logo

A Unified Thermodynamic Model of Flow-induced Crystallization of Polymer

ARTICLE

A Unified Thermodynamic Model of Flow-induced Crystallization of Polymer

Nie Cui
Peng Fan
Xu Ting-Yu
Sheng Jun-Fang
Chen Wei
Li Liang-Bin
Chinese Journal of Polymer ScienceVol.39, No.11pp.1489-1495Published in print 18 Sep 2021Available online 03 Aug 2021
6800

We propose a unified thermodynamic model of flow-induced crystallization of polymer (uFIC), which incorporates not only the conformational entropy reduction but also the contributions of flow-induced chain orientation, the interaction of ordered segments, and the free energy of crystal nucleus and crystal morphology. Specifically, it clarifies the determining parameters of the critical crystal nucleus size, and is able to account for the acceleration of nucleation, the emergence of precursor, different crystal morphologies and structures induced by flow. Based on the nucleation barrier under flow, we analyze at which condition precursor may occur and how flow affects the competition among different crystal forms such as orthorhombic and hexagonal phases of polyethylene. According to the uFIC model, the different crystal morphologies and structures in the flow-temperature space have been clarified, which give a good agreement with experiments of FIC.

pic

Considering the modified thermodynamic parameters of initial and final states under flow, we propose a uFIC model for flow-induced polymer crystallization. The model not only clarified the determinants of the critical nucleus size and accelerated nucleation, but also successfully explained the appearance of precursors and different crystal morphologies and structures under flow.

Flow-induced crystallizationUnified thermodynamic modelFlow-induced chain orientation

INTRODUCTION

Flow- or strain-induced crystallization (FIC or SIC) of polymers occurs not only in processing like film blowing, injection molding and fiber spinning, but also in service of some polymers like natural rubber and shape memory materials.[1-5] Though large efforts from both academic and industrial communities have been devoted for understanding FIC, it still remains as a classic and important challenge in polymer physics.[5,6] In recent two decades, new techniques such as in situ synchrotron radiation X-ray scattering and computer simulations provide more in-depth information about the kinetic pathway of FIC, but we still lack a thermodynamic theory of FIC to incorporate these new observations, such as the mesophase or precursor, shish-kebab, and the emergence of new crystal forms.[7-15]

At present, Flory’s conformational entropy reduction model (CERM) proposed in 1947 is the most widely recognized model for FIC, which was originated from uniaxial stretching of natural rubber.[16] CERM depicts a process of flow-induced chain straightening, which results in conformational entropy reduction (CER, ΔScon \Delta {S}_{\mathrm{c}\mathrm{o}\mathrm{n}} pic) of melt. Based on the classic nucleation theory (see Fig. 1a), CERM simply incorporates ΔScon \Delta {S}_{\mathrm{c}\mathrm{o}\mathrm{n}} pic (see Fig. 1b) and expresses the nucleation barrier under flow ΔGf\Delta {G}_{\rm{f}}^{*}pic as

1
Schematic illustration of the free energy barrier of nucleation (a) at quiescent condition and (b) under flow from the conformational entropy reduction (CERM) model. ΔG0 \Delta {G}_{0}^{*} pic and ΔGf \Delta {G}_{\rm{f}}^{*} pic are the nucleation barrier at quiescent condition and under flow; ΔScon \Delta {S}_{\mathrm{c}\mathrm{o}\mathrm{n}} pic is the conformational entropy reduction of melt induced by flow.
pic

1 ΔGf=ΔG0+TΔScon \Delta {G}_{\rm{f}}^{*}=\Delta {G}_{0}^{*}+T\Delta {S}_{\rm{c}\rm{o}\rm{n}} pic

where ΔG0 \Delta {G}_{0}^{*} pic represents the nucleation barrier at quiescent condition. As ΔScon \Delta {S}_{\rm{c}\rm{o}\rm{n}} pic is negative, thus ΔGf<ΔG0 \Delta {G}_{\rm{f}}^{*}< \Delta {G}_{0}^{*} pic, namely the nucleation barrier is reduced. Correspondingly, the equilibrium melting temperature Tmf0 {T}_{\rm{mf}}^{0} pic of the crystal increases under flow, which can be written as

2 1Tmf0=1Tm0ΔSconΔH \frac{1}{{T}_{\rm{m}\rm{f}}^{0}}=\frac{1}{{T}_{\rm{m}}^{0}}-\frac{\Delta {S}_{\rm{c}\rm{o}\rm{n}}}{\Delta H} pic

where Tm0 {{T}}_{\mathrm{m}}^{0} pic and ΔH \Delta H pic are the equilibrium melting temperature and the crystallization enthalpy of the crystal at quiescent condition, respectively. CERM can qualitatively account for flow-induced enhancement of nucleation rate and has been widely employed to explain FIC experiments and simulations.[12-14] Later on, although various expressions of ΔScon \Delta {S}_{\mathrm{c}\mathrm{o}\mathrm{n}} pic have been introduced by different groups and chain dynamics of polymer without crosslink is also considered, new models keep the essential physics of CERM and take ΔScon \Delta {S}_{\mathrm{c}\mathrm{o}\mathrm{n}} pic as the core contributor in understanding FIC.[17-19]

Conformational entropy is a global parameter on the individual chains, while nucleation and crystallization are packing processes involving local multi-chain segments, suggesting that ΔScon \Delta {S}_{\rm{c}\rm{o}\rm{n}} pic may not be the most important contributor in FIC. Indeed, CERM misses several important physics observed in experiments and simulations.[20-25] Fig. 2 presents a schematic illustration of the factors modified by flow. Starting from the classic nucleation theory with liquid and crystal states, we first examine what happen in the initial polymer liquid under flow. (i) Chain orientation. Chain straightening or ΔScon \Delta {S}_{\rm{c}\rm{o}\rm{n}} pic only considers individual chains and does not take chain orientation or parallel packing involving multi-chain segments into account, while parallel packing of segments is inevitably required in crystal nucleation. In uniaxial flow, the chain straightening and chain orientation occur synchronously, leading to little attention on the effect of orientation on FIC for a long period. However, a recent discovery about the frustrating SIC of natural rubber during biaxial stretching has suggested the crucial role of chain orientation on FIC.[20] Therefore, in addition to ΔScon \Delta {S}_{\rm{c}\rm{o}\rm{n}} pic, entropy change related to orientation ΔSori \Delta {S}_{\rm{o}\rm{r}\rm{i}} pic should be put into the theory of FIC separately. (ii) Chain interaction. Flow-induced preordering or precursor has been widely reported in FIC experiments, in which non-crystalline shish detected with in situ X-ray scattering is a representative example.[9,26] These experimental observations reveal that not only entropy but also enthalpy or molecular interactions are changed by flow. Thus flow-induced enthalpy change of polymer liquid ΔHfl \Delta {H}_{\rm{f}}^{\rm{l}} pic should be considered in FIC theory. After depicting the initial liquid, we then shift our attention to the final crystalline state. (iii) Crystal morphology and structure. Flow modifies crystal morphology (e.g., from lamellae to shish-kebab) and crystal form (e.g., from orthorhombic to hexagonal crystals for polyethylene (PE)) has been largely documented since the earliest FIC study,[27-33] but historically, these factors had been surprisingly overlooked in FIC theory. If flow induces new crystal forms, enthalpy difference ΔHfc \Delta {H}_{\rm{f}}^{\rm{c}} pic between crystals at quiescent and strained conditions should be introduced. Meanwhile, transforming crystal morphology from lamellae to shish should change the free energy density of end surface, that is, from folded-chain surface σe {\sigma }_{\rm{e}} pic to extended-chain surface σef {\sigma }_{\rm{ef}} pic. Evidently, in addition to ΔScon \Delta {S}_{\rm{c}\rm{o}\rm{n}} pic in CERM, the new FIC theory should consider the above three factors.

2
Flow-induced changes of the physical parameters in FIC. ΔScon \Delta {S}_{\rm{c}\rm{o}\rm{n}} pic and ΔSori \Delta {S}_{\rm{o}\rm{r}\rm{i}} pic are the entropy change of conformation and orientation of melt induced by flow; ΔHfl \Delta {H}_{\rm{f}}^{\rm{l}} pic is the enthalpy change of polymer melt due to the interaction of ordered segments; ΔHfc \Delta {H}_{\rm{f}}^{\rm{c}} pic is the enthalpy difference between two crystal forms like flow-induced orthorhombic and hexagonal phases of PE; σe {\sigma }_{\rm{e}} pic and σef {\sigma }_{\rm{ef}} pic are the free energy densities of crystal end surface at quiescent condition and under flow.
pic

Based on the experiment and simulation results, FIC models incorporating the effects of orientation, chain interaction, new crystal morphology and structure have been proposed in recent years, which, however, were considered separately. Chen et al.[20] introduced the orientation entropy and the chain interaction between the ordered segments, while Liu et al.[24] took the free energy change of the final crystalline state into account and proposed the entropy reduction-energy change (ER-EC) model. Although these efforts make a stepwise progress for understanding FIC, a unified FIC theory incorporating the above three factors has not been developed yet.

In this work, starting from classic nucleation theory, we propose a thermodynamic model for FIC to unify flow-induced changes of conformational entropy ΔScon \Delta {S}_{\rm{c}\rm{o}\rm{n}} pic, orientation entropy ΔSori \Delta {S}_{\rm{o}\rm{r}\rm{i}} pic, chain interaction of polymer liquid ΔHfl \Delta {H}_{\rm{f}}^{\rm{l}} pic, and modifications of crystal structure ΔHfc \Delta {H}_{\rm{f}}^{\rm{c}} pic and morphology σef {\sigma }_{\rm{ef}} pic. Our model is a thermodynamically phenomenological model but explicitly unifies all structural changes in FIC study. For the convenience of description, we name it as uFIC model hereafter. The possible calculation methods for the changes of conformational entropy ΔScon \Delta {S}_{\rm{c}\rm{o}\rm{n}} pic, orientation entropy ΔSori \Delta {S}_{\rm{o}\rm{r}\rm{i}} pic and chain interaction of polymer liquid ΔHfl \Delta {H}_{\rm{f}}^{\rm{l}} pic are given in the Appendix, as they have been discussed in early studies.[16,20]

THEORY AND METHOD

To deduce the uFIC model, we first incorporate flow-induced variations of entropy and enthalpy into the chemical potential μ=HTS \mu =H-TS pic. The chemical potential difference per unit volume between polymer crystal and melt (liquid) phases under flow can be written as

3 Δμf=ΔHfTΔSf=ΔH+ΔHfcΔHflT(ΔS(ΔScon+ΔSori)) \Delta {\mu }_{\rm{f}}=\Delta {H}_{\rm{f}}-T\Delta {S}_{\rm{f}}=\Delta H+\Delta {H}_{\rm{f}}^{\rm{c}}-\Delta {H}_{\rm{f}}^{\rm{l}}-T\left(\Delta S-\left(\Delta {S}_{\mathrm{c}\mathrm{o}\mathrm{n}}+\Delta {S}_{\mathrm{o}\mathrm{r}\mathrm{i}}\right)\right) pic

where ΔH \Delta H pic and ΔHf=ΔH+ΔHfcΔHfl \Delta {H}_{\rm{f}}=\Delta H+\Delta {H}_{\rm{f}}^{\rm{c}}-\Delta {H}_{\rm{f}}^{\rm{l}} pic are the crystallization enthalpies of polymer melt at quiescent condition and under flow, respectively. While ΔS \Delta S pic and ΔSf=ΔS(ΔScon+ΔSori) \Delta {S}_{\rm{f}}=\Delta S-\left(\Delta {S}_{\mathrm{c}\mathrm{o}\mathrm{n}}+\Delta {S}_{\mathrm{o}\mathrm{r}\mathrm{i}}\right) pic are the crystallization entropies of polymer melt at quiescent condition and under flow, respectively. Here, the contribution of conformation and orientation to the entropy change of the melt are considered. Note in current work, the symbols with subscript “f” represent the physical parameters under flow. In Eq. (3), we assume that the difference in the entropies of crystal phase between quiescent and strained conditions is negligible. At the equilibrium melting temperature under flow Tmf0 {T}_{\rm{mf}}^{0} pic, the chemical potential of these two phases is equal, Δμf=0 \Delta {\mu }_{\rm{f}}=0 pic, which results in

4 Tmf0=ΔHfΔSf=ΔH+ΔHfcΔHflΔS(ΔScon+ΔSori) {T}_{\rm{mf}}^{0}=\frac{\Delta {H}_{\rm{f}}}{\Delta {S}_{\rm{f}}}=\frac{\Delta H+\Delta {H}_{\rm{f}}^{\rm{c}}-\Delta {H}_{\rm{f}}^{\rm{l}}}{\Delta S-\left(\Delta {S}_{\mathrm{c}\mathrm{o}\mathrm{n}}+\Delta {S}_{\mathrm{o}\mathrm{r}\mathrm{i}}\right)} pic

Then, Eq. (3) can be written as

5 Δμf=ΔHf(1TTmf0)=ΔHfΔTfTmf0 \Delta {\mu }_{\rm{f}}=\Delta {H}_{\rm{f}}\left(1-\frac{T}{{T}_{\rm{mf}}^{0}}\right)=\frac{\Delta {H}_{\rm{f}}\Delta {T}_{\rm{f}}}{{T}_{\rm{mf}}^{0}} pic

where ΔTf\Delta {T}_{\rm{f}}pic is the supercooling under flow. Inserting Tm0=ΔHΔS{T}_{\rm{m}}^{0}=\dfrac{\Delta H}{\Delta S}pic into Eq. (4) and taking the reciprocal lead to the following equation

6 1Tmf0=ΔHΔHfTm0ΔScon+ΔSoriΔHf \frac{1}{{T}_{\rm{mf}}^{0}}=\frac{\Delta H}{\Delta {H}_{\rm{f}}{T}_{\rm{m}}^{0}}-\frac{\Delta {S}_{\mathrm{c}\mathrm{o}\mathrm{n}}+\Delta {S}_{\mathrm{o}\mathrm{r}\mathrm{i}}}{\Delta {H}_{\rm{f}}} pic

The CERM and other FIC models generally assume that the free energies of the critical nucleus and the final state under flow keep the same as those at quiescent condition. However, at the same crystallization temperature Tc {T}_{\rm{c}} pic, the size of the critical crystal nucleus should decrease as the equivalent supercooling under flow is expected to increase. Obviously, the free energy landscape changes under flow, which will be considered in our uFIC model (see Fig. 3a). Following the classical nucleation theory,[34] the free energy change for the formation of a cylindrical crystal nucleus from the liquid phase under flow is given by

3
Schematic illustration of the free energy barrier of nucleation under flow. (a) The uFIC model incorporates the entropy change of polymer melt originated from conformation and orientation ΔSfl=ΔScon+ΔSori \Delta {S}_{\rm{f}}^{\rm{l}}=\Delta {S}_{\mathrm{c}\mathrm{o}\mathrm{n}}+\Delta {S}_{\mathrm{o}\mathrm{r}\mathrm{i}} pic, enthalpy change of polymer melt ΔHfl \Delta {H}_{\rm{f}}^{\rm{l}} pic due to oriented segments and the free energy change of final nucleus with different crystal morphologies and structures induced by flow ΔHfc\Delta {H}_{\rm{f}}^{\rm{c}}pic. (b) The condition for the existence of a thermodynamically metastable intermediate phase under flow.
pic

7 ΔGf=πR2LΔμf+2πRLσsf+2πR2σef \Delta {G}_{\rm{f}}=\text{π} {R}^{2}L\Delta {\mu }_{\rm{f}}+2\text{π} RL{\sigma }_{\rm{sf}}+2\text{π} {R}^{2}{\sigma }_{\rm{ef}} pic

where σef {\sigma }_{\rm{ef}} pic and σsf {\sigma }_{\rm{sf}} pic are the free energy densities for the end and side surfaces of a cylindrical nucleus of length L L pic and radius R R pic under flow, Δμf \Delta {\mu }_{\rm{f}} pic is the chemical potential difference per unit volume between the crystal and melt phases under flow. Combining Eq. (5), the critical nucleus thickness Lcf {L}_{\rm{cf}} pic and the critical nucleus radius Rcf {R}_{\rm{cf}} pic are obtained by finding the maximum in ΔGf \Delta {G}_{\rm{f}} pic with respect to nucleus length L L pic and nucleus radius R R pic in Eq. (7),

8 pic

where we assume that the ratios of these two values to the length Lc {L}_{\rm{c}} pic and radius Rc {R}_{\rm{c}} pic of the critical nucleus at quiescent condition are KL {K}_{\rm{L}} pic and KR {K}_{\rm{R}} pic, respectively. Then the free energy densities of the end and side surfaces of crystals under flow can be obtained as

9 pic

Substituting these expressions of Lcf {L}_{\mathrm{c}\mathrm{f}} pic and Rcf {R}_{\mathrm{c}\mathrm{f}} pic into Eq. (7) leads to the free energy barrier of critical crystal nucleus under flow as

10 ΔGf=8πσef(σsf)2(Tmf0)2(ΔHf)2(ΔTf)2 \Delta {G}_{\rm{f}}^{*}=\frac{8\text{π} {\sigma }_{\rm{ef}}{\left({\sigma }_{\mathrm{s}\mathrm{f}}\right)}^{2}{\left({T}_{\rm{mf}}^{0}\right)}^{2}}{{\left(\Delta {H}_{\rm{f}}\right)}^{2}{\left(\Delta {T}_{\rm{f}}\right)}^{2}} pic

Inserting σef {\sigma }_{\rm{ef}} pic and σsf {\sigma }_{\rm{sf}} pic of Eq. (9) into Eq. (10), we have the nucleation barrier of critical nucleus under flow as

11 ΔGf=K(1+1Δμ0(ΔHfcΔHfl)+1Δμ0T(ΔScon+ΔSori))ΔG0 \Delta {G}_{\rm{f}}^{*}=K\left(1+\frac{1}{\Delta {\mu }_{0}}\left({\Delta H}_{\rm{f}}^{\rm{c}}-{\Delta H}_{\rm{f}}^{\rm{l}}\right)+\frac{1}{\Delta {\mu }_{0}}T(\Delta {S}_{\rm{con}}+\Delta {S}_{\rm{ori}})\right)\Delta {G}_{0}^{*} pic

where ΔG0 \Delta {G}_{0}^{*} pic and Δμ0 \Delta {\mu }_{0} pic are the nucleation barrier of critical nucleus and the chemical potential difference per unit volume between polymer crystal and melt (liquid) phases at quiescent condition, respectively. We also inform a shape factor K as

12 K=KLKR2=VcfVc K={K}_{\rm{L}}{K}_{\rm{R}}^{2}=\frac{{V}_{\rm{cf}}}{{V}_{\rm{c}}} pic

where Vc{V}_{\rm{c}}pic and Vcf{V}_{\rm{cf}}pic are the volume of critical cylindrical nucleus at quiescent condition and under flow, respectively. According to Eq. (3) and Eq. (9), Eq. (11) can be further written as

13 ΔGf=(σefσe)(σsfσs)2(Δμ0Δμf)2ΔG0 \Delta {G}_{\rm{f}}^{*}=\left(\frac{{\sigma }_{\rm{ef}}}{{\sigma }_{\rm{e}}}\right){\left(\frac{{\sigma }_{\rm{sf}}}{{\sigma }_{\rm{s}}}\right)}^{2}{\left(\frac{\Delta {\mu }_{0}}{\Delta {\mu }_{\rm{f}}}\right)}^{2}\Delta {G}_{0}^{*} pic

DISCUSSION

As the uFIC model takes both flow-induced modifications of liquid and crystal into account, it is expected that this model can explain all the experimental observations of FIC. To analyze how flow affects nucleation process, we compare the nucleation barriers at flow and quiescent conditions. Here we assume the free energy density of the crystal side surface under flow is unchanged, σsfσs=1\dfrac{{\sigma }_{\rm{sf}}}{{\sigma }_{\rm{s}}}=1pic, then the ratio of the nucleation barriers between these two conditions is

14 ΔGfΔG0=(σefσe)(Δμ0Δμf)2 \frac{\Delta {G}_{\rm{f}}^{*}}{\Delta {G}_{0}^{*}}=\left(\frac{{\sigma }_{\rm{ef}}}{{\sigma }_{\rm{e}}}\right){\left(\frac{\Delta {\mu }_{0}}{\Delta {\mu }_{\rm{f}}}\right)}^{2} pic

Taking the nucleation barrier expressed in Eq. (14) as the basic starting point, we are going to discuss the 4 characteristic observations in FIC experiments: (i) accelerating nucleation; (ii) changing the nucleus from lamellae to shish; (iii) preordering; and (iv) inducing new crystal form.

(i) Accelerating Nucleation

If we assume flow only changes the melt entropy, Eq. (14) can be re-written as

15 ΔGfΔG0=(σefσe)[1T(ΔScon+ΔSori)Δμ0+T(ΔScon+ΔSori)]2 \frac{\Delta {G}_{\rm{f}}^{*}}{\Delta {G}_{0}^{*}}=\left(\frac{{\sigma }_{\rm{ef}}}{{\sigma }_{\rm{e}}}\right){\left[1-\frac{T(\Delta {S}_{\mathrm{c}\mathrm{o}\mathrm{n}}+\Delta {S}_{\mathrm{o}\mathrm{r}\mathrm{i}})}{\Delta {\mu }_{0}+T(\Delta {S}_{\mathrm{c}\mathrm{o}\mathrm{n}}+\Delta {S}_{\mathrm{o}\mathrm{r}\mathrm{i}})}\right]}^{2} pic

Below the equilibrium melting temperature (ΔT>0 \Delta T>0 pic), Δμ0<0 \Delta {\mu }_{0}< 0 pic. The molecular chain is stretched and oriented by flow, (ΔScon+ΔSori)<0 (\Delta {S}_{\mathrm{c}\mathrm{o}\mathrm{n}}+\Delta {S}_{\mathrm{o}\mathrm{r}\mathrm{i}})< 0 pic, 1T(ΔScon+ΔSori)Δμ0+T(ΔScon+ΔSori)<1 1-\dfrac{T\left(\Delta {S}_{\mathrm{c}\mathrm{o}\mathrm{n}}+\Delta {S}_{\mathrm{o}\mathrm{r}\mathrm{i}}\right)}{\Delta {\mu }_{0}+T\left(\Delta {S}_{\mathrm{c}\mathrm{o}\mathrm{n}}+\Delta {S}_{\mathrm{o}\mathrm{r}\mathrm{i}}\right)}< 1 pic. Under a relative weak flow, free energy density of the end surface of crystal keeps the same as that at quiescent condition, namely σefσe=1\dfrac{{\sigma }_{\rm{ef}}}{{\sigma }_{\rm e}}=1pic, then ΔGfΔG0<1\dfrac{\Delta {G}_{\rm f}^{*}}{\Delta {G}_{0}^{*}} < 1pic. Here we should notice that besides decreasing the nucleation barrier, flow can also enhance the movement ability of polymer segments along the stretching direction, which would lead to the increase of the pre-factor of the nucleation rate. Both the two could lead to a higher nucleation rate as observed in FIC experiments and simulations.[10-13,35-37] As the flow strength increases, the degree of molecular chain stretching increases, which can transform the crystal nucleus from lamellae to fringed micellar structure.[38] Owning to the end surface crowding effect of the fringed micellar structure, it has a larger surface free energy, σefσe>1\dfrac{{\sigma }_{\rm{ef}}}{{\sigma }_{\rm e}} > 1pic, which would have some hindering effects on FIC. When the external flow is sufficient strong that the crowding effect on the crystal surface has the opportunity to be counteracted by the tensile stress on the molecular chain, the free energy density of end surface is reduced. Here, the crystal morphology and structure may change, which will be discussed below. At ΔT<0 \Delta T < 0 pic (Δμ0>0 \Delta {\mu }_{0}>0 pic), namely FIC occurs at Tc>Tm0{T}_{\rm c} > {T}_{\rm m}^{0}pic (but Tc<Tmf0{T}_{\rm c} < {T}_{\rm{mf}}^{0}pic), ΔGfΔG0<1\dfrac{\Delta {G}_{\rm f}^{*}}{\Delta {G}_{0}^{*}} < 1pic requires T(ΔScon+ΔSori)< T\left(\Delta {S}_{\mathrm{c}\mathrm{o}\mathrm{n}}+\Delta {S}_{\mathrm{o}\mathrm{r}\mathrm{i}}\right) <picΔμ0 -\Delta {\mu }_{0} pic, which supports that the deformation exceeding the critical strain is a necessary condition for FIC to occur at ΔT<0\Delta T < 0pic.

(ii) Changing the Nucleus from Lamellae to Shish

Considering that the free energy density of the crystal end surface increases due to the external flow, σefσe>1\dfrac{{\sigma }_{\rm{ef}}}{{\sigma }_{\rm e}} > 1pic, but the crystal form does not change, then the transformation from lamellar to shish nucleus can occur. The dimension of a nucleus in different directions is proportional to the free energy density of its perpendicular surface[34] and σsfσs1\dfrac{{\sigma }_{\rm{sf}}}{{\sigma }_{\rm s}}\approx 1pic, then according to Eq. (9), the shape factor K K pic can be written as

16 K=(σefσe)(Δμ0Δμf)3=(LfL0)(Δμ0Δμf)3=LfRf2L0R02 K=\left(\frac{{\sigma }_{\rm{ef}}}{{\sigma }_{\rm e}}\right){\left(\frac{\Delta {\mu }_{0}}{\Delta {\mu }_{\rm f}}\right)}^{3}=\left(\frac{{L}_{\rm f}}{{L}_{0}}\right){\left(\frac{\Delta {\mu }_{0}}{\Delta {\mu }_{\rm f}}\right)}^{3}=\frac{{L}_{\rm f}{{R}_{\rm f}}^{2}}{{L}_{0}{{R}_{0}}^{2}} pic

Here, combined with Eq. (3), we clarify the determinants of the critical crystal nucleus size, including conformational entropy change ΔScon\Delta {S}_{\rm{con}}pic, orientation entropy change ΔSori,\Delta {S}_{\rm{ori}},pic enthalpy change of polymer melt ΔHfl\Delta {H}_{\rm f}^{\rm l}pic and crystal ΔHfc\Delta {H}_{\rm f}^{\rm c}pic, and crystal morphology σef {\sigma }_{\mathrm{e}\mathrm{f}} pic. And compared with Eq. (14), under flow, Δμ0Δμf<1\dfrac{\Delta {\mu }_{0}}{\Delta {\mu }_{\rm f}} < 1pic, then K<ΔGfΔG0K < \dfrac{\Delta {G}_{\rm f}^{*}}{\Delta {G}_{0}^{*}}pic, which means that the critical crystal nucleus size is more reduced than the nucleation barrier under flow, and this is consistent with the results of simulation.[10]

Taking a lamellar nucleus with L0=20  nm {L}_{0}=20\; {\rm{nm}} pic and R0=10  nm {R}_{0}=10\; {\rm{nm}} pic at quiescent condition, if Δμf \Delta {\mu }_{\mathrm{f}} pic in FIC is twice as much as Δμ0 \Delta {\mu }_{0} pic at quiescent condition, Rf {R}_{\mathrm{f}} pic will be 3.5 nm. Here, owning to the surface crowding effect, we assume that σefσe=5\dfrac{{\sigma }_{\rm{ef}}}{{\sigma }_{\rm e}}=5pic, then K=58 K=\dfrac{5}{8} pic and the Lcf{L}_{\rm{cf}}pic increases to 100 nm in the scale of shish observed in FIC. This simple estimation is reasonable as compared with experimental observations.[24] Obviously, to meet the precondition of σefσe=5\dfrac{{\sigma }_{\rm{ef}}}{{\sigma }_{\rm e}}=5pic, the applied flow field must reach a certain strength, that is, only when the strain or strain rate is sufficiently high can shish-like crystalline structure be formed.

(iii) Preordering

If a precursor emerges in FIC with no change of the final crystal form, which may be either a transient kinetic state or a thermodynamic metastable phase corresponding to a free energy local minimum, the ratio of nucleation barriers is

17 ΔGfΔG0=(σefσe)[1+ΔHflT(ΔScon+ΔSori)Δμ0(ΔHflT(ΔScon+ΔSori))]2 \frac{\Delta {G}_{\rm f}^{*}}{\Delta {G}_{0}^{*}}=\left(\frac{{\sigma }_{\rm{ef}}}{{\sigma }_{\rm e}}\right){\left[1+\frac{{\Delta H}_{\rm f}^{\rm l}-T(\Delta {S}_{\mathrm{c}\mathrm{o}\mathrm{n}}+\Delta {S}_{\mathrm{o}\mathrm{r}\mathrm{i}})}{\Delta {\mu }_{0}-\left({\Delta H}_{\rm f}^{\rm l}-T(\Delta {S}_{\mathrm{c}\mathrm{o}\mathrm{n}}+\Delta {S}_{\mathrm{o}\mathrm{r}\mathrm{i}})\right)}\right]}^{2} pic

Obviously, when FIC occurs, Δμ0(ΔHflT(ΔScon+ΔSori))=\Delta {\mu }_{0}-\left({\Delta H}_{\rm f}^{\rm l}-T\left(\Delta {S}_{\mathrm{c}\mathrm{o}\mathrm{n}}+\Delta {S}_{\mathrm{o}\mathrm{r}\mathrm{i}}\right)\right)=picΔμf<0\Delta {\mu }_{\rm f} < 0pic. Here, the change of the free energy density of the crystal end surface under flow needs to be considered separately. (i) σefσe<1\dfrac{{\sigma }_{\rm{ef}}}{{\sigma }_{\rm e}} < 1pic. The chains in pre-ordered structure are arranged in parallel and retain some freedom of its own axes. Compared with the crystal-melt interface, the density and enthalpy difference between pre-ordered structure and the melt is smaller, which results in a smaller free energy density.[39] Therefore, the condition of ΔHflT(ΔScon+ΔSori)<0{\Delta H}_{\rm f}^{\rm l}-T\left(\Delta {S}_{\mathrm{c}\mathrm{o}\mathrm{n}}+\Delta {S}_{\mathrm{o}\mathrm{r}\mathrm{i}}\right) < 0pic may occur (see Fig. 3b), which supports that the precursor is a thermodynamically metastable phase. The occurence of this metastable precursor has been observed in FIC experiments on polyethylene and other polymers,[7,40] which can be qualitatively understood with current uFIC model as discussed above. ii) σefσe>1\dfrac{{\sigma }_{\rm{ef}}}{{\sigma }_{\rm e}} > 1pic. The fringed micellar structure of crystal nucleus is induced by strong flow, with an extended-chain surface of higher free energy density. So the condition for acceleration of nucleation (ΔGfΔG0<1)\left ( \dfrac{\Delta {G}_{\rm f}^{*}}{\Delta {G}_{0}^{*}} < 1 \right)pic requires ΔHflT(ΔScon+ΔSori)>0{\Delta H}_{\rm f}^{\rm l}-T\left(\Delta {S}_{\mathrm{c}\mathrm{o}\mathrm{n}}+\Delta {S}_{\mathrm{o}\mathrm{r}\mathrm{i}}\right) > 0pic. Here, the precursor is thermodynamically unfavorable as compared with polymer melt, which however may still occur as a transient kinetic state under flow.

(iv) Inducing New Crystal Form

Finally we consider the situation with the competition between two crystal forms like flow-induced orthorhombic and hexagonal phases of PE, where ΔHfc>0{\Delta H}_{\rm f}^{\rm c} > 0pic. According to Eq. (13), the ratio between their nucleation barriers is written as

18 ΔGhfΔGof=(σefhσefo)(ΔμofΔμhf)2=(σefhσefo)(1ΔHfcΔμhf)2 \frac{\Delta {G}_{\rm{hf}}^{*}}{\Delta {G}_{\rm {of}}^{*}}=\left(\frac{{\sigma }_{\rm {ef}}^{\rm h}}{{\sigma }_{\rm {ef}}^{\rm o}}\right){\left(\frac{\Delta {\mu }_{\rm {of}}}{\Delta {\mu }_{\rm {hf}}}\right)}^{2}=\left(\frac{{\sigma }_{\rm {ef}}^{\rm h}}{{\sigma }_{\rm {ef}}^{\rm o}}\right){\left(1-\frac{\Delta {H}_{\rm f}^{\rm c}}{\Delta {\mu }_{\rm {hf}}}\right)}^{2} pic

where ΔGof\Delta {G}_{\rm {of}}^{*}pic (ΔGhf\Delta {G}_{\rm {hf}}^{*}pic) is nucleation barrier of orthorhombic (hexagonal) phase under flow, Δμof\Delta {\mu }_{\rm {of}}pic (Δμhf\Delta {\mu }_{\rm {hf}}pic) is the chemical potential difference per unit volume between orthorhombic (hexagonal) crystal and polymer melt under flow, and σefo{\sigma }_{\rm {ef}}^{\rm o}pic (σefh)\left({\sigma }_{\rm {ef}}^{\rm h}\right)pic is the free energy density of orthorhombic (hexagonal) crystal end surface under flow. Obviously, Δμhf<0\Delta {\mu }_{\rm {hf}} < 0pic in FIC, as ΔHfc>0{\Delta H}_{\rm f}^{\rm c} > 0pic, and then 1ΔHfcΔμhf>11-\dfrac{\Delta {H}_{\rm f}^{\rm c}}{\Delta {\mu }_{\rm {hf}}} > 1pic. According to the relationship of ratio σeΔH\dfrac{{\sigma }_{\rm e}}{\Delta H}pic and the crystallization enthalpy ratio between orthorhombic phase and hexagonal phase, (σeΔH)o=3.5(σeΔH)h{\left(\dfrac{{\sigma }_{\rm e}}{\Delta H}\right)}_{\rm o}=3.5{\left(\dfrac{{\sigma }_{\rm e}}{\Delta H}\right)}_{\rm h}pic and ΔHhΔHo=13\dfrac{\Delta {H}_{\rm h}}{\Delta {H}_{\rm o}}=\dfrac{1}{3}pic,[41-43] then σefhσefo=110.5\dfrac{{\sigma }_{\rm {ef}}^{\rm h}}{{\sigma }_{\rm {ef}}^{\rm o}}=\dfrac{1}{10.5}pic. Therefore, the condition for favorable formation of the hexagonal phase (ΔGhfΔGof<1)\left(\dfrac{\Delta {G}_{\rm {hf}}^{*}}{\Delta {G}_{\rm {of}}^{*}} < 1 \right)pic requires Δμhf<0.45ΔHfc\Delta {\mu }_{\rm {hf}} < -0.45{\Delta H}_{\rm f}^{\rm c}pic. Combining Eq. (16), the relationship between Δμf \Delta {\mu }_{\mathrm{f}} pic and the critical nucleus size, it proves that for a significantly smaller critical nucleus size, the hexagonal phase is a thermodynamically stable phase, which is in good agreement with early reports.[41] In addition, the driving force Δμhf \left|\Delta {\mu }_{\mathrm{h}\mathrm{f}}\right| pic required for nucleation increases with temperature, indicating that there is a critical temperature Th {T}_{\mathrm{h}} pic, above which the hexagonal phase is more stable than the orthorhombic phase. Based on the above discussion, the crystal morphologies and structures in the flow-temperature space can be constructed, as shown in Fig. 4. As the temperature increases, the required deformation is increased for FIC to occur. And, taking PE as an example, the crystal morphology and structure gradually changes from orthorhombic lamellar crystal to orthorhombic shish crystal, and finally to hexagonal shish crystal at T>Th T>{T}_{\mathrm{h}} pic. This result agrees reasonably well with FIC experiments of PE.[24]

4
Schematic illustration of the crystal morphologies and structures of PE in the flow-temperature space.
pic

The above discussion is qualitative and based on some reasonable assumptions, which, however, demonstrates the capability of the uFIC model on understanding experimental observations. The CERM may be numerically applied to fit the flow accelerated nucleation rate, but experimental result suggests ΔScon \Delta {S}_{\mathrm{c}\mathrm{o}\mathrm{n}} pic may make a weak contribution on this effect.[20] Moreover, other observations of FIC, such as precursor, new crystal forms and morphologies have never been considered in the CERM. Through introducing the changes of orientation entropy ΔSori \Delta {S}_{\mathrm{o}\mathrm{r}\mathrm{i}} pic, chain interaction of polymer liquid ΔHfl\Delta {H}_{\rm f}^{\rm l}pic, the modifications of crystal structure ΔHfc\Delta {H}_{\rm f}^{\rm c}pic and morphology σef {\sigma }_{\mathrm{e}\mathrm{f}} pic, uFIC can not only specify the physical origins for accelerating nucleation rate, but also account for almost all experimental observations in FIC. The above discussion only refers to some specific cases, and the uFIC can be employed to explain other observations of FIC experiments and simulations.

CONCLUSIONS

Taking all the changes of chain conformation, chain orientation, interaction of ordered segments in the initial polymer melt, and the final crystal morphology and structure into account, we propose a uFIC model and rewrite the expression of nucleation free energy barrier under flow, as presented in Eq. (13). Our model can not only effectively explain the flow-accelerated polymer nucleation, but also offer understandings of other FIC experimental observations. It theoretically verifies that the intermediate phase or precursor is a metastable phase only when the free energy density of the interface between the preordering and the melt is smaller than that of the crystal-melt interface. Based on the nucleation barrier under flow, the competition of different crystal forms in the flow can be clarified. Taking PE as an example, the critical transition temperature between orthorhombic and hexagonal crystal forms can be determined by comparing the crystallization enthalpy and the end surface free energy density. In conclusion, our model has the potential to be applied to explain all the observations of FIC experiments which have been done before and can provide a theoretical guidance for further FIC study.

APPENDIX

The estimation of ΔScon \Delta {S}_{\mathrm{c}\mathrm{o}\mathrm{n}} pic has been proposed by different authors with different models.[44-46] Here we take the simple one proposed by Flory.[16] At quiescent condition, the relative coordination of chain ends of the equilibrium polymer melt is approximately considered to obey the Gaussian distribution. Under flow, the amorphous molecular chains are straightened and deviate from the random configurations, resulting in the decrease of conformational entropy. Based on Flory’s CERM model, the conformational entropy reduction ΔScon \Delta {S}_{\mathrm{c}\mathrm{o}\mathrm{n}} pic can be expressed as

19 pic

where x,y x,y pic and z z pic represent the coordinates of one end of the chain with respect to the other at quiescent condition, x,y {x}',{y}' pic and z {z}' pic are the relative coordination after stretching, kB{k}_{{\rm B}}pic is the Boltzmann constant, β=(3/2n)0.5/l \beta ={\left(3/2n\right)}^{0.5}/l pic represents the reciprocal chain displacement value that is the most probable, l l pic is the Kuhn length, n n pic is the segment number per chain, ξ \xi pic is the number of segments in the flow-induced crystalline phase, Nc{N}_{\rm{c}}pic is the total number of chains.

Considering the influence of orientation on the free energy of the initial melt, the parallel arrangement of the local stretched segments leads to the reduction in the number of possible conformations of the molecular chain, which is considered as the entropy effect. At the same time, there is an interaction between the parallel oriented segments, which manifests as an influence on enthalpy.[20] The entropy change of the melt caused by the oriented segments is defined as the orientation entropy, based on statistical mechanics and Maier-Saupe theory,[47] which can be expressed as

20   ΔSori=NskBlnf(ϕ)=aNV2NsTf2+NskBln{01exp[u(ν,f)kBT]dx} \;\Delta {S_{{\rm{ori}}}} = - {N_{\rm s}}{k_{\rm B}}\left\langle {{\rm{ln}}f\left( \phi \right)} \right\rangle = - \frac{{{a_{\rm N}}}}{{{V^2}}}\frac{{{N_{\rm s}}}}{T}{f^2} + {N_{\rm s}}{k_{\rm B}}{\rm{ln}}\left\{ {\int _0^1{\rm{exp}}\left[ { - \frac{{u\left( {\nu ,f} \right)}}{{{k_{\rm B}}T}}} \right]{\rm{d}}x} \right\} pic

where Ns {N}_{\rm{s}} pic is the number of the ordered segments, f(ϕ) f\left(\phi \right) pic is the distribution function of the angle between segments and the stretching direction, aN {a}_{\mathrm{N}} pic is a constant related to the mean field, V V pic is the molar volume of the polymer, f f pic is the Hermans’ orientation parameter, u(ν,f) u\left(\nu ,f\right) pic is the internal energy of a segment, and ν \nu pic is the segment vector. Considering that the change of volume of the polymer melt under deformation can be ignored, the enthalpy change of the melt induced by flow ΔHfl{\Delta H}_{\rm f}^{\rm l}pic equals the internal energy change due to Ns{N}_{\rm s}pic oriented segments,[20,47] which can be expressed as

21 ΔHfl=ΔUori=12NsaNV2f2 {\Delta H}_{\rm f}^{\rm l}=\Delta {U}_{\mathrm{o}\mathrm{r}\mathrm{i}}=\frac{1}{2}{N}_{\rm s}\frac{{a}_{\rm N}}{{V}^{2}}{f}^{2} pic

Through Taylor expansion, ignoring higher-order terms, according to the Eqs. (20) and (21), the contribution of orientation to the free energy of the melt can be written as follows

22 ΔGoril=ΔUoriTΔSori=af2+bf3+cf4 \Delta {G}_{\mathrm{o}\mathrm{r}\mathrm{i}}^{\rm l}=\Delta {U}_{\mathrm{o}\mathrm{r}\mathrm{i}}-T\Delta {S}_{\mathrm{o}\mathrm{r}\mathrm{i}}=a{f}^{2}+b{f}^{3}+c{f}^{4} pic

where a, b, and c are constants related to pressure and temperature.

The existence of crystals with different morphologies and structures under flow indicates that the existence of the enthalpy change of the crystal ΔHfc{\Delta H}_{\rm f}^{\rm c}pic induced by flow, which can be obtained from the melting enthalpy of the different crystals.

DATA AVAILABILITY

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

References
1 Chandran, S.; Baschnagel, J.; Cangialosi, D.; Fukao, K.; Glynos, E.; Janssen, L. M. C.; Müller, M.; Muthukumar, M.; Steiner, U.; Xu, J.; Napolitano, S.; Reiter, G.

Processing pathways decide polymer properties at the molecular level

Macromolecules 2019 52 7146 7156 10.1021/acs.macromol.9b01195

Chandran, S.; Baschnagel, J.; Cangialosi, D.; Fukao, K.; Glynos, E.; Janssen, L. M. C.; Müller, M.; Muthukumar, M.; Steiner, U.; Xu, J.; Napolitano, S.; Reiter, G. Processing pathways decide polymer properties at the molecular level. Macromolecules 2019, 52, 7146−7156.

Baidu ScholarGoogle Scholar
2 Cui, K.; Ma, Z.; Tian, N.; Su, F.; Liu, D.; Li, L.

Multiscale and multistep ordering of flow-induced nucleation of polymers

Chem. Rev. 2018 118 1840 1886 10.1021/acs.chemrev.7b00500

Cui, K.; Ma, Z.; Tian, N.; Su, F.; Liu, D.; Li, L. Multiscale and multistep ordering of flow-induced nucleation of polymers. Chem. Rev. 2018, 118, 1840−1886.

Baidu ScholarGoogle Scholar
3 Lendlein, A.; Kelch, S.

Shape-memory polymers

Angew. Chem. Int. Ed. 2002 41 2034 2057 10.1002/1521-3773(20020617)41:12<2034::AID-ANIE2034>3.0.CO;2-M

Lendlein, A.; Kelch, S. Shape-memory polymers. Angew. Chem. Int. Ed. 2002, 41, 2034−2057.

Baidu ScholarGoogle Scholar
4 Wang, Z.; Ma, Z.; Li, L.

Flow-induced crystallization of polymers: molecular and thermodynamic considerations

Macromolecules 2016 49 1505 1517 10.1021/acs.macromol.5b02688

Wang, Z.; Ma, Z.; Li, L. Flow-induced crystallization of polymers: molecular and thermodynamic considerations. Macromolecules 2016, 49, 1505−1517.

Baidu ScholarGoogle Scholar
5 Tang, X.; Chen, W.; Li, L.

The tough journey of polymer crystallization: battling with chain flexibility and connectivity

Macromolecules 2019 52 3575 3591 10.1021/acs.macromol.8b02725

Tang, X.; Chen, W.; Li, L. The tough journey of polymer crystallization: battling with chain flexibility and connectivity. Macromolecules 2019, 52, 3575−3591.

Baidu ScholarGoogle Scholar
6 Graham, R. S.

Understanding flow-induced crystallization in polymers: a perspective on the role of molecular simulations

J. Rheol. 2019 63 203 214 10.1122/1.5056170

Graham, R. S. Understanding flow-induced crystallization in polymers: a perspective on the role of molecular simulations. J. Rheol. 2019, 63, 203−214.

Baidu ScholarGoogle Scholar
7 Cui, K.; Liu, D.; Ji, Y.; Huang, N.; Ma, Z.; Wang, Z.; Lv, F.; Yang, H.; Li, L.

Nonequilibrium nature of flow-induced nucleation in isotactic polypropylene

Macromolecules 2015 48 694 699 10.1021/ma502412y

Cui, K.; Liu, D.; Ji, Y.; Huang, N.; Ma, Z.; Wang, Z.; Lv, F.; Yang, H.; Li, L. Nonequilibrium nature of flow-induced nucleation in isotactic polypropylene. Macromolecules 2015, 48, 694−699.

Baidu ScholarGoogle Scholar
8 Dukovski, I.; Muthukumar, M.

Langevin dynamics simulations of early stage shish-kebab crystallization of polymers in extensional flow

J. Chem. Phys. 2003 118 6648 6655 10.1063/1.1557473

Dukovski, I.; Muthukumar, M. Langevin dynamics simulations of early stage shish-kebab crystallization of polymers in extensional flow. J. Chem. Phys. 2003, 118, 6648−6655.

Baidu ScholarGoogle Scholar
9 Liu, D.; Tian, N.; Cui, K.; Zhou, W.; Li, X.; Li, L.

Correlation between flow-induced nucleation morphologies and strain in polyethylene: from uncorrelated oriented point-nuclei, scaffold-network, and microshish to shish

Macromolecules 2013 46 3435 3443 10.1021/ma400024m

Liu, D.; Tian, N.; Cui, K.; Zhou, W.; Li, X.; Li, L. Correlation between flow-induced nucleation morphologies and strain in polyethylene: from uncorrelated oriented point-nuclei, scaffold-network, and microshish to shish. Macromolecules 2013, 46, 3435−3443.

Baidu ScholarGoogle Scholar
10 Nicholson, D. A.; Rutledge, G. C.

Molecular simulation of flow-enhanced nucleation in n-eicosane melts under steady shear and uniaxial extension

J. Chem. Phys. 2016 145 244903 10.1063/1.4972894

Nicholson, D. A.; Rutledge, G. C. Molecular simulation of flow-enhanced nucleation in n-eicosane melts under steady shear and uniaxial extension. J. Chem. Phys. 2016, 145, 244903.

Baidu ScholarGoogle Scholar
11 Yamamoto, T.

Molecular dynamics simulation of stretch-induced crystallization in polyethylene: emergence of fiber structure and molecular network

Macromolecules 2019 52 1695 1706 10.1021/acs.macromol.8b02569

Yamamoto, T. Molecular dynamics simulation of stretch-induced crystallization in polyethylene: emergence of fiber structure and molecular network. Macromolecules 2019, 52, 1695−1706.

Baidu ScholarGoogle Scholar
12 Tang, X.; Yang, J.; Tian, F.; Xu, T.; Xie, C.; Chen, W.; Li, L.

Flow-induced density fluctuation assisted nucleation in polyethylene

J. Chem. Phys. 2018 149 224901 10.1063/1.5054273

Tang, X.; Yang, J.; Tian, F.; Xu, T.; Xie, C.; Chen, W.; Li, L. Flow-induced density fluctuation assisted nucleation in polyethylene. J. Chem. Phys. 2018, 149, 224901.

Baidu ScholarGoogle Scholar
13 Xie, C.; Tang, X.; Yang, J.; Xu, T.; Tian, F.; Li, L.

Stretch-induced coil–helix transition in isotactic polypropylene: a molecular dynamics simulation

Macromolecules 2018 51 3994 4002 10.1021/acs.macromol.8b00325

Xie, C.; Tang, X.; Yang, J.; Xu, T.; Tian, F.; Li, L. Stretch-induced coil–helix transition in isotactic polypropylene: a molecular dynamics simulation. Macromolecules 2018, 51, 3994−4002.

Baidu ScholarGoogle Scholar
14 Zhou, W.; Meng, L.; Lu, J.; Wang, Z.; Zhang, W.; Huang, N.; Chen, L.; Li, L.

Inducing uniform single-crystal like orientation in natural rubber with constrained uniaxial stretch

Soft Matter 2015 11 5044 5052 10.1039/C5SM00738K

Zhou, W.; Meng, L.; Lu, J.; Wang, Z.; Zhang, W.; Huang, N.; Chen, L.; Li, L. Inducing uniform single-crystal like orientation in natural rubber with constrained uniaxial stretch. Soft Matter 2015, 11, 5044−5052.

Baidu ScholarGoogle Scholar
15 Nie, Y.; Zhao, Y.; Matsuba, G.; Hu, W.

Shish-kebab crystallites initiated by shear fracture in bulk polymers

Macromolecules 2018 51 480 487 10.1021/acs.macromol.7b02357

Nie, Y.; Zhao, Y.; Matsuba, G.; Hu, W. Shish-kebab crystallites initiated by shear fracture in bulk polymers. Macromolecules 2018, 51, 480−487.

Baidu ScholarGoogle Scholar
16 Flory, P. J.

Thermodynamics of crystallization in high polymers. I. Crystallization induced by stretching

J. Chem. Phys. 1947 15 397 408 10.1063/1.1746537

Flory, P. J. Thermodynamics of crystallization in high polymers. I. Crystallization induced by stretching. J. Chem. Phys. 1947, 15, 397−408.

Baidu ScholarGoogle Scholar
17 Doufas, A. K.; Dairanieh, I. S.; McHugh, A. J.

A continuum model for flow-induced crystallization of polymer melts

J. Rheol. 1999 43 85 109 10.1122/1.550978

Doufas, A. K.; Dairanieh, I. S.; McHugh, A. J. A continuum model for flow-induced crystallization of polymer melts. J. Rheol. 1999, 43, 85−109.

Baidu ScholarGoogle Scholar
18 Coppola, S.; Grizzuti, N.; Maffettone, P. L.

Microrheological modeling of flow-induced crystallization

Macromolecules 2001 34 5030 5036 10.1021/ma010275e

Coppola, S.; Grizzuti, N.; Maffettone, P. L. Microrheological modeling of flow-induced crystallization. Macromolecules 2001, 34, 5030−5036.

Baidu ScholarGoogle Scholar
19 Read, D. J.; McIlroy, C.; Das, C.; Harlen, O. G.; Graham, R. S.

PolySTRAND model of flow-induced nucleation in polymers

Phys. Rev. Lett. 2020 124 147802 10.1103/PhysRevLett.124.147802

Read, D. J.; McIlroy, C.; Das, C.; Harlen, O. G.; Graham, R. S. PolySTRAND model of flow-induced nucleation in polymers. Phys. Rev. Lett. 2020, 124, 147802.

Baidu ScholarGoogle Scholar
20 Chen, X.; Meng, L.; Zhang, W.; Ye, K.; Xie, C.; Wang, D.; Chen, W.; Nan, M.; Wang, S.; Li, L.

Frustrating strain-induced crystallization of natural rubber with biaxial stretch

ACS Appl. Mater. Interfaces 2019 11 47535 47544 10.1021/acsami.9b15865

Chen, X.; Meng, L.; Zhang, W.; Ye, K.; Xie, C.; Wang, D.; Chen, W.; Nan, M.; Wang, S.; Li, L. Frustrating strain-induced crystallization of natural rubber with biaxial stretch. ACS Appl. Mater. Interfaces 2019, 11, 47535−47544.

Baidu ScholarGoogle Scholar
21 Nicholson, D. A.; Rutledge, G. C.

An assessment of models for flow-enhanced nucleation in an n-alkane melt by molecular simulation

J. Rheo. 2019 63 465

Nicholson, D. A.; Rutledge, G. C. An assessment of models for flow-enhanced nucleation in an n-alkane melt by molecular simulation. J. Rheo. 2019, 63, 465.

Baidu ScholarGoogle Scholar
22 Zhang, W.; Larson, R. G.

Effect of flow-induced nematic order on polyethylene crystal nucleation

Macromolecules 2020 53 7650 7657 10.1021/acs.macromol.0c01356

Zhang, W.; Larson, R. G. Effect of flow-induced nematic order on polyethylene crystal nucleation. Macromolecules 2020, 53, 7650−7657.

Baidu ScholarGoogle Scholar
23 Wang, Z.; Ju, J.; Yang, J.; Ma, Z.; Liu, D.; Cui, K.; Yang, H.; Chang, J.; Huang, N.; Li, L.

The non-equilibrium phase diagrams of flow-induced crystallization and melting of polyethylene

Sci. Rep. 2016 6 32968 10.1038/srep32968

Wang, Z.; Ju, J.; Yang, J.; Ma, Z.; Liu, D.; Cui, K.; Yang, H.; Chang, J.; Huang, N.; Li, L. The non-equilibrium phase diagrams of flow-induced crystallization and melting of polyethylene. Sci. Rep. 2016, 6, 32968.

Baidu ScholarGoogle Scholar
24 Liu, D.; Tian, N.; Huang, N.; Cui, K.; Wang, Z.; Hu, T.; Yang, H.; Li, X.; Li, L.

Extension-induced nucleation under near-equilibrium conditions: the mechanism on the transition from point nucleus to shish

Macromolecules 2014 47 6813 6823 10.1021/ma501482w

Liu, D.; Tian, N.; Huang, N.; Cui, K.; Wang, Z.; Hu, T.; Yang, H.; Li, X.; Li, L. Extension-induced nucleation under near-equilibrium conditions: the mechanism on the transition from point nucleus to shish. Macromolecules 2014, 47, 6813−6823.

Baidu ScholarGoogle Scholar
25 Balzano, L.; Kukalyekar, N.; Rastogi, S.; Peters, G. W.; Chadwick, J. C.

Crystallization and dissolution of flow-induced precursors

Phys. Rev. Lett. 2008 100 048302 10.1103/PhysRevLett.100.048302

Balzano, L.; Kukalyekar, N.; Rastogi, S.; Peters, G. W.; Chadwick, J. C. Crystallization and dissolution of flow-induced precursors. Phys. Rev. Lett. 2008, 100, 048302.

Baidu ScholarGoogle Scholar
26 Cui, K.; Meng, L.; Tian, N.; Zhou, W.; Liu, Y.; Wang, Z.; He, J.; Li, L.

Self-acceleration of nucleation and formation of shish in extension-induced crystallization with strain beyond fracture

Macromolecules 2012 45 5477 5486 10.1021/ma300338c

Cui, K.; Meng, L.; Tian, N.; Zhou, W.; Liu, Y.; Wang, Z.; He, J.; Li, L. Self-acceleration of nucleation and formation of shish in extension-induced crystallization with strain beyond fracture. Macromolecules 2012, 45, 5477−5486.

Baidu ScholarGoogle Scholar
27 Barham, P. J.; Keller, A.

High-strength polyethylene fibres from solution and gel spinning

J. Mater. Sci. 1985 20 2281 2302 10.1007/BF00556059

Barham, P. J.; Keller, A. High-strength polyethylene fibres from solution and gel spinning. J. Mater. Sci. 1985, 20, 2281−2302.

Baidu ScholarGoogle Scholar
28 Cui, K.; Ma, Z.; Wang, Z.; Ji, Y.; Liu, D.; Huang, N.; Chen, L.; Zhang, W.; Li, L.

Kinetic process of shish formation: from stretched network to stabilized nuclei

Macromolecules 2015 48 5276 5285 10.1021/acs.macromol.5b00819

Cui, K.; Ma, Z.; Wang, Z.; Ji, Y.; Liu, D.; Huang, N.; Chen, L.; Zhang, W.; Li, L. Kinetic process of shish formation: from stretched network to stabilized nuclei. Macromolecules 2015, 48, 5276−5285.

Baidu ScholarGoogle Scholar
29 Hu, W.; Frenkel, D.; Mathot, V.

Simulation of shish-kebab crystallite induced by a single prealigned macromolecule

Macromolecules 2002 35 7172 7174 10.1021/ma0255581

Hu, W.; Frenkel, D.; Mathot, V. Simulation of shish-kebab crystallite induced by a single prealigned macromolecule. Macromolecules 2002, 35, 7172−7174.

Baidu ScholarGoogle Scholar
30 Kanaya, T.; Matsuba, G.; Ogino, Y.; Nishida, K.; Shimizu, H. M.; Shinohara, T.; Oku, T.; Suzuki, J.; Otomo, T.

Hierarchic structure of shish-kebab by neutron scattering in a wide Q range

Macromolecules 2007 40 3650 3654 10.1021/ma062606z

Kanaya, T.; Matsuba, G.; Ogino, Y.; Nishida, K.; Shimizu, H. M.; Shinohara, T.; Oku, T.; Suzuki, J.; Otomo, T. Hierarchic structure of shish-kebab by neutron scattering in a wide Q range. Macromolecules 2007, 40, 3650−3654.

Baidu ScholarGoogle Scholar
31 Ma, Z.; Balzano, L.; Peters, G. W. M.

Dissolution and re-emergence of flow-induced shish in polyethylene with a broad molecular weight distribution

Macromolecules 2016 49 2724 2730 10.1021/acs.macromol.6b00333

Ma, Z.; Balzano, L.; Peters, G. W. M. Dissolution and re-emergence of flow-induced shish in polyethylene with a broad molecular weight distribution. Macromolecules 2016, 49, 2724−2730.

Baidu ScholarGoogle Scholar
32 Pennings, A. J.; Kiel, A. M.

Fractionation of polymers by crystallization from solution, III. On the morphology of fibrillar polyethylene crystals grown in solution

Kolloid-Zeitschrift und Zeitschrift für Polymere 1965 205 160 162

Pennings, A. J.; Kiel, A. M. Fractionation of polymers by crystallization from solution, III. On the morphology of fibrillar polyethylene crystals grown in solution. Kolloid-Zeitschrift und Zeitschrift für Polymere 1965, 205, 160−162.

Baidu ScholarGoogle Scholar
33 Pennings, A. J.; Lageveen, R.; de Vries, R. S.

Hydrodynamically induced crystallization of polymers from solution

Colloid Polym. Sci. 1977 255 532 542 10.1007/BF01549739

Pennings, A. J.; Lageveen, R.; de Vries, R. S. Hydrodynamically induced crystallization of polymers from solution. Colloid Polym. Sci. 1977, 255, 532−542.

Baidu ScholarGoogle Scholar
34 Muthukumar, M.

Nucleation in polymer crystallization

Adv. Chem. Phys. 2003 128 1 63 10.1002/0471484237.ch1

Muthukumar, M. Nucleation in polymer crystallization. Adv. Chem. Phys. 2003, 128, 1−63.

Baidu ScholarGoogle Scholar
35 Graham, R. S.; Olmsted, P. D.

Coarse-grained simulations of flow-induced nucleation in semicrystalline polymers

Phys. Rev. Lett. 2009 103 115702 10.1103/PhysRevLett.103.115702

Graham, R. S.; Olmsted, P. D. Coarse-grained simulations of flow-induced nucleation in semicrystalline polymers. Phys. Rev. Lett. 2009, 103, 115702.

Baidu ScholarGoogle Scholar
36 Hassan, M. K.; Cakmak, M.

Strain-induced crystallization during relaxation following biaxial stretching of PET films: a real-time mechano-optical study

Macromolecules 2015 48 4657 4668 10.1021/acs.macromol.5b00388

Hassan, M. K.; Cakmak, M. Strain-induced crystallization during relaxation following biaxial stretching of PET films: a real-time mechano-optical study. Macromolecules 2015, 48, 4657−4668.

Baidu ScholarGoogle Scholar
37 Wang, Z.; Su, F.; Ji, Y.; Yang, H.; Tian, N.; Chang, J.; Li, L.

Transition from chain- to crystal-network in extension induced crystallization of isotactic polypropylene

J. Chem. Phys. 2017 146 014901 10.1063/1.4973382

Wang, Z.; Su, F.; Ji, Y.; Yang, H.; Tian, N.; Chang, J.; Li, L. Transition from chain- to crystal-network in extension induced crystallization of isotactic polypropylene. J. Chem. Phys. 2017, 146, 014901.

Baidu ScholarGoogle Scholar
38 Nie, Y.; Gao, H.; Yu, M.; Hu, Z.; Reiter, G.; Hu, W.

Competition of crystal nucleation to fabricate the oriented semi-crystalline polymers

Polymer 2013 54 3402 3407 10.1016/j.polymer.2013.04.047

Nie, Y.; Gao, H.; Yu, M.; Hu, Z.; Reiter, G.; Hu, W. Competition of crystal nucleation to fabricate the oriented semi-crystalline polymers. Polymer 2013, 54, 3402−3407.

Baidu ScholarGoogle Scholar
39 Milner, S. T.

Polymer crystal–melt interfaces and nucleation in polyethylene

Soft Matter 2011 7 2909 2917 10.1039/c0sm00070a

Milner, S. T. Polymer crystal–melt interfaces and nucleation in polyethylene. Soft Matter 2011, 7, 2909−2917.

Baidu ScholarGoogle Scholar
40 Balzano, L.; Rastogi, S.; Peters, G. W. M.

Crystallization and precursors during fast short-term shear

Macromolecules 2009 42 2088 2092 10.1021/ma802169t

Balzano, L.; Rastogi, S.; Peters, G. W. M. Crystallization and precursors during fast short-term shear. Macromolecules 2009, 42, 2088−2092.

Baidu ScholarGoogle Scholar
41 Keller, A.; Hikosaka, M.; Rastogi, S.; Toda, A.; Barham, P. J.; Goldbeck-Wood, G.

An approach to the formation and growth of new phases with application to polymer crystallization: effect of finite size, metastability, and Ostwald's rule of stages

J. Mater. Sci. 1994 29 2579 2604 10.1007/BF00356806

Keller, A.; Hikosaka, M.; Rastogi, S.; Toda, A.; Barham, P. J.; Goldbeck-Wood, G. An approach to the formation and growth of new phases with application to polymer crystallization: effect of finite size, metastability, and Ostwald's rule of stages. J. Mater. Sci. 1994, 29, 2579−2604.

Baidu ScholarGoogle Scholar
42 Kwon, Y. K.; Boller, A.; Pyda, M.; Wunderlich, B.

Melting and heat capacity of gel-spun, ultra-high molar mass polyethylene fibers

Polymer 2000 41 6237 6249 10.1016/S0032-3861(99)00839-3

Kwon, Y. K.; Boller, A.; Pyda, M.; Wunderlich, B. Melting and heat capacity of gel-spun, ultra-high molar mass polyethylene fibers. Polymer 2000, 41, 6237−6249.

Baidu ScholarGoogle Scholar
43 Rastogi, S.; Kurelec, L.; Lemstra, P. J.

Chain mobility in polymer systems: on the borderline between solid and melt. 2. Crystal size influence in phase transition and sintering of ultrahigh molecular weight polyethylene via the mobile hexagonal phase

Macromolecules 1998 31 5022 5031 10.1021/ma980261h

Rastogi, S.; Kurelec, L.; Lemstra, P. J. Chain mobility in polymer systems: on the borderline between solid and melt. 2. Crystal size influence in phase transition and sintering of ultrahigh molecular weight polyethylene via the mobile hexagonal phase. Macromolecules 1998, 31, 5022−5031.

Baidu ScholarGoogle Scholar
44 Krigbaum, W. R.; Roe, R. J.

Diffraction study of crystallite orientation in a stretched polychloroprene vulcanizate

 J. Polym. Sci., Part A: Genr. Papers 1964 2 4391 4414 10.1002/pol.1964.100021010

Krigbaum, W. R.; Roe, R. J. Diffraction study of crystallite orientation in a stretched polychloroprene vulcanizate. J. Polym. Sci., Part A: Genr. Papers 1964, 2, 4391−4414.

Baidu ScholarGoogle Scholar
45 Yeh, G. S. Y.; Hong, K. Z.

Strain-induced crystallization, Part III: Theory

Polym. Eng. Sci. 1979 19 395 400 10.1002/pen.760190605

Yeh, G. S. Y.; Hong, K. Z. Strain-induced crystallization, Part III: Theory. Polym. Eng. Sci. 1979, 19, 395−400.

Baidu ScholarGoogle Scholar
46 Yeh, G. S. Y.; Hong, K. Z.; Krueger, D. L.

Strain-induced crystallization, Part IV: Induction time analysis

Polym. Eng. Sci. 1979 19 401 405 10.1002/pen.760190606

Yeh, G. S. Y.; Hong, K. Z.; Krueger, D. L. Strain-induced crystallization, Part IV: Induction time analysis. Polym. Eng. Sci. 1979, 19, 401−405.

Baidu ScholarGoogle Scholar
47 Maier, W.; Saupe, A.

Eine einfache molekular-statistische Theorie der nematischen kristallinflüssigen Phase. Teil II

Zeitschrift Naturforschung Teil A 1959 14 882 10.1515/zna-1959-1005

Maier, W.; Saupe, A. Eine einfache molekular-statistische Theorie der nematischen kristallinflüssigen Phase. Teil II. Zeitschrift Naturforschung Teil A 1959, 14, 882.

Baidu ScholarGoogle Scholar