Kazan Federal University, 420008 Kazan, Russia · 2018-11-09 · changes of the systems may do...

12
Morphology of Critically-Sized Crystalline Nuclei at Shear-Induced Crystal Nucleation in Amorphous Solid Bulat N. Galimzyanov * and Anatolii V. Mokshin Kazan Federal University, 420008 Kazan, Russia (Dated: November 9, 2018) In this work we study morphological characteristics of the critically-sized crystalline nuclei at initial stage of the shear-induced crystallization of a model single-component amorphous (glassy) system. These characteristics are estimated quantitatively through statistical treatment of the non- equilibrium molecular dynamics simulation results for the system under steady shear at various (fixed) values of the shear rate ˙ γ and at different temperatures. It is found that the sheared glassy system is crystallized through nucleation mechanism. From analysis of a time-dependent trajectories of the largest crystalline nuclei, the critical size nc and the nucleation time τc were defined. It is shown that the critically-sized nuclei in the system are oriented within the shear-gradient xy-plane at moderate and high shear rates; and a tilt angle of the oriented nuclei depends on the shear rate. At extremely high shear rates and at shear deformation of the system more than 60%, the tilt angle of the nuclei tends to take the value 45 respective to the shear direction. We found that this feature depends weakly on the temperature. Asphericity of the nucleus shape increases with increasing shear rate, that is verified by increasing value of the asphericity parameter and by the contour of the pair distribution function calculated for the particles of the critically-sized nuclei. The critical size increases with increasing shear rate according to the power-law, nc γτc) 1/3 , whereas the shape of the critically-sized nucleus changes from spherical to the elongated ellipsoidal. We found that the nc-dependencies of the nuclei deformation parameter evaluated for the system at different temperatures and shear rates are collapsed into unified master-curve. I. INTRODUCTION Due to disordered structure, the amorphous materi- als (metallic glasses, polymeric and colloidal amorphous solids) have unique mechanical and physical properties, which allow one to find their practical applications in photonics, medicine and electronics [1–7]. Furthermore, the microscopic structure of disordered systems under different mechanical deformations (compression, shear etc.) is of a special interest [8]. In particular, understand- ing of physical mechanisms of the steady shear influence on microscopic structure of the systems could provide a possibility to develop practical tools to control the struc- tural ordering process [2, 7, 9–14]. A large number of experimental and simulation stud- ies provide indications that the steady shear applied to metastable disordered systems (supercooled liquids and amorphous solids) generates a microscopic structure anisotropy in these systems [7, 13–22]. It is clear that the anisotropy arisen in the disordered systems under shear leads to change of morphological characteristics of emerging crystalline structures. Therefore, the mechani- cal and rheological properties of the systems can be sig- nificantly dependent on the anisotropy. This is confirmed by results of large number of studies (Refs. [1, 7, 14, 23– 32]). So, for example, Kumaraswamy and co-workers [22] have revealed experimentally that non-spherical ordered structures oriented along the shear direction occur for the case of polydisperse isotactic polypropylene melt un- der intensive shear. Moreover, they found that these * [email protected] structures appear as precursors of crystallization at a temperature below the melting point; and morphological properties of these precursors have a significant impact on the overall crystallization kinetics. On the basis of experimental and simulation studies done by Petekidis and co-workers [21, 32, 33], it was found that the mi- croscopic structure of polymeric colloidal glasses under shear changes with increasing shear rate. This is directly manifested in the shape of the evaluated density distribu- tion g(x, y). Namely, the function g(x, y) characterizing the particles distribution within the shear-gradient plane starts to change its shape under applied shear flow. Then, it is reasonably to expect that the observed structural changes of the systems may do impact on their rheologi- cal properties. It was found in Ref. [33] that the ordered structures forming under shear are not stable thermody- namically, and they may be characterized by rheologi- cal aging. These results are supplemented by simulation studies. For example, simulation results of Blaak et al. have revealed that the elongated crystalline nuclei ori- ented along the shear direction occur at shear-induced crystallization of colloidal suspensions [16]. Here, a tilt angle of the nucleus increases with increasing shear rate. On the other hand, Graham and Olmsted found that non-spherical elongated nuclei occur in sheared polymer melts, and such the shape has been mentioned as shish- kebab [26]. According to results of Ref. [26], the crystal nucleation may be accelerated by shear. Note, the similar results were also obtained for the case of crystallization of model glassy systems at homogeneous shear [7, 15, 34]. It was found in Refs. [7, 15, 34] that steady shear can both accelerate and suppress crystal nucleation process in the glassy systems. arXiv:1711.04989v1 [cond-mat.mtrl-sci] 14 Nov 2017

Transcript of Kazan Federal University, 420008 Kazan, Russia · 2018-11-09 · changes of the systems may do...

Page 1: Kazan Federal University, 420008 Kazan, Russia · 2018-11-09 · changes of the systems may do impact on their rheologi-cal properties. It was found in Ref. [33] that the ordered

Morphology of Critically-Sized Crystalline Nucleiat Shear-Induced Crystal Nucleation in Amorphous Solid

Bulat N. Galimzyanov∗ and Anatolii V. MokshinKazan Federal University, 420008 Kazan, Russia

(Dated: November 9, 2018)

In this work we study morphological characteristics of the critically-sized crystalline nuclei atinitial stage of the shear-induced crystallization of a model single-component amorphous (glassy)system. These characteristics are estimated quantitatively through statistical treatment of the non-equilibrium molecular dynamics simulation results for the system under steady shear at various(fixed) values of the shear rate γ and at different temperatures. It is found that the sheared glassysystem is crystallized through nucleation mechanism. From analysis of a time-dependent trajectoriesof the largest crystalline nuclei, the critical size nc and the nucleation time τc were defined. It isshown that the critically-sized nuclei in the system are oriented within the shear-gradient xy-planeat moderate and high shear rates; and a tilt angle of the oriented nuclei depends on the shear rate.At extremely high shear rates and at shear deformation of the system more than 60%, the tiltangle of the nuclei tends to take the value ' 45◦ respective to the shear direction. We found thatthis feature depends weakly on the temperature. Asphericity of the nucleus shape increases withincreasing shear rate, that is verified by increasing value of the asphericity parameter and by thecontour of the pair distribution function calculated for the particles of the critically-sized nuclei.The critical size increases with increasing shear rate according to the power-law, nc ∝ (γτc)

1/3,whereas the shape of the critically-sized nucleus changes from spherical to the elongated ellipsoidal.We found that the nc-dependencies of the nuclei deformation parameter evaluated for the systemat different temperatures and shear rates are collapsed into unified master-curve.

I. INTRODUCTION

Due to disordered structure, the amorphous materi-als (metallic glasses, polymeric and colloidal amorphoussolids) have unique mechanical and physical properties,which allow one to find their practical applications inphotonics, medicine and electronics [1–7]. Furthermore,the microscopic structure of disordered systems underdifferent mechanical deformations (compression, shearetc.) is of a special interest [8]. In particular, understand-ing of physical mechanisms of the steady shear influenceon microscopic structure of the systems could provide apossibility to develop practical tools to control the struc-tural ordering process [2, 7, 9–14].

A large number of experimental and simulation stud-ies provide indications that the steady shear appliedto metastable disordered systems (supercooled liquidsand amorphous solids) generates a microscopic structureanisotropy in these systems [7, 13–22]. It is clear thatthe anisotropy arisen in the disordered systems undershear leads to change of morphological characteristics ofemerging crystalline structures. Therefore, the mechani-cal and rheological properties of the systems can be sig-nificantly dependent on the anisotropy. This is confirmedby results of large number of studies (Refs. [1, 7, 14, 23–32]). So, for example, Kumaraswamy and co-workers [22]have revealed experimentally that non-spherical orderedstructures oriented along the shear direction occur forthe case of polydisperse isotactic polypropylene melt un-der intensive shear. Moreover, they found that these

[email protected]

structures appear as precursors of crystallization at atemperature below the melting point; and morphologicalproperties of these precursors have a significant impacton the overall crystallization kinetics. On the basis ofexperimental and simulation studies done by Petekidisand co-workers [21, 32, 33], it was found that the mi-croscopic structure of polymeric colloidal glasses undershear changes with increasing shear rate. This is directlymanifested in the shape of the evaluated density distribu-tion g(x, y). Namely, the function g(x, y) characterizingthe particles distribution within the shear-gradient planestarts to change its shape under applied shear flow. Then,it is reasonably to expect that the observed structuralchanges of the systems may do impact on their rheologi-cal properties. It was found in Ref. [33] that the orderedstructures forming under shear are not stable thermody-namically, and they may be characterized by rheologi-cal aging. These results are supplemented by simulationstudies. For example, simulation results of Blaak et al.have revealed that the elongated crystalline nuclei ori-ented along the shear direction occur at shear-inducedcrystallization of colloidal suspensions [16]. Here, a tiltangle of the nucleus increases with increasing shear rate.On the other hand, Graham and Olmsted found thatnon-spherical elongated nuclei occur in sheared polymermelts, and such the shape has been mentioned as shish-kebab [26]. According to results of Ref. [26], the crystalnucleation may be accelerated by shear. Note, the similarresults were also obtained for the case of crystallization ofmodel glassy systems at homogeneous shear [7, 15, 34].It was found in Refs. [7, 15, 34] that steady shear canboth accelerate and suppress crystal nucleation processin the glassy systems.

arX

iv:1

711.

0498

9v1

[co

nd-m

at.m

trl-

sci]

14

Nov

201

7

Page 2: Kazan Federal University, 420008 Kazan, Russia · 2018-11-09 · changes of the systems may do impact on their rheologi-cal properties. It was found in Ref. [33] that the ordered

2

In spite of large number of studies mentioned above,there are still a lot of unclear points related with in-fluence of a steady homogeneous shear on the mor-phology of crystalline structures emerging in amorphoussolids [15, 18, 35, 36]. For example, the physical mecha-nisms determining the crystalline nuclei asphericity atthe initial stage of crystallization, when the size ofthese nuclei is comparable with a critical size, are de-bated [7, 15, 16, 26]. According to the classical nucle-ation theory [37–39], only the critically-sized nuclei maydemonstrate a stable growth. Moreover, from experi-mental point of view it is not so easy to evaluate suchmorphological characteristics as the critical size and theshape of the crystalline nuclei in the atomistic and molec-ular systems under shear with different rates [13, 33, 40].This is especially appropriate for the case of deep super-cooling levels, when the temperatures are below the glasstransition temperature Tg.

In the present work, we study the microscopic mech-anisms of shear-induced crystallization of a single-component glassy system at different temperatures. Themain attention is paid to study morphological charac-teristics of the critically-sized crystalline nuclei in thesystem under a homogeneous shear with various (fixed)shear rates. These characteristics are considered as theterms dependent on the critical deformation γc ≡ γτc.Here, the dimensionless quantity γc is the deformation ofthe system under shear with the rate γ at the time mo-ment τc ≡ τ(nc), i.e. when the critically-sized nc nucleusappears. In other words, the quantity γc corresponds di-rectly in the time scale to the nucleation time τc. We findthat the homogeneous shear initiates the structural or-dering in the glassy system through the crystal nucleationmechanism, where the nuclei with pronounced aspheric-ity of their shape and oriented along the shear directionare formed. Simulation details and cluster analysis arepresented in Section II. The results are given and dis-cussed in Section III. Finally, the main results will besummarized in the conclusion.

II. SIMULATION DETAILS AND CLUSTERANALYSIS

Molecular dynamics simulations are performed for thesingle-component system, where the interaction betweenthe particles is defined through the short-ranged oscilla-tory Dzugutov potential (Dz-system) [41, 42]. The spe-cific shape of the potential allows one to reproduce ef-fectively the effect of electron screening for the ion-ioninteraction in metals. In addition, the system with suchthe potential is capable to generate a stable amorphousstate [7, 34, 43].

To realize the homogeneous shear, the SLLOD-algorithm is applied [12], where the shear velocity

~ui = γyi~ex (1)

is added to the x-component of intrinsic (thermal) veloc-

ity ~υi of each particle of the system. As a result, equa-tions of motion are [12]:

d~ridt

= ~υi + γyi~ex + ζ~ri, (2a)

d~υidt

=~Fim− γυyi~ex − (%+ ζ)~υi. (2b)

Here, ~ri(xi, yi, zi) and ~υi(υxi, υyi, υzi) are the positionand velocity of the ith particle (i = 1, 2, ... N , whereN is the number of particles); % and ζ are the param-eters of thermostat and barostat, respectively. In thepresent work, the homogeneous shear is applied with dif-ferent fixed shear rates: γ = 0.0001, 0.0005, 0.001, 0.002,0.005, 0.008, and 0.01 τ−1. Realization of the homoge-neous shear is schematically presented in Fig. 1.

Figure 1. (color online) Schematic representation of the ho-mogeneous shear realization with a linear velocity profile:green spheres show the particles of a system; the blue arrowscharacterize the shear directions with rate ~ui (i = 1, 2, ..., N).The axis OX is associated with the shear-direction; the axisOY corresponds to the gradient-direction; the axis OZ indi-cates the vorticity-direction.

The considered three-dimensional system consists ofN = 6912 identical particles located within the cubicsimulation cell. The Lees-Edwards periodic boundaryconditions are applied in both the gradient y- and vor-ticity z-directions [44]. The ordinary periodic boundaryconditions are also applied in the shear x-direction. Theinstantaneous velocities and trajectories of the particlesare determined by integration of the equations of mo-tion (2) using the Verlet algorithm with the time step∆t = 0.005 τ . All simulations are performed in theisobaric-isothermal ensemble, where the temperature Tand the pressure p are controlled by the Nose-Hooverthermostat and barostat, respectively. In the presentwork, the standard Lennard-Jones units are used: σ

Page 3: Kazan Federal University, 420008 Kazan, Russia · 2018-11-09 · changes of the systems may do impact on their rheologi-cal properties. It was found in Ref. [33] that the ordered

3

is the effective particle diameter; ε is the energy unit;τ = σ

√m/ε is the time unit, where m is the mass of the

particle; the temperature T in units of ε/kB , where kB isthe Boltzmann constant; the pressure p in units of ε/σ3;the shear rate γ in units of τ−1.

Molecular dynamics simulations are realized as follows.At first, a liquid sample is equilibrated at the temper-ature T = 2.3 ε/kB and at the pressure p = 14 ε/σ3,that is above the melting temperature of the systemTm ' 1.51 ε/kB at considered pressure (the phase dia-gram of the Dz-system can be found in Ref. [42]). Sec-ond, the glassy samples are prepared through rapid cool-ing of the equilibrated liquid with the cooling rate of0.001 ε/(kBτ) to the temperatures T = 0.05, 0.15, and0.5 ε/kB , that is much lower than the glass transitiontemperature Tg ' 0.65 ε/kB of the system. Thus, onehundred independent glassy samples are prepared foreach considered (p, T ) thermodynamic state, that arerequired to perform a statistical treatment of the simula-tion results. And, finally, each glassy sample is exposedto a homogeneous shear on the time scale ∼ 100 000 sim-ulation steps.

To detect the crystalline structures, the cluster anal-ysis based on the estimation of the bond orientationalorder parameters was applied [45, 46]. Namely, follow-ing to Steinhardt et al. [45], the local orientational orderparameters are defined as

ql(i) =

2l + 1

l∑m=−l

∣∣∣∣∣∣ 1

n(i)b

n(i)b∑j=1

Ylm(θij , ϕij)

∣∣∣∣∣∣2

1/2

, (3)

where

l = 4, 6, 8.

Also we estimate value of the global orientational orderparameter as an average of q6(i) (i.e. at l = 6) over allthe particles of the system [45]:

Q6 =

13

6∑m=−6

∣∣∣∣∣∣∑Ni=1

∑n(i)bj=1 Y6m(θij , ϕij)∑Ni=1 n

(i)b

∣∣∣∣∣∣2

1/2

. (4)

Here, n(i)b is the number of neighbors of the ith parti-

cle; Ylm(θij , ϕij) are the spherical harmonics; θij and ϕijare the polar and azimuthal angles, respectively. As arule, even spherical harmonics with the indexes l = 4,6 and 8 are sufficient to be evaluated recognize presenceof the ordered domains in the system. For example, thelocal orientational order parameters q4, q6, q8 take non-zero values for the particles forming sc (simple cubic),icos (icosahedral), bcc (base-centered cubic), fcc (face-centered cubic), and hcp (hexagonal close-packed) crys-talline structures [47]. For a disordered glassy system,the local orientational order parameters obey a normaldistribution, whereas the value of the parameter Q6 tends

to zero for this case [45, 48]. The order parameters qiand Qj , where i, j = 4, 6, 8, ..., take unique values forspecific crystalline structures, that allows one to identifythese structures in considered system [45]. In particular,according to original normalization of the parameter Q6,the parameter Q6 is 0.575 for the perfect fcc lattice; onehas Q6 = 0.485 for the perfect hcp lattice and Q6 = 0.663for the icosahedral lattice. The smaller values can be dueto the presents of defects.

One of the key conditions, by which the particle isidentified as entering into an ordered phase, is the so-called coherence condition [48]∣∣∣∣∣

6∑m=−6

~q∗6m(i) · ~q6m(j)

∣∣∣∣∣ ≥ 0.5, (5a)

~q6m(i) =q6m(i)√∑6

m=−6 |q6m(i)|2. (5b)

Namely, according to the scheme suggested in Ref. [48],the pair of particles i and j are considered as “solid-like”if the dot-product ~q∗6(i)·~q6(j) exceeds the threshold value0.5. In this case, the nearest neighborhood of the ith par-ticle must contain seven and more “solid-like” particles,for which condition (5) is satisfied.

For quantitative characterization of the crystalline nu-clei shape, we evaluate the asphericity parameter [46]

S0 =

⟨(Ixx − Iyy)2 + (Ixx − Izz)2 + (Iyy − Izz)2

2(Ixx + Iyy + Izz)2

⟩,

(6)where

Iαβ = m

nc∑i=1

(~r 2i δαβ − ~riα~riβ) (7)

are the moments of inertia (αβ ∈ {x, y, z}). Here, |~ri| isthe distance between the nucleus center-of-mass and theposition of ith “solid-like” particle; δαβ is the Kroneckerdelta. Thus, for the case of a perfect spherical shape,we have S0 = 0; the parameter is S0 → 1 for extremelyelongated shape.

By means of the cluster analysis, we obtain for eachαth simulation run the growth trajectories of the largestcrystalline nucleus, nα(t). Here the quantity nα(t) indi-cates that the nucleus of the size n appears at the timet during the αth simulation run, where α = 1, 2, ..., 100.These trajectories extracted from the different simula-tion runs are treated within the mean-first-passage-timemethod [36, 49]. According to this method, the curveτ(n) is defined, which is known as the mean-first-passage-time curve and which characterizes the average time ofthe first appearance of the largest nucleus with given sizen (for details, see Refs. [36, 49]). The critical size nc andthe average nucleation (or waiting) time for the nucleusτc, are defined from the analysis of the curve τ(n) andof the first derivative ∂τ(n)/∂n, according to the scheme

Page 4: Kazan Federal University, 420008 Kazan, Russia · 2018-11-09 · changes of the systems may do impact on their rheologi-cal properties. It was found in Ref. [33] that the ordered

4

suggested in Ref. [36]. In the case of activation type pro-cesses, the curve τ(n) characterizes by three regimes: (i)- the first regime is associated with pre-nucleation, wherethe small values of n correspond to τ(n) with zero value;(ii) - the second regime, in which the curve τ(n) has thepronounced non-zero slope, contains information about anucleation event. Namely, detected from the first deriva-tive ∂τ(n)/∂n location of an inflection point in the curveτ(n) for this regime defines the critical size nc, whereasτ(nc) ≡ τc is directly associated with the nucleation timeof the critically-sized nucleus; (iii) - the third regime,where the slope of τ(n) decreases corresponds to growthof the nucleus. Note that in the present work we focuson the characteristics for the largest crystalline nucleus.

III. RESULTS

A. Crystalline structures

Information about structural transformations in theglassy system at homogeneous shear is obtained throughevaluation of the orientational order parameters q4, q6,q8 and Q6. The most probable values of the order pa-rameters have been computed by averaging of the dataobtained from one hundred independent numerical exper-iments at different shear rates γ. As an example, Fig. 2ashows the time-dependent order parameter Q6 computedfor the system at the temperature T = 0.15 ε/kB and atthe shear rate of γ = 0.001 τ−1. Initially, the order pa-rameter Q6 takes the small value ∼ 0.023, which corre-sponds to a disordered system. Further, the glassy sys-tem is crystallized after initiation of the shear deforma-tion. Namely, the parameter Q6 begins to increase afterlag time ∼ 70 τ and at t > 250 τ the parameter goes tosaturation with value Q6 ' 0.47, which corresponds to acompletely ordered system. The sigmoidal shape of theQ6(t)-curve is typical for the activation processes when,for example, the crystallization occurs through the nucle-ation mechanism [15, 50]. A similar scenario of structuralordering is also observed at other shear rates and temper-atures with a difference only in the crystallization timescales.

Characteristic directions of the crystalline lattice wereinitially estimated for the critically-sized nucleus. As wefound, there is no alignment of crystal structure of thenuclei along the shear- and/or gradient-directions. More-over, the correspondence between orientation of the crys-tal planes and the specific shear directions was also notdetected even for the completely ordered systems, wherethe order parameter Q6 takes its largest value. This isseen from snapshots of the system at T = 0.15 ε/kB andshear rate 0.001 τ−1 [see Fig. 2(b)].

Fig. 3 shows the distributions P (q4), P (q6), and P (q8)for the system at the temperature T = 0.15 ε/kB and atthe shear rate γ = 0.001 τ−1, when the system is com-pletely crystallized (i.e. when the system correspondsto the state with Q6 ' 0.47). The normal distribu-

Figure 2. (color online) (a) Time-dependent orientational or-der parameter Q6(t) evaluated for the system at the temper-ature T = 0.15 ε/kB and at shear rate γ = 0.001 τ−1. In-sets: instantaneous configurations of a system correspondingto values Q6 ' 0.023 (left inset) and Q6 ' 0.47 (right inset).(b) Snapshots of the simulation cell corresponding to the sys-tem at the temperature T = 0.15 ε/kB and the shear rateγ = 0.001 τ−1, when an ordered state is achieved. Here, theorder parameter Q6 is ∼ 0.47. In fact, the system representsa nanocrystalline solid consisted of an assembly of crystalliteswith disordered orientations, and the degree of order achievedby the system is low in comparison to that of a perfect crystal.

tions P (q4), P (q6), and P (q8) corresponded to the glassysystem before initiation of the homogeneous shear arealso presented in Fig. 3 (here, the order parameter isQ6 ' 0.023). After initiation of the shear, the struc-tural ordering degree increases. This is also manifestedby a change of shapes of P (q4), P (q6), P (q8) and byappearance of peaks located at high values of the pa-rameters q4, q6, and q8. The locations of these peaks inthe distributions P (q4), P (q6), and P (q8) correspond to

Page 5: Kazan Federal University, 420008 Kazan, Russia · 2018-11-09 · changes of the systems may do impact on their rheologi-cal properties. It was found in Ref. [33] that the ordered

5

structures with the fcc and hcp-symmetries. The frac-tion of “fcc-particles” is ∼ 60-70 %, whereas the fractionof “hcp-particles” is ∼ 30-40 %. Moreover, locations ofthe peaks in the distributions P (q4), P (q6), and P (q8)corresponding to q4 ' 0.21, q6 ' 0.57, and q8 ' 0.41indicate on the fcc-symmetry. The peaks in the distri-butions at q4 ' 0.11, q6 ' 0.48, and q8 ' 0.31 corre-spond to the particles forming the hcp lattice [47]. Inaccordance to with the equilibrium phase diagram of theDzugutov system, the temperature region along the iso-bar p = 14 ε/σ3 includes the crystalline phases with thefcc and hcp lattices [41, 42]. In particular, at defor-mation of an amorphous system by shock waves, it wasobserved in Refs. [42, 51] a phase transition into thehigh-pressure states with fcc or hcp crystalline phase.Both the lattice types appear simultaneously and have aslight difference due to defects in the crystalline structurecaused by the shock waves.

Figure 3. (color online) Distributions P (q4), P (q6), and P (q8)evaluated for the Dz-system at the temperature T = 0.15 ε/kBwith Q6 ' 0.47 crystallized at the shear rate γ = 0.001 τ−1.

B. Critically-sized crystalline nuclei under shear

Results of the cluster analysis reveal that the struc-tural ordering in the glassy Dz-system occurs throughthe formation of small-sized crystalline nuclei consistingmainly of 30-50 particles. By the mean-first-passage-timemethod, the critical size nc and the nucleation time τchave been evaluated for the system at different temper-atures and shear rates [15, 49]. Fig. 4(a) shows the nu-cleation time τc as function of the critical deformation γcevaluated at different temperatures. We found differentregimes in γc-dependencies of τc. Namely, the nucleationtime decreases at small and moderate shear rates and itincreases at high shear rates. The nucleation time de-creases from τc ' 125 ± 15 τ to τc ' 40 ± 6 τ with in-creasing shear rate from γ = 0.0001 τ−1 (γc ' 0.0125)

to γ = 0.002 τ−1 (γc ' 0.08). The nucleation time in-creases from τc ' 41 ± 5 τ to τc ' 53 ± 10 τ at in-creasing shear rate from γ = 0.005 τ−1 (γc ' 0.205)to γ = 0.01 τ−1 (γc ' 0.53) in the considered temper-ature range T ∈ [0.05; 0.5] ε/kB . Such non-monotonicγc-dependence of the nucleation time τc is due to antag-onistic impact of shear flow on the nucleation process.This effect was also discussed in Refs. [7, 15]. As dis-cussed in Ref. [15], such non-monotonic behavior of τc isdue to impact of shear deformation on the kinetic andthermodynamic aspects of the crystal nucleation in thissystem. Namely, a slow shear-flow accelerates the nu-cleation through the attachment rate, whereas the highshear rates destabilize the critical nuclei and reduce theprobability of the particle attachment. On the otherhand, a shear flow gives rise to “pressure anisotropy” [52]manifested in the fact that the diagonal components ofthe pressure tensor begin to differ. This effect leads toanisotropy of the interfacial free energy. Moreover, theshear deformation introduces an elastic energy into thesystem and, thereby, it has an impact on the thermo-dynamic characteristics of crystal nucleation (the nucle-ation barrier, the interfacial free energy, etc.) [19, 53].

Although there are known experimental data charac-terizing the crystal nucleation in amorphous systems un-der shear, comparison of different experimental data andresults of numerical simulation is a non-trivial task. Infact, these systems are characterized by various physi-cal, chemical and rheological properties, and the systemsare studied at different thermodynamic and deformationconditions. Nevertheless, there are common regularitiesin structural ordering in these systems. Namely, the slowshear deformations accelerate the ordering, whereas thelarge shear rates suppress the crystallization. As a re-sult, the strain-dependencies of the crystallization (nucle-ation) time must be characterized by a minimum. There-fore, it is reasonable to propose that such the dependen-cies can be represented in the unified scaled form:

τcτm

= f

[(γcγm

)ξ](8)

Here, f [. . .] is a universal function, τm is the location ofthe minimum in the dependence τc(γc), whereas γm isthe value of the critical deformation γc at the time τm.Further, the dimensionless positive parameter ξ can beconsidered as the characteristic of glass-forming abilityof the system [36, 54, 55].

In Fig. 4(b), the scaled time τc/τm is presented as func-tion of the scaled critical deformation (γc/γm)ξ for differ-ent systems under shear: our simulation results for theDzugutov system (Dz); experimental data for colloidalsystem (CS) [56] and two polymer systems (PS) [57] and(sP) [58]. For the case of CS and PS systems, the datafor the induction time τind vs. the critical deformation γcwere taken. As seen from Fig. 4(b), all the data collapseonto unified curve. This is evidence that the scaling re-lation (8) is valid. Here, the parameter ξ takes the next

Page 6: Kazan Federal University, 420008 Kazan, Russia · 2018-11-09 · changes of the systems may do impact on their rheologi-cal properties. It was found in Ref. [33] that the ordered

6

Figure 4. (color online) (a) Nucleation time τc as function of critical deformation γc evaluated for Dz-system at differenttemperatures. (b) Scaled nucleation time τc/τm and induction time τind/τm as functions of scaled critical deformation (γc/γm)ξ.Here, τm and γm are the parameters defined the minimum of the dependence τc(γc) [and τind(γc)]. Symbols (◦), (2), and (4)indicate simulation results for the Dzugutov system at different temperatures, whereas symbols (5), (/), and (�) correspondto experimental data for colloidal system (CS) [56] and for two polymer systems: (PS) [57] and (sP) [58]. (c) Critical size ncvs. γc. Dashed curves are results of fit by Eq. (15).

values: ξ = 0.935 for the Dzugutov system, 1.282 for theCS-system, 2.33 for the PS-system and 1.89 for the sP-system. Thus, the parameter takes the lowest value forthe atomistic single-component Dzugutov system, and ittakes the larger values for the polymer and colloidal sys-tems with better glass-forming ability properties.

It should be noted that the crystal nucleation in amor-phous solid (glass) differs from crystal nucleation in su-percooled liquid. At low and moderate levels of super-cooling corresponding to a liquid (supercooled) phase,the nucleation rate is determined mainly by the thermo-dynamic factor [15, 36, 37]. For a glassy phase with deeplevels of supercooling, the nucleation is driven by the ki-netic features of the system. Competition between thethermodynamic and kinetic aspects is clearly manifestedin appearance of the characteristic maximum in the tem-perature dependence of the nucleation rate [37–39]. In-fluence of the shear on crystallization of the supercooledliquid and glass is also different. In particular, for thecase of a liquid, the shear deformation increases the vis-cosity and, thereby, nucleation is suppressed [19, 59]. Fora glassy system, the shear increases effectively the mo-bility of particles and, thereby, the shear decreases theviscosity.

The system evolves at temperatures much below theglass transition temperature Tg. And, therefore, detec-tion of the nucleation event for the system at zerothshear may seem surprising. Actually, features of the mi-croscopic kinetics of a glass changes with moving overphase diagram for the range of high pressures. Namely,at high pressures the structural relaxation as well as thetransition of glassy system into a state with the lowerfree energy proceeds over shorter time scales [60]. As wefound, the nucleation event are detectable within simula-tion time scales for the Dz-system at the thermodynamic

states with pressures p ≥ 12 ε/σ3 [36].In Fig. 4(c), the quantity nc as function of the criti-

cal deformation γc evaluated at different temperatures ispresented. At the absence of shear, for the consideredtemperature range T ∈ [0.05; 0.5] ε/kB , the critical sizeincreases with temperature by several tens of particles.This is in agreement with classical nucleation theory [37,38], according to which nc = (32πγ3∞)/(3ρc|∆µ|3). Here,γ∞ is the interfacial free energy, |∆µ| is the differenceof the chemical potential per phase unite in the melt(glass) and the crystal. Then, taking into account thatthe chemical potential difference |∆µ| is proportional thesupercooling, ∆T = Tm − T :

|∆µ| ∼ ∆T

Tm,

one has that the critical size nc increases with increasingtemperature. Further, with increase of shear rate withinthe range γ ∈ [0.0001; 0.01] τ−1, the critical size increasesfrom nc ' 88 to nc ' 130 particles. As seen fromFig. 4(c), at the low shear rates γ < 0.001 τ−1, the nucleiappear at critical deformation γc = 0.02, whereas at highshear rates formation of the nuclei occurs at the largerdeformations. Namely, at the shear rate γ = 0.01 τ−1,the nuclei appear at the critical deformation γc > 0.3.The large values of γc are due to that the high shearrates inhibit the crystal nucleation, that was previouslydiscussed in Refs. [15, 19].

For the crystalline nuclei with sizes ≤ 100 particles,it is quite reasonable within the statistical treatment tocharacterize their shape as a quantity averaged over dataof independent experiments. This averaging procedure isdenoted by 〈...〉 in Eq. (6). As seen from Fig. 5(a), for ashear-free system as well as for the system at small shearrates γ < 0.001 τ−1, the asphericity parameter takes the

Page 7: Kazan Federal University, 420008 Kazan, Russia · 2018-11-09 · changes of the systems may do impact on their rheologi-cal properties. It was found in Ref. [33] that the ordered

7

values S0 < 0.02. It indicates that the critically-sized nu-cleus approximate a symmetric shape close to be a spher-ical. Further, the quantity S0 increases from ' 0.005 toS0 ' 0.045 with the increasing critical deformation γc,that is due to increase of the nucleus shape aspheric-ity. Homogeneous steady-state shear promotes local re-arrangements of the particles and leads effectively to anincrease of particle mobility. Therefore, the system tem-perature takes a sense of the effective parameter [61],which increases with increase of the shear rate (Fig. 4in Ref. [15]). Moreover, there is anisotropy of interfacialfree energy due to shear. Then, the system with a specificcharacteristic interparticle interaction must be character-ized by a limit value of curvature of the solid (crystal)-liquid interface; while the geometries with larger valuesof curvature will be unstable. As a result, a plateau inthe γc-dependence of the asphericity parameter appearsand saturation of the γc-dependence of the critical size isobserved at high shear rates. Therefore, it is remarkablethat the γc-dependence of S0 is correlated with such thedependence of nc. One can see in Fig. 5(b) that boththe quantities S0 and nc are correlated, and a largervalue of the asphericity parameter S0 corresponds to alarger value of the critical size nc. This scenario dif-fers from crystal nucleation without external mechanicaldrive, where S0 ∝ 1/nc [46].

Projections of the pair-distribution function g(r) ontothe shear-gradient xy and the shear-vorticity xz planeswere defined for particles of the critically-sized nucleusas the next:

gc(x, y) =

⟨1

nc

nc∑i=1

nc∑j>i

δ (~r − ~rij(x, y))

⟩(9)

and

gc(x, z) =

⟨1

nc

nc∑i=1

nc∑j>i

δ (~r − ~rij(x, z))

⟩. (10)

Here, 〈...〉 means an averaging over data of indepen-dent simulation runs. At evaluation of the distributionsgc(x, y) and gc(x, z) the positions of particles were deter-mined with respect to geometric center of a nucleus.

The distributions gc(x, y) for the system at the temper-ature T = 0.15 ε/kB and at the shear rates γ = 0, 0.0005,0.005, and 0.01 τ−1 are shown in Fig. 6. Further, Fig. 7presents the distributions gc(x, z) for the shear-free sys-tem at the temperature T = 0.15 ε/kB and for the systemat the shear rate 0.01 τ−1. As seen, the closed lines cor-responding to the first, second and third coordinationsare well recognizable in these distributions. The lines arepresented by red, green and blue colors, respectively. Atthe absence of a shear, the contours of the distributionsgc(x, y) and gc(x, z) are circular, that is evidence of aspherical shape of the nucleus. As can be seen from con-tours of gc(x, y) and gc(x, z), the nucleus asphericity be-comes more pronounced with shear increase [Figs. 6(b),6(c), 6(d), and Fig. 7(b)]. This is also observed at other

Figure 5. (color online) (a) Asphericity parameter S0 as func-tion of the critical deformation γc. Insets show the critically-sized nuclei with various sizes and shapes. (b) Correlationbetween the asphericity parameter S0 and the critical size ncat different temperature of the system. Here, circles corre-spond to the system at T = 0.05 ε/kB , squares correspond toT = 0.15 ε/kB , and triangles correspond to T = 0.5 ε/kB .

considered temperatures, not presented in Figs. 6 and 7.

The contours of the distributions gc(x, y) and gc(x, z)change forms from circular to elliptical at increasingshear rate γ. The long axis of the elliptic contour ofgc(x, y) is oriented with respect to the gradient direction[see Figs. 6(b), 6(c) and 6(d)]. At the same time, theorientation of this ellipse within the xz-plane is not ob-served, that is seen, in particular, in Fig. 7. Namely,the long axis of the ellipse in the distribution gc(x, z)increases with increasing shear rate only in the sheardirection [see Fig. 7(b)], whereas the small axis of theellipse is practically unchanged both in the xy and xz-planes. This is direct evidence that at high shear ratesthe nuclei are characterized by an elongated ellipsoidalshape. Note that this is in agreement with results of

Page 8: Kazan Federal University, 420008 Kazan, Russia · 2018-11-09 · changes of the systems may do impact on their rheologi-cal properties. It was found in Ref. [33] that the ordered

8

Figure 6. (color online) Top: projections of the pair distribution function gc(x, y) for the particles of a critically-sized nucleusreceived for the system at the temperature T = 0.15 ε/kB and at the shear rates: (a) zeroth shear rate, γ = 0; (b) γ = 0.0005 τ−1

(critical deformation γc ∼ 3%); (c) γ = 0.005 τ−1 (γc ∼ 20%); (d) γ = 0.01 τ−1 (γc ∼ 50%). External boundaries of thedistribution gc(x, y) are shown by the dashed circle and ellipses. Bottom: projections of the pair distribution function ontoshear-gradient xy-plane obtained for hard-sphere glassy systems at different shear strains: (e) 1%, (f) 10%, (g) 20% and (h)60% (taken from Refs. [21, 32]).

Figure 7. (color online) Projections of the pair distributionfunction gc(x, z) onto the xz-planes for the particles of acritically-sized nucleus received for the system at the tem-perature T = 0.15 ε/kB and at the shear rates: (a) γ = 0; (b)γ = 0.01 τ−1. External boundaries of the distribution gc(x, z)are shown by the dashed circle and ellipse.

Refs. [16, 26, 40, 62].

The tilt angle ϕ between the gradient direction and thelongest axis of the (ellipsoid-like) nucleus was determineddirectly from contours of the distribution gc(x, y). InFig. 8, the γc-dependencies of ϕ obtained at different

shear rates and temperatures are shown. The tilt angleϕ increases with the increasing critical deformation γcaccording to the power-law dependence

ϕ(γc) = ϕ0 {1− exp (−Dγc)} . (11)

Here, D and ϕ0 are the positive fitting parameters. Asit appears, both the parameters are independent on thetemperature T for the considered temperature range andtake the values D = 6.0±0.05 and ϕ0 = 45◦±3◦, respec-tively. The function ϕ(γc) goes to saturation faster atthe higher value of the parameter D. The parameter ϕ0

is the limit value of ϕ. Namely, at the large deformations(in our case at γc > 0.6) the long axis of the elliptic con-tour in the distribution gc(x, y) is oriented at the angleϕ0 ≈ 45◦ with respect to the gradient direction (see alsoRefs. [11, 20, 32]). As it follows from Refs. [21, 32], apronounced anisotropy in microscopic structure of hard-sphere suspensions and glasses is observed with increas-ing shear strain. It was found in Refs. [21, 32] that withincreasing shear deformation the particles density startsto be dependent on direction, and the so-called stretchingand compression directions appear. It is remarkable thatthe contours of this distribution [see Fig. 6(e), 6(f), 6(g)and 6(h)] are similar to the contours of the distributiongc(x, y) [see Fig. 6(a), 6(b), 6(c) and 6(d)], where theangle between the stretching direction and the gradient

Page 9: Kazan Federal University, 420008 Kazan, Russia · 2018-11-09 · changes of the systems may do impact on their rheologi-cal properties. It was found in Ref. [33] that the ordered

9

Figure 8. (color online) Tilt angle ϕ of a critically-sized nu-cleus as a function of the critical deformation γc at differ-ent temperatures. The solid curve is fit by Eq. (11). Insets:schematic illustrations showing the change of orientation andshape of the nuclei at various γc.

direction goes to the limit value ≈ 45◦. On the otherhand, it was experimentally found by Nosenko et al. [20]for strongly coupled liquid that the microscopic structureof the system at influence of shear becomes anisotropic;and compressive and stretching axes tend to be orientedat the angle of ±45◦ respective to the direction of flow.Evidently, such ellipticity of the distribution gc(x, y) aris-ing at steady shear has an impact on the crystal nucle-ation processes and is also responsible for the critically-sized nuclei asphericity.

C. Size and Shape of Critically-Sized Nuclei

Let ac and bc be the small and large semiaxes of theelliptic contours in the distribution gc(x, y), respectively.The quantities ac and bc are evaluated from the distri-bution gc(x, y) at different critical deformation γc. Asseen from Fig. 9, the small semiaxis ac is practically un-changed with the increasing critical deformation γc, whilethe quantity bc increases with γc. Remarkably, the γc-dependencies of the quantities are well reproduced by

ac(γc) = a0 {1 +Aγκc } (12)

and

bc(γc) = a0 {1 + Bγκc } . (13)

Here, a0 is the radius of a spherical nucleus at γ = 0,and it is a0 ' 2.71, 2.81, and 2.87σ for temperaturesT = 0.05, 0.15, and 0.5 ε/kB , respectively. The dimen-sionless parametersA and B take the fixed positive valuesA ' 0.004±0.001 and B ' 0.42±0.04 for the considered

temperature range T ∈ [0.05; 0.5] ε/kB . The dimension-less parameter κ is defined by the thermodynamic prop-erties of a system and takes positive value κ = 1/3 forconsidered temperatures. We found that the ratio B/Ais ≈ 105 and, consequently, the asphericity of a nucleusshape increases mainly due to increase of the large semi-axis bc.

Figure 9. (color online) (a) Quantities ac and bc as functionsof the critical deformation γc. The solid curves are fit byEqs. (12) and (13). (b) Schematic image of the ellipsoidalshape, where ac and bc are small and large semiaxes deter-mined from the distribution gc(x, y).

From Eqs. (12) and (13) one obtains the γc-dependenceof the critical size nc:

nc(γc) =4

3πρc [ac(γc)]

2[bc(γc)] (14)

=4

3πρca

20b0{1 + Bγκc + 2Aγκc

+ 2ABγ2κc +A2γ2κc +A2Bγ2κc }.

Since the parameter A takes small values in the case ofthe Dz-system (A ' 0.004) and taking into account bc 'ac at γ = 0, Eq. (14) can be rewritten in the followingsimplified form:

nc(γc) = nc(γ = 0) {1 + Bγκc } , (15)

where

nc(γ = 0) =4

3πρca

30. (16)

Here, ρc ' 1.04±0.02σ−3 is the numerical density of thecrystalline phase. Fig. 4(c) shows the fit of the simulationresults by Eq. (15) at given values of B, κ, and a0. It canbe seen from Fig. 4(c), Eq. (15) reproduces correctly theγc-dependence of the critical size nc at the consideredtemperatures.

For the quantitative characterization of the nucleishape we define the deformation parameter

χ =bc − acbc + ac

. (17)

Page 10: Kazan Federal University, 420008 Kazan, Russia · 2018-11-09 · changes of the systems may do impact on their rheologi-cal properties. It was found in Ref. [33] that the ordered

10

0 0.2 0.4 0.6 0.80

0.02

0.04

0.06

0.08

0.1

0.12

0.14

0.16

0.18

γc

χ

1 2 3 4 5 6 7 80

0.2

0.4

0.6

0.8

1

nc/n0

χ

Dz, 0.05 ε/kB

Dz, 0.15 ε/kB

Dz, 0.5 ε/kB

GO

1 1.2 1.40

0.1

0.2

nc/n0

χ

0.05 ε/kB

0.15 ε/kB

0.5 ε/kB

master−curve

(b)(a)

Figure 10. (color online) (a) Quantity χ as function of the critical deformation γc. The solid curve is fit by Eq. (18). (b)Quantity χ as function of the scaled critical size nc/n0, where n0 ≡ nc(γ = 0) is the critical size at zeroth shear rate. The dashedcurve is fit by Eq. (19). The data for elongated critically-sized crystalline nuclei marked as GO are taken from Refs. [26, 27].

For the case bc ≥ ac the parameter χ takes the valuesfrom the range [0; 1]. Namely, if a nucleus is character-ized by a perfect spherical shape, we have χ = 0, whereasfor an elongated nucleus one has χ→ 1.

The γc-dependencies of the parameter χ at differenttemperatures are shown in Fig. 10(a). An increase ofχ with the critical deformation γc is mainly due to in-crease of the value bc. It should be noted that the γc-dependencies of S0 and χ are correlated [see Figs. 4(c)and 10(a)]. Further, the results reveal that the γc-dependencies of the parameter χ are reproduced by

χ(γc) =Bγκc

2 + Bγκc, (18)

which follows from Eqs. (12), (13) and (17). Here,B ' 0.42± 0.04 and κ ' 1/3 for the considered tempera-tures. We found that the γc-dependencies of the param-eter χ goes to saturation at high critical deformations,and extremely high shear rates do not lead to the highasphericity of a nucleus.

Moreover, Fig. 10(b) shows the parameter χ as func-tion of the scaled critical size nc/n0. Here, the quantityn0 is the critical size, which is evaluated for the system atabsence of a shear, n0 ≡ nc(γ = 0). This figures presentsour results as well as results of flow-induced nucleationin polymer melts taken from Refs. [26, 27]. Remarkably,all the data collapse into a master-curve reproducible by

χ(nc) =nc − n0nc + n0

. (19)

The Eq. (19) follows directly from Eqs. (15) and (18).

As seen from Fig. 10(b), values of nc/n0 and χ acces-sible to the considered Dz-system correspond to a nar-rower range in comparison with values for the polymersystem. The Dzugutov interaction potential is not ca-pable of forming elongated structures with large valuesof deformation parameter. Therefore, it is expected thatthe results for the Dzugutov pseudo-metallic system arelocated in the lower left part of the figure, whereas thedata for the soft systems (say, polymers) have to be lo-cated at higher values of parameters. Results presentedin Fig. 10(b) indicate on a possible universality in cor-relation between the characteristics of size and shapecritically-sized nucleus, that does not depend on specifictypes of the considered systems. Therefore, additionalstudies of this point would be useful.

IV. CONCLUSION

The main results of this study are the following:(i) The shear-induced structural ordering of the single-

component glassy system is initiated through crystal nu-cleation mechanism.

(ii) The critical size nc and the nucleation time τcwere evaluated at different fixed shear rates and tem-peratures. It is shown that both the size and the shapeof the critically-sized nuclei depend on the shear rate.The asphericity of the nuclei increases with the increas-ing shear rate γ (or with the increasing critical defor-mation γc ≡ γτc); the larger asphericity of the nucleuscorresponds to the larger critical size. It is verified by theincreasing asphericity parameter S0 and by the changes

Page 11: Kazan Federal University, 420008 Kazan, Russia · 2018-11-09 · changes of the systems may do impact on their rheologi-cal properties. It was found in Ref. [33] that the ordered

11

of the contours of gc(x, y) and gc(x, z) calculated for theparticles of the critically-sized nuclei.

(iii) The shapes of the critically-sized nuclei changewith increasing shear rate from spherical to ellipsoidal.The asphericity parameter and the critical size increaseaccording to the power-law ∝ (γτc)

1/3.

(iv) It is shown that the ellipsoidal nuclei are orientedwithin the shear-gradient plane at moderate and highshear rates. The tilt angle of the nuclei increases with theincreasing shear rate according to power-law dependence.

(v) It is found that the scaled nc-dependencies of thenuclei deformation parameter χ are collapsed into unified

master-curve.

ACKNOWLEDGMENTS

The work is supported in part by the grant MD-5792.2016.2 (program of support for young scientists inRF). Authors are grateful to the Ministry of Educationand Science of the Russian Federation for supporting theresearch in the framework of the state assignment (No3.2166.2017/4.6). The molecular dynamic simulationswere performed by using the computational clusters ofKazan Federal University and of Joint SupercomputerCenter of RAS.

[1] Janeschitz-Kriegl, H., “Some remarks on flow inducedcrystallization in polymer melts,” J. Rheol. 57, 1057-1064(2013).

[2] Scelsi, L., M. R. Mackley, H. Klein, P. D. Olmsted,R. S. Graham, O. G. Harlen, and T. C. B. McLeish,“Experimental observations and matching viscoelasticspecific work predictions of flow-induced crystallizationfor molten polyethylene within two flow geometries,” J.Rheol. 53, 859-876 (2009).

[3] Berthier, L., G. Biroli, “Theoretical perspective on theglass transition and amorphous materials,” Rev. Mod.Phys. 83, 587-645 (2011).

[4] Ediger, M. D., P. Harrowell, “Perspective: Supercooledliquids and glasses,” J. Chem. Phys. 137, 080901 (2012).

[5] Greer, L. A., “Metallic glasses,” Science 267, 1947-1953(1995).

[6] Lu, Z. P., and C. T. Liu, “Glass formation criterionfor various glass-forming systems,” Phys. Rev. Lett. 91,115505 (2003).

[7] Mokshin, A. V., and J.-L. Barrat, “Crystal nucleationand cluster-growth kinetics in a model glass under shear,”Phys. Rev. E 82, 021505 (2010).

[8] Lowen, H., “Colloidal soft matter under external con-trol,” J. Phys.: Condens. Matter 13, R415-R432 (2001).

[9] Haw, M. D., W. C. K. Poon, and P. N. Pusey, “Colloidalglasses under shear strain,” Phys. Rev. E 58, 4673-4687(1998).

[10] He, G.-L., R. Messina, H. Lowen, A. Kiriy, V. Bocharova,and M. Stamm, “Shear-induced stretching of adsorbedpolymer chains,” Soft Matter 5, 3014-3017 (2009).

[11] Dasgupta, R., H. George E. Hentschel, and I. Procaccia,“Microscopic mechanism of shear bands in amorphoussolids,” Phys. Rev. Lett. 109, 255502 (2012).

[12] Evans, D. J., G. P. Morriss, Statistical mechanics of non-equilibrium liquids (Cambridge Univ. Press, New York,2008).

[13] Fang, H., Y. Zhang, J. Bai, and Z. Wang, “Shear-inducednucleation and morphological evolution for bimodal longchain branched polylactide,” Macromolecules 46, 6555-6565 (2013).

[14] Butler, S., and P. Harrowell, “Kinetics of crystallizationin a shearing colloidal suspension,” Phys. Rev. E 52,6424-6430 (1995).

[15] Mokshin, A. V., B. N. Galimzyanov, and J.-L. Barrat,“Extension of classical nucleation theory for uniformlysheared systems,” Phys. Rev. E 87, 062307 (2013).

[16] Blaak, R., S. Auer, D. Frenkel, and H. Lowen, “Crystalnucleation of colloidal suspensions under shear,” Phys.Rev. Lett. 93, 068303 (2004).

[17] Kerrache, A., N. Mousseau, and L. J. Lewis, “Crystalliza-tion of amorphous silicon induced by mechanical sheardeformations,” Phys. Rev. B 84, 014110 (2011).

[18] Shao, Z., J. P. Singer, Y. Liu, Z. Liu, H. Li, M. Gopinad-han, C. S. O’Hern, J. Schroers, and C. O. Osuji, “Shear-accelerated crystallization in a supercooled atomic liq-uid,” Phys. Rev. E 91, 020301 (2015).

[19] Mura, F., and A. Zaccone, “Effects of shear flow on phasenucleation and crystallization,” Phys. Rev. E 93, 042803(2016).

[20] Nosenko, V., A. V. Ivlev, G. E. Morfill, “Microstructureof a liquid two-dimensional dusty plasma under shear,”Phys. Rev. Lett. 108, 135005 (2012).

[21] Koumakis, N., M. Laurati, S. U. Egelhaaf, J. F. Brady,and G. Petekidis, “Yielding of hard-sphere glasses duringstart-up shear,” Phys. Rev. Lett. 108, 098303 (2012).

[22] Kumaraswamy, G., J. A. Kornfield, F. Yeh, and B.S. Hsiao, “Shear-enhanced crystallization in isotacticpolypropylene. 3. Evidence for a kinetic pathway to nu-cleation,” Macromolecules 35, 1762-1769 (2002).

[23] Keller, A., M. J. Machin, “Oriented crystallization inpolymers,” J. Macromol. Sci., Phys. B1, 41-91 (1967).

[24] Roozemond, P., G. Peters, “Flow-enhanced nucleationof poly(1-butene): Model application to short-term andcontinuous shear and extensional flow,” J. Rheol. 57,1633-1653 (2013).

[25] Roozemond, P., M. van Drongelen, Z. Ma, A. Spoelstra,D. Hermida-Merino, and G. Peters, “Self-regulation inflow-induced structure formation of isotactic polypropy-lene,” Macromol. Rapid Commun. 36, 385-390 (2015).

[26] Graham, R. S., and P. D. Olmsted, “Coarse-grainedsimulations of flow-induced nucleation in semicrystallinepolymers,” Phys. Rev. Lett. 103, 115702 (2009).

[27] Graham, R., P. Olmsted, “Kinetic Monte Carlo simula-tions of flow induced nucleation in polymer melts,” Fara-day Discuss. 144, 71-92 (2010).

[28] McIlroya, C., and P. D. Olmsted, “Deformation of anamorphous polymer during the fused-filament-fabrication

Page 12: Kazan Federal University, 420008 Kazan, Russia · 2018-11-09 · changes of the systems may do impact on their rheologi-cal properties. It was found in Ref. [33] that the ordered

12

method for additive manufacturing,” J. Rheol. 61, 379-397 (2017).

[29] Viasnoff, V., and F. Lequeux, “Rejuvenation and over-aging in a colloidal glass under shear,” Phys. Rev. Lett.89, 065701 (2002).

[30] Rottler, J., and M. O. Robbins, “Shear yielding of amor-phous glassy solids: Effect of temperature and strainrate,” Phys. Rev. E 68, 011507 (2003).

[31] Wallace, M. L., and B. Joos, “Shear-induced overagingin a polymer glass,” Phys. Rev. Lett. 96, 025501 (2006).

[32] Koumakis, N., M. Laurati, A. R. Jacob, K. J. Mutch, A.Abdellali, A. B. Schofield, S. U. Egelhaaf, J. F. Brady, G.Petekidis, “Start-up shear of concentrated colloidal hardspheres: Stresses, dynamics, and structure,” J. Rheol.60, 603-623 (2016).

[33] Koumakis, N., A. B. Schofieldc, and G. Petekidis, “Ef-fects of shear induced crystallization on the rheology andageing of hard sphere glasses,” Soft Matter 4, 2008-2018(2008).

[34] Mokshin, A. V., and J.-L. Barrat, “Shear induced struc-tural ordering of a model metallic glass,” J. Chem. Phys.130, 034502 (2009).

[35] Shrivastav, G. P., P. Chaudhuri, J. Horbach, “Hetero-geneous dynamics during yielding of glasses: Effect ofaging,” J. Rheol. 60, 835-847 (2016).

[36] Mokshin, A. V., and B. N. Galimzyanov, “Scaling law forcrystal nucleation time in glasses,” J. Chem. Phys. 142,104502 (2015).

[37] Kashchiev, D., Nucleation: Basic theory with appplica-tions (Butterworth-Heinemann, Oxford, 2000).

[38] Kelton, K. F., A. L. Greer, “The classical theory,” Perg-amon Materials Series 15, 19-54 (2010).

[39] Kalikmanov, V. I., Nucleation theory; Lecture notes inphysics (Springer, New York, 2012), Vol. 860.

[40] Wang, X. W., J. Ouyang, and Z. Liu, “A phase field tech-nique for modeling and predicting flow induced crystal-lization morphology of semi-crystalline polymers,” Poly-mers 8, 230 1-20 (2016).

[41] Dzugutov, M., “Glass formation in a simple monatomicliquid with icosahedral inherent local order,” Phys. Rev.A 46, R2984-R2987 (1992).

[42] Roth, J., “Shock waves in materials with Dzugutov-potential interactions,” Phys. Rev. B 72, 014125 (2005).

[43] Mokshin, A. V., and J.-L. Barrat, “Shear-induced crys-tallization of an amorphous system,” Phys. Rev. E 77,021505 (2008).

[44] Lees, Q. W., and S. F. Edwards, “The computer study oftransport processes under extreme conditions,” J. Phys.C: Solid State Physics 5, 1921-1929 (1972).

[45] Steinhardt, P. J., D. R. Nelson, and M. Ronchetti,“Bond-orientational order in liquids and glasses,” Phys.Rev. B 28, 784-805 (1983).

[46] Reinhardt, A., J. P. K. Doye, “Free energy landscapesfor homogeneous nucleation of ice for a monatomic watermodel,” J. Chem. Phys. 136, 054501 (2012).

[47] Mickel, W., S. C. Kapfer, G. E. Schroder-Turk, and K.Mecke, “Shortcomings of the bond orientational order pa-

rameters for the analysis of disordered particulate mat-ter,” J. Chem. Phys. 138, 044501 (2013).

[48] ten Wolde, P. R., M. J. Ruiz-Montero, and D. Frenkel,“Numerical calculation of the rate of crystal nucleationin a Lennard-Jones system at moderate undercooling,”J. Chem. Phys. 104, 9932-9947 (1996).

[49] Mokshin, A. V., and B. N. Galimzyanov, “Steady-statehomogeneous nucleation and growth of water droplets:Extended numerical treatment,” J. Phys. Chem. B 116,11959-11967 (2012).

[50] Mokshin, A. V., and B. N. Galimzyanov, “Ordering inmodel metallic glass upon external homogeneous shear,”Bull. Russ. Acad. Sci. Phys. 77, 281-283 (2013).

[51] Roth, J., A. R. Denton, “Solid-phase structures of theDzugutov pair potential,” Phys. Rev. E 61, 6845-6857(2000).

[52] Reguera, D., J. M. Rubi, “Homogeneous nucleation ininhomogeneous media. II. Nucleation in a shear flow,” J.Chem. Phys. 119, 9888-9893 (2003).

[53] Lifshitz, E. M., L. P. Pitaevskii, Physical Kinetics,(Washington: Butterworth, 2006).

[54] Mokshin, A. V., and B. N. Galimzyanov, “Kinetics ofthe Crystalline Nuclei Growth in Glassy Systems,” Phys.Chem. Chem. Phys. 19, 11340-11353 (2017).

[55] N. B. Weingartner, C. Pueblo, F.S. Nogueira, K. F. Kel-ton, Z. Nussinov, “A Phase Space Approach to Super-cooled Liquids and a Universal Collapse of Their Viscos-ity,” Frontiers in Materials 3, 1-12 (2016).

[56] Holmqvist, P., M. P. Lettinga, J. Buitenhuis, and Jan K.G. Dhont, “Crystallization Kinetics of Colloidal Spheresunder Stationary Shear Flow,” Langmuir 21, 10976-10982 (2005).

[57] Tribout, C., B. Monasse, J. M. Haudin, “Experimen-tal study of shear-induced crystallization of an impactpolypropylene copolymer,” Colloid and Polymer Science274, 197-208 (1996).

[58] Coccorullo, I., R. Pantani, and G. Titomanlio,“Spherulitic Nucleation and Growth Rates in an iPP un-der Continuous Shear Flow,” Macromolecules 41, 9214-9223 (2008).

[59] Zaccone, A., H. Wu, D. Gentili, and M. Morbidelli, “The-ory of activated-rate processes under shear with applica-tion to shear-induced aggregation of colloids,” Phys. Rev.E 80, 051404 1-8 (2009).

[60] Shen, Y., S. B. Jester, T. Qi, and E. J. Reed, “Nanosec-ond homogeneous nucleation and crystal growth in shock-compressed SiO2,” Nature Materials 15, 60-65 (2016).

[61] Cugliandolo, L. F., J. Kurchan, L. Peliti, “Energy flow,partial equilibration, and effective temperatures in sys-tems with slow dynamics,” Phys. Rev. E 55, 3898-3914(1997).

[62] Somani, R. H., L. Yang, B. S. Hsiao, T. Sun, N. V. Pogo-dina, and A. Lustiger, “Shear-induced molecular orien-tation and crystallization in isotactic polypropylene: Ef-fects of the deformation rate and strain,” Macromolecules38, 1244-1255 (2005).