Sunteți pe pagina 1din 8

THE JOURNAL OF CHEMICAL PHYSICS 125, 224107 共2006兲

Electrostatic interactions in dissipative particle dynamics


using the Ewald sums
Minerva González-Melchora兲
Instituto de Física, Universidad Autónoma de Puebla, Apartado Postal J-48, 72570 Puebla, Mexico
Estela Mayoral and María Eugenia Velázquez
Centro de Investigación en Polímeros, grupo COMEX, Marcos Achar Lobatón No. 2, Tepexpan,
55885 Acolman, Estado de México, Mexico
José Alejandreb兲
Departamento de Química, Universidad Autónoma Metropolitana-Iztapalapa, Avenida San Rafael Atlixco
186, Colonia Vicentina, 09340 México DF, Mexico
共Received 7 August 2006; accepted 27 October 2006; published online 14 December 2006兲

The electrostatic interactions in dissipative particle dynamics 共DPD兲 simulations are calculated
using the standard Ewald 关Ann. Phys. 64, 253 共1921兲兴 sum method. Charge distributions on DPD
particles are included to prevent artificial ionic pair formation. This proposal is an alternative
method to that introduced recently by Groot 关J. Chem. Phys. 118, 11265 共2003兲兴 where the
electrostatic field was solved locally on a lattice. The Ewald method is applied to study a bulk
electrolyte and polyelectrolyte-surfactant solutions. The structure of the fluid is analyzed through the
radial distribution function between charged particles. The results are in good agreement with those
reported by Groot for the same systems. We also calculated the radius of gyration of a
polyelectrolyte in salt solution as a function of solution pH and degree of ionization of the chain.
The radius of gyration increases with the net charge of the polymer in agreement with the trend
found in static light scattering experiments of polystyrene sulfonate solutions. © 2006 American
Institute of Physics. 关DOI: 10.1063/1.2400223兴

I. INTRODUCTION Electrostatic interactions were recently included in DPD


by Groot.10 He proposed a method where the electrostatic
The dissipative particle dynamics 共DPD兲 simulation field is solved locally on a grid. The method was applied to
method1 was developed to accomplish the task of reaching study an electrolyte and a polyelectrolyte-surfactant solution
larger lengths and longer time scales than atomistic molecu- among other systems. He stated that the method was reason-
lar dynamics 共MD兲 simulations, resulting in a mesoscopic ably efficient, can treat local inhomogeneities in dielectric
description of the system.2,3 This approach has been success- permittivity, and it captures the most important features of
fully applied to study a wide variety of complex fluids in- electrostatic interactions.
cluding polymeric solutions, colloidal suspensions, surfac- On the other hand, the Ewald sum method11 is the most
tants, and biological membranes,4,5 where different lengths employed route to calculate electrostatic interactions in mi-
and time scales are involved. The DPD method was origi- croscopic molecular simulations. It is widely accepted that it
nally proposed to study repulsive soft interactions but it has is the best technique to treat Coulombic forces and extensive
been modified to include multibody effects6,7 which allow effort has been done to improve its efficiency.12–14 The
the inclusion of attractive interactions to simulate vapor- Ewald sum method can also be used in combination with
liquid equilibrium.8 We have recently shown that finite size techniques to include polarization, see Refs. 15 and 16 and
effects in DPD simulations9 are less important than in micro- references therein. The main problem in using the Ewald
scopic simulations, hence it is possible to simulate systems method in DPD simulations is that electrostatic interactions
using a small number of particles. between DPD particles are soft and atoms with opposite
For all the complex systems mentioned above, the elec- charge form artificial clusters of ions. The main goal of this
trostatic interactions play a key role in understanding phe- work is to show that the standard Ewald sum method com-
nomena that do not occur in noncharged systems. So the bined with charge distributions on particles to avoid the for-
inclusion of these interactions in DPD simulations is essen- mation of nondesirable ionic pairs can be used to calculate
tial to capture phenomena at mesoscopic level such as for- the electrostatic interactions in mesoscopic simulations. The
mation of polyelectrolyte-surfactant aggregates, charge stabi- method proposed in this work is easily implemented in stan-
lization of colloidal suspensions, and the formation of dard DPD programs. The inclusion of charge distributions
complexes driven by charged species in biological systems. does not increase the computational cost in the Ewald sum
procedure. However, the method might be computationally
a兲
Electronic mail: minerva@sirio.ifuap.buap.mx more demanding than that used by Groot.10 The efficiency
b兲
Electronic mail: jra@xanum.uam.mx can be improved by using techniques that allow fast calcula-

0021-9606/2006/125共22兲/224107/9/$23.00 125, 224107-1 © 2006 American Institute of Physics

Downloaded 08 Feb 2007 to 148.228.150.68. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp
224107-2 González-Melchor et al. J. Chem. Phys. 125, 224107 共2006兲

tion of the reciprocal part such as the smooth particle mesh with
Ewald.12 By using charge distributions on DPD particles it is
possible to use a polarization version of the Ewald sums to
simulate systems having regions with different dielectric
␻共r兲 = 再 1 − r/Rc , r 艋 Rc
0, r ⬎ Rc ,
冎 共4兲

constants. The proposed method was applied to study a bulk being Rc the cutoff distance. The intramolecular interaction
electrolyte and an aqueous solution of polyelectrolyte- in the polyelectrolyte and surfactant molecules, which have
surfactant mixture. These systems were studied by Groot and more than one particle in their structure, is given by har-
we make the corresponding comparison to validate our re- monic forces, so that if atoms i and j are bonded,
sults. In addition, the method was applied to calculate struc-
tural properties of a polyelectrolyte in a salt solution as a FBij = − K共r − r0兲rij/r, 共5兲
function of the net charge fraction on the polymer and solu- where K is the spring constant and r0 is the equilibrium bond
tion pH. distance.
The rest of this work is organized as follows: Section II The dissipative and random forces are related through
contains the description of the DPD method. Our proposal to the fluctuation-dissipation theorem 共␴2 = 2␥kBT兲, generating
treat electrostatic interactions is presented in Sec. III. The naturally the canonical distribution 共constant number of par-
comparison with Groot’s results on structural properties of ticles, N, volume, V, and temperature, T兲. The forces FD ij and
ionic systems is contained in Sec. IV. Section V gives the FRij act as an in-built thermostat. Here kB is the Boltzmann’s
end-to-end distance and radius of gyration of a polyelectro- constant. In this standard DPD formulation the conservative
lyte in solutions with salt added as a function of the charge forces consist of a short-range repulsive interaction modeling
fraction on the polymer and as a function of solution pH. the soft nature of neutral DPD particles. Electrostatic inter-
Finally, main conclusions are given. actions are conservative long-range forces that must be taken
into account to capture electrostatic phenomena. Once Cou-
lombic interactions are included, the total conservative
II. THE DPD APPROACH forces will determine the thermodynamic behavior of the
system. The next section describes our proposal to calculate
The DPD method was introduced by Hoogerbrugge and electrostatic interactions in DPD.
Koelman.1 The main conceptual difference with atomistic
dynamics is the coarse-graining procedure that allows the
mapping of several molecules from the real system into one III. THE EWALD SUMS AND CHARGE DISTRIBUTIONS
DPD particle. In this section we give a general description of
the method; more details are found in Ref. 3. In DPD simu- The method proposed in this work consists of a combi-
lations a set of interacting particles is considered and the nation of charge distributions on DPD particles and the
total force, Fij, between any pair of particles i and j is the Ewald sum technique, widely used to calculate electrostatic
sum of a conservative FCij , a dissipative FD R
ij , and a random Fij forces.
force, A set of N DPD particles, every one carrying a point
charge qi at positions ri in a cubic cell of side L and volume
Fij = FCij + FD R
ij + Fij . 共1兲 V = L3, is considered. Overall charge neutrality is assumed.
Charges interact according to Coulomb’s law, and the total
These forces are given by electrostatic energy for the periodic system is given by
1 qq
FCij = aij␻C共r兲êij , U共rN兲 = 兺兺兺 i j ,
4␲␧0␧r i j⬎i r 兩rij + nL兩
共6兲

where n = 共nx , ny , nz兲, nx , ny , nz are integer numbers and the


ij = ␥␻ 共r兲关êij · vij兴êij ,
FD D
prime means that terms with i = j are omitted when n = 0. The
variables ␧0 and ␧r are dielectric constants of vacuum and
FRij = ␴␻R共r兲êij␰ij , 共2兲 water at room temperature, respectively. The sum over n
takes into account the periodic images. In the Ewald treat-
ment the long-range electrostatic energy, Eq. 共6兲, is decom-
where êij = rij / r, rij = ri − r j, vij = vi − v j, r is the distance be-
posed in a real space and the reciprocal space
tween particles i and j and ri and vi are the position and
contributions.11,17 In this way, the real and the reciprocal
velocity of particle i, respectively. The random force is cal-
parts are both short-ranged sums written as

冋兺 兺
culated using a random number ␰ij which is uniformly dis-
tributed between 0 and 1 with Gaussian distribution, zero 1 erfc共␣r兲
mean, and unit variance. The constants aij, ␥, and ␴ deter- U共rN兲 = q iq j
4 ␲ ␧ 0␧ r r


i j⬎i
mine the strength of the conservative, dissipative, and ran-
dom forces, respectively. The weight functions ␻’s are given 2␲


N

by + 兺 Q共k兲S共k兲S共− k兲 − 冑␲ 兺i q2i ,
V k⫽0
共7兲

␻C共r兲 = ␻R共r兲 = 冑␻D共r兲 = ␻共r兲, 共3兲 with

Downloaded 08 Feb 2007 to 148.228.150.68. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp
224107-3 Electrostatic interactions in dissipative particle dynamics J. Chem. Phys. 125, 224107 共2006兲

2
e−k /4␣
2 N
2␲ inside a DPD particle, Vm = 18 cm3 mol−1 is the molar vol-
Q共k兲 = , S共k兲 = 兺 qie ik·ri
, k= 共mx,my,mz兲, ume of water, and NA is Avogadro’s number. Therefore, Rc
k2 L
i=1 = 4.48Nm1/3
Å and ⌫ = 20.08Nm −1/3
. Groot used Nm = 3, hence
共8兲 Rc = 6.46 Å and ⌫ = 13.87. Taking Nm = 3 and aii = 78.33 be-
tween solvent particles, the compressibility of pure water at
where ␣ is the parameter that controls the contribution in the room temperature is reproduced.21 In systems with salt, the
real space, k is the magnitude of the reciprocal vector k, molar concentration of salt was calculated10 using creal
whereas mx, my, and mz are integer numbers. Equation 共7兲 is = 共NNaCl / V*兲 / 共R3c NA兲 where NNaCl is the number of salt mol-
a good approach to 1 / r given by Eq. 共6兲, capturing the full ecules.
long-range nature of electrostatic interactions. The electro- The magnitude of the reduced force between two charge
static interactions between two DPD particles are calculated distributions is
whether the particles are bonded or not.
In DPD methodology, the conservative force appearing 4␲F*e Z iZ j * *
in Eq. 共2兲 is mathematically well defined at r = 0, allowing
ij
= *2 兵1 − e−2␤ r 关1 + 2␤*r*共1 + ␤*r*兲兴其. 共12兲
⌫ r
full overlap between particles. However, the electrostatic
contribution diverges at r = 0, leading to the formation of Equations 共11兲 and 共12兲 contain the Coulombic term which
artificial ionic pairs. To avoid this problem, Groot10 used a can be calculated in a simulation using the standard Ewald
charge distribution given by scheme. By including the charge distribution, the divergency
of the Coulomb interactions at r* = 0 was removed. In the
3 limit r* → 0, the energy and the force between two charged
␳共r兲 = 共1 − r/Re兲, 共9兲
␲R3e distributions are finite quantities, they are given by
for r ⬍ Re, where Re is a smearing radius. For r ⬎ Re, ␳共r兲 4␲u*共r兲
= 0. To calculate the electrostatic forces, Groot followed the lim = Z iZ j ␤ * 共13兲
r→0 ⌫
method of Beckers et al.,18 where the electrostatic field is
solved on a lattice. The charge distributions were spread out and
over the lattice nodes and the long-range part of the interac-
4␲F*e
ij
tion potential was calculated by solving the Poisson equation lim = 0, 共14兲
in real space. It was stated that the method works efficiently r*→0 ⌫
if the grid size is equal to the particle size. As the grid size is
respectively. In this work ␤* was chosen to match the inter-
reduced there is an improvement on the accuracy of the elec-
action between two charge clouds at r* = 0 proposed by
trostatic forces but the CPU time increases.
Groot,10 whose expression reduces in this limit to
In this work, in order to remove the divergency at r = 0
we consider a charge distribution on DPD particles of the 4␲u*Groot共r*兲 52 ZiZ j
form lim = . 共15兲
r*→0 ⌫ 35 R*e
q −2r/␭
␳共r兲 = e , 共10兲 Comparing Eqs. 共13兲 and 共15兲 we obtain
␲␭3
52
where ␭ is the decay length of the charge. This function is a ␤* = ⬇ 0.929. 共16兲
Slater-type charge density appearing in the context of quan- 35R*e
tum mechanical calculations. When Eq. 共10兲 is integrated Once ␤* is obtained the potential energy and forces are cal-
over the space the total charge is q. culated using Eqs. 共11兲 and 共12兲, respectively. The Coulom-
The functional form of the charge density might be com- bic term is then evaluated using the Ewald sums. Figure 1
pletely arbitrary but the forces cannot in general be calcu- shows the interaction between charged particles with the
lated analytically; however, for a distribution given by Eq. same sign as calculated in this work. The electrostatic inter-
共10兲, good approximated expressions for the force are actions used by Groot10 were slightly different and they were
available.19,20 In what follows, reduced units are employed to calculated from a lattice method in a real space. The theoret-
compare with Groot’s results. To make the length, energy, ical curve corresponding to the charge distributions studied
and mass dimensionless, we used Rc, kBT, and particle mass, by Groot are also shown in Fig. 1. The maximum value of
respectively. In what follows R*c = 共kBT兲* = m* = 1. The re- the force in both models is about the same, but the force used
duced interaction potential between two charged distribu- in this work is shifted to smaller distances making the repul-
tions separated by a distance r* from center to center is given sive interactions larger than in Groot’s model. An opposite
by19 behavior is observed for the attractive forces.
4␲u*共r*兲 ZiZ j * *
The total conservative force is thus given by adding F*eij
= * 关1 − 共1 + ␤*r*兲e−2␤ r 兴, 共11兲 from Eq. 共12兲 to FCij in Eqs. 共2兲. These forces will determine
⌫ r
the thermodynamic behavior of the system.
where Zi is the valence of ion i, e is the electron charge, ⌫ According to Ruelle’s test22 a system has a well defined
= e2 / 共kBT⑀0⑀rRc兲, and ␤* = Rc / ␭. The value Rc is obtained thermodynamic behavior if the total interaction energy satis-
from Rc = 共␳*NmVm / NA兲1/3 where ␳* is the reduced density of fies two requirements: first, the interaction between distant
DPD particles, Nm is the number of real water molecules particles must be negligible and second, it must satisfy the

Downloaded 08 Feb 2007 to 148.228.150.68. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp
224107-4 González-Melchor et al. J. Chem. Phys. 125, 224107 共2006兲

FIG. 2. Comparison between RDFs obtained by Groot 共Ref. 10兲 and the
FIG. 1. Electrostatic potential 共upper curves兲 and force 共lower curves兲 be- method proposed here. The repulsive parameters were set to aij = 25.0 and
tween two equal-sign charge distributions. The theoretical electrostatic in- Rc = 6.46 Å.
teraction and the force used by Groot 共Ref. 10兲 are shown with continuous
lines. The open circles represent the electrostatic interaction and the force
employed in this work, given by Eqs. 共11兲 and 共12兲. The vertical dashed line
is the distance Rcreal / Re = 1.875 at which the real part was truncated. Rcreal was represent the counterions. The total number density is ␳*
taken equal to Rcreal = 3.0Rc and Re = 1.6Rc. The dashed line is the Coulombic = 3.0. Mapping these quantities to real units, the system cor-
1 / r potential, which diverges at r = 0. responds to a salt concentration of 0.6M. The interaction
parameters for the conservative, dissipative, and random
so-called stability definition.22 The latter establishes the fact forces were aij = 25.0, ␭ = 4.5, and ␴ = 3.0, respectively. The
that the interaction must not cause the collapse of an infinite same value of aij was used for the interaction between any
number of particles into a bounded region. If a given poten- two DPD particles. We note that aij = 25.0 corresponds to
tial does not satisfy this condition it is called a catastrophic Nm = 1 instead of Nm = 3, but that was the value that Groot
potential. Different interactions have been analyzed in terms used in his work.
of its stability such as the Gaussian core model23 and the A reduced time step of ⌬t* = ⌬t共kBT / mR2c 兲1/2 = 0.02 was
electrostatic interaction between charge distributions. This used during the simulation. The real forces in the Ewald
latter was analyzed by Fisher and Ruelle.24 They concluded sums were truncated at Rreal c = 3.0Rc with ␣ = 0.15 Å . For
−1

that a system with charge density on particles, rather than the reciprocal part we considered the sum with a maximum
point charges, is stable. Furthermore, as stated in Refs. 22 vector kmax = 共5 , 5 , 5兲. The value of ␤* = 0.929, obtained in
and 24, if a positive pair interaction potential is added to a Eq. 共16兲, was used in all simulations. The average tempera-
stable one, the total interaction in the system remains stable. ture in all simulations is the same as the imposed external
Here it follows that the total potential energy used in this value.
work should give the right thermodynamic behavior. The structure of solvent and ions was determined by the
radial distribution functions 共RDFs兲. The results from this
IV. COMPARISON WITH GROOT’S RESULTS work and those obtained by Groot10 are shown in Fig. 2. A
general good agreement is observed. The first point to look at
To validate the method of calculating the electrostatic
is that there is no ionic formation at distances close to r = 0.
interactions proposed in this work, a comparison with
The structure between ions carrying the same charge is
Groot’s results was made for two systems: 共A兲 bulk electro-
slightly lower than those obtained by Groot because the in-
lyte and 共B兲 aqueous solution of polyelectrolyte-surfactant
teraction model used in this work is more repulsive at short
mixture. The DPD particles were moved using the modified
distances. One would expect that the maximum in the RDF
version of the velocity Verlet algorithm, DPD-VV.25
of the particles with opposite charge moved to shorter dis-
tances and had a higher value respect to Groot’s results.
A. Bulk electrolyte
However, the RDF is almost the same at short distances, this
As a first case of study, we present the simple electrolyte means that DPD repulsive interactions compensate the at-
studied by Groot.10 We used his same set of parameters to traction of ions with opposite charge. It is seen that water is
make a direct comparison with his results. Electrostatic in- slightly less structured than in Groot’s work.
teractions were calculated following the method described in In Ref. 10 the reduced quantities to calculate the electro-
Sec. III. The system consists of a total number of N = 3000 static interactions were obtained using Nm = 3 but he used the
DPD particles in a simulation cell of volume V* = 10⫻ 10 repulsion parameter aij = 25 which corresponds to Nm = 1.
⫻ 10. The solvent mimicking water at room temperature has Those values, aij = 25 and Nm = 3, do not change the conclu-
2804 DPD particles while 98 particles represent ions with net sions drawn in his work. The radial distribution functions
charge e and the same number, carrying a net charge of −e, obtained for the bulk electrolyte using aij = 78.3 共Nm = 3兲 are

Downloaded 08 Feb 2007 to 148.228.150.68. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp
224107-5 Electrostatic interactions in dissipative particle dynamics J. Chem. Phys. 125, 224107 共2006兲

FIG. 4. Comparison between RDFs obtained by Groot 共Ref. 10兲 and the
method proposed here. The system consists of a polyelectrolyte in solution,
FIG. 3. Comparison between RDFs for two different values of the repulsion surfactant molecules, and the corresponding counterions 共system B兲. The
parameter: aij = 25.0, shown with dotted, continuous and dashed lines; and symbols are read as follows: intramolecular polymer-polymer 共circles兲,
aij = 78.33 shown with the symbols displayed on the figure. polymer-head group 共squares兲, and head group-head group 共diamonds兲. The
intramolecular polymer-polymer RDF for the system with surfactant re-
placed by salt is shown with triangles. For comparison, the data obtained by
shown in Fig. 3 and they are compared with results from Groot 共Ref. 10兲 for the same systems are plotted with solid 共polyelectrolyte-
aij = 25. The position of the first maximum is around the surfactant-solvent兲 and dotted lines 共polyelectrolyte-salt-solvent兲.
same value for both coarse graining levels. Although the
density is the same, the fluid is more structured when Nm
aij = aii + ␹ij/0.306, 共17兲
= 3 because the interaction between particles is more repul-
sive. where ␹ij is the Flory-Huggins interaction parameter. The
parameters aij are given in Table I for completeness. The
time step used was ⌬t* = 0.04 and after equilibration the av-
B. Aqueous solution of polyelectrolyte-surfactant
erage properties were obtained from 2 ⫻ 105 time steps of
mixture
production time. It is convenient to mention again that aii
The system consists of a total number of N = 10 125 DPD = 25 for the solvent used by Groot corresponds to Nm = 1.
particles. The polymer 共50 monomers兲 corresponds to a mo- Figure 4 shows selected RDFs for a polyelectrolyte in
lecular weight M w = 8000; water 共9825 particles兲 and surfac- the presence of surfactant molecules. The agreement between
tant molecules 共75 particles兲 were allocated in a cubic box of results from this work and those from Ref. 10 is excellent for
volume V* = 15⫻ 15⫻ 15. The interaction parameters were polymer-polymer 共pp兲 and polymer-head 共ph兲 interactions.
the same as those used by Groot. Each monomer on the The RDF from this work for the head-head 共hh兲 particles is
polymer carries a net charge of e / 2. Anionic surfactant mol- systematically lower than those reported by Groot due to
ecules were modeled as head-tail dimers. The head groups higher repulsion between those particles at short distances.
carry a net charge −e and tails are represented by neutral The effect is magnified at large distances. In this work the
particles. To preserve charge neutrality in the system 25 electrostatic interactions do not include polarization effects
polymer counterions of charge −e and 75 surfactant ions of in contrast with the results reported in Ref. 10 where water
net charge e were added. A harmonic force between bonded and head groups carried out large polarizabilities. That is
particles in the polyelectrolyte and surfactant molecules was another difference in our model that might explain the dif-
used. The spring constant was K = 4.0 and the bond equilib- ferences found for the head-head RDF. The RDF for poly-
rium distance was zero. The repulsive interaction parameters electrolyte particles without surfactant is also in excellent
aij for the cross interactions in a system with ␳* = 3 were agreement with Groot’s data. The slopes on RDF between
obtained10 through 1 nm ⬍r ⬍ 2.8 nm obtained in this work for the polymer-

TABLE I. Repulsive interaction parameters between DPD particles for the polyelectrolyte-surfactant system.

aij Solvent Tail Head Monomer Surfactant ion Counterion

Solvent 25.0 83.8 25.0 27.1 25.0 25.0


Tail 83.8 25.0 44.6 44.6 25.0 25.0
Head 25.0 44.6 25.0 25.0 25.0 25.0
Monomer 27.1 44.6 25.0 25.0 25.0 25.0
Surfactant ion 25.0 25.0 25.0 25.0 25.0 25.0
Counterion 25.0 25.0 25.0 25.0 25.0 25.0

Downloaded 08 Feb 2007 to 148.228.150.68. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp
224107-6 González-Melchor et al. J. Chem. Phys. 125, 224107 共2006兲

polymer pairs were −1.3 and −1.9 for systems with and with-
out surfactant, respectively. These values compared well
with those reported by Groot, −1.18 and −1.90, respectively.

V. RESULTS OF POLYELECTROLYTE IN SOLUTION


WITH SALT ADDED

The main results of this section are the end-to-end dis-


tance, 具Ree 典 , and the radius of gyration, 具R2g典1/2, of a poly-
2 1/2

electrolyte in solution as a function of dissociation degree.


The radius of gyration is a property that has been widely
studied, both experimentally26,27 and theoretically,28,29 to cor-
relate changes in polymer conformation with different physi-
cochemical and biological properties. It is known that in a
polyelectrolyte, 具R2g典1/2 depends on the solution pH because
the net charge on the polymer changes with it. 具R2g典1/2 is also
dependent on counterion concentration, for this reason most
of studies are carried out with a fixed amount of salt added or FIG. 5. Typical conformation of a neutral polyelectrolyte in a salt solution
constant ionic strength. of 0.14M.

A. End-to-end distance and radius of gyration Conformational properties of the polyelectrolyte as a


function of the fraction charge ␪ are summarized in Table II.
In this section we present DPD simulation results of a The end-to-end distance and 具R2g典1/2 for this linear chain in-
single polyelectrolyte in water as function of the polymer
crease monotonically with the degree of ionization. Experi-
charge in the presence of salt. Five charge fractions, defined
mentally it is well known29 that a neutral polymer in solution
by ␪ = Nq / N p, were considered. Here Nq is the number of
with salt has the smallest radius of gyration compared either
charged beads on the chain and N p is the total number of
with a partially or a fully ionized chain. Figures 5 and 6
beads conforming the polymer. Going from a neutral to a
show instantaneous snapshots of a neutral and a fully ionized
fully ionized polymer, ␪ takes the values 0, 0.24, 0.48, 0.76,
polyelectrolyte in solution with salt. It is seen that ionization
and 1.0, containing Nq = 0, 12, 24, 38, and 50 charged beads,
elongates the chain in agreement with experimental findings.
respectively. The charge distributions of magnitude e / 2 were
The value of 具Ree 典 and 具R2g典1/2 for the fully ionized chain is
2 1/2
assigned uniformly to different beads on the chain. Counter
1.7 and 1.5 times larger than that found for the the neutral
ions of charge density −e were added to preserve charge
polymer, respectively. These results are evidence that the
neutrality. All systems contained 75 salt molecules which
method used in this work for calculating the electrostatic
give 75 cations 共net charge e兲 and 75 anions 共net charge −e兲.
forces gives the right trend on 具R2g典1/2 for a charged polymer.
The total number of beads in the system was N = 10 125 in all
For a 1:1 electrolyte at room temperature the Debye
cases.
Each simulation involved N p = 50 polymer beads in a
simulation cell of volume V* = 15⫻ 15⫻ 15 in the presence
of solvent, salt, and counterions. The solution corresponds to
a salt concentration of 0.14M. The total density was fixed at
␳* = 3. The equilibrium bond distance r0 was set to zero and
the spring constant was K = 4.0.
The repulsive interaction between beads was obtained
trough aij = aii + ␹ij / 0.306, where aii = 78.3. The parameter ␹ij
for polymer and water was ␹ pw = 0.65, taken from Ref. 10. In
this way, for polymer and water apw = 80.42 and aij = 78.3 for
the other species.

TABLE II. Reduced end-to-end distance and radius of gyration as function


of the fraction of charge ␪ going from a neutral to a fully ionized polymer.

␪ 具Ree典
2 1/2
具R2g典1/2

0.00 6.5± 1.0 2.9± 0.2


0.24 7.2± 0.7 3.1± 0.2
0.48 8.3± 1.0 3.5± 0.2
0.76 9.6± 1.7 3.9± 0.4
1.00 10.9± 1.7 4.4± 0.4 FIG. 6. Typical conformation of a fully ionized polyelectrolyte in a salt
solution of 0.14M.

Downloaded 08 Feb 2007 to 148.228.150.68. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp
224107-7 Electrostatic interactions in dissipative particle dynamics J. Chem. Phys. 125, 224107 共2006兲

screening length is given by lD = 3.045C−1/2 Å, where C is


the molar concentration. For systems simulated in Secs. IV A
and V A, the Debye screening lengths are 3.9 and 8.1 Å,
respectively. At these concentrations the Debye-Hückel
theory predicts that the interactions between charged par-
ticles separated further than 3.9 and 8.1 Å become screened
and the interaction is no longer long range. However, the
Debye-Hückel theory is valid for very diluted electrolytes
and is not expected to be valid for highly concentrated sys-
tems. A more recent generalized Debye-Hückel theory has
been developed where the screening length has become spa-
tially dependent, in this way a local effect of the charges30,31
is introduced. The method presented here has the advantage FIG. 7. Scheme of the polyelectrolyte used in this work, A can take the
that any salt concentration can be studied. The system stud- forms A− 共ionized兲 and A0 共neutral兲.
ied in Sec. IV A was chosen for comparison purposes while
that in Sec. V A was chosen as a simple polymer model at a
the polymer, at a constant amount of salt. Experimentally, the
moderate concentration. It is very likely that a smaller cut-off
net charge of the polyelectrolyte can be fixed by changing
radius for the real forces could have been used but details
the system’s pH. We can relate the pH with the fraction of
concerning this sort of optimization were not studied in this
net charge by
work, nevertheless they are necessary to justify the choice of
parameter values used in the Ewald method.
An important issue is the range of concentrations where
explicit salt ions can be replaced by a screened interaction
pH = log10 冉 冊 ␪
1−␪
+ pKa , 共18兲

potential between particles. This topic is beyond the scope of where pKa is the acidity constant; in this work pKa was
the present work. It deserves special attention and certainly is chosen to be 5.5 which is a typical value of a weak polyelec-
a study that will be undertaken. trolyte. This equation was obtained considering that initially
Concerning the decay of correlations in systems with the polymer is fully deprotonated. As H+ ions are added, the
electrostatic interactions, some simulation works and several polyelectrolyte starts to protonate reaching an equilibrium
theoretical calculations have been done in terms of the pa- where there is a balance of dissociated species.
rameter x defined by x = ␬a, where ␬ is the inverse Debye In our model the polyelectrolyte is a branched chain that
screening length ␬ = 1 / lD and a is the diameter of the contains two different groups, one of this is neutral and the
particle.32,33 For a certain value of x, called the Kirkwood other one is a monoprotic acid. This molecule can be mapped
parameter, xK, the behavior of the structural correlation func- into 12 DPD beads, as shown in Fig. 7. As the pH is modi-
tion crosses from a simple screened monotonic decay for x fied, the four A0 beads remain neutral while the other eight
⬍ xK, to an oscillatory decay for x ⬎ xK. We have not ana- 共A in Fig. 7兲 can be A = A0 or A = A− if they are ionized
lyzed the decay behavior of correlations in the DPD fluids according to Eq. 共18兲. So for a given value of pH we fix a
presented here. Focus on this issue requires a more detailed charge distribution of net charge −e on the DPD polymer
analysis and is left for future work. beads which are randomly chosen and that charge remains
fixed during the simulation.
We consider that each DPD particle contains three real
B. Effect of solution pH on radius of gyration
water molecules, Nm = 3. The reduced density was ␳* = 3 in a
In addition, we studied the behavior of the radius of cubic box of reduced box length of 8.5. The repulsive inter-
gyration of a weak monoprotic polyacid as a function of pH action parameters34 between two particles were aij = 78.33
and fraction of net charge ␪ 共the number of charged mono- except for A0-water, A0-CI+, and A0-A− where aij = 79.33,
mers divided by the total amount of ionizable monomers兲 on respectively. The number of particles of each species used in

TABLE III. Number of particles used in simulations, A− = A ionized, A0 = A neutral, I* = polymer counterion.
The salt concentration is 0.01M, a typical experimental value.

␪ pH Na+ Cl− I* A− A0 H 2O

0.0 1.00 1 1 0 0 12 1816


0.125 4.62 1 1 1 1 11 1816
0.25 4.99 1 1 2 2 10 1816
0.375 5.25 1 1 3 3 9 1816
0.5 5.47 1 1 4 4 8 1816
0.625 5.69 1 1 5 5 7 1816
0.75 5.95 1 1 6 6 6 1816
0.875 6.32 1 1 7 7 5 1816
1.0 14.00 1 1 8 8 4 1816

Downloaded 08 Feb 2007 to 148.228.150.68. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp
224107-8 González-Melchor et al. J. Chem. Phys. 125, 224107 共2006兲

FIG. 8. 具R2g典1/2 as a function of net charge fraction on


the polymer 共left side兲 and as a function of pH 共right
side兲.

our simulations, is listed in Table III. After equilibration, the thanks R. D. Groot for helpful correspondence on this topic.
averages were obtained over 350 000 time steps. When beads The authors also thank T. A. Darden and A. Gama for useful
are bonded the equilibrium bond distance r0 was zero and the discussions.
spring constant was K = 100. The reduced time step was
⌬t* = 0.04.
1
P. J. Hoogerbrugge and J. M. V. A. Koelman, Europhys. Lett. 19, 155
共1992兲.
Figure 8 on the left-hand side shows results for 具R2g典1/2 as 2
P. Español and P. B. Warren, Europhys. Lett. 30, 191 共1995兲.
a function of ␪ from simulations of this work. The trend 3
R. D. Groot and P. B. Warren, J. Chem. Phys. 107, 4423 共1997兲.
4
agrees with the experimental results for a weak R. D. Groot, Langmuir 16, 7493 共2003兲.
5
polyelectrolyte.29 On the right-hand side of Fig. 8 are also T. Murtola, E. Falck, M. Patra, M. Karttunen, and I. Vattulainen, J. Chem.
Phys. 121, 9156 共2004兲.
shown the results obtained for radius of gyration as a func- 6
S. Y. Trofimov, E. L. F. Nies, and M. A. J. Michels, J. Chem. Phys. 117,
tion of pH. We can see an inflexion point around the pH 9383 共2002兲.
7
value of the polyelectrolyte’s pKa, which agrees qualitatively I. Pagonabarraga and D. Frenkel, J. Chem. Phys. 115, 5015 共2001兲.
8
with observed experimental results.26 It is important to point P. B. Warren, Phys. Rev. E 68, 066702 共2003兲.
9
M. E. Velázquez, A. Gama-Goicochea, M. González-Melchor, M. Neria,
out that the pH values of 1 and 14 correspond to the fully and J. Alejandre, J. Chem. Phys. 124, 084104 共2006兲.
protonated and fully ionized polyelectrolyte structures, re- 10
R. D. Groot, J. Chem. Phys. 118, 11265 共2003兲.
11
spectively. P. P. Ewald, Ann. Phys. 64, 253 共1921兲.
12
U. Essmann, L. Perera, M. L. Berkowitz, T. Darden, H. Lee, and L. G.
Pedersen, J. Chem. Phys. 103, 8577 共1995兲.
VI. CONCLUSIONS 13
M. Deserno and C. Holm, J. Chem. Phys. 109, 7678 共1998兲.
14
C. Sagui and T. Darden, J. Chem. Phys. 114, 6578 共2001兲.
An alternative method to Groot’s10 implementation to 15
C. Haibo Yu, T. Hansson, and W. F. van Gunsteren, J. Chem. Phys. 118,
treat electrostatic interactions in DPD simulations is pro- 221 共2003兲.
16
posed. Charge distributions are assigned to charged DPD A. Toukmaji, C. Sagui, J. Board, and T. Darden, J. Chem. Phys. 113,
10913 共2000兲.
particles to avoid the formation of artificial ion pairs. The 17
D. Frenkel and B. Smit, Understanding Molecular Simulations, From
long-range term 1 / r is calculated via the standard Ewald sum Algorithms to Applications 共Academic, New York, 1996兲.
18
scheme. The proposed method makes use of conservative J. V. L. Beckers, C. P. Lowe, and S. W. de Leeuw, Mol. Simul. 20, 369
interactions with a well defined stable thermodynamic be- 共1998兲.
19
M. Carrillo-Tripp, Ph.D. thesis, Universidad Autónoma del Estado de
havior and it is easily implemented in DPD simulation pro- Morelos, 2005.
grams. It can also be combined with fast Fourier transform 20
H. Saint-Martin, J. Hernández-Cobos, M. I. Bernal-Uruchurtu, I. Ortega-
techniques to improve the calculation of the reciprocal space Blake, and H. J. Berendsen, J. Chem. Phys. 113, 10899 共2000兲.
21
part. Electrostatic interactions with different dielectric con- B. Hafskjold, C. Liew, and W. Shinoda, Mol. Simul. 30, 879 共2004兲.
22
D. Ruelle, Statistical Mechanics: Rigorous Results 共World Scientific,
stants might be calculated using the Ewald sum method com- Singapore/Imperial College Press, London, 1999兲.
bined with techniques to include polarization. The method 23
A. A. Louis, P. G. Bolhuis, and J. P. Hansen, Phys. Rev. E 62, 7961
proposed in this work was compared with the results ob- 共2000兲.
24
tained by Groot and good agreement was found for the radial M. E. Fisher and D. Ruelle, J. Math. Phys. 7, 260 共1966兲.
25
I. Vattulainen, M. Karttunen, G. Besold, and J. M. Polson, J. Chem. Phys.
distribution functions of charged particles in a bulk electro- 116, 3967 共2002兲.
lyte solution and in a polyelectrolyte-surfactant mixture. 26
P. C. Griffiths, A. Paul, Z. Khayat, K.-W. Wan, S. M. King, I. Grillo, R.
When it was applied to study polyelectrolytes in solution the Schweins, P. Ferruti, J. Franchini, and R. Duncan, Biomacromolecules 5,
obtained electrostatic forces give the right experimental trend 1422 共2004兲.
27
E. Nordmeier, Polym. J. 25, 1 共1993兲.
on radius of gyration as a function of polymer ionization and 28
D. Stigter and K. A. Dill, Macromolecules 28, 5325 共1995兲.
as a function of solution pH as that found in experiments. 29
D. Stigter and K. A. Dill, Macromolecules 28, 5338 共1995兲.
30
B. P. Lee and M. E. Fisher, Phys. Rev. Lett. 76, 2906 共1996兲.
31
H. Frusawa, R. Hayakawa, Phys. Rev. E 61, R6079 共2000兲.
ACKNOWLEDGMENTS 32
R. J. F. Leote de Carvalho, R. Evans, and Y. Rosenfeld, Phys. Rev. E 59,
1435 共1999兲.
This work was supported by Centro de Investigación en 33
B. P. Lee and M. E. Fisher, Europhys. Lett. 39, 611 共1997兲.
Polímeros, Grupo COMEX. One of the authors 共M.G.M.兲 34
R. D. Groot and K. L. Rabone, Biophys. J. 81, 725 共2001兲.

Downloaded 08 Feb 2007 to 148.228.150.68. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

S-ar putea să vă placă și