You are on page 1of 11

pubs.acs.

org/Macromolecules Article

Charge-Density-Specific Response of Grafted Polyelectrolytes to


Electric Fields: Bending or Tilting?
Turash Haque Pial, Mihirkumar Prajapati, Bhargav Sai Chava, Harnoor Singh Sachar,
and Siddhartha Das*
Cite This: https://doi.org/10.1021/acs.macromol.2c00237 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via UNIV OF MARYLAND COLG PARK on March 31, 2022 at 17:17:23 (UTC).

ABSTRACT: The response of polyelectrolytes (PEs) to applied


electric fields drives applications in energetics, diagnostics,
materials development, and many more. Here we employ all-
atom molecular dynamics (MD) simulations to probe the response
of grafted PE brushes to axial electric fields. For PEs with large
charge densities, the electric field triggers a left−right asymmetric
distribution of counterions around the PE backbone: consequently,
depending on the location (left or right), there is an unequal
screening and an unequal force on the PE functional groups
causing a bending-driven brush height reduction. However, for the
weakly charged PEs, the electric field causes a uniform distribution
of the counterions across the brush, enforcing a uniform (without left−right asymmetry) partially unscreened PE brush: therefore, all
the brush segments experience similar force causing a brush tilting-driven brush height reduction. Such bending versus tilting
responses is commensurate with the electric-field-driven increase (decrease) in the flexibility of strongly (weakly) charged PEs.

■ INTRODUCTION
Polyelectrolyte (PE) molecules, when densely grafted, occupy
simulation-based (employing coarse-grained methods18−22)
approaches that have so far been employed to study the
“brush”-like configurations. Such PE brushes are present in electric-field-response of the PE brush molecules. There are
biological interfaces like articular cartilage, mineralized only a handful of related all-atom simulation studies for PE
collagen in bone, etc.1,2 In terms of engineering applications, molecules (nonbrush systems),23 and new insight, to be
synthetic nanochannels grafted with PE brushes, in the revealed by all-atom MD simulations, is needed for better
presence of applied electric fields, can be used as nanofluidic understanding the responses of PE brush molecules to external
ionic circuits, nanofluidic sensors, and flow controlling devices electric fields.
In this paper, we bridge this knowledge gap: we employ all-
and gates.3−8 All these applications rely on the manner in
atom molecular dynamics (MD) simulations and study the
which the applied electric field affects the distribution of the
responses of densely grafted poly(acrylic acid) (PAA) and
ions around the PE brushes (these ions are typically those that
poly(styrene sulfonate) (PSS) brushes to an axial electric field.
are screening the PE brush charges) and the resulting
These two PE molecules vary significantly in terms of their
alterations in the behavior/configurations of the PE brushes.
charge densities. Therefore, our paper seeks answers to the
The electric-field-mediated changes in the brush configu-
following questions: What is the charge-density-dependent
ration/orientation influence the lubrication action of inter-
electric-field response of PE brush molecules, and how well-
penetrating the PE brush bilayer,9,10 triggering a layer-by-layer
resolved atomistic simulations can capture such responses and
assembly of grafted PE system,11−13 and possible soft robotics
the associated responses of the PE brush supported counter-
application.14−16
ions and solvent molecules? Our all-atom MD simulations, in
These applications make it fundamentally important to
the process of answering these questions, lead to the following
develop a precise atomistic-level understanding on the manner
key discovery: we discover that the significant disparity in the
in which the imposed electric field affects the behavior of the
charged monomers of the PE, the counterions screening the
charges of these monomers, and the water molecules present Received: January 30, 2022
within the solvation shell of the PE monomer and the Revised: March 1, 2022
counterions and the manner in which the variation in the PE
charge density affects these behaviors. Such precise, atomisti-
cally resolved understanding, unfortunately, has been missing
either in experimental (often limited by its scale)17 or

© XXXX American Chemical Society https://doi.org/10.1021/acs.macromol.2c00237


A Macromolecules XXXX, XXX, XXX−XXX
Macromolecules pubs.acs.org/Macromolecules Article

charge densities of these two PEs (PAA and PSS) leads to 0.03/σ2. As the pendant group of the PAA and the PS chains are of
massively varied responses of the corresponding PE-counterion different sizes (large benzene group in the PSS), the same grafting
complex to the electric field, and the result is a bending- density for the PAA and the PSS brushes would imply a different
induced and a tilting-induced brush height reduction for the degree of confinement for the two cases. To nullify any effect of the
grafting density, we have modeled the PMAA with a grafting density
PAA and PSS brushes, respectively. Such responses are also of 0.03/σ2. The results for the PMAA (with same grafting density as
commensurate with the disparate electric-field-induced PSS) are similar to that of PAA: this indicates that the results are
changes in the intrinsic flexibilities of these two PE systems. mostly grafting density independent. We have also checked the
Our findings will be significant in a vast number of applications density of water inside the PE to quantify the degree of confinement
that rely on enforcing a particular configuration of the PE (see Figure S5). If we have used the same PE chains, a lower grafting
system (e.g., friction reduction24−26) or on the specifics of the density and a consequent lesser degree of confinement would result in
response of the PEs to the applied electric field (e.g., a higher density of water inside the brush layer. Here, although we
fabrication of electrically responsive soft devices14−16). have used a significantly smaller grafting density for the PSS brushes


as compared to the PAA brushes, the water density is actually higher
inside the PAA brush layer (see Figure S5). Such a finding implies
METHODS that the strategy of considering different grafting densities for the PAA
Molecular Dynamics Simulations Parameters. Our main and the PSS is reasonable, and it ensures that the degree of
simulation contains fully ionized poly(acrylic acid) (PAA), poly- confinement (for the two cases) becomes comparable. Also, these
(methacrylic acid) (PMAA), and poly(styrene sulfonate) (PSS) as the grafting density values used for the PAA and PSS chains are within the
polyelectrolytes and SPC/E water27 molecules. Sodium (Na+) range of experimentally reported grafting densities.40,41
counterions screen the PE brush charges. In addition, we add 0.2 Electric Field Strength. In our simulations, we have used electric
M NaCl salt. PAA chains are grafted with a grafting density of 0.05/σ2 fields in the range of 0.1 V/nm to 1 V/nm. These electric fields are
(σ = 3.5 Å, is the LJ distance parameter of backbone carbon atoms), higher than that usually employed in the experiments. A few
whereas PSS and PMAA chains are grafted with a grafting density of considerations will be discussed justifying the application of such
0.03/σ2 (see the justification in the end of this section for details). high electric fields. As we have performed all-atom MD simulations, a
Each chain has 49 backbone carbon atoms. Purely repulsive walls are high electric field is needed to observe the different results associated
placed at the top and the bottom of the system to prevent the mobile with the electric field. As the trends remain the same for 0.1 V/nm
ions and water molecules from escaping the system. 36 (25) PAA and 1 V/nm, we can say that our observed results will remain
(PSS, PMAA) PE chains are grafted on the bottom wall in a 6 × 6 (5 qualitatively similar even for smaller electric field strengths.
× 5) array. The particle trajectories are calculated using the Velocity− An important concern associated with the employment of large
Verlet algorithm, with a time step of 2 fs. Nonbonded interactions are electric field strengths is that for very large electric fields water might
modeled via a shifted-truncated 12-6 Lennard−Jones potential (ULJ) dissociate. The reported electric field strength value corresponding to
with a cutoff of 13 Å. Long range Columbic interactions are calculated the initiation of water dissociation is around 2.5 V/nm, and for
using a PPPM (particle−particle particle−mesh) algorithm.28 The frequent dissociation, one needs an applied electric field of 3.5 V/
bonds and angles of water molecules are conserved by using the nm:42,43 these numbers are several times higher than the highest
SHAKE algorithm.29 Simulations are performed in LAMMPS,30 and electric field (1 V/nm) used in our simulation.
OVITO31 is used to visualize the simulation system. We also want to discuss the practical applicability of the high
We use the OPLS-AA32 force field to model the brush molecules external electric field strengths employed in our simulations.
and employ the study by Joung et al.33 for calculating the potentials Experimental systems involving water have employed electric fields
for the mobile ions. These vastly used parameters for monovalent as high as 1−5 V/nm.44 In fact, electric fields on the order of 0.1−1
ions33 were adjusted to the solvation free energy of ions in water and V/nm are used in several experimental studies related to biopolymers.
the lattice energy of ionic crystals. The OPLS force field, which is used For example, electric fields were calculated to be in the range of 1−2.5
to model the PE brush molecule, has been used for modeling a variety V/nm in a study where protein cytochrome c was immobilized on
of polymer systems34−36 and is considered as one of the most accurate silver electrodes in the presence of electric fields.45 Similarly, an
force field parameters. Geometric mixing rules are used for the LJ electric field of 0.31 V/nm was used for probing the electromechanics
interactions between dissimilar atoms, except for the ion−ion and of a DNA molecule in a synthetic nanopore.46
ion−water interactions. For these ion−ion and ion−water, we have Fully Ionized PAA. In our study, we tried to quantify the effect of
used Lorentz−Berthelot mixing rules to remain consistent with the the PE charge density on the electric-field response of the PAA and
approach of Joung et al.33 These parameters are discussed in section PSS brushes. For this, a similar degree of ionization for both the PAA
S1 of the Supporting Information. and PSS molecules was used. PAA is a weakly dissociating polymer,
Molecular Dynamics Simulations Protocol. The system is first and its degree of ionization depends on the local pH, salt
run in the NPzT ensemble (the subscript z signifies that only the concentration, etc. For a pH value above its pKa (∼5), PAA is likely
system height is allowed to change) to obtain the correct simulation to be highly ionized,40 which justifies our consideration of fully
box height at 300 K and 1 atm, by applying the Nosé−Hoover ionized PAA. Experiments with a similar degree of polymerization and
thermostat and barostat.37,38 Then, the system is equilibrated in the grafting density (see ref 40; PAA molar mass of 2 kDa and 0.30 chain/
NVT ensemble to obtain the correct equilibrium configuration of the nm2) shows the behavior of highly charged PE chains.


system by applying the Langevin thermostat.39 We performed each
simulation until the brush height starts fluctuating around a constant RESULTS
value. Then the production period is performed for 15 ns for all the
cases of the simulations. After initial equilibration, we apply the axial Our all-atom MD simulation system consists of densely grafted
electric fields under the NVT ensemble. Again, we performed each fully ionized PAA (and PMAA) and PSS brushes. PAA and
simulation until the end-to-end height and distance starts fluctuating PMAA monomers both contain carboxylate group (COO−)
around a constant value. Then the final production period is whose total charge is −0.9e (please note that 1.0e represents
considered to be 15 ns. Further details of the simulation parameters, the charge of a proton). This stems from the fact that each O
protocols, equilibration, and simulation times are given in the atom of the COO− group has a charge of −0.8e and the C
Supporting Information.
Justifications of Using Different Simulation Parameters. atom of the COO− group has a charge of 0.7e. On the other
Grafting Density. PAA chains are grafted with a grafting density of hand, the total charge of the SO3− group (of the PSS
0.05/σ2 (σ = 3.5 Å, is the LJ distance parameter of backbone carbon molecule) is −0.56e. This stems from the fact that each O
atoms), whereas PSS chains are grafted with a grafting density of atom of the SO3− group has a charge of −0.68e, while the S
B https://doi.org/10.1021/acs.macromol.2c00237
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules pubs.acs.org/Macromolecules Article

atom of the SO3− group has a charge of 1.48e. The effect of this observations, namely, the shrinking (or bending) of the PAA
disparity in the overall charge of each functional group (or the brush and the tilting (in a direction opposite to the direction of
overall charge density) between the PAA (or PMAA) and the the applied electric field) of the PSS brush.
PSS brushes in the electric-field-response of these brushes will In Figure 2, we probe the response of the PAA brushes to
be probed here. the axial electric field. First, in Figure 2a, we show how a
We study the responses of the PAA (and PMAA) and PSS
brushes to an external axial (in the x direction) electric field E
(of strength 0.1 V/nm and 1 V/nm) (see Figure 1a). An
electric field of strength E induces a force of qiE on the atom i
having a charge of qi.

Figure 1. (a, left) Simulation system with grafted PE (PAA) brush,


counterions, co-ions, and water molecules for elucidating the effect of
an axially employed electric field on the PE molecule. Green, yellow,
and blue circles represent the carbon, oxygen, and hydrogen of the
PE. Red and purple circles indicate Na+ and Cl− ions. Teal dots
indicate water. (a, center) Chemical structures of the studied PEs
(PAA and PSS). (a, right) Considering the black line as a
representative PE chain, we define the end-to-end height and the Figure 2. (a) Distribution of counterions around an OCarboxylate for the
end-to-end distance. (b) Equilibrium (i.e., in the absence of an case of (i) E = 0 and (ii) E = 1 V/nm. Progression from lighter to
applied electric field) and steady-state (i.e., in the presence of the darker red indicates an increase in the counterion distribution. (a-iii)
applied electric field of different strengths) end-to-end brush heights Distribution profile of the counterion around an OCarboxylate along the
and end-to-end distances for PAA and PSS. (c) Equilibrium (i.e., in electric field direction. (b-ii) Counterion distribution profile around
the absence of an applied electric field) and steady-state (i.e., in the OCarboxylate for two different regions under a 1 V/nm electric field. (b-
presence of an applied electric field of strength 1 V/nm) profile for i) Schematic of the left side of the PE backbone (with respect to the
the PAA and PSS brushes. grafting position) and a cartoon representation of the uniform
counterion distribution around an OCarboxylate. (b-iii) Schematic of the
right side of the PE backbone and a cartoon representation of the
Depending on the type of the PE, the electric field changes nonuniform counterion distribution. (c-i) Schematic showing the
the configuration of the PE in two separate ways. In the expulsion of counterions from the left side of the PE chain under an
presence of the electric field, the reduction in average end-to- applied axial electric field. (c-ii) Schematic showing the manner in
end brush height (defined in Figure 1a) is similar for both the which the electric field can localize the counterions to the right side of
types of PE. Despite that, the corresponding variation of the the monomer without exclusion. (c-iii) Symmetric and asymmetric
end-to-end distance (defined in Figure 1a) is significantly number of counterions per OCarboxylate. (c-iv) Schematic showing the
different between the cases of the two PEs (Figure 1b). For mechanism of the reduction of the brush height as a result of the
PAA, for example, the electric field reduces both the end-to- bending. The distances appearing in all the figures have units of Å.
end height and end-to-end distance (Figure 1b). On the
contrary, for the PSS brush, there is little change in the end-to- positive counterion surrounds a negatively charged carboxylate
end distance with the electric field (in fact, the end-to-end monomer (of the PAA polymer) in absence and in the
distance even increases a little for the highest electric field) presence of the electric field. We assign a (0,0,0) position to
(Figure 1b). The above findings point to an end-to-end height every OCarboxylate atom (charged oxygen atom) of the PAA,
reduction via the bending-driven and tilting-driven mecha- identify the solvation shell of this OCarboxylate atom, and quantify
nisms for the PAA and the PSS brushes, respectively. Figure 1c the locations of the counterions within this solvation shell
(also see section S2 in the Supporting Information for the (please see section S3 in the Supporting Information for the
brush profiles for E = 0.1 V/nm) providing the average brush corresponding OCarboxylate-counterion radial distribution func-
profiles along the direction of the electric field hints at these tion). The ensemble averages of counterion locations for the
C https://doi.org/10.1021/acs.macromol.2c00237
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules pubs.acs.org/Macromolecules Article

Figure 3. (a) Distribution of counterions around an OSulfonate (OSulfonate acts as the center of the sphere) for the case of (i) E = 0 and (ii) E = 1 V/
nm. Progression from lighter to darker red indicates an increase in the counterion distribution. (a-iii) Distribution of the counterion around an
OSulfonate in the absence and in the presence (along the electric field direction) of the applied electric field. (b) Number of counterions per OSulfonate
as a function of the distance of the OSulfonate from the grafting position in the absence and in the presence of the applied electric field. (c) Schematic
representation of the mechanism of the tilting of the PSS brush (and the resulting reduction of the brush height) in the presence of the electric
field. The distances appearing in all the figures have units of Å.

cases of E = 0 and E = 1 V/nm are shown in Figure 2a-i,a-ii. complex implies that for the monomers that themselves are
Counterions are positively charged; therefore, an electric field on the right side of the PE backbone, the counterion will move
along the positive x-direction enforces the positively charged to the right of the monomers, i.e., away from the PE backbone
counterions to move in the positive direction (i.e., to the right (as they are subjected to no retarding effect from the PE
of the OCarboxylate atom) causing a larger concentration of backbone). Consequently, there is an asymmetrically larger
counterions for x > 0 (Figure 2a-ii). Figure 2a-iii confirms this counterion distribution on the right side of the OCarboxylate (or
observation by depicting a higher value of counterion of the monomer) that themselves are on the right side of the
distribution at a distance of 2 Å away from an OCarboxylate in PAA PE backbone (see Figure 2b-ii,b-iii).
the positive x-direction in the presence of the applied electric As the counterions on the right side of the PE backbone are
field. The space integral of this counterion distribution is equal not impeded in their responses to the applied axial electric field
to the average number of counterions in the first solvation shell (Figure 2c-ii), their numbers per monomer remain almost
of an OCarboxylate; in other words, if we denote the counterion similar to and without the electric field (Figure 2c-iii,c-iv).
distribution (as shown in Figure 2a-iii) as Y and the average However, the counterions on the left side of the PE backbone,
number of counterions in the first solvation shell of an being impeded by the PE backbone in their (migratory)
OCarboxylate as N, we can write ∫ r−rY dr = N (where r represents responses to the axial electric field, are forced to leave the
the solvation shell size of the OCarboxylate atom). solvation shell of the PE monomer (Figure 2c-i) ensuring that
We next divide the monomers of the PAA PE brush into two at such locations their numbers per monomer become much
categories: monomers on the left and monomers on the right smaller than that with no electric field (Figure 2c-iii). It should
of the PE backbone (with respect to the grafting position). be noted that the average number of Na+ ions per OCarboxylate
Figure 2a confirms that in the presence of the applied electric atom is 1.4−1.6 (for E = 0): this stems from the different
field, the counterions prefer to localize on the right side of the bridging behaviors (please see ref 47 and Figure 7 and
monomers giving rise to a directional nature to the monomer- associated discussion) resulting in a given Na+ ion being
counterion (or OCarboxylate-counterion) complex. However, for present in the first solvation shell of more than one OCarboxylate
those monomers that are themselves on the left side of the PE atom.
backbone, this directional migratory tendency of the complex Since in the presence of an axial electric field there are lesser
is impeded due to the steric hindrance imparted by the number of counterions on the left side of the PAA backbone,
presence of the PE backbone and the repulsion (to the as compared to that on the right side of the PAA backbone,
counterions) exerted by CCarboxylate (partially positively this particular section (on the left of the PAA backbone)
charged) atoms of the PE backbone (Figure 2c-i). As a behaves like a locally electronegative segment. As a
consequence, there is an almost uniform counterion distribu- consequence, this segment, in the presence of the positive
tion around an OCarboxylate (of the monomer) that is on the left electric field, attempts to move in the negative x-direction, gets
side of the PE backbone (see Figure 2b-i,b-ii). This same compressed (and hence undergoes bending and not tilting),
directional migratory nature of the OCarboxylate-counterion and ultimately results in an overall brush height reduction
D https://doi.org/10.1021/acs.macromol.2c00237
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules pubs.acs.org/Macromolecules Article

(Figure 2c-iv). The brush profiles (see Figure 1c) and the significantly change the CDF. On the other hand, as OSulfonate
brush monomer probability distributions (see Figure 5a and has a lesser charge density, not all the counterions are inside
associated discussions) along the electric field direction the first solvation shell; therefore, in the presence of the
confirm these observations. electric field, the counterions become even less condensed on
Things become significantly different when we change the the monomers thereby making the PE monomers locally
polymer type from PAA to PSS. Counterion distribution electronegative. These results indicate the possibility of
around the OSulfonate atom (charged oxygen atom) of the PSS independent movement of monomers and Na+ in PSS and
brushes, very much like the counterion distribution around the lack of it in PAA grafted systems. Results for CDF also justifies
OCarboxylate atom of the PAA, is directional (see Figure 3a-i,a-ii). the ion distribution in Figure 3b: in the presence of the electric
Interestingly, however, the value of the counterion distribution field, the ion distribution flattens (i.e., the peaks witnessed for
around the OSulfonate atom of the PSS is reduced to almost one- E = 0 are no longer present) and becomes uniform (without
third of the corresponding value of the counterion distribution left−right asymmetry) across the PE backbone.
around the OCarboxylate atom of the PAA (see Figure 3a-iii and An understanding of the variation of the monomer
compare it with Figure 2a-iii). Given that the space integral of distribution of the two different types of the PEs (PAA and
this counterion distribution around the OSulfonate atom is equal PSS) in response to the applied electric field is important as
to the average number of counterions in the first solvation shell that will help verify our explanations (see above) on the
of an OSulfonate atom (see the discussions for Figure 2), Figure charge-density dependent changes in the configurations of the
3a-iii confirms a lesser average number of counterions in the PEs in response to the applied electric field. Such variation in
first solvation shell of OSulfonate than OCarboxylate. This can be the monomer distribution is studied by quantifying the
attributed to a weaker charge density of the PSS. A lesser probability distributions of OCarboxylate (for the PAA brush
number of counterions (0.4−0.5) per OSulfonate is also visible molecule) and OSulfonate (for the PSS brush molecule) around
from Figure 3b (on the contrary, the average number of their respective grafting position (along the direction of the
counterions per OCarboxylate was 1.4−1.6). Also, in the presence axial electric field) (see Figure 5a-i,a-ii). For the PAA brush
of the electric field, the number of counterions per OSulfonate
reduces uniformly (i.e., unlike the case of PAA, there is no
left−right asymmetry in the counterion numbers per O, with
respect to the polymer grafting position). As the number of
Na+ counterions is less (as well as nearly uniform) in all
portions of the PSS brush molecule irrespective of the position
of a monomer from the grafting location, all the monomers act
as locally electronegative segments with nearly identical net
negative (unbalanced) charge. Accordingly, the application of
the electric field tries to move all of these locally electro-
negative segments in a direction opposite to the direction of
the electric field. This leads to the tilting of the PSS brush
molecule in the direction opposite to the direction of the
applied electric field (the mechanism is shown in Figure 3c and
is also confirmed by the PSS profile shown in Figure 1c),
eventually causing a decrease in the height of the PSS brush
molecule.
Figure 4 shows the cumulative distribution function (CDF)
of counterions around PAA and PSS monomers. CDF is Figure 5. (a) Monomer probability distribution for (i) PAA and (ii)
calculated by considering all the counterions inside the brush PSS in the presence and in the absence of the axial electric field. (b)
layer and finding the nearest O of a monomer from it (see Orientational correlation function of chemical distance s for different
details in ref 47). For the PAA brush molecule, all the PEs in the absence and in the presence of the applied electric field. (c)
counterions remain within the first solvation shell of OCarboxylate, Dissimilarity matrix for the PAA and PSS brush molecules in the
and accordingly, the application of the electric field does not absence and in the presence of the applied axial electric field. Blue to
yellow in the colorbar indicates the occurrence from similar chains to
dissimilar chains.

molecule, we find very little shift in the OCarboxylate probability


distribution in the presence of the electric field: this indicates
that the PAA monomer does not prefer to align in any
particular direction in the presence of the applied electric field.
On the other hand, there is a significant axial shift (toward the
negative x direction) in the OSulfonate probability distribution for
the PSS in the presence of the applied electric field. Such an
electric-field-driven shift of the OSulfonate probability distribu-
tion toward the negative x direction is commensurate with the
electric-field-driven tilting of the PSS chains in the negative x
Figure 4. Cumulative distribution function (CDF) of counterions direction.
around the PE monomer for different PEs in the absence and in the In order to quantify the electric-field-induced changes in the
presence of the applied electric field. intrinsic flexibility of the PAA and PSS chains (which might
E https://doi.org/10.1021/acs.macromol.2c00237
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules pubs.acs.org/Macromolecules Article

affect the brush conformational changes in the presence of the analyzed the end-to-end brush height and the distance for the
applied electric field), we calculate the backbone angle−angle PMAA brush molecule in the presence and in the absence of
or orientational correlation as a function of the chemical the applied axial electric field. It is evident that, with the
distance (s) which is defined as application of the electric field, both the end-to-end height and
the end-to-end distance of the PMAA brush molecule reduces
⟨cos θ(s)⟩ = ⟨ai⃗ . aj⃗ ⟩/lb 2 (1) in a manner that is very much similar to that of the PAA
brushes (see Figure 6b,c). Furthermore, no significant tilt is
where s = | i − j|, a⃗i = r⃗i − r⃗i−1, lb = | a⃗i|, and r⃗i is the position of observed in the average PMAA brush profile under the electric
each backbone carbon.48 For more flexible chains, this field (see Figure 6b,c). These results again reinforce our
orientational correlation shows a larger decay (from 1) with inference that the charge density of the pendant group is the
s. Figure 5b shows that with no electric field, both types of PEs main factor contributing to the eventual response of the PE
have similar flexibility. However, in the presence of the electric brushes to the electric field. Since, PMAA systems have the
field, there is a quick loss of orientational correlation for the same grafting density as PSS, it also signifies that grafting
PAA chains, confirming increased flexibility: this is commen- density does not affect our main interpretations significantly.


surate with the electric-field-driven bending (and the resulting
height reduction) of the PAA brushes. For the PSS chains, on DISCUSSION
the other hand, the orientational correlation increases with the
In our simulations, we have considered 36 PE chains with each
electric field. Therefore, the presence of the electric field makes
containing 49 backbone carbons. Although this degree of
the PSS chains stiffer, justifying its behavior as a tilted (and not
polymerization is achievable in experiment (see ref 40 for the
bent) system (and tilting-driven height reduction).
case where the PAA molar mass is 2 kDa), this lies in the lower
To check for the similarity among the chains, we have next
range of the chain size. As the PE layer is highly charged, we
calculated the dissimilarity matrix based on the cosine criteria
need to use enough solution above the brush layer to ensure
(see Figure 5c).49 For the calculation of the cosine
that we get appropriate ion distribution in the bulk. Also, this
dissimilarity, we have used the distance of each backbone
implies that the appropriate size of the simulation box scales as
carbon from their grafting position in the x direction. From
N (degree of polymerization). Given this requirement of
Figure 5c, we can see that the PAA chains are very much
considering an appropriately large size of the simulation box,
dissimilar from each other. The application of the electric field
coupled with the fact that we are considering an explicit water,
does not introduce any similarity in the PAA PE chains. The
we are forced to use a relatively small N. In the current study,
chains are also dissimilar for the PSS in equilibrium (i.e., in the
we tried to capture the effect of the applied electric field on the
absence of the applied electric field). However, in the presence
conformations of the PE layer. As such conformations might
of the applied axial electric field, the PSS PE chains tilt in the
depend on the flexibility of the PE chains, we have checked the
negative x direction, and their dissimilarity reduces drastically.
flexibility of the PE chains (in the presence and in the absence
This drastic reduction in the dissimilarity for the PSS PE
of the electric field) by using an orientational correlation
chains confirms that all the PSS chains responds to the electric
function (see Figure 5b). We have found that while there is not
field similarly, and their final configurations are similar to one
much of a difference in the intrinsic flexibility of PAA and PSS
another.
chains, their flexibilities (on account of the large differences in
Main Results for the PMAA PE Brushes. To be more
the charge density of their monomers) varied significantly in
confident in our simulations and inferences, we simulated a
the presence of an external electric field. This indicated that
fully ionized poly(methacrylic acid) (PMAA) brush. PMAA
the effect of the variation in the charge density of the
has the same charged functional group (COO−) as the PAA
monomers of a PE chain might be significantly more important
with an additional methyl group (see Figure 6a). The
than the initial flexibility of the PE chains in determining their
parameters for the PMAA are similar to that of the PAA.
external electric field induced conformational changes. Having
The grafting density for the PMAA brush is similar to that of
said that, we do agree that probing the electric-field-induced
the PSS brushes. To compare the findings associated with the
changes in the conformation of a given PE brush with larger N
PMAA brush with those reported in Figure 1, we have
is an important issue: a longer chain (with a different initial
flexibility as compared to the same brush molecule with a
smaller N) might show a quantitatively different behavior when
subjected to an applied electric field. Unfortunately, simulating
such chains with large N is difficult in the current all-atom MD
simulation setting and hence is beyond the scope of the
present paper.
Another important issue is the grafting density. We did
consider two different types of brushes (PAA and PMAA) with
identical charge densities (of their monomers) but different
grafting density values. We find that despite differences in their
grafting densities, they demonstrate similar electric-field-
Figure 6. (a) Chemical structures of the studied PMAA. (b) induced conformational changes. This suggests that monomer
Equilibrium (i.e., in the absence of an applied electric field) and
charge density or the charge density of the pendant group plays
steady-state (i.e., in the presence of the applied electric field of
different strengths) end-to-end heights and end-to end distances for a much more important role than the grafting density in
PMAA brushes. (c) Equilibrium (i.e., in the absence of an applied determining the qualitative nature of the electric-field-response
electric field) and steady-state (i.e., in the presence of an applied of the PE brushes. In this context, it is useful to point out that
electric field of strength 1 V/nm) profiles for the PMAA brush the grafting density values that we consider here are in the
molecule. regime of moderate grafting density values (please see ref 40,
F https://doi.org/10.1021/acs.macromol.2c00237
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules pubs.acs.org/Macromolecules Article

compare it with the grafting density of 0.30 chain/nm2); cases the salt, and we find Δq ≈ 3 in the bulk (confirming this near
for other values of grafting densities (e.g., a very small grafting perfect electroneutrality) in the presence of the applied electric
density causing the grafted polymer molecules to occupy the field. Given that almost all of the Cl− ions from the salt is
mushroom regime or a very large grafting density causing the within the brush-free bulk, this near electroneutrality in the
brushes to become fully stretched) might show different brush-free bulk implies that almost all the Na+ ions added from
electric-field-induced behavior of the PE brushes due to the the salt must also remain within the brush-free bulk (i.e., these
combination of different factors (e.g., grafted polymers Na+ ions from the salt will rarely enter the brush layer either in
behaving like free chains for weak grafting density or grafted presence or in absence of the electric field). Therefore, we can
polymers having very little room to undergo conformational infer that given that the majority of the salt ions remain within
changes for the case of large grafting density): however, such the brush-free bulk and do not enter the PE brush layer, the
studies delineating the effect of the grafting density values, due salt concentration will have an insignificant effect in
to the extreme computational cost associated with simulating determining the configuration of the PE brush molecules in
all the case for a given grafting density value, will necessitate a the presence of the applied electric field.
separate study and is hence beyond the scope of the present The above findings do not imply that under no circum-
study. stances the presence of the salt will have no effect on the brush
Effect of the Added Salt in the Electric-Field- configuration in the presence of the electric field. Some of the
Response of the PE Brushes. In this paper, we have made situations where the added salt might have noticeable
the following inference: in the presence of the applied axial influences are as follows:
electric field, the counterion distribution around an oxygen of a 1. Significantly high (several molar) salt concentration,
respective monomer (OCarboxylate or OSulfonate) of the PE dictates where the number of added salt ions significantly
the conformation of the PE layer. Given that the changed overwhelm the number of counterions screening the
brush conformation is a result of this local arrangement of ions, negative charges of the PE brush molecules.
instead of global, the effect of salt concentration (on the brush
configuration) should be minimal as long as the majority of the 2. The salt cation is different from the counterion and have
salts ions remain within the brush-free bulk. energetically more favorable interactions with the
In our simulations, we have 864 counterions (Na+ ions) to negative charge centers (e.g., OCarboxylatefor the PAA
neutralize the anionic brush (for the case of PAA brushes). We brushes and OSulfonatefor the PSS brushes): for such
have also added 0.2 M NaCl salt which corresponds to 170 cases, it might be possible that some of the original
Na+ and 170 Cl− ions. In Figure 7, we present the ion counterions (screening the PE brush charge centers) are
getting replaced by the cations from the salt.
3. The negatively charged PE brush molecules are grafted
on the inner walls of a nanochannel: under such
circumstances, we can see an overscreening (OS) effect
inside the brush layer (our recent publication6 has
probed this problem). This OS effect occurs even for salt
ion concentrations that are much smaller than the
counterion concentrations.
Our current paper considers none of these above-identified
situations. Also, we have established through Figure 7 and the
associated discussions that there is little presence of the salt
ions inside the brush layer with or without the electric field.
Figure 7. (a) Mobile ion distribution along simulation box height.
Left axis shows the number density and right axis shows number of Hence, we can confidently infer that the PE brush
ions in a slab of 5 Å along the Z axis in the simulation box. Number configuration in the presence of the applied electric field is
densities as well as the total number of Cl− ions inside and near the solely determined by the corresponding responses of the
brush layer are zoomed in the inset. The inset axes have the same counterions around the OCarboxylate or OSulfonate of the PE brush
units as the main figure. (b) Variation of the net charge Δq = total molecule, with very little influence of the added salt.
number of positive charges − total number of negative charges inside Given the weak ionization nature of the PAA PE molecules,
and outside the PE brush layer. In consideration of the total number it is important to understand the findings of Figure 7 (in terms
of charges, we account for the charges of the PE brushes, counterions, of the effect of the added salt) in the context of existing
and salt ions. Perfect electroneutrality would imply Δq = 0. Both the experimental studies. For a weak PE like PAA, experiments
figures provide the results for the PAA brushes. Shaded vertical lines demonstrate that salt concentration is indeed extremely
in part a indicate the average brush height.
important in dictating the overall brush configuration:40 the
added salt provides the necessary counterions to neutralize the
distributions in the simulation system. Figure 7a shows that PE charge, and without enough counterions (coming from
there are very few Cl− ions inside the brush layer both in the added salt) deprotonation is unfavorable. The setup of our
presence and in the absence of the applied electric field. In system (as is the case with the standard molecular dynamics
Figure 7b, we quantify the variation of the net charges (Δq) for simulation system probing charged PE molecules) is such
the entire domain (for this purpose, we also account for the where we are adding enough counterions to neutralize all the
local charges of the PE brush molecules), i.e., both inside the negative charges of the PE brushes. In other words, we are not
brush layer and inside the brush-free bulk. We find nearly relying on the added salt to provide the counterions to
perfect electroneutrality in the brush-free bulk both in the neutralize the charges of the PE brushes. In fact, if these added
presence and in the absence of the applied electric field. Please counterions (introduced in our MD set up to neutralize the PE
note that there are as many as 170 Na+ and 170 Cl− ions from brush charges) were to come directly from the added salt, the
G https://doi.org/10.1021/acs.macromol.2c00237
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules pubs.acs.org/Macromolecules Article

corresponding salt concentration would have been very large. the same chain, the interactions are denoted as intrachain
It is also observed in the experiment that after the PE bridging, while the interactions are denoted as interchain
monomer charges are completely screened by the counterion bridging if the non-neighboring monomers are from different
coming from the salt, the brush height variation is no longer chains. Higher interchain bridging is attributed to higher
significant with further addition of salt to the solution. Also, in frictions which hinder the applicability of the PE brushes in
our previous paper, we have performed simulation without lubricity applications.25 Bridging can also dictate the ion
added salt and with 0.4 M of added salt (ref 47): we observe distribution inside the PE brush layer and the surrounding
only little changes in the brush height for the fully charged solution, which can be important in ion sensing, flow inside the
PAA brushes and that the counterions remain localized within nanochannel, etc. Here we have checked the variation of
the brush layer for both the cases of with and without the interchain bridging for all of the cases, and the results are
added salt (of concentration of 0.4 M) stays in the brush layers provided in Figure 8. For this calculation, we considered all the
for both of the cases.
The experimental study performed in ref 40 considered the
behavior of the PAA brushes at pH = 3, 7, and 9. They found
that at pH = 7, the brush height is very similar to the brush
height at pH = 9 (pH = 9 represents the case, where the PAA
brushes will be fully or highly charged). Therefore, one can
consider that the PAA brushes studied in ref 40 will be fully
charged (or highly charged) at pH = 7 as well. In the present
study, we have implicitly considered pH = 7. Also, the size of
the polymer chain (N ∼ 24) and the grafting density (0.05/σ2)
considered in our simulations is very close to the chain size
(molar mass of 2 kDa, which corresponds to N ∼ 28 for the
PAA molecules) and the grafting density (0.3 chain/nm2,
which corresponds to a grafting density of ∼0.04/σ2)
considered in ref 40. Accordingly, we can consider that the Figure 8. Counterion mediated interchain bridging interactions for
PAA brushes considered in our simulations are actually fully the PAA, PMAA, and PSS brushes in the absence and in the presence
charged (or highly charged). Hence, such localization of the of the applied axial electric fields.
counterions within the brush layer (which happens for highly
charged PE brushes) and the resulting exclusion of the added counterions that are present inside the brush layer and checked
salt ions from within the brush layer are justifiable their first solvation shell for the O from different monomers
considerations in our paper for both the PSS and PAA brushes. (see ref 47 for the calculation details). With the application of
The main focus of this study is to understand the effect of the electric fields, interchain bridging increases for the PAA
charge density of the PE monomer in the electric-field-induced and PMAA brushes. On the other hand, no such augmentation
conformational changes of the PE brushes. The variation in the of interchain bridging is observed for the PSS brushes.
charge density (e.g., between the PAA and PSS monomers) Drummond10 had shown that the application of the electric
changes the counterion distribution within the solvation shell field can enhance the lubrication properties of the PE layers;
of these monomers and such changes eventually affect the for example, the friction coefficient can sometimes become
electric-field response of these brushes. Given the fact that the almost undetectable in the presence of an external electric field.
counterions for both PSS and PAA brushes remain mostly This is because of the electric-field-mediated reduction of the
localized within the brush layer itself (as discussed above), only PE brush height which creates a salt-water layer in between
the counterions that are confined within the brush layer two originally interpenetrating PE layers. In our study, the
participate in such changes in the counterion distribution; brush height reduces for all cases too, and such a brush height
accordingly, the added salt, which is mostly excluded out of the reduction, following the principle proposed by ref 10, can be
brush layer, has little effect on the electric-field-induced harnessed for better lubrications in a nanochannel (despite the
alterations of the PE brush configuration. fact that our study considers the presence of an axial applied
Relevance of the Present Study. This paper demon- electric field, while ref 10 considers a transverse electric field).
strates that a variation in the charge density of the grafted PE Lubrication properties50,51 might also depend on the alignment
brushes and the resulting response of the counterions (to an of the friction-inducing particles with respect to the applied
external axial electric field) ensures a brush height reduction by electric field. If a PE chain or other type of connected beads are
virtue of either bending (for cases of PEs with higher charge perpendicular to the electric field, larger friction will be
densities, e.g., PAA and the PMAA) or tilting (for the PE with induced9,50 as compared to the case when these connected
lesser charge density, i.e., PSS). Such specific PE responses to beads are aligned parallelly to the applied electric field. Our
the electric field have the potential to be employed in a myriad study shows that depending on the charge density of the PE
of applications. In the following sections, we will try to briefly brushes, there can be significant differences (with respect to
summarize a few of the possible impacts of our findings. the direction of the applied electric field) in the overall
Counterion mediated bridging interactions is an important alignment of the PE chains: this, therefore, can significantly
concept that dictates the conformation of PE brushes and alter the lubrication properties as a function of the nature of
regulates the use of the PE brushes in several applications. the PE brushes. PE chain stiffness can also be important when
Bridging interactions occur when a counterion accommodates designing a system for a lower friction coefficient. A recent
(in its solvation shell) two or more monomers from non- study45 has shown that a stiffer chain can be linked to lower
neighboring monomers. These non-neighboring monomers friction. In our calculations (see Figure 5b), we showed that
can be from the same or different chains; in case they are from the application of the axial electric field will decrease the
H https://doi.org/10.1021/acs.macromol.2c00237
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules


pubs.acs.org/Macromolecules Article

stiffness of the PSS chains and increase the stiffness of PAA ASSOCIATED CONTENT
chains. Therefore, our findings also indicate the possibilities of *
sı Supporting Information
regulating friction in PE grafted systems by leveraging the The Supporting Information is available free of charge at
electric-field-mediated alteration of the stiffnesses of the PE https://pubs.acs.org/doi/10.1021/acs.macromol.2c00237.
chains.
Force field parameters and other details for the MD
Researchers11−13 have proposed an electric field enabled
simulations (section S1), brush profile for 0.1 V/nm
method for fabricating drug and gene carrying multilayered PE (section S2), ionic properties (section S3), and water
complex (PEC) films (the electric field maintains a particular density (section S4) (PDF)


configuration of the PEs and speeds up the process). Our
simulations establish that depending upon the charge density
of the PE, the electric field can enforce a particular
AUTHOR INFORMATION
configuration (a compressed state or a tilted state) to the Corresponding Author
PE: this knowledge has the potential to guide researchers to Siddhartha Das − Department of Mechanical Engineering,
design a more efficient method of electric field enabled PEC University of Maryland, College Park, Maryland 20742,
film fabrication. United States; orcid.org/0000-0002-1705-721X;
In a soft device, a soft structure should respond to a given Email: sidd@umd.edu
signal by changing its shape to perform a specified task. The Authors
study shows that ensuring electric field actuated controlled and Turash Haque Pial − Department of Mechanical Engineering,
specified tilting of PE molecules (constituting a PE gel) is University of Maryland, College Park, Maryland 20742,
essential in developing soft devices based on such PE United States
gels.14,15,26 On the other hand, electric field driven periodic Mihirkumar Prajapati − Department of Mechanical
bending and shrinking is observed in a PE gel made of PAA Engineering, University of Maryland, College Park,
and Na+ counterions.16 Given that depending on the PE Maryland 20742, United States; orcid.org/0000-0003-
charge density, we observe both of such behaviors (tilting or 0524-7523
shrinking) in the system we study (in fact, our results on the Bhargav Sai Chava − Department of Mechanical Engineering,
shrinking and bending of the grafted PAA brushes screened University of Maryland, College Park, Maryland 20742,
with Na+ counterions provide the first possible atomistically United States
Harnoor Singh Sachar − Department of Mechanical
resolved explanation of the response of the PAA to an applied
Engineering, University of Maryland, College Park,
electric field), our findings will be significant for a better
Maryland 20742, United States
understanding of a myriad of different soft device applications.


Complete contact information is available at:
https://pubs.acs.org/10.1021/acs.macromol.2c00237
CONCLUSIONS
In this article, we have demonstrated the effect of the Author Contributions
monomer charge of PE brushes on the conformational change S.D. conceived the problem. T.H.P. ran the simulations.
as a result of the external electric field. We have considered two T.H.P., M.P., B.S.C., and H.S.S. analyzed the data. T.H.P. and
different pendant groups, carboxylate (PAA and PMAA) and S.D. wrote the paper. All authors commented on the paper.
sulfonate (PSS), with carboxylate having a higher charge Notes
density. The axial electric field reduced the total end-to-end The authors declare no competing financial interest.
brush height for all PE brushes; interestingly, the end-to-end
distance tells a different story. For PAA and PMAA, the end-to-
end distance decreases with the application of an external
■ ACKNOWLEDGMENTS
This work has been supported by the Department of Energy
electric field; this indicates a bending of the PE chains Office of Science Grant DE-SC0017741. The authors also
(bending is also visible from the brush profile). On the other gratefully acknowledge the Deepthought2 High-Performance
hand, with the application of the applied axial electric field, the Computing cluster at the University of Maryland for providing
end-to-end distance of the PSS brush remains similar, and the the necessary computational resources.
average chain profile shows that the brush tilts along the
electric field. Stronger interaction of highly charged carboxylate
monomer and counterion initiates an asymmetric ion
■ REFERENCES
(1) Lee, S.; Spencer, N. D. Sweet, hairy, soft, and slippery. Science
distribution; this is associated with the electric field driven 2008, 319, 575−576.
asymmetric migration of the counterion-monomer complexes, (2) Fantner, G. E.; Hassenkam, T.; Kindt, J. H.; Weaver, J. C.;
Birkedal, H.; Pechenik, L.; Cutroni, J. A.; Cidade, G. A. G.; Stucky, G.
which results in a bent brush conformation. For PSS, weaker D.; Morse, D. E.; Hansma, P. K. Sacrificial bonds and hidden length
monomer−counterion interaction implies that PE chains can dissipate energy as mineralized fibrils separate during bone fracture.
be affected by the electric field independently and move along Nat. Mater. 2005, 4, 612−616.
the direction of the electric field. The cumulative distribution (3) Vilozny, B.; Wollenberg, A. L.; Actis, P.; Hwang, D.; Singaram,
function of the counterion around the PE chains, the average B.; Pourmand, N. Carbohydrate-actuated nanofluidic diode: Switch-
able current rectification in a nanopipette. Nanoscale 2013, 5, 9214−
brush profile, and the similarity analysis of the chain supports
9221.
our findings. This study will expand our understanding of PE (4) Ali, M.; Ramirez, P.; Nasir, S.; Cervera, J.; Mafe, S.; Ensinger, W.
under an external field which can also influence future Ionic circuitry with nanofluidic diodes. Soft Matter 2019, 15, 9682−
innovations. 9689.

I https://doi.org/10.1021/acs.macromol.2c00237
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules pubs.acs.org/Macromolecules Article

(5) Ali, M.; Neumann, R.; Ensinger, W. Sequence-Specific (26) Ma, S.; Zhang, X.; Yu, B.; Zhou, F. Brushing up Functional
Recognition of DNA Oligomer Using Peptide Nucleic Acid (PNA)- Materials. NPG Asia Mater. 2019, 11, 24.
Modified Synthetic Ion Channels: PNA/DNA Hybridization in (27) Berendsen, H. J. C.; Grigera, J. R.; Straatsma, T. P. The Missing
Nanoconfined Environment. ACS Nano 2010, 4, 7267−7274. Term in Effective Pair Potentials. J. Phys. Chem. 1987, 91, 6269−
(6) Pial, T. H.; Sachar, H. S.; Desai, P. R.; Das, S. Overscreening, 6271.
Co-Ion-Dominated Electroosmosis, and Electric Field Strength (28) Hockney, R. W.; Eastwood, J. W. Computer Simulations Using
Mediated Flow Reversal in Polyelectrolyte Brush Functionalized Particles; McGraw-Hill International Book Co., New York, 1981.
Nanochannels. ACS Nano 2021, 15, 6507−6516. (29) Ryckaert, J.-P.; Ciccotti, G.; Berendsen, H. J. Numerical
(7) Buchsbaum, S. F.; Nguyen, G.; Howorka, S.; Siwy, Z. S. DNA- Integration of the Cartesian Equation of Motion of a System with
Modified Polymer Pores Allow pH- and Voltage-Gated Control of Constraints: Molecular Dynamics of n-Alkanes. J. Comput. Phys. 1977,
Channel Flux. J. Am. Chem. Soc. 2014, 136, 9902−9905. 23, 327−341.
(8) Zhou, C.; Mei, L.; Su, Y.-S.; Yeh, L.-H.; Zhang, X.; Qian, S. (30) Plimpton, S. Fast Parallel Algorithms for Short-range Molecular
Gated ion transport in a soft nanochannel with biomimetic Dynamics. J. Comput. Phys. 1995, 117, 1−19.
polyelectrolyte brush layers. Sens. Actuators, B 2016, 229, 305−314. (31) Stukowski, A. Visualization and Analysis of Atomistic
(9) Manzato, C.; Foster, A. S.; Alava, M. J.; Laurson, L. Friction Simulation Data with OVITO−the Open Visualization Tool. Model.
control with nematic lubricants via external fields. Phys. Rev. E 2015, Simul. Mater. Sci. Eng. 2010, 18, No. 015012.
91, No. 012504. (32) Jorgensen, W. L.; Maxwell, D. S.; Tirado-Rives, J. Development
(10) Drummond, C. Electric-Field-Induced Friction Reduction and and Testing of the OPLS All-Atom Force Field on Conformational
Control. Phys. Rev. Lett. 2012, 109, 154302. Energetics and Properties of Organic Liquids. J. Am. Chem. Soc. 1996,
(11) Zhang, P.; Qian, J.; Yang, Y.; An, Q.; Liu, X.; Gui, Z. 118, 11225−11236.
Polyelectrolyte Layer-by-Layer Self-Assembly Enhanced by Electric (33) Joung, I. S.; Cheatham, T. E., III Determination of Alkali and
Field and their Multilayer Membranes for Separating Isopropanol− Halide Monovalent Ion Parameters for Use in Explicitly Solvated
Water Mixtures. J. Membr. Sci. 2008, 320, 73−77. Biomolecular Simulations. J. Phys. Chem. B 2008, 112, 9020−9041.
(12) Zhang, P.; Qian, J.; An, Q.; Liu, X.; Zhao, Q.; Jin, H. Surface (34) Odegard, G. M.; Clancy, T. C.; Gates, T. S. Prediction of
Morphology and Pervaporation Performance of Electric Field Mechanical Properties of Polymers with Various Force Fields. In 46th
Enhanced Multilayer Membranes. J. Membr. Sci. 2009, 328, 141−147. AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and
(13) Yin, M.; Qian, J.; An, Q.; Zhao, Q.; Gui, Z.; Li, J. Materials Conference, Austin, TX, 2005.
Polyelectrolyte Layer-by-Layer Self-Assembly at Vibration Condition (35) Smith, M. D.; Rao, J. S.; Segelken, E.; Cruz, L. Force-Field
and the Pervaporation Performance of Assembly Multilayer Films in Induced Bias in the Structure of Aβ21−30: A Comparison of OPLS,
Dehydration of Isopropanol. J. Membr. Sci. 2010, 358, 43−50. AMBER, CHARMM, and GROMOS Force Fields. J. Chem. Inf.
(14) Li, Y.; Sun, Y.; Xiao, Y.; Gao, G.; Liu, S.; Zhang, J.; Fu, J.
Model. 2015, 55, 2587−2595.
Electric Field Actuation of Tough Electroactive Hydrogels Cross-
(36) Sharma, P.; Roy, S.; Karimi-Varzaneh, H. A. Validation of Force
linked by Functional Triblock Copolymer Micelles. ACS Appl. Mater.
Fields of Rubber through Glass-Transition Temperature Calculation
Interfaces 2016, 8, 26326−26331.
by Microsecond Atomic-Scale Molecular Dynamics Simulation. J.
(15) Zolfagharian, A.; Kaynak, A.; Khoo, S. Y.; Kouzani, A. Z.
Phys. Chem. B 2016, 120, 1367−1379.
Polyelectrolyte Soft Actuators: 3D Printed Chitosan and Cast Gelatin.
(37) Hoover, W. G. Canonical Dynamics: Equilibrium Phase-Space
3D Print Addit. Manuf. 2018, 5, 138−150.
Distributions. Phys. Rev. A 1985, 31, 1695−1697.
(16) Blyakhman, F. A.; Safronov, A. P.; Shklyar, T. F.; Filipovich, M.
(38) Nosé, S. A Unified Formulation of the Constant Temperature
A. To the Mechanism of Polyelectrolyte Gel Periodic Acting in the
Constant DC Electric Field. Sens. Actuator A Phys. 2015, 229, 104− Molecular Dynamics Methods. J. Chem. Phys. 1984, 81, 511−519.
(39) Schneider, T.; Stoll, E. Molecular-Dynamics Study of a Three-
109.
(17) Roy, T.; Szuttor, K.; Smiatek, J.; Holm, C.; Hardt, S. Electric- Dimensional One-Component Model for Distortive Phase Tran-
field-induced Stretching of Surface-tethered Polyelectrolytes in a sitions. Phys. Rev. B 1978, 17, 1302−1322.
Microchannel. Phys. Rev. E 2017, 96, No. 032503. (40) Hollingsworth, N. R.; Wilkanowicz, S. I.; Larson, R. G. Saltand
(18) Netz, R. R. Nonequilibrium Unfolding of Polyelectrolyte pH-Induced Swelling of a Poly (acrylic Acid) Brush via Quartz Crystal
Condensates in Electric Fields. Phys. Rev. Lett. 2003, 90, 128104. Microbalance w/Dissipation (QCM-D). Soft Matter 2019, 15, 7838−
(19) Suma, A.; Di Stefano, M.; Micheletti, C. Electric-Field-Driven 7851.
Trapping of Polyelectrolytes in Needle-Like Backfolded States. (41) Tran, Y.; Auroy, P. Synthesis of Poly (styrene sulfonate)
Macromolecules 2018, 51, 4462−4470. Brushes. J. Am. Chem. Soc. 2001, 123, 3644−3654.
(20) Yuan, J.; Antila, H. S.; Luijten, E. Structure of Polyelectrolyte (42) Saitta, A. M.; Saija, F.; Giaquinta, P. V. Ab Initio Molecular
Brushes on Polarizable Substrates. Macromolecules 2020, 53, 2983− Dynamics Study of Dissociation of Water under an Electric Field.
2990. Phys. Rev. Lett. 2012, 108, 207801.
(21) Lee, C. J.; Wu, H.; Hu, Y.; Young, M.; Wang, H.; Lynch, D.; (43) Stuve, E. M. Ionization of Water in Interfacial Electric Fields:
Xu, F.; Cong, H.; Cheng, G. Ionic Conductivity of Polyelectrolyte An Electrochemical View. Chem. Phys. Lett. 2012, 519, 1−17.
Hydrogels. ACS Appl. Mater. Interfaces 2018, 10, 5845−5852. (44) Scovell, D. L.; Pinkerton, T. D.; Medvedev, V. K.; Stuve, E. M.
(22) Radhakrishnan, K.; Singh, S. P. Collapse of a Confined Phase Transitions in Vapor-Deposited Water Under the Influence of
Polyelectrolyte Chain Under an AC Electric Field. Macromolecules High Surface Electric Fields. Surf. Sci. 2000, 457, 365−376.
2021, 54, 7998−8007. (45) Khoa Ly, H.; Sezer, M.; Wisitruangsakul, N.; Feng, J.-J.;
(23) Mahinthichaichan, P.; Tsai, C. C.; Payne, G. F.; Shen, J. Kranich, A.; Millo, D.; Weidinger, I. M.; Zebger, I.; Murgida, D. H.;
Polyelectrolyte in Electric Field: Disparate Conformational Behavior Hildebrandt, P. Surface-Enhanced Vibrational Spectroscopy for
Along an Aminopolysaccharide Chain. ACS Omega 2020, 5, 12016− Probing Transient Interactions of Proteins with Biomimetic
12026. Interfaces: Electric Field Effects on Structure, Dynamics and Function
(24) Kobayashi, M.; Terada, M.; Takahara, A. Polyelectrolyte of Cytochrome C. FEBS J. 2011, 278, 1382−1390.
Brushes: A Novel Stable Lubrication System in Aqueous Conditions. (46) Heng, J. B.; Aksimentiev, A.; Ho, C.; Marks, P.; Grinkova, Y. V.;
Faraday Discuss. 2012, 156, 403−412. Sligar, S.; Schulten, K.; Timp, G. The Electromechanics of DNA in a
(25) Yu, J.; Jackson, N. E.; Xu, X.; Morgenstern, Y.; Kaufman, Y.; Synthetic Nanopore. Biophys. J. 2006, 90, 1098−1106.
Ruths, M.; De Pablo, J. J.; Tirrell, M. Multivalent Counterions (47) Pial, T. H.; Sachar, H. S.; Das, S. Quantification of Mono-and
Diminish the Lubricity of Polyelectrolyte Brushes. Science 2018, 360, Multivalent Counterion-Mediated Bridging in Polyelectrolyte
1434−1438. Brushes. Macromolecules 2021, 54, 4154−4163.

J https://doi.org/10.1021/acs.macromol.2c00237
Macromolecules XXXX, XXX, XXX−XXX
Macromolecules pubs.acs.org/Macromolecules Article

(48) Hsu, H.-P.; Paul, W.; Binder, K. Standard Definitions of


Persistence Length Do Not Describe the Local “Intrinsic” Stiffness of
Real Polymer Chains. Macromolecules 2010, 43, 3094−3102.
(49) Xia, P.; Zhang, L.; Li, F. Learning Similarity with Cosine
Similarity Ensemble. Information Science 2015, 307, 39−52.
(50) Cheng, C. H. A.; Kellogg, L. H.; Shkoller, S.; Turcotte, D. L. A
liquid-Crystal Model for Friction. Proc. Natl. Acad. Sci. U.S.A. 2008,
105, 7930−7935.
(51) Yu, Y.; Yao, Y.; van Lin, S.; de Beer, S. Specific Anion Effects on
the Hydration and Tribological Properties of Zwitterionic Phosphor-
ylcholine-Based Brushes. Eur. Polym. J. 2019, 112, 222−227.

K https://doi.org/10.1021/acs.macromol.2c00237
Macromolecules XXXX, XXX, XXX−XXX

You might also like