Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

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

net/publication/226221631

A New Method for Computation of Long Ranged


Coulomb Forces in Computer Simulation of Disordered
Systems

Article in Journal of Low Temperature Physics · January 2005


DOI: 10.1007/s10909-005-5451-5

CITATIONS READS

30 406

2 authors:

Eugene Yakub Claudio. Ronchi


Odesa National Economic University European Commission
45 PUBLICATIONS 438 CITATIONS 134 PUBLICATIONS 2,998 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Eugene Yakub on 21 May 2014.

The user has requested enhancement of the downloaded file.


Journal of Low Temperature Physics, Vol. 139, Nos. 5/6, June 2005 (© 2005)
DOI: 10.1007/s10909-005-5451-5

A New Method for Computation of Long Ranged


Coulomb Forces in Computer Simulation of
Disordered Systems

E. Yakub1 and C. Ronchi2

1
European Commission, Joint Research Centre, Institute for
Transuranium Elements, P.O. Box 2340, D-76125 Karlsruhe, Germany;
2
Computer Science Dept., Odessa State Economic University,
Preobrazhenskaya 8, 65026, Odessa, Ukraine
E-mail: yakub@unive.odessa.ua

Applications of a new method for computation of Coulomb forces in Monte


Carlo or molecular dynamics simulation of a wide class of disordered sys-
tems including plasmas, ionic fluids and amorphous solids is discussed. This
method, based on angular averaging of Ewald sums over all orientations of the
reciprocal lattice under conditions of computer simulation, eliminates peri-
odicity artifacts imposed by conventional Ewald scheme and provides much
faster computation of electrostatic energy in computer simulations of dis-
ordered condensed systems.

1. Introduction

Computer simulations of disordered systems like plasmas, ionic fluids,


amorphous solids, etc., require an accurate account for long ranged Coulomb
forces which determine to the largest extent the correctness in the predicted
stability of disordered condensed phases and strongly affect the predictions of
their thermodynamic and transport properties. An essential problem both
in conventional1 and ab initio 2 computer simulations of such systems3 is
how to combine the accurate account of long ranged Coulomb forces with
periodic boundary conditions (PBC). Despite the great progress achieved in
the microscopic description of condensed matter since the pioneering works
of Madelung4 and Ewald,5 the correct description of Coulomb forces under
conditions of computer simulation still remains a topical issue.6 Usual Ewald
summation procedure when applied to simulation of disordered Coulomb

633

0022-2291/05/0600-0633/0 © 2005 Springer Science+Business Media, Inc.


634 E. Yakub and C. Ronchi

systems invokes a non-isotropic electric field having the cubic symmetry of


the crystalline lattice composed of main cells as elementary units. It results
in an artificial “crystalline field” effects in simulation of fluids, amorphous
solids and other spatially uniform condensed phases.7
The mean Coulomb interaction energy of two particles at given distance
r in uniform systems is independent of orientation of vector r. This is obvi-
ously not the case in the Ewald scheme, and the example presented in Ref. 7
shows how important may be these effects of artificial “crystalline field” in
real simulations.
In ab initio computer simulations,3 the number of particles in the main
cell is very limited, mainly by supercomputer facilities available. Periodicity
artifacts imposed by conventional Ewald summation procedure may appear
a major issue here. The absolute value of the periodicity artifacts in effective
Coulomb interaction is almost negligible at small distances, even for a relat-
ively small number N of charged particles in the cell, but becomes essential
at larger distances. For instance, for N = 200, which is characteristic for
up-to-date ab initio simulations, using the best available computing codes
and facilities,9,10 the maximum value of the Ewald artifact is about 10% in
energy in third coordination sphere and reaches ∼ 100% at the edge of the
cell.7
Accurate computer simulations of biochemical and other systems having
complex elemental composition require sometimes up to a million particles8 in
the main Monte Carlo or molecular dynamics computation cell1 . Obviously,
the larger the number of charged particles in the main cell, the more acute is
the problem of the effective evaluation of the electrostatic contribution. Fast
and accurate evaluation of electrostatic fields is important therewith for any
size of the cell. Periodicity artifacts in this case are small but heavy processor
load imposed by conventional Ewald summation procedure is crucial in such
simulations.
We apply here the recent modification of the Ewald scheme7 to a few
simplest systems both spatially uniform and ordered. In Section 2. we out-
line the method, in Section 3. we present its generalization on one-component
plasma (OCP) and new Monte Carlo simulation results on OCP and restric-
ted primitive ionic model (RPIM) fluid, as well as calculations of Madelung
constant for three ordered structures, in Section 4. we discuss possible applic-
ations of the method in study of a wide class of disordered Coulomb systems
including highly compressed fluids and amorphous phases of cryogenic solids
and present some concluding remarks.

1
Thereinafter simply called main cell.
Computation of Coulomb Forces in Disordered Systems 635

2. PRE-AVERAGED EFFECTIVE POTENTIALS

Let us consider a standard cubic main cell of edge L and volume V = L3 ,


containing N = M α=1 Nα charged particles of M sorts. The electrostatic
P

forces acting between the i-th and the j-th particles obey the Coulomb law:
Qi Qj
Fij = 2 .
4πε0 rij

Here rij = |r i −r j | is the distance between i-th and j-th charged particles and
Qi is the value of the i-th point charge of type α : Q(α) = {Q(1) , . . . , Q(M ) }.
We shall here assume that the electro-neutrality condition
N M
Nα Q(α) = 0
X X
Qi = (1)
i=1 α=1

is satisfied, and the standard PBC are imposed as described in Refs. 1,11.
The total Coulomb energy of N charges in the main cell is:
(C) X
UN = Qi ϕ(r i ) ,
1≤i≤N

where ϕ(r i ) is the electrostatic potential at the position r i of i-th charge.


According to Ewald,5 this contribution is, in turn, the sum of one- and two-
particle terms:
N
1X
ϕ(r i ) = ϕ1 (r i ) + ϕ2 (r i , r j ) .
2 j6=i

In the absence of external field, the unary potential is a constant:


" # !
Qi 1 X 1 π 2 n2 δ
ϕ1 = 2
exp − 2
−√ (2)
4πε0 L 2π n>0 n δ π

and the binary contribution can be written as follows:5,11


(
Qj erfc(δ(rij /L))
ϕ2 (r ij ) =
4πε0 rij
" # )
1 X 1 π 2 n2 2π

+ exp − cos n · r ij . (3)
πL n>0 n2 δ2 L

Here δ/L is the conventional Ewald parameter,5,11 n/L is the three-dimensio-


nal reciprocal lattice site vector (n = |n|), and erfc(x) is the complementary
error function.
636 E. Yakub and C. Ronchi

Keeping in mind that all orientations of the main cell in an isotropic me-
dia should be equivalent, we can average both sides of (3) over all directions
of the vector n at a fixed distance rij . If the brackets
Z +1 Z π
1
h. . .i = d(cos ϑ) dψ . . . ,
4π −1 −π

indicate averaging, whereas ψ, ϑ are the polar and azimuthal angles defining
the direction of the vector n (n · r = nr cos ϑ), we can determine the pre-
averaged (effective) potential as ϕ2 (rij ) ≡ hϕ2 (r ij )i. Integration of (3) over
all orientations of the vector n gives immediately:
Qj rij
  
ϕ2 (rij ) = erfc δ
4πε0 rij L
" # )
1 X 1 π 2 n2 2π
 
+ exp − sin nrij . (4)
2π 2 n>0 n3 δ2 L

The pre-averaged charge–charge potential (4) is a continuous function


of the inter-particle distance rij and can be expanded in converging power
series in terms of this distance. Since both erfc(x) − 1 and sin(x) are odd
functions  
Qj  X
2k+1 
ϕ2 (rij ) = 1+ Ck rij . (5)
4πε0 rij k≥0

The coefficients Ck in (5) are found in Ref. 7 by direct expansion of (4) in


a MacLaurin’s series. The procedure is straightforward: by applying the
Euler–MacLaurin formula — generalized for summation over three-dimen-
sional integers n,12 — the following result holds:
" #
1 X 1 π 2 n2 2δ
C0 = 2
exp − 2 − √ ,
π n>0 n δ π

C1 = ,
3L3
Ck = 0, for k > 1 .

By taking into account the electro-neutrality condition (1), it can be


seen that the term in (5), which is independent of distance (proportional to
C0 ) and the one-particle contribution (2) cancel one another. This implies
that the total Coulomb energy of N charged particles in the main cell can
be described by the sum:
N N X N
(C) X 3Q2i 1X
UN = − + φ̃(rij ) , (6)
i=1
16πε0 rm 2 i=1 j=1
j6=i
Computation of Coulomb Forces in Disordered Systems 637

where φ̃(rij ) is an effective potential defined by:


      2 
 Qi Qj 1 + 1 r r
− 3 r < rm
φ̃(r) = 4πε0 r 2 rm rm (7)
0 r ≥ rm

and rm = (3/4π)1/3 L is the radius of the volume-equivalent sphere of the


main cell: 43 πrm
3 = L3 .

The pair effective (pre-averaged) potential (7) has the following properties:

• It tends to the pure Coulomb pair potential at small distances.


• It is zero at r = rm and remains zero at r > rm .
• Its first derivative is zero at r = rm .
• Its range is related to the size of the main cell: rm = 0.62035 L.

The last property entails some inconsistency with the initial main cell con-
figuration. If the range of interaction does not exceed L/2 each particle in
the cell contributes (or not) to the sum of interactions with the selected one
just one time: either as original object (inside the main cell) or as one of its
“images.” This is not the case for potential (7) because rm > L/2.
The subsequent modification of the simulation algorithm was described
in Ref. 7. Two different zones exist within the effective sphere surrounding an
arbitrary charge in the main cell.7 The first zone contains charged particles
(inside the main cell) or their images (outside it) which contribute to the
sum of pair interactions of the chosen charge just one time. The second zone
contains those charged particles which contribute twice to the interaction en-
ergy with the given charge — both as original charges and as their additional
“phantom” images.7

3. MONTE CARLO SIMULATION

3.1. One-Component Plasma

One-component plasma (OCP), i.e. a system of point ions placed in


neutralizing background is a text-book example of classical Coulomb system
for which the standard Ewald approach was successfully applied.13,14 OCP
may be regarded as a limiting case of a two-component system when the
number of e.g. negative ions increases and the charge per ion decreases, thus
maintaining the electroneutrality. Separating the interactions of the negative
(background) charge distribution in (6) and replacing summation over neg-
ative ions by integration over uniformly distributed background charge, one
638 E. Yakub and C. Ronchi

gets the following generalization of (6) for OCP:


N X N
(OCP) N Q2 1X
UN = −0.9 + φ̃(rij ) , (8)
4πε0 rm 2 i=1 j=1
j6=i

where N is the number of positive ions. It should be noted that both terms
in (8) are size-dependent but not directly proportional to N . The first con-
tribution in (8) at given temperature T may be expressed in terms of plasma
parameter Γ = Q2 N 1/3 (4πε0 rm kT )−1 as −0.9kT Γ N 2/3 .

(OCP)
Table 1. Comparison of predicted Monte Carlo internal energies −UMC /N kT
for one-component plasma.

Γ N = 64 (This work) N = 216 13 N = 686 14


1 0.599 ± 0.010 0.580
10 7.895 ± 0.020 7.996
50 43.073 ± 0.019 43.094
100 87.555 ± 0.033 87.480 87.52a)
160 141.00 ± 0.016 140.890 141.72
200 176.62 ± 0.012 176.77b)
a)
Best fit value
b)
Fluid initial conditions

We performed two series of Monte Carlo simulations using the method


outlined above and the algorithm described in Ref. 7 to prove that (8) gives
correct estimations of the OCP energy at relatively small N and the mean
(OCP)
value of UN /N converges for large N . Detail of Monte Carlo simulation
procedure is the same as reported earlier.16 The results are presented in
Table 1 for Γ = 1 . . . 200 at fixed N = 64 and compared in Fig. 1 to res-
ults obtained within standard Ewald scheme by Hansen,14 and Stringfellow,
DeWitt and Slattery14 at fixed Γ = 100 for wide range of N .

3.2. Restricted Primitive Ionic Model

Another simple model system suitable for testing of the method is the
Restricted Primitive Ionic Model (RPIM), i.e. two-component fluid build
from oppositely charged hard spheres of one size (diameter σ). We per-
formed an additional set of Monte Carlo simulations for RPIM fluid at
Computation of Coulomb Forces in Disordered Systems 639

(MC)
Fig. 1. Thermal contribution to internal energy ∆E (th) /N kT = (UN −
(bcc)
UN )/N kT of one-component plasma at fixed Γ = 100 and different sizes of the
main cell. Comparison of our Monte Carlo simulation results (open circles) with
data of Hansen13 (solid square), and the “best fit” value recommended in Ref. 14
(solid triangle). Cubic spline interpolation of data presented in Table 3.3. was util-
(bcc)
ized to calculate the Madelung bcc-contribution UN .

one reduced temperature T ∗ = 4πε0 kT σQ−2 and different packing fractions


η = πN σ 3 (6V )−1 for wide range of N to test the convergence for large N .
In Fig. 2 we compare our calculations with results of Larsen.15

3.3. Ordered Solid Structures

Perfect crystalline structures are the worst application case for a pre-
averaged effective potential essentially devised for spatially uniform systems.
Nevertheless, the test on the perfect NaCl structure presented in Ref. 7 proves
the ability of the method to predict accurate values of the electrostatic energy
even in ordered condensed systems (see Table 1 in Ref. 7). In this work we
performed additional tests of the method using three other cubic crystalline
structures.
The values of the Madelung constant for bcc-structure of Coulomb crys-
tal formed in OCP, CsCl lattice and fluorite (CaF2 ) structure are computed
from (6)–(7) and presented in Tables 3.3., 3.3. and 3.3. at different numbers
of charged particles in the main cell.
640 E. Yakub and C. Ronchi

Fig. 2. Deviations of Monte Carlo internal energy from data of Larsen:15 ∆E/kT =
(MC) (MC)
(UN − UN )/N kT at fixed temperature T ∗ = 4πε0 kT σQ−2 = 0.595 and four
different packing fractions plotted against the size of the main cell: 1 – η = 0.005023
(squares), 2 – η = 0.02058 (triangles), 3 – η = 0.04842 (circles), 4 – η = 0.09525
(diamonds).

4. CONCLUDING REMARKS

Pre-averaging the Ewald sums over all spatial directions leads to an en-
ergy formulation which eliminates periodicity artifacts imposed by conven-
tional Ewald scheme and provides a fast method for computation of electro-
static contribution to the energy of disordered dense systems in Monte Carlo
or molecular dynamics computer simulations.7 Additional calculations and
Monte Carlo simulation results presented in Section 3. prove the ability of the
method to produce accurate simulation results at low computational cost.
Monte Carlo simulations of one-component plasma (small cell size N = 64)
presented in Table 1 are in good agreement with other simulation data.13,14
At the same time the comparison presented in Fig. 1 shows that it is very
likely coincidental. The convergence to a constant value at large N is non-
monotonous. Predicted thermal contribution to internal energy ∆E (th) /N kT
at Γ = 100 shown in Fig. 1 is in accordance with data of Hansen13 but is
higher than predicted in Ref. 14. This rather small (about 0.5% of the total
energy) difference is well beyond the total estimated error of both simulations.
This discrepancy may be due to periodicity artifacts and requires additional
study because it may affect the location of OCP crystallization point.
Computation of Coulomb Forces in Disordered Systems 641

Table 2. Predicted values of the Madelung constant for OCP Coulomb crystal (bcc
lattice). L/a is the ratio of the cell size to the lattice constant, in the second and
third columns are numbers of charged particles in the cell and in the equivalent
sphere. The fourth column show their difference (the actual net charge of the
equivalent sphere). The last two columns present the predicted Madelung constants
and their deviations from the exact value.
Nr of particles in: Deviation Madelung constant
L/a
Cell Sphere ∆(N ) This work ∆,%
2 16 15 −1 −0.89230 −0.41
3 54 59 5 −0.90361 +0.86
4 128 137 9 −0.89985 +0.44
5 250 259 9 −0.89921 +0.37
6 432 411 −21 −0.89595 +0.00
7 686 701 15 −0.89523 −0.08
8 1024 965 −59 −0.89411 −0.20
9 1458 1459 1 −0.89510 −0.09
10 2000 1989 −11 −0.89591 +0.00
11 2662 2685 23 −0.89656 +0.07
12 3456 3479 23 −0.89648 +0.06
13 4394 4477 83 −0.89633 +0.04
14 5488 5577 89 −0.89621 +0.03
15 6750 6735 −15 −0.89579 −0.02
Exact = −0.89593

Our simulation of restricted primitive ionic model fluid presented in


Fig. 2 is in good agreement with the results of Larsen.15 The smaller the
packing fraction the faster is the convergence.
As it was found in Ref. 7, the pre-averaged effective potential originally
devised for spatially uniform systems provides surprisingly fast convergence
of predicted energy to exact values in ordered crystalline structures, despite
the inevitable non-zero net charge inside the equivalent sphere. New calcula-
tions of Madelung constant provided in Tables 3.3.–3.3. confirm this feature
for three other periodic structures. Note that the calculated electrostatic
energy oscillates upon approaching the exact value. The non-monotonous
convergence to the large-N limit is also characteristic for OCP and RPIM
models. This feature must be taken into account in precision calculations for
obtaining results in this limit.
We believe the implementation of pre-averaged potentials in existing
and upcoming conventional and ab initio simulation packages may essentially
speed up the simulations and decrease the ab initio simulation errors caused
by periodicity artifacts, especially near the phase transition points.
642 E. Yakub and C. Ronchi

Table 3. Predicted values of the Madelung constant for CsCl lattice. Columns are
the same as in the Table 2, except for the deviations in number of ions and net
charge (columns four and five) differ.

Numb. inside: Deviations of: Madelung constant


L/a
Cell Sphere Number Charge This work ∆,%
2 16 15 −1 −1 1.7568 −0.33
3 54 59 5 −5 1.7081 −3.10
4 128 137 9 25 1.8137 +2.89
5 250 259 9 −7 1.7419 −1.18
6 432 411 −21 −5 1.7687 +0.34
7 686 701 15 21 1.7719 +0.52
8 1024 965 −59 5 1.7527 −0.57
9 1458 1459 1 19 1.7711 +0.48
10 2000 1989 −11 53 1.7612 −0.08
11 2662 2685 23 29 1.7604 −0.13
12 3456 3479 23 7 1.7669 +0.24
13 4394 4477 83 −67 1.7577 −0.28
14 5488 5577 89 25 1.7636 +0.05
15 6750 6735 −15 79 1.7633 +0.03
Exact = 1.76267

Table 4. Predicted values of the Madelung constant for the fluorite structure.
Columns 3 to 6 show the deviations in number of ions in equivalent spheres centered
on Ca2+ and F− ions, and the net charges within that spheres.

Nsphere − Ncell Sphere net charge Madelung constant


L/a Ncell
Ca2+ F− Ca2+ F− This work ∆, %
2 12 −3 −1 −6 1 4.8351 −4.06
4 96 3 13 30 −25 5.1141 +1.47
6 324 −29 −17 −34 5 5.0442 +0.09
8 768 −39 −43 18 −5 5.0223 −0.35
10 1500 −1 5 94 −53 5.0424 +0.05
12 2592 31 19 38 −7 5.0531 +0.26
14 4116 93 61 18 23 5.0454 +0.11
16 6144 −19 −61 154 −35 5.0370 −0.06
18 8748 −17 −89 −34 89 5.0379 −0.04
20 12000 21 37 −30 −1 5.0422 +0.05
22 15972 5 83 298 −227 5.0407 +0.02
Exact = 5.0398
Computation of Coulomb Forces in Disordered Systems 643

REFERENCES
1. R.W. Hockney and J.W. Eastwood, Computer Simulation Using Particles,
McGraw-Hill (1981).
2. R. Car and M. Parrinello, Phys. Rev. Lett. 55, 2471 (1985).
3. See, e.g., S.A. Bonev, B. Militzer, and G. Galli, Phys. Rev. B 69, 01 4101
(2004), and references therein.
4. E. Madelung, Phys. Z. 19, 524 (1918).
5. P.P. Ewald, Ann. Phys.(Leipzig) 64, 253 (1921).
6. G.J. Martyna and M.E. Tuckerman, J. Chem. Phys. 110, 2810 (1999);
Z.-H. Duan and R. Krasny, J. Chem. Phys. 113, 3492 (2000).
7. E. Yakub and C. Ronchi, J. Chem. Phys. 119, 11 556 (2003).
8. H.-Q. Ding, N. Karasawa, and W.A. Goddard, J. Chem. Phys. 97, 4309 (1992).
9. F. Gygi, Ab-initio Molecular Dynamics Code GP 1.20.0, Lawrence Livermore
National Laboratory (1999–2003).
10. W. Andreoni and A. Curioni, Parallel Computing 26, 819 (2000).
11. D.C. Rappaport, The Art of Molecular Dynamics Simulation, Cambridge
University Press (1995).
12. E.M. Lifshitz and L.P. Pitaevskii, Electrodynamics of Continuous Media.
Course of Theoretical Physics Vol. 8. 2. ed., Butterworth-Heinemann,
Oxford... (2000).
13. J.P. Hansen, Phys. Rev. A 8, 3110 (1973).
14. G.S. Stringfellow, H.E. DeWitt, and W.L. Slattery, Phys. Rev. A 41,
1105 (1990).
15. B. Larsen, Chem. Phys. Lett. 27, 47 (1974).
16. E.S. Yakub, J. Low Temp. Phys. 122, 559 (2001).

View publication stats

You might also like