3.1.1. Configurational Microphase Separation and General Flow Behavior
The mean fractional extension of chains,
, is defined as the ensemble-averaged end-to-end distance normalized by the theoretical maximum end-to-end distance
(
Å for
).
Figure 1a displays the average fractional extension as a function of the Hencky strain,
, for startup of uniaxial elongational flow (solid lines) at various values of
. Flows of
were not simulated (except for the quiescent case of
) because of the excessive computational resources required for such large simulation cells in UEF at very low
and because the results of the prior PEF simulations [
7,
8,
16,
17] revealed that only the expected behavior arising from reptation theory manifested at such small strain rates. For comparison, the values obtained from prior PEF simulations are also presented in
Figure 1a using dotted lines of the same color for similar values of
. Note that PEF simulations at
and
were not performed in the prior work due to limitations of computational resources [
7,
8,
16,
17]; hence these two
values for PEF are not included in
Figure 1 and other figures throughout the paper.
The mean fractional extension increases very slowly at all
during the first Hencky strain of deformation. At higher strains,
begins to grow at a faster rate with a slope that increases with
. The transient response of
afterward can be classified into three categories of extension rates, namely,
,
, and
. Simulations of UEF at
were not feasible from a computational standpoint, but results are expected to be very similar to those of PEF for
[
7,
8,
17,
18]. In the first region,
, the fractional extension undergoes an overshoot roughly at
–5.5, followed by a gradual decay to an eventual quasisteady-state plateau. However, the mean fractional extension at
fluctuates such that an indisputable steady state cannot be achieved even after
units of deformation. This leads to some ambiguity in determining the steady-state fractional extension. Indeed, the steady-state fractional extension in this region could change significantly depending on the cutoff Hencky strain for calculating the ensemble averages. As discussed below (with reference to Panel (e) of
Figure 1), this is caused by individual macromolecules gradually transitioning from relatively coiled to more stretched configurations, and vice versa. Consequently, steady-state values throughout this article were calculated as ensemble averages of the respective data at
. The lower limit Hencky strain of 8 was chosen because this value is roughly the highest deformation that can be achieved using common rheological measurements of polymeric liquids for reporting steady-state properties, such as extensional viscosity. The fractional extension exhibited in the UEF simulations at these
is qualitatively similar to that of PEF, although quantitatively there is a greater degree of extension in the UEF simulations due to the higher degree of stretching in UEF imparted by the compression in both directions perpendicular to the direction of flow.
In the second region, , the overshoot in mean fractional extension vanishes, and smoothly approaches its steady-state value at . Unlike the lower range, the fluctuation in at steady state is insignificant in this range of extension rates, indicating that individual macromolecules mostly remain in either relatively stretched or coiled configurations. Furthermore, the steady-state fractional extension monotonically increases as increases.
The behavior of is fairly similar at higher extension rates in the third region, i.e., , except that the steady-state exhibits a jump compared to the second region, resembling a discontinuity in steady-state fractional extension, and remains nearly independent of . This discontinuity is associated with flow-induced crystallization, which will be discussed later. The fractional extension under PEF is qualitatively and quantitatively similar to that of UEF within this range of . Furthermore, at , which is at the lower limit of the FIC range, the initial increase of with is commensurate with that of lower values of . However, at , it experiences an increase in slope before settling at its steady-state value at . Hence the initial response at is that of a stretched liquid state, which begins to evidence FIC only once the molecules are sufficiently extended at .
The steady-state probability distribution functions (PDFs) for the molecular radius of gyration,
, and fractional extension,
, are plotted in
Figure 1b–d for both UEF and PEF. Under quiescent conditions, the PDFs depict essentially coiled macromolecules distributed almost symmetrically about the peak values (i.e., possessing Gaussian-like characteristics), with means of 57.9 Å and 141.8 Å, respectively, as indicated in
Table 1. The three characteristic extension rate regions described above can also be recognized in these PDF plots. It is evident from these distributions that the flow response in the first regime,
, is quite different and somewhat unexpected compared to the other two flow strength regimes. Specifically, the PDFs are broad and bimodal, with one peak positioned at values approximately corresponding to the radius of gyration and end-to-end distance of the molecules under quiescent conditions, mostly assuming coiled configurations, and a second peak appearing at relatively high values of
and
x, corresponding to more highly stretched configurations. Note that the coiled-peak position is somewhat independent of flow strength, whereas the stretched-peak moves to the right (higher chain extension) as
increases. This behavior was attributed to a coil-stretch transition in entangled melts, as reported in the past for PEF simulations of
and
PE melts using NEMD [
7] and DPD [
10]. In this range of flow strength, the chain-like macromolecules can exist in either of the two stable configurational states (i.e., coiled and stretched). Moreover, individual chains can occasionally transition from one configurational state to the other. Such transitions may be inferred from the small but nonzero population of the chains between the two peaks of the distributions—see the orange curves (
) of
Figure 1b,c, for instance.
Figure 1e depicts the transition of a few selected molecules between the coiled and stretched states at
by plotting their individual chain fractional extension as a function of Hencky strain. These plots indicate that the transitioning of molecules between the two states is a random and gradual process, taking about 10 Hencky stain units of deformation (or dimensionless time). Such random exchanges between the coiled and stretched populations give rise to significant fluctuations in the mean fractional extension of the chains, as discussed earlier—see the blue, orange, and green curves of
Figure 1a corresponding to
, respectively. These fluctuations of individual chain extension could manifest as fluctuations (or even anomalies) in other properties of the system, such as stress and extensional viscosity, which will be discussed later. Both UEF and PEF simulations exhibit similar behavior, although the stretched-peaks of UEF are narrower and shifted to higher extensions compared to those of PEF.
In the second region,
, the distributions
and
become unimodal and sharply peaked, shifting to the right as
is increased, implying a highly stretched liquid state at higher extension rates. The PDFs remain unimodal at even higher extension rates, i.e.,
; however, the peak positions shift discontinuously to higher extension at
and asymptotically approach that of fully stretched molecules (
). The observed discontinuity resembles a first-order phase transition, and it can be associated with flow-induced crystallization, as briefly mentioned earlier and discussed in
Section 3.1.2 in greater detail. Note that the UEF and PEF distributions at
are virtually identical to each other.
A rough sketch of a configurational phase diagram can be constructed by plotting the peak positions of the steady-state
at various
, as shown in
Figure 1f. Blue symbols denote UEF data, whereas orange symbols specify PEF data. Furthermore, squares represent coiled configurational states and triangles denote stretched states. These data points were obtained from the positions of the relevant peaks of the
distributions, as displayed in
Figure 1c,d. At low
(PEF data only), the melt is composed of coiled molecules that are barely discernible from those of the quiescent state. At
, the liquid is comprised of molecules that are relatively coiled and others that are relatively extended. At
, the steady-state liquid is comprised solely of relatively extended molecules, whereas at
one observes a predominantly monoclinic flow-induced semicrystalline state.
Individual molecular trajectories of fractional extension are plotted vs. Hencky strain in
Figure 2 for startup of UEF at various
. At
, the distribution of individual molecular configurations is quite wide, as indicated by the broad PDFs of
Figure 1a. There are molecules that rapidly become relatively highly extended upon application of flow, and others that remain relatively coiled even after 20 Hencky strain units. Other molecules gradually transition from one configurational state to the other over multiple
units. (This is hard to visualize from
Figure 2 because of the large number of molecules in the UEF simulations; see
Figure 1e for an easier visualization of some selected chains). At
, all macromolecules eventually assume relatively stretched configurations, although some molecules rapidly increase in extension after inception of flow, whereas others slowly transition from a relatively coiled configuration over a much longer period of time. At
, many molecules rapidly rise to a pseudosteady state at
, corresponding to a stretched liquid state, and then quickly transition to an even more extended state at
, which is a manifestation of FIC. Even so, there remain some relatively coiled molecules that only begin to transition to the stretched state after
.
Snapshots of the simulation cell of the PE melt undergoing startup of UEF at
are displayed in
Figure 3 at various values of Hencky strain. Cooler colors are assigned to molecules that are relatively coiled whereas warmer colors denote relatively stretched molecules. Under quiescent conditions (
), the molecules are exclusively coiled. With the onset of flow, certain individual molecules begin to elongate in the direction of flow, whereas others remain relatively coiled, retaining configurations that are essentially similar to those they possessed under quiescent conditions. As time elapses, the stretched molecules approach steady-state configurations and the coiled molecules mostly remain unaffected by the flow; however, after a substantial amount of deformation has occurred, a bicontinuous phase structure is developed wherein the stretched chains form into a mesh-like network of highly elongated molecules that enwrap ellipsoidal domains of relatively coiled chains. The bicontinuous inhomogeneous phase appears to be formed by a migration of stretched and coiled chains into domains of similarly configured molecules rather than a spontaneous heterogeneous development at certain locations within the simulation cell. This statement is supported by the lower right snapshot, which is the same snapshot at
but with the assignment of individual molecular colors according to the final steady-state configuration rather than the corresponding one at
; this snapshot provides evidence that the chains forming specific domains at steady-state are more or less randomly distributed throughout the simulation cell at
.
The spatial inhomogeneity of the biphasic structure exhibited under UEF is distinctly different than that exhibited for the same PE melt undergoing PEF, as presented in preceding publications [
8,
10,
17]. In PEF, a biphasic system is also formed over a similar range of
, but there the heterogeneous liquid is comprised of sheet-like regions of highly stretched molecules that interweave between spheroidal domains of relatively coiled molecules. This difference seems intuitively to be a manifestation of the fact that PEF maintains a neutral direction, whereas UEF implies compression in both directions perpendicular to that of flow. It is not clear, in either PEF or UEF, if the coiled chains are migrating to regions near the flow stagnation point, or vice versa; at first glance, this does not appear to be the case, but the nature of applying the KRBC or DHBC complicates the analysis considerably.
Snapshots of the PE melt undergoing steady-state UEF at various
are displayed in
Figure 4. Individual chain molecules are again colored according to the relative end-to-end extension, with tightly coiled chains displayed in a cool blue color and highly extended chains in a warm red color. The bicontinuous liquid microstructure is observed throughout the range
, where a 3-dimensional mesh-work of elongated chains essentially envelope spatial domains of concentrated coiled molecules. At lower
, the numbers of molecules comprising the two phases are similar (see
Figure 2), but as
increases, the fraction of coiled chains diminishes relative to the fraction of stretched chains. However, even at
, vestigial traces can be discerned of a spatially inhomogeneous bicontinuous liquid phase (see the bottom left snapshot of
Figure 4), although the difference in extension of the coiled and stretched phase molecules is rather slight and the
distribution of
Figure 1 remains practically unimodal. At higher
, the inhomogeneous microstructural has essentially dissolved into a homogeneous liquid phase composed of highly stretched molecules.
As discussed briefly above,
Figure 1f displays the average steady-state fractional extension for different states of the liquid as a function of
for the UEF and PEF simulations. Hence, in the range
, where the distributions are bimodal, two values are plotted corresponding to the coiled and stretched states. For other flow strengths (
), where
is unimodal, the peak position is virtually the same as the mean fractional extension of the highly stretched liquid. Note that PEF and UEF data are practically the same, indicating that this biphasic microstructure is possibly a universal phenomenon of elongational flows. Due to the configurational bistability of the system at intermediate extension rates,
Figure 1f effectively exhibits a hysteresis loop in the fractional extension profile of the liquid at steady state. It should be noted that exhibiting a hysteresis means that either purely coiled or stretched states within the bistable region should be achievable by some appropriate manipulation of the flow kinematics. This hysteresis phenomenon has not been directly studied herein; i.e., the flow conditions under which a coiled-only or stretched-only configurational state could appear in the range
. However, it was demonstrated in previous work [
7,
10] on PEF of
that a metastable stretched-only state could be attained by decreasing the flow strength of a system composed of stretched molecules at steady state at a higher value of
(e.g.,
) to a lower
located within the bistable region (e.g.,
). In other words, the purely stretched state of the bistable region could be generated by moving from the right branch (high
) of the sigmoidal curve in
Figure 1f toward the left (low
). Nevertheless, it was also demonstrated that such purely stretched states were not globally stable and eventually transitioned into biphasic states similar to those depicted in
Figure 3 and
Figure 4 for sufficiently long simulation times [
7,
10].
As noted in prior simulations of the same PE melt undergoing steady-state PEF [
8], the extensional stress and temperature of the bicontinuous phase remain homogeneous throughout the simulation cell—see Supplementary Materials of Ref. [
8]. This implies that the average stress state is not affected by spatial location within the simulation cell, indicating a homogeneous stress state in both space and time. Consequently, the segregation of the macromolecules observed does not seem to be driven by stresses developed within the system under UEF or PEF. This observation points to a microphase separation that is thermodynamically induced by architecturally identical chain molecules whose sole point of difference lies in their adopted individual steady-state configurations which are developed under flow.
Topological properties of the
melt are displayed in
Figure 5 for both UEF and PEF at various
as calculated using the Z1 code [
69,
70].
Figure 5a displays the ensemble-averaged number of kinks per chain,
, plotted as a function of
under both types of flow. At low
(<0.3; PEF only),
remains close to its quiescent value of 24.3, roughly twice the number of entanglements. At intermediate
within the biphasic region (
), however,
drops precipitously to a low value less than 2.5, which subsequently remains relatively close to zero at all higher
.
Figure 5b displays the steady-state PDFs of the number of entanglement kinks,
, of individual chains at various
for the
melt undergoing UEF. Within the biphasic region, these PDFs are decidedly bimodal, with peaks located at both small and large values of
. The high-
peak is associated with the coiled liquid phase, whereas the low-
peak is associated with the stretched phase. Indeed, both phases tend to maintain the same level of entanglements that would be expected from either a homogeneous coiled or stretched liquid phase hypothetically existing under the same flow conditions. At higher
where only the homogeneous stretched liquid state exists (
) as well as the FIC phase (
), the
are unimodal, possessing tall, narrow peaks at very small values of
, indicating almost completely unentangled systems. (Note that the PDFs of the FIC phases all overlap each other in the plot).
The positions of the peaks,
, of the PDFs displayed in
Figure 5b (as well as those of the corresponding PDFs under PEF; not shown) are plotted versus
in
Figure 5c. This entanglement phase diagram makes clear the biphasic nature of the intermediate
(0.3–1.0) regime with its characteristic inverse-S shape, wherein higher values of
are associated with the coiled phase and lower values with the stretched phase. It also implies that metastable entanglement phases of either homogeneous coiled or stretched states could be induced by gradually increasing or decreasing
from a monophasic
state point, leading to a hysteresis effect; such stretched metastable states were previously observed for the same
melt under PEF at
[
17].
The tube stretch is plotted versus
for the
melt under steady-state UEF in
Figure 5d. At low
, the tube stretch increases only slightly from its quiescent value of unity, and this trend continues for
as high as 0.6, which is well within the nonlinear viscoelastic regime. At
, which corresponds to the applied strain rate,
, equalling the reciprocal of the Rouse time,
, the tube stretch parameter begins to climb substantially with
, indicating strong stretching of the polymer entanglement strands. This increase in chain extension is commensurate with the decrease in chain entanglements discussed above. The rate of increase of
with
eventually plateaus for
, where FIC has evidently occurred. Note that the plateau value of ≈2.4 observed in panel (d) is close to the theoretical maximum value of 2.77 [
64].
The elongational viscosity upon startup (
) of UEF at various
is displayed in
Figure 6a as a function of applied Hencky strain. At low values of
, a rapid initial increase in
eventually leads to a maximum that gradually fades to a steady-state plateau at large
. At all larger
, the overshoot is not apparent, with the initial increase eventually plateauing at its steady-state value. The sole exception is the case of
, which is at the lower limit of the FIC range of
. In this case, there is an initial plateau that appears in
at
, and then a rapid increase to a larger value of
at
. This behavior is consistent with that discussed above with reference to the molecular mean fractional extension evident upon startup of UEF at
in
Figure 1a (as well as the bottom right panel of
Figure 2) and is associated with the initial formation of a stretched liquid phase upon startup of flow, which eventually transitions into a semicrystalline state after the molecules have become sufficiently stretched and aligned along the flow direction.
Steady-state values of the elongational viscosity are displayed in
Figure 6b for both UEF and PEF. These quantities are defined as
where
is a diagonal component of the Irving-Kirkwood stress tensor [
72] with the 1-component denoting the flow direction. At low
and within the biphasic region, the PEF data are subject to large statistical error because of the small size of these simulations; however, the much larger UEF simulations have much greater statistical certainty. At low
(<0.1) (where only PEF data are available), the viscosity is consistent with a linear viscoelastic behavior exhibiting a constant viscosity, within statistical limitations. At intermediate
(0.3–1.5) within the biphasic region, however, the viscosity appears to decrease to a plateau value, and then decreases again at higher
(>3.0). The steady-state viscosity data of the PE melt under UEF and PEF evidently overlap, within statistical uncertainty, over the entire range of simulated
. This observation is quite remarkable physically, and it also speaks to the consistency of the simulations since the PEF simulations were drastically smaller in size than the UEF simulations, and given the fact that the PEF simulations at
were performed in the
ensemble, whereas all UEF simulations were performed in the
ensemble. Furthermore, the PEF simulations were performed with KRBC, whereas the UEF simulations implemented the DHBC.
The steady-state elongational stress,
for PEF and
for UEF (where the index 1 is associated with the flow direction), is plotted versus
in
Figure 7 for both UEF and PEF. Again, the results of both types of flow simulations appear to overlap to within statistical error. As expected, the stress increases monotonically at low
where only a coiled liquid phase exists (C), but within the biphasic region (C & S) where coiled and stretched liquid states coexist, the trend is not definitive. At higher
(> 1.5), however, the trend is once again monotonically increasing (S), even throughout the semicrystalline FIC regime (M).
3.1.2. Flow-Induced Crystallization at High Deborah Number
Typical semicrystalline polyethylenes have normal melting points that vary greatly from low density to high density materials (0.88–0.96
), ranging roughly from 385–420 K for common varieties. The heat of fusion,
, is generally accepted at roughly 290 J/g [
73,
74,
75], and many studies indicate that single crystals formed from a quiescent melt assume orthorhombic or hexagonal lattices [
46,
76], whereas crystals formed under application of a mechanical tension tend to arrange into a monoclinic structure [
77,
78,
79,
80]. Indeed, a wide array of interesting semicrystalline morphologies can be generated under the application of a flow field, including the “shish-kebab” fibrillar structure of transverse lamellae emanating from a central axial fibril [
81,
82,
83], at temperatures several decades above the quiescent melting point [
24,
84].
Toda [
85] observed a melting point (
) of 398–403 K for single crystals of a linear PE of molar mass 13,000 g/mol grown under quiescent conditions. This is very close to the molar mass of the linear
melt (14,002 g/mol) simulated in this study. To date, it has been unfeasible computationally to simulate quiescent crystallization of entangled polyethylenes to determine a precise value of
from molecular dynamics simulations, but the accrued evidence suggests a reasonable range would be 385–400 K [
28,
32,
33,
59,
86]. Therefore, the simulated melt temperature employed in the present simulations is likely greater than 50 K than the value of
determined via experiment or equilibrium MD simulation.
The observed discontinuities in fractional extension profiles at high
in
Figure 1a, as well as the two-stage behavior at
evidenced in the same figure, and also
Figure 2 (bottom right panel) and
Figure 6a at
, serve to indicate the possibility of a phase transition occurring within the PE liquid after an initial rapid formation of a highly stretched homogeneous liquid state. In fact, prior work has demonstrated that
undergoes flow-induced crystallization when subjected to strong PEF at
at a temperature of 450 K [
17,
18]. In this subsection, the FIC phase produced under UEF is examined in detail and compared with the same FIC phase developed under PEF at comparable
.
A variety of mathematical variables have been used to distinguish between liquid and crystalline phases in molecular simulations, as well as to quantify the degree of order induced in the FIC phase. These quantities aim to estimate the global or local ordering of polymer molecules or constituent segments within the simulation cell. One of the most common ones is the order parameter defined as
, where
is the eigenvalue of the segmental order parameter tensor,
, with the largest absolute value. The segmental order parameter tensor is defined herein as
where
is a unit vector connecting adjacent even-numbered and odd-numbered carbon atoms along the chain backbone (which are separated by a single monomeric unit). The brackets
denote an ensemble average over all unit vectors, and
is the unit tensor. This definition and similar order parameters, which describe an ensemble average of the local degree of ordering within the sample, are somewhat arbitrary and are usually made based on
a priori assumptions about the state of the polymeric material obtained from the simulation at specific state points. For example, Nafar Sefiddashti et al. [
16,
17,
18] used the ranges
to define a semicrystalline state,
for an ordered pretransitional liquid state,
for a stretched liquid, and
for a coiled liquid; however, these points were defined rather arbitrarily from a preconstructed nonequilibrium phase diagram [
17].
The transient startup behavior of the order parameter is displayed in
Figure 8a as a function of Hencky strain for both UEF and PEF at various
. These data are qualitatively similar to their counterparts in
Figure 1a for the mean fractional extension. Specifically, the same three flow regions described in
Section 3.1.1 are noticeable here as well. Characteristic overshoots and pronounced long-lived fluctuations are evident within the biphasic range
, followed by smoothly and monotonically ascending
q within the range
, with
q values indicating a stretched liquid state. At
, a discontinuity is evident as the steady-state values of the curves jump significantly from those at lower
, similarly to the plots of
in
Figure 1a. All
display steady-state values greater than the critical value of
that indicates a transition from a stretched liquid state to a semicrystalline one.
The startup behavior of
q at
is a special case that requires additional deliberation. Indeed,
is the critical flow strength at the lower limit of the transition from the stretched liquid phase to the semicrystalline phase. This transition is tracked by observing the evolution of the order parameter, which initially increases with
in a similar fashion as those of
and
. Specifically,
q exhibits a stretched liquid-like response by approaching a steady-state value of about
, well below the critical order parameter of 0.75, up to a Hencky strain of
. At this point, the
q curve passes through an inflection point and rises at a high rate for nearly two Hencky strains until it attains a new steady-state of roughly
. Note that for PEF, a similar qualitative behavior with an inflection point at
was observed at
, which was the lower limit of the FIC regime under PEF noted in Ref. [
18]. This provides an indication that UEF is a stronger flow for orienting and extending the chain-like molecules than PEF due to the simultaneous compression along the two perpendicular axes with respect to the flow direction under UEF.
The steady-state PDFs of the order parameter (see
Figure 8b) of the PE melt under both UEF and PEF are bimodal in the first
(< 1.5) region, with a low-
q peak at
corresponding to essentially random coils and a high-
q peak at
–0.4, corresponding to moderately aligned and stretched chains. The high-
q peak position shifts to the right (i.e., higher
q values) as
increases. Within the range
, the distributions become unimodal and approximately Gaussian, varying roughly within the range
, which is consistent with a purely stretched liquid state. At higher
, the PDFs significantly shift to the right and become narrower at
q values greater than 0.75, demonstrating a highly ordered semicrystalline phase. Although differences between UEF and PEF in the PDFs of
are difficult to discern in
Figure 1c, the PDFs of
q are more clearly evident, where it is obvious that UEF is a stronger flow, producing the same effects as PEF but at slightly lower
. For instance, the biphasic region for UEF occurs within the range
, whereas for PEF it occurs at
. Furthermore, the FIC region begins at
under UEF but at
under PEF [
18], although once inside the FIC regime, the PDFs of UEF and PEF practically coincide.
The peak positions of the
are plotted as functions of
in
Figure 8c. Note that outside the biphasic region (C & S), the peak positions and the ensemble averages are virtually identical. This figure is very similar to that of fractional extension displayed in
Figure 1f, except that the distinction between the liquid and semicrystalline phases is significantly more pronounced here. This plot effectively serves as a nonequilibrium phase diagram of the PE melt under elongational flow, and a hysteresis loop for the
liquid in the biphasic region is evident at the constant temperature of 450 K, rendering this
region as a possibly universal feature of extensional flow.
A plot of
q versus time after cessation of steady-state UEF at
and 450 K is presented as
Figure 9. In these virtual experiments, the temperature can be quenched at the moment that the flow is turned off. If temperature is held fixed at 450 K,
q rapidly relaxes back to its quiescent value associated with a liquid composed of coiled macromolecules, indicating melting of the semicrystalline phase that was formed under flow. Interestingly, if the temperature is simultaneously quenched to 425 K or 435 K at the moment of flow cessation,
q remains essentially fixed at the same steady-state value (≈0.83), indicating that the semicrystalline phase remains indefinitely. This provides evidence that the semicrystalline phase formed under UEF has a melting point somewhere between 420–450 K, which is somewhat higher than the case for PEF of the same
liquid that melted between 415 K and 420 K in prior work [
17]. Hence the FIC semicrystalline phase formed under UEF appears to be more stable than a similar FIC phase formed under PEF. The effective melting point is at least 35 K above the melting point of the quiescently grown crystal of 385–400 K, which implies a greater thermal stability of the FIC phase. This stability was examined by Baig and Edwards from the perspective of configurational temperature [
28,
29,
87] and found that the FIC phase corresponded to a configurational state of the material, as imposed by the flow, that could be as much as several decades below the kinetic temperature. In other words, the configurational state adopted by the system in response to the flow implied a molecular ordering of the chains that was consistent with a much lower effective temperature, possibly even lower then the nominal melting point.
The enthalpy change of the
melt under an applied UEF at several
is presented in
Figure 10, where the enthalpy was calculated as
and
is the ensemble-averaged sum of the potential and kinetic energies. At
(and similarly for lower values), the enthalpy gradually decreases upon startup of flow, but this descent accelerates after one Hencky strain unit of deformation and eventually settles into a lower enthalpy state associated with the stretched liquid phase. This decrease in
is due to the changing configuration of the chains from random coils within the quiescent melt to highly elongated and aligned molecules in the stretched liquid phase, which represents a more favorable energy state once entropic forces have been overcome by the applied flow. At higher
within the FIC regime, the rapid descent of enthalpy is much larger in magnitude before eventually levelling off to a steady-state value. Note that at
, the decrease in enthalpy follows a two-step process in which there is an initial decrease to a stretched liquid state (
), followed by a further decrease to the FIC phase at
. Evidently, a sufficient degree of ordering and stretching of the macromolecules is necessary before the FIC phase can begin to develop. At
higher than 9.0, the ordering and alignment is sufficiently fast that the first step of the process blends in with the second. This is related to the formation and propagation of flow-enhanced nucleation events, which will be discussed later.
The heat of fusion of the semicrystalline phase grown under elongational flow can be estimated from the steady-state enthalpy at various
and compared to the quiescent value (≈707.7 J/g) according to
. Several steady-state values of enthalpy in the FIC region are computed as
J/g,
J/g, and
J/g, which provide
values of
,
, and
J/g, respectively. A similar value of
J/g was calculated at
in the PEF simulations of Nafar Sefiddashti et al. [
17]; note that this value was erroneously reported by the authors in Ref. [
17] as
J/g. These values effectively quantify the amount of heat released upon stretching and packing the molecules into the semicrystalline phase. At
, the enthalpy change associated with the stretched liquid state comprises about 50% of the total steady-state
. This indicates that a substantial amount of the imposed elongational force is required to deform and orient the macromolecules, which subsequently can more readily pack into the monoclinic lattices of the semicrystalline phase (with a commensurate density reduction), as discussed below. The reported values are of similar magnitude and range to experimental heats of fusion for typical (linear) PEs, which range from
to
J/g for PEs with degrees of crystallinity covering a wide range of values (roughly, 60–90%) [
73,
74,
75,
88].
Specific heat capacity at constant pressure,
, is plotted versus Hencky strain in
Figure 11 under startup of UEF (panel a) and PEF (panel b) at several values of
in the FIC regime. The quantity was calculated from the energy dispersion relationship used in
equilibrium molecular dynamics [
89],
although there is no explicit guarantee that this expression remains valid away from equilibrium. As evident from the figure, this quantity fluctuated markedly, although the
y-axes in the panels are magnified to accentuate the differences between the various curves, which serves to exaggerate the actual magnitude of the fluctuation. At low
, all curves emanate from the quiescent value of about 1.305 J/(g K). Note that for
less than 9.0, where the PE melt remains in a liquid state,
retains its quiescent value even under nonequilibrium conditions, as illustrated in panel (a) for UEF at
. However, at
inside the FIC regime,
decreases slightly upon startup of flow in both UEF and PEF, ultimately approaching steady-state plateaus with values that decrease with increasing
. Nevertheless, from the quiescent state to the steady-state value at
(about 1.28 J/(g K)), the change in heat capacity is minimal (about 2%). Typical experimental values of specific heat capacities for polyethylenes generally range from 1.3–2.7 J/(g K) [
88,
90,
91,
92] at temperatures up to about 450 K. Hence the values computed from the NEMD simulations are at the low end of the range.
Ensemble-averaged measures of order, such as
and
q, cannot provide any information concerning the heterogeneous, localized microstructure development that occurs under application of elongational flow. Recently, local thermodynamic-like variables at the atomistic level, such as entropy and enthalpy, were introduced as a means of quantifying the spatial and temporal evolution of flow-enhanced nucleation that spontaneously occurred in the
melt under PEF at 450 K [
16]. The advantage of these variables is that they are associated with each atom in the simulation, defined within a very small environment of that particular particle. They are thus defined based on very localized spatial averages, but no temporal averaging is applied. This implies that these variables can change from one time step to the next, and their time evolution can thus be tracked, rendering an efficient mechanism for observing flow-enhanced nucleation (FEN) events as they appear and grow in size or wane and fade away over the course of a simulation. A key aspect of these variables is their ability to distinguish between small liquid-like and crystalline-like regions within the bulk sample. Of course, these variables can also be ensemble-averaged, if desired, to provide a gross measure of the macroscopic entropy and enthalpy of the bulk liquid.
The local or atomistic entropy,
, is estimated from the atomic radial distribution (pair correlation) function defined relative to particle
i according to
where
(
in LJ units for the quiescent liquid) is the overall particle number density,
is Boltzmann’s constant,
r is the spatial coordinate, and
Å (5.1 in LJ units) is the cut-off limit for the integration [
18].
is the radial distribution function (RDF) centered at the
i-th atom. (Note that the effect of particle density on atomic entropy was discussed extensively in Appendix A of Ref. [
18]. The effects of the other parameters were analyzed in Appendix B of Ref. [
18]. In addition, note that
in LJ units for the liquid state and higher for the semicrystalline state; its value was adjusted in accordance with the
simulations). The RDF is approximated as
where
is the distance between atoms
i and
j (i.e., the neighbors of atom
i) and
(=0.07 in LJ units) is a broadening parameter [
18]. The average local atomistic entropy,
, is defined as
and it is used herein in lieu of
to increase the resolution and continuity in distinguishing between different highly localized phases. In this expression,
is the number of neighbors of atom
i within a cut-off distance of
Å (or
). One advantage of
and
over other measures of entropy is that they are defined only over a small neighborhood of a given particle, not ensemble-averaged over all particles within the simulation cell [
18]. Therefore, they can distinguish local regions of liquid-like or solid-like behavior that coexist within an inhomogeneous sample. This can be a great aid to examining flow-enhanced nucleation and flow-induced crystallization, as discussed below.
The local or atomistic enthalpy is defined as
, where
p and
V are the system pressure and volume, respectively, and
N is the total number of particles. (An extensive discussion of this definition was provided in Appendix A of Ref. [
18]).
is the potential energy of atom
i, which is readily calculated from the simulation output. The average local enthalpy,
, is defined in a similar manner as the corresponding average local entropy,
in Equation (
13). It should be noted that the atomistic entropy and enthalpy were first introduced and used by Piaggi et al. [
93] to distinguish between liquid and various solid crystalline phases of sodium and aluminium in molecular dynamics simulations. See Refs. [
18,
93] for additional details of the method and analysis of the numerical values of the model parameters in Equations (
11)–(
13).
By examining the 3-dimensional probability distribution functions of
and
for liquid and semicrystalline phases of the
melt at 450 K (9.5745 in LJ units), the liquid-to-crystalline state threshold values for the average local entropy,
, and enthalpy,
, were determined as
and
, both expressed in reduced LJ units— see Figure 2 of Ref. [
18]. In other words, particles occupying liquid-like local locations most likely possess a local entropy greater than
and a local enthalpy greater than 7.0, whereas the local entropy and enthalpy for particles in crystal-like neighborhoods would be less than
and less than 7.0, respectively. These threshold values provide a simple and direct criterion for distinguishing liquid and crystalline phases in MD simulations of polyethylene melts [
18].
The local thermodynamic-like variables defined above can be ensemble averaged (with respect to space and time) to provide some bulk-scale measures of the atomistic energetics and local ordering of the PE melt under elongational flow. Note, however, that these quantities should not be interpreted as inherent thermodynamic properties of the fluid. For instance, the long-range configurational entropy of the polymer chains is not accounted for in the atomistic entropy expression for , nor the kinetic energy of the atoms included in the atomistic enthalpy calculation, and so on. Rather, these variables are in essence microstructural variables, such as q, whose magnitude provides a rough measure of system order. The advantage of using these ensemble-averaged variables over q is that they possess a clear discontinuity at phase transitions, which is completely lacking in q.
Figure 12a displays the steady-state ensemble-averaged local atomistic entropy,
, and enthalpy,
, as functions of
for the PE melt under both UEF and PEF. Both quantities practically overlap for the UEF and PEF simulations, suggesting a minor difference between uniaxial and planar extension concerning the short-range or local configurational properties and energetic responses of entangled linear melts to these purely extensional flow fields. The average local entropies are practically independent of the flow strength in the range
, but exhibit a dramatic discontinuous drop at
, again remaining constant at higher flow strengths. This discontinuity arises from a first-order type transition of a liquid to semicrystalline phase induced by the extensional flow. It is evident that the local ordering of the polymer is solely a function of the material phase and does not change with the flow strength and long-range orientation and stretching of the molecules. The local enthalpy behaves somewhat similarly in the sense that it exhibits a discontinuous drop at
where the phase transition occurs. However, unlike
,
is a relatively weak decreasing function of
in the liquid phase and a moderate descending function in the semicrystalline phase. This decreasing trend is attributed to the intermolecular LJ and torsional energies that continuously decrease even at very high
as molecules stretch and the inherent dihedral angles assume more
trans conformations, which consequently, allow segments of individual chains to approach each other and pack into definite crystal lattices, thereby reducing the intermolecular LJ energy—see also Figure 11b of Ref. [
18].
Inspired by a thermodynamic Legendre transformation, an average local Gibbs free energy can be defined as
.
Figure 12b depicts the ensemble-average local atomic Gibbs free energy,
, for UEF and PEF as a function of
. Despite the large error bars, it is evident from this figure that the liquid and semicrystalline phases possess very different local Gibbs free energies, rendering this quantity a valuable phase assessment tool. Note that at high temperatures (such as 450 K),
is dominated by the local entropy and seemingly not particularly informative; however, it was demonstrated previously [
18] that the transient
response exhibits a minimum at the initiation of FIC, which is an indication of FIC not manifested by either
or
individually. Using the stated values of
and
, and the dimensionless temperature corresponding to 450 K (9.5745), the threshold value of atomic Gibbs free energy is
in LJ units. This quantity then provides a rather definitive value for distinguishing between liquid-like and solid-like behavior of the PE material, which can be exploited in the transient response of the melt under elongational flow.
To illustrate the use of local Gibbs energy in determining the state of the local microstructure in the PE material,
Figure 13 displays snapshots of the simulation cell at various Hencky strain units for startup of UEF at
. The atoms comprising the simulation box are colored in the instantaneous snapshots based on their individual values of
, with cooler colors representing low values of
and the opposite for warmer colors. Under quiescent conditions (
), almost all simulation atoms have
values below the threshold value of
; however, there are a number of randomly-located localized regions (of roughly 5–10 Å in size) that possess
, indicating a degree of crystal-like ordering. These nucleation sites tend to appear spontaneously for a brief period of time and then gradually fade away, like fireflies flashing over a grassy field at dusk. At
after inception of flow, these crystal-like nucleation sites continue to appear spontaneously and then fade away at random locations under flow, but in greater number and possibly a larger overall dimension. (To date, work continues on a viable method to quantify the size and kinetics of this flow-enhanced nucleation.) At
, the nucleates stabilize and begin to increase in size rapidly, ultimately forming into relatively large crystallites on the order of 50 Å in diameter.
We hypothesize that the onset of the phase transition behavior at
of the
simulation discussed above originates from random (with respect to space and time) flow-enhanced nucleation events of local crystal-like order that appear within small localized regions of the sample. At
, these fluctuating nucleates are unstable, possessing very short lifetimes, and disappear rapidly after formation before they can stabilize into localized semicrystalline regions. It is worth emphasizing that these random FEN events occur at all flow strengths, even under quiescent conditions (see the first snapshot at
of
Figure 13). At higher
, once a critical value of strain deformation has been attained (
for
, but around two strain units for higher
), these nucleates stabilize and begin to grow in size, developing into stable crystalline domains that eventually enlarge to a global steady-state FIC phase (at
).
An important aspect of FIC from both theoretical and practical perspectives is the kinetics of nucleation and crystallization. The threshold values of the local thermodynamic-like variables provide a means to delineate and quantify phase transitions in atomistic simulations. Since these quantities are computed for individual atomic units, they can be conveniently used to determine an average concentration or size for the liquid and semicrystalline phases at each time step of simulation. Such information conceptually facilitates the quantification of the dynamic process of crystallization (or melting), hence enabling measurement of the kinetics of the FIC process. Using these threshold values, the number of particles occupying liquid-like, , and crystalline-like, , neighborhoods, can be calculated, respectively, as the number of particles whose and . The fraction of atoms occupying liquid and semicrystalline states are given by and , where is the total number of particles (united atoms) in the simulation cell.
In prior work [
18], a first-order reversible reaction (
) between the liquid and semicrystalline states (or localized domains within the simulation cell) was proposed [
18], and an expression was derived for quantifying the fraction of liquid-like particles within the simulation box as
where
and
are the forward and reverse reaction rate constants with
where
is the fraction of atoms occupying local liquid-like environments at steady state and
is the equilibrium constant—see Ref. [
18] for more details. The overall degree of crystallinity can be approximated as
, which is essentially the steady-state fraction of particles occupying a semicrystalline local state.
The fraction for the liquid-like phase upon startup of the
melt under UEF at various
is depicted in
Figure 14 as a function of the Hencky strain (dimensionless time). It is apparent that the system remains liquid-like at
with a value of
at all times, in agreement with all other evidence discussed in this section. At
, however,
only up to
, at which time the fraction of liquid-like particles begins to drop precipitously, plateauing at a low value of approximately 0.15 once steady state has been attained at
. The behavior of
at higher
is qualitatively similar, although the onset of crystallization shifts to lower Hencky strains ranging from 2.0–2.5. These results are in agreement with the observed FIC onset in terms of
of prior PEF simulations [
18]. The longer delay of FIC at
arises from the fact that the dominant phase of the system at this critical flow strength is liquid up to
when the crystallization initiated via flow-enhanced nucleation events (that are randomly located throughout the sample) begins to develop from stable nuclei—see
Figure 8a,
Figure 10 and
Figure 13.
Table 2 summarizes the forward,
, and reverse,
, rate constants obtained from a nonlinear least-squares fitting of
data versus time (after excluding the initial lag times) to Equation (
14) for
under steady-state UEF. The equilibrium constants,
K, densities of the semicrystalline phase,
, and degrees (percentages) of crystallinity,
, are also listed. The corresponding values from the PEF simulations (see Table 2 of Ref. [
18]) are displayed in parentheses for comparison. It is evident from these values that the forward rate constants increase almost linearly with flow strength, implying a faster FIC at higher
. The reverse rate constants, equilibrium constants, and
also increase with flow strength, although with significantly smaller slope values than that of
. Furthermore, the corresponding values of PEF agree reasonably well with the UEF values: the PEF rate constants are slightly higher than those of UEF at all
, however, their ratio, and therefore the equilibrium constants and degrees of crystallinity under PEF are slightly lower than the corresponding UEF values. Consequently, direct evidence once again indicates that UEF has stronger orientation and ordering capability than PEF. The values of
and
are quite reasonable based on observations from experiments of stretched PE samples [
77,
78,
79,
80].
The semicrystalline phase developed under flow via FIC using NEMD simulations can be characterized in terms of the radial distribution (pair correlation) function and structure factor, which allow direct comparison with spectroscopic experimental data.
Figure 15 displays the total RDF,
, and intermolecular pair correlation function,
(inset) of the
melt at equilibrium (
) and under UEF (solid lines) and PEF (dashed lines) for two representative extension rates, the lower (
) depicting a stretched liquid state and the larger (
) a semicrystalline phase. The first point to notice in this figure is the virtually identical curves of UEF and PEF at corresponding values of
, indicating negligible influence of the geometry of extensional flows on the short to medium length-scale ordering and microstructure of the molecules; this observation agrees very well with the overlapping curves of
,
, and
observed in
Figure 12.
As discussed by Nafar Sefiddashti et al. [
18], there are five prominent peaks in each total RDF (
) at positions 1.54, 2.58, 3.14, 4.04, and 5.16 Å (relative to the test atom at
), respectively denoting the bond distance (1,2 carbon atoms), bond angle (1,3 atoms), a
gauche configuration, and
trans configurations (1,4 and 1,5 atoms)—see Figure 9 of Ref. [
41]. These peaks are present at all
, but the relative heights of the last three change with increasing
: the peak at 3.14 Å shrinks as the number of
gauche dihedral angles diminishes as the molecules extend, whereas the last two peaks grow in height due to the increasing number of
trans dihedral angles. At higher distances from the test atom (
), the equilibrium and
distributions practically overlap since neither liquid possess any long-range order; however, at
, evident long-range ordering persists out to 20–25 Å.
The intermolecular distributions (
) at
(shown in the inset of
Figure 15) possess two distinct peaks at 5.46 and 10.34 Å, but practically no long-range order beyond the second nearest neighbor shell is observed. This suggests significant local longitudinal ordering of the stretched liquid-phase molecules along the flow direction at
, but very little lateral ordering of the individual chains perpendicular to their common axis of orientation. At
, however, the first two peaks of the intermolecular RDF shift slightly to the left at positions 5.10 and 9.48 Å, indicating a relatively higher density in the semicrystalline phase. The
ensemble simulations allow calculation of the density change due to crystallization, as the simulation cell dimensions are adjusted by the applied barostat. The simulation results show a 15–17.5% increase in the system density in the range
for the semicrystalline phase compared to the quiescent melt at 450 K, as tabulated in
Table 2. Furthermore, the
curve at
has several additional peaks beyond the second one, indicating long-range intermolecular packing of the semicrystalline structure. These additional peaks decay after approximately 20–25 Å, indicating crystallites on the order of 50 Å in diameter. With the degree of crystallinity at roughly 90% at steady state, this implies that the various nucleation sites formed at small strains and random locations grow in size under continued strain and eventually encroach upon each other (see
Figure 13), but do not merge together to form a single larger crystallite. It is therefore likely that there exist dislocation (and possibly disclination) defects lying between them.
The global ordering of the FIC state is depicted in
Figure 16a for a random slab of width 11.8 Å (oriented perpendicularly to the direction of flow) of the PE semicrystalline phase produced under steady-state UEF at
. In this snapshot, taken at a random time step within the steady-state regime, atoms occupying different molecules have been assigned various assorted colors, whereas atoms belonging to the same molecule all possess the same color. Most of the molecules shown in the snapshot appear as single dots because the molecules are almost fully aligned axially, and the snapshot is perpendicular to the direction of alignment. Throughout the sample, semicrystalline order is apparent, interlaced with small regions of amorphous character.
The pair correlation function for a random slab of 3 Å width (oriented perpendicularly to the direction of flow) of the semicrystalline state at
is displayed in
Figure 16b. A very narrow slab was chosen to minimize the peak-broadening effects on
from the axial (
z) dimension; this allows for a much more accurate calculation of the crystal lattice parameters in the
x–
y plane, as discussed shortly. The first tall, narrow peak at 1.54 Å is associated with the carbon-carbon (1,2) bond distance. The second peak corresponds to the (1,3) carbon distance of 2.58 Å. The reason this peak is so small is because the slab is only 3 Å thick, and the molecules are mostly aligned axially such that these chain units are perpendicular to the slab; therefore, very few of these units lie within the slab chosen for the calculation. This distance represents the lattice vector
b. The next two peaks are associated with lattice vectors
c (4.73 Å) and
a (8.72 Å), respectively. The fact that all three vectors are of unequal length, and since angles
and
are orthogonal whereas
is not, indicates a monoclinic lattice structure. Seto et al. [
79] indexed a linear PE material formed under mechanical stress to a monoclinic lattice with
,
, and
Å, which agrees reasonably well with the simulation results. The PEF simulations generated lattice parameters of
,
, and
Å [
17], which also agree well with the UEF and experimental values.
A method for characterizing and distinguishing liquid-like and crystal-like microstructures using MD data that offers direct comparison to experimental data is to calculate the static structure factor from the Fourier transform of the total radial distribution function,
. The structure factor can then be directly compared to experimental crystal X-ray diffraction (XRD) data.
Figure 17 presents the structure factors for the simulated
melt at 450 K under equilibrium conditions (
; blue curve) and steady-state UEF at
(orange curve). In addition, there are the structure factors determined using XRD experiments for a n-eicosane liquid at 315 K (green curve) and at 298 K for a quiescently grown crystal [
27]. A comparison of the structure factors for the
melt at
and 450 K and eicosane at 315 K reveals a direct overlap of the two curves at essentially all wavenumbers, one simulated and the other obtained from experiment. This provides reasonable evidence for the accuracy of the SKS potential model for alkane and PE materials. Both structure factors display characteristics of liquid alkane materials, as discussed elsewhere [
27]. Of more importance, however, are the
profiles for the simulated melt at
and the experimental XRD data of the eicosane crystal, where the simulated PE exhibits distinct Bragg peaks at low wavenumbers that correspond closely to similar peaks in the experimental eicosane crystal, which was indexed to a triclinic crystal arrangement with lattice vector dimensions
,
, and
, where the final value lies along the axis of the fully extended eicosane molecules constituting the crystalline material. The overall degree of similarity between the simulated and experimental
curves provide additional evidence of the occurrence of FIC in the simulated PE melt under UEF. Note that a more detailed comparison between simulation and experimental
S(k) can be found for PEF in Ref. [
17].