Sie sind auf Seite 1von 11

Fluid Phase Equilibria 286 (2009) 17–27

Contents lists available at ScienceDirect

Fluid Phase Equilibria


journal homepage: www.elsevier.com/locate/fluid

Modeling of thermodynamic behavior of PVT properties and cloud point


temperatures of polymer blends and polymer blend + carbon dioxide
systems using non-cubic equations of state
Pedro F. Arce ∗ , Martín Aznar
School of Chemical Engineering, State University of Campinas (UNICAMP), P.O. Box 6066, 13083-970 Campinas, SP, Brazil

a r t i c l e i n f o a b s t r a c t

Article history: In recent years, many factors influencing phase behavior of polymer blends have been studied because
Received 4 June 2009 of their widely technological importance, as a simple method of formulating new materials with tai-
Received in revised form 6 August 2009 lored properties which make them suitable for a variety of applications. This work has three main goals
Accepted 7 August 2009
which were reached by using the Perturbed Chain Statistical Associating Fluid Theory (PC-SAFT) and the
Available online 15 August 2009
Sanchez–Lacombe (SL) non-cubic equations of state (EoS), which in previous works have shown their
ability to handle long chain and associating interactions. First, both equations of state were tested with
Keywords:
the correlation of the specific volumes of pure blends (PBD/PS, PPO/PS, PVME/PS, PEO/PES) and the pre-
Blends
Polymer
diction of the specific volumes for blends; second, the modeling of blend miscibilities in the liquid–liquid
Liquid–liquid equilibria equilibria (LLE) of PBD/PS, PPG/PEGE, PVME/PS, PEO/PES, and PnPMA/PS blends; third, the modeling of
Non-cubic the phase behavior of PS/PVME blends at various compositions in the presence of CO2 . PC-SAFT and SL
Equation of state pure-component parameters were regressed by fitting pure-component data of real substances (liquid
pressure–volume–temperature, PVT, data for polymers and vapor pressure and saturated liquid molar
volume for CO2 ) and the fluid phase behavior of blend systems were simulated fitting one binary interac-
tion parameter (kij ) by regression of experimental data using the modified likelihood maximum method.
Results were compared with experimental data obtained from literature and an excellent agreement
was obtained with both EoS, which were also capable of predicting the fluid phase behavior correspond-
ing to the critical solution temperatures (LCST: lower critical solution temperature, UCST: upper critical
solution temperature) of blends.
© 2009 Elsevier B.V. All rights reserved.

1. Introduction blends give the chances to understand and predict the relation-
ship between structures and physical properties [1]. On the other
During the last decades, several theoretical and experimental hand, the effect of solvent on polymer blends and block copoly-
approaches were used to investigate the structure and thermo- mers has important consequences for material processing, so much
dynamic properties of polymer and their blends, of there that in so numerous theoretical and experimental studies have demon-
recent years the use of polymer blends has been increased due strated that the addition of solvents can either suppress or induce
to several reasons which have stimulated the interest to inves- polymer phase separation in blends, depending on the relative
tigate physical blends of different polymers. For instance, for quality of the solvent for each of the constituents [2]. In recent
engineering applications, polymer blends are a cost-effective way years, the study of the miscibility and phase behavior of polymer
to produce improved polymer properties; many times, polymer blends has attracted a large amount of theoretical and experimen-
tal interest [3]. For instance, for the polymer pairs of PEO and PVAc
[poly(vinyl acetate)], there are several publications in the scientific
literature about theoretical prediction and experimental results.
Abbreviations: PBD, polybutadiene; PEGE, polyethylene glycol mono-methyl
ether; PPG, polyethylene glycol; PPO, 2,6-dimethyl phenylene oxide; PS, Kalfoglou et al. [4,5], Munoz et al. [6], Martuscelli et al. [7–9] and
polystyrene; PVME, poly(vinyl methyl ether); PEO, poly(ethylene oxide); PES, Chen et al. [10] had discussed the miscibility of this blend system
poly(ether sulfone); PnPMA, poly(n-methyl methacrylate); CO2 , carbon dioxide; from many aspects including morphology, dynamic mechanical
SBS, styrene-butadiene styrene; PoClS, poly(o-chlorostyrene); PBrS, brominated
properties, rheological properties, thermal behavior, crystallization
polystyrene; SBR, styrene-butadiene rubber; SAN, styrene acrylonitrile; NBR,
butadiene-acrylonitrile rubbers. and viscosity in dilute solution. In previous communication [11,12],
∗ Corresponding author. Tel.: +55 19 3521 3962; fax: +55 19 3521 3965. interaction parameters of PEO–PVAc blends with different compo-
E-mail address: pfarcec@yahoo.com.br (P.F. Arce). sitions and temperatures were calculated based on Flory solution

0378-3812/$ – see front matter © 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.fluid.2009.08.002
18 P.F. Arce, M. Aznar / Fluid Phase Equilibria 286 (2009) 17–27

(PPO/PS, PVME/PS, PBD/PS, and PEO/PES), the cloud point temper-


ature data of the liquid–liquid equilibria in polymer blends (PBD/PS,
PPG/PEGE, PVME/PS, PEO/PES, and PnPMA/PS) and the fluid phase
equilibria of blend + solvent systems (PVME/PS blend + CO2 ). The
results obtained with both models are compared with experimental
data [2,35–44] and reports in terms of temperature deviations. All
experimental data used here correspond to monodispersed poly-
mers, which polydispersity indices are less than 1.1, which is not
expected to have a significant influence on blend miscibility.

2. Thermodynamic models

2.1. Perturbed-chain statistical fluid theory EoS (PC-SAFT)

The PC-SAFT EoS [31,32] calculates the dimensionless


Helmholtz free energy, ã, from the summation of three terms: a
reference hard-sphere-chain (ãhc ), a perturbation contribution,
ãpert , and associating contribution, ãassoc (in this paper the associ-
ation term is neglected because all the species considered in this
Fig. 1. Generic phase diagram for liquid mixtures with upper and lower critical
work are non-associating), where ã = A/NkT and A, N, k and T are
solution temperature, UCST and LCST, respectively.
the Helmholtz free energy, total number of molecules, Boltzmann
constant and absolute temperature, respectively. The hard-sphere-
theory modified by Hamada et al. [13]. Results showed that interac- chain contribution was provided by Chapman et al. [45,46] and
tion parameters of PEO–PVAc blends were negative and increased was based on the first-order thermodynamic perturbation theory,
with enhancing the content of PEO and the temperature. Further,
cloud points and heats of mixing of this polymer pair were detected 
nc

by using DSC on the assumption that values of heats of mixing were ãhc = m̄ãhs − xi (mi − 1) ln giihs (ii ) + ãideal (1)
considered to be heats of demixing with opposite sign [14,15]. i
A typical generic phase diagram of the UCST and LCST poly-
mer blends is shown in Fig. 1. There are three regions of different where ghs is the radial pair distribution function for segments in the
degree of miscibility: (1) single-phase miscible region between hard-sphere system and m̄ is the average segment number, which
the two binodals, (2) four fragmented metastable regions between are calculated as:
binodals and spinodals, and (3) two-phase separated “spinodal”    2
1 di dj 32 di dj 222
regions of immiscibility, bordered by the spinodals. The diagram gijhs = + + (2)
(1 − 3 ) di + dj (1 − 3 )
2 di + dj (1 − 3 )
3
also shows two types of phase transitions which have been reported
on the basis of the temperature dependence of the segmental
interaction parameters (critical solution temperatures), the lower,

nC

m̄ = xi mi (3)
LCST, the temperature above which two polymers phase separate
i
(at higher temperature), and the upper, UCST, the temperature
above which two polymers mix (at lower temperature) [16,17]. The where x is the mole fraction and m is the segment number per
phase diagram with two critical points is a rule for mixtures with chain. The hard-sphere contribution (ãhs ) and the radial distribu-
low-molecular-weight component(s), whereas the polymer blends tion function (gijhs ) depend on the auxiliary variable  k (k = 0.3) and 
usually show either LCST (most) or UCST [18]. Few blends having depends on temperature-dependent segment diameter, d, and the
UCST are shown and exemplified by: PS with SBS, PoClS, PBrS, or total number density of molecules, ; thus,
poly(methyl-phenyl siloxane), BR with SBR, SAN with NBR, etc. [19].
 
nc
For several years, different thermodynamic models have been
developed and reported in the literature to describe liquid–liquid k =  xi mi dik , k = (0, 1, 2, 3) (4)
6
non-ideality. The primary theoretical background in early theo- i=1
ries was based on incompressible-lattice models, generalized van
while the temperature-dependent segment diameter (di ) depends
der Waals partition-function theories, compressible-lattice mod-
on directly the segment diameter () and the depth pair potential
els, and their improvements. These include Flory–Huggins theory
(ε) in according to
[20,21], Flory and Patterson equation [22,23], perturbed hard-
  εii

chain theory [24,25], lattice fluid theory [26,27], lattice-cluster
di = ii 1 − 0.12 exp −3 (5)
theory [28], and so on. However, these theories and equations of kT
state ignore the continuous nature of real polymer configurations.
Recently, an active research area to develop equations of state The perturbation contribution (ãpert ) [47] the sum of the first
for chain-like fluids based on off-lattice model has emerged [29]. (ã1 ) and second-order (ã2 ) perturbation terms, which are defined
Several equations of state for homopolymer systems have been as
summarized briefly by Hino et al. [30]. Many of them have their ã1 = −2I1 (, m̄)mi,j ε1ij i,j
2 (6)
origin based on the perturbation theory; one of these is the PC-
SAFT [31,32] EoS, whose performance have been tested in modeling  −1
of binary and ternary systems involving polymers and copolymers, ∂Z hc
ã2 = −m̄ 1 + Z hc +  3
I2 (, m̄)mi,j ε2i,j i,j (7)
solvent and co-solvent at low and high temperatures and pressures ∂
[33,34]. n n l
The purpose of this paper is to use the PC-SAFT and SL EoS where mi,j εli,j i,j
k =
i
c
j
c
xi xj mi mj (εij /kT ) ijk and are the PC-
to correlate and predict the PVT fluid behavior of several blends SAFT mixing rules (van der Waals one-type). Conventional
P.F. Arce, M. Aznar / Fluid Phase Equilibria 286 (2009) 17–27 19

combining rules are used to determine the cross parameters: where kij is the binary interaction parameter. Volume fraction of
component i, ϕi , is defined as
1

ij = ( + jj ) and εij = εii εjj (1 − ij ) (8)


2 ii mi /i vi
ϕi = (18)
where ij is the adjustable interaction parameter which evaluates j
(mj /j vj )
the interaction between segments. In the previous equations, m, 
and ε are the pure-component parameters for the PC-SAFT model. where mi is the mass fraction of component i. Density can be
The compressibility term is calculated as expressed as
  1 V
hc∂Z hc 8 − 22 20 − 272 +123 − 24 vR = = ∗ (19)
Z + = m̄ 4
+(1−m̄) 2
R V
∂ (1 − ) [(1 − )(2 − )]
(9) where V ∗ = N(r v∗ ), V* is close-packed mixture volume and N is the
total number of molecules. Mixing rule for v∗mix is
where the packing fraction () also depends on temperature- 
dependent segment diameter (d). Abbreviations (I1 (, m̄), I2 (, m̄)) v∗mix = ϕi ϕj v∗ij (20)
defined in Eqs. (6) and (7) also depend on the packing fraction () i j
and the mean segment number (m̄) [31,32].
Final form for the fugacity coefficient is: where the cross term, v∗ij , is the arithmetic mean of the two pure-
component characteristic volumes
ˆ res

ˆi =
ln
i
− ln Z (10)
kT 1 ∗
v∗ij = (v + v∗jj ) (21)
2 ii
    
ˆ res
∂ãres 
nc
∂ãres Mixing rule for the number of sites that a mixture occupies and
i
= ãres +(Z − 1) + − xj the mixture characteristic parameter are
kT ∂xi ∂xj
T,V,xj =
/ i j=1 T,V,xk =
/ j
1 ϕ
i
(11) = (22)
rmix ri
i
ˆ i is the chemical potential of component i in the mixture.
where

RTmix

Pmix = ∗ (23)
2.2. Sanchez–Lacombe (SL) EoS [26,27,48] Vmix

The SL lattice-gas model, which was developed to describe the The chemical potential is obtained from the thermodynamic
thermodynamic properties of small molecules and polymers, is relationship
composed of a van der Waals-type attractive term with a lattice-gas  
∂G
repulsive term. In reduced form, this EoS becomes as i = (24)
∂ni
  1
  T,P,ni =
/ j
R2 + PR + TR ln(1 − R ) + 1 − R = 0 (12)
r where the Gibbs free energy of the mixture (G) is derived using the
where PR , TR and R are the reduced pressure, temperature and SL EoS. Note that ni is the number of moles of component i. With
density, respectively and are defined as the SL model, the chemical potential of component i in the mixture
is
T P  ⎡ ⎛ ⎞
TR = , PR = , R = (13)   
T∗ P∗ ∗ ri r 2 
i = RT ln ϕi + 1 − − i ⎣ ⎝ ϕj v∗ij (ε∗ij − ε∗mix )⎠
where T, P and  are the absolute temperature, system pressure and rmix R v∗mix
j=1
solution density, respectively, and T*, P* and * are called the three
⎤ ⎧
characteristic parameters and are defined as ⎨  
RT R
ε∗ ε∗ MW + ε∗mix ⎦ + ri (1 − R )ln(1 − R ) + ln R

T = , P =∗
,  = ∗
(14) ⎩ R ri
R v∗ r v∗
⎡ ⎤⎫
where ε*, v∗ , R, MW and r are the interaction energy, close-packed ⎬
P 
molar volume, gas constant, molecular weight and the number of + ⎣2 ϕj v∗ij − v∗mix ⎦ (25)
lattice sites occupied by a molecule., respectively. For mixtures, it is R ⎭
j=1
necessary to define a characteristic mixture temperature, pressure
and close-packed molar volume. Characteristic mixture tempera-
ture is


ε∗mix 3. Phase equilibria
Tmix = (15)
R
The coexistence curve in liquid–liquid equilibria (LLE) is found
where the mixing rule for ε∗mix and the cross term, ε∗ij , are from the following conditions:

1  Li I = iLII (26)
ε∗mix = ϕi ϕj ε∗ij v∗ij (16)
v∗mix
i j where is the change in chemical potential upon isothermally

transferring component i from the pure state to the mixture, LI and
ε∗ij = εii εjj (1 − kij ) (17)
LII denote the two liquid phases at equilibrium.
20 P.F. Arce, M. Aznar / Fluid Phase Equilibria 286 (2009) 17–27

Table 1
Polymer properties and PHSC pure-component parameters.

Polymer Temperature (K) Pressure (MPa) Polydispersity (Mw/Mn) PHSC pure-component parameters

A* (×108 m2 /mol) V* (×103 m3 /mol) E* (MPa m3 /mol)

PEGE 300.00–400.00 0–100 1.03 7.8526 23.4120 5.4523


a
PPG 300.00–400.00 0–100 6.1452 25.7415 5.8542
PES 320.00–450.00 0–200 1.05 6.2563 24.5412 6.1524
PnPMA 300.00–550.00 0–100 1.02 7.4152 26.5842 6.8569
a
Polydispersitiy index is not available. In this case, it was used Mw/Mn = 1.00.

 
4. Results and discussions  
NP vL,exp − vL,calc 
1  i i
vL = (30)
In this work, the PC-SAFT and SL EoS were used to correlate NP
i
vL,exp
i
and predict the thermodynamic behavior of pure blends (PPO/PS,
PVME/PS, PBD/PS, and PEO/PES), the correlation of LLE of polymer where NP represents of experimental data points.
blends (PBD/PS, PPG/PEGE, PVME/PS, PEO/PES, and PnPMA/PS), as For PEGE, PPGE, PES, and PnPMA polymers, liquid PVT exper-
well as the fluid phase behavior of PVME/PS blends at various com- imental data are not available. We applied a method to estimate
positions in the presence of CO2 were modeled by using the PC-SAFT liquid PVT data, which was applied successfully to generate
and SL EoS. PVT data of poly(lactide-co-glycolide) copolymers at several
compositions [52]. This method was proposed by Song et al.
4.1. PC-SAFT and SL pure-components parameters of pure [53,54] and extended by Elvassore et al. [55] which applied a
polymers and CO2 group-contribution method coupled with the perturbed-hard-
sphere-chain (PHSC) EoS to predict the PHSC pure-component
PC-SAFT and SL pure-component parameters for each polymer parameters of low-molecular-weight compounds by estimat-
were obtained by fitting liquid pure-component PVT data [49] of ing PVT data. They reported values of the group-contribution
real substances over a pressure range suitable for engineering cal- parameters with which it is possible to extrapolate the PHSC pure-
culations. The minimization function (FO) used in the modified component parameters to high-molecular-weight compounds
maximum likelihood [50,51] was (polymers) obtaining secondarily PVT data of that compound. The
only input required for the model is the structure in terms of func-
 
1  Pl − Pl 
NP exp calc tional groups and the molecular weight. In Table 1 some physical
FO = exp (27) properties of PEGE, PPGE, PES, and PnPMA polymers are shown,
NP Pl
l as well as the three pure-component parameters of PHSC EoS
(A*: characteristic surface area parameter, E*: characteristic cohe-
The fitted characteristic parameters of each thermodynamic sive energy parameter, and V*: characteristics volume parameter)
model were then used to calculate the average deviation from obtained for each polymer which are calculated in a temperature
experimental specific volume, range shown also in this table. PVT data obtained with the PHSC EoS
  are used with the PC-SAFT, and SL models to predict their respective
1  vl − vl 
NP exp calc
pure-component parameters over the same temperature ranges
(28) displayed in Table 1 and all of which are suitable for engineer-
NP vexp
l
l ing calculations. In this table, the polydispersity is also shown as
a relation between the mass average molecular weight (Mw) and
where NP is the number of experimental data points. The liquid
number average molecular weight (Mn). Table 2 lists the PC-SAFT
pressure and volume deviations of these polymers were calculated
and SL segment parameters for the polymers used in this work.
using the following criteria:
For CO2 , PC-SAFT and SL pure-component parameters were
 
 
NP P L,exp − P L,calc 
obtained by fitting vapor pressure and molar volume data for sat-
1  i i urated pure liquid [56]. Table 2 also shows the PC-SAFT and SL
P L = L,exp
(29) pure-component parameters for CO2 . In this table it is also shown
NP Pi
i the deviations in saturated vapor pressure ( PL,sat ) and liquid vol-

Table 2
PC-SAFT and SL pure-component parameters for polymers and CO2 .

Component PC-SAFT EoS SL EoS


−3 −1 10 L L
m/MW (×10 kg/mol)  (×10 m) ε/k (K) P v T* (K) P* (MPa) * (kg/m3 ) PL vL

PBD 0.060552 2.9845 240.2132 0.06 0.30 562.13 412.41 972.41 1.54 1.80
PEGE 0.043561 3.1526 290.5413 0.08 0.39 630.21 502.32 1020.12 2.05 1.75
PEO 0.080017 2.5125 230.1245 0.22 0.47 663.40 487.44 1168.23 1.86 2.13
PES 0.023433 3.6148 356.1422 0.10 0.31 881.14 645.10 1389.18 0.26 0.22
PnPMA 0.024136 3.8415 334.4023 0.09 0.39 681.26 483.35 1120.46 0.89 1.23
PPG 0.040123 3.3342 310.5026 0.07 0.46 615.12 520.41 1140.63 1.23 1.12
PPO 0.022392 3.8245 361.4521 0.06 0.19 650.41 460.23 1120.32 1.42 1.08
PS 0.033238 3.5022 320.1372 0.02 0.29 731.25 361.23 1112.36 0.57 1.22
PVME 0.024245 3.9145 372.4584 0.18 0.24 580.52 392.45 1082.41 0.62 1.16

CO2 0.048237 2.7352 166.21 0.01a 0.26b 301.23 585.61 1532.53 0.19a 0.25b
a L,sat
This valor corresponds to P .
b
This valor corresponds to vL,sat .
P.F. Arce, M. Aznar / Fluid Phase Equilibria 286 (2009) 17–27 21

Table 3
Some physical characteristic of blends.

Blends Composition NP Temperature (K) Pressure (MPa) Volume (×10−3 m3 /kg) MW Polymer 1 MW Polymer 2

90/10 303.65–603.75 0.8911–1.0742


80/20 303.55–603.85 0.8939–1.0817
70/30 303.45–605.05 0.8980–1.0936
60/40 303.65–604.25 0.9012–1.0962
PPO/PSa 50/50 275 303.05–603.85 0–200 0.9030–1.1016 114,200 244,000
40/60 303.65–604.35 0.9042–1.1059
30/70 303.25–602.45 0.9061–1.1095
20/80 303.35–604.05 0.9105–1.1135
10/90 303.65–604.05 0.9141–1.1191

90/10 303.15–470.95 0.9036–1.0818


80/20 302.95–471.25 0.9025–1.0777
70/30 303.25–471.75 0.9017–1.0768
60/40 302.85–472.45 0.9065–10.767
PVME/PSb 50/50 242 302.55–471.65 0–200 0.9035–1.0636 114,200 99,000
40/60 303.15–471.15 0.9058–1.0609
30/70 303.05–470.95 0.9049–1.0491
20/80 302.65–470.35 0.9119–1.0504
10/90 303.45–472.25 0.9162–1.0532

90/10 1.13–1.25
76/24 1.09–1.20
60/40 1.06–1.16
PBD/PSc 66 383.15–533.15 0–97.86 330,000 600,000
40/60 1.03–1.10
30/70 1.00–1.05
10/90 0.97–1.01

60/40 303.85–470.85 0.7614–0.9053


PEO/PESd 130 0–200 200,000 20,000
80/20 347.15–451.35 0.8200–0.9385
a
Ref. [39].
b
Refs. [37–39].
c
Refs. [35,40].
d
Ref. [39].

Table 4
Correlated results for modeling the PVT properties of polymer blends.

Blends Composition PC-SAFT EoS SL EoS


a
kij v (%) v (%) kij v (%) v (%)a

90/10 0.6235 0.21 1.89 0.6123 0.84 3.89


80/20 0.3512 0.18 2.05 0.5684 0.75 4.12
70/30 0.2641 0.15 1.96 0.5023 0.92 4.76
60/40 0.2412 0.23 2.15 0.6123 1.05 5.03
PPO/PS 50/50 0.2487 0.12 1.75 0.4815 0.84 3.95
40/60 0.2532 0.20 2.06 0.5212 0.92 4.15
30/70 0.2845 0.09 1.86 0.4125 1.07 4.39
20/80 0.3015 0.05 1.92 0.5816 1.12 4.75
10/90 0.4125 0.15 1.78 0.5341 1.26 4.08

90/10 0.3013 0.21 2.08 0.6523 1.15 6.41


80/20 0.2351 0.20 2.15 0.6241 1.09 5.84
70/30 0.2132 0.18 2.31 0.7452 0.98 6.03
60/40 0.2075 0.20 1.96 0.5632 0.93 4.84
PVME/PS 50/50 0.2045 0.15 1.89 0.4860 1.06 4.92
40/60 0.0842 0.12 2.04 0.5314 1.23 5.06
30/70 0.2562 0.13 2.08 0.5650 1.74 5.18
20/80 0.3541 0.11 1.78 0.4289 1.82 4.87
10/90 0.5641 0.20 2.13 0.6120 0.96 5.03

90/10 0.2036 0.12 2.08 0.4952 1.56 4.63


76/24 0.1953 0.15 1.86 0.5658 0.96 4.89
60/40 0.1874 0.20 1.96 0.4953 1.23 5.06
PBD/PS
40/60 0.2032 0.18 1.84 0.6025 1.38 5.89
30/70 0.2365 0.16 2.06 0.5742 1.08 6.14
10/90 0.2428 0.19 2.07 0.7012 0.98 6.28

60/40 0.2641 0.08 2.32 0.8412 1.08 6.84


PEO/PES
80/20 0.2243 0.07 2.07 0.7362 1.23 7.41

Total mean deviation (TMD)b 0.17 2.01 1.13 5.32


a
These deviations correspond for kij = 0.0.

1
b
TMD = NP
NPi vi .
i=1
22 P.F. Arce, M. Aznar / Fluid Phase Equilibria 286 (2009) 17–27

Fig. 2. PVT behavior of PPO/PS blend (90/10). Experimental data [26]: () 0 MPa,
Fig. 4. PVT behavior of PBD/PS blends at 98.07 MPa. Experimental data [22,27]: ()
(䊉) 40 MPa, () 80 MPa, () 120 MPa, () 160 MPa, and () 200 MPa.
PBD/PS = 90/10, (䊉) PBD/PS = 76.5/23.5, and () PBD/PS = 40/60.

ume ( vL,sat ) for CO2 obtained with the PC-SAFT and SL models,
which were calculated in similar form than Eqs. (29) and (30). Figs. 2 and 3 show the specific volumes of PPO/PS and PVME/PS
blends (90/10) against temperature under different pressures,
4.2. Correlation and prediction of the PVT fluid behavior of pure respectively, corresponding to the experimental data [37–39] and
polymer blends those obtained by the PC-SAFT and SL EoS, while Fig. 4 shows the
specific volumes of PBD/PS blends against temperature for several
In Table 3, some physical characteristics of polymer blends stud- PBD/PS compositions at 98.07 MPa. PVT fluid behavior of PEO/PES
ied in this work are shown. In this table, each blend has two (80/20) blends correlated and predicted with the PC-SAFT and SL
numbers; the first one corresponds to the mass percent of the first EoS is shown in Fig. 5. In all of cases, the PC-SAFT and SL models
polymer in the blend, while the second one stands for the mass are capable to correlate and predict the PVT fluid phase behavior of
percent of the second polymer in the blend. these blends with high performance at several pressure and tem-
PC-SAFT and SL EoS are used for calculating and predicting perature conditions, but in terms of volume deviations, it was found
of the specific volumes for blends of polymers by adjusting only that PC-SAFT EoS gives better correlation than the SL model. Total
one interaction parameter to fit PVT experimental data of blends mean deviations for 26 blends are 0.17% for the PC-SAFT model
[39]. Table 4 lists the correlated results obtained with the PC-SAFT and 1.13% for SL model. Table 4 also shows the predicted results
and SL models in terms of the binary parameters which take into (kij = 0.0) being that the total mean deviations for 26 blends are
account the interactions between the monomers of polymers, and 2.01% and 5.32% for the PC-SAFT and SL Eos, respectively. From this,
the specific volume deviations, calculated in the same way as in Eq. obviously the PC-SAFT model can obtain a better prediction.
(30).

Fig. 3. PVT behavior of PVME/PS blend (90/10). Experimental data [24–26]: () Fig. 5. PVT behavior of PEO/PES blend (80/20). Experimental data [26]: () 0 MPa,
0 MPa, (䊉) 40 MPa, () 80 MPa, () 120 MPa, () 160 MPa, and () 200 MPa. (䊉) 40 MPa, () 80 MPa, () 120 MPa, () 160 MPa, and () 200 MPa.
P.F. Arce, M. Aznar / Fluid Phase Equilibria 286 (2009) 17–27 23

Table 5
Results of the cloud point temperatures obtained with the PC-SAFT and SL EoS.

Blends MWPolymer 1 /MWPolymer 2 PC-SAFT EoS SL EoS


a
kij T kij Ta

1100/1340 0.0123 0.0645


1100/1670 0.0112 0.1087
PBD/PS 1100/4370 0.0140 1.03 0.0840 3.43
2350/1900 0.0095 0.1021
2350/3300 0.0125 0.0963

99,000/50,000 0.0094 0.1143


99,000/100,000 0.0081 0.0815
PVME/PS 1.11 2.81
95,000/67,000 0.0120 0.1016
95,000/106,000 0.0106 0.0915

2000/550 0.0083 0.1028


PPG/PEGE 1.14 3.04
2000/750 0.0093 0.0643

20,000/4000 0.0103 0.0443


PEO/PES 20,000/20,000 0.0126 0.98 0.0506 2.59
20,000/200,000 0.0096 0.0613

7570/6610 0.0105 0.0814


7570/6835 0.0123 0.1043
PnPMA/PS 1.26 4.54
7570/6960 0.0097 0.0943
7570/7110 0.0104 0.0840
 exp 

NP
Ti −Ticalc 
1
a
T = NP exp .
T
i
i

4.3. Cloud points temperature of blends molecular-weight is 50,000 (blend 4), the value of its LCST is higher
than the blend 2 when PS molecular-weight is 100,000. Similar con-
Table 5 lists the results obtained for modeling the polymer blend clusion is applied when are compared the LCSTs of blends 1 and 3.
miscibility (cloud point temperature) for five blends in which the Value of the LCST is higher than when the PS molecular-weight is
polymers have different molecular weights. In this table the binary 67,000 against 106,000 in blend 1.
parameters are shown, which take into account the different inter- Both thermodynamic models managed to correlate and pre-
actions between the monomers of polymers in the LLE, and the dict with high precision the fluid phase behavior (LLE) of PVME/PS
cloud point temperature deviations, which were calculated in the blends as well as also the solution critical point (LCST), even the PC-
same way as in Eq. (30). It can be also seen that the binary parame- SAFT EoS became to be more efficient in terms of the cloud point
ters depend on the molecular weights of two polymers which form temperature deviations (1.11% for the PC-SAFT EoS and 2.81% for
the blend. From these temperature deviations, it can be possible the SL model).
to deduce that both thermodynamic models are able to correlate
the fluid phase behavior of the cloud point temperatures of these 4.3.3. PPG/PEGE blends
polymer blend systems with reasonable accuracy. Fig. 8 shows the phase diagrams of experimental data [36] of
PPG/PEGE systems and those obtained with the PC-SAFT EoS. In
4.3.1. PBD/PS blends
Fig. 6 shows the comparison of the fluid phase behavior
diagrams of PBD/PS systems with different molecular weights
corresponding to experimental data [35] and the polymer blend
miscibility diagrams obtained with the PC-SAFT and SL models.
In this figure it is also observed a fluid phase behavior of typical
UCST type for all molecular weights. The quality of correlation and
prediction made by the PC-SAFT and SL EoS are good in terms of
temperature deviations (1.03% for the PC-SAFT EoS and 3.43% for
the SL EoS). Thus, both models describe successfully these polymer
blends on the whole concentration and temperature range.
In all cases for these polymer blends, when the molecular weight
of PBD segments remains constant and the molecular weight of
the PS segments increases, the value of UCST tends to increase, in
others words the UCST value depends directly on the molecular
weight of one of the polymer segments which forms the blend, in
this particular case the PS polymer.

4.3.2. PVME/PS blends


Fig. 7 shows the polymer blend miscibility diagrams of PVME/PS
systems corresponding to the experimental data [37,38] and the
phase diagrams obtained with the PC-SAFT and SL EoS. These sys-
Fig. 6. PC-SAFT and SL correlations and predictions of cloud point temperature
tems have the oriented interaction between the chains of PVME and
behavior of PBD/PS blends. Experimental data [22]: () blend 1 [PBD (1100)/PS
exhibit a LCST behavior. In Fig. 7, it is also noticed that for blends 2 (1340)]; (䊉) blend 2 [PBD (1100)/PS (1670)]; () blend 3 [PBD (1100)/PS (4370)];
and 4, if PVME molecular-weight remains constant (99,000) and PS () blend 4 [PBD (2350)/PS (1900)]; () blend 5 [PBD (2350)/PS (3300)].
24 P.F. Arce, M. Aznar / Fluid Phase Equilibria 286 (2009) 17–27

Fig. 9. PC-SAFT and SL correlations and predictions of cloud point temperature


behavior of PEO/PES blends. Experimental data [28]: () blend 1 [PEO (4000)/PES
(20,000)]; (䊉) blend 2 [PEO (20,000)/PES (20,000)]; () blend 3 [PEO (200,000)/PES
(20,000)].

Fig. 7. PC-SAFT and SL correlations and predictions of cloud point tempera-


ture behavior of PVME/PS blends. Experimental data [24,25]: () blend 1 [PVME 4.3.4. PEO/PES blends
(95,000)/PS (106,000)]; (䊉) blend 2 [PVME (99,000)/PS (100,000)]; () blend 3 In Fig. 9, it is shown three PEO/PES blend miscibility behavior
[PVME (95,000)/PS (67,000)]; () blend 4 [PVME (99,000)/PS (50,000)].
predicted by the PC-SAFT and SL models and experimental data
[41]. This PEO/PES blend which is completely miscible at low tem-
these blends, there are two different polymer chains which inter-
perature due to specific interaction between the polymer segments
act strongly, so there is a proper orientation to each other. In similar
but partially miscible at higher temperatures, where these interac-
form that for PBD/PS blends, the PPG/PEGE blends present also the
tions weaken and free-volume effects dominate [41,57].
fluid phase behavior of UCST type. In Fig. 8 two blends are shown,
From this figure it is also possible to observe the phase behav-
both blends have the PPG segments which molecular weight is con-
ior of LCST type that three PEO/PES blends, differentiated just by
stant (2000) and PEGE segments (550 and 750 for both blends).
the change of the molecular weight of PEO segments, while the
Blend which contains PEGE segments with high-molecular weights,
molecular weight of PES segments remains constant. Effect of the
presents high value of its UCST than other PPG/PEGE blend with
molecular weight of the segments of one of the polymers on the
low-molecular weight of PEGE segments. Cloud point temperature
LCST of polymer blend is also noticed for PEO/PES blends. While
deviations obtained by the PC-SAFT and SL models were 1.14% and
the molecular weight of PEO segments (4000–200,000) increases,
3.04%, respectively.
the value of critical point temperature (LCST) decreases remain-
ing the molecular weight of PES segments constant (20,000). Even
though both thermodynamic models managed to correlate in a rea-
sonable way the LLE of PEO/PES blends, the PC-SAFT EoS managed
to predict with higher accuracy the polymer blend miscibilities in
terms of the cloud point temperature deviations than the SL EoS
(0.98% and 2.59% for the PC-SAFT and SL, respectively).

4.3.5. PnPMA/PS blends


Cloud point temperatures for PnPMA/PS blends with four dif-
ferent molecular weights of PS segments corresponding to the
experimental data [42] and predicted values obtained with the PC-
SAFT and SL EoS are shown in Fig. 10. For lower molecular weights of
PS segments (blends 1, 2 and 3), both UCST and LCST are observed.
Thus, blends become homogeneous at temperatures between the
UCST and the LCST. For these three systems when the PS molec-
ular weight increases, the UCST increases and the LCST decreases.
For the highest PS molecular weight, the UCST and the LCST curves
join together to form a special fluid phase behavior of hourglass
type [42]. Correlation results are excellent and the effect of the
polymer molecular weights on LLE is predicted by each thermo-
dynamic model, especially for the PC-SAFT EoS, which in terms of
Fig. 8. PC-SAFT and SL correlations and predictions of cloud point temperature
behavior of PPG/PEGE blends. Experimental data [23]: () blend 1 [PPG (2000)/PEGE
cloud point temperature deviations, became more efficient than
(750)]; (䊉) blend 2 [PPG (2000)/PEGE (550)]. the SL EoS (1.26% against 4.54%).
P.F. Arce, M. Aznar / Fluid Phase Equilibria 286 (2009) 17–27 25

Fig. 10. Cloud point temperature behavior of PnPMA/PS blends predicted with the
PC-SAFT and SL EoS. Experimental data [29]: () blend 1 [PnPMA (7570)/PS (6610)];
() blend 2 [PnPMA (7570)/PS (6835)]; (䊉) blend 3 [PnPMA (7570)/PS (6960)]; ()
blend 4 [PnPMA (7570)/PS (7110)].

Fig. 12. (a) CO2 sorption isotherms obtained with the PC-SAFT EoS in PS/PVME
blends (T = 293.15 K) at different blend compositions. Experimental data [31]: ()
PVME/PS blend (100/0); (䊉) PVME/PS blend (75/25); () PVME/PS blend (50/50);
() PVME/PS blend (30/70); () PVME/PS blend (0/100). (b) CO2 sorption isotherms
obtained with the SL EoS in PS/PVME blends (T = 293.15 K) at different blend com-
Fig. 11. CO2 sorption isotherms using the PC-SAFT and SL EoS in PS/PVME blends. positions. Experimental data [31]: () PVME/PS blend (100/0); (䊉) PVME/PS blend
Experimental data [2]: () 293.15 K; (䊉) 313.15 K. (75/25); () PVME/PS blend (50/50); () PVME/PS blend (30/70); () PVME/PS blend
(0/100).

Table 6
Pressure deviations and binary interaction parameters for CO2 + polymer, blends and blend + CO2 systems.

EoS Temperature (K) Systemsa

CO2 + PSb CO2 + PVMEc PVME/PS (50/50)d PVME/PS (50/50) + CO2 e

kij 0.0634 0.0080 0.2047


293.15 2.58
P 1.02 1.85 1.60
PC-SAFT
kij 0.0603 0.0066 0.2059
313.15 2.89
P 1.06 1.36 1.18

kij −0.0048 0.0254 0.4821


293.15 5.12
P 3.05 3.40 3.85
SL
kij −0.0029 0.0285 0.5041
313.15 4.89
P 2.99 4.02 4.03
a
MWPS = 100,000; MWPVME = 99,000.
b
Ref. [58].
c
Experimental data shown by [43].
d
See Section 4.2.
e
P.
26 P.F. Arce, M. Aznar / Fluid Phase Equilibria 286 (2009) 17–27

obtained by regression of PVT data of pure polymers. It was found


that both models can be satisfactorily used to correlate and pre-
dict the specific volumes of pure blends by introducing only one
interaction parameter by adjusting the PVT experimental data. Pre-
dicted results of the cloud point temperature for blends using the
PC-SAFT and SL EoS were also acceptable and satisfactory predic-
tions of the molecular weight effects on polymer miscibility of both
the UCST and LCST type are also possible with the PC-SAFT and SL
models. There is an influence of the molecular weight on the loca-
tions of the critical point (CSTs, critical solutions temperatures).
For PBD/PS and PPG/PEGE blends, when the PEGE and PS molecular
weight increases, the UCST increases and for PVME/PS and PEO/PES
blends, when the PS and PEO molecular weight increases, the LCST
decreases. PC-SAFT and SL EoS were also able to predict the fluid
phase behavior of PnPMA/PS blends of UCST, LCST, and hourglass
type. A rigorous analysis using the PC-SAFT and SL EoS was used
to know qualitatively the phase stability of the pseudo binary sys-
tem PS/PVME + CO2 . Results indicate that the sorption of CO2 in
PS/PVME blends can be correlated by the PC-SAFT and SL models
Fig. 13. Stability analysis for the pseudo binary system PS/PVME blend + CO2 using with high accuracy and that CO2 solubilities increase with PVME
the PC-SAFT and SL EoS. Experimental data [2]: (䊉) 313.15 K. concentration in PVME/PS blends.

4.4. Fluid phase behavior of PS/PVME blend + CO2 systems Acknowledgments

Since each of the three binary interaction parameters is known The financial support of Fundação de Amparo à Pesquisa do
independently (Table 6), the calculation can in principle be con- Estado de São Paulo, FAPESP (Brazil), through grant 05/53685-4,
ducted without any additional parameters. It is instructive to first is gratefully acknowledged.
assess how the EoS formalism performs for predicting the solubil-
ity of CO2 in the pseudo binary PS/PVME + CO2 system. In Fig. 11, References
it is shown the results and experimental data for the CO2 sorp-
tion isotherms obtained with the PC-SAFT and SL EoS for PVME/PS [1] B.H. Chang, Y.C. Bae, Molecular thermodynamics approach for liquid–liquid
(50/50) blend + CO2 at 293.15 and 313.15 K. Results indicate that equilibria of the symmetric polymer blend systems, Chem. Eng. Sci. 58 (2003)
2931–2936.
calculations are correct, providing precise agreement with the [2] V.S. Ramachandra Rao, J.J. Watkins, Phase separation in polystyrene-poly(vinyl
experimental data. As shown in Fig. 12a and b, the experimental methyl ether) blends dilated with compressed carbon dioxide, Macromolecules
sorption isotherms for CO2 are too linear with the equilibrium pres- 33 (2000) 5143–5152.
[3] X. Chen, Z. Sun, J. Yin, L. An, Thermodynamics of blends of PEO with PVAc:
sure [44]. In these figures, it is easy to notice that CO2 solubility
application of the Sanchez–Lacombe lattice-fluid theory, Polymer 41 (2000)
increases with PVME concentration. Perhaps it can be explained by 5669–5674.
the effect of polarity on the PVME segment which is higher than [4] N.K. Kalfoglou, Compatibility of poly(ethylene oxide)–poly(vinyl acetate)
blends, J. Polym. Sci. Polym. Phys. Ed. 20 (1982) 1259–1267.
that of the PS ones, thus the dipole–dipole interactions with CO2
[5] N.K. Kalfoglou, D.D. Sotiropoulou, A.G. Margaritis, Thermal and morphology
increase with the concentration in PVME. Form theses figures it characterization of blends of poly(ethylene oxide) with poly(vinyl acetate), Eur.
can also observe that isotherms for high PS concentrations show a Polym. J. 4 (1988) 389–394.
slight concavity towards the pressure axis even if the correlation [6] E. Munoz, M. Calahorra, A. Santamaria, Melt behaviour of poly(ethylene
oxide)–poly(vinyl acetate) blends, Polym. Bull. 7 (1982) 295–301.
obtained with the PC-SAFT and SL are linear, as shown in Fig. 12a [7] E. Martuscelli, C. Silvestre, C. Gismondi, Morphology, crystallization and ther-
and b, respectively. mal behaviour of poly(ethylene oxide)/poly(vinyl acetate) blends, Makromol.
The stability criterion for the ternary mixture PS/PVME/CO2 can Chem. 186 (1985) 2161–2176.
[8] C. Silvestre, F.E. Karasz, W.J. Macknight, E. Martuscelli, Morphology of
be evaluated directly by differentiation of the Gibbs free energy [2] poly(ethylene oxide)/poly(vinyl acetate) blends, Eur. Polym. J. 23 (1987)
from the following equation: 745–751.
  [9] M.L. Addonizio, E. Martuscelli, C. Silvestre, Isothermal and non-isothermal crys-
 Gxx Gxy Gxṽ  tallisation of poly(ethylene oxide)/poly(viny acetate) blend, J. Polym. Mater. 7
 
 Gyx Gyy Gyṽ  > 0 (31) (1990) 63–70.
G G G 
[10] X. Chen, H. Hu, J. Yin, C. Zheng, Miscibility studies of poly(ethylene
ṽx ṽy ṽṽ oxide)–poly(vinyl acetate) blends by viscometry methods, J. Appl. Polym. Sci.
56 (1995) 247–252.
where x and y are two independent compositions variables and ṽ is [11] X. Chen, J. Yin, G.C. Alfonso, E. Pedemonte, A. Turturro, E. Gattiglia, Thermody-
the reduced volume of the system [17]. Prediction of the stability namics of blends of poly(ethylene oxide) with poly(methyl methacrylate) and
of the pseudo binary system using the PC-SAFT EoS and SL is given poly(vinyl acetate): prediction of miscibility based on Flory solution theory
modified by Hamada, Polymer 39 (1998) 4929–4935.
in Fig. 13. Both models accurately predict a phase split upon CO2
[12] X. Chen, L. An, L. Li, J. Yin, Z. Sun, Mixing enthalpy and phase behavior of
sorption (no change of the sign in the determinant of Eq. (31)), PEO/PVAc blends, Macromolecules 32 (1999) 5905–5910.
concluding that the prediction of phase separation pressure of CO2 [13] F. Humada, T. Shiomi, K. Fujisawa, A. Nakajima, Statistical thermodynamics of
polymer solutions based on free volume theory, Macromolecules 13 (1980)
made by the PC-SAFT and SL EoS is reasonable and correct.
729–734.
[14] M. Ebert, R.W. Garbella, J.H. Wendorff, Studies on the heat of demixing of
5. Conclusions poly(ethylene acrylate)–poly(vinylidene fluoride) blends, Macromol. Chem.
Rapid. Commun. 7 (1986) 65–70.
[15] S. Shen, J.M. Torkelson, Miscibility and phase separation in poly(methyl
PC-SAFT and SL EoS were used to correlate and predict specific methacrylate)/poly(vinyl chloride) blends: study of thermodynamics by ther-
volumes of pure blends (PBD/PS, PPO/PS, PVME/PS, and PEO/PES), mal analysis, Macromolecules 25 (1992) 721–728.
cloud point temperature of blend miscibilities of several polymer [16] D.R. Paul, S. Newman (Eds.), Polymer Blends, Academic Press, New York, 1978.
[17] I.C. Sanchez, Polymer Compatibility and Incompatibility Principles and Prac-
blends (PBD/PS, PPG/PEGE, PVME/PS, PEO/PES, and PnPMA/PS) and tices, in: K. Solc (Ed.), MIMIM Press Symposium Series, vol. 2, Hardwood
the sorption of CO2 in PVME/PS blends. Segment parameters were Academic Publishers, Chur, Switzerland, 1982.
P.F. Arce, M. Aznar / Fluid Phase Equilibria 286 (2009) 17–27 27

[18] L.A. Utracki, Polymer Blends Handbook, vol. I, Kluwer Academic Publishers, The distribution mixtures of polystyrene and poly(vinyl methyl ether), Macro-
Netherlands, 2002. molecules 22 (1989) 3395–3399.
[19] L.A. Utracki, Polymer Alloys and Blends, Hanser Publishers, Munich, 1989. [39] P. Zoller, D. Walsh, Standard Pressure–Volume–Temperature Data for Poly-
[20] P.J. Flory, Thermodynamics of high polymer solutions, J. Chem. Phys. 10 (1942) mers, Technomic Publishing Co. Inc., Lancaster, Basel, 1995.
51–61. [40] Y.Z. Wang, K.H. Hsieh, L.W. Chen, H.C. Tseng, Effect of compatibility on specific
[21] M.L. Huggins, Some properties of solutions of long-chain compounds, J. Phys. volume of molten polyblends, J. Appl. Polym. Sci. 53 (1994) 1191–1201.
Chem. 46 (1942) 151–158. [41] D.J. Walsh, The role of specific interactions in polymer miscibility, in: L.A. Klein-
[22] P.J. Flory, R.A. Orwoll, A. Vrij, Statistical thermodynamics of chain molecule tjens, P.S. Lemstra (Eds.), Integration of Fundamental Polymer Science and
liquids. I. An equation of state for normal paraffin hydrocarbons, J. Am. Chem. Technology, Elsevier, New York, 1986.
Soc. 86 (1964) 3507. [42] D.Y. Ryu, M.S. Park, S.H. Chae, J. Jang, J.K. Kim, Phase behavior of polystyrene and
[23] D. Patterson, Free volume and polymer solubility, a qualitative view, Macro- poly(n-pentyl methacrylate) blend, Macromolecules 35 (2002) 8676–8680.
molecules 2 (1969) 672–677. [43] Y. Wang, Molecular Modeling Applied to CO2 -Soluble Molecules and Confined
[24] S. Beret, J.M. Prausnitz, Perturbed hard-chain theory—equation of state for flu- Fluids, PhD Thesis, University of Pittsburgh, Pittsburgh, PA, USA, 2006.
ids containing small or large molecules, AIChE J. 21 (1975) 1123–1132. [44] A. Mokdad, A. Dubault, L. Monnerie, Sorption and diffusion of carbon diox-
[25] M.D. Donohue, J.M. Prausnitz, Perturbed hard chain theory for fluid ide in single-phase polystyrene/poly(vinylmethylether) blends, J. Polym. Sci.
mixtures—thermodynamic properties for mixtures in natural-gas and B: Polym. Phys. 34 (1996) 2723–2730.
petroleum technology, AIChE J. 24 (1978) 849–860. [45] W.G. Chapman, G. Jackson, K.E. Gubbins, Phase equilibria of associating fluids,
[26] I. Sanchez, R. Lacombe, An elementary molecular theory of classical fluids. Pure Mol. Phys. 65 (1988) 1057–1079.
fluids, J. Phys. Chem. 80 (1976) 2352–2362. [46] W.G. Chapman, K.E. Gubbins, G. Jackson, M. Radosz, New reference equation of
[27] R.H. Lacombe, I. Sanchez, Statistical thermodynamics of fluid mixtures, J. Phys. state for associating liquids, Ind. Eng. Chem. Res. 29 (1990) 1709–1721.
Chem. 80 (1976) 2568–2580. [47] J.A. Barker, D. Henderson, Perturbation theory and equation-of-state for fluids:
[28] J. Dudowicz, K.F. Freed, Effect of monomer structure and compressibility on the square-well potential, J. Chem. Phys. 47 (1967) 2856–2861.
the properties of multicomponent polymer blends and solutions. 3. Applica- [48] I. Sanchez, R. Lacombe, Statistical thermodynamics of polymer solutions,
tion to deuterated polystyrene [PS(D)]poly(vinyl methyl ether) (PVME) blends, Macromolecules 11 (1978) 1145–1156.
Macromolecules 24 (1991) 5112–5123. [49] P.A. Rodgers, Pressure–volume–temperature relationships for polymer liquids:
[29] C. Peng, H. Liu, Y. Hu, Equations of state for copolymer systems based on dif- a review of equations of state and their characteristic parameters for 56 poly-
ferent perturbation terms, Fluid Phase Equilib. 206 (2003) 147–162. mers, J. Appl. Polym. Sci. 48 (1993) 1061–1080.
[30] T. Hino, Y.H. Song, J.M. Prausnitz, Equation-of-state analysis of binary copoly- [50] L. Stragevitch, S.G. d’Avila, Application of a generalized maximum likelihood
mer systems. 2. Homopolymer and copolymer mixtures, Macromolecules 28 method in the reduction of multicomponent liquid–liquid equilibrium data,
(1995) 5717–5724. Br. J. Chem. Eng. 14 (1997) 41–52.
[31] J. Gross, G. Sadowski, Application of perturbation theory to a hard-chain refer- [51] V.G. Niesen, V.F. Yesavage, Application of a maximum likelihood method using
ence fluid: an equation of state for square-well chains, Fluid Phase Equilib. 168 implicit constraints to determine equation of state parameters from binary
(2000) 183–199. phase behavior data, Fluid Phase Equilib. 50 (1989) 249–266.
[32] J. Gross, G. Sadowski, Perturbed-chain SAFT: an equation of state based on [52] P. Arce, M. Aznar, Modeling the thermodynamic behavior of poly(lactide-co-
a perturbation theory for chain molecules, Ind. Eng. Chem. Res. 40 (2001) glycolide) + supercritical fluid mixtures with equations of state, Fluid Phase
1244–1260. Equilib. 244 (2006) 16–25.
[33] P. Arce, M. Aznar, Modeling the phase behavior of commercial biodegradable [53] Y. Song, S.M. Lambert, J.M. Prausnitz, A perturbed hard-sphere-chain equation
polymers and copolymer in supercritical fluids, Fluid Phase Equilib. 238 (2005) of state for normal fluids and polymers, Ind. Eng. Chem. Res. 33 (1994) 1047.
242–253. [54] Y. Song, T. Hino, S.M. Lambert, J.M. Prausnitz, Liquid–liquid equilibria for
[34] P. Arce, S. Mattedi, M. Aznar, Vapor–liquid equilibrium of copolymer + solvent polymer solutions and blends, including copolymers, Fluid Phase Equilib. 117
mixtures: thermodynamic modeling by two theoretical equations of state, Fluid (1996) 69.
Phase Equilib. 246 (2006) 52–63. [55] N. Elvassore, A. Bertucco, M. Fermeglia, Phase-equilibria calculation by group-
[35] S. Rostami, D.J. Walsh, Simulation of upper and lower critical phase diagrams for contribution perturbed-hard-sphere-chain equation of state, AIChE J. 48 (2002)
polymer mixtures at various pressures, Macromolecules 18 (1985) 1228–1235. 359.
[36] K. Takahashi, T. Kyu, Q. Trancong, O. Yano, T. Soen, Phase-separation in mix- [56] DIPPR Information and Data Evaluation Manager, Version 1.2.0, 2000.
tures of poly(ethylene glycol monomethylether) and poly(propylene glycol) [57] E.P. Voutsas, G.P. Pappa, C.J. Boukouvalas, K. Magoulas, D.P. Tassios, Miscibility
oligomers, J. Polym. Sci. B: Polym. Phys. 29 (1991) 1419–1425. in binary polymer blends: correlation and prediction, Ind. Eng. Chem. Res. 43
[37] Y.C. Bae, J.J. Shim, D.S. Soane, J.M. Prausnitz, Representation of vapor–liquid (2004) 1312–1321.
and liquid–liquid equilibria for binary systems containing polymers: applica- [58] P. Arce. Modelling and Computation of Multiphase Equilibrium of Critical Flu-
bility of an extended Flory–Huggins equation, J. Appl. Polym. Sci. 47 (1993) ids and Phenomena in Polymer Solubilities in Supercritical CO2 + Co-solvent
1193–1206. Mixtures, D.Sc. Thesis, State University of Campinas, Campinas, SP, Brazil, 2005
[38] D.J. Walsh, G.T. Dee, J.L. Halary, J.M. Ubiche, M. Millequant, J. Lesec, L. Monnerie, (in Portuguese).
The application of equation-of-state theories to narrow molecular-weight

Das könnte Ihnen auch gefallen