Polymeric complex coacervation occurs when two oppositely charged polyelectrolytes undergo an associative phase separation in aqueous salt solution, resulting in a polymer-dense coacervate phase and a polymer-dilute supernatant phase. This phase separation process represents a powerful way to tune polymer solutions using electrostatic attraction and is sensitive to environmental conditions such as salt concentration and valency. One area of particular research interest is using this to create nanoscale polymer assemblies, via (for example) block copolymers with coacervate-forming blocks. The key to understanding coacervate-driven assembly is the formation of the interface between the coacervate and supernatant phases and its corresponding thermodynamics. In this work, we use recent advances in coacervate simulation and theory to probe the nature of the coacervate-supernatant interface. First, we show that self-consistent field theory informed by either Monte-Carlo simulations or transfer matrix theories is capable of reproducing interfacial features present in large-scale molecular dynamics simulations. The quantitative agreement between all three methods gives us a way to efficiently explore interfacial thermodynamics. We show how salt affects the interface, and we find qualitative agreement with literature measurements of interfacial tension. We also explore the influence of neutral polymers, which we predict to drastically influence the phase behavior of coacervates. These neutral polymers can significantly alter the interfacial tension in coacervates; this has a profound effect on the design and understanding of coacervate-driven self-assembly, where the equilibrium structure is tied to interfacial properties.

Oppositely charged polyelectrolytes can undergo an associative liquid-liquid phase separation in salt solution in a process known as complex coacervation.1–4 This phase behavior has been widely exploited in both industry and research due to its sensitivity to the ionic environment; for example, the food and personal care industries have long used coacervates as encapsulants and viscosity modifiers,5,6 and research has considered coacervation as a route to controlled drug release vehicles,7–9 as tissue engineering scaffolds,10,11 and in synthetic underwater adhesives.12,13 More recently, coacervation-driven block copolymer assembly has been used for multimodal probes and reactive sensors.14 This diverse range of applications stems from the ability to readily tune the strength of coacervation by changing the chemical and physical properties of the coacervate-forming polyelectrolytes and their environment.

Building on longstanding interest in coacervate materials,2,3 there has been a resurgent effort to characterize their equilibrium and dynamic behavior on a fundamental level. These include extensive experimental efforts; for example, the phase behavior of coacervation has been explored for a number of different polyelectrolyte systems, as a function of salt concentration,15–19 salt valency and identity,20 temperature,16 chirality,21,22 pH,16,20 molecular weight,15,17 monomer sequence,23 and polymer architecture.24 These largely rely on turbidity as a proxy for phase separation since while phase separation is macroscopic, initial droplets are resistant to coarsening due to a low surface tension.25,26 Further investigations have characterized a number of physical properties in these same systems, such as chain conformation,27,28 bulk coacervation thermodynamics,29,30 bulk rheology,31,32 and surface tension.25,33,34 Coacervates are promising because they are affected by a large number of chemical and physical parameters; however, this also highlights the need for good theoretical models to aid the exploration of such a large parameter space.

Bulk coacervate studies have gone hand in hand with efforts to understand oppositely charged block copolyelectrolytes that assemble via coacervation. Building on longstanding efforts to develop coacervate-core micelles for drug delivery,7–9 both diblock and triblock coacervates self-assemble into a variety of (for example) BCC and hexagonally packed cylinder ordered morphologies.35–37 Initial phase behavior has been mapped for a few simple systems;35 however, there has yet to be a full physical understanding of coacervate-driven assembly akin to the developed theory of χ-driven block copolymer solution assembly.38–41 This is in part due to the rapidly evolving state of our understanding of bulk coacervates,4 but also due to a still-nascent understanding of the interface between coacervate and supernatant. Indeed, existing theory on standard χ-driven assembly relies on classic work on polymer-polymer interfaces;40–43 however, there have so far been only a few attempts to understand this interface in coacervates.33,44

Most recent attempts to understand coacervate interfaces are built around the traditional theoretical understanding of complex coacervation, known as Voorn-Overbeek theory.15,33,45,46 The free energy of the system is given by

(1)

where α is an electrostatic interaction parameter, σi is the charge density of species i (polycation, polyanion, cation, anion, or water), and ϕi is the volume fraction of species i. The first term on the right side of the equation accounts for the mixing entropy for each species. Electrostatic interactions are captured in the second term using the results of Debye-Hückel theory,47 and the third term accounts for short-range interactions via the Flory χ-parameter.48 This free energy has been used to determine phase diagrams,45 which can match some experimental phase diagram measurements.15 Likewise, experimental measurements of interfacial tension exhibit scaling behaviors predicted from Voorn-Overbeek-based calculations by Qin et al.33 Despite this success, it is widely understood that Voorn-Overbeek theory does not adequately capture the physics of coacervation.3,4,49–52 This has long been appreciated because this theory (1) uses Debye-Hückel theory that is only accurate in the low-salt limit53 and (2) neglects polymer connectivity,4,49–51 instead of treating salt charges and polymer charges equivalently. Recent work by the authors and others has demonstrated using simulation and theory that these effects are important50,51,54–56 and that matching to Voorn-Overbeek is due to a fortuitous cancellation of errors due to the neglect of both the charge connectivity in the polyelectrolytes and the finite size of the charges.18,49 There have been a number of efforts to systematically incorporate these effects using more sophisticated field theories, such as the random phase approximation (RPA) and field theoretic simulations.44,51,54–60 However, these techniques still have limitations; recent particle-based simulation results have shown that local charge correlations play a primary role in coacervation.18 This limitation can be overcome by either (1) using fully fluctuating fields using, e.g., complex Langevin methods44,59,60 or (2) using theory or simulations in order to resolve local charge correlations and thus inform field theories.61 The first approach has resulted in investigations using fully fluctuating field theories,44,59 and in this article, we will pursue the second approach.

Recent studies by the authors have developed a new theoretical approach to coacervation that is inspired by the concept of counterion release, which is a widely used concept in the pair complexes of dilute polyelectrolytes.62–64 Here, counterions condense onto polyelectrolytes to neutralize highly charge-dense polyelectrolytes at the cost of their translational entropy.65–67 When complexing polyelectrolytes condense onto one another, the counterions are released increasing their translational entropy, driving complex formation.62,63 This concept has been included into a number of coacervate models,68–70 for example, via “doping” of solid complexes19,30 or in combination with Voorn-Overbeek arguments.50 Recently, the authors have developed a transfer matrix (TM) theory that formalizes these ideas by mapping complex coacervation onto a one-dimensional adsorption model.71,72 The substrate is a test polyelectrolyte chain, which can adsorb the various other oppositely charged species. The statistics of this adsorption can be parameterized by Monte Carlo (MC) simulation to yield a free energy expression71,72

(2)

Here, a transfer matrix M was developed that captures the Boltzmann factors of the potential adsorption states along a test polyelectrolyte of length N. The vector ψ1 sets the initial Boltzmann factors for the first monomer, and ψ0 is a vector of ones. Additional parameters include κ, which sets the magnitude of a phenomenological third-order term that captures excluded volume, and Λ, which reflects the difference in excluded volume between the polymer and monomer. This theory can nearly quantitatively match simulation results with reasonable choices for fitting parameters71 and shows consistency between the local adsorption statistics (i.e., the local charge correlations) as well as the phase behavior.71 This model can be straightforwardly adapted to reflect changes in molecular features such as charge spacing, polyelectrolyte rigidity and architecture, solvent dielectric constant, and salt valency.71,72

In this article, we start by describing a range of theoretical and computational methods (Sec. II), demonstrating how we use both Monte Carlo (MC) simulation and our recent transfer matrix (TM) approach to inform self-consistent field theory (MC-SCFT61 and TM-SCFT). In Sec. III A, we show that both TM-SCFT and MC-SCFT can quantitatively match both coacervate bulk thermodynamics and the structure of the interface seen in particle-based molecular dynamics (MD) simulations, demonstrating the efficacy of field theory techniques developed by the authors.61 We subsequently demonstrate in Sec. III B that the two SCFT-based methods, informed by theory or simulation, exhibit the same interfacial thermodynamics; this includes the surface tension, interfacial width, and the surface excess of salt ions. In Secs. III C–III D, we modify both MC-SCFT and TM-SCFT to include a neutral polymer and to understand interfaces similar to those found in block copolyelectrolytes. We show that increasing neutral polymer density drastically increases the immiscible region of the phase diagram due to the interplay between the excluded volumes of the polymeric species. Indeed, this can lead to large regions of small, but positive, surface tension; this understanding is crucial to understand the thermodynamics of coacervate-driven assembly.

We use a combination of theoretical and computational methods to study complex coacervate interfaces, in part to explore their limitations and capabilities. Specifically, the MD simulations require the fewest assumptions because all chains and salt ions are explicitly represented; however, this comes at a significant computational cost. SCFT is far more efficient, but comes with the trade-off that all chains are treated at a mean-field level. However, we use methods (both TM and MC) that provide information about the local charge correlations that are crucial to understanding coacervate thermodynamics.49,73 TM and MC provide different routes to this correlation information, with MC accounting for charge correlations directly from simulation18,24,61 while TM reflects a theoretical model developed by the authors.71,72 We outline all three methods here, leaving details for  Appendixes A and  B.

There are approximations present in all of our methods. In particular, all methods consider the same restricted primitive model (RPM) representation of all the charges in the system.74 The RPM is a standard approximation that removes atomistic detail, leaving charged species and their corresponding hydration shells as charged beads in an implicit solvent medium.73,74 The charged beads represent both the small-molecule ions as individual beads and polyelectrolytes as connected chains of beads. By neglecting atomistic detail and solvent degrees of freedom, we are unable to describe specific ion and Hofmeister effects,75 as well as any water structure effects that can become important at high polymer concentrations.76–78 These effects are largely thought to prevent quantitative, but not qualitative matching to experiment, and indeed most comparisons between theory and simulation do not rely on these details.4 This model also neglects dielectric effects,79–85 which also may play a role in the thermodynamics of coacervation. We similarly do not expect this to play anything more than a quantitative role.

We consider a system composed of polyelectrolytes, neutral polymers, salt, and water and use SCFT methods standard in the literature.86–90 To simplify our SCFT calculation, we assume both local electroneutrality and charge stoichiometry between both polymers and ions; therefore, we only need one polyelectrolyte species representing both polycations and polyanions and one salt species representing both cations and anions. This is a standard assumption common to most coacervate models and is used to simplify experiments;3,4,45 this can in principle be relaxed.71 For this paper, the degree of polymerization is N = 100 and is the same for both the polyelectrolyte and the neutral polymer. Species-specific quantities are denoted with subscripts; A for polyelectrolytes, B for neutral polymers, S for salt, and W for water. We thus define a set of volume fractions ϕi(x), where i = A, B, S, W that is a function of a spatial position x.

The homogeneous free energy F of a coacervate system can be written as51,61

(3)

Here, we introduce the per-volume free energy f. The first four terms on the right side account for the translational entropies of all the species. The fifth term is an excess free energy that accounts for all remaining contributions to F and is generally a function of the set of volume fractions ϕj. fEXC{ϕj} is the excess free energy, which accounts for the local charge interactions that drive coacervation. We will describe how to obtain explicit forms for values and derivatives of fEXC{ϕj} in Secs. II B and II C as well as  Appendix A (for MC) or  Appendix B (for TM).

The homogeneous free energy F in Eq. (3) corresponds to a Hamiltonian functional H,90 

(4)

The Hamiltonian is a functional of the position-dependent density fields ϕj(r) and auxiliary fields ωj(r) for all of the species.90,ρ0 is the bulk number density. The single chain partition functions QA and QB, along with the single particle partition functions QS and QW, are functionals of their respective auxiliary fields j(r).90 The first term in the integral is the local excess free energy, to be calculated from MC or TM, and the second term in the integral ensures that the sum of the volume fractions is 1 with a coefficient ζ that is set to be large. In writing this Hamiltonian for inhomogeneous coacervate systems, we introduce an ansatz that a homogeneous value of the excess free energy fEXC({ϕj}) can be used locally at points r, leading to a position-dependent fEXC({ϕjr}). This term is thus approximated as a local contribution to the Hamiltonian, based on the observation that coacervation takes place at high salt/polymer concentrations where we expect local charge correlations to dominate.15,18,52 We do not justify this “local homogeneity” ansatz a priori because electrostatic correlations in MC simulations are longer-range than the field theory grid spacing; we will instead justify a posteriori through comparison to particle-based MD. This theory thus differs from alternative field theoretic formalisms that directly include the Coulomb interaction in the Hamiltonian,44,59,91 simplifying the SCFT calculation but introducing an assumption that we can use a free energy associated with charge correlations while neglecting how they may be affected by concentration gradients that arise in inhomogeneous polymer systems.42,90

The goal of SCFT is to find the “saddle point” or sets of fields ϕj*(r) and ωj*(r) where the Hamiltonian is at an extremum,90 

(5)

These functional derivatives with respect to the auxiliary field lead to the expression for the density in terms of the propagators,86–90 

(6)

For the polymer components, this becomes86–90 

(7)

For the water and salt components, this becomes86–90 

(8)

Here, we have defined a field Wj = j, and the quantity qj(r, s; [Wj]) in Eq. (7) is a propagator that represents the probability that a sub-chain of length s has its terminal chain end located at a position r in a field Wj. These are calculated via the diffusion equation, which in one dimension (rx),86–90 

(9)

We also consider the functional derivatives of the Hamiltonian [in Eq. (5)] with respect to the density fields,86–90 

(10)

The partial derivative is the local excess chemical potential μj,EXC(ϕi(r)), calculated with respect to the number density ρj of species j. This leads to the result86–90 

(11)

In this paper, the value of μj,EXC(ϕi(r)) is determined from either MC or TM using the set of values ϕi(r). A self-consistent scheme is used to calculate the saddle point. This consists of calculating the fields Wj(r) using the species densities ϕj(r) via Eq. (11). These fields are then used to calculate the propagator for the polymer species qjx,s;[Wj] in Eq. (9) and subsequently the species densities ϕj(r) using Eqs. (7) and (8). These densities can once more be used to calculate Wj(r), and this process repeats until all of the fields have converged to their saddle point values Wj*(r) and ϕj*(r). For this paper, we use the value ζ = 1000kBT.

We can use simulation to obtain fEXC so that detailed charge correlation information is incorporated into our SCFT calculations. In this work, we use a standard Monte Carlo (MC) simulation based on the restricted primitive model (RPM) of charged polyelectrolyte systems.18,23,24,61 The RPM considers all charges to be hard spheres in an implicit solvent, which interact through a Coulomb potential. Polyelectrolyte charges are connected via bond and angle potentials. In this article, we call the hybrid method MC-SCFT; related methods have been used extensively in prior work by the authors,23,24,61 so we relegate the model in detail to  Appendix A. The key feature of these simulations is that standard Widom insertion methods can be used to determine the excess chemical potential,61,92,93

(12)

Thermodynamic integration of the excess chemical potential enables calculation of fEXC, which in turn serves as an input to SCFT.

We can also use theory to obtain fEXC, and in particular, we use the transfer-matrix (TM) theory developed by the authors.71,72 This model maps the local charge environment around a test chain to a one-dimensional adsorption model.71,72 This is motivated by the correlations observed in simulation, which exhibit prominent neighbor peaks in an otherwise highly charge screened system.18,23 Monomer “sites” on the test chain can adsorb oppositely charged species, including both the oppositely charged polyelectrolyte or the counterion and occasionally may remain without an adsorbed species. A transfer matrix M can be written to account for the Boltzmann weight associated with the local environment around each monomer and can be used to calculate a grand canonical partition function that is related to the interaction free energy.71 With the inclusion of a phenomenological cubic term that accounts for excluded volume, the analytical expression for fEXC is71,72

(13)

Here, the function Θ is71,72

(14)

We have used the analytical expression for the interaction free energy, where Θ is the largest eigenvalue of M and the quantities A0, B0, and ϵ̃ can be parameterized directly from molecular simulation.71 We use this form for this paper, but more generally it is possible to numerically calculate the interaction free energy via Eq. (2) if M is no longer analytically tractable.72 See  Appendix B for a more detailed discussion.

Molecular dynamics (MD) simulations were performed on a system consisting of nA+ polycations and nA polyanions, modeled as connected monovalently charged beads with diameter σ. Both polymers have the same degree of polymerization, N. nS+ cations and nS anions are modeled as monovalently charged beads, also with a diameter σ. Water is included as an implicit solvent with a relative dielectric constant, ϵr = 78.5. These simulations were performed using the Large-scale Atomic/Molecular Massively Parallel Simulator (LAMMPS) with a Langevin thermostat.94 See Fig. 1 for a schematic of our model. The overall potential energy is given by

(15)

UE models the electrostatic interactions,

(16)

where qi is the charge on bead i, ϵ0 is the permittivity of free space, and rij is the separation between beads i and j. Standard Ewald summation is used to account for the long-range interactions under periodic boundary conditions.92 The polymers are bound together with a bond potential UB,

(17)

For this contribution, κB is the strength of the bond potential and r0 is the equilibrium bond distance. In addition to the bond potential, the polymers have a bending potential Uθ,

(18)

Here, κθ is the strength of the bending potential, θi,i−1,i−2 is the angle between the two bond vectors, and θ0 is the equilibrium angle between bond vectors. The excluded volume of the beads is modeled by a Lennard-Jones potential ULJ,

(19)

where ϵLJ is the depth of the potential well, σLJ is the interparticle separation at which ULJ becomes zero, and rc is the cutoff distance of this potential.92 

FIG. 1.

Schematic demonstrating the features of both our MD and MC simulation models. All species are represented as beads of diameter σ (including the neutral polymer beads, green), and charged species interact via a Coulomb potential UE. Polymer charges are connected by a bonding potential UB and have a bending potential Uθ. Only minor differences exist between the MD and MC models; in particular, MC has a hard-sphere potential that keeps charges from overlapping, while MD uses a corresponding Lennard-Jones potential parameterized so that both simulations exhibit nearly the same pair correlation functions (see Fig. 2). There are also minor differences in the bonding potential UB.

FIG. 1.

Schematic demonstrating the features of both our MD and MC simulation models. All species are represented as beads of diameter σ (including the neutral polymer beads, green), and charged species interact via a Coulomb potential UE. Polymer charges are connected by a bonding potential UB and have a bending potential Uθ. Only minor differences exist between the MD and MC models; in particular, MC has a hard-sphere potential that keeps charges from overlapping, while MD uses a corresponding Lennard-Jones potential parameterized so that both simulations exhibit nearly the same pair correlation functions (see Fig. 2). There are also minor differences in the bonding potential UB.

Close modal

For this simulation, the degree of polymerization is kept at N = 100. We choose parameters consistent with our previous efforts;18,61κB = 250kBT, r0 = 1.05σ, κθ = 1.65kBT, and θ0 = π. Lennard-Jones parameters to model excluded volume were ϵLJ = 10.75kBT, σLJ = σ, and rc = σ. This mimics the hard-sphere potential in the MC simulations by having the interparticle excluded volume potential at zero until the particles overlap. The bead diameter σ was taken as 4.25 Å.

We use MD simulations and SCFT to capture the physics of this coacervate system, with the latter informed by both MC simulations and TM theory. We can show that all methods, while they provide different levels of resolution and assumptions, exhibit nearly quantitative matching for interfacial properties. Importantly, we motivate our use of MC-SCFT and TM-SCFT by comparing with a completely particle-based simulation. To ensure that we have comparable models, we show that the MC used to inform the SCFT yields the same radial distribution functions when compared to the full-particle MD simulations. Figure 2 shows radial distribution functions from both methods (MC and MD) using nA+ = nA = 6 in a cubic simulation box for ϕA = 0.08 and ϕA = 0.12 [Figs. 2(a) and 2(b), respectively]. Both MC and MD simulations agree nearly quantitatively. Some differences do occur due to differences in the interaction potentials; the MC uses hard-core interactions that are not possible in MD simulations, where a Lennard-Jones potential is used instead. These disparities are apparent at the contact peak near r/σ = 1, where the differences in these potentials are most pronounced.

FIG. 2.

Radial distribution functions (RDF), gijσ, as a function of separation, r/σ, from Monte Carlo (MC, solid lines) simulations and molecular dynamics (MD, dashed lines) simulations. (a) RDF at ϕA = 0.12 and ϕS = 0.02. (b) RDF at ϕA = 0.08 and ϕS = 0.08. The green line is cation-anion (S+S−) pairs, the blue line is polycation-polyanion (A+A−) pairs, the orange line is polycation-anion (A+S−) pairs, the pink line is cation-cation (S+S+) pairs, the black line is polycation-cation (A+S+) pairs, and the red line is polycation-polycation (A+A+) pairs. Qualitatively, the MC and MD simulations show the same structure. However, some differences are present due to the use of a Lennard-Jones potential for excluded volume interactions as opposed to a hard sphere potential.

FIG. 2.

Radial distribution functions (RDF), gijσ, as a function of separation, r/σ, from Monte Carlo (MC, solid lines) simulations and molecular dynamics (MD, dashed lines) simulations. (a) RDF at ϕA = 0.12 and ϕS = 0.02. (b) RDF at ϕA = 0.08 and ϕS = 0.08. The green line is cation-anion (S+S−) pairs, the blue line is polycation-polyanion (A+A−) pairs, the orange line is polycation-anion (A+S−) pairs, the pink line is cation-cation (S+S+) pairs, the black line is polycation-cation (A+S+) pairs, and the red line is polycation-polycation (A+A+) pairs. Qualitatively, the MC and MD simulations show the same structure. However, some differences are present due to the use of a Lennard-Jones potential for excluded volume interactions as opposed to a hard sphere potential.

Close modal

The phase behavior of a coacervate with no neutral polymer is shown in Fig. 3(a) (calculated from MC-SCFT), showing the two-phase coacervation regime at low salt ϕS and polymer ϕA concentrations. Sloped tie lines are observed connecting the large-ϕA coacervate phase at a slightly lower ϕS than the low-ϕA supernatant phase. This is due to the partitioning of salt to the supernatant, which has been demonstrated by the previous computational, theoretical, and experimental work.18,30,49,71,72 We also show MD simulation snapshots from phase-separating mixtures along the indicated tie lines, with the orange and blue polymers representing the polycations and polyanions, respectively. This structure becomes more diffuse as salt is increased. This is apparent in Figs. 3(b)–3(g), where we plot the interfacial profiles (ϕi versus distance x/σ) that form along these same tie lines [Fig. 3(a)]. These plots include full-particle MD simulations, along with interfacial profiles calculated with both MC-SCFT and TM-SCFT. MD simulations were performed with nA+ = nA = 15 in a rectangular prism simulation box with the x-dimension 3 times the length of the y- and z-dimension. Starting volume fractions for each simulation were taken as the midpoint of the tie lines in Fig. 3(a).

FIG. 3.

(a) The phase diagram in the ϕS versus ϕA plane, determined from MC-SCFT calculations. This phase diagram is consistent with prior theoretical and computational efforts.71 The labeled tie lines correspond to MD simulation snapshots shown on the right, showing that phase separation occurs within the simulation box. (b)–(g) To quantify this phase separation behavior, we show a comparison of interfacial profiles by plotting ϕi versus distance x/σ from MD simulations (blue points), MC-SCFT (black lines), and TM-SCFT (red lines). Filled points and solid lines correspond to the volume fraction of polyelectrolyte ϕA, and open points and dashed lines correspond to the volume fraction of salt ϕS. MD interfacial profiles have been shifted along the x-axis to overlay the interfaces with SCFT calculations. Nearly quantitative matching is demonstrated between all three techniques for the tie lines considered.

FIG. 3.

(a) The phase diagram in the ϕS versus ϕA plane, determined from MC-SCFT calculations. This phase diagram is consistent with prior theoretical and computational efforts.71 The labeled tie lines correspond to MD simulation snapshots shown on the right, showing that phase separation occurs within the simulation box. (b)–(g) To quantify this phase separation behavior, we show a comparison of interfacial profiles by plotting ϕi versus distance x/σ from MD simulations (blue points), MC-SCFT (black lines), and TM-SCFT (red lines). Filled points and solid lines correspond to the volume fraction of polyelectrolyte ϕA, and open points and dashed lines correspond to the volume fraction of salt ϕS. MD interfacial profiles have been shifted along the x-axis to overlay the interfaces with SCFT calculations. Nearly quantitative matching is demonstrated between all three techniques for the tie lines considered.

Close modal

All 3 techniques exhibit nearly quantitative matching in all interfacial profiles shown. This a posteriori justifies the approximations we made in the TM-SCFT and MC-SCFT, which is that we have a separation of length scales between the charge correlations (which are treated as a homogeneous system at each SCFT grid point) and the polymer length scales (which span multiple SCFT grid points).61 We do note that correlations in Fig. 2 have a similar range to the grid spacing used in our simulations (0.75σ), which is set relatively small to resolve the interfacial structure. However, this correlation length scale seems to still be sufficiently small compared to the length scales of variation in ϕA and ϕS such that square gradient corrections to fEXC are unnecessary to match SCFT with MD. The increasingly diffuse interface at higher salt concentrations poses a practical challenge for the full-particle MD, where it is difficult to obtain a good average of the interface; this is why we focus on tie lines far from the critical point. We thus demonstrate that MC-SCFT and TM-SCFT capture the correct interfacial profile, motivating our use of these field-based methods to calculate interfacial properties such as interfacial tension, interfacial width, and the surface excess of salt.

We can use the results of both MC-SCFT and TM-SCFT to determine the interfacial tension of a coacervate. We use the expression95–99 

(20)

We have used tildes to denote normalization of energy scales by kBT and length scales by σ. The second and third terms in the integrand are square gradient expressions for the polyelectrolyte and the neutral polymer, approximated using the result from the Random Phase Approximation (RPA).98,99 The first term in the integrand is the free energy of a homogeneous system with the volume fractions at a point x relative to the bulk free energy of the system, given by the expression96,97

(21)

Here, μi is the chemical potential of species i in the bulk phase.

We show a calculation of this interfacial tension γ̃ for a coacervate without the presence of a neutral polymer, shown in Fig. 4(a). Consistent with the previous experimental and theoretical literature,25,26,33,34,44 we plot the decrease in γ̃ as a function of salt concentration, where we choose the supernatant salt concentration ϕSβ. We will typically denote phases by a superscript α (coacervate) and β (supernatant). Indeed, we demonstrate that both MC-SCFT and TM-SCFT exhibit nearly identical values of γ̃, with only a small horizontal offset of the γ̃ versus ϕSβ curve. We can also show that our results exhibit the scaling law described by Qin et al.,33 for the interfacial tension γ̃/γ̃0(1ϕSβ/ϕScrit)3/2 as ϕSβ approaches its critical point value ϕScrit (Fig. 5). Here, interfacial tension is normalized by its zero-salt value, γ̃0. Qin demonstrated matching between this prediction and with experimental data25,26,33,44 despite the use of Voorn-Overbeek theory in the derivation of the scaling behavior. Our results are consistent with the Qin prediction, showing a 3/2 slope in a log-log plot in Fig. 5 for both TM-SCFT and MC-SCFT; our results are also consistent with the experimental results from the literature,25,26,33,44 which are also included in Fig. 5 as open symbols. We note that these results are based on mean-field SCFT models; a different power law will likely be observed as the critical point is approached, which would require beyond-mean-field polymer field theories.

FIG. 4.

Interfacial properties for the coacervate-forming system calculated using both MC-SCFT (black) and TM-SCFT (red) as a function of supernatant salt concentration, ϕSβ. Lines are guides for the eye. (a) The interfacial tension decreases with increasing salt concentration. (b) The interfacial width increases with increasing salt concentration. (c) The interfacial excess of salt initially increases with increasing salt concentration, but at large enough salt concentrations, the interfacial excess of salt decreases. Observed trends are seen using both MC-SCFT and TM-SCFT. Both techniques qualitatively agree.

FIG. 4.

Interfacial properties for the coacervate-forming system calculated using both MC-SCFT (black) and TM-SCFT (red) as a function of supernatant salt concentration, ϕSβ. Lines are guides for the eye. (a) The interfacial tension decreases with increasing salt concentration. (b) The interfacial width increases with increasing salt concentration. (c) The interfacial excess of salt initially increases with increasing salt concentration, but at large enough salt concentrations, the interfacial excess of salt decreases. Observed trends are seen using both MC-SCFT and TM-SCFT. Both techniques qualitatively agree.

Close modal
FIG. 5.

Log-log plot showing the scaling of interfacial tension γ̃/γ̃0 versus 1ϕSβ/ϕScrit using both MC-SCFT and TM-SCFT. This exhibits the scaling predicted by Qin et al.,33γ̃/γ̃0(1ϕSβ/ϕScrit)3/2, with the black fit line showing the exponent to be 1.50 ± 0.01. For MC-SCFT, ϕScrit=0.0997, and for TM-SCFT, ϕScrit=0.1065. We also plot the experimental data (open symbols) used in the work of Qin et al.,33 compiled from the studies of Priftis et al. (magenta),26 Riggleman et al. (green),44 and Cohen Stuart et al. (blue).25 Our predictions are consistent with these experimental results.

FIG. 5.

Log-log plot showing the scaling of interfacial tension γ̃/γ̃0 versus 1ϕSβ/ϕScrit using both MC-SCFT and TM-SCFT. This exhibits the scaling predicted by Qin et al.,33γ̃/γ̃0(1ϕSβ/ϕScrit)3/2, with the black fit line showing the exponent to be 1.50 ± 0.01. For MC-SCFT, ϕScrit=0.0997, and for TM-SCFT, ϕScrit=0.1065. We also plot the experimental data (open symbols) used in the work of Qin et al.,33 compiled from the studies of Priftis et al. (magenta),26 Riggleman et al. (green),44 and Cohen Stuart et al. (blue).25 Our predictions are consistent with these experimental results.

Close modal

We consider other interfacial properties, such as the interfacial width D. D is defined using the interfacial profile for the polyelectrolyte,95 

(22)

Here, the derivative in the denominator is taken at the midpoint concentration between the polyelectrolyte concentrations in the coacervate phase ϕAα and the supernatant phase ϕAβ. We plot D/σ as a function of ϕSβ for both MC-SCFT and TM-SCFT in Fig. 4(b). Similar to Fig. 4(a), we again see mainly small differences in this quantity between the two methods. The exception to this comes at large values of ϕSβ, which is close to the critical point such that D/σ becomes large. The curves do not become large at the same rate, consistent with Fig. 4(a) which shows that the critical value of ϕSβ appears to be shifted. The physical observation that D increases is consistent with traditional polymer theory,42,43,98,100 where a decrease in the driving force for phase separation broadens the interface. In non-charged systems, this is usually controlled by temperature; however we are instead weakening the electrostatic interaction by using salt concentration ϕSβ instead.

This system has effectively 3 species, the polyelectrolytes, salt ions, and water; thus, we can calculate the surface excess of salt at the interface as the overall salt concentration is increased. The surface excess of any species i can be calculated via the relationship96,97

(23)

Here, D is defined in Eq. (22) and Dα is the size of the phase α. This definition, however, depends on the choice of interface position; a related value is independent of this choice,96,97

(24)

This value of Γi(j) describes the surface excess of i, using a reference of ji such that it does not depend on the choice of ji. This physically represents the deviation in excess ϕi [captured by the first term in Eq. (24)] away from any excess ϕj calculated from the normalized interfacial structure in the other species j [the second term in Eq. (24)]. We plot ΓS(A), the surface excess of salt, as a function of ϕSβ in Fig. 4(c).96,97 We once more observe only small differences between MC-SCFT and TM-SCFT and show that the surface excess approaches ΓS(A)=0 at the limits of ϕSβ (where there is no salt) and at the critical value of ϕSβ where the system becomes homogeneous. At intermediate values, the surface excess reaches a maximum, indicating that salt preferentially partitions to the interface. The surface excess salt remains quite small, however, with ΓS(A)7×104 indicating that any deviations of ϕS from what would be expected from the ϕA profile are between one or two orders of magnitude smaller than the overall variations in ϕS.

Interfacial tension plays a significant role in the classical theoretical work on the self-assembly of block copolymers.40–43,100,101 The form of the interfacial free energy follows that of a simple polymer-polymer interface and is driven by a positive short-range Flory χ parameter.42,43,90,98 Self-assembled structures formed due to coacervation are driven by long-range electrostatic attractions that are not well represented by χ;102 instead, we use the analogous interface between a coacervate and a neutral polymer to determine the interfacial tensions that would arise in the coacervate-driven self-assembly. Furthermore, the presence of a neutral polymer has implications for biological coacervates, which are often in crowded macromolecular environments.103,104 Recent reports from experiment demonstrate that “crowding” can have a significant impact on the formation of bio-inspired coacervates.104 

First, we show that the addition of a neutral polymer species significantly affects the phase diagram depending upon the concentration of the neutral polymer [Figs. 6(a) and 6(b)]. To plot the effect of the neutral polymer, we note that it primarily partitions into the supernatant phase. We thus adjust the neutral polymer concentration by fixing its value in this phase ϕBβ. The corresponding neutral polymer concentration in the coacervate phase ϕBα is typically negligible, except near the critical point. We thus emphasize that Fig. 6 is the projection of a three-dimensional phase diagram onto the ϕS versus ϕA plane.

FIG. 6.

Phase diagrams and salt partitioning including a neutral polymer species. Lines are a guide to the eye. (a) MC-SCFT phase diagram as a function of polyelectrolyte volume fraction ϕA and salt volume fraction ϕS for a number of supernatant neutral polymer volume fractions ϕBβ. (b) TM-SCFT phase diagram as a function of polyelectrolyte volume fraction ϕA and salt volume fraction ϕS for a number of supernatant neutral polymer volume fractions ϕBβ. Qualitatively, both methods show a large increase in the immiscible region with larger increases corresponding to more supernatant neutral polymer volume fraction. However, TM-SCFT does not show the same critical point behavior as MC-SCFT. This is probably due to the use of a simple cubic term for excluded volume, which does not fully capture the complex interplay of all species excluded volume. (c) We demonstrate the importance of this excluded volume term, by removing the effect of the excluded volume effect for the B-polymer. We see that, without the finite excluded volume of B, only minor shifts in the phase boundary are observed with increasing ϕBβ. (d) Salt partitioning, λ=ϕSα/ϕSβ, as a function of supernatant salt volume fraction ϕSβ for MC-SCFT. (e) Salt partitioning as a function of ϕSβ for TM-SCFT. Differences between panels (d) and (e) are observed, which we attribute to the use of a third-order term to approximate the excluded volume in the TM theory. (f) Similar to (c), the removal of the finite excluded volume of the neutral B-polymer leads to only small changes in the partitioning of other components.

FIG. 6.

Phase diagrams and salt partitioning including a neutral polymer species. Lines are a guide to the eye. (a) MC-SCFT phase diagram as a function of polyelectrolyte volume fraction ϕA and salt volume fraction ϕS for a number of supernatant neutral polymer volume fractions ϕBβ. (b) TM-SCFT phase diagram as a function of polyelectrolyte volume fraction ϕA and salt volume fraction ϕS for a number of supernatant neutral polymer volume fractions ϕBβ. Qualitatively, both methods show a large increase in the immiscible region with larger increases corresponding to more supernatant neutral polymer volume fraction. However, TM-SCFT does not show the same critical point behavior as MC-SCFT. This is probably due to the use of a simple cubic term for excluded volume, which does not fully capture the complex interplay of all species excluded volume. (c) We demonstrate the importance of this excluded volume term, by removing the effect of the excluded volume effect for the B-polymer. We see that, without the finite excluded volume of B, only minor shifts in the phase boundary are observed with increasing ϕBβ. (d) Salt partitioning, λ=ϕSα/ϕSβ, as a function of supernatant salt volume fraction ϕSβ for MC-SCFT. (e) Salt partitioning as a function of ϕSβ for TM-SCFT. Differences between panels (d) and (e) are observed, which we attribute to the use of a third-order term to approximate the excluded volume in the TM theory. (f) Similar to (c), the removal of the finite excluded volume of the neutral B-polymer leads to only small changes in the partitioning of other components.

Close modal

The inclusion of the neutral polymer results in a large extension of the two-phase region at high ϕS and low ϕA, which extends further to the right as the concentration ϕBβ increases. This trend is observed in both MC-SCFT and TM-SCFT. We interpret this trend as due to the matching of pressure between the supernatant and coacervate phases; the coacervate phase in the absence of a neutral polymer is limited to low ϕA and ϕS due in part to the high pressure due to excluded volume in the polymer-dense coacervate.18 This phase would expand except this would dilute the combinatoric entropy of polymer-polymer interactions that primarily drives phase separation.71,72 With the addition of the neutral polymer, this polymer partitions to the supernatant and applies a counterpressure. When ϕAϕB, the pressures of the two phases balance essentially removing the pressure-based driving force for miscibility; this results in an extended two-phase region. To demonstrate that the changes in phase behavior are driven primarily by the excluded volume of the neutral polymer, we remove the ϕB contribution to the excluded volume in the TM calculation [we remove ϕB from the second term of Eq. (B5)]. The resulting phase diagram is plotted in Fig. 6(c) and exhibits only minor changes in the binodal as ϕBβ is increased, contrasting significantly with the phase diagrams seen in Figs. 6(a) and 6(b).

These observations are further supported by the examination of the salt partitioning λ=ϕSα/ϕSβ, where ϕSα is the volume fraction of salt in the coacervate phase and ϕSβ is the volume fraction of salt in the supernatant phase. This is plotted as a function of ϕSβ in Figs. 6(d) and 6(e), corresponding to parts a and b directly above. In the limit of no neutral polymer, the salt partitions such that the salt concentration in the supernatant is greater than that in the coacervate phase (λ < 1). This is consistent with the previous experiment and simulation,18,71,72,105 and we attribute this to the excluded volume of the polymer in the coacervate phase. Upon addition of a neutral polymer, the salt partitioning behavior is altered, in different ways depending on the value of ϕSβ. At low ϕSβ, salt partitioning becomes increasingly even between the two phases (λ → 1). This is likely due to the neutral polymer in the supernatant, which begins to equalize the excluded volume in the two phases as its concentration increases. This occurs below the critical salt concentration of the phase diagram in the absence of a neutral polymer; however at high ϕSβ, the opposite trend is observed. Here, larger values of ϕBβ lead to smaller values of λ. We attribute this “switch” to the large extent that the coacervate phase is increasingly pushed to large ϕA and ϕS, where excluded volume effects become enhanced. We also plot λ versus ϕSβ in Fig. 6(d) when excluded volume of the neutral polymer is removed from the TM calculation. Similar to Fig. 6(c), we find that only minor changes are observed when ϕBβ is increased, consistent with the argument that the excluded volume of the neutral polymer is what drives the marked differences seen in Figs. 6(d) and 6(e).

Our excluded volume argument is consistent with the differences between TM-SCFT and MC-SCFT, particularly with respect to the value of λ. While some aspects of the qualitative trends are similar, there are some differences, namely, the lack of a critical point in the region tested for TM-SCFT and the different magnitudes for the plots of λ. We attribute this to the phenomenological argument for the excluded volume term in Eq. (2), which is chosen to approximate the free energy associated with packing the species.71,72 We also note that Fig. 6(d) becomes non-monotonic at large values of ϕSβ, with an increase in partitioning with increasing ϕSβ at intermediate salt concentrations. This does not occur for the TM-SCFT predictions in Fig. 6(e). We note that this non-monotonicity occurs at a high-ϕAand high-ϕS region of the phase diagram in Figs. 6(a) and 6(b), where there are also different shapes in the binodal curves. We once more attribute these differences to the phenomenological excluded volume term in TM-SCFT, which should become significant in this region.

We choose a point in the phase diagrams in Fig. 6 to demonstrate that these large shifts in the phase diagram are indeed observed in the full MD simulation. We selected a value of ϕBβ=0.03 and average concentrations of ⟨ϕA⟩ = 0.025 and ⟨ϕS⟩ = 0.12 for all simulation/theory methods (MD, TM-SCFT, MC-SCFT). We plot the resulting volume fraction profiles in Fig. 7, demonstrating that all three methods exhibit near-quantitative matching. Indeed, this choice of salt and polymer concentrations will not phase separate without the presence of the neutral polymer. Snapshots at the bottom of Fig. 7 show the MD simulation corresponding to the interfacial profile. Both show the same system; however, the top snapshot includes the ions, and the bottom snapshot removes the ions to show the polymer components more clearly. Consistent with the expectations from the SCFT calculations, the neutral polymer (green) is almost completely in the supernatant phase and the polycation and polyanion (blue and orange) are completely in the coacervate phase.

FIG. 7.

Interfacial profiles for a system with neutral polymers plotted with the volume fractions ϕi as a function of x, for MD, MC-SCFT, and TM-SCFT. ϕBβ=0.03, ⟨ϕA⟩ = 0.025, and ⟨ϕS⟩ = 0.12, which in the absence of a neutral polymer does not undergo phase separation. We indeed observe phase separation, and our theoretical and simulation-informed SCFT models are capable of matching all-particle MD calculations nearly quantitatively. We show snapshots from simulations on the bottom, from the same simulation. In the top snapshot, salt is included, while in the lower snapshot, it is removed to show that the neutral polymer (green) has phase separated completely from the charged polymers (orange and blue).

FIG. 7.

Interfacial profiles for a system with neutral polymers plotted with the volume fractions ϕi as a function of x, for MD, MC-SCFT, and TM-SCFT. ϕBβ=0.03, ⟨ϕA⟩ = 0.025, and ⟨ϕS⟩ = 0.12, which in the absence of a neutral polymer does not undergo phase separation. We indeed observe phase separation, and our theoretical and simulation-informed SCFT models are capable of matching all-particle MD calculations nearly quantitatively. We show snapshots from simulations on the bottom, from the same simulation. In the top snapshot, salt is included, while in the lower snapshot, it is removed to show that the neutral polymer (green) has phase separated completely from the charged polymers (orange and blue).

Close modal

We use both MC-SCFT and TM-SCFT to determine how the interfacial tension γ̃ changes as ϕBβ is increased. These results correspond to the phase diagrams in Fig. 6 and are shown in Figs. 8(a) and 8(d) for the two different methods. For both sets of results, the interfacial tension monotonically increases with ϕBβ, including both above and below the critical ϕSβ at ϕB = 0. Interestingly, for small quantities of ϕBβ, the surface tension remains very close to γ̃0 for a large range of ϕSβ. This trend is also apparent in the interfacial width D [Figs. 8(b) and 8(e)], which exhibits significantly broader interfaces at small values of ϕBβ. After the initial addition of ϕBβ, the width of the interface begins to decrease to what appears to be a limiting value of D at high ϕBβ. Finally, an increase in ϕBβ drastically increases the surface excess of salt ΓS(A) [Figs. 8(c) and 8(f)], which we attribute to the increase of excluded volume in both the coacervate and supernatant phases. In this case, the salt preferentially partitions to the interface where excluded volume is minimized. This is directly observed in Fig. 7, where the profile shows a significant increase of salt at the interface for all methods (MD, MC-SCFT, and TM-SCFT). The calculated value of ΓS(A)3×103 is similar to the magnitude of the positive deviation in ϕS at the interface.

FIG. 8.

Interfacial properties calculated including a neutral polymer species. Lines are a guide to the eye. (a) The interfacial tension, γ̃, calculated using MC-SCFT as a function of supernatant salt volume fraction ϕSβ for a number of supernatant neutral polymer volume fractions ϕBβ. With increasing ϕSβ, the interfacial tension decreases, consistent with the trends in Fig. 4. With increasing ϕBβ, the interfacial tension increases due to an increase in the density of interactions. (b) The interfacial width, D/σ, calculated using MC-SCFT. The width increases with increasing ϕSβ because the phase separation is weaker making the interface more diffuse. Increasing ϕBβ causes the width to decrease. (c) The interfacial excess of salt, ΓSA, calculated using MC-SCFT. This is a non-monotonic function of ϕSβ. Increasing ϕBβ causes an increase in the amount of excess salt at the interface. This is due to the supernatant phase being rich in the neutral polymer causing both phases to have a large excluded volume. Salt adsorbs to the interface where there is excluded volume. (d) The interfacial tension, γ̃, calculated using TM-SCFT. (e) The interfacial width, D/σ, calculated using TM-SCFT. (f) The interfacial excess of salt, ΓSA, calculated using TM-SCFT. Qualitatively, MC-SCFT and TM-SCFT show similar trends for all of the interfacial properties. However, differences between the two methods are observed at higher ϕBβ due to the approximate form of the excluded volume in the TM theory.

FIG. 8.

Interfacial properties calculated including a neutral polymer species. Lines are a guide to the eye. (a) The interfacial tension, γ̃, calculated using MC-SCFT as a function of supernatant salt volume fraction ϕSβ for a number of supernatant neutral polymer volume fractions ϕBβ. With increasing ϕSβ, the interfacial tension decreases, consistent with the trends in Fig. 4. With increasing ϕBβ, the interfacial tension increases due to an increase in the density of interactions. (b) The interfacial width, D/σ, calculated using MC-SCFT. The width increases with increasing ϕSβ because the phase separation is weaker making the interface more diffuse. Increasing ϕBβ causes the width to decrease. (c) The interfacial excess of salt, ΓSA, calculated using MC-SCFT. This is a non-monotonic function of ϕSβ. Increasing ϕBβ causes an increase in the amount of excess salt at the interface. This is due to the supernatant phase being rich in the neutral polymer causing both phases to have a large excluded volume. Salt adsorbs to the interface where there is excluded volume. (d) The interfacial tension, γ̃, calculated using TM-SCFT. (e) The interfacial width, D/σ, calculated using TM-SCFT. (f) The interfacial excess of salt, ΓSA, calculated using TM-SCFT. Qualitatively, MC-SCFT and TM-SCFT show similar trends for all of the interfacial properties. However, differences between the two methods are observed at higher ϕBβ due to the approximate form of the excluded volume in the TM theory.

Close modal

For all interfacial properties in Fig. 8, MC-SCFT and TM-SCFT both capture the same qualitative trends. However, there are differences connected to the lack of critical points in TM-SCFT seen in Fig. 6. This manifests as a decrease in the interfacial tension γ̃ and interfacial excess of salt ΓS(A) to 0, along with a sharp increase in D for the MC-SCFT results. Similar to the phase diagram, we primarily attribute these differences to our approximate treatment of the excluded volume in the TM theory.71,72

In this manuscript, we show that we can use a range of computational and theoretical routes to understand the interfacial properties of polymeric complex coacervates. We can directly simulate interfaces using molecular dynamics, using a restricted primitive model representation that considers all polymeric and salt species as explicit particles. We match these results to SCFT predictions, which capture charge correlations using either a Monte-Carlo based method (MC-SCFT) or a transfer matrix theory of coacervation (TM-SCFT); we demonstrate nearly quantitative matching among the interfacial profiles of all three methods, as well as matching between the pair correlation functions in the MC and MD simulations. We subsequently use both MC-SCFT and TM-SCFT to probe the interfacial thermodynamics of coacervates, demonstrating trends in the interfacial tension γ consistent with previous observations in the literature.25,26 Similarly, we provide predictions for the interfacial width D and surface excess of salt ΓS(A).

We subsequently consider the effect of adding a neutral polymer due to its relevancy for the interface of self-assembling coacervate block copolymers. The coacervation phase behavior changes drastically with the neutral polymer partitioning strongly to the supernatant phase and “balancing” the excluded volume pressure of the coacervate phase. This results in a large increase in the region of coacervation, seen in both MC-SCFT and TM-SCFT calculations. This change in the phase behavior is commensurate with large changes in the interfacial properties, with a large region of low-γ observed upon initially including the neutral polymer. There are corresponding changes in interfacial width D and significant increases in the surface-excess salt ΓS(A) that are due to the presence of significant excluded volume in both phases.

Nearly quantitative agreement with full-particle MD, as well as field theories informed by both simulation and theory, provides insight into the important features needed to understand the physics of coacervation. First, the separation of length scales between the polymer conformation and the charge correlations is reflected in the SCFT calculations and does not seem to adversely affect the predictive power of the theory. Second, quantitative prediction hinges on the accuracy of the excluded volume model, which plays a large role near the critical point when there is a neutral polymer present; small differences in this term lead to large changes in the observed phase behavior. Finally, the complexity of the phase behavior when the neutral polymer is present poses a challenge for developing intuitive scaling arguments; in contrast to χ-driven phase separation,42,43,98 it will likely be difficult to distill quantities like γ down to compact scaling expressions.

While the agreement between the different methods is excellent, we note the limitations of our current approach. In particular, all of the polymers considered possessed a high charge density. This is crucial for the transfer matrix theory, which assumes that most of the charges along the chain are matched with a condensed counterion or an oppositely charged chain.71 It may be possible to extend the theory to low charge densities; however in these limits, we approach regions where field theoretic approaches may be more apt.33,44,51,54–58,106 Nevertheless, we use realistic parameters in our model that are roughly consistent with standard polyelectrolytes used in complex coacervates and have exhibited qualitative matching to experiment in a number of circumstances.18,23,24 We further note that all methods used in this paper (MC-SCFT, TM-SCFT, and MD) in principle have challenges in describing systems where both ϕS and ϕA are small. For MD and MC, we can easily go to low salt concentrations but do not see a large enough simulation box to explicitly capture the polymer concentrations seen on the low-ϕA branch of the binodal curve. We instead rely on the extrapolation of fEXC from finite concentrations of A to this limit. Similarly, TM does not consider changes in polymer conformations that are known to play a role in polyelectrolyte thermodynamics at low polymer concentrations.56,68,106,107 These may limit the accurate determination of the low-ϕA branch of the binodal curve, but we expect this to have little quantitative effect on our predictions of the high-ϕA branch.

We anticipate that this work will inform the further development of an understanding of coacervates, in particular, the self-assembly of coacervate-based block copolymers. Models of block copolymer self-assembly incorporate surface free energies via interfacial tension γ,40 and we have provided an initial picture of how this term would behave in the case of block copolyelectrolyte solution self-assembly.

This material is based upon work supported by the National Science Foundation under Grant No. DMR-1654158. The authors also acknowledge helpful conversations with Sarah L. Perry.

MC simulations were performed on systems with the same species as in the MD simulations, shown in Sec. II D. However, neutral polymer chains, nB with the same degree of polymerization as the polyelectrolytes was added in order to model its effect on the phase separation and interfacial profiles of complex coacervates. The total potential energy in these simulations includes an electrostatic potential, a bonding potential, a bending potential, and a hard-sphere potential,

(A1)

where UE and Uθ are the same potentials, as described in Sec. II D. Potentials used in the MC simulations and the MD simulations differ only by how the bonding and excluded volume potentials are modeled. Bond potentials in MC simulations are described as

(A2)

where the asterisk over the summation indicates that only bonded species are considered. This describes the bonds between polymer beads as a square well potential, whereas the MD simulations described bonds as a Gaussian potential. Excluded volume is described as a hard sphere potential UHS,

(A3)

In comparison to MD simulations, the hard sphere potential is stronger than the Lennard-Jones potential used to model excluded volume. Differences seen in the radial distribution functions shown in Fig. 2 are attributed to differences in how the bond and excluded volume potentials are modeled in MC and MD.

In this manuscript, in addition to being used to calculate radial distribution functions in Fig. 2, MC simulations were used to determine the excess free energy landscape as a function of salt, polyelectrolyte, and neutral polymer concentration. Widom insertion was performed for each species, i, to calculate excess chemical potentials, μEXC,i. This was done using joint insertion of charge neutral pairs and chain-end monomer addition to calculate per-monomer μEXC,i for the polymeric species. μEXC,i allows calculation of the excess free energy from a reference state ϕS0, ϕA0, and ϕB0,

(A4)

These fEXC values can be fit with a third-order polynomial using orthogonal distance regression. Simulations to determine chemical potential values in the limit of ϕB = 0.00 were performed with nA+ = nA = 6, and simulations to determine chemical potential values with the neutral polymer used nB = 12.

Previous work by the authors has demonstrated that a molecularly informed transfer matrix theory of complex coacervation nearly quantitatively matches both phase diagrams and the molecular structure for a number of systems with different molecular features.71,72 An overview of the theory will be included here, but detailed descriptions can be found in previous work.71,72

Polymers considered in this manuscript are in the large charge density limit, so a majority of monomers can be assumed to be adjacent to a condensed species of opposite charge.62,65,71 Both the entropic release of counterions and the combinatoric entropy of many chain-chain interactions drive complex coacervation.67,71,72 The entropic contributions from these effects depend on the concentration of polyelectrolytes and salt ions. These condensed states can be mapped to an adsorption model by considering each monomer as an adsorption site.71 Possible states for each adsorption site are a counterion, C, a different oppositely charged polyelectrolyte, P′, the oppositely charged polyelectrolyte continuing its run, P, or the monomer can be unpaired, 0. Transfer matrix formulation allows for calculation of the grand canonical partition function,

(B1)

where ψ is a column vector of ones and M is a matrix of Boltzmann weight for each adsorption state. Raising the matrix M to the Nth power allows enumeration of the Boltzmann weight of each monomer adsorption state to the partition function.

The matrix M is given by

(B2)

Each matrix entry is the weight associated due to an adsorbed species at monomer si given the adsorbed species at monomer si−1. Adjacent pairs of monomers are assigned to factors A=eμ̃S, B=eμ̃P, and D=eϵ̃, where the tilde denotes normalization by kBT. μ̃S and μ̃P are the chemical potentials for salt ions and polyelectrolyte monomers. ϵ̃ is a fitting parameter which captures the electrostatic driving force for oppositely charged species to condense onto the chain. For this work, we assume that it is constant for all values of ϕA, ϕB, and ϕS; however, more generally, it may be important to consider variations in ϵ̃ due to changes in the local electrostatic environment, especially in inhomogeneous polymer systems such as those described in this paper.108–110 The parameter E is the single-monomer partition function describing the confinement of adsorbed monomers after the initial monomer is adsorbed to the chain. The transfer matrix M can be multiplied by an arbitrary constant without affecting the thermodynamic properties, so we are free to choose E = 1.

The partition function, Eq. (B1), is analytically solvable by assuming the largest eigenvalue dominates,

(B3)

A free energy of interaction, Fint, can be calculated as

(B4)

where A and B can be determined using simple expressions for the chemical potentials μ̃S=μ̃S0+lnϕS=lnA0ϕS and μ̃P=μ̃P0+lnϕA=lnB0ϕA. μ̃S0 and μ̃P0 are reference chemical potentials for the salt and polyelectrolyte, and A0=expμ̃S0 and B0=expμ̃P0.

The excess free energy of the system, FEXC, can be calculated using Eq. (B4) with a phenomenological expression capturing the excluded volume of the non-water species,71,72

(B5)

where κ determines the strength of the excluded volume interaction and Λ = 0.6875 accounts for the smaller excluded volume of the polymer relative to the salt ions. Substitution of Eq. (B4) and the chemical potential expressions into Eq. (B5) gives the excess free energy as

(B6)

where

(B7)
(B8)
(B9)

For the species in the system, we can write the chemical potentials as

(B10)
(B11)
(B12)
(B13)

Here we have defined

(B14)

These expressions for μj,EXC can be used in Eq. (11) to incorporate the transfer matrix theory into the SCFT calculation. Parameters used in this manuscript are A0 = 20.5, B0 = 12.2, ϵ = 0.0, and κ = 19.0, consistent with previous work.71 

1.
H. G.
Bungenberg de Jong
and
H. R.
Kruyt
, “
Coacervation (partial miscibility in colloid systems)
,”
Proc. K. Ned. Akad. Wet.
32
,
849
856
(
1929
).
2.
J.
van der Gucht
,
E.
Spruijt
,
M.
Lemmers
, and
M. A.
Cohen Stuart
, “
Polyelectrolyte complexes: Bulk phases and colloidal systems
,”
J. Colloid Interface Sci.
361
,
407
422
(
2011
).
3.
S.
Srivastava
and
M. V.
Tirrell
, in
Advances in Chemical Physics
, edited by
S. A.
Rice
and
A. R.
Dinner
(
John Wiley and Sons
,
Hoboken, NJ
,
2016
).
4.
C. E.
Sing
, “
Development of the modern theory of polymeric complex coacervation
,”
Adv. Colloid Interface Sci.
239
,
2
16
(
2017
).
5.
S. L.
Turgeon
,
C.
Schmitt
, and
C.
Sanchez
, “
Protein-polysaccharide complexes and coacervates
,”
Curr. Opin. Colloid Interface Sci.
12
,
166
178
(
2007
).
6.
C.
Schmitt
and
S. L.
Turgeon
, “
Protein/polysaccharide complexes and coacervates in food systems
,”
Adv. Colloid Interface Sci.
167
,
63
70
(
2011
).
7.
M. A.
Cohen Stuart
,
N. A. M.
Besseling
, and
R. G.
Fokkink
, “
Formation of micelles with complex coacervate cores
,”
Langmuir
14
,
6846
6849
(
1998
).
8.
D. V.
Pergushov
,
A. H. E.
Müller
, and
F. H.
Schacher
, “
Micellar interpolyelectrolyte complexes
,”
Chem. Soc. Rev.
41
,
6888
6901
(
2012
).
9.
A.
Harada
and
K.
Kataoka
, “
Formation of polyion complex micelles in an aqueous milieu from a pair of oppositely-charged block copolymers with poly(ethylene glycol) segments
,”
Macromolecules
28
,
5294
5299
(
1995
).
10.
H.
Chu
,
J.
Gao
,
C. W.
Chen
,
J.
Huard
, and
Y.
Wang
, “
Injectable fibroblast growth factor-2 coacervate for persistent angiogenesis
,”
Proc. Natl. Acad. Sci. U. S. A.
108
,
13444
13449
(
2011
).
11.
K. A.
Black
,
D.
Priftis
,
S. L.
Perry
,
J.
Yip
,
W.
Byun
, and
M.
Tirrell
, “
Protein encapsulation via polypeptide complex coacervation
,”
ACS Macro Lett.
3
,
1088
1091
(
2014
).
12.
D. S.
Hwang
,
H.
Zeng
,
A.
Srivastava
,
D. V.
Krogstad
,
M.
Tirrell
,
J. N.
Israelachvili
, and
H. J.
Waite
, “
Viscosity and interfacial properties in a mussel-inspired adhesive coacervate
,”
Soft Matter
6
,
3232
3236
(
2010
).
13.
R. J.
Stewart
,
C. S.
Wang
, and
H.
Shao
, “
Complex coacervates as a foundation for synthetic underwater adhesives
,”
Adv. Colloid Interface Sci.
167
,
85
93
(
2011
).
14.
Y.
Fan
,
S.
Tang
,
E. L.
Thomas
, and
B. D.
Olsen
, “
Responsive block copolymer photonics triggered by protein-polyelectrolyte coacervation
,”
ACS Nano
8
,
11467
11473
(
2014
).
15.
E.
Spruijt
,
A. H.
Westphal
,
J. W.
Borst
, and
M. A.
Cohen Stuart
, “
Binodal compositions of polyelectrolyte complexes
,”
Macromolecules
43
,
6476
6484
(
2010
).
16.
D.
Priftis
and
M.
Tirrell
, “
Phase behaviour and complex coacervation of aqueous polypeptide solutions
,”
Soft Matter
8
,
9396
9405
(
2012
).
17.
R.
Chollakup
,
J. B.
Beck
,
K.
Dirnberger
,
M.
Tirrell
, and
C. D.
Eisenbach
, “
Polyelectrolyte molecular weight and salt effects on the phase behavior and coacervation of aqueous solutions of poly(acrylic acid) sodium salt and poly(allylamine) hydrochloride
,”
Macromolecules
46
,
2376
2390
(
2013
).
18.
M.
Radhakrishna
,
K.
Basu
,
Y.
Liu
,
R.
Shamsi
,
S. L.
Perry
, and
C. E.
Sing
, “
Molecular connectivity and correlation effects on polymer coacervation
,”
Macromolecules
50
,
3030
3037
(
2017
).
19.
Q.
Wang
and
J. B.
Schlenoff
, “
The polyelectrolyte complex/coacervate continuum
,”
Macromolecules
47
,
3108
3116
(
2014
).
20.
S. L.
Perry
,
Y.
Li
,
D.
Priftis
,
L.
Leon
, and
M.
Tirrell
, “
The effect of salt on the complex coacervation of vinyl polyelectrolytes
,”
Polymers
6
,
1756
1772
(
2014
).
21.
S. L.
Perry
,
L.
Leon
,
K. Q.
Hoffmann
,
M. J.
Kade
,
D.
Priftis
,
K. A.
Black
,
D.
Wong
,
R. A.
Klein
,
C. F.
Pierce
 III
,
K. O.
Margossian
,
J. K.
Whitmer
,
J.
Qin
,
J. J.
de Pablo
, and
M.
Tirrell
, “
Chirality-selected phase behaviour in ionic polypeptide complexes
,”
Nat. Commun.
6
,
6052
(
2015
).
22.
K. Q.
Hoffman
,
S. L.
Perry
,
L.
Leon
,
D.
Priftis
,
M.
Tirrell
, and
J. J.
de Pablo
, “
A molecular view of the role of chirality in charge-driven polypeptide complexation
,”
Soft Matter
11
,
1525
1538
(
2015
).
23.
L.-W.
Chang
,
T. K.
Lytle
,
M.
Radhakrishna
,
J. J.
Madinya
,
J.
Velez
, and
C. E.
Sing
, “
Sequence and entropy-based control of complex coacervates
,”
Nat. Commun.
8
,
1273
(
2017
).
24.
B. M.
Johnston
,
C. W.
Johnston
,
R.
Letteri
,
T. K.
Lytle
,
C. E.
Sing
,
T.
Emrick
, and
S. L.
Perry
, “
The effect of comb architecture on complex coacervation
,”
Org. Biomol. Chem.
15
,
7630
7642
(
2017
).
25.
E.
Spruijt
,
J.
Sprakel
,
M. A.
Cohen Stuart
, and
J.
van der Gucht
, “
Interfacial tension between a complex coacervate phase and its coexisting aqueous phase
,”
Soft Matter
6
,
172
178
(
2010
).
26.
D.
Priftis
,
R.
Farina
, and
M.
Tirrell
, “
Interfacial energy of polypeptide complex coacervates measured via capillary adhesion
,”
Langmuir
28
,
8721
8729
(
2012
).
27.
E.
Spruijt
,
F. A. M.
Leermakers
,
R.
Fokkink
,
R.
Schweins
,
A. A.
van Well
,
M. A.
Cohen Stuart
, and
J.
van der Gucht
, “
Structure and dynamics of polyelectrolyte complex coacervates studied by scattering of neutrons, x-rays, and light
,”
Macromolecules
46
,
4596
4605
(
2013
).
28.
A. B.
Marciel
,
S.
Srivastava
, and
M. V.
Tirrell
, “
Structure and rheology of polyelectrolyte complex coacervates
,”
Soft Matter
14
,
2454
2464
(
2018
).
29.
D.
Priftis
,
N.
Laugel
, and
M.
Tirrell
, “
Thermodynamic characterization of polypeptide complex coacervation
,”
Langmuir
28
,
15947
15957
(
2012
).
30.
J.
Fu
and
J. B.
Schlenoff
, “
Driving forces for oppositely charged polyion association in aqueous solutions: Enthalpic, entropic, but not electrostatic
,”
J. Am. Chem. Soc.
138
,
980
990
(
2016
).
31.
Y.
Liu
,
H. H.
Winger
, and
S. L.
Perry
, “
Linear viscoelasticity of complex coacervates
,”
Adv. Colloid Interface Sci.
239
,
46
60
(
2016
).
32.
D.
Priftis
,
X.
Xia
,
K. O.
Margossian
,
S. L.
Perry
,
L.
Leon
,
J.
Qin
,
J. J.
de Pablo
, and
M.
Tirrell
, “
Ternary, tunable polyelectrolyte complex fluids driven by complex coacervation
,”
Macromolecules
47
,
3076
3085
(
2014
).
33.
J.
Qin
,
D.
Priftis
,
R.
Farina
,
S. L.
Perry
,
L.
Leon
,
J.
Whitmer
,
K.
Hoffmann
,
M.
Tirrell
, and
J. J.
de Pablo
, “
Interfacial tension of polyelectrolyte complex coacervate phases
,”
ACS Macro Lett.
3
,
565
568
(
2014
).
34.
S.
Lim
,
D.
Moon
,
H. J.
Kim
,
S. H.
Seo
,
I. S.
Kang
, and
H. J.
Cha
, “
Interfacial tension of complex coacervated mussel adhesive protein according to the Hofmeister series
,”
Langmuir
30
,
1108
1115
(
2014
).
35.
D. V.
Krogstad
,
S.-H.
Choi
,
N. A.
Lynd
,
D. J.
Audus
,
S. L.
Perry
,
J. D.
Gopez
,
C. J.
Hawker
,
E. J.
Kramer
, and
M. V.
Tirrell
, “
Small angle neutron scattering study of complex coacervate micelles and hydrogels formed from ionic diblock and triblock copolymers
,”
J. Phys. Chem.
118
,
13011
13018
(
2014
).
36.
D. J.
Audus
,
J. D.
Gopez
,
D. V.
Krogstad
,
N. A.
Lynd
,
E. J.
Kramer
,
C. J.
Haker
, and
G. H.
Fredrickson
, “
Phase behavior of electrostatically complexed polyelectrolyte gels using an embedded fluctuation model
,”
Soft Matter
11
,
1214
1225
(
2015
).
37.
D. V.
Krogstad
,
N. A.
Lynd
,
S.-H.
Choi
,
J. M.
Spruell
,
C. J.
Hawker
,
E. J.
Kramer
, and
M. V.
Tirrell
, “
Effects of polymer and salt concentration on the structure and properties of triblock copolymer coacervate hydrogels
,”
Macromolecules
46
,
1512
1518
(
2013
).
38.
T. P.
Lodge
,
B.
Pudil
, and
K. J.
Hanley
, “
The full phase behavior for block copolymers in solvents of varying selectivity
,”
Macromolecules
35
,
4707
4717
(
2002
).
39.
T. P.
Lodge
,
K. J.
Hanley
,
B.
Pudil
, and
V.
Alahapperuma
, “
Phase behavior of block copolymers in a neutral solvent
,”
Macromolecules
36
,
816
822
(
2003
).
40.
L.
Leibler
,
H.
Orland
, and
J. C.
Wheeler
, “
Theory of critical micelle concentration for solutions of block copolymers
,”
J. Chem. Phys.
79
,
3550
3557
(
1983
).
41.
A.
Halperin
, “
Polymeric micelles: A star model
,”
Macromolecules
20
,
2943
2946
(
1987
).
42.
E.
Helfand
and
Y.
Tagami
, “
Theory of the interface between immiscible polymers
,”
J. Polym. Sci., Part B: Polym. Lett.
9
,
741
746
(
1971
).
43.
E.
Helfand
and
Y.
Tagami
, “
Theory of the interface between immiscible polymers. II
,”
J. Chem. Phys.
56
,
3592
3601
(
1972
).
44.
R. A.
Riggleman
,
R.
Kumar
, and
G. H.
Fredrickson
, “
Investigation of the interfacial tension of complex coacervates using field-theoretic simulations
,”
J. Chem. Phys.
136
,
024903
(
2012
).
45.
I.
Michaeli
,
J. T. G.
Overbeek
, and
M. J.
Voorn
, “
Phase separation of polyelectrolyte solutions
,”
J. Polym. Sci.
23
,
443
449
(
1957
).
46.
J. T. G.
Overbeek
and
M. J.
Voorn
, “
Phase separation in polyelectrolyte solutions: Theory of complex coacervation
,”
J. Cell. Comp. Physiol.
49
,
7
26
(
1957
).
47.
P.
Debye
and
E.
Hũckel
, “
Theory of electrolytes
,”
Phys. Z.
24
,
185
(
1923
).
48.
P. J.
Flory
,
Principles of Polymer Chemistry
(
Cornell University Press
,
Ithaca, NY
,
1953
).
49.
S. L.
Perry
and
C. E.
Sing
, “
Prism-based theory of complex coacervation: Excluded volume versus chain correlation
,”
Macromolecules
48
,
5040
5053
(
2015
).
50.
A.
Salehi
and
R. G.
Larson
, “
A molecular thermodynamic model of complexation in mixtures of oppositely charged polyelectrolytes with explicit account of charge association/dissociation
,”
Macromolecules
49
,
9706
9719
(
2016
).
51.
J.
Qin
and
J. J.
de Pablo
, “
Criticality and connectivity in macromolecular charge complexation
,”
Macromolecules
49
,
8789
8800
(
2016
).
52.
L.
Li
,
S.
Srivastava
,
M.
Andreev
,
A. B.
Marciel
,
J. J.
de Pablo
, and
M. V.
Tirrell
, “
Phase behavior and salt partitioning in polyelectrolyte complex coacervates
,”
Macromolecules
51
,
2988
2995
(
2018
).
53.
D. A.
McQuarrie
,
Statistical Mechanics
(
University Science Books
,
Sausalito
,
2000
).
54.
A.
Kudlay
and
M. O.
de la Cruz
, “
Precipitation of oppositely charged polyelectrolytes in salt solutions
,”
J. Chem. Phys.
120
,
404
412
(
2004
).
55.
A.
Kudlay
,
A. V.
Ermoshkin
, and
M. O.
de la Cruz
, “
Comoplexation of oppositely charged polyelectrolytes: Effect of ion pair formation
,”
Macromolecules
37
,
9231
9241
(
2004
).
56.
K.
Shen
and
Z.-G.
Wang
, “
Polyelectrolyte chain structure and solution phase behavior
,”
Macromolecules
51
,
1706
1717
(
2018
).
57.
M.
Castelnovo
and
J. F.
Joanny
, “
Complexation between oppositely charged polyelectrolytes: Beyond the random phase approximation
,”
Eur. Phys. J. E
6
,
377
386
(
2001
).
58.
V. Y.
Borue
and
I. Y.
Erukhimovich
, “
A statistical theory of globular polyelectrolyte complexes
,”
Macromolecules
23
,
3625
3632
(
1990
).
59.
J.
Lee
,
Y. O.
Popov
, and
G. H.
Fredrickson
, “
Complex coacervation: A field theoretic simulation study of polyelectrolyte complexation
,”
J. Chem. Phys.
128
,
224908
(
2008
).
60.
K. T.
Delaney
and
G. H.
Fredrickson
, “
Theory of polyelectrolyte complexation—Complex coacervates are self-coacervates
,”
J. Chem. Phys.
146
,
224902
(
2017
).
61.
T. K.
Lytle
,
M.
Radhakrishna
, and
C. E.
Sing
, “
High charge-density coacervate assembly via hybrid Monte Carlo-single chain in mean field theory
,”
Macromolecules
49
,
9693
9705
(
2016
).
62.
S.
Liu
and
M.
Muthukumar
, “
Langeving dynamics simulation of counterion distribution around isolated flexible polyelectrolyte chains
,”
J. Chem. Phys.
116
,
9975
(
2002
).
63.
Z.
Ou
and
M.
Muthukumar
, “
Entropy and enthalpy of polyelectrolyte complexation: Langevin dynamics simulations
,”
J. Chem. Phys.
124
,
154902
(
2006
).
64.
V. S.
Rathee
,
A. J.
Zervoudakis
,
H.
Sidky
,
B. J.
Sikora
, and
J. K.
Whitmer
, “
Weak polyelectrolyte complexation driven by associative charging
,”
J. Chem. Phys.
148
,
114901
(
2018
).
65.
G. S.
Manning
, “
Limiting laws and counterion condensation in polyelectrolyte solutions. I. Colligative properties
,”
J. Chem. Phys.
51
,
924
933
(
1969
).
66.
M.
Muthukumar
, “
Theory of counter-ion condensation on flexible polyelectrolytes: Adsorption mechanism
,”
J. Chem. Phys.
120
,
9343
9350
(
2004
).
67.
M. T.
Record
,
C. F.
Anderson
, and
T. M.
Lohman
, “
Thermodynamic analysis of ion effects on the binding and conformational equilibria of proteins and nucleic acids: The roles of ion association or release, screening, and ion effects on water activity
,”
Q. Rev. Biophys.
11
,
103
178
(
1978
).
68.
M.
Muthukumar
, “
50th anniversary perspective: A perspective on polyelectrolyte solutions
,”
Macromolecules
50
,
9528
9560
(
2017
).
69.
M. A.
Leaf
, “
Conducting polyelectrolyte complexes: Assembly, structure, and transport
,” Ph.D. thesis,
University of Massachusetts Amherst
,
2017
.
70.
R.
Zhang
and
B. I.
Shklovskii
, “
Phase diagram of solution of oppositely charged polyelectrolytes
,”
Phys. A
352
,
216
238
(
2005
).
71.
T. K.
Lytle
and
C. E.
Sing
, “
Transfer matrix theory of polymer complex coacervation
,”
Soft Matter
13
,
7001
7012
(
2017
).
72.
T. K.
Lytle
and
C. E.
Sing
, “
Tuning chain interaction entropy in complex coacervation using polymer stiffness, architecture, and salt valency
,”
Mol. Syst. Des. Eng.
3
,
183
196
(
2018
).
73.
M.
Radhakrishna
and
C. E.
Sing
, “
Charge correlations for precise, coulombically driven self assembly
,”
Macromol. Chem. Phys.
217
,
126
136
(
2016
).
74.
J. P.
Hansen
and
I. R.
McDonald
,
Theory of Simple Liquids
(
Elsevier
,
Boston
,
2006
).
75.
A.
Salis
and
B. W.
Ninham
, “
Models and mechanisms of Hofmeister effects in electrolyte solutions, and colloid and protein systems revisited
,”
Chem. Soc. Rev.
43
,
7358
7377
(
2014
).
76.
R.
Zhang
,
Y.
Zhang
,
H. S.
Antila
,
J. L.
Lutkenhaus
, and
M.
Sammalkorpi
, “
Role of salt and water in the plasticization of PDAC/PSS polyelectrolyte assemblies
,”
J. Phys. Chem. B
121
,
322
333
(
2017
).
77.
Y.
Zhang
,
F.
Li
,
L. D.
Valenzuela
,
M.
Sammalkorpi
, and
J. L.
Lutkenhaus
, “
Effect of water on the thermal transition observed in poly(allylamine hydrochloride)-poly(acrylic acid) complexes
,”
Macromolecules
49
,
7563
7570
(
2016
).
78.
E.
Yildirim
,
Y.
Zhang
,
J. L.
Lutkenhaus
, and
M.
Sammalkorpi
, “
Thermal transitions in polyelectrolyte assemblies occur via a dehydration mechanism
,”
ACS Macro Lett.
4
,
1017
1021
(
2015
).
79.
I.
Nakamura
,
N. P.
Balsara
, and
Z.-G.
Wang
, “
Thermodynamics of ion-containing polymer blends and block copolymers
,”
Phys. Rev. Lett.
107
,
198301
(
2011
).
80.
R.
Wang
and
Z.-G.
Wang
, “
Effects of ion solvation on phase equilibrium and interfacial tension of liquid mixtures
,”
J. Chem. Phys.
135
,
014707
(
2011
).
81.
Z.-G.
Wang
, “
Fluctuation in electrolyte solutions: The self energy
,”
Phys. Rev. E
81
,
021501
(
2010
).
82.
I.
Nakamura
and
Z.-G.
Wang
, “
Effects of dielectric inhomogeneity in polyelectrolyte solutions
,”
Soft Matter
9
,
5686
5690
(
2013
).
83.
R.
Kumar
,
B. G.
Sumpter
, and
M.
Muthukumar
, “
Enhanced phase segregation induced by dipolar interactions in polymer blends
,”
Macromolecules
47
,
6491
6502
(
2014
).
84.
Y.-Z.
Wei
,
P.
Chiang
, and
S.
Sridhar
, “
Ion size effects on the dynamic and static dielectric properties of aqueous alkali solutions
,”
J. Chem. Phys.
96
,
4569
4573
(
1992
).
85.
A.
Levy
,
D.
Andelman
, and
H.
Orland
, “
Dielectric constant of ionic solutions: A field-theory approach
,”
Phys. Rev. Lett.
108
,
227801
(
2012
).
86.
K. M.
Hong
and
J.
Noolandi
, “
Theory of inhomogeneous multicomponent polymer systems
,”
Macromolecules
14
,
727
736
(
1981
).
87.
K.
Shull
, “
Mean-field theory of block copolymers: Bulk melts, surfaces, and thin films
,”
Macromolecules
25
,
2122
2133
(
1992
).
88.
K. R.
Shull
and
E. J.
Kramer
, “
Mean-field theory of polymer interfaces in the presence of block copolymers
,”
Macromolecules
23
,
4769
4779
(
1990
).
89.
K. R.
Shull
, “
Theory of end-adsorbed polymer brushes in polymeric matrices
,”
J. Chem. Phys.
94
,
5723
5738
(
1991
).
90.
G.
Fredrickson
,
The Equilibrium Theory of Inhomogeneous Polymers
(
Oxford University Press
,
2013
).
91.
Q.
Wang
,
T.
Taniguchi
, and
G. H.
Fredrickson
, “
Self-consistent field theory of polyelectrolyte systems
,”
J. Phys. Chem. B
108
,
6733
6744
(
2004
).
92.
D.
Frenkel
and
B.
Smit
,
Understanding Molecular Simulation: From Algorithms to Applications
(
Academic Press
,
San Diego, CA
,
2002
).
93.
S. K.
Kumar
,
I.
Szleifer
, and
A. Z.
Panagiotopoulos
, “
Determination of the chemical potentials of polymeric systems from Monte Carlo simulations
,”
Phys. Rev. Lett.
66
,
2935
2938
(
1991
).
94.
S.
Plimpton
, “
Fast parallel algorithms for short-range molecular dynamics
,”
J. Comput. Phys.
117
,
1
19
(
1995
).
95.
J. W.
Cahn
and
J. E.
Hilliard
, “
Free energy of a nonuniform system. I. Interfacial free energy
,”
J. Chem. Phys.
28
,
258
267
(
1958
).
96.
C.
Huang
,
M.
Olvera de la Cruz
, and
B. W.
Swift
, “
Phase separation of ternary mixtures: Symmetric polymer blends
,”
Macromolecules
28
,
7996
8005
(
1995
).
97.
C.
Huang
and
M.
Olvera de la Cruz
, “
Adsorption of a minority component in polymer blend interfaces
,”
Phys. Rev. E
53
,
812
819
(
1996
).
98.
H.
Tang
and
K. F.
Freed
, “
Interfacial studies of incompressible binary blends
,”
J. Chem. Phys.
94
,
6307
6322
(
1991
).
99.
C. E.
Sing
,
J. W.
Zwanikken
, and
M.
Olvera de la Cruz
, “
Theory of melt polyelectrolyte blends and block copolymers: Phase behavior, surface tension, and microphase periodicity
,”
J. Chem. Phys.
142
,
034902
(
2015
).
100.
M. W.
Matsen
and
F. S.
Bates
, “
Unifying weak- and strong-segregation block copolymer theories
,”
Macromolecules
29
,
1091
1098
(
1996
).
101.
M. W.
Matsen
and
F. S.
Bates
, “
Origins of complex self-assembly in block copolymers
,”
Macromolecules
29
,
7641
7644
(
1996
).
102.
C. E.
Sing
and
M.
Olvera de la Cruz
, “
Polyelectrolyte blends and nontrivial behavior in effective flory-huggins parameters
,”
ACS Macro Lett.
3
,
698
702
(
2014
).
103.
R. J.
Ellis
, “
Macromolecular crowding: An important but neglected aspect of the intracellular environment
,”
Curr. Opin. Struct. Biol.
11
,
114
119
(
2001
).
104.
A. M.
Marianelli
,
B. M.
Miller
, and
C. D.
Keating
, “
Impact of macromolecular crowding on RNA/spermine complex coacervation and oligonucleotide compartmentalization
,”
Soft Matter
14
,
368
378
(
2018
).
105.
P.
Schaaf
and
J. B.
Schlenoff
, “
Saloplastics: Processing compact polyelectrolyte complexes
,”
Adv. Mater.
27
,
2420
2432
(
2015
).
106.
P.
Zhang
,
N. M.
Alsaifi
,
J.
Wu
, and
Z.-G.
Wang
, “
Salting-out and salting-in of polyelectrolyte solutions: A liquid-state theory study
,”
Macromolecules
49
,
9720
9730
(
2016
).
107.
M.
Muthukumar
, “
Double screening in polyelectrolyte solutions: Limiting laws and crossover formulas
,”
J. Chem. Phys.
105
,
5183
(
1996
).
108.
J. P.
Mahalik
,
Y.
Yang
,
C.
Deodhar
,
J. F.
Ankner
,
B. S.
Lokitz
,
S. M.
Kilbey
,
B. G.
Sumpter
, and
R.
Kumar
, “
Monomer volume fraction profiles in ph responsive planar polyelectrolyte brushes
,”
J. Polym. Sci. Part B: Polym. Phys.
54
,
956
964
(
2016
).
109.
R.
Nap
,
P.
Gong
, and
I.
Szleifer
, “
Weak polyelectrolytes tethered to surfaces: Effect of geometry, acid-base equilibrium and electrical permittivity
,”
J. Polym. Sci. Part B: Polym. Phys.
44
,
2638
2662
(
2006
).
110.
E. B.
Zhulina
and
O. V.
Borisov
, “
Poisson-Boltzmann theory of ph-sensitive (annealing) polyelectrolyte brush
,”
Langmuir
27
,
10615
10633
(
2011
).