handbook of surface and colloid chemistry fourth edition

49
This article was downloaded by: 10.3.98.104 On: 13 Jan 2022 Access details: subscription number Publisher: CRC Press Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: 5 Howick Place, London SW1P 1WG, UK Handbook of Surface and Colloid Chemistry Fourth Edition K.S. Birdi Thermodynamics of Polymer Solutions Publication details https://www.routledgehandbooks.com/doi/10.1201/b18633-4 M. Kontogeorgis Georgios, von Solms Nicolas Published online on: 25 Jun 2015 How to cite :- M. Kontogeorgis Georgios, von Solms Nicolas. 25 Jun 2015, Thermodynamics of Polymer Solutions from: Handbook of Surface and Colloid Chemistry Fourth Edition CRC Press Accessed on: 13 Jan 2022 https://www.routledgehandbooks.com/doi/10.1201/b18633-4 PLEASE SCROLL DOWN FOR DOCUMENT Full terms and conditions of use: https://www.routledgehandbooks.com/legal-notices/terms This Document PDF may be used for research, teaching and private study purposes. Any substantial or systematic reproductions, re-distribution, re-selling, loan or sub-licensing, systematic supply or distribution in any form to anyone is expressly forbidden. The publisher does not give any warranty express or implied or make any representation that the contents will be complete or accurate or up to date. The publisher shall not be liable for an loss, actions, claims, proceedings, demand or costs or damages whatsoever or howsoever caused arising directly or indirectly in connection with or arising out of the use of this material.

Upload: others

Post on 14-Jan-2022

16 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Handbook of Surface and Colloid Chemistry Fourth Edition

This article was downloaded by: 10.3.98.104On: 13 Jan 2022Access details: subscription numberPublisher: CRC PressInforma Ltd Registered in England and Wales Registered Number: 1072954 Registered office: 5 Howick Place, London SW1P 1WG, UK

Handbook of Surface and Colloid Chemistry Fourth Edition

K.S. Birdi

Thermodynamics of Polymer Solutions

Publication detailshttps://www.routledgehandbooks.com/doi/10.1201/b18633-4

M. Kontogeorgis Georgios, von Solms NicolasPublished online on: 25 Jun 2015

How to cite :- M. Kontogeorgis Georgios, von Solms Nicolas. 25 Jun 2015, Thermodynamics ofPolymer Solutions from: Handbook of Surface and Colloid Chemistry Fourth Edition CRC PressAccessed on: 13 Jan 2022https://www.routledgehandbooks.com/doi/10.1201/b18633-4

PLEASE SCROLL DOWN FOR DOCUMENT

Full terms and conditions of use: https://www.routledgehandbooks.com/legal-notices/terms

This Document PDF may be used for research, teaching and private study purposes. Any substantial or systematic reproductions,re-distribution, re-selling, loan or sub-licensing, systematic supply or distribution in any form to anyone is expressly forbidden.

The publisher does not give any warranty express or implied or make any representation that the contents will be complete oraccurate or up to date. The publisher shall not be liable for an loss, actions, claims, proceedings, demand or costs or damageswhatsoever or howsoever caused arising directly or indirectly in connection with or arising out of the use of this material.

Page 2: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

199

3 Thermodynamics of Polymer Solutions

Georgios M. Kontogeorgis and Nicolas von Solms

3.1 INTRODUCTION

Knowledge of phase behavior (thermodynamics) of polymer solutions is important for the design of many processes and products, including many specific applications in colloid and surface chemistry.

Among the many applications, we can mention the following:

1. The design of separations for the removal of unreacted monomers, colorants, by-products, and additives after solution or emulsion polymerizations1

CONTENTS

3.1 Introduction ....................................................................................................................... 1993.2 Choice of Solvents .............................................................................................................200

3.2.1 Rules of Thumb Based on Solubility Parameters .................................................. 2013.2.2 Rule of Thumb Based on the Infinite Dilution Activity Coefficient .....................2023.2.3 Rule of Thumb Based on the Flory–Huggins Model ............................................203

3.3 Free-Volume Activity Coefficient Models .........................................................................2063.3.1 The Free-Volume Concept .....................................................................................2063.3.2 Entropic-FV Model ...............................................................................................2073.3.3 Results and Discussion ..........................................................................................2083.3.4 Correlative Versions of the Entropic-FV model .................................................... 213

3.3.4.1 Entropic-FV/UNIQUAC Model .............................................................. 2133.3.4.2 Extension of Free-Volume Models to Gas Solubilities in Elastomers .... 214

3.4 SAFT Family of Equations of State .................................................................................. 2153.4.1 Introduction ...........................................................................................................2153.4.2 SAFT Equation of State—History and Introduction .............................................2173.4.3 Original SAFT .......................................................................................................2173.4.4 “Original” SAFT ...................................................................................................2203.4.5 Simplified SAFT ....................................................................................................2233.4.6 LJ-SAFT ................................................................................................................2233.4.7 SAFT-VR ...............................................................................................................2253.4.8 PC-SAFT ...............................................................................................................2263.4.9 Simplified PC-SAFT .............................................................................................2273.4.10 Extensions to SAFT-Type Equations .....................................................................2293.4.11 Applications of SAFT ............................................................................................230

3.5 Concluding Remarks and Future Challenges .................................................................... 2393.A Appendix ...........................................................................................................................2403.A.1 An Expression of the FH Model for Multicomponent Mixtures .......................................240Abbreviations ................................................................................................................................. 241References ......................................................................................................................................242

Page 3: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

200 Handbook of Surface and Colloid Chemistry

2. Selection of appropriate mixed solvents for paints and coatings, which can meet strict pro-duction and environmental criteria (fewer VOCs, water-based products)2,3

3. Numerous applications in the coatings industry, for example, in the control of emissions during production and application, the effect of temperature, and swelling of film or sorp-tion of gases and chemicals from the atmosphere3–6

4. Novel recycling methods for polymer waste based on physicochemical methods, the so-called selective dissolution7,8

5. Design of advanced materials based on compatible polymer blends9 or novel structures, for example, star or hyberbranched polymers10,11

6. Permeabilities of gases in the flexible polymeric pipelines used, for example, in the North Sea for transporting hydrocarbons from the seabed to the surface12,13

7. Use of CO2 as user-friendly solvent for many polymer-related applications including those involving paints and coatings14–16

8. Simultaneous representation of bulk and surface thermodynamic properties17

9. Migration of plasticizers from PVC18–21

10. Separation of proteins based on the so-called aqueous two-phase systems using polymers like PEG or dextran1,22,23

11. Choice of suitable solvents for polymers and especially copolymers used as stabilizers in colloidal dispersions

The aforementioned list shows some of the many applications where polymer thermodynamics plays a key role. Polymer solutions and blends are complex systems: frequent existence of liquid–liquid equilibria (LLE) (UCST, LCST, closed loop, etc.), the significant effect of temperature and polymer molecular weight including polydispersity in phase equilibria, free-volume (FV) effects, and other factors may cause difficulties. For this reason, many different models have been developed for poly-mer systems and often the situation may seem rather confusing to the practicing engineer.

The choice of a suitable model will depend on the actual problem and depends, specifically, on

• Type of mixture (solution or blend, binary or multicomponent, etc.)• Type of phase equilibria (VLE, LLE, SLLE, gas solubility, etc.)• Conditions (temperature, pressure, concentration)• Type of calculations (accuracy, speed, yes/no answer or complete design, etc.)

This chapter will present key tools in polymer thermodynamics at three different levels from the “simpler” but also “more easy-to-use” methods up to the “more advanced” but also “more complex” and potentially more accurate approaches:

1. Rules of thumb for choosing solvents including a discussion of the Flory–Huggins (FH) approach (Section 3.2)

2. FV activity coefficient models based on UNIFAC (Section 3.3), which are often predictive, and have found widespread applications at low pressures

3. The advanced statistical-associating fluid theory (SAFT) (Section 3.4), which is a theoretically based equation of state, which can be rigorously applied, unlike the activity coefficient models, at both low and high pressures

3.2 CHOICE OF SOLVENTS

A summary of some rules of thumb for predicting polymer–solvent miscibility, with focus on the screening of solvents for polymers, is presented here. These rules are based on well-known concepts of thermodynamics (activity coefficients, solubility parameters) and some specific ones for polymers

Page 4: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

201Thermodynamics of Polymer Solutions

(the FH parameter). It can be roughly stated that a chemical (1) will be a good solvent for a specific polymer (2), or in other words, the two compounds will be miscible if one (or more) of the following “rules of thumb” are valid:

1. If the polymer and the solvent have similar hydrogen bonding degrees:

δ δ1 2

3 1 21 8− ≤ . ( ) /cal/cm (3.1)

where δ is the solubility parameter. 2. If the polymer and the solvent have very different hydrogen bonding degrees:

4 1 2

21 2

21 2

2( ) ( ) ( )δ δ δ δ δ δd d p p h h R− + − + − ≤ (3.2)

where R is the Hansen-solubility parameter sphere radius and the subscripts d, p, h in the solubility parameter denote the dispersion, polar, and hydrogen bonding contributions, respectively.

3. Ω1 6∞ ≤ (the lower the infinite dilution activity coefficient of the solvent, the greater is the miscibility of a chemical). Values of the infinite dilution activity coefficient above 10 indicate nonmiscibility. In the intermediate region, it is difficult to conclude if the specific chemical is a solvent or a nonsolvent.

4. χ12 ≤ 0.5 (the lower the FH parameter value, the greater is the miscibility). Values much above 0.5 indicate nonmiscibility.

3.2.1 Rules of Thumb based on solubiliTy PaRameTeRs

They are widely used. The starting point (in their derivation and understanding) is the equation for the Gibbs free energy of mixing:

Δ Δ ΔG H T Smix mix mix= − (3.3)

A negative value implies that a solvent–polymer system forms a homogeneous solution, that is, the two components are miscible. Since the contribution of the entropic term (−TΔS) is always negative, it is the heat of mixing term that determines the sign of the Gibbs energy. The heat of mixing can be estimated from various theories, for example, the Hildebrand regular solution theory for nonpolar systems, which is based on the concept of the solubility parameter. For a binary solvent (1)–polymer (2) system, according to the regular solution theory,

ΔH Vmix = −ϕ ϕ δ δ1 2 1 22( ) (3.4)

where φi is the so-called volume fraction of component i. This is defined via the mole fractions xi and the molar volumes Vi as (for binary systems):

ϕi

i i

i i j j

x VxV x V

=+

(3.5)

According to Equation 3.4, the heat of mixing is always positive. For some systems with specific interactions (hydrogen bonding), the heat of mixing can be negative and Equation 3.4 does not hold.

Page 5: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

202 Handbook of Surface and Colloid Chemistry

Thus, the regular solution theory is strictly valid for nonpolar–slightly polar systems, without any specific interactions.

According to Equations 3.3 and 3.4, if solvent and polymer have the same solubility parameters, the heat of mixing is zero and they are therefore miscible at all proportions. The lower the solubility parameter difference, the larger the tendency for miscibility. Many empirical rules of thumb have been proposed based on this observation. Seymour24 suggests that if the difference of solubility parameters is below 1.8 (cal/cm3)1/2, then polymer and solvent are miscible (Equation 3.1).

Similar rules can be applied for mixed solvent–polymer systems, which are very important in many practical applications, for example, in the paints and coatings industry and for the separation of biomolecules using aqueous two-phase systems.

The solubility parameter of a mixed solvent is given by the equation:

δ ϕ δ=∑ i i

i

(3.6)

Barton25,26 provides empirical methods based on solubility parameters for ternary solvent systems.Charles Hansen introduced the concept of 3D solubility parameters, which offers an exten-

sion of the regular solution theory to polar and hydrogen bonding systems. Hansen observed that when the solubility parameter increments of the solvents and polymers are plotted in 3D plots, then the “good” solvents lie approximately within a sphere of radius R (with the poly-mer being in the center). This can be mathematically expressed as shown in Equation 3.2. The quantity under the square root is the distance between the solvent and the polymer. Hansen found empirically that a universal value 4 should be added as a factor in the dispersion term to approximately attain the shape of a sphere. This universal factor has been confirmed by many experiments. Hansen27 in his book provides a review of the method together with extensive tables of parameters.

The Hansen method is very valuable. It has found widespread use particularly in the paints and coatings industry, where the choice of solvents to meet economical, ecological, and safety constraints is of critical importance. It can explain some cases in which polymer and solvent solu-bility parameters are almost perfectly matched and yet the polymer will not dissolve. The Hansen method can also predict cases where two nonsolvents can be mixed to form a solvent. Still, the method is approximate; it lacks the generality of a full thermodynamic model for assessing mis-cibility and requires some experimental measurements. The determination of R is typically based on visual observation of solubility (or not) of 0.5 g polymer in 5 cm3 solvent at room temperature. Given the concentration and the temperature dependence of phase boundaries, such determination may seem a bit arbitrary. Still the method works out pretty well in practice, probably because the liquid–liquid boundaries for most polymer–solvent systems are fairly “flat.”

3.2.2 Rule of Thumb based on The infiniTe diluTion acTiviTy coefficienT

Since in several practical cases concerning polymer–solvent systems, the “solvent” is only present in very small (trace) amounts, the so-called infinite dilution activity coefficients are important. On a molar and weight basis, they are defined as follows:

γ γ

γγ

i i

ii i

i

pol

solv

xi

wi

xw

MM

∞ ∞

=

=⎛

⎝⎜

⎠⎟ =

lim

lim

0

01Ω

(3.7)

Page 6: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

203Thermodynamics of Polymer Solutions

The latter part of the equation is valid for a binary solvent–polymer solution and γ1∞ is the infinite

dilution activity coefficient of the solvent.The weight-based infinite dilution activity coefficient, Ω1

∞, which can be determined experimen-tally from chromatography, is a very useful quantity for determining good solvents. As mentioned in the previous section, low values (typically below 6) indicate good solvents, while high values (typically above 10) indicate poor solvents according to rules of thumb discussed by several investi-gators.28,29 The derivation of this rule of thumb is based on the FH model.

This method for solvent selection is particularly useful, because it avoids the need for direct liquid–liquid measurements and it makes use of the existing databases of solvent infinite dilution activity coefficients, which is quite large (e.g., the DECHEMA and DIPPR databases30,31).

Moreover, in the absence of experimental data, existing thermodynamic models (such as the FH, the Entropic-FV, and the UNIFAC-FV discussed later) can be used to predict the infinite dilution activity coefficient. Since, in the typical case today, existing models perform much better for VLE and activity coefficient calculations than directly for LLE calculations, this method is quite valuable and successful, as shown by sample results in Table 3.1.

This rule of thumb makes use of either experimental or predicted, by a model, infinite dilution activity coefficients. However, the results depend not only on the accuracy of the model, but also on the rule of thumb, which in turns depends on the assumptions of the FH approach. A thermo-dynamically more correct method is to employ the activity–concentration diagram. The maximum indicates phase split, while a monotonic increase of activity with concentration indicates a single liquid phase (homogeneous solutions).

3.2.3 Rule of Thumb based on The floRy–huggins model

The FH model for the activity coefficient, proposed in the early 1940s by Flory and Huggins,32,33 is a famous Gibbs free energy expression for polymer solutions. For binary solvent–polymer solu-tions and assuming that the parameter of the model, the so-called FH interaction parameter χ12, is constant, the activity coefficient is given by the equation:

ln ln lnγ

ϕ ϕχ ϕ

ϕϕ χ ϕ1

1

1

1

112 2

2 1

12 12 2

21 1 1= + − + = + −

⎝⎜

⎠⎟ +

x x x r (3.8)

whereφi can be volume or segment fractionsr is the ratio of the polymer volume to the solvent volume V2/V1 (approximately equal to the

degree of polymerization)

Appendix 3.A presents the general expression for the FH model suitable for multicomponent systems.

Using standard thermodynamics and Equation 3.8, it can be shown that for high molecular weight polymer–solvent systems, the polymer critical concentration is close to zero and the interac-tion parameter has a value equal to 0.5. Thus, a good solvent (polymer soluble in the solvent at all proportions) is obtained if χ12 ≤ 0.5, while values greater than 0.5 indicate poor miscibility. Since the FH model is only an approximate representation of the physical picture and particularly the FH parameter is often not a constant at all, this empirical rule is certainly subject to some uncertainty. Nevertheless, it has found widespread use and its conclusions are often in good agreement with experiment.

Page 7: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

204 Handbook of Surface and Colloid Chemistry

TABLE 3.1Choice of Suitable Solvents Using the ΩΩ∞∞

1 -Rule of Thumb for PBMA Systems (See Section 3.2)

Solvent S/NS EFV UFV GCFL

Hexane NS 7.1 7.0 10.7

n-Octane NS 6.7 6.3 10.4

n-Decane NS 6.5 6.0 10.7

n-Dodecane NS 6.6 6.0 11.3

n-Hexadecane NS 6.8 6.1 13.2

Toluene S 3.2 4.4 4.7

Xylene S 2.3 3.6 5.7

Methylene dichloride S 3.3 2.5 3.0

Chloroform S 1.9 2.1 1.7

Carbon tetrachloride S 2.2 2.2 2.9

Ethylene dichloride S 3.5 3.0 —

Trichloroethylene S 2.5 2.9 33.9

Chlorobenzene S 2.5 3.0 3.0

o-Dichlorobenzene S 1.3 2.5 2.7

Acetone S 10.9 14.2 11.1

MEK S 8.4 10.5 8.2

MIBK S 6.3 7.7 5.7

Acetophenone S 8.1 9.3 8.6

Ethyl acetate S 6.7 6.7 60.3

Butyl acetate S 5.3 5.1 31.4

Diethyl ether S 5.2 5.8 11.6

THF S 3.8 4.0 —

1,4-Dioxane S 4.1 4.4 159.4

Methanol NS 43.7 57.7 35.7

Ethanol NS 29.2 31.3 17.3

1-Butanol NS 18.1 17.1 8.1

Cyclohexanol NS 24.3 20.1 3.0

Ethylene glycol NS 277.8 — 15,947.0

Propylene glycol NS 212.6 — 1,879.2

1,3-Butanediol NS 158.5 — 525.5

Glycerol NS 294.6 — 2,282.4

Isopropanol NS 23.4 21.6 10.6

Isobutanol S 19.0 17.9 7.9

Diethylene glycol NS 240.1 — 2,470.4

Dipropylene glycol NS 127.9 945.7 287.8

Nitromethane NS 16.7 17.2 —

1-Nitropropane S 4.7 5.2 —

N,N-dimethylformamide S 3.8 — —

Source: Modified from Lindvig, Th. et al., AIChE J., 47(11), 2573, 2001.Notes: EFV, Entropic-FV; UFV, UNIFAC-FV; GCFL, GC-Flory; S, solvent; NS, nonsolvent (experimental observation).

Page 8: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

205Thermodynamics of Polymer Solutions

There are several, still rather obscure issues about the FH model, which we summarize here together with some recent developments:

1. The FH parameter is typically not a constant and should be estimated from experimental data. Usually, it varies with both temperature and concentration, which renders the FH model useful only for describing experimental data. It cannot be used for predicting phase equilibria for systems for which no data is available. Moreover, when fitted to the critical solution temperature, the FH model cannot yield a good representation of the whole shape of the miscibility curve with a single parameter.

2. Accurate representation of miscibility curves is possible with the FH model using suitable (rather complex) equations for the temperature and the concentration dependence of the FH parameter.34

3. In some cases, a reasonable value of the FH parameter can be estimated using solubility parameters via the equation:

χ χ χ δ δ12

11 2

20 35= + = + −s hVRT

. ( ) (3.9)

Equation 3.9, without the empirical 0.35 factor, is derived from the regular solution the-ory. The constant 0.35 is added for correcting for the deficiencies of the FH combina-torial term. These deficiencies become evident when comparing experimental data for athermal polymer and other asymmetric solutions to the results obtained with the FH model. A systematic underestimation of the data is observed, as discussed extensively in the literature,28,35 which is often attributed to the inability of the FH model in accounting for the FV differences between polymers and solvents or between compounds differing significantly in size such as n-alkanes with very different chain lengths. The term, which contains the “0.35 factor,” corrects in an empirical way for these FV effects. However, and although satisfactory results are obtained in some cases, we cannot generally rec-ommend Equation 3.9 for estimating the FH parameter. Moreover, for many nonpolar systems with compounds having similar solubility parameters, the empirical factor 0.35 should be dropped.

4. Lindvig et  al.36 proposed an extension of the FH equation using the Hansen solubility parameters for estimating activity coefficients of complex polymer solutions.

ln ln

. ( ) . (

γϕ ϕ

χ ϕ

χ δ δ δ δ

11

1

1

112 2

2

121

1 22

1

1

0 6 0 25

= + − +

= − + −

x x

VRT d d p p22

21 2

20 25) . ( )+ −⎡⎣ ⎤⎦δ δh h

(3.10)

In order to achieve that, Lindvig et al.36 have, as shown in Equation 3.10, employed a universal correction parameter, which has been estimated from a large number of polymer–solvent VLE data. Very good results are obtained, especially when the vol-ume-based combinatorial term of FH is employed, also for ternary polymer–solvent systems.37

5. Based on the FH model, several techniques have been proposed for interpreting and for cor-relating experimental data for polymer systems, for example, the so-called Schultz–Flory

Page 9: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

206 Handbook of Surface and Colloid Chemistry

(SF) plot. Schultz and Flory38 have developed, starting from the FH model, the following expression, which relates the critical solution temperature (CST), with the theta tempera-ture and the polymer molecular weight:

1 1 1 1 1 12CST

= + +⎛

⎝⎜

⎠⎟

⎣⎢

⎦⎥Θ ψ r r

(3.11)

where ψ = ((1/2)−χs) is the entropic parameter of the FH model (Equation 3.9)

r is the ratio of molar volumes of the polymer to the solvent

This parameter is evidently dependent on the polymer’s molecular weight. The SF plot can be used for correlating data of critical solution temperatures for the same polymer–solvent system, but at different polymer molecular weights. This can be done, as anticipated from Equation 3.11, because the plot of 1/CST against the quantity in parentheses in Equation 3.11 is linear. The SF plot can also be used for predicting CST for the same system but at different molecular weights than those used for correlation as well as for calculating the theta temperature and the entropic part of the FH parameter. It can be used for correlating CST/molecular weight data for both the UCST and LCST areas. Apparently different coef-ficients are needed.

3.3 FREE-VOLUME ACTIVITY COEFFICIENT MODELS

3.3.1 The fRee-volume concePT

The FH model provides a first approximation for polymer solutions. Both the combinatorial and the energetic terms require substantial improvement. Many authors have replaced the random van-Laar energetic term by a nonrandom local-composition term such as those of the UNIQUAC, NRTL, and UNIFAC models. The combinatorial term should be modified to account for the FV differences between solvents and polymers.

The improvement of the energetic term of FH equation is important. Local-composition terms like those appearing in NRTL, UNIQUAC, and UNIFAC provide a flexibility, which cannot be accounted for by the single-parameter van Laar term of FH. However, the highly pronounced FV effects should always be accounted for in polymer solutions.

The concept of FV is rather loose, but still very important. Elbro28 demonstrated, using a simple definition for the FV (Equation 3.12), that the FV percentages of solvents and polymers are differ-ent. In the typical case, the FV percentage of solvents is greater (40%–50%) than that of polymers (30%–40%). There are two notable exceptions to this rule; water and PDMS: water has lower FV than other solvents and closer to that of most of the polymers, while PDMS has quite a higher FV percentage, closer to that of most solvents. LCST is, as expected, related to the FV differences between polymers and solvents. As shown by Elbro,28 the larger the FV differences, the lower is the LCST value (the larger the area of immiscibility). For this reason, PDMS solutions have a LCST, which are located at very high temperatures.

Many mathematical expressions have been proposed for the FV. One of the simplest and suc-cessful equations is

V V V V Vf w= − = −* (3.12)

originally proposed by Bondi39 and later adopted by Elbro et  al.35 and Kontogeorgis et  al.40 in the so-called Entropic-FV model (described later). According to this equation, FV is just

Page 10: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

207Thermodynamics of Polymer Solutions

the “empty” volume available to the molecule when the molecules’ own (hard-core or closed-packed V*) volume is subtracted.

The FV is not the only concept, which is loosely defined in this discussion. The hard-core volume is also a quantity difficult to define and various approximations are available. Elbro et al.35 suggested using V* = Vw, that is, equal to the van der Waals volume (Vw), which is obtained from the group increments of Bondi and is tabulated for almost all existing groups in the UNIFAC tables. Other investigators41 interpreted somewhat differently the physical meaning of the hard-core volume in the development of improved FV expressions for polymer solutions, which employ Equation 3.12 as basis, but with V* values higher than Vw (about 1.2–1.3Vw).

3.3.2 enTRoPic-fv model

The original UNIFAC model does not account for the FV differences between solvents and polymers and, as a consequence of that, it highly underestimates the solvent activities in polymer solutions. One of the most successful and earliest such models for polymers is the UNIFAC-FV by Oishi and Prausnitz.42 The UNIFAC-FV model was developed for solvent activities in polymers, but it cannot be successfully applied to LLE.

A similar to UNIFAC-FV but somewhat simpler approach, which can be readily extended to multicomponent systems and LLE, is the so-called Entropic-FV model proposed by Elbro et al.35 and Kontogeorgis et al.40:

ln ln ln

ln ln

γ γ γ

γφ φ

φ

i icomb FV

ires

icomb FV i

FV

i

iFV

i

iFV

x x

= +

= + −

=

− 1

xxVx V

x V Vx V V

i i FV

j j FVi

i i wi

j j wjj

ires

,

,

( )( )

ln

∑ ∑=

→γ UNIFAC

(3.13)

As can been seen from Equation 3.13, the FV definition given by Equation 3.12 is employed. The combinatorial term of Equation 3.13 is very similar to that of FH. However, instead of volume or seg-ment fractions, FV fractions are used. In this way, both combinatorial and FV effects are  combined into a single expression. The combinatorial–FV expression of the Entropic-FV model is derived from statistical mechanics, using a suitable form of the generalized van der Waals partition function.

The residual term of Entropic-FV is taken by the so-called new or linear UNIFAC model, which uses a linear-dependent parameter table43:

a a a T Tmn mn mn o= + −, , ( )1 2 (3.14)

This parameter table has been developed using the combinatorial term of the original UNIFAC model. As with UNIFAC-FV, no parameter re-estimation has been performed. The same group parameters are used in the “linear UNIFAC” and in the Entropic-FV models.

A common feature for both UNIFAC-FV and Entropic-FV is that they require values for the volumes of solvents and polymers (at the different temperatures where application is required). This can be a problem in those cases where the densities are not available experimentally and have to be estimated using a predictive group-contribution or other method, for example, GCVOL44,45 or van Krevelen methods. These two estimation methods perform quite well and often similarly even for low molecular weight compounds or oligomers such as plasticizers.

Page 11: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

208 Handbook of Surface and Colloid Chemistry

Both UNIFAC-FV and Entropic-FV, especially the former, are rather sensitive to the density values used for the calculations of solvent activities.

As already mentioned, the Entropic-FV model has been derived from the van der Waals partition function. The similarity of the model with the van der Waals equation of state P = RT/(V − b) − a/V 2 becomes apparent if the van der Waals equation of state is written (when the classical van der Waals one fluid mixing and classical combining rules are used) as an activity coefficient model:

ln ln ln ln (γ γ γϕ ϕ

δ δi icomb FV

ires i

FV

i

iFV

i

iix x

VRT

= + = + −⎛

⎝⎜

⎠⎟ + −− 1 ii j

iFV i i i

j j jj

ii

i

x V bx V b

aV

)

( )( )

2 2ϕ

ϕ

δ

⎝⎜

⎠⎟

=−

=

∑ (3.15)

where φi is the volume fraction as defined in Equation 3.5. The first term in Equation 3.15 is the same as in Entropic-FV with Vw = b, while the latter term is a regular solution theory or van Laar-type term.

3.3.3 ResulTs and discussion

Table 3.2 presents an overview of the results with Entropic-FV model for different applications together with the corresponding references. Selected results are shown in Tables 3.3 through 3.9 and Figures 3.1 and 3.2.

The most important general conclusions can be summarized as following:

1. Satisfactory predictions are obtained for solvent activities, even at infinite dilution, and for nonpolar, as well as for complex polar and hydrogen bonding systems including solutions of interest to paints and coatings. Rather satisfactory predictions are also achieved when mixed solvents and copolymers are present.

TABLE 3.2Applications of the Entropic-FV Model

Application References

VLE binary solutions [40,41]

VLE complex polymers–solvents [48]

VLE ternary polymer–solvents [37]

Paints [61]

Dendrimers

VLE copolymers [47]

VLE athermals systems [28,35,50,62]

SLE hydrocarbons [51,52]

SLE polymer–solvents [49]

Comparison with other models [40,46,53,60]

LLE polymer–solvents [58]

LLE ternary polymer–mixed solvents [8]

Polymer blends [59]

EFV + UNIQUAC

Page 12: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

209Thermodynamics of Polymer Solutions

TABLE 3.3Prediction of the Solubility for Characteristic Polymer–Solvent Systems Using the ΩΩ∞∞

1 Rule of Thumb and Two FV Models for Solvent Selection

System Experiment Entropic-FV UNIFAC-FV

PBMA/nC10 NS 6.5 (–) 6.1 (–)

PBMA/xylene S 2.3 (S) 3.6 (S)

PBMA/CHCl3 S 1.9 (S) 9.1 (NS)

PBMA/acetone S 0.2 (NS) 14.1 (NS)

PBMA/ethyl acetate S 6.7 (–) 6.7 (–)

PBMA/ethanol NS 29.2 (NS) 31.3 (NS)

PMMA/acetone S 10.0 (NS) 16.5 (NS)

PMMA/ethyl acetate S 6.6 (–) 8.4 (NS)

PMMA/butanol NS 26.8 (NS) 14.4 (NS)

PEMA/MEK S 8.1 (NS) 11.7 (NS)

PEMA/diethyl ether S 5.8 (S) 7.6 (–)

PEMA/nitropropane NS 4.5 (S) 1.4 (S)

PVAc/hexane NS 38.7 (NS) 38.6 (NS)

PVAc/methanol S 18.9 (NS) 19.4 (NS)

PVAc/ethanol NS 15.2 (NS) 38.9 (NS)

PVAc/nitromethane S 3.9 (S) 3.8 (S)

PVAc/THF S 8.4 (NS) 5.6 (S)

Notes: S, good solvent; NS, nonsolvent; –, no answer according to the rule of thumb.

TABLE 3.4Prediction of Infinite Dilution Activity Coefficients for Polyisoprene (PIP) Systems with Two Predictive Group Contribution Models

PIP Systems Exper. Value Entropic-FV UNIFAC-FV

+Acetonitrile 68.6 47.7 (31%) 52.3 (24%)

+Acetic acid 37.9 33.5 (12%) 17.7 (53%)

+Cyclohexanone 7.32 5.4 (27%) 4.6 (38%)

+Acetone 17.3 15.9 (8%) 13.4 (23%)

+MEK 11.4 12.1 (6%) 10.1 (12%)

+Benzene 4.37 4.5 (2.5%) 4.4 (0%)

+1,2-Dichloroethane 4.25 5.5 (29%) 6.5 (54%)

+CCl4 1.77 2.1 (20%) 1.8 (0%)

+1,4-Dioxane 6.08 6.3 (4%) 5.9 (2%)

+Tetrahydrofurane 4.38 4.9 (14%) 3.9 (10%)

+Ethylacetate 7.47 7.3 (2%) 6.6 (11%)

+n-Hexane 6.36 5.1 (20%) 4.6 (27%)

+Chloroform 2.13 3.00 (41%) 2.6 (20%)

Experimental values and calculations are at 328.2 K.

Page 13: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

210 Handbook of Surface and Colloid Chemistry

TABLE 3.5Average Absolute Deviations between Experimental and Calculated Activity Coefficients of Paint-Related Polymer Solutions Using the FH/Hansen Method and Two FV Models

Model % AAD (Systems in

Database) % AAD

Araldite 488 % AAD Eponol-55

FH/Hansen, Volume (Equation 3.10) 22 31 28

Entropic-FV 35 34 30

UNIFAC-FV 39 119 62

Source: Adapted from Lindvig, Th. et al., Fluid Phase Equilib., 203, 247, 2002.The second column presents the systems used for optimization of the universal parameter (solutions containing acrylates and acetates). The last two columns show predictions for two epoxy resins.

TABLE 3.6Mean Percentage Deviations between Experimental and Calculated Activity Coefficients of Solvents in Various Nearly Athermal Solutions

% AAD Infinite Dilution Conditions γγ∞∞1 (ΩΩ∞∞

1 for Polymers) Entropic-FV UNIFAC-FV Flory-FV

Short n-alkanes/long alkanes 8 15 20

Short branched, cyclic alkanes/long alkanes 10 17 20

Alkanes/polyethylene 9 23 19

Alkanes/polyisobutylene 16 12 38

Organic solvent/PDMS, PS, PVAc 20 29 26

Overall 13 19 25

Source: Kouskoumvekaki, I. et al., Fluid Phase Equilib., 202(2), 325, 2002. With permission.

TABLE 3.7Mean Percentage Deviations between Experimental and Calculated Activity Coefficients of Heavy Alkanes Solutes in Alkane Solvents

Entropic-FV UNIFAC-FV Flory-FV

% AAD infinite dilution conditions, γ2∞

Symmetric long alkanes/short alkanes 36 47 10

Medium asymmetric long alkanes/short alkanes 34 48 12

Asymmetric long alkanes/short alkanes 44 54 37

Overall 38 50 20

% AAD finite concentrations, γ2

Symmetric long alkanes/short alkanes 14 17 6

Medium asymmetric long/short alkanes 23 31 11

Asymmetric long/short alkanes 40 55 16

Overall 26 34 11

Source: Kouskoumvekaki, I. et al., Fluid Phase Equilib., 202(2), 325, 2002. With permission.

Page 14: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

211Thermodynamics of Polymer Solutions

TABLE 3.8Average Absolute Logarithmic Percentage Deviations between Experimental and Predicted Equilibrium Pressures and Average Absolute Deviation (×100) between Calculated and Experimental Vapor Phase Compositions (Mole Fractions) for Various Ternary Polymer–Mixed Solvent Systems

Sys. No. Variable SAFT EFV/UQ FH Pa-Ve EFV/UN UFV GCFL GCLF FHHa

1 P 11 6 6 6 2 1 21 8 2

y 4 3 3 3 3 3 3 3 3

2 P 4 2 14 8 2 2 — 2 11

y 5 2 5 1 2 2 — 4 5

3 P — — — — 3 3 12 5 18

y — — — — 3 3 2 4 4

4 P — — — — 4 4 13 9 5

y — — — — 18 18 19 19 15

5 P 14 16 11 4 17 19 93 5 52

y 17 17 16 11 17 13 2 18 14

Sources: Katayama, T. et al., Kagaku Kogaku, 35, 1012, 1971; Matsumara, K. and Katayama, T., Kagaku Kogaku, 38, 388, 1974; Tanbonliong, J.O. and Prausnitz, J.M., Polymer, 38, 5775, 1997.

Based on the results shown by Lindvig et al.37

Notes: 1, PS–toluene–ethylbenzene at 303 K; 2, PS–toluene–cyclohexane at 303 K; 3, PVAc–acetone–ethyl acetate at 303 K; 4, PVAc–acetone–methanol at 303 K; 5, PS–chloroform–carbon tetrachloride at 323.15 K.

TABLE 3.9Prediction of Liquid–Liquid Equilibria for Polymer–Solvent Systems with Various Thermodynamic Models

Polymer System Molecular Weight Entropic-FV N0065w UNIFAC GC-Flory

PS/acetone 4,800 84 21 75

PS/acetone 10,300 98 8 42

PS/cyclohexane 20,400 38 62 —

PS/cyclohexane 37,000 26 59 —

PS/cyclohexane 43,600 24 63 —

PS/cyclohexane 89,000 11 60 —

PS/cyclohexane 100,000 15 62 —

PS/cyclopentane 97,200 27 105 —

PS/cyclopentane 200,000 12 103 —

HDPE/n-butyl acetate 13,600 10 82 72

HDPE/n-butyl acetate 20,000 22 97 70

HDPE/n-butyl acetate 61,100 29 107 71

PMMA/1-chloro butane 34,760 53 — —

PBMA/n-pentane 11,600 Hourglass — —

PBMA/n-octane 11,600 155 — —

Absolute difference between experimental and predicted UCST for several polymer solutions using various models (based on results from Ref. [58]). All results are predictions. The three last models are based on group contributions. The new UNIFAC model is a combination of FH with the UNIFAC residual term.Notes: PS, polystyrene; HDPE, high-density polyethylene; PMMA, poly(methyl methacrylate); PBMA, poly(butyl

methacrylate).

Page 15: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

212 Handbook of Surface and Colloid Chemistry

2. Qualitatively correct and occasionally also quantitatively satisfactory representation of LLE for binary and ternary polymer–solvent systems is achieved, especially for the UCST-type phase behavior. A single investigation for SLLE also shows good results.

3. Less satisfactory results are obtained for polymer blends, where FV effects are not as dominant as for polymer solutions.

Some additional observations for specific cases are hereafter presented:

1. Athermal solutions, for example, polyolefins with alkanes offer a way of testing FV terms and numerous such investigations have been presented. FV models perform generally bet-ter than those that do not contain volume-dependent terms.60 Better FV terms than that of

600.00

19,800

10,300

4,800

4,800

exp. dataexp. dataexp. data

19,80010,300

500.00

400.00

300.00

200.00Tem

pera

ture

(K)

100.00

0.000.00 0.20 0.40

Polymer weight fraction0.60 0.80

FIGURE 3.1 Experimental and predicted LLE diagram for the system polystyrene/acetone at three polymer molecular weights (4,800, 10,300, 19,800). The points are the experimental data and the lines are the predic-tions with the Entropic-FV model. (From Kontogeorgis, G.M. et al., Ind. Eng. Chem. Res., 34, 1823, 1995. With permission.)

Methyl cyclohexane

Acetone PS

Experimental dataEntropic-FV modelHA model

FIGURE 3.2 Ternary LLE for PS(300,000)/methyl cyclohexane/acetone, T = 298.15 K. (Reprinted from Pappa, G.D. et al., Ind. Eng. Chem. Res., 36, 5461, 1997. With permission.)

Page 16: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

213Thermodynamics of Polymer Solutions

Entropic-FV (Equation 3.13) have been proposed41,51,52,57,60 and they may serve as the basis for future developments resulting in even better activity coefficient models for polymer solutions. A rigorous test for newly developed FV expressions is provided by athermal alkane systems, especially the activity coefficients of heavy alkanes in short-chain ones.

2. A difference of 10°–30° should be expected in UCST predictions with Entropic-FV, see Table 3.9.

3. An alternative to Entropic-FV successful approach is the FH/Hansen model presented previously (Equation 3.10). In this case, best results are obtained when the original FH combinatorial rather than a FV term is used together with HSP. It seems that HSP incorpo-rate some FV effects.

3.3.4 coRRelaTive veRsions of The enTRoPic-fv model

Both UNIFAC-FV and Entropic-FV are group contribution models. This renders the models truly predictive, but at the same time with little flexibility if the performance of the models for specific cases is not satisfactory. Two interesting alternative approaches are discussed here, which still main-tain the FV terms but use different residual terms.

3.3.4.1 Entropic-FV/UNIQUAC ModelThe first approach is to employ the UNIQUAC expression for the residual term. This Entropic-FV/UNIQUAC model has been originally suggested by Elbro et al.35 and has shown to give very good results for polymer solutions if the parameters are obtained from VLE data between the solvent and the low molecular weight monomer (or the polymer’s repeating unit).

The Entropic-FV/UNIQUAC model has been recently further developed and extended indepen-dently by two research groups.55–57 Both VLE and LLE equilibria are considered, but the emphasis is given to LLE. Very satisfactory results are obtained as can be seen for two typical systems in Figures 3.3 and 3.4. It has been demonstrated that the Entropic-FV/UNIQUAC approach can

440

420

400

380

360

T (K

)

340

320

3000.00 0.04 0.08 0.12

Weight fraction of polymer0.16 0.20

FIGURE 3.3 Correlation of liquid–liquid equilibria for the PVAL/water system with the Entropic-FV/UNIQUAC model. (•) Exp. data (Mn = 140,000 g/mol); (—) correlation. (From Bogdanic, G. and Vidal, J., Fluid Phase Equilib., 173, 241, 2000. With permission.)

Page 17: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

214 Handbook of Surface and Colloid Chemistry

correlate both UCST/LCST and closed-loop behavior and even show the pressure dependency of critical solution temperatures (UCST and LCST).

3.3.4.2 Extension of Free-Volume Models to Gas Solubilities in ElastomersThe second approach proposed by Thorlaksen et al.63 is based on a combination of the Entropic-FV term with Hildebrand’s regular solution theory and developed a model for estimating gas solubilities in elastomers. The so-called Hildebrand–Entropic-FV model is given by the equation:

ln ln lnγ γ γi iR

iC FV= + + (3.16)

ln ( )

ln ln

γδ δ

γ

22 1 2

212

22

2

2

21

RL

C FVFV FV

VR T

x x

=⋅ − ⋅

= + −+

Φ

Φ Φ (3.17)

whereδ1 is the solvent solubility parameterδ2 is the gas solubility parameterx2 is the gas mole fraction in liquid/polymerΦ2 is the “apparent” volume fraction of solvent, given by

Φ2

2 2

1 1 2 2=

⋅⋅ + ⋅x V

x V x V

L

L L

475

450

425

400

375

350

T (K

)

325

300

275

250

2250.00 0.05 0.10 0.15

Weight fraction of polymer0.20 0.25 0.30 0.35

FIGURE 3.4 Correlation and prediction of liquid–liquid equilibria for the PBD/1-octane system with the Entropic-FV/UNIQUAC model. (◾) Exp. data (Mv = 65,000 g/mol), (—) correlation; (▴) Exp. data (Mv = 135,000 g/mol), (⎯ ⎯) prediction; and (⦁) Exp. data (Mw = 44,500 g/mol), (- - - -) prediction. (From Bogdanic, G. and Vidal, J., Fluid Phase Equilib., 173, 241, 2000. With permission.)

Page 18: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

215Thermodynamics of Polymer Solutions

Φ2FV

is the “FV” fraction given by

Φ22 2 2

2 2 2 1 1 1

FVL VW

L VW L VW

x V Vx V V x V V

=⋅ −( )

⋅ −( ) + ⋅ −( )V L

2 is a hypothetical liquid volume of the (gaseous) solute.

12 2

2 1 22 2

1

xff

VR T

l

g

L= ⋅

⋅ − ⋅⋅

2ˆ exp ( )δ δ Φ

(3.18)

f̂ g2 is the fugacity of the gasf l2 is the fugacity of the hypothetical liquid, which can be estimated from the equation:

ln . . . . .fP T

T T Tl

c rr r

2 23 54811 4 74547 1 60151 0 87466 0 10971= − + ⋅ − ⋅ + ⋅ rr3 (3.19)

Finally, the gas solubility in the polymer is estimated from the equation:

1 12

2

2

2 1 22

12

2

2

2

2xff

VR T x x

l

g

L FV FV= ⋅

⋅ −( ) ⋅⋅

+ + −⎛

⎜⎜

⎠ˆ exp ln

δ δ Φ Φ Φ ⎟⎟⎟

(3.20)

Calculations showed that the hypothetical gas “liquid” volumes are largely independent to the  polymer used, and moreover, for many gases (H2O, O2, N2, CO2 and C2H2), these are related to the critical volume of the gas by the equation:

V VLc2 1 776 86 017= −. . (3.21)

Very satisfactory results are obtained as shown in Table 3.10.A final note for these “classical” activity coefficient models is that, despite the advent of advanced

SAFT and other equations of state discussed next (Section 3.4), they are still quite popular and widely used in practical applications. They are also well cited in literature. For example, the histori-cal articles by Flory and Huggins (Refs. [32,33]) are cited 998 (13.5) and 1034 (14) and the citations of the articles by Elbro et al.35: 164 (6.6), Lindvig et al.36: 44 (3.4), Kontogeorgis et al.40: 121 (5.5), and Oishi and Prausnitz42: 353 (9.5). The citations are per May 2014 and the numbers in parenthesis are citations per year.

3.4 SAFT FAMILY OF EQUATIONS OF STATE

3.4.1 inTRoducTion

“Statistical mechanics is that branch of physics which studies macroscopic systems from a microscopic or molecular point of view. The goal of statistical mechanics is the understanding and prediction of macroscopic phenomena and the calculation of macroscopic properties from the prop-erties of the individual molecules making up the system.”

This is the opening paragraph of “Statistical Mechanics” written by McQuarrie,64 already in 1976. Attempts to achieve this goal of statistical mechanics have been around for a long time. For example, Wertheim65 was the first to derive an equation of state for hard-sphere systems. Carnahan  and

Page 19: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

216 Handbook of Surface and Colloid Chemistry

Starling66 made an empirical modification to Wertheim’s solution based on molecular simulation data to arrive at what is by now the famous Carnahan–Starling equation of state. As well as being an early attempt to arrive at an engineering model using results from “hard” science (or what we might cynically call “impractical” science), this work also showed the usefulness of using results from molecular simulation. In a sense, molecular simulation fulfils the goal of statistical mechanics, in that it predicts (some) macroscopic properties using only molecular properties as input. However, molecular simulation is system specific, time consuming, and ultimately can only be as successful as the molecular model it is based upon. Thus, while it is certainly a step forward to be able to predict the properties of a system of hard spheres, the hard sphere as a model is itself an incomplete descrip-tion of a real molecule. Nevertheless, increasing computer power and ever more detailed knowledge of molecular properties, extending even to the quantum level, means that molecular simulation will continue to be an important tool in engineering thermodynamics. See Economou67 for a review of current industrial applications of molecular simulation.

However, equations of state, too, will be an essential component of chemical engineering theory and practice for the foreseeable future, and as ever, the balance will need to be struck between rigorous theory and engineering applicability. One equation of state, which seems to have done an admirable job of bridging the gap between molecular theory and engineering application, is statis-tical associating fluid theory (SAFT) and it is with this equation of state and its spin-offs that the remainder of this discussion is concerned.

TABLE 3.10Summary of the Performance of the Models Tested at T = 298 K; P = 101.3 kPa

Polymer Gas Michaels/Bixler Tseng/Lloyd Hildebrand/Scott Hildebrand

Entropic-FV-1 Hildebrand

Entropic-FV-2

PIP N2 14.7 73 3.9 −7.9 −4.6

O2 −16.1 −4 14 10.8 11.8

Ar −32.5 −23 — 29.4 −22.2

CO2 −3.2 13 4.5 8.7 4.6

PIB N2 −2.5 — 6.8 3.1 5.0

O2 −6.1 — −1.7 −8.3 1.7

Ar — — — — —

CO2 32.8 — −1.9 41.1 35.2

PBD N2 22.3 — 8.1 8.1 12.6

O2 14.9 — −6 8.7 10.8

Ar 12.1 — — 111.1 24.0

CO2 −9.7 — −4.6 0.4 −4.0

PDMB N2 — — −23 −7.5 −3.1

O2 — — −32 −16.8 −15.9

Ar — — — — —

CO2 — — −24 2.3 −2.2

PCP N2 58.1 — 49 −7.0 −4.2

O2 43.7 — 60 −1.4 −1.4

Ar — — — — —

CO2 8.8 — 27 −13.3 −17.1

AAD 19.8 28 18 16.8 10.6

Errors associated with models for predicting gas solubilities in polymers.Hildebrand Entropic-FV-1: The liquid volume of the gas is determined from its relationship with the critical volume, Equation 3.26. Hildebrand Entropic-FV-2: The average hypothetical liquid volume of a gas is used.

Page 20: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

217Thermodynamics of Polymer Solutions

3.4.2 safT equaTion of sTaTe—hisToRy and inTRoducTion

A series of four seminal papers once again written by Wertheim68–71 appeared in 1984 and 1986. These papers laid the foundation for the associating theory (or thermodynamic perturbation the-ory [TPP]), which was to become the key feature in the novel molecular-based equation of state known as SAFT. The theories presented in Wertheim’s papers are highly complex and almost intractable—“essentially incomprehensible,” as one author has put it.1 However, in the period 1988–1990 at Cornell University, Chapman and coworkers72–75 performed the monumental task of transforming the abstruse theory of Wertheim into workable equations, and finally into an engi-neering equation of state. SAFT is by no means the only molecular-based equation of equation of state out there—PHSC is another76 but it differs from the vast majority of other similar equations of state in one important respect—it is extensively used. In their review of SAFT published in 2001, Müller and Gubbins77 estimate that 200 articles dealing with SAFT had appeared. Since that review appeared, a further 70 or so articles have appeared dealing directly with SAFT or one of its variants. Significantly, SAFT is also now available in industrial process simulators such as ASPEN, PRO/II, and ChemCad as well as in the SPECS thermodynamics package of IVC-SEP at the Technical University of Denmark.

One of the drawbacks arising from the extensive use of SAFT is that many versions have appeared, with the result that the literature is complex and can be confusing. Here, we will try to outline the theoretical development in some detail of the original SAFT model (this too is not unambiguous, since “original SAFT” is often used to describe the version due to Huang and Radosz,78,79 which is slightly different from that of Chapman), as well as some of its modified ver-sions. We will then summarize some of the more interesting results obtained using SAFT and its variants.

3.4.3 oRiginal safT74

The foundation for what was to become SAFT was laid in two papers72,75 which appeared in the journal Molecular Physics in 1988. The first of these papers developed the theory required for associating fluids, while the second focused on chain formation. However, the first paper to contain an equation of state that can realistically be called SAFT appeared in the August, 1990, issue of Industrial and Engineering Chemistry Research.74 It is interesting that the Huang and Radosz paper appeared in the November issue of the same journal.78

To understand exactly what occurs in SAFT, we refer to Figure 3.5, taken from Fu and Sandler.80 Initially, a pure fluid is assumed to consist of equal-sized hard spheres (b). Next a dispersive poten-tial is added to account for attraction between the spheres (c). Typical potentials are the square-well or Lennard–Jones potential. Next each sphere is given two (or more) “sticky” spots, which enables the formation of chains (d). Finally, specific interaction sites are introduced at certain positions in the chain, which enable the chains to associate through some attractive interaction (hydrogen bonding)  (e). This interactive energy is often taken to be a square-well potential. The final sin-gle molecule is shown in Figure 3.5a. Each of these steps contributes to the Helmholtz energy. The residual Helmholtz energy is given by

a a a ares seg chain assoc= + + (3.22)

whereaseg is the Helmholtz energy of the segment, including both hard-sphere reference and dispersion

termsachain is the contribution from chain formationaassoc is the contribution from association

Page 21: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

218 Handbook of Surface and Colloid Chemistry

It is this separation of the Helmholtz energy into additive components that is partly respon-sible for the fact that SAFT has so many variants—each of the contributions can be considered (and modified) separately, or new terms may be added (such as polar, electrostatic, or other con-tributions). The individual terms may also be used outside of the context of SAFT. Thus, the term that accounts for association has been combined (with minor modification) with the SRK equation to create CPA, an equation of state, which has had remarkable success in a number of applications.81

It is worth noting that both the chain formation and the association term derive from Wertheim’s TPT. However, chains (i.e., covalent bonds) are formed in the limit of complete association. It is these two terms that make up the innovative development in SAFT—the first because now we have available a rational method for considering polymer molecules (linear chains with very many bonded segments) and the second because we can now consider associating molecules in a way that more closely resembles the actual physical picture. The calculation of useful thermodynamic prop-erties from aassoc was initially a complex procedure, although Michelsen and Hendriks82 simplified the computations considerably by recasting the equations in a simpler, although mathematically equivalent form.

We now consider each of the terms individually. We follow the original notation of Chapman et  al.74 Each pure component is characterized by a chain length m, a size-parameter σ, and an energy parameter, ε. If the molecule is self-associating, there are two further parameters,

(b)

(e) (d)

(c)

Chainside

(a)

Associationside

FIGURE 3.5 Procedure to form a molecule in the SAFT model. (a) The proposed molecule. (b) Initially the fluid is a hard-sphere fluid. (c) Attractive forces are added. (d) Chain sites are added and chain molecules appear. (e) Association sites are added and molecules form association complexes through association sites. (From Fu, Y.-H. and Sandler, S.I., Ind. Eng. Chem. Res., 34, 1897, 1995. With permission.)

Page 22: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

219Thermodynamics of Polymer Solutions

which characterize the volume (κAiBi) and energy (εAiBi) of association. The association term is given for mixtures by

aRT

X X X Massoc

iA

A

iAi

ii

i

= −⎡

⎣⎢

⎦⎥ +

⎣⎢⎢

⎦⎥⎥∑∑ ln

212

(3.23)

whereXAi is the mole fraction of molecules i not bonded at site AMi is the number of association sites on molecule i

We have

X N XAAV j

B A B

Bj

i j i j

j

= +⎡

⎢⎢

⎥⎥∑∑−

1

1

ρ Δ (3.24)

whereNAV is the Avogadro’s numberρj is the molar density of jΔAiBj is the association strength given by

ΔA Bij ij ij

seg A BA B

i j i ji j

d g dkT

=⎛

⎝⎜⎜

⎠⎟⎟ −

⎣⎢⎢

⎦⎥⎥

3 1( ) expκε

(3.25)

Here dij is a temperature-dependent size-parameter related to σij by

d f kT m=

⎝⎜

⎠⎟σ

ε, (3.26)

where

f kT m kT

kT f m kTεε

ε ε, .

. ( )( )⎛

⎝⎜

⎠⎟ =

++ +

1 0 29771 0 33163 2

// /

(3.27)

and

f m m

m( ) . .= +

−0 0010477 0 025337 1 (3.28)

This temperature dependence is incorporated to account for the fact that real molecules are not hard spheres, but rather there is some degree of interpenetration between molecules, particularly at high temperatures. Thus, the “effective” hard-sphere diameter of a segment is smaller at higher temperatures. The radial distribution function in Equation 3.25 is given by the mixture version of the Carnahan–Starling equation of state for hard-sphere mixtures:

g d g d d dd d

dijseg

ij ijHS

iji j

i j

i( )( )

≈ ( ) =−

++

⎝⎜

⎠⎟

−++ 1

13

13

2

32ζ

ζζ

ddd d

j

i j+

⎝⎜

⎠⎟

222

33

21ζζ( )

(3.29)

Page 23: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

220 Handbook of Surface and Colloid Chemistry

where

ζπ

ρkAV

i i iik

i

N X m d= ∑6 (3.30)

For the chain term in Equation 3.22, we have

aRT

X m g dchain

i i ii iiHS

i

= −∑ ( ) ln( ( ) )1 (3.31)

Finally, for the segment term in Equation 3.22, we have

a a X mseg segi i

i

= ∑0 (3.32)

where the 0 subscript indicates a nonassociated segment. The segment energy consists of a hard-sphere reference and a dispersion contribution:

a a aseg HS disp0 0 0= + (3.33)

The Carnahan–Starling equation66 is used for both pure components and mixtures to give

aRT

HS0

2

24 3

1=

−−η η

η( ) (3.34)

where for mixtures η = ζ3 as defined by Equation 3.30. The dispersion term is given by

a Rk

a aT

disp dispdisp

R0 01

02= +⎛

⎝⎜

⎠⎟

ε (3.35)

where

adisp R R R R01

2 30 85959 4 5424 2 1268 10 285= − − − +⎡⎣ ⎤⎦ρ ρ ρ ρ. . . . (3.36)

adisp R R R R02

2 31 9075 9 9724 22 216 15 904= − − − − +⎡⎣ ⎤⎦ρ ρ ρ ρ. . . (3.37)

The reduced quantities are given by TR = kT/ε and ρR = (6/20.5π)η.Most of the results of this initial paper are comparisons with simulation data for chains with

various parameters, although pure-component parameters for six hydrocarbons and two associating fluids were fitted. No results for mixtures of real fluids are presented.

3.4.4 “oRiginal” safT78

Probably the main contribution of the Huang and Radosz78 version of SAFT was the regression of pure-component parameters for over 100 different fluids. There are also some notational

Page 24: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

221Thermodynamics of Polymer Solutions

differences. Thus, instead of a size parameter σ, they use a volume parameter v00, which is related through the equation

v NAV00 3

6=πτ

σ (3.38)

Here τ = 0.74048 is the highest possible packing fraction for a system of pure hard spheres. They also use the notation u0 instead of ε for the energy parameter, although these terms are completely equivalent.

The rather complex temperature dependence of the hard-sphere diameter given by Equations 3.26 through 3.28 was simplified by Huang and Radosz, following Chen and Kreglewski83 to

d ukT

= −−⎡

⎣⎢

⎦⎥

⎣⎢⎢

⎦⎥⎥

σ 1 0 12 3 0. exp (3.39)

The dispersion term is also different from that of Chapman et al.74 and is given by

aRT

D ukT

disp

ij

i

j

j

i

0 =⎡

⎣⎢⎤

⎦⎥⎡

⎣⎢⎤

⎦⎥∑∑ ητ

(3.40)

whereDij are universal constantsu is the temperature-dependent energy parameter given by u = u0(1 + e/kT) and e/k is a constant

set to −10, with a few exceptions

Another important contribution of Huang and Radosz is the presentation of detailed tables discuss-ing bonding schemes for different associating fluids. These schemes are presented as Tables 3.11 and 3.12 and have been widely adopted in the literature of SAFT and other equations of state for associating fluids.

In their paper on mixture properties,79 Huang and Radosz also use the full mixture version of the Carnahan–Starling equation for the hard-sphere mixtures reference system:

aRT

HS=

−+

−+ −⎛

⎝⎜

⎠⎟ −

⎣⎢

1 31 1

10

1 2

3

23

3 32

23

32 0 3

ζζ ζζ

ζζ ζ

ζζ

ζ ζ( )

ln( )⎢⎢

⎦⎥⎥

(3.41)

One of the reasons that the Huang and Radosz version of SAFT has been adopted (and is widely referred to as “original” SAFT) is that they undertook an extensive pure-component parameter-ization for over 100 pure fluids. This meant that their equation of state could be used imme-diately for real fluids of industrial interest without any intermediate steps. Both Huang and Radosz were employed by Exxon during the development of SAFT. The fact that the model had the backing of a major oil company may also help explain its rapid adoption and use as an engineering tool.

Besides it is interesting to note that the first paper to appear describing PC-SAFT also had pure-component parameters for about 100 species.84 This fact, coupled with the success of the model, has certainly been partly responsible for the rapid adoption of PC-SAFT in the few years it has been in existence, both in industry and in academia.

Page 25: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

222 Handbook of Surface and Colloid Chemistry

TABLE 3.11Unbonded Site Fractions XA for Different Bonding Types

Type Δ Approximations XA Approximations XA

1 ΔAA ≠ 0− + +1 1 4

2

1 2( ) /ρρ

ΔΔ

2A ΔAA = ΔAB = ΔBB ≠ 0 XA = XB− + +1 1 8

4

1 2( ) /ρρ

ΔΔ

2B ΔAA = ΔBB = 0ΔAB ≠ 0

XA = XB − + +1 1 42

1 2( ) /ρρ

ΔΔ

3A ΔAA = ΔAB = ΔBB = ΔAC = ΔBC = ΔCC ≠ 0 XA = XB = XC− + +1 1 12

6

1 2( ) /ρρ

ΔΔ

3B ΔAA = ΔAB = ΔBB = ΔCC = 0ΔAC = ΔBC ≠ 0

XA = XB

XC = 2XA − 1− − + + +( ) (( ) ) /1 1 4

4

2 1 2ρ ρ ρρ

Δ Δ ΔΔ

4A ΔAA = ΔAB = ΔBB = ΔAC = ΔBC = ΔCC = ΔAD = ΔBD = ΔCD = ΔDD ≠ 0

XA = XB = XC = XD − + +1 1 168

1 2( ) /ρρ

ΔΔ

4B ΔAA = ΔAB = ΔBB = ΔAC = ΔBC = ΔCC = ΔDD = 0ΔAD = ΔBD = ΔCD ≠ 0

XA = XB = XC

XD = 3XA − 2− − + + +( ) (( ) ) /1 2 1 2 4

6

2 1 2ρ ρ ρρ

Δ Δ ΔΔ

4C ΔAA = ΔAB = ΔBB = ΔCC = ΔCD = ΔDD = 0ΔAC = ΔAD = ΔBC = ΔBD ≠ 0

XA = XB = XC = XD − + +1 1 84

1 2( ) /ρρ

ΔΔ

Source: Huang, S.H. and Radosz, M., Ind. Eng. Chem. Res., 29, 2284, 1990. With permission.

TABLE 3.12Types of Bonding in Real Associating Fluids

Species Formula Rigorous Type Assigned Type

Acid O

OCC

HO

OH

1 1

AlkanolO

BC H

A 3B 2B

WaterO H c

AB

H D

4C 3B

AminesTertiary

NA 1 Non-self-associating

SecondaryN

A

H B

2B 2B

PrimaryN

C

H A

H B3B 3B

AmmoniaN

D

H C

H BA H4B 3B

Source: Huang, S.H. and Radosz, M., Ind. Eng. Chem. Res., 29, 2284, 1990. With permission.

Page 26: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

223Thermodynamics of Polymer Solutions

3.4.5 simPlified safT80

The key idea in the work of Fu and Sandler is the simplification of the dispersion term. All other terms from Huang and Radosz are retained. Since the dispersion term given by Huang and Radosz, Equation 3.40, contains 24 constants, it seems reasonable to attempt to simplify this term. For mix-tures, the dispersion Helmholtz free energy is

aRT

mZ vv v Y

disp

Ms

s=

+

⎜⎜

⎟⎟

ln * (3.42)

where“average” chain length m x mi i

i=∑

ZM = 36 is the maximum coordination numbervs = 1/ρm is the total molar volume of a segment

v YN x x m m d u kT

x x m m

AV i j i j ij ijji

i j i jji

* exp( )=

( ) −⎡⎣ ⎤⎦∑∑∑∑

3 2 1/ (3.43)

The remaining terms have the same meaning as in Huang and Radosz.78 Generally, simplified SAFT performs as well as Huang and Radosz SAFT, although it requires refitting all the pure- component parameters. Fu and Sandler provide parameters for 10 nonassociating and 8 associating fluids. Table 3.13 is reproduced from Fu and Sandler. It is interesting because it presents different types of cross-association in some detail. This scheme is completely general and applicable to any equation of state incorporating association. Combining rules for cross-association also need to be introduced, however, and are far from self-evident.

3.4.6 lJ-safT85,86

The main change in the SAFT version of Kraska and Gubbins85,86 is that they use Lennard–Jones (LJ) spheres for the reference term, rather than hard spheres. The remaining terms are unchanged, except that the radial distribution function used in the calculation of the chain and association contributions in Equations 3.25 and 3.31 is the radial distribution function for LJ spheres rather than hard spheres. Thus, an equation of state for LJ spheres is required. The equation used is that of Kolafa and Nezbeda.87 The Helmholtz energy for the reference (LJ) system is (for a pure fluid)

A m A T B C Tseg HS hBH iji j

i j

= + − +⎛

⎜⎜

⎟⎟∑exp( * ) *,

/

,

γρ ρ ρ22

2Δ (3.44)

where

A THS = − +− +−

⎝⎜

⎠⎟

53

1 34 33 46 1

2

2ln( ) ( )( )

ηη η η

η (3.45)

ΔB CThBH ii

i2

2

7

0

,/=

=−∑ (3.46)

Page 27: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

224 Handbook of Surface and Colloid Chemistry

ρ* = mb

Vm (3.47)

η

πρ σ=

63* BH (3.48)

σBH ii

i

DT D T= +=−∑ /

ln ln2

2

1

(3.49)

Apart from using a different reference system, the notation in Kraska and Gubbins does not follow the customary SAFT notation, nor are the pure-component parameters defined in the same way (e.g., the energy parameter with units of 1/K is defined inversely as a temperature parameter with units K) and care should be taken in using it. They also incorporate a term to account for dipole–dipole interactions.

TABLE 3.13Type of Association for Cross-Associating Mixtures

Mixture Component 1 Component 2 Association Type

Acid–acid O

O HC Site A1

H O

OCSite A2

∈ ≠

∈ ≠

∈ ≠

A A

A A

A A

1 1

2 2

1 2

000

Alcohol–alcohol

ÖC

HSite A1

Site B1

ÖC

HSite A2

Site B2 ∈ ≠ ∈ ≠

∈ =∈ ≠

∈ =∈ =∈ =∈ =

A B A B

A B A B

A A A A A A A A

B B

1 1 2 2

1 2 2 1

1 1 2 2 1 2 2 1

1 1

0 00

0

,

==∈ =∈ =∈ =B B B B B B2 2 1 2 2 1 0

Acid–alcoholO

O HC Site A1

H

ÖCSite B2

Site A2

∈ ≠ ∈ ≠

∈ =∈ ≠

∈ =∈ =

A B A B

A A A B

A A B B

1 1 2 2

1 2 1 2

2 2 2 2

0 000

,

Water–acid

ÖH

H

Site C1

Site A1

Site B1

H O

O

CSite A2

∈ =∈ ≠ ∈ ≠

∈ =∈ =∈ ≠

∈ =∈ =∈ =

A C B C A A

A A B A C A

A A B B C C

1 1 1 1 2 2

1 2 1 2 1 2

1 1 1 1 1 1

0 00

,

∈∈ =A B1 1 0

Water–alcohol

ÖH

H

Site C1

Site B1

Site A1

Ö

H

C

Site B2

Site A2

∈ =∈ ≠ ∈ ≠

∈ =∈ ∈ ≠

∈ =∈ =∈ =∈

A C B C A B

A B B B C A

A A B B C C

1 1 1 1 2 2

1 2 1 2 1 2

1 1 1 1 1 1

0 00

,

AA B

A A B B

A A B A C B

1 1

2 2 2 2

1 2 1 2 1 2

=

∈ =∈ =

∈ =∈ =∈ =

00

0

Source: Fu, Y.-H. and Sandler, S.I., Ind. Eng. Chem. Res., 34, 1897, 1995. With permission.

Page 28: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

225Thermodynamics of Polymer Solutions

3.4.7 safT-vR88

SAFT-VR is the version of SAFT developed by George Jackson and coworkers first at the University of Sheffield and currently at Imperial College (Gil-Vilegas et al.88; McCabe et al.89). SAFT-VR is identical to the Huang and Radosz version except in the dispersion contribution. This term incor-porated attraction in the form of a square-well potential. Thus, in addition to a segment being characterized by a size and an energy parameter, the square-well width (λ) is also included as a pure-component parameter. Thus, changing the parameter λ changes the range of attraction of the segment (hence the name VR for “variable range”). It is the introduction of this extra term that gives SAFT-VR greater flexibility, since we now have an extra pure component to “play with.” Although it is generally desirable to describe pure-component liquid densities and vapor pressure with the minimum number of parameters, the extra variable-range parameter may be necessary for the description of certain anomalous behaviors in systems containing water. The Helmholtz energy for the dispersion energy is given by

a a

kTakT

disp = +1 22( )

(3.50)

where

a x x gs s i s j ijVDW HS

x xeff

ji1 = − ⎡⎣ ⎤⎦∑∑ρ α σ ζ, , ; (3.51)

The subscript s refers to segment rather than molecule properties. We have

α

πε σ λijVDW ij ij ij

=−( )2 1

3

3 3

(3.52)

and gHS is the radial distribution function for hard spheres as before except that the arguments are different:

ζ ζ ζ ζxeff

x x xc c c= + +1 22

33 (3.53)

σ σx s i s j ijji

x x3 3= ∑∑ , , (3.54)

ζ

πρ σx s x=

63 (3.55)

The constants ci in Equation 3.53 are given by

ccc

1

2

3

2 25855 1 50349 0 2494340 66927 1 40049 0

⎜⎜⎜

⎟⎟⎟=

− −

. . .. . .8827739

10 1576 15 0427 5 30827

1

2. . .−

⎜⎜⎜

⎟⎟⎟

⎜⎜⎜

⎟⎟⎟

λ

λij

ij

(3.56)

Page 29: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

226 Handbook of Surface and Colloid Chemistry

The second-order term in Equation 3.50 is given by

a x x K as i s j HS ij s

sj

n

i

n

21

11

12

=∂∂

==∑∑ , , ε ρ

ρ (3.57)

where

KHS =

−− + − +

ζ ζζ ζ ζ ζ ζ ζ

0 34

0 32

1 2 3 23

11 6 1 9

( )( ) ( )

(3.58)

3.4.8 Pc-safT84

A recent version of SAFT that has appeared is that due to Gross and Sadowski84 developed at the Technical University of Berlin. Once again, most of the terms in PC-SAFT are the same as those in the Huang and Radosz version. The term that is different is the dispersion term. However, it is not simply a different way of expressing the dispersion attraction between segments, but rather it tries to account for dispersion attraction between whole chains. Referring to Figure 3.5 should make this clear. Instead of adding the dispersion to hard spheres and then forming chains, we first form hard-sphere chains and then add a chain dispersion term, so the route in Figure 3.5 would be (b)–(d)–(c)–(e). To do this, we require interchain rather than intersegment radial distribution func-tions. These are given by O’Lenick et al.90 The Helmholtz energy for the dispersion term is given as the sum of a first- and second-order term:

AkTN

AkTN

AkTN

disp= +1 2 (3.59)

where

AkTN

mkT

u x g m xd

x dxhc1 2 3 2

1

2= −⎛

⎝⎜

⎠⎟

⎝⎜

⎠⎟

∫πρε

σσ�( ) ; (3.60)

AkTN

m Z Z mkT

u x ghchc

hc21

22

3 21= − + +∂∂

⎝⎜

⎠⎟

⎝⎜

⎠⎟

∂∂

πρ ρρ

εσ

ρρ �( ) mm x

dx dx; σ⎛

⎝⎜

⎠⎟

⎢⎢

⎥⎥

∫ 2

1

(3.61)

wherex = r/σ�u x u x( ) ( )/= ε is the reduced intermolecular potential

The radial distribution function ghc is now an interchain function rather than a segment function as before. This is a key point in PC-SAFT. The term involving compressibilities is given by

1 1 8 21

1 20 27 12 22

4

2 3 4+ +

∂∂

⎝⎜

⎠⎟ = +

−−

+ −− + −Z Z m mhc

hcρ

ρη η

ηη η η η

( )( )

((( )( ))1 2 2− −

⎝⎜

⎠⎟

η η (3.62)

Page 30: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

227Thermodynamics of Polymer Solutions

We still need to solve the integrals in Equations 3.60 and 3.61. Setting

I u x g m xd

x dxhc1

2

1

=⎛

⎝⎜

⎠⎟

∫ �( ) ; σ (3.63)

I u x g m xd

x dxhc2

2 2

1

=∂∂

⎝⎜

⎠⎟

⎢⎢

⎥⎥

∫ρρ

σ�( ) ; (3.64)

we can substitute the Lennard–Jones potential and the radial distribution function of O’Lenick et al.90 This was done for the series of n-alkanes and the integrals were fit as a power series:

I ai i

i1

0

6

==∑ η (3.65)

I bi i

i2

0

6

==∑ η (3.66)

with

a a m

ma m

mmm

ai i i i= +−

+− −

0 1 21 1 2

(3.67)

b b m

mb m

mmm

bi i i i= +−

+− −

0 1 21 1 2

(3.68)

Equations 3.67 and 3.68 require a total of 42 constants, which are adjusted to fit experimen-tal pure-component data of n-alkanes. This direct fitting to experimental data to some extent accounts for errors in the reference equation of state, the perturbing potential, and the radial distribution function, which appear in the integrals of Equations 3.63 and 3.64. The dispersion potential given by Equations 3.60 and 3.61 is readily extended to mixtures using the van der Waals one-fluid theory.

Since this first PC-SAFT paper appeared, the authors rapidly published a series of further papers applying PC-SAFT to polymers,91,92 associating fluids93 and copolymers.94

3.4.9 simPlified Pc-safT95

Two simplifications to PC-SAFT have been proposed, which simplify phase equilibrium calcula-tions substantially for mixtures. For pure components, this simplified PC-SAFT becomes original PC-SAFT, so the simplifications may be considered as a particular set of mixing rules. The advan-tage of this is that existing pure-component parameters can be used in simplified PC-SAFT—no refitting is required. The targets of the simplifications are Equations 3.29 and 3.41. In other words, this simplified PC-SAFT targets the hard-sphere reference equation of state. The remaining terms are the same, except as mentioned previously, the simpler radial distribution function will affect both the chain and association terms, since the radial distribution function appears in both of these terms.

Page 31: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

228 Handbook of Surface and Colloid Chemistry

By setting η ≡ ζ3, Equations 3.29 and 3.40 reduce to

g dHS ( )

( )+ =

−−

1 21 3ηη/

(3.69)

and

�aHS = −

−4 3

1

2

2η η

η( ) (3.70)

respectively. In the initial paper,95 use of Equation 3.69 only is called modification 1, while use of both Equations 3.69 and 3.70 is called modification 2. This and subsequent work has shown that there is no loss of accuracy using the most simplified version of PC-SAFT, so modification 2 is used throughout and is called “simplified PC-SAFT.” Simplified PC-SAFT has since been applied in our group to several polymer systems including: polymer VLE,96 polymer–solvent binary LLE,97 ternary polymer–solvent and blend systems,98 and high-pressure gas solubility in polymers.99,100

One of the interesting points about SAFT in general is the ability to extrapolate the properties of higher molecular weight substances from knowledge of similar shorter chain compounds. This is most evident for the n-alkane series. Pure-component parameters for polyethylene can in principle be predicted by extrapolating from the properties of the n-alkanes. In practice, this is problematic, since for very long chains, effects such as entanglement of the individual polymer chains start to influence the behavior. Figure 3.6 shows PC-SAFT parameters for the alkanes up to eicosane (C20). The parameters m, mσ3, and mε/k are all linear with molecular weight. Tables 3.14 and 3.15 show computing times for simplified PC-SAFT compared with original PC-SAFT, as well as with other equations of state.

2000

mε/kmσ3

m

9

1800

1600

1400

1200

1000

800

600

400

200

00 100 200

Molecular weight300

8

7

6

5

4

3

2

1

0

FIGURE 3.6 The groups m, mσ3, and mε/k vs. molecular weight for linear alkanes up to eicosane. Points are PC-SAFT parameters, lines are linear fits to these points, excluding methane. (From von Solms, N. et al., Ind. Eng. Chem. Res., 42, 1098, 2003. With permission.)

Page 32: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

229Thermodynamics of Polymer Solutions

3.4.10 exTensions To safT-TyPe equaTions

The additive contributions within SAFT mean that the equation of state is quite versatile— contributions can be added to account for effects not included in the discussion earlier. Thus, papers have appeared accounting for polar,101 quadrupolar,102 and dipolar103 molecules, as well as polariz-able dipoles.104

Versions of SAFT have also appeared, which include an electrostatic contribution, the intention being to develop a version of SAFT suitable for modeling electrolytes.105,106

Care should be taken when adding very many contributions, since each extra contribution will almost always require one or more pure-component parameters. The actual physical picture should always be borne in mind when considering which contributions to include.

Another recent development is the application of group contribution methods to SAFT (see, e.g., Le Thi et al.107 and references therein). Rather than the addition of extra terms, which then require more pure-component parameters, the use of group contribution methods is an attempt to generalize the parameters in SAFT (extending even to binary interaction parameters). The hope is that in this way SAFT becomes a more predictive tool, relying less on fitting of parameters to experimental data.

TABLE 3.14Comparison of Computing Times for 36-Component Phase-Envelope Calculation with PC-SAFT and with Simplified PC-SAFT

Calculation Computing Time (ms)

Phase-envelope calculation, 36-component mixture, full PC-SAFT 48

Phase-envelope calculation, 36-component mixture, PC-SAFT modification 2 32

Source: von Solms, N. et al., Ind. Eng. Chem. Res., 42, 1098, 2003. With permission.Computations were performed on a 2.0 GHz Pentium IV machine with DVF compiler.

TABLE 3.15Comparison of Computing Times for Various Models and Mixtures

Mixture Model

Computing Time (μs)

Fugacity Coefficients OnlyAll Derivatives (T, P, Composition

Residual Heat Capacity)

SRK, 6 components 1.9 3.4

SRK, 6 components 3.1 5.6

CPA, 15 components, 0 sites 15 20.1

CPA, 15 components, 2 sites 18.5 24.8

CPA, 15 components, 4 sites 24.4 32.2

CPA, 15 components, 6 sites 32.3 40.9

CPA, 15 components, 8 sites 40.7 50.9

Modification 2, 15 components, 0 sites 15.6 23.1

Modification 2, 15 components, 6 sites 39.1 54.4

Source: von Solms, N. et al., Ind. Eng. Chem. Res., 42, 1098, 2003. With permission.Number of sites refers to number of association sites on a molecule when employing an equation of state with association.

Page 33: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

230 Handbook of Surface and Colloid Chemistry

3.4.11 aPPlicaTions of safT

The remainder of this discussion looks at some polymer applications of SAFT. Figure 3.7 shows the results of using parameters extrapolated based on Figure 3.6. The system is methane– tetratetracontane (C44), where the C44 parameters are obtained by extrapolation of the lines in Figure 3.6. While this is not a polymer system, the method of obtaining parameters by extrapola-tion is applicable to polymer systems. A more sophisticated method for finding polymer parameters based on extrapolation of the monomer properties and polymer density data has been published recently.108

Figures 3.8 and 3.9 show a comparison of the various modifications and original PC-SAFT for VLE in the systems polystyrene–propyl acetate (Figure 3.8) and polypropylene–diisopropyl ketone (Figure 3.9). In general, PC-SAFT and simplified PC-SAFT performed similarly, as can be seen from Table 3.16, which gives the errors in prediction for a large number of polymer–solvent VLE systems. Figure 3.10 is a pressure–weight fraction diagram (VLE) for the polymer poly(vinyl acetate) in the associating solvent 2-methyl-1-propanol. A small value of the binary interaction parameter correlates the data well.

Figure 3.11 is an illustration of a novel method developed by von Solms et al.97 for finding LLE in polymer systems, known as the method of alternating tangents. This figure shows the Gibbs energy of mixing for two binary systems as a function of the mole fraction of component 1. The method will be illustrated with reference to the system methanol(1)–cyclohexane(2), since this curve clearly shows the existence of two phases. The composition of methanol in each phase is found by locating a single line, which is a tangent to the curve in two places (the common tangent). In Figure 3.11, these compositions are given by xeq1

1 and xeq12. In fact the curve for the system PS(1)–acetone(2) also

shows the existence of two phases, although this is not visible. The first step in the procedure is to determine whether a spinodal point exists (this is a necessary condition for phase separation). In the figure, the two spinodal points are given by the compositions xsp1

1 and xsp12. The spinodal condi-

tion is given by ((∂2gmix/RT)/∂x2) = 0, that is, an inflection point on the curve. Once a spinodal point has been found (using a Newton–Raphson method), the next step is to find the point of tangent of a

8

7

Darwish et al.

6

5

4

3

P (M

Pa)

2

1

00 0.1 0.2

X1

0.3

kij = 0.0

kij = 0.04

0.4

FIGURE 3.7 Vapor pressure curve for the system methane(1)–tetratetracontane(2) at T = 423.2 K. The dashed line is PC-SAFT, the solid line is PC-SAFT with a binary interaction parameter kij = 0.04. The points are experimental data. (From von Solms, N. et al., Ind. Eng. Chem. Res., 42, 1098, 2003. With permission.)

Page 34: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

231Thermodynamics of Polymer Solutions

40

35

30

25

20

15Pres

sure

(kPa

)

Modification 1PC-SAFTBawn and Wajid

Modification 210

5

00 0.2 0.4

Propyl acetate weight fraction0.6 0.8 1

FIGURE 3.8 Vapor pressure curve for the system propyl acetate(1)–polystyrene(2) at T = 343.15 K. The dashed line is PC-SAFT (the lowest line on the plot), the solid line is modification 1, and the dotted line is modification 2. Points are experimental data. All lines are pure predictions (kij = 0). (From von Solms, N. et al., Ind. Eng. Chem. Res., 42, 1098, 2003. With permission.)

6

5

4

3

2Pres

sure

(kPa

)

1

00 0.2 0.4

Solvent weight fraction0.6 0.8

Brown et al. (1964)Simplified PC-SAFTPC-SAFT

1

FIGURE 3.9 Pressure–weight fraction plot of polypropylene(1)–diisopropyl ketone(2) at T = 318 K. Polypropylene molecular weight = 20,000. Comparison of experimental data with the predictions of original (solid line) and the simplified version (dotted line) of PC-SAFT. In both curves, the interaction parame-ter kij = 0. (From Fluid Phase Equilib., 215, Kouskoumvekaki, I.A., von Solms, N., Michelsen, M.L., and Kontogeorgis, G.M., Application of a simplified perturbed chain SAFT equation of state to complex polymer systems, 71–78, Copyright 2004, with permission from Elsevier.)

Page 35: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

232 Handbook of Surface and Colloid Chemistry

line originating at the spinodal point. In the figure, this is the line connecting xsp11 and x1. This point

is just to the left of xeq12 (i.e., we are not yet quite at the equilibrium concentration after one step).

Once the first tangent has been found, the point of tangent opposite is then found in a similar way. This process is repeated until the change in the composition at the tangent point is within a specified tolerance. At this point, the equilibrium values have been calculated.

TABLE 3.16Comparison of the Performance of the Simplified against the Original PC-SAFT in Predicting Vapor–Liquid Equilibria of Polymer Solutions (kij = 0 in All Cases)

% AAD

PC-SAFT

Simplified Version Original Version

Cyclic hydrocarbons

PS–cyclohexane 13 16

PS–benzene 28 13

PS–ethyl benzene 3 6

PS–m-xylene 25 16

PS–toluene 18 7

PVAc–benzene 6 10

Chlorinated hydrocarbons

PS–carbon tetrachloride 18 12

PS–chloroform 30 11

PP–dichloromethane 59 74

PP–carbon tetrachloride 55 47

Esters

PS–propyl acetate 5 21

PS–butyl acetate 3 25

PVAc–methyl acetate 3 2

PVAc–propyl acetate 19 18

Ketones

PS–acetone 6 26

PS–diethyl ketone 7 28

PS–methyl ethyl ketone 14 12

PVAc–acetone 4 7

PP–diethyl ketone 16 27

PP–diisopropyl ketone 4 11

PVAc–methyl ethyl ketone 7 6

Amines

PVAc–propylamine 4 3

PVAc–isopropyl amine 17 16

Alcohols

PVAc–1-propanol 56 54

PVAc–2-propanol 84 73

PVAc–1-butanol 59 59

PVAc–2-butanol 39 36

PVAc–2-methyl-1-propanol 29 29

Overall 23 24

Source: Kouskoumvekaki, I.A. et al., Fluid Phase Equilib., 215, 71, 2004. With permission.Average percentage deviation between experimental and predicted equilibrium pressure curves.

Page 36: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

233Thermodynamics of Polymer Solutions

12

10

8

6

Wibawa et al. (2002)

kij = –0.012kij = 04

2

00 0.1 0.2 0.3

Solvent weight fraction0.4 0.5

Pres

sure

(kPa

)

FIGURE 3.10 Pressure–weight fraction plot of poly(vinyl acetate)(1)–2-methyl-1-propanol(2) at T = 313 K. Comparison of experimental data with prediction (kij = 0) and correlation (kij = −0.012) results of simplified PC-SAFT. Poly(vinyl acetate) molecular weight = 167,000. (From Fluid Phase Equilib., 215, Kouskoumvekaki, I.A., von Solms, N., Michelsen, M.L., and Kontogeorgis, G.M., Application of a simpli-fied  perturbed chain SAFT equation of state to complex polymer systems, 71–78, Copyright 2004, with permission from Elsevier.)

0

Methanol–cyclohexanePS–acetone

0

–0.02

1eq1

–0.04

–0.06

–0.08

–0.1

–0.12

–0.14

–0.160 0.2

gmix

/RT

0.4Mole fraction (x1)

0.6 0.8 1

–0.5

–1

–1.5

–2

–2.5

1eq2

1

1sp2

1sp1

FIGURE 3.11 Illustration of the method of alternating tangents. The solid line is the system methanol(1)–cyclohexane(2). The dotted line is the system PS(1)–acetone. The two spinodal points are indicated by xsp1

1 and xsp1

2. The equilibrium (binodal) points are indicated by xeq11 and xeq1

2. Starting from a spinodal point, the equilibrium values can be calculated by solving for only one point at a time. (From Fluid Phase Equilib., 222–223, von Solms, N., Kouskoumvekaki, I.A., Lindvig, T., Michelsen, M.L., and Kontogeorgis, G.M., A novel approach to liquid-liquid equilibrium in polymer systems with application to simplified PC-SAFT, 87–93, Copyright 2004, with permission from Elsevier.)

Page 37: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

234 Handbook of Surface and Colloid Chemistry

Figures 3.12 through 3.15 are examples of binary LLE for polymer systems. Figure 3.12 shows results for the system polystyrene–methylcyclohexane for different molecular weights of polysty-rene. The experimental data are from the classic work of Dobashi et al.109 The lines are simpli-fied PC-SAFT correlations with kij = 0.0065 for polystyrene molecular weights 10,200, 46,400 and 719,000 in order of increasing critical solution temperature. The data are reasonably well cor-related over a very large range of molecular weight with a single value of the binary interaction parameter, kij. The binary interaction parameter was adjusted to give the correct upper critical solution temperature. However, the correct critical solution concentration is not obtained, although the experimental trends are correctly predicted by the model: The critical solution temperature increases, and the polymer weight fraction at the critical solution temperature decreases with increasing molecular weight.

Figure 3.13 shows results for the system polyisobutylene–diisobutyl ketone at different polymer molecular weights. The experimental data are from Shultz and Flory.110 The lines are simplified PC-SAFT correlations. A single binary interaction parameter (kij = 0.0053) was used for all three systems, although it seems that there is a weak dependence of molecular weight on kij. Incorporating a functional dependence of kij on molecular weight (e.g., a linear fit) would improve the correlation. It should also be noted that these three systems represent a very large range of molecular weights.

Figure 3.14 shows the results for a single molecular weight of HDPE in five different n-alkanol solvents from n-pentanol up to n-nonanol. The results are well correlated using simplified PC-SAFT with a small value of the binary interaction parameter kij. A kij value of around 0.003 gives a good cor-relation for all the systems, except HDPE–n-pentanol. In the figure, a small value (kij = 0.0006) was used to correlate the data, although the data is also well predicted by simplified PC-SAFT (kij = 0), giving an error in the upper critical solution temperature of 3 K in the case of HDPE–n-pentanol.

Figure 3.15 shows the results for the system HDPE–butyl acetate. This system displays both UCST and LCST behaviors. A single binary interaction parameter (kij = 0.0156) was used to cor-relate the data for both molecular weights shown. The binary interaction parameter was adjusted to

340

719,000

330

320

310

300

290

280Tem

pera

ture

(K)

270Dobashi et al., (1980)Dobashi et al., (1984) (MW 17,200)260

2500 0.1 0.2 0.3

Weight fraction polymer0.4 0.5 0.6

181,000109,00046,40034,90020,20017,30017,20016,10010,200

FIGURE 3.12 Liquid–liquid equilibrium in the system polystyrene–methyl cyclohexane for different molecular weights of polystyrene. The experimental data are from Dobashi et  al.109 The lines are simpli-fied PC-SAFT correlations with kij = 0.0065 for polystyrene molecular weights 10,200, 46,400 and 719,000 in order of increasing critical solution temperature. (From Fluid Phase Equilib., 222–223, von Solms, N., Kouskoumvekaki, I.A., Lindvig, T., Michelsen, M.L., and Kontogeorgis, G.M., A novel approach to liquid-liquid equilibrium in polymer systems with application to simplified PC-SAFT, 87–93, Copyright 2004, with permission from Elsevier.)

Page 38: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

235Thermodynamics of Polymer Solutions

340

330

6,000,000

320

310

300

290

280Tem

pera

ture

(K)

270

260

2500 0.1 0.2 0.3 0.4

Weight fraction polymer

285,00022,700

FIGURE 3.13 Liquid–liquid equilibrium in the system polyisobutylene–diisobutyl ketone. PC-SAFT param-eters for diisobutyl ketone were obtained by fitting to experimental liquid density and vapor pressure data in the temperature range 260–600 K. This data was taken from the DIPPR database. The parameters were: m = 4.6179, ε/k = 243.72 K, and σ = 3.7032 Å. Average percent deviations were 1.03% for vapor pressure and 0.64% for liquid density. Lines are simplified PC-SAFT correlations with kij = 0.0053, the same at all three molecular weights. (From Fluid Phase Equilib., 222–223, von Solms, N., Kouskoumvekaki, I.A., Lindvig, T., Michelsen, M.L., and Kontogeorgis, G.M., A novel approach to liquid–liquid equilibrium in polymer systems with application to simplified PC-SAFT, 87–93, Copyright 2004, with permission from Elsevier.)

450

440

430

420

410

400 Pentanol

390

3800 0.1 0.2 0.3 0.4

Tem

pera

ture

(K)

Weight fraction polymer

HexanolHeptanolOctanolNonanol

FIGURE 3.14 Liquid–liquid equilibrium for HDPE with n-alkanols. Lines are simplified PC-SAFT corre-lations for each of the five solvents (pentanol highest, nonanol lowest). Polymer molecular weight is 20,000. Binary interaction parameters are as follows: pentanol: 0.0006; hexanol: 0.003; heptanol: 0.0025; octanol: 0.0033; and nonanol: 0.0029. (From Fluid Phase Equilib., 222–223, von Solms, N., Kouskoumvekaki, I.A., Lindvig, T., Michelsen, M.L., and Kontogeorgis, G.M., A novel approach to liquid–liquid equilibrium in polymer systems with application to simplified PC-SAFT, 87–93, Copyright 2004, with permission from Elsevier.)

Page 39: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

236 Handbook of Surface and Colloid Chemistry

give a good correlation for the UCST curve at the higher molecular weight (64,000). As mentioned earlier, the LCST curve is rather insensitive to kij. Nevertheless, the LCST curve is reasonably well correlated using this value. The prediction (kij = 0) is almost as good for the LCST curve, although the UCST will then be substantially underpredicted.

Figure 3.16 shows a ternary phase diagram for the system polystyrene–acetone– methylcyclohexane. The binary interaction parameters were obtained by fitting to the individual  binary systems.

540

520

500

480

460

13,600

440

420

400

3800 0.05 0.1 0.15

64,000

Tem

pera

ture

(K)

Weight fraction HDPE

FIGURE 3.15 Liquid–liquid equilibrium in the system HDPE–butyl acetate. The system displays both upper and lower critical solution temperature behaviors. The experimental data for molecular weights 13,600 and 64,000. Lines are simplified PC-SAFT correlations with kij = 0.0156 for both molecular weights. (From Fluid Phase Equilib., 222–223, von Solms, N., Kouskoumvekaki, I.A., Lindvig, T., Michelsen, M.L., and Kontogeorgis, G.M., A novel approach to liquid–liquid equilibrium in polymer systems with application to simplified PC-SAFT, 87–93, Copyright 2004, with permission from Elsevier.)

100

90

80

70

60

50

40

30

20

10

00 10

Methylcyclohexane(3)20 30 40 50 60 70 80 90

Acetone(2)

k12 = –0.005k13 = 0.006

Polystyrene(1)

100

FIGURE 3.16 Ternary phase diagram for the system polystyrene–acetone–methylcyclohexane. The binary interaction parameters were obtained by fitting to the individual binary systems. The ternary coexistence curves are predictions.

Page 40: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

237Thermodynamics of Polymer Solutions

The ternary coexistence curves are predictions. The algorithm for finding ternary LLE in systems containing polymers is an extension of the binary algorithm discussed earlier and was developed by Lindvig et al.98

Figure 3.17 from Gross et al.94 shows high pressure equilibrium for mixtures of poly(ethylene-co-methyl acrylate) (EMA) and propylene for different repeat unit compositions of EMA. As repeat units of methyl acrylate (MA) are added to the polyethylene chain, the demixing pressure at first declines, but then increases as the composition of MA increases. This effect is correctly predicted by PC-SAFT.

Finally, Figure 3.18 shows a comparison of SAFT-VR and simplified PC-SAFT in a recent study100 where the two models were compared in their ability to model multicomponent phase equilibrium in systems typical of real polyethylene reactors. The system shown here is polyethylene/nitrogen/1-butene. The results of the simplified PC-SAFT and SAFT-VR calculations for this ternary are consistent. From the figure, one can see that as butene in the vapor is replaced by nitrogen, the calculated absorption of butene decreases (not surprisingly—there is less of it to absorb). However, it is also clear that as nitrogen in the vapor is replaced by butene, absorption of nitrogen increases, even though there is less nitrogen to absorb. This suggests that there may be some enhancement/inhibition of absorption effect.

In the time since the previous edition of this work (2008), around 150 articles have appeared applying SAFT to various polymer systems. Certain newer themes become apparent here:

The use of density functional theory combined with equations of state to examine surface and structural properties is a relatively new phenomenon, resulting in equations such as iSAFT.111,112 For example, Jain et al.111 used SAFT and density functional theory to probe the structure of tethered polymer chains.

Perhaps the most frequently occurring, relatively new idea is that of group-contribution SAFT, where there has been a great deal of activity.107,113–122

Part of the motivation for a group-contribution method is the need for an equation that can be readily extended to complex molecules, where pure-component properties may be unavailable, and for which SAFT types of equation are well suited. Examples of complex systems where SAFT has been applied are biopolymers,123 hyperbranched polymers,124 and asphaltenes.125

1500

1000

500

P (b

ar)

075 95 115 135

T (C)155

68% MA

175

58% MA0% MA

25% MA

FIGURE 3.17 High-pressure equilibrium for mixtures of poly-(ethylene-co-methyl acrylate) (EMA) and propylene for different repeat-unit compositions of the EMA. Comparison of experimental cloud-point measurements to calculation results of the PC-SAFT equation of state. (EMA [0% MA] is equal to LDPE: open diamonds and dashed line.) (From Gross, J. et al., Ind. Eng. Chem. Res., 42, 1266, 2003. With permission.)

Page 41: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

238 Handbook of Surface and Colloid Chemistry

150

Solu

bilit

y (g/

100

g PE

) 120

90

60

30

(a)

041 32

p (MPa)0

Solu

bilit

y (g/

100

g PE

) 50% N2

N2

50% nC4=

(c)

150

120

90

60

30

041 320

Solu

bilit

y (g/

100

g PE

)

N2

20% nC4 =80% N2

nC4 =

(e)

4.5

3

1.5

0

6

41 320

Solu

bilit

y (g/

100

g PE

)

N2

50% N2

50% nC4 =

(d)

4.5

3

1.5

0

6

1 4320

Solu

bilit

y (g/

100

g PE

)

100% N2

N2

(f)

4.5

3

1.5

0

6

1 432

PCVR

0

Solu

bilit

y (g/

100

g PE

)

nC4 = 25% N2

N2

75% nC4=

Psat

(b)

150

120

90

60

30

041 32

p (MPa)

p (MPa)p (MPa)

p (MPa) p (MPa)

0

VRPC

100% nC4=

nC4 =

nC4 =nC4 =

Psat

FIGURE 3.18 Gas absorptions in amorphous PE calculated with SAFT-VR and simplified-PC-SAFT for a range of vapor compositions of the ternary mixture of (but-1-ene + nitrogen + the reference PE [MW = 12,000 g mol−1]) at T = 80°C: (a) 100 vapor mol% butene (binary mixture); (b) 75% but-1-ene, 25% nitrogen; (c) 50% but-1-ene, 50% nitrogen (with a vertical scale chosen to highlight butene absorption); (d) 50% but-1-ene, 50% nitrogen (with a vertical scale chosen to highlight nitrogen absorption); (e) 20% butene, 80% nitrogen; and (f) 100% nitrogen (binary mixture). In each case, solid curves represent SAFT-VR calculations and dashed curves represent simplified-PC-SAFT calculations. (From Haslam, A.J. et al., Fluid Phase Equilib., 243, 74, 2006. With permission).

Page 42: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

239Thermodynamics of Polymer Solutions

Another encouraging feature is the use of SAFT-type models in industrial applications. In addi-tion to the asphaltene example noted earlier,125 work has appeared relevant to refrigeration126,127 and catalytic polymerization processes.128

Finally, it can be mentioned that since the review article by Müller and Gubbins,77 several mono-graphs reviewing various aspects of SAFT have appeared.129–131

3.5 CONCLUDING REMARKS AND FUTURE CHALLENGES

Attempting to summarize in a few words the current status in polymer thermodynamics, we could state

1. Many databases (some available in computer form) and reliable group-contribution meth-ods are available for estimating many pure polymer properties and phase equilibria of polymer solutions such as densities, solubility parameters, glass and melting temperatures, and solvent activity coefficients.

2. Simple group-contribution methods based on UNIFAC, containing corrections for the FV effects, satisfactorily predict the solvent activities and vapor–liquid equilibria for binary and ternary polymer solutions. They are less successful for the prediction of liquid–liquid equilibria if the parameters are based on VLE. They are much more successful if the parameters are based on LLE data. The combination of a simple FV expression such as that employed in the Entropic-FV model and a local-composition energetic term such as that of UNIQUAC seems to be a very promising tool for both VLE and LLE in polymer solutions. We expect that such tools may find widespread use in the future for practical applications.

3. The SAFT model will continue to be a very successful tool for polymer systems. The low-pressure and especially the high-pressure results for systems including solvents and nonpolar polymers (with emphasis to those of interest to the polyolefin industry) are very satisfactory. However, the extension to polar systems represents so far a limitation of the model. In many ways, PC-SAFT has fulfilled the early promise of SAFT—it is reasonably simple to implement (compared with many versions of SAFT), it is relatively undemanding computationally, and it has proved successful in predicting and correlating phase equilibria in many systems containing polymers and/or associating compounds. There is still work to be done, however. One area is in finding parameters for pure polymers. Since polymers are nonvolatile, one cannot use vapor-pressure regression and the energy parameter is rela-tively insensitive within the range of experimental P–V–T data. Regressing pure- component parameters from binary systems is problematic as this leads to nonunique pure-component parameters—a situation best avoided. Another area where research is needed is in the modeling of water. There was been a great deal of effort in this area, but as yet no satisfac-tory model, fitting within the existing SAFT framework, has been developed for modeling water-containing systems. Examples of such deficiencies are predicting the density maxi-mum for pure water, as well as modeling water–hydrocarbon mutual solubilities.

4. Most theoretical/modeling studies in polymer thermodynamics are limited to a. Organic polymers b. Binary systems often involving monodisperse polymers and single solvents c. Rather “simple” polymers (polyolefins, polystyrene, PVC, etc.) d. Vapor–liquid equilibria and activity coefficients e. “Rules of thumb” estimates of miscibility (solubility parameters, theta parameters, etc.)

Some of the future challenges in the area of polymer thermodynamics will involve

1. More emphasis to multicomponent systems including both mixed solvents, blend-solvent systems as well as the effect of polydispersity.

Page 43: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

240 Handbook of Surface and Colloid Chemistry

2. Better treatment of condensed phases especially liquid–liquid and liquid–solid equilibria as well as water-soluble polymer systems and polyelectrolytes.

3. Emphasis to high-pressure systems involving both the typical nonpolar and polar polymers as well.

4. New directions such as description of “special materials” including those involving oligomers, copolymers, new structures (star-like and dendrimers), “inorganic” polymers (e.g., tin-based antifouling paints).

5. Proper account for the effects of crystallinity and cross-linking with special attention to swelling phenomena.

6. Closer collaboration with industry, for example, for testing existing theories for poly-mers having novel structures, for commercial polymers for which so far the structure is not revealed to academic researchers and for many other applications of practical inter-est. Many industrial systems are much more complex than the ones studied in academia. Closer collaboration in the future between academia and the polymer and paint/adhesives industries may further help the advancements in the area of polymer thermodynamics in the coming years.

3.A APPENDIX

3.A.1 AN EXPRESSION OF THE FH MODEL FOR MULTICOMPONENT MIXTURES

The FH model was originally developed as a model for the entropy of mixing for mixtures contain-ing molecules of different size, but it was soon modified also to account for energetic interactions. The model can be formulated in terms of the excess Gibbs energy as follows (Lindvig et al.37):

G G G

GRT

nx

GRT

a

E E comb E res

E comb

ii

ii

N

E res

i j ijj

= +

=

=

=

=

, ,

,

,

ln ϕ

ϕ ϕ

1

111

2

N

i

N

ij ij ia v

∑∑=

Using basic thermodynamics, the following expression for the activity coefficient is obtained:

ln ln lnγ γ γi icomb

ires= +

where the combinatorial term is given by

ln lnγ

ϕ ϕicomb i

i

i

ix x= + −1

and the residual term is

ln γ ϕ ϕ ϕires

i j ij i j k jkk

NC

j

NC

j

NC

v a v a= −===∑∑∑2

111

Page 44: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

241Thermodynamics of Polymer Solutions

The aforementioned formulation of the FH model is slightly different from the conventionally used formulation using the FH interaction parameter (χ12), although there is an interrelationship based on the simple equation shown above.

For a binary mixture, the multicomponent equation reduces to the traditional FH residual term:

ln γ χ ϕ1 12 22res =

ABBREVIATIONS

BR butadiene rubberCPA cubic plus associationCST critical solution temperatureEAC ethyl acetateEoS equation of stateEFV entropic-free volumeEMA poly(ethylene-co-methyl acrylate)FH Flory–Huggins (model/equation/interaction parameter)FV free-volumeGC group contribution (method/principle)GC Fl(ory) group contribution Flory equation of stateGCVOL group contribution volume (method for estimating the density)HDPE high density polyethyleneLCST lower critical solution temperatureLJ-SAFT Lennard–Jones SAFTLLE liquid–liquid equilibriaMA methyl acrylateMCSL Mansoori–Carnahan–Starling–LelandPBMA polybutyl methacrylatePC-SAFT perturbed chain-SAFTPDMS polydimethylsiloxanePE polyethylenePEMA polyethyl-methacrylatePEO polyethylene oxidePHSC perturbed hard-sphere chainPIB polyisobutylenePS polystyrenePVAC polyvinyl acetatePVC polyvinyl chlorideSAFT statistical associating fluid theorySAFT-VR SAFT-variable rangeSLE solid–liquid equilibriaSLLE solid–liquid–liquid equilibriaSRK Soave–Redlich–KwongTPT thermodynamic perturbation theoryUCST upper critical solution temperatureUNIFAC UNIQUAC functional activity coefficient (a method for estimating activity coefficients)UFV UNIFAC-FVvdW1f van der Waals one fluid (mixing rules)VLE vapor–liquid equilibriaVOC volatile organic content

Page 45: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

242 Handbook of Surface and Colloid Chemistry

REFERENCES

1. Prausnitz, J.M., Lichtenthaler, R.N., and Azevedo, E.G.D., 1999. Molecular Thermodynamics of Fluid Phase Equilibria, 3rd edn. Prentice-Hall International, Upper Saddle River, NJ.

2. Holten-Andersen, J., 1987. Physical models and coatings technology. Report T12-87, Scandinavian Paint and Printing Ink Research Institute, Hørsholm, Denmark. Also Presented at the International Conference in Organic Coatings Science and Technology, Athens, Greece, pp. 13–30, 1986.

3. Holten-Andersen, J., 1986. Heat balance and solvent balance in the drying of coatings. JOCCA, 69(12): 324–331. 4. Holten-Andersen, J. and Hansen, C.M., 1983. Solvent and Water Evaporation from coatings, Prog. Org.

Coat., 11(3): 219. 5. Holten-Andersen, J. and Eng, K., 1988. Activity coefficients in polymer solutions. Prog. Org. Coat., 16: 77. 6. Doong, S.J. and Ho, W.S., 1991. Sorption of organic vapors in polyethylene. Ind. Eng. Chem. Res., 30:

1351–1361. 7. Kampouris, E.M., Diakoulaki, D.C., and Papaspyrides, C.D., 1986. Solvent recycling of rigid poly(vinyl

chloride) bottles. J. Vinyl Technol., 8(2): 79. 8. Pappa, G.D., Kontogeorgis, G.M., and Tassios, D.P., 1997. Prediction of ternary liquid-liquid equilibria

in polymer-solvent-solvent systems. Ind. Eng. Chem. Res., 36: 5461. 9. Coleman, M.M., Graf, J.F., and Painter, P.C., 1995. Specific Interactions and the Miscibility of Polymer

Blends. Technomic Publishing Co., Inc., Lancaster, PY. 10. Mio, C., Kiritsov, S., Thio, Y., Brafman, R., Prausnitz, J.M., Hawker, C., and Malmstrom, E.E., 1998.

Vapor-liquid equilibria for solutions of dendritic polymers. J. Chem. Eng. Data, 43: 541–550. 11. Lieu, J.G., Liu, M., Frechet, J.M.J., and Prausnitz, J.M., 1999. Vapor-liquid equilibria for dendritic-

polymer solutions. J. Chem. Eng. Data, 44: 613–620. 12. von Solms, N., Nielsen, J.K., Hassager, O., Rubin, A., Dandekar, A.Y. Andersen, S.I., and Stenby, E.H.,

2004. Direct measurement of gas solubilities in polymers with a high-pressure microbalance. J. Appl. Polym. Sci., 91: 1476–1488.

13. Patterson, R., Yampolskii, Y.P., Fogg, P.G.T., Bokarev, A., Bondar, V., Illinich, O., and Shishatskii, S., 1999. IUPAC-NIST solubility data series 70. The solubility of gases in glassy polymers. J. Phys. Chem. Ref. Data, 28(5): 1255–1452.

14. Leitner, W., May 2000. Designed to dissolve. Nature, 405: 129–130. 15. Teja, A.S. and Eckert, C.A., 2000. Commentary on supercritical fluids: Research and applications. Ind.

Eng. Chem. Res., 39: 4442–4444. 16. Perrut, M., 2000. Supercritical fluid applications: Industrial developments and economic issues. Ind. Eng.

Chem. Res., 39: 4531–4535. 17. Hansen, C.M., 1967. The three-dimensional solubility parameter and solvent diffusion coefficient, PhD

Thesis. Technical University of Denmark, Copenhagen, MI. 18. Shashoua, Y., 2000. Permanence of plasticizers in polyvinylchloride objects in the museum environment.

Polym. Preprints, 41(2): 1796–1797. 19. Su, C.S., Patterson, D., and Schreiber, H.P., 1976. Thermodynamic interactions and the properties of the

PVC-plasticiser systems. J. Appl. Polym. Sci., 20: 1025–1034. 20. Demertzis, P.G., Riganakos, K.A., and Akrida-Demertzi, K., 1990. Study of compatibility of PVC and

polyester-type plasticizer blends by inverse gas chromatography. Eur. Polym. J., 26(2): 137–140. 21. Demertzis, P.G., Riganakos, K.A., and Akrida-Demertzi, K., 1991. An inverse gas chromatographic study

of the compatibility of food grade PVdC copolymer and low volatility plasticisers. Polym. Int., 25: 229–236. 22. Prausnitz, J.M., 1995. Some new frontiers in chemical engineering thermodynamics. Fluid Phase

Equilib., 104: 1–20. 23. Prausnitz, J.M., 1989. Biotechnology: A new frontier for molecular thermodynamics. Fluid Phase

Equilib., 53: 439–451. 24. Seymour, R.B., 1982. Plastics vs. Corrosives. SPE Monograph Series. Wiley, New York. 25. Barton, A.F.M., 1983. Handbook of Solubility Parameters and Other Cohesion Parameters. CRC Press,

Boca Barton, FL. 26. Barton, A.F.M., 1990. CRC Handbook of Polymer–Liquid Interaction Parameters and Solubility

Parameters. CRC Press, Boca Raton, FL. 27. Hansen, C.M., 2000. Hansen Solubility Parameters. A User’s Handbook. CRC Press, Boca Raton, FL. 28. Elbro, H.S., 1992. Phase equilibria of polymer solutions—With special emphasis on free volumes, Ph.D.

Thesis. Department of Chemical Engineering, Technical University of Denmark, Copenhagen, MI. 29. Klein, J. and Jeberien, H.E., 1980. Chainlength dependence of thermodynamic properties of polyethyl-

ene (glycol). Makromol. Chem., 181: 1237.

Page 46: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

243Thermodynamics of Polymer Solutions

30. Wen, H., Elbro, H.S., and Alessi, P., 1991. Polymer Solution Data Collection. Dechema Chemistry Data Series, Frankfurt, Germany.

31. High, M.S. and Danner, R.P., 1992. Polymer Solution Handbook; DIPPR 881 Project. Design Institute for Physical Property Data, PY.

32. Flory, P.J., 1941. Thermodynamics of high polymer solutions. J. Chem. Phys., 9: 660. 33. Huggins, M.L., 1941. Solutions of long chain compounds. J. Chem. Phys., 9: 440. 34. Qian, C., Mumby, S.J., and Eichinger, B.E., 1990. Application of the theory of phase diagrams to binary

polymer solutions and blends. Polym. Preprints, 31: 621. 35. Elbro, H.S., Fredenslund, Aa., and Rasmussen, P., 1990. A new simple equation for the prediction of

solvent activities in polymer solutions. Macromolecules, 23: 4707. 36. Lindvig, Th., Michelsen, M.L., and Kontogeorgis, G.M., 2002. A Flory–Huggins model based on the

Hansen solubility parameters. Fluid Phase Equilib., 203: 247. 37. Lindvig, Th., Economou, I.G., Danner, R.P., Michelsen, M.L, and Kontogeorgis, G.M., 2004. Modeling of

multicomponent vapour-liquid equilibria for polymer-solvent systems. Fluid Phase Equilib., 220: 11–20. 38. Schultz, A.R. and Flory, P.J., 1953. Phase equilibria in polymer-solvent systems. II. Thermodynamic

interaction parameters from critical miscibility data. J. Am. Chem. Soc., 75: 496. 39. Bondi, A., 1968. Physical Properties of Molecular Crystals, Liquids and Glasses. John Wiley & Sons,

New York. 40. Kontogeorgis, G.M., Fredenslund, Aa., and Tassios, D.P., 1993. Simple activity coefficient model for the

prediction of solvent activities in polymer solutions. Ind. Eng. Chem. Res., 32: 362. 41. Kouskoumvekaki, I., Michelsen, M.L., and Kontogeorgis, G.M., 2002. An improved entropic expression

for polymer solutions. Fluid Phase Equilib., 202(2): 325. 42. Oishi, T. and Prausnitz, M., 1978. Estimation of solvent activities in polymer solutions using a group-

contribution method. Ind. Eng. Chem. Proc. Des. Dev., 17(3): 333. 43. Hansen, H.K., Coto, B., and Kuhlmann, B., 1992. UNIFAC with lineary temperature-dependent group-

interaction parameters, IVC-SEP Internal Report 9212, Technical University of Denmark. 44. Elbro, H.S., Fredenslund, Aa., and Rasmussen, P., 1991. Group contribution method for the prediction of

liquid densities as a function of temperature for solvents, oligomers, and polymers. Ind. Eng. Chem. Res., 30: 2576.

45. Tsibanogiannis, I.N., Kalospiros, N.S., and Tassios, D.P., 1994. Extension of the GCVOL method and application to some complex compounds. Ind. Eng. Chem. Res., 33: 1641.

46. Bogdanic, G. and Fredenslund, Aa., 1994. Revision of the GC-Flory EoS for phase equilibria in mixtures with polymers. 1. Prediction of VLE for polymer solutions. Ind. Eng. Chem. Res., 33: 1331.

47. Bogdanic, G. and Fredenslund, Aa., 1995. Prediction of VLE for mixtures with co-polymers. Ind. Eng. Chem. Res., 34: 324.

48. Lindvig, Th., Hestkjær, L.L., Hansen, A.F., Michelsen, M.L., and Kontogeorgis, G.M., 2002. Phase equi-libria for complex polymer solutions. Fluid Phase Equilib., 663: 194.

49. Harismiadis, V.I. and Tassios, D.P., 1996. Solid-liquid-liquid equilibria in polymer solutions. Ind. Eng. Chem. Res., 35: 4667.

50. Kontogeorgis, G.M., Coutsikos, Ph., Tassios, D.P., and Fredenslund, Aa., 1994. Improved models for the prediction of activity coefficients in nearly athermal mixtures. Part I. Empirical modifications of free-volume models. Fluid Phase Equilib., 92: 35.

51. Coutinho, J.A.P., Andersen, S.I., and Stenby, E.H., 1995. Evaluation of activity coefficient models in prediction of alkane SLE. Fluid Phase Equilib., 103: 23.

52. Polyzou, E.N., Vlamos, P.M., Dimakos, G.M., Yakoumis, I.V., and Kontogeorgis, G.M., 1999. Assessment of activity coefficient models for predicting solid-liquid equilibria of asymmetric binary alkane systems. Ind. Eng. Chem. Res., 38: 316–323.

53. Kontogeorgis, G.M., Fredenslund, Aa., Economou, I.G., and Tassios, D.P., 1994. Equations of state and activity coefficient models for vapor-liquid equilibria of polymer solutions. AIChE J., 40: 1711.

54. Bogdanic, G. and Vidal, J., 2000. A segmental interaction model for liquid-liquid equilibrium calcula-tions for polymer solutions. Fluid Phase Equilib., 173: 241–252.

55. Bogdanic, G., 2001. The FV-UNIQUAC segmental interaction model for liquid-liquid equilibrium calculations for polymer solutions. Part 2. Extension to solutions containing polystyrene. Fluid Phase Equilib., 4791: 1–9.

56. Panagou, E., Vidal, J., and Bogdanic, G., 1998. A segmental interaction model for LLE correlation and prediction—Application to the poly(vinyl alcohol)/water system. Polym. Bull., 40: 117.

57. Pappa, G.D., Voutsas, E.C., and Tassios, D.P., 2001. Liquid-liquid phase equilibrium in polymer-solvent systems: Correlation and prediction of the polymer molecular weight and the pressure effect. Ind. Eng. Chem. Res., 40(21): 4654.

Page 47: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

244 Handbook of Surface and Colloid Chemistry

58. Kontogeorgis, G.M., Saraiva, A., Fredenslund, Aa., and Tassios, D.P., 1995. Prediction of liquid-liquid equilibrium for binary polymer solutions with simple activity coefficient models. Ind. Eng. Chem. Res., 34: 1823.

59. Harismiadis, V.I., van Bergen, A.R.D., Saraiva, A., Kontogeorgis, G.M., Fredenslund, Aa., and Tassios, D.P., 1996. Miscibility of polymer blends with engineering models. AIChE J., 42: 3170.

60. Pappa, G.D., Voutsas, E.C., and Tassios, D.P., 1999. Prediction of solvent activities in polymer solutions with simple group-contribution models. Ind. Eng. Chem. Res., 38: 4975.

61. Lindvig, Th., Michelsen, M.L., and Kontogeorgis, G.M., 2001. Thermodynamics of paint related systems with engineering models. AIChE J., 47(11): 2573–2584.

62. Kontogeorgis, G.M., Nikolopoulos, G.I., Tassios, D.P., and Fredenslund, Aa., 1997. Improved models for the prediction of activity coefficients in nearly athermal mixtures. Part II. A theoretically based GE-model based on the van der Waals partition function. Fluid Phase Equilib., 127: 103.

63. Thorlaksen, P., Abildskov, J., and Kontogeorgis, G.M., 2003. Prediction of gas solubilities in elastomeric polymers for the design of thermopane windows. Fluid Phase Equilib., 211: 17.

64. McQuarrie, D.A., 1976. Statistical Mechanics. Harper Collins, New York. 65. Wertheim, M.S., 1963. Exact solution of Percus-Yevick integral equation for hard spheres. Phys. Rev.

Lett., 10: 321. 66. Carnahan, N.F. and Starling, K.E., 1970. Thermodynamic properties of a rigid-sphere fluid. J. Chem.

Phys., 53: 600. 67. Economou, I.G., 2003. Molecular simulation for industrial applications. In: Kontogeorgis, G. and Gani,

R., eds., Computer Aided Property Estimation for Process and Product Design, Elsevier, Amsterdam, the Netherlands. pp. 279–308.

68. Wertheim, M.S., 1984. Fluids with highly directional attractive forces. I. Statistical thermodynamics. J. Stat. Phys., 35: 19.

69. Wertheim, M.S., 1984. Fluids with highly directional attractive forces. II. Thermodynamic perturbation theory and integral equations. J. Stat. Phys., 35: 35.

70. Wertheim, M.S., 1986. Fluids with highly directional attractive forces. III. Multiple attraction site. J. Stat. Phys., 42: 459.

71. Wertheim, M.S., 1986. Fluids with highly directional attractive forces. IV. Equilibrium polymerization. J. Stat. Phys., 42: 477.

72. Chapman, W.G., Jackson, G., and Gubbins, K.E., 1988. Phase equilibria of associating fluids: Chain molecules with multiple bonding sites. Mol. Phys., 65: 1057.

73. Chapman, W.G., Gubbins, K.E., Jackson, G., and Radosz, M., 1989. SAFT: Equation-of-state solution model for associating liquids. Fluid Phase Equilib., 52: 31.

74. Chapman, W.G., Gubbins, K.E., Jackson, G., and Radosz, M., 1990. New reference equation of state for associating liquids. Ind. Eng. Chem. Res., 29, 1709.

75. Jackson, G., Chapman, W.G., and Gubbins, K.E., 1988. Phase equilibria of associating fluids: Spherical molecules with multiple bonding sites. Mol. Phys., 65: 1.

76. Donohue, M.D. and Prausnitz, J.M., 1978. Perturbed hard chain theory for fluid mixtures—Thermodynamic properties for mixtures in natural-gas and petroleum technology. AIChE J., 24: 849–860.

77. Müller, E.A. and Gubbins, K.E., 2001. Molecular-based equations of state for associating fluids: A review of SAFT and related approaches. Ind. Eng. Chem. Res., 40: 2193.

78. Huang, S.H. and Radosz, M., 1990. Equation of state for small, large, polydisperse and associating molecules. Ind. Eng. Chem. Res., 29: 2284.

79. Huang, S.H. and Radosz, M., 1991. Equation of state for small, large, polydisperse and associating molecules: Extension to fluid mixtures. Ind. Eng. Chem. Res., 30: 1994.

80. Fu, Y.-H. and Sandler, S.I., 1995. A simplified SAFT equation of state for associating compounds and mixtures. Ind. Eng. Chem. Res., 34: 1897.

81. Kontogeorgis, G.M., Voutsas, E.C., Yakoumis, I.V., and Tassios, D.P., 1996. An equation of state for associating fluids. Ind. Eng. Chem. Res., 35: 4310.

82. Michelsen, M.L. and Hendriks, E.H., 2001. Physical properties of association models. Fluid Phase Equilib., 180: 165.

83. Chen, S.S. and Kreglewski, A., 1977. Applications of augmented van der Waals theory of fluids. 1. Pure fluids. Ber. Bunsen Phys. Chem., 81: 1048.

84. Gross, J. and Sadowski, G., 2001. Perturbed-chain SAFT: An equation of state based on a perturbation theory for chain molecules. Ind. Eng. Chem. Res., 40: 1244.

85. Kraska, T. and Gubbins, K.E., 1996. Phase equilibria calculations with a modified SAFT equation of state. 1. Pure alkanes, alkanols and water. Ind. Eng. Chem. Res., 35: 4727.

Page 48: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

245Thermodynamics of Polymer Solutions

86. Kraska, T. and Gubbins, K.E., 1996. Phase equilibria calculations with a modified SAFT equation of state. 2. Binary mixtures of n-alkanes, 1-alkanols and water. Ind. Eng. Chem. Res., 35: 4738.

87. Kolafa, J. and Nezbeda, I., 1994. The Lennard-Jones fluid: An accurate analytic and theoretically-based equation of state. Fluid Phase Equilib., 100: 1.

88. Gil-Vilegas, A., Galindo, A., Whitehead, P.J., Mills, S.J., Jackson, G., and Burgess, A.N., 1997. Statistical associating fluid theory for chain molecules with attractive potentials of variable range. J. Chem. Phys., 106: 4168.

89. McCabe, C., Gil-Vilegas, A., and Jackson, G., 1999. Gibbs ensemble computer simulation and SAFT-VR theory of non-conformal square-well monomer-dimer mixtures. Chem. Phys. Lett., 303: 27.

90. O’Lenick, R., Li, X.J., and Chiew, Y.C., 1995. Correlation functions of hard-sphere chain mixtures: Integral equation theory and simulation results. Mol. Phys., 86: 1123.

91. Gross, J. and Sadowski, G., 2002. Modeling polymer systems using the perturbed-chain statistical asso-ciating fluid theory equation of state. Ind. Eng. Chem. Res., 41: 1084.

92. Tumakaka, F., Gross, J., and Sadowski, G., 2002. Modeling of polymer phase equilibria using the perturbed-chain SAFT. Fluid Phase Equilib., 194–197: 541.

93. Gross, J. and Sadowski, G., 2002. Application of the perturbed-chain SAFT equation of state to associat-ing systems. Ind. Eng. Chem. Res., 41: 5510.

94. Gross, J., Spuhl, O., Tumakaka, F., and Sadowski, G., 2003. Modeling copolymer systems using the perturbed-chain SAFT equation of state. Ind. Eng. Chem. Res., 42: 1266.

95. von Solms, N., Michelsen, M.L., and Kontogeorgis, G.M., 2003. Computational and physical perfor-mance of a modified PC-SAFT equation of state for highly asymmetric and associating mixtures. Ind. Eng. Chem. Res., 42: 1098.

96. Kouskoumvekaki, I.A., von Solms, N., Michelsen, M.L., and Kontogeorgis, G.M., 2004. Application of a sim-plified perturbed chain SAFT equation of state to complex polymer systems. Fluid Phase Equilib., 215: 71–78.

97. von Solms, N., Kouskoumvekaki, I.A., Lindvig, T., Michelsen, M.L., and Kontogeorgis, G.M., 2004. A novel approach to liquid-liquid equilibrium in polymer systems with application to simplified PC-SAFT. Fluid Phase Equilib., 222–223: 87–93.

98. Lindvig, T., Michelsen, M.L., and Kontogeorgis, G.M., 2004. Liquid-liquid equilibria for binary and ternary polymer solutions with PC-SAFT. Ind. Eng. Chem. Res., 43: 1125–1132.

99. von Solms, N., Michelsen, M.L., and Kontogeorgis, G.M., 2005. Prediction and correlation of high-pressure gas solubility in polymers with simplified PC-SAFT. Ind. Eng. Chem. Res., 44: 3330.

100. Haslam, A.J, von Solms, N., Adjiman, C.S., Galindo, A., Jackson, G., Paricaud, P., Michelsen, M.L., and Kontogeorgis, G.M., 2006. Predicting enhanced absorption of light gases in polyethylene using simpli-fied PC-SAFT and SAFT-VR. Fluid Phase Equilib., 243: 74.

101. Jog, P.K., Sauer, S.G., Blaesing, J., and Chapman, W.G., 2001. Application of dipolar chain theory to the phase behavior of polar fluids and mixtures. Ind. Eng. Chem. Res., 40: 4641.

102. Gross, J., 2005. An equation-of-state contribution for polar components: Quadrupolar molecules. AIChE J., 51: 2556–2568.

103. Gross, J. and Vrabec, J., 2006. An equation-of-state contribution for polar components: Dipolar molecules. AIChE J., 52: 1194–1204.

104. Kleiner, M. and Gross, J., 2006. An equation of state contribution for polar components: Polarizable dipoles. AIChE J., 52: 1951–1961.

105. Behzadi, B., Patel, B.H., Galindo, A., and Ghotbi, C., 2005. Modeling electrolyte solutions with the SAFT-VR equation using Yukawa potentials and the mean-spherical approximation. Fluid Phase Equilib., 236: 241–255.

106. Cameretti, L.F., Sadowski, G., and Mollerup, J.M., 2005. Modeling of aqueous electrolyte solutions with perturbed-chain statistical associated fluid theory. Ind. Eng. Chem. Res., 44, 3355–3362.

107. Le Thi, C., Tamouza, S., Passarello, J.P., Tobaly, P., de Hemptinne, J.-C., 2006. Modeling phase equilibrium of H-2+n-alkane and CO2+n-alkane binary mixtures using a group contribution statisti-cal association fluid theory equation of state (GC-SAFT-EOS) with a k(ij) group contribution method. Ind. Eng. Chem. Res., 45: 6803–6810.

108. Kouskoumvekaki, I.A., von Solms, N., Lindvig, T., Michelsen, M.L., and Kontogeorgis, G.M., 2004. Novel method for estimating pure-component parameters for polymers: Application to the PC-SAFT equation of state. Ind. Eng. Chem. Res., 43: 2830–2838.

109. Dobashi, T., Nakata, M., and Kaneko, M., 1984. Coexistence curve of polystyrene in methylcyclohexane. 3. Asymptotic-behavior of ternary-system near the Plait point. J. Chem. Phys., 80: 948–953.

110. Shultz, A.R. and Flory, P.J., 1952. Phase equilibria in polymer-solvent systems. J. Am. Chem. Soc., 74: 4760–4767.

Page 49: Handbook of Surface and Colloid Chemistry Fourth Edition

Dow

nloa

ded

By:

10.

3.98

.104

At:

17:4

7 13

Jan

202

2; F

or: 9

7814

6659

6689

, cha

pter

3, 1

0.12

01/b

1863

3-4

246 Handbook of Surface and Colloid Chemistry

111. Jain, S., Jog, P., Weinhold, J., Srivastava, R., and Chapman, W.G., 2008. Modified interfacial statistical associating fluid theory: Application to tethered polymer chains. J. Chem. Phys., 128: 154910.

112. Bymaster, A. and Chapman, W.G., 2010. An iSAFT density functional theory for associating polyatomic molecules. J. Phys. Chem. B, 114: 12298.

113. Tamouza, S., Passarello, J.P., Tobaly, P., de Hemptinne, J.C., 2004. Group contribution method with SAFT EOS applied to vapor liquid equilibria of various hydrocarbon series. Fluid Phase Equilib., 222–223: 67.

114. Grandjean, L., de Hemptinne, J.-C., and Lugo, R., 2014. Application of GC-PPC-SAFT EoS to ammonia and its mixtures. Fluid Phase Equilib., 367: 159.

115. Tihic, A., von Solms, N., Michelsen, M.L., Kontogeorgis, G.M., and Constantinou, L., 2009. Application of sPC-SAFT and group contribution sPC-SAFT to polymer systems—Capabilities and limitations. Fluid Phase Equilib., 281: 70.

116. Tihic, A., von Solms, N., Michelsen, M.L., Kontogeorgis, G.M., and Constantinou, L., 2009. Analysis and applications of a group contribution sPC-SAFT equation of state. Fluid Phase Equilib., 281: 60.

117. Tihic, A., Kontogeorgis, G.M., von Solms, N., and Michelsen, M.L., 2008. A predictive group- contribution simplified PC-SAFT equation of state: Application to polymer systems. Ind. Eng. Chem. Res., 47: 5092.

118. Peng, Y., Goff, K.D., dos Ramos, M.C., McCabe, C., 2010. Predicting the phase behavior of polymer systems with the GC-SAFT-VR approach. Ind. Eng. Chem. Res., 49: 1378.

119. Clark, G.N.I., Galindo, A., Jackson, G., Rogers, S., and Burgess, A.N., 2008. Modeling and under-standing closed-loop liquid-liquid immiscibility in aqueous solutions of poly(ethylene glycol) using the SAFT-VR approach with transferable parameters. Macromolecules, 41: 6582.

120. Sanchez, F.A., Pereda, S., and Brignole, E.A., 2011. GCA-EoS: A SAFT group contribution model-extension to mixtures containing aromatic hydrocarbons and associating compounds. Fluid Phase Equilib., 306: 112.

121. Peters, F.T., Laube, F.S., Sadowski, G., 2012. Development of a group contribution method for polymers within the PC-SAFT model. Fluid Phase Equilib., 324: 70.

122. Papaioannou, V., Lafitte, T., Avendano, C., Adjiman, C.S., Jackson, G., Muller, E.A., and Galindo, A., 2014. Group contribution methodology based on the statistical associating fluid theory for heteronuclear molecules formed from Mie segments. J. Chem. Phys., 140: 054107.

123. Cameretti, L.F. and Sadowski, G., 2008. Modeling of aqueous amino acid and polypeptide solutions with PC-SAFT. Chem. Eng. Proc., 47: 1018.

124. Kozlowska, M.K., Jurgens, B.F., Schacht, C.S., Gross, J., and de Loos, T.W., 2009. Phase behavior of hyperbranched polymer systems: Experiments and application of the perturbed-chain polar SAFT equa-tion of state. J. Phys. Chem. B, 113: 1022.

125. Vargas, F.M., Gonzalez, D.L., Hirasaki, G.J., and Chapman, W.G., 2009. Modeling asphaltene phase behavior in crude oil systems using the perturbed chain form of the statistical associating fluid theory (PC-SAFT) equation of state. Energy Fuels, 23: 1140.

126. von Solms, N. and Kristensen, J., 2010. Refrigeration plants using carbon dioxide as refrigerant: Measuring and modelling the solubility, diffusivity and permeability of carbon dioxide in polymers used as packing and sealing materials. Int. J. Refrig., 33: 19.

127. Neela, V. and von Solms, N., 2014. Permeability, diffusivity and solubility of carbon dioxide in fluoropolymers: An experimental and modeling study. J. Polym. Res., 21: 401.

128. Krallis, A. and Kanellopoulos, V., 2013. Application of Sanchez-Lacombe and perturbed-chain statisti-cal associating fluid theory equation of state models in catalytic olefins (co)polymerization industrial applications. Ind. Eng. Chem. Res., 52: 9060.

129. Tan, S.P., Adidharma, H., and Radosz, M., 2008. Recent advances and applications of statistical associat-ing fluid theory. Ind. Eng. Chem. Res., 47: 8063.

130. Sadowski, G., 2011. Modeling of polymer phase equilibria using equations of state. In: Enders, S. and Wolf, B.A., eds. Polymer Thermodynamics: Liquid Polymer-Containing Mixtures. Advances in Polymer Science, Vol. 238, p. 389, Springer-Verlag, Berlin, Heidelberg.

131. Kleiner, M., Turnakaka, F., and Sadowski, G., 2009. Thermodynamic modeling of complex systems. In: Lu, X. and Hu, Y., eds. Molecular Thermodynamics of Complex Systems. Structure and Bonding, Vol. 131, p. 75, Springer-Verlag, Berlin, Heidelberg.

132. Tanbonliong, J.O. and Prausnitz, J.M., 1997. Vapour-liquid equilibria for some binary and ternary poly-mer solutions. Polymer, 38: 5775.

133. Matsumara, K. and Katayama, T., 1974. Katayama, T. Mateumura, K. and Urahama, Y.1971. Measurement of Vapor-Liquid Equilibria of Binary and, ternary Solutions containing polystyrene as a Component. Kagaku Kogaku, 38: 388.

134. Tanbonliong, J.O. and Prausnitz, J.M., 1997. Vapor-Liquid Equilibria of Binary and Ternary Solutions Containing Polyvinylacetate as a Component. Polymer, 38: 5775.