FPE Final Edited Joda

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/270338935

Application of NRTL, UNIQUAC and NRTL-NRF models for Ternary Liquid-


Liquid Equilibrium Systems Including Ionic Liquids

Article · December 2011

CITATION READS

1 2,859

3 authors:

Ali Haghtalab Aliakbar Paraj


Tarbiat Modares University Tarbiat Modares University
136 PUBLICATIONS   1,515 CITATIONS    14 PUBLICATIONS   34 CITATIONS   

SEE PROFILE SEE PROFILE

Babak Mokhtarani
Chemistry & Chemical Engineering Research Center of Iran
115 PUBLICATIONS   1,422 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Correlation between Foaming Behaviour of Polyehtylene Blends with their Thermal and Rheological Properties View project

Synthesis of PES-based cation exchange membrane and enhancement of their properties by adding functionalized graphene oxide Nano sheets View project

All content following this page was uploaded by Ali Haghtalab on 05 July 2015.

The user has requested enhancement of the downloaded file.


Fluid Phase Equilibria 278 (2009) 20–26

Contents lists available at ScienceDirect

Fluid Phase Equilibria


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

Modification of NRTL-NRF model for computation of liquid–liquid equilibria in


aqueous two-phase polymer–salt systems
Ali Haghtalab ∗ , Marzieh Joda
Department of Chemical Engineering, Tarbiat Modares University (TMU), P.O. Box 14115-111, Tehran, Iran

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

Article history: The aqueous two-phase system (ATPS), a polymer and a salt, is widely used as an extraction process for
Received 20 September 2008 separation of biomolecules in down stream processing. At this work, using local area fraction, the local
Received in revised form 5 December 2008 composition non-random factor (NRF) model of Haghtalab and Vera is extended for aqueous binaries of
Accepted 18 December 2008
electrolyte, polymer and ternaries of aqueous polymer–salt solutions. Using area fraction of species, a new
Available online 27 December 2008
function of excess Gibbs energy has been developed for multicomponent aqueous polymer–electrolyte
systems. The modified NRTL-NRF model is applied to correlate the phase behavior of aqueous two-phase
Keywords:
polymer–salt systems of PEG + Na2 SO4 + H2 O, PEG + K2 HPO4 + H2 O and PEG + (NH4 )2 SO4 + H2 O at 25 ◦ C.
Activity coefficient
Excess Gibbs energy
The effect of pH solution and molecular weight of polyethylene glycol (PEG) is investigated and the tie
Electrolyte solutions lines with binodal curves for the fifteen systems are shown. The comparison of the results of the present
Aqueous two-phase system model with the experiment shows that the modified NRTL-NRF model is more accurate.
NRTL-NRF © 2008 Elsevier B.V. All rights reserved.
Polyethylene glycol
Sodium sulfate
Potassium sulfate
Ammonium sulfate

1. Introduction logical materials [1]. A wide range of biomolecules (protein, lipids,


nucleic acids, viruses and whole cells) have been separated using
Presently, industrial demands and economic downstream pro- this technique [2,3], specially for proteins, since it takes less time
cesses for extraction and purification of biomolecules with high consuming and has the potential to give high yield and high purity,
yield purity of the product are growing fast [1,2]. The aqueous two- involving low investment, less energy, and lower labor costs. On
phase system (ATPS) provides a powerful method to separating the other hand, a liquid–liquid extraction process requires knowl-
mixtures of biomolecules by extraction in down stream processing. edge of the phase behavior of the system for engineering design
Using ATPS as a practical process allows one to integrate clarifi- and process optimization.
cation, concentrating, and partial purification of biomolecules in To design and operate an aqueous two-phase extraction pro-
one step. The ATPS consists of two immiscible aqueous solutions cess, availability of the predictive models to construct the phase
that it contains two different polymers such as polyethylene gly- diagrams is of great importance [4,5]. As a consequence, some suc-
col (PEG) and Dextran or one polymer and one salt such as PEG cessful models in this specialized area have been developed. The
and ammonium sulfate. Also polymer–salt aqueous two-phase sys- first group of the models, based on osmotic virial expansion, was
tems have several advantages such as low price, low viscosity, and derived from knowledge of the osmotic pressure of a solvent in a
short time for phase separation. The basis of partitioning depends solution. King et al. [6], Sassi et al. [7], Wu et al. [8] and Kabiri-
upon surface properties of the particles and molecules that include Badr and Cabezas [9] developed a virial model which was based on
size, charge, and hydrophobicity. Moreover, the most characteris- Hill theory. Haynes et al. [10,11] developed a McMillan-Mayer type
tic feature of a two-phase system is that the water content in such model that was derived in the semi-grand ensemble at constant
system is as high as 85–99%, when is complemented with suitable temperature and pressure. The second group of the models, based
buffers and salts allows one to provide a suitable milieu for bio- on the extension of the local composition concept, includes the
NRTL, UNIQUAC, UNIFAC [12–16], UNIQUAC-NRF and UNIFAC-NRF
models [17,18] that have been used for correlation and prediction
liquid–liquid equilibria and phase diagrams of polymer–polymer
∗ Corresponding author at: Department of Chemical Engineering, Tarbiat Modares
and polymer–salt ATPS.
University (TMU), P.O. Box 14115-143, Tehran, Iran. Tel.: +98 21 82883313;
fax: +98 21 82883381. Haghtalab and Vera [19] proposed the NRTL-NRF model for
E-mail address: haghtala@modares.ac.ir (A. Haghtalab). mean activity coefficient of the binary aqueous electrolyte sys-

0378-3812/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.fluid.2008.12.006
A. Haghtalab, M. Joda / Fluid Phase Equilibria 278 (2009) 20–26 21

tems. At this model, the random state of species is considered i and j. The Gibbs energy of the same cell in the reference state,
as a reference state and the nonrandom factor,  , is used to random state, is expressed as
show the deviation of local composition from bulk composition.

n

Using the assumptions of NRTL-NRF model, Haghtalab and Asadol- gi0 = j gji (8)
lahi [20] developed the UNIQUAC-NRF model which was used for
j=1
polymer–polymer ATPS. Haghtalab and Mokhtarani [17] applied the
j=
/ i for i = 1, 2
UNIQUAC-NRF model for polymer–salt ATPS. Based on the group
contribution method, Haghtalab and Mokhtarani developed the where j is the effective area fraction of species j and ji = zj ji for
UNIFAC-NRF model [21,22] for ATPS of polymer–salt with and with- ionic species and ji = ji for molecular species, where z is the charge
out proteins. At this work on extension of the NRTL-NRF model, the number of an ion. As one can see, we assume the random state as
new version of local area fraction model is developed for aqueous a reference state. So using the two-fluid theory, the molar residual
polymer–salt solutions. excess Gibbs energy of a solution is written as
   gE
2. The modified NRTL-NRF model with area fraction gE i
= xi (9)
RT RT
R i
At this work, we follow Haghtalab and Vera’s model, NRTL-NRF
[19], with assumption of the existence of four types of cells for At this study, the local area fraction is represented by means of
polymer–salt solutions that it depends on number of species in a nonrandom factors (NRF). Thus, the effective local area fraction of
solution. The excess Gibbs energy and activity coefficient functions i–j interactions can be written as
are derived on the basis of the assumptions of the NRTL-NRF model ji = j ji (10)
and local area fraction of the species. Using the general approach of
the local composition models, the excess Gibbs energy is written as Similarly,
a sum of three contributions,
        ii = i ii (11)
gE gE gE gE
= + + (1) where “ ” is called the nonrandom factor. So by dividing the above
RT RT RT RT relations,
DH C R

where the first term accounts for the contribution of long-range ji j
electrostatic interactions; the second and third terms denote the = ji (12)
ii i
contribution of combinatorial and residual parts, respectively. So by
proper differentiation of the above equation, the activity coefficient where
of species i can be expressed as ji
ji = (13)
ln i = ln iDH + ln iC + ln iR (2) ii
On the other hand, using Boltzmann factor as was proposed by
The long-range activity coefficient, i.e., Debye-Hückel equation,
Wilson [23],
is given in Appendix A. Using Guggenheim equation, the combina-  a 
torial activity coefficient can be written as ji
ji = exp − (14)
Z
  
3    
3
i z i i where
ln iC = ln + q ln + li − xj lj (3)
xi 2 i i xi gji
i=1 j=1
aji = (15)
RT
where
where aij is the binary interaction adjustable parameter of i–j pair
z
li = (ri − qi ) − (ri − 1) (4) and Z is the coordination number. By combining Eqs. (6–8) and
2
using Eqs. (10–13), the residual molar excess Gibbs energy, Eq. (9),
and for a multicomponent system is obtained as
ri xi      
i = 3 (5) gE
n
n ji
rx
i=1 i i =− Zxi j (ji − 1)ln (16)
RT j=1 ii
R i=1
The short-range contribution of activity coefficient can be
developed as following. For a multicomponent system the config- Consequently, by proper differentiation of Eq (16), the equa-
urational excess Gibbs energy for each central cell can be written tion of activity coefficient for a multicomponent system can be
obtained:
as

n   
n

n
ji qi ij
giE = gi − gi0 (i = a, c, w, p) (6) ln i = Z(1 − i ) j ln +
qj
ln −Z
ii jj
j=1 j=1 k=1
where gi and gi0 are Gibbs energy of the cell “i” in the nonrandom
j=
/ i j=
/ i k=/ i
and random states, respectively. for k = j, k = ion
Also the configurational Gibbs energy of the each cell in the
nonrandom state is expressed as   
n

n
ji qi ij qi   kj
j k ji ln + ij kj ln +Z   ln (ij kj − 1) (17)
qj qj j k

n ii jj
j=1 k=1
jj

gi = ij gji (7)


j=
/ i k=
/ i, j
j=1
j=
/ i for j = 1, 2 where
x qi
where ji
is effective local area fraction of species j surrounding i = n i (18)
xq
j=1 j j
central species i, and gji is interaction energy between two species
22 A. Haghtalab, M. Joda / Fluid Phase Equilibria 278 (2009) 20–26

ji 1 Table 1
ji = ; ii = (19) The group parameters for segments of PEG and water [15].

n 
n
j .ji j .ji Group H2 O CH2 OH CH2 OCH2
j=1 j=1 R 0.92 1.674 1.593
i=
/ j for ion i=
/ j for ion Q 1.400 1.740 1.320

For a binary electrolyte system with three species, an anion (1),


a cation (2) and a molecular solvent (3), using Eq. (16) the molar and “Q” parameters for the ions were calculated by Bondi’s equa-
excess Gibbs energy can be written as [24]: tion [26]. The structural parameters for all components are given
  at Table 2. The mean activity coefficient data for binary aqueous
gE
= qE xE E (E − 1)aE − q3 x3 3 (3 − 1)aE3 (20) systems of K2 HPO4 , (NH4 )2 SO4 and Na2 SO4 are given by Stokes and
RT
R Robison [25].
where the area parameter of an electrolyte pair-ion is expressed as The computation of liquid–liquid equilibria for the
ternary systems PEG–Na2 SO4 –H2 O, PEG–K2 HPO4 –H2 O and
qE = 1 q1 + 2 q2 (21) PEG–(NH4 )2 SO4 –H2 O is investigated for various molecular weights
The unsymmetrical mean activity coefficient for a binary elec- of PEG and different pH at 25 ◦ C. Gao et al. [15,16], Peng et al. [28]
trolyte system is derived as [24]: and Haghtalab and Mokhtarani [17] reported the experimental

qE 

 qE
 data for those systems. Using computation of the LLE with the
(ln ±∗ )R =  aE (E − 1)(1 + 3 ) +E2 E 3 −1 equilibrium experimental data, the interaction parameters of the
 E E ZE E qE species are obtained by optimization of the following objective
qE  2 qE

2ZE E qE 3
 function [15,16]:
(32   3 − 1)
3
+  3 (3 − 1) − aE3 1 −
  qE 2
(22)

n
4 (xi i∗ )Ij − (xi i∗ )˘
j
1 = (27)
where aE and aE3 are the binary adjustable parameters for a binary
i=1 (xi i∗ )IIj
j=1
electrolyte system. The subscripts “E” and “3” denote electrolyte
where I and II represent upper and lower phases, respectively. The
and solvent, respectively.
subscripts i and j denote the species and tie lines, respectively. The
interaction parameters are used for calculation of the composition
3. Aqueous two-phase system
of species in the two phases. However, to obtain the accurate values
of the parameters, those again should be optimized by using the
At this work, the extended NRTL-NRF model is applied for
following objective function:
computation of liquid–liquid equilibria of the aqueous two-phase
2  2
systems. The system consists of four species as anion (1), cation (2), 
M

3 
exp cal
polyethylene glycol (PEG) (3) and water (4). The standard state is 2 = Min xijk − xijk (28)
chosen in a conventional way, i.e., the pure liquid is taken for water k i j
and infinite dilution for ions and PEG. So the activity coefficient of
the ionic species and polymer compound is easily normalized to where xexp and xcal are the experimental and calculated mol frac-
those based on the infinite dilution standard state. Thus, the activ- tions, respectively. The subscripts j, i and k denote the numbers of
ity coefficient of the anion and cation at infinite dilution is obtained phases, species and tie lines, respectively. The calculated values of
as the compositions in the two equilibrium phases are obtained by
q1 flash calculation through Newton-Rafson algorithm.
ln 1∞ = Z (1 − 14 )ln 14 (23) The present model has six adjustable parameters aE, aE3, aE4, a3E,
q4
a4E and a34 as follows:
q2
ln 2∞ = Z (1 − 24 )ln 24 (24)
q4 a12 = a21 = aE (29)
Similarly the activity coefficient of the polymer at infinite dilu- with using assumption:  13 =  23 , one can write
tion is expressed as
 2  a13 = a23 + aE3 + aE3 and a31 = a32 = a3E (30)
43 q3
ln 3∞ = −Z −1 ln 43 − Z (34 − 1)ln 34 (25) Similarly, one can conclude,
E3 q4

Finally, the unsymmetrical residual molar excess Gibbs energy a14 = a24 + aE4 + aE3 and a41 = a42 = a4E (31)
can be expressed as
    Table 2
g E,∗ gE
= − x1 ln 1∞ − x2 ln 2∞ − x3 ln 3∞ (26) The volume and surface parameters of the different compounds.
RT RT
R R Component r (pm) q (cm2 /mole) Reference

H2 O 0.92 1.40 [15]


4. Results and discussion
(NH4 )2 SO4 2.70 2.226 [27]
Na2 SO4 1.68 1.914 [27]
The structural parameters, i.e., the surface and volume of ions, K2 HPO4 3.144 2.362 [27]
polymer, and water, are needed in the MNRTL-NRF equation. So PEG(1000) 37.26 31.59 [15]
dividing the molecular form of PEG into two groups: CH2 OH and PEG(1500) 55.34 46.58 [15]
PEG(1450) 53.53 44.71 [15]
CH2 OCH2 , the surface and volume parameters of PEG and water PEG(1540) 56.79 47.77 [15]
molecules are obtained using the group division that was used by PEG(2000) 73.42 61.55 [15]
Gao et al. [15,16]. The volume and surface area parameters of the PEG(3350) 122.4 102.2 [15]
groups were denoted by “R” and “Q”, respectively, which are shown PEG(4000) 145.7 121.5 [15]
PEG(6000) 218.7 181.4 [15]
at Table 1. The radii of ions are given by Marcus [27] and the “R”
A. Haghtalab, M. Joda / Fluid Phase Equilibria 278 (2009) 20–26 23

Fig. 1. The activity of water in the PEG–H2 O system versus weight fraction of the polymer at different molecular weight of polymer, experimental data [29].

It should be noted that the parameters a34 and a43 are not algorithm, the values of interaction parameters were optimized
equivalent; however, one can obtain a43 in terms of the other with regression of the experimental data. These parameters are
parameters as shown in Table 5, respectively, for the systems PEG–Na2 SO4 –H2 O,
PEG–K2 HPO4 –H2 O and PEG–(NH4 )2 SO4 –H2 O. The average relative
a43 = a34 + a4E + aE3 − a3E − aE4 (32)
deviation is defined as

1  (xi − xi )
In correlation of ATPS, the binary interaction parameters for n exp
cal
the salt–polymer–water systems are obtained. Before applying the ε= exp × 100% (33)
MNRTL-NRF model for calculation of LLE in ATPS, the activity of n xi
i=1
water was correlated for the binary PEG–H2 O systems at different
molecular weight of PEG [29]. Table 3 shows the binary interac- where n is the number of experimental points; the subscripts ‘exp’
tion parameters for aqueous PEG solution at different molar masses. and ‘cal’ denote the experimental and calculated values, respec-
Also Fig. 1 shows the change of activity of water respect to the mole tively. The overall average deviation is defined as follows:
fraction of PEG at various molar masses. As one may consider the 1
results of correlation of the water activity is in very good agreement OARD = εj (34)
3
with experiment. j
Using Eq. (22), the mean activity coefficient of the binary elec-
where εj presents the relative deviation for each component. It
trolyte systems of (NH4 )2 SO4 –H2 O, K2 HPO4 –H2 O and Na2 SO4 –H2 O
should be noted that the missing pHs at Table 5 were not reported
were correlated and the two binary interaction parameters aE and
in the original papers [28,15].
aE4 were adjusted as shown in Table 4.
The results of four systems at different molecular weight and
By fixing four parameters and adjusting only two parameters,
pH are presented at Table 6. As one can see the results are in very
attempts to correlate the ternary systems in ATPS were not suc-
good agreement with experiment. Also one may observe that the
cessful because the data for the binaries were from different source.
overall average relative deviation of the calculated mole fraction
So by fixing the two electrolyte parameters, only four parameters
from the experiment is very low for the systems PEG–Na2 SO4 –H2 O,
were optimized for ternary systems at each molecular weight of
PEG–K2 HPO4 –H2 O and PEG–(NH4 )2 SO4 –H2 O at different molecu-
PEG and for given pH. Using Simplex method of the Nelder-Mead
lar weight and pH. The results of the present model are compared
with those obtained using the UNIFAC, UNIQUAC and UNIFAC-NRF
Table 3
The interaction parameters for binary PEG(3)–H2 O(4) systems at 25 ◦ C using MNRTL-
NRF, experimental data [29]. Table 4
The interaction parameters for binary electrolyte systems at 25 ◦ C using MNRTL-NRF,
MW of PEG a34 a43 RMSDa
experimental data [25].
1000 25.01 −0.998 0.218
1450 26.404 −1.098 0.095 Binary electrolyte solution aE aE4 RMSDa
3350 26.634 −1.248 0.067 (NH4 )2 SO4 –H2 O −2.000 4.000 0.359
6000 26.104 −1.648 0.079 K2 HPO4 –H2 O 18.058 63.906 0.327
 1/2 Na2 SO4 –H2 O 74.000 59.800 0.199
 2
 1/2
exp 2
a
RSMD = 1/N (ai − acal
i
) × 100; ai = activity of water. a
RSMD = 1/N
exp
(ln i − ln ical ) × 100;  i = mean activity coefficient of
i electrolyte.
24 A. Haghtalab, M. Joda / Fluid Phase Equilibria 278 (2009) 20–26

Table 5
The interaction parameters for the ternary PEG(3)–Salt(E)–H2 O(4) systems at different molecular weight and pH, at 25 ◦ C.

MW of PEG pH aE3 a3E a4E a34

PEG–Na2 SO4 –H2 O [17]


1500 4.7 −69.975 −59.975 −113.975 45.017
4000 9.2 −530.983 −106.983 −140.983 66.709
4000 5.3 −560.983 −126.983 −178.983 80.709

PEG–K2 HPO4 –H2 O [17]


1500 9.1 −85.983 −96.983 −127.983 54.709
4000 7.2 −519.983 −89.983 −129.983 57.009
4000 9.1 −500.983 −94.983 −130.115 56.709
4000 10.8 −460.983 −101.083 −125.083 62.709

PEG–K2 HPO4 –H2 O [28]


1000 – −887.993 −130.000 −1.001 139.730
2000 – −839.975 −48.950 −4.277 11.017
4000 – −880.030 −44.120 −1.002 0.999
6000 −896.405 −96.647 −1.265 0.998

PEG–(NH4 )2 SO4 –H2 O [15]


1000 – −394.975 −260.977 146.024 203.017
1540 – −11.938 20.061 146.061 101.054
2000 – −14.975 16.038 108.024 136.017
4000 – −24.975 76.024 105.024 72.017

Table 6 Table 8
The percent of average relative deviation (ARD%) of the calculated mole fraction from The overall average relative deviation for the K2 HPO4 –PEG–H2 O and
the experiment using MNRT-NRF model for PEG, salt and water. Na2 SO4 –PEG–H2 O systems using the UNIFAC-NRF and MNRTL-NRF models.

MW of PEG PEG Salt Water MW of PEG pH UNIFAC-NRF [21] MNRTL-NRF

PEG–Na2 SO4 –H2 O [17] PEG–Na2 SO4 –H2 O


1500 0.010 2.732 3.262 1500 4.7 4.300 1.316
4000 0.153 0.970 2.826 4000 9.2 4.240 2.002
4000 0.100 0.171 2.200 4000 5.3 5.800 4.123

PEG–K2 HPO4 –H2 O [17] PEG–K2 HPO4 –H2 O


1500 0.269 0.177 1.849 1500 9.1 2.920 0.764
4000 0.574 1.833 0.869 4000 7.2 2.840 1.090
4000 0.458 0.794 1.399 4000 9.1 5.100 1.552
4000 0.219 1.470 1.579 4000 10.8 2.370 1.089

PEG–K2 HPO4 –H2 O [28]


1000 5.815 2.160 0.172
2000 0.053 1.651 1.210 be noted that experimental data of LLE from different sources might
4000 0.012 1.597 3.180 induce more deviation. This may result from the difference in the
6000 9.909 0.298 1.280 characteristics of the polymer, such as polydispersity and distribu-
PEG–(NH4 )2 SO4 –H2 O [15] tion of molecular weight (Table 8).
1000 0.085 0.940 3.572 To present the accuracy of the model, a comparison of the tie
1540 0.104 1.691 2.162 lines for the systems were performed. Figs. 2–4 show the binodal
2000 0.118 3.242 2.866
curves and the tie lines for the systems PEG2000–(NH4 )2 SO4 –H2 O,
4000 0.140 1.465 2.645
PEG4000–Na2 SO4 –H2 O and PEG4000–K2 HPO4 –H2 O, respectively.

models. As shown at Table 7, the results of the MNRTL-NRF model


show a better accuracy, particularly at higher molecular weight of
PEG, in correlation of the experimental data of liquid–liquid equi-
libria. The missing deviations at Table 7 have not been reported for
the system PEG–K2 HPO4 –H2 O using the UNIQUAC model. It should

Table 7
The overall average relative deviation for the PEG–(NH4 )2 SO4 –H2 O and
PEG–K2 HPO4 –H2 O systems using the UNIQUAC, UNIFAC and MNRTL-NRF
models.

MW of PEG UNIQUAC [15] UNIFAC [16,28] MNRTL-NRF

PEG–(NH4 )2 SO4 –H2 O


1000 3.686 2.07 1.513
1540 2.793 1.85 1.319
2000 3.656 1.73 2.043
4000 2.906 – 1.417

PEG–K2 HPO4 –H2 O


1000 – 2.160 2.716
2000 – 2.890 0.971
4000 – 2.913 1.597
Fig. 2. The binodal curve and tie lines for the PEG2000–(NH4 )2 SO4 –H2 O system at
6000 – 4.486 3.829
25 ◦ C using MNRTL-NRF model, the symbols are experimental data [15].
A. Haghtalab, M. Joda / Fluid Phase Equilibria 278 (2009) 20–26 25

Fig. 6. The equilibrium phase diagram of the PEG–Na2 SO4 –H2 O system at different
Fig. 3. The binodal curve and tie lines for the PEG4000–Na2 SO4 –H2 O system at 25 ◦ C
pH, the solid lines and symbols show the calculated and experiment, respectively
and pH: 9.2 using NRTL-NRF model, the symbols are experimental data [17].
[17].

One can see the calculated tie lines are in excellent agreement
with the experimental data. The advantage of the present model
demonstrates a very good agreement between the calculated and
experiment and coincidence of feed composition at respected point
on the tie line between the top and bottom compositions.
Fig. 5 shows the phase diagram of the PEG–(NH4 )2 SO4 –H2 O sys-
tem and the effect of molecular weight of PEG on the ATPS. As one
can observe, at higher molecular weight of the PEG, the lower con-
centration of the polymer is required for phase separation. Fig. 6
shows the binodal curve of the PEG–Na2 SO4 –H2 O system and the
effect of pH on the phase diagram, so with increasing of pH, i.e., at
higher pH more concentration of the polymer is required for phase
separation. The points are the experiment and the curves represent
the calculated values using the MNRTL-NRF model. As a result, using
the binary interaction parameters from binary aqueous electrolyte
systems, only four adjustable parameters are needed to correlate
the phase equilibria in the ATPS of ternary aqueous-polymer–salt
systems.

Fig. 4. The binodal curve and tie lines for the PEG1500–K2 HPO4 –H2 O system at 25 ◦ C 5. Conclusion
and pH 9.1 using NRTL-NRF model, the symbols are experimental data [17].
At this study, a new version of local composition model, MNRTL-
NRF, based on area fraction has been developed for binaries of
electrolyte–solvent, polymer–solvent and ternary aqueous two-
phase polymer–salt systems. The extended model for binaries of
aqueous electrolyte and polymer systems has only two adjustable
parameters and four adjustable parameters were used to corre-
lation of aqueous ternary polymer–salt systems. The local area
fraction, expressed in terms of nonrandom factor, is used to rep-
resent the effect of short-range interactions in aqueous two-phase
polymer–salt systems. The Debye-Hückel theory based on Fowler-
Guggenheim was applied for long-range interactions of ionic
species. Calculation of the LLE for the systems PEG–Na2 SO4 –H2 O,
PEG–K2 HPO4 –H2 O and PEG–(NH4 )2 SO4 –H2 O was performed and
the effects of pH and molecular weight of polyethylene glycol (PEG)
were investigated. Also the tie lines with binodal curves for those
systems were constructed. In addition, the binodal curves and tie
lines at different molecular weight and pH for those systems were
obtained in which the estimated results are in very good agreement
with the experiment. As a result, the MNRTL-NRF model could be
considered as a correlation model that can be successfully applied
Fig. 5. The equilibrium phase diagram of the PEG–(NH4 )2 SO4 –H2 O system at differ-
for calculation of phase behavior of ATPS with very good accuracy.
ent molecular weight of PEG, the solid lines and symbols show the calculated and Finally, the results showed that the present model is more accurate
experiment, respectively [15]. than the UNIQUAC and UNIFAC-NRF models.
26 A. Haghtalab, M. Joda / Fluid Phase Equilibria 278 (2009) 20–26


List of symbols −zi2 A I
aij interaction parameter between molecule i and j ln iLR = √ (A.1)
1+b I
A Debye-Hückel constant
B distance of closest approach parameter where the subscript of i stands for the ion species, zi the absolute
D dielectric constant of water charge number of ionic species i and “I”, the ionic strength of the
di density of substance i mixture, is defined as
d density of a mixed solution 
n

gE molar excess Gibbs energy I= mi zi2 (A.2)


I ionic strength i=1
m molality
where mi is the molality of species i.
MW molecular weight
For neutral species, the contribution of the long-range forces to
n number of experimental points
the activity coefficient takes the form [13]:
qi surface parameter of molecule i
2AVm d

√ 1 √ 
ri volume parameter of molecule i LR
ln m = 3
1+b I− √ − 2 ln(1 + b I) (A.3)
R gas constant b 1+b I
RMSD root mean square deviation
where m stands for the neutral species, vm is the molar volume of
T absolute temperature
the species m, d is the density of the water. The constants of A and
V total volume of the solution
b are the Debye-Hückel parameters that were given as [13]:
v̄j partial molar volume of component j
x mole fraction A = 1 · 327757 × 105 × d0.5 /(D · T )1.5
(A.4)
Z coordination number b = 6 · 359696 × d0.5 × d0.5 /(D · T )0.5
ZE absolute value of ionic charge
where D is the dielectric constant of water that its value is 78.3.
Greek letters
References
 nonrandom factor
i activity coefficient for component i [1] H. Walter, G. Johansson, Method Enzymol, vol. 228, Academic Press, London,
± mean ionic activity coefficient for electrolyte 1994.
ε error percent [2] M.W. Beijernick, Ueber une Engentumlichketit der Loslichen starke, Zentbl.
Bakteriol. 2 (1896) 627–697.
i surface area fraction [3] P.-A. Albertsson, Partition of cell Particles and Macromolecules, 3rd ed., Wiley
i effective surface area fraction Interscience, New York, 1986.
 ij local area fraction [4] E. Anderson, L. Persson, P. Albertson, B. Hahn-Hagerdal, Biotechnol. Bioeng. 27
(1985) 1036–1043.
ij effective local area fraction
[5] G. Birkenmeier, E. Usbeck, G. Kopperschlager, Anal. Biochem. 136 (1984)
1, 2 objective function 264–271.
 Boltzmann factor [6] R.S. King, H.W. Blanch, J.M. Prausnitz, AIChE J. 34 (1988) 1585–1593.
[7] A.P. Sassi, J. Wu, H.W. Blanch, J.M. Prausnitz, Polymer 37 (1996) 2151–2164.
[8] Y.-T. Wu, D.-Q. Lin, Z.-Q. Zhu, L.-H. Mei, Fluid Phase Equilibr. 124 (1996) 67–79.
Superscripts [9] M. Kabiri-Badr, H. Cabezas Jr., Fluid Phase Equilibr. 115 (1996) 39–58.
[10] C.A. Haynes, H.W. Blanch, J.M. Prausnitz, Fluid Phase Equilibr. 53 (1989)
C combinatorial
463–474.
cal calculation [11] C.A. Haynes, F.J. Benitez, H.W. Blanch, J.M. Prausnitz, AIChE J. 39 (1993)
exp experimental 1539–1557.
LR Long-range [12] Y.T. Wu, Z.-Q. Zhu, D.-Q. Lin, L.-H. Mei, Fluid Phase Equilibr. 121 (1996) 125–139.
[13] Y.-T. Wu, D.-Q. Lin, Z.-Q. Zhu, Fluid Phase Equilibr. 147 (1998) 25–43.
R Residual [14] Y.-T. Wu, Z. Qiangzhu, D.-Q. Lin, Mianli, Fluid Phase Equilibr. 154 (1999) 109–122.
I, II two aqueous phase in equilibria [15] Y.-L. Gao, Q.-H. Peng, Z.-C. Li, Y.-G. Li, Fluid Phase Equilibr. 63 (1991) 157–171.
* unsymmetrical convention [16] Y.-L. Gao, Q.-H. Peng, Z.-C. Li, Y.-G. Li, Fluid Phase Equilibr. 63 (1991) 173–182.
[17] A. Haghtalab, B. Mokhtarani, Fluid Phase Equilibr. 215 (2004) 151–161.
∞ infinite dilution [18] A. Haghtalab, B. Mokhtarani, Fluid Phase Equilibr. 180 (2001) 139–149.
[19] A. Haghtalab, J.H. Vera, AIChE J. 34 (1988) 803–813.
[20] A. Haghtalab, M.A. Asadollahi, Fluid Phase Equilibr. 171 (2000) 77–90.
Subscripts
[21] A. Haghtalab, B. Mokhtarani, J. Chem. Thermodyn. 37 (3) (2005) 289–295.
i, j, k component index [22] A. Haghtalab, B. Mokhtarani, G. Maurer, J. Chem. Eng. Data 48 (5) (2003)
m neutral species 1170–1177.
s salt [23] G.M. Wilson, J. Am. Chem. Soc. 86 (1964) 127–130.
[24] A. Haghtalab, M. Joda, Int. J. Thermophys. 28 (3) (2007) 876–890.
± mean ionic [25] R.A. Robison, R.H. Stokes, Electrolyte Solutions, 2nd ed., Butterworths, London,
1, 2, 3, 4 anion, cation, polyethylene glycol and water 1959.
[26] A. Bondi, Physical Properties of Molecular Crystals Liquids, and Glasses, John
Wiley & Sons, Inc., 1967.
Appendix A [27] Y. Marcus, Ion Properties, Marcel Dekker, Inc., 1997.
[28] Q. Peng, Z.-C. Li, Y.-G. Li, Fluid Phase Equilibr. 95 (1994) 341–357.
[29] L. Ninni, M.S. Camargo, A.J.A. Meirelles, Thermochim. Acta 328 (1999) 169–176.
Using Fowler and Guggenheim [30], the contribution of long- [30] R.H. Fowler, E.A. Guggenheim, Statistical Thermodynamics, Cambridge Univ.
range electrostatic interactions for ionic species is written as Press, 1949.

View publication stats

You might also like