Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2020 Oct 2;117(41):25407–25413. doi: 10.1073/pnas.2006703117

Dynamic phase transitions in freestanding polymer thin films

Robert J S Ivancic a, Robert A Riggleman b,1
PMCID: PMC7568329  PMID: 33008880

Significance

As a bulk supercooled liquid approaches the glass transition temperature, spatiotemporal fluctuations in dynamics increase by orders of magnitude. While this effect is fundamental to the glass transition, few have studied how confinement affects this dynamic heterogeneity. Here, we examine this effect in thin films through the lens of the dynamic phase transition from high- to low-mobility dynamic states in a set of model freestanding thin films. We find that changes in this dynamic transition are similar to deviations in thermodynamic transitions under confinement due to capillary condensation. Furthermore, these changes are reminiscent of several atypical features of the glass transition observed in ultrathin film experiments, suggesting a possible link between our simulations and experiments.

Keywords: glass, thin film, trajectory phase transition, wetting transition

Abstract

After more than two decades of study, many fundamental questions remain unanswered about the dynamics of glass-forming materials confined to thin films. Experiments and simulations indicate that free interfaces enhance dynamics over length scales larger than molecular sizes, and this effect strengthens at lower temperatures. The nature of the influence of interfaces, however, remains a point of significant debate. In this work, we explore the properties of the nonequilibrium phase transition in dynamics that occurs in trajectory space between high- and low-mobility basins in a set of model polymer freestanding films. In thick films, the film-averaged mobility transition is broader than the bulk mobility transition, while in thin films it is a variant of the bulk result shifted toward a higher bias. Plotting this transition’s local coexistence points against the distance from the films’ surface shows thick films have surface and film-center transitions, while thin films practically have a single transition throughout the film. These observations are reminiscent of thermodynamic capillary condensation of a vapor–liquid phase between parallel plates, suggesting they constitute a demonstration of such an effect in a trajectory phase transition in the dynamics of a structural glass former. Moreover, this transition bears similarities to several experiments exhibiting anomalous behavior in the glass transition upon reducing film thickness below a material-dependent onset, including the broadening of the glass transition and the homogenization of surface and bulk glass transition temperatures.


There are numerous applications of ultrathin amorphous films, including flexible integrated circuits (1), tissue engineering (2), and semiconductor manufacturing (3). However, after more than 25 years of study, the dynamics of amorphous ultrathin film materials are still not well understood. As film thickness (h) decreases in systems with weak substrate interactions, the glass transition temperature (Tg) in both polymeric and small-molecule glass films drops by 15 K to 30 K (410). This effect is more substantial in freestanding polystyrene films where decreases of up to 70 K are reported independent of molecular weight (1113). In the thick film limit (h>hc 30 nm) or when one probes time scales shorter than ∼1 s where the confinement effects are relatively modest, the changes in Tg can be understood as a consequence of a mobile layer propagating from the free surface into the film (14).

However, reducing film thickness below hc and probing longer time scales causes several new features to emerge. Measurements of the gradient of Tg (Tg) in supported polystyrene films indicate that the changes in Tg cannot be explained as a film with a bulk-like Tg being supplanted by a surface layer with a lower glass transition temperature (Tg,s=TgΔT) as h decreases (15). While Tg,sTg in thick films, Tg,s anomalously increases dramatically upon reduction of h until it joins the film-averaged Tg for h<hc. At this point, the Tg measured at the surface can no longer be distinguished from that of the entire film. In the case of freestanding polystyrene films, the suppression of Tg occurs through the film-averaged Tg decreasing with h (16). In contrast, Tg,s remains constant. For h<2hc, Tg,s and Tg join and continue to decrease with h. These significant results suggest that the film dynamics as a function of the distance from the interface homogenize for thin films.

While it is conceivable that having a distinct surface and bulk Tg that converge to a single value upon reduction of h would lead to a narrowing of the glass transition, this view does not agree with results in supported thin films with weak substrate interactions. In these cases, the reduction of Tg is accompanied by a broadening of Tg for both polymeric and molecular glass films, suggesting the material still samples a broad distribution of dynamics in ultrathin films (9, 10). This transition occurs in two stages. As h decreases in thick films, the lower Tg onset decreases independent of the constant bulk-like upper Tg onset, suggesting an increasing amount of fast modes are becoming available due to the effect of the free surfaces. For h<hc, these lower and upper onsets decrease sharply and in parallel, implying that bulk-like dynamics begin to disappear at this length scale. The dynamics of freestanding and supported polymer films have been studied directly using the dynamics of fluorescent dye molecules and demonstrate the emergence of a fast decay process independent of the single, slow decay process available for thick films (17, 18). This population of fast decaying processes may account for the increased dynamic heterogeneity in thin films.

These corresponding transitions indicate a qualitative difference in the dynamics of glassy films below hc, but an explanation for these transitions remains elusive. Molecular dynamics simulations have studied the mobility of ultrathin films extensively in the weakly supercooled regime (1922) but only observe a weak increase in the size of the mobile surface layer with decreasing temperature (23, 24). However, this may be due to the limited timescales available to molecular dynamics simulations. The fastest rate typically used in experimental cooling rate-dependent Tg measurements is ∼1 Ks1. At this cooling rate, Tg only weakly depends on h (9), while the longest timescale reached in direct simulations of glassy solids is ∼1 μs (24, 25). Moreover, while many of these molecular dynamics studies focus on the average local relaxation time on the approach to a film’s interface, we have not found a study that looks at the spatiotemporal fluctuations of the dynamics in ultrathin films. This oversight seems conspicuous considering that this is a fundamental feature of bulk supercooled liquids on approach to the glass transition (26) and given the widening distribution of relaxation rates seen in thin-film experiments.

One approach to studying dynamic heterogeneity in bulk glasses is the use of space- and time-integrated dynamic order parameters defined over entire trajectories. This approach contrasts with equilibrium thermodynamics that uses configurational order parameters. A typical dynamic order parameter may count the number of particle “hops” in a trajectory of a given number of particles and observation time. Applying a large deviations theory framework (27, 28) to such dynamic order parameters has allowed for the determination of a first-order phase transition in the dynamics of supercooled liquids from an “active,” high-mobility state to an “inactive,” low-mobility state in the ensemble of trajectories. Examples of these transitions include both model (2932) and experimental (33) materials. For a finite-sized system, this transition indicates an excess of low-mobility trajectories in the distribution of a dynamic order parameter in equilibrium compared to the expected Gaussian distribution. This excess of low-mobility trajectories is the dynamic analogy to excess low-density fluctuations as a liquid approaches a liquid–vapor transition. The so-called fat tails of the dynamic distributions are typically enhanced by using a biasing field that couples linearly with the dynamic order parameter in analogy to a reduced pressure biasing a liquid toward low-density states. Importantly, while initially observed in kinetically constrained models (34), the phenomenology behind such dynamic transitions is independent of any particular theory. Recent simulations (32, 35) indicate that this transition could be related to the ideal glass transition predicted by the random first-order transition (RFOT) theory (36, 37).

Here, we use the “s-ensemble” described above to analyze the dynamic phase transition in model low-molecular-weight polymer freestanding thin films. We find evidence that the active phase “wets” the free surface of the interface in equilibrium (s=0) at the temperatures we study. In thick films, we find a broad, rounded transition. Locally, this transition splits into two parts: one that is near bulk-like in the film’s center and one that dewets the film’s surface at higher field strengths. In thin films, the entire film transitions nearly uniformly. We interpret these observations as a dynamic version of capillary condensation analogous to thermodynamic capillary condensation in Ising magnets (38, 39) and binary polymer films (40). We also discuss what such a transition would mean in experimental cases and whether our results relate to the fundamental transition in dynamics for h<hc.

Methods

In this work, we model oligomeric bead-spring polymers each with 10 beads of diameter σ and mass m. The nonbonded interactions are taken via a Lennard-Jones potential with a well depth of ϵ that is truncated at 2.5σ, and the bonded interactions are taken via a finite extensible nonlinear elastic potential (41). We study systems of N = 1,250 monomers at two temperatures: T=0.5ϵ/kB and 0.48ϵ/kB. At each temperature, we simulate two ensembles of freestanding films, thick films with h=20σ and thin films with h=12.5σ, in the NVT ensemble with a timestep of δt=0.002τLJ where τLJ=mσ2/ϵ is the unit Lennard-Jones time scale. The temperature is maintained through an Andersen thermostat (42) using a massive stochastic collision every Δt=656δt. At T=0.5ϵ/kB, we also simulate bulk trajectories using periodic boundary conditions in all directions and modify Andersen’s algorithm so that it is run in the NPT ensemble at a pressure of P=0. We generate trajectories by harvesting configurations every Δt. Except for our finite-sized scaling results where we vary tobs, we maintain trajectories of duration tobs17.4τα,b where τα,b is the bulk relaxation time, which we define as the time at which the self part of the intermediate scattering function intersects 1/e. Precisely, we analyze trajectories in which tobs=270Δt and 576Δt for our high- and low-temperature cases, respectively. All simulations are performed with LAMMPS (43).

The dynamic order parameter of interest in this study measures the mobility of a trajectory and is defined as

Kx=Δti=1Nobs1j=1NΔrjti1,ti2, [1]

where x=rj(t)1jN,0t<tobs is a trajectory, ti=iΔt is a time within x, and Δrjt,t=rj(t)rj(t) is the distance a monomer j travels between times t and t within x. Here, rj(t) is the position of particle j at time t within x, and we take Nobs=tobs/Δt to be the number of observed frames within x. This order parameter has been shown to exhibit a dynamic phase transition in the bulk (29) and has been linked to a dynamic transition from a state with a large number of particle hops (large K[x]) to a state with a small number of particle hops (small K[x]) (31). Following refs. 29 and 44, we may now consider the probability distribution of trajectories at a fictitious field (s) that is linearly coupled to this order parameter

PsxP0xesKx [2]

in which P0x is the equilibrium probability distribution of trajectories and Psx is the probability distribution of trajectories at a given field strength s. While we could obtain this distribution at all s by generating a sufficient number of trajectories in the equilibrium distribution (s=0), this is not computationally efficient because of the rarity of the low-mobility (small K[x]) trajectories at the temperatures at which we simulate. Instead, we perform Monte Carlo in trajectory space, transition path sampling (TPS) (45), at multiple values of s. All averages are calculated using the multistate Bennett acceptance ratio (46). We implement replica-exchange Monte Carlo across windows in s (47) and use waste recycling (48, 49) to enhance our sampling. Obtaining converged statistics requires more than 105 trajectories across all windows in s for a given (h,T) combination. Complete simulation and TPS details can be found in SI Appendix.

Global Dynamic Transition

We begin our analysis by considering the average value of our mobility order parameter Kx at a given field strength (Ks=K[x]s). To better visualize this order parameter’s behavior, we plot the difference between the mean mobility at s and its equilibrium value (ΔKs=KsK0) for T=0.5ϵ/kB and 0.48ϵ/kB in Fig. 1 A and B. We include plots of K0 in SI Appendix, Fig. S1. The ΔKs plots show sigmoidal behavior in all cases. We include plots that demonstrate this behavior sharpens with increased observation time, consistent with previous investigations in the bulk in SI Appendix, Fig. S2 (29). This result indicates that the sigmoid behavior will sharpen into a discontinuous transition with a coexistence point, s=s*, in the large system limit (N,tobs). Thus, we are observing a first-order trajectory phase transition in Kx. This transition is equivalent to a fat tail in the ΔKx probability distribution in equilibrium (P0ΔK[x]) at low ΔKx compared to the expected Gaussian, as shown in Fig. 1 A and B, Insets. Larger versions of these insets are in SI Appendix, Fig. S3. These insets reveal that low-mobility trajectories become less probable under confinement, and the low-mobility tails of the distribution become more Gaussian in ultrathin films.

Fig. 1.

Fig. 1.

Evidence of changing dynamics in amorphous thin films. In all plots, the gray, blue, and red curves correspond to the bulk, thick film, and thin film results. Error bars show bootstrap error across our two configurations and are indicated by the shaded regions. (A and B) Average space–time mobility order parameter minus the equilibrium mobility (ΔKs=KsK0) as a function of field strength at two temperatures, T=0.5ϵ/kB and 0.48ϵ/kB. The sigmoidal behavior in these plots demonstrates a first-order dynamic phase transition. This transition indicates fat tails in the equilibrium probability distribution of the our dynamic order parameter (ΔK[x]=K[x]K0) compared to the expected Gaussian distribution (black dashed lines), as shown in the insets. Increased confinement reduces this deviation from Gaussian behavior, suggesting decreased spatiotemporal dynamic heterogeneity in thin films. (C and D) Susceptibility, χs=Kss, at all confinements. These plots demonstrate that confinement shifts the peak in χs, a finite-sized estimate coexistence point of the dynamic transition (s*), to higher field strengths. They also demonstrate a rounding of the thick film transition, which becomes sharper with increased confinement. (E and F) Alpha-relaxation times (τα) measured as the time at which the self-intermediate scattering function averaged at a field strength of s intersects 1/e. We plot local relaxation times (τα(z)) for z=6σ away from the interface as dashed lines for both film thicknesses. The thick and thin film lines track each other until τα(z) of each film sharply increases near its respective coexistence point. If the changes in τα(z) were described by overlapping dynamic gradients, we would expect that a sharper increase in τα(z) occurs in thick films as the thin film begins to transition to compensate for the homogenization of the thin film. Instead, τα(z) in the thick film increases less sharply. We do not include error bars in the inset or local relaxation rate plots for clarity.

Next, we analyze the susceptibility of our Kx order parameter

χs=Kss=KxKs2s, [3]

which we plot in Fig. 1 C and D. Considering the peak of this function to be the finite system estimate of the coexistence point (s*) between high- and low-mobility phases, we find that low temperatures have decreased s* and sharper peaks compared to high temperatures, as seen in ref. 29. These changes in s* indicate changes in equilibrium dynamic behavior, with lower values of s* indicating more significant deviance from Gaussian, consistent with the behavior in Fig. 1 A and B, Insets. We also find s* increases with increasing confinement at both temperatures. Again, this effect is related to the more Gaussian distribution of thin films’ dynamics demonstrating decreased low mobility fluctuations. We also observe that the thick film has a broader transition than the bulk or thin film. This broadness is consistent with thick films falling out of dynamic equilibrium in stages rather than all together.

This dynamic transition corresponds to a dramatic increase in the alpha-relaxation time (τα) defined as the time at which the self part of the intermediate scattering function evaluated at a field strength s,

Fs(t)=jsinqΔrjt,t+tNqΔrjt,t+tt,s, [4]

intersects 1/e, as shown in Fig. 1 E and F. Here we take q=6.96σ1. While this increase is evident in all instances, the growth in the films’ relaxation rates is significantly less than the apparent relaxation rate of the bulk in the low-mobility, inactive phase. Thus, although the films can reach a low mobility dynamic phase, the effect of this phase on the relaxation rates appears to be significantly lessened.

We also plot the local alpha-relaxation time (τα(z)) a distance of z=6σ from the free surface for the thick and thin films in Fig. 1E. While these curves track each other for small fields (s<0.003σ2Δt1), the two curves separate from each other at s values where the thick film begins to transition. Further analysis of τα(z) suggests that this difference is not due to linear additive effects of two h-independent mobility profiles (SI Appendix, Fig. S6). While the thin and thick films’ local τα(z) maintain the roughly double exponential form found in typical molecular dynamics simulations (1922) for s>0, the parameters of this fit do not obey the thickness-independent scaling relationship found in ref. 24. This result is robust to the definition of the local alpha-relaxation time. Additional details of this calculation are available in SI Appendix.

Local Dynamic Transition

We now explore the local dynamics of the film in greater detail by measuring our mobility order parameter as a function of the distance away from the interface

K(z)[x]=Δt2πδ2i=1Nobs1j=1NΔrjti1,ti2e(zzj(ti1))22δ2, [5]

where zj(t) is the shortest distance of monomer j to the interface at time t. We find that our results are qualitatively insensitive to the choice of δ so long as δ0.05σ. For each trajectory, we define the interface as the slab that the local film density reaches 1/2 its average interior value. We start by examining the average value of the local dynamics at a given field strength (Ks(z)=K(z)[x]s) for thick and thin films at T=0.5ϵ/kB in Fig. 2. As is typical in simulations of freestanding films, we find an increase in mobility as measured by Ks(z) near the free surface. Similar to other simulation results, we find the dynamics of the equilibrium systems (s=0) are nearly identical in the thick and thin film cases. As s increases, however, the profiles of our local mobility order parameters in thick and thin films differ. At s=0.008σ2Δt1, the h=20σ films experience a substantial decrease in the mobility in the films’ center while a smaller change occurs in the dynamics of the h=12.5σ films. As we increase field strength to s=0.011σ2Δt1, we find thick films’ mobility profile from 2σ<z<4σ joins the thin films’ mobility profile while separation still occurs in the centers of the films. When both films reach the inactive state (s=0.013σ2Δt1), mobility profiles again overlap.

Fig. 2.

Fig. 2.

Average local mobility in films as a function of the distance away from the interface at selected field strengths (Ks(z)=K(z)[x]s) for thick (solid) and thin (dashed) films. While thick and thin films have similar profiles at high (s=0.013σ2Δt1) and low (s=0.000σ2Δt1) field strengths, they separate at intermediate field strengths in the films’ centers. (Inset) A close-up of the same profiles.

To further investigate this effect, we show the local mean mobility as a function of field strength minus the equilibrium local mean mobility for a selection of distances from the interface z (ΔKs(z)=Ks(z)K0(z)) in Fig. 3 A and B for thick and thin films, respectively. At each z, our local order parameter ΔKs(z) exhibits a clear sigmoidal transition associated with a local dynamic phase transition. These plots illuminate two trends with z. First, the decrease in dynamics at the local coexistence points (s*(z)) decays on approach to the surface, indicating the active and inactive basins are moving closer together. Second, these local transitions’ coexistence points shift smoothly as a function of z in the thick film. In the thin film, these transitions occur at a similar high s* throughout the entire film except for the plane closest to the free surface.

Fig. 3.

Fig. 3.

Average local mobility in films as a function of field strength for selected distances from the interface. (A and B) Plots of the difference between the mean local mobility at s and equilibrium at selected distances away from the interface z (ΔKs(z)=Ks(z)K0(z)) for thick and thin films. These plots demonstrate that while the local coexistence point in thick films shifts to higher field strengths on approach to the interface, the thin films seem to transition at a single high coexistence point.

We can see the homogeneity in the thin film transition more clearly by plotting s*(z) in both the h=20σ and 12.5σ films in Fig. 4 A and B for T=0.5ϵ/kB and 0.48ϵ/kB, respectively. We extract this value by fitting the profiles in Fig. 3 to the sum of a linear function and a hyperbolic tangent at each z. In each case, it appears that the thick films have two distinct transitions: one on the interior of the film and one at the interface with a smooth transition between the two, and this is consistent with the broadening of the susceptibility peak for the thick films shown above in Fig. 1. For thin films, this splitting of the surface and interior transitions dramatically decreases (T=0.5ϵ/kB) or disappears entirely (T=0.48ϵ/kB), indicating that sufficiently thin films only have a single transition that has a coexistence point near that of the surface transition in the thick films. This observation demonstrates that, while the equilibrium distribution of local dynamics in thick films transitions from being very non-Gaussian at the film’s center to slightly non-Gaussian at the film’s surface, the equilibrium distribution of local dynamics in thin films is homogeneously slightly non-Gaussian throughout the film.

Fig. 4.

Fig. 4.

Evidence of two local dynamic transitions in thick films that merge. (A and B) This figure shows the local mobility order parameters’ coexistence points (s*(z)) as a function of distance from the interface (z) at T=0.5ϵ/kB and 0.48ϵ/kB. Here, we see that the thick films have two transitions, one that is bulk-like and one that is on the surface of the film. The thin films only have a single transition that is surface-like. This effect is more pronounced at lower temperatures. The dotted line in the inset of A represents our model of the thin film as a linear combination of mobility in the thick film. While this model homogenizes the film, it does so at the bulk-like rather than the surface-like coexistence point.

We now investigate whether we can describe the merging of the surface and bulk transitions with a simple model in which we assume the dynamics in a thin film as simple linear combinations of the enhanced mobility at each interface. To do this, we assume that we can write the local mobility in the h=12.5σ films as a linear combination of the local mobility in the h=20σ films from both interfaces, which are distances z and 12.5σz away in the h=12.5σ films. Here, we suppose that interfaces in h=20σ films are separated enough to leave local mobility unaffected by the interference of multiple interfaces. Thus, we model Ksh=12.5(z)=Ksh=20(z)+Ksh=20(12.5z) and find the local coexistence points of this model as the dotted line in Fig. 4A. While this model homogenizes the thin films, it does so at the film center s* rather than the surface s*, in contrast to our simulation results. This observation indicates that the merging of s*(z) is not a result of a linear combination of the enhanced dynamics at the films’ interfaces. Thus, the homogenization in thin films is not a result of overlapping mobility gradients.

Discussion

In conclusion, we have applied “s-ensemble” methods to show evidence of changes in the nonequilibrium, trajectory phase transition in the dynamics of supercooled freestanding films. Our thick films have a much more rounded transition than bulk or thin films as evidenced by a broader peak in the susceptibility. In the thinner films, the peak sharpens and requires stronger biasing (larger s*) to induce the phase transition, consistent with the strong suppression of low-mobility fluctuations in the thin films. Moreover, we have looked at the local mobility profiles in these cases and have determined that the thin films experience a single transition while thick films experience two: a film center and a surface transition. This result implies a homogenization of the dynamic transition under confinement.

These features are reminiscent of capillary condensation between parallel plates (3840). Capillary condensation occurs when a binary fluid is in contact with two symmetric walls in which one fluid species prefers the walls. Such a scenario can cause the phase boundary between the blended and single phases to shift to lower temperatures and larger exchange potentials (Δμ). The shift occurs because the local decrease in Δμ at the interface due to interactions with the walls causes the free energy to favor the surface species. For films with thicknesses much larger than the bulk correlation length (ξ), the films divide into three regions: 1) a wetted region close to the wall with width hw, 2) a transition region with width ξ, and 3) a bulk-like region. When h2(hw+ξ), the transition begins to round. For h<2hw, the transition becomes a shifted version of the bulk. The parallels between the evidence we have presented and these features point to a dynamic version of capillary condensation occurring in the freestanding films we have studied.

The establishment of such an effect in freestanding films leads to the question: Do thin films on a substrate also see dynamic versions of confined wetting transitions? Near a rough substrate, dynamics tend to slow down, suggesting that these surfaces could enhance fluctuations into the low-mobility, inactive phase. Such a scenario would lead to the dynamic analog of a delocalization–localization phase transition due to the antisymmetric wetting conditions (40, 5052). We would expect that such a transition also broadens and is pushed to lower temperatures and higher field strengths as in the capillary condensation case, and we speculate that one may also observe a merging of the near-substrate and near-surface transitions in thin films. In this view, the suppression of Tg in equilibrium (15, 16) is equivalent to the suppression of s* shown in Fig. 4 as we reduce film thickness, and hc2hw. Moreover, the broadening of Tg (9, 10) in supported films may be explained via a broadening of dynamic delocalization-localization transition.

Due to the difficulty in probing freestanding films, few experiments address them directly. However, there are several features in previous experiments that are consistent with this point of view. The most compelling evidence is the suppression of Tg in freestanding films upon reducing the total film thickness (16). Moreover, the larger decrease in Tg in freestanding polystyrene films than supported polystyrene films (1113) is consistent with larger shifts in the phase boundary for capillary condensation phenomena as compared to delocalization–localization phenomena (40). Rationalizing our current simulations and the fluorescence dye rotation measurements of ref. 17 that show the persistence of a bulk relaxation mode coexisting with a fast mode in freestanding films as thin as 14 nm remains an outstanding question. This bulk process is particularly unexpected because coarse-grained simulations of thin freestanding films, including ref. 22 and our thin films in Fig. 1E, show deviations from bulk behavior in relaxation times in the films’ centers at this level of confinement. Given that the mobile free surface layer is expected to grow upon cooling, this would suggest even more significant deviations from bulk behavior in experiments that regularly reach significantly larger timescales and higher mobility gradients. Our biased film-averaged intermediate scattering functions logarithmically decay in the caged region rather than exhibiting a two-step decay (SI Appendix, Fig. S5). Noting that the dynamic phase transition leaves the film’s surface relaxation rates nearly unaffected (SI Appendix, Fig. S6), we speculate that using a longer observation time may lead to such a separation of timescales and that the fast mode may be interpreted as a layer near the free surface, while the rest of the film exhibits bulk-like fluctuations into the low-mobility phase.

The homogenization of the transition point in thin films also provides further evidence of interfacial tension between the active and inactive phases studied herein. However, we emphasize that these results do not require a view of the glass transition through the lens of dynamic facilitation (34). Given that it is known that the inactive phase preferentially samples low-energy states (53) with locally preferred packing motifs (32), we cannot rule out the possibility that our results are a manifestation of the predicted interfacial tension between amorphous states predicted from the RFOT theory (54). Some attempts at describing the dynamics near free surfaces have relied on accounting for simple changes in the number of particles forming a cage near the free surface (55) and propagating this effect into the film (56). Our results suggest that it may be important to take into account the spatial variation in the materials’ ability to sample low-energy amorphous states and the interfacial energy between them. This accounting would manifest as a term that penalizes the formation of large gradients in mobility and not a simple additive effect of the two interfaces involved, as shown in Fig. 4A and SI Appendix, Fig. S6. As emphasized above, it will be of interest to observe how this manifests in films with asymmetric and competing boundary conditions.

Supplementary Material

Supplementary File
Supplementary File
Download video file (8.3MB, mov)
Supplementary File
Download video file (8MB, mov)

Acknowledgments

We thank Zahra Fakhraai and Sanat K. Kumar for helpful discussions regarding experimental work and wetting transitions in thin-film materials, and Connie B. Roth for helpful discussions. The NSF supported this research through the University of Pennsylvania Materials Research Science and Engineering Center (DMR-1720530), including its shared computational equipment facility. This work also benefited from computational resources provided through Extreme Science and Engineering Discovery Environment allocation DMR-150034.

Footnotes

The authors declare no competing interest.

This article is a PNAS Direct Submission.

This article contains supporting information online at https://www.pnas.org/lookup/suppl/doi:10.1073/pnas.2006703117/-/DCSupplemental.

Data Availability.

Code to run transition path sampling simulations and analysis of trajectories have been deposited in Open Science Framework at http://doi.org/10.17605/OSF.IO/TGFV3. Some study data are available.

References

  • 1.Salvatore G. A., et al. , Wafer-scale design of lightweight and transparent electronics that wraps around hairs. Nat. Commun. 5, 2982 (2014). [DOI] [PubMed] [Google Scholar]
  • 2.Liu X., Ma P. X., Polymeric scaffolds for bone tissue engineering. Ann. Biomed. Eng. 32, 477–486 (2004). [DOI] [PubMed] [Google Scholar]
  • 3.Stoykovich M. P., et al. , Directed assembly of block copolymer blends into nonregular device-oriented structures. Science 308, 1442–1446 (2005). [DOI] [PubMed] [Google Scholar]
  • 4.Keddie J. L., Jones R. A. L., Cory R. A., Size-dependent depression of the glass transition temperature in polymer films. Europhys. Lett. 27, 59–64 (1994). [Google Scholar]
  • 5.Forrest J. A., Dalnoki-Veress K., The glass transition in thin polymer films. Adv. Colloid Interface Sci. 94, 167–195 (2001). [Google Scholar]
  • 6.Roth C. B., Dutcher J. R., Glass transition and chain mobility in thin polymer films. J. Electroanal. Chem. 584, 13–22 (2005). [Google Scholar]
  • 7.Fakhraai Z., Forrest J. A., Probing slow dynamics in supported thin polymer films. Phys. Rev. Lett. 95, 025701 (2005). [DOI] [PubMed] [Google Scholar]
  • 8.Yang Z., Fujii Y., Lee F. K., Lam C.-H., Tsui O. K. C., Glass transition dynamics and surface layer mobility in unentangled polystyrene films. Science 328, 1676–1679 (2010). [DOI] [PubMed] [Google Scholar]
  • 9.Glor E. C., Fakhraai Z., Facilitation of interfacial dynamics in entangled polymer films. J. Chem. Phys. 141, 194505 (2014). [DOI] [PubMed] [Google Scholar]
  • 10.Zhang Y., et al. , Effect of substrate interactions on the glass transition and length-scale of correlated dynamics in ultra-thin molecular glass films. J. Chem. Phys. 149, 184902 (2018). [DOI] [PubMed] [Google Scholar]
  • 11.Forrest J. A., Dalnoki-Veress K., Stevens J. R., Dutcher J. R., Effect of free surfaces on the glass transition temperature of thin polymer films. Phys. Rev. Lett. 77, 2002–2005 (1996). [DOI] [PubMed] [Google Scholar]
  • 12.Mattsson J., Forrest J. A., Börjesson L., Quantifying glass transition behavior in ultrathin free-standing polymer films. Phys. Rev. E 62, 5187 (2000). [DOI] [PubMed] [Google Scholar]
  • 13.Pye J. E., Roth C. B., Two simultaneous mechanisms causing glass transition temperature reductions in high molecular weight freestanding polymer films as measured by transmission ellipsometry. Phys. Rev. Lett. 107, 235701 (2011). [DOI] [PubMed] [Google Scholar]
  • 14.Schweizer K. S., Simmons D. S., Progress towards a phenomenological picture and theoretical understanding of glassy dynamics and vitrification near interfaces and under nanoconfinement. J. Chem. Phys. 151, 240901 (2019). [DOI] [PubMed] [Google Scholar]
  • 15.Ellison C. J., Torkelson J. M., The distribution of glass-transition temperatures in nanoscopically confined glass formers. Nat. Mater. 2, 695–700 (2003). [DOI] [PubMed] [Google Scholar]
  • 16.Kim S., Torkelson J. M., Distribution of glass transition temperatures in free-standing, nanoconfined polystyrene films: A test of de Gennes’ sliding motion mechanism. Macromolecules 44, 4546–4553 (2011). [Google Scholar]
  • 17.Paeng K., Swallen S. F., Ediger M. D., Direct measurement of molecular motion in freestanding polystyrene thin films. J. Am. Chem. Soc. 133, 8444–8447 (2011). [DOI] [PubMed] [Google Scholar]
  • 18.Paeng K., Richert R., Ediger M. D., Molecular mobility in supported thin films of polystyrene, poly(methyl methacrylate), and poly(2-vinyl pyridine) probed by dye reorientation. Soft Matter 8, 819–826 (2012). [Google Scholar]
  • 19.Baschnagel J., Varnik F., Computer simulations of supercooled polymer melts in the bulk and in confined geometry. J. Phys. Condens. Matter 17, R851–R953 (2005). [Google Scholar]
  • 20.Kob W., Roldán-Vargas S., Berthier L., Non-monotonic temperature evolution of dynamic correlations in glass-forming liquids. Nat. Phys. 8, 164–167 (2012). [Google Scholar]
  • 21.Hocky G. M., Berthier L., Kob W., Reichman D. R., Crossovers in the dynamics of supercooled liquids probed by an amorphous wall. Phys. Rev. E 89, 052311 (2014). [DOI] [PubMed] [Google Scholar]
  • 22.Shavit A., Riggleman R. A., Physical aging, the local dynamics of glass-forming polymers under nanoscale confinement. J. Phys. Chem. B 118, 9096–9103 (2014). [DOI] [PubMed] [Google Scholar]
  • 23.Peter S., Meyer H., Baschnagel J., Thickness-dependent reduction of the glass-transition temperature in thin polymer films with a free surface. J. Polym. Sci. B Polym. Phys. 44, 2951–2967 (2006). [Google Scholar]
  • 24.Diaz-Vela D., Hung J.-H., Simmons D. S., Temperature-independent rescaling of the local activation barrier drives free surface nanoconfinement effects on segmental-scale translational dynamics near Tg. ACS Macro Lett. 7, 1295–1301 (2018). [DOI] [PubMed] [Google Scholar]
  • 25.Hung J.-H., Patra T. K., Meenakshisundaram V., Mangalara J. H., Simmons D. S., Universal localization transition accompanying glass formation: Insights from efficient molecular dynamics simulations of diverse supercooled liquids. Soft Matter 15, 1223–1242 (2019). [DOI] [PubMed] [Google Scholar]
  • 26.Ediger M. D., Harrowell P., Perspective: Supercooled liquids and glasses. J. Chem. Phys. 137, 080901 (2012). [DOI] [PubMed] [Google Scholar]
  • 27.Touchette H., The large deviation approach to statistical mechanics. Phys. Rep. 478, 1–69 (2009). [Google Scholar]
  • 28.Elmatad Y. S., Jack R. L., Chandler D., Garrahan J. P., Finite-temperature critical point of a glass transition. Proc. Natl. Acad. Sci. U.S.A. 107, 12793–12798 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Hedges L. O., Jack R. L., Garrahan J. P., Chandler D., Dynamic order-disorder in atomistic models of structural glass formers. Science 323, 1309–1313 (2009). [DOI] [PubMed] [Google Scholar]
  • 30.Speck T., Malins A., Royall C. P., First-order phase transition in a model glass former: Coupling of local structure and dynamics. Phys. Rev. Lett. 109, 195703 (2012). [DOI] [PubMed] [Google Scholar]
  • 31.Speck T., Chandler D., Constrained dynamics of localized excitations causes a non-equilibrium phase transition in an atomistic model of glass formers. J. Chem. Phys. 136, 184509 (2012). [DOI] [PubMed] [Google Scholar]
  • 32.Turci F., Royall C. P., Speck T., Nonequilibrium phase transition in an atomistic glassformer: The connection to thermodynamics. Phys. Rev. X 7, 031028 (2017). [Google Scholar]
  • 33.Pinchaipat R., et al. , Experimental evidence for a structural-dynamical transition in trajectory space. Phys. Rev. Lett. 119, 028004 (2017). [DOI] [PubMed] [Google Scholar]
  • 34.Merolle M., Garrahan J. P., Chandler D., Space–time thermodynamics of the glass transition. Proc. Natl. Acad. Sci. U.S.A 102, 10837–10840 (2005). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Berthier L., Jack R. L., Evidence for a disordered critical point in a glass-forming liquid. Phys. Rev. Lett. 114, 205701 (2015). [DOI] [PubMed] [Google Scholar]
  • 36.Biroli G., Bouchaud J.-P., “The random first-order transition theory of glasses: A critical assessment” in Structural Glasses and Supercooled Liquids: Theory, Experiment, and Applications, Wolynes P. G., Lubchenko V., Eds. (Wiley, 2012), pp. 31–113. [Google Scholar]
  • 37.Parisi G., Urbani P., Zamponi F., Theory of Simple Glasses: Exact Solutions in Infinite Dimensions (Cambridge University Press, 2020). [Google Scholar]
  • 38.Fisher M. E., Nakanishi H., Scaling theory for the criticality of fluids between plates. J. Chem. Phys. 75, 5857–5863 (1981). [Google Scholar]
  • 39.Nakanishi H., Fisher M. E., Critical point shifts in films. J. Chem. Phys. 78, 3279–3293 (1983). [Google Scholar]
  • 40.Müller M., Interplay between wetting and miscibility in thin binary polymer films. Comput. Phys. Commun. 147, 292–297 (2002). [Google Scholar]
  • 41.Kremer K., Grest G. S., Dynamics of entangled linear polymer melts: A molecular? dynamics simulation. J. Chem. Phys. 92, 5057–5086 (1990). [Google Scholar]
  • 42.Allen M. P., Tildesley D. J., Computer Simulation of Liquids (Oxford University Press, 2017). [Google Scholar]
  • 43.Plimpton S., Fast parallel algorithms for short-range molecular dynamics. J. Comput. Phys. 117, 1–19 (1995). [Google Scholar]
  • 44.Lecomte V., Appert-Rolland C., van Wijland F., Thermodynamic formalism for systems with Markov dynamics. J. Stat. Phys. 127, 51–106 (2007). [Google Scholar]
  • 45.Dellago C., Bolhuis P. G., Chandler D., Efficient transition path sampling: Application to Lennard-Jones cluster rearrangements. J. Chem. Phys. 108, 9236–9245 (1998). [Google Scholar]
  • 46.Shirts M. R., Chodera J. D., Statistically optimal analysis of samples from multiple equilibrium states. J. Chem. Phys. 129, 124105 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Calvo F., All-exchanges parallel tempering. J. Chem. Phys. 123, 124106 (2005). [DOI] [PubMed] [Google Scholar]
  • 48.Frenkel D., Speed-up of Monte Carlo simulations by sampling of rejected states. Proc. Natl. Acad. Sci. U.S.A. 101, 17571–17575 (2004). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Athènes M., Marinica M.-C., Jourdan T., Estimating time-correlation functions by sampling and unbiasing dynamically activated events. J. Chem. Phys. 137, 194107 (2012). [DOI] [PubMed] [Google Scholar]
  • 50.Binder K., Landau D. P., Ferrenberg A. M., Thin ising films with competing walls: A Monte Carlo study. Phys. Rev. E 51, 2823–2838 (1995). [DOI] [PubMed] [Google Scholar]
  • 51.Binder K., Evans R., Landau D. P., Ferrenberg A. M., Interface localization transition in Ising films with competing walls: Ginzburg criterion and crossover scaling. Phys. Rev. E 53, 5023–5034 (1996). [DOI] [PubMed] [Google Scholar]
  • 52.Binder K., Modeling of wetting in restricted geometries. Annu. Rev. Mater. Res. 38, 123–142 (2008). [Google Scholar]
  • 53.Jack R. L., Hedges L. O., Garrahan J. P., Chandler D., Preparation and relaxation of very stable glassy states of a simulated liquid. Phys. Rev. Lett. 107, 275702 (2011). [DOI] [PubMed] [Google Scholar]
  • 54.Kirkpatrick T. R., Thirumalai D., Wolynes P. G., Scaling concepts for the dynamics of viscous liquids near an ideal glassy state. Phys. Rev. A Gen. Phys. 40, 1045–1054 (1989). [DOI] [PubMed] [Google Scholar]
  • 55.Stevenson J. D., Wolynes P. G., On the surface of glasses. J. Chem. Phys. 129, 234514 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Phan A. D., Schweizer K. S., Theory of the spatial transfer of interface-nucleated changes of dynamical constraints and its consequences in glass-forming films. J. Chem. Phys. 150, 044508 (2019). [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary File
Supplementary File
Download video file (8.3MB, mov)
Supplementary File
Download video file (8MB, mov)

Data Availability Statement

Code to run transition path sampling simulations and analysis of trajectories have been deposited in Open Science Framework at http://doi.org/10.17605/OSF.IO/TGFV3. Some study data are available.


Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES