Using analytical considerations and particle-based simulations of a coarse-grained model, we study the relaxation of a density modulation in a polymer system without nonbonded interactions. We demonstrate that shallow density modulations with identical amplitudes and wavevectors that have been prepared by different processes exhibit different nonexponential decay behaviors. Thus, in contrast to the popular assumption of dynamic self-consistent field theory, the density alone does not suffice to characterize the configuration of the polymer system. We provide an analytic description within Linear-Response Theory (LRT) and the Rouse model that quantitatively agree with the results of the particle-based simulations. LRT is equivalent to a generalized model-B dynamics with an Onsager coefficient that is nonlocal in space and time. Alternatively, the Rouse description can be cast into a dynamic density-functional theory that uses the full probability distribution of single-chain configurations as a dynamic variable and yields a memory-free description of the dynamics that quantitatively accounts for the dependence on the preparation process. An approximate scheme that only considers the joint distribution of the first two Rouse modes—the ellipsoid model—is also explored.

The free-energy landscape of dense polymer systems can be accurately described by Self-Consistent Field Theory (SCFT).1–12 In the limit of a large invariant degree of polymerization, N̄(ρ0Re3/N)2 (with ρ0, N, and Re being the segment number density, the number of segments per polymer, and the polymer’s root mean squared end-to-end distance, respectively), one chain molecule interacts with many neighbors, and the mean-field theory provides a quantitatively accurate description of thermal equilibrium. This provides an excellent starting point for studying the kinetics of structure formation.13,14

In this paper, we study a deceptively simple problem—the temporal decay of a shallow density modulation in an ideal gas of polymers. Understanding this basic phenomenon is a prerequisite for studying more complex dynamic phenomena, such as phase separation in homopolymer blends, self-assembly of block copolymers, or polymer crystallization.

Traditionally, one characterizes the one-component polymer system by the local segment density, ρ(r), as a function of space, r, that can be computed from the segment positions, rji, that completely specify the microscopic configuration of the system at a given time,
ρ(r)=j=1ni=1Nδ(rrji)=j=1nN01dsδ(rrj(s)),
(1)
where the sum runs over all polymers, j = 1, …, n, in a system of volume, V. Each polymer is discretized into N segments. The spatial position of the ith segment along polymer j is denoted by rji. In the latter expression, we have taken the continuum limit of fine discretization of the molecular contour, N → ∞, introduce a continuous contour variable, 0 ≤ s = (i − 1/2)/N ≤ 1, and characterize the single-chain configurations by rj(s). The spatial Fourier transform of the collective segment density is defined by
ρk=1Vdreikrρ(r),
(2)
where ρ0 = nN/V denotes the spatial average of the segment density.
Within SCFT, the free-energy landscape is quantified by a functional, F[ρk], of the local segment density. ΔF does not depend on the sign and complex phase of ρk. Thus, a Landau expansion [Random-Phase Approximation (RPA)] of ΔF begins with |ρk|2. The thermal average of segment-density fluctuations defines the collective structure factor, Skcoll=|Vρk|2/(nN), that is experimentally accessible by scattering experiments. Thus, a shallow density fluctuation with wavevector k increases the free energy by an amount6,15
ΔF[ρk]=kBTV2ρ0k|ρq|2Skcoll+O(|ρk|3),
(3)
where kBT denotes the thermal energy. Since polymers in an ideal gas do not interact, individual molecules independently contribute to the collective structure factor, i.e., Skcoll=Sk, where Sk denotes the static, single-chain structure factor. For a continuous Gaussian chain, it is given by the Debye function, Sk = Ng(x), with x=(kRe)2/6,6,15
g(x)=2[(kRe)2/6]2e(kRe)261+(kRe)26.
(4)
Since the segment number is a conserved quantity, its density obeys the continuity equation16,
dρkdt=ikjk,
(5)
where j denotes the density current. For a shallow density fluctuation, ρkρ0, Linear-Response Theory (LRT) gives rise to a linear relation between the current and the gradient of the chemical potential. In Dynamic Self-Consistent Field Theory (D-SCFT),7,10,17–31 this relation is assumed to be local in time,
jk=ikΛkμkkBTwithμk=Nρ0VFρkkBTρkSk/N,
(6)
where Λk denotes the wavevector-dependent Onsager coefficient.10,16,20,28,29,31–33 The wavevector dependence accounts for the spatially extended chain structure. Thus, the accurate free-energy functional [Eq. (3)] in conjunction with this model-B dynamics16 predicts an exponential decay of a density fluctuation in time,
dρkdt=k2ΛkρkSk/N,
(7)
with a wavevector-dependent relaxation rate, 1τk=k2ΛkSk/N. Equation (6) ignores thermal fluctuations, and therefore, ρk(t) corresponds to the average over different realizations of the stochastic relaxation process.
In the following, we create an initial, shallow density modulation with a specific wavevector, q, by a weak external preparation field of strength, h, for times t < 0. Let γ(s) = 1 characterize all segments on which the initial external fields acts, whereas γ(s) = 0 for all other segments along the molecular contour. For t < 0, the external preparation field contributes
ΔFkBT=Nhdrcos(qr)ρh(r)
(8)
to the excess free energy, with ρh(r)=j=1ndsγ(s)δ(rrj(s)). A weak external preparation field gives rise to a spatial modulation of the density ρ(r) ∼ ρh(r) ∼ −h cos(qr) for t ≤ 0.

At t ≥ 0, the field is switched off, and the density modulation relaxes. In Fig. 1, we present the time evolution of the normalized amplitude, a(t) ≡ 2[ρ±q(t)]/ρ0, for qRe = 4π. This wavevector corresponds to a spatial period of L = Re/2 ≈ 1.22Rg, where Rg=Re/6 denotes the radius of gyration of the Gaussian polymer. The relaxation of such a density modulation with a length scale that is comparable to or smaller than the molecular extension does not only involve the diffusion of the molecule’s center of mass but, additionally, the dynamics of (internal) molecular configurations. The data are obtained by particle-based simulations of an ideal gas of Gaussian polymer chains (cf.  Appendix B for a description of the model and simulation technique). []ρq(0) denotes the average over independent realizations of the stochastic relaxation process. One can clearly appreciate that the amplitude does not exponentially decay in time in marked contrast to Eq. (7).

FIG. 1.

Relaxation of the amplitude of shallow density modulations in an ideal gas of homopolymers. Time, t, is measured in units of the Rouse time, τR. Symbols mark results of particle-based simulations. The dashed line corresponds to the predictions of D-SCFT [Eq. (7)] with an Onsager coefficient, Λk = D, where D=Re23π2τR stands for the polymer’s self-diffusion coefficient. Solid lines show the predictions of LRT/D-RPA [Eq. (21)]. The black solid line coincides with the decay of the dynamic structure factor [Eq. (10)].

FIG. 1.

Relaxation of the amplitude of shallow density modulations in an ideal gas of homopolymers. Time, t, is measured in units of the Rouse time, τR. Symbols mark results of particle-based simulations. The dashed line corresponds to the predictions of D-SCFT [Eq. (7)] with an Onsager coefficient, Λk = D, where D=Re23π2τR stands for the polymer’s self-diffusion coefficient. Solid lines show the predictions of LRT/D-RPA [Eq. (21)]. The black solid line coincides with the decay of the dynamic structure factor [Eq. (10)].

Close modal
Since the description in terms of the collective segment density, ρq(t), involves a projection from the individual segment coordinates [cf. Eq. (1)], one should expect on general grounds34,35 that this projection imparts memory onto the time evolution of the coarse-grained order parameter, ρq, resulting in a nonexponential decay in time. Generally, a typical shallow density fluctuation, ρq(0), decays like the collective dynamic structure factor of density fluctuations. Distinguishing between the average, []ρq(0), over realizations of the stochastic relaxation process, ensuing from configurations with an initial ρq(0), and the thermal average ⟨⋯⟩ over initial configurations, we obtain [see Eq. (21) for a more detailed derivation within LRT]
Sqcoll(t)=[ρq(t)]ρq(0)ρq*(0)=[ρq(t)]ρq(0)ρq(0)ρq(0)ρq*(0)=[ρq(t)]ρq(0)ρq(0)Sqcoll.
(9)
Here, we have assumed that [ρq(t)]ρq(0)/ρq(0) only is a function of t and q but independent from the microscopic configuration at t = 0. For an ideal gas, the time-dependent collective and single-chain structure factors coincide, Sqcoll(t)=Sq(t), and an explicit expression is known within the Rouse model (see Appendix 4.III of Ref. 36),
Sq(t)=N01ds01dse(qRe)2|ss|6(qRe)2t3π2τR×e2(qRe)23π2p>01ep2t/τRp2cosπpscosπps.
(10)
The Rouse time, τR, and the self-diffusion coefficient of the molecule’s center of mass are related by D=Re2/(3π2τR).

Indeed, Fig. 1 shows that if the shallow density modulation has been prepared by an external spatially varying field that acts on all segments uniformly, γ = 1 for all s, the time evolution observed in the simulations (filled black circles) quantitatively agrees with the decay of a density fluctuation [Eqs. (9) and (10)] (black line) in equilibrium without an adjustable parameter. Note that the normalized amplitude of the density modulation is only 0.02; thus, LRT is accurate.

Figure 1, however, also reveals that if the density modulation has been prepared by a different process, its time evolution differs. If we use an external field for t < 0 that only acted on an end of a macromolecule, γ(s) = δ(s)/N (blue squares), we observe that the density modulation decays faster than predicted by Eq. (9) after the field has been switched off at t = 0. If the initial preparation field only acted on the segments in the middle of a chain, i = N/2 and N/2 + 1 [or γ(s) = δ(s − 1/2)/N, green triangles], the density modulation decays slower than predicted by the collective structure factor, Sq(t), of density fluctuations at the wavevector, q, of the density modulation.

Both observations demonstrate that the collective density, ρq, at a given time t = 0, alone cannot appropriately characterize the initial state of the system. Therefore, by using ρq(t) as the only order parameter, traditional D-SCFT7,10,17–31 fails to accurately describe the short-time dynamics of the segment density even in an ideal gas of homopolymers.37 

In this paper, we explore two avenues to improve the description of the short-time dynamics:

  • Exactly rewriting LRT with an ideal gas as the reference system, i.e., D-RPA,38–42 in Sec. II, we generalize the model-B dynamics [Eq. (7)] via an Onsager coefficient that is nonlocal in space and time,31,33,40,43
    dρq(t)dt=tdtq2Λq(tt)ρq(t)Sq/N.
    (11)
    This spatiotemporal nonlocality results from the extended shape of the macromolecules and their spectrum of relaxation times. The nonlocality in time quantifies memory effects that extend up to the Rouse time, τR,36 and accounts for the fact that the chain configurations need not to be in equilibrium with the instantaneous density distribution, ρq(t). In the case that the density modulation has been prepared by an external field that acted only on a portion of the chain contour (e.g., a chain end or middle segment), it turns out to be necessary to describe the time evolution of both, ρq(t) and ρhq(t). We obtain an explicit expression of the Onsager coefficients with memory (or rather their Laplace transforms) in terms of dynamic single-chain structure factors in accord with prior work on incompressible, multicomponent polymer systems.31,33
  • Rather than projecting the degrees of freedom associated with the chain configurations onto the segment density, ρq(t), subsequently in Sec. III, we use the probability distribution, P, of single-chain configurations—the molecular density—as an order parameter that characterizes the system. Such a more detailed description is routinely employed in classical Density-Functional Theory (DFT) of molecules with orientational degrees of freedom or other internal degrees of freedom that characterize extended molecular configurations.44–47 The equilibrium properties of the molecular density, P, of polymer systems are described by Polymer Density-Functional Theory (PDFT),9,48–63 which has attracted abiding interest to accurately predict liquid-like packing effects of polymer fluids. In  Appendix A, we show that static SCFT and PDFT yield the same description of equilibrium properties for continuous Gaussian chains.

Exactly rewriting the Rouse model64 of an ideal gas of chain molecules, we obtain a Dynamic Density-Functional Theory (D-DFT)65,66 that uses the free-energy functional of PDFT. This D-DFT provides a memory-free description of the dynamics and accurately accounts for the dependence on the preparation process observed in Fig. 1. A simplified but approximate approach—the ellipsoid model—that characterizes a molecular configuration only via the chain’s center of mass and overall shape of the segment distribution is also briefly discussed.

This paper concludes with a summary and a brief outlook.

Even if the weak external field acted at times, t < 0, on all segments, i.e., γ = 1 for all s, traditional D-SCFT7,10,17–31 with a wavevector-dependent Onsager coefficient fails to describe the short-time dynamics. Only within the generalized model-B [Eq. (11)]31,33,40 with an Onsager coefficient that encodes the memory of the chain configurations, can we accurately describe the relaxation of a shallow density fluctuation. Since such a density fluctuation decays like the dynamic single-chain structure factor [Eq. (9)] one can invert the relation31 
ddtSq(t)=0tdtq2NΛ(tt)Sq(t)Sq
(12)
to construct the time-dependent Onsager coefficient.

In the limit of long wavelengths, qRe → 0, a density fluctuation decays by the diffusion of the chain’s center of mass, and Sq(t) → N exp(−q2Dt).32 In this limit, the above equation yields an Onsager coefficient that is local in time, Λ(t) → (t). Applying this expression at finite qRe, we obtain an exponential shown by the dashed-red line in Fig. 1.

Because of the convolution in time, in general, we can only provide an analytic expression for the Laplace transform of the Onsager coefficient (vide infra). Numerically, we split the Onsager coefficient into a local, ultra-short time contribution and a nonlocal contribution in time,31,
Λq(t)=Λ0δ(t)Λnlq(t)/τR,
(13)
with
Λ0=1q2NdSqdtt=0=Re23π2τR=D,
(14)
and define an integrated memory kernel K(t) according to Λ(t) = dK/dt.67 This integrated memory kernel can be iteratively constructed as K(t)=n=0Kn(t) with Sq(0)K0(t) = dSq/dt and Kn+1(t)=0tdtKn(t)K0(tt).67 

Having numerically determined the memory kernel for various values of qRe, we present the nonlocal contribution of the Onsager coefficient, Λnlq(t), in Fig. 2. It is similar to the memory kernel of the generalized model-B dynamics for incompressible, symmetric diblock copolymer melts.31 We observe that the memory, Λnlq, extends up to the Rouse time, τR, encompassing the entire spectrum of relaxation times of the Rouse chain. The data in Fig. 2 approach the limit of small length scales, qRe ≫ 1, and short time scales, tτR, characteristic for the universal scale-invariant dynamics inside a Gaussian polymer coil.64 In this limit, we expect that wavevector- and time-dependent, nonlocal contribution Λnlq(t) is a function of a single scaling variable, t*=(qRe)4t/(3π2τR)=(qRe)2q2Dt, for Rouse dynamics, extending up to tτR or t*(qRe)4/(3π2). This expectation is confirmed by the data in Fig. 2. We hypothesize that the scale invariance of the dynamics inside a Gaussian polymer coil gives rise to a power-law behavior of Λnlq(t) in this regime.

FIG. 2.

Dimensionless, nonlocal contribution, Λnlq(t)=τRδ(t)Λq(t)/D, to the Onsager coefficient, required for a generalized model-B description31,33,40 [Eq. (11)] if the external preparation field acts on all segments, γ(s) = 1 for all s. The data for different wavevectors, q, are scaled according to the prediction of the Rouse model64 in the short-wavelength limit, qRe → ∞. Vertical lines indicate the Rouse time, τR.

FIG. 2.

Dimensionless, nonlocal contribution, Λnlq(t)=τRδ(t)Λq(t)/D, to the Onsager coefficient, required for a generalized model-B description31,33,40 [Eq. (11)] if the external preparation field acts on all segments, γ(s) = 1 for all s. The data for different wavevectors, q, are scaled according to the prediction of the Rouse model64 in the short-wavelength limit, qRe → ∞. Vertical lines indicate the Rouse time, τR.

Close modal

Since the Onsager coefficient in the generalized model-B dynamics31,33,40 accounts for memory effects, the dynamics does require information of the system at times t < 0 when there is a distinction between segments that are subjected to the external field and others that are not. Thus, we characterize the system configuration at a given time by both, the total segment density, ρ(r), and the density of segments, ρh(r), on which the external field has been applied for t < 0, in order to incorporate the dependence on the preparation process. If the field was applied to all segments, the generalized model-B description only requires ρq. If the field only acted on a portion of a chain, we employ two order-parameter fields, ρhq, and the density of segments, ρnqρqρhq, on which the external field did not act for t < 0, i.e., we describe the system as a block copolymer.

Combining these two order-parameter field into ρq=(ρhq,ρnq)T, we obtain a quadratic free-energy functional within RPA [in analogy to Eq. (3)],15,
ΔF[ρq]=kBTV2ρ0qρqTSq1ρq,
(15)
where Sq is the 2 × 2 matrix of partial, static, single-chain structure factors. The chemical potentials of the two segment species take the form μq=kBTNSq1ρq, resulting in the generalized model-B dynamics,31,33,40
dρqdt=q20tdtq(tt)μq(t)kBT=q2N0tdtq(tt)Sq1ρq(t),
(16)
where q(t) denotes the symmetric matrix of nonlocal Onsager coefficients. The Laplace transform (or one-sided temporal Fourier transform) of the above equation reads
iωρ̃q(ω)ρq(0)=q2Ñq(ω)Sq1ρ̃q(ω),
(17)
with the transform ρ̃q(ω)=0dteiωtρq(t).
These Onsager coefficients, q(t), can be obtained by comparing Eq. (17) with the predictions of D-RPA,38–42 i.e., the LRT for an ideal gas of block copolymers subjected to a time-dependent external field, Vq(t),
ρq(t)ρ0=tdtxq(tt)Vq(t),
(18)
where the matrix of linear-response functions, xq, is related to the matrix of partial dynamic structure factors, Sq(t), of the ideal-gas reference system by
xq(t)=dSq(t)dtfort0
(19)
and xq(t)=0 for t < 0.
In the present case, the external field is constant for t < 0, Vq(t)V=Nh2(1,0)T, and vanishes for t ≥ 0. Thus, Eq. (18) can be simply integrated to
ρq(t)ρ0=Sq(t)V.
(20)
The result for the segment density ρq(t)=ρhq(t)+ρnq(t)=Shhq(t)+Snhq(t)ρ0Nh/2 takes the form
ρ±q(t)ρ0=Nh201ds01dsγ(s)e(qRe)2|ss|6(qRe)2t3π2τR×e2(qRe)23π2p>01ep2t/τRp2cosπpscosπps.
(21)
This prediction of LRT is plotted in Fig. 1 as colored solid lines for the different preparation processes described by γ(s), and it quantitatively agrees with the simulations of the particle-based model without an adjustable parameter.
Differentiating and eliminating V, we obtain
dρqdt=dSq(t)dtρ0V=dSq(t)dtSq1(t)ρq(t).
(22)
The comparison of the Laplace transform of this equation
iωρ̃q(ω)ρq(0)=iωS̃q(ω)SqS̃q1(ω)ρq(t)=x̃q(ω)S̃q1(ω)ρq(t)
(23)
and Eq. (17) identifies an explicit, analytic expression for the matrix of Onsager coefficients,33,43,68
q2Ñq(ω)=x̃q(ω)S̃q1(ω)Sq=iωx̃q1(ω)Sq11.
(24)
Since these Onsager coefficients in conjunction with the generalized model B [Eq. (16)] for both, ρhq and ρnq, are equivalent to Eq. (20), they quantitatively describe the short-time dynamics in Fig. 1 even if the initial, external field only acted on a chain portion.

This equivalence of the generalized model-B dynamics (with the nonlocal Onsager coefficients, q, given by this combination of single-chain, linear-response functions, xq(t) and static, single-chain structure factors, Sq) and LRT/D-RPA38–42 is not restricted to an ideal gas but generally holds for interacting systems, such as incompressible, multicomponent polymer systems (within linear response).33 

Formally, Eq. (24) provides an explicit, analytic expression for the Laplace transform of the Onsager coefficient. The accurate determination of q(t) in the time domain, however, is a nontrivial challenge, and the memory kernel, q(t), in Eq. (16) increases the numerical complexity of integrating the generalized model-B equation. Therefore, it is worth to explore an approach that utilizes a more microscopic description of the system configuration but results in a simpler memory-free dynamics.

Whereas SCFT uses the segment density to specify the configuration of a homopolymer system, PDFT employs the molecular density, P, that quantifies the probability of a single-chain configuration. A single-chain configuration can either be specified by the N segment positions, ri with i = 1, …, N or, equivalently, by the N Rouse modes, Xp with p = 0, …, N − 1.64 The transformation between segment positions and Rouse modes reads
Xp1Ni=1NricosπpNi12,
(25)
ri=X0+2p>0N1XpcosπpNi12.
(26)
The zeroth Rouse mode denotes the center of mass of a polymer, and in the limit of fine chain discretization, we obtain X0 = dsr(s). For p ≥ 1, Eq. (25) becomes
Xp=01dsr(s)cos(πps)=01dsdrdssin(πps)πp,
(27)
where the latter expression shows that Rouse modes with p > 0 characterize the distribution of bond vectors and do not depend on the absolute position of the macromolecule in space.

In the following, we use the Rouse modes, X, instead of the segment positions, R, to specify a single-chain configuration because, within the framework of the Rouse model,64 the dynamics of the different Rouse modes are independent and are simply described by Ornstein–Uhlenbeck processes with relaxation times, τR/p2.

First, we study the equilibrium distribution of Rouse modes in the presence of a weak external preparation field that acts on a portion of the chain. This description is the analog of static RPA and provides information about the initial P(X, t) at the time, t = 0, when the external field that has prepared the density modulation is switched off. Second, we study the dynamics of P(X, t) in the absence of the external field, t > 0, and project onto the time evolution of the segment density. This strategy reproduces the predictions of LRT/D-RPA.38–42 Third, we demonstrate that the time evolution of P(X, t) according to the Rouse model64 is identical to a deterministic D-DFT65,66 using the free-energy functional of PDFT and a simple local mobility. Fourth, we explore an approximate description in terms of the first two Rouse modes, X0 and X1, only—the ellipsoid model.

In the absence of an external field or nonbonded interactions, the zeroth Rouse mode, X0, is uniformly distributed in the volume, V. In this case, the Cartesian components of the higher Rouse modes, Xp with p > 0, are statistically independent and Gaussian distributed with average, ⟨Xp⟩ = 0, and variance, Xp2σp2Re2/(6π2p2), for N → ∞.36 

In the following, we consider the response of P(X) to an external preparation field of strength, kBTh, acting on a portion of a chain with γ(s) = 1. In equilibrium, the distribution of the Rouse modes in the canonical ensemble is given by
P(X)ep=13π2p2Xp2Re2Nh01dsγ(s)cosqX0+2p=1qXpcosπps.
(28)
The first term in the Boltzmann factor represents the bonded energy, Hb [also see Eq. (A3)], whereas the second term quantifies the potential energy of the segment with contour index s in the external field. In the canonical ensemble, the free-energy functional F[P] does not depend on the normalization of the molecular density [vide infra Eq. (A24)]. Thus, we choose DXP(X) = 1, where DX=VdX0p=1dXp integrates over all molecular configurations.
The segment density, ρ(r), is obtained by projection
ρ(r)=nDXP(X)i=1Nδ(rri)=ni=1Nδ(rri)P,
(29)
with [⋯]P standing for the average with respect to the molecular density, P. The molecular density, P, however, cannot be reconstructed from the segment density, ρ(r).
The spatial Fourier transform of the segment density takes the form
ρkρ0=DXP(X)dseikX0+2p>0Xpcosπps.
(30)
To make progress, we assume that the external preparation field is weak, Nhγ(s) ≪ 1, and approximate
P(X)ep>03π2p2Xp2Re21Nh01dsγ(s)×cosqX0+2p>0qXpcosπps+O(h2).
(31)
While different Rouse modes are statistically independent in the absence of an external field, h = 0, they become correlated even within this linear, weak-field approximation. Whereas, for p > 0, the averages
XpP=0andXp2P=Re22π2p2
(32)
over P(X) remain unaltered, a weak, external, spatially periodic field induces a spatial variation of the moments of the Rouse modes, i.e., cross-correlations of the form
q̂Xpsin(qX0)Pand(q̂Xp)2cos(qX0)P,
(33)
where q̂=q/|q| denotes the unit vector of the periodic, spatial modulation, may not vanish (vide infra Fig. 5).
Within this linear, weak-field approximation [Eq. (31)], we obtain
ρ±qρ0Nh2dsdsγ(s)p=1dXpe3π2p2Xp2Re2(Re2/3πp2)3/2×e±2p>0iqXpcos(πps)cos(πps)+O(h2),
(34)
=Nh2dsdsγ(s)e(qRe)262p>0cos(πps)cos(πps)πp2=Nh2dsdsγ(s)e(qRe)26|ss|,
(35)
which agrees with the value of Eq. (21) at t = 0. Other Fourier components, ρk, of the segmental density with k ≠ 0, ±q vanish in linear order.
If the preparation field uniformly acts on all segments, γ(s) = 1 for all s, we obtain the well-known RPA relation
ρ±qρ0Nh2gD(qRe)=h2Sq
(36)
or
ρ(r)ρ01hSqcos(qr).
(37)
If the field only acts on an end segment with strength, he, i.e., Nhγ(s) = heδ(s), we find
ρ±qρ0he26(qRe)21e(qRe)26.
(38)
If the field only acts on the middle segment with strength, hm, i.e., Nhγ(s)=hmδ(s1/2), we obtain
ρ±qρ0hm212(qRe)21e(qRe)212.
(39)
Equations (36), (38), and (39) quantify the strengths of the external fields necessary to bring about a segment-density variation of a given amplitude and wavevector, q. These results coincide with the prediction of static RPA or D-RPA [Eq. (20)] evaluated at time t = 0. They are also confirmed by the particle-based simulations (see Table I).
TABLE I.

Amplitude of the initial density modulation and strength of the external preparation field in units of the thermal energy kBT used in the particle-based simulation with SOMA.

All segmentsEnd segmentTwo middle segments
2ρ±qρ0=0.0207 0.0202 0.0205 
h=0.285128 he=64128=0.5 hm=17.564=0.273 
All segmentsEnd segmentTwo middle segments
2ρ±qρ0=0.0207 0.0202 0.0205 
h=0.285128 he=64128=0.5 hm=17.564=0.273 
Equations (38) and (39) have a simple interpretation: Since the external field only acts on a single segment, sh = 0 or 1/2, respectively, it does not distort the internal distances, i.e., the joint distribution, P̃(r,d), of segment sh being at location r and segments s and s′ being separated by a distance vector d factorizes
P̃(r,d)=DXP(X)δrX02p>0Xpcos(πpsh)×δd2p>0Xpcos(πps)cos(πps)
(40)
1hcos(qr)V32πRe2|ss|3/2e3d22Re2|ss|.
(41)
Unfortunately, however, the fact that the internal-distance distribution adopts its field-free equilibrium form, independently from the position of the segment on which the external preparation field is applied, does not imply that the Rouse modes with p > 0 are uncorrelated or are distributed like in the absence of the external preparation field [cf. Eq. (31)]. Therefore, also a simple factorization of the relaxation of a density modulation into a time-dependent dynamics of the distribution, ρh(r, t), of the segment, sh, on which the external preparation field has acted, and an equilibrium (i.e., time-independent) density distribution of all segments around this segment, sh, is not exact because the dynamics of segment, sh, and the internal distances of the chain molecule are coupled.

These static considerations within the Rouse model show that even if the variation of the segment density, ρ(r), is identical, the probability distribution of the Rouse modes, Xp, given by Eq. (28) or Eq. (31), and the density profile of specific segments, ρ(r, s), at position s along the molecular contour do depend on the process by which the density modulation has been prepared. This observation explicitly illustrates the failure of the basic assumption of D-SCFT7,10,17–31 that the segment density, ρ(r), is the only slow variable and that all other characteristics of the system, such as the probability distribution of the Rouse modes, Xp, or the density profile of specific segments, ρ(r, s), adopt unique values that minimize the free energy given the constrained segment density, ρ(r).

Whereas the initial distribution, P(X, t), at the time t = 0, when the weak external field is switched off, is nontrivial, the ensuing dynamics of the Rouse modes in the absence of the external field is simple. Each Rouse mode independently relaxes toward their equilibrium distribution according to an Ornstein–Uhlenbeck process.36,64 The pth Rouse mode relaxes according to the Langevin equation of the Rouse model,
dXp=p2τRXpdt+21+δp,06π2τRRe2dW,
(42)
where dWα (with α = x, y, z denoting the Cartesian components) are the increments of the Wiener process, i.e., ⟨dWα⟩ = 0 and ⟨dWα(t)dWβ(t′)⟩ = δα,βδ(tt′)dt. The factor 1 + δp,0 could be easily eliminated by redefining the pth Rouse mode as Xp=2Xp for p > 0.
The molecular density, P(X, t), evolves according to the corresponding Fokker–Planck equation,
P(X,t)t=Xp2τRXP+1+δp,06π2τRRe2XP.
(43)
This time evolution conserves the normalization of the single-chain probability distribution, i.e., DX P(X, t) = 1 also for t > 0.
The conditional probability of observing the molecular configuration, X, at time, t, given the molecular configuration X(0) at time t = 0, is given by Green’s function of the Fokker–Planck equation (43). For an ideal gas (Rouse model), the Green’s function factorizes
G(X,t|X(0))=G0(X0,t|X0(0))p>0Gp(Xp,t|Xp(0))
(44)
=4Re2t3πτR3/2exp3π2τR4Re2tX0X0(0)2×p>0Re23πp21e2p2t/τR3/2×exp3π2p2Re2XpXp(0)p2t/τR21e2p2t/τR.
(45)
The convolution with the initial distribution, P(X, 0), given by Eq. (31), yields the molecular density at time t > 0,
P(X,t)=DX(0)G(X,t|X(0))P(X(0),0)
(46)
=P0(X)1Nhdsγ(s)×e(qRe)2t3π2τR(qRe)23π2p>01e2p2t/τRcos2(πps)p2×cosqX0+2p>0qXpep2t/τRcos(πps),
(47)
where
P0(X)=1Vp>0Re23πp23/2exp3π2p2Re2Xp2
(48)
denotes the equilibrium distribution of Rouse modes in the absence of an external field. The time evolution of the segment density, in turn, is obtained from P(X, t) by projection [cf. Eq. (29)],
ρ±q(t)ρ0=DXP(X,t)dse±iqX0+2p=1Xpcos(πps).
(49)
After straightforward calculation, the integrals over X can be performed and the prediction [Eq. (21)] of LRT/D-RPA is obtained, as expected. In fact, the dynamics of the Rouse modes is typically used to derive this explicit expression for the partial, dynamic, single-chain structure factors that are employed in LRT/D-RPA.36 

We use this analytic prediction to calculate the ratio of time-dependent amplitudes, a*(t) = ρ±q(t)/ρ±q(0), with * = a denoting a field that acts on all segments and * = e corresponding to an external field that only acts on one chain end. The results for different wavevectors are presented in Fig. 3. In the limit of small wavevectors, q → 0, the density modulation decays by the diffusion of the center of mass, and it becomes independent from the process by which the density modulation has been prepared. If the density variation occurs on the length scale of the molecular extension, qRe ≳ 2π, or smaller, however, the preparation process matters because the initial distribution of the internal Rouse modes, Xp with p > 0, differs. Density modulations that have been prepared by an external field on one chain end decay faster than those, which have been fabricated by a field on all segments, and this discrepancy increases with qRe. In the ultimate long-time limit, both amplitudes decay like exp(q2Dt)=exp[(qRe)2t/(3π2τR)] and the ratio of amplitudes tends to a wavevector-dependent constant.

FIG. 3.

Ratio of the time-dependent amplitudes, aa(t) and ae(t), of decaying density modulations that have been prepared by a weak spatially modulated external field with wavevector q, acting on all segments or only one end segment, respectively. The arrows on top mark τR for qRe = π and 2π. For tτR, the ratio tends toward a constant.

FIG. 3.

Ratio of the time-dependent amplitudes, aa(t) and ae(t), of decaying density modulations that have been prepared by a weak spatially modulated external field with wavevector q, acting on all segments or only one end segment, respectively. The arrows on top mark τR for qRe = π and 2π. For tτR, the ratio tends toward a constant.

Close modal
The time evolution of the molecular density, P(X, t), in an ideal gas of homopolymers follows the Fokker–Planck Eq. (43). This dynamic equation can be exactly rewritten in form of a deterministic D-DFT equation as65,66
P(X,t)t=1+δp,06π2τRRe2XPXσp2+XlnPP(X,t)t=ΓXPXδδPF[P]nkBT.
(50)
F[P] takes the form of the canonical free-energy functional of PDFT [vide infra Eq. (A24)],
F[P]kBT=lnn!+E[P]kBT+nDXP(X)lnP(X),
(51)
where E[P]=nDXP(X)Hb(X)=kBTρ0VNDXP(X)p>0Xp22σp2 denotes the average, bonded energy in the case of an ideal gas of chain molecules. Comparing Eqs. (43) and (50), we identify the mobility, Γ,
Γ=1+δp,06π2τRRe2.
(52)
The key observation is that the Fokker–Planck equation (43) for the Rouse modes has the form of a deterministic D-DFT65,66 [Eq. (50)] with the free-energy functional of equilibrium PDFT in the absence of nonbonded interactions. Thus, by using the molecular density, P(X), instead of the segment density, ρ(r), for the macromolecular liquid, we obtain a simple, memory-free time evolution.

The simple form of Eq. (50) suggests that it can be generalized to interacting multichain systems by additionally including nonbonded interactions in E[P] of Eq. (51). The free-energy functional including nonbonded interactions, ΔF, and its variation with respect to P are given in Eqs. (A24) and (A28), respectively. This D-DFT form of the polymer dynamics also allows us to establish connections to the wealth of knowledge about the dynamics of simple and complex fluids.65,66

Starting from the Langevin equation of interacting polymer segments, Fredrickson and Helfand,69 Grzetic, Wickham, and Shi,70 and Fredrickson and Orland71 employed the Martin–Siggia–Rose formalism72 to derive a Fokker–Planck equation for the molecular density as an intermediate step toward dynamic mean-field theory (DMFT),
P(R,t)t=kBTζR2P1ζRK,P
(53)
where ζ stands for the segmental friction and K=Kb+KnbR3N denotes the sum of bonded and nonbonded forces that act on all segments. The nonbonded forces, Knb, are self-consistently calculated from the molecular density, vide infra. Here, we also demonstrate that this Fokker–Planck equation for the interacting system can be cast in the form of the deterministic D-DFT [Eq. (50)], with the canonical free-energy functional, F[P], of equilibrium PDFT for the interacting system according to Eq. (A24) and a simple mobility
P(R,t)t=kBTζRPRlnPKkBT
(54)
=Re2N3π2τRRPRδδPkBTnlnn+kBTnDRP(lnP1)+E[P]=F[P]nkBT.
(55)
Thus, we identify the mean force with K=1nRδEδP, where E[P]=nDRPHb(R)+ΔF[ρ[P]] is given by Eq. (A17). ΔF[ρ] is the excess free energy of nonbonded interactions. This relation immediately clarifies the equilibrium statistical mechanics.65,66 For the three examples considered, we have solved the Fokker–Planck equation and obtained an explicit, analytic result, Eq. (47). Moreover, the Fokker–Planck equation also serves as basis for systematically deriving approximate schemes, like the ellipsoid model in Subsection III D.

Grzetic et al.70 did not have a strategy for tackling the high-dimensional Fokker–Planck equation71 and applied the scheme to monomers, where the segment density and the molecular density are identical, ρ(r) = P(R). In the monomer case, (i) the distinction between D-SCFT, using a free-energy functional of the local segment density, and D-DFT, as a functional of the molecular density, as well as (ii) the memory effects that stem from the spectrum of relaxation times in a single molecule are irrelevant.

Fredrickson and Orland71 proposed to solve the equations numerically by explicitly evolving chain replica configurations in time, i.e., the molecular density is represented by an infinitely large ensemble of independent single-chain configurations that obey Langevin equations, involving the mean force, K. This scheme was adopted in later applications.73,74 The resulting particle-based simulations are essentially identical to SCMF simulations75,76 in the mean-field limit, N̄. SCMF simulations use force-biased MC simulations,77,78 whereas DMFT uses a Langevin equation for the individual segment positions.71,73,74 Both schemes explicitly account for the strong, bonded interactions, Hb. SCMF simulations represent the weak nonbonded interactions of a segment with its surrounding by quasi-instantaneous fields, ω(r)=δΔF/δρ according to Eq. (A29)75,76 whereas DMFT uses mean-forces, −∇ω(r).71,73,74 Both quantities are calculated from the segment densities of the ensemble of chain replica configurations on a collocation lattice.

Therefore, D-DFT and DMFT/SCMF simulations for N̄ will give rise to essentially the same results, albeit by solving a deterministic partial differential equation for the molecular density, P(R), in the case of D-DFT or an infinite set of ordinary stochastic differential equations for the explicit segment positions in case of DMFT/SCMF simulations for N̄.

There is, however, one important difference between SCMF simulations for finite N̄< and DMFT/SCMF simulations for N̄. In the case of a finite ensemble of chain replica configurations, the quasi-instantaneous fields fluctuate and these fluctuations are physically meaningful.75,76 Using an appropriate discretization of space and molecular contour and frequently updating the quasi-instantaneous fields,76 SCMF simulations capture fluctuation effects, such as the shift of the transition of a symmetric diblock copolymer from second-order as predicted by Leibler’s mean-field theory15 to a fluctuation-induced first-order transition,79 capillary waves at interfaces,80 or the power-law decay of the bond–bond autocorrelation of chains in a nearly incompressible melt.81 Whereas SCMF simulations at finite N̄ capture these fluctuation and correlation effects,76 neither D-DFT nor DMFT/SCMF simulations for N̄ are expected to describe these effects for they vanish in the limit, N̄.

Qualitatively, the detailed molecular density, P(X, t), employed in D-DFT, results in a memory-free description of the dynamics because all characteristics of a single molecule are explicitly considered. P, however, is a high-dimensional probability distribution that contains much more detailed information than the segment density, ρ, used in D-SCFT7,10,17–31 or ρ and ρh employed in the generalized model-B dynamics. Moreover, the concomitant free-energy functional, F[P], may be difficult to obtain for systems with interacting polymers. Therefore, it is worth to explore simpler yet approximate descriptions.

Recently, we have made efforts to obtain the free energy as a functional of the segment density, ρ, and the variance, X12, of the first Rouse modes.82,83 This approach has the advantage to reproduce the accurate thermodynamic description of SCFT1–12 while retaining some additional information about nonequilibrium molecular conformations. Here, we take a related but different approach by using X0 and X1 to characterize the system.

The molecule’s center of mass, X0, is the slowest diffusing degree of freedom and related to the conserved number of polymer chains. The other Rouse modes with p > 0 relax with different p-dependent time scales, τp = τR/p2, according to Eq. (44). The first Rouse mode, X1, is the next slowest degree of freedom that characterizes the overall shape of a molecule around its center of mass. Progressively higher Rouse modes describe smaller-scale details of the internal molecular configurations and relax faster. Although there is no sharp time-scale separation between the Rouse modes, we try to approximate the molecular density by assuming that all Rouse modes with p > 1 instantaneously relax to their (constraint) equilibrium values, Peq(X2, X3, …|X0, X1), given X0 and X1. Thus, we approximate
P(X,t)Pell(X0,X1,t)Peq(X2,X3,|X0,X1)
(56)
and only follow the dynamics of Pell(X0, X1, t). We refer to this approximate description as the ellipsoid model because Pell(X0, X1, t) describes the system by soft particles. Each particle represents an entire polymer molecule and is characterized by its center-of-mass position, X0, and the anisotropic, overall shape, X1. In a similar vein, such a description has previously been devised by Murat and Kremer84 and others85–89 for highly coarse-grained computer simulations of polymer solutions, blends, and copolymers.
In the case of an ideal gas of macromolecules without an external field, h = 0, Peq, adopts the particularly simple form
Peq(X2,X3,|X0,X1)=p>1Re23πp23/2e3π2p2Re2Xp2,
(57)
which is independent from X0 and X1. In this case, the molecular density of the ellipsoid model coincides with the marginal of P(X, t),
Pell(X0,X1,t)=p>1dXpP(X,t)p>1dXpPeq(X2,X3,|X0,X1)=p>1dXpP(X,t).
(58)
Inserting the approximate molecular density [Eq. (56)] into the Fokker–Planck equation (43), we obtain the reduced Fokker–Planck equation of the ellipsoid model,
τRPellt=X0Re23π2X0Pell+X1X1Pell+Re26π2X1Pell,
(59)
with the concomitant Green’s function
Gell(X0,X1,t|X0(0),X1(0))=3πτR4Re2t3/2e3π2τR4Re2tX0X0(0)2×3πRe21e2t/τR3/2e3π2Re2X1X1(0)t/τR21e2t/τR.
(60)
Again, the Green’s function of the Rouse model [Eq. (44)] and that of the ellipsoid model are simply related by
Gell(X0,X1,t|X0(0),X1(0))=p>1dXpG(X,t|X(0)),
(61)
and we obtain consistently
Pell(X0,X1,t)=p>1dXpDX(0)G(X,t|X(0))P(X(0),0)
(62)
=dX0(0)dX1(0)Gell(X0,X1,t|X0(0),X1(0))×p>1dXp(0)P(X(0),0).
(63)
Performing the integrals, we derive90 
Pell(X0,X1,t)=1V(3π)3/2Re3e3π2Re2X121Nh01dsγ(s)×cosqX0+2qX1et/τRcosπs×e(qRe)2t3π2τR(qRe)23π21e2t/τRcos2(πs)p>11p2cos2(πps).
(64)
Importantly, for the dynamics in an ideal gas without an external field, Pell(X0, X1, t) can be either obtained as marginal of P(X, t), given in Eq. (47), or as the limit of Eq. (47) when the relaxation times of the higher internal Rouse modes vanish, τp → 0 for p > 1.
The projection onto the segment density results in
ρell±q(t)ρ0=Nh201ds01dsγ(s)e(qRe)2|ss|6(qRe)2t3π2τR×e2(qRe)23π21e2t/τRcosπscosπs+p>11p2cosπpscosπps.
(65)
In Fig. 4, we compare this result of the ellipsoid model to the exact Rouse result [Eq. (21)] for the decay of a density modulation with wavevector qRe = 4π. The strength of the preparation field is identical for the Rouse and the ellipsoid model. The instantaneous relaxation of the higher internal Rouse modes, p > 1, at time t = 0+, reduces the amplitude of the density modulation at t = 0+. Only after the relaxation time of the second Rouse mode, t > τ2 = τR/22, do the ellipsoid model and the Rouse model start to agree. This observation suggests that the description at early times can be systematically improved by retaining the time evolution of higher Rouse modes, X3, X4, …, instead of assuming that they instantaneously relax to their equilibrium values. In turn, if we only accounted for the zeroth Rouse mode, i.e., the center of mass, and assumed all internal Rouse modes to equilibrate instantaneously, we would obtain a single-exponential decay with rate 1/τq = q2D. Indeed, this relaxation due to center-of-mass diffusion is compatible with the ultimate long-time decay, tτR.
FIG. 4.

Comparison of the temporal decay of a shallow density modulation with qRe = 4π, obtained by the Rouse model64 [solid lines, Eq. (21)] and the ellipsoid model [dashed lines with symbols, Eq. (65)], which assumes that higher internal Rouse modes with p > 1 instantaneously equilibrate. Three preparation processes are indicated in the key. The red dashed-dotted vertical line marks the relaxation time of the second Rouse mode, τ2 = τR/4—the slowest mode assumed to be instantaneous relaxed. The red dashed line indicates an exponential decay with rate 1/τq=q2D=(qR)23π2τR.

FIG. 4.

Comparison of the temporal decay of a shallow density modulation with qRe = 4π, obtained by the Rouse model64 [solid lines, Eq. (21)] and the ellipsoid model [dashed lines with symbols, Eq. (65)], which assumes that higher internal Rouse modes with p > 1 instantaneously equilibrate. Three preparation processes are indicated in the key. The red dashed-dotted vertical line marks the relaxation time of the second Rouse mode, τ2 = τR/4—the slowest mode assumed to be instantaneous relaxed. The red dashed line indicates an exponential decay with rate 1/τq=q2D=(qR)23π2τR.

Close modal

The reduction of the ellipsoid model and the observation that its probability distribution corresponds to the marginal of the molecular density for an ideal gas also help us to visualize the time evolution of the molecular density. In Fig. 5, we plot the relative deviation R(X0, X1, t) = Pell(X0, X1, t)/Pell(X0, X1, ∞) − 1 ≈ ln Pell(X0, X1, t)/Pell(X0, X1, ∞) of the joint distribution function of X0 and X1 from its equilibrium value as a function of the first two Rouse modes, X0=q̂X0 and X1=q̂X1, along the spatial modulation. The top row displays the initial molecular distribution at time t = 0+ for the three different preparation processes. One can appreciate that the joint distribution of the molecule’s center-of mass-position along the density modulation, X0, and its overall orientation and extension, X1, greatly differ although all three preparation processes result in a periodic density modulation with a single characteristic wavevector, qRe = 4π.

FIG. 5.

Relative deviation R(X0, X1, t) = Pell(X0, X1, t)/Pell(X0, X1, ∞) − 1 of the time-dependent joint distribution of qX0/2π (horizontal axis) and 6π2X1/Re (vertical axis) along the direction, q̂, of the density modulation. qX0/2π ≈ 0 and 1 correspond to low density regions, whereas the density is maximal at qX0/2π = 1/2. The three columns represent the cases that the external preparation field that created a density modulation with the wavevector qRe = 4π for t < 0 acted on all segments (left column), on one end chain (middle column), or on the middle segment of the polymer (right column). In the first case, we plot R/(hN), whereas in the latter two cases, when the field only acts on one segment, the graph displays R/h. The three rows show the results for times t = 0 (top row), t = 0.05τR (center row), and t = 0.1τR (bottom row).

FIG. 5.

Relative deviation R(X0, X1, t) = Pell(X0, X1, t)/Pell(X0, X1, ∞) − 1 of the time-dependent joint distribution of qX0/2π (horizontal axis) and 6π2X1/Re (vertical axis) along the direction, q̂, of the density modulation. qX0/2π ≈ 0 and 1 correspond to low density regions, whereas the density is maximal at qX0/2π = 1/2. The three columns represent the cases that the external preparation field that created a density modulation with the wavevector qRe = 4π for t < 0 acted on all segments (left column), on one end chain (middle column), or on the middle segment of the polymer (right column). In the first case, we plot R/(hN), whereas in the latter two cases, when the field only acts on one segment, the graph displays R/h. The three rows show the results for times t = 0 (top row), t = 0.05τR (center row), and t = 0.1τR (bottom row).

Close modal

If the external preparation field acted on all segments, molecules with a center of mass, X0, in a high density region, qX0/2π = 1/2, are characterized by a narrower distribution of X1, indicating that the chains are less extended than those, whose center of mass is localized in a low density region. If the external preparation field acted only on the middle segment of a polymer, the distribution of X1 is largely independent for the molecule’s center-of-mass position. In these two cases, the probability distribution is symmetric for X1 ↔ −X1. This symmetry, however, is broken if the external preparation field, γ(s) = δ(s)/N, acted only on one chain end. Given these pronounced differences in the initial molecular density, the significant influence of the preparation process on the dynamics even in this simple system can be rationalized.

The subsequent rows depict the molecular density of the ellipsoid model at times, t/τR = 0.05 and 0.1. Whereas the amplitude of the variations, R(X0, X1, t), decreases in time, the pattern does not qualitatively change with time. Only for the middle column, γ(s) = δ(s)/N, the pattern rotates clockwise in the X0X1 plane. Again, one can clearly observe that the structure that has been prepared by an external field acting on an end segment (middle column) decays much faster than that prepared by an external field on the middle segment.

We have studied the temporal decay of a density modulation with wavevector q in an ideal gas of homopolymers. The thermodynamics of a shallow density modulation can be accurately described by SCFT or PDFT that result in identical predictions for noninteracting, continuous Gaussian chains. The dynamics of the molecules is described by the Rouse model.64 The center of mass, X0, of a molecule diffuses, whereas the internal degrees of freedom, quantified by higher Rouse modes, Xp with p > 0, exhibit a spectrum of relaxation times up to the Rouse time, τR.

We prepare the density modulation with length scale ≲Re by an external field that acts on a portion, γ(s), of a polymer chain for t < 0. The initial distribution of the center of mass and the internal Rouse modes (Fig. 5) depends on the preparation process, specified by γ(s), even if the projected segment density, ρ(r) is identical.

Therefore, the basic assumption of D-SCFT7,10,17–31 that the segment density is the only slow order parameter and all other degrees of freedom are in instantaneous equilibrium with this segment density, ρ(r, t), is overtly inaccurate. Whereas D-SCFT with a wavevector-dependent Onsager coefficient predicts an exponential decay of the amplitude of the density modulation in time, a qualitatively different short-time evolution is predicted by exact analytic calculations within the Rouse model64 and observed in particle-based simulations. Thus, while SCFT has proven extremely successful in providing an exquisite level of insights into the structure and thermodynamics of multicomponent polymer systems, its generalization to the dynamics remains a challenge.

We have used LRT/D-RPA38–42 to predict the decay of a density modulation for t > 0. Within this framework, we have to describe the time evolution of both, the density of segments, ρh, on which the external preparation field acted at t < 0, and the density of other segments, ρn = ρρh. LRT can be exactly rewritten in terms of a generalized model-B dynamics31,33,40 [Eq. (16)] with Onsager coefficients that are nonlocal in space and time. Equation (24) relates the Laplace transform of these Onsager coefficients to dynamic, single-chain properties and agrees with previous studies of incompressible, multicomponent systems.33 In the special case that the external preparation field acted on all segments, ρ = ρh, only a single Onsager coefficient, Λ(t), is required, and its time dependence is presented in Fig. 2—this memory effect accounts for the spectrum of relaxation times of the internal Rouse modes. One can recover the D-SCFT-form by approximating the Laplace transform Λ̃(ω) by its zero-frequency limit, i.e., Λk(t)Λ̄kδ(t), with Λ̄k=0dtΛk(t)=Sk/(Nk2τk), where τk defines a single wavevector-dependent relaxation time.28,29,31,33

In analogy to equilibrium PDFT, we can alternatively utilize the entire molecular density, P, of single-chain configurations as order parameter. Here, we use the Rouse modes, Xp with p = 0, …, N − 1, to completely specify a single-chain configuration because their dynamics in an ideal gas decouple and are described by independent Ornstein–Uhlenbeck processes according to the Rouse model. The corresponding Fokker–Planck equation (43) is equivalent to a deterministic D-DFT [Eq. (50)]65,66 with the free-energy functional of PDFT and a local mobility [Eq. (52)].

Additionally, we have explored the ellipsoid model that only considers the joint probability distribution of the first two Rouse modes, X0 and X1, but assumes that all higher internal Rouse modes instantaneously relax to their equilibrium distribution. Thereby, we describe the polymer configuration by its center of mass, X0, and its overall orientation and extension, X1. Such an approximate description of the molecular density turns out to be fairly accurate for the studied example (at times larger than τ2 = τR/4) and provides some intuition for the time evolution of the molecular density.

In the present paper, we have only focused on the simplest case of short-time dynamics in an ideal gas of homopolymers. The more relevant cases that the polymers are subjected to an external field or mutually interact remain to be explored. Given related approaches70,71 discussed in Sec. III C, we expect that the D-DFT description also can be transferred to systems of interacting chains. Additionally, the noncrossability of the molecular backbones results in entangled dynamics for long, mutually interacting polymers. Whereas these effects can be captured by slip-link78,91–93 and slip-spring models94–96 in a particle-based description, such as DMFT and SCMF simulations, the generalization of D-DFT toward entangled dynamics remains an important challenge.97–99 

Generally, the kinetics of structure formation and transformation in polymer systems is important because these complex macromolecular fluids rarely reach full thermodynamic equilibrium and, often, the experimentally observed state depends on the process history.100,101 Understanding the short-time dynamics is important, for instance, to predict the early stages of spinodal decomposition,20,32 and the formation or broadening of interfaces27,102 in polymer blends, or process-directed self-assembly14 after a sudden quench of thermodynamic condition (e.g., pressure103), mechanical deformation,104,105 or a chemical reaction106 in a periodically modulated phase of copolymers. After the quench, the copolymer system spontaneously relaxes on the free-energy landscape into one of a multitude of metastable, periodic mesostructures that have no analog in the equilibrium phase diagram. Since the structural changes often only involve the scale, Re, of a single unit cell of the periodically modulated structure, the relaxation toward the metastable state occurs on the time scale, τR. More generally, the short-time kinetics of structure formation may template the final morphology.

Another potential application that requires the quantitative prediction of the short-time dynamics is the Heterogeneous Multiscale Method (HMM)107–110 or alternate projective integration methods where the dynamics of slow variables, e.g., the segment density, is inferred from the observation of the short-time dynamics. These techniques are promising strategies for extending the simulation time scale of these fascinating complex fluids.

M.M. acknowledges stimulating discussions with Kostas Ch. Daoulas, Jörg Rottler, David Steffen, and De-Wen Sun. Financial support by the Deutsche Forschungsgemeinschaft (DFG) Mu 1674/15-2 is gratefully acknowledged. We thank the von Neumann Institute for Computing (NIC) for access to the supercomputer JUWELS at the Jülich Supercomputing Centre (JSC). Additional computational resources at the HLRN Berlin/Göttingen were also made available to the project.

The author has no conflict of interests to declare.

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Within the grand canonical ensemble, we explicitly verify that PDFT and SCFT yield the same equilibrium distribution.9 We consider a system with polymer activity, z, volume, V, and temperature, T. Whereas the basic degree of freedom of the free-energy functional, F[ρ], of SCFT is the local density, ρ(r), PDFT provides a description in terms of the molecular density, i.e., the unnormalized probability distribution, P, of single-chain configurations. A single-chain configuration is specified either by the N segment positions, R = {r1, …, rN} or, alternatively but equivalently, by the N Rouse modes, {X0, …, XN−1}. Since the molecular density, P(R), provides more information than the segment density, ρ(r), one can obtain the segment density by a projection of the molecular density,
ρ(r)=DRzP(R)i=1Nδ(rri).
(A1)
Integrating Eq. (A1) over space, we obtain the normalization of the molecular density, DRzP = ⟨n⟩. The starting point of PDFT is the grand canonical partition function,
Zgcn=0znn!j=1nDRjej=1NHb(Rj)+ΔF[ρ̂]kBT,
(A2)
where Hb stands for the bonded interactions that are additive for the different chains. For the standard Gaussian chain model, bonded interactions along a single macromolecule can be expressed in terms of the segment positions or the Rouse modes,6,36
HbkBT=i=1N13(N1)2Re2ri+1ri2=p=1N1Xp22σp2,
(A3)
with σp2Re2/(6π2p2) for N → ∞. The expression in terms of the Rouse modes is more convenient because Hb(X) is also additive for the different Rouse modes, Xjp of a given polymer j.
ΔF[ρ] denotes the free energy of nonbonded interactions that depends on the segment coordinates only via the microscopic segment density, ρ̂(r), that is calculated from the microscopic segment coordinates [cf. Eq. (1)]. The hat highlights that this quantity is calculated from the segment coordinates. For an ideals gas, ΔF=0, whereas for a system with the zero-ranged, soft pairwise repulsion potential, Φ=kBTκnN/Vδ(rr), between two segments at positions, r and r′, the nonbonded interaction free energy takes the form
ΔF[ρ]=kBTκ2nNVdrρ2(r),
(A4)
where the model parameter κN is related to the inverse, isothermal compressibility.76 
Analog to ρ̂(r), we define a microscopic molecular density zP̂(R)=j=1ni=1Nδ(rirji), where the hat again indicates the dependence on all segment coordinates. With this definition, we can rewrite the microscopic segment density in a form that is the analog of Eq. (A1),
ρ̂(r)=DRzP̂(R)i=1Nδ(rri).
(A5)
Using the standard procedure of SCFT, we introduce the auxiliary field variable, W(R), that depends on the 3N degrees of freedom, R, of a single polymer,
Zgcn=0znn!DWDPj=1nDRjej=1nHb(Rj)+ΔF[ρ̂]kBTeDRWkBTzPP̂,
(A6)
n=0znn!DWDPeΔF[ρ]kBT+zDRPWkBTQn,
(A7)
with
Q[W]=DReHb(R)+W(R)kBT
(A8)
ZgcDWDPeG[W,P]kBT,
(A9)
with
G=ΔF[ρ]zDRPWkBTzQ.
(A10)
Approximating the functional integral over the auxiliary field by the saddlepoint method, we obtain
ZgcDPeΩ[P]kBTwithΩ[P]=G[W*,P],
(A11)
and the saddlepoint condition reads
0=δG[W,P]δW(R)W*=zP(R)kBTzδQ[W]δW(R)W*
(A12)
or
W*(R)=Hb(R)kBTlnP(R).
(A13)
This relation allows us to rewrite Eq. (A8) in the form of an equation of state that relates the activity, z, to the average number of molecules, ⟨n⟩,
Q[W*]=DRP(R)=nz.
(A14)
Within this mean-field approximation, the grand potential, Ω[P], as a functional of the molecular density, P(R), takes the form
Ω[P]kBT=E[P]kBT+zDRP(lnP1)
(A15)
with the nonbonded and bonded interaction (free) energy
E[P]=ΔF[ρ[P]]+zDRPHb(R)
(A16)
=ΔF[ρ[P]]+nDRPHb(R)DRPHbP,
(A17)
where we have introduced the average, [⋯]P, of a single-chain quantity with respect to the molecular distribution, P.
Minimizing this density functional, Ω[P] with respect to the molecular density, P(R), we obtain the second saddlepoint condition
0=δΩ[P]δP(R)P*=δE[P]δP(R)P*+kBTzlnP*.
(A18)
This condition yields
P*(R)=eHb(R)kBT1kBTdrδΔFδρ(r)i=1Nδ(rri)
(A19)
=expHb(R)+i=1Nω(ri)kBT,
(A20)
with ω(r)=δΔF[ρ]δρ(r). At this saddlepoint value of P, the auxiliary field takes the form W*(R)=i=1Nω(ri). Inserting this molecular distribution in equilibrium into the density functional, we calculate the equilibrium grand potential, Ω* = Ω[P*],
Ω*=kBTn+ΔF[ρ[P*]]zDRP*i=1Nω(ri).
(A21)
Starting from the grand potential, Ω[P], in Eq. (A15), we obtain the canonical free-energy functional, F[P], at given n = ⟨n⟩, V, and T by Legendre transformation in conjunction with the equation of state [Eq. (A14)],
F[P]kBT=Ω[P]kBT+nlnz
(A22)
=nlnne+E[P]+nDRPlnPDRPDRP
(A23)
=lnn!+E[P]kBT+nlnPDRPP.
(A24)
Note that the free-energy functional, F[P], of the canonical ensemble does not depend on the normalization DRP(R) of the molecular density. The change of the normalization is equivalent to multiplying a constant factor to the activity, z, or adding a constant to the chemical potential, respectively. For the considerations in the canonical ensemble, we choose DRP(R) = 1 in the main text.
The variation of the free-energy functional, F[P], yields the molecular chemical potential
δF[P]δP(R)=δE[P]δP(R)+kBTnlnPDRPDRPlnPDRP2
(A25)
=drω(r)δρ(r)δP(R)+nHbDRPDRPHbDRP2+kBTnDRPlnP[lnP]P
(A26)
=drω(r)1DRPni=1Nδ(rri)ρ(r)+nDRPHb[Hb]P+kBTlnP[lnP]P
(A27)
=nDRPi=1Nω(ri)+Hb+kBTlnPi=1Nω(ri)+Hb+kBTlnPpindependent fromR.
(A28)
In equilibrium, the molecular chemical potential is independent from R, δF[P]δP(R)P*= const, resulting in P*expHb+i=1Nω(ri)kBT, in accord with Eq. (A20). The saddlepoint of the auxiliary field is given by
ω(r)=δΔFδρ(r)ρ*(r),
(A29)
and the equilibrium density is given by Eq. (29). Equations (A20)(29) and (A29) coincide with the self-consistent set of equations that describe equilibrium in SCFT.
Inserting this equilibrium molecular density into the canonical free-energy functional [Eq. (A24)], we obtain the canonical free energy, F* = F[P*],
F*kBT=E[P*]kBT+nlnne+nlnP*P*nlnDRP*
(A30)
=nlnne+E[P*]kBTnHb+i=1Nω(ri)kBTP*nlnDRexpHb+i=1Nω(ri)kBT
(A31)
=nlnnVe+ΔF[ρ*]kBTnkBTi=1Nω(ri)P*nln1VDRexpHb+i=1Nω(ri)kBT,
(A32)
which agrees with the standard expression of the free energy in SCFT.

To verify the predictions of LRT/D-RPA38–42 and the Rouse model,64 we use a coarse-grained particle model75,76,111,112 in conjunction with the Graphics Processing Unit (GPU)-accelerated simulation code SOft coarse grained Monte carlo Acceleration (SOMA).112 We discretize the molecular contour into N = 128 segments. The interactions are given by Eq. (A3), and we use a spatially periodic, external field of strength kBTh per monomer to prepare the initial density modulation. This coarse-grained model will also allow us to extend the study to cases with soft, nonbonded interactions for which no analytic solution is available.

We study a system of spatial extension, Re, and apply periodic boundary conditions along all Cartesian directions. The simulation box contains n = 65 536 polymers, corresponding to a high invariant degree of polymerization, N̄. In the absence of nonbonded interactions, however, the chains are decoupled and no mean-field approximation needs to be invoked in SCFT. In the present study, the high density of the ideal gas merely allows for efficient parallel simulations. We use Smart Monte Carlo (SMC) moves that use the bonded forces to propose trial displacements of the segments. This Monte Carlo (MC) dynamics closely mimics the Rouse model.93 The self-diffusion coefficient is D=9.48106Re2/MCS, where each segment is attempted to be once displaced in one Monte Carlo Step (MCS). Length scales are measured in units of Re, and time scales are reported in units of the Rouse time, τR=Re23π2D=3563 SMC.

The amplitude of the initial density modulation and the strength of the external preparation field are reported in Table I. The relation between the field strength, h, and the amplitude of the initial density modulation is in quantitative agreement with the analytical predictions for the continuous Gaussian chain in the weak-field limit [Eqs. (36) and (39)] if the external preparation field acts on all segments or the two middle segments. In the case of the field acting on a single end segment, there is a small deviation of 6% that stems from the fact that (i) he = 0.5 is not small and (ii) chain discretization effects are particularly pronounced if the field acts on a single end segment.

Simulation data in Fig. 1 have been averaged over more than 400 stochastic realizations of the relaxation process. The density profiles at a given time, t, are averaged over the realizations and, subsequently, the amplitude, ρ±q(t), is extracted from the averaged profiles.

1.
E.
Helfand
and
Y.
Tagami
, “
Theory of interface between immiscible polymers
,”
J. Polym. Sci., Part B: Polym. Lett.
9
,
741
(
1971
).
2.
J. M. H. M.
Scheutjens
and
G. J.
Fleer
, “
Statistical-theory of the adsorption of interacting chain molecules. 1. Partition-function, segment density distribution, and adsorption-isotherms
,”
J. Phys. Chem.
83
,
1619
1635
(
1979
).
3.
K. M.
Hong
and
J.
Noolandi
, “
Theory of inhomogeneous multicomponent polymer systems
,”
Macromolecules
14
,
727
736
(
1981
).
4.
M. W.
Matsen
and
M.
Schick
, “
Stable and unstable phases of a diblock copolymer melt
,”
Phys. Rev. Lett.
72
,
2660
2663
(
1994
).
5.
A.-C.
Shi
,
J.
Noolandi
, and
R. C.
Desai
, “
Theory of anisotropic fluctuations in ordered block copolymer phases
,”
Macromolecules
29
,
6487
6504
(
1996
).
6.
M. W.
Matsen
, “
The standard Gaussian model for block copolymer melts
,”
J. Phys.: Condens. Matter
14
,
R21
R47
(
2002
).
7.
G. H.
Fredrickson
, “
Dynamics and rheology of inhomogeneous polymeric fluids: A complex Langevin approach
,”
J. Chem. Phys.
117
,
6810
6820
(
2002
).
8.
G. H.
Fredrickson
,
The Equilibrium Theory of Inhomogeneous Polymers
(
Clarendon Press
,
Oxford
,
2006
).
9.
M.
Müller
, “
Comparison of self-consistent field theory and Monte Carlo simulations
,” in
Soft Matter
(
John Wiley & Sons
,
2005
), pp.
179
281
.
10.
M.
Müller
and
F.
Schmid
, “
Incorporating fluctuations and dynamics in self-consistent field theories for polymer blends
,”
Adv. Polym. Sci.
185
,
1
58
(
2005
).
11.
C. A.
Tyler
and
D. C.
Morse
, “
Orthorhombic Fddd network in triblock and diblock copolymer melts
,”
Phys. Rev. Lett.
94
,
208302
(
2005
).
12.
A.-C.
Shi
and
B.
Li
, “
Self-assembly of diblock copolymers under confinement
,”
Soft Matter
9
,
1398
1413
(
2013
).
13.
M.
Müller
and
J. J.
de Pablo
, “
Computational approaches for the dynamics of structure formation in self-assembling polymeric materials
,”
Annu. Rev. Mater. Sci.
43
,
1
34
(
2013
).
14.
M.
Müller
, “
Process-directed self-assembly of copolymers: Results of and challenges for simulation studies
,”
Prog. Polym. Sci.
101
,
101198
(
2020
).
15.
L.
Leibler
, “
Theory of microphase separation in block co-polymers
,”
Macromolecules
13
,
1602
1617
(
1980
).
16.
P. C.
Hohenberg
and
B. I.
Halperin
, “
Theory of dynamic critical phenomena
,”
Rev. Mod. Phys.
49
,
435
479
(
1977
).
17.
J. G. E. M.
Fraaije
,
B. A. C.
van Vlimmeren
,
N. M.
Maurits
,
M.
Postma
,
O. A.
Evers
,
C.
Hoffmann
,
P.
Altevogt
, and
G.
Goldbeck-Wood
, “
The dynamic mean-field density functional method and its application to the mesoscopic dynamics of quenched block copolymer melts
,”
J. Chem. Phys.
106
,
4260
4269
(
1997
).
18.
T.
Kawakatsu
,
M.
Doi
, and
R.
Hasegawa
, “
Dynamic density functional approach to phase separation dynamics of polymer systems
,”
Int. J. Mod. Phys. C
10
,
1531
1540
(
1999
).
19.
G. J. A.
Sevink
,
A. V.
Zvelindovsky
,
B. A. C.
van Vlimmeren
,
N. M.
Maurits
, and
J. G. E. M.
Fraaije
, “
Dynamics of surface directed mesophase formation in block copolymer melts
,”
J. Chem. Phys.
110
,
2250
2256
(
1999
).
20.
E.
Reister
,
M.
Müller
, and
K.
Binder
, “
Spinodal decomposition in a binary polymer mixture: Dynamic self-consistent-field theory and Monte Carlo simulations
,”
Phys. Rev. E
64
,
041804
(
2001
).
21.
H.
Morita
,
T.
Kawakatsu
, and
M.
Doi
, “
Dynamic density functional study on the structure of thin polymer blend films with a free surface
,”
Macromolecules
34
,
8777
8783
(
2001
).
22.
J. G. E. M.
Fraaije
,
A. V.
Zvelindovsky
, and
G. J. A.
Sevink
, “
Computational soft nanotechnology with mesodyn
,”
Mol. Simul.
30
,
225
238
(
2004
).
23.
E.
Reister
and
M.
Müller
, “
Formation of enrichment layers in thin polymer films: The influence of single chain dynamics
,”
J. Chem. Phys.
118
,
8476
8488
(
2003
).
24.
X.
He
and
F.
Schmid
, “
Using prenucleation to control complex copolymeric vesicle formation in solution
,”
Macromolecules
39
,
8908
8910
(
2006
).
25.
L.
Zhang
,
A.
Sevink
, and
F.
Schmid
, “
Hybrid lattice Boltzmann/dynamic self-consistent field simulations of microphase separation and vesicle formation in block copolymer systems
,”
Macromolecules
44
,
9434
9447
(
2011
).
26.
X.
Wan
,
T.
Gao
,
L.
Zhang
, and
J.
Lin
, “
Ordering kinetics of lamella-forming block copolymers under the guidance of various external fields studied by dynamic self-consistent field theory
,”
Phys. Chem. Chem. Phys.
19
,
6707
6720
(
2017
).
27.
S.
Qi
and
F.
Schmid
, “
Dynamic density functional theories for inhomogeneous polymer systems compared to Brownian dynamics simulations
,”
Macromolecules
50
,
9831
9845
(
2017
).
28.
T.
Ghasimakbari
and
D. C.
Morse
, “
Dynamics and viscoelasticity of disordered melts of symmetric diblock copolymers
,”
Macromolecules
52
,
7762
7778
(
2019
).
29.
S.
Mantha
,
S.
Qi
, and
F.
Schmid
, “
Bottom-up construction of dynamic density functional theories for inhomogeneous polymer systems from microscopic simulations
,”
Macromolecules
53
,
3409
3423
(
2020
).
30.
F.
Schmid
and
B.
Li
, “
Dynamic self-consistent field approach for studying kinetic processes in multiblock copolymer melts
,”
Polymers
12
,
2205
(
2020
).
31.
J.
Rottler
and
M.
Müller
, “
Kinetic pathways of block copolymer directed self-assembly: Insights from efficient continuum modeling
,”
ACS Nano
14
,
13986
13994
(
2020
).
32.
K.
Binder
, “
Collective diffusion, nucleation, and spinodal decomposition in polymer mixtures
,”
J. Chem. Phys.
79
,
6387
6409
(
1983
).
33.
G.
Wang
,
Y.
Ren
, and
M.
Müller
, “
Collective short-time dynamics in multicomponent polymer melts
,”
Macromolecules
52
,
7704
7720
(
2019
).
34.
R.
Zwanzig
, “
Memory effects in irreversible thermodynamics
,”
Phys. Rev.
124
,
983
992
(
1961
).
35.
H.
Mori
, “
Transport, collective motion, and Brownian motion
,”
Prog. Theor. Phys.
33
,
423
455
(
1965
).
36.
M.
Doi
and
S. F.
Edwards
,
The Theory of Polymer Dynamics
(
Oxford University Press
,
New York
,
1994
).
37.

Short-time dynamics refers to the universal dynamics of the polymers on the scale of a fraction of the coil size, Re. This dynamics is dictated by the self-similar structure and dynamics of Gaussian polymer chains64 and not by the dynamics on the scale of a monomeric repeating units that is determined by the specific chemistry of the monomers.

38.
F.
Brochard
and
P. G.
de Gennes
, “
Dynamics of compatible polymer mixtures
,”
Physica A
118
,
289
299
(
1983
).
39.
A. Z.
Akcasu
,
M.
Benmouna
, and
H.
Benoit
, “
Application of random phase approximation to the dynamics of polymer blends and copolymers
,”
Polymer
27
,
1935
1942
(
1986
).
40.
A.
Semenov
, “
Relaxation of long-wavelength density fluctuations in a concentrated polymer solution
,”
Zh. Eksp. Teor. Fiz.
90
,
1230
1235
(
1986
).
41.
M.
Benmouna
,
H.
Benoit
,
M.
Duval
, and
Z.
Akcasu
, “
Theory of dynamic scattering from ternary mixtures of two homopolymers and a solvent
,”
Macromolecules
20
,
1107
1112
(
1987
).
42.
M.
Benmouna
,
H.
Benoit
,
R.
Borsali
, and
M.
Duval
, “
Theory of dynamic scattering from copolymer solutions using the random phase approximation
,”
Macromolecules
20
,
2620
2624
(
1987
).
43.
B.
Li
,
K.
Daoulas
, and
F.
Schmid
, “
Dynamic coarse-graining of polymer systems using mobility functions
,”
J. Phys.: Condens. Matter
33
,
194004
(
2021
).
44.
A. M.
Somoza
and
P.
Tarazona
, “
Density functional approximation for hard-body liquid crystals
,”
J. Chem. Phys.
91
,
517
527
(
1989
).
45.
E.
Velasco
,
L.
Mederos
, and
D. E.
Sullivan
, “
Density-functional theory of inhomogeneous systems of hard spherocylinders
,”
Phys. Rev. E
62
,
3708
3718
(
2000
).
46.
A.
Chrzanowska
,
P. I. C.
Teixeira
,
H.
Ehrentraut
, and
D. J.
Cleaver
, “
Ordering of hard particles between hard walls
,”
J. Phys.: Condens. Matter
13
,
4715
4726
(
2001
).
47.
R. C.
Hidalgo
,
D. E.
Sullivan
, and
J. Z. Y.
Chen
, “
Smectic ordering of homogeneous semiflexible polymers
,”
Phys. Rev. E
71
,
041804
(
2005
).
48.
C. E.
Woodward
and
A.
Yethiraj
, “
Density functional theory for inhomogeneous polymer solutions
,”
J. Chem. Phys.
100
,
3181
3186
(
1994
).
49.
A.
Yethiraj
and
C. E.
Woodward
, “
Monte Carlo density functional theory of nonuniform polymer melts
,”
J. Chem. Phys.
102
,
5499
5505
(
1995
).
50.
A.
Yethiraj
, “
Density functional theory of polymers: A Curtin-Ashcroft type weighted density approximation
,”
J. Chem. Phys.
109
,
3269
3275
(
1998
).
51.
M.
Müller
and
L. G.
MacDowell
, “
Interface and surface properties of short polymers in solution: Monte Carlo simulations and self-consistent field theory
,”
Macromolecules
33
,
3902
3923
(
2000
).
52.
Y.-X.
Yu
and
J.
Wu
, “
Density functional theory for inhomogeneous mixtures of polymeric fluids
,”
J. Chem. Phys.
117
,
2368
2376
(
2002
).
53.
A. L.
Frischknecht
,
J. D.
Weinhold
,
A. G.
Salinger
,
J. G.
Curro
,
L. J.
Douglas Frink
, and
J. D.
McCoy
, “
Density functional theory for inhomogeneous polymer systems. I. Numerical methods
,”
J. Chem. Phys.
117
,
10385
10397
(
2002
).
54.
A. L.
Frischknecht
,
J. G.
Curro
, and
L. J.
Douglas Frink
, “
Density functional theory for inhomogeneous polymer systems. II. Application to block copolymer thin films
,”
J. Chem. Phys.
117
,
10398
10411
(
2002
).
55.
M.
Müller
,
L. G.
MacDowell
, and
A.
Yethiraj
, “
Short chains at surfaces and interfaces: A quantitative comparison between density-functional theories and Monte Carlo simulations
,”
J. Chem. Phys.
118
,
2929
2940
(
2003
).
56.
M.
Müller
and
L. G.
MacDowell
, “
Wetting of polymer liquids: Monte Carlo simulations and self-consistent field
,”
J. Phys.: Condens. Matter
15
,
R609
R653
(
2003
).
57.
J.
Forsman
and
C. E.
Woodward
, “
An improved density functional description of hard sphere polymer fluids at low density
,”
J. Chem. Phys.
119
,
1889
1892
(
2003
).
58.
N.
Patel
and
S. A.
Egorov
, “
Interactions between colloidal particles in polymer solutions: A density functional theory study
,”
J. Chem. Phys.
121
,
4987
4997
(
2004
).
59.
P.
Bryk
,
K.
Bucior
,
S.
Sokołowski
, and
G.
Żukociński
, “
Phase transition of short linear molecules adsorbed on solid surfaces from a density functional approach
,”
J. Phys. Chem. B
109
,
2977
2984
(
2005
).
60.
J.
Wu
, “
Density functional theory for chemical engineering: From capillarity to soft materials
,”
AIChE J.
52
,
1169
1193
(
2006
).
61.
J.
Wu
and
Z.
Li
, “
Density-functional theory for complex fluids
,”
Annu. Rev. Phys. Chem.
58
,
85
112
(
2007
).
62.
P.
Bryk
and
L. G.
MacDowell
, “
Solvation effects for polymers at an interface: A hybrid self-consistent field–density functional theory approach
,”
J. Chem. Phys.
135
,
204901
(
2011
).
63.
S. A.
Egorov
,
A.
Milchev
,
P.
Virnau
, and
K.
Binder
, “
A new insight into the isotropic–nematic phase transition in lyotropic solutions of semiflexible polymers: Density-functional theory tested by molecular dynamics
,”
Soft Matter
12
,
4944
4959
(
2016
).
64.
P. E.
Rouse
, “
A theory of the linear viscoelastic properties of dilute solutions of coiling polymers
,”
J. Chem. Phys.
21
,
1272
1280
(
1953
).
65.
U. M. B.
Marconi
and
P.
Tarazona
, “
Dynamic density functional theory of fluids
,”
J. Chem. Phys.
110
,
8032
8044
(
1999
).
66.
M.
Te Vrugt
,
H.
Löwen
, and
R.
Wittkowski
, “
Classical dynamical density functional theory: From fundamentals to applications
,”
Adv. Phys.
69
,
121
247
(
2020
).
67.
H.
Meyer
,
P.
Pelagejcev
, and
T.
Schilling
, “
Non-Markovian out-of-equilibrium dynamics: A general numerical procedure to construct time-dependent memory kernels for coarse-grained observables
,”
Europhys. Lett.
128
,
40001
(
2020
).
68.

Taking the zero-frequency limit of Eq. (24), we find q2NΛ̃q(ω)q2NΛ̃q(0)=SqSq1(0)Sq and obtain an approximation for a wavevector-dependent Onsager coefficient Λq(t)δ(t)0dtΛq(t)=δ(t)Λ̃q(0)=δ(t)Sq/(Nq2τq),29,31 with a single relaxation time, τq.

69.
G. H.
Fredrickson
and
E.
Helfand
, “
Collective dynamics of polymer solutions
,”
J. Chem. Phys.
93
,
2048
2061
(
1990
).
70.
D. J.
Grzetic
,
R. A.
Wickham
, and
A.-C.
Shi
, “
Statistical dynamics of classical systems: A self-consistent field approach
,”
J. Chem. Phys.
140
,
244907
(
2014
).
71.
G. H.
Fredrickson
and
H.
Orland
, “
Dynamics of polymers: A mean-field theory
,”
J. Chem. Phys.
140
,
084902
(
2014
).
72.
P. C.
Martin
,
E. D.
Siggia
, and
H. A.
Rose
, “
Statistical dynamics of classical systems
,”
Phys. Rev. A
8
,
423
437
(
1973
).
73.
H.
Chao
,
J.
Koski
, and
R. A.
Riggleman
, “
Solvent vapor annealing in block copolymer nanocomposite films: A dynamic mean field approach
,”
Soft Matter
13
,
239
249
(
2017
).
74.
D. J.
Grzetic
and
R. A.
Wickham
, “
Dynamical self-consistent field theory captures multi-scale physics during spinodal decomposition in a symmetric binary homopolymer blend
,”
J. Chem. Phys.
152
,
104903
(
2020
).
75.
M.
Müller
and
G. D.
Smith
, “
Phase separation in binary mixtures containing polymers: A quantitative comparison of single-chain-in-mean-field simulations and computer simulations of the corresponding multichain systems
,”
J. Polym. Sci., Part B: Polym. Lett.
43
,
934
958
(
2005
).
76.
K. C.
Daoulas
and
M.
Müller
, “
Single chain in mean field simulations: Quasi-instantaneous field approximation and quantitative comparison with Monte Carlo simulations
,”
J. Chem. Phys.
125
,
184904
(
2006
).
77.
M.
Müller
and
C.
Pastorino
, “
Cyclic motion and inversion of surface flow direction in a dense polymer brush under shear
,”
Europhys. Lett.
81
,
28002
(
2008
).
78.
M.
Müller
and
K. C.
Daoulas
, “
Single-chain dynamics in a homogeneous melt and a lamellar microphase: A comparison between smart Monte Carlo dynamics, slithering-snake dynamics, and slip-link dynamics
,”
J. Chem. Phys.
129
,
164906
(
2008
).
79.
G. H.
Fredrickson
and
E.
Helfand
, “
Fluctuation effects in the theory of microphase separation in block copolymers
,”
J. Chem. Phys.
87
,
697
705
(
1987
).
80.
A.
Werner
,
F.
Schmid
,
M.
Müller
, and
K.
Binder
, “‘
Intrinsic’ profiles and capillary waves at homopolymer interfaces: A Monte Carlo study
,”
Phys. Rev. E
59
,
728
738
(
1999
).
81.
J. P.
Wittmer
,
H.
Meyer
,
J.
Baschnagel
,
A.
Johner
,
S.
Obukhov
,
L.
Mattioni
,
M.
Müller
, and
A. N.
Semenov
, “
Long range bond-bond correlations in dense polymer solutions
,”
Phys. Rev. Lett.
93
,
147801
(
2004
).
82.
M.
Müller
,
P.
Sollich
, and
D.-W.
Sun
, “
Nonequilibrium molecular conformations in polymer self-consistent field theory
,”
Macromolecules
53
,
10457
10474
(
2020
).
83.
M.
Müller
, “
Dynamics of nonequilibrium single-chain conformations in triblock copolymers
,”
Macromolecules
54
,
6296
6311
(
2021
).
84.
M.
Murat
and
K.
Kremer
, “
From many monomers to many polymers: Soft ellipsoid model for polymer melts and mixtures
,”
J. Chem. Phys.
108
,
4340
4348
(
1998
).
85.
F.
Eurich
and
P.
Maass
, “
Soft ellipsoid model for Gaussian polymer chains
,”
J. Chem. Phys.
114
,
7655
7668
(
2001
).
86.
F.
Eurich
,
A.
Karatchentsev
,
J.
Baschnagel
,
W.
Dieterich
, and
P.
Maass
, “
Soft particle model for block copolymers
,”
J. Chem. Phys.
127
,
134905
(
2007
).
87.
T.
Vettorel
,
G.
Besold
, and
K.
Kremer
, “
Fluctuating soft-sphere approach to coarse-graining of polymer models
,”
Soft Matter
6
,
2282
2292
(
2010
).
88.
G.
D’Adamo
,
A.
Pelissetto
, and
C.
Pierleoni
, “
Polymers as compressible soft spheres
,”
J. Chem. Phys.
136
,
224905
(
2012
).
89.
G.
Zhang
,
K. C.
Daoulas
, and
K.
Kremer
, “
A new coarse grained particle-to-mesh scheme for modeling soft matter
,”
Macromol. Chem. Phys.
214
,
214
224
(
2013
).
90.
We note in passing that the argument of the last exponential is related to the mean-squared distance between the sth segment and the center of mass in an unperturbed Gaussian chain along a Cartesian direction,
12q{r(s)X0}2P=p>0(qRe)23π2p2cos2(πps).
91.
A. E.
Likhtman
, “
Single-chain slip-link model of entangled polymers: Simultaneous description of neutron spin-echo, rheology, and diffusion
,”
Macromolecules
38
,
6128
6139
(
2005
).
92.
D. M.
Nair
and
J. D.
Schieber
, “
Linear viscoelastic predictions of a consistently unconstrained Brownian slip-link model
,”
Macromolecules
39
,
3386
3397
(
2006
).
93.
A.
Ramirez-Hernandez
,
F. A.
Detcheverry
,
B. L.
Peters
,
V. C.
Chappa
,
K. S.
Schweizer
,
M.
Müller
, and
J. J.
de Pablo
, “
Dynamical simulations of coarse grain polymeric systems: Rouse and entangled dynamics
,”
Macromolecules
46
,
6287
(
2013
).
94.
V. C.
Chappa
,
D. C.
Morse
,
A.
Zippelius
, and
M.
Müller
, “
Translationally invariant slip-spring model for entangled polymer dynamics
,”
Phys. Rev. Lett.
109
,
148302
(
2012
).
95.
T.
Uneyama
and
Y.
Masubuchi
, “
Multi-chain slip-spring model for entangled polymer dynamics
,”
J. Chem. Phys.
137
,
154902
(
2012
).
96.
A. F.
Behbahani
,
L.
Schneider
,
A.
Rissanou
,
A.
Chazirakis
,
P.
Bačová
,
P. K.
Jana
,
W.
Li
,
M.
Doxastakis
,
P.
Polińska
,
C.
Burkhart
,
M.
Müller
, and
V. A.
Harmandaris
, “
Dynamics and rheology of polymer melts via hierarchical atomistic, coarse-grained, and slip-spring simulations
,”
Macromolecules
54
,
2740
2762
(
2021
).
97.
G.
Szamel
and
K. S.
Schweizer
, “
Reptation as a dynamic mean-field theory: Self and tracer diffusion in a simple model of rodlike polymers
,”
J. Chem. Phys.
100
,
3127
3141
(
1994
).
98.
K.
Miyazaki
and
A.
Yethiraj
, “
Entanglement effects in mode coupling theories of polymers
,”
J. Chem. Phys.
117
,
10448
10451
(
2002
).
99.
D. M.
Sussman
and
K. S.
Schweizer
, “
Microscopic theory of the tube confinement potential for liquids of topologically entangled rigid macromolecules
,”
Phys. Rev. Lett.
107
,
078102
(
2011
).
100.
V.
Abetz
,
K.
Kremer
,
M.
Müller
, and
G.
Reiter
, “
Functional macromolecular systems: Kinetic pathways to obtain tailored structures
,”
Macromol. Chem. Phys.
220
,
1800334
(
2019
).
101.
S.
Chandran
,
J.
Baschnagel
,
D.
Cangialosi
,
K.
Fukao
,
E.
Glynos
,
L. M. C.
Janssen
,
M.
Müller
,
M.
Muthukumar
,
U.
Steiner
,
J.
Xu
,
S.
Napolitano
, and
G.
Reiter
, “
Processing pathways decide polymer properties at the molecular level
,”
Macromolecules
52
,
7146
7156
(
2019
).
102.
C.
Yeung
and
A.-C.
Shi
, “
Formation of interfaces in incompatible polymer blends: A dynamical mean field study
,”
Macromolecules
32
,
3637
3642
(
1999
).
103.
M.
Müller
and
D.-W.
Sun
, “
Directing the self-assembly of block copolymers into a metastable complex network phase via a deep and rapid quench
,”
Phys. Rev. Lett.
111
,
267801
(
2013
).
104.
M.
Müller
and
J.
Tang
, “
Alignment of copolymer morphology by planar step elongation during spinodal self-assembly
,”
Phys. Rev. Lett.
115
,
228301
(
2015
).
105.
D.-W.
Sun
and
M.
Müller
, “
Fabrication of ellipsoidal mesostructures in block copolymers via a step-shear deformation
,”
Macromolecules
51
,
275
281
(
2018
).
106.
D.-W.
Sun
and
M.
Müller
, “
Process-accessible states of block copolymers
,”
Phys. Rev. Lett.
118
,
067801
(
2017
).
107.
W.
E
,
B.
Engquist
, and
Z.
Huang
, “
Heterogeneous multiscale method: A general methodology for multiscale modeling
,”
Phys. Rev. B
67
,
092101
(
2003
).
108.
W.
E
,
B.
Engquist
,
X.
Li
,
W.
Ren
, and
E.
Vanden-Eijnden
, “
The heterogeneous multiscale method: A review
,”
Commun. Comput. Phys.
2
,
367
450
(
2007
).
109.
E.
Vanden-Eijnden
, “
On HMM-like integrators and projective integration methods for systems with multiple time scales
,”
Commun. Math. Sci.
5
,
495
505
(
2007
).
110.
W.
E
,
W. Q.
Ren
, and
E.
Vanden-Eijnden
, “
A general strategy for designing seamless multiscale methods
,”
J. Comput. Phys.
228
,
5437
5453
(
2009
).
111.
M.
Müller
, “
Studying amphiphilic self-assembly with soft coarse-grained models
,”
J. Stat. Phys.
145
,
967
(
2011
).
112.
L.
Schneider
and
M.
Müller
, “
Multi-architecture Monte-Carlo (MC) simulation of soft coarse-grained polymeric materials: SOft coarse grained Monte-Carlo acceleration (SOMA)
,”
Comput. Phys. Commun.
235
,
463
(
2019
).
Published open access through an agreement with University of Göttingen Faculty of Physics