This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Equilibrium phases and domain growth kinetics of calamitic liquid crystals

Nishant Birdi srz188382@sire.iitd.ac.in School of Interdisciplinary Research, Indian Institute of Technology, Hauz Khas, New Delhi 110016, India.    Tom L. Underwood tlu20@bath.ac.uk Department of Chemistry, University of Bath, Bath BA2 7AY, United Kingdom.    Nigel B. Wilding nigel.wilding@bristol.ac.uk H.H. Wills Physics Laboratory, University of Bristol, Royal Fort, Bristol BS8 1TL, United Kingdom.    Sanjay Puri purijnu@gmail.com School of Interdisciplinary Research, Indian Institute of Technology, Hauz Khas, New Delhi 110016, India. School of Physical Sciences, Jawaharlal Nehru University, New Delhi 110067, India.    Varsha Banerjee varsha@physics.iitd.ac.in School of Interdisciplinary Research, Indian Institute of Technology, Hauz Khas, New Delhi 110016, India. Department of Physics, Indian Institute of Technology, Hauz Khas, New Delhi 110016, India.
Abstract

The anisotropic shape of calamitic liquid crystal (LC) particles results in distinct values of energy when the nematogens are placed side-by-side or end-to-end. This anisotropy in energy which is governed by a parameter κ\kappa^{\prime} has deep consequences on equilibrium and non-equilibrium properties. Using the Gay-Berne (GB) model, which exhibits the nematic (Nm) as well as the low temperature smectic (Sm) order, we undertake large-scale Monte Carlo and molecular dynamics simulations to probe the effect of κ\kappa^{\prime} on the equilibrium phase diagram and the non-equilibrium domain growth following a quench in the temperature TT or coarsening. There are two transitions in the GB model: (i) isotropic (I) to Nm at Tc1T_{c}^{1} and (ii) Nm to Sm at Tc2<Tc1T_{c}^{2}<T_{c}^{1}. κ\kappa^{\prime} decreases Tc1T_{c}^{1} significantly, but has relatively little effect on Tc2T_{c}^{2}. Domain growth in the Nm phase exhibits the well-known Lifshitz-Allen-Cahn (LAC) law, L(t)t1/2L(t)\sim t^{1/2} and the evolution is via annihilation of string defects. The system exhibits dynamical scaling that is also robust with respect to κ\kappa^{\prime}. We find that the Sm phase at the quench temperatures TT (T>Tc1T<Tc2T>T_{c}^{1}\rightarrow T<T_{c}^{2}) that we consider has SmB order with a hexatic arrangement of the LC molecules in the layers (SmB-H phase). Coarsening in this phase exhibits a striking two-time-scale scenario: first the LC molecules align and develop orientational order (or nematicity), followed by the emergence of the characteristic layering (or smecticity) along with the hexatic bond-orientational-order (BOO) within the layers. Consequently, the growth follows the LAC law L(t)t1/2L(t)\sim t^{1/2} at early times and then shows a sharp crossover to a slower growth regime at later times. Our observations strongly suggest that L(t)t1/4L(t)\sim t^{1/4} in this regime. Interestingly, the correlation function shows dynamical scaling in both the regimes and the scaling function is universal. The dynamics is also robust with respect to changes in κ\kappa^{\prime}, but the smecticity is more pronounced at larger values. Further, the early-time dynamics is governed by string defects, while the late-time evolution is dictated by interfacial defects. We believe this scenario is generic to the Sm phase even with other kinds of local order within the Sm layers.

I Introduction

Liquid crystals (LCs) are a state of matter that is intermediate between liquids and crystals as they manifest partial orientational and/or translational order Stephen and Straley (1974); de Gennes and Prost (1995); Singh (2000); Priestly (2012); Andrienko (2018). The LC mesophases emerge in response to changes in temperature or concentration. Thermotropic LCs are usually pure compounds of anisotropic organic molecules which exhibit phase changes by variation of temperature. Depending on their structure, the molecular shape can be rod like (calamitic), disc like (discotic) or banana shaped (bent-core). Lyotropic LCs are often mixtures of mesogens in a solvent and exhibit phase changes as the concentration of one of the components is varied. Amongst these different types, calamitic LCs are the most well studied due to their simplicity and wide applicability. At high temperatures, they exhibit an isotropic (I) phase where the rod-shaped molecules are randomly oriented. At low temperatures, the molecules align statistically parallel along a locally preferred axis to yield the nematic (Nm) phase with purely orientational order. This I-Nm phase transition is first order. If the Nm phase is uniaxial, it is described by a sign-invariant unit vector known as the director 𝒏\bm{n}.

As the temperature is further reduced, some LCs exhibit another transition to the smectic (Sm) phase characterized by partial translational order due to emergence of stacks of layers in addition to the lamellar order along 𝒏\bm{n}. The Sm nomenclature depends on the ordering in the layers. In the smectic A (SmA) phase, the layers are fluid-like. SmB has local order in the layers: e.g., SmB-H has a six-fold or hexatic bond-orientational-order (BOO) within the layers and SmB-C has long-ranged translational or crystalline order in the layers. These smectic phases have been observed experimentally in various LC compounds, either singly or together Moncton and Pindak (1979); Goodby (1981); Pindak et al. (1981); Davey et al. (1984); Voronov et al. (2020). Depending on the coupling between the Nm and the Sm order parameters, the Nm-Sm transitions could be either first order or second order McMillan (1971); Kralj et al. (2007). Several commercially used LCs such as the n-alkyl cyanobiphenyl (nCB) compounds indeed exhibit these twin transitions for n8\text{n}\geq 8, as revealed by light scattering and differential scanning calorimetry experiments Coles and Strazielle (1979); Chaban et al. (2020).

LCs are an important topic of research not only because of their enormous variety of applications, but also because they provide a platform for addressing a variety of fundamental problems in physics. The Nm phase is extensively used in liquid crystal displays, and the search for newer LCs with improved sensitivity and stability remains an ongoing activity Chen et al. (2018). The SmA mesophase provides a general template for striped systems such as biological membranes and flexible polymer crystallization Li and de Jeu (2003). It also shares symmetries with certain types of self-assembled block copolymer films which have applications in photolithography Harrison et al. (2000, 2002); Ruiz et al. (2007); Stoykovich et al. (2007). In recent years, there has been considerable interest in the study of the hexatic phase which was first predicted as an intermediate state between a crystal and a liquid in the theory of two-dimensional (2d2d) melting Halperin and Nelson (1978). The SmB-H phase provides the 3dd analog for the 2dd hexatic phase Davey et al. (1984); Voronov et al. (2020) Furthermore, LCs are experimentally accessible continuous symmetry systems. They have provided the framework for development of the theory of topological defects Kleman and Friedel (2008). The latter are relevant for a wide range of fields encompassing condensed matter physics, cosmology and biology Chuang et al. (1991); Pargellis et al. (1991); Lavrentovich et al. (2001); Wang et al. (2016); Kim et al. (2018); Sandford O’Neill et al. (2020).

Experimental measurements to probe various equilibrium and non-equilibrium responses, especially in LC mesophases with lower symmetry, remain a challenge because length scales (\sim nm) of morphologies and timescales (\sim ns) of evolution are often too small to be accessible. Consequently, computer simulations have emerged as a powerful tool for these investigations. In this context, the most important ingredient is the inter-particle potential which takes into account the anisotropy in the shape of the LC molecules as well as the attractive forces between them. The form proposed by Gay and Berne in 1981, based on the Gaussian overlap model of Berne and Pechukas Berne and Pechukas (1972), is one of the most popular pair potentials for anisotropic entities Gay and Berne (1981). The Gay-Berne (GB) model takes into account the aspect ratio κ\kappa of the mesogens and their energy anisotropy κ\kappa^{\prime}. The latter is defined as the ratio of the potential energies when a pair of mesogens are placed side-by-side (ss) and end-to-end (ee), see Fig. 1. Further, the model exhibits I, Nm and Sm phases, and the computed quantities agree well with the corresponding experimental measurements Luckhurst and Simmonds (1993); Berardi et al. (1993); Bates and Luckhurst (1999); Emerson et al. (1994); Cienega-Cacerez et al. (2014). The GB model has therefore become a prototype for investigations of LC systems Adams et al. (1987); Luckhurst et al. (1990); Miguel et al. (1991); Hashim et al. (1995); Satoh et al. (1996); La Penna et al. (1996); Cleaver et al. (1996); Neal et al. (1997); de Miguel and Vega (2002); Zannoni (2001); Martínez-Haya and Cuetos (2007).

Laboratory experiments generally require application of external fields that drive the system out-of-equilibrium. The system re-equilibrates, and the approach to equilibrium critically depends on the complexity of the free energy landscape. An important non-equilibrium study in this context is the kinetics of domain growth or coarsening, initiated by a sudden quench of the system from the disordered phase to the ordered phase Bray (2002); Puri (2004); Puri and Wadhawan (2009). The domains grow in size via annihilation of defects. The subsequent domain growth, characterized by a growing length scale L(t)L(t), is monitored with time. The growth law depends upon several factors such as symmetry of the order parameter, conservation laws, hydrodynamics, etc. It also provides important insights on the barriers to coarsening and relaxation time-scales. Phase ordering in the d=3d=3 Nm LCs is well studied using coarse-grained free energy models Bray et al. (1993); Wickham (1997) and lattice models Blundell and Bray (1992); Zhang et al. (1993); Bradac et al. (2011); Birdi et al. (2020). The domain growth obeys the Lifshitz-Allen-Cahn (LAC) law, L(t)t1/2L(t)\sim t^{1/2} Allen and Cahn (1979), with strings as the dominant defects. However, work in the context of this important non-equilibrium phenomenon for the smectic mesophases remains limited. There have been few experimental Harrison et al. (2002) and computational Abukhdeir and Rey (2008) studies on coarsening in the d=2d=2 SmA phase. They indicated that the orientational correlation length obeys an unusual L(t)t1/4L(t)\sim t^{1/4} law. Similar growth law, with speculations about logarithmic corrections, has been predicted for the d=3d=3 SmA phase using coarse-grained free energy models Liarte et al. (2015). Surprisingly, none of these studies address the significant role of the energy anisotropy that is a key feature of calamitic LCs.

Motivated to augment the above studies, we undertake large-scale simulations of the d=3d=3 GB model to understand the consequences of the energy anisotropy on equilibrium and non-equilibrium properties. There are two significant aspects of our study. Firstly, using Monte Carlo (MC) simulations, we identify the phase transition temperatures Tc1T_{c}^{1} (I\rightarrowNm) and Tc2T_{c}^{2} (Nm\rightarrowSm) for a range of κ\kappa^{\prime} values. These estimates equip us to perform temperature quenches in the Nm and Sm mesophases for the second part of our study. Subsequent to the quench, we study phase ordering kinetics via molecular dynamics (MD) simulations, which are better suited to monitor the systemic evolution as compared to MC simulations. The main results of our paper are as follows:
a) The Nm and Sm phases are observed for all values of κ\kappa^{\prime}. The Nm phase shrinks with increasing values of κ\kappa^{\prime} due to a (substantial) decrease in Tc1T_{c}^{1} and a (marginal) increase in Tc2T_{c}^{2}.
(b) When quenched from the I\rightarrowNm phase, domains with orientational order or nematicity emerge and grow with time. The correlation function C(r,t)C(r,t) vs. rr exhibits dynamical scaling indicating the presence of a unique length scale. The scaling function is universal for different values of κ\kappa^{\prime}.
(c) The tail of the structure factor obeys the generalized Porod law, S(k,t)k5S(k,t)\sim k^{-5} indicating scattering off string defects. The growth law in the Nm phase is the usual LAC law, L(t)t1/2L(t)\sim t^{1/2} characteristic of systems with non-conserved dynamics.
(d) For the quenches T>Tc1T<Tc2T>T_{c}^{1}\rightarrow T<T_{c}^{2}, we access the SmB-H phase. The coarsening in this phase is a two stage process: first there is emergence of nematicity, followed by the layering of mesogens or smecticity along with the development of hexatic order within the layers. The latter is enhanced by increasing values of κ\kappa^{\prime}.
(e) The correlation function C(r,t)C(r,t) exhibits dynamical scaling. However the scaling functions show small variations at short distances in the two regimes. These are reflected in the tails of the structure factor: S(k,t)k5S(k,t)\sim k^{-5} (early time nematicity) indicating scattering off strings; S(k,t)k4S(k,t)\sim k^{-4} (late time smecticity) implying scattering off interfaces. The mechanism of domain growth is thus distinct in the two regimes. The growth law exhibits a previously unreported cross-over from t1/2t^{1/2} to t1/4t^{1/4} as time evolves.

Our paper is organized as follows. Sec. II provides a detailed discussion of the Gay-Berne (GB) model. In Sec. III, we present the numerical details and results from our MC simulations for the equilibrium phase diagram of the system for a range of energy anisotropy values. Sec. IV presents numerical details and results from our MD simulations for the phase ordering kinetics of the system. The various tools for analyzing the coarsening morphologies are also discussed. In Sec. V, we conclude with a summary and discussion of our results.

II Gay-Berne Model

The Gay-Berne (GB) model is specially developed to mimic the interactions between ellipsoidal LC molecules, which are of equal size Gay and Berne (1981). Besides Lennard-Jones (LJ) potential with attractive and repulsive parts which decrease with the intermolecular separation rr as r6r^{-6} and r12{r^{-12}} respectively, the GB potential includes terms with additional dependence on the orientations of the LC molecules. The potential is anisotropic, and can model the orientational order observed in systems with anisotropic constituents.

Let us consider two uniaxial LC molecules ii and jj, with orientations defined by the unit vectors 𝒖i{\bm{u}_{i}} and 𝒖j{\bm{u}_{j}}, and centers separated by 𝒓{\bm{r}}. The GB potential for this prototypical pair is defined by Gay and Berne (1981):

Eij(𝒖i,𝒖j,𝒓)=4ϵ(𝒖i,𝒖j,𝒓^){[σ0rσ(𝒖i,𝒖j,𝒓^)+σ0]12[σ0rσ(𝒖i,𝒖j,𝒓^)+σ0]6},E_{ij}({\bm{u}_{i},\bm{u}_{j},\bm{r}})=4\epsilon({\bm{u}_{i},\bm{u}_{j},\bm{\hat{r}}})\left\{\left[\frac{\sigma_{0}}{r-\sigma({\bm{u}_{i},\bm{u}_{j},\bm{\hat{r}}})+\sigma_{0}}\right]^{12}-\left[\frac{\sigma_{0}}{r-\sigma({\bm{u}_{i},\bm{u}_{j},\bm{\hat{r}}})+\sigma_{0}}\right]^{6}\right\}, (1)

where σ0\sigma_{0} scales the distance and 𝒓^{\bm{\hat{r}}} is the unit vector along 𝒓{\bm{r}}. The other terms in Eq. (1) are as follows:
(a) The orientation-dependent range parameter σ(𝒖i,𝒖j,𝒓^)\sigma({\bm{u}_{i},\bm{u}_{j},\bm{\hat{r}}}) contains information about the shape of the LC molecules and is given by:

σ(𝒖i,𝒖j,𝒓^)=σ0{1χ2[(𝒖i𝒓^+𝒖j𝒓^)21+χ(𝒖i𝒖j)+(𝒖i𝒓^𝒖j𝒓^)21χ(𝒖i𝒖j)]}1/2.\sigma({\bm{u}_{i},\bm{u}_{j},\bm{\hat{r}}})=\sigma_{0}\left\{1-\frac{\chi}{2}\left[\frac{({\bm{u}_{i}\cdot\bm{\hat{r}}+\bm{u}_{j}\cdot\bm{\hat{r}}})^{2}}{1+\chi({\bm{u}_{i}\cdot\bm{u}_{j}})}+\frac{({\bm{u}_{i}\cdot\bm{\hat{r}}-\bm{u}_{j}\cdot\bm{\hat{r}}})^{2}}{1-\chi({\bm{u}_{i}\cdot\bm{u}_{j}})}\right]\right\}^{-1/2}. (2)

The shape anisotropy parameter χ\chi determines the system’s capability to form an orientationally ordered phase and is given by:

χ=κ21κ2+1,\chi=\frac{\kappa^{2}-1}{\kappa^{2}+1}, (3)

where κ\kappa is the aspect ratio of the LC molecules. If σe\sigma_{e} and σs\sigma_{s} are the length and breadth of the molecule, then κσe/σs\kappa\equiv\sigma_{e}/\sigma_{s}. More precisely, σe\sigma_{e} and σs\sigma_{s} are the contact distances or intermolecular separations at which the attractive and the repulsive terms in the potential cancel out each other when the LC molecules are in the end-to-end (ee) and the side-by-side (ss) configurations.
(b) The energy term ϵ(𝒖i,𝒖j,𝒓^)\epsilon({\bm{u}_{i},\bm{u}_{j},\bm{\hat{r}}}) is defined as:

ϵ(𝒖i,𝒖j,𝒓^)=ϵ0ϵ1μϵ2ν,\epsilon({\bm{u}_{i},\bm{u}_{j},\bm{\hat{r}}})=\epsilon_{0}\epsilon_{1}^{\mu}\epsilon_{2}^{\nu}, (4)

where ϵ0\epsilon_{0} scales the energy. The parameters ϵ1\epsilon_{1} and ϵ2\epsilon_{2} are defined follows:

ϵ1\displaystyle\epsilon_{1} =\displaystyle= 1χ2{(𝒓^𝒖i+𝒓^𝒖j)21+χ(𝒖i𝒖j)+(𝒓^.𝒖i𝒓^𝒖j)21χ(𝒖i𝒖j)},\displaystyle 1-\frac{\chi^{\prime}}{2}\left\{\frac{({\bm{\hat{r}}\cdot\bm{u}_{i}+\bm{\hat{r}}\cdot\bm{u}_{j}})^{2}}{1+\chi^{\prime}({\bm{u}_{i}\cdot\bm{u}_{j}})}+\frac{({\bm{\hat{r}}.\bm{u}_{i}-\bm{\hat{r}}\cdot\bm{u}_{j}})^{2}}{1-\chi^{\prime}({\bm{u}_{i}\cdot\bm{u}_{j}})}\right\}, (5)
ϵ2\displaystyle\epsilon_{2} =\displaystyle= [1χ2(𝒖i𝒖j)2]1/2.\displaystyle\left[1-\chi^{2}\left({\bm{u}_{i}\cdot\bm{u}_{j}}\right)^{2}\right]^{-1/2}. (6)

The energy anisotropy parameter χ\chi^{\prime}, analogous to Eq. (3), is defined as:

χ=κ1/μ1κ1/μ+1,\chi^{\prime}=\frac{\kappa^{\prime 1/\mu}-1}{\kappa^{\prime 1/\mu}+1}, (7)

where κ\kappa^{\prime} is the energy anisotropy which if greater than 1 promotes orientational order characteristic of LCs. If ϵe\epsilon_{e} and ϵs\epsilon_{s} are the well depths for the ee and the ss configurations, κϵs/ϵe\kappa^{\prime}\equiv\epsilon_{s}/\epsilon_{e}. The parameters μ\mu and ν\nu modify the well depths of the potential, and hence their impact on the nematicity and smecticity is very subtle. For instance in the ee configuration, κ=3.0\kappa=3.0 and κ=5.0\kappa^{\prime}=5.0 yields ϵ/ϵ0=1/3\epsilon/\epsilon_{0}=1/3 for μ=2\mu=2, ν=1\nu=1, and 25/27 for μ=1\mu=1, ν=3\nu=3. Similarly in the ss configuration, ϵ/ϵ0=5/3\epsilon/\epsilon_{0}=5/3 for μ=2\mu=2, ν=1\nu=1, and 125/27 for μ=1\mu=1, ν=3\nu=3. Fig. 1 shows variation of the GB potential Eij/ϵ0E_{ij}/\epsilon_{0} vs. rij/σ0r_{ij}/\sigma_{0} in the ee configuration for different values of (a) the energy anisotropy κ\kappa^{\prime} and (b) the exponents μ\mu, ν\nu. The corresponding variation in the ss configuration is shown in the insets. From the inset of Fig. 1(a), it is clear that the ss configuration is energetically favorable for all values of κ\kappa^{\prime} (the curves for different values of κ\kappa^{\prime} are coincident as the energy in the ss configuration depends only on κ\kappa and ν\nu).

Summing up, the GB model contains four essential parameters: κ\kappa, κ\kappa^{\prime}, μ\mu and ν\nu. Clearly, there is a large variety of GB homologues which differ from each other in terms of the values chosen for the four parameters Gay and Berne (1981); Berardi et al. (1993); Bates and Luckhurst (1999); Zannoni (2001). A frequently employed choice of GB parameters is κ=3\kappa=3, κ=5\kappa^{\prime}=5, μ=1\mu=1 and ν=3\nu=3 due to Berardi et al. Berardi et al. (1993). The choice of κ=3.0\kappa=3.0 is elementary as for real LC systems, the length-to-breadth ratio of the constituent molecules must be equal to or greater than about 3:1. The parameters used by Berardi et al. have two important features. Firstly, they provides diverse phases, viz. isotropic, nematic and smectic. The nematic phase is observed over a wide range of temperatures unlike the narrow region observed with the original GB parameterization: κ=3\kappa=3, κ=5\kappa^{\prime}=5, μ=2\mu=2 and ν=1\nu=1 Gay and Berne (1981). Secondly, the simulation results exhibit convergence with experiments. For instance, the computed temperature variation of the orientational order parameter is in agreement with the experimental data in many real systems Berardi et al. (1993). Consequently, the parameters κ=3\kappa=3, κ=5\kappa^{\prime}=5, μ=1\mu=1 and ν=3\nu=3 are frequently chosen for simulations of the GB model. We use this set of values in our present work and plan to do a comparative study with other choices of parameters at a later stage.

III Equilibrium studies using Monte Carlo simulations

III.1 Simulation Details

Prior to investigating the phase ordering kinetics, we first visit the problem of equilibrium phase transitions in the GB model to precisely identify the quench temperatures for the Nm and the Sm regimes. To sample the available phase space, MC simulations are performed in the canonical (NVTNVT) ensemble. We use the simulation program DL_MONTE Purton et al. (2013); Brukhno et al. (2021) for this purpose. DL_MONTE is a general-purpose MC program which supports a wide range of MC simulation techniques and interatomic potentials. However, prior to this work it was not applicable to particles with implicit orientations and anisotropic interaction potentials, for instance the GB particles. Our work is a step in that direction since it involved extending DL_MONTE to make it suitable for such systems. The latest version of DL_MONTE which includes our improvements is available at DL_ . This code is a beneficial resource for the research community to study uniaxial GB systems with a number of MC simulation techniques, including grand-canonical and constant-NPTNPT ensembles Frenkel and Smit (2002). In this work, we only present results using MC in the NVTNVT ensemble. Our DL_MONTE input files with all the necessary commands, parameters and comments are provided in the Supplementary Material for interested readers together with further details regarding the improvements we have made to DL_MONTE to treat the GB and similar models.

We consider a system of NN ellipsoidal particles interacting via the GB potential specified in Eq. (1) and use κ=3.0\kappa=3.0, κ=5.0\kappa^{\prime}=5.0, μ=1.0\mu=1.0 and ν=3.0\nu=3.0 unless specified. It is convenient to define scaled variables E=E/ϵ0=i,j;i<jN(Eij/ϵ0E^{\ast}=E/\epsilon_{0}=\sum_{i,j;i<j}^{N}(E_{ij}/\epsilon_{0}), E1=E/NE_{1}^{\ast}=E^{\ast}/N, T=kBT/ϵ0T^{\ast}=k_{B}T/\epsilon_{0}, r=r/σ0r^{\ast}=r/\sigma_{0} and ρ=Nσ03/V\rho^{\ast}=N\sigma_{0}^{3}/V. Simulations have been performed on cubic lattices with N=512N=512 and 10001000 particles using periodic boundary conditions and ρ\rho^{\ast} is set to 0.30. The initial configuration is a perfectly aligned state. The MC moves are performed according to the Metropolis algorithm Frenkel and Smit (2002). In these moves, a randomly selected LC molecule is either translated or rotated with translations and rotations attempted in each move with equal probability. The maximum angle to rotate an LC molecule is kept as 1515^{\circ}. In order to compute the energy efficiently, a spherical cutoff of rc=4.0σ0r_{c}=4.0\sigma_{0} is employed in conjunction with the Verlet neighbor list scheme Frenkel and Smit (2002). The simulations are performed for 10610^{6} MC cycles (one MC cycle corresponds to NN MC moves). The initial 5×1055\times 10^{5} cycles are necessary for equilibration. The remaining 5×1055\times 10^{5} cycles are used for thermal (block) averaging of various thermodynamic quantities of interest and the ensemble averaging is performed over 500 configurations.

III.2 Phase Diagram

The GB phase diagram is obtained by studying the average energy per particle E1\langle E_{1}^{\ast}\rangle vs. TT^{\ast} for values of the energy anisotropy κ\kappa^{\prime} = 1.25, 2.5, 5.0, 10.0 and 20.0. For each value, the transition temperatures Tc1T_{c}^{1} (I\rightarrowNm transition) and Tc2T_{c}^{2} (Nm\rightarrowSm transition) are identified from the discontinuity in E1\langle E_{1}^{\ast}\rangle. It is pertinent to point out here that for our coarsening experiments, we only need approximate boundaries as we consider the quenches far away from these. The I and Nm phases are confirmed by evaluation of the orientational order parameter, 𝒮\mathcal{S} given by:

𝒮=P2(cosθi)=3cos2θi12,\mathcal{S}=\langle P_{2}\left(\cos\theta_{i}\right)\rangle=\left\langle\frac{3\cos^{2}\theta_{i}-1}{2}\right\rangle, (8)

where cosθi=𝒖i𝒏\cos\theta_{i}=\bm{u}_{i}\cdot\bm{n} and the angular brackets \langle\cdot\cdot\cdot\rangle indicate an ensemble average. 𝒮=0\mathcal{S}=0 in the I phase, while 𝒮=1\mathcal{S}=1 in the perfectly aligned Nm phase. The defects correspond to regions with 𝒮0\mathcal{S}\simeq 0, even if the defect cores are biaxial Callan-Jones et al. (2006).

The smectic (layered) phases can be distinguished from the I and Nm phases by evaluating the translational order parameter Bates and Luckhurst (1999):

𝒯=|τ(𝒓𝒊)|=|exp(2iπ𝒓𝒊𝒏/dl)|,\mathcal{T}=\lvert\langle\tau(\bm{r_{i}})\rangle\rvert=\lvert\left\langle\exp(2i\pi\bm{r_{i}}\cdot\bm{n}/d_{l})\right\rangle\rvert, (9)

where 𝒓𝒊\bm{r_{i}} is the position vector of the LC molecule ii and dld_{l} is the layer spacing. In simulations, 𝒯\mathcal{T} is determined by first separately performing the ensemble averaging of the real and imaginary terms: cos(2iπ𝒓𝒊𝒏/dl)\cos(2i\pi\bm{r_{i}}\cdot\bm{n}/d_{l}) and sin(2iπ𝒓𝒊𝒏/dl)\sin(2i\pi\bm{r_{i}}\cdot\bm{n}/d_{l}), followed by calculation of the modulus for the ensemble average and further maximizing it with respect to dld_{l} Bates and Luckhurst (1999). In the I phase, 𝒯\mathcal{T} \rightarrow 0 as the layers are not well defined. On the other hand, 𝒯\mathcal{T} \rightarrow 1 in a perfectly layered structure.

A clear distinction between the SmA and SmB phases can be made by evaluating the hexatic bond-orientational order parameter given by Bates and Luckhurst (1999):

𝒞6=|ψ6(𝒓𝒊)|=|(kw(rik)exp(6iϕik))kw(rik))|,\mathcal{C}_{6}=\lvert\langle\psi_{6}(\bm{r_{i}})\rangle\rvert=\left\lvert\left\langle\left(\frac{\sum_{k}w(r_{ik}^{\ast})\exp(6i\phi_{ik}))}{\sum_{k}w(r_{ik}^{\ast})}\right)\right\rangle\right\rvert, (10)

where the summation is over the nearest neighbours (nn) kk of the LC molecule ii, ϕik\phi_{ik} is the angle between the vector (𝒓𝒊𝒓𝒌\bm{r_{i}}-\bm{r_{k}}) projected onto the plane normal to the director and a fixed reference axis (xx axis say) and w(rik)w(r_{ik}^{\ast}) is a cutoff function to select the nn for evaluation of ψ6(𝒓𝒊)\psi_{6}(\bm{r_{i}}). (Like 𝒯\mathcal{T}, for 𝒞6\mathcal{C}_{6} also first the ensemble averaging is done separately for the real and imaginary terms followed by calculation of the modulus for the average.) It is important to use the cutoff function as the number of nn might not be 6 and could be 7, 5 or 4 when the local translational order is imperfect. In our work, we used the procedure in reference Bates and Luckhurst (1999) to evaluate this cutoff function: w(rik)w(r_{ik}^{\ast}) is unity for rikr_{ik}^{\ast} below 1.4, zero for rikr_{ik}^{\ast} above 1.8 and with a linear interpolation in between these two extremes. If there is hexatic order, 𝒞6\mathcal{C}_{6} has an appreciable non-zero value and also the ensemble average for the cutoff function, w(rik)6\left\langle w(r_{ik}^{\ast})\right\rangle\rightarrow 6. In it’s absence, 𝒞6\mathcal{C}_{6} vanishes and w(rik)<6\left\langle w(r_{ik}^{\ast})\right\rangle<6.

The radial distribution function g(r)g(r) is routinely obtained in scattering experiments, and measures the probability of finding two molecules separated by distance rr relative to that in an ideal gas Frenkel and Smit (2002). It is a useful tool to distinguish between the local order in the different phases Bates and Luckhurst (1999), and is given by:

g(r)=ρ(r)¯ρ0,g(r^{\ast})=\frac{\langle\overline{\rho(r^{\ast})}\rangle}{\rho_{0}}, (11)

where ρ0=N/V\rho_{0}=N/V is the density of the ideal gas and ρ(r)¯\overline{\rho(r^{\ast})} is the average density of the system around rr^{\ast}. The numerical evaluation is facilitated by the following formula Frenkel and Smit (2002); Billeter and Pelcovits (1998):

g(r)=1Nρ0iji,jNδ(rrij)(4/3)π[(r+Δr)3(r)3].g(r^{\ast})=\frac{1}{N\rho_{0}}\bigg{\langle}\sum_{\stackrel{{\scriptstyle i,j}}{{i\neq j}}}^{N}\frac{\delta(r^{\ast}-r_{ij}^{\ast})}{(4/3)\pi[(r^{\ast}+\Delta r^{\ast})^{3}-(r^{\ast})^{3}]}\bigg{\rangle}. (12)

The δ\delta function is unity if rijr_{ij}^{\ast} falls within the shell centered on rr^{\ast} and is zero otherwise. The division by NN is done to normalize g(r)g(r^{\ast}) to a per-molecule function. By construction, g(r)=1g(r^{\ast})=1 for an ideal gas and any deviation implies correlations between the particles due to the intermolecular interactions. In the Nm phase, it has a noticeable maximum for small-rr^{\ast} and is 1 for large-rr^{\ast} indicating short-range order. In the SmA phase, g(r)g(r^{\ast}) exhibits a large nn peak and small oscillations thereafter with periodicity dl\sim d_{l} due to the layering. In the SmB-H phase on the other hand, g(r)g(r^{\ast}) shows an oscillatory behaviour with many sharp peaks separated by r1r^{*}\simeq 1 and a split-peak at r2r^{*}\simeq 2 characteristic of hexatic order within the layers. In the case of SmB-C order, the peak intensities do not decay till large distances.

We now proceed to evaluate the phase diagram in (T,κ)(T^{\ast},\kappa^{\prime})-space for κ\kappa^{\prime} = 1.25, 2.5, 5.0, 10.0 and 20.0. In Fig. 2(a), we show the variation of the per-particle equilibrium energy E1\langle E_{1}^{\ast}\rangle vs. TT^{\ast} for N=512N=512 (open up triangles) and N=1000N=1000 (open down triangles) at κ=5.0\kappa^{\prime}=5.0. (For brevity, we do not show the data sets for other values of κ\kappa^{\prime}.) The angular brackets indicate thermal averages. The simulation data coincide, indicating that the average energies are independent of the system size. We also plot the corresponding benchmarking results of Berardi et al. for N=512N=512 (white circles) and N=1000N=1000 during heating (white up triangles) and cooling (white down triangles) protocols Berardi et al. (1993). There is an excellent agreement with our simulation results obtained using DL_MONTE, even though the initial conditions in the two sets of simulation experiments are quite distinct. Furthermore, the effects of system size on the average energy are negligible (except near the transition in some cases). The left and right edges of the Nm region (in green) provide the scaled transition temperatures Tc2T^{\ast 2}_{c} and Tc1T^{\ast 1}_{c} as indicated. Both the transitions are discontinuous (first-order).

To confirm the phase corresponding to each data point, we have also evaluated for the N=1000N=1000 system, the temperature variation of the average order parameters 𝒮\langle\mathcal{S}\rangle, 𝒯\langle\mathcal{T}\rangle and 𝒞6\langle\mathcal{C}_{6}\rangle, or g(r)g(r^{\ast}) vs. rr^{\ast} as appropriate. Fig. 2(b) shows the typical behaviour of the order parameters in various phases. Note the rise in 𝒮\langle\mathcal{S}\rangle as the temperature crosses Tc1T^{\ast 1}_{c} from above, indicating the onset of nematic order. Similarly, note the rise in translational order parameter 𝒯\langle\mathcal{T}\rangle as the temperature crosses Tc2T^{\ast 2}_{c} indicating the development of SmA order (while evaluating 𝒯\mathcal{T}, we have observed that the layer spacing dld_{l} which maximizes 𝒯\mathcal{T} is less than the aspect-ratio κ=3.0\kappa=3.0 due to interdigitation of the layers Bates and Luckhurst (1999)). At slightly lower values of TT, the hexatic order parameter 𝒞6\langle\mathcal{C}_{6}\rangle becomes significant suggesting that the phase is SmB-H. Accurate determination of the phase boundaries will require careful evaluations. We refrain from going in this direction as the focus of the present study is on coarsening and approximate phase boundaries are sufficient for this purpose. (We wish to mention here that though the equilibrated configurations, their energies in various phases and the order parameter 𝒮\langle\mathcal{S}\rangle were obtained using DL_MONTE, the codes for evaluation of the order parameters 𝒯\mathcal{T} and 𝒞6\mathcal{C}_{6} were written separately.)

To evaluate g(r)g(r^{\ast}), we have used a set of 745745 bins with a separation cut-off of r=7.45r^{\ast}=7.45 which yields Δr=0.01\Delta r^{\ast}=0.01 (see Supplementary Material for details regarding specification of parameters for DL_MONTE). Fig. 2(c) shows the g(r)g(r^{\ast}) vs. rr^{\ast} behaviour for the Nm phase (green line, T=2.5T^{\ast}=2.5) and the Sm phase (blue line, T=1.0T^{\ast}=1.0). It is characterized by multiple sharp peaks with a split peak at r2r^{\ast}\simeq 2, characteristic of the SmB-H order. Further, it can be seen that g(r)g(r^{\ast}) decays as rr^{\ast} increases, implying that the translational order is lost at larger distances and hence the SmB phase is not of crystal type.

Fig. 2(d) depicts the variation of Tc1T^{\ast 1}_{c} (white up triangles) and Tc2T^{\ast 2}_{c} (white down triangles) for a range of κ\kappa^{\prime} values. To be noted here is that Tc1T^{\ast 1}_{c} decreases considerably with increasing κ\kappa^{\prime} thereby shrinking the Nm phase. On the other hand, Tc2T^{\ast 2}_{c} increases only slightly. With the original GB parameterization, de Miguel et al. also reported similar observations de Miguel et al. (1996).

IV Kinetic properties from Molecular Dynamics Simulations

IV.1 Methodology

We now turn to a study of domain growth kinetics in the d=3d=3 GB model via MD simulations. To the best of our knowledge, they have not been addressed. There have been some coarsening studies for the Nm phase of the GB model in d=3d=3. A relevant contribution in this context is by Billeter et al. Billeter et al. (1999). They observed the usual defect structures (e.g. disclinations and monopoles), but could not extract a reliable growth law due to the small system size considered (\sim60000 particles). Using large systems (\sim260000 particles), we perform deep quenches from the I phase to the Nm and the Sm phases [shown in Fig. 2(d)] and allow the system to evolve for long times. Our main interest is to determine the novel defects and growth laws in the Sm phase, and the impact of the energy anisotropy κ\kappa^{\prime} on ordering. We also undertake analogous studies for the Nm phase to complement previous works Billeter et al. (1999).

All our MD simulations have been performed in the NVTNVT ensemble using the LAMMPS software package Plimpton (1995); Lam . The details regarding implementation of the model into LAMMPS 111Note that LAMMPS technically implements the generalized Gay-Berne model Berardi_1995 in which interactions of the ellipsoidal particles are characterized via a (diagonal) shape matrix S=S= diag(σa,σb,σc\sigma_{a},\sigma_{b},\sigma_{c}) and a (diagonal) energy matrix E=E= diag(ϵa,ϵb,ϵc\epsilon_{a},\epsilon_{b},\epsilon_{c}), where σa,σb,σc\sigma_{a},\sigma_{b},\sigma_{c} are the lengths and ϵa,ϵb,ϵc\epsilon_{a},\epsilon_{b},\epsilon_{c} are the relative well depths of interaction along the three semi-axes of an ellipsoid. The GB model discussed in Sec. II and employed in Sec. III was the potential originally presented for ellipsoidal particles of equal size. The generalized GB model was developed to represent dissimilar biaxial ellipsoids Berardi_1995. It reduces to the GB model described in Sec. II when the molecules become uniaxial. Hence, for our study, we work with the generalized GB model in the uniaxial limit. and also the analytical expressions for the forces and torques, have been described in Brown et al. (2009). We consider N=262144N=262144 uniaxial ellipsoidal particles confined in a cubic box of linear size Lsσ0L_{s}\sigma_{0} with periodic boundary conditions in all three coordinate directions. Hence the volume V=V= (Lsσ0L_{s}\sigma_{0})3 = (95.6033σ095.6033\sigma_{0})3 such that ρ0.3\rho^{\ast}\simeq 0.3. The parameters in the GB potential of Eq. (1) are those employed by Berardi et al. Berardi et al. (1993). We also study the effects of varying the energy anisotropy parameter κ\kappa^{\prime}. The MD runs are carried out using the standard velocity Verlet algorithm. In LAMMPS, the dimensionless MD time unit t0=m0σ02/ϵ0=1.0t_{0}=\sqrt{m_{0}\sigma_{0}^{2}/\epsilon_{0}}=1.0. We choose the reduced MD integration time step Δt=Δt/t0=0.001\Delta t^{\ast}=\Delta t/t_{0}=0.001. The temperature TT^{\ast} is controlled and maintained constant via the Nosé-Hoover thermostat, which is known to preserve hydrodynamics Frenkel and Smit (2002); Nosé (1984); Binder et al. (2004). The homogeneous initial configurations are prepared by equilibrating the system at a high temperature (T=6.0T=6.0) for about 10510^{5} MD steps. To initiate the coarsening process (t=0t=0), the system is quenched to the indicated temperatures in Fig. 2(b). The evolution of the system is then monitored. All statistical quantities of interest are averaged over 1919 independent initial conditions. Our input files and parameters used in LAMMPS are provided in the Supplementary Material.

IV.2 Characterization Tools

For a translationally invariant system, the usual probe to characterize configurational morphologies is the equal-time correlation function Puri and Wadhawan (2009):

C(r,t)=ψ(r1,t)ψ(r2,t)ψ(r1,t)ψ(r2,t),C(\vec{r},t)=\langle\psi(\vec{r_{1}},t)\psi(\vec{r_{2}},t)\rangle-\langle\psi(\vec{r_{1}},t)\rangle\langle\psi(\vec{r_{2}},t)\rangle, (13)

where ψ(r,t)\psi(\vec{r},t) is a suitable order parameter, r=r2r1\vec{r}=\vec{r_{2}}-\vec{r_{1}} and \langle\cdot\cdot\cdot\rangle represents the ensemble average. Small-angle scattering experiments yield the structure factor:

S(k,t)=𝑑reikrC(r,t),S(\vec{k},t)=\int d\vec{r}~{}e^{i\vec{k}\cdot\vec{r}}~{}C(\vec{r},t), (14)

where k\vec{k} is the wave-vector of the scattered beam. A characteristic length scale L(t)L(t) is usually defined as the distance at which C(r,t)C(\vec{r},t) decays to, say, 0.20.2 times its maximum value. If the domain growth is characterized by a unique length scale L(t)L(t), then C(r,t)C(\vec{r},t) and S(k,t)S(\vec{k},t) show the dynamical scaling property Puri (2004); Puri and Wadhawan (2009): C(r,t)=g(r/L)C(\vec{r},t)=g(r/L); S(k,t)=Ldf(kL)S(\vec{k},t)=L^{d}f(kL). The asymptotic (large-kk) tail of S(k,t)S(\vec{k},t) contains information about the defects in the system. Continuous O(n)O(n) spin models exhibit the generalized Porod law, with the asymptotic form: S(k,t)k(d+n)S(k,t)\sim k^{-(d+n)} Porod et al. (1982); Oono and Puri (1988); Bray and Puri (1991). For n=1n=1, the defects are interfaces, and the corresponding scattering function exhibits the usual Porod law: S(k,t)k(d+1)S(k,t)\sim k^{-(d+1)}. For n>1n>1, the different topological defects are vortices (n=2,d=2n=2,\ d=2), strings (n=2,d=3n=2,\ d=3), and monopoles or hedgehogs (n=3,d=3n=3,\ d=3). So in d=3d=3, S(k,t)k5S(k,t)\sim k^{-5} or k6\sim k^{-6} depending on whether strings or monopoles dominate in the defect dynamics.

In LC mesophases such as the Sm phase, an appropriate measure of the orientational and translational order is provided by the longitudinal pair correlation function g(r)g_{\parallel}(r_{\parallel}) (parallel to the long axis of the LC molecules) and the transverse pair correlation function g(r)g_{\perp}(r_{\perp}) (perpendicular to the long axis of the LC molecules) Richter and Gruhn (2006). Evaluation of g(r)g_{\parallel}(r_{\parallel}) employs a cylindrical volume to probe the LC molecules aligned in the ee configuration and is given by Richter and Gruhn (2006); Cañeda-Guzmán et al. (2014):

g(r)=ijNδ(rrij,i)θ(σ0rij,i)θ(σ0rij,j)Nρπ(σ0/2)2h.g_{\parallel}(r_{\parallel})=\bigg{\langle}\sum_{i\neq j}^{N}\frac{\delta(r_{\parallel}-r_{ij,\parallel_{i}})\theta(\sigma_{0}-r_{ij,\perp_{i}})\theta(\sigma_{0}-r_{ij,\perp_{j}})}{N\rho^{\ast}\pi(\sigma_{0}/2)^{2}h}\bigg{\rangle}. (15)

The Heaviside step function θ(x)=1\theta(x)=1 when x0x\geq 0 and θ(x)=0\theta(x)=0 otherwise, \langle\cdot\rangle indicates an ensemble averaging over different initial (independent) conditions, hh is the cylinder height used to discretize the volume, rij,i=|𝒓ij,i|=|𝒓ij𝒖i|r_{ij,\parallel_{i}}=|\bm{r}_{ij,\parallel_{i}}|=|\bm{r}_{ij}\cdot\bm{u}_{i}| is the center-of-mass separation along the director of molecule ii (the director for molecule jj of the pair could also be considered in this δ\delta-function evaluation as the value remains almost the same in the ee configuration). rij,i=|𝒓ij,i|=|𝒓ij𝒓ij,i|r_{ij,\perp_{i}}=|\bm{r}_{ij,\perp_{i}}|=|\bm{r}_{ij}-\bm{r}_{ij,\parallel_{i}}| and rij,j=|𝒓ij,j|=|𝒓ij𝒓ij,j|r_{ij,\perp_{j}}=|\bm{r}_{ij,\perp_{j}}|=|\bm{r}_{ij}-\bm{r}_{ij,\parallel_{j}}| are the corresponding transverse separations from 𝒖i\bm{u}_{i} and 𝒖j\bm{u}_{j}. The quantity g(r)g_{\parallel}(r_{\parallel}) probes the average orientation of the LC molecules and the layering or smecticity in the system.

Evaluation of g(r)g_{\perp}(r_{\perp}) employs hollow, concentric cylinders to probe LC molecules aligned in the ss configuration and is given by Richter and Gruhn (2006); Cañeda-Guzmán et al. (2014):

g(r)=ijNδ(rrij,i)θ(δL/2rij,i)θ(δL/2rij,j)Nρπ((r+δL)2r2)h.g_{\perp}(r_{\perp})=\bigg{\langle}\sum_{i\neq j}^{N}\frac{\delta(r_{\perp}-r_{ij,\perp_{i}})\theta(\delta L_{\perp}/2-r_{ij,\parallel_{i}})\theta(\delta L_{\perp}/2-r_{ij,\parallel_{j}})}{N\rho^{\ast}\pi((r_{\perp}+\delta L_{\perp})^{2}-r_{\perp}^{2})h}\bigg{\rangle}. (16)

In the above equation, δL\delta L_{\perp} represents the thickness of the hollow cylinder and as for the g(r)g_{\parallel}(r_{\parallel}) case, we can use the director for molecule jj of the pair in the δ\delta-function evaluation since the value remains almost the same in the ss configuration. The quantity g(r)g_{\perp}(r_{\perp}) probes the translational structure about the LC molecules and their arrangement within layers.

IV.3 Morphologies, Textures and Growth Laws

Let us first discuss the kinetics of domain growth following a quench from the I phase to a temperature T=2.5T^{\ast}=2.5 in the Nm phase. The left panel of Fig. 3, shows representative snapshots from the time evolution of the configurational structure for (a) κ=5.0\kappa^{\prime}=5.0, t=8192t=8192; (b) κ=5.0\kappa^{\prime}=5.0, t=53248t=53248; and (c) κ=10.0\kappa^{\prime}=10.0, t=53248t=53248. For clarity, we have shown only a 20320^{3} corner of the entire box. These corners on average consist of about N=2400N=2400 particles. The right panel shows the corresponding top surface (d=2d=2 cross-sections). There is emergence and growth of orientational Nm order, and the energy anisotropy does not affect the coarsening phenomenon. We characterize the morphologies and their texture by evaluating the correlation function and the structure factor. These are obtained by a coarse-graining procedure in which the system is divided into non-overlapping sub-boxes of size (3.0σ03.0\sigma_{0})3. The sub-box size is carefully chosen to ensure that each one contains about 8 to 10 particles. The continuum LC configurations are thus mapped onto a simple cubic lattice of size (32σ032\sigma_{0})3. The relevant order parameter is the orientational order parameter defined in Eq. (8), i.e., P2(cosθi)¯\overline{P_{2}(\cos\theta_{i})}. Here θi\theta_{i} is the angle made by a molecule located in the ii-th box with the global director 𝒏\bm{n} of the system (determined by taking average of orientations for all the NN particles present in the system), and the overline implies an average over all the particles in the sub-box.

Fig. 4(a) shows the scaled correlation function, C(r,t)C(r,t) vs. r/L(t)r/L(t) at tt = 8192, 16384 and 32768 for κ\kappa^{\prime} = 5.0 and at t=32768t=32768 for κ\kappa^{\prime} = 2.5 and 10.0. The data exhibits dynamical scaling as well as super-universality with respect to the energy anisotropy κ\kappa^{\prime}. The dynamical scaling property demonstrates that the coarsening patterns are statistically self-similar in time. The property of super-universality indicates that the morphologies in the Nm phase are independent of the energy anisotropy. This is expected because the GB energy between a pair of ellipsoids in the ss arrangement depends only on κ\kappa and ν\nu. A log-log plot of the corresponding scaled structure factors, S(k,t)L(t)3S(k,t)L(t)^{-3} vs. kL(t)kL(t) is shown in Fig. 4(b). In the asymptotic large-kk limit, the structure factor follows the generalized Porod law: S(k,t)k5S(k,t)\sim k^{-5}, indicating scattering off string defects Bray et al. (1993); Birdi et al. (2020). Next, we study the time-dependence of the domain size. Fig. 4(c) shows the variation of L(t)L(t) vs. tt on a log-log scale for κ\kappa^{\prime} = 2.5, 5.0 and 10.0 respectively. It can be clearly observed that after an initial transient, the evolving systems are consistent with the t1/2t^{1/2} growth regime (LAC law) characteristic of systems with non-conserved order parameter. The system size and time scales of our simulation are sufficient to establish the LAC domain growth law for the Nm phase of the GB model although there is onset of finite-size effects at late times.

We now come to the primary focus of our paper: kinetics of domain growth in the Sm mesophase. The coarsening is initiated by a quench from T=6.0T^{\ast}=6.0 (I phase) to T=1.0T^{\ast}=1.0 (Sm phase). Fig. 5 shows the prototypical evolution morphologies for: (a) κ=5.0\kappa^{\prime}=5.0, t=4096t=4096; (b) κ=5.0\kappa^{\prime}=5.0, t=98304t=98304; and (c) κ=10.0\kappa^{\prime}=10.0, t=98304t=98304. As in Fig. 3, we have shown only a 20320^{3} corner of the entire box, consisting of about 2400 particles on average. The frames on the right show the corresponding top surface. It is interesting to note the initial onset of nematicity at the earlier time (t=4096t=4096) followed by the development of smecticity at later time (t=98304t=98304). Additionally, as observed in Fig. 5(c), the smecticity is significantly enhanced for κ=10.0\kappa^{\prime}=10.0.

To characterize the Sm order, Fig. 6 shows for the specified values of tt and κ\kappa^{\prime}: (a) the longitudinal pair correlation function g(r/σ0)g_{\parallel}(r_{\parallel}/\sigma_{0}) vs. r/σ0r_{\parallel}/\sigma_{0} evaluated using Eq. (11) and (b) the transversal pair correlation function g(r/σ0)g_{\perp}(r_{\perp}/\sigma_{0}) vs. r/σ0r_{\perp}/\sigma_{0} evaluated using Eq. (12). Notice that in Fig. 6(a), the early-time behaviour at t=4096t=4096 predominantly exhibits a single peak at 3σ03\sigma_{0} - the length of the LC molecule. It is characteristic of Nm order with molecular alignment along an average direction or the director. As time evolves, there is emergence of newer peaks at approximately 6σ06\sigma_{0}, 9σ09\sigma_{0}, etc. as the system coarsens. This signifies development of long-range longitudinal order or layers with an inter-layer spacing of 3σ0\sim 3\sigma_{0}. (The average separation between the peaks can be an approximate measure of the inter-layer spacing dld_{l}. In Fig. 6(a), dl2.5σ0d_{l}\approx 2.5\sigma_{0} due to interdigitation of the neighbouring layers.) Notice that increase in κ\kappa^{\prime} reduces the intensity variation between the first and second peaks implying enhancement of smecticity. (Recall that perfect translational order is characterized by peaks of equal intensity in the pair distribution function). On the other hand, g(r/σ0)g_{\perp}(r_{\perp}/\sigma_{0}) vs. r/σ0r_{\perp}/\sigma_{0} in Fig. 6(b) exhibits peaks around multiples of σ0\sigma_{0} - the width of the ellipsoidal LC molecule. The splitting of the peak at r2.0σ0r_{\perp}\simeq 2.0\sigma_{0}, a signature of BOO within the layers, is clearly seen at later times. Further, the translational order within the layers is not long-ranged as consecutive peaks have decreasing intensity. These characteristics, along with the non-zero value of the hexatic order parameter 𝒞6\langle\mathcal{C}_{6}\rangle in Fig. 2(b), confirm the presence of the SmB-H phase. The intra-layer BOO is not affected by κ\kappa^{\prime}.

The observations from Fig. 6 suggest the following scenario for coarsening of the SmB-H phase: it is a two-timescale process, with the onset of Nm order followed by SmB-H order. To confirm this, we have also performed a similar study for the Nm\rightarrowSmB-H quenches and evaluated these distribution functions. Except for the emergence of multiple peaks at earlier time in the longitudinal pair correlation function (it’s now a one-timescale process as Nm order is already present and hence with time, the system exhibits only layering with BOO), the behavior of the correlation functions is qualitatively similar to that observed in Fig. 6 and hence we do not present them separately. Quantitatively, the peaks in both the functions have much higher intensity compared to the corresponding peaks in Fig. 6, indicating faster layering as well as development of BOO. We do not present these data sets to prevent repetition. The two-time-scale process is the most significant outcome from our study of the GB model. It is further reiterated by the growth laws which will be discussed shortly.

We next focus on the scaling functions that describe the time-dependent morphologies. To evaluate them, we follow the same coarse-graining approach, but now the system is divided into non-overlapping sub-boxes of size (6.0σ06.0\sigma_{0})3, which maps the system onto a simple cubic lattice of size (16.0σ016.0\sigma_{0})3. This procedure gives us a continuous order parameter field and eliminates any molecular-level anisotropies. We have changed the system size to ensure that even at later times, the sub-boxes contain about 60 to 70 particles. In Fig. 7(a), we plot C(r,t)C(r,t) vs. r/L(t)r/L(t) for three specified values of tt and κ\kappa^{\prime}. The typical scalar nematic order parameter P2(cosθi,t)¯\overline{P_{2}(\cos\theta_{i},t)} is determined at each site ii of this discretized lattice (as for the nematic regime earlier) and subsequently the standard probes are evaluated. The data sets for κ=5.0\kappa^{\prime}=5.0 at different times neatly collapse onto a single master function, showing that the scaling regime has been reached. This data collapse indicates the existence of dynamical scaling. Furthermore, the excellent data collapse for different κ\kappa^{\prime} values at time t=98304t=98304 suggests that the scaling functions are robust with respect to the anisotropy in energy. Fig. 7(d) shows the corresponding scaled structure factor, S(k,t)L(t)3S(k,t)L(t)^{-3} vs. kL(t)kL(t) for a range of tt and κ\kappa^{\prime} values. There is an excellent data collapse, confirming both dynamical scaling and super-universality. The tail decays as S(k,t)k5S(k,t)\sim k^{-5} due to the presence of string defects.

It is also possible to determine the correlation between the Sm layers by evaluating the translational correlation function Cl(r,t)C_{l}(r,t). This can be done by evaluating the translational order parameter, |τ(𝒓𝒊,t)¯|=|exp(2iπ𝒓𝒊𝒏/dl)¯|{\lvert\overline{\tau(\bm{r_{i}},t)}\rvert}={\lvert\overline{\exp(2i\pi\bm{r_{i}}\cdot\bm{n}/d_{l})}\rvert}. Here, 𝒓𝒊\bm{r_{i}} is the position vector of the LC molecule (located in the sub-box with index ii), 𝒏\bm{n} is the global director of the system and the layer spacing dld_{l} is evaluated from the average separation between the peaks in the longitudinal pair correlation function at a given tt. The overline implies an average over all the particles located in the sub-box with index ii. Fig. 7(b) shows the scaled translational correlation function Cl(r,t)C_{l}(r,t) vs. r/Ll(t)r/L_{l}(t) for the specified values of tt and κ\kappa^{\prime}. The corresponding structure factors Sl(k,t)Ll(t)3S_{l}(k,t)L_{l}(t)^{-3} vs. kLl(t)kL_{l}(t) are shown in Fig. 7(e). The high quality of the data collapse again confirms dynamical scaling and super-universality of the scaling functions. The structure factor tail exhibits the Porod decay, Sl(k,t)k4S_{l}(k,t)\sim k^{-4}, characteristic of scattering off sharp interfaces arising between the Sm layers.

We have also determined the correlation function CH(r,t)C_{H}(r,t) using the bond-orientational order parameter |ψ6(𝒓𝒊,t)¯|{\lvert\overline{\psi_{6}(\bm{r_{i}},t)}\rvert}. Here, 𝒓𝒊\bm{r_{i}} is the position vector of the LC molecule located in the sub-box with index ii, and the overline implies an average over all the particles located in the sub-box with index ii. Fig. 7(c) shows the scaled correlation function CH(r,t)C_{H}(r,t) vs. r/LH(t)r/L_{H}(t) for the specified values of tt and κ\kappa^{\prime}. The corresponding structure factors SH(k,t)LH(t)3S_{H}(k,t)L_{H}(t)^{-3} vs. kLH(t)kL_{H}(t) are shown in Fig. 7(f). The good data collapse re-confirms dynamical scaling and super-universality of the scaling functions.

Finally, we study the domain growth laws for the SmB-H mesophase. Fig. 8(a) shows L(t)L(t) vs. tt, Fig. 8(b) shows Ll(t)L_{l}(t) vs. tt and Fig. 8(c) shows LH(t)L_{H}(t) vs. tt on a log-log scale for the three typical values of κ\kappa^{\prime}. (There are small differences in the pre-factors of the data sets, but they are not evident on the log-log scale.) The data sets exhibit an initial LAC growth regime t1/2\sim t^{1/2} characteristic of the Nm phase and then a crossover to a slower t1/4t^{1/4} growth regime. The dashed lines have been shown for reference. (Larger system sizes and longer simulation times will be required to remove the finite size effects observed at late times.) These observations emphasize the two-time-scale scenario identified in Fig. 6, and are the second novel aspect of our study.

V Summary and Discussion

We conclude with a summary and discussion of our results. We have undertaken a comprehensive numerical investigation of the equilibrium and non-equilibrium phenomena in the d=3d=3 Gay-Berne model. This model is known to exhibit isotropic (I), nematic (Nm), and smectic (Sm) phases and yields satisfactory comparisons with experimental observations in liquid crystal systems. There are two important parameters in this model: (i) the shape anisotropy parameter κ\kappa which is the length-to-breadth ratio of the ellipsoidal molecules; and (ii) the energy anisotropy parameter κ\kappa^{\prime} which is the ratio of energies when molecules are in the side-by-side (ss) and in the end-to-end (ee) configurations, see Fig. 1. In all our studies, we make a standard choice of κ=3.0\kappa=3.0 and vary κ\kappa^{\prime} over a wide range of values from 1.25 to 20. The primary focus of our work is to understand domain growth in the Sm mesophase. To the best of our knowledge, this is the first such study.

Equilibrium studies have been performed using canonical (NVTNVT) ensemble Monte Carlo (MC) simulations on systems with 512 and 1000 particles. Equilibration was achieved in approximately 5×1055\times 10^{5} MC cycles and the observations were made in the window of 5×1051065\times 10^{5}-10^{6} MC cycles. We confirmed the presence of I, Nm and Sm phases, and two distinct phase transitions: (i) I\rightarrowNm at Tc1T_{c}^{\ast 1} and (ii) Nm\rightarrowSm at Tc2<Tc1T_{c}^{\ast 2}<T_{c}^{\ast 1}. Our numerics indicate that Tc1T_{c}^{\ast 1} decreases substantially as κ\kappa^{\prime} increases, but Tc2T_{c}^{\ast 2} increases only slightly.

We studied the non-equilibrium phenomenon of coarsening via MD simulations of the Gay-Berne model with \sim260000 particles in the NVTNVT ensemble. An initially disordered and homogeneous state was rapidly quenched from (a) I\rightarrowNm and (b) I\rightarrowSm phases, see Fig. 2(d). The system was then allowed to evolve till late times, and we identified the morphology textures and growth laws through this evolution.
(a) Our results for the nematic quench (I\rightarrowNm) are as follows: the equal-time spatial correlation function C(r,t)C(r,t) exhibits dynamical scaling and is robust with respect to κ\kappa^{\prime}. Coarsening is due to annihilation of the string defects [S(k,t)k5S(k,t)\sim k^{-5}], and the domain growth obeys the Lifshitz-Allen-Cahn (LAC) law: L(t)t1/2L(t)\sim t^{1/2}. Earlier results in the literature were inconclusive due to small system sizes used in simulations Billeter et al. (1999).
(b) In our I\rightarrowSm quenches, the low temperature phase has a Sm B hexatic (SmB-H) order. With regard to our novel results on coarsening in this phase, the domain growth exhibits a two-time-scale scenario: With the onset of coarsening, the LC molecules align and develop orientational order (nematicity). The arrangement in layers (smecticity) with hexatic bond-orientational-order (BOO) within the layers follows thereafter. Consequently, the growth follows the LAC law, L(t)t1/2L(t)\sim t^{1/2} at early times and then a crossover to a slower growth at later times. Interestingly, the correlation functions and corresponding structure factors show dynamical scaling in both the regimes with universal scaling functions. These are also robust with respect to κ\kappa^{\prime}, and the smecticity and BOO are pronounced at larger values. Furthermore, the early-time dynamics is governed by string defects [S(k,t)k5S(k,t)\sim k^{-5}], while the late-time evolution is dictated by interfacial defects [Sl(k,t)k4S_{l}(k,t)\sim k^{-4} and SH(k,t)k4S_{H}(k,t)\sim k^{-4}]. We believe these results to be valid for other classes of the Sm phase as well.

In conclusion, we believe that the novel results presented in this paper reveal many unusual aspects of ordering in the Sm phase. Although of consequence in the LC arena, it has received little attention due to experimental and computational difficulties. The methodology presented in this work to study equilibrium and non-equilibrium phenomena in calamitic LCs could also be applied to study the mesophases that occur in discotic and bent-core LC systems. Another significant aspect of our work has been incorporation of the Gay-Berne potential in the general-purpose MC program, DL_MONTE DL_ . This new functionality should prove useful for future MC studies of uniaxial as well as biaxial and bent core LC systems.

Acknowledgements.
NB acknowledges UGC, India for a senior research fellowship. VB acknowledges DST, India for a core research grant. NB, TU, NBW and VB also acknowledge DST-UKIERI for a research grant which has facilitated this collaboration. NB and TU acknowledge Prof. Steve Parker at the University of Bath (UK) for the kind hospitality provided during the developmental stages of this work. NB and VB gratefully acknowledge the High Performance Computing (HPC) facility at IIT Delhi for computational resources.

References

  • Stephen and Straley (1974) M. J. Stephen and J. P. Straley, Rev. Mod. Phys. 46, 617 (1974).
  • de Gennes and Prost (1995) P. G. de Gennes and J. Prost, The Physics of Liquid Crystals (Oxford: Oxford University Press, 1995).
  • Singh (2000) S. Singh, Physics Reports 324, 107 (2000).
  • Priestly (2012) E. Priestly, Introduction to Liquid Crystals (Springer, 2012).
  • Andrienko (2018) D. Andrienko, Journal of Molecular Liquids 267, 520 (2018).
  • Moncton and Pindak (1979) D. E. Moncton and R. Pindak, Phys. Rev. Lett. 43, 701 (1979).
  • Goodby (1981) J. Goodby, Molecular Crystals and Liquid Crystals 72, 95 (1981).
  • Pindak et al. (1981) R. Pindak, D. E. Moncton, S. C. Davey,  and J. W. Goodby, Phys. Rev. Lett. 46, 1135 (1981).
  • Davey et al. (1984) S. C. Davey, J. Budai, J. W. Goodby, R. Pindak,  and D. E. Moncton, Phys. Rev. Lett. 53, 2129 (1984).
  • Voronov et al. (2020) V. P. Voronov, A. R. Muratov, S. N. Sulyanov, D. S. Molodenskiy, P. V. Dorovatovskii, V. V. Grebenev,  and B. I. Ostrovskii, Liquid Crystals 47, 1366 (2020).
  • McMillan (1971) W. L. McMillan, Phys. Rev. A 4, 1238 (1971).
  • Kralj et al. (2007) S. Kralj, G. Cordoyiannis, A. Zidanšek, G. Lahajnar, H. Amenitsch, S. Žumer,  and Z. Kutnjak, The Journal of Chemical Physics 127, 154905 (2007).
  • Coles and Strazielle (1979) H. J. Coles and C. Strazielle, Molecular Crystals and Liquid Crystals 55, 237 (1979).
  • Chaban et al. (2020) I. Chaban, C. Klieber, R. Busselez, K. A. Nelson,  and T. Pezeril, The Journal of Chemical Physics 152, 014202 (2020).
  • Chen et al. (2018) H.-W. Chen, J.-H. Lee, B. Y. Lin, S. Chen,  and S. T. Wu, Light: Sci. Appl. 7, 17168 (2018).
  • Li and de Jeu (2003) Li and W. H. de Jeu, Macromolecules 36, 4862 (2003).
  • Harrison et al. (2000) C. Harrison, D. H. Adamson, Z. Cheng, J. M. Sebastian, S. Sethuraman, D. A. Huse, R. A. Register,  and P. M. Chaikin, Science 290, 1558 (2000).
  • Harrison et al. (2002) C. Harrison, Z. Cheng, S. Sethuraman, D. A. Huse, P. M. Chaikin, D. A. Vega, J. M. Sebastian, R. A. Register,  and D. H. Adamson, Phys. Rev. E 66, 011706 (2002).
  • Ruiz et al. (2007) R. Ruiz, R. Sandstrom,  and C. Black, Advanced Materials 19, 587 (2007).
  • Stoykovich et al. (2007) M. P. Stoykovich, H. Kang, K. C. Daoulas, G. Liu, C.-C. Liu, J. J. de Pablo, M. Müller,  and P. F. Nealey, ACS Nano 1, 168 (2007).
  • Halperin and Nelson (1978) B. I. Halperin and D. R. Nelson, Phys. Rev. Lett. 41, 121 (1978).
  • Kleman and Friedel (2008) M. Kleman and J. Friedel, Rev. Mod. Phys. 80, 61 (2008).
  • Chuang et al. (1991) I. Chuang, R. Durrer, N. Turok,  and B. Yurke, Science 251, 1336 (1991).
  • Pargellis et al. (1991) A. Pargellis, N. Turok,  and B. Yurke, Phys. Rev. Lett. 67, 1570 (1991).
  • Lavrentovich et al. (2001) O. D. Lavrentovich, P. Pasini, C. Zannoni,  and S. Žumer, Defects in Liquid Crystals: Computer Simulations, Theory and Experiments (Springer, 2001).
  • Wang et al. (2016) X. Wang, D. S. Miller, E. Bukusoglu, J. J. de Pablo,  and N. L. Abbott, Nature Materials 15, 106 (2016).
  • Kim et al. (2018) D. S. Kim, S. Copar, U. Tkalec,  and D. K. Yoon, Science Advances 4 (2018).
  • Sandford O’Neill et al. (2020) J. J. Sandford O’Neill, P. S. Salter, M. J. Booth, S. J. Elston,  and S. M. Morris, Nature Communications 11, 2203 (2020).
  • Berne and Pechukas (1972) B. J. Berne and P. Pechukas, The Journal of Chemical Physics 56, 4213 (1972).
  • Gay and Berne (1981) J. G. Gay and B. J. Berne, The Journal of Chemical Physics 74, 3316 (1981).
  • Luckhurst and Simmonds (1993) G. Luckhurst and P. Simmonds, Molecular Physics 80, 233 (1993).
  • Berardi et al. (1993) R. Berardi, A. P. J. Emerson,  and C. Zannoni, J. Chem. Soc., Faraday Trans. 89, 4069 (1993).
  • Bates and Luckhurst (1999) M. A. Bates and G. R. Luckhurst, The Journal of Chemical Physics 110, 7087 (1999).
  • Emerson et al. (1994) A. Emerson, G. Luckhurst,  and S. Whatling, Molecular Physics 82, 113 (1994).
  • Cienega-Cacerez et al. (2014) O. Cienega-Cacerez, J. A. Moreno-Razo, E. Díaz-Herrera,  and E. J. Sambriski, Soft Matter 10, 3171 (2014).
  • Adams et al. (1987) D. Adams, G. Luckhurst,  and R. Phippen, Molecular Physics 61, 1575 (1987).
  • Luckhurst et al. (1990) G. R. Luckhurst, R. A. Stephens,  and R. W. Phippen, Liquid Crystals 8, 451 (1990).
  • Miguel et al. (1991) E. D. Miguel, L. F. Rull, M. K. Chalam,  and K. E. Gubbins, Molecular Physics 74, 405 (1991).
  • Hashim et al. (1995) R. Hashim, G. R. Luckhurst,  and S. Romano, J. Chem. Soc., Faraday Trans. 91, 2141 (1995).
  • Satoh et al. (1996) K. Satoh, S. Mita,  and S. Kondo, Liquid Crystals 20, 757 (1996).
  • La Penna et al. (1996) G. La Penna, D. Catalano,  and C. A. Veracini, The Journal of Chemical Physics 105, 7097 (1996).
  • Cleaver et al. (1996) D. J. Cleaver, C. M. Care, M. P. Allen,  and M. P. Neal, Phys. Rev. E 54, 559 (1996).
  • Neal et al. (1997) M. P. Neal, A. J. Parker,  and C. M. Care, Molecular Physics 91, 603 (1997).
  • de Miguel and Vega (2002) E. de Miguel and C. Vega, The Journal of Chemical Physics 117, 6313 (2002).
  • Zannoni (2001) C. Zannoni, J. Mater. Chem. 11, 2637 (2001).
  • Martínez-Haya and Cuetos (2007) B. Martínez-Haya and A. Cuetos, The Journal of Physical Chemistry B 111, 8150 (2007).
  • Bray (2002) A. J. Bray, Advances in Physics 51, 481 (2002).
  • Puri (2004) S. Puri, Phase Transitions 77, 407 (2004).
  • Puri and Wadhawan (2009) S. Puri and V. Wadhawan, Kinetics of phase transitions (CRC Press, Boca Raton, 2009).
  • Bray et al. (1993) A. J. Bray, S. Puri, R. E. Blundell,  and A. M. Somoza, Phys. Rev. E 47, R2261 (1993).
  • Wickham (1997) R. A. Wickham, Phys. Rev. E 56, 6843 (1997).
  • Blundell and Bray (1992) R. E. Blundell and A. J. Bray, Phys. Rev. A 46, R6154 (1992).
  • Zhang et al. (1993) Z. Zhang, O. G. Mouritsen,  and M. J. Zuckermann, Phys. Rev. E 48, 2842 (1993).
  • Bradac et al. (2011) Z. Bradac, S. Kralj,  and S. Zumer, The Journal of Chemical Physics 135, 024506 (2011).
  • Birdi et al. (2020) N. Birdi, V. Banerjee,  and S. Puri, EPL (Europhysics Letters) 132, 66002 (2020).
  • Allen and Cahn (1979) S. M. Allen and J. W. Cahn, Acta Metallurgica 27, 1085 (1979).
  • Abukhdeir and Rey (2008) N. M. Abukhdeir and A. D. Rey, New Journal of Physics 10, 063025 (2008).
  • Liarte et al. (2015) D. B. Liarte, M. Bierbaum, M. Zhang, B. D. Leahy, I. Cohen,  and J. P. Sethna, Phys. Rev. E 92, 062511 (2015).
  • Purton et al. (2013) J. Purton, J. Crabtree,  and S. Parker, Molecular Simulation 39, 1240 (2013).
  • Brukhno et al. (2021) A. V. Brukhno, J. Grant, T. L. Underwood, K. Stratford, S. C. Parker, J. A. Purton,  and N. B. Wilding, Molecular Simulation 47, 131 (2021).
  • (61) https://gitlab.com/dl_monte.
  • Frenkel and Smit (2002) D. Frenkel and B. Smit, Understanding Molecular Simulation: From Algorithms to Applications, 2nd ed. (Academic Press, San Diego, 2002).
  • Callan-Jones et al. (2006) A. C. Callan-Jones, R. A. Pelcovits, V. A. Slavin, S. Zhang, D. H. Laidlaw,  and G. B. Loriot, Phys. Rev. E 74, 061701 (2006).
  • Billeter and Pelcovits (1998) J. Billeter and R. Pelcovits, Computers in Physics 12, 440 (1998).
  • de Miguel et al. (1996) E. de Miguel, E. Martín del Rio, J. T. Brown,  and M. P. Allen, The Journal of Chemical Physics 105, 4234 (1996).
  • Billeter et al. (1999) J. L. Billeter, A. M. Smondyrev, G. B. Loriot,  and R. A. Pelcovits, Phys. Rev. E 60, 6831 (1999).
  • Plimpton (1995) S. Plimpton, Journal of Computational Physics 117, 1 (1995).
  • (68) https://lammps.sandia.gov.
  • Note (1) Note that LAMMPS technically implements the generalized Gay-Berne model Berardi_1995 in which interactions of the ellipsoidal particles are characterized via a (diagonal) shape matrix S=S= diag(σa,σb,σc\sigma_{a},\sigma_{b},\sigma_{c}) and a (diagonal) energy matrix E=E= diag(ϵa,ϵb,ϵc\epsilon_{a},\epsilon_{b},\epsilon_{c}), where σa,σb,σc\sigma_{a},\sigma_{b},\sigma_{c} are the lengths and ϵa,ϵb,ϵc\epsilon_{a},\epsilon_{b},\epsilon_{c} are the relative well depths of interaction along the three semi-axes of an ellipsoid. The GB model discussed in Sec. II and employed in Sec. III was the potential originally presented for ellipsoidal particles of equal size. The generalized GB model was developed to represent dissimilar biaxial ellipsoids Berardi_1995. It reduces to the GB model described in Sec. II when the molecules become uniaxial. Hence, for our study, we work with the generalized GB model in the uniaxial limit.
  • Brown et al. (2009) W. M. Brown, M. K. Petersen, S. J. Plimpton,  and G. S. Grest, The Journal of Chemical Physics 130, 044901 (2009).
  • Nosé (1984) S. Nosé, The Journal of Chemical Physics 81, 511 (1984).
  • Binder et al. (2004) K. Binder, J. Horbach, W. Kob, W. Paul,  and F. Varnik, Journal of Physics: Condensed Matter 16, S429 (2004).
  • Porod et al. (1982) G. Porod, O. Glatter,  and O. Kratky, by O. Glatter and O. Kratky, Academic Press, London , 17 (1982).
  • Oono and Puri (1988) Y. Oono and S. Puri, Mod. Phys. Lett. B 2, 861 (1988).
  • Bray and Puri (1991) A. J. Bray and S. Puri, Phys. Rev. Lett. 67, 2670 (1991).
  • Richter and Gruhn (2006) A. Richter and T. Gruhn, The Journal of Chemical Physics 125, 064908 (2006).
  • Cañeda-Guzmán et al. (2014) E. Cañeda-Guzmán, J. Moreno-Razo, E. Díaz-Herrera,  and E. Sambriski, Molecular Physics 112, 1149 (2014).
Refer to caption
Figure 1: (a) Variation of the scaled GB energy between a pair of ellipsoidal molecules as a function of their scaled separation (Eij/ϵ0E_{ij}/\epsilon_{0} vs. rij/σ0r_{ij}/\sigma_{0}) in the ee configurations with κ=3.0\kappa=3.0, μ=1.0\mu=1.0, ν=3.0\nu=3.0 for energy anisotropy κ\kappa^{\prime} = 2.5 (red line), 5.0 (blue line) and 10.0 (green line). The inset shows the corresponding variation in the ss configuration. Notice that the curves for different values of κ\kappa^{\prime} overlap. It can be observed that the energy anisotropy parameter affects the well depth for the ee configuration but not for the the ss configuration. (b) Similar variation for μ,ν\mu,\nu = 1.0, 3.0 (blue lines) and 2.0, 1.0 (magenta lines) with κ=3.0\kappa=3.0, κ=5.0\kappa^{\prime}=5.0. The exponents μ\mu and ν\nu affect the well depth in the ee as well as the ss configurations.
Refer to caption
Figure 2: (a) Variation of the scaled per-particle energy for the GB model system, E1\langle E_{1}^{\ast}\rangle with respect to the scaled temperature TT^{\ast} with κ=3.0\kappa=3.0, κ=5.0\kappa^{\prime}=5.0, μ=1.0\mu=1.0 and ν=3.0\nu=3.0 for the two system sizes N=512N=512 (open up triangles) and N=1000N=1000 (open down triangles). Corresponding results from Berardi et al. Berardi et al. (1993) for N=512N=512 (white circles), and for N=1000N=1000 during heating (white up triangles) and cooling (white down triangles) protocols are also shown for comparison. (b) Variation of 𝒮\langle\mathcal{S}\rangle (scalar nematic), 𝒯\langle\mathcal{T}\rangle (translational) and 𝒞6\langle\mathcal{C}_{6}\rangle (hexatic) order parameters with respect to TT^{\ast} at κ=5.0\kappa^{\prime}=5.0 for the N=1000N=1000 system. (c) Radial distribution function g(r)g(r^{\ast}) vs. rr^{\ast} at TT^{\ast} = 1.0 (Sm phase) and TT^{\ast} = 2.5 ( Nm phase) for κ=5.0\kappa^{\prime}=5.0 and N=1000N=1000. (d) Phase diagram in the TκT^{\ast}-\kappa^{\prime} plane indicating the isotropic (I) phase (yellow), nematic (Nm) phase (green) and Sm phase (blue). The asterisks and crosses indicate the quenches in the Nm and Sm regimes to study coarsening. The arrows indicate two typical quenches from I\rightarrowNm and I\rightarrowSm.
Refer to caption
Figure 3: Evolution snapshots after after a quench from the I phase to the Nm phase (T=2.5T^{\ast}=2.5) for different values of tt and κ\kappa^{\prime}: (a) κ=5.0\kappa^{\prime}=5.0, t=8192t=8192, (b) κ=5.0\kappa^{\prime}=5.0, t=53248t=53248 and (c) κ=10.0\kappa^{\prime}=10.0, t=53248t=53248. Left: 20320^{3} corner portions of the entire simulation box allowing for visibility of individual particles. Right: the corresponding top d=2d=2 cross-sections.
Refer to caption
Figure 4: (a) Scaled correlation functions C(r,t)C(r,t) vs. r/L(t)r/L(t) for the Nm phase at the specified values of κ\kappa^{\prime} and tt. The data collapse indicates that the system exhibits dynamical scaling, and the morphologies are invariant with respect to κ\kappa^{\prime} and tt apart from a scale factor. (b) Corresponding scaled structure factors S(k,t)L(t)3S(k,t)L(t)^{-3} vs. kL(t)kL(t) on a log-log plot. The dashed line of slope -5 denotes the expected generalized Porod tail. (c) Domain length scale L(t)L(t) vs. tt for different values of κ\kappa^{\prime}. The growth obeys the LAC law, L(t)t1/2L(t)\sim t^{1/2} (dashed line) characteristic of systems with non-conserved order parameter and is slower for larger values of κ\kappa^{\prime}.
Refer to caption
Figure 5: Evolution snapshots after quenches from the I to the Sm phase [T=1.0T^{\ast}=1.0, see Fig.2(d)] for (a) κ=5.0\kappa^{\prime}=5.0, t=4096t=4096; (b) κ=5.0\kappa^{\prime}=5.0, t=98304t=98304 and (c) κ=10.0\kappa^{\prime}=10.0, t=98304t=98304. Left: 20320^{3} corner portions of the entire simulation box allowing for visibility of individual particles. Right: the corresponding top d=2d=2 cross-sections. Notice that smecticity is enhanced by larger values of κ\kappa^{\prime}. The layers are interdigitated and within a layer, the particles exhibit hexagonal order.
Refer to caption
Figure 6: (a) Time evolution of the longitudinal radial distribution functions g(r/σ0)g_{\parallel}(r_{\parallel}/\sigma_{0}) vs. (r/σ0r_{\parallel}/\sigma_{0}) for the I\rightarrowSm quench at the specified values of κ\kappa^{\prime} and tt signaling layering in the system which occurs only at later times (t104t\gtrsim 10^{4}). As time evolves, subsequent peaks emerge at separations corresponding to multiples of the ellipsoid length (3σ03\sigma_{0}) indicating ee alignment for the LC molecules (the separations between the peaks is 3σ0\lesssim 3\sigma_{0} due to interdigitation of the layers). With increase in κ\kappa^{\prime}, the intensity of the first peak increases and the variation between the intensities for the first and second peaks decreases. Hence, the degree of smecticity increases as κ\kappa^{\prime} increases. (b) Time evolution of the corresponding transversal radial distribution functions g(r/σ0)g_{\perp}(r_{\perp}/\sigma_{0}) vs. (r/σ0r_{\perp}/\sigma_{0}). These functions indicate alignment within the layers. The peaks in these functions indicate SmB-H ordering, see corresponding text for details.
Refer to caption
Figure 7: Scaled correlation functions for the SmB-H phase at specified values of κ\kappa^{\prime} and tt: (a) C(r,t)C(r,t) vs. r/L(t)r/L(t), (b) Cl(r,t)C_{l}(r,t) vs. r/Ll(t)r/L_{l}(t) and (c) CH(r,t)C_{H}(r,t) vs. r/LH(t)r/L_{H}(t) (see text for evaluation of these correlation functions). The data collapse in all the three evaluations suggests that the system exhibits dynamical scaling, and the morphologies are invariant with respect to tt and κ\kappa^{\prime} although the coarsening scenario is distinct at early and late times. The corresponding structure factors are shown on a log-log plot: (d) S(k,t)L(t)3S(k,t)L(t)^{-3} vs. kL(t)kL(t), (e) Sl(k,t)Ll(t)3S_{l}(k,t)L_{l}(t)^{-3} vs. kLl(t)kL_{l}(t) and (f) SH(k,t)LH(t)3S_{H}(k,t)L_{H}(t)^{-3} vs. kLH(t)kL_{H}(t). The dashed line of slope -5 in (d) indicates the Porod tail characteristic of scattering off string defects. Similarly, the dashed lines with slope -4 in (e) and (f) indicate interfacial scattering.
Refer to caption
Figure 8: Growth laws (log-log scale) for the SmB-H phase at different values of κ\kappa^{\prime}: (a) L(t)L(t) vs. tt, (b) Ll(t)L_{l}(t) vs. tt and (c) LH(t)L_{H}(t) vs. tt. Dashed lines of slope 1/2 and 1/4 are a guide to the eye. Early time data obeys the LAC law. The data indicates crossover to a slower growth at late times and hints at t1/4\sim t^{1/4} growth. The two-time-scale scenario seen is reconfirmed by these distinct evaluations.