Accepted Manuscript From thermodynamics to kinetics: Theoretical study of CO2 dissolving in poly (lactic acid) melt
Kesong Yu, Hongfu Zhou, Xiangdong Wang, Zhongjie Du, Jianguo Mi PII: DOI: Reference:
S0167-7322(18)34348-4 https://doi.org/10.1016/j.molliq.2019.02.005 MOLLIQ 10405
To appear in:
Journal of Molecular Liquids
Received date: Revised date: Accepted date:
22 August 2018 1 December 2018 1 February 2019
Please cite this article as: K. Yu, H. Zhou, X. Wang, et al., From thermodynamics to kinetics: Theoretical study of CO2 dissolving in poly (lactic acid) melt, Journal of Molecular Liquids, https://doi.org/10.1016/j.molliq.2019.02.005
This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT From Thermodynamics to Kinetics: Theoretical Study of CO2 Dissolving in Poly (lactic acid) Melt
b
IP
School of Materials and Mechanical Engineering, Beijing Technology and Business University, Beijing, China
CR
a
T
Kesong Yua,b, Hongfu zhoua, Xiangdong Wanga*, Zhongjie Dub, and Jianguo Mi b*
State Key Laboratory of Organic-Inorganic Composites, Beijing University of Chemical Technology, Beijing,
US
China
AN
Abstract: Under the framework of dynamic density functional theory, the dissolution and diffusion behaviors
AC
CE
PT
ED
M
of CO2
AC
CE
PT
ED
M
AN
US
CR
IP
T
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT 1. Introduction Poly (lactic acid) (PLA) is an aliphatic thermoplastic polyester with the biodegradable attribute and comparable mechanical properties [1,2]. PLA foam has the potential to replace traditional petroleum-based polymer foams in a wide array of applications [3-5]. Moreover, PLA has relatively high CO2 solubility, which provides a good prerequisite for developing nanocellular cells with uniform cell morphology to further enhance
T
its impact resistance, toughness, and thermal and sound insulation [6-8]. It is known that the cell morphology is
IP
generally governed by the following thermodynamic and kinetic properties: CO2 dissolution, PLA swelling,
CR
CO2-PLA interfacial tension, CO2 diffusion, and PLA viscosity. In order to confirm the condition to produce
US
nanocellular PLA foam, these fundamental issues should be clearly identified in advance. It is not easy to handle and ensure the dissolution equilibrium of CO2 in highly PLA melt in experiment, thus
AN
the available experimental data on CO2 solubility are restricted in a narrow range of temperature and pressure [9-11]. Meanwhile, the swelling ratio of PLA depends on how much CO2 dissolves in PLA [12], and the
M
corresponding interfacial structure and tension of CO2-PLA has to be regulated due to PLA plasticization
ED
[13,14]. The swelling and interfacial properties are more difficult to measure via experimental approaches. A reliable thermodynamic model, capable of accurately describing CO2 solubility, the swelling ratio of glassy
PT
PLA, and the interfacial structure and property between two phases over a wide range of temperature and
AC
CE
pressure, is an essential prerequisite for developing a good transport model.
The diffusion of CO2 in PLA is a complex process as
well. It can be influenced by concentration, pressure, temperature, PLA viscosity, external force and so on. For these reasons, the measurement of diffusivity of supercritical CO2 in molten polymer has always been a challenging research topic for the foam processing [15]. T 16,17] Theoretical prediction is an alternative approach to reduce the workload of experiment and to provide more efficient data aimed at predicting the cell size and density of foaming products.
ACCEPTED MANUSCRIPT A multitude of equations of state were applied to calculate and predict gas solubility in polymer systems [1822]. Through the proper choice of specific model parameters regressed from experimental data, these equations are capable of semi-quantitatively evaluating the dissolving behavior. However, the physical meaning of these parameters is obscure, because
T
restrict the usefulness of equations as the predictive tools.
IP
Another important type of equations was derived from the statistical associating fluid theory (SAFT) [23],
CR
which was widely applicated to describe the dissolution behaviors of gases in macromolecular systems [24-28]. Based on the statistical mechanics, the theory has advantageous over those semi-empirical equations by
US
providing molecular-level interpretation. Although SAFT has yielded encouraged results, its further
AN
improvement is difficult, since its semi-quantitative parameters are often confined to a certain region. Two different empirical equations of state, the Sanchez-Lacombe [29] and Simha-Somocynsky [30] types,
M
were used to calculate the swelling property of polymer melts that was induced by gas permeation. These
ED
equations are also semi-quantitatively and the model parameters rely completely on the available specific pressure–volume–temperature data. It has been recognized that the equations are difficult to predict the swelling
PT
properties given the extreme shortage of these fundamental data. This failure is a result of the inability to
CE
accurately predict the chemical potentials of the gas and polymer melt arising from the asymmetric interactions among CO2-CO2, CO2-polymer and polymer-polymer.
AC
For calculation of gas diffusion in polymer melt,
PT
ED
M
AN
US
CR
IP
T
ACCEPTED MANUSCRIPT
CE
In this work, we try to engage the dissolution and diffusion of CO2 molecules in PLA melt using the dynamic
AC
density functional theory (DDFT) [36]. Given the reliable free-energy function, the theory allows us to quantitatively describe the dynamic density evolution under nonequilibrium condition. In particular, DDFT returns to DFT when the nonlocal density distribution not changes with time. In formulation of the free-energy expression for addressing the CO2-PLA binary system, we integrate the interactions of CO2-CO2, CO2-PLA, and PLA-PLA to reflect the contribution of enthalpy, and the chain connectivity and bond-angle bending energy to account for the contribution of entropy. Unlike previous investigations, where the solubility of polymer in CO2 and the swellings of polymer were overlooked, the present theory simultaneously provides the coexistent curves of CO2 and PLA-CO2 phases and the corresponding swelling properties. During the kinetic diffusion, a
ACCEPTED MANUSCRIPT
introduce the idea of including the nonlinear coupling in
M
AN
US
CR
IP
T
expressing the transport coefficients on the basis of clear thermodynamic evidences.
ED
Fig. 1. Schematic diagram of repetitive units in a coarse-grained PLA chain. The black, red, and blue balls
PT
denote carbon, oxygen, and hydrogen atoms, respectively.
CE
2. Theoretical section
connected
AC
In the theoretical approach, CO2 is modeled as spherical molecule, and PLA chains are constituted with s. The atomic structure of lactic acid unit in a PLA chain is shown in Fig. 1.
The Lennard-Jones parameters for CO2, CO 3.66 Å, CO / kB 235.56 K, are taken directly from literature [37]. 2
2
The force field parameters for the coarse-grained beads are chose from MARTINI force field [38,39] with
p 4.7
Å
p / kB 481.12
In the CO2(1)-PLA(2) binary system, the DDFT equation derived from Langevin equation can be generalized as the following time-dependent free-energy functional form [40,41]
ACCEPTED MANUSCRIPT (r , t ) t
F [ 1 (r, t ), 2 (r, t )] D (r, t ) (r, t )
where (r,t ) is time-evolution density in real space, D is diffusion coefficient, represents a CO2 molecule or PLA monomer, and F[ 1 (r, t ), 2 (r, t )]
CR
IP
T
the free-energy
US
F[ 1 (r, t ), 2 (r, t )]
F [ 1 (r, t ), 2 (r, t )] kBT dr (r )[ln (r , t ) 1] F ex [ 1 (r, t ), 2 (r, t )]
AN
M
where k B is Boltzmann constant, and the first term on the right hand side denotes the free-energy of ideal-gas.
ED
The excess term F ex [ 1 (r, t ), 2 (r, t )] includes the contributions of hard-sphere repulsion, dispersive attraction, and chain connectivity and configuration over the ideal gas state.
PT
The fundamental measure theory [42] is commonly used to describe the contribution of hard-sphere repulsion,
CE
which is expressed as
F hs [ 1 (r, t ), 2 (r, t )] kBT hs [ni (r )]dr ,
(3)
where hs [ni (r )]
AC
fundamental geometric measures
free-energy density of component , which is connected to the averaged hs [ni (r )]
,
F att [ 1 (r, t ), 2 (r, t )] k BT (r ) a1[ (r )] a 2 [ (r )] dr
(4)
ACCEPTED MANUSCRIPT in which (r )
The detailed expr
a[ (r)]
The free-energy contribution arising from the chain connectivity [46] and configuration of PLA is expressed as
(5)
where y ' (r1 , r2 ) is the cavity correlation function [46] between
CR
IP
T
F
1 ( r1 r2 ' ) ' 1 y (r1 , r2 ) 2 (r2 ) 2 (r, t ) dr1 (r, t ) ln dr2 ' 2 2 4 ( ) 2 1 , 1 m 1 ( r1 r2 ' ) 1 dr1 (r, t ) ln dr2 g 0 2 (r2 ) 2 4 ( ' ) 2 2 =1 m
chain
and , and '
US
' =( + ' ) / 2 . The second term on the right hand side belongs to the
AN
contribution of chain configuration given by the bond-angle bending formula, and g 0 is triplet function [47]. After introducing CO2 into the melted PLA for enough time, the CO2 phase and CO2-PLA mixed phase reach
M
equilibrium at given temperature and pressure. Without consideration of the nonlocal density fluctuation, the
ED
free-energy expression is simplified to its local type, and can be easily applied to the phase equilibrium
S
CE
S 0 / [ L (1 xCO2 )]
PT
calculation for the binary system to derive the solubilities of CO2 in PLA and PLA in CO2. Accordingly, t
xCO
2
AC
where 0 is the density of pure polymer, L is the total density of polymer phase after dissolving CO2, and .
On the other hand, the nonlocal density distributions of CO2 and PLA at the CO2-PLA interface can be derived by minimizing the excess free-energy
F ex 1 (r, t ), 2 (r, t ) (r, t ) exp (r, t )
(7)
ACCEPTED MANUSCRIPT with an ordinary Picard iteration procedure. Here is the chemical potential of component at bulk state. The iteration is repeated until the average fractional difference over any grid point between the old and new
(r ) is less than 1.0 104 . The CO2-PLA interfacial tension is the excess free-energy at the interface of polymer phase and CO2 phase
/ A [ 1 (r), 2 (r)] ( 1 , 2 ) / A
T
(8)
dr (r, t)
(9)
US
[ 1 (r), 2 (r)] F[ 1 (r), 2 (r)]
CR
IP
where A is the interfacial area, [ 1 (r), 2 (r)] is the grand potential, and can be calculated by
AN
Under the framework of dynamic DDFT, an individual CO2 molecule is chosen as a test particle located at time t 0 . The initial density distribution is numerically specified by a Gaussian distribution
T (r, 0) ( / )3/ 2 exp( r )
M
2
(10)
ED
where is set as 200. Other CO2 molecules and PLA monomers are distributed around the labeled particle. It is
PT
assumed that the labeled CO2 is in equilibrium, the density distribution of the surrounding molecules is approximated to their radial distribution functions
AC
respectively. D0
CE
During the dynamic calculations, the length and time are nondimensionalized by P and B P 2 / D0 ,
dr 0.01 P D0
dt 105 B
ACCEPTED MANUSCRIPT
R2
Do ,0 lim R 2 /(6t )
T
t
IP
After derivation of the time-dependent local density distribution of the test particle, the mean square
CR
displacement R 2 can be extracted from a simple statistical formula R 2 drT (r, t ) r . As such, the 2
local diffusion coefficient is determined by the Stokes-Einstein relation, D / Do lim R 2 /(6t )
M
AN
US
t
ED
the extended test-particle method [51], we allow
PT
one bead from an arbitrarily selected block to be fixed at the origin location, and in a symmetric external field [41]. The mean square
CE
displacement of a motion chain is calculated from the statistical average of all beads. Similarly, the diffusion
As a result,
AC
coefficient is calculated by D p / Do lim R 2 /(6t ) , and D0 t
kBT / 6 RD p
R
Dp
CR
IP
T
ACCEPTED MANUSCRIPT
US
Fig. 2. Comparison of the radial distribution function of PLA given by the theoretical calculation and
AN
molecular simulation.
M
3. Results and discussion
CE
PT
ED
The macroscopic properties of polymers are always related to their microscopic structure.
1.25 ∙ F ex [ 1 (r, t ), 2 (r, t )]
AC
F ex [ (r, t )]
The theoretical prediction is generally in accordance with the simulation result, indicating that the theory can grasp the essential features of chain conformation.
CR
IP
T
ACCEPTED MANUSCRIPT
US
Fig. 3. Solubilities of CO2 in PLA at various temperatures and pressures. x 1 is the mass fraction (wt%) of
AN
CO2 in PLA melt. The scatter dots are the experimental data.
We then calculate the phase equilibria of CO2-PLA using the free-energy equation without consideration of
M
the nonlocal density distribution. Fig. 3 presents the measured and calculated solubilities of CO2 in PLA melt.
ED
Although PLA is a semi-crystalline polymer, we choose the temperature below the cold crystallization temperature or above the melting temperature of PLA to avoid the generation of crystals, i.e., PLA can be
PT
regarded as an amorphous polymer. It can be known that CO2 solubility in PLA increases with decreasing
CE
temperature and increasing pressure The average deviations between calculated and experimental values [9-11] are 8.46% (308 K), 7.58% (333 K), and 9.69% (463 K), respectively. The general agreement between
AC
theoretical calculations and experimental data has been achieved, which indicates that the theoretical model provides a reasonable free-energy expression for the asymmetric interactions between PLA chains and CO2 molecules. Another advantage of phase equilibrium calculation is to provide the solubilities of PLA in supercritical CO2 phase according to the chemical potential equilibrium in the two phases. As shown in Fig. 4, it is difficult to detect PLA at low pressure due to its high cohesive energy. At middle pressure 50 solubilities of PLA in CO2 can be predicted by theoretical prediction, although the values are
he
M
AN
US
CR
IP
T
ACCEPTED MANUSCRIPT
Fig. 4. Solubilities of PLA in CO2 at various temperatures and pressures. y1 is the mass fraction (wt%) of
PT
ED
PLA in CO2 phase.
According to the definition of polymer expansion in supercritical CO2, the swelling ratios can be calculated
CE
through the equilibrium densities of PLA before and after the swelling process. From Fig. 5 one can see that the
AC
swelling of PLA increases with elevated pressure or depressed temperature. When the temperature is constant, the swelling ratio increases monotonously with pressure, and shows a significant slowdown of slop at high pressure. By comparing the swelling curves with the pressure-composition curves, one can find that the swelling ratios are coherently associated with CO2 solubilities. For instance, the slowing enhanced solubility at high pressure results in small changes in PLA density and swelling ratio. In general, the averaged relative deviation is 4.52%, showing a good agreement between the predicted and experimental swelling ratios.
CR
IP
T
ACCEPTED MANUSCRIPT
US
Fig. 5. Calculated swelling ratios for PLA in the presence of CO2 at elevated pressures. The scatter dots are the
CE
PT
ED
M
AN
experimental data.
Fig. 6. Density distributions of CO2 and PLA at the phase interface region at (a) 463 K, 1 MPa; (b) 463 K, 10
AC
MPa; (c) 333K, 10 MPa.
After the phase equilibria have been achieved, the interfacial density distributions of CO2 and PLA can be determined by the density functional method, and the results are shown in Fig. 6. From the three subFig.s, one can immediately see the enrichments of CO2 in the interfacial region. The decrease of PLA density leads to reduced short-range repulsion of CO2-PLA, which is beneficial for the enhancement of CO2 concentration. At low pressure, the stacking of CO2 at the interface is pretty low (Fig. 6a). When the pressure is high enough, the
ACCEPTED MANUSCRIPT CO2 stack in the interface region becomes notably due to high solubility, since the attractions between PLA and CO2 are much stronger than that among CO2 molecules. As for influence of temperature, it could be seen from the subFig.s 6b and 6c that the density of CO2 at the interface increases at lower temperature, because low
M
AN
US
CR
IP
T
temperature can result in stronger asymmetric attractions.
Fig. 7. Variation of interfacial tension with respect to temperature and pressure. The scatter dots are the
PT
ED
experimental data.
CE
According to the interfacial density distributions under different temperature and pressure effects, the interfacial tensions can be easily calculated and presented in Fig. 7. The approximate agreement of experimental
AC
values [55,56] and theoretical predictions manifests that interfacial structure analysis is reliable. When the temperature is constant, the interfacial tension decreases as pressure increases at a different rate. An increase of pressure results in high CO2 solubility, polymer swelling, and the reduced free volume, therefore leads to low interfacial tension. It is also shown that the lower temperature sample has a higher interfacial tension. Under low pressure, solubility also declines, and the influence of temperature becomes more obvious. At high pressure, however, solubility is large enough and the influence of dissolved CO2 becomes trivial. Therefore, the differential of interfacial tension between different temperatures decreases. Figure 7 also indicates that the predicted and experimental values are in good consistence.
CR
IP
T
ACCEPTED MANUSCRIPT
US
Fig. 8. Nonlocal density relaxation process of the test CO2 molecule.
t 0.1 B , the density distribution is described as r, t 0 / g(r ) . T
M
initial stage
AN
Fig. 8 shows the density evolution of the test CO2 molecule at elevated time during diffusion process. In the
PT
ED
t 5.0 B
t 50.0 B
and approaches to its bulk density
AC
CE
As time goes by, the density peaks gradually decrease
The corresponding free-energy evolution of the test CO2 molecule is given in Fig. 9. One sees a monotone decline over the whole diffusion process. In it obvious that the declining free-energy provides the driving force for diffusion. Nevertheless, the slopes of these curves are not constant. At the intermediate region, those second
ACCEPTED MANUSCRIPT derivatives on the three curves change from minus to positive, and then return to minus. This phenomenon means that the diffusion rate of CO2 slows down in this region. Due to the frequent collisions with its neighboring CO2 molecules and PLA chains, the test molecule is confined to a small region but this restriction is not strong enough. Consequently, it escapes from the confinement. It can also be seen that, the temperature
T
has a perceivable influence on free-energy. As temperature decreases, the concentration of CO2 increases, and
PT
ED
M
AN
US
CR
IP
the time cost to cross the confinement increases, resulting in the dropdown of driving force at low temperature.
AC
CE
Fig. 9.
According to the time-dependent density evolutions, the diffusion coefficients can be easily derived and the results are shown in Fig. 10. At high pressure, more CO2 molecules are dissolved into the PLA melt, thus the collision frequency increases, leading to slower diffusivity. Under the pressure of 20 MPa, the diffusivity at 453 K is almost two orders of magnitude higher than the value at 323 K. It is unsurprised sine the free-energy driving force has been depressed at low temperature. One can also see that the predicted diffusion coefficients
ACCEPTED MANUSCRIPT are overestimated when compared with the corresponding experimental data [57,58]. It could be that there exist some entanglements and crystal regions in PLA during experimental test at low temperature. This effect has been ignored in theoretical study. Because CO2 molecules can hardly permeate through the entanglements and crystals, the actual diffusion has been retarded to some extent. In addition, the calculated diffusivity at high
AC
CE
PT
ED
M
AN
US
CR
IP
T
temperature deviates obviously from
. The scatted dots denote experimental data
Fig. 11 depicts the calculated viscosities of PLA at 383K and 453 K. It can be seen that the viscosity of PLA is reduced rapidly with increasing pressure, especially before the supercritical state of CO2. When PLA melt absorbs CO2 molecules, the chains rearrange themselves towards a new equilibrium conformation. For small CO2 concentration equivalent to low pressure, the plasticizing effect is reflected in the rapid rising swelling
ACCEPTED MANUSCRIPT ratio as shown in Fig. 5. Thus effect leads to enhanced ease of hole formation and chain motion. After that, the variation tendency of the curve becomes planus. Although the plasticizing is prominent at high pressure, the slope gradient of increasing swelling ratio decreases. Fig. 11 also shows that the predicted viscosities deviate from the experimental values with about one order of magnitude, indicating that description of chain dynamic
T
given by the theoretical model is only qualitatively reasonable due to some crude approximations, such as chain
. The scatted dots
denote experimental data.
AC
CE
PT
ED
M
AN
US
CR
IP
block definition and regardlessness of hydrodynamic interactions given by dissolved CO2 molecules.
4. Conclusion We have presented a DDFT approach to quantitatively predict the CO2 solubilities and diffusivities, PLA swelling ratios and viscosities, and CO2-PLA interfacial tensions. Before its application, the theoretical approach has been fully tested by the computational radial distribution of PLA, some experimental solubilities, interfacial tensions, and diffusion coefficients. It has been shown that the predictions of solubilities, swelling
ACCEPTED MANUSCRIPT ratios, interfacial tensions, diffusivities, and viscosities are generally reasonable. Therefore, we conclude that such theoretical approach has some advantages over previous thermodynamic or dynamic equations: (i) the solubilities of CO2 in PLA and PLA in CO2 as well as the swelling ratios of PLA can be determined simultaneously by the phase equilibrium calculations for the binary system; (ii) the local and nonlocal density distributions can be derived by the same theory under equilibrium condition; (iii) the time-dependent diffusion
T
process is performed strictly on the density and free-energy evolution, such that the kinetic process is coherently
IP
associated with the thermodynamic state, and the driven force can be clearly identified from the free-energy
CR
variation. Therefore, this work provides a comprehensive model for in-depth understanding the role of various factors in influencing PLA foaming that could find important applications for the control of cell density and
AN
US
morphology.
Acknowledgements
M
This work has been supported by the Natural Science Foundation of Beijing (2162012), the National Science
AC
CE
PT
ED
Foundation of China (51673004) , and the 2017 Special Commercial Projects (19005757053).
ACCEPTED MANUSCRIPT References [1] O. Martin, L. Averous, Poly (lactic acid): plasticization and properties of biodegradable multiphase systems, Polymer. 42 (2001) 6209-6219. [2] R. Datta, S.P.Tsai, P. Bonsignore, S.H. Moon, J.R. Frank, Technological and economic potential of poly
T
(lactic acid) and lactic acid derivatives, FEMS Microbiol. Rev. 16 (1995) 221-231.
IP
[3] M. Nofar, C.B. Park, Poly (lactic acid) foaming, Prog. Polym. Sci. 39 (2014) 1721-1741.
CR
[4] G. Harikrishnan, D.V. Khakhar, Modeling the dynamics of reactive foaming and film thinning in polyurethane foams, AIChE J. 56 (2010) 522-530.
US
[5] L.T. Lim, R. Auras, M. Rubino, Processing technologies for poly (lactic acid), Prog. Polym. Sci. 33 (2008)
AN
820-852.
[6] A.R. Kakroodi, Y. Kazemi, W.D. Ding, A. Ameli, C.B. Park, Poly (lactic acid)-based in situ microfibrillar
M
composites with enhanced crystallization kinetics, mechanical properties, rheological behavior, and foaming
ED
ability, Biomacromolecules. 16 (2015) 3925-3935.
[7] D. Li, T. Liu, L. Zhao, W. Yuan, Controlling sandwich‐structure of PET microcellular foams using
PT
coupling of CO2 diffusion and induced crystallization, AIChE J. 58 (2012) 2512-2523.
CE
[8] X. Shi, G. Zhang, Y. Liu, Z. Ma, Z. Jing, X. Fan, Microcellular foaming of polylactide and poly (butylene adipate‐co‐terphathalate) blends and their CaCO3 reinforced nanocomposites using supercritical carbon
AC
dioxide, Polym. Advan. Technol. 27 (2016) 550-560. [9] R. Pini, G. Storti, M. Mazzotti, H. Tai, K.M. Shakesheff, S.M. Howdle, Sorption and swelling of poly (dl‐ lactic acid) and poly (lactic‐co‐glycolic acid) in supercritical CO2: An experimental and modeling study, J. Polym. Sci. Pol. Phys. 46 (2008) 483-496. [10] D. Liu, D. L. Tomasko, Carbon dioxide sorption and dilation of poly (lactide-co-glycolide), J. Supercrit. Fluid. 39 (2007) 416-425. [11] S. H. Mahmood, M. Keshtkar, C. B. Park, Determination of carbon dioxide solubility in polylactide acid
ACCEPTED MANUSCRIPT with accurate PVT properties, J. Chem. Thermodyn. 70 (2014) 13-23. [12] T.H. Mareci, S. Doenstrup, A. Rigamonti, NMR imaging and relaxation study of polymer swelling and chain dynamics, J. Mol. Liq. 38 (1988) 185-206. [13] H. Marubayashi, S. Akaishi, S. Akasaka, S. Asai, M. Sumita, Crystalline structure and morphology of poly (L-lactide) formed under high-pressure CO2, Macromolecules. 41 (2008) 9192-9203.
T
[14] E. Aionicesei, M. Škerget, Ž. Knez, Measurement of CO2 solubility and diffusivity in poly (l-lactide) and
IP
poly (d, l-lactide-co-glycolide) by magnetic suspension balance, J. Supercrit. Fluid. 47 (2008) 296-301.
CR
[15] L. Urbanczyk, C. Calberg, C. Detrembleur, C. Jérôme, M. Alexandre, Batch foaming of SAN/clay nanocomposites with scCO2: a very tunable way of controlling the cellular morphology, Polymer. 51 (2010)
US
3520-3531.
AN
[16] S. Areerat, E. Funami, Y. Hayata, D. Nakagawa, M. Ohshima, Measurement and prediction of diffusion coefficients of supercritical CO2 in molten polymers, Polym. Eng. Sci. 44 (2004) 1915-1924.
M
[17] T. Komatsuka, A. Kusakabe, K. Nagai, Characterization and gas transport properties of poly (lactic acid)
ED
blend membranes, Desalination. 234 (2008) 212-220.
[18] A. Bertucco, C. Mio, Prediction of vapor-liquid equilibrium for polymer solutions by a group-contribution
PT
Redlich-Kwong-Soave equation of state, Fluid Phase Equilibr. 117 (1996) 18-25.
CE
[19] J.O. Valderrama, The state of the cubic equations of state, Ind. Eng. Chem. Res. 42 (2003) 1603-1618. [20] D. Liu, H. Li, M.S. Noon, D.L. Tomasko, CO2-induced PMMA swelling and multiple thermodynamic
AC
property analysis using Sanchez- Lacombe EOS, Macromolecules. 38 (2005) 4416-4424. [21] K.S. Schweizer, J.G. Curro, Integral equation theories of the structure, thermodynamics, and phase transitions of polymer fluids, Adv. Chem. Phys. 98 (1997) 1-142. [22] Y. Tang, B.C.Y. Lu, A study of associating Lennard–Jones chains by a new reference radial distribution function, Fluid Phase Equilibr. 171 (2000) 27-44. [23] W.G. Chapman, K.E. Gubbins, G. Jackson, G. Radosz, New reference equation of state for associating liquids, Ind. Eng. Chem. Res, 29 (1990) 1709-1721. [24] N. I. Diamantonis, G. C. Boulougouris, E. Mansoor, D.M. Tsangaris, I.G. Economou, Evaluation of cubic,
ACCEPTED MANUSCRIPT SAFT, and PC-SAFT equations of state for the vapor–liquid equilibrium modeling of CO2 mixtures with other gases, Ind. Eng. Chem. Res. 52 (2013) 3933-3942. [25] M.C. Arndt, G. Sadowski, Modeling poly (N-isopropylacrylamide) hydrogels in water/alcohol mixtures with PC-SAFT, Macromolecules. 45 (2012) 6686-6696. [26] S.K. Maity, Correlation of solubility of single gases/hydrocarbons in polyethylene using PC‐SAFT, Asia-
T
Pac J. Chem. Eng. 7 (2012) 406-417.
IP
[27] I. Stoychev, F. Peters, M. Kleiner, S. Clerc, F. Ganachaud, M. Chirat, B. Fournel, G. Sadowski, P. Lacroix-
CR
Desmazes, Phase behavior of poly (dimethylsiloxane)-poly (ethylene oxide) amphiphilic block and graft copolymers in compressed carbon dioxide, J. Supercrit. Fluid. 62 (2012) 211-218.
US
[28] Y. Peng, K. D. Goff, M.C. dos Ramos, C. McCabe, Predicting the phase behavior of polymer systems with
AN
the GC-SAFT-VR approach, Ind. Eng. Chem. Res. 49 (2009) 1378-1394. [29] Q. Xu, J. Mi, C. Zhong, Integral Equation Theory for Gas Sorption and Swelling of Glassy Atactic
M
Polystyrene, Ind. Eng. Chem. Res. 49 (2010) 4914-4922.
ED
[30] I.C. Sanchez, R.H. Lacombe, An elementary molecular theory of classical fluids. Pure fluids, J. Phys. Chem. 80 (1976) 2352-2362.
PT
[31] B.D. Freeman, Basis of permeability/selectivity tradeoff relations in polymeric gas separation membranes,
CE
Macromolecules. 32 (1999) 375-380.
[32] A.Y. Alentiev, Y.P. Yampolskii, Free volume model and tradeoff relations of gas permeability and
AC
selectivity in glassy polymers, J. Membrane Sci. 165 (2000) 201-216. [33] A. Alentiev, Y. Yampolskii, Correlation of gas permeability and diffusivity with selectivity: orientations of the clouds of the data points and the effects of temperature, Ind. Eng. Chem. Res. 52 (2013) 8864-8874. [34] L.M. Robeson, Z.P. Smith, B. D. Freeman, D.R. Paul, Contributions of diffusion and solubility selectivity to the upper bound analysis for glassy gas separation membranes, J. Membrane Sci. 453 (2014) 71-83. [35] F.H. Stillinger, P.G. Debenedetti, Glass transition thermodynamics and kinetics, Annu. Rev. Conden. Ma P. 4 (2013) 263-285. [36] M. Rex, H. Löwen, Dynamical density functional theory with hydrodynamic interactions and colloids in
ACCEPTED MANUSCRIPT unstable traps, Phys. Rev. Lett. 01 (2008) 148302. [37] C. Avendano, T. Lafitte, A. Galindo, C.S. Adjiman, G. Jackson, E.A. Muller, SAFT-γ force field for the simulation of molecular fluids. 1. A single-site coarse grained model of carbon dioxide, J. Phys. Chem. B. 115 (2011) 11154-11169. [38] S.J. Marrink, H.J. Risselada, S. Yefimov, P. Tieleman, A.H. Vries, The MARTINI force field: coarse
T
grained model for biomolecular simulations, J. Phys. Chem. B. 111 (2007) 7812-7824.
IP
[39] X. Li, T. Xiao, N. Xiao, The application of nonlocal theory method in the coarse-grained molecular
CR
dynamics simulations of long-chain polylactic acid, Acta Mech. Solida Sin. 30 (2017) 630-637. [40] U.M.B. Marconi, P. Tarazona, Dynamic density functional theory of fluids, J. Chem. Phys. 110 (1999)
US
8032-8044.
AN
[41] Y. Ye, Z. Du, M. Tian, L. Zhang, J. Mi, Diffusive dynamics of polymer chains in an array of nanoposts, Phys. Chem. Chem. Phys. 19 (2017) 380-387.
M
[42] Y. Rosenfeld, Free-energy model for the inhomogeneous hard-sphere fluid mixture and density-functional
ED
theory of freezing, Phys. Rev. Lett. 63 (1989) 980.
Phys. 117 (2002) 10156-10164.
PT
[43] Y.X. Yu, J. Wu, Structures of hard-sphere fluids from a modified fundamental-measure theory, J. Chem.
116 (2012) 3042-3049.
CE
[44] D. Zhou, J. Mi, C. Zhong, Theoretical study of dissolved gas at a hydrophobic interface, J. Phys. Chem. C.
AC
[45] M. Zeng, J. Mi, C. Zhong, Wetting behavior of spherical nanoparticles at a vapor-liquid interface: a density functional theory study, Phys. Chem. Chem. Phys. 13 (2011) 3932-3941. [46] S. Tripathi, W.G. Chapman, Microstructure of inhomogeneous polyatomic mixtures from a density functional formalism for atomic mixtures, J. Chem. Phys. 122 (2005) 094506. [47] E.A. Müller, K.E. Gubbins, Simulation of hard triatomic and tetratomic molecules: A test of associating fluid theories, Mol. Phys. 80 (1993) 957-976. [48] J.G. Curro, A.L. Frischknecht, The structure of poly(ethylene oxide) liquids : comparison of integral equation theory with molecular dynamics simulations and neutron scaling, Polymer. 46 (2005) 6500-6506.
ACCEPTED MANUSCRIPT [49] D. Stopper, K. Marolt, R. Roth, H. Hansen-Goos, Modeling diffusion in colloidal suspensions by dynamical density functional theory using fundamental measure theory of hard spheres, Phys. Rev. E. 92 (2015) 022151. [50] L. Ambrosone, C.D. Volpe, G. Guarino, R. Sartorlo, V. Vitagliano, Free diffusion data in some polymersolvent systems at 20°C, J. Mol. Liq. 50 (1991) 187-196.
T
[51] Y.X. Yu, J. Wu, Extended test-particle method for predicting the inter-and intramolecular correlation
IP
functions of polymeric fluids, J. Chem. Phys. 118 (2003) 3835-3842.
CR
[52] M. Mihai, M.A. Huneault, Favis, B. D. Rheology and extrusion foaming of chain‐branched poly (lactic acid), Polym. Eng. Sci. 50 (2010) 629-642.
US
[53] J. Mi, Y. Tang, C. Zhong, Theoretical study of Sutherland fluids with long-range, short-range, and highly
AN
short-range potential parameters, J. Chem. Phys. 128 (2008) 054503.
[54] X. Li, N.S. Murthy, R.A. Latour, Structure of Hydrated Poly (d, l-lactic acid) Studied with X-ray
M
Diffraction and Molecular Simulation Methods, Macromolecules. 45 (2012) 4896-4906.
ED
[55] K. Sarikhani, K. Jeddi, R. B. Thompson, C.B. Park, P. Chen, Effect of pressure and temperature on interfacial tension of poly lactic acid melt in supercritical carbon dioxide, Thermochim. Acta. 609 (2015) 1-6.
PT
[56] S.H. Mahmood, A. Ameli, N. Hossieny, C.B. Park, The interfacial tension of molten polylactide in
CE
supercritical carbon dioxide, J. Chem. Thermodyn. 75 (2014) 69-76. [57] G. Li, H. Li, L.S. Turng, C. Zhang, Measurement of gas solubility and diffusivity in polylactide, Fluid
AC
Phase Equilibr. 246 (2006) 158-166.
[58] E. Aionicesei, M. Škerget, Ž. Knez, Measurement of CO2 solubility and diffusivity in poly (l-lactide) and poly (d, l-lactide-co-glycolide) by magnetic suspension balance, J. Supercrit. Fluid. 47 (2008) 296-301. [59] C. A. Kelly, S. M. Howdle, K. M. Shakesheff, M. J. Jenkins, G. A. Leeke, Viscosity studies of poly (DL‐ lactic acid) in supercritical CO2. J. Polym. Sci. Pol. Phys. 50 (2012) 1383-1393.
AC
CE
PT
ED
M
AN
US
CR
IP
T
ACCEPTED MANUSCRIPT