22 l
with j (r N) constrained to be equal to j 0 and j0

performing the unbiased average


where the first term is a quantity known along a
212 ^ N Þlj 0 X 1 ›j 2 simulation, and easy to compute, while the rest has
kZ Oðr
Ocond ¼ ; Z¼ ð15Þ still to be simplified. Observing that by time reversal
212
kZ lj 0 i m i ›r i symmetry the average value of a total derivative is
zero, one can rewrite
where it appears a reasonably-simple-to-compute     
unbiasing factor, originated by the metric factor 1 d 21 j d 21 j
kZ 22 p_ j lj 0 ¼ ½Z 2 p  2 Z 2 p
related to the momenta integration on the hypersur- dt j0 dt j0
face determined by the constraints in phase space.    
The effective mean force in Eq. (13) can now be 1 1 1 ›Z
¼ Z 22 _ j
·up
expressed after some algebra [16,23] by the con- 2 Z ›u j0
strained average  
212 212 kB T › ln Z
D 1 E ¼ 2kZ kB TGlj 0 þ Z ;
2 ›j j 0 ð19Þ
Z 22 ››Hj
dW j0
¼
dj 0 1
kZ 22 l j0 with the observable G a completely explicit quantity
defined by
  
1 1
Z 22 ››j V N 2 kB T ln ðjJjZ Þ
2
j0 1 X N
1 ›j › 2 j ›j
¼ ð16Þ G¼ · · : ð20Þ
1
kZ 22 lj 0 Z 2 i;j¼1 mi mj ›r i ›r i ›r j ›r j

Again to use this results one has to make explicit the Recombining all terms together we finally find
coordinate transformation with its Jacobian determi-
nant and the vectors needed to differentiate with k 2 ››Hj d ðj 2 j 0 Þl 1
kZ 22 ½l þ kB TGlj 0
Fj0 ¼ ¼ ; ð21Þ
respect to j in the configuration space. kd ðj 2 j 0 Þl 1
kZ 22 lj 0
Instead of this “brute force” approach, an
alternative, simpler way, has been proposed [24] on the sought for expression containing no implicit
the basis of the equations of motion written for the quantities and using the constraint forces which are
generalized coordinates {u; p u } where one has, in automatically provided by SHAKE [25].
turn, to consider explicitly the Lagrangian multiplier Three comments are in order. First, the Lagran-
l associated to the constraint on j gian multiplier l appearing in the above equations
›H is exactly the same in the cartesian equations of
u_ ¼ ; ð17Þ motion. Second, in a previous work, Frenkel and
›p u
Ruiz [26] have found on a sound physical basis,
›H although not completely general mathematically, a
p_ u ¼ 2 2 ldju : closely related result which also permits one to
›u
avoid the definition of the other coordinates q and
As can be seen from Eq. (13), the difficulty was to compute explicitly both jJj and ð›=›j Þ: Finally
brought about when performing integration over the notice that the simplification occurs when using
momenta. We could instead try to keep the the constraint forces. The expression using the
momentum-dependent observable ð›HÞ=ð›j Þ and generalized force ð›HÞ=ð›j Þ in the constrained
explicitly compute the difference between the ensemble is still implicit and complex. The idea
configurationally unbiased constrained average of to use the constraint force to simplify the
our quantity and the corresponding conditional calculation of the reversible work has been put
average. Remembering that the transformations forward by Mulders et al. [27]. However, these
{r N ; p N } $ {u; p u } is canonical and that the value of authors have forgotten that constrained averages
the Lagrangian multiplier appearing in our have to be unbiased since the reversible work is
equations is independent of the chosen coordinates not directly related to constrained averages but
(this can be demonstrated by a tiring but otherwise only to conditional ones.
BLUE MOON APPROACH 791

TIME CORRELATION FUNCTIONS BY BLUE


MOON ENSEMBLE

As already recalled in section “Reactive Flux


Correlation Formulas”, the two quantities relevant
to the understanding of activated processes are the FIGURE 3 Schematic representation of Blue Moon sampling. The
bold line represents the constrained ðj ðrÞ ¼ j 0 Þ dynamical
reversible work (a time-independent quantity) and evolution in phase space. Open circles represent common points
time correlation functions of the form, in configuration space which are the initial conditions of the
activated trajectory sampling. Note that these points are not real
crossings in phase space since the two trajectories differ in the
kOðr N ð0Þ;p N ð0ÞÞOðr N ðtÞ;p N ðtÞÞd ðj ðr N ð0ÞÞ 2 j 0 Þl momentum space. The dynamics represented by the light lines in
Cj 0 ðtÞ ¼ ;
kd ðj ðr N ð0ÞÞ 2 j 0 Þl the vicinity of the crossing points gives the dynamical information
needed in Eq. (28).
ð22Þ

(the transmission coefficient in chemical reactions is the argument a bit more, let us now formulate in
typically of this form). When j ðr N Þ ¼ j 0 is a rare event statistical terms the choice of the sample of initial
these quantities are very difficult to obtain for exactly conditions. We have to sample from the distribution
the same reasons that occur in the calculation of the 1
reversible work. Furthermore, we now need to Pðr N ; p N Þd ðj ðr N Þ 2 j 0 Þ ¼ exp { 2 bHðr N ; p N Þ}
Q
integrate the real dynamics of the system and not the ð25Þ
constrained one. These difficulties can be overcome d ðj ðr N Þ 2 j 0 Þ ¼ Pr ðr N ÞPp ðp N jr N Þd ðj ðr N Þ 2 j 0 Þ;
by sampling independently initial conditions from
the appropriate ensemble, instead of using points where Pr ðr N Þ / exp{ 2 bV N ðr N Þ} and Pp ðp N jr N Þ does
along a dynamical trajectory, and then integrating not depend on r N and is a simple product of
the correct dynamics for the (short) times needed to Maxwellians. Since j 0 is infrequent we are sampling
be followed to construct the relevant correlation 1
functions. This is what we call the Gibbs definition of Pj 0 ðr N Þ / Z 2 exp { 2 bV N ðr N Þ}d ðj ðr N Þ 2 j 0 Þ: ð26Þ
time correlation functions:
We disregard Pj 0 ðp N jr N Þ and we sample p N from
N N N 0 N N
kOð0;r ;p Þd ðj ð0;r Þ 2 j ÞOðt;r ;p Þl 1
Pp(p N). Multiplying the probability used by Z 22 ; we
ð obtain as probability distribution
1
¼ dr N dp N exp{ 2 bHðr N ;p N Þ}
Q 1
Z 22 Pj 0 ðr N ÞPp ðp N Þ / Pðr N ; p N Þd ðj ðr N Þ 2 j 0 Þ; ð27Þ
N N N N 0
£ Oðr ;p Þd ðj ðr ;p Þ 2 j Þ which is proportional to the correct one. This is what
N N we call the Blue Moon ensemble, from which
£ ½Uðt;0ÞOðr ;p Þ; ð23Þ
kOð0; r N ; p N ÞOðt; r N ; p N Þd ðj ð0; r N Þ 2 j 0 Þl
where Uðt;0Þ is the evolution operator of the system
from time 0 to time t. This should be contrasted with kd ðj ð0; r N Þ 2 j 0 Þl
the Boltzmann definition 1
kZ 22 Oð0; r N ; p N ÞOðt; r N ; p N ÞlBlue Moon
¼ 1 : ð28Þ
N N
kOð0;r ;p ÞOðt;r ;p Þl N N kZ 22 lBlue Moon

ðT Equation (28) extends the theorem in Eq. (15) to


1 dynamical properties. Practically, what is required is
¼ lim dt Oðt;r N ;p N ÞOðt þ t;r N ;p N Þ: ð24Þ
T !1T 0 to take an average over a Blue Moon ensemble of
initial conditions of trajectories propagated by
What we have to do then, is to sample correctly following the free time evolution of the system
initial positions and momenta corresponding to the after release of the constraint j (see Fig. 3 and Ref. [28]
infrequent value of the reaction coordinate. This can for a specific application in the case of the ion
be done extracting a sample of space configurations, dissociation reaction).
1
corrected for the unbias factor Z 22 ; from a trajectory
constrained to satisfy the “rare” value j 0 and
sampling the momenta from the unbiased velocity A MORE GENERAL CASE AND CONCLUSIONS
distribution and, then, integrating the full (uncon-
strained) dynamics starting from these initial In the most interesting cases, the “true” reaction
conditions (see Fig. 3). Looking at Eq. (24) we see coordinate is far from being a well known expression
that, since we use the full unconstrained evolution in terms of the cartesian coordinates of the system. A
operator Uðt; 0Þ; if the initial conditions are pro- more exploratory attitude is really needed to
perly chosen, our procedure is correct. To formalize investigate the behavior of the reversible work
792 G. CICCOTTI AND M. FERRARIO

expressed in terms of averages over a molecular


dynamics constrained trajectory
1
kjJj22 Oðr
^ N ; p N Þlh 0 ;...;h 0
Ocond ¼ 21
1 L
ð30Þ
kjJj 2 lh 01 ;...;h 0L

in which the role of the factor Z is now taken by the


determinant of the L £ L matrix J
X 1 ›ha ›hb
Jab ¼ ·
i m i ›r i ›r i

An example is given in the paper by Maragliano et al.


[31] with explicit formulas carried out for the case of
the dissociation of protein dimers. Armed with such
general vectorial formulation of Blue Moon, one can
fully explore the reduced dimensionality space of
FIGURE 4 A multidimensional reaction coordinate space with
interest to build up amazing free energy landscapes,
two minima identified by jA and jB and separated by a region of but at the expense of a very cumbersome procedure.
high free energy. The minimum free energy path defining the Indeed, there are accelerated methods that are
reaction coordinate j is shown together with its orthogonal planes
at specific points.
already available and in use [32,33], based on the
idea of Landau’s filling potential [34,35], which
permit to tackle the problem in exactly the same
spirit. An even more promising direction, however,
function. One possibility is to set up an “order appears to be the one recently being explored of new,
parameter” space of small dimension L and construct smarter and numerically more efficient approaches
in it a free energy (hyper)surface to locate both [36 –40], where constraints can play the role of very
the transition state and the path defining the reaction useful implementation tools, that attempt to find out
coordinate as a function of the L variables the “reaction path” without any previous knowledge
ðh1 ; h2 ; . . .; hL Þ: Keeping notation short, let us use of the reaction coordinate, whose explicit functional
L ¼ 2 and name the two order parameters h1(r N) and dependence from the coordinates would be no longer
h2(r N ). The reversible work Wðh1 ; h2 Þ can be needed a priori.
expressed as a conditional average

Wðh 01 ; h 02 Þ ¼ 2kB T ln Pðh 01 ; h 02 Þ Acknowledgements

¼ 2kB T ln kd ðh1 2 h 01 Þd ðh2 2 h 02 Þl ð29Þ We are grateful to R. Kapral, D. Frenkel and M. Sprik,
for stimulating collaboration on this subject over
In Fig. 4, a graphical solution is given for an the years.
imaginary case in which the use of a single coordinate,
say h1 is not sufficient to describe the rare events.
Here, one needs to compute the values of the References
reversible work function W on enough points in the [1] Yamamoto, T. (1960) “Quantum statistical mechanical theory
ðh1 ; h2 Þ plane to compute the contour lines and locate of the rate of exchange chemical reactions in the gas phase”,
the transition state. The reaction path in the drawing J. Chem. Phys. 33, 281.
[2] Chandler, D. (1978) “Statistical mechanics of isomerization
is nothing but the graphical rendering of the “reaction dynamics in liquids and the transition state approximation”,
coordinate” j ¼ j ðh1 ; h2 Þ ¼ j ðr N Þ: In order to tackle J. Chem. Phys. 68, 2959.
such a case, the Blue Moon approach we sketched in [3] Kapral, R. (1981) “Kinetic theory of chemical reactions in
liquids”, Adv. Chem. Phys. 48, 71.
the previous sections needs to be generalized to a [4] Chandler, D. (1998) “A story of rare events: from barriers
multidimensional case. The full derivation is given by to electrons to unknown pathways”, In: Berne, B.J.,
Sergi et al. [29] in the simple case when there are no Ciccotti, G. and Coker, D.F., eds, Classical and Quantum
Dynamics in Condensed Phase Simulations (World Scientific,
molecular constraints involving the same cartesian Singapore), p 3.
coordinates appearing in the multidimensional [5] Kapral, R., Consta, S. and McWhirter, L. (1998) “Chemical rate
coordinate h1 ðr N Þ; . . .; hL ðr N Þ; that, therefore, can be laws and rate constants”, In: Berne, B.J., Ciccotti, G. and
Coker, D.F., eds, Classical and Quantum Dynamics In Condensed
independently constrained, and, more recently, by Phase Simulations (World Scientific, Singapore), p 583.
Coluzza et al. [30] in the most general case. The [6] Torrie, G.M. and Valleau, J.P. (1977) “Nonphysical sampling
derivations are essentially the same, although much distributions in Monte Carlo free energy estimation: umbrella
sampling”, J. Comput. Phys. 23, 187.
more cumbersome, as the one outlined here with the [7] Keck, J. (1962) “Statistical investigation of dissociation cross-
key point that the conditional averages are now sections for diatoms”, Disc. Far. Soc. 33, 173.
BLUE MOON APPROACH 793

[8] Keck, J. (1967) “Variational theory of reaction rates”, Adv. [23] Ciccotti, G. and Ferrario, M. (2000) “Rare events by
Chem. Phys. 13, 85. constrained molecular dynamics”, J. Mol. Liq. 89, 1.
[9] Anderson, J.B. (1973) “Statistical theories of chemical [24] Sprik, M. and Ciccotti, G. (1998) “Free energy from
reactions, distributions in the transition region”, J. Chem. constrained molecular dynamics”, J. Chem. Phys. 109, 7737.
Phys. 58, 4684. [25] Ciccotti, G. and Ryckaert, J.P. (1986) “Md simulation of rigid
[10] Bennett, C.H. (1975) “Exact defect calculations in model molecules”, Comput. Phys. Rep. 4, 345.
substances”, In: Nowick, A. and Burton, J., eds, Diffusion in [26] Ruiz-Montero, M.J., Frenkel, D. and Brey, J.J. (1997) “Efficient
Solids: Recent Developments (Academic Press), p 73. schemes to compute diffusive barrier crossing rates,
[11] Bennett, C.H. (1977) “Molecular dynamics and transition appendix d”, Mol. Phys. 90, 925.
state theory: the simulation of infrequent events”, In: [27] Mülders, T. and Krüger, P. (1996) “Free energy as the potential
Christofferson, R.E., ed., Algorithms for Chemical Computation of mean constraint force”, J. Chem. Phys. 104, 4869.
ACS Symposium Series No. 46, (American Chemical Society, [28] Ciccotti, G., Ferrario, M., Hynes, J.T. and Kapral, R. (1990)
Washington, DC), p 63. “Dynamics of ion pair interconversion in a polar solvent”,
[12] Ryckaert, J.P., Ciccotti, G. and Berendsen, H.J.C. (1977) J. Chem. Phys. 93, 7137.
“Numerical integration of the cartesian equations of motion [29] Sergi, A., Ciccotti, G., Falconi, M., Desideri, A. and Ferrario,
of a system with constraints: molecular dynamics of M. (2002) “Effective binding force calculation in a dimeric
n-alkanes”, J. Comput. Phys. 23, 327. protein by molecular dynamics simulation”, J. Chem. Phys.
[13] Car, R. and Parrinello, M. (1985) “Unified approach for 116, 6329.
molecular dynamics and density-functional theory”, Phys. [30] Coluzza, I., Sprik, M. and Ciccotti, G. (2003) “Constrained
Rev. Lett. 55, 2471. reaction coordinate dynamics for systems with constraints”,
[14] Car, R. and Parrinello, M. (1988) “The unified approach Mol. Phys. 101, 2885.
for molecular dynamics and density functional theory”, [31] Maragliano, L., Ferrario, M. and Ciccotti, G. (2004) “Effective
In: Polian, A., Loubeyre, P. and Boccara, N., eds, Simple binding force calculation in dimeric proteins”, Mol. Simul.
Molecular System at Very High Density (Plenum Press, 30 (11–12), 809– 818.
New York), p 455. [32] Laio, A. and Parrinello, M. (2002) “Escaping free energy
[15] Galli, G. and Pasquarello, A. (1993) “First-principle molecular minima”, Proc. Natl Acad. Sci. USA 99, 12562.
dynamics”, In: Allen, M.P. and Tildesley, D.J., eds, Computer [33] lannuzzi, M., Laio, A. and Parrinello, M. (2003) “Efficient
Simulation in Chemical Physics NATO ASI Series No. 397, exploration of reactive potential energy surfaces using
(Kluwer, Dordrecht), p 261. car-parrinello molecular dynamics”, Phys. Rev. Lett. 90,
[16] Carter, E.A., Ciccotti, G., Hynes, J.T. and Kapral, R. (1989) 238302.
“A new ensemble for molecular dynamics simulation of rare [34] Wang, F. and Landau, D.P. (2001) “Efficient, multiple-range
events”, Chem. Phys. Lett. 156, 472. random walk algorithm to calculate the density of states”,
[17] Ciccotti, G., Ferrario, M., Laria, D. and Kapral, R. (1995) Phys. Rev. Lett. 86, 2050.
“Simulation of classical and quantum activated processes in [35] Wang, F. and Landau, D.P. (2001) “Determining the density of
the condensed phase”, In: Reatto, L. and Manghi, F., eds, states for classical statistical models: a random walk
Progress in Computational Physics of Matter: Methods, Software algorithm to produce a flat histogram”, Phys. Rev. E 64,
and Applications (World Scientific, Singapore), p 150. 056101.
[18] Voth, G.A., Chandler, D. and Miller, W.H. (1989) “Time [36] Olender, R. and Elber, R. (1996) “Calculation of classical
correlation function and path integral analysis of quantum trajectories with a very large time step: formalism and
rate constants”, J. Phys. Chem. 93, 7009. numerical examples”, J. Chem. Phys. 105, 9299.
[19] Miller, S.S.W.H. and Tromp, J. (1983) “Quantum mechanical [37] Henkelman, G. and Jonsson, H. (1999) “A dimer method for
rate constants for bimolecular reactions”, J. Chem. Phys. 79, finding saddle points on high dimensional potential surfaces
4889. using only first derivatives”, J. Chem. Phys. 111, 7010.
[20] Hynes, J.T. (1985) “The theory of reactions in solutions”, [38] Johannesson, G.H. and Jonsson, H. (2001) “Optimization of
In: Baer, M., eds, The Theory of Chemical Reaction Dynamics hyperplanar transition states”, J. Chem. Phys. 115, 9644.
(CRC, Boca Raton, FL) Vol. IV, p 171. [39] E, W., Ren, W. and Vanden-Eijnden, E. (2002) “String method
[21] Hänggi, P., Talkner, P. and Borkovec, M. (1990) “Reaction-rate for the study of rare events”, Phys. Rev. B 66, 052301.
theory: fifty years after kramers”, Rev. Mod. Phys. 62, 251. [40] E, W. and Vanden-Eijnden, E. (2004) “Conformation
[22] Chandler, D. (1987) Introduction to Modern Statistical Mechanics dynamics and transition pathways in complex systems”,
(Oxford University Press, Oxford). LNCSE (Springer Verlag, Berlin) Vol. 39.

View publication stats

You might also like