diffusion of aromatic solutes in aliphatic polymers above glass...

49
Page | 1 Diffusion of aromatic solutes in aliphatic polymers above glass 1 transition temperature 2 Xiaoyi Fang 2 , Sandra Domenek 2 , Violette Ducruet 1 , Matthieu Refregiers 3 , Olivier Vitrac 1 * 3 1 INRA, UMR 1145 Ingénierie Procédés Aliments, F-91300 Massy, France 4 2 AgroParisTech, UMR 1145 Ingénierie Procédés Aliments, F-91300 Massy, France 5 3 Synchrotron SOLEIL, l’Orme des Merisiers, F-91192 Gif-sur-Yvette, France 6 Abstract 7 The paper presents a harmonized description of the diffusion of solutes with repeated aromatic 8 jumping units (JU) in entangled aliphatic polymers above their Tg. It is shown that the trace 9 diffusion coefficients, D, are scaled with the number of jumping units or equivalently with solute 10 molecular mass, M, as 1 K T Tg K M M , where K α and K β are temperature-equivalent 11 parameters related to Williams-Landel-Ferry (WLF) ones. K α is almost a generic constant for 12 aliphatic polymers. The scaling of diffusion behaviors of linear aliphatic and aromatic solutes 13 appear separated by a temperature shift, K β , of ca. 91 K. The effects of the number of JU and the 14 distance between two JU were specifically probed in several aliphatic polymers (polypropylene, 15 * Author to whom correspondence should be addressed. E-mail: [email protected] Tel. +33 (0)169935063

Upload: others

Post on 04-Feb-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

  • Page | 1

    Diffusion of aromatic solutes in aliphatic polymers above glass 1

    transition temperature 2

    Xiaoyi Fang2, Sandra Domenek

    2, Violette Ducruet

    1, Matthieu Refregiers

    3, Olivier Vitrac

    1* 3

    1INRA, UMR 1145 Ingénierie Procédés Aliments, F-91300 Massy, France 4

    2AgroParisTech, UMR 1145 Ingénierie Procédés Aliments, F-91300 Massy, France 5

    3Synchrotron SOLEIL, l’Orme des Merisiers, F-91192 Gif-sur-Yvette, France 6

    Abstract 7

    The paper presents a harmonized description of the diffusion of solutes with repeated aromatic 8

    jumping units (JU) in entangled aliphatic polymers above their Tg. It is shown that the trace 9

    diffusion coefficients, D, are scaled with the number of jumping units or equivalently with solute 10

    molecular mass, M, as 1 K T Tg KM M

    , where Kα and Kβ are temperature-equivalent 11

    parameters related to Williams-Landel-Ferry (WLF) ones. Kα is almost a generic constant for 12

    aliphatic polymers. The scaling of diffusion behaviors of linear aliphatic and aromatic solutes 13

    appear separated by a temperature shift, Kβ, of ca. 91 K. The effects of the number of JU and the 14

    distance between two JU were specifically probed in several aliphatic polymers (polypropylene, 15

    * Author to whom correspondence should be addressed.

    E-mail: [email protected]

    Tel. +33 (0)169935063

    mailto:[email protected]

  • Page | 2

    polylactide and polycaprolactone) at different temperatures above Tg with two homologous 16

    solute series: short oligophenyls and diphenyl alkanes. An extended free volume theory for many 17

    JU was accordingly inferred to account for the observed statistical independence between the 18

    fluctuations of the free volumes probed by each JU and the probability of the collective 19

    displacement of the center-of-mass of the solute. Outstanding properties of short oligophenyls 20

    series provided further insight on the underlying molecular mechanism of translation. Their 21

    activation energies grow differently according to the number of phenyl rings, NPh, being odd or 22

    even. Constrained molecular dynamics demonstrated that such a parity effect could be 23

    remarkably reproduced when the translation of each JU (i.e. phenyl ring) was randomly 24

    controlled by a combination of short and long-lived contacts. 25

    Keywords: diffusion, aromatic molecules, aliphatic polymers, scaling exponents, free volume, 26

    activation energy, molecular dynamics 27

    28

    1 INTRODUCTION 29

    No general diffusion model is available to predict the broad range of trace diffusion 30

    coefficients (D) of organic solutes such as oligomers, additives, residues, contaminants or 31

    degradation products in polymer materials at solid state, which corresponds to their conditions of 32

    service. By analyzing D values in polyolefins, a strong dependence of D values has been 33

    highlighted for several categories of organic solutes: resembling the polymer (e.g. linear or 34

    branched solutes) or not (e.g. aromatic molecules, hindered antioxidants)1. The exponents, α, 35

    scaling D with molecular mass (M) as D∝M-α, were found strictly greater than 1 and typically 36

    above 2, which reflects variations of D over several decades with only small change of M, 37

    between 102 and 10

    3 g⋅mol-1 for most of the molecules of technological interest such as 38

  • Page | 3

    antioxidants, UV stabilizers, plasticizers,… Those variations were related mainly to three 39

    geometric factors: molecular volume, shape factor and gyration radius. D values reported by 40

    Berens2 suggest that mass/volume dependence is even greater when below the polymer glass 41

    transition temperature (Tg). On the opposite, in melts, scaling exponents for homologous alkane 42

    series in polyethylene were experimentally assessed close to 1 in agreement with the Rouse 43

    theory3. Molecular dynamics simulations of generic polydisperse systems above Tg found also 44

    exponents of 1 for a wide range of thermodynamical conditions4. The absence of temperature 45

    effect on α was, however, not verified experimentally. Thus, Kwan et al.5 reported a strong effect 46

    of temperature on α values – with α falling from 4.7 to 2.1 between 23°C and 85°C – for n-47

    alkanes dispersed in a lightly cross-linked amorphous polyamide, suggesting a continuous but 48

    sharp evolution from Tg to higher temperatures and without a significant contribution of the 49

    possible crystalline phase. 50

    Such a high mass dependence, with α values varying between 26 and 3

    7 have also been 51

    reported for the self-diffusion of entangled polymer chains subjected to reptation and strong 52

    reptation translation mechanisms respectively, with intermediate values close to 2.4 when 53

    reptation is combined with a constraint release mechanism8. The formal analogy is, nevertheless, 54

    of a limited use for solutes with molecular masses ranging between 102 and 10

    3g⋅mol-1, as their 55

    gyration radii are much smaller than the typical entanglement length of polymer segments. For 56

    organic solutes larger than voids between polymer segments and smaller than entanglement 57

    length, only a partial and local coupling between the reorientation and local relaxation modes of 58

    the polymer can be expected9. It was thus shown that aromatic molecules remain non-oriented in 59

    the amorphous phase of polyethylene when the material was stretched uniaxially10, 11

    , whereas 60

    anisotropic diffusion of toluene was observed in compressed natural rubber12

    . 61

  • Page | 4

    The main goal of this study is to provide a consistent description of molecular diffusion 62

    mechanisms of aromatic solutes mimicking molecules of technological interest in linear 63

    polymers and to provide a polymer-independent description of the strong dependence of D with 64

    M at any temperature greater than Tg. In a first approximation, such solutes can be described as 65

    the repetition of rigid jumping units subjected to independent displacements on short time scales. 66

    An extended version of free volume theory is accordingly introduced to account for the weak 67

    coupling between polymer relaxation, controlling the fast displacements of each jumping unit, 68

    and the collective reorientation and translation of the whole solute. It was validated on two 69

    independent homologous series of solutes. The first series of diphenyl alkanes with close size 70

    was used to probe polymer effects (i.e. independent motions of jumping units) at a similar 71

    reference temperature. The second series, short linear oligophenyls, probed specifically solute-72

    related effects (i.e. the collective motions of repeated jumping units) and their activation by 73

    temperature. As both series started with a common first molecule, biphenyl, entropic effects 74

    could be reliably scaled for both series. 75

    This paper is organized as follows. Scaling laws for linear aliphatic and aromatic solutes in 76

    aliphatic polymers are discussed in section two in the framework of a coarse-grained theory of 77

    diffusion of solutes consisting in a linearly repeated jumping units or blobs. The existence of at 78

    least two correlation time-scales between the displacements of jumping units and surrounding 79

    host particles is used to justify the major deviation to the Rouse theory13

    and to conventional free 80

    volume theories in dense entangled aliphatic polymers (i.e. far below their melting points). An 81

    extended free volume theory is proposed based on a formal separation of thermal expansion 82

    effects acting on the displacements of each jumping unit, so called “host effect”, and barrier 83

    effects, so called “guest effects”. A simple theory of activation barriers for the diffusion of 84

  • Page | 5

    oligophenyls is accordingly proposed. Section three reports the methodologies used to determine 85

    trace diffusion coefficients, ranged between 10-17

    m2⋅s-1 and 10-12 m2⋅s-1, in three different 86

    polymers: polypropylene (PP), polycaprolactone (PCL) and polylactide (PLA) for both 87

    homologous solute series. A fourth polymer, plasticized and unplasticized polyvinyl alcohol 88

    (PVA), was used as external validation for arbitrary polymers above Tg. The proposed free 89

    volume theory is tested against experimental in section four. The scaling of activation energies 90

    and entropies of oligophenyls is discussed according to results obtained in constrained molecular 91

    dynamics simulation. The likely mechanism of translation of aromatic molecules in linear 92

    polymers above Tg and its consequence on solute mass dependence is finally proposed in the last 93

    section. 94

    95

    2 THEORY 96

    This section discusses the possible causes responsible for the deviation of solutes larger than 97

    voids to existing diffusion theories in linear polymers at solid state. By relying on the diffusion 98

    properties of solutes including linearly repeated jumping units (or also called blobs), two major 99

    arguments are proposed: i) the increase in polymer density with decreasing temperature down to 100

    Tg affects the individual displacements of each jumping unit, ii) and the resulting displacement 101

    of the center-of-mass is strongly affected by the heterogeneous dynamics of individual jumping 102

    unit. 103

    104

    2.1 Scaling of D with the number of jumping units and temperature for linear solutes 105

    The strong slowdown of D values with M (respectively the number of jumping units in solutes 106

    with linearly repeated patterns) is major characteristic of organic solutes in polymers at solid 107

  • Page | 6

    state (e.g. semi-crystalline polymers or polymers below their flow thresholds). Published values 108

    are scarce, mainly due to the necessity to work with homologous structure series and a wide 109

    range of D values. Scaling of D in polyethylene far above its Tg is reported in Fig. 1 for two 110

    solute series artificially split into two categories : i) linear aliphatic solutes (Fig. 1a), including 111

    linear n-alkanes (n=12..60)3,

    14

    , n-alcohols (n=7..18)15

    and esterified phenols containing a 3(3,5-112

    di-tert-butyl-4-hydroxyphenyl) head and a long n-alkyl tail (n=6..18)16

    , and ii) aromatic solutes 113

    (Fig. 1b), including n-alkylbenzene containing short n-alkyl chains (n=0..4)17

    , substituted and 114

    hindered phenols (M= 94-545 g⋅mol-1)15, respectively. For each series, D values are scaled as a 115

    power law of M (ranging between 70 and 103 g⋅mol-1), whose exponents are much greater than 116

    unity and tend to decrease when temperature is increasing. α values assessed up to 4 or 5 117

    constitute a major deviation to the Rouse theory that predicts values close to unity instead. 118

    According to the Rouse theory, unitary values are related to a purely single chain relaxation with 119

    a uniform friction factor and without any chain end effects13

    . Mutual diffusion coefficients of 120

    n-alkanes in polyethylene melts3 extrapolated to infinite dilution suggest that α values close to 121

    unity should be recovered far from Tg. In agreement with experimental results and free volume 122

    theories, we propose a scaling exponent deviation to the Rouse theory, denoted Δα, which 123

    depends mainly on the temperature difference T-Tg: 124

    1T Tg T TgM MD M

    (1) 125

  • Page | 7

    126

    Figure 1. Log-log plots of trace diffusion coefficients in polyethylene (PE) either with low 127

    density (LDPE) or high density (HDPE) of a) linear aliphatic and b) aromatic solutes. Dref was 128

    chosen as the diffusion coefficient of the first solute in the considered solute series. 129

    130

    For solutes consisting in a small number of linearly repeated jumping units or blobs, N, the 131

    deviation to the Rouse theory close to Tg can be thought as the consequence of irregular jumping 132

    unit displacements with time. To justify this argument, we adopt the coarse-grained model of 133

    Herman et al.18-21

    initially proposed to describe the dynamics of flexible linear chains in the melt. 134

    Similarly, by neglecting end-effects, we assume that the mean-square-displacement of a single 135

    jumping unit increases with time as: 136

    6 blobg t D t t (2) 137

    where Dblob(t) is the time-dependent diffusion coefficient of a blob/jumping unit. For any solute 138

    with identical jumping units indexed i=1..N, Dblob is assumed to decrease with the amount of 139

  • Page | 8

    cumulated pair correlations between the displacements of jumping unit i, denoted Δri(t), and the 140

    displacements of all particles within the system (i.e. including host and solute), denoted Δrj(t), 141

    as: 142

    0 0 0

    ,

    2

    ·blob

    i j

    all j including i

    all j including i

    D D DD t

    C t C tt t

    t

    i j

    i

    Δr Δr

    r

    (3) 143

    where D0 is a scaling constant. In this simplified description, jumping units have a similar 144

    spherical shape and volume regardless the detailed chemical structure of the solute and host 145

    polymer chains. An increasing C(t) with time leads to a sub-diffusive regime. The mean-square-146

    displacement of the center-of-mass (CM) is accordingly given by the covariance of the averaged 147

    displacements of all jumping units: 148

    , ,2 21 , : , :

    1 1 22

    N N N

    CM i j i j

    i i j i i j jij

    g t g t C t g t C t g tN N N

    (4) 149

    For a linear and flexible solute and by neglecting torsional constraints, only the N-1 150

    correlations between the displacements of connected jumping units are significant. Eq. (4) 151

    becomes accordingly: 152

    2

    12

    t

    CM connect

    g t Ng t C g t

    N N

    (5) 153

    wherein tconnectC is the normalized correlation between the displacements of two connected 154

    jumping units. Incorporating Eqs. (2) and (3) yields finally: 155

    00

    61 1 1 12 1 6 t

    connect

    t

    connectCM C C t

    C D tg t D t

    N C t N C t N C t

    (6) 156

    By assuming that t

    connectC is small comparatively to correlations with the displacements of host 157

    polymer segments in a dense system, the mean-square-displacement of CM and consequently the 158

  • Page | 9

    tracer diffusion coefficient, lim 6t CMD g t t , is scaled as 1/N, in agreement with the 159

    Rouse theory13

    . 160

    161

    In presence of strong coupling between the lateral displacements of solute jumping units and 162

    host particles (e.g. large jumping unit), the translation of each connected jumping unit is 163

    expected to be highly heterogeneous and controlled by a combination of short and long-lived 164

    dynamic contacts as discussed in general terms in22

    and23

    . In the followings, we will assume the 165

    simplest case where two relaxation modes resulting of many body interactions can be 166

    independently applied to each jumping unit: C(t) and ttrappedC , with respective probabilities 1-p 167

    and p. ttrappedC ≫C(t) is the total correlation when a jumping unit is almost “blocked” or 168

    significantly hindered by host segments. From this simplified description, gCM is governed by the 169

    superposition of all possible partitions between “trapped” (i.e. long-lived contacts with host) and 170

    “untrapped” (i.e. short-lived contacts with host) jumping units. By assuming that g(t) still obey to 171

    Eqs. (2) and (3), Eq. (6) is replaced by: 172

    2 21

    0

    11

    61

    t

    trapped

    N j jumping units translate according to C t

    all jumping units j jumping units translate according to Ctranslate according to C t

    NN j

    tj trappedCM

    N N j jp p

    NC t j N C t Ng

    p

    t C

    D t

    1

    1

    2

    1

    1

    2 1 1 1 1

    2 12 1

    2 1

    t

    trapped

    NN j

    j

    N

    Np N

    CN

    N

    C t

    Np

    j

    N

    N C t C tN

    (7) 173

  • Page | 10

    wherein N

    j

    is the number of combinations to choose exactly j jumping units among N with 174

    long-lived contacts (governed by ttrappedC ) and p

    j is the related probability by assuming that j 175

    jumping units are blocked independently. 176

    Eq. (7) stresses that a higher dependence to N (α>1), and therefore to M, is expected for small 177

    N (typically lower than 5) as soon as p→1 and while several populations of contacts between 178

    jumping units and surrounding coexists. Considering only two populations provides a rough 179

    scaling of D with the number of jumping units, but its main interest is to highlight that the 180

    scaling with N or M would be the result of the heterogeneous dynamics of the coupling between 181

    the fluctuations of free volume (i.e. associated to the displacements of a single jumping unit) and 182

    collective displacements of many jumping units. 183

    As the value p is related to the probability to find a free volume close to each jumping unit, we 184

    propose to test against experimental values a phenomenological temperature superposition model 185

    where the temperature shift factor Δα is a function of the reciprocal fractional free volume in the 186

    polymer, written as for any T>Tg-Kβ: 187

    1 1K

    T Tg T TgT Tg K

    (8)

    188

    where Kα is a scaling constant and Kβ is a constant depending on the size of the jumping unit. 189

    Accordingly Kβ is expected to be larger for aromatic solutes than for aliphatic ones. It is 190

    important to note that Eq. (8) is a special form of the Williams-Landel-Ferry (WLF) equation24

    191

    applied to the scaling relationship in Eq. (1). Indeed, by following the suggestions of Ehlich et 192

    al.25

    and Deppe et al.26

    , D(T), relative to its value at Tg and to a reference molecular mass, M0, 193

    can be written as: 194

  • Page | 11

    0 0

    ln ln ( ) ln lnTD T KM T Tg M

    a Tg TD Tg M T Tg K MK

    (9)

    195

    which can be identified to the standard WLF equation 1

    2

    ln TC T Tg

    aC T Tg

    , with 196

    1

    0

    lnK M

    CMK

    and 2C K . The absolute value of Kβ appears in Eq. (9) to keep the 197

    consistency D(T)>D(Tg) for any T>Tg-Kβ. More rigorously, a temperature greater than Tg-Kβ 198

    should be chosen instead of Tg. Besides, M0 should be chosen as the unity or as a multiple of the 199

    molecular mass of one single jumping unit. 200

    When applied to a broad range of temperatures, Eq. (9) implies a significant deviation of 201

    activation of diffusion from Arrhenius’ law and a log-dependence of the apparent activation 202

    energy, denoted Ea, with M: 203

    2

    2

    0

    ln ,, K ln

    1 Tg+K

    D T M RT MEa T M R

    T MT

    (10) 204

    Similar deviations to Arrhenius behavior was already proposed by Deppe et al (see Eq. (7) in26

    ) 205

    for aromatic solutes in rubbery poly(isobutyl methacrylate) near its Tg. An Arrhenius behavior is 206

    expected to occur only when T≫Tg-Kβ. A similar log-type scaling of Ea with M is consistently 207

    inferred for n-alkanes in low-density polyethylene from activation energies reported in15

    , with 208

    values of 57, 66 and 107 kJ⋅mol-1 for dodecane, octadecane and dotriacontane respectively. From 209

    D values collected in17

    , the same trend is also drawn for alkylbenzene in polyethylene. As 210

    discussed in25

    , the dependence of WLF parameter C1 to 0

    lnM

    M could be related either to a 211

    stronger coupling with matrix mobility or to an increase in the volume of the solute jumping unit. 212

    213

  • Page | 12

    2.2 Conventional free volume theories 214

    Free volume theories argue that the translation of the center-of-mass of the solute is controlled 215

    in time by the redistribution of voids around the solute due to thermal fluctuations. However, 216

    early models derived from viscosity theories27

    , such as the one proposed by Cohen and Turbull28

    , 217

    tend to underestimate the mass dependence by assuming that the whole solute translates as a 218

    single jumping unit, which yields a scaling exponent close to 1. Theories modified by the 219

    Vrentas and Duda model29, 30

    for polymer-solvent mixtures incorporate two additional features: i) 220

    an internal energy change is required to initiate a translation of the center-of-mass and ii) an 221

    increase in free volume due to polymer thermal expansion. The corresponding mutual diffusion 222

    coefficient of the solute indexed 1 within the polymer indexed 2 is given by Eqs (11)-(13): 223

    1 1 2 2

    1 0

    V VD exp exp

    V̂ /FH

    ED

    RT

    (11)

    224

    P SE E E

    (12) 225

    11 21 1 12 22 2

    1 2

    1 2

    ˆ K K Tg K K TgFH

    T TV

    (13) 226

    where 1 and 2 are the mass fractions in species 1 (solute with glass transition temperature 227

    1Tg ) and 2 (polymer with glass transition temperature 2Tg ) respectively. 0D is a constant. *E 228

    is the effective energy barrier per mole to overcome attractive forces and defined as a balance 229

    between the barrier in dilute state ( 1 0 ), EP, and in pure solvent ( 1 1 ), Es. *

    1V and *

    2V are 230

    the specific hole free volume of solute and polymer required for a jump respectively. ˆFHV is 231

    the effective (including overlaps) average hole free volume per unit of mass of mixture. ξ is the 232

    solute fractional jumping unit. In the case of our studied linear aromatic solutes, it could be 233

  • Page | 13

    envisioned as the ratio of the solute jumping unit (i.e. phenyl ring) to the polymer jumping unit. 234

    1,2i i

    , 1,2ii i

    K

    , 1,2; ,2; 1i iij j j

    K

    , are parameters that account for overlapping factors, free-235

    volume parameters and of their interactions respectively. 236

    At diluted state ( 1 0 ), which is of technological interest for most of the additives and 237

    polymer residues, Eqs. (11)-(13) can be recast in a simpler model: 238

    2 21 0

    12 22 2

    0

    2

    VD exp exp

    K K Tg

    = D exp expTg

    a

    b

    ED

    RT T

    KE

    RT TK

    (14) 239

    Original constants ξ, γ2, K12 and K22 are recast into two lumped parameters named Ka and Kb 240

    respectively to their counter parts Kα and Kβ in Eq. (8). In the remainder of the paper, we drop 241

    the indices 1 and 2, and Tg2 is replaced by Tg. As Eq. (8), Eq. (14) is also related to a special 242

    form of WLF equation24

    , with an additional temperature-related translation term induced by the 243

    energy barrier E*: 244

    1 1

    ln ln aTb b

    D T KE T Tga

    D Tg R T Tg K T Tg K

    (15) 245

    When E*=0, Eqs. (15) and (9) are equivalent with 0lnaK K M M and bK K . According 246

    to Vrentas et al.31

    , E* is related to the variation of internal energy when a solute is introduced at 247

    infinite solution in the polymer. This quantity is expected to be low for any solute with good 248

    solubility in the polymer at the considered temperature. 249

    250

    2.3 Extended free-volume models for aromatic solutes in aliphatic polymers 251

    The need to extend conventional free volume theories for large solute in amorphous polymers 252

    has been already discussed for plasticizers in PVC 32-36

    and for flexible and semi-flexible solutes 253

  • Page | 14

    that move in a segmentwise manner in25, 26

    . Indeed, translation of additive-type solutes invoke 254

    relatively high activation energies, ranging typically from 50 to 150 kJ·mol-1

    9, 37

    . As such values 255

    are close to the activation energies of polymer relaxations, it is thought that large or bulky 256

    solutes involve the same cooperative motions of polymer segments as observed in viscous flow 257

    38,39. Conventional free volume models do not provide any indication on the strong scaling of D 258

    with the size of solutes, on the origin of temperature dependence of α with M for linearly 259

    repeated solutes and on the size of the elemental jumping unit. All solute-related effects are 260

    gathered in Eq. (14) into a single parameter, independent of temperature, , lumped within a 261

    single parameter 2 2 12*a VK K combined with the contribution of the polymer thermal 262

    expansion. According to this description, the translation of the solute center-of-mass and 263

    polymer relaxation would be simultaneous and interrelated phenomena at rubbery state: one 264

    translation of polymer segments causing necessarily the translation of center-of-mass of the 265

    whole solute regardless the size of the solute. It is however well established that though the 266

    activation volumes of translational diffusion of additive-type molecules are significant, ranging 267

    from 80 to 250 Å3, they are smaller than the solute itself

    40. As a result, a piece-wise translation 268

    based on the translation of elemental flexible units appears more likely. Besides, the possible 269

    absence of exact match between the shape of free volumes freed by the polymer and the shape of 270

    elemental jumping units should promote a delay (i.e. long-lived contacts) between the effective 271

    translation of the jumping unit and un-concerted motions of polymer segments. This last 272

    assumption is very likely in the light of experimental results showing that aromatic solutes 273

    remained non-oriented with aliphatic polymer segments10, 11

    . 274

    In this work, we adopt the following description. The displacements of each individual 275

    jumping unit obeys to conventional free volume theories on short time scales but the collective 276

  • Page | 15

    displacement of the center-of-mass of the solute (CM) is controlled by the necessity of concerted 277

    motions of all jumping units on long time-scales. As a result, the joint probability of a solute 278

    translation of CM was associated to two favorable events: a favorable fluctuation of the contour 279

    of polymer segments and the concerted displacements of all jumping units. As both events are 280

    assumed to be independent, Eq. (14) was refactored for solutes made of repeated jumped units as 281

    two independent exponential factors (i.e. associated to two independent marginal probabilities): 282

    .

    , ,, ,

    , ,

    , exp

    solute polymerexcess

    asoluteexcess b

    D T Tg soluteD T solute D T Tg

    D T Tg reference solute

    KD T solute

    K T Tg

    (16) 283

    where soluteexcess

    D is a polymer-independent dimensionless factor enabling the extrapolation to solutes 284

    comprising the same type of jumping units as the reference one, but with a different combination 285

    or repetition. Although initial considerations are different, Eq. (16) resembles Eq. (9) used in17

    to 286

    describe the diffusion of n-alkylbenzene in polyethylene via a hybrid model. In our case, the 287

    concept of “polymer effect” is related to the probing of free volume effects by a single jumping 288

    unit as discussed in25

    . For solutes based on the same jumping unit (i.e. phenyl ring in linear 289

    aromatic solutes), this contribution is expected to depend only on the polymer host and not on 290

    the solute. 291

    292

    Eqs. (2)-(7) suggests a complex dependence of D with the collective displacements of jumping 293

    units. The chosen solute series enabled to test two effects: contribution of the distance between 294

    two jumping units (diphenyl alkanes series) and contribution of the number of repeated jumping 295

    units (oligophenyls series). 296

    297

  • Page | 16

    By assuming that the probability of concerted motions between two jumping units decreases 298

    exponentially with the distance between two jumping units, Eq. (16) was written for diphenyl 299

    alkanes according to the number of carbons between two phenyl rings, NC, (from 0 to 2), as: 300

    10

    , ,exp ln10 exp

    , 0

    C C a

    C C b

    D T Tg N N K

    D T N N K T Tg

    (17) 301

    where 10C

    N is the number of carbons required to decrease D by 10. The first exponential in Eq. 302

    (17) emphasizes that increasing the distance between two phenyl rings (higher NC) strengthens 303

    considerably C(t) (and possibly Ctrapped). In particular, the polymer host is thought to fill the 304

    space between two phenyl rings as the consequence of a reduction of the solute excluded 305

    volume. As the number of phenyl rings are the same in diphenyl alkanes, Eq. (17) offered an 306

    opportunity to assess with a good accuracy of polymer host effects by fitting Ka and Kb on D 307

    values obtained in different polymers (i.e. with different Tg) and at a similar reference 308

    temperature. 309

    310

    Regardless the host polymer, Eq. (10) predicts that the change in α values (assessing the 311

    collective motions of jumping units) with temperature should be followed by an increase of the 312

    apparent activation energy of D with N. From Eq. (5), cooperative displacements of jumping 313

    units are expected significant only for small N values and if tconnectC and C(t) are of the same 314

    magnitude order. Such effects were accounted by considering a solute related free energy barrier, 315

    which varies with the number of phenyl rings, NPh: solute Ph solute Ph solute PhA N Ea N TS N , 316

    where Easolute, and Ssolute are the solute translation activation energy and entropy respectively. In 317

    the canonical ensemble, soluteA is the Helmholtz free energy associated to the probability to find 318

  • Page | 17

    a significant translation of CM when all phenyl rings are subjected to correlated displacements 319

    with the surroundings. As in Eq. (17), Eq. (16) was written relatively to biphenyl as: 320

    0

    2, ,

    exp exp2

    exp

    solute PhexcessPh a

    bsolute Ph

    A ND T Tg N K

    RT K T TgA ND

    RT

    (18) 321

    with 2 2solute Ph solute Ph solute Phexcess

    A N A N A N for NPh≥2 and 2solute Phexcess

    A N =0 for 322

    NPh=2. 323

    324

    2.4 Modeling of activation terms for oligophenyl solutes 325

    The assumption of solute PhA N controlled by a partitioning of the correlation with the 326

    displacements of surrounding jumping units as C(t) and Ctrapped(t), with probabilities 1-p and p 327

    respectively, was tested by constrained molecular dynamics for the simplest scenario: 328

    Ctrapped(t)→∞ and C(t)>0. The advantage of this scenario is that it can be simulated directly 329

    molecular dynamics simulation on isolated molecules by assuming that one or several rings have 330

    their positions fixed. 331

    Easolute and Ssolute versus NPh were calculated respectively in order to enable a comparison with 332

    experimental values. Easolute was defined according to the typical temperature-dependent 333

    translational time of CM, expring Ph Pu hsol teEa RN TN , to cross a distance equal to the 334

    diameter of a phenyl ring ∅ring when 0 to NPh rings are blocked. As in Eq. (7), ring PhN is 335

    obtained by averaging over all possible configurations to block phenyl rings: 336

  • Page | 18

    1..

    1

    0

    1 1 1 1

    1 1

    1 , ,

    1 1

    Ph

    Ph Ph

    Ph Ph

    Ph

    Ph Ph

    Ph Ph

    j N rings are trapped

    N Nno ring trappedj jN N

    N j j

    j k j k

    ring Ph trappedN NN NPh Phj

    P

    j

    j j

    hp p j k p j k

    NN N

    p p p

    N

    pj j

    (19) 337

    where τ(j,k) is set the minimum time to induce a displacement of CM equal to ∅ring with exactly j 338

    rings blocked within the kth

    configuration chosen among the PhN

    j

    possibilities: 339

    2

    2

    ,

    , mi 1n 1ring

    j

    P

    k

    hj k for jt t

    N

    CMΔr

    (20) 340

    The special cases, where no ring is blocked and where all rings are blocked, were associated to 341

    τ0 and τ(NPh,1) respectively. τ(NPh,1) was expected to be large but finite, so that it can be set to a 342

    constant τtrapped. Beyond τ(j,k), CM is likely to displace to a distance larger than one ring so that a 343

    different combination of trapped and untrapped rings is expected to occur. 344

    According to Eqs. (19) and (20), translation of CM occurs mainly as a sequence of macrostates 345

    where 1..NPh-1 rings are randomly blocked. The transition from one macrostate to the next one 346

    occurs at the fastest rate enabled by the dynamics of the CM when the solute is subjected to 347

    topological constraints. 348

    349

    The translational entropy, Ssolute, which measures the number of microstates associated to a 350

    macrostate was calculated analytically according to the same framework but at atomistic scale. 351

    All possible displacements of atoms were described as the superposition of 3NA-6 quantum 352

    harmonic oscillators, with NA the number of heavy atoms (i.e. carbons for tested molecules) in 353

    the considered solute. Absolute entropies were calculated according to Eq. (21), as justified in41

    : 354

  • Page | 19

    3 6

    1

    ln 1 expexp 1

    A

    i Bi B

    i

    N

    conformation

    B

    al B

    i

    kS k k

    k

    TT

    T

    (21) 355

    where the quasiharmonic frequencies Bi ik T were calculated from the eigenvalues, λi, of 356

    the covariance matrix of the fluctuations of atom positions: 357

    1 1 1 22 2,i i i i i ir r r r (22) 358

    where i1 and i2 are carbon coordinate indices chosen among 1…3NA. 359

    The translational entropy was calculated by the entropy difference when the position of CM 360

    was fixed or not, as described in42

    : 361

    conformational conformational, , ,fixed CMS j k S j k S j k

    (23) 362

    Finally, Ssolute was obtained by averaging over all possibilities to block any combination of 363

    phenyl rings in the solute: 364

    0

    1 1

    1

    1 ,

    1

    Ph

    Ph

    Ph

    Ph

    Ph

    N

    jNN j

    ring

    j k

    solute Ph NN Ph

    P

    j

    j

    hp S p S j k

    S NN

    N

    p pj

    (24) 365

    where S(NPh,1)=0 due to the absence of available degree of freedoms. S0(NPh) is the entropy 366

    when no phenyl ring is blocked. 367

    368

    3 EXPERIMENTAL SECTION 369

    3.1 Materials 370

    Tables 1 and 2 list the studied aromatic solutes and polymers respectively. The two tested 371

    series included phenyl rings as elemental jumping units and shared biphenyl as common 372

    molecule. In details, the diphenyl alkanes series enabled to assess the effect of the distance 373

  • Page | 20

    between two elemental jumping units for different T-Tg values while keeping T close to a same 374

    reference temperature. By contrast, the oligophenyls series was used to assess the effect of the 375

    number of jumping units and the effect of collective barriers (Easolute, Ssolute) by shifting both T 376

    and Tg values. 377

    Table 1. List of studied aromatic solutes 378

    Series Solutes

    Diphenyl alkanes biphenyla Diphenylmethaneb Bibenzylb

    Structure

    M (g⋅mol-1) 154.2 168.2 182.3

    NC= 0 1 2

    Oligophenyls biphenyla p-terphenyla p-quaterphenyla

    structure

    M (g⋅mol-1) 154.2 230.3 306.4

    NPh= 2 3 4 asupplied by Sigma-Aldrich Chemistry (Germany) with purity of 99.5 %.

    bsupplied by Acros 379

    Organics (France) with purity of 99 %. 380

    381

    Effects of T-Tg ranging from 10 K to 110 K on solute diffusion coefficients were investigated 382

    in four different polymer hosts above their respective Tg, including: polylactide (PLA), 383

    polypropylene (PP), polycaprolactone (PCL) and polyvinyl alcohol (PVA). Tg of PVA was 384

    modulated by using it both at dry state and equilibrated at an intermediate relative humidity of 385

    21% with 2.4 wt% of water content. Ka and Kb values used in Eqs. (17) and (18) were 386

    exclusively fitted on D values of diphenyl alkanes obtained in PLA, PP and PCL at 343 K, 333 K 387

    and 323 K respectively. PVA data were used exclusively for external validation purposes. 388

    389

    Table 2. Information and characterization of processed films 390

  • Page | 21

    Polymer Tg (°C)

    Crytallinity %

    Thickness (mm) × width or diameter (mm)

    Film processing

    Supplier/product reference

    Polylactide (PLA)a 60 23.6 0.02×600 Extrusion Treofan (Germany)/ BiophanTM

    Polypropylene (PP)a 0 55.5 0.2×800 Extrusion blowing

    Borealis (Austria)/ HD621CF

    Polycaprolactone (PCL)b

    -60 50.3 0.01-0.04×200 Solution casting

    Creagif Biopolyméres (France)/ CAPA 6800

    Polyvinyl alcohol (PVA)c

    55 82

    50.0 0.01-0.03×200 Solution casting

    Sigma-Aldrich (USA)/ Mowiol® 20-98

    aFilms were processed at industrial scale and used as received.

    bPCL films with molecular 391

    weight of 8⋅104 g⋅mol-1 were processed at laboratory scale as described in43. cMolecular weight 392

    of PVA is 125000 g⋅mol-1 with 98.0-98.8 % of hydrolysis degree. Tg of PVA films were of 82°C 393

    and of 55°C, at dry state and when the films were equilibrated at a relative humidity of 21% 394

    respectively. All films were processed at laboratory scale as described in44

    . 395

    396

    3.2 Film processing and formulation 397

    PLA and PP were supplied as films and used as received. PCL and PVA films were processed 398

    by solution casting, according to43,44

    respectively. PCL dissolved in dichloromethane with 399

    concentration of 2.4 wt% was poured into a glass petri dish. After 24 h of evaporation at room 400

    temperature, PCL films were peeled off from the dish and dried in an oven at 30 °C for three 401

    days. PVA films were formed by applying the similar process by using deionized water as 402

    solvent with PVA concentration of 2 wt%. The evaporation of water took at least three days. 403

    Then, the films were conditioned under controlled relative humidity. 404

    Tg and crystallinity degree of each polymer were measured by differential scanning 405

    calorimetry (model Q100, TA Instruments, USA) at a heating rate of 10 °C/min within 406

    temperature limits adapted to each polymer. The crystallinity degree was calculated from the 407

  • Page | 22

    melting endotherm in the first heating scan with the help of the theoretical melting enthalpy of 408

    the 100 % crystalline polymer: 93 J⋅g-1 for PLA45, 165 J⋅g-1 for PP46, 139 J⋅g-1 for PCL47, 138.6 409

    J⋅g-1 for PVA48. Glass transition temperatures of all polymers except PVA were measured in the 410

    second heating scan and taken at the mid-point of the heat capacity step. In the case of both dry 411

    and plasticized PVA, Tg values were determined from the first heating scan. Determinations 412

    were triplicated. 413

    414

    3.3 Methods 415

    Diffusion coefficient determination 416

    According to expected values of diffusion coefficients, roughly below and above 10-14

    m2⋅s-1, 417

    two complementary solid-contact methods operating at two different length scales were used to 418

    reach contact times shorter than two weeks. It was checked that both methods gave similar 419

    diffusion coefficients for the solute common to both series: biphenyl. Films acting as sources of 420

    solutes were formulated with each solute either by soaking films in a 0.05 g⋅ml-1

    solute-ethanol 421

    solution during a minimal duration of one week at 60°C (cases of PP, PVA) or by adding the 422

    desired solute to the casting solution at a concentration of 0.2 wt% (case of PCL). Due to the 423

    difficulty of absorbing bulky aromatic solutes in PLA films below or close to its Tg (to avoid 424

    recrystallization), PP films were used as sources instead in PLA experiments. All processed films 425

    were stored stacked to prevent solute losses and to facilitate the internal homogenization of 426

    concentration profiles. The uniformity of concentration profiles in sources was tested over the 427

    cross section of microtomed films by fluorescence imaging. 428

    429

  • Page | 23

    A modified method originally proposed by49

    was used for high diffusion coefficients (above 430

    10-14

    m2⋅s-1, i.e. mainly in a temperature range of Tg+90 K and Tg+110 K). It consisted in 431

    stacking twelve virgin films with two source films formulated with the considered solute. 432

    Theoretically, by positioning sources in positions 5 and 10 in a stack consisting of 14 films with 433

    approximately the same thickness, the solute concentration profile evolve with diffusion time 434

    from a bimodal one to a monomodal one (with one single maximum located between films 7 and 435

    8). The variations in shape of the profile combined with the variations in concentration improved 436

    dramatically the comparison with the corresponding theoretical profiles expressed in function of 437

    the dimensionless position x/l and dimensionless time Fo=Dt/l2, where x is the position and l the 438

    average film thickness. In our experiment, films cut as disks were folded in aluminum foil and 439

    inserted in a copper cylinder of a same diameter. The cylinder was closed by two Teflon disks 440

    and the whole stack was packed with pressure by a screw system. Such conditions ensured good 441

    contact between films without mass transfer resistance and impervious conditions at both ends of 442

    the stack and on its lateral surface. After contact time at a constant temperature, solute 443

    concentration in each film was measured after solvent extraction in double beam UV/VIS 444

    spectrophotometer (model UVIKON 933, KONTRON Instruments, France). Extraction solvents 445

    were dichloromethane for PP and PCL, and deionized water for PVA. D values were retrieved by 446

    fitting numerically dimensionless theoretical concentration profiles (including the real geometry 447

    of each film) to measured concentration profiles. When several contact times or stack geometries 448

    were used for the same solute, D was defined as the regression slope of Fo versus x/l2. 449

    450

    For low diffusion coefficients, similar principles but at microscopic scale were used to reach 451

    similar contact times. A source film was sandwiched between two virgin films during a 452

  • Page | 24

    prescribed time. Diffusion was stopped by quenching films in liquid nitrogen, The concentration 453

    profile along the section of each film was subsequently determined, after microtoming (model 454

    LKB 2218 HistoRange, LKB-Produkter AB, Sweden) with cutting thickness of 15 µm, by 455

    deep-UV fluorescence microspectroscopy on an inverted microscope (model IX71, Olympus, 456

    Japan) in epi-configuration mode. A synchrotron source with a specific excitation wavelength 457

    ranging from 275 nm to 295 nm was used according to the tested solute (DISCO beamline, 458

    synchrotron Soleil, France). Bi-dimensional fluorescence emission spectra were acquired with a 459

    spectral resolution of 0.5 nm between 280 nm and 480 nm (LSM 710, Zeiss, Germany), and a 460

    spatial resolution of 0.5 or 1 µm. Concentrations were inferred from the surface area between the 461

    spectrum and the baseline. D values were identified similarly by a fitting procedure with a 462

    numerical model incorporating partitioning effects when required. Three to five concentration 463

    profiles along the same section were used in the fitting procedure. 464

    All experiments were triplicated or duplicated. All numerical models relied on finite volume 465

    difference scheme taking into account the real thickness of each film and a solver optimized for 466

    diffusion problems and distributed as an open-source project50

    . 467

    468

    Constrained molecular dynamics simulation 469

    Displacements of solute atoms available in constrained oligophenyls solutes were studied 470

    regardless the polymer host by constrained molecular dynamics. Starting from a random 471

    configuration, explicit constraints were applied by fixing the positions of all atoms of one or 472

    several phenyl rings. Long-term molecular dynamics simulations of isolated molecules subjected 473

    to such constraints were performed in the NVT ensemble at 298 K under vacuum boundary 474

    conditions using a Nosé-Hoover thermostat. For covalent and non-covalent interactions, 475

  • Page | 25

    COMPASS forcefield51

    was employed without any cutoff and dynamic simulations were 476

    conducted with Discover program (Accelrys, USA). Mean-square displacements of the center-of-477

    mass and covariances of atom displacements were averaged over several initial configurations. 478

    479

    4 RESULTS AND DISCUSSION 480

    4.1 Comparison of the scaling of D between linear aliphatic solutes and aromatic solutes 481

    The assumption of the scaling exponent α following a temperature dependence as proposed in 482

    Eq. (8) is tested in Fig. 2 for α values inferred from diffusion coefficients collected at different 483

    temperatures on a semi-crystalline polyethylene (PE)3, 14-16 and on amorphous polyamide, based 484

    upon poly-(ε-caprolactam) lightly cross-linked with diglycidyl ether of bisphenol A5. It is 485

    important to notice that reported α values are associated to diffusion coefficients measured only 486

    within the same study (with the same polymer and the same measurement protocol) and for 487

    solutes larger than 70 g⋅mol-1. Tg values involved in Eq. (8) were either the reported ones or 488

    derived from the host polymer mass. As applied in Fig. 1, solute series were categorized as 489

    linear aliphatic and aromatic solutes. The “linear aliphatic” series merged linear n-alkanes, 490

    n-alcohols and esterified phenols with long n-alkyl chains (n=6..18) whereas the “aromatic” 491

    series combined short n-alkylbenzenes (n=0..4), substituted and hindered phenols (M= 94-545 492

    g⋅mol-1). Values of α generated by this study for oligophenyls were excluded from the fitting 493

    procedure and included only a posteriori as external validation of Eq. (8). 494

  • Page | 26

    495 Figure 2. Scaling exponents of trace diffusion coefficients versus T-Tg. Symbols plot 496

    determinations of α inferred from experimental D values reported in references enclosed within 497

    brackets. Values for oligophenyls measured by this study were not used to fit Eq. (8). The 498

    classification of solutes as linear aliphatic and aromatic is based on the same one as used in Fig. 499

    1. 500

    501

    Eq. (8) fitted well α values of both linear aliphatic solutes and aromatic solutes and also offers 502

    a continuous reconciliation between the diffusion behavior in solids and melts. Master curves 503

    looked homothetic with a positive temperature shift of 91 K from linear aliphatic solutes to 504

    aromatic solutes, corresponding to a Kβ value of 40 K and -51 K respectively. At high 505

    temperature, a unitary coefficient α, as predicted by the Rouse theory13

    , has been found by von 506

    Meerwall et al.3 for linear aliphatic solutes, but it has not still been described for aromatic 507

    solutes. The similar sharp decrease observed of aromatic solutes suggests, however, that an 508

    exponent close to unity could be universal for both linear aliphatic and aromatic solutes in 509

  • Page | 27

    aliphatic polymers and by extension in elastomers and thermosets far above their Tg. At 510

    intermediate temperatures (in semi-crystalline or below the flow threshold of the polymer), 511

    exponents increase dramatically when the temperature is decreasing, with literature values 512

    reported up to 4.75. The results obtained in this study on oligophenyls enabled an external 513

    validation of Eq. (8) applied to aromatic solutes near Tg where no data has been published. The 514

    comparison showed a good agreement whatever the considered aliphatic polymer, PP and PCL, 515

    and demonstrated the existence of α values even larger than previously reported ones. 516

    517

    As no existing theory holds for exponents α larger than 1 and solutes that are non-entangled 518

    with polymer segments, some analogies with the scaling of self-diffusion in monodisperse 519

    systems are first discussed. In monodisperse mixtures of non-entangled n-alkanes (M=114-844 520

    g⋅mol-1), scaling exponents were described to decrease almost linearly from 2.72 to 1.85 when 521

    temperature was increasing from 303 K to 443 K52, 53

    . Tg of corresponding liquid n-alkanes are 522

    expected to be lower than polyethylene (theoretically 200 K for an infinitely long 523

    polyethylene54

    ), with values ranged between 51 K and 186 K (according to the equation of Fox 524

    and Loshaek55

    and parameters fitted in56, 57

    ). According to Fig. 2, self-diffusion could be 525

    envisioned equivalently as trace diffusion in a host with a very low Tg. In a small range of 526

    temperatures centered around a reference temperature θ, the linear decrease of α(T) is 527

    particularly granted by the asymptotic behavior of Eq. (8) towards the melt region. When 528

    max 0,T Tg K , Eq. (8) is indeed approximated by529

    2

    1T K T Tg K T T Tg . The absence of effect of Kβ far from Tg 530

    suggests that only thermal expansion effects of the polymer host dominate. They are controlled 531

    by the value of Kα, which was found very similar for both linear and aromatic solutes, 144 K and 532

  • Page | 28

    156 K respectively. Similar arguments were used to explain the thermal dependence on α in non-533

    entangled monodisperse systems52, 53

    .They may be nevertheless approximate because not only 534

    the static properties (i.e. density, free volumes distribution… of the polymer) are affected by the 535

    host molecular mass but also the dynamic ones (i.e. Tg is increasing with M)3, 4

    . 536

    Near Tg, the variation of static polymer effects cannot be invoked alone to justify the large 537

    values of α. We related phenomenologically this additional effect in Eq. (8) to a guest parameter, 538

    formally -Kβ, which can be envisioned as a critical temperature deviation to Tg to translate an 539

    elemental jumping unit. For solutes consisting in the repetition of a similar jumping unit, Kβ was 540

    thought to be constant: positive when the jumping unit resembles polymer segments and can 541

    easily accommodate the fluctuations of the contour of polymer segments; and negative 542

    otherwise. The concepts of accommodation between a bulky guest molecule (e.g. aromatic 543

    fluorescent dye) and host aliphatic chains has been studied in low molecular weight alkanes58

    . 544

    Coarse grained simulation of molecular dynamics of spherical solutes larger than the polymer 545

    beads39

    confirmed further the proposed description involving both polymer static and dynamic 546

    effect: trace diffusion coefficients were found scaled as a power law of the volume of the bead 547

    with a scaling exponent increasing from 0.8 to 1.43 when the stiffness of the polymer host 548

    increased. 549

    550

    4.2 Scaling diffusion coefficients according to Eqs. (1), (17)-(18) 551

    Phenomenological scaling of D with M at different temperatures for both tested aromatic 552

    solute series is depicted in Fig. 3 along with the predictions according to Eqs. (17) and (18). One 553

    important goal is to demonstrate that the temperature shift factor associated to D depends on 554

    some solute contributions and that proposed equations prolong naturally the conventional 555

  • Page | 29

    Williams-Landel-Ferry model (see Eqs. 9 and 15). To test the proposed free volume theory, the 556

    following fitting procedure was applied. Polymer related parameters, Ka and Kb, were 557

    exclusively fitted from the D values of diphenyl alkanes series at constant temperature (i.e. from 558

    D values obtained in different polymers with different Tg). The determined values of Ka and Kb 559

    values were directly applied to each oligophenyl solute to extract Easolute(NPh) and Ssolute(NPh). 560

    External validation was finally achieved by predicting D values of biphenyl in plasticized and 561

    unplasticized PVA, where the experimental D values of PVA were evidently not included in the 562

    fitting procedure. 563

    Fig. 3 shows the very strong mass dependence on D values for both diphenyl alkanes and 564

    oligophenyls, with α decreasing with increasing temperature from 24 to 20 (95% confidence 565

    interval ∓ 4.5), and from 5.3 to 4.3 (95% confidence interval ∓ 1.6), respectively. Such values 566

    were far from values previously reported for aromatic solutes (see Fig. 2) and varied in a small 567

    extent with tested temperatures and considered polymers. As depicted in Fig. 4, Eqs. (17) and 568

    (18) fitted well the broad distribution of D over five decades for all tested polymers, with a 569

    fitting error distributed normally and in the range of experimental errors. Predictions of D values 570

    of biphenyl in an external polymer (PVA) by Eq. (17) were also in good agreement with 571

    experimental data for both dry and plasticized PVA. The predictions for dry and plasticized PVA 572

    confirmed the temperature-humidity induced plasticization superposition assessed with 573

    fluorescent diffusion probes in polyamide59

    . Such preliminary comparisons between 574

    experimental values and model ones justify globally the separation of the polymer and solute 575

    contributions for aromatic solutes in aliphatic polymers. Both effects are analyzed separately 576

    hereafter. 577

  • Page | 30

    578

    Figure 3. a) Log-Log plot of trace diffusion coefficients of diphenyl alkanes and b) oligophenyls 579

    in various polymers. Symbols are experimental values. Continuous straight lines are scaling 580

    relationships fitted according to Eq. (1). Dashed lines are values fitted from Eqs. (17) and (18) 581

    for diphenyl alkanes and oligophenyls respectively. 582

  • Page | 31

    583

    Figure 4. Comparison between calculated (with Eqs. (17) and (18) respectively) and measured 584

    diffusion coefficients of a) diphenyl alkanes series and b) oligophenyls series in PLA, PP, 585

    PCL(×) and PVA(▲).The continuous lines plot the straight line y=x. The corresponding 586

    distribution of relative fitting errors values and fitted Gaussian distribution (continuous curve) 587

    are plotted within insets. Values in PVA are external validation predictions not used in the fitting 588

    procedure. 589

    590

    4.3 Polymer effects as probed with diphenyl alkanes 591

    Diphenyl alkanes series presented several remarkable features to probe polymer effects. 592

    Firstly, solutes corresponding to a small range of NC values (here 0, 1, 2) are of similar size so 593

  • Page | 32

    that they are probing almost the same size of free volume pockets. Secondly, increasing the 594

    distance between phenyl rings enabled to assess the effect of expected higher correlations with 595

    the surroundings, C(t). Finally, the large spread of diffusion coefficients with NC improved the 596

    accuracy on polymer parameter estimates, Ka and Kb, used in Eqs. (17) and (18). 597

    By noting , 0 exp apolymer Cb

    KD D T N

    K T Tg

    , Fig. 5 plots both the scaling of D and 598

    the solute contribution defined as D/Dpolymer versus the number of carbon atoms, NC, at a constant 599

    temperature. The inferred scaling did not depend on the considered polymer and was associated 600

    to a 10C

    N value of 1.3 in Eq. (17), which implies that D decreases 10-folds when 1.3 carbons is 601

    added between the two phenyl rings. 602

    603

  • Page | 33

    Figure 5. Scaling diffusion coefficients of diphenyl alkanes with the number of carbons, NC, 604

    between phenyl rings: a) raw diffusion coefficients measured at 343 K for PLA, 333 K for PP, 605

    and 323 K for PCL b) diffusion coefficients relative to biphenyl , 0CD T N . Eq. (17) 606

    predictions are plotted as continuous lines. 607

    608

    The corresponding polymer contribution was assessed as D/Dsolute with 609

    10

    , 0 exp ln10 Csolute CC

    ND D T N

    N

    . D and D/Dsolute in polymers used for fitting (PLA, PP, 610

    PCL) and validation (PVA) are plotted versus T-Tg in Figs. 6a and 6b respectively. Results 611

    showed accordingly an evolution of D/Dsolute that was independent of the considered solute and 612

    where polymer effects were finally reduced to an effect of the distance to Tg. Values of Ka and 613

    Kb were found equal to 600 K and 58 K respectively, and predicted independently with an 614

    acceptable accuracy of the D values of biphenyl in both dry and plasticized PVA. The estimated 615

    value of Kb is of the same magnitude order as the value of 50 K for the polymer-related 616

    parameter K22 reported in30

    , when polystyrene is probed with toluene and ethylbenzene. As 617

    reported in32

    , such ranges of Kb were assumed to be generic for linear polymers and therefore 618

    also valid for oligophenyls series too. According to Eqs. (9) and (15), it could be thought that 619

    (Ka, Kb) (i.e. fitted on our diphenyl alkanes data) and (Kα, K ) (i.e. fitted exclusively from 620

    literature data on different aromatic solutes) might be also related together. However, it is 621

    expected to be true only if no additional energy barrier exists. In our study, it is very likely for a 622

    solute comprising only one single jumping unit. By taking the molecular mass of benzene 623

    (M/M0=78 with M0=1 g⋅mol-1

    ) and the average value of Kα reported in section 4.1 (150 K), we 624

  • Page | 34

    get a value Kα⋅lnM of 654 K close to the value of Ka (ca. 600 K) reported here. The agreement 625

    between Kb and K is even more convincing with values of 58 K and 51 K, respectively. 626

    627

    Figure 6. Experimental diffusion coefficients of diphenyl alkanes in PLA, PP, PCL (empty 628

    symbols) and PVA (filled symbols) versus T-Tg a) raw diffusion coefficients, b) D/Dsolute. Eq. 629

    (17) fitted on empty symbols is plotted as continuous lines. Filled symbols are used for external 630

    validation purposes. 631

    632

    4.4 Solute activation parameters of oligophenyls 633

    Diffusion coefficients of oilgophenyls exhibited much a lower dependence with molecular 634

    mass, which was associated in Eq. (18) to a free energy barrier, which is also a function of the 635

  • Page | 35

    number of phenyl rings. Fig. 7 plots the dependence of D (Figs. 7a and 7b) and of D scaled by 636

    the polymer exponential factor, exp a

    b

    K

    K T Tg

    , (Figs. 7c and 7d) versus the number of 637

    phenyl rings, NPh. Whatever the considered temperature and tested polymer, the exponential 638

    decrease of D with NPh was non-regular, suggesting a non-monotonous variation of the energy 639

    barrier to translation with NPh. The trend is confirmed by plotting both D and D scaled by 640

    polymer effects on a van’t Hoff plot in Fig. 8. While a significant deviation to an Arrhenius 641

    behavior was observed on raw diffusion coefficients in polymers close to their Tg as shown in 642

    Fig. 8a and 8b, a pure Arrhenius behavior was recovered by contrast once the effects of density 643

    were corrected. It was particularly interesting to notice that the slope of the van’t Hoff plot was 644

    systematically much higher for p-terphenyl than those of the former and following solutes in the 645

    series. 646

  • Page | 36

    647

    Figure 7. Scaling of diffusion coefficients of oligophenyls with the number of phenyl rings, NPh, 648

    a,c) in PCL and b,d) in PP at different temperatures: a,b) raw diffusion coefficients, c,d) scaled 649

    diffusion coefficients with polymer effects removed. Predictions according to Eq. (18) are 650

    plotted as continuous lines. 651

  • Page | 37

    652

    Figure 8. Normalized van’t Hoff plots of oligophenyls a,b) raw diffusion coefficients and c,d) 653

    scaled diffusion coefficients with polymer effect removed a,c) in PCL and b,d) in PP. 654

    Predictions according to Eq. (18) are plotted as continuous lines. 655

    656

    The non-monotonous variations of solute activation energies, Easolute, and entropies, Ssolute, are 657

    specifically captured in Figs. 9b and 9d. Raw activation values, as estimated from Figs. 8a and 658

    8b, are also given in Figs. 9a and 9c. Regardless polymer effects were included or not, activation 659

    parameters exhibited systematically a hat shape with a maximum for NPh=3. Such 660

    non-monotonous variations appeared for both tested polymers (PP and PCL) with apparent 661

    activation energies and entropies of PP shifted from PCL ones by approximately 35 kJ⋅mol-1 and 662

  • Page | 38

    95 J⋅mol-1⋅K-1, respectively. PP and PCL activation values were exactly reconciled once 663

    polymer effects were removed by Eq. (18). It is worth to notice that the reconciliation was 664

    obtained by removing the polymer contribution, whereas it was inferred independently from the 665

    D values of diphenyl alkanes. Since both enthalpy and entropy exhibited similar shape, an 666

    apparent position correlation between both quantities was found as plotted in Fig. 9e. Enthalpy-667

    entropy compensation is often considered to be a statistical artifact due to correlations between 668

    errors on each estimate and because entropy must be extrapolated to infinite temperature. As 669

    polymer effects were removed in our case, it could however be thought that the extrapolation of 670

    solute effects regardless the true physical state of the polymer at an infinite temperature can have 671

    a reasonable meaning. Such kind of discussions can be found in60

    . The corresponding internal 672

    free energy barriers at 298 K required to induce a translation of oligophenyls are represented in 673

    Fig. 9f. As activation energy and entropy contribute with opposite signs to the free energy 674

    barrier, free energy was found monotonous with NPh. 675

  • Page | 39

    676

    Figure 9. a,c) Raw and b,d) solute activation energies and diffusion entropy of oligophenyls in 677

    PP and PCL versus the number of phenyl rings, NPh; e) correlation between solute activation 678

    parameters; f) related free barrier energy to diffusion at 298 K. 679

    680

    4.5 Mechanisms of translation of oligophenyls in aliphatic polymers 681

    Diffusion of large organic solutes in polymer hosts cannot be directly studied by molecular 682

    dynamics close to their Tg at atomistic scale. Our experimental results highlighted however two 683

  • Page | 40

    noticeable properties to derive tractable simulations and to gain further insights on the translation 684

    mechanisms of linear aromatic molecules in rubber polymers: 685

    - The displacements of each jumping unit (i.e. “polymer effects” in the text) and “solute 686

    effects” (i.e. collective displacements of jumping units) on D are separable; 687

    - Activation energies associated to the displacement of several jumping units vary with the 688

    parity of the number of jumping units. Such a feature can be directly tested by simulation. 689

    The main idea was to test whether two correlation modes between the displacements of phenyl 690

    rings and surrounding atoms could explain the effect of the parity of NPh on activation terms. 691

    Since neither the partitions between both modes nor the correlation times are known, our strategy 692

    consisted in studying the translation of CM in constrained long-term molecular dynamics (10 ns 693

    or more), where the positions of one or several phenyl rings (i.e. jumping units) are kept fixed, 694

    under vacuum boundary conditions (i.e. without polymer host). The minimum times to enable a 695

    translation of CM longer than a phenyl ring diameter (i.e. length of one jumping unit) and the 696

    related translational entropy were analyzed according to Eqs. (19) and (24) respectively and 697

    finally compared to their experimentally inferred counter-parts: Easolute and Ssolute. By fixing a 698

    priori different values to p, it was in particular possible to assess which value could reproduce a 699

    non-monotonous variation with the number of jumping units. 700

    The results obtained by averaging over a wide range of initial configurations are plotted in Fig. 701

    10 and compared to experimental values reported in Fig. 9. The typical “hat shape” of activation 702

    energies was particularly reproduced without significant bias with RTln(τring-τtrapped) calculated 703

    from Eqs. (19) and (20) when p→1 (Fig. 10a). Ssolute values derived from Eqs. (21)-(24) led also 704

    approximately to the similar trend when p→1 (Fig. 10b). The theoretical translational entropy 705

    underestimated however systematically the real one due to the loss of fluctuations information 706

  • Page | 41

    caused by the fixed positions of some phenyl rings. It is worth to notice that the reported “hat-707

    shape” of activation terms could not be predicted with the simple coarse-grained theory 708

    supported either in Eq. (5) or in Eq. (7) because they assume that the displacements of all phenyl 709

    rings are equivalent. In reality, the gyration radius and mean-square-displacement of 710

    oligophenyls are dramatically affected by the combination of phenyl rings that are fixed. Among 711

    all tested oligophenyls (NPh=2,3,4), p-terphenyl is the one that offers the highest ratio of 712

    possibilities to reduce dramatically the fluctuations of CM and other atoms. Particularly efficient 713

    configurations (i.e. five over the 23 possibilities): blocking the central ring alone or blocking 714

    randomly two or three rings. Such effects could not be captured by a generic flexible model that 715

    assume a uniform relaxation model along the chain (see Eq. (3)) or uniform covariances between 716

    connected jumping units (see Eq. (4)) and were consequently better reproduced by atomistic 717

    simulations. It is particularly noticeable that increasing p in simulations induced enthalpy-718

    entropy compensation as experimentally assessed so that the partitioning between short and long-719

    lived contacts could be proposed as the main cause of the phenomenon. Corollary, the suggested 720

    high value of p (i.e. at least one jumping unit has long-lived contacts with surrounding host) does 721

    not depend on the length of studied solutes and could be general for all linear aromatic solutes 722

    within aliphatic polymers. 723

  • Page | 42

    724

    Figure 10. a) Comparison of relative solute activation energy (continuous lines, left scale), 725

    calculated as ring rinPh Phg trappedN N for different p values according to Eqs. (19)-(20), 726

    with experimental values (dotted lines, right scale) reported in Fig. 9b. The horizontal dashed 727

    line represents the average value of Easolute for studied oligophenyls. b) Comparison of solute 728

    entropy (continuous lines, left scale), calculated for different p values according to Eqs. (21)-729

    (24), with experimental values (dotted line, right scale) reported in Fig. 9d. 730

    731

  • Page | 43

    5 CONCLUSIONS 732

    The presented work introduces several invariant scaling relationships for the trace diffusion 733

    coefficients (D) of bulky solutes in thermoplastic polymers above their Tg and presents a first 734

    molecular interpretation of corresponding translation mechanisms. D values over five decades 735

    were measured for two homologous series of short aromatic solutes, diphenyl alkanes and 736

    oligophenyls, in different polymers and at various temperatures, T. The collected values 737

    completed the available picture from literature by integrating D values for low T-Tg values. The 738

    analysis of the scaling with molecular mass, as D∝M-α(T-Tg), shows that aromatic solutes have a 739

    parallel behavior to linear aliphatic solutes but shifted by ca. +91 K. The proposed scaling is 740

    notably able to reconcile α values experimentally assessed in melts3 and in solids. It could be 741

    thought that the temperature shift would include the behavior all organic solutes larger than 742

    voids, with left and right bounds set by linear aliphatic and aromatic solutes respectively. 743

    However, α values derived for diphenyl alkanes showed that aromatic solutes including a 744

    flexible segment between two phenyl rings were associated to a much larger temperature shift. 745

    Such dramatic effect of the solute chemical structure was found independent of the considered 746

    aliphatic polymer and associated only to the size and type of the solute jumping unit. It would 747

    explain both the logarithm dependence of the apparent activation energy with molecular mass 748

    and the additional deviation to the Arrhenius behavior close to Tg, as specifically discussed in4. 749

    Because the specific volume of polymers is also a function of T-Tg, a modification of the 750

    distribution of free volume in the polymer must be also invoked, as already suggested for the 751

    self-diffusion of non-entangled n-alkanes52

    . 752

    To reach a consistent description of all effects, an extended free volume theory is proposed 753

    assuming that the free volume fluctuations of polymer host segments, which control the short-754

  • Page | 44

    range translation of each jumping unit and finally the displacements of many jumping units obey 755

    to two independent statistical distributions with different activations by temperature. Static 756

    polymer effects were associated to a single exponential term, exp a bK T Tg K . It was 757

    verified thus that the values of 600 K and 58 K, proposed for of Ka and Kb respectively, enabled 758

    the extrapolation of known diffusion coefficients from one aliphatic polymer to another one (e.g. 759

    unplasticized to plasticized, apolar to polar), at least at a similar temperature. If the diffusion 760

    coefficient needs to be extrapolated to a different temperature, the solute-related activation needs 761

    to be accounted specifically. Solute effects were found to follow an Arrhenius’ law for 762

    oligophenyls including from two to four phenyl rings. Related activation energies and entropies 763

    were, however, highlighted to vary non-monotonously with M. 764

    This outstanding effect of the parity of the number of ring was used to assess, via constrained 765

    molecular dynamics simulations, the probability of phenyl rings (jumping units) to behave as 766

    anchors in the translation mechanisms of aromatic solutes. It is argued that, though the polymer 767

    at rubbery state does not control the reorientation frequency of the entire molecule (i.e. each 768

    phenyl ring is displacing independently while being limited by the connectivity of rings), it tends 769

    to hinder the reorientation of individual phenyl rings. Our simulations show that constraining 770

    randomly one or several phenyl rings slows down the mean-square-displacement of the center-771

    of-mass (CM) of solutes in a different way according to the number of rings, NPh, being odd or 772

    even. For example, when NPh=3, it is twice more likely to block a ring at one end than in the 773

    middle; so that the slowdown is stronger than for NPh=2 or 4. As our interpretation matched 774

    remarkably the relative energies experimentally determined on oligophenyls series, it is thought 775

    that the proposed anchor effect of phenyl rings is universal in aliphatic polymers at least between 776

  • Page | 45

    Tg+51 K and Tg+150 K, where α was estimated to be much larger than unity (as shown in Fig. 777

    2). 778

    Presented results should find applications in many domains where diffusion coefficients of 779

    aromatic molecules and polymer residues are particularly critical: contamination by substances 780

    leeched from polymer materials, loss of additives during physical ageing of polymers, 781

    reactivities in polymers in processing and use conditions. In particular, reported results open the 782

    way to design of substances with low diffusion coefficients with an odd number of rings and/or 783

    with flexible segments close to CM. The behavior of branched aromatic molecules will be 784

    presented in a companion paper. 785

    786

    6 ACKNOWLEDGEMENTS 787

    The first authors would like to acknowledge the support of the PhD grant from Région Île-de-788

    France. We also gratefully thank Dr. Frédéric Jamme for his technical assistance. This work was 789

    supported by the DISCO beamline of the synchrotron Soleil (Proposal 20100909). 790

    791

    7 REFERENCES 792

    (1) Vitrac, O.; Lézervant, J.; Feigenbaum, A. Journal of Applied Polymer Science 2006, 101, 793

    (4), 2167-2186. 794

    (2) Berens, A. R. Pure Appl. Chem. 1981, 53, (2), 365-375. 795

    (3) von Meerwall, E. D.; Lin, H.; Mattice, W. L. Macromolecules 2007, 40, (6), 2002-2007. 796

    (4) Durand, M.; Meyer, H.; Benzerara, O.; Baschnagel, J.; Vitrac, O. The Journal of 797

    Chemical Physics 2010, 132, (19), 194902. 798

    (5) Kwan, K. S.; Subramaniam, C. N. P.; Ward, T. C. Polymer 2003, 44, (10), 3061-3069. 799

    (6) Gennes, P. G. d. The Journal of Chemical Physics 1971, 55, (2), 572-579. 800

    (7) Bueche, F. The Journal of Chemical Physics 1968, 48, (3), 1410-1411. 801

    (8) Lodge, T. P. Physical Review Letters 1999, 83, (16), 3218-3221. 802

  • Page | 46

    (9) Quijada-Garrido, I.; Barrales-Rienda, J. M.; Frutos, G. Macromolecules 1996, 29, (22), 803

    7164-7176. 804

    (10) Phillips, P. J. Chemical Reviews 1990, 90, (2), 425-436. 805

    (11) Wirtz, A. C.; Hofmann, C.; Groenen, E. J. J. ChemPhysChem 2011, 12, (8), 1519-1528. 806

    (12) Fechete, R.; Demco, D. E.; Blümich, B. Macromolecules 2003, 36, (19), 7155-7157. 807

    (13) Rouse, J. P. E. The Journal of Chemical Physics 1953, 21, (7), 1272-1280. 808

    (14) Vitrac, O.; Mougharbel, A.; Feigenbaum, A. Journal of Food Engineering 2007, 79, (3), 809

    1048-1064. 810

    (15) Koszinowski, J. Journal of Applied Polymer Science 1986, 31, (8), 2711-2720. 811

    (16) Möller, K.; Gevert, T. Journal of Applied Polymer Science 1994, 51, (5), 895-903. 812

    (17) Doong, S. J.; Ho, W. S. W. Industrial & Engineering Chemistry Research 1992, 31, (4), 813

    1050-1060. 814

    (18) Herman, M. F.; Tong, P. Macromolecules 1993, 26, (15), 3733-3737. 815

    (19) Herman, M. F. The Journal of Chemical Physics 1995, 103, (10), 4324-4332. 816

    (20) Herman, M. F.; Panajotova, B.; Lorenz, K. T. The Journal of Chemical Physics 1996, 817

    105, (3), 1153-1161. 818

    (21) Herman, M. F.; Panajotova, B.; Lorenz, K. T. The Journal of Chemical Physics 1996, 819

    105, (3), 1162-1174. 820

    (22) Gniewek, P.; Kolinski, A. The Journal of Chemical Physics 2011, 134, (5), 056101. 821

    (23) Skolnick, J.; Yaris, R.; Kolinski, A. The Journal of Chemical Physics 1988, 88, (2), 1407-822

    1417. 823

    (24) Williams, M. L.; Landel, R. F.; Ferry, J. D. Journal of the American Chemical Society 824

    1955, 77, (14), 3701-3707. 825

    (25) Ehlich, D.; Sillescu, H. Macromolecules 1990, 23, (6), 1600-1610. 826

    (26) Deppe, D. D.; Miller, R. D.; Torkelson, J. M. Journal of Polymer Science Part B: 827

    Polymer Physics 1996, 34, (17), 2987-2997. 828

    (27) Doolittle, A. K. Journal of Applied Physics 1951, 22, (8), 1031-1035. 829

    (28) Cohen, M. H.; Turnbull, D. The Journal of Chemical Physics 1959, 31, (5), 1164-1169. 830

    (29) Vrentas, J. S.; Duda, J. L. Journal of Polymer Science: Polymer Physics Edition 1977, 831

    15, (3), 403-416. 832

    (30) Vrentas, J. S.; Vrentas, C. M. Macromolecules 1994, 27, (17), 4684-4690. 833

    (31) Vrentas, J. S.; Vrentas, C. M. European Polymer Journal 1998, 34, (5–6), 797-803. 834

    (32) Mauritz, K. A.; Storey, R. F.; George, S. E. Macromolecules 1990, 23, (2), 441-450. 835

    (33) Mauritz, K. A.; Storey, R. F. Macromolecules 1990, 23, (7), 2033-2038. 836

    (34) Coughlin, C. S.; Mauritz, K. A.; Storey, R. F. Macromolecules 1990, 23, (12), 3187-837

    3192. 838

  • Page | 47

    (35) Coughlin, C. S.; Mauritz, K. A.; Storey, R. F. Macromolecules 1991, 24, (7), 1526-1534. 839

    (36) Coughlin, C. S.; Mauritz, K. A.; Storey, R. F. Macromolecules 1991, 24, (8), 2113-2116. 840

    (37) Hayashi, H.; Matsuzawa, S. Journal of Applied Polymer Science 1992, 46, (3), 499-505. 841

    (38) Solunov, C. A. Journal of Physics: Condensed Matter 2002, 14, (31), 7297. 842

    (39) Budzien, J.; McCoy, J. D.; Rottach, D.; Curro, J. G. Polymer 2004, 45, (11), 3923-3932. 843

    (40) Kovarskiĭ, A. L., Molecular Dynamics of Additives in Polymers. VSP: 1997. 844

    (41) Andricioaei, I.; Karplus, M. The Journal of Chemical Physics 2001, 115, (14), 6289-845

    6292. 846

    (42) Schafer, H.; Mark, A. E.; Gunsteren, W. F. v. The Journal of Chemical Physics 2000, 847

    113, (18), 7809-7817. 848

    (43) Marras, S. I.; Kladi, K. P.; Tsivintzelis, I.; Zuburtikudis, I.; Panayiotou, C. Acta 849

    Biomaterialia 2008, 4, (3), 756-765. 850

    (44) Otsuka, E.; Suzuki, A. Journal of Applied Polymer Science 2009, 114, (1), 10-16. 851

    (45) Fischer, E. W.; Sterzel, H. J.; Wegner, G. Colloid & Polymer Science 1973, 251, (11), 852

    980-990. 853

    (46) Wunderlich, B., Macromolecular physics. Academic Press: 1980. 854

    (47) Crescenzi, V.; Manzini, G.; Calzolari, G.; Borri, C. European Polymer Journal 1972, 8, 855

    (3), 449-463. 856

    (48) Peppas, N. A.; Merrill, E. W. Journal of Applied Polymer Science 1976, 20, (6), 1457-857

    1465. 858

    (49) Roe, R.-J.; Bair, H. E.; Gieniewski, C. Journal of Applied Polymer Science 1974, 18, (3), 859

    843-856. 860

    (50) Vitrac, O. FMECA engine: software developed in the framework of the project 861

    SafeFoodPack Design https://github.com/ovitrac/FMECAengine (July 17, 2012), 862

    (51) Sun, H. The Journal of Physical Chemistry B 1998, 102, (38), 7338-7364. 863

    (52) von Meerwall, E.; Beckman, S.; Jang, J.; Mattice, W. L. The Journal of Chemical Physics 864

    1998, 108, (10), 4299-4304. 865

    (53) Harmandaris, V. A.; Doxastakis, M.; Mavrantzas, V. G.; Theodorou, D. N. The Journal 866

    of Chemical Physics 2002, 116, (1), 436-446. 867

    (54) Miller, A. A. Journal of Polymer Science Part A-2: Polymer Physics 1968, 6, (1), 249-868

    257. 869

    (55) Fox, T. G.; Loshaek, S. Journal of Polymer Science 1955, 15, (80), 371-390. 870

    (56) Waheed, N.; Ko, M.; Rutledge, G., Atomistic Simulation of Polymer Melt Crystallization 871

    by Molecular Dynamics. In Progress in Understanding of Polymer Crystallization, 872

    Reiter, G.; Strobl, G., Eds. Springer: Berlin, Heidelberg, 2007; Vol. 714, pp 457-480. 873

    (57) Reiter, G.; Strobl, G. R., Progress in Understanding of Polymer Crystallization. Springer: 874

    2007. 875

  • Page | 48

    (58) Park, H. S.; Chang, T.; Lee, S. H. The Journal of Chemical Physics 2000, 113, (13), 876

    5502-5510. 877

    (59) Hatzigrigoriou, N. B.; Vouyiouka, S. N.; Joly, C.; Dole, P.; Papaspyrides, C. D. Journal 878

    of Applied Polymer Science 2012, 125, (4), 2814-2823. 879

    (60) Starikov, E. B.; Nordén, B. The Journal of Physical Chemistry B 2007, 111, (51), 14431-880

    14435. 881

    882 883

  • For Table of Contents use only

    Page | 49

    Diffusion of aromatic solutes in aliphatic polymers above glass 884

    transition temperature 885

    Xiaoyi Fang2, Sandra Domenek

    2, Violette Ducruet

    1, Matthieu Refregiers

    3, Olivier Vitrac

    1* 886

    887

    888

    889