Complex Dynamics of Photo-Switchable Guest Molecules in All

13 hours ago - We study theoretically the kinetics of non-interacting photo-switchable guest molecules (model azo-dye) dispersed at low concentration ...
0 downloads 17 Views 2MB Size
Subscriber access provided by READING UNIV

Article

Complex Dynamics of Photo-Switchable Guest Molecules in AllOptical Poling Close to Glass Transition: Kinetic Monte Carlo Modelling Wojciech Radosz, Grzegorz Pawlik, and Antoni Mitus J. Phys. Chem. B, Just Accepted Manuscript • DOI: 10.1021/acs.jpcb.7b11949 • Publication Date (Web): 15 Jan 2018 Downloaded from http://pubs.acs.org on January 16, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry B is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Complex Dynamics of Photo-Switchable Guest Molecules in All-Optical Poling Close to Glass Transition: Kinetic Monte Carlo Modelling W. Radosz, G. Pawlik, and A.C. Mitus∗ Department of Theoretical Physics, Wroclaw University of Science and Technology, Wybrzeze Wyspianskiego 27, 50–370 Wroclaw, Poland E-mail: [email protected]

1

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Abstract We study theoretically the kinetics of non-interacting photo-switchable guest molecules (model azo-dye) dispersed at low concentration in host (model polymer matrix) in the all-optical poling process close to the glass transition temperature Tg . We modify kinetic Monte Carlo model used in our previous studies of nonlinear optical processes in host-guest systems. The polymer matrix is simulated using the bond-fluctuation model. The kinetics of multiple trans-cis-trans cycles is formulated in terms of transition probabilities which depend on local free volume in the matrix and its dynamics. Close to Tg the build-up of polar order, monitored in terms of angular probability density functions, follows a power-law in time while the evolution of the nonlinear susceptibilities related to second harmonic generation effect follows the stretched-exponential law. This complex dynamics of guest molecules implies the presence of dynamic heterogeneities of the matrix in space and time which spread the complexity from the matrix to the otherwise simple dynamics of non-interacting guest molecules. A qualitative physical picture of mosaic-like states – intertwined areas of free- and hindered angular motion of guest molecules – is proposed and the role of related short and longer scales in space for the promotion of complex dynamics of guest molecules is discussed. A brief comparison of the theory to available experimental data is given.

1

Introduction: physical picture

Host-guest systems composed of a host matrix (polymer, biopolymer or liquid crystal) with chemically or physically attached guest molecules (chromophores, dipoles, octupoles, coated spheres and others) are commonly used in nonlinear optical (NLO) applications and techniques like, e.g., second (SHG) or third harmonic generation, sum and difference frequency generation, all-optical poling, four-wave mixing, self focusing, self-phase modulation, and many others. 1–3 A wide spectrum of those effects, which offer numerous applications in modern material sciences, makes both experimental and theoretical studies of host-guests system

2

ACS Paragon Plus Environment

Page 2 of 32

Page 3 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

of key importance. In this paper we model theoretically one of those effects – SHG in a system of photoswitchable chromophores (azo-dyes) dispersed in a polymer matrix. The necessary condition for this effect is a build-up of a non-vanishing global polar order for the distribution of long axes of azo-dyes in the space, which manifests itself in non-zero macroscopic secondorder nonlinear susceptibilities. Polar order can be promoted either by an external electric field poling, photo-assisted electric field poling or by all-optical poling techniques. 4 While the electric-field driven build-up was studied theoretically in our paper, 5 current study is related to the latter method. In the all-optical poling experiment the chromophores interact with two linearly polarized light beams (with frequencies ω and 2 ω) which result, firstly, in multiple one- and two-photon-driven photoisomerization cycles trans-cis-trans of azo-dyes and, secondly, in forming a constant electric field acting on the dipolar moments of the choromphores. 4 The former effect promotes axial symmetry, the latter – polar symmetry. This simple physical picture of light-matter interaction-driven build-up of polar order is fundamentally modified by the equilibrium thermal fluctuations of the local structure of the host matrix in the nano-scale, typical for the size of, say, Disperse Red 1 (DR1) azo-benzene molecule. 4 Unluckilly, little is known, on the quantitative level, about this kind of local structure of a polymer matrix and about its spatial and temporal correlations. The main challenge is the characterization of static local inhomogeneities in the distribution of the monomers and their dynamic evolution, driven by local jamming/clustering effects, which modify the dynamics of the guest molecules. The quantification of those topics is in its very early stage. 6,7 Local jamming/clustering events constitute the microscopic origin of complex macroscopic polymer dynamics, particularly well expressed in the close vicinity of the glass transition (GT) temperature Tg . The slowing down of polymer dynamics 8,9 is usually ascribed to an appearance of long-tailed (power-law) distributions of waiting times for local structural changes, triggered by jamming effects which ultimately cause dynamic heterogeneities 10–12 in

3

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

space and time (this scenario occurs for two-dimensional Lennard-Jones liquid 13 ). Thus, the dynamics of nonlinear optical phenomena in host-guest systems close to Tg belongs to the class of complex systems’ dynamics and its theoretical analysis, which encounters methodological difficulties, constitutes a challenge. The objective of this paper is to study theoretically the influence of glass transition in the host polymer matrix on the dynamics of guest (azo-dye) molecules dispersed at low concentration, in a model host-guest system mimicking all-optical poling experiment. Host-guest systems can be studied theoretically using kinetic Monte Carlo (MC) modelling, combining standard MC bond fluctuation simulations of a polymer matrix with suitable kinetic models which account for the interaction of guest molecules with the matrix and with external fields as well as for their intermolecular interactions. An advantage of this kind of modelling is that it grants a deeper insight into ”microscopic” processes promoting macroscopic effects. More specifically, it makes possible a detailed characterization of ”microscopic” orientational order of guest molecules and its temporal evolution which is hard (if at all possible!) to infer indirectly from experimental data. This approach was used by some of us to study chosen NLO phenomena in host-guest systems with photochromic and dipolar molecules serving as guest molecules: inscription and erasure of diffraction gratings, 14–16 diffraction gratings in DNA based biomaterials, 17 SHG in electrically poled systems, 5 inscription of surface relief gratings 18 and photoinduced reorientation of nematic liquid crystals, 19,20 see also review paper. 21 Those studies were performed for temperatures below Tg ; moreover, no systematic analysis of temperature effects was done. In this study we use the same approach with a suitably generalized MC kinetic model onto the case of all-optical poling. Quite recently kinetic Monte Carlo simulations 22 were used to study the dynamics of needle-like photo-switchable molecules on a surface, motivated by a power-law decay of initial birefringence in a dense monolayer of photo-switchable dye methyl-red molecules demonstrated experimentally. 23 Authors report the complex dynamics of the molecules in

4

ACS Paragon Plus Environment

Page 4 of 32

Page 5 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

low-temperature and high density limit, traced to dynamic heterogeneity originating from intermolecular interactions; the polymer matrix is not present in their model. The paper is organized as follows. In the next chapter we present the basic features of MC modelling of all-optical poling in a host-guest system. The third chapter is devoted to the characterization of the process of build-up of polar order in this system in a wide temperature interval around Tg . The discussion of complex kinetics of guest molecules close to Tg constitutes the main part of the fourth chapter in which the topics of the localization of the glass transition temperature and the relation of our results to experimental data are also briefly discussed.

2

MC modelling of all-optical poling

Monte Carlo modelling of all optical poling in host-guest systems addresses three topics: (i), characterization of orientational order of guest azo-dyes molecules and its relation to macroscopic SHG susceptibilities; (ii), MC simulation of equilibrium polymer matrix and an analysis of local parameters which condition the local dynamics of azo-dyes and (iii), the kinetic model for the interaction of azo-dyes with optical field and with the matrix. Those topics are discussed in the next subsections.

2.1

Polar order and nonlinear susceptibilities

Efficiency of the SHG in poled thin films consisting of quasi one-dimensional (1D) charge transfer chromophores in a polymer matrix depends on the degree of polar order. For such quasi 1D rod-like molecule the orientation of its long axis is described by angle Ω = (θ, φ), where θ and φ denote, respectively, the polar and azimuthal angle defined wrt. an external direction Z being the poling field direction (linearly polarized poling/probing optical field perpendicular to the thin film surface), see Fig. 1. The distribution of orientations of long axes of the molecules is given by the normalized probability density function (PDF) ρ(Ω).

5

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 32

w

Z q

2w

Y

X

Figure 1: Schema of SHG model setup.

This PDF makes possible the calculation of the statistical averages h...i of an arbitrary function f (θ, φ): Z hf (θ, φ)i =



Z

Z

1

f (θ, φ) ρ(θ, φ) d cos θ.



f (θ, φ) ρ(Ω)dΩ = 0

(1)

−1

The second order NLO susceptibility χ(2) for a set of rod-like molecules is characterized by two tensor components: 24,25 the diagonal

(2)

χZZZ (−2ω; ω, ω) = N F βzzz hcos3 θi,

(2)

1 (2) χXXZ (−2ω; ω, ω) = N F βzzz hsin2 θ cos θi. 2

(3)

and the off diagonal one

Here, X denotes a direction perpendicular to Z, N – number density of azo-dyes in trans state, βzzz – the dominant component of hyperpolarizability tensor for a rod-like molecule in trans state (molecular first hyperpolarizability), and F – the local field factor. Monte Carlo modelling yields the PDF ρ(θ, φ; t), where t denotes the Monte Carlo ”time” – number of Monte Carlo Steps, see below – which fully characterizes the dynamics of build(2)

(2)

up of polar order and of NLO susceptibilities χZZZ , χXXZ .

6

ACS Paragon Plus Environment

Page 7 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

2.2

MC modelling of polymer host

We have used the same standard MC lattice bond-fluctuation model of the equilibrium polymer matrix 26 as in our previous studies of host-guest systems (Sect. 1). The polymer chain consisted of a set of beads (monomers) located on a simple cubic (s.c.) lattice, connected √ √ √ by bonds with lengths 2, 5, 6, 3, 3, 10 (in units of lattice constant); six non-equivalent bond orientations were used. The dimensionless energies  of the bonds are  = 1 for the first three cases, and  = 0 for the last three lengths. N0 = 24000 polymer chains, each consisting of L = 20 monomers, were placed on a Vp = 200 × 200 × 200 lattice; the reduced density ρ = 8 N0 L/Vp = 0.48 corresponds to the concentrated polymer solution. 26 In spite of its simplicity, this model plays one of central roles in modelling of polymer systems. 27

Figure 2: A snapshot of a polymer system (10% of all chains are displayed) (a) and scheme of calculation of parameter V (b).

Simulations were made at constant reduced temperature T (its unit is related to the dimensionless energy scale ). A single Monte Carlo Step (MCS) consisted of trial moves of unit length of all the monomers along one of the three axes x, y, z of s.c. lattice. Each move was accepted if: (a), the trial location (lattice node) was empty; (b), the lengths of emerging new bonds were acceptable; and (c), the standard Metropolis acceptance criterion 26 7

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

was satisfied. The starting configuration consisted of isotropically oriented bonds. Periodic boundary conditions were used. A typical instantaneous configuration of the polymer system is shown in Fig. 2 (a).

Figure 3: (color online) Instantaneous 2D cross-section of the polymer system: map of V (large map) and part of this map scaled up (small map).

Local static and dynamic features of the polymer matrix which affect the dynamics of the guest molecules are characterized by two parameters, V (~r) and C (~r), introduced in our earlier papers, 15,16 see also in. 21 Those quantities enter the expressions for the transition rates in the kinetic model described in Sect. 2.4. The first one, V (~r), describes the local inhomogeneities (local voids) of the distribution of the monomers around lattice point ~r. The value of V (~r) is calculated using a 3 × 3 × 3 cube, centered around ~r (Fig. 2 (b)). If the distance of a monomer from the lattice node ~r is 0 or 1, then V = 0. In other cases V = 7 − k, where 0 ≤ k ≤ 7 stands for the number of monomers in the cube. High values of V indicate that there are few monomers in a close vicinity of point ~r; low values imply that its closest environment is crowded with monomers. An eye-inspection of instantaneous maps of parameter V (~r) (Fig. 3) reveals a spatial heterogeneity in its distribution. Clearly, some complex spatial correlations are present, nevertheless the pattern is too complicated to draw, at this point, any far-reaching quantitative conclusions. The characterization of 8

ACS Paragon Plus Environment

Page 8 of 32

Page 9 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

local static and dynamic heterogeneity constitutes one of the challenges for the statistical physics of model polymer systems. It amounts to the characterization of the random spatiotemporal fields V (~r, t) and C(~r, t) (see below), which requires the introduction of various kinds of probability density functions and correlation functions. This task goes far beyond the scope of current paper. Some time independent correlation functions which quantify the concept of voids were calculated in Ref. 6,7 The second parameter, C (~r), describes a short-time reorganization of the monomers in a local neighbourhood of point ~r and affects the rotational dynamics (rotational mobility) of a guest molecule at this point, see Sect. 2.4. Its value is equal 15,16,21 to the total number of changes δni (in a single MCS) of the occupation state ni of a node by the monomers (ni = 1 P for site i occupied by a monomer; ni = 0 when site i is monomer-free): C (~r) = i |δni |. The summation extends over the nodes of a 9 × 9 × 9 cube centered around the investigated cell ~r. The large size of the cube is dictated by the fact that monomers are relatively sparse - in our simulations one monomer is localized, on the average, in a volume composed of 16 lattice sites.

2.3

Glass transition temperature Tg

For further analysis an estimation of the glass transition temperature Tg is necessary. Its localization is more ambiguous than for equilibrium I- or II-order phase transitions, where thermodynamics plays an important role. 28,29 Some parameters, mostly geometric in nature, but also of thermodynamic origin, like mean bond energy , Flory parameter, specific heat, or internal temperature, were proposed and studied for the bond-fluctuation model in the seminal paper. 30 Those parameters clearly mark the glass transition but the values of Tg are not localized in a narrow temperature interval. The geometric parameters studied were: mean squared bond length, radius of gyration or the mean free volume (MFV) parameter 30 v, originally introduced for 2D systems. 31 Similar in spirit, approach using dynamically available volume, 32 was used in Ref. 33 MFV parameter 9

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

v (0 ≤ v ≤ 1) is equal to the normalized average (over polymer chains and configurations) number of lattice sites available for monomers in a single MC step. The lattice site is available if the displacement of a monomer is not blocked and the emerging new bond length is acceptable. Fig. 4 shows the temperature dependence of v. Three regimes are present: linear at low (T ≤ 0.15) and linear at high (T ≥ 0.35) temperatures and a transient regime 0.15 < T < 0.35. Linear regimes correspond, respectively, to glassy and melt polymer phases; the crossover between them takes place in the transient regime. Clearly, no precise estimation of Tg is possible. Some estimation can be obtained using the linear fits; their intersection yields Tg ≈ 0.25. Similar conclusions are drawn from the plot of average bond energy (T ); we find Tg ≈ 0.23. On the other hand, the temperature dependence of the radius of gyration, Fig. 4 (inset), displays a well-defined crossover between linear regimes at T = 0.26. To conclude, we estimate that Tg lies in the interval 0.23 ≤ Tg ≤ 0.26.

Figure 4: (color online) Plot of MFV parameter v (black points, left axis) and of mean bond energy  (blue circles, right axis) against reduced temperature T . Inset: plot of mean squared radius of gyration of polymer chain against T . Straight lines represent linear fits.

10

ACS Paragon Plus Environment

Page 10 of 32

Page 11 of 32

cis R N

)

p(c is

R

ns ) tra

p(trans ®

®

ci s

N

R

DW

R

N N

angular diffusion

R

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

N N

trans

R

Figure 5: Illustration of trans-cis-trans cycle and angular diffusion process.

2.4

Kinetic model for all-optical poling in low concentration limit

The kinetic MC modelling uses analytical expressions for transition probabilities p in unit time (transition rates) for various constituent processes. For host-guest systems the kinetic model accounts for the interaction of azo-dyes with an optical field (light-matter interaction), with the matrix as well as for intermolecular dye-dye interactions. This study concerns the case of a low concentration of the azo molecules, leading to a simplification of the general kinetic model. Namely, both static (excluded volume) and dynamic (photoisomerization cycles) effects of the model azodyes on the polymer matrix are neglected. In other words, we study a system of noninteracting azo molecules in dynamical external field (due to optical field and the matrix) which is independent on the dynamics of those molecules. The motivation for this approximation is as follows. We are interested in the ”transfer” of complexity from the matrix close to glass transition onto a system of noninteracting azodyes which, in the absence of the matrix, display a simple (bi-exponential)

11

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

dynamics. 14 In this context, the concentration of the azodyes has to be sufficiently low to avoid steric and electrostatic dye-dye interactions. Those interactions become important for sufficiently high concentration of the dyes and lead to their complex dynamics, even in the absence of the matrix, as shown experimentally 23 and through simulations. 22 So, in a general case, there are two very different origins/mechanisms of complex dynamics of the guest molecules. While the ultimate goal might be to study a general case, where both mechanisms contribute, in various ways, to the overall complex dynamics, in this paper we study only the first one (matrix induced) – the second mechanism (interaction induced in the absence of the matrix), was studied theoretically. 22 The validity of the assumedly weak impact of azo molecules on the matrix in the low concentration case follows from an analysis of SHG experiments in host-guest systems. Consider, e.g., SGH in doped PMMA/DCNP system 34 with 2 wt % concentration of DCNP molecules. The molar masses of DCNP and of PMMA repeat unit (r.u.) are, respectively, 222.25 g/mol and 100.12 g/mol. 2 wt % concentration corresponds to 1 DCNP molecule per 109 PMMA r.u. PMMA has 8.8 r.u. per Kuhn segment 35 and thus a single DCNP molecule corresponds to 12.4 Kuhn segments. In our simulations the chain consists of 20 Kuhn segments which yields approximately 1.6 azo molecules per modelled chain. Actually, SHG signal can be measured at 1 wt % or even lower concentrations, which corresponds in our simulations to less than a single azo molecule per chain. Those data give a sound ground to assume that the effect of the azo molecules on the matrix can be weak/negligible while SHG signal can be still monitored. The kinetic MC modelling in the case of pure light-matter interaction was studied by us some time ago. 14 The transition rates in the presence of the matrix, in the low concentration limit when dye-dye intermolecular interactions and the impact of the isomerization dynamics on the matrix are negligible were proposed in our papers. 15,16 Some applications of MC kinetic modelling to other physical systems are reviewed in Ref. 36 In this paper we generalize the transition rates 15,16 onto the case of all–optical poling in low concentration limit.

12

ACS Paragon Plus Environment

Page 12 of 32

Page 13 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

The light-driven trans-cis-trans cycle of azo-dye molecules consists of three events: transitions trans → cis, cis → trans and angular diffusion (see Fig. 5). The transition rate for one-photon trans → cis transition p(θ, φ) ∝ cos2 θ is independent of angle φ. In the case of all-optical poling additional transition rates have to be accounted for, 4 which correspond to (i), two-photon absorption, where p(θ) ∝ cos4 θ, and (ii), the presence of electric field which breaks the centrosymmetry: p(θ) ∝ cos3 θ. An overall transition rate is a weighted sum of those rates. Upon optimization requirement (equal excitation probabilities through one- and two-photon absorptions) the rates p(trans → cis) and p(cis → trans) read: 4

p(trans → cis) = V I ptc (cos2 θ + cos4 θ + 2 cos3 θ),

(4)

p(cis → trans) = V I pct ,

(5)

where I – light intensity, ptc , pct denote the probabilities of photoisomerization in a single act of interaction with light. The parameter R = ptc /pct has a strong impact on the kinetics of host-guest systems. 17,22,37 The transition rate for rotational diffusion reads: 15,16    pdif f V C , ∆Ω < ∆Ω0 p(Ω → Ω + ∆Ω) =   0 , ∆Ω > ∆Ω0 ,

(6)

where pdif f is a constant. Eqs. (4)-(6) imply that the azo-dye molecules couple to the polymer matrix by the factors V and C. In the simulations we have used the values pct = 0.0001, ptc = 0.1 pct and pdif f = 0.005 (as in our previous papers). We assume 14,38 that the orientation of a long axis of a trans molecule after cis → trans transition can be chosen at random. All-optical poling process consists of two periods: the writing period of length 2 · 105 − 5 · 105 MCS when the intensity of linearly polarized light is constant and the decay period performed in darkness (I = 0, 106 MCS), where the dynamics is due only to the angular diffusion. The detailed structure and size of the model azo molecules are not relevant because in the

13

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

low concentration limit they do not influence the dynamics of the matrix (and, correspondingly, no interactions are included in the Metropolis conditions). The azo molecules react to the local fields (optical and matrix-induced), are fixed in space and occupy in a random way the nodes of the same s.c. lattice on which the monomers are located. This choice follows from the following argumentation. Since in the low concentration limit azo molecules do not influence the dynamics of the matrix, the statistics of the data for dynamics of the azo molecules can be increased by a simple trick: for a single instantaneous configuration of the matrix the azo molecules can be safely placed in all nodes of the lattice. Consequently, in the limit of low density the only effect of the translation in the space would be to probe another node. Since all nodes are probed, as explained above, the translation becomes inessential in this case.

3 3.1

Results Onset and evolution of polar order

Figure 6: (color online) (a): polar plot of angular distribution ρ(θ; t)/ sin θ for t = 0 (thick solid line, black), t = 104 MCS (thin solid line, blue) and t = 2 · 105 MCS (dashed line, red) for T = 0.3. (b): angular distribution ρ(θ)/ sin θ at the end of the inscription phase for T = 0.1 (thin solid line, red), T = 0.275 (thick solid line, black) and T = 0.5 (dashed line, blue). Excerpt: Part of the plot close to θ = 0◦ .

14

ACS Paragon Plus Environment

Page 14 of 32

Page 15 of 32

0-20

/Nmax

1.0

0.8

0.6

N

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

0.4

(a)

0.0

0.1

0.2

0.3

0.4

0.5

0.6

T

Figure 7: (color online) (a): temperature dependence of normalized average number N0−20 /Nmax of molecules in trans state with 0◦ ≤ θ ≤ 20◦ at the end of inscription phase. (b): log-log plot of N0−20 (t) for a few temperatures.

The transition rates discussed in the previous Section are independent on angle φ and the PDF ρ(θ, φ; t) reduces to ρ(θ; t). In Fig. 6 (a) we present the polar plot of ρ(θ; t)/ sin θ for t = 0, 104 and 2 · 105 MCS. Here, the factor sin θ ensures that an isotropic distribution of orientations of dyes is represented as a circle on the polar plot of ρ(θ)/ sin θ. Clearly, the symmetry breaking from isotropic to polar order occurs: the plot, initially (t = 0) consisting of a circle, develops a maximum around θ = 180◦ and a minimum at θ = 0◦ . The latter, so-called angular hole-burning (AHB) effect, has a double origin: axial symmetry promoted by cos2 θ and cos4 θ terms and polar symmetry promoted by cos3 θ term in Eq. (4). We conclude that our kinetic model correctly describes the effect of centrosymmetry breaking and an onset of polar order in all-optical poling phenomenon. The AHB effect depends on the temperature. Fig. 6 (b) shows the plots of ρ(θ)/ sin θ at the end of the inscription phase for three temperatures: well below Tg (T = 0.1), close to Tg (T = 0.275) and well above Tg (T = 0.5). The inset shows the plot for angles from θ = 0◦ to θ = 60◦ . In each case a well-developed polar order is present. An eye inspection shows that the order of appearance of the plots does not show any systematic dependence on T . To quantify this statement we plot in Fig. 7 (a) the temperature dependence of the normalized 15

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 32

average number N0−20 /Nmax of particles in trans state for which 0◦ ≤ θ ≤ 20◦ , at the end of the inscription phase. Here, Nmax = N0−20 (T = 0.075). A well-defined minimum which corresponds to the strongest AHB effect occurs between T = 0.275 and T = 0.3. At higher temperatures a linear regime sets in. Moreover, the temporal evolution of the number N0−20 close to this interval of temperatures has a very different character than away from it, see Fig. 7 (b). Namely, the double logarithmic plot is linear for T = 0.25 (and nearly linear for T = 0.275) which indicates a power-law behavior, while for two other temperatures away from Tg , T = 0.2 and 0.4, the power-law is absent.

3.2

Nonlinear susceptibilities (2)

(2)

The nonlinear susceptibilities χZZZ , χXXZ depend on two parameters calculated in MC modelling: number density N of trans molecules and averages of trigonometric functions of angle θ. We have found that the former was practically constant during the simulation: (2)

|N (t) − N (0)|/N (0) ≈ 5 · 10−3 . Susceptibility χZZZ becomes then hcos3 θi scaled by a constant factor. For this reason in what follows we analyze hcos3 θi factor only. Inset in Fig. 8 shows the time dependence of hcos3 θi for temperatures: T = 0.1, 0.25, 0.3 and 0.5. A qualitative conclusion is that different types of kinetics are present close to Tg and away from it: in the first case the plot does not saturate after t = 2 · 105 MCS while it does for T = 0.5 and shows a weak growth effect for T = 0.1. Moreover, the absolute value of hcos3 θi is larger close to Tg than away from it. (2)

The build-up of susceptibility χXXZ follows the same pattern and in the remaining part (2)

of the paper we present the results for χZZZ only. The curves hcos3 θi(t) were fitted (Origin 9.5) using single exponential fit (3 parameters), bi-exponential fit (4 parameters) and a stretched exponential fit (3 parameters):

h i d hcos3 θi(t) = A 1 − e−(t/τ ) ,

16

ACS Paragon Plus Environment

(7)

Page 17 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 8: (color online) Plot of parameter hcos3 θi(t) in the inscription phase. Inset: temperatures T = 0.1 (thin red solid line), 0.25 (thick black line), 0.3 (dashed blue line) and 0.5 (dotted green line). Main plot: MC results (solid line, black) and stretched exponential fit (dashed line, red). T = 0.15.

where τ denotes the characteristic time. The visual inspection showed that the single exponential fit failed while both stretched and bi-exponential fits satisfactorily reproduced the data. However, the parameters of the stretched exponential fit were with good accuracy independent on the length of simulation interval (from 2 · 105 MCS to 5 · 105 MCS) while the bi-exponential fit has failed in this respect. We conclude that stretched exponential fit represents a statistical model for the simulated data while bi-exponential - does not. This conclusion holds for both inscription and erasure phases. For this reason we interpret the simulated data using stretched exponential fit, see Fig. 8. The results of fitting hcos3 θi (t) for different temperatures are shown in Fig. 9. The plot of characteristic time τ (T ) (panel (a)) in the inscription phase is not monotonic and displays a distinct maximum around T = 0.25, in the temperature interval where the glass transition occurs. There is a substantial increase – say, 2 times – of relaxation time τ close to T = 0.25 as compared to temperatures away from it. In the decay phase, on the contrary, there is

17

ACS Paragon Plus Environment

The Journal of Physical Chemistry

80 000 18 000

0.9

(a)

70 000

(b)

16 000

50 000

d

12 000

0.8

decay

60 000 14 000

growth

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 32

0.7

40 000 10 000 30 000

0.6

8 000 20 000

0.5 0.0

6 000

0.0

0.1

0.2

0.3

0.4

0.5

0.1

0.2

0.3

0.4

0.5

T

T

Figure 9: Parameters of stretched exponential fits, Eq. (7), of hcos3 θi (t): τ (T ), panel (a) and d(T ), panel (b). Inscription phase: black circles, decay phase: red squares.

no maximum close to the glass transition temperature and the characteristic times decrease nearly linearly with the temperature (a maximum occurs at low temperatures but it is not related to the glass transition - this effect is not of interest for the present study). Panel (b) shows the plot of exponent d(T ). As in the case of τ (T ), in the inscription phase the plot is not monotonic and displays a minimum in the temperature range 0.2 < T < 0.225, slightly below the assumed glass transition temperature. In the decay phase the parameter d has higher values than in the inscription phase (around 0.8-0.9) and displays larger fluctuations (not shown in Fig. 9). The emerging physical picture of complex dynamics of guest molecules is discussed in the next Section.

4

Discussion

4.1

Glass transition, complex dynamics and mosaic-like states

The results presented above clearly show an influence of the glass transition on the kinetics of azo-dyes, expressed in a number of anomalous physical effects in temperature and time. The angular hole burning effect is the strongest for the temperatures close to the GT and

18

ACS Paragon Plus Environment

Page 19 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

becomes weaker for higher and lower temperatures (Fig. 6 (b) and Fig. 7 (a)). The process of the build-up of the polar phase and emergence of nonlinear susceptibilities is clearly marked by the closeness to the GT. Firstly, the characteristic time τ displays a maximum close to the GT (Fig. 9 (a)). This effect can be also observed in the time dependence of the polar order parameter which saturates more slowly close to the GT than away from it (Fig. 8). Secondly, the exponent d displays a minimum close to the GT (Fig. 9 (b)). Still more important is the fact that the kinetics of guest molecules close to the glass transition displays features typical for complex systems: power laws and non-exponential relaxation. The former appears in the angular hole burning effect, Fig. 7 (b), where the number of trans molecules N0−20 (t) displays a power law at T = 0.25:

N0−20 (t) ∼ tα ,

(8)

with the exponent α ≈ 0.3. The power law, inherently related to a lack of a characteristic time scale, breaks down upon a small departure from T = 0.25. Complex relaxation is present in the process of build-up of the polar phase (nonlinear susceptibilities): the stretched exponential function, Eq. (7), constitutes a reliable model for the simulated data, in contrast with single- or double-exponential functions. The exponent d close to the GT has a value around d = 0.55, well away from d = 1 which corresponds to a simple kinetics. For temperatures close to the GT exponent d keeps its low value. At larger departures from GT exponent d increases but in different ways for lower and higher temperatures. In the former case it keeps its value well below d = 1, indicating a complex dynamics. With increasing temperatures d displays a strong systematic increase towards d = 1 (e.g., d ≈ 0.85 for T = 0.5), approaching a simple, non-complex dynamics. We conclude that while in the interval of temperatures studied in this paper the kinetics of guest molecules is complex, this effect is the strongest close to the GT. The complex kinetics of guest molecules close to the glass transition temperature is a direct consequence of complex dynamics of polymer matrix. Namely (i), the intermolecular 19

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

interactions are negligible due to the low concentration of the dyes and (ii), in the absence of polymer matrix non-interacting dyes display bi-exponential kinetics. 14 Our studies based on local analysis of the structure of the matrix offer a deeper insight into its local dynamics close to Tg . Namely, the origin of stretched exponential complexity is usually ascribed 39–43 to a broad (power-law) distribution of trapping periods of time when the dynamics of the guest molecules is blocked or hindered. In the case of a polymer matrix the trapping events are governed by local voids characterized by the time-dependent field V (~r, t). A typical instantaneous configuration of V (~r, t) (Fig. 3) has a form of an intertwining mosaic (speculated on a long time ago 44 ) of blocked (red) and free (green) spaces. In the former the dynamics is strongly hindered, in the latter it is free from steric limitations. The temporal evolution of the matrix changes the field V (~r, t) and modifies the spatial distribution of trapping events, leading to complex behavior of guest molecules. Let us point out an analogy to this hypothetical scenario. For the 2D Lennard-Jones liquid close to melting the temporal correlation functions describing local structure display stretch-exponential behavior while away from it simple exponential dynamics sets in; the origin of this behavior was ascribed to observed power-law distribution of trapping times related to a change of some topological characteristics of mosaic-like local arrangements of the atoms. 13 Characterization of the mosaic-like states V (~r, t) is a challenge and constitutes one of the most important topics in modelling of polymer systems. First quantitative results, related to static characterization of the mosaic, were presented in Ref. 6,7 Present study yields some qualitative conclusions concerning the local scales and local dynamics of the mosaic-like states, originating from an analysis of a very different temperature dependence of time τ in the inscription and decay phases (Fig. 9 (a)). In the former case the maximum of τ reflects, as expected, the slowing-down effect due to the matrix close to GT. Surprisingly, the slowingdown has no visible impact on relaxation time τ in the decay phase. The origin of this effect is bound up with the type of the motions of the guest molecules, in particular with the way they probe the angular phase space. In the decay phase there are no photoisomerization

20

ACS Paragon Plus Environment

Page 20 of 32

Page 21 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

transitions (I = 0) and the molecules perform a diffusive-like orientational motion with a small amplitude; the transition rate depends both on V and C, see Eq. (6). The angular kinetics is weakly complex close to glass transition (exponent d = 0.8 − 0.9 is close to d = 1). In the d = 1 limit the normal angular diffusion is the result of random, uncorrelated in time and space, collisions with the molecules of the environment, and has thermal origin. Thus, the observed kinetics of dye molecules is mainly due to the dynamics of parameter C, while the weak complexity results from a correlated dynamics of local voids. Since the amplitude of trial diffusive steps is small, the molecules probe only some small parts of the angular phase space and the contribution of spatiotemporal correlations of the field V (~r, t) to the kinetics is weak. The situation in the inscription phase is very different. The guest molecules in the trans state probe the whole angular phase space because of the random orientation of a trans molecule after the cis → trans transition. The probability of this transition depends on the local structure of the mosaic at some elapsed time after the trans → cis transition took place. In this case the spatiotemporal correlations of the field V (~r, t) become important; in particular they govern the probability of trapping events. In turn, trapping of a guest molecule in a cis state leads to a wide spectrum of waiting times and well-developped complex dynamics. This qualitative physical picture does not suffice to explain the origin of a rather counterintuitive effect: the largest efficiency of the AHB close to the glass transition (Fig. 7 (a)). Intuitively one expects that the AHB process should be the least effective in this case because of the slowing down effect. This process is governed by the spatiotemporal correlations of V (~r, t) but its kinetics depends on the time scales related to trans → cis and cis → trans transitions. Preliminary results show that the minimum in the plot N0−20 (T ) gradually vanishes as R increases. Nevertheless, parameter N0−20 always differentiates between glassy and melt phases of the matrix. The quantitative analysis of the origin of complex dynamics of guest molecules has to take into account an interplay of time scales related to pct , ptc and of the characteristic times

21

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

for the spatiotemporal correlations of the field V (~r, t). This analysis is beyond the scope of this paper.

4.2

Localization of glass transition via all-optical poling

Our study provides tools for the localization of glass transition on the basis of some anomalies of directly or indirectly measurable physical quantities and not of geometrical (like MFV) parameters. The strongest anomaly corresponds to the order parameter hcos3 θi directly related to the intensity of SHG. The maximum of characteristic time τ in the inscription phase (Fig. 9 (a)) is at T = 0.25. The well-defined minimum of exponent d is localized in the interval T = 0.225 − 0.25 (Fig. 9 (b)). The AHB effect, indirectly measured in SHG experiments, is the strongest in an interval T = 0.275 − 0.3 (Fig. 6 (b), Fig. 7 (a)). Power law is present at T = 0.25 and vanishes already at 0.275 (Fig. 7 (b)). Those anomalies lie in a rather wide interval of temperatures: 0.225 ≤ T ≤ 0.3. We ascribe the origin of the scatter of those estimates for Tg obtained on the basis of the anomalies of physical parameters to the fact that the glass transition manifests itself through singularities of various physical quantities in different ways which reflect their specific dynamics.

4.3

Prediction of MC model vs. experiment

Experimental study of the photoinduced SHG in all-optical poling on thin polymer films of PMMA grafted copolymer with DR1 guest molecules was reported in Ref., 45 for three temperatures: T = 50◦ C, 90◦ C and 110◦ C. The Authors have analyzed only the decay process. We have analyzed the time evolution of χ(2) in the inscription phase using the data retrieved from original data in Ref. 45 The plots of χ(2) (t) and their fits are shown in Fig. 10. The results of the fitting procedure were inconclusive – both single exponential and stretched exponential fits reproduced the data in a satisfactory way and gave approximately the same values of characteristic times τ (inset in Fig. 10). However, in the experiment the ratio of 22

ACS Paragon Plus Environment

Page 22 of 32

Page 23 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 10: (color online) χ(2) amplitude for PMMA grafted copolymer with DR1: experimental data for 50◦ C (#), 90◦ C (4) and 110◦ C (). Red solid line and green dashed line represent, correspondingly, exponential and stretched exponential fits. Inset shows the corresponding relaxation times τ (black circles and empty circles) obtained from fitting procedures; the size of the error bars are of the size of graphical symbols. On the basis of data from Ref. 45 the time of measurement to maximal value of τ was approximately equal 3. We have shown in Sect. 3.2 that reliable conclusions concerning the functional form of χ(2) (t) can be drawn when this ratio is at least one order of magnitude larger. Experimental data clearly show a non-monotonic behavior of τ on the temperature (inset). At maximum, for T = 90◦ C, those times are approximately twice larger (τ = 2000 s) than at lower (T = 50◦ C) and higher (T = 110◦ C) temperatures (τ = 1100 s). This observation remains in a close analogy to slowing-down behavior found in the simulations, see Fig. 9 (a). While in the simulations the maximum was located close to the glass temperature Tg , the situation in experiment is not clear: while the authors of Ref. 45 state that glass temperature of the investigated system was Tg = 130◦ C, in the literature 46 a wide range of possible glass temperature values for such systems is reported, from 85◦ C up to 165◦ C.

23

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

All optical poling of DR1 molecules dispersed in PMMA was studied at room temperature in Ref. 47 The statistics of the data for the inscription phase was not as good as in the above case and no reliable conclusions could be drawn. We conclude that much longer all-optical poling experiments are necessary to reveal the complex nature of underlying physical processes.

5

Summary

We have proposed and studied a simple kinetic Monte Carlo model for the theoretical analysis of the kinetics/dynamics of all-optical poling process in a host (polymer matrix) – guest (azodye molecules) system in the limit of low concentration of guest molecules. We have found that the kinetics/dynamics of guest molecules close to the glass transition temperature Tg belongs to the class of complex systems. The ”increase” of complexity close to Tg constitutes an analogy to the increase of complexity in a coexistence phase for 2-dimensional Lennard-Jones liquid, 13 and sheds some light on the structure/dynamics of microscopic/mesoscopic inhomogeneities in mosaic-like states of the matrix, which (hypothetically) constitute an origin of complexity in locally ordered but globally disordered systems. We suggest that the large-scale correlations in mosaic-like states play an important role in the inscription phase leading to ”strong” complexity while the decay phase is mostly driven by short scale collisions of kinetic origin, leading to ”weak”complexity. The analysis of both spatial and temporal correlations in mosaic-like states is under progress now and the results will be published elsewhere. Finally, let us discuss briefly some perspectives for a modification of the model onto the case of higher concentration of guest molecules and some related physical effects. The influence of the dynamics of azo molecules on the chain, neglected in this paper, becomes important for sufficiently high concentration. The generalization of the model onto that case is possible but not straightforward. The dye-dye interaction can be studied following the lines

24

ACS Paragon Plus Environment

Page 24 of 32

Page 25 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

of Ref. 22 The impact of photoisomerization on the dynamics/diffusion of the chain 48 can be incorporated into the model following the lines of our paper, 18 where a high concentration of azo molecules chemically attached to the chain was studied. In this case a photoisomerization act grants an additional non-thermal trial MC move of neighboring monomer/s. Preliminary results in the two-dimensional case published recently 49 and previous studies 50 clearly indicate that isomerization changes the character of the diffusion of the chain as a whole. Next, for high concentration of the azo molecules their translation has to be accounted for because it would dynamically modify the local environment of polymer chains and thus would have an impact on their dynamics. However, the modelling of the translation induced by isomerization is a non-trivial problem - a sound solution to it was proposed in Ref. 51 An interesting topic which could be studied using the extended model is the modification of the glass transition temperature due to the photoisomerization. 23,48 Qualitatively, two main competing effects emerge. Firstly, isomerization increases locally the mobility of the chain, thus decreasing the Tg . On the other hand, the steric effects may hinder this mobility and increase Tg . Let us point out that in the case of high concentration of the azo molecules the mosaiclike states will undergo a change as compared to the case of the matrix without the dyes and the evaluation of additional correlation functions will be necessary. Finally, a replacement in the extended model of azo molecules by dipolar molecules oriented using corona poling method would make possible a theoretical analysis of recent measurements. 52

ACKNOWLEDGMENTS This paper is partially supported in the form of scholarship by founds from purpose donation, given to Faculty of Fundamentals Problems of Technology of Wrocaw University of Science and Technology by Ministry of Science and Higher Education in 2016 for scientific research 25

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

of young scientists and PhD students. We thank A. Miniewicz for helpful discussions.

References (1) Boyd, R.W. Nonlinear optics. Boston Academic Press 1992. (2) Stegeman, G.I.; Stegeman, R.A. Nonlinear optics. Wiley 2012. (3) Kuzyk, M. Polymer fiber optics: materials, physics, and applications. CRC Press 2006. (4) Sekkat, Z.; Knoll, W. Photoreactive organic thin films. Academic Press 2002. (5) Pawlik, G.; Rau, I.; Kajzar, F.; Mitus, A.C. Second-harmonic generation in poled polymers: pre-poling history paradigm. Opt. Express 2010, 18, 18793-804. (6) Pawlik, G.; Radosz, W.; Mitus, A.C.; Mysliwiec, J.; Miniewicz, A.; Kajzar, F.; Rau, I. Holographic grating inscription in DR1: DNA-CTMA thin films: the puzzle of time scales. Cent. Eur. J. Chem. 2014, 12, 886892. (7) Pawlik, G.; Orlik, R.; Radosz, W.; Mitus, A.C.; Kuzyk, M.G. Towards understanding the photomechanical effect in polymeric fibers: analysis of free volume in a model polymeric matrix. Proc. SPIE 2012, 84740A1. (8) Binder, K.; Kob, W. Glassy Materials and Disordered Solids. World Scientific 2005. (9) Berthier, L.; Biroli, G. Theoretical perspective on the glass transition and amorphous materials. Rev. Mod. Phys. 2011, 83, 587-645. (10) Berthier, L. Trend: dynamic heterogeneity in amorphous materials. Physics 2011, 4, 42-48. (11) Berthier, L.; Biroli, G.; Bouchaud, J.-P.; Cipelletti, L.; van Saarloos, W. Dynamical heterogeneities in glasses, colloids, and granular media. Oxford University Press 2011.

26

ACS Paragon Plus Environment

Page 26 of 32

Page 27 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(12) Ediger, M.D. Spatially heterogeneous dynamics in supercooled liquids. Annu. Rev. Phys. Chem. 2000, 51, 99-128. (13) Patashinski, A.Z.; Ratner, M.A.; Grzybowski, B.A.; Orlik, R.; Mitus, A.C. Heterogeneous structure, heterogeneous dynamics, and complex behavior in two-dimensional liquids. J. Phys. Chem. Lett. 2012, 3, 2431-2435. (14) Pawlik, G.; Mitus, A.C.; Miniewicz, A.; Kajzar, F. Kinetics of diffraction gratings formation in a polymer matrix containing azobenzene chromophores: Experiments and Monte Carlo simulations. J. Chem. Phys. 2003, 119, 6789-6801. (15) Pawlik, G.; Mitus, A.C.; Miniewicz, A.; Kajzar, F. Monte Carlo simulations of temperature dependence of the kinetics of diffraction gratings formation in a polymer matrix containing azobenzene chromophores. J. Non. Opt. Phys. & Mat. 2004, 13, 481-489. (16) Pawlik, G.; Mitus, A.C.; Miniewicz, A.; Sobolewska, A.; Kajzar, F. Temperature dependence of the kinetics of diffraction gratings formation in a polymer matrix containing azobenzene chromophores: Monte Carlo simulations and experiment. Mol. Cryst. Liq. Cryst. 2005, 426, 243-252. (17) Pawlik, G.; Mitus, A.C.; Mysliwiec, J.; Miniewicz, A.; Grote, J.G. Photochromic dye semi-intercalation into DNA-based polymeric matrix: computer modeling and experiment. Chem. Phys. Lett. 2010, 484, 321-323. (18) Pawlik, G.; Sobolewska, A.; Miniewicz, A.; Mitus, A.C. Generic stochastic Monte Carlo model of the photoinduced mass transport in azo-polymers and fine structure of surface relief gratings. Europhys. Lett. 2014, 105, 26002. (19) Pawlik, G.; Mitus, A.C.; Miniewicz, A. Modelling of enhanced photoinduced reorientation of nematic liquid crystal molecules in twisted geometry: Monte Carlo approach. Mol. Cryst. Liq. Cryst. 2012, 554, 56-64.

27

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(20) Pawlik, G.; Mitus, A.C.; Karpinski, P.; Miniewicz, A. Laser light-induced molecular reorientation in 90◦ twisted nematic liquid crystal: classic approach, Monte Carlo modeling and experiment. Opt. Materials 2012, 34, 1697-1703. (21) Pawlik, G.; Mitus, A.C.; Rau, I.; Kajzar, F. Monte Carlo modeling of chosen non-linear optical effects for systems of guest molecules in polymeric and liquid-crystal matrices. Nonlinear Opt. Quantum Opt. 2009, 38, 227-244. (22) Tavarone, R.; Charbonneau, P.; Stark, H. Kinetic Monte Carlo simulations for birefringence relaxation of photo-switchable molecules on a surface. J. Chem. Phys 2016, 144, 104703. (23) Fang, G.; Maclennan, J.; Yi, Y.; Glaser, M.; Farrow, M.; Korblova, E.; Walba, D.; Furtak, T.; Clark, N. Athermal photofluidization of glasses. Nat. Commun. 2013, 4, 1521. (24) Woo, J.W. Birefringent and electro-optic effects in poled polymer films: steady-state and transient properties. J. Opt. Soc. Am. B 1991, 8, 142-152. (25) Rau, I.; Kajzar, F. Second harmonic generation and its applications. Nonl. Opt. Quant. Opt. 2008, 38, 99140. (26) Deutsch, H.P.; Binder, K. Interdiffusion and selfdiffusion in polymer mixtures: a Monte Carlo study. J. Chem. Phys. 1991, 94, 2294-2304. (27) Wittmer, J.P.; Cavallo, A.; Kreer, T.; Baschnagel, J.; Johner, A. A finite excluded volume bond-fluctuation model: Static properties of dense polymer melts revisited. J. Chem. Phys. 2009, 131, 064901. (28) Frenkel, D.; Smit, B. Understanding molecular simulation. Academic Press 2002. (29) Allen, M.P.; Tildesley, D.J. Computer simulation of liquids. Oxford University Press 1987. 28

ACS Paragon Plus Environment

Page 28 of 32

Page 29 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(30) Baschnagel, J.; Binder, K.; Wittmann, H.-P. The influence of the cooling rate on the glass transition and the glassy state in three-dimensional dense polymer melts: a Monte Carlo study. J. Phys.: Condens. Matter 1993, 5, 1597-1618. (31) Wittmann, H.-P.; Kremer, K.; Binder, K. Glass transition of polymer melts: a twodimensional Monte Carlo study in the framework of the bond fluctuation method. J. Chem. Phys. 1992, 96, 6291-6306. (32) Lawlor, A.; Reagan, D.; McCullagh, G.D.; De Gregorio, P.; Tartaglia, P.; Dawson, K.A. Universality in lattice models of dynamic arrest: introduction of an order parameter. Phys. Rev. Lett. 2002, 89 245503. (33) Molina-Mateo, J.; Meseguer-Duenas, J.M.; Gomez-Ribelles, J.L. Geometric distribution of dynamically accessible volume in the bond fluctuation model. Polymer 2007, 48, 3361-3366. (34) Manea, A.-M.; Rau, I.; Tane, A.; Kajzar, F.; Sznitko, L.; Miniewicz, A. Poling kinetics and second order NLO properties of DCNP doped PMMA based thin film. Opt. Mat. 2013, 36, 69-74. (35) van Krevelen, D.W.; te Nijenhuis, K. Properties of polymers. Elsevier Science Amsterdam 2009. (36) Battaile, C.C. The kinetic Monte Carlo method: foundation, implementation, and application. Comput. Methods Appl. Mech. Eng. 2008, 197, 3386-3398. (37) Kiselev, A.D.; Chigrinov, V.G.; Kwok, H.-S. Kinetics of photoinduced ordering in azodye films: two-state and diffusion models. Phys. Rev. E 2009, 80, 011706. (38) Dumont, M.; Sekkat, Z.; Loucif-Saibi, R.; Nakatani, K.; Delaire, J. Photoisomerization and second harmonic generation in disperse red one-doped and -functionalized poly(methyl methacrylate) films. Nonlinear Opt. 1993, 5, 229236. 29

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(39) Stanislavsky, A.; Weron, K. Stochastic tools hidden behind the empirical dielectric relaxation laws. Rep. Prog. Phys. 2017, 80, 036001. (40) Metzler, R.; Klafter, J. From stretched exponential to inverse power-law: franctional dynamics, Cole-Cole relaxation processes, and beyond. J. Non-Cryst. Solids 2002, 305, 81-87. (41) Vainstein, M.H.; Costa, I.V.; Morgado, R.; Oliveira, F.A. Non-exponential relaxation for anomalous diffusion. EPL 2006, 73, 726-732. (42) Bohmer, R.; Ngai, K.; Angell, C.; Plazek, D. Nonexponential relaxations in strong and fragile glass formers. J. Chem. Phys. 1993, 99, 4201-4209. (43) Shlesinger, M.F. Fractal time in condensed matter, Annu. Rev. Phys. Chem. 1988, 39, 269-290. (44) Adam, G.; Gibbs, J.H. On the temperature dependence of cooperative relaxation properties in glassforming liquids. J. Chem. Phys. 1965, 43, 139-146. (45) Fiorini, C.; Nunzi, J.M. Dynamics and efficiency of all-optical poling in polymers. Chem. Phys. Letters 1998, 286, 415-420. (46) Ashby, M.F. Materials selection in mechanical design. Elsevier 2005. (47) Chan, S.W.; Nunzi, J.M.; Quatela, A.; Casalboni, M. Retardation of orientation relaxation of azo-dye doped amorphous polymers upon all-optical poling. Chem. Phys. Lett. 2006, 428, 371-375. (48) Teboul, V.; Saiddine, M.; Nunzi, J.M. Isomerization-induced dynamic heterogeneity in a glass former below and above Tg . Phys. Rev. Lett. 2009, 103, 265701. (49) Wysoczanski, T.; Mitus, A.C.; Radosz, W.; Pawlik, G. Photoinduced (anomalous) dynamics of functionalized polymer chains: applications for surface relief grating modelling. Proc. SPIE 2017, 101010W. 30

ACS Paragon Plus Environment

Page 30 of 32

Page 31 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(50) Pawlik, G.; Kordas, W.; Mitus, A.C.; Sahraoui, B.; Czaplicki, R.; Kajzar, F. Model kinetics of surface relief gratings formation in organic thin films: experimental and Monte Carlo studies. Proc. SPIE 2007, 674007. (51) Lefin, P.; Fiorini, C.; Nunzi, J.M. Anisotropy of the photoinduced translation diffusion of azo-dyes. Optical Materials 1998, 9, 323-328. (52) Umezawa, H.; Jackson, M.; Lebel, O.; Nunzi, J.M.; Sabat, R.G. Second-order nonlinear optical properties of mexylaminotriazine-functionalized glass-forming azobenzene derivatives. Opt. Mat. 2016, 60, 258-263.

31

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 11

32

ACS Paragon Plus Environment

Page 32 of 32