You are on page 1of 11

On standardised moments of force distribution in simple liquids

On standardised moments of force distribution in simple liquids


Jonathan Utterson1, a) and Radek Erban1, b)
Mathematical Institute, University of Oxford, Radcliffe Observatory Quarter, Woodstock Road, Oxford, OX2 6GG,
United Kingdom
(Dated: 30 August 2021)
The force distribution of a tagged atom in a Lennard-Jones fluid in the canonical ensemble is studied with a focus on its
dependence on inherent physical parameters: number density (n) and temperature (T ). Utilising structural information
from molecular dynamics simulations of the Lennard-Jones fluid, explicit analytical expressions for the dependence of
standardised force moments on n and T are derived. Leading order behaviour of standardised moments of the force
distribution are obtained in the limiting cases of small density (n → 0) and low temperature (T → 0), while the variations
in the standardised moments are probed for general n and T using molecular dynamics simulations. Clustering effects
are seen in molecular dynamics simulations and their effect on these standardised moments is discussed.

and the second moment F12 would be sufficient to




I. INTRODUCTION
parametrize the force distribution. However, the force distri-
Understanding the moments and measures of a distribution for butions in simple liquids have been reported to deviate from
a fully atomistic molecular dynamics (MD) simulation allow Gaussian distribution12–14 . Thus our analysis will also tell us
us to better fit coarser models that reproduce these1–3 . It is of- how non-Gaussian the force distribution is.
ten the case in model coarse graining that we wish to directly Much work has been done in the area of force distribu-
reconcile the energy landscape of the fully atomistic system tions of many-body systems: with seminal work from Chan-
to a more basic representation that allows us to maintain as drasekhar15 that employed Markov’s theory of random flights
many physical properties of the system of interest, with as lit- to give an expression for the force distribution of a many-body
tle computational cost as possible4 . Though, it is also natural system interacting through a 1/r gravitational potential. More
to match forces between the high and low resolution systems recent work has been done with the help of MD by Gabrielli
in an effort to reproduce the force distribution which will in- et al16 , who derived an expression for the kurtosis of the force
herently give rise to the energy landscape5–10 . distribution for a lattice system of atoms interacting through
Let F = [F1 , F2 , F3 ] denote a force on a tagged atom in a the gravitational potential. Further, using the density func-
liquid. Considering an isotropic system, the equilibrium dis- tional theory, an expression for the probability distribution of
tribution of each force coordinate is the same. We define the force for a system interacting through an arbitrary weakly re-
standardised moment of the force distribution by using its first pulsive potential was derived by Rickayzen et al17,18 .
coordinate as
k In this paper, we study the number density and tempera-
F ture dependence of the force distribution for a many-body sys-
αk =
1 k/2 , (1)
F12 tem interacting through a Lennard-Jones 12-6 potential19,20 ,
which is ubiquitously used and has been shown to model ho-
where F1k is the k-th moment of the force distribution and αk


mogeneous systems of interacting (Argon) atoms well21–23 .
standardises the k-th moment by scaling it with the k-th power In Section III, an in depth investigation is given to the sim-
of the standard deviation of the force distribution. In a ho- ple two-body system in one spatial dimension, which provides
mogeneous fluid, the force distribution will exhibit symmetry the ideal platform to illustrate the underlying methods while
around the origin and thus all odd standardised moments van- retaining interesting dynamical behaviour. From first princi-
ish, i.e. 0 = α1 = α3 = α5 = . . . . As α2 ≡ 1 by definition (1), ples we derive first-order partial differential equations (PDEs)
the first non-trivial standardised moment is kurtosis, denoted describing the dependence of the standardised moments of the
α4 , which provides a measure of spread that details how tailed force distribution has on parameters. In doing so we further
the force distribution is relative to a normal distribution11 . In derive an analytic expression for the partition function of a
this paper, we study how the force distribution depends on the two-body system that depends solely on the standardised mo-
number density of a homogeneous many-body system, and the ments of the force distribution whereupon the expression is
temperature of the same system in a canonical ensemble. We exact in an asymptotic limit of the infinitely dilute system
will do this by studying
the behaviour of the second moment (n → 0). Similarly, an expression is derived relating the av-
of the force distribution F12 and standardised even moments erage energy of the system to standardised moments of force
α4 , α6 , α8 , . . . . If the force distribution was Gaussian, then from the temperature dependent PDE. In parameter regimes
the even standardised moments would be where long-range forces between atoms dominate, we use a
k/2 truncated Taylor series expansion to derive the leading order
αk = (k − 1)!! = ∏ (2i − 1), for k = 2, 4, 6, 8, 10, . . . , (2) behaviour of the kurtosis of the force distribution in the limit
i=1 n → 0. Finally, we utilise a Laplace integral approximation
to ascertain the leading order behaviour of the standardised
moments of force at low temperatures (T → 0). Results from
a) Electronic mail: utterson@maths.ox.ac.uk simple MD simulation are presented to provide evidence for
b) Electronic mail: erban@maths.ox.ac.uk the efficacy of these methods and underlying assumptions.
On standardised moments of force distribution in simple liquids 2

This is followed by Section IV, where the natural idea that the classical Hamiltonian H(q, p) = K(p) +U(q) with kinetic
long range force calculations dictate asymptotic behaviour is energy K(p) = |p|2 /2 (where the usual factor of mass is unity
extended from the 1D model to many-body systems of arbi- under reduced units) and a general potential U(q). The statis-
trary size in three spatial dimensions. These systems exhibit tical average of a quantity X for this N-body system is given
the physical properties of standard MD simulations: i.e cu- by
bic geometry with periodic boundary conditions that employ
1
ZZ
the minimum image convention. In this way, we perform cal- hXi = X exp[−β H(q, p)] d3 q d3 p , (6)
culations on a principal computational cubic cell of a theo- ZN h3N N!
Ωq ×Ωp
retically infinite simple fluid. In Section IV B, MD results are
displayed for many-body systems. We present the dependence where the Boltzmann factor acts as a statistical weighting for
of the standardised force moments on density, n, and temper- a configuration {q, p} ∈ R6N , normalised such that h1i = 1.
ature, T , and discuss the parameters and integrator schemes We label atoms so that the first one is the tagged atom. De-
utilised in producing the results of MD simulations. noting the force on the tagged atom produced from the j-th
atom by F j = [Fj,1 , Fj,2 , Fj,3 ] ∈ R3 , for j = 2, 3, . . . , N, the to-
tal force F = [F1 , F2 , F3 ] on the tagged atom is
II. NOTATION
N
We consider a system of N identical atoms interacting via the F= ∑ F j.
Lennard-Jones 12-6 potential19 . This is a ubiquitous inter- j=2

atomic pairwise potential; here the potential between atoms We define


labelled i, j = 1, 2, . . . , N positioned at qi , q j ∈ R3 is given (in
reduced units24 ) by the expression
!k
Z N
! fk = ∑ Fj,1 exp[−β U(q)] d3 q (7)
Ωq
1 1 j=2
Ui j (ri j ) = 4 12 − 6 , (3)
ri j ri j for k = 0, 1, 2, . . . . Then we have

where ri j = qi − q j is the distance between atoms. The *
N
!k +
fk
Lennard-Jones potential (3) between two atoms has a unique = ∑ Fj,1 = hF1k i.
minima obtained at ri j = r∗ = 21/6 . f0 j=2
We employ the framework of statistical mechanics for this
closed many-body system and describe atom i = 1, 2, . . . , N Then the k-th standardised moment (1) is given by
by phase space coordinates {qi , pi } ∈ R6 , were pi denotes the k/2−1
momentum of the i-th atom. We work in the canonical en- f0 fk
αk = , (8)
semble with temperature T ; the partition function therefore k/2
f2
becomes
1
ZZ where we are interested in cases k = 4, 6, 8, . . . .
ZN (T,V ) = exp[−β H(q, p)] d3 q d3 p , (4) In order to study how the force distribution depends on the
h3N N!
Ωq ×Ωp physical parameters of interest it is useful to identify how
changes in these parameters will manifest themselves in the
where V is the volume of our closed system, and q = system. Indeed, we choose to work in the canonical ensemble
(q1 , q2 , . . . , qN )T and p = (p1 , p2 , . . . , pN )T are vectors con- with a target temperature of T : this is accomplished with the
taining the positions and momenta of all atoms in the system. use of a thermostat which is discussed further in Section IV B
Our integration domain is given by Ωq × Ωp ⊂ R3N × R3N . and Appendix B. It is more illuminating to see that if we have
This denotes the phase space of our system. For systems of a system with a fixed number of free interacting atoms N in a
interest Ωp ≡ R3N . The underlying geometry of the system cubic box of side L; the (reduced) number density is given by
(and principle simulation cell) is a cubic box of size L > 0, n = N/L3 . Therefore the approach we employ in this paper
therefore Ωq ≡ (−L/2, L/2] × · · · × (−L/2, L/2]. The phase to ascertain how values of standardised moments depend on
space volume elements in equation (4) are denoted by number density, will be to keep the number of atoms fixed but
vary the box width L - this will manifest as a change in den-
N N
d3 q = ∏ d3 qi and d3 p = ∏ d3 pi . (5) sity n. Similarly one could keep the volume of the cubic box
i=1 i=1 the same and vary the number of atoms though this is a point
of discussion in Section IV B.
Throughout this work we make use of reduced units24 , utilis- For the remainder of the paper we will study systems of
ing Argon parameters25 . In particular, all instances of T in this varying dimensions and size, using equation (8) as a crucial
work can be translated back to SI units with the transforma- starting point in each calculation. We will naturally proceed
tion T → kB T where kB is the Boltzmann factor. Therefore, by investigating systems of increasing complexity; starting
in the partition function (4), we have β = 1/T and h is the from a cartoon one-dimensional model culminating to a gen-
Planck constant (≈ 2.17 in reduced units). Finally, H(q, p) is eral many-body system of arbitrary size.
On standardised moments of force distribution in simple liquids 3

III. ONE ATOM IN A POTENTIAL WELL B. Far-field integral approximation

We now go on to illustrate three approaches to obtain the de- To further analyze integrals (10), we introduce a cutoff c,
pendence of the force distribution on parameters n and T . It which satisfies that r∗ < c < L/2, where r∗ = 21/6 is a unique
is useful to note that, as we are now working in one spatial maximum of exp[−β U(z)], which can be Taylor expanded as
dimension, density n is proportional to 1/L, i.e. we have β (1 + 4z−6 + 4z−12 − 16/3z−18 + 8z−24 . . . ). Considering suf-
n ∝ 1/L. We will consider a simple system in one spatial ficiently large L, we can choose the cutoff c, so that
dimension consisting of two atoms interacting through the  
Lennard-Jones potential (3) in interval [0, L] with periodic
Zc ZL/2
4
f0 (L) − 2  exp[−β U(r)] dr + β 1 + dr ≤ ε , (13)
boundary conditions. One of the atoms is considered to be r 6
fixed at position q0 = L/2 ∈ [0, L] and the other atom is free 0 c

to move, therefore, we have N = 1 free atom. Its position
where tolerance ε is chosen to be 10−4 in our illustrative com-
is denoted x ∈ [0, L]. Therefore, the inter-atomic distance is
putations. This splitting allows us to numerically calculate the
r = |x − q0 |. Using our simplified one-dimensional set up,
bulk of the integral (10) as a constant independent of L and
F1 = F and Ωq = (0, L), equation (7) reduces to
then use the second term to give an analytic expression for αk
ZL with dependence on L, and ultimately on n.
fk (L) = F k (|x − q0 |) exp[−β U(|x − q0 |)] dx, (9) The range of values of T that are of typical use are chosen in
0
order to maintain the liquid state of Argon during simulation.
These are approximately temperatures in the interval 0.70 <
which is the marginalised expected value of the k-th moment T < 0.73 under ambient conditions26 . Therefore, as volume
of force F(x) = −dU/dx, where we have dropped subscripts is varied we are in a regime where β = o(1), for convenience
in the Lennard-Jones potential (3) and we write it as U(z) = we set β = 1. Though given that the density of our system
4(z−12 − z−6 ). Utilising the symmetry of the potential (and changes between each simulation some systems will be in a
therefore the force) we are left with liquid phase and others in a gaseous phase, this is a point of
ZL/2 discussion in Section IV B.
fk (L) = 2 F k (r) exp[−β U(r)] dr. (10) Splitting the integration domain [0, L/2] of integral (10)
into [0, c] and [c, L/2], we use the exact form of the integrand
0
in [0, c] to obtain a ‘near-field’ contribution. Utilising an ap-
In what follows, we will assume that we are in a regime where proximate form for the integrand given by the truncated Taylor
the box width L satisfies L ≫ r∗ , where r∗ = 21/6 minimizes expansion f (z) in the domain [c, L/2] gives rise to a density
the Lennard-Jones potential U. dependent ‘far-field’ contribution. Combining these we arrive
at the approximate form for f0 (L). Using cutoff c = 2, equa-
tion (13) is satisfied with ε = 10−4 . Therefore, upon numer-
A. Differential equation for standardised moments ically calculating the bulk contribution for the integral with
domain [0, 2], we get
We consider a perturbation of the form L → L + δ L. Using
equation (10) and considering terms to the order O(δ L), we ZL/2  
obtain f0 (L) = 2 exp[−β U(r)] dr = b0 + L + o L−6 (14)
0
fk (L + δ L) = fk (L) + fk′ (L) δ L + o δ L2

with b0 = −0.71832, which depends on our choice of cut-
= fk (L) + F k (L/2) exp[−β U(L/2)] δ L + o δ L2 .

off c = 2. Similarly, we can calculate far-field integral ap-
Using equation (8), we approximate αk (L + δ L) by proximations of integrals (10) for general values of k =
2, 4, 6, 8, 10, 12. The integrand F k (r) exp[−β U(r)] has max-
αk (L) + αk (L) υk (L) exp[−β U(L/2)] δ L + o δ L2 ,

ima when r = r∗ = 21/6 or when kU ′′ (r) = β (U ′ (r))2 . This
where our notation αk (L) highlights the dependence of the forms a cubic in r6 that can be solved. For the values of k
standardised moments of force, αk , on L, and function υk (L) used in this work, this sometimes results in a global maxi-
is given by mum, that always lies at a distance less than r < r∗ from the
origin. Therefore r∗ = 21/6 is the furthest maximum of the
k−2 F k (L/2) k F 2 (L/2) integrand from the origin.
υk (L) = + − . (11)
2 f0 (L) fk (L) 2 f2 (L) Splitting integral (10) into a near-field and far-field contri-
Taking the limit δ L → 0, we obtain the derivative of the k-th bution, using the general cutoff c = 2, we find
standardised moment of force, with respect to L, as  
fk (L) = bk + o L−7k (15)
∂ αk
(L) = υk (L) exp[−β U(L/2)] αk (L), (12) The near-field contributions, bk , generally increase vastly if
∂L
we increase the value of k, for example
where υk (L) are expressed in terms of integrals (10) as given
by equation (11). b0 = −0.71832, b2 = 130.64 and b4 = 2.5727 × 105 , (16)
On standardised moments of force distribution in simple liquids 4

By utilising the far field integral approximation (14), we arrive


250 at f0 (L) ∼ (b0 + L), where b0 = b0 (c) is a constant term that
depends on cutoff parameter c. With this, our leading order
approximation of the k-th standardised moment, αk0 , obeys
200
∂ αk0 k−2
(L) = α 0 (L) .
∂L 2 (b0 + L) k
150
Finally this gives us that
k/2−1
αk0 (L) = Ck (b0 + L)k/2−1 = Ck b0 + n −1 , (19)
100
where n = 1/L is the reduced number density and Ck is a con-
stant. Equation (19) gives the same leading order behaviour
n1−k/2 in the limit n → 0 as equation (17): the same behaviour
50
0.1 0.15 0.2 0.25
is also seen for the Lennard-Jones fluid in Section IV. Though
the method above is more generally applicable to include po-
tentials that monotonically decay as r−a as r → ∞ for a > 0.
We next make the observation that equation (4) in 1D can be
FIG. 1. Plot of α4 as a function of n = 1/L for the illustrative
one-atom system. Results of MD simulations are compared with
written as:
α4 = −10.828 + 15.074 n−1 obtained by using equation (17) with ZL Z∞
β p2
 
b0 , b2 and b4 given by (16) (blue dashed line). MD simulation re- 1
Z1 (T,V ) = exp[−β U(q)] dq exp − dp , (20)
sults utilising Langevin dynamics27 described in equation (B1), with h 2
0 −∞
friction parameter γ = 0.1 are represented by red dots. The MD sim-
ulation length was a total of 1.1 × 108 time steps with the first 107 where the Planck factor of 1/h arises instead of 1/h3 due to
time steps used for initialisation. the fact that we are in one-dimensional physical space. Us-
ing (10), we obtain
r
while the dependence on L decreases more rapidly for larger β
values of k. Therefore, the non-negligible density contribu- f0 (L) = h Z1 (T,V ) . (21)

tions to αk (L) in the low density limit come exclusively from
Considering the low density limit n → 0 (i.e. L → ∞) in equa-
the normalisation f0 (L) given by (14).
tion (12) and using (18) and (21), we obtain
Substituting equations (14) and (15) in equation (8), we ob-
tain an expression for the general k-th standardised moment √ "  −1 #
(k − 2) 2π ∂ αk
of force Z1 (T,V ) ∼ p αk (L) (L) , (22)
h2 β ∂L
k/2−1  k/2−1
b bk L 
as L → ∞. In particular, we can obtain the partition func-
αk (L) = 0 k/2 1 + + o L−6 . (17)
b2 b0 tion (20) in the dilute (low density) limit by using information
about the moments of the force distribution. The accuracy of
Using the values of b0 , b2 and b4 given by (16), we obtain equation (22) is illustrated in Figure 2, where we use k = 4.
the dependence of the kurtosis of the force distribution on the We use MD simulations of a single atom, using a range of
reduced number density n = 1/L in the dilute limit n → 0 as simulation box widths L. We estimate the values of kurtosis of
α4 = −10.828 + 15.074 n−1 + o(n6 ). Figure 1 compares this the force distribution, its derivative with respect of L and use
result with the results obtained by MD simulation of the one the right hand side of equation (22) to estimate the Z1 (T,V ).
atom system. We observe that MD is in good agreement with Considering L ≥ 10, the result is within 5% error when com-
the results obtained by formula (17). pared with the exact result (20), while for larger values of box
width L the error decreases to around 1%, confirming that the
formula (22) is valid in the asymptotic limit L → ∞.
C. Leading order behaviour for differential equation (12)

Since L/2 > r∗ , the force F(L/2) monotonically decreases as D. Temperature dependence of standardised moments
a function of L. When looking at leading order approxima-
tions in the low density limit n → 0 (equivalent to limit L → ∞) One can perform a similar analysis as in Section III A, viewing
to equation (12), we need to analyse υk (L). The second and the moments αk = αk (T ) as a function of temperature T =
third term in equation (11) converge to zero more rapidly than 1/β . To do that, we consider the moment definition (10) as a
the first term as L → ∞, therefore the leading order behaviour function of temperature T , namely, we define
is given by the first term
ZL/2  
U(r)
k−2 fk (T ) = 2 F k (r) exp − dr. (23)
υk (L) ∼ as L → ∞. (18) T
2 f0 (L) 0
On standardised moments of force distribution in simple liquids 5

Since the inter-atomic potential U(r) has a global minimum at


r = r∗ in interval [0, L/2], integrals of the form (10) and (23)
can be approximated by Laplace’s method in the limit β → ∞
10 3 and T → 0, respectively. A general discussion of Laplace’s
method is given in Chapter 6 of the book by Bender and
Orszag28 . We calculate the asymptotic expansion of f0 (T )
by applying Laplace’s method to integral (23) for k = 0. We
approximate the integration limits of integral (23) to lie within
Z1

10 2 the domain r ∈ (r∗ − ε , r∗ + ε ), where ε ≪ 1, and we Taylor


expand U(r) at r = r∗ . Using U ′ (r∗ ) = 0, we have
U(r) ≈ U(r∗ ) + (r − r∗ )2U ′′ (r∗ )/2
+ (r − r∗ )3U (3) (r∗ )/6 + (r − r∗ )4U (4) (r∗ )/24 ,
1
10
10 1 10 2 10 3 where we denote the mth derivative of U as U (m) for m ≥ 3.
Substituting into integral (23), we arrive at the asymptotic ex-
pansion
FIG. 2. Approximation of the partition function Z1 (T,V ) obtained √
using the right hand side of equation (22) with k = 4 and values of π T exp[−U(r∗ )/T ] h i
kurtosis (α4 ) estimated from MD simulation (blue dashed line). The f0 (T ) ∼ p 1 + A1 T + o(T ) , (27)
exact values obtained by (20) are plotted as the red dots. 2U ′′ (r∗ )
as T → 0, where constant A1 is given by28
Considering small perturbations of these functions with re- 5 (U (3) (r∗ ))2 U (4) (r∗ )
spect to T → T + δ T , while fixing the domain length L, and A1 = − .
24 (U ′′ (r∗ ))3 8 (U ′′ (r∗ ))2
collecting terms up to first order in δ T , we obtain
To apply Laplace’s method to integral (23) for k = 2, 4, 6, . . . ,
∂ αk we note that F k (r) = (U ′ (r))k for even values of k. Using the
(T ) = νk (T ) αk (T ) , (24) truncated Taylor expansion around r = r∗ and noting U ′ (r∗ ) =
∂T
0, we have
where  k 
F k (r) ≈ (r −r∗ )k U ′′ (r∗ ) + (r − r∗ )Ck,1 + (r − r∗ )2 Ck,2 ,
f0 (T ) fk′ (T )
  ′   ′
k k f2 (T )
νk (T ) = −1 + − . (25)
2 f0 (T ) fk (T ) 2 f2 (T ) where Ck,1 and Ck,2 are constants, which can be expressed (see
Appendix A) in terms of the derivatives of potential U(r) at
Combining equations (24) and (25) with equation (21) where r = r∗ as
β = 1/T , we obtain
k (U ′′ (r∗ ))k−1U (3) (r∗ )
 2 k Ck,1 = (28)
αk f2
 
∂ ∂ 1 2
log = (k − 2) ln(Z1 ) − . k
∂T fk2 ∂T 2T Ck,2 = (U ′′ (r∗ ))k−2 (29)
24 
Since −∂ /∂ β (ln Z1 ) is equal to the average energy of the sys-  2
tem, hEi, we have × 3(k − 1) U (3) (r∗ ) + 4U ′′ (r∗ )U (4) (r∗ ) .

T T2 ∂
 2 k
α f This gives the asymptotic expansion
hEi = + log k 2 2 , (26) √
2 k−2 ∂T fk π T exp[−U(r∗ )/T ]
fk (T ) ∼ p
2U ′′ (r∗ ) (30)
where the first term on the right hand side of equation (26) h  i
is the average kinetic energy of our one-atom system. Sub- × Bk,1 T k/2 + Bk,2 T k/2+1 + o T k/2+1 .
stituting equation (8) into the second term on the right hand
side, it can be rewritten as T 2 ∂ (ln f0 )/∂ T . Thus, using equa- as T → 0, where constants Bk,1 and Bk,2 are given by
tion (6), we confirm that the second term on the right hand k/2
side of equation (26) is the average potential energy. Bk,1 = U ′′ (r∗ ) (k − 1)!!
and
E. Low temperature limit Ck,2 (k + 1)!! U (3) (r∗ )Ck,1 (k + 3)!!
Bk,2 = −
|U ′′ (r∗ )|k/2+1 6 |U ′′ (r∗ )|k/2+2
Next, we consider the behaviour of the k-th standardised mo- !
ment of force, αk (T ), given by equation (8), in the low tem- (U ′′ (r∗ ))k/2−2 (k + 3)!! (k + 5)U (3) (r∗ ) (4) ∗
+ −U (r ) .
perature limit, T → 0, which is equivalent to the limit β → ∞. 24 3U ′′ (r∗ )
On standardised moments of force distribution in simple liquids 6

Substituting (27)-(30) into (8) gives the expression as T → 0:


10
   
Bk,1 k Bk,2 k B2,2
αk ∼ k/2 1 + T − 1 A1 + − . (31) 9
B 2 Bk,1 2 B2,1
2,1
8
In particular, we have
7
α2 ∼ 1 + O(T 2 ),
(32) 6
α4 ∼ 3.000 + 45.084 T + O(T 2 ),
5
Therefore, Laplace’s method predicts that the kurtosis, α4 (T ),
tends to 3 in the low temperature limit, and to leading order 4
the perturbation away from this value scales linearly in tem-
perature. This limiting behaviour is to be expected as dur- 3
ing the Laplace approximation we use a Gaussian distribution 0 0.05 0.1 0.15
to approximate the Boltzmann factor, and the kurtosis of any
Gaussian distribution is identically equal to 3. We can inter-
pret this approach as approximating the force distribution as FIG. 3. Kurtosis, α4 , as a function of temperature, T , for T ≤ 0.15.
Gaussian and perturbations of the system around small tem- The linear behaviour is estimated as α4 (T ) ∼ 2.9388 + 37.002 T
for T ∈ (0.01, 0.10) (using the MD computed data visualized as red
peratures give rise to non-Gaussian contributions to the stan-
dots). We compare this to the theoretical result predicted by equa-
dardised moments. tion (32) (illustrated by the blue dashed line).
Results from MD simulation are illustrated in Figure 3. Due
to the rapid nature of the growth of F k , one has to restrict to a
smaller range of temperature to exhibit the linear growth pre- A. Dependence of αk on density for N = 2 interacting atoms
dicted. This holds for the range of k = 2, 4, 6, 8, 10, 12 tested
in this work. This can be explained by the fact that the in-
In Section III, we have considered two atoms in the one-
tegrand of equation (23) sometimes admits 2 maxima in the
dimensional spatial domain, where one atom was fixed at
range (0, ∞). The global maximum of the integrand does not
position q0 , i.e. we have effectively studied a single atom
always correspond with the global minimum of the potential
in a one-dimensional potential well. Here, we will consider
U(r) and therefore the global maximum of exp[−β U(r)]. We
N = 2 interacting atoms in the three-dimensional cubic do-
therefore need to probe smaller temperature ranges to more
main [0, L]3 with periodic boundary conditions. We calcu-
accurately match the predicted behaviour in equation (32),
late the k-th standardised moment of force according to equa-
though maintaining the canonical ensemble at such minute
tion (8). To do so, we consider equation (7), where we have
temperature ranges necessitates coupling the simulation ther-
d3 q = d3 q1 d3 q2 , U(q) = U(r12 ), F1 (q) = F1 (r12 ) and we in-
mostat tightly, which may affect dynamics29 .
tegrate over the domain Ω = [0, L]3 × [0, L]3 to get
Z
IV. MANY-BODY SYSTEMS fk = F1k (r12 ) exp[−β U(r12 )] d3 q1 d3 q2 . (34)

In this section we employ the far field approximation ap- It is useful to introduce a change of coordinates ξ ℓ = qℓ1 − qℓ2
proach introduced in Section III B and we will vary the num- and η ℓ = qℓ1 + qℓ2 for ℓ = x, y, z. We note that r12 is only de-
ber density of the system by changing the size L of the inte- pendent on the ξ ℓ variables, therefore one can trivially inte-
gration domain, which will be given as the three-dimensional grate (34) through the η ℓ variables as the integrand has no
cube [0, L]3 . Using notation introduced in Section II, the dis- dependence on these to obtain
tance between atoms labelled i, j = 1, 2, . . . , N positioned at
ZL ZL ZL
qi , q j ∈ R3 is denoted by ri j = qi − q j . Taking
into account L3
fk = F1k (r12 ) exp[−β U(r12 )] dξ x dξ y dξ z ,
the periodic boundary conditions, the distance qi − q j is the 8
−L −L −L
minimum image inter-atomic distance given by
where r12 is the minimum image inter-atomic distance (33).
2 2 2 1/2
 
|qi − q j | = (qxi − qxj ) + (qyi − qyj ) + (qzi − qzj ) , (33) This integral can be written in terms of standard Euclidean
distance r2 = (ξ x )2 + (ξ y )2 + (ξ z )2 as
where the overline denotes ζ = ζ − L [ζ /L] for ζ ∈ R and [.] ZL/2 ZL/2 ZL/2
rounds a real number to the nearest integer. For an interact- fk = 8 L 3
F1k (r) exp[−β U(r)] dξ , (35)
ing N-body system the dimensionality of the integral given by
0 0 0
equation (7) is 3N. We first present an illustrative calculation
with N = 2 interacting atoms in Section IV A and then we where dξ = dξ x dξ y dξ z .
In order to analyse fk further by
study systems with larger values of N in Section IV B. implementing a far field approximation, we need to make sure
On standardised moments of force distribution in simple liquids 7

we are in a regime where the integrand is small - we do this N tsim L0 n0


by introducing a cutoff γ , which will divide the cube [0, L/2]3
2 109 5 0.016
into 8 cuboid subdomains, including
8 107 3 1/64
Ω1 = [0, γ ]3 , Ω2 = [0, γ ]2 × [γ , L/2], 64 106 5 1/64
512 104 10 1/64
Ω3 = [0, γ ] × [γ , L/2]2 , Ω4 = [γ , L/2]3 .
Utilising the symmetry of the problem, we can rewrite inte- TABLE I. The length of MD simulation, tsim , the (smallest) box
gral (35) as width, L0 , used for simulations with N atoms and density n0 for MD
Z Z Z Z ! simulations with varying temperatures.

f k = 8 L3 +3 +3 + F1k (r) exp[−β U(r)] dξ .


Ω1 Ω2 Ω3 Ω4 dependency. In all cases we utilise a time step ∆t = 0.01. In
(36) the case of the simulation with N = 2 atoms, we initialise the
Considering (36) for k = 0, the integral over Ω1 is independent positions of atoms by setting q1 = 0 and q2 = (L/2, L/2, L/2),
of L and provides a bulk contribution to f0 that will depend whereas for the N = 8, 64, 512 atom systems, we choose to
on γ . The remaining three terms have integration domains initialise these on a uniform cubic lattice.
that allow the integrand to be accurately described by a Taylor The MD simulation parameters are summarised in Table I,
expansion giving the leading order contribution in the asymp- where tsim is the total simulation time used for calculating the
totic limit L → ∞ as f0 ∝ L6 , which can be rewritten in terms required statistics, which is preceded by the initial simulation
of the density, n, in the form of length tsim /10 used for equilibrating the system. When
f0 ∝ n−2 as n → 0. (37) investigating the number density dependence, we perform
20 simulations each with a box width of L = L0 × (6/5)i−1 ,
Considering fk for k 6= 0, the integral over Ω1 in equation (36) where i = 1, 2, . . . , 20 labels the simulation number and L0 is
is again independent of L. However in the far field expansion the smallest cubic box width. We simulate the N = 8, 64, 512-
the integrals over Ω2 , Ω3 and Ω4 all decay with L due to the atom systems with L0 = 3, 5, 10, respectively. This enables
force factor. As the integration domain has essentially been direct comparison because we can identify triplets of simu-
transformed into that of inter-atomic distances about the three lated systems corresponding to systems of the same number
coordinates, when we increase the domain length, the inter- densities. The two-atom system however is simulated in a
atomic force necessarily decays to 0. Therefore in the limit sparser regime with L0 = 5. We calculate statistics on the fly
L → ∞ the dominant term arises from integrating over Ω1 , for every time step, for every atom and for each coordinate
and we see that, for k = 2, 4, 6, 8, . . . , - therefore we average the computed results over the number
of time steps (tsim /∆t) and atom coordinates (3N). In partic-
fk ∝ n−1 as n → 0 . (38) ular, the statistics are calculated over 3 N tsim /∆t data points.
This is equal to 6 × 1011 (resp. 1.536 × 109 ) data points in the
This leaves us with the final result that in the low density
simulation with N = 2 (resp. N = 512) atoms.
limit n → 0, combining equation (8) with asymptotic expres-
Since we use periodic boundary conditions, we are effec-
sions (37) and (38),
tively looking at an infinite fluid with a prescribed number
αk ∝ n1−k/2 as n → 0 . (39) density n = N/L3 and we can calculate the behaviour of kur-
tosis α4 as n varies. The results are presented in Figure 4.
While this result has been calculated for N = 2 interacting We see general agreement between behaviour of each of the
atoms, it is also confirmed for larger values of N by estimating four systems. We see when n is equal, the values of kurto-
the k-th standardised moments using MD simulations, as it is sis are larger for N = 2 than for the many-body systems with
shown in Figure 5. N = 8, 64, 512, which agree well amongst themselves.
The results in Figure 4 enable us to test the asymptotic
expression (39) for k = 4 derived in the limit n → 0. Util-
B. MD simulations with N interacting atoms ising similar log-log plots for MD data, we estimate the
power law behaviour of each standardised moment, αk , for
In this section we present the results from MD simulations of k = 4, 6, 8, 10, 12. Figure 5 illustrates the results. All systems
many-body systems in three spatial dimensions using differ- agree well with the predicted asymptotic behaviour (39), in
ent values of N, including the case N = 2 (analyzed in Sec- particular the N = 512 atom system. There is a slight devia-
tion IV A). Atoms are subject to pairwise interactions gov- tion between the results due to the fact that the smaller atom
erned by a Lennard-Jones potential, given in equation (3). For systems require a larger tsim in order to converge fully to the
each system we use a velocity-Verlet23 integrator and main- predicted value. This discrepancy is amplified when looking
tain the system in the canonical ensemble by incorporating at higher standardised moments due to the fact that we are cal-
a Nosé-Hoover thermostat38 , see Appendix B. We perform culating statistics resulting from F112 (i.e. for α12 ) compared
two types of MD simulation studies: those that are used for to F14 (i.e. for α4 ), for example.
studying how the number density, n, of a system affects stan- The dependence of kurtosis α4 on temperature T is pre-
dardised moments, and those that aim to probe temperature sented in Figure 6, where we keep the density fixed at n = n0
On standardised moments of force distribution in simple liquids 8

10 6

10 4

10 2

10 0
10 -5 10 -4 10 -3 10 -2 10 -1 10 0

FIG. 4. Dependence of kurtosis α4 on density n. Each of the larger atomic systems (N = 8, 64, 512) is simulated over the same domain of
number densities, while the N = 2 system is simulated in a sparser domain. We truncate the results of the N = 2 simulation in the plot, though
the additional data points are used in calculating asymptotic results displayed in Figure 5.

5 800

4
600
3
400
2
200
1

0 0
4 6 8 10 12 0 0.5 1 1.5

FIG. 5. Comparison of the results of N-atom simulations for a range FIG. 6. Dependence of kurtosis α4 on temperature T . Each atomic
of values of the total number of atoms, N, found within the principal system is simulated at approximately the same density n = n0 given
cubic domain. After long time simulation, we compute the asymptotic in Table I.
behaviour αk ∝ n−κ and compare the leading order power scalings
for each system. We compare this with the theoretical result (39)
(denoted with a blue dashed line) that in the limit n → 0 we expect
ier tails and therefore distributions which become increasingly
the universal behaviour κ = k/2 − 1.
leptokurtic.
In Figure 6, we observe that there is a qualitative difference
between the results for N = 2 and larger atom systems. We
given in Table I. We observe that as temperature increases so see a bifurcation for the N = 64, 512 system at some critical
does the kurtosis of the force distribution associated with each temperature Tc ∈ (0.6, 0.65), where a steady increase in kur-
system. This can be explained in terms of the dynamics of the tosis meets a large jump resulting in a rapid increase. This
interacting atom system. If we maintain each system in the difference is due to a clustering mechanism which has been
canonical ensemble, we expect on average that each atom will seen in MD simulations of Lennard-Jones fluids30 . From our
have a kinetic energy equivalent to 3 T /2 (when in reduced results we see that the N = 2 system has missed this behaviour
units). As we increase this target temperature, the atoms be- completely. Snapshots of the N = 512-atom system at some
come more energetic and thus are able to probe closer inter- T = 0.6 < Tc , and T = 0.66 > Tc are displayed in Figure 7.
atomic distances before a large repulsive force overcomes this For T = 0.6, we see a large cluster has formed in the many-
inertial attraction. The range of forces on the tagged parti- atom system. There would be far fewer outlier force results
cle widens as temperature increases and therefore contributes in this case due to the fact that the large majority of atoms
to more outlier results in the distribution - leading to heav- are moving as a collective and effectively have fixed inter-
On standardised moments of force distribution in simple liquids 9

atomic forces. Compared to the T = 0.66 snapshot, where (a) (b)


we see that the atoms are too kinetically unstable to form
these larger stable cluster structures, this results in more out-
lier forces felt between atoms due to the fact that the system
is intrinsically more disordered. It is useful to note that this
bifurcation point is located on the vapour-liquid coexistence
boundary, the mechanisms of which have been studied on di-
lute Lennard-Jones fluids31 ; here we see that this results in a
bifurcation on standardised moments of the force distribution.
To understand the underlying variations of kurtosis, α4 ,
with respect to changes in temperature and density, we use
12 × 16 MD simulations with N = 512 atoms, varying sim-
ulation parameters (n, T ), where n = 10−2 + (i − 1)/10, for FIG. 7. Snapshots32 of the MD simulation are taken for the system
i = 1, 2, . . . , 12, and T = j/10, for j = 1, 2, . . . , 16. Each sim- with N = 512 atoms at time t = 7.5 × 105 for: (a) T = 0.6 < Tc ; and
(b) T = 0.66 > Tc . Density is n = 1/64.
ulation was performed with similar initialisation as described
previously in Section IV B with tsim = 106 . The results for ex-
cess kurtosis (α4 − 3) are displayed in Figure 8. Here a bifur-
cation can be seen when using the smallest density n = 0.01,
as the change in colour is prominent in this vertical strip, indi-
cating a large change of kurtosis. This occurs around T = 0.6, 1.5
which is consistent with the result in Figure 6, where we saw
the bifurcation similarly located, though the slight shift in
temperature is accounted for by the shift in density parame-
ters used in each simulation (namely n = 0.01 and n = 1/64). 1
In general, this low density strip contains the largest val-
ues of kurtosis, and covers much of the purely gas phase of
the Lennard-Jones fluid. This paper has so far probed the low
density limit in an attempt to understand why the standardised 0.5
moments of force are so large, though Figure 8 gives a good
overview that in general, regardless of phase, a decrease in
temperature, or an increase in density, systematically lead to 0 0.2 0.4 0.6 0.8 1
a lower value of standardised moments. In this case as n → ∞
or T → 0, we expect the α4 → 3 (excess kurtosis tends to
zero). This limiting regime corresponds to the solid phase of
FIG. 8. The excess kurtosis, α4 − 3, calculated as a function of den-
a Lennard-Jones system, where the force variations are min- sity n and temperature T for n ≤ 1.11 and T ≤ 1.6. The white dot-
imal and the distribution is Gaussian. There is not enough ted lines describe coexistence lines of different phases of a Lennard-
space, nor energy, that lead to (many) outlier forces experi- Jones fluid taken from the literature33–36 . The solid black dots in-
enced by any atom, so the force distribution becomes less and dicate (from left to right), the critical point and vapour-liquid-solid
less skewed from Gaussian, the deeper we probe in these re- triple points.
gions. This intuition was demonstrated analytically in Sec-
tion III E when we showed this limiting behaviour on a 1D
cartoon model with equation (32). It is interesting to note that
the asymptotic expression (39). Though results are derived
these changes in values of α4 appear smooth about changes in
with a two-body system in a principal cubic domain, utilising
temperature and density (in absence of the bifurcation point
the minimum image convention implies that we are studying
for larger values of n), regardless of phase transitions.
an infinite fluid at a prescribed number density. Therefore an-
alytic results are contrasted with MD simulations of four sys-
tems of N = 2, 8, 64, 512 interacting Lennard-Jones atoms and
V. DISCUSSION AND CONCLUSIONS these are compared. The results agree well with theoretical
predictions though the results for systems with larger values
In Section III we have demonstrated use of a variety of meth- of N are seen to converge more readily to the theoretically pre-
ods to study the standardised moments of the force distribu- dicted results. In particular, rich dynamics such as clustering
tion in order to probe both their temperature and number den- of Lennard-Jones fluids is completely missed by the systems
sity dependence. This gave way to a rich structure where we with smaller values of N, but captured for systems with N as
show that the partition function for a 1D system can be cal- small as N = 64 atoms. In general, as temperature increases
culated entirely from these standardised moments. Extending αk increases due to energetic nature of atoms allowing them
the far field method introduced in Section III B to a system to push closer together and experience larger forces. Cluster-
with N atoms in three-dimensional physical space, Section IV ing exhibited at the vapour-liquid coexistence phase incurs a
studies the dependence of αk on number density n, deriving bifurcation point whereby a large increase is seen in the stan-
On standardised moments of force distribution in simple liquids 10

dardised moments of force in Figure 6, though a general in- no closed form for the number of solutions. In particular,
crease in temperature, or decrease in number density, results simplifying equation (A2) by solving equations (A3)–(A4) is,
in an increase in a4 regardless of the temperature/number den- in general, not possible. However, noting the specific prop-
sity domain studied, as shown in Figure 8. erty that F(r∗ ) = 0 = U ′ (r∗ ), we see that all terms that have
α0 6= 0 will vanish when evaluated at this unique minimum
r = r∗ . In particular, we will obtain relatively simple forms
ACKNOWLEDGMENTS of the sum (A2) for m = k + 1 and m = k + 2 by considering
equations (A3)–(A4) with α0 = 0.
This work was supported by the Royal Society [grant number First, let us consider that m = k + 1. Using α0 = 0, there
RGF\EA\180058] and by the Engineering and Physical Sci- is only one solution of equations (A3)–(A4) in non-negative
ences Research Council [grant number EP/V047469/1]. integers, namely α1 = k − 1, α2 = 1 and α3 = α4 = · · · = 0.
Therefore, equation (A2) implies
 k−1
Appendix A: Derivation of equations (28) and (29) F k, (k+1) (r∗ ) = C(0, k − 1, 1, 0, . . . , 0) F (1) (r∗ ) F (2) (r∗ ).

The constants used in equation (30), namely Ck,1 and Ck,2 , are Using the general Leibniz rule37 , we evaluate the combina-
given in equations (28) and (29), which can be derived in the torial prefactor as C(0, k − 1, 1, 0, . . . , 0) = k(k + 1)!/2. Sub-
following manner. Using F k (r) = (U ′ (r))k for even values of stituting into Ck,1 = F k, (k+1) (r∗ )/(k + 1)! and using F(r∗ ) =
k and U ′ (r∗ ) = 0, we first note that −U ′ (r∗ ) and that k is an even integer, we obtain formula (28).
k Second, we will consider the case m = k + 2. Using α0 =
F k, k (r∗ ) = k! U ′′ (r∗ ) ,
0, there are two solutions of equations (A3)–(A4) in non-
F k, m (r∗ ) = 0, for m ≤ k − 1, negative integers. The first solution is α1 = k − 1, α2 = 0,
α3 = 1 and α4 = α5 = · · · = 0.The second solution is α1 =
where F k, m
denotes the m-th derivative of F k , i.e. the m-th k − 2, α2 = 2 and α3 = α4 = · · · = 0. Therefore, equation (A2)
derivative of the k-th power of F. Therefore, the first three implies
non-zero terms of the Taylor expansion of F k (r) around r = r∗
are F k, (k+2) (r∗ )
k−1
F k, (k+1) (r∗ )

= C(0, k − 1, 0, 1, 0, . . . , 0) F (1) (r∗ ) F (3) (r∗ )
k ∗ k ′′ ∗
k ∗ k+1
F (r) ≈ (r − r ) U (r ) + (r − r )
(k + 1)!  k−2  2
F k, (k+2) (r∗ ) +C(0, k − 2, 2, 0, 0, . . . , 0) F (1) (r∗ ) F (2) (r∗ ) .
+(r − r∗ )k+2 . (A1)
(k + 2)!
Using the general Leibniz rule37 , we evaluate these combina-
Therefore, we have Ck,1 = F k, (k+1) (r∗ )/(k + 1)!
and Ck,2 = torial prefactors as
F k, (k+2) (r∗ )/(k + 2)! and, to derive equations (28) and (29), k
we need to express derivatives F k, m (r∗ ) for m = k + 1 and C(0, k − 1, 0, 1, 0, . . . , 0) = (k + 2)!,
m = k + 2 in terms of derivatives of U(r) at r = r∗ . Using the 6
product rule, the m-th derivative of F k (r) can be, in general, k(k − 1)
C(0, k − 2, 2, 0, 0, . . . , 0) = (k + 2)!.
written as a finite sum of the form 8
k  m αi Substituting into formula Ck,2 = F k, (k+2) (r∗ )/(k + 2)! and us-
F k, m
(r) = ∑ C(α0 , α1 , . . . , αm ) ∏ F (i) (r) , (A2) ing F(r∗ ) = −U ′ (r∗ ) and that k is an even integer, we obtain
α0 ,α1 ,...,αm =0 i=0 equation (29). Thus, we have arrived at the the expressions
for Ck,2 and Ck,2 that are used in equation (30).
where F (i) (r) is the i-th derivative of function F(r) and
C(α0 , α1 , . . . , αm ) are constants, many of them equal to zero.
In fact, all terms in the expansion (A2) have multiplicities that Appendix B: Thermostats used in MD simulations
sum to k, that is we can only sum over sequences satisfying
m Considering 3D simulations in Section IV B, we use a Nosé-
∑ αi = k, (A3) Hoover thermostat. Its parameter, originally38 denoted Q, is
i=0 the relaxation time of the thermostat. It is a measure of how
and all terms in the expansion (A2) have m derivatives, that is, strongly the thermostat is attached to the dynamics of the sys-
we have tem. We choose a cautious value of Q = 10 T for each simula-
tion; this linear scaling with T is necessary as we need to more
m
tightly couple the thermostat at lower temperatures in order to
∑ i αi = m, (A4)
accurately maintain the system in the canonical ensemble39 .
i=0
For 1D simulations in Section III, we maintain the canon-
where αi ∈ {0, 1, . . . , k} for i = 0, 1, 2, . . . , m. Equation (A4) ical ensemble at a target (reduced) temperature T by imple-
is of the form of a finite Diophantine equation, which has menting a Langevin thermostat. This is due to problems with
On standardised moments of force distribution in simple liquids 11

ergodicity utilising the Nosé-Hoover thermostat for small sys- 19 J. Jones. On the determination of molecular fields. — II. From the equation
tems40,41 . Here the evolution of the free particle is modelled of state of a gas. Proceedings of the Royal Society A, 106:738 (1924)
20 H. Watanabe, N. Ito and C. Hu. Phase diagram and universality of the
(in reduced units) as42,43
Lennard-Jones gas-liquid system. Journal of Chemical Physics 136, 204102
(2012)
dU p 21 J-P. Hansen and L. Verlet. Phase transitions of the Lennard-Jones system.
ẍ = − − γ ẋ + 2 γ T R(t) , (B1) Physical Review 184:151 (1969)
dx 22 A. Rahman. Correlations in the motion of atoms in liquid argon. Physical

Review A 136(2) (1964)


where R(t) is standard white noise, and γ acts as a friction pa- 23 L. Verlet. Computer "experiments" on classical fluids. I. Thermodynamical
rameter. We choose γ = 0.1 when calculating our illustrative properties of Lennard-Jones molecules. Physical Review 159(1) (1967)
results presented Figures 1 and 3. 24 D. Frenkel and B. Smit. Understanding molecular simulation: from algo-

rithms to applications. 2nd ed. Academic Press (1996)


25 L. Rowley, D. Nicholson and N. G. Parsonage. Monte Carlo grand canon-

ical ensemble calculation in a gas-liquid transition region for 12-6 Argon.


Journal of Computational Physics 17, p401 (1975)
REFERENCES 26 D. Lide. Properties of the elements and inorganic compounds; melting,

boiling, triple, and critical temperatures of the elements. CRC Handbook


of Chemistry and Physics (86th ed.), CRC Press, Chapter 4 (2005)
27 A. Giró, E. Guardia and J. A. Padró. Langevin and molecular dynamics sim-
1 S. Joshi and S. Deshmukh, A review of advancements in coarse- ulations of Lennard-Jones liquids. Journal of Molecular Physics 55, Issue 5
grained molecular dynamics simulations. Molecular Simulation, DOI: (1985)
28 C. M. Bender and S. A. Orszag. Advanced mathematical methods for sci-
10.1080/08927022.2020.1828583 (2020)
2 Y. Wang et al. Effective force coarse-graining. Journal of Physical Chem- entists and engineers: asymptotic methods and perturbation theory: v. 1.
istry Chemical Physics 11, p2002 (2009) Springer (1999)
3 R. Erban and S. J. Chapman. Stochastic modelling of reaction-diffusion pro- 29 S. Kirkpatrick, C. Gelatt Jr and M. Vecchi. Optimization by simulated an-

cesses, Cambridge Texts in Applied Mathematics, Cambridge University nealing. Science 220, p671 (1983)
30 N. Yoshii and S. Okazaki. Molecular dynamics study of structure of clusters
Press (2020)
4 H. Ingólfsson et al. The power of coarse graining in biomolecular simula- in supercritical Lennard–Jones fluid. Fluid Phase Equilibria 144(1-2), p225
tions. Wiley Interdisciplinary Reviews: Computational Molecular Science (1998)
31 J. Jung, J. Lee and J. Kim. Cluster growth mechanisms in Lennard-Jones
4(3), p225 (2014)
5 A. Davtyan et al. Dynamic force matching: A method for constructing dy- fluids: a comparison between molecular dynamics and Brownian dynamics
namical coarse-grained models with realistic time dependence. Journal of simulations. Chemical Physics 449, p1 (2015)
32 W. Humphrey, A. Dalke and K. Schulten. VMD - Visual Molecular Dynam-
Chemical Physics 142, 154104 (2015)
6 R. Erban. Coupling all-atom molecular dynamics simulations of ions ics. Journal of Molecular Graphics 14, p33 (1996)
33 A. Schultz and D. Kofke. Erratum: "Comprehensive high-precision
in water with Brownian dynamics. Proceedings of the Royal Society A
472(2186):20150556 (2016) high-accuracy equation of state and coexistence properties for classi-
7 D. Wales. Exploring energy landscapes. Annual Review of Physical Chem- cal Lennard-Jones crystals and low-temperature fluid phases". Journal of
istry 69, p401 (2018) Chemical Physics 153, 059901 (2020)
8 R. Gunaratne et al. On short-range and long-range interactions in multi- 34 S. Stephan et al. Thermophysical properties of the Lennard-Jones fluid:

resolution dimer models. Interface Focus 9 (3), rsfs.2018.0070 (2019) database and data assessment. Journal of Chemical Information and Mod-
9 E. Rolls, Y. Togashi and R. Erban. Varying the resolution of the Rouse model eling 2019 59 (10), 4248 (2019)
35 S. Stephan, J. Staubach and H. Hasse. Review and comparison of equations
on temporal and spatial scales: application to multiscale modelling of DNA
dynamics. Multiscale Modeling and Simulation 15(4), p1672 (2017) of state for the Lennard-Jones fluid. Fluid Phase Equilibria 523, 112772
10 R. Erban. From molecular dynamics to Brownian dynamics, Proceedings of (2020)
36 E. Mastny and J. de Pablo. Melting line of the Lennard-Jones system, infinite
the Royal Society A 470(2167): 20140036 (2014)
11 L. DeCarlo. On the meaning and use of kurtosis. Psychological Methods size, and full potential. Journal of Chemical Physics 127, 104504 (2007)
37 A. Traheem and A.Laradji. Classroom note: a generalization of Leibniz
2(3), p292 (2014)
12 A. Carof, R. Vuilleumier and B. Rotenberg. Two algorithms to compute rule for higher derivatives. International Journal of Mathematical Educa-
projected correlation functions in molecular dynamics simulations. Journal tion in Science and Technology 34(6): 905 (2003)
38 S. Nosé. A molecular dynamics method for simulations in the canonical
of Chemical Physics 140, 124103 (2014)
13 H. Shin et al. Brownian motion from molecular dynamics. Chemical Physics ensemble. Molecular Physics 52(2), p255 (1984)
39 P. H. Hünenberger. Thermostat algorithms for molecular dynamics simu-
375, p316 (2010)
14 R. Erban. Coarse-graining molecular dynamics: stochastic models with lations. In: Dr. Holm C., Prof. Dr. Kremer K. (eds) Springer Advanced
non-Gaussian force distributions. Journal of Mathematical Biology 80, Computer Simulation. Advances in Polymer Science 173, p105 (2005)
40 M. E. Tuckerman et al. Non-Hamiltonian molecular dynamics: generaliz-
p457 (2020)
15 S. Chandrasekhar. Stochastic problems in physics and astronomy. Review ing Hamiltonian phase space principles to non-Hamiltonian systems. Jour-
of Modern Physics 15:1 (1943) nal of Chemical Physics 115, p1678 (2001)
16 A. Gabrielli et al. Force distribution in a randomly perturbed lattice of iden- 41 P. F. Tupper. Ergodicity and the numerical simulation of Hamiltonian sys-

tical atoms with 1/r2 pair interaction. Physical Review E 74:021110 (2006) tems. SIAM Journal on Applied Dynamical Systems 4(3), p563 (2005)
17 G. Rickayzen et al. Single atom force distributions in simple fluids. Journal 42 T. Schlick. Molecular modeling and simulation. Springer (2002)
43 B. Leimkuhler and C. Matthews. Molecular dynamics with determinis-
of Chemical Physics 137, 094505 (2012)
18 A. C. Branka, D. M. Heyes and G. Rickayzen. Pair force distributions in tic and stochastic numerical methods. Springer Interdisciplinary Applied
simple fluids. Journal of Chemical Physics 135, 164507 (2011) Mathematics 39 (2015)

You might also like