ARTICLE
pubs.acs.org/Macromolecules
Polyelectrolytes in Salt Solutions: Molecular Dynamics Simulations
Jan-Michael Y. Carrillo and Andrey V. Dobrynin*
Polymer Program, Institute of Materials Science and Department of Physics, University of Connecticut, Storrs, Connecticut 06269,
United States
ABSTRACT: We present results of the molecular dynamics simulations of salt solutions of
polyelectrolyte chains with number of monomers N = 300. Polyelectrolyte solutions are
modeled as an ensemble of bead spring chains of charged Lennard-Jones particles with
explicit counterions and salt ions. Our simulations show that in dilute and semidilute
polyelectrolyte solutions the electrostatic induced chain persistence length scales with the
solution ionic strength as I 1/2. This dependence of the chain persistence length is due to
counterion condensation on the polymer backbone. In dilute polyelectrolyte solutions the
chain size decreases with increasing the salt concentration as R I 1/5. This is in agreement
with the scaling of the chain persistence length on the solution ionic strength, lp I 1/2. In
semidilute solution regime at low salt concentrations the chain size decreases with increasing
polymer concentration, R cp 1/4, while at high salt concentrations we observed a weaker
dependence of the chain size on the solution ionic strength, R I 1/8. Our simulations also
confirmed that the peak position in the polymer scattering function scales with the polymer
concentration in dilute polyelectrolyte solutions as cp1/3. In semidilute polyelectrolyte solutions at low salt concentrations the
location of the peak in the scattering function shifts toward the large values of q* cp1/2 while at high salt concentrations the peak
location depends on the solution ionic strength as I 1/4. Analysis of the simulation data throughout the studied salt and polymer
concentration ranges shows that there exist general scaling relations between multiple quantities X(I) in salt solutions and
corresponding quantities X(I0) in salt-free solutions, X(I) = X(I0)(I/I0)β. The exponent β = 1/2 for chain persistence length
lp, β = 1/4 for solution correlation length ξ, and β = 1/5 and β = 1/8 for chain size R in dilute and semidilute solution regimes,
respectively.
polyelectrolyte chain forms an unusual necklace-like structure of
dense polymeric beads connected by strings of monomers.1,16 30
Similar necklace-like structure can be formed in hydrophobically
modified polyelectrolytes in which a hydrophobic side chains are
attached to the polyelectrolyte backbone.31
Addition of salt leads to screening of the electrostatic interactions between ionized groups reducing the polyelectrolyte effect
(see for review refs 1 and 2). At high salt concentrations
properties of polyelectrolyte solutions are similar to those of
neutral polymers with effective second virial coefficient between
monomers which strength is determined by the salt concentration and by the fraction of the ionized groups along the polymer backbone. While there is a significant number of experimental studies of polyelectrolytes in salt solutions (see for review
refs 1, 2, and 6) the computational studies of the salt effect on the
properties of polyelectrolyte solutions are lagging behind.3,32 41
The computer simulations of the polyelectrolyte solutions in the
presence of salt were limited to investigation of the salt effect on
the single chain properties.33 37 The salt ions in these simulations were taken into account either explicitly or implicitly by
modeling the screening effect of the salt ions by representing
the electrostatic interactions between charged monomers by the
1. INTRODUCTION
Polyelectrolytes are macromolecules with ionizable groups. In
aqueous solutions charged groups dissociate, leaving charges on
the chain and releasing counterions into solution. Common
polyelectrolytes include poly(acrylic acid) and poly(methacrylic
acid) and their salts, poly(styrenesulfonate), DNA, RNA, and
other polyacids and polybases.1 9
Polyelectrolytes play an important role in a diverse number of
fields ranging from materials science and colloids to biophysics.
These polymers are used as rheology modifiers, adsorbents,
coatings, biomedical implants, colloidal stabilizing agents, and
suspending agents for pharmaceutical delivery systems.
Electrostatic interactions between charges lead to the rich
behavior of these polymeric systems (see for review refs 1 5
and 7). For example, in salt-free polyelectrolyte solutions the
electrostatic interactions between charged groups on the polymer backbone result in a strong chain elongation with chain size
scaling almost linearly with the chain degree of polymerization.
Because of this strong dependence of the chain size on the chain
degree of polymerization, the crossover to semidilute polyelectrolyte solution regime occurs at much low polymer concentrations than in solutions of neutral polymers.1,5,6,8,10 The main
contribution to the osmotic pressure in polyelectrolyte solutions
comes from the ionic component.9,11 15 Polyelectrolyte conformations are sensitive to the solvent quality for the polymer
backbone. In poor solvent conditions for the polymer backbone a
r XXXX American Chemical Society
Received:
Revised:
A
April 5, 2011
June 6, 2011
dx.doi.org/10.1021/ma2007943 | Macromolecules XXXX, XXX, 000–000
Macromolecules
ARTICLE
Debye Huckel potential. A comparison between simulations
with explicit and implicit salt ions was performed by Stevens and
Plimpton.41 Simulations with explicit salt ions35 37 were restricted
to a dilute solution regime where the intrachain electrostatic
interactions dominate over interchain ones and system behavior
can be judged from single chain simulations. The general conclusion
of these simulations was that the salt ions screen electrostatic
interactions, resulting in reduction of the chain size with increasing
the salt concentration. Unfortunately, the studied interval of salt
concentrations prevented identification of pure scaling regimes in
chain size dependence on the solution ionic strength.
In this paper we use molecular dynamics simulations to study
effect of the salt concentration on the polyelectrolyte solution
properties in dilute and semidilute solution regimes. We elucidated the effect of salt concentration on the chain size, solution
correlation length, chain overlap concentration, counterion condensation on the polymer backbone, and solution pressure. To
establish the effect of the chain rigidity on the solution properties,
we have performed simulations of polyelectrolyte chains with
different bending rigidities.
Figure 1. Monomer radial density distribution, F(r), with respect to
chain center of mass along the direction connecting centers of mass of two
neighboring chains located at r = 0 and r = D for fully charged polyelectrolyte
chains with the degree of polymerization N = 300, chain bending constant
K = 0, at polymer concentration cp = 10 3 σ 3 and at different salt concentrations: cs = 5 10 5 σ 3 (black bars), cs = 5 10 4 σ 3 (red bars), cs =
5 10 3 σ 3 (gray bars), and cs = 5 10 2 σ 3 (magenta bars).
2. MOLECULAR DYNAMICS SIMULATIONS
We performed molecular dynamics simulations42 of polyelectrolyte solutions in dilute and semidilute solution regimes in the
presence of salt. The simulation details are given in Appendix A,
below we briefly describe our model. In our simulations we used a
coarse-grained representation of polyelectrolyte chains, counterions, and salt ions.32,41,43,44 In this representation monomers and
small ions are modeled by the charged Lennard-Jones particles
with diameter σ. The value of the Lennard-Jones interaction
parameter εLJ for polymer polymer pairs was set to 0.3 kBT (where
kB is the Boltzmann constant and T is the absolute temperature).
The selection of the parameters for the LJ interaction potential
between polymer ion pairs corresponds to pure repulsive interactions (see Appendix A for details). The connectivity of the monomers into a polymer chain was maintained by the FENE potential.
To introduce a chain bending rigidity, we have imposed a bending
potential between neighboring along the polymer backbone bond
vectors. The value of the chain bending constant was set to K = 0, 3,
and 6. The chain degree of polymerization was equal to N = 300 for
all simulations. The solvent was treated implicitly as a medium with
the dielectric permittivity ε. All monomers on the polymer backbone were charged. This corresponds to the fraction of charged
monomers f = 1. The electrostatic interactions between all charges in
a system were treated explicitly through the Coulomb potential. The
value of the Bjerrum length was lB = 1.0 σ, where lB = e2/εkBT is
defined as the length scale at which the Coulomb interaction
between two elementary charges, e, in a medium with the dielectric
constant, ε, is equal to the thermal energy, kBT. In our simulations
we varied the polymer concentration, cp, between 10 4 and 10 1
σ 3 and salt concentration, cs, between 5 10 5 and 5 10 2 σ 3.
The simulations were performed in the NVT ensemble with
periodic boundary conditions. A mapping of coarse-grained parameters onto real polymeric systems is given in Appendix A.
distances D larger than their size. At such low polymer concentrations the intrachain electrostatic interactions and interactions with
the surrounding polyelectrolyte chain ions determine the chain
conformations. At polymer concentrations higher than the overlap
concentration, cp > cp*, the interchain electrostatic interactions with
surrounding chains control chain conformations.
Figure 1 shows radial distribution of the monomer density
with respect to the chain center of mass along the line connecting
centers of mass of two neighboring polyelectrolyte chains. This
distribution function of the monomer density was obtained by
averaging the monomer density within spherical shells of size r and
thickness Δ = 2σ during the production run. The average distance
between chain centers of mass D was calculated by using a 3-D
tessellation procedure. This allowed us to determine the list of the
nearest-neighbor chains and to obtain distances between them
during simulation runs. These distances were averaged over all
system configurations. This analysis shows that the average distance
between chains is on the order of D ≈ (6N/πcp)1/3. In semidilute
solution regime the monomer clouds surrounding chain center of
mass overlap (see Figure 1). With increasing the salt concentration
the monomer density distribution around the chain center of mass
shrinks, eventually leading to nonoverlapping monomer clouds. The
overlap between monomer clouds begins at D ≈ 4ÆRg2æ1/2 where
the mean-square chain radius of gyration is defined as
ÆRg 2 æ ¼
1 N
Æ ðr
N 2 i < j Bi
∑
r j Þ2 æ
B
ð1Þ
Br i is the radius vectors of the ith monomer on the polymer backbone
and average is calculated over all chain configurations. Note that at
these separations between chain centers of mass the average
monomer density in the overlapping region is less than 1% of the
maximum value of the monomer density. These monomer densities
are lower than the density fluctuation values. At separations between
chains on the order of D ≈ 3ÆRg2æ1/2 the monomer density in the
overlap region is larger than the 5% of its maximum value and is
3. OVERLAP CONCENTRATION
We begin discussion of properties of polyelectrolyte solutions
by establishing location of the crossover between dilute and
semidilute solution regimes by determining a chain overlap concentration cp*.45,46 In the dilute solution regime, cp < cp*, the
polyelectrolyte chains are separated from each other by average
B
dx.doi.org/10.1021/ma2007943 |Macromolecules XXXX, XXX, 000–000
Macromolecules
ARTICLE
Figure 3. Diagram of different solution regimes. Filled symbols correspond to semidilute solution regime, and open symbols show dilute
solution regime.
Figure 2. Dependence of the mean-square radius of gyration ÆRg2æ on
salt concentration for fully charged polyelectrolyte chains with the degree of
polymerization N = 300, chain bending constant K = 0 at different polymer
concentrations: cp = 10 4 σ 3 (open circles), cp =10 3 σ 3 (filled circles),
cp = 5 10 3 σ 3 (half-filled circles, right), cp = 10 2 σ 3 (half-filled
circles, left), cp = 5 10 2 σ 3 (half-filled circles, top), and cp = 10 1 σ 3
(half-filled circles, bottom). Inset shows location of the overlap concentration evaluated from ÆRg2æ = D2/9, which separates dilute and semidilute
solution regimes for two polymer concentrations cp = 10 4 σ 3 (dashed
line) and cp = 10 3 σ 3 (solid line).
larger than the density fluctuations. We selected polymer concentration at which the distance between chain’s center of mass
D ≈ 3ÆRg2æ1/2 as a crossover polymer concentration to the semidilute solution regime cp*. Note that this selection is not unique. For
example, one can determine the chain overlap concentration as
concentration at which the distance between chains D is on the
order of the square-root of the mean square end-to-end distance,
D ≈ (ÆRe2æ)1/2. This will lead to a different numerical coefficient in
the expression for an overlap concentration as a function of the chain
degree of polymerization.
In Figure 2, we plot dependence of the mean-square value of
the chain radius of gyration on the salt concentration cs for
polymer concentration cp varied between 10 4 and 0.1 σ 3. The
horizontal lines on this figure correspond to the following
equation D2/9 ≈ (2/9π)2/3cp 2/3. At overlap concentration
the polymer density is on the order of
cp cp
2N
9πÆRg 2 æ3=2
Figure 4. Dependence of the system pressure P on salt concentration for fully charged polyelectrolyte chains with the degree of polymerization N = 300, different values of the chain bending constants:
K = 0 (black circles), K = 3 (green squares), and K = 6 (blue rhombs);
and at different polymer concentrations. Black triangles with solid
line show pressure dependence on salt concentration in polymerfree systems.
ð2Þ
4. SYSTEM PRESSURE AND OSMOTICALLY ACTIVE
COUNTERIONS
In polyelectrolyte solutions the pressure is dominated by
contribution from small ions.1 7,15,47,48 Figure 4 shows dependence
of the system pressure P on the salt concentration at different
polymer concentrations. At low salt concentrations the system
pressure has a plateau. The magnitude of the plateau shifts toward
larger pressure values with increasing the polymer concentration. At
high salt concentrations, 2cs > cp, the system pressure is a linear
function of the salt concentration, indicating that it is controlled by
the pressure generated by salt ions. For systems with polymer
concentrations cp < 5 10 2 σ 3 in the high salt concentration
regime the pressure curves approach those obtained from simulations of the polymer-free systems. The crossover between high and
low salt concentration regimes shifts toward high salt concentrations
with increasing the polymer concentration and occurs when the
It follows from this plot that for our lowest polymer concentration cp
= 10 4 σ 3 the system is just below the chain overlap concentration.
For polymer concentration cp = 10 3 σ 3 the salt-free solution and
salt solutions with cs < 10 3 σ 3 are in a semidilute solution regime
while polyelectrolyte solutions with cs > 10 3 σ 3 are in a dilute
solution regime. Polyelectrolyte solutions with polymer concentrations cp > 10 3 σ 3 are in semidilute solution regime throughout the
entire interval of the salt concentrations studied in our simulations.
Figure 3 shows location of the dilute and semidilute solution
regimes as a function of the polymer and salt concentrations.
Note that with increasing a chain bending constant K the size
of the polyelectrolyte chain increases. The analysis of the
simulation data for K = 3 and 6 shows that all these systems
show a slightly narrower interval of polymer and salt concentrations corresponding to a dilute solution regime.
C
dx.doi.org/10.1021/ma2007943 |Macromolecules XXXX, XXX, 000–000
Macromolecules
ARTICLE
where parameter R satisfies the following nonlinear equation
Rcell
2ð2fcell 1ÞR
ð5Þ
¼
tan 2R log
σ
ð2fcell 1Þ2 R2
and parameter γ0 is equal to the ratio of the Bjerrum length lB to the
projection distance between charged monomers dc on the local chain
orientation direction, γ0 = lB/dc. The parameter γ0 is also known as
Manning Oosawa counterion condensation parameter.1 The critical value of the parameter γcrit
0 = log(Rcell/σ)/(1 + log(Rcell/σ))
determines a counterion condensation threshold. At γ0 = γcrit
0 the
value of the parameter R is equal to zero. Equations 4 and 5 can
be used to find values of the parameters R and γ0 as a function of the
system osmotic coefficient f and cell size Rcell. Note that the cell size
Rcell is related to the polymer concentration cp and Manning
Oosawa parameter γ0. The monomer concentration within a
cylindrical cell is on the order of cp ≈ (πRcell2dc) 1 ≈ (γ0/πRcell2lB).
Solving it for Rcell, we can derive an expression for the cell size in
terms of the polymer concentration cp and Manning Oosawa
parameter γ0, Rcell ≈ (γ0/πlBcp)1/2. The dashed line in Figure 5
corresponds to the boundary of the counterion condensation regime
(or saturated condensation regime in classification of refs 1 and 15),
crit
γ0 = γcrit
0 , which was obtained by substituting into eq 4 R = 0, γ0 = γ0
and solving numerically the resultant nonlinear equation. As one can
see, all our data points are inside the counterion condensation regime.
Solving eqs 4 and 5 for the values of the osmotic coefficient in
the plateau regime, we obtain for average values of γ0 ≈ 1.55 (K = 0),
γ0 ≈ 1.41 (K = 3), and γ0 ≈ 1.31 (K = 6). The value of the
Manning Oosawa counterion condensation parameter γ0 increases
with decreasing the chain bending rigidity K, indicating crumpling of
more flexible chains at short length scales. This is manifested in
shorter projection distances between charged monomers for flexible
chains dc ≈ 0.65σ (K = 0) in comparison with that for our stiffest
chains with K = 6, dc ≈ 0.77σ.
With increasing the polymer concentration the distance
between chains, Rcell, decreases, leading to increase of the value of
the parameter R. This leads to increase of the value of the osmotic
coefficient seen in Figure 5. The detailed comparison of the
predictions of the Katchalsky’s cell model with the results of the
computer simulations of flexible and rodlike chains in salt-free
solutions is given in ref 15. This analysis showed that the polymeric
contribution begins to influence the system osmotic pressure at
polymer concentrations cp > 0.05 σ 3. In this concentration range
the osmotic coefficient exceeds unity. The same trend is seen for our
simulation data. Thus, to minimize polymeric contribution to the
counterion activity coefficient, we only included data with cp e 0.05
σ 3 for our scaling analysis below.
It also worth pointing out that at salt concentration cs = 0.05 σ 3 in
the polymer-free case the system pressure is about 20% higher than
one would expect from the ideal gas contribution of the salt ions. This
points out that in addition to polymeric effects at high concentrations
of polymers and salt ions the excluded volume interactions can also
be an important factor leading in increase of the system pressure.15,48
At finite salt concentrations we can define a fraction of the
osmotically active counterions or counterion activity coefficient as
Figure 5. Dependence of the osmotic coefficient f on polymer concentration in salt-free solutions of fully charged polyelectrolyte chains
with the degree of polymerization N = 300 and different values of the
chain bending constants: K = 0 (black circles), K = 3 (green squares),
and K = 6 (blue rhombs). The dashed line shows the boundary of the
counterion condensation regime of the Katchalsky’s cell model (see text
for details). Inset shows a schematic representation of a chain in the
Katchalsky’s cell model.
salt concentration becomes on the order of half of the polymer
concentration, cP/2. It also follows from Figure 4 that the data sets
are shifted by a factor which magnitude depends on polymer
concentration. To establish a shift factor, we plot dependence of
the osmotic coefficient
f ¼ f ¼
P
kB Tcp
ð3Þ
in salt-free solutions (cs = 0) on polymer concentration cp (see
Figure 5). Note that only in salt-free solutions the system pressure is
equivalent to the osmotic pressure of the system. The value of the
osmotic coefficient in salt-free solutions is equal to the faction f * of
the osmotically active counterions.1,5,13,15,47 In some publications
the osmotically active counterions are referred to as “free” counterions that can explore the system volume and provide contribution to
the system pressure (see for review ref 1). One can also think of the
fraction of osmotically active counterions f * as a counterion activity
coefficient. At polymer concentrations cp < 10 2 σ 3 the fraction of
the osmotically active counterions f * is constant. It begins to
increase with increasing the polymer concentration for cp > 10 2
σ 3. At polymer concentration cp = 0.1 σ 3 the value of f * is larger
than unity, indicating that the polymeric part of the system pressure
start to provide a sizable contribution to the system pressure (see for
details refs 1 and 15).
Following ref 15 we can explain results of Figure 5 by using
Katchalsky’s cell model.49,50 In the framework of this approach each polyelectrolyte chain locally is approximated
by a rod surrounded by a cylindrical cell with size Rcell (see
inset in Figure 5). The distribution of electrostatic potential
throughout the cell is obtained by solving nonlinear Poisson
Boltzmann equation (see for details refs 1 and 15). According
to Katchalsky’s cell model, the value of the osmotic coefficient
is equal to
fcell ¼
1+R
2γ0
f ¼
P
2kB Tcs
kB Tcp
ð6Þ
It is important to point out that the definition of the fraction of
osmotically active counterions given by eq 6 is only warranted in the
concentration range where the pressure of the polyelectrolyte
solutions is dominated by the linear terms in the virial expansion.
2
ð4Þ
D
dx.doi.org/10.1021/ma2007943 |Macromolecules XXXX, XXX, 000–000
Macromolecules
ARTICLE
Figure 7. Dependence of the reduced pressure P(I)/P(I0) on the ratio
of the ionic strengths I/I0 for fully charged polyelectrolyte chains with
the degree of polymerization N = 300 at polymer concentrations: cp =
10 4 σ 3 (open symbols), cp = 5.0 10 4 σ 3 (half-filled symbols,
hourglass), cp =10 3 σ 3 (filled symbols), cp = 5.0 10 3 σ 3 (halffilled symbols, right), cp = 10 2 σ 3 (half-filled symbols, left), and cp =
5.0 10 2 σ 3 (half-filled symbols, top), and different values of the
chain bending constants: K = 0 (black circles), K = 3 (green squares),
and K = 6 (blue rhombs).
Figure 6. Dependence of fraction 1 f* of the condensed counterions
on the ratio of the Debye screening length to the bond length, rD/b, for
polyelectrolyte solutions of chains with K = 0 (black circles), K = 3
(green squares), and K = 6 (blue rhombs). The solid lines are the best fit
to eq 8 with the values of the fitting parameters: A = 0.566, B = 1.26 for
K = 0, A = 0.532, B = 0.937 for K = 3, and A = 0.498, B = 0.710 for K = 6.
The data points correspond to polymer concentrations cp e 0.05 σ 3
and salt concentrations cs < 0.05 σ 3.
Analysis of the pressure data shows that we can represent the
fraction of the condensed counterions (osmotically inactive counterions) 1 f* in terms of the ratio of the Debye screening length
rD
2
rD 2 ðcp , cs Þ ¼ 4πlB ðf cp + 2cs Þ
In Figure 7 we plot reduced value of the system pressure P(I)/P(I0)
as a function of the ratio of the solution ionic strengths. For this plot
we excluded data points corresponding to the highest polymer concentration cp = 0.1 σ 3 and points corresponding to salt concentration cs = 0.05 σ 3 (see discussion above).
Since the salt ions and osmotically active counterions control
screening of the electrostatic interactions between charged
monomers on the polymer backbone, one can expect a more
general scaling relation between a quantity X(I) in salt solutions
and that in a salt-free solution X(I0)
β
I
ð11Þ
XðIÞ XðI0 Þ
I0
ð7Þ
to the average bond length b (see Figure 6). Note that in definition
of the Debye screening length we only included osmotically active
counterions by multiplying counterion concentration by activity
coefficient, f *. All data sets in Figure 6 have initial linear increase
followed by saturation regime in the limit of large rD/b ratios. Flexible
chains with K = 0 have the highest value of the fraction of the
condensed counterions 1 f * in the plateau regime. This is in agreement with our observation that the flexible chains are more crumpled
at short length scales in comparison with the chains having a finite K
value. The simulation data points can be fitted by a function
1
f ¼ A½1
expð BðrD ðcp , cs Þ=b
1ÞÞ
where β is a scaling exponent. It is important to point out that in
the concentration range where f * ≈ f*0 eq 11 reduces to the wellknown scaling relation X(I) ≈ X(I0)(1 + 2cs/f *cp)β (see for review
refs 1 and 5). The invariance of the fraction of the condensed
counterions, 1 f *, with solution ionic strength (plateau regime in
Figure 6) was used for analysis of the experimental data on solution
osmotic pressure,1 solution viscosity,1 chain’s relaxation time,1 force
acting on the DNA during pore translocation,51 and swelling of the
DNA molecule in dilute solutions.52 A new feature observed in our
simulations is the decrease of the fraction of the condensed
counterions at high ionic strengths. We hope that future experiments
will test our observation.
The scaling relationships in dilute and semidilute polyelectrolyte solutions are discussed in details in the following sections.
ð8Þ
Using estimated values of the fraction f * of the osmotically active
counterions, we can collapse all data sets shown in Figure 4 into one
universal plot. At low polymer and salt concentrations the system
pressure is dominated by ideal gas-like contribution from osmotically
active counterions and salt ions
!
f cp + 2cs
PðIÞ kB Tðf cp + 2cs Þ PðI0 Þ
f0 cp
PðI0 Þ
I
I0
ð9Þ
5. DILUTE POLYELECTROLYTE SOLUTIONS
where P(I0) ≈ kBTcpf*0 is a pressure in a salt-free solution, f0* is the
fraction of osmotically active counterions in a salt-free solution
I Iðcp , cs Þ ¼ f cp + 2cs
5.1. Chain Persistence Length. In dilute polyelectrolyte
solutions the electrostatic interactions are screened at the length
scales larger than the Debye screening length, rD. The total effect
of the electrostatic interactions on the chain’s conformations is
reduced to the local chain stiffening which is manifested in
renormalization of the chain’s persistence length (known as the
ð10Þ
is the solution ionic strength, and I0 is the solution ionic strength in
salt-free solutions. Thus, one can conclude that the system pressure
is a universal function of the ratio of the solution ionic strengths.
E
dx.doi.org/10.1021/ma2007943 |Macromolecules XXXX, XXX, 000–000
Macromolecules
ARTICLE
~ (~l) of coarse-grained
Figure 9. Bond bond correlation function G
chains with fraction of charged monomers f = 1, degree of polymerization N = 300, and chain bending constant K = 0 at polymer concentration cp = 10 4 σ 3 and at different salt concentrations. Inset illustrates a
coarse-graining procedure for ng = 3.
Figure 8. Bond bond correlation function G(l) for fully charged
polyelectrolyte chains with the degree of polymerization N = 300 and
chain bending constant K = 3 at polymer concentration cp = 10 4 σ 3
and different salt concentrations.
electrostatic persistence length53 55) and to the additional chain
swelling which is due to interactions between remote along the
polymer backbone charges.52,55 59 We begin discussion of the
effect of the salt on chain conformations in dilute polyelectrolyte
solutions by analyzing orientational correlations between unit
bond vectors along the polymer backbone.
The bond bond correlation function describing decay of the
orientational memory along the polymer backbone between unit
n s+l pointing along the chain bonds and
bond vectors B
n s and B
separated by l-bonds is defined as
GðlÞ ¼
Nb
1
Nb
l
l
∑
s¼0
dimensionless force fe by setting (λe/fe)1/2 = λ2 and (λefe)1/2 =
βw 1. The effective chain bending rigidity at short length scales is
equal to λe = λ2/βw and the effective force is defined as fe = 1/λ2βw.
Thus, the first term in the right-hand side of eq 13 describes longlength scale bond bond orientational correlations while the second
one characterizes chain tension at short-length scales.
This approach works well for polyelectrolyte chains with
bending constants K = 3 and 6. However, for flexible chains with
K = 0, eq 13 fails to fit simulation data at short length scales due to
fast decay of the orientational memory between bond vectors
separated by several bonds. To minimize effect of the short-length
scale orientational fluctuations, we applied a coarse-graining procedure by grouping original bonds and representing a chain by a set of
the end-to-end vectors B
b g of the bond groups consisting of ng bonds
each (see inset in Figure 9). One can think of this coarse-graining
procedure as representation of the chain in terms of the electrostatic
blobs and treating electrostatic blobs as new effective monomers. The
optimal number of bonds ng for this coarse- graining procedure was
found self-consistently by minimizing the difference between the
simulation data and eq 10. The effective bond length of the group of
Bg|æ.
ng bonds was calculated as be = Æ|b
The results of the fitting procedure of the bond bond
correlation function are shown in Figure 10. As one can see
from the plot, the value of the chain’s bending constant λe is
almost independent of the Debye screening length for polyelectrolyte chains with the bending constants equal to K = 3 and
K = 6. However, it decreases with decreasing the value of the
Debye screening length for chains with K = 0. The decrease in
the effective chain bending constant λe is due to decrease in the
number of bonds in the coarse-grained unit from 3 at low salt
concentrations to 2 at high salt concentrations. The long-length
scale bending constant (correlation length) λ1 increases linearly
with increasing the value of the Debye radius. For the large values
of the Debye screening length the parameter λ1 is a universal
function of the Debye screening length and is independent of the
initial value of the chain’s bending constant K. It is important
to point out that the linear scaling of the chain bending constant λ1 on the Debye screening length is qualitatively different from a quadratic dependence observed for semiflexible
1
Æð n
Bs 3 nBs+l Þæ
ð12Þ
where Nb is the number of bonds in polyelectrolyte chain, Nb =
N 1, and brackets Ææ denote averaging over chain conformations. To minimize the end effects in obtaining the average values
of the bond bond correlation function, we have only considered
200 bonds in the middle of the chain during an averaging
procedure. Figure 8 shows evolution of the bond bond correlation function with salt concentration in dilute solution regime.
To extract information about chain bending rigidity, we fitted our
simulation data by combination of two exponential functions
jlj
jlj
+ βw exp
ð13Þ
GðlÞ ¼ ð1 βw Þ exp
λ1
λ2
This form of the bond bond correlation function indicates that
there are two different characteristic length scales λ1 and λ2 that
control bending rigidity of a polyelectrolyte chain at long and
short length scales respectively (see for details ref 60 ). The
values of the parameters λ1 describing the chain orientational
correlations at long length scales are always large such that we can
expand eq 13 in the power series of l/λ1. This results in the
following modification of the eq 13
jlj
jlj
GðlÞ 1 ð1 βw Þ + βw exp
1
ð14Þ
λ1
λ2
Note that the second term in the right-hand side of this equation has
a form characteristic of a semiflexible chain under tension. We can
introduce an effective chain bending constant λe and effective
F
dx.doi.org/10.1021/ma2007943 |Macromolecules XXXX, XXX, 000–000
Macromolecules
ARTICLE
Figure 10. Dependence of the bending rigidities λe and λ1 on the ratio
rD/be (where effective bond length be was set to b for chains with the
bending constants K = 3 and 6) for fully charged polyelectrolyte chains
with the degree of polymerization N = 300 in dilute polyelectrolyte
solutions at polymer concentrations: cp = 10 4 σ 3 (open symbols), cp =
5.0 10 4 σ 3 (half-filled symbols, hourglass), cp =10 3 σ 3 (filled
symbols), cp = 5.0 10 3 σ 3 (half-filled symbols, right), and different
values of the chain bending constants: K = 0 (black circles), K = 3 (green
squares), and K = 6 (blue rhombs).
Figure 11. Dependence of the normalized value of the electrostatic
bending constant Kpebe/lB on the parameter f *rD/~b in dilute polyelectrolyte solutions of chains with the degree of polymerization N = 300 at
polymer concentrations: cp = 10 4 σ 3 (open symbols), cp = 5.0 10 4
σ 3 (half-filled symbols, hourglass), cp =10 3 σ 3 (filled symbols),
cp = 5.0 10 3 σ 3 (half-filled symbols, right), and different values of
the chain bending constants: K = 0 (black circles), K = 3 (green squares),
and K = 6 (blue rhombs).
polyelectrolyte chains interacting via the screened Debye
Huckel potential.53 55,60 63 At low salt and polymer concentrations
the data points for λ1 begin to deviate from a straight line. This
deviation is due to a finite size effect. The number of charged pairs
contributing to stiffening of a chain at the length scales smaller than
or on the order of the Debye screening length saturates as the Debye
screening length becomes comparable with the chain size.
The linear dependence of the bending constant λ1 is weaker
than expected for the electrostatic induced chain bending rigidity. In
accordance with the Odijk Sckolnik Fixman (OSF) theory53,54 a
chain bending constant is equal to the sum of the chain’s bare
bending constant K and electrostatic induced chain bending constant Kpe
λ1 K + Kpe
l B f r D 2
K+C
b b
It follows from our Figure 6 that the fraction f * of the osmotically
active counterions increases with decreasing the Debye screening
length rD (increasing the solution ionic strength, I). This indicates
that the weaker dependence of the√
chain bending constant λ1 on the
solution ionic strength, λ1 1/ I, could be due to counterion
condensation on the polymer backbone.
To test this hypothesis in Figure 11, we plot dependence of the
reduced value of the electrostatic contribution to the chain bending
constant Kpebe/lB as a function of the parameter (f *rD/~b), where ~b
and be were equal to b for chains with bending constants K = 3 and 6,
and ~b = be/ng for flexible chains with K = 0. (The parameter
(f *rD/~b) is a ratio of the Debye screening length rD to the distance
between uncompensated (osmotically active) charged monomers
along the polymer backbone, ~b/f*.) In Figure 11, one can clearly
identify two different scaling regimes. In the range of small values of
the parameter (f *rD/~b), when the Debye screening length is only
marginally larger than the distance between uncompensated charges
along the polymer backbone, the reduced value of the electrostatic
bending constant Kpe demonstrate quadratic dependence on the
value of parameter (f *rD/~b) (see eq 15). Unfortunately, the errors
in the data and the covered range of parameters did not allow us to
draw a definite conclusion that the observed discrepancy in Figure 11
is only due to counterion condensation. Another factor that can
result in a weaker concentration dependence of the electrostatic
contribution to the chain’s bending constant is the nonuniform
distribution of the ions around a polymer backbone. There is an
excess of positive ions near the polymer backbone due to lower value
of the electrostatic potential. This leads to a better screening of the
polymeric charge which in turn could lead to a weaker dependence
of the chain bending rigidity on the solution ionic strength.
Note that in the original OSF calculations it was assumed that
charges on a chain interact through a screened electrostatic
(Debye Huckel) potential. However, this approximation is
only correct if polymeric charge results in a weak perturbation
of the uniform ion distribution, and it breaks down in the case
of strong electrostatic interactions. It is interesting to point
out that the weaker than quadratic dependence of the chain
ð15Þ
where C is a numerical constant equal to 1/4 in the original OSF
theory. Computer simulations of polyelectrolyte chains interacting
through the screened Debye Huckel potential show a slightly
larger value of a numerical constant C = 0.264.60 In the case of the
coarse-grained polyelectrolyte chain consisting of coarse-grained
units each having ng bonds and effective bond length be, we have to
substitute in eq 15 the bond length b by the effective bond length be,
b f be, and effective monomer valence f * by that of the coarse
grained unit f *ng, f * f f *ng. (Here and below we assume that the
effective fraction of charged monomers is equal to the fraction of
the osmotically active counterions. In reality, they are different by a
numerical coefficient.)
It follows from eq 15 that the electrostatic contribution to the
chain bending constant should demonstrate a quadratic dependence on the Debye screening length or be inversely proportional to the solution ionic strength, λ1 1/I. However, this
scaling behavior is only correct if the effective fraction of the
charged monomers on the polymer backbone f * is concentration-independent, and electrostatic interactions between charged
monomers can be approximated by the Debye Huckel potential.
G
dx.doi.org/10.1021/ma2007943 |Macromolecules XXXX, XXX, 000–000
Macromolecules
ARTICLE
persistence length on the Debye screening length was also
reported by Le Bret64 and Fixman.61 They studied contributions from nonlinear effects in ion and electrostatic potential
distributions around a charged chain to a persistence length.
In the range of parameters where (f *rD/~b) is larger than unity
the reduced electrostatic bending constant, Kpebe/lB, scales
linearly with magnitude of the parameter (f *rD/~b). This interval
of the Debye screening lengths corresponds to low salt concentrations where screening of the polymeric charge is controlled by counterions. Also in this salt concentration regime
the fraction of the condensed counterion is almost constant
(see Figure 6).
In order to obtain a persistence length of a chain which
orientational memory correlations are described by a bond bond
correlation function given by eq 13, we have to calculate the mean~ e2æ. The calculations
square value of the chain end-to-end distance, ÆR
are similar to those used for calculation of the mean-square value of
the end-to-end distance of a chain with a fixed bond angle (see for
details ref 45) and are reduced to summations of the geometric series.
In the large Nb approximation such that both Nb/λ1 and Nb/λ2 are
larger than unity one can show that
2
~ e æ Nb b2 ðð1
ÆR
βw Þhðλ1 Þ + βw hðλ2 ÞÞ
Figure 12. Dependence of the chain persistence length lp on the solution
ionic strength in dilute polyelectrolyte solutions of chains with the
degree of polymerization N = 300 at polymer concentrations cp =
10 4 σ 3 (open symbols), cp = 5.0 10 4 σ 3 (half-filled symbols,
hourglass), cp =10 3 σ 3 (filled symbols), cp = 5.0 10 3 σ 3 (halffilled symbols, right), and different values of the chain bending
constants K = 0 (black circles), K = 3 (green squares), and K = 6
(blue rhombs). Dashed line shows dependence of the Debye screening length, r D, on the solution ionic strength.
ð16Þ
where we introduced function h(x)
hðxÞ ¼ ð1 + expð x 1 ÞÞ=ð1
expð x 1 ÞÞ
ð17Þ
where lp is the chain persistence length which includes electrostatic
contributions (see previous section) and Bel is the electrostatic second
virial coefficient. In the case of the strong electrostatic interactions,
when the energy of electrostatic repulsion, U(g) kBT(lBg2/rD)
between g charges within the Debye screening length, g f *rD/b, is
much larger than the thermal energy U(g) . kBT (or uf*2 > b/rD),
the connectivity of the charged monomers into a chain plays an
important role. This is manifested in appearance of the correlation hole with size on the order of the Debye screening length
surrounding the polymer backbone. The electrostatic second virial
coefficient between charged monomers is estimated as (see for details
refs 52 and 56)
A chain persistence length lp is defined as
2
lp ¼
~e æ b
ÆR
ðð1
2Nb b 2
βw Þhðλ1 Þ + βw hðλ2 ÞÞ
ð18Þ
Note that in the limit when the chain long-range bending constant
λ1 . λ2 eq 18 can be simplified and rewritten as lp ≈ b(1 βw)λ1. In
this regime the chain persistence length is determined by the value of
the long-range correlation length (bending constant) λ1.
In Figure 12, we show dependence of the chain persistence
length lp, calculated by using eq 18, on the solution ionic strength.
The chain persistence length decreases
with increasing the
√
solution ionic strength as lp 1/ I . The deviation from the
scaling regime occurs at low solution ionic strengths corresponding to salt-free solutions where only counterions are
contributing to the screening of the polymer charge and
Debye screening length becomes comparable with a chain
size. Thus, in the high salt concentration regime where the
screening of the electrostatic interactions is dominated by the
salt ions, persistence length scales linearly with the Debye
screening length. In the next section we will show how this
scaling dependence of the chain persistence length manifests
itself in the concentration dependence of the chain size in
dilute polyelectrolyte solutions.
5.2. Chain Size. In addition to renormalization of the chain
persistence length, the electrostatic interactions between
charged monomers also result in chain swelling. This happens due to electrostatic repulsion between remote along the
polymer backbone charged monomers. The scaling analysis
of the salt effect on the polyelectrolyte chain size can be done
by using a Flory-like approach by optimizing the elastic and
monomer monomer interaction contributions to the free
energy of a chain with size R
F
R 2 Bel N 2
+ 3
kB T
blp N
R
Bel b2 rD
ð20Þ
Minimizing eq 19 with respect to a chain size R, we obtain
R bðlp rD =b2 Þ1=5 N 3=5 N 3=5 rD 2=5
ð21Þ
In rewriting eq 21, we took into account the linear relationship
between chain’s persistence length and the Debye screening
length, lp I 1/2 rD (see Figure 12). It follows from eq 21
that the chain size decreases with solution ionic strength as
R I 1/5. Figure 13 shows dependence of the mean-square value
of the chain radius of gyration ÆRg2æ on the solution ionic strength.
Our simulation data confirm scaling dependence ÆRg2æ I 2/5,
which is expected at high salt concentrations (see eq 21). The inset
in Figure 13 shows dependence of the normalized value of the
mean-square value of the chain radius of gyration on the ratio of
the solution ionic strengths.
2=5
ÆRg 2 ðIÞæ
I
ð22Þ
I0
ÆRg 2 ðI0 Þæ
All data points have collapsed into one universal line confirming the validity of the scaling assumption. Note that observed
dependence of the chain size on the solution ionic strength
ð19Þ
H
dx.doi.org/10.1021/ma2007943 |Macromolecules XXXX, XXX, 000–000
Macromolecules
ARTICLE
Figure 13. Dependence of the mean-square radius of gyration ÆRg2æ on
the solution ionic strength in dilute polyelectrolyte solutions of chains
with the degree of polymerization N = 300 at polymer concentrations
cp = 10 4 σ 3 (open symbols), cp = 5.0 10 4 σ 3 (half-filled symbols,
hourglass), cp = 10 3 σ 3 (filled symbols), and different values of the
chain bending constants K = 0 (black circles), K = 3 (green squares), and
K = 6 (blue rhombs). Inset shows dependence of the reduced value of
mean-square radius of gyration ÆRg2(I)æ/ÆRg2(I0)æ on the ratio of the
solution ionic strengths I/I0.
Figure 14. Dependence of 2π/q* on polymer concentration cp in dilute
polyelectrolyte solutions of chains with the degree of polymerization N =
300 at polymer concentrations cp = 10 4 σ 3 (open symbols), cp = 5.0
10 4 σ 3 (half-filled symbols, hourglass), cp =10 3 σ 3 (filled symbols),
and different values of the chain bending constants K = 0 (black circles),
K = 3 (green squares), and K = 6 (blue rhombs).
the characteristic features of the polyelectrolyte solutions is the
presence of the peak in a scattering function (see next section and
Appendix B for details). This peak appears due to suppression of
the polymer density fluctuations by the requirement of the
charge neutrality. In the dilute polyelectrolyte solutions the peak
is located at wavenumbers q* 1/D cp1/3 and is independent
of the salt concentration (see Figure 14).65,66,9
At wave vectors q larger than the inverse distance between
chains 1/D the scattering function S(q) is proportional to the
chain form factor P(q)
ÆRg2æ I 2/5 is weaker than one observed in simulations of
polyelectrolyte chains interacting through the screened Debye
Huckel potential, ÆRg2æ I 3/5.55,62 The main reason for
discrepancy between simulations is the counterion condensation
on the polymer backbone which weakens dependence the chain
persistence length on the solution ionic strength.
Using scaling relation for a chain size eq 22, we can obtain
dependence of the chain overlap concentration cp* on the
solution ionic strength. The overlap concentration cp* is defined
as concentration at which the average concentration in a solution
is equal to the monomer concentration inside polymer coil
!3=5
Iðcp , cs Þ
N
N
ð23Þ
cp
RðIÞ3 RðI0 Þ3 Iðcp , 0Þ
SðqÞ ¼ NPðqÞ
where the chain form factor is defined as follows:
Pð B
qÞ
∑
Æexpð ið B
q 3ð B
rj
r k ÞÞÞæ
B
N
∑ Æexpð
j, k ¼ 1
*
N
∑
j, k ¼ 1
ið B
q 3ð B
rj
sinðqj B
rj
qj B
rj
+
r k jÞ
B
r kj
B
r k ÞÞÞæ
B
ð26Þ
The summation in eq 26 is performed over all pairs of monomers
j and k, and averaging is performed over chain’s configurations.
At these length scales each chain is contributing independently to
the total system scattering intensity. The chain form factor has
the following asymptotic forms
8
pffiffiffiffiffiffiffiffiffiffi
< 1 q2 ÆRg 2 æ=3, for q ÆRg 2 æ > 1
pffiffiffiffiffiffiffiffiffiffi
ð27aÞ
PðqÞ
: ðqbÞ df ,
for 1= ÆRg 2 æ < q < 1=b
where df is a fractal dimension of a chain at the length scales l ∼
1/q. For a rodlike chain conformations df = 1.
In Figure 15a, we show evolution of the chain form factor P(q)
with the solution ionic strength in dilute polyelectrolyte solutions. At high q values all curves collapse, indicating that at
short length scales chain are stretched with P(q) 1/q.
Experimentally, the chain form factor P(q) is used to evaluate
a chain persistence length67 69 by fitting experimental data
Nm
j, k ¼ 1
1
N2
1
¼ 2
N
At low salt concentrations (f *cp . 2cs) eq 23 reduces to the
expression for overlap concentration of rodlike polyelectrolyte
chains (R N) in a salt-free solution c*p N 2. However, at high
salt concentrations (f *cp , 2cs) one obtains c*p cs3/5N 4/5, with
the same scaling dependence on the chain degree of polymerization N as in the case of neutral polymers in a good solvent for
the polymer backbone.
5.3. Scattering Function and Chain Form Factor in Dilute
Solutions. Information about chain structure in dilute solutions
can be obtained from scattering experiments. The scattering
intensity I(q
q is proportional to
B)at given scattering wavevector B
the scattering function S(q
)
B
1
Sð B
qÞ
Nm
ð25Þ
ð24Þ
where Nm = NchN is the total number of monomers in a system,Br j
is the position of the jth monomer and brackets Æ...æ denote an
ensemble average. In isotropic system the scattering function
only depends on the magnitude of the scattering vector q. One of
I
dx.doi.org/10.1021/ma2007943 |Macromolecules XXXX, XXX, 000–000
Macromolecules
ARTICLE
Table 1
cp = 1 10
cs = 5 10
5
cs = 5 10
4
cs = 5 10
3
σ
3
lp from G(l)
lp from P(q) using eq 27b
σ
3
115.3 σ
35.6 σ
σ
3
36.8 σ
19.9 σ
18.4 σ
12.2 σ
11.1 σ
8.6 σ
3
cs = 2.5 10
4
σ
σ
3
3
single-exponential decay of the bond bond correlation function used for derivation of the eq 27b. As we showed in section
5.1, this assumption fails for polyelectrolyte chains for which
the bond bond correlation function is a multiscale function.
6. SEMIDILUTE POLYELECTROLYTE SOLUTIONS
6.1. Solution Correlation Length and Scattering Function.
The important length scale above the overlap concentration, cp >
cp*, is the correlation length ξ—the average mesh size of the
semidilute polyelectrolyte solution.1,5,8 It separates two different
regimes in chain behavior. At the length scales smaller than the
solution correlation length, the conformations of the chain sections
with size smaller than the solution correlation length ξ are similar to
those in dilute polyelectrolyte solutions. At the length scales larger
than the solution correlation length the chain conformations are
those of the random walk of the correlation blobs.
In salt free-solutions a solution correlation length can be
estimated by optimizing the elastic and electrostatic parts of
the chain free energy by dividing a chain into sections each
containing gξ monomers and having size ξ
!
!
F
N ξ2 lB f 2 gξ 2
1
+
+ l B f 2 c p ξ2
N
kB T gξ bl0p gξ
ξ
bl0p cp 2 ξ4
Figure 15. (a) Chain form factor P(q) in dilute polyelectrolyte solutions of chains with the degree of polymerization N = 300 at polymer
concentrations, cp = 10 4 σ 3 and salt concentrations: cs = 5 10 5 σ 3
(black line), cs = 5 10 4 σ 3 (red line), cs = 2.5 10 3 σ 3 (blue
line), and cs = 5 10 3 σ 3 (green line). (b) Plot q2P(q) vs q for cs = 5
10 5 σ 3 (black open circles), cs = 5 10 4 σ 3 (red open squares),
cs = 2.5 10 3 σ 3 (blue open rhombs), and cs = 5 10 3 σ 3 (green
open triangle). The lines are the best fit to eq 27b considering the value
of the chain persistence length lp as a fitting parameter.
ð28Þ
In rewriting eq 28, we took into account that the correlation
blobs are space-filling, cp gξ/ξ3, and neglected logarithmic
correction to the electrostatic energy of a chain section within the
correlation blob (see for details refs 1, 5, and 8). Minimizing
eq 28 with respect to correlation blob size ξ, we obtain for the
correlation blob size
points by the function
expð xÞ + x
PðxÞ ¼ 2
x2
ξ bðuf 2 Þ
1
2lp 4
7
+
+
bN 15 15x
11 7
+
expð xÞ
15 15x
1=6 0
ðlp =bÞ 1=6 ðcp b3 Þ 1=2
cp 1=2
ð29aÞ
and for the number of monomers per correlation blob
g ξ c p ξ3
ð27bÞ
where parameter x = q2blpN/3. Figure 15b shows the results of
the fitting procedure of the simulation data to eq 27b. The
values of the fitting parameters are listed in Table 1 together
with values of the persistence length evaluated from the
bond bond correlation function. As one can see from the
Figure 15b, the agreement between eq 27b and simulation
data is reasonably good at high salt concentrations. At low salt
concentrations eq 27b completely misses the data points in
the interval of large q values. Also, it follows from the Table 1
that the values of the chain persistence length obtained from
the fitting procedure are always smaller than ones obtained
from the analysis of the bond bond correlation function. The
main reason for the discrepancy is the assumption of the
ðuf 2 Þ
1=2 0
ðlp =bÞ 1=2 ðcp b3 Þ 1=2
cp
1=2
ð29bÞ
where u is the ratio of the Bjerrum length lB to bond length b, u =
lB/b. Comparing eqs 29a and 29b, one can conclude that the
number of monomers in correlation blobs scales linearly with the
blob size. Thus, chains are stretched on the length scales on the
order of the solution correlation length.
In salt-solutions we can use a scaling approach to estimate
correlation blob size dependence on the polymer and salt
concentrations. In the framework of the scaling approach it is
assumed that on the length scales on the order of the solution
correlation length ξ a chain section with gξ monomer has the
same conformation as a chain in dilute solutions at the same salt
J
dx.doi.org/10.1021/ma2007943 |Macromolecules XXXX, XXX, 000–000
Macromolecules
ARTICLE
Figure 16. Scattering function S(q) for fully charged polyelectrolyte
chain with the degree of polymerization N = 300 and chain bending
constant K = 0 at polymer concentration cp = 10 2 σ 3 and different salt
concentrations.
Figure 17. Dependence of the correlation length ξ on the solution ionic
strength in semidilute polyelectrolyte solutions of chains with the degree of
polymerization N = 300 at polymer concentrations cp = 5.0 10 3 σ 3
(half-filled symbols, right), cp = 10 2 σ 3 (half-filled symbols, left), cp =
2.5 10 2 σ 3 (open symbols, cross), cp = 5.0 10 2 σ 3 (half-filled
symbols, top), and different values of the chain bending constants K = 0
(black circles), K = 3 (green squares), and K = 6 (blue rhombs). Inset shows
dependence of the reduced value of the correlation length ξ(I)/ξ(I0) on the
ratio of ionic strengths I/I0.
concentrations (see for details refs 1, 5, and 8). Taking this into
account, we can write the following expression for the solution
correlation length as a function of ionic strength
1=4
I
ξðIÞ ξðI0 Þ
I0
ð30Þ
solution of neutral polymers. The saturation of the scattering
intensity occurs at q ≈ 2π/ξ. Thus, we can use location of the
intersection describing the crossover between two different
scaling regimes in S(q) dependence on the magnitude of the
wavevector q for evaluation of the solution correlation length.
Figure 17 shows the dependence of the solution correlation
length on the ionic strength I. All lines converge to ξ I 1/2 with
increasing the solution ionic strength, which represents a scaling
dependence of the solution correlation length in low salt concentration regime. In the inset to Figure 17 we test the scaling prediction
eq 30 by plotting the reduced solution correlation length as a function
of the reduced ionic strength. All data points have collapsed into one
universal plot. However, the scaling exponent corresponding to the
high salt concentration regime has a value slightly smaller than 1/4
anticipated from eq 30. We can attribute this difference to the finite
size effect. With increasing the salt concentration the system moves
closer to the overlap concentration, and for our chain degree of polymerizations we do not have a sufficient range of the salt concentrations to observe a pure semidilute high salt concentration regime.
The change of the chain conformations on the length scales on
the order of the solution correlation length can be monitored by
plotting dependence of the correlation length ξ on the number of
monomers gξ within correlation length. In order to obtain gξ, we
used dependence of the square root of the mean-square value of
the end-to-end distance ÆRe2(n)æ1/2 of a section of a chain
consisting of n monomers. The results are shown in Figure 18.
At low salt concentrations and salt-free solution regime there is a
linear relationship between the solution correlation length ξ and
the number of monomers within the correlation length gξ. The
linear relationship between ξ gξ confirms the scaling assumption that the chain is stretched on the length scales smaller than
the solution correlation length (see eqs 29a and 29b). With
increasing salt concentration the number of monomers gξ within
correlation length increases. This increase is accompanied by the
change in power law dependence of ξ as a function of gξ. In this
salt concentration regime the data points approach a scaling law
In low salt concentration regime (f *cp . 2cs) the solution
correlation length ξ ≈ ξ(I0) scales with polymer concentration
as ξ cp 1/2 (see eq 29a) and is proportional to the Debye
screening length. At high salt concentrations (f *cp , 2cs) the
concentration dependence of the solution correlation length is
similar to that in solution of neutral polymers, ξ cp 3/4.
Experimentally, solution correlation length is obtained from the
peak location in the scattering function S(q).65,66,69 73 As in the case
of the dilute solutions, the peak appears due to suppression of the
polymer density fluctuations at the length scales larger than the
solution correlation length by the osmotic compressibility of
the counterions and is manifestation of the electroneutrality condition (Donnan equilibrium) at these length scales. Below we used
location of the peak q* in the scattering function S(q) to determined
solution correlation length ξ = 2π/q*.8,65,66,69 72
Note that experimental studies of polyelectrolyte solutions at
low ionic strengths show an abrupt upturn in the scattering
intensity at small q values.71,74 This behavior was confirmed by
neutron and light scattering studies.71,74 Unfortunately, in our
simulations we have not been able to corroborate the upturn
observed in scattering experiments. Our systems are too small to
separate small q upturn from the finite size effect imposed by
periodic boundary conditions.
Figure 16 shows dependence of the system scattering function
on the salt concentration in semidilute solution regime. It follows
from this figure that with increasing the salt concentration the
peak position in the scattering function shifts toward smaller
values of the wavevectors. This confirms that the correlation
length of the solution increases and system moves closer to the
overlap concentration. At the same time the magnitude of the
plateau located at small q interval increases making the peak less
pronounced. Finally, at high salt concentrations the peak completely disappears. In this range of salt concentrations the
scattering function has a form characteristic of that for semidilute
K
dx.doi.org/10.1021/ma2007943 |Macromolecules XXXX, XXX, 000–000
Macromolecules
ARTICLE
Figure 18. Dependence of the correlation length ξ on the number of
monomers within the correlation length gξ in semidilute polyelectrolyte
solutions of chains with the degree of polymerization N = 300 at polymer
concentrations cp = 5.0 10 3 σ 3 (half-filled symbols, right), cp = 10 2
σ 3 (half-filled symbols, left), cp =2.5 10 2 σ 3 (open symbols,
cross), cp = 5.0 10 2 σ 3 (half-filled symbols, top), and different
values of the chain bending constants K = 0 (black circles), K = 3 (green
squares), and K = 6 (blue rhombs). Inset shows dependence of the
reduced value of the correlation length ξ(I)/ξ(I0) on the reduced
number of monomers within the correlation length gξ(I)/gξ(I0).
Figure 20. Dependence of the reduced chain persistence length lp(I)/
lp(I0) on the ratio of the solution ionic strengths I/I0 in semidilute
polyelectrolyte solutions of chains with the degree of polymerization N =
300 at polymer concentrations cp = 5.0 10 3 σ 3 (half-filled symbols,
right), cp = 10 2 σ 3 (half-filled symbols, left), cp =2.5 10 2 σ 3
(open symbols, cross), cp = 5.0 10 2 σ 3 (half-filled symbols, top),
and different values of the chain bending constants K = 0 (black circles),
K = 3 (green squares), and K = 6 (blue rhombs).
solution correlation length, lp ξ cp 1/2.1,5,75 In this salt concentration interval the bending of the polyelectrolyte chain is caused by
repulsion from neighboring chains because the Debye screening
length due to free counterions and salt ions is larger than the solution
correlation length ξ. It is also important to point out that in this salt
concentration regime the sections of the chain are also contributing to
the screening of the electrostatic interactions setting up the electrostatic screening length to be on the order of the solution correlation
length. For detailed discussion of the screening effects in semidilute
polyelectrolyte solutions see refs 1 and 5. The deviation from the
linear relationship between persistence length and correlation length
occurs at high polymer concentrations when correlation length
becomes comparable with the bare chain persistence length of a
chain l0p. This can be clearly seen in the inset in Figure 19 which shows
data for polyelectrolyte chains with the bending constant K = 6.
At high salt concentrations, f*cp < 2cs, the chain persistence
length decreases with increasing solution correlation length ξ. In this
range of salt and polymer concentrations the persistence length is
inversely proportional to the square of the solution correlation length,
lp ξ 2. Taking into account the scaling relation between correlation
length ξ and ionic strength I, ξ I1/4(see eq 30), one can show that
the chain persistence length lp is inversely proportional to the square
root of the solution ionic strength I. Thus, in this high salt concentration regime we see the same salt concentration dependence
of the chain persistence length as in dilute polyelectrolyte solutions
(see Figure 12). In Figure 20, we plot lp(I)/lp(I0) as a function of the
ratio of the solution ionic strengths. The data sets group into three
groups according to the values of the bare chain persistence length.
For systems with K = 0 and K = 3 the high salt data sets approach a
scaling law
1=2
I
ð31Þ
lp ðIÞ lp ðI0 Þ
I0
Figure 19. Dependence of the chain persistence length lp on correlation
length ξ in semidilute polyelectrolyte solutions of chains with the degree
of polymerization N = 300 at polymer concentrations cp = 5.0 10 3
σ 3 (half-filled symbols, right), cp = 10 2 σ 3 (half-filled symbols, left),
cp = 2.5 10 2 σ 3 (open symbols, cross), cp = 5.0 10 2 σ 3 (halffilled symbols, top), and different values of the chain bending constants
K = 0 (black circles), K = 3 (green squares), and K = 6 (blue rhombs).
ξ gξν with exponent ν between 3/5 and 1/2 (see inset in
Figure 18). Unfortunately, the sections of a chain within correlation length ξ are too short to distinguish between 3/5 and 1/2
exponents.
6.2. Chain Persistence Length. The chain persistence length
in semidilute solution regime was obtained from fitting a chain
bond bond correlation function (see eq 12) by sum of two
exponentials (see eq 13) and using relation eq 18 between fitting
parameters and chain persistence length. The results of this procedure
are summarized in Figure 19. At low salt concentrations, f*cp > 2cs,
when the screening of the electrostatic interactions is dominated by
counterions, the chain persistence length is proportional to the
6.3. Chain Size Scaling. The scaling model of a polyelectrolyte chain in semidilute solutions is based on the assumption
L
dx.doi.org/10.1021/ma2007943 |Macromolecules XXXX, XXX, 000–000
Macromolecules
ARTICLE
Figure 22. Dependence of the mean-square radius of gyration ÆRg2æ on
the solution ionic strength in semidilute polyelectrolyte solutions of
chains with the degree of polymerization N = 300 at polymer concentrations cp = 5.0 10 3 σ 3 (half-filled symbols, right), cp = 10 2 σ 3
(half-filled symbols, left), cp = 2.5 10 2 σ 3 (open symbols, cross),
cp = 5.0 10 2 σ 3 (half-filled symbols, top), and different values of the
chain bending constants K = 0 (black circles), K = 3 (green squares), and
K = 6 (blue rhombs). Inset shows dependence of the reduced value
of mean-square radius of gyration ÆRg2(I)æ/ÆRg2(I0)æ on the ratio of the
solution ionic strengths I/I0.
Figure 21. Dependence of the ratio of the square root of the meansquare radius of gyration and correlation length ÆRg2æ1/2/ξ on the ratio
of the degree of polymerization N and the number of monomers within
the correlation length gξ in semidilute polyelectrolyte solutions of chains
with the degree of polymerization N = 300 at polymer concentrations
cp = 5.0 10 3 σ 3 (half-filled symbols, right), cp = 10 2 σ 3 (half-filled
symbols, left), cp = 2.5 10 2 σ 3 (open symbols, cross), cp = 5.0
10 2 σ 3 (half-filled symbols, top), and different values of the chain
bending constants K = 0 (black circles), K = 3 (green squares), and K = 6
(blue rhombs).
chain radius of gyration
of the existence of a single length scale—the correlation length
ξ.1,5,8 At the length scales larger than the solution correlation
length ξ, other chains and counterions screen electrostatic interactions, and the statistics of the chain are those of a Gaussian chain
with the effective persistence length on the order of the correlation
length ξ. Thus, the polyelectrolyte chain is assumed to be flexible at
the length scales on the order of the correlation length ξ. According
to the scaling model, a chain in the semidilute polyelectrolyte
solution is a random walk of correlation blobs with size
!1=2
N
ð32Þ
Rξ
gξ
I
ÆRg ðIÞæ ÆRg ðI0 Þæ
I0
2
N
ξðI0 Þ
ÆRg 2 ðI0 Þæ
gξ ðIÞ
ξðIÞ
1=4
ð34Þ
In low salt concentration regime eq 33 reduces to ÆRg2æ
N1/2cp 1/2, while at high salt concentration limit the chain size scales
with polymer concentration as ÆRg2æ N1/2cp 1/4 recovering the
polymer concentration of the chain size in solutions of neutral
polymers.1,5 Figure 22 combines our simulation data for dependence
of the mean-square value of the chain radius of gyration in semidilute
solution regime. As expected at low salt concentration limit the data
follow ÆRg2æ I 1/2 scaling dependence while at high salt concentrations we see ÆRg2æ I 1/4. To collapse all our simulation data into
one universal plot in the inset, we plot the ratio of the mean-square
value of the chain radius of gyration ÆRg2(I)æin salt solutions to
that in a salt-free solution as a function of the ratio of the solution
ionic strengths (see eq 34). Data collapse into one universal line with
slope 1/4, thus confirming eq 34.
To test scaling hypothesis for the chain size dependence on the
number of correlation blobs per chain, the plot of the normalized
chain size (ÆRg2æ)1/2/ξ as a function of the number of correlation
blobs per chain N/gξ is shown in Figure 21. All points for chains at
different polymer and salt concentrations collapse onto one universal
line, with the slope 1/2 as expected for Gaussian chains with N/gξ
correlation blobs. There is a slight deviation from universal behavior
in the range of small number of the correlation blobs per chain.
Using expression for the chain size dependence on the number
of blobs per chain and solution correlation length ξ, we can
derive expression for the chain size as a function of the solution
ionic strength. For a mean-square value of the chain radius of
gyration we have
ÆRg 2 ðIÞæ ξðIÞ2
2
7. SUMMARY
We performed molecular dynamics simulations of salt solutions of polyelectrolyte chains with the degree of polymerization
N = 300. Polyelectrolyte chains were modeled by Lennard-Jones
particles connected by the finite extension nonlinear elastic
bonds. The mutual orientation of the neighboring along the
polymer backbone bonds was maintained by imposing an angular
potential. The counterions and salt ions were included explicitly
into our simulations.
Our simulations show that in salt-free solutions in polymer
concentration range cp g 10 4 σ 3 osmotic coefficient show a
plateau then increases with increasing polymer concentration. At
polymer concentrations cp larger than 5 10 2 σ 3 the value of
the osmotic coefficient exceeds unity, indicating that polymeric
and excluded volume effects begin to contribute to the system
pressure. For lower polymer concentrations we related the value
ð33Þ
In rewriting the last equation, we took into account that the
correlation blobs are space-filling and cp ≈ gξ(I)/ξ(I)3 ≈ gξ(I0)/
ξ(I0)3. Using eq 30, we can express the ratio of the solution
correlation lengths in terms of the ratio of ionic strengths. This
results in the following expression for a mean-square value of the
M
dx.doi.org/10.1021/ma2007943 |Macromolecules XXXX, XXX, 000–000
Macromolecules
ARTICLE
N/gξ (see Figure 21). Thus, in salt-free solutions in semidilute
solution regime a polyelectrolyte chain can be considered as a
random walk of correlation blobs. Note that in this salt concentration
range the number of monomers within correlation length scales
linearly with its size. At high salt concentrations the chain size R
decreases with increasing the solution ionic strength as R I 1/8
(see Figure 22). This scaling relation provides additional evidence for
inverse square root dependence of the chain persistence length on
the solution ionic strength, lp I 1/2.
In our study of the effect of the salt on the polyelectrolyte
solution properties we only considered the effect of monovalent
ions. The behavior of polyelectrolytes in the presence of divalent
ions is qualitatively different from that in solutions of monovalent
salts.76 81 In the case of divalent ions the counterion condensation on the polymer backbone can lead to chain associations and
phase separation.78,79,81 We are planning to consider these effects
in a separate publication.
Table 2. Scaling Relations for Dilute and Semidilute Polyelectrolyte Solutions
dilute solutions
P(I)
P(I0)I/I0
ÆRg2(I)æ
ξ(I)
ÆRg2(I0)æ(I/I0)
lp(I)
lp(I0)(I/I0)
semidilute solutions
P(I0)I/I0
1/2
2/5
ÆRg2(I0)æ(I/I0)
ξ(I0)(I/I0)1/4
lp(I0)(I/I0)
1/4
1/2
of the osmotic coefficient with the fraction of osmotically active
counterions. In the salt solutions the fraction of osmotically
active counterions was obtained from the difference between
system pressure P and ideal pressure of salt ions, 2kBTcs (see
eq 6). The fraction of osmotically active counterions f * shows a
weak dependence on the chain bending rigidity, K. It has a
smaller value for flexible chains with bending constant K = 0
and increases with increasing the chain’s bending rigidity (see
Figure 6). This is manifestation of the crumpling of a chain at
short length scales and higher linear charge density for chains
with K = 0 in comparison with that for more rigid chains.
Using fraction of osmotically active counterions f * obtained
from counterion contribution to the system pressure, we
defined ionic strength of the solution I (see eq 10) and verified
a general scaling relation between a quantity X(I) in salt solutions
and that in a salt-free solution X(I0). The results are summarized
in Table 2.
In dilute and semidilute solution regimes the electrostatic
interactions between charged monomers result in chain stiffening. Our analysis of the bond bond correlation function shows
that chain persistence length lp is inversely proportional
√ to the
square root of the solution ionic strength, lp 1/ I. This
dependence of the chain persistence length is weaker than
I 1, and observed
predicted by the OSF model,53 55,60,61 lOSF
p
in computer simulations of the polyelectrolyte chains interacting via
screened Coulomb potential.60,62,63 We showed that the weaker
dependence of the chain persistence length on the solution ionic
strength can be explained by counterion condensation on the polymer backbone. The square root dependence of the chain persistence
length on the solution ionic strength in dilute solution regime is in
agreement with a chain size R decrease with ionic strength as R
I 1/5 (see Figure 13).
The correlation length ξ is an important length scale in
semidilute solution regime, which separates two different length
scales. At the length scales smaller than the solution correlation
length the chain statistics is the same as in a dilute solution while at
the length scales larger than the solution correlation length the
interactions between monomers are screened and chain behaves as a
Gaussian chain of correlation blobs. The value of the solution correlation length was obtained from the peak position in the polymer
scattering function S(q). At low salt concentrations we observed the
exponent for the concentration dependence of the solution correlation length to be close to cp 1/2, in agreement with the scaling model
of semidilute polyelectrolyte solutions. In the high salt concentration
regime, f *cp < 2cs, our data confirm a general scaling relation ξ(I) ≈
ξ(I0)(I/I0)1/4 (see Figure 17). This dependence of the solution
correlation length is in agreement with the scaling of the chain
persistence length with ionic strength as lp I 1/2. We have also
tested a scaling assumption that in salt-free solutions the chain
persistence length is proportional to the solution correlation length
(see Figure 19). This was further corroborated by the scaling of the
reduced chain size R/ξ on the number of correlation blobs per chain
’ APPENDIX A. SIMULATION DETAILS
We performed molecular dynamics simulations of polyelectrolyte chains with explicit counterions and salt ions. Polyelectrolytes were modeled by chains of charged Lennard-Jones (LJ)
particles (beads) with diameter σ and degree of polymerization
N = 300. Each monomer on the polymer backbone was charged.
Counterions and salt ions were modeled as LJ particles (beads)
with diameter σ. The solvent was treated implicitly as a dielectric
medium with dielectric constant ε.
All particles in the system interacted through truncated-shifted
Lennard-Jones (LJ) potential:
ULJ ðrij Þ ¼
8
>
>
<
2
σ
4εLJ 4
rij
>
>
:0
!12
σ
rij
!6
σ
rcut
3
12 6
σ 5
+
rcut
r e rcut
r > rcut
ðA:1Þ
where rij is the distance between ith and jth beads and σ is the
bead diameter chosen to be the same regardless of the bead type.
The cutoff distance, rcut = 2.5σ, was set for polymer polymer
interactions, and rcut = 21/6σ was selected for all other pairwise
interactions. The interaction parameter εLJ was equal to kBT
for polymer ion and ion ion interactions. The value of the
Lennard-Jones interaction parameter for the polymer polymer
pairs was set to 0.3kBT, which is close to a theta solvent condition
for the polymer backbone. By selecting the strength of the
polymer polymer interactions close to the θ-point, we minimized the effect of the short-range interactions on polyelectrolyte solution properties.
The connectivity of monomers into polymer chains was
maintained by the finite extension nonlinear elastic (FENE)
potential:
!
1
r2
2
kspring Rmax ln 1
ðA:2Þ
UFENE ðrÞ ¼
2
Rmax 2
with the spring constant kspring = 30kBT/σ2 and the maximum
bond length Rmax = 1.2σ. The repulsive part of the bond potential
was represented by the truncated-shifted LJ potentials with εLJ =
1.5kBT and rcut = 21/6σ.
The chain bending rigidity was introduced into the model
through a bending potential controlling the mutual orientations
between two neighboring along the polymer backbone unit bond
N
dx.doi.org/10.1021/ma2007943 |Macromolecules XXXX, XXX, 000–000
Macromolecules
ARTICLE
Table 3. List of Studied Systems
cp [σ 3]
0.0
0.0001
0.0005
0.001
Figure 23. Snapshot of the simulation box containing fully charged
polyelectrolyte chains, counterions, and salt ions. The polyelectrolyte
chains are shown in blue, chain counterions are shown in red, and cyan
and orange colored beads represent negatively and positively charged
salt ions, respectively.
0.005
vectors B
n i and B
n i+1
Ui,bend
i + 1 ¼ kB TKð1
ðn
Bi 3 n
Bi+1 ÞÞ
ðA:3Þ
The bending constant K was equal to 0, 3, and 6.
Interaction between any two charged particles with charge
valences qi and qj, and separated by a distance rij, was given by the
Coulomb potential
UCoul ðrij Þ ¼ kB T
l B qi qj
rij
0.01
ðA:4Þ
where lB = e2/εkBT is the Bjerrum length. In our simulations, the
value of the Bjerrum length lB was equal to 1.0σ. The particle
particle particle mesh (PPPM) method implemented in
LAMMPS82 with the sixth-order charge interpolation scheme
and estimated accuracy 10 5 was used for calculations of the
electrostatic interactions between all charges in the system.
Simulations were carried out in a constant number of particles,
volume, and temperature ensemble (NVT) with periodic boundary conditions. The snapshot of the simulation box is shown in
Figure 23. The simulation box sizes, number of particles in
simulation box, and covered polymer and salt concentrations are
summarized in Table 3.
The simulations were performed at a constant temperature,
which was maintained by coupling the system to the Langevin
thermostat. The motion of beads was described by the following
equation
d v ðtÞ
¼ B
F i ðtÞ
m Bi
dt
ξB
v i ðtÞ +
R
B
F i ðtÞ
0.025
0.05
0.1
ðA:5Þ
where m is the bead mass, Bv i(t) is the bead velocity, and B
F i(t)
denotes the net deterministic force acting on the ith bead. The
BRi (t)æ = 0 and
stochastic force B
F Ri (t) has a zero average value ÆF
O
cs [σ 3]
Lx [σ]
Nch
Ncion
Nsalt
Ntotal
0.00005
0.0005
0.0025
0.005
0.025
0.05
0
0.00005
0.0005
0.0025
0.005
0
0.00005
0.0005
0.0025
0.005
0.025
0
0.00005
0.0005
0.0025
0.005
0.025
0.05
0
0.00005
0.0005
0.0025
0.005
0.025
0.05
0
0.00005
0.0005
0.0025
0.005
0.025
0.05
0
0.00005
0.0005
0.0025
0.005
0.025
0
0.00005
0.0005
0.0025
0.005
0.025
0.05
0
0.00005
0.0005
0.0025
0.005
0.025
0.05
368.40
171.00
100.00
79.37
46.42
36.84
493.24
432.68
275.89
181.71
144.23
288.45
278.50
228.94
161.34
133.89
84.34
228.94
225.06
200.83
148.88
128.06
84.34
66.94
133.89
133.89
129.27
117.45
106.27
74.89
62.14
106.27
106.27
104.46
106.27
106.27
106.27
59.44
78.30
78.30
77.64
75.60
73.43
62.14
62.14
62.14
62.14
61.09
60.00
54.51
49.32
49.32
49.32
49.32
48.91
48.49
45.79
43.27
0
0
0
0
0
0
40
27
7
2
1
40
36
20
7
4
1
40
38
27
11
7
2
1
40
40
36
27
20
7
4
40
40
38
40
40
40
7
40
40
39
36
33
20
40
40
40
38
36
27
20
40
40
40
39
38
32
27
0
0
0
0
0
0
12000
8100
2100
600
300
12000
10800
6000
2100
1200
300
12000
11400
8100
3300
2100
600
300
12000
12000
10800
8100
6000
2100
1200
12000
12000
11400
12000
12000
12000
2100
12000
12000
11700
10800
9900
6000
12000
12000
12000
11400
10800
8100
6000
12000
12000
12000
11700
11400
9600
8100
5000
5000
5000
5000
5000
5000
0
8100
21000
30000
30000
0
2160
12000
21000
24000
30000
0
1140
8100
16500
21000
30000
30000
0
240
2160
8100
12000
21000
24000
0
120
1140
6000
12000
60000
21000
0
48
468
2160
3960
12000
0
24
240
1140
2160
8100
12000
0
12
120
585
1140
4800
8100
5000
5000
5000
5000
5000
5000
24000
24300
25200
31200
30600
24000
23760
24000
25200
26400
30600
24000
23940
24300
23100
25200
31200
30600
24000
24240
23760
24300
24000
25200
26400
24000
24120
23940
30000
36000
84000
25200
24000
24048
23868
23760
23760
24000
24000
24024
24240
23940
23760
24300
24000
24000
24012
24120
23985
23940
24000
24300
dx.doi.org/10.1021/ma2007943 |Macromolecules XXXX, XXX, 000–000
Macromolecules
ARTICLE
solution. It is defined as
Sð B
qÞ ¼
1 Nm
Æ expðið B
q 3ð B
rk
Nm k, j
∑
r j ÞÞÞæ
B
ðB:1Þ
The summation in eq B.1 is taken over all pairs of monomers
Nm = NchN in a system, Br k is the radius vector of the kth
monomer, and brackets Æ...æ denote the ensemble average. In
performing calculations of the eq B.1, it is useful to introduce a
Fourier transform of the monomer density in a given system
configuration
Fð B
qÞ ¼
∑ expðið Bq 3 Br kÞÞ
ðB:2Þ
k¼1
In terms of the F(q
B) the scattering function can be rewritten as
Figure 24. Evolution of the mean-square value of the radius of gyration
during simulation run in salt-free solution of fully charged polyelectrolyte chains with the degree of polymerization N = 300, chain bending
constant K = 0, at polymer concentration cp =1 10 4 σ 3.
ÆF
BRi (t)F
BRi (t0 )æ
Nm
Sð B
qÞ ¼
1
ÆFð B
q ÞFð
Nm
q Þæ
B
ðB:3Þ
Thus, calculation of the scattering function is reduced to
calculation of the F(q
B) in a given system configuration. The
FFT is an efficient method in calculating of the Fourier transform of the special distributed function. To apply an FFT
method to our system, we represented the actual monomer
distribution by distribution over Ng = L/Δ grid points. The
number of the grid points was varied between 128 and 256
depending on the system sizes. The distance between the grid
points was selected from the interval b/2 e Δ < b to satisfy the
condition that the number of the grid points was a power of 2.
We used a linear interpolation scheme to smear a monomer
between neighboring grid points. In this representation the
function F(q
B) was transformed to
0
δ-functional correlations
= 6kBTξδ(t t ). The
friction coefficient ξ was set to ξ = 0.143m/τLJ, where τLJ is the
standard LJ time τLJ = σ(m/kBT)1/2. The velocityVerlet algorithm with a time step Δt = 0.01τLJ was used for
integration of the equations of motion eq A.5. All simulations
were performed using LAMMPS.82
Simulations were performed using the following procedure: at
the beginning of each simulation run polyelectrolyte chains in the
self-avoiding walk conformation, counterions and salt ions were
randomly distributed over the volume of the simulation box. This
followed by the equilibration step. The equilibration step continued until the mean-square value of the chain radius of gyration
averaged over all chains in the simulation box reached saturation.
This followed by a production run lasting 2 106 integration
steps. Figure 24 shows evolution of the mean-square average
value of the radius of gyration. At each time step the average was
calculated over all chains in a system. Note that the total duration
of the simulation runs was varied between 4 106 and 8 106
integration steps, depending on how long it took for the system
to reach equilibrium.
Let us relate the parameters used in our coarse-grained MD
simulations of polyelectrolyte solutions with charged polymeric
systems. In the bead spring representation of a polymer chain,
each bead represents several chemical units. For example, if we
assume that the value of the Bjerrum length, lB = 1σ, used in our
simulations is equal to the Bjerrum length in water at room
temperature (T = 298 K), lB = 0.714 nm, the monomer size is
equal to 0.714 nm. This corresponds approximately to 2.9
monomers of sodium poly(styrenesulfonate) with monomer
size 0.25 nm and leads to a polymer chain with the degree of
polymerization on the order of 870 monomers. The polymer
concentrations in our simulations correspond to polyelectrolyte
solutions with polymer concentrations 4.6 10 4 e cp e 4.6
10 1 M. The salt concentrations used in our simulations
correspond to salt concentrations of NaCl within the range 2.3
10 4 e cs e 2.3 10 1 M.
Fð B
q Þ ¼ Fðl, m, nÞ
Ng
1
∑
¼
s, j, h ¼ 0
!
2π
exp i ðls + mj + nhÞ nðs, j, hÞ
Ng
ðB:4Þ
where n(s,j,h) is a number of monomers in a grid point with
coordinates s, j, and h. Note that wavevector B
q has coordinates
qx = 2πl/L, qy = 2πm/L, qz = 2πn/L, and q = (qx2 + qy2 + qz2)1/2.
The FFT method was used to calculate the complex function
F(l,m,n). Knowing function F(l,m,n), we have calculated the
scattering function in grid points
Sðl, m, nÞ ¼
1
ÆðRe Fðl, m, nÞÞ2 + ðIm Fðl, m, nÞÞ2 æ
Nm
ðB:5Þ
where Re F(l,m,n) and Im F(l,m,n) are real and imaginary parts
of the complex number F(l,m,n). Figure 25 shows results of the
3-D FFT of the polyelectrolyte system.
For isotropic systems one can use a 1-D FFT by pointing
vector B
q along x-axis, q = qx. Taking this into account we can
simplify eq B.4
FðqÞ ¼ Fðq, 0, 0Þ ¼ Fðl, 0, 0Þ
!
Ng 1
2πls
¼
exp i
nðsÞ
Ng
s¼0
∑
’ APPENDIX B. CALCULATIONS OF THE SCATTERING
FUNCTION
Scattering function S(q
B) provides information about conformations of polyelectrolyte chains and structural properties of the
ðB:6Þ
where n(s) represents a one-dimensional monomer distribution
over the grid points along the x-axis. It is obtained by sorting
monomers only according to their x-coordinates. Note that it can
P
dx.doi.org/10.1021/ma2007943 |Macromolecules XXXX, XXX, 000–000
Macromolecules
ARTICLE
’ REFERENCES
(1) Dobrynin, A. V.; Rubinstein, M. Theory of polyelectrolytes in
solutions and at surfaces. Prog. Polym. Sci. 2005, 30, 1049–1118.
(2) Colby, R. H. Structure and linear viscoelasticity of flexible
polymer solutions: comparison of polyelectrolyte and neutral polymer
solutions. Rheol. Acta 2010, 49 (5), 425–442.
(3) Dobrynin, A. V. Theory and simulations of charged polymers:
From solution properties to polymeric nanomaterials. Curr. Opin.
Colloid Interface Sci. 2008, 13 (6), 376–388.
(4) Holm, C.; Joanny, J. F.; Kremer, K.; Netz, R. R.; Reineker, P.;
Seidel, C.; Vilgis, T. A.; Winkler, R. G. Polyelectrolyte theory. Adv.
Polym. Sci. 2004, 166, 67–111.
(5) Dobrynin, A. V.; Colby, R. H.; Rubinstein, M. Scaling theory of
polyelectrolyte solutions. Macromolecules 1995, 28 (6), 1859–1871.
(6) Forster, S.; Schmidt, M. Polyelectrolytes in Solution. Adv. Polym.
Sci. 1995, 120, 51–133.
(7) Barrat, J. L.; Joanny, J. F. Theory of polyelectrolyte solutions.
Adv. Chem. Phys. 1996, 94, 1–66.
(8) de Gennes, P. G.; Pincus, P.; Brochard, F.; Velasco, R. M.
Remarks on polyelectrolyte conformation. J. Phys. (Paris) 1976,
37, 1461–1476.
(9) Yethiraj, A. Liquid State Theory of Polyelectrolyte Solutions.
J. Phys. Chem. B 2009, 113 (6), 1539–1551.
(10) Boris, D. C.; Colby, R. H. Rheology of sulfonated polystyrene
solutions. Macromolecules 1998, 31 (17), 5746–5755.
(11) Takahashi, A.; Kato, N.; Nagasawa, M. The osmotic pressure of
polyelectrolyte in neutral salt solutions. J. Phys. Chem. 1970, 74, 944–946.
(12) Koene, R. S.; Nicolai, T.; Mandel, M. Scaling relations for
aqueous polyelectrolyte-salt solutions. 3. Osmotic pressure as a function
of molar mass and ionic strength in the semidilute regime. Macromolecules 1983, 16, 231–236.
(13) Raspaud, E.; da Conceicao, M.; Livolant, F. Do free DNA
counterions control the osmotic pressure? Phys. Rev. Lett. 2000, 84 (11),
2533–2536.
(14) Nagy, M. Thermodynamic study of aqueous solutions of
polyelectrolytes of low and medium charge density without added salt
by direct measurement of osmotic pressure. J. Chem. Thermodyn. 2010,
42 (3), 387–399.
(15) Liao, Q.; Dobrynin, A. V.; Rubinstein, M. Molecular dynamics
simulations of polyelectrolyte solutions: Osmotic coefficient and counterion condensation. Macromolecules 2003, 36 (9), 3399–3410.
(16) Dobrynin, A. V.; Rubinstein, M.; Obukhov, S. P. Cascade of
Transitions of Polyelectrolytes in Poor Solvent. Macromolecules 1996,
29, 2974–2979.
(17) Dobrynin, A. V.; Rubinstein, M. Hydrophobic polyelectrolytes.
Macromolecules 1999, 32 (3), 915–922.
(18) Solis, F. J.; de la Cruz, M. O. Variational approach to necklace
formation in polyelectrolytes. Macromolecules 1998, 31 (16), 5502–5506.
(19) Limbach, H. J.; Holm, C. Conformational properties of poor
solvent polyelectrolytes. Comput. Phys. Commun. 2002, 147 (1 2),
321–324.
(20) Limbach, H. J.; Holm, C.; Kremer, K. Conformations and solution
structure of polyelectrolytes in poor solvent. Macromol. Symp. 2004,
211, 43–53.
(21) Limbach, H. J.; Holm, C. Single-chain properties of polyelectrolytes in poor solvent. J. Phys. Chem. B 2003, 107 (32), 8041–8055.
(22) Chodanowski, P.; Stoll, S. Monte Carlo simulations of hydrophobic polyelectrolytes: Evidence of complex configurational transitions. J. Chem. Phys. 1999, 111 (13), 6069–6081.
(23) Liao, Q.; Carrillo, J. M. Y.; Dobrynin, A. V.; Rubinstein, M.
Rouse dynamics of polyelectrolyte solutions: Molecular dynamics study.
Macromolecules 2007, 40, 7671–7679.
(24) Jeon, J.; Dobrynin, A. V. Necklace globule and counterion
condensation. Macromolecules 2007, 40, 7695–7706.
(25) Lyulin, A. V.; Dunweg, B.; Borisov, O. V.; Darinskii, A. A.
Computer simulation studies of a single polyelectrolyte chain in poor
solvent. Macromolecules 1999, 32 (10), 3264–3278.
Figure 25. 3-D FFT from solution of fully charged polyelectrolyte
chains with the degree of polymerization N = 300, chain bending
constant K = 0, at polymer concentration cp =5 10 3 σ 3 in saltfree regime.
Figure 26. Scattering function S(q) of fully charged polyelectrolyte
chains with the degree of polymerization N = 300, chain bending
constant K = 0, at polymer concentration cp =5 10 3 σ 3 in saltfree solution calculated by using 3-D FFT (open circles) and 1-D FFT
(red squares).
also be obtained from a 3-D monomer distribution function
n(s,j,k) by performing summations over indexes j and k. To
improve accuracy, it is useful to perform FFT calculations for x, y,
and z directions and average the final result for S(q).
In Figure 26 we combined results for S(q) obtained from 3-D
FFT and 1-D FFT calculations. The results are close. However, it
is important to point out that in the range of large q (q 2π/b)
the FFT-based methods introduce errors due to finite grid size.
’ ACKNOWLEDGMENT
The authors are grateful to the National Science Foundation
for the financial support under Grant DMR-1004576.
Q
dx.doi.org/10.1021/ma2007943 |Macromolecules XXXX, XXX, 000–000
Macromolecules
ARTICLE
(51) Keyser, U. F.; Koeleman, B. N.; S, v. D.; Krapf, D.; Smeets,
R. M. M.; Lamy, S. G.; Dekker, N. H.; Dekker, C. Direct force
measurements on DNA in a solid-stae nanopore. Nature Phys. 2006,
2, 473–477.
(52) Dobrynin, A. V.; Carrillo, J. M. Y. Swelling of biological and
semiflexible polyelectrolytes. J. Phys.: Condens. Matter 2009, 21 (42),
424112.
(53) Skolnick, J.; Fixman, M. Electrostatic persistence length of a
wormlike polyelectrolyte. Macromolecules 1977, 10 (5), 944–948.
(54) Odijk, T. Polyelectrolytes near the rod limit. J. Polym. Phys., Part
B: Polym. Phys. 1977, 15, 477–483.
(55) Ullner, M. Comments on the scaling behavior of flexible
polyelectrolytes within the Debye-Huckel approximation. J. Phys. Chem.
B 2003, 107 (32), 8097–8110.
(56) Fixman, M.; Skolnick, J. Polyelectrolyte excluded volume
paradox. Macromolecules 1978, 11, 863–867.
(57) Odijk, T.; Houwaart, A. C. On the theory of the excludedvolume effect of a polelectrolyte in a 1 1 electrolyte solution. J. Polym.
Sci., Polym. Phys. Ed. 1978, 16, 627–639.
(58) Muthukumar, M. Adsorption of a polyelectrolyte chain to a
charged surface. J. Chem. Phys. 1987, 86 (12), 7230–7235.
(59) Muthukumar, M. Double screening in polyelectrolyte solutions: Limiting laws and crossover formulas. J. Chem. Phys. 1996, 105
(12), 5183–5199.
(60) Gubarev, A.; Carrillo, J. M. Y.; Dobrynin, A. V. Scale-Dependent Electrostatic Stiffening in Biopolymers. Macromolecules 2009, 42
(15), 5851–5860.
(61) Fixman, M. The flexibility of polyelectrolyte molecules. J. Chem.
Phys. 1982, 76, 6346–6353.
(62) Everaers, R.; Milchev, A.; Yamakov, V. The electrostatic
persistence length of polymers beyond the OSF limit. Eur. Phys. J. E
2002, 8 (1), 3–14.
(63) Nguyen, T. T.; Shklovskii, B. I. Persistence length of a polyelectrolyte in salty water: Monte Carlo study. Phys. Rev. E 2002, 66 (2),
021801–1 7.
(64) Le Bret, M. Electrostatic contribution of the persistence length
of a polyelectrolyte. J. Chem. Phys. 1982, 76, 6243–6255.
(65) Borsali, R.; Rinaudo, M.; Noirez, L. Light-scattering and smallangle neutron-scattering from polyelectrolyte solutions - the succinoglycan. Macromolecules 1995, 28 (4), 1085–1088.
(66) Vallat, P.; Catala, J. M.; Rawiso, M.; Schosseler, F. Flexible
conjugated polyelectrolyte solutions: A small angle scattering study.
Macromolecules 2007, 40 (10), 3779–3783.
(67) Nishida, K.; Urakawa, H.; Kaji, K.; Gabrys, B.; Higgins, J. S.
Electrostatic persistence length of NaPSS polyelectrolytes determined
by a zero average contrast SANS technique. Polymer 1997, 38 (24),
6083–6085.
(68) Buhler, E.; Boue, F. Chain persistence length and structure in
hyaluronan solutions: Ionic strength dependence for a model semirigid
polyelectrolyte. Macromolecules 2004, 37 (4), 1600–1610.
(69) Essafi, W.; Spiteri, M. N.; Williams, C. E.; Boue, F. Hydrophobic polyelectrolytes in better polar solvent. Structure and chain
conformation as seen by SAXS and SANS. Macromolecules 2009,
42, 9568–9580.
(70) Prabhu, V. M.; Amis, E. J.; Bossev, D. P.; Rosov, N. Counterion
associative behavior with flexible polyelectrolytes. J. Chem. Phys. 2004,
121 (9), 4424–4429.
(71) Wang, D. L.; Lal, J.; Moses, D.; Bazan, G. C.; Heeger, A. J. Small
angle neutron scattering (SANS) studies of a conjugated polyelectrolyte
in aqueous solution. Chem. Phys. Lett. 2001, 348 (5 6), 411–415.
(72) Boucard, N.; David, L.; Rochas, C.; Montembault, A.; Viton, C.;
Domard, A. Polyelectrolyte microstructure in chitosan aqueous and
alcohol solutions. Biomacromolecules 2007, 8 (4), 1209–1217.
(73) Combet, J.; Isel, F.; Rawiso, M.; Boue, F. Scattering functions of
flexible polyelectrolytes in the presence of mixed valence counterions:
Condensation and scaling. Macromolecules 2005, 38 (17), 7456–7469.
(74) Ermi, B. D.; Amis, E. J. Domain structures in low ionic strength
solutions. Macromolecules 1998, 31, 7378–7384.
(26) Baigl, D.; Ober, R.; Qu, D.; Fery, A.; Williams, C. E. Correlation
length of hydrophobic polyelectrolyte solutions. Europhys. Lett. 2003, 62
(4), 588–594.
(27) Spiteri, M. N.; Williams, C. E.; Boue, F. Pearl-necklace-like
chain conformation of hydrophobic polyelectrolyte: a SANS study of
partially sulfonated polystyrene in water. Macromolecules 2007, 40 (18),
6679–6691.
(28) Aseyev, V. O.; Klenin, S. I.; Tenhu, H. Conformational changes
of a polyelectrolyte in mixtures of water and acetone. J. Polym. Sci., Part
B: Polym. Phys. 1998, 36 (7), 1107–1114.
(29) Aseyev, V. O.; Klenin, S. I.; Tenhu, H.; Grillo, I.; Geissler, E.
Neutron scattering studies of the structure of a polyelectrolyte globule in
a water-acetone mixture. Macromolecules 2001, 34 (11), 3706–3709.
(30) Morawetz, H. Revisiting some phenomena in polyelectrolyte
solutions. J. Polym. Sci., Part B: Polym. Phys. 2002, 40 (11), 1080–1086.
(31) Dobrynin, A. V.; Rubinstein, M. Hydrophobically modified
polyelectrolytes in dilute salt-free solutions. Macromolecules 2000, 33
(21), 8097–8105.
(32) Dobrynin, A. V. Molecular simulations of charged polymers. In
Simulation Methods for Polymers; Kotelyanskii, M., Theodorou, D. N.,
Eds.; Marcel Dekker: New York, 2004; pp 259 312.
(33) Sarraguca, J. M. G.; Skepo, M.; Pais, A.; Linse, P. Structure of
polyelectrolytes in 3: 1 salt solutions. J. Chem. Phys. 2003, 119 (23),
12621–12628.
(34) Wei, Y. F.; Hsiao, P. Y. Role of chain stiffness on the
conformation of single polyelectrolytes in salt solutions. J. Chem. Phys.
2007, 127 (6), 064901.
(35) Wei, Y. F.; Hsiao, P. Y. Effect of chain stiffness on ion
distributions around a polyelectrolyte in multivalent salt solutions.
J. Chem. Phys. 2010, 132 (2), 024905.
(36) Hsiao, P. Y.; Luijten, E. Salt-induced collapse and reexpansion
of highly charged flexible polyelectrolytes. Phys. Rev. Lett. 2006, 97 (14),
148301.
(37) Liu, S.; Ghosh, K.; Muthukumar, M. Polyelectrolyte solutions
with added salt: A simulation study. J. Chem. Phys. 2003, 119 (3),
1813–1823.
(38) Klos, J.; Pakula, T. Lattice Monte Carlo simulations of threedimensional charged polymer chains. J. Chem. Phys. 2004, 120 (5),
2496–2501.
(39) Klos, J.; Pakula, T. Computer simulations of a polyelectrolyte
chain with a mixture of multivalent salts. J. Phys.: Condens. Matter 2005,
17 (37), 5635–5645.
(40) Chang, R. W.; Yethiraj, A. Brownian dynamics simulations of
salt-free polyelectrolyte solutions. J. Chem. Phys. 2002, 116 (12),
5284–5298.
(41) Stevens, M. J.; Plimpton, S. J. The effect of added salt on
polyelectrolyte structure. Eur. Phys. J. B 1998, 2, 341–345.
(42) Frenkel, D.; Smit, B. Understanding Molecular Simulations;
Academic Press: New York, 2002.
(43) Stevens, M. J.; Kremer, K. The nature of flexible linear polyelectrolytes in salt-free solution - a molecular-dynamics study. J. Chem.
Phys. 1995, 103 (4), 1669–1690.
(44) Liao, Q.; Dobrynin, A. V.; Rubinstein, M. Molecular dynamics
simulations of polyelectrolyte solutions: Nonuniform stretching of
chains and scaling behavior. Macromolecules 2003, 36 (9), 3386–3398.
(45) Grosberg, A. Y.; Khokhlov, A. R. Statistical Physics of Macromolecules; AIP Press: New York, 1994.
(46) Rubinstein, M.; Colby, R. H. Polymer Physics; Oxford University
Press: New York, 2003.
(47) Oosawa, F. Polyelectrolytes; Marcel Dekker: New York, 1971.
(48) Chang, R.; Yethiraj, A. Osmotic pressure of salt-free polyelectrolyte solutions: A Monte Carlo simulation study. Macromolecules 2005,
38 (2), 607–616.
(49) Fuoss, R. M.; Katchalsky, A.; Lifson, S. The potential of an
infinite rod-like molecule and the distribution of counterions. Proc. Natl.
Acad. Sci. U.S.A. 1951, 37, 579–586.
(50) Alfrey, T.; Berg, P. V.; Morawetz, H. The counterion distribution
in solutions of rod-shaped polyelectrolytes. J. Polym. Sci. 1951, 7, 543–551.
R
dx.doi.org/10.1021/ma2007943 |Macromolecules XXXX, XXX, 000–000
Macromolecules
ARTICLE
(75) Spiteri, M. N.; Boue, F.; Lapp, A.; Cotton, J. P. Polyelectrolyte
persistence length in semidilute solution as a function of the ionic
strength. Physica B 1997, 234, 303–305.
(76) Chang, R. W.; Yethiraj, A. Brownian dynamics simulations of
polyelectrolyte solutions with divalent counterions. J. Chem. Phys. 2003,
118 (24), 11315–11325.
(77) Ermoshkin, A. V.; de la Cruz, M. O. Polyelectrolytes in the
presence of multivalent ions: Gelation versus segregation. Phys. Rev. Lett.
2003, 90 (12), 125504-1-4.
(78) Ermoshkin, A. V.; De la Cruz, M. O. Gelation in strongly
charged polyelectrolytes. J. Polym. Sci., Part B: Polym. Phys. 2004, 42 (5),
766–776.
(79) Gonzalez-Mozuelos, P.; de la Cruz, M. O. Association in
electrolyte solutions: Rodlike polyelectrolytes in multivalent salts.
J. Chem. Phys. 2003, 118 (10), 4684–4691.
(80) Huang, C. I.; de la Cruz, M. O. Polyelectrolytes in multivalent
salt solutions: Monomolecular versus multimolecular aggregation.
Macromolecules 2002, 35 (3), 976–986.
(81) Raspaud, E.; de la Cruz, M. O.; Sikorav, J. L.; Livolant, F.
Precipitation of DNA by polyamines: A polyelectrolyte behavior.
Biophys. J. 1998, 74 (1), 381–393.
(82) Plimpton, S. J. Fast parallel algorithms for short-range molecular dynamics. J. Comput. Phys. 1995, 117 (1), 1-19. http://lammps.
sandia.gov.
S
dx.doi.org/10.1021/ma2007943 |Macromolecules XXXX, XXX, 000–000