Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Article

Cite This: Macromolecules 2019, 52, 6116−6125 pubs.acs.org/Macromolecules

Increased Polymer Diffusivity in Thin-Film Confinement


James F. Pressly,† Robert A. Riggleman,*,‡ and Karen I. Winey*,†,‡

Department of Materials Science & Engineering and ‡Department of Chemical & Biomolecular Engineering, University of
Pennsylvania, Philadelphia, Pennsylvania 19104, United States
*
S Supporting Information

ABSTRACT: The behavior of polymer melts under planar


confinement was investigated using molecular dynamics
simulations. Polymers with a range of lengths, from
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

unentangled to highly entangled (N = 25−400), were


confined between two discrete-bead parallel walls to create
thin films with thicknesses (h = 3.5−40σ, where σ is the unit
Downloaded via YUAN ZE UNIV on December 14, 2019 at 21:38:11 (UTC).

length) ranging from much larger to much smaller than the


polymer size. These simulations were used to measure
polymer-chain conformations, entanglement densities, and
center-of-mass diffusion, and the results were compared with
previous simulations of polymer melts under cylindrical confinement. Changes to the entanglement density and radius of
gyration in thin films follow the same behavior as in cylindrical confinement: decreased entanglements per chain, decreased Rg
perpendicular to the confining wall, and increased Rg parallel to the confining wall, though the deviations from bulklike
conformations are smaller. Despite similarities in conformation and entanglement behavior between thin-film and cylindrical
confinements, the diffusion behavior differs. Under planar confinement, the diffusion coefficient increases monotonically up to 6
times the bulk diffusivity with decreasing film thickness, while it behaves nonmonotonically in cylinders. This is due to the
increased degrees of freedom afforded to the polymer chains in a thin film compared to those in a cylinder, allowing chains to
diffuse around one another rather than through, as found in cylindrical confinement. Normalized diffusion coefficients, Dnorm,
can be well described by a master curve with an exponential dependence on confinement after scaling Dnorm by a thickness-
dependent term.

■ INTRODUCTION
Understanding the effects of confinement on polymer behavior
direction, the conformation increases slightly.4−6,13,16,34,36,37
Entanglement molecular weight has also been shown to
has long been a focus of polymer physics research and has increase as film thickness or cylindrical pore radius is
implications in fields ranging from semiconductor manufactur- decreased.15−19,37 Similarly, polymer simulations also indicate
ing to polymer nanocomposites.1 Thin-film and cylindrically faster segmental motion in thin-film and cylindrical confine-
confined systems provide a simple and uniform confining ments relative to the bulk.15,22,35,38,39 Regarding chain-scale
geometry that is ideal for developing theories about the dynamics, previous simulations of unentangled polymers have
fundamental behavior of confined polymers, which can be shown that polymer diffusivity increases within the plane of the
applied to more complex confining environments, e.g., polymer thin film and slows in the out-of-plane direction.4,19,29,40 This
nanocomposites.2 Experiments, simulations, and theoretical behavior has also been observed in cylindrical confinement,35
models have extensively probed static properties of thin films, though recent simulations by our group indicate a non-
including chain conformations3−13 and entanglement net- monotonic change in diffusivity as the confining radius is
works,14−19 as well as dynamic properties, particularly decreased.37 A subset of the diffusion results from this previous
segmental relaxations15,20−24 and glass-transition behav- study is shown in Figure 1, with representative images of two
ior.25−33 Similar studies have examined polymer behavior in simulation systems, r = 2.5σ and 7σ and N = 200, where r is
cylindrical confinement, and in some cases, direct comparisons the cylindrical pore radius and σ is the bead diameter. The
have been made between the effects of thin-film and cylindrical nonmonotonic behavior is the result of chain segregation at
confinements.16,34,35 high degrees of confinement, in this case, when reff < 4σ. Chain
In general, simulations have shown polymer behavior in segregation has also been observed in ultrathin, strictly two-
thin-film and cylindrical confinements to be qualitatively dimensional (2D) polymer films,41−44 but it is unclear if the
similar, with properties increasingly deviating from bulk as level of confinement in our thin films with small but nonzero
confinement increases, though the effects are typically more thickness ∼3−40σ is high enough for polymer chains to begin
pronounced in cylindrical confinement. Several studies in thin
films and cylinders have indicated that polymer conformation Received: May 15, 2019
perpendicular to a confining interface, as measured by Rg, is Revised: July 19, 2019
substantially reduced relative to the bulk, while in the parallel Published: August 7, 2019

© 2019 American Chemical Society 6116 DOI: 10.1021/acs.macromol.9b01001


Macromolecules 2019, 52, 6116−6125
Macromolecules Article

Figure 2. Representative image showing a confined polymer


simulation system (h = 7σ, N = 200, L = 43.77σ). The confining
wall (gray beads) is partially cut away to expose the polymer. Different
colors represent different polymer chains.
Figure 1. Normalized polymer diffusion coefficient for N = 200 beads
as a function of effective pore radius. The two representative images of explanations of the pore wall generation and subsequent polymer
the cylindrically confined simulation systems correspond to r = 2.5σ equilibration can be found in our previous work.16
(upper) and r = 7σ (lower), both for N = 200. The segregation of the Thin films were generated in multiple sizes, with nominal
chains in the r = 2.5σ system creates large entropic barriers to polymer thicknesses of h = 3.5σ, 5σ, 7σ, 10σ, 14σ, 20σ, 28σ, and 40σ. The
motion that decrease the diffusion coefficient. Figure adapted from effective thickness of the films is defined by the accessible volume, vacc,
previous work by Pressly et al.37 where heff = vacc/L2, in which L is the side length of the thin
film.16,35,37 The accessible volume is determined by randomly
generating points within the simulation box and determining if the
to segregate and affect diffusivity. The extent to which this point is accessible (inside the film) or inaccessible (inside the pore
segregation behavior occurs in thin films and the resulting wall or within a minimum distance of a pore wall bead). The
effect on polymer diffusivity is the focus of this paper. minimum distance was determined by measuring the minimum
In this study, we focus on polymer behavior under rigid distance between two monomers in a Lennard-Jones fluid under the
same simulation conditions, i.e., ϕ = 0.85σ−3, T = 1, and repulsive
symmetric thin-film confinement and compare changes in Lennard-Jones potential. A detailed explanation of the accessible
polymer conformation and diffusivity in thin-film confinement volume calculation can be found in our previous publication.16 Due to
to cylindrical confinement over a range of chain lengths and excluded volume effects, the effective film thicknesses, heff, are
thicknesses. We first express the confined chain conformation approximately 0.8−0.9σ less than the nominal thicknesses. The side
using the relative shape anisotropy parameter, κ2, as a function lengths of the films are longer than the average polymer-chain end-to-
of the confinement parameter δfilm = 2Rg,bulk/heff. We find that end distance, Ree, to limit chains from interacting with themselves
κ2(δfilm) varies less in thin films than in cylindrical pores and is across periodic boundaries. We examined bulk and thin-film systems
constrained by the relative shape anisotropy of a 2D random with polymer-chain lengths of N = 25, 50, 100, 200, and 400, where N
walk for large δfilm. Additionally, we show that polymer is the number of monomers. Table 1 summarizes the average effective
diffusivity continually increases as film thickness decreases,
unlike polymers in cylindrical confinement and that this Table 1. Average Effective Film Thickness for Each Nominal
behavior can be described by an empirical master curve. The Film Thickness
increase in diffusivity is attributed mainly to a decrease in the nominal thickness, h (σ) average effective thickness, heff (σ)
number of entanglements per chain, which we describe using a 3.5 2.66
volume fraction model. Finally, we discuss why the diffusive 5 4.18
behavior of thin-film and cylindrically confined polymers 7 6.14
differs by examining the entropic barriers to diffusion. 10 9.15
Throughout this work, we will demonstrate the strong links 14 13.15
between polymer conformation and chain-scale dynamics and 20 19.15
how knowledge of both is necessary to fully understand 28 27.15
polymer behavior under confinement.


40 39.15

SIMULATION METHODS
The simulations are performed using the Kremer−Grest model45 for film thicknesses for each nominal film thickness; detailed information
coarse-grained molecular dynamics (MD) simulations with non- is provided in Table S1. The confinement parameter, δfilm, is equal to
bonded monomer interactions governed by the repulsive portion of 2Rg,bulk/heff and varies from approximately 0.1−10.
the Lennard-Jones (LJ) potential. The simulation units are After equilibration, MD simulations were run using a Langevin
normalized to the monomer size, σ, potential strength, ϵ, and time, thermostat, T = 1, with a time step of 0.006τ, and the mean-squared
τ = σ(m/ϵ)1/2, where m is the monomer mass. All simulations were displacement (MSD, ⟨r2⟩) was calculated from the chain center of
run under an NVT ensemble using the LAMMPS MD simulation mass using a moving time origin. In the thin-film systems, MSD was
package46,47 with the velocity-Verlet algorithm and a monomer calculated in the plane of the film (xy-plane) and the diffusion
density of ϕ = 0.85σ−3. The films were confined by two parallel plates coefficient, D, was calculated from the derivative of the diffusive
made of discrete immobile particles with the same size, mass, and region of the MSD such that D = (d⟨r2⟩/dt)/4. In the bulk systems,
repulsive LJ potential as the polymer monomers, Figure 2. The wall MSD was calculated in three dimensions and D = (d⟨r2⟩/dt)/6. Plots
was made of discrete particles to prevent the strong, long-range of (d⟨r2⟩/dt) as a function of t are provided in Figure S1 and show a
ordering of the polymer monomers at the interface. Detailed plateau at long times, indicative of diffusive motion.

6117 DOI: 10.1021/acs.macromol.9b01001


Macromolecules 2019, 52, 6116−6125
Macromolecules Article

Entanglement properties were studied using the CReTA


algorithm,48 which uses hard-core interactions and random string-
aligning moves to reduce the chain conformations to a primitive path
network. The entanglement network statistics were calculated by
taking multiple frames from each system, with the time between
successive frames on the order of the time required to reach the
diffusive regime, ensuring that conformations were sufficiently
decorrelated. The average number of entanglements per chain, ⟨Z⟩,
was determined using the S-kink method. The same-chain
conformations were also analyzed using the Z1 primitive path
algorithm.49−51 Despite previous simulations indicating that the
CReTA and Z1 algorithms provide similar results,16,51,52 we found
discrepancies between the algorithms in the thinnest films. Based on
these results, we use the CReTA algorithm to calculate the
entanglement networks presented in this study. More details and a
discussion of this observation are provided in the Supporting
Information (Figure S2). The polymer radius of gyration, Rg, was
calculated from similarly spaced frames.
In addition to the Rg calculated for the MD simulations, Rg,RW was
calculated for polymer conformations generated using a random walk
(RW) model. In this model, random walk polymer-chain
conformations of N steps of length 1.3σ were generated in
confinement and in the bulk. The step size was chosen to match
the bulk Rg from the random walk model to the MD simulations. In
the RW simulations, the confinement is defined by smooth parallel
plates separated by the same thicknesses used in the MD systems.
When generating the RW conformations, the first end of the polymer
chain was randomly placed within the film thickness and new bead
positions were rejected if they fell outside the confining walls. For
comparison with the bulk MD simulations, the RW simulations were
performed without confinement. Inclusion of the random walk model
allows us to distinguish the effect of excluded volume on the polymer
conformation in confinement.

■ RESULTS AND DISCUSSION


Chain Conformation in Planar Confinement. The
average radii of gyration in the unconfined (in-plane, Rg,xy)
and confined (out-of-plane, Rg,z) dimensions are plotted as a
function of chain length for both the MD simulations (Figure
3a) and the confined random walk model (Figure 3b). The
average one-dimensional radii of gyration for the bulk, Rg,bulk1D
= Rg,bulk/√3, are also shown. In both the MD simulations and
RW model, Rg,z is always less than Rg,bulk1D and the difference
between Rg,z and Rg,bulk1D increases with increasing chain length
and decreasing h. We also observe that Rg,z appears to
asymptote as chain length increases. This asymptote
corresponds to the radius of gyration of the cross section of Figure 3. Average Rg perpendicular (Rg,z, out-of-plane, filled markers)
a rectangular slit, heff/2√3. and parallel (Rg,xy, in-plane, open markers) to the confining walls,
Within the plane of the thin film, we observe very different calculated using (a) MD simulations and (b) a confined random walk
behavior between the MD simulations and RW model. In the model. The solid line corresponds to Rg,bulk1D. In both the MD
RW model, Rg,xy equals Rg,bulk1D, while in the MD simulations, simulations and RW model, Rg,z decreases significantly, while Rg,xy
Rg,xy is greater than Rg,bulk1D and the difference increases with increases in the MD simulations and remains bulklike in the RW
increasing chain length and decreasing h. This is due to the model with the points lying on the Rg,bulk1D line.
lack of excluded volume interactions in the RW model that are
present in the MD simulations. The presence of excluded
volume interactions in the MD simulations means that Rg,xy cylindrical confinements, respectively. The normalized Rg,⊥
must increase as the film thickness is decreased. behavior is nearly identical in thin-film and cylindrical
Figure 4 compares the normalized Rg perpendicular and confinements, while Rg,∥ increases more dramatically and at
parallel to the confining wall (Rg,⊥ and Rg,∥) for polymers in lower confinement in cylindrical confinement. These findings
planar and cylindrical confinements as a function of their are in agreement with the previous models by Sussman et al.
respective confinement parameters, δfilm = 2Rg,bulk/heff for comparing chain conformations in thin-film and cylindrical
planar confinement and δcyl = Rg,bulk/reff for cylindrical confinements.16
confinement, where reff is the effective pore radius. We see We further examined chain conformation in thin-film
that the changes in Rg are well described by the confinement confinement by computing the average shape anisotropy
parameters; Rg(δfilm) and Rg(δcyl) collapse to single curves in parameter, ⟨κ2⟩, of each simulated system. The shape
the perpendicular and parallel directions for planar and anisotropy parameter is defined as
6118 DOI: 10.1021/acs.macromol.9b01001
Macromolecules 2019, 52, 6116−6125
Macromolecules Article

Figure 4. Normalized Rg perpendicular and parallel to the confining


walls, for planar (blue, filled) and cylindrical (green, open)
confinements as a function of the confinement parameters, δfilm =
2Rg,bulk/heff for planar confinement and δcyl = Rg,bulk/reff for cylindrical
confinement. The change in Rg,⊥ is the same in planar and cylindrical
confinements, while Rg,∥ increases more and at lower levels of
confinement for cylindrical confinement. Cylindrical confinement
data is taken from the previous work by Pressly et al.37

4 4 4
3 λ1 + λ 2 + λ3 1
κ2 = −
2 (λ12 + λ 22 + λ32)2 2 (1)

where λ21, λ22, and λ23 are the principal moments of the radius of
gyration tensor, and their sum is equal to R2g.53 A κ2 value of 1
indicates that the monomers lie on a line, while a value of 0
means the monomer distribution is spherically symmetric. For
an unconfined polymer chain, ⟨κ2⟩ ≈ 0.4. Relative shape
anisotropy calculations for the MD simulations (Figure 5a)
and RW model (Figure 5b) are in good agreement over all
δfilm. There is a small difference in ⟨κ2⟩ near δfilm = 1, where
⟨κ2⟩ for short chains (N ≤ 50) in the MD simulations remain
bulklike, while for larger chains and for all chains in the RW Figure 5. Average relative shape anisotropy, ⟨κ2⟩, of polymer chains in
planar (blue, filled) and cylindrical (green, open) confinements for
model, there is a slight decrease of ∼5% in ⟨κ2⟩. The (a) the MD simulations and (b) the RW model plotted against their
nonmonotonic behavior of ⟨κ2⟩ is due to the orientation of respective confinement parameters, δfilm = 2Rg,bulk/heff for planar and
polymer chains and is discussed in detail in our previous δcyl = Rg,bulk/reff for cylindrical confinements. The dashed line is ⟨κ2⟩
work.37 for bulk polymer chains and the dash-dotted line is for a 2D random
Figure 5 also plots ⟨κ2⟩ for polymer chains under cylindrical walk. Cylindrical confinement data is taken from the previous work by
confinement as a function of δcyl. We can see that ⟨κ2⟩ has the Pressly et al.37
same nonmonotonic behavior in cylindrical and planar
confinements, transitioning from a bulklike value for small δ
to a minimum near δ = 1 before increasing. However, the due to a decrease in entanglement density near the
magnitude of the changes in ⟨κ2⟩ is much smaller in thin-film interface.17,19,37 Our current findings agree with these previous
confinement than in cylindrical due to the reduced number of studies. Here, we adapt our empirical model, previously
confining dimensions. Additionally, it appears that ⟨κ2⟩ begins developed to describe Z as a function of pore radius in
to asymptote in thin-film confinement around δfilm = 6, or cylindrical confinement,37 to a thin-film geometry. The model
when Rg,bulk ∼ 3heff. Based on relative shape anisotropy simplifies the change in entanglement density moving away
calculations for a random walk confined to a two-dimensional from the pore wall to a step function, with an entanglement
plane, we can estimate that ⟨κ2⟩ ≈ 0.55 as the average relative density of zero within a distance td of the pore wall and a
shape anisotropy at the extreme limit of thin-film confinement, bulklike density otherwise. The average number of entangle-
as indicated by the dash-dotted line in Figure 5. The ments per confined chain is given by
asymptotic behavior near δfilm = 6 indicates that the polymer (heff − 2td)
conformation is transitioning from a three-dimensional (3D) Zconf = Z bulk
heff (2)
to pseudo-2D behavior. Finally, the observed values of ⟨κ2⟩ fall
within a small range, 0.35−0.55, indicating that thin-film where Zconf and Zbulk are the average number of entanglements
confinement does not significantly alter overall chain per chain in confinement and bulk, respectively, and td is the
conformation compared to cylindrical confinement where width of the entanglement depletion region near the interface.
⟨κ2⟩ values of 0.3−0.8 were observed.37 Figure 6a plots Zconf, calculated from the MD simulations, as
Entanglement Networks in Confinement. The number a function of heff for planar confinement and 2reff for cylindrical
of entanglements per chain, Z, has been shown to decrease confinement for the entangled systems (N ≥ 100). The solid
with increasing confinement in thin films and cylindrical pores lines are fits to the corresponding empirical model, eq 2, for
6119 DOI: 10.1021/acs.macromol.9b01001
Macromolecules 2019, 52, 6116−6125
Macromolecules Article

supported by the depletion width in cylindrical confinement


being greater than in thin-film confinement, for the same N.
While previous MD simulations of entangled polymers under
planar confinement observed entanglement depletion near the
interface, the relationship between chain length and depletion
thickness was not explored.17
Chain Segregation. In cylindrical confinement, chain
segregation has been shown to slow polymer diffusion, Figure
1.37 To accurately compare diffusive behavior in thin-film and
cylindrical confinements, it is necessary to determine if the
polymer chains segregate in thin-film confinement. To quantify
the chain segregation, we calculated the 2D radial self-chain
monomer density about the chain center of mass, ρself(rxy). The
calculation was limited to a 2.66σ slice (heff of the thinnest
system) centered on the chain center of mass. This method
gives comparable results to the ρself(r) calculation taken in
three dimensions while maintaining a simple normalization
volume and minimizing the effects of film thickness;
comparisons between 2D and 3D monomer density profiles
can be found in Figure S3. Figure 7a shows 2D radial self-chain
monomer density profiles for select systems; the distance from
the center of mass is normalized by the bulk polymer radius of
gyration, rxy/Rg,bulk. From Figure 7a, it is evident that self-chain
monomer density increases as film thickness decreases and that
this increase is greater for longer polymer chains.
Figure 7b plots the normalized self-chain monomer density
at the center of mass, ρself(rxy = 0)/ρself,bulk(rxy = 0), as a
function of δfilm. We observe that ρself(rxy = 0) is bulklike, while
δfilm < 1.7, and when δfilm > 1.7, ρself(rxy = 0) increases
approximately as δfilm0.75, though significantly more data is
needed to verify that this is truly power law behavior. In the
most confined system (N = 400, h = 3.5σ), ρself(rxy = 0) is 3
times greater than in the bulk. While defining a cutoff value for
a dispersed to segregated chain transition based on ρself(rxy =
Figure 6. (a) Entanglements per chain, Z, calculated from the MD 0) would be arbitrary, the data indicate that the polymer chains
simulations for thin-film (blue, filled) and cylindrical (green, open) are becoming weakly segregated as confinement increases.
confinements, plotted as a function of effective film thickness, heff, or Previous simulations of 2D polymer melts suggest that, in the
twice the effective pore radius, 2reff, respectively. The solid lines
represent fits to eq 2, for planar confinement, and to the eq Zconf =
ultrathin limit, the polymer chains will segregate.41−44 The
Zbulk (reff − td)2/reff2 for cylindrical confinement, where td is the fitting effect of planar confinement and chain segregation on polymer
parameter.37 The dashed lines represent the number of entanglements diffusivity is discussed in the next section.
per chain in the bulk. (b) Width of the entanglement-depleted region, Polymer Diffusion in Confinement. The diffusion
td, as a function of chain length, N, for thin-film (blue, filled) and coefficients of the confined polymers, normalized by the bulk
cylindrical (green, open) confinements. In thin-film confinement, td polymer diffusion coefficient, are plotted as a function of δfilm
varies approximately as N0.2, while in cylindrical confinement, it varies in Figure 8. It is clear that as the polymer chains become more
approximately as N0.5. Cylindrical confinement data is taken from the confined, the normalized diffusion coefficient, Dnorm = Dconf/
previous work by Pressly et al.37 Dbulk, increases monotonically. This differs from the diffusion
behavior in cylindrical confinement where Dnorm behaves
thin films and Zconf = Zbulk(reff − td)2/reff2 for cylinders,37 with nonmonotonically with changing pore radius.37 The increase in
td as the sole fitting parameter. The dashed lines correspond to diffusivity is mainly attributed to the decrease in entanglement
Zbulk. In both planar and cylindrical confinements, Zconf density as the film thickness decreases. Tung et al. calculate
decreases with decreasing confining length, though the Dnorm using both the mean-squared displacement and the
decrease is greater in cylindrical pores for similar confining reptation model, defining the reptation diffusion coefficient,
lengths (heff ≈ 2reff). In Figure 6b, the width of the depletion Drep, as
region, td, is plotted as a function of N. The depletion width
scales differently from chain length in these two confining Drep ∝ R ee2τ0−1Z −1 (3)
geometries, scaling approximately with N0.2 in planar and N0.5
in cylindrical confinement. The difference in scaling is most where Ree is the polymer end-to-end distance and τ0 is the
likely due to differences in curvature between the two monomer relaxation time.35 From eq 3, we see that the
interfaces; the negative mean curvature of the pore wall in diffusion coefficient is inversely proportional to the number of
cylindrical confinement extends the region where there is a entanglements per chain, Z, and that a decrease in Znorm =
high concentration of same-chain monomers further from the Zconf/Zbulk will lead to a corresponding increase in Dnorm.
wall compared to that of the flat wall in thin-film confinement, The decrease in the normalized number of entanglements
resulting in a stronger dependence on N for td. This is further per chain, Znorm, fully accounts for the increase in Dnorm for N =
6120 DOI: 10.1021/acs.macromol.9b01001
Macromolecules 2019, 52, 6116−6125
Macromolecules Article

Dconf
* (δfilm) =
Dnorm
f (heff )D bulk (4)

where f(heff) is a thickness-dependent shift factor, we find that


* has an exponential dependence on the confinement
Dnorm
parameter, for the range of confinements studied here (Figure
8b,c). Figure 8d plots f(heff), and we observe that f follows the
form
heff − 2ti
f (heff ) =
heff (5)

where ti is the thickness of a monomer layer near the polymer−


wall interface. In other words, f is the fraction of monomers in
the thin film that are not near the interface and presumably
behave in a bulklike manner. Note the similarities between eqs
5 and 2. Simultaneously fitting eqs 4 and 5 to the data reveals
* = e0.29δfilm, as shown in Figure 8b−8d.
that ti = 0.49σ and Dnorm
While the fitted equations above are purely empirical, note
that the thickness dependence is independent of N. This
suggests that the slowing down of Dnorm, relative to D*norm, is
due to polymer−wall interactions, possibly a frictional force.
The interaction between the polymer chains and the confining
walls slows diffusion in thinner films more than that in thicker
films (Figure 8d), even when the chain-scale confinement is
similar (i.e., similar δfilm). This is clearly observed in Figure 8a
where, for a given confinement, polymer chains in thicker films
tend to have greater diffusivities. This behavior is reminiscent
of the simulations of linear polymers in a diamond network
confinement by Zhang et al. who scaled the confined polymer
diffusivity by a tortuosity-dependent term.54 In this study, we
are scaling by a thickness-dependent term.
It would be interesting to compare the results of this study
to similar systems with free polymer interfaces. One would
Figure 7. (a) Self-monomer density, ρself(rxy), profiles as a function of expect polymer−wall interactions, frictional or otherwise, to be
radial, in-plane distance from the chain center of mass normalized by
Rg,bulk. Results are shown for N = 100 and 400 in bulk and h = 3.5σ greatly reduced or nonexistent in a system with free interfaces,
and 5σ in thin films. Self-monomer density increases with decreasing suggesting that Dnorm(δfilm) could lie on a single curve without
chain length and decreasing film thickness. (b) Normalized self- shifting. Simulations of freestanding thin films have been
monomer density near the chain center of mass, ρself(rxy = 0)/ performed for large-chain lengths, showing a similar entangle-
ρself,bulk(rxy = 0), as a function of confinement, δfilm. The normalized ment behavior; however, the diffusive behavior of polymer
self-monomer density is bulklike for approximately δfilm < 1.7, vertical chains was not examined.17 This discussion of polymer−wall
dashed line, and increases as approximately δfilm0.75 for greater interaction is most relevant for temperatures where glassy
confinement. polymer physics are not present, i.e., T ≫ Tg.
Entropic Barriers to Polymer Diffusion. Despite
polymer-chain segregation developing under increased confine-
ment, there is no evidence of polymer diffusion slowing as
100 polymer-chain lengths. For longer-chain lengths, Znorm confinement is increased. This is in contrast to previous
underpredicts the increase in Dnorm by up to 50% (Figure S5). simulations of cylindrically confined polymers where chain
The remaining increase in Dnorm can be attributed to changes segregation caused a dramatic decrease in diffusivity that could
in chain conformation and segmental relaxation (as expressed be observed even before chains fully segregated (as seen in
in eq 3), as well as other effects that are unique to confined Figure 1). The difference in diffusive behavior stems from the
polymer systems (e.g., polymer−wall friction). As chain length difference in the number of degrees of freedom polymer chains
increases, Znorm increasingly underpredicts Dnorm due to the have in each geometry. In cylindrical confinement, the polymer
increasing deviation of these parameters from their bulk values. chains can only diffuse in one dimension, along the cylinder,
Interestingly, despite the polymer chains tending toward while in thin-film confinement, polymers can diffuse in two
segregation, Dnorm monotonically increases with decreasing film dimensions, greatly reducing the entropic barrier to polymer
thickness. As explained in the next section, this is due to small diffusion. In other words, segregated polymers in cylindrical
entropic barriers to diffusion in thin films, even for very strong confinement must become highly elongated and diffuse
confinement. through each other, while segregated polymers in planar
In Figure 8a, it appears that Dnorm(δfilm) follows a similar confinement can diffuse around each other.
curve for each film thickness, suggesting that the data could be A potential of mean force analysis was performed on the
shifted to create a single curve. Shifting Dnorm by a factor of f bulk and h = 3.5σ systems to quantify the strength of the
such that entropic barrier in both unconfined and highly confined
6121 DOI: 10.1021/acs.macromol.9b01001
Macromolecules 2019, 52, 6116−6125
Macromolecules Article

Figure 8. (a) Normalized diffusion coefficients, Dnorm, plotted as a function of confinement, δfilm. (b, c) Dnorm
* , plotted as a function of δfilm, showing
that the diffusion coefficients can be shifted to a single curve described by an exponential function. (d) Shift factor, f, as a function of heff. Fits to eqs
4 and 5 are shown in (b)−(d).

systems. The potential of mean force was calculated using the in thin-film confinement are minimal and easily overcome due
equation to increased degrees of freedom.
w(r ) = −kT ln g (r ) = u(r ) − s(r )T
where w(r) is the potential of mean force between two
(6)
■ CONCLUSIONS
We used molecular dynamics simulations to examine the chain
polymer chains with centers of mass separated by distance r, conformations, entanglement densities, and diffusion coef-
g(r) is the chain center-of-mass distribution function, and u(r) ficients of polymer chains under athermal thin-film confine-
and s(r) are the average interaction energy and entropy ment. The simulation systems covered a range of chain lengths,
between two chains separated by distance r, respectively. In the N = 25−400, spanning the unentangled and entangled regimes.
bulk systems, the separation distance rCM is calculated in three The variety of film thicknesses, h = 3.5−40σ, created
dimensions, while in the confined systems, rxy,CM is calculated confinement conditions ranging from δfilm = 0.1 to 10, with
in the plane of the thin film. Additionally, T = 1, and the the most confined systems tending toward chain segregation.
Boltzmann constant, k, is 1. To calculate the entropic barrier, Additionally, a random walk (RW) model was applied over the
u(r) was calculated between pairs of polymer chains and same range of chain length and film thicknesses to isolate the
binned by the distance between the chain centers of mass. This effects of excluded volume and geometric confinement on
distribution was then subtracted from w(r) to determine −s(r). chain conformations.
Figure 9 shows the entropic contribution to the potential of The Rg perpendicular and parallel to the confining walls and
mean force for two chains separated by a distance r for the bulk the relative shape anisotropy of the polymer chains were
(Figure 9a) and h = 3.5σ confined systems (Figure 9b). For N calculated to examine chain conformations in the MD
= 400 in the bulk, −s(rCM) decreases steadily as rCM simulations and confined random walk models. In both the
approaches 0, whereas −s(rxy,CM) has a slight increase near 0 MD simulations and RW model, Rg,⊥ was significantly reduced
in the h = 3.5σ confined system. This slight increase is the relative to Rg,bulk1D; meanwhile, Rg,∥ was slightly greater than
entropic barrier for in-plane diffusion and is approximately 1− Rg,bulk1D for the MD simulations and equivalent to Rg,bulk1D in
2kT, an easily overcome barrier when the thermal energy, kT, the confined RW model, consistent with previous studies. As a
of the system is 1. Previous simulations of cylindrically function of confinement, Rg,⊥ was shown to behave similarly in
confined polymers showed entropic barriers on the order of planar and cylindrical confinements, while Rg,∥ increased more
18kT when N = 200 and r = 2.5σ, resulting in a 75% reduction rapidly in cylindrical confinement. The relative shape
in diffusivity.37 It is clear that the entropic barriers to diffusion anisotropy, ⟨κ2⟩, behavior as a function of confinement was
6122 DOI: 10.1021/acs.macromol.9b01001
Macromolecules 2019, 52, 6116−6125
Macromolecules Article

cylindrically confined systems where chain segregation


significantly decreased diffusivity, we observe no slowing of
diffusion in the thin-film systems studied, despite the polymer
chains tending toward segregation as the film thickness
decreased. This is a result of the increased degrees of freedom
afforded to the chains due to only being confined in one
dimension, resulting in minimal entropic barriers to diffusion,
even in the most confined systems.
This work demonstrates the competing effects of confining
geometry on confined polymer behavior. Despite qualitatively
similar conformational behavior, planar and cylindrical confine-
ments produce a markedly different diffusive behavior. These
findings provide an important foundation for understanding
increasingly complex confinement geometries.


*
ASSOCIATED CONTENT
S Supporting Information
The Supporting Information is available free of charge on the
ACS Publications website at DOI: 10.1021/acs.macro-
mol.9b01001.
Plots of (d⟨r2⟩/dt) as a function of t (Figure S1);
comparison of the average number of entanglements per
chain calculated using the Z1 and CReTA algorithms
(Figure S2); comparison of the self-monomer density
calculated in 3D and using the 2D slice method (Figure
S3); plots of Dnorm as a function of heff (Figure S4);
contribution of chain disentanglement to the increase in
polymer diffusivity (Figure S5); film dimensions, chain
lengths, number of chains, and confinement parameters
for each simulation system (Table S1) (PDF)

Figure 9. Entropic contribution to the potential of mean force for two


■ AUTHOR INFORMATION
Corresponding Authors
polymer chains separated by a distance r in (a) the bulk and (b) a thin *E-mail: rrig@seas.upenn.edu (R.A.R.).
film of thickness 3.5σ. In the bulk, −s(rCM) continually decreases as *E-mail: winey@seas.upenn.edu (K.I.W.).
chains get closer together, while in the thin film, the initial decrease in ORCID
−s(rxy,CM) is followed by a slight increase of approximately 1−2kT for
rxy,CM < 10σ.
James F. Pressly: 0000-0002-9365-9562
Robert A. Riggleman: 0000-0002-5434-4787
Karen I. Winey: 0000-0001-5856-3410
similar for both the MD simulations and RW model; both Notes
showed a slight dip at δ = 1 and asymptotically approached The authors declare no competing financial interest.


≈0.55, which is the ⟨κ2⟩ value of a 2D random walk. The
variation of ⟨κ2⟩ in thin-film confinement is much smaller than ACKNOWLEDGMENTS
in cylindrical confinement.
We adapted our previously developed, two-layer entangle- J.F.P., K.I.W., and R.A.R. acknowledge support from DOE-BES
ment model from a cylindrical to a thin-film geometry (eq 2) via DE-SC0016421. Computational resources were made
and found that it accurately describes the number of available at the National Energy Research Scientific Comput-
entanglements per chain as a function of film thickness and ing Center through Project 60059 and the Pittsburgh
chain length. The width of the entanglement depletion region, Supercomputing Center through XSEDE Award TG-
DMR150034.


td, was shown to scale with N0.2 compared to that of the N0.5
scaling observed in cylindrical confinement, likely due to
changes in the curvature of the interface. Additionally, the REFERENCES
number of entanglements per chain was shown to decrease (1) Russell, T. P.; Chai, Y. 50th Anniversary Perspective: Putting the
more in cylindrical confinement for similar confining lengths. Squeeze on Polymers: A Perspective on Polymer Thin Films and
Confined polymer diffusive behavior was examined by Interfaces. Macromolecules 2017, 50, 4597−4609.
(2) Rittigstein, P.; Priestley, R. D.; Broadbelt, L. J.; Torkelson, J. M.
normalizing the in-plane diffusion coefficient to the bulk Model Polymer Nanocomposites Provide an Understanding of
diffusion coefficient, and it was shown that Dnorm monotoni- Confinement Effects in Real Nanocomposites. Nat. Mater. 2007, 6,
cally increases with increasing confinement, δfilm = 2Rg,bulk/heff. 278.
Furthermore, the normalized diffusion coefficients can be (3) Kumar, S. K.; Vacatello, M.; Yoon, D. Y. Off-Lattice Monte
shifted by a thickness-dependent term to produce a master Carlo Simulations of Polymer Melts Confined between Two Plates. J.
curve with an exponential dependence on δfilm. Finally, unlike Chem. Phys. 1988, 89, 5206−5215.

6123 DOI: 10.1021/acs.macromol.9b01001


Macromolecules 2019, 52, 6116−6125
Macromolecules Article

(4) Bitsanis, I.; Hadziioannou, G. Molecular Dynamics Simulations (25) Forrest, J. A.; Dalnoki-Veress, K.; Dutcher, J. R. Interface and
of the Structure and Dynamics of Confined Polymer Melts. J. Chem. Chain Confinement Effects on the Glass Transition Temperature of
Phys. 1990, 92, 3827−3847. Thin Polymer Films. Phys. Rev. E 1997, 56, 5705−5716.
(5) Kumar, S. K.; Vacatello, M.; Yoon, D. Y. Off-Lattice Monte (26) Dalnoki-Veress, K.; Forrest, J. A.; Murray, C.; Gigault, C.;
Carlo Simulations of Polymer Melts Confined between Two Plates. 2. Dutcher, J. R. Molecular Weight Dependence of Reductions in the
Effects of Chain Length and Plate Separation. Macromolecules 1990, Glass Transition Temperature of Thin, Freely Standing Polymer
23, 2189−2197. Films. Phys. Rev. E 2001, 63, No. 031801.
(6) Pakula, T. Computer Simulation of Polymers in Thin Layers. I. (27) Tsui, O. K. C.; Zhang, H. F. Effects of Chain Ends and Chain
Polymer Melt between Neutral Walls - Static Properties. J. Chem. Entanglement on the Glass Transition Temperature of Polymer Thin
Phys. 1991, 95, 4685−4690. Films. Macromolecules 2001, 34, 9139−9142.
(7) Shuto, K.; Oishi, Y.; Kajiyama, T.; Han, C. C. Aggregation (28) Ellison, C. J.; Torkelson, J. M. Sensing the Glass Transition in
Structure of Two-Dimensional Ultrathin Polystyrene Film Prepared Thin and Ultrathin Polymer Films via Fluorescence Probes and
by the Water Casting Method. Macromolecules 1993, 26, 6589−6594. Labels. J. Polym. Sci., Part B: Polym. Phys. 2002, 40, 2745−2758.
(8) Jones, R. L.; Kumar, S. K.; Ho, D. L.; Briber, R. M.; Russell, T. P. (29) Varnik, F.; Baschnagel, J.; Binder, K. Reduction of the Glass
Chain Conformation in Ultrathin Polymer Films. Nature 1999, 400, Transition Temperature in Polymer Films: A Molecular-Dynamics
146. Study. Phys. Rev. E 2002, 65, No. 021507.
(9) Brûlet, A.; Boué, F.; Menelle, A.; Cotton, J. P. Conformation of (30) Roth, C. B.; Dutcher, J. R. Glass Transition and Chain Mobility
Polystyrene Chain in Ultrathin Films Obtained by Spin Coating. in Thin Polymer Films. J. Electroanal. Chem. 2005, 584, 13−22.
Macromolecules 2000, 33, 997−1001. (31) Roth, C. B.; Torkelson, J. M. Selectively Probing the Glass
(10) Kraus, J.; Müller-Buschbaum, P.; Kuhlmann, T.; Schubert, D. Transition Temperature in Multilayer Polymer Films: Equivalence of
W.; Stamm, M. Confinement Effects on the Chain Conformation in Block Copolymers and Multilayer Films of Different Homopolymers.
Thin Polymer Films. Europhys. Lett. 2000, 49, 210. Macromolecules 2007, 40, 3328−3336.
(11) Jones, R. L.; Kumar, S. K.; Ho, D. L.; Briber, R. M.; Russell, T. (32) Ellison, C. J.; Torkelson, J. M. The Distribution of Glass-
P. Chain Conformation in Ultrathin Polymer Films Using Small- Transition Temperatures in Nanoscopically Confined Glass Formers.
Angle Neutron Scattering. Macromolecules 2001, 34, 559−567. Nat. Mater. 2003, 2, 695.
(12) Cavallo, A.; Müller, M.; Binder, K. Unmixing of Polymer Blends (33) Jin, K.; Torkelson, J. M. Tg-Confinement Effects in Strongly
Confined in Ultrathin Films: Crossover between Two-Dimensional Miscible Blends of Poly(2,6-Dimethyl-1,4-Phenylene Oxide) and
and Three-Dimensional Behavior. J. Phys. Chem. B 2005, 109, 6544− Polystyrene: Roles of Bulk Fragility and Chain Segregation. Polymer
6552. 2017, 118, 85−96.
(13) Cavallo, A.; Müller, M.; Wittmer, J. P.; Johner, A.; Binder, K. (34) Carrillo, J.-M. Y.; Sumpter, B. G. Structure and Dynamics of
Single Chain Structure in Thin Polymer Films: Corrections to Flory’s Confined Flexible and Unentangled Polymer Melts in Highly
and Silberberg’s Hypotheses. J. Phys.: Condens. Matter 2005, 17, Adsorbing Cylindrical Pores. J. Chem. Phys. 2014, 141, No. 074904.
S1697. (35) Tung, W.-S.; Composto, R. J.; Riggleman, R. A.; Winey, K. I.
(14) Si, L.; Massa, M. V.; Dalnoki-Veress, K.; Brown, H. R.; Jones, R. Local Polymer Dynamics and Diffusion in Cylindrical Nanoconfine-
A. L. Chain Entanglement in Thin Freestanding Polymer Films. Phys. ment. Macromolecules 2015, 48, 2324−2332.
Rev. Lett. 2005, 94, No. 127801. (36) Martínez-Haya, B.; Gordillo, M. C. Effect of Macromolecular
(15) Meyer, H.; Kreer, T.; Cavallo, A.; Wittmer, J. P.; Baschnagel, J. Crowding on the Conformation of Confined Chain Polymers.
On the Dynamics and Disentanglement in Thin and Two-Dimen- Macromol. Theory Simul. 2005, 14, 421−427.
sional Polymer Filmsl. Eur. Phys. J.: Spec. Top. 2007, 141, 167−172. (37) Pressly, J. F.; Riggleman, R. A.; Winey, K. I. Polymer Diffusion
(16) Sussman, D. M.; Tung, W.-S.; Winey, K. I.; Schweizer, K. S.; Is Fastest at Intermediate Levels of Cylindrical Confinement.
Riggleman, R. A. Entanglement Reduction and Anisotropic Chain and Macromolecules 2018, 51, 9789−9797.
Primitive Path Conformations in Polymer Melts under Thin Film and (38) Varnik, F.; Baschnagel, J.; Binder, K.; Mareschal, M.
Cylindrical Confinement. Macromolecules 2014, 47, 6462−6472. Confinement Effects on the Slow Dynamics of a Supercooled
(17) Sussman, D. M. Spatial Distribution of Entanglements in Thin Polymer Melt: Rouse Modes and the Incoherent Scattering Function.
Free-Standing Films. Phys. Rev. E 2016, 94, No. 012503. Eur. Phys. J. E 2003, 12, 167−171.
(18) Lee, N.-K.; Diddens, D.; Meyer, H.; Johner, A. Local Chain (39) Vladkov, M.; Barrat, J.-L. Local Dynamics and Primitive Path
Segregation and Entanglements in a Confined Polymer Melt. Phys. Analysis for a Model Polymer Melt near a Surface. Macromolecules
Rev. Lett. 2017, 118, No. 067802. 2007, 40, 3797−3804.
(19) Li, S.; Ding, M.; Shi, T. Spatial Distribution of Entanglements (40) Baschnagel, J.; Varnik, F. Computer Simulations of Supercooled
and Dynamics in Polymer Films Confined by Smooth Walls. Polymer Polymer Melts in the Bulk and in Confined Geometry. J. Phys.:
2019, 172, 365−371. Condens. Matter 2005, 17, R851−R953.
(20) Hall, D. B.; Torkelson, J. M. Small Molecule Probe Diffusion in (41) Carmesin, I.; Kremer, K. Static and Dynamic Properties of
Thin and Ultrathin Supported Polymer Films. Macromolecules 1998, Two-Dimensional Polymer Melts. J. Phys. 1990, 51, 915−932.
31, 8817−8825. (42) Müller, M. Chain Conformations and Correlations in Thin
(21) Yang, Z.; Fujii, Y.; Lee, F. K.; Lam, C.-H.; Tsui, O. K. C. Glass Polymer Films: A Monte Carlo Study. J. Chem. Phys. 2002, 116,
Transition Dynamics and Surface Layer Mobility in Unentangled 9930−9938.
Polystyrene Films. Science 2010, 328, 1676−1679. (43) Meyer, H.; Kreer, T.; Aichele, M.; Cavallo, A.; Johner, A.;
(22) Shavit, A.; Riggleman, R. A. Physical Aging, the Local Dynamics Baschnagel, J.; Wittmer, J. P. Perimeter Length and Form Factor in
of Glass-Forming Polymers under Nanoscale Confinement. J. Phys. Two-Dimensional Polymer Melts. Phys. Rev. E 2009, 79, No. 050802.
Chem. B 2014, 118, 9096−9103. (44) Wittmer, J. P.; Meyer, H.; Johner, A.; Kreer, T.; Baschnagel, J.
(23) Mirigian, S.; Schweizer, K. S. Theory of Activated Glassy Algebraic Displacement Correlation in Two-Dimensional Polymer
Relaxation, Mobility Gradients, Surface Diffusion, and Vitrification in Melts. Phys. Rev. Lett. 2010, 105, No. 037802.
Free Standing Thin Films. J. Chem. Phys. 2015, 143, No. 244705. (45) Kremer, K.; Grest, G. S. Dynamics of Entangled Linear Polymer
(24) Hesami, M.; Steffen, W.; Butt, H.-J.; Floudas, G.; Koynov, K. Melts: A Molecular-Dynamics Simulation. J. Chem. Phys. 1990, 92,
Molecular Probe Diffusion in Thin Polymer Films: Evidence for a 5057−5086.
Layer with Enhanced Mobility Far above the Glass Temperature. ACS (46) Plimpton, S. Fast Parallel Algorithms for Short-Range
Macro Lett. 2018, 7, 425−430. Molecular Dynamics. J. Comput. Phys. 1995, 117, 1−19.

6124 DOI: 10.1021/acs.macromol.9b01001


Macromolecules 2019, 52, 6116−6125
Macromolecules Article

(47) Auhl, R.; Everaers, R.; Grest, G. S.; Kremer, K.; Plimpton, S. J.
Equilibration of Long Chain Polymer Melts in Computer Simulations.
J. Chem. Phys. 2003, 119, 12718−12728.
(48) Tzoumanekas, C.; Theodorou, D. N. Topological Analysis of
Linear Polymer Melts: A Statistical Approach. Macromolecules 2006,
39, 4592−4604.
(49) Kröger, M. Shortest Multiple Disconnected Path for the
Analysis of Entanglements in Two- and Three-Dimensional Polymeric
Systems. Comput. Phys. Commun. 2005, 168, 209−232.
(50) Shanbhag, S.; Kröger, M. Primitive Path Networks Generated
by Annealing and Geometrical Methods: Insights into Differences.
Macromolecules 2007, 40, 2897−2903.
(51) Hoy, R. S.; Foteinopoulou, K.; Kröger, M. Topological Analysis
of Polymeric Melts: Chain-Length Effects and Fast-Converging
Estimators for Entanglement Length. Phys. Rev. E 2009, 80,
No. 031803.
(52) Baig, C.; Mavrantzas, V. G.; Kröger, M. Flow Effects on Melt
Structure and Entanglement Network of Linear Polymers: Results
from a Nonequilibrium Molecular Dynamics Simulation Study of a
Polyethylene Melt in Steady Shear. Macromolecules 2010, 43, 6886−
6902.
(53) Theodorou, D. N.; Suter, U. W. Shape of Unperturbed Linear
Polymers: Polypropylene. Macromolecules 1985, 18, 1206−1214.
(54) Zhang, T.; Winey, K. I.; Riggleman, R. A. Polymer
Conformations and Dynamics under Confinement with Two Length
Scales. Macromolecules 2019, 52, 217−226.

6125 DOI: 10.1021/acs.macromol.9b01001


Macromolecules 2019, 52, 6116−6125

You might also like