Está en la página 1de 505

INSTITUT FRANCAIS DU PETROLE PUBLICATIONS

Jean VlDAL
Associate Director of Research at IFP
Professor at IFP School

THERMODYNAMICS
APPLICATIONS IN CHEMICAL ENGINEERING
AND THE PETROLEUM INDUSTRY
Translated from the French by Thomas S. Pheney
and Eileen M. McHugh, TCNY

2003

t Editions TECHNIP

27 rue Cinoux, 75737 PARIS Cedex IS,FRANCE

FROM THE SAME PUBLISHER

Rtroleum Refining Series


1. Crude Oil. Petroleum Products. Process Flowsheets J.P.
2. Separation Processes J.P. WAUQUIER, Ed.
3. Conversion Processes P. LEPRINCE, Ed.
4. Materials and Equipment P. TRAMBOUZE,Ed.
5. Refinery Operation and Management J.P. FAVENNEC, Ed.

WAUQUIER,

Ed.

Petrochemical Processes.Technical and Economic Characteristics.


1. Synthesis-Gas Derivatives and Major Hydocarbons
2. Major Oxygenated, Chlorinated and Nitrated Derivatives
A. CHAUVEL, C. LEFEBVRE

Technology of Catalytic Oxidations (The)


1. Chemical, Catalytic and EngineeringAspects
2. Safety Aspects
P. ARPENTINIER, F. CAVANI, F. TRlFlRb

Scale-up Methodology for Chemical Processes


J.P. EUZEN, P. TRAMBOUZE, J.P. WAUQUIER

Computational Fluid DynamicsApplied to Process Engineering


P. TRAMBOUZE,

Ed.

Petroleum Process Thermodynamics

E. BEHAR,

Ed.

Permeability of Gases in Polymer Materials M.H. KLOPFFER


Chemical Reactors. Design. Engineering. Operation
P. TRAMBOUZE, H. VAN LANDECHEM, J.P. WAUQUIER

Combustion and Flames. Chemical and Physical Principles


R. BORCHI, M. DESTRIAU

Principlesof Turbulent Fired Heat G.

MONNOT

Translation of Thermodynamique.Application
au g h i e chimique et B Iindustrie p&roli&e. J. Vidal
0 1997, Editions Technip, Paris

0 2003, EditionsTechnip, Paris


All rights reserved. No part of this publication may be reproduced or
transmitted in any form or by any means, electronic or mechanical,
includingphotocopy, recording, or any informationstorage and retrieval
system, without the prior written permission of the publisher.

ISBN 2-7108-0800-5

Preface

The petroleum industry has such an extensive range of application for the principles of
thermodynamics that a study such as this one could not possibly cover the subject in its
entirety. For catagenesis,successive migration in the formation of petroleum fluids, reservoir exploitation, transport of natural gas or crude oils, refining and petrochemical
processes,or energy applications,we avail ourselves of existing equations for properties as
diverse as density, energy, and equilibrium conditions between phases. These equations
were developed during what is called the golden age of classic thermodynamics.We are
left with finding the most appropriate way to apply them.
We are faced with many obstacles.
First of all, the composition of petroleum fluids is poorly understood. At best, we are
aware of their complexity.The sheer number of components and the poorly defined structure of some of these components requires simplifications,which are more or less justifiable. The pressure and temperature conditions of some natural gas or light oil deposits
place these fluids close to their critical conditions. In addition, we note the particularly
high pressure levels of recently discovered fields.
The treatment of natural gas, the separation by extractive distillation or liquid-liquid
extraction, the synthesis of compounds such as ethers that are used in place of tetraethyl
lead in the formulation of gasoline,and the very diversified field of petrochemistry lead to
the treatment of mixtures in which hydrocarbons and compoundswith heteroatomic structures coexist, and give rise to more complex molecular interactions.
The simulation of processes and their optimization assumes that the properties of the
mixtures concerned are known. Although this knowledge is still based on experimental
measurements for the most part, it is also the result of calculation methods that have been
developed. These methods owe their value to the laws of thermodynamics, which assure
them a wide range of application. For example, an equation of state allows of course for
the density calculation of a fluid as a function of pressure, temperature, and composition,
but it also allows for the calculation of phase change conditions,and the energy exchanges
that result from the imposed transformations.However, these methods are still approximations, and remain to be perfected.The need for advancement makes thermodynamics a
living discipline. It is based on the relationships that exist between experimentation and
the notion of models.The engineer must keep abreast of progress and evaluate it intelligently.
For the most part, this study is dedicated to the description of our methods for the calculation and prediction of the thermodynamic properties of the processed fluids during
reservoir exploitation,refining, or petrochemical processing. The evolution of these methods that we have just emphasized, the limits of our experience, and the need to choose,
make for numerous deficiencies.We offer only an introduction to the literature in the field,
which is treated in a far from exhaustive manner. The proliferation of literature proposing

VI

Preface

new models, or sometimes real improvements to existing methods, is such that the bibliographical references provided are incomplete, or, even worse, the methods described in
this text probably will be abandoned. However, we believe that a discussion of these methods will facilitate understanding of future developments. At the end of this preface, we
mention a number of books that were the source of constant inspiration during our
research, and that have, more or less knowingly,influenced the writing of this study.
Some sections are merely necessary updates of an older text, published over twenty
years ago [Vidal, 1973,19741.The chapters dealing with phase equilibria duplicate, in part,
a text that is published in a more general work on petroleum refining. We have made no
attempt to introduce unrealistic modifications in this text, but we have, of course,
expanded it.
While preparing the text, we were very much inspired by our teaching experience at the
&Cole Nationale Suptrieure du Pttrole et des Moteurs (ZFP-School),and by the questions
from students.We may hope that this very same teaching has benefited from our research
work. It is certainly at the source of some of its outcome.
Suffice it to say that the use of different conventionswithin the field gives rise to endless
discussions about nomenclature and notation. As much as possible, we have conformed to
the most common usage. For the units, we have taken some liberties with the International
System, using the Celsius and Kelvin scale, and preferring to express pressures in bar,
which is much more day-to-day than the pascal and its multiples. Only rarely have we
retained data in the Anglo-Saxon unit systems.
I would like to thank several of my colleagues who have contributed to this study to
varying degrees: MUeA. Boutrouille, MM. J. Ch. de Hemptinne, C1. Jaffret and B. Tavitian,
and especially M.L. Asselineau, whose knowledge, experience, and friendship I valued
each day.
Not to have dedicated a section, no matter how brief, to polymer solutions would have
been an important omission. I owe thanks to Mme G. Bogdanic (INA R&D) for having
instructed me in this area and guided the writing of the chapter, which is dedicated to it.
Finally, I wish to acknowledge Professor Renon (&Cole des Mines de Paris) who welcomed me into the research group that he formed at the Znstitut Franeais du Pttrole (ZFP),
and inspired my first works in the field of calculating thermodynamic properties, and
Professor PCneloux (Universitt dAix-Marseille ZZ) with whom I enjoyed constant and
fruitful exchanges.

Editions Technip would like to give special thanks to Mr. Jean-Charles de Hemptinne
from Division Chimie physique of Znstitut franeais du pttrole for participating in the
realisation of this book with his valuable comments, suggestions and corrections on the
text.

Preface

VI I

REFERENCES
Lewis GN, Randall M, Pitzer KS, Brewer L (1961) Thermodynamics,2nd edition. McGraw-Hill, New
York.
Prigogine I, Defay R (1950) Thermodynamique chimique. Desoer, LiBge.
Pkneloux A, Cours de Thermodynamique,Universitk d'Aix-Marseille.
Prausnitz JM, Lichtenthaler RN, de Azevedo EG (1986) Molecular Thermodynamics of Fluid-Phase
Equilibria. Prentice-Hall, Englewood Cliffs, New Jersey.
Reid RC, Prausnitz JM, Sherwood TZlK (1977) The Properties of Gases and Liquids, 3rd edition.
McGraw-Hill, New York.
Reid RC, Prausnitz JM, Poling BE (1987) The Properties of Gases and Liquids, 4th edition. McGrawHill, New York.
Rowlinson JS, Swinton FL (1982) Liquids and Liquid Mixmres. Butterworth, London.
Sandler SI (1989) Chemical Engineering Thermodynamics.Wiley, New York.
Tassios DP (1993) Applied Engineering Thermodynamics.Springer-Verlag,Berlin.
Vidal J (1973, 1974) Thermodynamique. Mtthodes appliqutes au raffinage et au gtnie chimique.
Editions Technip,Paris.

Table of Contents

Preface ..........................................................................................................................
Symbols ........................................................................................................................

V
XIX

Principles Thermodynamic Functions The Ideal Gas


1.1 Definitions .............................................................................................................
1.2 The First Law ........................................................................................................

1.2.1 The Energy of a System............................................................................................


1.2.2 Energy Exchanges during a Transformation....................................................
1.2.3 Statement of the First Law Applied to a Closed System..................................

..................................................................................................
.........................................................................
............................................................................
.........................................................
1.3 Application of the First Law to an Open System..............................................
1.2.3.1
1.2.3.2
1.2.3.3
1.2.3.4

General
IsochoricTransformations
Adiabatic Compression
Transformationsat Constant Pressure

1.3.1 General ...........................................................................................................


1.3.2 Steady-StateSystems ......................................................................................

1.4 The Second Law ....................................................................................................


1.4.1
1.4.2
1.4.3
1.4.4

Entropy ...........................................................................................................
Relationshipbetween Internal Energy and Entropy .......................................
Application of the Equilibrium Condition ......................................................
Statistical Significanceof Entropy ...................................................................

1.5 Helmholtz Energy and Gibbs Energy ................................................................


1.6 ThermodynamicFunction:Internal Energy, Enthalpy, Entropy,
Helmholtz Energy and Gibbs Energy ................................................................
1.6.1 Dependence on Temperature,Volume,or Pressure .........................................
1.6.2 CharacteristicFunctions..................................................................................

6
8
8
9
10
11
12
12
19
19

22

1.7 The Ideal Gas ........................................................................................................

23
23
25
26

References....................................................................................................................

26

1.7.1 Equation of State and Thermodynamic Properties ..........................................


1.7.2 Heat Capacity .................................................................................................
1.7.3 DataTables .....................................................................................................

Table of Contents

Properties of Pure Substances

..

2.1 The Relationship between Pressure.Volume and Temperature


Liquid-Vapor E q ~ i b rurn
i ...................................................................................
2.2 Vapor Pressure ......................................................................................................

2.2.1 Liquid and Vapor States ..................................................................................


2.2.2 Vapor Pressure Equations ...............................................................................

.................................................................................
.............................................................................
2.3 Enthalpy Diagram and Heat of Vaporization ...................................................
2.3.1 Dependence of Enthalpy on Pressure and Temperature..................................
2.3.2 Heat of Vaporization.......................................................................................
2.4 Calculation ofThermodynamic Properties .......................................................
2.4.1 Residual Enthalpy...........................................................................................
2.4.2 Residual Gibbs Energy ...................................................................................
2.4.3 Fugacity .........................................................................................................
2.4.4 Calculation of Thermodynamic Properties in the Liquid Phase ......................
2.4.5 General Equations .........................................................................................
Conclusion....................................................................................................................
References ....................................................................................................................
2.2.2.1 Clapeyron Equation
2.2.2.2 Empirical Correlations

29
34
34
36
36
37

44
44
46
49
50
51
53
54
58
59
59

Predicting Thermodynamic Properties of Pure Substances


General Principles Corresponding States
Group Contributions

3.1 Techniques of Molecular Simulation..................................................................

63

3.2 The Corresponding States Principle...................................................................

64
69
69
70
71
72
74
75
81
81
82
83

3.2.1 Correlations Using the Critical CompressibilityFactor ...................................

.......................................................................................
....................................................................................
3.2.2 Correlations Using the Acentric Factor ...........................................................
3.2.2.1 Prediction of the Second Virial Coefficient ..................................................
3.2.2.2 Properties at Liquid-Vapor Equilibrium......................................................
3.2.2.3 Lee and Kesler Method ............................................................................
3.2.3 Extensions to the CorrespondingStates Principle ...........................................
3.2.3.1 Extension to Polar Compounds .................................................................
3.2.3.2 Extension to Mixtures ..............................................................................
3.2.4 Conclusion Concerning the CorrespondingStates Principle............................
3.2.1.1 Watson Method
3.2.1.2 Rackett Equation

Table of Contents

3.3 Structure Property Correlations........................................................................

3.3.1 Properties of the Ideal Gas .............................................................................


3.3.2 Critical Coordinates ........................................................................................
3.3.3 Calculation of Molar Volume in the Liquid Phase ...........................................

3.4 Examples of the Relationships Between ThermodynamicProperties...........


3.4.1 Calculation of Critical Properties from Vapor Pressure and Density Data ......
3.4.2 Calculation of the Heat of Vaporization: Watson Equation .............................
3.4.3 Empirical Equations Developed from the Normal Boiling Point and Density .

Conclusion....................................................................................................................
References ....................................................................................................................

XI

84
85
88
90
90
90
92
93
94
95

Equations of State
4.1 Equations of State Derived from the Vial Development ..............................
4.1.1
4.1.2
4.1.3
4.1.4

Volume Virial Equation of State Truncated after the Second Term .................
Volume Virial Equation of State Truncated after the Third Term ....................
Pressure Virial Equation of State Truncated after the Second Term ................
The Benedict. Webb. and Rubin Equation .......................................................

4.2 Equations of State Derived from the Van Der Waals Theory .........................

4.2.1 The Soave-Redlich-Kwong and Peng-Robinson Equations of State ................


4.2.2 Recent Developments of Cubic Equations of State .........................................
4.2.2.1 Dependence of Attraction Parameter a on Temperature ..............................
4.2.2.2 Modifications of the AttractionTerm
4.2.2.3 Application of the Concept of Group Contribution ......................................
4.2.2.4 Equations of State for Rigid Spheres and Hard Chains..................................

..........................................................

102
105
106
106
111
112
113
124
125
126
130
134

4.3 Specific Equations of State for Certain Pure Substances................................

136

4.4 The Tait Equation.................................................................................................


References ....................................................................................................................

138
139

Characterization of Mixtures
5.1 Partial Molar Values in the Homogeneous Phase ............................................

144
144
145

5.2 Chemical Potential................................................................................................

149
149
150

5.1.1 Definitions. Main Equations ............................................................................


5.1.2 Determination of Partial Molar Values ...........................................................
5.2.1 Definition .......................................................................................................
5.2.2 Equilibrium Condition Between Phases ..........................................................

XI1

Table of Contents

5.2.3 Relationshipsbetween the Chemical Potential and the Other Thermodynamic


151
Properties .......................................................................................................

.....................................................................

5.2.3.1 The Gibbs-Duhem Equation


5.2.3.2 Dependence of Chemical Potential on Pressure and Temperature
5.2.3.3 Relationships between the Chemical Potential and the Other Thermodynamic.
Functions

...................

...............................................................................................
5.3 Fugacity..................................................................................................................
5.3.1 Definition .......................................................................................................
5.3.2 Dependence of Fugacity on Temperature, Pressure. and Composition ............
5.4 Mixing Values Activity .........................................................................................
5.4.1 Definitions ......................................................................................................
5.4.2 Dependence of Activity on Temperature. Pressure, and Composition .............
5.5 The Ideal Solution ................................................................................................
5.6 Calculation of Fugacity ........................................................................................
5.7 Excess Values and Activity Coefficients............................................................
5.7.1 Definitions .......................................................................................................
5.7.2 Dependence of Excess Values on Temperature,Pressure. and Composition ....
5.7.3 Activity Coefficients .......................................................................................
5.7.4 Dependence of Activity Coefficients on Temperature,Pressure, and
Composition ...................................................................................................
5.8 Comparison of ' h o Methods for Calculating Fugacity....................................
5.9 Asymmetric Convention:the Henry Constant..................................................
Reference......................................................................................................................

151
151
152
152
152
153
154
154
155
156
157
159
159
160
160
161
163
165
166

6
Mixtures: Liquid-Vapor Equilibria
6.1 Description of the Vaporization or Condensation Phenomena......................

6.1.1 Isobaric Liquid-Vapor Equilibrium Diagram ..................................................


6.1.2 Isothermal Liquid-Vapor E uilibrium Diagrams.Evolution with Temperature
Critical Point and Retrogra e Condensation...................................................
6.1.3 Azeotropic Systems.........................................................................................

6.2 The Liquid-Vapor Equilibrium Condition The Equilibrium Coefficient .....

..

6.3 Dependence of the Equilibrium Conditions on Temperature,Pressure,


and Composibon...................................................................................................

6.3.1 Dependence of Bubble Pressure on Composition ...........................................


6.3.2 Dependence of Bubble Pressure on Temperature. Clapeyron Equation
Applied to a Mixture.......................................................................................
6.3.3 Coherence Tests ..............................................................................................
6.3.4 Stability and Critical Point Conditions ............................................................

168
168
171
176
179
184
184
186
187
188

Xlll

Table of Contents

6.4 Liquid-Vapor Equilibrium Problems .................................................................

190
192
192
192

6.5 Calculation Algorithms ........................................................................................

192
193
193
194
194
195
195
195
196
201
201
201
203

6.4.1 At Given Temperature (or Pressure) and Vaporized Fraction .........................


6.4.2 At Given Temperature and Pressure ...............................................................
6.4.3 Case Where One of the Data is a ThermodynamicProperty ...........................
6.5.1 Calculation of the Bubble Point ......................................................................

............................................................
.......................................................
6.5.2 Calculation of the Dew Point ..........................................................................
6.5.2.1 Calculation of the Dew Pressure ................................................................
6.5.2.2 Calculation of the Dew Temperature ..........................................................
6.5.3 Partial Vaporization ........................................................................................
6.5.4 Application to Ideal Solutions.........................................................................
6.5.5 Non-Ideal Solutions ........................................................................................
6.5.5.1 Non-Ideal Solutionsat Low Pressure ..........................................................
6.5.5.2 General Case ..........................................................................................
6.5.6 General Calculation Method of Liquid Vapor Equilibria ................................
6.6 Solubility of Gases in Liquids..............................................................................
References ....................................................................................................................
6.5.1.1 Calculation of the Bubble Pressure
6.5.1.2 Calculation of the Bubble Temperature

204
206

Deviations from Ideality in the Liquid Phase

..

7.1 Excess Quanbties..................................................................................................

7.1.1 Excess Volume. Excess Heat Capacity.............................................................


7.1.2 Heat of Mixing ................................................................................................
7.1.3 Excess Gibbs Energy and Activity Coefficients...............................................

210
210
212
214

7.2 Correlation of Liquid Vapor Equilibria at Low Pressure Coherence Test ...

218

7.3 Influence of Varying Molar Volume: the Combinatorial Term .......................

223

7.4 The Concept of Local Composition ...................................................................

227
229
229
230

7.5 Regular Solutions..................................................................................................

231

7.6 Empirical Models Based on the Concept of Local Composition....................

238
238
239
244
246

7.4.1 The.Lattice Model...........................................................................................


7.4.2 The Quasi Chemical Model.............................................................................
7.4.3 General Remarks............................................................................................

7.6.1
7.6.2
7.6.3
7.6.4

The Wilson Equation.......................................................................................


The NRTL Equation .......................................................................................
The UNIQUAC Model ...................................................................................
The Wilson, NRTL. UNIQUAC Models.Conclusion .......................................

XIV

Table of Contents

7.7 Group ContributionMethods .............................................................................

7.7.1 The ASOG Method .........................................................................................


7.7.2 The UNIFAC Method ......................................................................................
7.7.3 Group Contribution Method .Conclusion .......................................................

248
250
251
257

7.8 Associated Solutions ............................................................................................

258

7.9 Ionic Solutions.......................................................................................................

263

References....................................................................................................................

265

Application of Equations of State to Mixtures


Calculation of Liquid-Vapor Equilibria Under Pressure
8.1 Extensions ofthe Corresponding States Principle ...........................................

8.1.1 Calculation Rules for Pseudocritical Points .....................................................


8.1.2 Calculation of Thermodynamic Properties and Fugacity Coefficients in a
Mixture ...........................................................................................................

270
271
275

8.2 Virial Equations of State lhncated after the Second Term............................

277

8.3 Equations of State Derived from the vm der Wads The0ry...........................

279
279
282
283
286
290
290
291
292

8.3.1
8.3.2
8.3.3
8.3.4
8.3.5

The Classical Mixing Rules .............................................................................


Calculation of Chemical Potentials and Fugacity Coefficients .........................
Application Range and Results .......................................................................
The Binary Interaction Parameter ...................................................................
Alternatives on the Classical Mixing Rules .....................................................

.............................
.............................
8.3.6 Calculation of the Thermodynamic Properties of the Mixture .........................
Mixing Rules and Excess Functions ...................................................................
8.3.5.1 Dependence of the Attraction Parameter on Composition
8.3.5.2 Application of a Quadratic Mixing Rule to the Covolume

8.4

8.4.1 Calculation of Excess Quantities Using Equations of State: The Problem


of Reference States .........................................................................................
8.4.2 Mixing Rules Derived from Excess Gibbs Energy at Infinite Pressure ............
8.4.3 Mixing Rules and Excess Functions at Constant Packing Fraction ..................
8.4.3.1 Formulation of Equations of State Derived from the van der Waals Theory
in Terms of Packing Fraction
8.4.3.2 Calculation of the Helmholtz Energy A
8.4.3.3 Application to a Mixture and its Components
8.4.3.4 Results: Abdoul Group Contributions Method

8.4.4
8.4.5
8.4.6

.....................................................................
......................................................
..............................................
.............................................
The MHV2Method .....................................................................................
The Wong and Sandler Method .......................................................................
Advantages and Disadvantages of Mixing Laws Derived from Models and
Excess Functions .............................................................................................

292
293
298
302
302
303
304
307
309
313
315

XV

Table of Contents

8.5 Calculation of Liquid-Vapor Equilibria.............................................................

8.5.1 Newton Method ..............................................................................................


8.5.2 Tangent Plane Method ....................................................................................

316
319
321

Conclusion....................................................................................................................

324

References ....................................................................................................................

324

Liquid-Liquid and Liquid-Liquid-VaporEquilibria


9.1 Liquid-Liquid Equilibria and Deviations from Ideality...................................

330

9.2 General Description of Liquid-Liquid Equilibria.............................................

331
331
334

9.3 Selectivity ofthe Liquid-Liquid Equilibrium ....................................................

337

9.4 Liquid-Liquid-VaporEquilibria..........................................................................

339

9.5 Calculation Methods ............................................................................................

345

9.6 Water, Hydrocarbon Systems ..............................................................................

348
348

9.2.1 Binary Systems................................................................................................


9.2.2 Ternary Systems..............................................................................................

9.6.1 Total Immiscibility Hypothesis: Calculation of the Three-Phase Equilibrium ..


9.6.2 Application of Equations of State to the Calculation of Phase Equilibria for
Water Hydrocarbon Systems...........................................................................

351

Conclusion....................................................................................................................

352

References ....................................................................................................................

352

10

Fluid-Solid Equilibria Crystallization Hydrates


10.1 Liquid-Solid Equilibrium Diagram ..................................................................

356

10.2 Calculation of CrystallizationEquilibria .........................................................

359
359
363
365

10.3 Hydrates ..............................................................................................................

10.3.1 Generalities................................................................................................
10.3.2 Phase Diagrams..........................................................................................
10.3.3 Calculation of Hydrate Formation Equilibria .............................................

367
367
368
370

References ....................................................................................................................

372

10.2.1 General Equations .....................................................................................


10.2.2 Paraffin Crystallization...............................................................................
10.2.3 Fluid-Solid Phase Transition at High Pressure ............................................

XVI

Table of Contents

11

Polymer Solutions and Alloys


11.1 Polymers in Solution ..........................................................................................

377
377
379
382
383
386
389

11.2 Polymer Mixtures ...............................................................................................


Conclusion....................................................................................................................

389

References ....................................................................................................................

392

11.1.1 The Flory-Huggins Model ..........................................................................


11.1.2 The Influence of Free Volume ....................................................................
11.1.3 The Entropic-FV Model .............................................................................
11.1.4 The GC-Flory Model ..................................................................................
11.1.5 The GCLF Equation of State (Group Contribution Lattice Fluid) .............
11.1.6 Extension to Liquid-Liquid Equilibria........................................................

392

12

Multicomponent Mixtures
12.1 Pseudocomponents............................................................................................
12.1.1
12.1.2
12.1.3
12.1.4

Complex Mixture Analysis .........................................................................


Lumping.....................................................................................................
Thermodynamic Properties of Pseudocomponents.....................................
Representing the Heavy Fraction of Natural Gases ....................................

12.2 ContinuousThermodynamics...........................................................................

12.2.1 Definition ...................................................................................................


12.2.2 Chemical Potential, Fugacity Coefficient.and Equilibrium Condition
Between Phases ..........................................................................................
12.2.3 Application Examples ................................................................................

.......................................
...............................................................................
...............................................
Conclusion....................................................................................................................
References....................................................................................................................
12.2.3.1 Liquid-Vapor Equilibrium in an Ideal Solution
12.2.3.2 Excess Gibbs Energy of a Polymer Solution in Semicontinuous
Thermodynamics
12.2.3.3 Retrograde Condensation of a Natural Gas. Application of the
Soave-Redlich-KwongEquation of State

396
396
398
400
402
404
404
405
406
406
408
409
411
411

13

Chemical Reactions
13.1 ThennochemicalData .......................................................................................

13.1.1 Standard Enthalpy of Formation. Standard Gibbs Energy of Formation ....


13.1.2 Application of Group Contribution Methods .............................................

414
414
416

Table of Contents

13.1.3 CoherentEnthalpy Data .........................................................................


13.1.4 Standard Enthalpy and Gibbs Energy of Reaction .....................................
13.1.4.1 Definition and Calculation from Standard Enthalpies and Gibbs Energies
of Formation .....................................................................................
13.1.4.2 Dependence of Enthalpy of Reaction and Gibbs Energy of Reaction
on Temperature .................................................................................

XVI I

419
419
419
420

13.2 Heat of Reaction and Energy Balance ............................................................


13.2.1 Heat of Reaction ........................................................................................
13.2.2 Energy Balance of a Reactor or a Reacting Section....................................

421
421
424

13.3 Chemical Equhbria ...........................................................................................


13.3.1 The Equilibrium Condition ........................................................................
13.3.2 The Law of Mass Action .............................................................................
13.3.3 The Laws of Equilibrium Displacement .....................................................

425
425
426
429

13.4 Calculationof Simultaneous Chemical Equilibria .........................................

431

References....................................................................................................................

433

Appendix 1
Database .......................................................................................................................

435

Appendix 2
Lee and Kesler Method .Compressibility Factor. Residual Terms for Enthalpy.
Entropy. and Heat Capacity at Constant Pressure. and Fugacity Coefficient ......

443

Appendix 3
Surface Volume and Interaction Parameters Applied in the UNIFAC Method ..

477

Appendix 4
Properties of the Ethane (1) Propane (2) System at 45Cand at 2.5 MPa as a
Function of Composition ...........................................................................................

479

Appendix 5
Detailed Analysis of a Straight-Run Gasoline Cut .................................................

483

Appendix 6
Units .............................................................................................................................

487

INDEX .........................................................................................................................

489

..

Principles
Thermodynamic Functions
The Ideal Gas

It would be inaccurate to say that the development of thermodynamics is not based on


experimentation. It is from observation that the concepts of quantity of heat, of temperature, of energy of a system, and of irreversibility have been developed. However, there was
an important turning point when the statements of the first and second laws allowed us not
only to condense the observations that produced them, but also to establish apriori laws
that subsequent observations have verified, to generalize from them and diversify the field
of their application. The greatest diversity exists, however, within the field of thermodynamics. Mechanical engineers, energy specialists, and chemists apply the same principles,
but they have constructed their own conventions reflecting their practical concerns.
Many statements of these principles can be set forth and it is not the goal of this chapter
to enumerate them or to demonstrate their equivalence. For the most part, they are familiar to the reader who may, if he so wishes, consult the general references listed in the bibliography. In this chapter, we will limit ourselves to a few topics emphasizing the points that
seem most important to us for the rest of this study.
First, we must review the terminology proper to thermodynamics. We derive our inspiration from the teachings of A. PCneloux [1992].

1.1

DEFINITIONS

Thermodynamics applies to a physical entity, the system,possibly composed of distinct


parts, or subsystems. The system is defined only if its physical limitations or boundaries are specified, as well as the nature of the exchanges that it may maintain with the
rest of the universe (the surroundings). The system is termed closedor openaccording to whether or not exchanges of matter are possible. For example, its properties may
change due to differences of temperature or pressure between the system and the surroundings. Its boundaries on the other hand may resist such transformations. The energy
exchanges with the surroundings will therefore have to be specified according to the
exchanges of matter. Similarly, certain changes may be excluded by virtue of internal

1. Principles. Thermodynamic Functions. The Ideal Gas

constraints; first and foremost, those changes defining the possible boundaries of subsystems that may be fixed or mobile, adiabatic or diathermic, impermeable or porous. It is also
well known that certain chemical reactions occur only with a catalyst or an initiator, that
other reactions may be inhibited and, for example, that we may apply the laws of thermodynamics to the solubility of air in hydrocarbons without being concerned about the possibility of combustion. As with boundaries, these constraints must also be specified in order
to describe a system and the changes that we may expect or exclude.
The most natural subsystems that we can define are made up of the phases into which
matter is organized. In particular, we shall study the equilibria between the liquid and
vapor phases. A phaseforms a homogeneousphysical entity in the sense that all the
parts of equal volume have the same properties (they have the same quantity of matter,
the same composition, etc.).
With a defined system, in principle we may describe its stateby determining its properties, or state values, meaning the entirety of what is observable: temperature, pressure, volume, quantity of matter, and composition, for example. These properties are not
totally independent, and we know very well that the volume occupied by a system is fixed
from the moment we know the quantities of each component, the pressure, and the temperature.
Within the scope of this text, these properties can be defined only for steady-statesystems whose state does not change over the course of time (and whose state does not
depend on its development). In fact, this steady-state is often due to the existence of
appropriate boundaries or constraints, and insofar as these boundaries are precisely
described, we will state that the system is in a state of equilibrium.This means that the
system will return to this state after any infinitesimal disturbance with respect to its boundaries. A state of equilibrium depends on the constraints imposed on the system; constraints
which we will take into account when defining and applying the conditions of equilibrium
stated in the laws of thermodynamics.
Some of these properties are additive in the sense that if we naturally or artificially
divide the system into several parts, such a property of the whole is calculated by using the
sum of the values of this property in each of the parts. For a homogeneous system, this
value is proportional to the size of the system, namely the quantity of matter. Volume is the
simplest example of such a property, but it is the same for internal energy, entropy, etc.,
which we will define later. Such properties are termed extensive and we shall consistently denote them by the capital letter symbols V U, S, etc.
On the other hand, other properties in a homogeneous system are independent of the
size of the system: density, pressure, temperature, etc. They are termed intensiveand on
this subject, we recall the definition given by PCneloux [1992]: phase refers to the entirety
of intensive properties. Among these properties are some that govern the equilibrium
among the various parts of a system. If two parts are separated by a mobile boundary, the
pressures must be the same in each part. It is the same with temperatures if the boundary
is diathermic, and with the chemical potential (which we shall define in Chapter 5 ) of each
component if the boundary does not resist exchanges of matter. Such intensive properties
are potentials and intervene in conjunction with corresponding extensive properties:
volume, entropy, and quantities of matter. There are other intensive properties defined by

1. Principles. fiermodynamic Functions. The Ideal Gas

the value taken by an extensive property for the unit of matter (one mole): molar volume,
molar heat capacity,etc. These are sometimes referred to as densities.We shall give them
the same symbols,but in lower case: v, molar volume, u, internal molar energy, etc.
If certain constraints that assure the equilibrium of a system are removed, the system
generally undergoes a transformation, and changes via a series of intermediate states to
a new state of equilibrium. This change occurs by variation of its properties, and, in general, by energy exchangeswith the surroundings, such as the work done by pressure, if its
volume changes, for example. In some cases, we can imagine that the opposite progression
might be possible. It would then be a reversible transformation, and in particular, it
might occur via a series of closely related states of equilibrium in which the properties of
the system do not change from one state to the next except in a finite manner. We are then
speaking about a quasistatictransformation. To accomplish this, the motive agents of this
transformation will continually adapt to the state of the system. Such is the case with a
compression process during which the external pressure remains practically equal to that
of the system, or with a transformation due to an infinitesimal difference in temperature.

1.2

THE FIRST LAW

1.2.1 The Energy of a System


The first law is based on attributing an extensive property to any closed material system,
namely energy, and on the establishment of an exact balance between the variations of this
property during the course of a transformation on the one hand, and on the other, the
work accomplished by the surrounding environment due to mechanical, electrical, and
magnetic forces, as well as so-called heat exchanges caused by differences in temperature.
Certain components of energy are familiar and independent of the internal structure of
the system. In the first place, they involve potential energy in the gravitational field EpOt,
which is expressed as a function of mass m,the acceleration of gravity g, and the calculated
elevation compared to an arbitrary reference 2.It is the same for the kinetic energy of the
entire system, Ekinr
that is related to mass, to the moment of inertia, and to the speed of the
system.
Mechanics is a function of these components. In thermodynamics we also take into
account internal energy, U,which originates from:
0 First and foremost, the kinetic energy that accompanies the random movement of the
molecules that make up the system (translation, rotation of all or part of the molecule, longitudinal or transversal vibrations of the interatomic bonds). It determines
the properties of the ideal gas.
0 Secondly, the intermolecular cohesive energy related to the forces of attraction and
repulsion between the molecules. Opposing the disordered kinetic agitation of molecules, they contribute to the organization of matter in different phases.
0 And finally, the interatomic cohesive energy that assures the stability of the molecular structure and is evident during chemical reactions.

1. Principles. ThermodynamicFunctions. The Ideal Gas

Of course, it would be appropriate to quantify the intra-atomic cohesive energies that


are much more numerous, but they do not occur in the phenomena studied here that
preserve the structure of the atom.
As with the potential energy in the gravitational field, internal energy can be calculated
only in relation to an arbitrary reference.
We therefore write:
(1.1)
E = Epot+ Ekin+ U
0

In the problems we present here, most often the variations of potential energy in the
gravitational field and in the kinetic energy of the system will be zero or negligible. The
variations of system energy will be reduced to variations of internal energy.

1.2.2 Energy Exchanges during a Transformation


When the system in question changes from one state of equilibrium to a new state of equilibrium due to the removal of certain constraints that insured the stability of the initial
state, this transformation is accompanied by energy exchanges with the surrounding environment. We shall now attempt to specify the nature of these exchanges. For the most part,
they are attributable to external forces acting on the system, and in particular, to the forces
the work supplied to the
of pressure. If the pressure acting on the system is designated Pext,
system can be expressed by the equation:

For this transformation to be reversible, it is necessary that the pressure acting on the
system Pextbe equal to the prevailing pressure within the system P, and the work dW,,,
will then be equal to:
6Wr,, = -P dV
(1.3)
If we take into account the sign of the variation of volume dV that is related to the Values of Pextand P respectively,we can easily convince ourselves that:

6Ws mY,,,

or

6Ws-PdV

(1.4)

It is similarly appropriate to point out, for example, the work due to the presence of
electrical and magnetic fields surrounding the system. In general they will be absent from
the transformations that we will study, and as a reminder we shall designate them here as
W.
We also know that, due to the difference in temperature between the surrounding environment and the system, the state of the system may change. It is customary to state that
the surrounding environment provides heat to the system. Certainly this statement is
incorrect as neither the surroundings nor the system possesses heat. Nevertheless, we
retain this term as it has passed into practice and is unambiguous. These heat exchanges
shall be designated by Q, or for an infinitesimal transformation, by 6Q.

1. Principles. Thermodynamic Functions. The Ideal Gas

1.2.3 Statement of the First l a w Applied to a Closed System


1.2.3.1 General
The first law, applied to a closed system, establishes an exact balance between the variation
of the total energy of the system during any transformation, reversible or irreversible, and
the total work, and the quantities of heat absorbed by the system:

AE= A(E,,, + ELio+ U )= W + W + Q

(1.5)

the symbol A designates energy variation: E,, - Ehi~d.


Very often, when retaining only the internal energy variations, the work done by pressure and the thermal exchanges,we shall write the equation in the form:
AU=W+Q

(1.6a)

dU=@+6W

(1.6b)

or, for an infinitesimal change, as:

At this point, it is appropriate to consider a few particular cases that correspond to this
simplified version of the first law.

1.2.3.2

lsochoricTransformations

If the transformation in question takes place at constant volume and the work done by
pressure is equal to zero, the preceding equation is reduced to:

Auv = Qv

(1.7a)

dUv = @,

(1.7b)

or, for an infinitesimal change, to:


The thermal exchanges accompanying an isochoric transformation therefore correspond to the variation of the internal energy of the system.
If this transformation has as its only effect a variation in the temperature of the system
(there is neither a change in phase nor a chemical transformation),the heat exchangesmay
then be expressed as a function of the heat capacity at constant volume, Cv, and of the
temperature variation:
dUv = Cv dT
(1.8)

1.2.3.3 Adiabatic Compression


In this case, thermal exchanges equal zero, and the work done by adiabatic compression is
expressed as:
Wadiabatic = "
(1.9)

1. Principles. Thermodynamic Functions. The Ideal Gas

1.2.3.4

Transformationsat Constant Pressure

If the external pressure is constant, and the system is in pressure equilibrium with the surroundings at both the onset and at the end of this transformation, the work done by pressure is expressed by the equation:
Wp=-Pext AV
(1.10)
and, if we take into account the pressure equilibrium conditions at both the beginning and
at the end of the transformation, we can write:
AV = A ( P V ) ,

cxt

where P stands for the system pressure, such that the first law allows us to calculate the
quantity of heat exchanged:
(1.11)
Q p= A( U + P V ) ,
The sum U + PV is the enthalpy of the system. It is an extensive property just like
internal energy and volume. It is more commonly used in practical calculations because
the previous equation relates these variations to isobaric heat exchanges. It is also used in
the expression of the first law applied to open systems, as we shall see subsequently. It is
designated by the symbol H
H=U+PV
(1.12)
such that we can write:
Qp

=m

(1.13)

Therefore the quantities of heat necessary for fusion, for the vaporization of a pure substance, or for a chemical reaction performed at constant pressure will often be called
enthalpy of fusion (negative of the enthalpy of melting), enthalpy of vaporization or
enthalpy of reaction. If the only effect of the transformation is to cause a variation in
system temperature, it would seem that the elementary enthalpy variation is expressed as
a function of heat capacity at constant pressure, and may then be written as:
dHp = C, d T

1.3

(1.14)

APPLICATION OF THE FIRST LAW TO AN OPEN SYSTEM

1.3.1 General
We have made it clear that the preceding equations relate to a closed system that does not
exchange matter with the surroundings. In practice, however, such exchanges are the rule
and it is appropriate to account for the energy contained within the flows of matter that
enter into or exit from the system in question, just as we will have to include their mass in
establishing a balance of mass.
Examples of open systems are varied: a balloon that we blow up, a reservoir that we fill,
a section of a pipeline, or a turbine. Several of these examples can be combined to make up
a less elementary open system. A distillation column, a chemical reactor, or a moving car

1. Principles. Thermodynamic Functions. The Ideal Gas

are other cases. In such systems, neither the pressure nor the temperature are uniform;
they change in space and time.
During a short interval of time, dt, the kinetic energy, the potential energy in the gravitational field, the internal energy of the system, and the mass of the system will vary. They
vary according to the type of content, mechanical work, and the heat and modifications of
the system properties (pressure, temperature, nature of components).
The system can be defined initially by the position of its boundaries (certain of which
are real, others imaginary). For example, a unit of pipeline will be marked between two
clearly indicated sections. We will denote flows of matter dmi, where m is mass, and assign
them a positive value if matter enters into the system, and a negative one if flows are leaving the system. We can also account for instantaneous outflows: dmi = Di dt. Each flow is
characterized by a certain number of properties: potential energy Epot,i,kinetic energy
Ekin,i,
temperature Ti,
pressure Pi,and, given here per unit of mass: volume wi , internal
energy ui,and enthalpy hi.
Heat transfers emanating from sources with temperature Ti will, however, be denoted
by Se,.
k will also be excluded, for example, the work done by presOther energy transfers, m
sure accompanying the possible variation in volume of the system (a balloon being blown
up), the energy supplied by the motor of a pump, etc. However, at this level we shall not
exclude the work accompanying the transfer of matter.
The total energy of the system Eo is:
EO

= EO,pot

(1.1)

+ EO,kin 0
+ '

Transfer of matter is broken down into two steps. Mass dmi occupies a volume w idmi ,
and possesses an internal energy ui dm, .
First, we shall unite the system itself with the mass dmi.This step therefore increases the
mass of the system by dmi, and its energy by:
dE, = epot,idmi + ekin,i dmi + uidmi
This step is not accompanied by work done by pressure; the volume of the system has,
however, varied by + v idmi.
In the second step, we will reduce the new system formed in this way to its initial boundaries. The variation in volume will therefore be -wi dm, and the work that we must supply
Piwi dmi. The transfer of matter therefore is accompanied by a total amount of energy
equivalent to:
epot,idmi + ekin,i
dmi + uidmi + Piwi dmi = (hi + epot,i+ ekin,i)dmi
In the total balance that describes the energy variation of the system, we will sum these
elementary transfers, as well as those corresponding to the work performed on the system
and the quantities of heat:
dEo =

c (hi+
i

epot,i+ ekin,j)dmi

c
i

@j

m
k

(1.15)

This very general expression of the first law obviously must be adapted to each particular case.

1. Principles. Thermodynamic Functions. The Ideal Gas

1.3.2 Steady-State Systems


We say that an open system is in a steady-state when there is no variation on any point in
the quantity of matter or energy, over time. Temperature, pressure, and composition may
vary from one point to the next within the system, but remain on all points invariable over
time.
The previous equation therefore may be integrated over any time period At with
dE, = 0. We denote by Q the quantity of total heat supplied to the system during this time
span, by W the sum of work (other than that corresponding to the transfer of matter), by
H2 the enthalpy of the flows of matter leaving the system, and by H I the enthalpy of the
flows of matter entering into the system:
H2 = - z h i D i At

if

D, d 0 (dm, G 0)

Hl = + x h i D i At

if

D, 3 0 (dm, 2 0)

However, we will exclude the variations in kinetic and potential energy of these flows of
matter. Under these conditions, the energy balance is written as:

H2-Hl=Q+W

or

AH=Q+W

(1.16)

Applied to an open system in steady-state, and, inasmuch as the variations of potential


energy in the gravitational field and the kinetic energy of the flow of matter may be
excluded,the first law establishes an exact balance between the variation of enthalpy from
the entry into, to the exit from the system on the one hand, and on the other hand the
quantities of heat and the work absorbed by the system excluding those accompanying
exchanges of matter.
This equation is closely related to the equation expressing the first law for a closed system. Enthalpy has been substituted for internal energy (to include the work of transfer).
Moreover, the operator A here corresponds to the difference produced between exit from
and entry into the system, while in a closed system, it corresponds to the difference
between the final state and the initial state. We must also recognize the importance of the
function of enthalpy. The energy balances of open systems in steady-state are enthalpic
balances. In this way, for example, the variation in enthalpy will correspond to the work of
adiabatic compression,or even to the heat balance of an item in the absence of work other
than transfer work. Finally,flow that takes place without exchange of heat or work is called
isenthalpic.

1.4

THE SECOND LAW

The first law in no way permits us to specify how the energy exchanges are distributed
(work done by pressure,heat, etc.) during the course of a transformation.It expresses only
the total energy exchange. However, we have seen that the work done by pressure is at
least equal to an attained limit in cases of reversible change. Are heat transfers also limited?

1. Principles. Thermodynamic Functions. The Ideal Gas

We know that the steady-state of a system is assured only by the existence of appropriate boundaries. If some are removed, the system will evolve toward a new state and the
reverse change will not be possible. Given two states, it is therefore important to know
which one is favored. It is equally important to specify the conditions that the properties of
the system must satisfy in order for its state to be favored in relation to all transformations
respecting certain conditions (transformations at constant temperature or pressure, for
example), and that it is in a state of equilibrium.
It is to such questions that the second law of thermodynamics responds. Its statement
varies considerably from one author to another, and an in-depth analysis shows that it is in
fact inseparable from the first law.

1.4.1 Entropy
The statement of the second principle, the definition of the function of entropy and that of
the absolute temperature scale, are inseparable:
Given a system described by its volume, internal energy, and the quantity of matter for
each of its components, there exists a property of the system, entropy S:
It is an extensive function.
0 During a transformation its variations are broken down into two terms:
ds=ds,+ds,
(1.17)

The term dSeis related to heat exchanges by the equation:


SQ
a,=T

(1.18)

where T is a property of the system that depends solely on its temperature, and we designate this property as absolute temperature.
dsi is a term related to internal modifications of the system. It is always positive for
spontaneous transformations, and zero for reversible transformations.

Consequently:

dSt20

(1.19)

ds>-SQ

(1.20)

The previous definitions leave the signs undetermined (which are related) for absolute
temperature and for entropy. As a convention, we state T > 0.
We see later on that the properties of low density fluid allow us to determine the relationship of two absolute temperatures. The absolute scale will therefore be fixed by adopting a fixed point; by convention we state:
at the triple point of water
T = 273.16

It appears that for an isolated system, and therefore in the absence of heat exchange
with the surroundings, entropy can increase only during irreversible processes and stabi-

10

1. Principles. ThermodynamicFunctions. The Ideal Gas

lizes itself at a maximum value when equilibrium is attained. Of course, in the case of heat
exchange, entropy may decrease.
Moreover, the Equation (1.20) may be written as:
6QcTdS

(1.21)

and thus corresponds to the limit mentioned earlier (Eq. 1.4) for the work done by pressure:

6W36Wre,

or

6Wa-PdV

(1.4)

1.4.2 Relationship between Internal Energy and Entropy


The variation of internal energy between two states may be arrived at by observing any
transformation connecting the two states. We may do this using a quasistatic transformation during which, if the work done to the system is limited to the work done by pressure,
we will always have:
6Wr=-PdV

and

6Qr=TdS

(1.22)

We therefore obtain the fundamental expressions that tie the first and second law
together:
dU=TdS-PdV
(1.23a)
1
P
dS=-dU+-dV
T
T

and:

(1.24a)

These equations generally pertain to one mole:


du = Tds - P dv
1
P
ds=-du+-dv
T
T

(1.23b)
(1.24b)

Of course, if other forms of energy transfer intervene (for example, due to the existence
of an electromotive force), it is appropriate to take them into account and we would write:
dU = T dS - P dV +

zFidli

(1.25)

Within these expressions appear the intensive values, or potentials, P, T, Fi, and the
extensive values with which they are associated S, V, and li .
These equations in no way allow for the calculation of work done on the system or the
heat transfers during the course of any process. We can state only:

Q.1

2
1

TdS

and

W s I 21 - P d V

(1.26)

1. Principles. Thermodynamic Functions. The Ideal Gas

11

1.4.3 Application of the Equilibrium Condition


This equilibrium condition shall be illustrated using an isolated system, namely one that
exchanges neither heat nor work (of any kind) with the surroundings:
dW= 0, dQ = 0
The internal energy and the volume are therefore constant:
dU = 0, dV= 0
and, according to the second law, entropy can only increase or be at maximum:
dS 3 0
We will split this system into two subsystemsA and B, using a partition that will be fixed
initially,adiabatic,and impervious to all flow of matter such that each of these two subsystems is also isolated. On either side of the partition temperature, pressure, and composition may be different.
Let us now suppose that the partition is diathermic,that is to say that it allows for heat
flow. As internal energy and the entropy are extensive functions and the system is isolated,
we have:
dU= dUA+ dUB= 0 and dS = dSA+ dS,
We shall express the entropy variations of each part A and B by applying Equation 1.24a. Since the volumes VA and V, are unchanged and the total internal energy is
constant, we obtain:

As entropy can only increase, the result is that dUAand TA- TBare opposite in sign. If
we now suppose that the partition is at the same time diathermic and mobile, we must take
into account the variations in volume dVAand dV,, which are related:

dVA+ dV, = 0
and would write:
and thus:

and, finally:

The maximum condition imposed on the system by the second law thus yields the
equality of temperatures and pressures.

12

1. Principles. Thermodynamic Functions. The Ideal Gas

Finally, if we had considered that the partition separating the two subsystems was not
impermeable,we would have had to use the derivatives of entropy as related to quantities
of matter, and would have established the condition of equality of chemical potentials.This
condition will be examined later (Chapter 5).

1.4.4 Statistical Significance of Entropy


In statistical thermodynamics we can calculate the number of configurations Wthat a system can assume, given the energy and volume. It is possible to show that this number is
related to entropy by the equation:
S=klnW
(1.27)
6 being the Boltzmann constant.
In this way entropy is a measure of the disorder of the system, of its indetermination.
This equation must be retained in order to interpret the sign of the variations of entropy
with volume, the entropy of mixing (Chapter 5), and to establish the expression of combinatorial entropy (Chapter 7).

1.5

HELMHOLTZ ENERGY AND GIBBS ENERGY

We have examined the condition of equilibrium and change in an isolated system. Any
possible change may be characterized as internal rearrangement, and is possible only with
an increase in entropy. In practice, we will deal with transformations during which the system exchanges heat and work with the surroundings. In order to take these exchanges into
account,we relate the two equations:
d U = 6Q+ 6 W + 6W'
in which 6W represents the work done by pressure and 6W' the other work (electrical,
etc.), and:
6QsTdS
(1.21)
Substituting 6Q we obtain:
4

6W + 6W' b dU - T dS

During the course of a change at constant temperature, this equation may be written as:
6W + 6W' 3 d( U - TS ),
or W + W' A( U - TS ) T
(1.28)
Furthermore, if the system changes at constant volume, then the work of pressure force
is zero and this equation becomes:
sW'>d(U-TS),,
or W'bA(U-TS)
at constant temperature and volume
The function:
(1.29)
A=U-TS
is called the Helmholtz energy.

1. Principles. Thermodynamic Functions. The Ideal Gas

13

The preceding equations are therefore written as:

6W3 dA,, or W 3 AA
at constant temperature and volume

(1.30)

A change at constant temperature and volume is therefore possible only if the


Helmholtz energy decreases. In this case, the system may give off a quantity of energy
(-W)to the surroundingsthat is at most equal to the decrease in the Helmholtz energy. If
the Helmholtz energy increases, for the change to be possible, it would be necessary to
supply to the system (in the form of electrical energy, for example) a quantity of energy
that is at least equal to the increase in the Helmholtz energy. Since we are presuming that
there are no such exchanges here, we shall retain the condition of change:
dAT,V

(1.31)

dAT,V=o

(1.32)

and the condition of equilibrium:


If the system changes at constant temperature and pressure, and in pressure equilibrium
with the surroundings,the work done by pressure is then expressed by the equations:

6W=-PdV

or

W=-PAV

such that we would write Equation 1.28 as:


or W 3 A(U + PV- TS)
at constant temperature and pressure

6W3 d(U + P V - TS),,,


The function:

G = U + P V - TS = H- TS = A

+PV

(1.33)

6W3dGT,,
or W3AG
at constant temperature and pressure

(1.34)

is called the Gibbs energy.


The previous equations therefore may be written as:

Thus, a change at constant temperature and pressure is possible only if the Gibbs
energy decreases. In this case, the system may give off a quantity of energy (-W)to the
surroundings that is at most equal to this decrease. If the Gibbs energy increases, for the
change to be possible, it would be necessary to supply to the system (in the form of electrical energy, for example) a quantity of energy that is at least equal to the increase in the
Gibbs energy.
Since we presume that there are no such exchanges, we shall retain the condition of
change:
(1.35)
dGT3 < 0
and the condition of equilibrium:
dG,

=0

or more precisely, minimum G at constant temperature and pressure.

(1.36)

14

1. Principles. Thermodynamic Functions. The Ideal Gas

EXAMPLE 1 .I

Condensation by adiabatic expansion


To illustrate the use of thermodynamic properties in the calculation of compression
work and heat exchanges, we will examine an ethylene condensation process that
functions according to the steps that follow.
Ethylene, available at atmospheric pressure and at 100F (37.8"C), is compressed in
an isothermal fashion to 50 bar. The fluid then undergoes adiabatic expansion to
return to atmospheric pressure. During this expansion, there is partial condensation;
the vapor phase is recycled.
It is necessary to calculate the proportion of liquefied ethylene, the work and heat
exchanges accompanying the compression step, and the work that may be collected
during expansion. To do this, we make use of a diagram (Fig. 1.1) that provides the
pressure variation, shown on the ordinate, as a function of enthalpy, shown on the
abscissa, along the isothermal or isentropic curves. These calculations will force us to
make certain hypotheses relative to the reversibility of each step. Most often, the laws
of thermodynamics do not allow us to estimate the exchanges of heat and work
except in the case of reversibility.

Figure 1.1 Diagram of ethylene enthalpy pressure


[Canjar and Manning, 19671.

15

1. Principles. Thermodynamic Functions. The Ideal Gas

First, a few brief comments on the diagram: given its origin [Canjar and Manning,
19671, the units are British and the necessary conversions will be done using the factors supplied in Appendix 6.
We note that on the diagram the two-phase envelope curve takes the shape of a
dome. It is made up of two parts that relate to the two phases, liquid and vapor in
equilibrium. At a given temperature, the enthalpy of the liquid phase is less than that
of the vapor, and we may therefore easily describe the two parts that meet tangentially at the critical point (T, = 282.4 K, P, = 50.3 bar).
In the vapor phase zone, at low pressure, the isotherms are practically parallel to the
pressure axis, conveying the fact that these vapors are close to the state of an ideal gas,
the state at which enthalpy does not depend upon pressure, as we shall see at the end
of this chapter. If the temperature is less than the critical temperature, the isotherm
curve drops upon crossing the two-phase envelope to a constant pressure plateau that
is equal to the vapor pressure at the temperature in question. In effect, under these
conditions, the system is monovariant.
As for the necessary hypotheses, we will assume that the compression and expansion
steps are reversible. As it is adiabatic, expansion is isentropic.
On the diagram we locate the points that are characteristic of the process. We note
that at the end of isentropic expansion, the point representative of the system is found
within the two-phase zone. The liquid and vapor phases in equilibrium are represented by two points situated on the isobaric and isothermal plateau, and on the twophase envelope. The table below gives the properties corresponding to each of these
points.
Table 1.1
Properties of ethylene during a liquefaction cycle
State
Vapor
Vapor
Two-phase
Liquid
Vapor

Temperature
(OF)

100
100
-154.6
-154.6
-154.6

Pressure
(Psi)

Enthalpy
(Btu/lb)

Entropy
(Btdlbl'R)

14.7
735
14.7
14.7
14.7

1100
1 060
990
815
1 020

1.89
1.55
1.55
1.02
1.70

We note that the properties of the saturated phases are related by the equilibrium
equation:
gvP-gL'J = 0 + (hVP- hLP) - T(sVK'- &'J) = 0
(1.37)
The proportion of condensed ethylene will be determined by application of an
enthalpy balance to the two-phase state. Exponents Lo and Vo respectively denote
the liquid and vapor saturated states. For enthalpy and mass, we use the following balances:
m = mv,'J+ mL.a
(1.38)
m .h = mvPhv.0 + mL.'JhLP

16

1. Principles. ThermodynamicFunctions. The Ideal Gas

and thence derive:


mL.0

--

hv*a- h
1020 - 990
= 0.146
hv~a-hL,"- 1020-815

(1.39)

To determine energy exchanges during the isothermal compression, we apply the first
law considering that the system is an open system in steady-state:
(1.16)

AH=Q+W
2

and the second law:

Q s j ' TdS

(1.40)

As compression is reversible and isothermal, we have:


and

Q=TAS

(1.41)

WT = AH- Q = AGT

(1.42)

Calculationswill be performed using the units provided in the diagram. In particular, we


note that the temperatures are expressed in degrees Rankine ("R).Toconvert to Kelvin:
T("R) = 1.8. T(K) = @ ( O F )

+ 460

As the unit of energy is Btu/lb, we have:

1lb = 453.6 g,

1Btu = 252.cal= 1054 J

and

1Btu/lb = 2.324 J.g-'

So, we arrive at:


T = 560"R
Q = 560*(1.55- 1.89) = -190.4 Btdlb = -442.5 J-g-'
W = (1060 - 1100) - 560 *(1.55- 1.89) = 150.4 Btu/lb = 350 J-g-'
In order for the expansion to be reversible, it is appropriate to recover the work of
expansion. We will apply the previous Equation (1.16) to calculate this work, while
taking into account that there is no exchange of heat, and therefore:
W=AH
(1.43)
for an adiabatic transformation
or:

W = 990 - 1060 = -70 Btdlb = -163 J-g-'

In reality, compression and expansion are irreversible to a certain degree, and the Values above represent the limits. Furthermore, compression is accomplished in two (or
more) adiabatic steps, and therefore is accompanied by an increase in temperature.
Each increase is followed by cooling to restore the initial temperature. Thus the first
step can be substituted with two isentropic compressions going from atmospheric
pressure to 7 bar, and then from 7 to 50 bar. The calculation shows that in this case the
work that must be provided is in the neighborhood of 400 J .g-l and the heat to be
released approximately -500 J * g-'.
Expansion may also take place without recovering work. We would have Q = 0 and
W = 0 at the same time. Expansion would therefore be isenthalpic. With the help of
the diagram, we may verify that it would not allow for ethylene condensation if the
compression step were not modified, and that the compression should reach approximately 120 bar to achieve the same rate of condensation.

1. Principles. Thermodynamic Functions. The Ideal Gas

17

EXAMPLE 1.2

Methanol battery
The electromotive force of a methanol battery operating in an alkaline environment
at atmospheric pressure and under reversible conditions is theoretically equal to
1.33V at 25C.
In practice, the following results were observed in the laboratory under the same
conditions of temperature and pressure: a potential difference of 0.2 V and a heat
release of 721 kJlmol from oxidized methanol.
The reaction is described by the stoichiometry:

CH30H+80H--6e
3
-00,+3H20+6e
2
CH30H +

CO;-+6H20

60H-

3
0, + 2 OH-+
2

CO 3"-

+ 3 H,O

The reactants and the products are in an aqueous liquid solution with the exception of
oxygen, which is gaseous.
These data allow us to determine the variations of the thermodynamic functions
accompanying the oxidation reaction, as well as the exchanges of heat in the case of
reversibility or free combustion.
First, we note that the variations of the thermodynamic functions will be the same
regardless of the degree of irreversibility of the process. Indeed, the initial and final
states are the same.
The work done by pressure can be easily calculated since the pressure of the surroundings is constant.
Wp = -P AV = -A(PV)
The variation in volume results from the consumption of oxygen.We apply the equation for the ideal gas state:
A ( P V ) = RTAv
where Av denotes the variation in the number of moles (-1.5 mol) and R
(8.3145 J-mol-l K-') denotes the ideal gas constant. We find that for a mole of
methanol oxidized: W, = 3.7 kJ.
Furthermore, we can calculate the electrical energy involved: If E designates the
electromotive force, N Avogadro's number (6.022 x
e the electron charge
(e = 1.602 x
C), and An the number of electrons transferred, on the basis of the
preceding stoichiometry (An = 6), then:
W' = N AneE = 96472 AnE

We thus find that W' = -770 kJ in the case of reversibility (E = 1.33 V) and
W' = -115.7 kJ from the results observed (E = 0.2 V). This energy is considered negative as the setup produces energy (electromotiveforce).

18

1. Principles. ThermodynamicFunctions. The Ideal Gas

By applying Equation (1.34) in a reversible case, we derive:


AG = W:ev = - 770 kJ
at constant temperature
In the case of an actual working battery, we first determine that the electrical energy
given off to the surroundings is effectively less than the decrease in the Gibbs energy:
-W < -AG. We can further group the energy exchanges: W = 3.7 kJ, Q = -721 kJ and
-W = 115.7 kJ. We can therefore calculate the variation of internal energy:
A U = W + W+Q=-833kJ
from which we derive the variation of enthalpy:
AH = A U

+ A(PV) = -836.7 kJ

and of the Helmholtz energy:


AA = AG - A(PV) = -766 kJ
The entropy variation is derived from the values of AH and AG:
AS =

AH - AG
= -223.7 J. K-
T

It is from this value that we can calculate the heat exchanges in the case of reversibility:
Q = T AS = -66.7 kJ
Finally, if there is no production of electrical energy (free combustion):

Q = AH = -836.7 kJ
Table 1.2 below summarizes these results (units: kilojoule, kelvin).

Table 1.2
Variation of thermodynamic properties, heat exchanges,
and work during the oxidation of methanol

AU
AH
AS

AA
AG

Q
W
W

Reversible Battery

Experimental Battery

Free Combustion

-833
-836.7
-223.7
-766
-770

-833
-836.7
-223.7
-766
-770

-833
-836.7
-223.7
-766
-770

-66.7
3.7
-770

-721
3.7
-115.7

-836.7
3.7
0

19

1. Principles. Thermodynamic Functions. The ideal Gas

1.6

THERMODYNAMIC FUNCTION: INTERNAL ENERGY, ENTHALPY,


ENTROPY, HELMHOLTZ ENERGY AND GIBBS ENERGY

1.6.1 Dependence on Temperature, Volume, or Pressure


Helmholtz energy,A, and Gibbs energy, G, are, like internal energy, enthalpy, and entropy,
extensive functions. A change in each one of these properties allows us to judge the possibility or impossibility of a transformation, in accordance with the constraints that are
maintained over the course of this transformation. Thermodynamic equilibrium is related
to an extreme condition of any one of these functions. It is for this reason that the functions may be called thermodynamic potentials. It is therefore essential to know their
variation as a function of the properties of the system.
As independent variables we will consider temperature, and either volume (for internal
energy, entropy, and Helmholtz energy), or pressure (for enthalpy, entropy, and Gibbs
energy). To arrive at the elementary variations of the thermodynamic functions we will use
the fundamental Equation 1.23:
dU=TdS-PdV
(1.23a)
Taking into account the definition of Helmholtz energy and Gibbs energy (Eqs. 1.29
and 1.33) one may write:
(1.44a)
dA = -P dV- S d T
dG = V d P - S d T
(1.45a)
or for one mole:

(1.44b)
(1.45b)

da = -Pdv --s d T
dg = v d P - s d T

Since the internal energy, the Helmholtz energy and the Gibbs energy are state functions, and dU, dA,and dG, are exact, total differentials, we derive the following equations
from them:
a2A - a2A
that is
(1.46)
aTav avaT
V

(g)T=
(g)

(1.47)
Similarly,we obtain the derivatives of internal energy and enthalpy as they relate to the
volume and to the pressure respectively:
U =A

+ TS

gives ( g ) T = ( g ) T + T ( $ ) T

and:
H = G + TS gives

au

andthus i a v ) , = T ( g )

(g)T=
(g)T+
T (g)Tand thus

aH

(%);

V- T

V- P

(g)

(1.48)

(1.49)

We have already seen that the elementary variations of internal energy at constant volume, or of enthalpy at constant pressure as a function of temperature, was expressed, for a
homogeneous mixture, using the corresponding heat capacities Cv or C, (Eqs. 1.8 and
1.14). The same is true of course, for the elementary variations of entropy.

20

1. Principles. ThermodynamicFunctions. The Ideal Gas

We list below (Table 1.3) the expressions for the thermodynamic function differentials
represented by TV for internal energy, entropy, and Helmholtz energy, and by T P for
enthalpy, entropy, and Gibbs energy. These expressions do not include the variable for
quantity of matter, which must be considered for systems of variable composition. They
will be given subsequently (Chapter 5 ) . Since these equations are all related to extensive
properties, if we apply them to a pure substance,or to a homogeneous mixture of constant
composition, these equations will be preserved while showing not the property of the system (U,H, S, A, G, V),but the molar property: u, h, s, a, g, v.
Table 1.3
Variation of thermodynamic functions with temperature and either volume or pressure

dT
dS=Cv---+(;)
T

dV

or

du=c,dT+

(1.50)

dP

or

dh=c,dT+

(1.51)

dV

or

dT
ds=c,- T

(1.52)

+&(!),

(1.53)

dA=-PdV-SdT
dG=VdP-SdT

or
or

da=-pdv-sdT
dg=vdP-sdT

These equations are fundamental. Indeed, it is with the help of the derivatives of the
thermodynamic functions with regard to volume or pressure that we shall arrive at the
deviations from the ideal gas law. Here we invoke the Gibbs-Helmholtz equation
applied to Helmholtz energy and Gibbs energy:

(1.54)

(1.55)

Their proof is very simple.For example, for the second equation we can write:
a-G
H
- - G + - 1 aG
- = - - (1H - T S ) - - S =1- T (3T)p
T2
T
T2

(&),=i2

1. Principles. Thermodynamic Functions. The Ideal Gas

21

They should come close to a very common graphical representation in physical chemistry: we use the variation of the logarithm of a value such as vapor pressure, the
liquidhapor equilibrium constant, the solubility of a gas, the equilibrium constant of a
chemical reaction, or the rate constant, as a function of the inverse of absolute temperature. In fact, for each of these values there exists an equation relating it to the variation of
Gibbs energy, AG, accompanying the described phenomenon: Gibbs energy of condensation, of reaction, or of formation of an activated complex, etc. and the graphical representation, allows us to obtain the variation of enthalpy, AH, of this same phenomenon.
This enthalpy variation itself is most frequently only slightly sensitive to temperature,
and the graph obtained is linear.
If the variation of temperature causes a phase change or a chemical reaction, it is then
no longer possible to state that dUv = C , dT or dHp = Cp dT and it is necessary.to introduce the latent heat associated with this phase change or with the heat of reaction. On
the other hand we may, with the help of Equations 1.51 or 1.55, calculate the variation of
the heat of reaction or the Gibbs energy of reaction with the temperature, while pressure
remains constant, provided that the systems formed by (1)the reactants, and (2) the products, do not themselves undergo a phase change. We will then write:

(1.56)

(1.57)

where AG and AH are, respectively, Gibbs energy and enthalpy of reaction, and AC, represents the difference between the heat capacity of the products and reactants at constant
pressure, taking into account the stoichiometry (see Chapter 13).
The heat capacities at constant volume and pressure are related, as are the internal
energy and the enthalpy.We may thus write:
dU = d(H- PV) = dH- P dV- VdP
which is to say that by application of Equation 1.50:

[ (31

dU=CvdT+ T - - P dV=dH-PdV-VdP
or:

CvdT+T - dV=dH-VdP
(

If we are looking at an isobaric process, we would write:


CvdT+T($)

and thus:

(E)P d T = d H p = C p d T

v aT

Cp-Cv=T($)

v (E)
P

(1.58)

22

1. Principles. ThermodynamicFunctions. The Ideal Gas

We can only use the derivatives at constant pressure or at constant volume by applying
the equations:

or:

dV= - d T + - d P
(zF)p

(zF)T

andthus

Thus is obtained:
(1.59)

1.6.2 Characteristic Functions


We return here to the particular role played by each of the functions that we have defined:
the second law may be expressed using any one function from among them, according to
the constraints applied to the system:
Internal energy, U:minimization at constant entropy and volume
Enthalpy, H:minimization at constant entropy and pressure
Entropy, S: maximization at constant internal energy and volume
Helmholtz energy, A: minimization at constant temperature and volume
Gibbs energy, G minimization at constant temperature and pressure
Furthermore, knowing one of these functions as a function of the variables thus related
to it, in fact provides all the necessary information.We shall use the example of Helmholtz
energy.
If the function A (T,V) is known, one disposes also by applying Equation 1.44 on the
pressure expressions (equation of state) and the entropy as a function of these variables:
(1.60)
(1.61)
and the internal energy, enthalpy, and Gibbs energy are obtained immediately:

(3,

U(T,V) = A + TS =A(T,V) - T -

(1.62)

1. Principles. Thermodynamic Functions. The Ideal Gas

23

H(T,V)=U+PV=A(TV)-T - -V -

(1.63)

V):(

( g ) T

MT

G(T,V)=A+PV=A(T,V)-V -

(1.64)

To obtain the heat capacity at constant volume, we can apply Equation 1.52:

Cv = T ( g )
V

and thus, according to Equation 1.61:


(1.65)
Heat capacity at constant pressure is obtained from the heat capacity at constant volume
by application of Equation 1.59,which utilizes only the derivatives of pressure expressed in
1.60, compared to volume at constant temperature or temperature at constant volume.
This calculation of all of the thermodynamicproperties from the characteristic function
A (T,V) may be used for other characteristic functions, G (T,P),for example.
In fact, more often than not, we do not use the specified characteristic function as a
function of its own variables, but rather the equation of state P (TV) or V (T,P).We may
also use the expression for the residual or configuration property that expresses the difference between the characteristicfunction and the value that this function would have if the
fluid were an ideal gas. As we shall see, these data (equation of state and residual property)
are equivalent,but insufficientin the sense that it is also necessary to know the variation of
heat capacity of the ideal gas with temperature.
Note that all the preceding equations, written for a system containing N moles and involving the extensive values U,H, S, A, G, Cp,etc, may be used for molar values: v, 4 h, s, etc

1.7

THE IDEAL GAS

1.7.1 Equation of State and Thermodynamic Properties


It is important to recall the definition and properties of the ideal gas. Firstly, it represents
the limit state of the real fluid when the density (mass/volume) approaches zero (often
misstated as when pressure approaches zero). It also makes up, as we shall soon see, a
constant step in evaluating the thermodynamic properties.
In the ideal gas state, the forces of intermolecular cohesion are null, and the effective
volume of the molecules is also null. Knowing this fact, the equation of state that will form
the definitive equation is especially simple.From now on, we shall denote the fluid properties in the ideal gas state by the exponent #, with N representing the number of moles, and
therefore would write:
PV' = NRT
(1.66a)

24

1. Principles. ThermodynamicFunctions. The Ideal Gas

or for one mole:

Pv' = RT

(1.66b)

R, the ideal gas constant, is expressed using the unit that is most appropriate to the calculation being performed:

R = 8.314 J-rnol-l*K-'
or:
R = 1.987 cal.rnol-l.K-l= 83.145 b a ~ c r n ~ . r n o l - ~ .=K82.058
- ~ atrn*~rn~*rnol-~.K-~
If we apply the equations in Table 1.3, we end up with the results listed in Table 1.4 (for
one mole of ideal gas).
Table 1.4
Variation of thermodynamic functions of the ideal gas
with temperature and either volume or pressure

du# = C; dT

(1.67)

= c! d T

(1.68)

dh'
&#=c[-

da# = -RT

d T + R -dv =c& d T -R- d P


T
v
T
P

dv

- - s# d T

(1.69)

or

da; = -RT d In v

(1.70)

or

dg:=-RTdInP

(1.71)

dg'=RT-

dP
-s#dT
P

In this way, internal energy and enthalpy of the ideal gas depend solely on temperature.
The same is true for heat capacities at constant volume or constant pressure between
which exists the Mayer equation:
C;-C:=
R
(1.72)
But entropy, Helmholtz energy, and Gibbs energy depend on both temperature and volume (or pressure). However, we note the particularly simple formula that relates Gibbs
energy and pressure. We know that for ideal gas, pressure provides us with a convenient
scale of Gibbs energy. Several simple expressions of equilibrium conditions result from
this equation; notably Raoult's law for liquid-vapor equilibria, and the Guldberg and
Waage law for chemical equilibria,where partial pressures play a role.
Thermodynamic properties in the ideal gas state are often related to an arbitrary reference temperature To,and a "standard" pressure of 1bar, designated as Po.
At this standard pressure we can write:

+I

ho(T,Po)= ho(To,Po)

TO

c;(T,Po) dT

(1.73)

25

1. Principles. Thermodynamic Functions. The Ideal Gas

s"(T,P") =s"(To,P0)+

c,"(T,P")

To

and at pressure P:

dT

(1.74)

c,"(T,P") dT

(1.75)

+I

h#(T,P) = h"(T,P") =h" (To,P")

TO

s#(T,P) = s"(To,P") +

dT-Rln-

P
P"

(1.76)

1.7.2 Heat Capacity


It is thus clear that the properties for the ideal gas can be known as soon as we have the
values for heat capacity c,". We know that the kinetic theory of gases yields particularly
simple results in the case of mono and diatomic compounds. Heat capacities are practically
invariant with temperatures close to 5 cal mol-' K-' for noble gases, and 7 cal mol-' * K-'
for diatomic compounds under ordinary conditions of temperature. For more complex
molecules, heat capacity depends on the nature of the compound in question and on temperature. In this calculation, we must take into account the rotational movement of the
groups composing the molecule (and possibly the barriers to rotation), and the longitudinal and transversal vibrations of the interatomic bonds. This calculation is supported by
spectroscopic data. It generally results in complex expressions that are difficult to integrate (which is a major inconvenience when calculating the variations of enthalpy and
entropy), and it is generally preferable to apply more empirical equations that permit us to
obtain the results of rigorous calculations with all the precision required.
Polynomial expressions:
c;= a + b T + cT2 + dT3 + ...
(1.77)
in which a, b, c, d are empirical variables dependent on the component under examination,
are quite often used. However, it is advisable to avoid all extrapolation beyond the temperature range for which these expressions have been formulated, a range that should be
known. In general, they may not be used at low temperature (below 25C). Therefore, the
expression proposed by Aly and Lee [1981] is preferable:

"=

-t

D/T
(D/T)

[ sinh

F/T

] [ cosh (F/T) ]

(1.78)

The expressions corresponding to enthalpy and entropy are:


h"(T) = h"(T,,)
s"(T) = so( To) +

I. [s
In T + C

coth

(g)

-In sinh (;)]-E

(1.79)

;[

tanh (;)-In

cosh

(:)]Lo

(1.80)

26

1. Principles. Thermodynamic Functions. The Ideal Gas

In these equations, B, C, 0,
E, F a r e the empirical coefficients that depend on the nature
of the compound in question.
As an example, we can also cite the formula applied by Younglove and Ely [1987] to the
calculation of heat capacity for light hydrocarbons (methane, ethane, propane, n-butane,
and isobutane):
(1.81)

1.7.3 Data Tables


The literature supplies the values of thermodynamic properties for a large number of pure
substances in the standard state, that is, for most substances that we encounter, in the
ideal gas state, and at the standard pressure of one bar (0.1 MPa). The principal exceptions
are carbon, which is in the graphite state, and sulfur, whose state varies according to the
tables.
These properties are often supplied for a number of temperatures, and at least for
298.15 K. They contain molar heat capacity, enthalpy, and entropy. For enthalpy, the starting point is arbitrary: sometimes we find the values for h+- hO, ,that is, the enthalpy relative to a temperature of 0 K. Any other starting point is acceptable. For entropy, most of
the tables adhere to the third law of thermodynamics, or to the Nernst theorem, according to which the entropy of any pure, solid, crystallized substance is null at absolute zero.
There is no disadvantage in selecting an arbitrary starting point as long as the calculations
are not for a chemical reaction (see Chapter 13).
Such data, for a very limited number of compounds, are found in Appendix 1.

REFERENCES
General Works
Bett KE, Rowlinson JS, Saville G (1975) Thermodynamics for Chemical Engineers. The Athlone
Press, London.
Callen HB (1960) Thermodynamics. John Wiley and sons, New York.
Dodt M (1956) Bases fondamentales et applications de la thermodynamique chimique. SociCtC dBdition denseignement supBrieur,Paris.
Model1 MM, Reid RC (1983) Thermodynamics and its Applications. Prentice-Hall, Englewood Cliffs,
New Jersey.
PBneloux A (1992) Cours de Thermodynamique, UniversitC d Aix-Marseille.

Specific References
Aly FA, Lee LL (1981) Self consistent equations for calculating the ideal gas heat capacity, enthalpy
and entropy. Fluid Phase Equilibria, 6,169-179.

1. Principles. Thermodynamic Functions. The Ideal Gas

27

Canjar LN, Manning FS (1967) Thermodynamic Properties and Reduced Correlations for Gases.Gulf
Publishing Corporation, Houston,Texas.
Younglove BA, Ely JF (1987) Thermophysical properties of fluids. 11. Methane, ethane, propane, isobutane and normal butane.J. Phys. Chem. Re$ Data, 16,577-797.

Properties of Pure Substances

The equations presented in the previous chapter show that the determination of equilibria
or of energy balances implies the evaluation of thermodynamic properties.
Due to the existence of intermolecular cohesion forces, the calculation of these properties in the ideal gas state is only a first step. Since the systems we deal with in real life are
themselves mixtures,a knowledge and understanding of the behavior of pure substances is
itself but a second step, and generally an essential one. To study the behavior of a mixture,
we often apply simple rules of weighting to the properties of its components.Any imprecision during this second step inevitably affects the final result. For example,we cannot hope
to correctly calculate the liquid-vapor equilibria of mixtures if the vapor pressures for pure
substances are in error.
Although sometimes merely a review of well known information,knowledge of some of
the values is almost indispensable to applying the methods that will be developed later on.
To a certain extent, these values make up the minimal databasethat, although limited,
nevertheless allows us to understand the thermodynamicproperties across a wide range of
temperature, pressure, and composition. They are the critical points, vapor pressure, and
heat of vaporization.
We also stress the continuity that exists between the liquid and gas states. This continuity warrants, sometimes even imposes, that the same calculation methods be applied to
these states; it helps to understand the behavior of fluids at high pressure that we sometimes find extraordinary.
In this chapter, we will repeatedly point out deviations from the ideal gas laws. For
properties such as volume or the thermodynamic functions, we will gather the expressions
that relate these deviations and allow us to calculate them.

2.1

THE RELATIONSHIP BETWEEN PRESSURE, VOLUME


A N D TEMPERATURE. LIQUID-VAPOR EQUILIBRIUM

If the distance between molecules is very large and the molar volume approaches infinity,
the cohesion forces become negligible and the properties of the real fluid approach those
of the ideal gas. The first step with calculation of the thermodynamic properties of the ideal
gas is then a reasonable approximationfor gases at low pressure.The extent of the range of

30

2. Properties of Pure Substances

application depends on temperature, and of course on the desired precision, as we shall see.
It is thus in particular at low density, that the pressure tends toward the value calculated for
the ideal gas and that, regardless of temperature, the ratio Pv/RT approaches 1:
if v + w

then P + -

RT

or

Pv

-+l
RT

This does not mean that the difference between the molar volume of the real fluid and
the molar volume of the ideal gas cancel each other out. When the molar volume
approaches infinity, the value limit of this difference is not zero. It is called the second virial coefficient, and increases from very negative values at low temperature to low, but positive values at high temperature, as shown in Table 2.1 and Figure 2.1. Knowing the values
of the second virial coefficient allows for an exact evaluation of the thermodynamicproperties at low density. These values were the subject of compilations [Dymond and Smith,
19801 and of predictive correlations that we shall discuss later (Chapter 3).
Table 2.1
Variation of the second virial coefficient, B, of ethane with temperature (units: cm3 . mol-)
T
(K)
173.15
198.15
223.15
248.15
273.15
298.15
323.15

B
-539
408
-321
-260
-215
-180
-152

T
(K)
348.15
373.15
398.15
423.15
448.15
473.15
523.15

B
-130
-112
-97
-84
-73
-63
47

T
(K)
573.15
623.15
673.15
723.15
773.15
823.15
873.15

B
-34
-24
-15
-8
-2
3
8

By following,at constant temperature, the evolution of the pressure as a function of volume for a given quantity (one mole for example) of a compound, we can qualitatively
characterize the different ranges of the pressure, volume, temperature space and determine certain characteristic properties such as vapor pressure, molar volumes of saturated
phases, and critical points.
Let us consider the case of ethane, for example (Figs. 2.2 and 2.3,Table 2.2). When the
system is homogeneous and if the amount of matter is known, two variables, such as temperature and either pressure or volume, are necessary to determine its properties.At 20C
(293.15 K) and at atmospheric pressure (0.101325 MPa), ethane is in a gaseous state and
the occupied volume is close to 23861 cm3 * mol-l, a value slightly less (24054 cm3 * mol-l)
than the volume obtained by applying the equation for the ideal gas. The compressibility
factor defined by the equation:
z = Pv
-=-v
RT v #
is slightly less than one. We state that the deviation from the ideal gas is negative, at least
as regards the calculation of the volume. It is the result of intermolecular attraction forces,
which will be more pronounced if we decrease the average distance between the mole-

31

2. Properties of Pure Substances

1oc

-1 00

-I

al -200

-300

400
200

400

600

800

Temperature (K)

Figure 2.1 Variation of the second virial coefficient of


ethane with temperature.

cules. Indeed, when we decrease volume and increase pressure, the compressibilityfactor
decreases.We follow on the 293.15 K isotherm in Figures 2.2 or 2.3 the curve AB. When the
pressure reaches 3.76 MPa and the molar volume 345 cm3* mol-, we see that a more dense
phase appears, the liquid phase. We speak of a dew point. The state of the system is represented by point B with coordinates vvO and Pa on Figure 2.2, and Po and Zva on
Figure 2.3, point C with coordinates
(88.3 cm3/mol) and Pa on Figure 2.2, and Pa and
ZL*on Figure 2.3 corresponding to the liquid phase. If we continue to decrease the volume, this liquid phase will accumulate at the expense of the vapor phase. The equilibrium
condition reduces the number of independent intensive variables. The specific or molar
volumes of each one of the phases vva and .,as well as pressure Pa will remain constant
for the duration of the condensation process. This pressure, characteristic for a pure substance in a state of equilibrium between the two phases of liquid and vapor, is called vapor
pressure. Representative point M of the heterogeneous system will lie on segment BC of
Figure 2.2, marked by pressure Pa, and the molar volumes of the equilibrium phases vva

32

2. Properties of Pure Substances

Figure 2.2 Pressure, volume, temperature diagram for ethane.

-:isotherms; - - - - - :two-phase envelope;

:critical point.

0.25-

2.5

7.5

Pressure (MPa)

Figure 2.3 Variation of the compressibility factor for ethane with


pressure and temperature.
: isotherms;
:two-phase envelope;
:critical point.

- ----

33

2. Properties of Pure Substances

Table 2.2
Volumetric properties of ethane
T = 293.15 K

T = 298.1 K

P (MPa)

v (cm3.mol-')

P (MPa)

v (cm3.mol-')

0.101 325
3.763 4
3.763 4
5.5

23 861
345
88.3
82.5

0.990
0.533
0.136
0.186

0.101 325
4.1876
4.1876
5.5

24 278
283
95
87

0.992
0.478
0.160
0.193

T = 303.15 K

T = 305.34

P (MPa)

v (crn3.mol-')

P (MPa)

v(cm3.mol-')

0.101 325
4.650 8
4.6508
5.5

24 694
215
107.9
93.6

0.993
0.397
0.199
0.204

0.101 325
4.871 4
4.871 4
5.5

24 877
145.5
145.5
98

0.993
0.279
0.279
0.212

P (MPa)

v (cm3.mol-')

0.101325
3.5
4.5
5.5

25 111
513.8
316.9
107.6

0.999
0.702
0.557
0.262

and vL,", and their compressibility factors Z"" and ZL," of Figure 2.3. For a volume balance, we can know the proportion of each of the phases by:

This equation is expressed by applying the lever rule to segment BC of Figures 2.2 or 2.3.
The last trace of vapor phase disappears when the volume is reduced to a value of vL'"
(88.3 cm3/mol).We may then speak of a "bubble" point. It corresponds to point C with
coordinates vL>"(88.3 cm3/mol)and P" on Figure 2.2, and pb and ZL," on Figure 2.3. If we
continue to decrease the volume, since the liquid phase is not very compressible,the pressure rises rapidly (curve CD on Fig. 2.2) and the compressibilityfactor of the liquid phase
is practically proportional to the pressure (curve CD on Fig. 2.3). At a higher temperature,
298.15 K for example,the change in the system state and in the pressure is qualitatively the
same. However, we observe that vapor pressure is higher and that the difference between
the properties of the two equilibrium phases diminishes: densities approach each other
and the liquid phase becomes more compressible.Beyond 305.4 K (at 308.15 K for example) and regardless of the volume, we cannot detect a phase change, and pressure increases
in a regular fashion (curve B'D' in Figs. 2.2 and 2.3) when the volume is decreased. The
temperature above which the phase change phenomenon disappears is called the critical
temperature, 305.4 K in the case of ethane. The "threshold of condensation" is then
reduced to a point of inflection, which is called the critical point. The coordinates are the
critical volume v, and the critical pressure P,. The points representing saturated vapor

34

2. Properties of Pure Substances

phase B and saturated liquid phase C describe the two curves of the two-phase envelope.These curves are joined tangentially at the critical point.
Through the critical constraints:

(E)
=(E)
=O
av
av2 T

for T = T , a n d P = P ,

knowing the T,, P,, and v, coordinates often will allow for the identification of the variables of the equation of state (see Chapter 4). We will need it to apply the predictive correlations based on the corresponding states principle (see Chapter 3). For this reason, these
data may be considered characteristic properties of the components and put into any database. They are not always accessible to experimental determination. Therefore in paraffin
hydrocarbons with a high number of carbon atoms (greater than approximately 16) the
phenomenon of thermal decomposition hinders their determination [Teja, 19891.They are
then calculated using structure correlation properties (Chapter 3), or from other measurable variables such as boiling temperature at atmospheric pressure, or density. Without
physical reality, the critical points in this case represent only characteristic parameters
related to the correlations that produced them, and to the models that make use of them.

2.2
2.2.1

VAPOR PRESSURE
liquid and Vapor States

The vapor pressure curve is limited at low temperature by the triple point that represents
the conditions at which the solid, liquid, and vapor phases coexist, and by the critical point
at high temperature.
Figure 2.4 shows the vapor pressure curves for the alkanes from methane through
decane,for carbon dioxide, and for methanol.Note that the hydrocarbon critical points follow a regular variation with the number of atoms in the paraffin chain, at least above
ethane: increase in critical temperature, decrease in critical pressure. We also note that the
critical pressures of more polar compounds (carbon dioxide has a quadropolar structure),
or compounds autoassociated by hydrogen bonding such as methanol, are clearly higher
than those of the hydrocarbons. The critical point for water is: T, = 647 K, P, = 21.7 MPa.
Familiarity with the vapor pressure curve helps evaluate the states of matter, liquid or
vapor in particular.
Identification is, however, not without ambiguity. For example, if we have a pressure
between the triple point pressure and the critical point pressure at a temperature that is
higher than the melting temperature, a rise in temperature will allow us to observe the
phenomenon of vaporization. We will have then correctly identified the corresponding
ranges pertaining to the liquid and vapor states. On the other hand, if the pressure is higher
than the critical pressure, we will then detect no change of phase, and the attribution of
qualifiers and limits to the ranges involved is subjective.It is the same for isothermal compression according to whether it lies above or below the critical temperature.
It is useful to recall the experiment termed critical point tour shown in figures 2.2
and 2.5. Departing from point B characteristicof a saturated vapor (for ethane,Pa = 3.76 MPa,

35

2. Properties of Pure Substances

10

*CH,OH
I
I

*co,
I
I
I
I

E
E

I
I
I
I

i 5

0
200

300

400

500

600

Temperature (K)

Figure 2.4 Vapor pressure curves and critical points ( *)of paraffins
C, - C,,, of carbon dioxide, and of methanol.

vv'=
345 cm3/mol),at a temperature of 293.15 K and constant pressure, we increase this temperature to 308.15 K (path BB'), and then at constant temperature we compress the system to
a pressure higher than the critical pressure along B'D. We then return to the initial temperature (293.15 K) at constant pressure along D'D, and, at this temperature, decrease the pressure to a value equal to the vapor pressure (P"=3.76 MPa) along DC.We find ourselvesat the
saturated liquid state (v"" = 88.3 cm3/mol).Along this path BB'D'DC, we never crossed the
vapor pressure curve, but nevertheless went from the saturated vapor state to the saturated
liquid state. We must conclude that there is continuity between the two states no matter how
marked the property differences such as molar volume are, or how natural the distinction
between liquid and vapor is along the vapor pressure curve.As we have seen, these property
differences decrease when temperature increases, and disappear at the critical point where
liquid and vapor are identical.We conclude by saying that this distinction between the liquid
and vapor states makes sense only if both states coexist.
We must also point out certain features of the homogeneous fluid phase within the critical zone. As shown in Figure 2.2, the molar volume varies very quickly with pressure or
temperature. It is also the case with molecular interactions as well as with some thermodynamic properties such as enthalpy. Under these conditions, fluid, while remaining homogeneous, demonstrates intermediate properties between the liquid and vapor phases and its
properties are particularly sensitive to relatively small variations of temperature and pressure. Among these properties is solvent power, put to good use in supercritical solvent
extraction. We also know that it is particularly difficult to evaluate accurately within this
zone certain properties, such as heat capacity for example (see Chapter 4, Fig. 4.3).

36

2. Properties of Pure Substances

Dt- - - - - YD
I
I
I
I
I
I
I
I

I
I
I
I

/ i

3
??

i 4
2

(I

2
280

300
Temperature (K)

320

Figure 2.5 Continuity of liquid and vapor states: the criticalpoint tour.

The vapor pressures of the most common hydrocarbons and a large number of compounds have been determined, and the results have been collected in many databases such
as the ones assembled by Reid et al. [1976,1987],Daubert and Danner [1986],or Boublik
et al. [1984]. However, for compounds with large molar mass, in particular of heavy hydrocarbons, experimental data are scarce and sometimes inaccurate. A real effort is being
made to determine them, motivated by the problems posed in calculating the thermodynamic properties of petroleum fluids.

2.2.2
2.2.2.1

Vapor Pressure Equations


Clapeyron Equation

Applying the equilibrium condition allows for the derivation of the Clapeyron equation.
The transfer of dn moles from the liquid phase to the vapor phase is accompanied by a

37

2. Properties of Pure Substances

variation of Gibbs energy equal to +gv" dn for the vapor phase, and -gL," dn for the liquid
phase. Total variation,which must be zero by virtue of the equilibrium condition,is written
as:
dG,, = (gv"-gL,a) dn = 0
ant thus:

gva = gL.a

(2.5)

Note that the change of phase happens without discontinuity of Gibbs energy contrary
to what we observe for volume or enthalpy.Furthermore, the enthalpy and entropy variations accompanying vaporization are related: if gv" = $.."then:

hvo0-Ts"=

hL,O- T ~ L , u

or

h"-hLsO=

~(~vu-~L.a)

(2.6)

This equation between the enthalpy and entropy variations accompanying the phase
change has already been used in the previous chapter in the example relating to a liquefaction cycle (Eq. 1.37).
Equation 2.5 remains valid as we move along the liquid-vapor equilibrium curve. The
variations of vapor pressure dPO and temperature dT are themselves related by the equation:
d(gv" - gL,a)= 0
or by the application of Equation 1.45:
(vv"-vL*a)

or:

dp-(sv"-sL,a)

dpU svU-sL.0
dT vV"-vL,a

--

dT=O
(2.7)

and by taking into account Equation 2.6, we may substitute the variations of enthalpy for
the variations of entropy, that is to say the heat of vaporization Ah":

The Clapeyron equation that we have just derived embodies no approximation whatsoever, it is a simple and especially important example of equations that thermodynamics
imposes on properties such as enthalpy, volume, and equilibrium characterization.
However, it involves a differential equation whose integration requires, at least in principle, knowledge of variables (heat of vaporization,molar volume) that are more difficult to
acquire than vapor pressure itself. It is therefore applied with the help of restrictive
hypotheses and results in empirical correlations.

2.2.2.2 Empirical Correlations


A large number of equations have been proposed to express the change in vapor pressure
with temperature. They differ in the extent of the range covered, their ability to allow for
accurate extrapolation outside the range (in particular to low temperatures), their theoretical basis, and of course their complexity.It is not within the scope of this text to present an
exhaustive review of all of them; we will provide but a few examples.

38

2. Properties of Pure Substances

The Clausius-Clapeyron Equation


In order to integrate the Clapeyron equation, we may consider the molar volume of the
liquid phase to be negligible compared to that of the vapor phase, and that the vapor phase
behaves like an ideal gas. We arrive at the expression:
dPO - LW
dT
RT2
PO
or:

If we further assume that the heat of vaporization is constant, we can then integrate the
preceding equation and obtain the well known form:
(2.10)
that may be effectively applied to the calculation of vapor pressures in the neighborhood
of the boiling temperature at atmospheric pressure.
More generally, we can write:
B
(2.11)
lnP=A+T
where the values of A and B are considered variables adjusted to experimental data and
may depend on the range of temperatures under examination.
It is interesting to note that the graphical representation suggested by the Equations 2.10 and 2.11 is in fact applied to a very large temperature range. Figure 2.6 shows the
change of vapor pressure for ethane and methanol between the boiling temperature (Tb)
at atmospheric pressure (Pb) and the critical temperature. We observe that the experimental data are found, with reasonable approximation, on the straight line of the equation:
1

In P'-ln Pb
In Pc- In Pb

1
Tb

Tc

Tb

(2.12)

Table 2.3 allows for a comparison between experimental and calculated values. We note
that interpolation of the calculation yields results of acceptable precision. Yet, below the
normal boiling temperature, error increases rapidly due in part to the extrapolation, and in
part to the fact that the errors are evaluated as relative values. Figure 2.7, where the deviations between experimental interpolated values have been deliberately exaggerated,
demonstrates this phenomenon. We maintain that the expression 2.12 should not be
applied except for interpolation.

39

2. Properties of Pure Substances

1000lT(K)

Figure 2.6 Variation of the logarithm of vapor pressure for ethane


and methanol as a function of the inverse of absolute temperature.
Table 2.3
Vapor pressure for ethane and methanol. Calculation by interpolation between
the boiling point at atmospheric pressure and the critical point (Application of Eq. 2.12);
extrapolation to low temperatures; comparison with experimental data
Ethane

T
90.3
100
120
140
160
180
184.6
190
200
210
220
230
250
270
290
300
305.4

Pexp
(bar)
0.000011
0.00011
0.003 5
0.038
0.214
0.79
1.01325
1.347
2.174
3.340
4.922
7.004
13.01
22.1
35.14
43.54
48.8

Methanol

Peal
(bar)
0.000037
0.00025
0.0052
0.045
0.225
0.79
1.01325
1.338
2.154
3.313
4.900 5
7.006
13.14
22.45
35.63
43.87
48.8

APIP
(/)
223
130
46
17
5
0.18
0
-0.63
-0.91
-0.8
-0.44
0.02
1
1.6
1.41
0.75
0

T
257.7
273.15
293.15
313.15
323.15
333.15
337.9
343.15
353.15
373.15
393.15
413.15
433.15
453.15
473.15
493.15
512.6

Pexp
(bar)
0.0133
0.04
0.1275
0.350 9
0.551 3
0.8397
1.01325
1.243
1.794
3.529
6.396
10.88
21.95
27.14
40.43
58.44
81

Peal
(bar)
0.018 5
0.047 7
0.142 1
0.366 1
0.562 4
0.8419
1.01325
1.231
1.762
3.407
6.160
10.52
21.43
26.61
39.91
57.91
81

APIP
(/)
39
19
11
4.3
2
0.26
0
-0.95
-1.8
-3.45
-3.68
-3.33
-2.35
-1.95
-1.30
-0.65
0

40

2. Properties of Pure Substances

z
e

C
-

-2

-4

5
1000/T(K)

Figure 2.7 Variation of the logarithm of ethane vapor pressure as


a function of the inverse of absolute temperature (the deviations
from linearity have been exaggerated).

Table 2.4
Calculation of vapor pressure by linear interpolation between
the boiling point at atmospheric pressure and the critical point

Compound
Methane
Ethane
Propane
n-Butane
Isobutane
n-Pentane
n-Hexane
n-Heptane
n-Octane

Compound
Average

Maximum

1.7
1

2.7
1.6
1.7
2.2
3
2
3.1
3.7
4

0.84

0.94
1.5
1
1.46
1.28
1.61

Cyclohexane
Benzene
Toluene
Methanol
Ethanol
Acetone
Carbon Dioxide
Hydrogen Sulfide
Water

Error (%)
Average

Maximum

1.58
1.55
1.30
2.57
4.26
2.11
0.54
3.32
2.7

3.7
2.7
2.2
3.7
7.3
4.8
0.9
7.5
4.6

2. Properties of Pure Substances

41

Table 2.4 shows the average errors obtained for some compounds, always using interpolation, between the boiling temperature at atmospheric pressure and the critical point.
However, it is quite evident that none of the three approximations used above to derive
the Clausius-Clapeyron equation is valid within such a wide temperature range. As we
approach the critical point, the molar volume of the liquid phase is not negligible compared to that of the vapor phase since they both converge toward a common value. The
vapor phase deviates significantly from the behavior of the ideal gas, as shown by the Values in Table 2.2. Finally, the heat of vaporization decreases steadily toward zero. In fact, the
linear correlation obtained results from the compensation between deviations, and it is
only valid if vapor pressure is low and the temperature range is narrow.
If we denote by AZ the difference between the vapor and liquid compressibility factors
at equilibrium:
(2.13)
the Clapeyron equation (Eq. 2.8) can then be written:
(2.14)

The experimental data show that the ratio Ah/AZa varies little with temperature,
which explains the near linearity of the vapor pressure plotted against the temperature
when using logarithmic scales of In ( P ) and 1/T.We see here an example of simple and relatively precise correlations arising from a fortunate coincidence of imprecise hypotheses. There are other, less obvious ones. In this particular case, it would not make sense to
derive the heat of vaporization from the slope of the line represented by Equation 2.9,
since it would be necessary to determine the temperature to which this quantity corresponds.
The Antoine Equation
If we limit ourselves to a relatively narrow range, the Antoine equation:
(2.15)
(where 8 denotes the temperature), can achieve excellent precision; the variables u, b, c are
specific to the substance involved. It should be emphasized that, according to the authors,
the units of temperature, pressure, and the logarithmic base will vary.
As an example, Table 2.5 shows the values of these variables for several compounds
[Boublik, 19841.They have been determined from experimental results covering pressure
intervals, which vary according to the nature of the substance in question, yet are always
near atmospheric pressure. The use of these data is not recommended if the vapor pressure is not in the interval of 20-200 kPa (0.2 to 2 atmospheres). Other very extensive tabulations of these variables can be easily found in the literature [Reid ef uf., 1987; Yaws,
19891.

42

2. Properties of Pure Substances

Table 2.5
Parameters of the Antoine equation (Eq. 2.15).
The temperature is given in "C, the pressure in kPa, and the logarithms are base 10
Constituent

n-Pentane
n-Hexane
n-Heptane
n-Octane
Cyclohexane
Benzene
Toluene
Ethylbenzene

5.990 28
6.01098
6.027 01
6.043 94
6.005 69
6.019 05
6.084 36
7.078 50

1071.187
1176.102
1267.592
1351.938
1223.273
1204.637
1347.620
1421.653

232.766
224.899
216.796
209.120
225.089
220.069
219.787
212.676

Constituent
o-Xylene
m-Xylene
p-Xylene
Methanol
Ethanol
Acetone
Water

6.126 99
6.13232
6.11356
7.206 6
7.242 22
6.250 16
7.062 52

1476.753
1460.805
1452.215
1582.698
1595.811
1214.208
1650.270

213.911
214.895
215.158
239.765
226.448
230.002
226.346

# EXAMPLE2.1
Comparison of the vapor pressures of cyclohexane and benzene
As an example of the use of the Antoine equation, we shall compare the vapor pressures of cyclohexane and benzene.
First, let us calculate the boiling points under atmospheric pressure (101.325 kPa) for
these two compounds by writing Equation 2.15 in the form:

Using the numerical values from Table 2.5, we find:


cyclohexane: = 80.73"C;
benzene: = 80.089"C.
The very similar volatilities of these two hydrocarbons can be further verified by calculating the vapor pressures at 25C and at 80"C, again using Equation 2.15: the
results are presented in Table 2.6. It will be noted that, at 25"C, the vapor pressure of
cyclohexane is higher than that of benzene, while at 80C the order of volatility is
reversed.
Table 2.6
Comparison of the volatilities of cyclohexane and benzene
Hydrocarbon
Cyclohexane
Benzene

(in oc)

Pain kPa
(at 25C)

Pain kPa
(at 80C)

80.73
80.089

13.011
12.692

99.012
101.045

eb

In this range, a temperature exists at which the vapor pressures of these two compounds
are identical. To find this temperature, we shall write the Antoine equation for cyclohexane in the form 2.15; and for benzene, by using different letters for the variables:

2. Properties of Pure Substances

43

b
cyclohexane : In Po = a - C+8

e
benzene : In Po = d - -

f+8

The equality of the vapor pressures would mean that:


(a - d)02 + [(a - d ) ( c + f ) - ( b - e ) ] 8 + (a - d ) c f + ec - bf= 0

The numerical values of the variables a, b, ... f lead to two solutions: 8 = -1 890C
and 8 = 50.84"C. Using this second solution, the corresponding vapor pressure is
37.35 kPa.
Of course, we must allow for the uncertainty of the values arrived at in this way. Even
so, the inversion of the vapor pressures of cyclohexane and benzene at about 50C can
be taken as certain.

The point of intersection of the vapor pressure curves of two compounds is known as
the Bancroft point.

The Frost-Kalkwarf Equation


Without adopting the very restrictive hypotheses that led to the Clausius-Clapeyron
equation, we can consider that the molar volumes of the liquid and vapor phases at saturation obey the van der Waals equation of state (see Chapter 4), and that the heat of vaporization is a linear function of the temperature. We then arrive at the expression [Harlacher,
19701:
B
Po
In P'=A - - - C l n T + D (2.16)
T
T
The parameters of this equation, while connected in principle to those of the van der
Waals equation of state and to the heat of vaporization, are in fact determined by regression from experimental data.Their values are given in the work of Reid et al. [1987].Their
number can be reduced if we make the equation pass through the critical point, or through
the boiling point under atmospheric pressure. These possibilities have been examined
quite recently by Rogalski et al. [1991], who propose several methods of predicting these
parameters in the case of hydrocarbons.
The fact that this equation is not explicit in either pressure or temperature is a minor
disadvantage.
When studying the problem of calculating vapor pressures at low temperature, we may
have to perform a more rigorous analysis of the variation of the heat of vaporization with
temperature. This analysis sheds light on the heat capacities of the liquid and of the vapor
at saturation. It has been developed by King and Al-Najjar [1974] and applied by Ambrose
[1980].

44

2. Properties of Pure Substances

The Cox Equation

This equation has only three parameters that need to be determined from experimental data,
but it also assumes that the boiling point Tbat atmosphericpressure is known. It is written as:

PO

exp(Ao+A1T+A2T2)

(2.17)

Ruzicka and Mayer [1994] recommend it for representing the vapor pressures of hydrocarbons.
The Wagner Equation

The precision with which experimental data can be interpolated by various expressions
has been investigated by Ambrose [1978a]. Given the number of variables that need to be
determined, preference is given to the expression proposed by Wagner [1973]:
Po T,
T
where z = 1- In - = - (AT+Bz1j5+ Cz3 + Or6)
(2.18)
pc
T
Tc
The qualities of this same equation when extrapolating to elevated temperatures and
when calculating the critical pressure, have also been studied [Ambrose, 1978bl. The
numerical values of the parameters are set out in the work of Reid et al. [1987].

2.3

ENTHALPY DIAGRAM AND HEAT OF VAPORIZATION

2.3.1 Dependence of Enthalpy on Pressure and Temperature


In order to present the change in enthalpy of a pure substance with pressure and temperature, we shall make use of the enthalpy diagram for ethane (Fig. 2.8) and describe it along
several isobaric paths. The variations in enthalpy then correspond to the quantities of heat
absorbed by the system along these paths (Eq. 1.13).We note first of all that the enthalpy
is evaluated only with respect to an arbitrary origin (Chapter 1). This origin can be the
enthalpy in the standard state at a temperature of 0 K, and we would then consider the difference h , - hO,.More commonly,we will choose a reference enthalpy,such as that of the
liquid at its boiling point at atmospheric pressure. Finally we must recall that, since the
choice of the origin remains open for each of the constituents of the system under consideration, it can be adapted to any application, but obviously it cannot be modified during
the course of a given computation. It should also be recalled that the numerical values of
enthalpy cannot be compared without considering the origins selected!
At a pressure below the critical pressure and starting from a temperature where the
fluid is unquestionably in the liquid state (point A, T = 285 K, P = 3.76 MPa), we can distinguish three sections along an isobaric variation, which can be found in Figure 2.8: curve
AB represents the change in the liquid from 285 to 293.15 K, segment BC ( T = 293.15 K)
pertains to vaporization at the particular pressure, and finally curve CD represents the

45

2. Properties of Pure Substances

12.5 P = O . 1 MPa

280

285

290

295

300

310

305

315

320

Temperature (K)

Figure 2.8 Enthalpy diagram for ethane.


-. .isobars;
:two-phase envelope;

-- ---

* critical point.
:

change in the vapor between 293.15 and 315 K. Using Equation 1.14, the variations in
enthalpy can be evaluated as a function of the heat capacities of the liquid cp", of vapor c:,
and of the heat of vaporization Ah'? Several isobars can thus be traced, noting that the pressure has little influence on the enthalpy of the liquid, at least as long as the temperature is
well below the critical temperature, and these isobars are practically the same in the liquid
range. By connecting the points representing the liquid and vapor phases at saturation,we
trace two curves which meet at the critical point, and which constitute the "two-phase envelope" on this diagram. It corresponds,point by point, to the two-phase envelope drawn on
the pressure/volume diagram (Fig. 2.2). Although they describe the variation in enthalpy
with temperature, the slopes of the curves that make up the two-phase envelope do not correspond to the heat capacity at constant pressure. We may speak of the heat capacity at saturation, $:",or &,' related to the heat capacity at constant pressure by the equation:
dT
or:

(2.19)
(2.20)

Since the term dhv/dP is negative, the curve for the enthalpy of the saturated vapor
passes through a maximum. Because of the very low value of the term dhL/dP,the curve
for the saturated liquid is practically merged with the isobars.
At a pressure greater than the critical pressure (curve A'D' in Fig. 2.2), no change in
phase will be observed, and the variation of enthalpy will be evaluated solely as a function

46

2. Properties of Pure Substances

of the heat capacity of the fluid at constant pressure. However, it will be noted that this
heat capacity maximizes at approximately the critical temperature, being more pronounced as the pressure approaches the critical pressure. If the pressure is equal to the
critical pressure, the critical isobar has an inflection point with a tangent parallel to the
enthalpy axis and the heat capacity at constant pressure takes on an infinite value.
In the vapor domain, the isobars pertaining to low pressures are practically merged with
the curve (independent of the pressure) that can be drawn for a perfect gas. For a given temperature, the diagram thus presented will bring out the deviations from the perfect gas laws,
with regard to the enthalpy. We find that these are negative: the enthalpy of the real fluid is
less than that of the perfect gas, especially since the pressure is higher for the vapor state.
This phenomenon reverses itself however when the temperature is very high as compared to
the critical temperature. This deviation for the liquid state is very substantial (in absolute
value) and is made up of two terms: the deviation with respect to the saturated vapor, whose
magnitude depends on the vapor pressure, and the heat of vaporization (sign reversed).

2.3.2 Heat of Vaporization


As we have mentioned repeatedly, the heat of vaporization decreases when temperature
increases, and cancels itself out at the critical temperature, when the liquid and vapor
phases are the same. Figure 2.9 shows some examples of this change.

Figure 2.9 Variation of heat of vaporization with temperature:


the n-alkanes C , - C,,, carbon dioxide and methanol.

47

2. Properties of Pure Substances

Experimental measurements of heat of vaporization are relatively rare, and most often
its value is derived from vapor pressure measurements by applying the Clapeyron equation. Its value is essential for a precise calculation of the enthalpic properties in the liquid
phase. It is appropriate to recall here the Trouton equation according to which, at atmospheric pressure, the entropy of vaporization, as related to the heat of vaporization at that
temperature, is equal to 21 cal .mol-' * K-' This equation provides a very useful, but rather
rough approximation, as shown in Table 2.7, that gives the values of the adimensional quotient AhVRT at boiling temperature at atmospheric pressure. It must be remembered that
the Trouton equation is not applied to autoassociated components.
Equation (2.14) may be written as:
(2.21)

A precise application of the vapor pressure values with the goal of calculating the heat
of vaporization by application of Equation 2.21 requires a correlation of these values with
the aid of one of the previously described empirical equations, as well as data for the compressibility factor (or the molar volume) of the vapor phase (equation of state for example) and of the liquid phase. For example, at moderate pressure we can apply the equation
of the virial state (see Chapter 3, Section 3.2.2.1, and Chapter 4, Section 4.1) to the vapor
phase, and calculate the molar volume in liquid phase using the Rackett equation
(Chapter 3, Section 3.2.1.2).
Variation in the heat of vaporization with temperature can be evaluated using the
empirical equation proposed by Watson:

(2.22)

The precision of this equation has been improved by Thek and Stiel [1966, 19671 by
adapting the value of the exponent to the nature of the component.
Table 2.7
Entropy of adimensional vaporization, AhVRT,
at atmospheric pressure
Compound
Methane
Ethane
Propane
n-Butane
Isobutane
n-Pentane
n-Hexane
n-Heptane
n-Octane

Compound

AhOIRT
8.79
9.57
9.78
9.90
9.84
10.1
10.2
10.3
10.5

Cyclohexane
Benzene
Toluene
Methanol
Ethanol
Acetone
Acetonitrile
Water
II

AhOIRT
10.2
10.5
10.5
12.5
13.5
10.9
10.3
13.2

48

2. Properties of Pure Substances

0 EXAMPLE2.2
Calculation of the heat of vaporization
for benzene using the clapeyron equation

Around the boiling temperature at atmospheric pressure, the vapor pressure of benzene may be calculated by applying the Antoine equation:
b

log,,PO=a--

(2.15)

c+e

The numerical values of the parameters are listed in Table 2.5: a = 6.01905
b = 1204.637 c = 220.069, with the units "C, kPa, and the base 10 logarithms.
Additionally we will use the following data:
molar volume in liquid phase: v L = 95.60 cm3/mol
second virial coefficient at 80C: B = -943 cm3/mol.
In order to be able to apply the Clapeyron equation more easily, we shall convert the
Antoine equation to SI units and use natural (Neperian) logarithms (In x = 2.30258
1oglO.x)so that Equation 2.15 becomes:
In P"(Pa) =a'-

b'
c ' + T(K)

= 20.767 13 -

2 773.779
T (K) - 53.081

The pressure is equal to atmospheric pressure: Pa = 101325 Pa, and the temperature
is derived by application of the Antoine equation in the form:
T=

b'
~

a'- In P

- C'

We find: T = 353.24 K.
The Clapeyron equation is written as:
dP'
-- dT

Ah0

T(v-'v

v',?

T(vV'J-vLU dPa

and thus:

)dT

We will express the derivative dPs/dT from the Antoine equation:

-dPa
dT

b'pa

( c ' + T)2

= 3 120 Pa/K

It is also necessary to calculate the molar volume of the saturated vapor:


RT + B = 8.314.353.24
vvu= -943.10-6 = 28041.10-6 m3 mol-'
Po
101325

(2.8)

49

2. Properties of Pure Substances

We end up with:
dP'
dT

Ahu= T(v~'-v"u) -=353.24(28041

-95.6).10-6.3120 = 30794 Jmmol-'

This result is very close to the experimental value (30,760 J * mol-'). If we had compared the vapor phase to an ideal gas and neglected the molar volume of the liquid,
we would have arrived at 31,938 J .mol-' with an error in excess of 3.7%.

2.4

CALCULATION OF THERMODYNAMIC PROPERTIES

The method most generally applied for determining the thermodynamic properties of a
system consists in separating the contribution corresponding to the translational, rotational, and vibrational energies of all or part of the molecule and the contribution resulting
from the molecular interaction forces; the deviations from the ideal gas laws, or the "residual properties", correspond to this second contribution. These properties cannot be
considered corrective terms. We may not disregard them, except at low pressure or more
exactly, at low density, since a condensed liquid or solid phase may be stable at low pressure.
To evaluate these residual properties, the real fluid and the ideal gas may be considered
either under the same conditions of temperature and pressure, or under the same conditions of temperature and volume. For enthalpy, for example:
(2.23)
(2.24)
Note that if the temperature and the pressure are the same for the real fluid and the
ideal gas, the molar volume is not the same, and if the temperature and the molar volume
are the same, the pressure is different.
The residual properties at a given temperature and pressure, on the one hand, and at a
given temperature and volume, on the other hand, are not identical in all cases, as we shall
see. Thermodynamic statistics first yield the residual properties at the given temperature
and volume, and it is the same for most equations of state. Yet the conditions imposed are
most often temperature and pressure and we refer most often to the residual properties at
a given temperature and pressure, which we will do from now on except for specific situations.
To calculate them we generally call upon the relationships that exist between the variations of the thermodynamic properties at constant temperature, dU,, dH,, etc., and
depending on the case, the variations in volume dV, or pressure dP. If the behavior of the
system is represented by an equation of state, and if this equation is explicit in volume or
pressure, we arrive at the residual properties at the given temperature and pressure, or at
the given temperature and volume respectively. It is therefore important to specify the differences between these two types of residual properties and the equations that link them.
We shall do this by calculating residual enthalpy and Gibbs energy.

50

2 . Properties of Pure Substances

Generally, X stands for any thermodynamic property (internal energy, enthalpy,


entropy, etc.). The calculation of the residual value from the given pressure,volume, temperature is based on:
(2.25)

(2.26)

and:

2.4.1

Residual Enthalpy

The derivative for molar enthalpy as it relates to pressure at constant temperature is


expressed by the equation:

(Z),=.-T($)

(1.51)

shown in Chapter 1, Section 1.6.1.


For the ideal gas, the variation of enthalpy with pressure is zero (Chapter 1, Eq. 1.68);
and we have:
(2.27)

Applying this equation supposes that we may express the volume as a function of pressure, that is, using a volume explicit equation of state. This is not always the case. If the
equation of state is simultaneously applied to both the liquid and vapor phase, it will for
certain values of pressure yield at least two molar volume values. One value is valid for the
liquid phase and the other for the vapor phase.The equation has to be pressure explicit. Of
course, the preceding integration may be performed by parts. It should be remembered
that isothermal variations in enthalpy are expressed, preferably from pressure variations;
variations in internal energy are expressed as a function of volume variations. So we will
first calculate the residual term that relates to internal energy at a given temperature and
volume. To explain this calculation in further detail, we shall specify the values for temperature, volume, and pressure. The derivative for internal energy as it relates to volume at
constant temperature is expressed by the equation:

(&p($)
V- P

(1S O )

It is zero for ideal gas (Chapter 1, Section 1.7.1,Eq. 1.67), and the residual term at the
given temperature and volume are calculated using:
(2.28)

2. Properties of Pure Substances

51

To obtain the residual internal energy at the given temperature and pressure, we write:

However, with regard to the internal energy of ideal gas, the difference:

is zero. We therefore obtain:

To arrive at the residual term for enthalpy, it suffices to remember the equations:
h = u + Pv

and

h# = u # + RT

that is ( h - h#) = (u - u # )+ Pv - RT

We then obtain the equation:


h(T,P)-h#(T,P) =

2.4.2

1: [ (9,
1

T - -P dv+Pv-RT

(2.29)

Residual Cibbs Energy

We shall use the same process to derive the equation for residual Gibbs energy, but we
shall note important differences.
The derivative for Gibbs energy with respect to pressure at constant temperature is
equal to the volume:

(1.45)
For ideal gas, it is not zero:
(1.71)

We therefore write:
(2.30)

If we cannot calculate the volume as a function of pressure (pressure explicit equation


of state), then as done previously, we will first calculate the residual term at the given temperature and volume that relates to Helmholtz energy. To explain this calculation in further detail, we shall specify the values for temperature, volume, and pressure. The deriva-

52

2. Properties of Pure Substances

tive for Helmholtz energy with respect to volume at constant temperature is expressed by
the equations:

(E),=-P

(1.44)

(1.70)

and
The residual term is calculated according to the expression:

(2.31)
To obtain residual Helmholtz energy at given temperature and pressure we have:

as opposed to internal energy, the variation in Helmholtz energy of an ideal gas with
respect to volume is not zero:
(1.70)
so that:

We therefore obtain:

To obtain the residual Gibbs energy term, it suffices to recall the equations:

g = a + Pv

and

g # = a # + RT that is ( g - g #) = (a - a # ) + Pv - RT

We then obtain the equation:

g (I;P) - g #( I;P) =

1: (- + 5)
P

Pv

dv - RT In - + Pv - RT
RT

(2.32)

Table 2.8 lists the residual function expression at given temperature and pressure for
the principal thermodynamic properties. For internal energy and Helmholtz energy, temperature and volume are the most natural independent variables. For enthalpy and Gibbs
energy, these variables are .the temperature and the pressure. Finally, residual entropy is
expressed within the two systems.

2. Properties of Pure Substances

53

Table 2.8
Expression of residual values

(2.28)

(2.27)

hres(TP)= h ( T 4 v )- h#

p)=%
[T( g ) v - P ] d v + P v - R T

v #= -

(2.29)

(2.33)

(2.34)

ures(Tw)= a ( l ; e v ) - a #

,v)=

1[ %]
-P

dv

(2.31)

(2.30)

2.4.3

Fugacity

The potential that determines the equilibrium of a pure substance between different
phases is the molar Gibbs energy as shown in Equation (2.5). In mixtures,it is the chemical
potential, meaning the contribution of each component to the Gibbs energy of the mixture. From a purely formal point of view, resorting to these properties has certain disadvantages: molar Gibbs energy cannot be calculated in relation to an arbitrary origin, and
tends toward --oo if the pressure nears zero. Furthermore, it is regrettable that between
pressure and Gibbs energy there exists no equation that is as simple as the one previously
derived for ideal gas:
dg,#=RTdlnP
(1.71)
Lewis [1923] proposed representing Gibbs energy using an auxiliary property, fugacity
(initially referred to as escaping tendency), whose isothermal variables are defined by
the equation:
dg,= RTdlnf
(2.35)

54

2. Properties of Pure Substances

As we can see, for Gibbs energy, fugacity plays the same role for a real fluid as pressure
plays for ideal gas. This analogy is completed by defining the limiting condition:

f
(2.36)
ifP+O
P
Fugacity presents itself as an effective pressure, that is to say, the value that must be substituted for pressure in order to preserve the expressions for ideal gas, always with respect
to calculation of the variations in the Gibbs energy. For residual Gibbs energy, we can substitute the fugacity coefficient:
f
cp= (2.37)
P
-+1

and write:

g(T,P)-g#(T,P) = RTln - = RTln cp


P

So, by applying the Equations (2.30) or (2.32):

R T In cp =

jop(v

y)

dP

(2.38)

(2.39)

R T l n c p = j l ( - P + $ ) d v - R T l n - + PPv
v-RT
RT

(2.40)

For a pure substance at low pressure, fugacity is numerically close to the pressure for
the gas phase and the vapor pressure for the liquid phase, as we shall see.

2.4.4

Calculation of Thermodynamic Properties in the liquid Phase

We have emphasized the continuity that exists between the liquid and vapor phases, and the
ambiguity of these terms. The fact remains that the distinction between these states is often
important. If we apply the equations that allow for the calculation of residual values for a pure
substance under conditions where it is characterized as a liquid, these equations then incorporate the variations of the thermodynamic functions accompanying the change of phase.
Figure 2.10 lists five states in the diagram (temperature versus pressure). For points 1,2,3,
and 4, the temperature is less than the critical temperature. Point 1designates a condensable
vapor and the pressure is less than the vapor pressure. Points 2 and 3 are merged at coordinates T,Pwhere the pressure is equal to the vapor pressure. The first of these points (point 2)
represents a saturated vapor; the second point, a saturated 1iquid.Theyare situated on either
side of the vapor pressure curve. Point 4 represents a liquid phase that is obtained, for example, by compression of the saturated liquid. Finally point 5, located at a temperature higher
than the critical temperature represents a fluid state that we may call supercritical.
Calculation of the residual enthalpy or Gibbs energy terms that relate to states 1,2, and 5 by
application of Equations (2.27 to 2.34) does not pose any problem. The integration path
encounters no change in phase. On the other hand, for the saturated liquid (point 3), or under
pressure (point 4), we are led to perform the integration within the range of phase change.
This forces us to suppose that the expression (equation of state) that we are using provides
the derivatives for the thermodynamic function in question as it relates to volume (expres-

2. Properties of Pure Substances

55

6-

t4

t5

H
v
p!

4-

p!

b
n

,
I

2.

280

300

320

Temperature (K)

Figure 2.10 Calculation of thermodynamic properties as a function of


temperature and pressure.

sion 2.26) or pressure (expression 2.25);values (without physical reality) whose integral correctly describes the variation of the function during condensation. Such is the implicit postulate to which we shall return with regard to the equations of state. We can also break down
the calculation of a thermodynamic function into several steps while preserving the physics
and the precision. We shall spec@ these steps for enthalpy, Gibbs energy, and fugacity.

Calculation of Enthalpy
We can first evaluate the residual enthalpy term relative to the saturated vapor (point 2)
by application of Equation 2.27 (or 2.29), and derive the enthalpy of this vapor:
(2.41)
h " ( 7 y ) = h#(?;P? + h y e , ( z y )
In order to move to the saturated liquid (point 4), we subtract the heat of vaporization:
hL( I;P? = h#(z

y ) + hYeJ 7 y )- Ahu

(2.42)

56

2. Properties of Pure Substances

and finally, for the pressurized liquid (point 4), we include the variation of the enthalpy of
the liquid with pressure (Eq. 1.51):

This last term is generally low compared to the preceding terms, in particular compared
to the heat of vaporization.
Calculation of Gibbs energy and Fugacity

As with enthalpy, we first calculate the Gibbs energy or the fugacity of the saturated vapor:
gv(TP? = g#(TP? + gres(TP?

(2.44)

or, by expressing residual Gibbs energy using the fugacity coefficient:

gv( TP? = g#( T,P) + RT In rp'

(2.45)

and thus, by applying Equation (2.39):


(2.46)
Observing that the equilibrium condition imposes equality of molar Gibbs energy or
fugacity in the vapor and liquid phase, we have:
g L ( T P 9 = g V ( T P )= g o

(2.47)

Finally, for a pressurized liquid phase (point 4), we will include the variation of Gibbs
energy or fugacity with pressure:
R T l n (fL(T,:p)
T)=j

vLdP
F

or:

(2.49)

The vL/RT ratio is generally low since it is equal to the ratio of the molar volume of the
compound in question to the molar volume of ideal gas at atmospheric pressure. The
fugacity of a dense phase therefore varies with pressure and the term:
T = e x p ( I pa
P eRT
) d P

(2.50)

called the Poynting correction, is close to unity at moderate pressure. To calculate it, we
often acknowledge that the liquid phase is not compressible, and we can write:
VL(P-

T = exp(

RT

P)

(2.51)

57

2. Properties of Pure Substances

EXAMPLE 2.3

Calculation of the fugacity of ethane at 25C


Using the data presented in Table 2.9 (first and second columns), we will calculate the
fugacity coefficient of ethane as a function of pressure at 25C.
Figure 2.11 shows the variation of the difference v - RT/P (Table 2.9, third column) as
a function of pressure, and permits the graphical integration necessary for the calculation of fugacity in the vapor phase (Eq. 2.39). The results of this integration are
shown in Table 2.9 for several values of pressure. They are used to calculate the fugacity coefficient and fugacity.This value follows the pressure; the difference is greater as
the pressure rises.

300

200

rt

100

I
i
.

0
0

2
3
Pressure (MPa)

Figure 2.11 Calculation of the fugacity coefficient for ethane at 25C.

Table 2.9
Calculation of the fugacity of ethane in vapor phase at 25C
v - RTIP

P
(MPa)
0.1

0.5
1
1.5
2
2.5
3
3.5
4
4.1876

(cm3.mol-')

24 602
4 767
2 282
1449
1028
770
592.2
456.5
337.4
286.85

-186
-190.6
-196.8
-203.5
-211.4
-221.5
-234
-251.7
-282.3
-305

rp

(MPa .cm3.mol-')

-18.6
-94.2
-191
-291
-395
-503
-616
-737
-869
-924

0.992
0.963
0.926
0.889
0.853
0.816
0.780
0.743
0.704
0.689

0.992
4.81
9.26
13.3
17.1
20.4
23.4
26
28.2
28.9

58

2. Properties of Pure Substances

Extending experimental data to saturated vapor (P = 41.876 bar), we obtain fugacity


at saturation, 28.9 bar, relative to both vapor and liquid in equilibrium. If we wish to
calculate the fugacity at a higher pressure, we apply Equation 2.51 that requires the
knowledge of molar volume of the saturated liquid. Using &'= 95 cm3 * mol-' and
assuming that the liquid is not compressible, we obtain, for example (Table 2.10):
Table 2.10
Poynting correction applied to ethane
between 50 and 100 bar, at 25C

I
2.4.5

50

1.031
1.135
1.25

29.76
32.75
36

General Equations

To evaluate a thermodynamic property such as enthalpy or entropy, we are led to proceed


according to the following steps:
0 Choose an origin for which the values (arbitrary) will be given as standard enthalpy
or entropy (ideal gas, pressure Po = 0.1 MPa): i.e. at To,P" we have h" (To,P"),
s o (TOP")
0 Calculate the values of enthalpy and entropy, always at standard state, but at temperature T: ?;Po,h" (T,Po),so(T,P")
0 Calculate the values for enthalpy and entropy in the ideal gas state at temperature T
andatpressurep T,P,h#(T,P),s#(T,P)
0 Finally add the residual property (h - h#)rp
,or (s thus stating for enthalpy:
h = h"(T0,P")+ [h"(T,P")-h"(T~,P")]
+ [h#(T,P)-h"(T,P")]+ (h-h#),p

(2.52)

and for entropy:


s = S " ( T ~ , P " )[ +
s " ( T , P " ) - s " ( T ~ , P " [s#(T,P)-s"(T,P")]+
)]+
(S-S')~~

(2.53)

The calculations used for the ideal gas may be expressed (bearing in mind the fact that
the enthalpy of the ideal gas is independent of pressure) as:

h = h"(To,P")+

cp" d T + (h - h#)rp

(2.54)

TO

s=so(To,Po)+
\TO

P
d T - R l n - +(S-S#)~.
T
P"

(2.55)

2. Properties of Pure Substances

59

CONCLUSION
We have just seen that the calculation of thermodynamic properties is based on the understanding of the equations relating pressure, volume, and temperature, illustrating the
importance of the development of the equations of state. We have also emphasized the
particular role played by certain properties such as the critical points, vapor pressure, and
heat of vaporization. They make up our minimum database.
We must, of course, go beyond this minimum database. It is rare to have at our disposal
for a compound a complete set of values for molar volume, enthalpy, heat capacity,
entropy, etc., for the entire range of pressure and temperature. We must therefore apply
methods that are as predictive as possible.The following chapters are devoted to this subject.

REFERENCES
Ambrose D (1978a) The correlation and estimation of vapour pressures. I. A comparison of three
vapour pressure equations.J. Chem. Thermodynamics, 10,765-769.
Ambrose D (1978b) The correlation and estimation of vapour pressures. 1I.A new procedure for estimation and extrapolation. J. Chem. Thermodynamics, 10,765-769.
Ambrose D (1980) The correlation and estimation of vapour pressures. 111. Reference values for low
pressure estimations. J. Chem. Thermodynamics, 12,871-879.
Ambrose D (1985) The Evaluation of Vapour Pressure Data. Private communication.
Boublik T, Fried V, Hala E (1984) The Vapour Pressures of Pure Substances. Elsevier,Amsterdam.
Carruth GF, Kobayashi R (1972) Ind. Eng. Chem. Fundam., 11,509.
Dymond JH, Smith EB (1980) The Virial Coefficients of Pure Gases and Mixtures. Clarendon Press,
Oxford.
Daubert R, Danner (1986) DIPPR Data Compilation. AIChE J., New York.
Harlacher EA, Braun WG (1970) A four parameter extension of the theorem of corresponding states. Ind. Eng. Chem. Process Des. Develop., 9,479-483.
King MB, Al-Najjar H (1974) A method for correlating and extending vapour pressure data to lower
temperatures using thermal data: vapour pressure equations for some n-alkanes below the normal
boiling point. Chem. Eng. Sci., 29,1003-1011.
Lewis GN, Randall M (1923) Thermodynamics and the Free Energy of Chemical Substances.
McGraw-Hill Book Co., first edition.
Mosselman C, van Vugt WH, Vos H (1982) J. Chem Eng. Data, 27,246.
Reid RC, Prausnitz JM, Shenvood Th K (1976) The Properties of Gases and Liquids, third edition.
McGraw-Hill Book Co.
Reid RC, Prausnitz JM, Poling BE (1987) The Properties of Gases and Liquids, fourth edition.
McGraw-Hill Book Co.
Scott SW, Osborn AG (1979) Representation of vapor pressure data. J. Phys. Chem., 83,2714-2723.
Teja AS, Gude M, Rosenthal DJ (1989) Novel methods for the measurement of the critical properties
of thermally unstable fluids. Fluid Phase Equilibria, 52,193-200.

60

2. Properties of Pure Substances

Thek RE, Stiel LI (1966) A new reduced vapor pressure equation. AZChE J., 12,599.
Thek RE, Stiel LI (1967) (Erratum to preceding reference). AZChE J., 13,626.
Rogalski M, Mato Chain FA, Carrier B (1991) Vapour pressures of hydrocarbons, modelisation,
extrapolation and prediction. Chem. Eng. Sci.
Ruzicka K, Majer V (1994) Simultaneous treatment of vapor pressures and related thermal data
between the triple and normal boiling temperatures for n-alkanes C,-(&.J. Phys. Chem. Re$ Data,
23,l-39.
Yaws CL,Yang HC (1989) To estimate vapor pressure easily. Hydrocarbon Processing, 68, (lo), 65-68.

PredictingThermodynamic
Properties of Pure Substances
General Principles
Corresponding States
Group Contributions

In reviewing the behavior of fluids, we have emphasized that knowing certain properties
such as the critical points and vapor pressure, for example, are indispensable to the calculation of thermodynamic quantities. We have also specified the steps that this calculation
would most often follow: calculation of the properties in the ideal gas state, then the residual values. These steps are based on the equations that link pressure, volume, and temperature. It must be stressed that with few exceptions, these equations have not been determined for most pure substances, and aforfiori,for mixtures. We must therefore use predictive methods. With limited initial knowledge, these methods provide access to all thermodynamic properties. These methods exist, and their limitations cause them to change. The
simplicity and efficiency of some methods lead to misuse. We are in difficult territory
where any conclusion is obsolete from the moment it is drawn.
In essence, the problem facing us is one of accounting for the consequences of molecular interaction forces at the macroscopic level. Limiting ourselves to non-electrolyte
solutions, it is therefore appropriate to review a number of simple concepts.
Essentially, the forces of intermolecular attraction fall into four categories
[Emschwiller, 1951,p. 1363-13671.
0 Orientation forces are present between molecules having a permanent dipole
moment. This dipole moment reflects the asymmetry of electrical charges due to the presence of electronegative atoms.

The interaction energy depends of course on the respective orientations of the dipoles
but, according to the Boltzmann statistics, averaging may take into account all orientations. Using pi for the dipole moment, r for the distance between dipoles, and k as the
Boltzmann constant, we find that:

62

3. Predicting Thermodynamic Properties of Pure Substances

Induction forces: a permanent dipole may induce an electrical asymmetry in a neighboring molecule that is not polar, but polarizable. Interaction energy between a permanent
dipole and an induced dipole is established, which depends on the polarizability a of the
molecule with the induced dipole. Its is expressed as:

0 Dispersion forces: we must take into account the cohesion forces that exist between
non-polar molecules such as hydrocarbons. We may imagine that spontaneous dipoles
exist that are of no consequence in time or space, but that are themselves susceptible to
interaction by induction. A simplified calculation yields the following expression for the
energy of dispersion as a function of polarizability a,the oscillation frequencies v, and
Plancks constant h:

0 Hydrogen bonding: finally, we know that two molecules may be bound by a hydrogen
bond. Such is the case with alcohols undergoing polymer chain autoassociation and, of
course, with water. These bonds may form between molecules of a similar nature (autoassociation) or molecules of a dissimilar nature (complexation) as with water/alcohol solutions or ether/alcohol solutions.

Besides the intermolecular attraction forces, there are repulsion forces as may be
observed by the low compressibility of the liquid phase. These forces prevent molecules
from approaching one another beyond a certain limit, such as hard spheres (or chains).
It is very difficult to tell what boils down to the forces of orientation, induction, and
dispersion in a macroscopic property even though it represents molecular interactions
such as heat of vaporization. We need to remember that if we are talking about complex
molecules, molecular interaction is the result of interactions at work among the
component groups of two neighboring molecules. For example, Figure 3.1 shows the
variation of cohesion energy E, (itself derived from the heat of vaporization) for linear
paraffins and some nitriles as a function of the number of carbon atoms [Meyer, 19711. For
long chain hydrocarbon nitriles, the dipolar interactions are somewhat diluted as the
length of the chain increases such that the amount of orientation energy, E,,, decreases.The
shift of this asymptote compared to straight line E,(n) which refers to the paraffins
represents induction energy. We note that the dispersion forces, E d , which are solely
responsible for the hydrocarbon cohesion forces, also make up the main cohesion forces of
the nitriles.
Most predictive methods that we shall encounter were developed for non-polar molecules, in particular the hydrocarbons. It is possible that when applied to polar compounds
they lead to good results, but prudence dictates that we treat such results as fortunate coincidences and refrain from an application that would most often yield gross errors.
Nevertheless, we shall point out some widely applied extensions.

3. Predicting Thermodynamic Properties of Pure Substances

63

Number of carbon atoms

Figure 3.1 Comparison of the cohesion energy of paraffins and


nitriles [Meyer et al., 19711.

3.1

TECHNIQUES OF MOLECULAR SIMULATION

It is not within the scope of this study to introduce molecular simulation techniques. They
are in and of themselves a true specialty.We shall mention them only briefly in order to
emphasize their potential and rapid development.
Our knowledge of molecular interactions and the progress shown in the field of scientific calculators allow us to calculate the macroscopic properties for a number of molecules. A molecular model and a representation of the energy of intermolecular interaction
are necessary. We apply either the Monte-Carlo method or molecular dynamics.
The Monte-Carlo method randomly produces several molecular groupings and retains
only those that obey the Boltzmann statistics for a given volume and temperature. Molecular
dynamics uses newtonian mechanics and solves the equations for molecular movement. In
both cases, the macroscopic properties are estimated by obtaining appropriate averages.
A fundamental question is posed when we apply these molecular simulation techniques.
What is the validity of the applied potential? In many cases, we have an empirical model,
such as those of Lennard Jones or Kihara (see Eqs. 3.4 and 3.8). We may also consider that
the intermolecular interaction forces result from interactions between the groups that
make up the complex molecules. It is therefore advisable to remain suspicious of the
results of these calculations and compare them to experimental values.
Furthermore, they generally require very complicated calculation methods. For this reason, they may not be considered operational substitutes for the empirical models applied

64

3. Predicting ThermodynamicProperties of Pure Substances

by chemical engineers. On the other hand, molecular simulation is a rapidly evolving discipline [Gubbins, 19891. In the future, it will undoubtedly produce reliable values and supplant experimental data to a certain extent.

3.2

THE CORRESPONDING STATES PRINCIPLE

We have repeatedly mentioned intermolecular cohesion potentials.Many expressions proposed for them are, as we shall see, used for the calculations for molecular simulation.
Figure 3.2 provides two classic examples of such expressions.The first is the Lennard Jones
potential, resulting from a combination of attraction forces acting at an average distance,
and the repulsion forces that predominate at short distances. Its expression:
E

= 4%

[(

:)12

I):(
-

(3.4)

shows two intrinsic parameters of the compound in question: the value for the distance
between the centers of attraction at which the potential cancels itself out, r,,, and the minimal value of potential %. The second example is the square well example.The molecules
are treated as rigid spheres and the distance between the centers of the molecules always
remains greater than at a certain value. In an interval defining the width of the well, the
potential takes on a constant value and beyond it cancels itself out. This very schematic
model allows for relatively simple calculations and itself introduces the two characteristic
parameters: the width and depth of the well.
In general, if we state that:
the potentials related to the degrees of internal freedom of molecules (translation,rotation, vibration), and to external freedom (cohesion) may be expressed independently;

r/ro

Figure 3.2 Lennard Jones potential and Square Well potential.

3. Predicting Thermodynamic Properties of Pure Substances

65

the cohesion energy of a system is expressed using the sum of the binary interactions
between the centers of attraction;
0 these interactions are reduced to the dispersion forces and may be calculated using an
expression in the form E = EO F (r/ro),in which the function F is a universal function,
since only E and ro are related to the nature of the substance. It is well understood
that this function F may differ from that of Lennard Jones.
It is then possible to demonstrate that the macroscopic effects of molecular interactions
are expressed as a function of (1) two parameters characteristic only of the substance
(characteristics corresponding to EO and r o ) ,and (2) variables of state such as pressure and
temperature. The corresponding states principle is therefore not an axiom, but a demonstrable theorem.
However, it is appropriate to point out that this law was stated by van der Waals well
before studies of intermolecular potentials.
The equation of state that is derived from the configuration integral by applying the
preceding hypotheses has the following form:

where Z is the compressibility factor, is the Boltzmann constant, and I\I is Avogadros
number. Stating that it verifies the critical conditions (Chapter 2, Eq. 2.4), we obtain the
values for Z,, dkT,, and v,/No3,which are universal, meaning that they are the same for
all substances satisfying the initial hypotheses.
In particular, we may define a characteristic temperature T* = d k , and a characteristic
volume o* = No3,or any other group of macroscopic parameters. The choice is generally
taken from the critical points themselves, preferably the critical temperature and the critical pressure due to the imprecision distorting the critical volume values.
We define the reduced variables:
reduced temperature: T, = T/T,
(3.5)
reduced pressure: P, = P/P,
and the corresponding states principle may be stated in a very simple form:
Deviations from the ideal gas laws (or residual values) depend only on reduced
variables.
Figure 3.3 shows that the change in compressibility factor Z is actually a function of
reduced pressure P, at the given reduced temperature (here, 0.96,1.02, and 1.2) regardless
of the nature of the compound in question (here, methane, ethane, propane, and n-butane,
represented respectively by the symbols A, X, 0and 0).
If we include this principle in the calculation of the compressibility factor, it is then also
applied to the determination of the fugacity coefficient, the adimensional residual terms
for enthalpy (h - h#)/RT,entropy (s - s#)/R,and heat capacity (cp-c;)/Z?. We shall demonstrate its use for the residual enthalpy and for the fugacity coefficient.
To calculate the residual enthalpy we apply Equation 2.27:
(2.27)

66

4t

a
a
0

In

4t

a
0

it

;a

3. Predicting Thermodynamic Properties of Pure Substances

N
7

II

a a

I
u!

a
0

*
a

a
0

*
a

a
0

a
a
0

*
a

a
0

*
a

a.

*0
a
0 0

a .

a.

aa

* I

* *

**

00

a.

aa

*0*0

08

I
In

a
0

a a
.a
o m

* **

a a
000

*w

aa
oQ3

.a

**

I
7

3. Predicting Thermodynamic Properties of Pure Substances

67

We express the volume by the equation:

RT
V=ZP
and so:
Thus we obtain:

Similarly, to calculate the fugacity coefficient we write:


In rp=

(L- L) d P =
RT

1"
0

( Z - 1) dPr
Pr

(3.7)

It can be seen that if the compressibility factor obeys the corresponding states principle,
that is to say, it is a function only of T, and of P, then it is the same for adimensional residual enthalpy and the fugacity coefficient.
Yet, this corresponding states principle provides only a first approximation, sometimes
a rough one, of residual terms.The deviations are more particularly marked when we apply
it around the critical point or in the liquid phase, and finally to the calculation of vapor
pressure. The critical compressibility factor at T, = 1and P, = 1,varies in fact within a relatively wide range, as shown in the data in appendix 1.Similarly, the vapor pressure curves
for hydrocarbons drawn for a system with reduced variables T J T and In P/Pc (Fig. 3.4) do
not merge into a unique line, but into a bundle of quasi-linear curves. The values or the
critical compressibility factor and the slopes of the vapor pressure curves shift in a regular
fashion with the length of the paraffin chain, or with molecular volume, which is in this
case equivalent. In order to better characterize a compound, it seems necessary to introduce a third characteristic parameter that takes these variations into account.
This is indeed what is done at the level of interaction energy calculation when we use
the Kihara potential. It differs from that of Lennard Jones in that it introduces the term d
to represent the molecular diameter:
&=4%

r-d

r-d

From this statement, several correlations have been developed that are distinctive in
their choice of this third parameter and the database from which they were established.
Indeed, it is appropriate to specifically state that if the basis for these correlations relies on
an acceptable theoretical basis, their development assumes the acquisition of experimental
data for, generally, volumetric and vapor pressure behavior, and sometimes enthalpic
behavior, for compounds whose critical points are known, as well as the value of this third
parameter.

68

3. Predicting ThermodynamicProperties of Pure Substances

-0.5

-1

.....................

z
UJ
0
-1.5

-2

T,= 0.7

-2.5
1

1.2

1.4

1.6

1.8

Tc f l

Figure 3.4 Vapor pressure curves of n-paraffins (C, - C,) in reduced


variables.

Although in principle these correlations apply to the estimation of residual terms in liquid and vapor phase, the size of the terms as they relate to the liquid phase (heat of vaporization,compressibilityfactor, etc.) decreases the precision of the calculation for this state.
In general, we will have an advantage in calculating these values using correlations that are
specific to the dense phase. We will give a few examples later on. As for the most general
methods, in particular we cite those developed by adopting the following as the third
parameter:
0 the critical compressibility factor
0 the acentric factor, defined by the equation:
0 = -loglo

P,- 1

(3.9)

where P, is the reduced vapor pressure for T, = 0.7.


This last parameter was introduced by Pitzer [1955] as a measure of the acentricity of
a molecule. Indeed, it is practically zero in compounds where the molecule is spherical
(rare gases, methane) and in paraffinic hydrocarbons, for example, it increases with the

69

3. Predicting ThermodynamicProperties of Pure Substances

length of the chain. The idea to introduce a third parameter related to the vapor pressure
curve in applying the corresponding states principle is not limited to Pitzer. Riedel [1954]
proposed a parameter that has become known as the Riedel factor:
dln Po
d In T

a=-

for T = T,

(3.10)

However, the calculation of the Riedel factor requires values that are not readily available, and this proposition has not lead to any important developments.
The acentric factor and the critical compressibility factor are related as shown in
Figure 3.5, and the equation proposed by Pitzer to represent this dependence is:
2, = 0.291 - 0.08 o

(3.11)

We note, however, that this correlation is not perfect.There is no similarity between the
methods that is based on one or the other of these two parameters (Z, or w).

0.3

No 0.29 -

--

0.28-

fn

0.27 -

.2.
.0
fn
p!

. .

8 0.26.d
c
0.25 0.24

0.1

0.2

0.3

0.4

0.5

Acentric factor w

Figure 3.5 Relationship between the critical compressibility factor Z,,


and the acentric factor o.

3.2.1 Correlations Using the Critical Compressibility Factor


3.2.1.1

Watson Method

A number of correlations pertaining to the compressibility factor, the residual terms


of internal energy (u - u#)/T,,enthalpy (h - h#)/T,,entropy (s - s#)/T,,heat capacity at
constant pressure ( c p- c!)/ T,, and the fugacity coefficient have been established [Hougen

70

3. Predicting ThermodynamicProperties of Pure Substances

et al., 19591 based on data relating to 82 components, taking into account a critical

compressibility factor range from 0.23 to 0.30. For values of this parameter that are less
than 0.26, some polar, and even some autoassociative compounds such as water, were
included in the database. They certainly do not conform to the strict framework in which
the corresponding state correlations should be applied, but form a rather important extension of it.
In general, the residual term is first calculated for a compound whose critical compressibility factor is equal to 0.27. A second order correction is then applied, which is proportional to the deviation between the actual value of the critical compressibility factor and
0.27. The proportionality coefficient is a function of the reduced variables T, and P,. For
properties of saturated phases, the residual terms are given as a function of the reduced
temperature for several values of the critical compressibility factor. From this correlation
we come away with two characteristics: primarily that the reference value for the critical
compressibility factor essentially corresponds to an average value that is valid at first
approximation for the most common apolar compounds. The application of the correlation
at low critical compressibility factor values was made using data for polar substances.
However, we should emphasize that this extension has no theoretical basis as the corresponding states principle is applicable to apolar compounds only, and results in inaccuracies that are difficult to evaluate.

3.2.1.2

Rackett Equation

In addition to the general correlations developed by Watson, it is appropriate to mention


an equation for the calculation of molar volume of the saturated liquid phase proposed by
Rackett [1970]. It is simple, precise, and in its initial form is written as:
In - = (1- T ) 2 l n Z c
(v::)

(3.12)

If we express the critical volume as a function of the critical temperature, pressure, and
compressibility factor, we arrive at the equivalent form:
(3.13)
Its application already results in satisfactory precision of 2.4% for a database of
36 hydrocarbons. However, it was proposed [Spencer and Danner, 19721 that the critical
compressibility factor be substituted with a variable specific to this method, the Racket
compressibility factor, ZRa,resulting in the expression:
(3.14)
Table 3.1 shows the values of this parameter for several hydrocarbons. It may also be
estimated by reversing the preceding equation using a value for the molar volume in the
liquid phase that has been obtained from elsewhere.

3. Predicting ThermodynamicProperties of Pure Substances

71

Table 3.1
Rackett compressibility factor, ZRa[Spencer and Danner, 19721
Compound
Methane
Ethane
Propane
n-Butane
Isobutane
n-Pentane
n-Hexane
n-Heptane
n-Octane
n-Nonane
n-Decane

Compound
0.287 6
0.278 9
0.276 3
0.272 8
0.275 0
0.268 5
0.263 5
0.261 1
0.256 7
0.263 7
0.246 6

Ethylene
Propylene
2-Butene (trans)
Isobutene
Cyclohexane
Methylcyclohexane
Benzene
Toluene
Ethylbenzene
m-xylene
p-xylene

'Ra

0.281 0
0.278 5
0.272 1
0.272 7
0.272 9
0.269 9
0.269 6
0.264 6
0.262 6
0.259 3
0.258 9

3.2.2 Correlations Using the Acentric Factor


Compared to the correlations that we have just discussed, the correlations based on the
corresponding states principle and utilizing the acentric factor as the third parameter have
been reviewed and developed constantly since the introduction of the Pitzer method
[1955]. They are applied to the calculation of residual terms, as well as to the equations of
state, as we shall see later.
The acentric factor is defined from the average slope of the vapor pressure curve compared to a straight line in a system with variables T c / T and In ( P J P ) . Its calculation
requires the knowledge of the critical point and the vapor pressure at a reduced temperature T, = 0.7. For most compounds, the data are available and the vapor pressure is generally known at this reference temperature with satisfactory precision. The fact remains that
the definition of the acentric factor relies on data whose values may vary as a function of
the precision of the experimental techniques. According to the database, we may find values for this parameter that are themselves slightly different.
If vapor pressure actually varies in a linear fashion in a system with variables T c / Tand
In ( P c / P ) ,the acentric factor may be arrived at using any vapor pressure data, especially
boiling temperature Tb at atmospheric pressure Pb.We may then write:
D

(3.15)

This equation has the advantage of using only values that are readily available in any
database, but provides only one value approaching that of the acentric factor. Similarly, to
the calculation of vapor pressure we may apply the equation:
Po 7
log - = -- (1 + w)
- 1)
(3.16)
pc
3
and obtain a precision that is equivalent to that of Equation 2.12.

(;

72

3. Predicting Thermodynamic Properties of Pure Substances

We shall provide some examples of methods using the acentric factor. Later on, they
will provide us with a complete practical estimation of thermodynamicproperties.

EXAMPLE 3.1

Comparison of values from three databases


It may happen that the data we find in the literature is contradictory.Such is the case for
the data for pentene 1as shown in Table 3.2 below, which gives the critical properties, the
acentric factor, and the boiling temperature at atmospheric pressure. The values in the
third database differ considerablyfrom those provided by the two other databases for the
critical pressure and, to a lesser degree,for the acentric factor. Can we make a choice?

Database

Tc
(K)

pc
(at4

(1)
(2)
(3)

464.78
464.8
464.7

34.81
34.68
40

0.233
0.233
0.245

Tb

(K)
303.11
303.1
303.1

The inherent consistency of each database will be tested by applying Equation (3.16),
at a value of the temperature equal to T,,. We find that the first two databases yield
results close to one atmosphere (1.0167 atm and 1.0121 atm respectively), while we
obtain 1.1304 atm using the third database, which will therefore be removed.

3.2.2.1

Prediction of the Second Virial Coefficient

We know (Chapter 2) that the second virial coefficient, B, is defined by the equation:
(3.17)
Its value may be predicted using a correlation proposed by Pitzer and Curl [1957] and
modified by Tsonopoulos [1974] who also expanded it to include polar compounds. It
takes the form:
(3.18)
with

0.330 0.1385 0.0121 0.000607


F(O)(T,.)= 0.1445 - -- -- -Tr
T,2
T)
T,8

(3.19)

and:

0.331 0.423 0.008


F(')( T,.)= 0.063 7 + -- -- T:
T)
T,8

(3.20)

73

3. Predicting Thermodynamic Properties of Pure Substances

-1.0--

I
I

I
I

I
I
I

I
I
I

I
I
I

I
I

I
I

I
I

I
I
I

I
I
I

I
I

I
I

I
I
I

I
I
I

I
I
I

I
I
I

I
I
I
I

2.2

2.4

I
I
I

I
I
I

I
I
I

I
I
I

I
I

- -

I
I
I

I
I
I

-1.5
0

0.2

0.4

0.6

0.8

1.2

1.4

1.6

1.8

Reduced temperature

Figure 3.6 Calculation of the second virial coefficient using


the Tsonopoulos correlation (simple fluid).
0.5

0.6

0.8

0.2

I
0.4

1.2

1.4

1.6

1.8

2.2

I
1

I
2.4

Reduced temperature

Figure 3.7 Calculation of the second virial coefficient using


the Tsonopoulos correlation (correction term).

Figures 3.6 and 3.7 represent the functions F(O) and F(').
For polar components that are not autoassociated by hydrogen bonding, an additional
term is necessary to take into account the variation of the second virial coefficient with
temperature, which is more pronounced. The development of a predictive correlation in

74

3. Predicting ThermodynamicProperties of Pure Substances

which the components are characterized by their dipole moments is, however, hindered by
the scarcity of experimental data. It is the same, a fortiori, for autoassociative compounds.

3.2.2.2

Properties at Liquid-Vapor Equilibrium

Table 3.3 allows for the calculation of vapor pressure, the compressibility factor of the saturated vapor phase, and entropy of vaporization as a function of reduced temperature.The
following equations are used:
logloPp= (log,oP~)(o)+
OA(lOg,,P~)
(3.21)
ZVU= (ZVU

A s u = (Asa)(')

1(0) + W M V U

(3.22)

+ wAl(Asa)+ o 2 4 ( A s U )

(3.23)

In particular, it corresponds to an equation for vapor pressure in reduced variables:


4.318 1.454 0.3456
- logloPp = 3,209 - -+ -- T,
T,2
TP
2.008 2.524 0.3981
0.1175 + -- (3.24)
Tr
TF

'7)

Table 3.3
Correlation of properties at saturation [Pitzer et al., 19551
- (log P,?)'')

1
0.99
0.98
0.97
0.96
0.95
0.94
0.92
0.90
0.88
0.86
0.84
0.82
0.80
0.78
0.76
0.74
0.72
0.7
0.68
0.66
0.64
0.62
0.60
0.58
0.56

0
0.025
0.050
0.076
0.102
0.129
0.156
0.212
0.270
0.330
0.391
0.455
0.522
0.592
0.665
0.742
0.823
0.909
1
1.096
1.198
1.308
1.426
1.552
1.688
1.834

I -A (log ,,P,?)
0
0.021
0.042
0.064
0.086
0.109
0.133
0.180
0.230
0.285
0.345
0.405
0.475
0.545
0.620
0.705
0.8
0.895
1
1.12
1.25
1.39
1.54
1.70
1.88
2.06

0.291
0.43
0.47
0.51
0.54
0.565
0.59
0.63
0.67
0.70
0.73
0.756
0.781
0.804
0.826
0.846
0.864
0.881
0.897
0.911
0.922
0.932
0.94
0.947
0.953
0.959

A Z v0

(As")(')

-0.08
-0.03
0
0.02
0.035
0.045
0.055
0.075
0.095
0.110
0.125
0.135
0.140
0.144
0.144
0.142
0.137
0.131
0.122
0.113
0.104
0.097
0.090
0.083
0.077
0.070

0
2.83
3.38
4
4.52
5
5.44
6.23
6.95
7.58
8.19
8.79
9.37
9.97
10.57
11.20
11.84
12.49
13.19
13.89
14.62
15.36
16.12
17.02
17.74
18.64

0
2.83
3.91
4.72
5.39
5.96
6.51
7.54
8.53
9.39
10.3
11.2
12.1
13
13.9
14.9
16
17
18.1
19.3
20.5
21.8
23.2
24.6
26.2
27.8

0
0.6
0.9
1.1
1.3
1.4
1.5
1.8
2
2.2
2.4
2.5
2.6
2.7
2.8
2.9
2.9
2.8
2.8
2.7
2.6
2.5
2.4
2.3
2.2
2.1

75

3. Predicting Thermodynamic Properties of Pure Substances

3.2.2.3

l e e and Kesler Method

The general method proposed by Pitzer rests on the hypothesis that the residual properties
are, at a given reduced pressure and temperature, linear functions of the acentric factor.
This hypothesis was adopted by Lee and Kesler [1975] who developed a linear interpolation of residual properties based on:
0 those of the simple fluid (o= 0), which were calculated using data for argon, krypton, and methane;
0 those of a heavy reference fluid (o= 0.3978), n-octane, for which volumetric and
enthalpy data were available.
The volumetric properties of these fluids are represented by a modified equation of
state from Benedict, Webb, and Rubin (see Chapter 4):
(3.25)
where v rrepresents an adimensional form of the volume:
(3.26)
and is therefore found to be related to the compressibility factor Z by the equation:
Vr
z = Pr
-

(3.27)

Tr
Variables B, C, D are temperature dependent:
(3.28)

c = c , -c2- 2

(3.29)

D = d , + -d2
Tr

(3.30)

3,

Tr

Variables bi,ci,di,p, and yof this equation are of course different according to whether
we are dealing with methane or octane (Table 3.4).
Table 3.4
Parameters of the Benedict, Webb, and Rubin equation of state applied
by Lee and Kesler to a simple fluid (o= 0), and a reference fluid ( w = 0.3978)

Parameter

Simple Fluid

Reference Fluid

Parameter

Simple Fluid

Reference Fluid

bl
b2
b3
b4

0.118 1193
0.265 728
0.15479
0.030 323
0.023 674 4
0.018 6984

0.202 657 9
0.331 511
0.027 655
0.203 488
0.031 3385
0.050 361 8

c3
c4
104.d,
104.d2

0
0.042 724
0.155488
0.623689
0.653 92
0.060 167

0.016 901
0.041 577
0.487 36
0.074 033 6
1.226
0.037 54

C1

c2

76

3. Predicting ThermodynamicProperties of Pure Substances

The method applies to residual properties at given pressure and temperature, compressibility factor Z , enthalpy (h - h#)/RT,entropy (s - s#)/R,heat capacity at constant
pressure (cp- c!)/R, and the fugacity coefficient.The range covered for pressure and temperature according to the authors is as follows:

0,3 G T, G 4

and

0 s P, s 10.

Calculations are carried out by:


determining the reduced variables of the fluid to which we want to apply the
method:
determining, using the equation of state for residual properties (compressibility factor, for example) of the simple fluid Z(O)and of the reference fluid Z(,) corresponding
to the reduced variables:
0 performing linear interpolation:

(3.31)
As an example, Figures 3.8 and 3.9 show the compressibility factor of the simple fluid
and the coefficient AZ. Tables A2.1 to A2.10 (Appendix 2) provide the corresponding values for residual quantities (enthalpy, entropy, heat capacity at constant pressure) and for
the fugacity coefficient.
The calculation of vapor pressure may be accomplished by application of the equilibrium equation (equality of the fugacity in the vapor and liquid phase).

.2

Figure 3.8 Calculation of the compressibility factor for the simple fluid Z(O)
using the Lee and Kesler method.

77

3. Predicting Thermodynamic Properties of Pure Substances

I
I
I

I
I
I

I
I

I
I
I

I
I

Figure 3.9 Calculation of the compressibility factor


using the Lee and Kesler method: correction term AZ.

However, the authors suggest using the following equation:


lnPP=5.92714-

15.2518-

6.096 48

___ - 1.28862 In

Tr

15.6875
~

Tr

Tr + 0.168347T:
(3.32)

- 13.4721 In T, + 0.43577 T:

This last equation may also be applied to the calculation of the acentric factor if we
know data for the vapor pressure.
There is not an exact equivalence between the vapor pressure resulting from the equal
fugacity condition with fugacities calculated using the Lee and Kesler method, and the
result from Equation 3.32. At low reduced temperature, the differences are small and
inconsequential. On the other hand, once the reduced temperature surpasses approximately 0.98, the value resulting from Equation 3.32 does not allow for the calculation of
saturated vapor densities as the equation of state for the reference fluid has no root corresponding to that phase.
The Lee and Kesler method may also be applied to mixtures by defining pseudocritical variables (see Section 3.2.3.2). The precision is very satisfactory for the calculation of
residual properties and, while maintaining its principle, we can adapt it to other problems.
However, prudence is advisable when we are talking about fluids whose acentric factor is
clearly higher than that of the reference fluid (0.3978).We then find ourselves extrapolating with respect to the database that allowed us to develop the method, and it is difficult to
appreciate the accuracy of the results. Of course, the method is not valid for polar compounds, at least in its original form.

78

3. Predicting ThermodynamicProperties of Pure Substances

EXAMPLE 3.2

Calculation of the isochoric change of a fluid


We shall determine the pressure change in a bottle of liquefied gas as a function of
temperature.We consider two cases: first, a bottle, with a capacity of 1 liter, containing
314.6 grams of ethane; and second, a bottle with the same capacity that contains only
113.24 grams of ethane. The range of temperature goes from approximately 0" to
60C.
The first point to examine is the state of ethane in the bottle for temperatures less
than the critical temperature. If two phases are present, the pressure is equal to the
vapor pressure, and we can calculate the proportion of liquid phase. If there is only
one phase, the pressure will then be determined using a Z( T,, P,, o)correlation. As
the range of temperature extends beyond that of the critical temperature, we shall
certainly encounter this last correlation.
For a two-phase system, the proportion of liquid phase is determined by balancing the
volume, which is the given property. We will use the following notations:
Nt :total number of moles
NL :number of moles in liquid phase
N' :number of moles in vapor phase
V :total volume
w :overall molar volume (w = V/NJ
wL~":molar volume in liquid phase
WK" :molar volume in vapor phase
X :proportion of liquid phase
We may write:
N~+N'=N~
N L ~ L , ~NvwV:O=
+
V

and derive from it:

The proportion X must be between 0 and 1. If it is negative, the system is homogeneous vapor. If it is greater than 1, it is homogeneous liquid.
What remains is to choose a way to calculate the vapor pressure, and the molar volume in the liquid and gas phase. For the vapor pressure, we use the Lee and Kesler
correlation (Eq. 3.32). The compressibilityfactors and the molar volumes will also be
calculated using this method for the vapor phase, but by using the simpler Rackett
correlation (Eq. 3.14) for the saturated liquid phase.
We shall use the following data:

T, = 305.4 K, P, = 4.88 MPa, o = 0.099, Z , = 0.2789, M = 30.1 g mol-l


The total number of moles is equal to N l = 314.6/30.1 = 10.45 moles for the first bottle, and N2 = 113.24/30.1 = 3.76 moles for the second bottle. The given volumes are
respectively wl = 1,000 /10.45 = 95.69 cm3/mol and w 2 = 1,000/3.76= 265.6 cm3/mol.

79

3. Predicting Thermodynamic Properties of Pure Substances

As the applied methods are based on the reduced temperatures, the calculations will
be performed between T, = 0.9 and T, = 1.1.
We shall specify the calculation of the compressibility factor of the saturated vapor
phase at T, = 0.94 ( T = 287.1 K). By applying the vapor pressure correlation (3.32) we
find that P," = 0.6758, that is P = 3.30 MPa. After solving the Equation (3.25) applied
to a simple fluid and then to a reference fluid, we find:
Z(O)= 0.61103, Z@)= 0.52950

and

A Z = -0.20494

and therefore:
Z = Z(O)+ oAZ = 0.61103 - 0.099 * 0.20494 = 0.59074
and vKc = ZRT/P = 427.6 cm3 . mol-I .
Table 3.5 below summarizes the results of the calculations.After the liquid mole fractions X I and X,, we have shown the liquid volume fractions Y l ,Yz,
which are more
representative of the height of the liquid vapor interface in one or the other bottles.
They are calculated by the equation:

xiv

y.= -,

i = 1,2

Table 3.5
Isochoric change (two-phase system)
Tr

vKU

(cm3.
mol-')

(MPa)

0.90
0.92
0.94
0.96
0.98

274.88
281
287.1
293.2
299.3

0.5101
0.5885
0.6758
0.7727
0.8804

2.49
2.87
3.30
3.77
4.30

0.082
0.096
0.113
0.135
0.165

74.91
78.03
81.95
87.23
95.59

0.674
0.636
0.591
0.536
0.458

619.1
517.1
427.6
346.4
265.5

I I
x2

y1

0.18
0.17
0.15
0.10
0

We note that the fraction of liquid in the bottles was calculated such that the twophase region extends for one as it does for the other, to T = 299.3 K. At this temperature, the first bottle is full of liquid, the second full of vapor, and beyond that the calculations apply to a single-phase system. We continue by applying the Lee and Kesler
method offering a detailed example for T, = 1.1, with the bottle filled with vapor. We
must verify the equation:

in which Z(T,, P,, 0) stands for the compressibility factor expression provided by the
Lee and Kesler method. The calculation is iterative. Since the reduced temperature is
given, we make an assumption for the value of the reduced pressure. We can also calculate the compressibility factor and see if the preceding equation is valid. Table 3.6
below gives the details of the calculation.

80

3. Predicting Thermodynamic Properties of Pure Substances

Table 3.6
Calculation of pressure as a function of temperature and volume by application of the Lee and
Kesler method: 113.24 grams of ethane in a one-liter bottle at a temperature of 336K (Tr= 1.1)

1.2
1.22
1.24
1.26
1.28

0.598
0.589
0.579
0.569
0.559

0.634
0.627
0.619
0.611
0.604

0.089 7
0.095 1
0.101 0
0.1066
0.1125

0.607 3
0.598 1
0.588 9
0.579 6
0.5702

0.5567
0.5660
0.575 3
0.584 5
0.593 8

Through interpolation we arrive at the result: P, = 1.2548,and P = 6.12 MPa.


For the bottle filled with liquid, the calculations would be done in an analogous manner.
Table 3.7 summarizes the results of the calculation for the two bottles between 300 K
and 335 K.
Table 3.7
Isochoric change (single-phase system)
First Bottle

0.98
0.99
1
1.02
1.04
1.06
1.08
1.1

Pr

299.3
302.4
305.4
311.5
317.6
323.7
329.8
336

0.88
1.031
1.171
1.457
1.746
2.036
2.327
2.619

4.3
5.03
5.71
7.11
8.52
9.94
11.36
12.78

7
Second Bottle
Pr
P
pr

0.88
0.914
0.946
0.914
1.009
1.071
1.132
1.194
1.255

4.3
4.46
4.92
t:i6
4.62
4.92
5.22
5.53
5.83
6.12

We note that in the case of the bottle that fills with liquid, two distinct methods have
been applied: the Rackett equation for the calculation of molar volume of the saturated liquid phase up to Tr = 0.98, on the one hand, and on the other hand, the Lee
and Kesler method for the compressed liquid, as the Rackett method is not valid in
the compressed state, beyond T, = 0.98. The two methods yield results that differ by
1% at their common point (T, = 0.98).This divergence introduces a small discontinuity in the calculated change of pressure as a function of temperature.
Figure 3.10 shows the pressure dependency on temperature for the two bottles. Up to
T, = 0.98, this change corresponds to the vapor pressure curve. Beyond, we enter the
homogeneous zones following two isochoric lines, which we note are practically linear. The increase in pressure is moderate in the bottle filled with vapor and more
rapid in the one filled with liquid. This example is therefore representative of a very
simple security problem concerning the filling limit in the case of liquefied gas.

3. Predicting Thermodynamic Properties of Pure Substances

280

300

320

81

340

Temperature (K)

Figure 3.10 Isochoric change of ethane in a two-phase system,then


a single-phase system.

We could compare the results of this calculation to the graph presented in Figure 2.2.The
results would have been very close. On Figure 2.2, the isochoric evolution investigated
here is represented by vertical lines at the two values for the overall molar volume.
In this context, we must mention the experiment called the Natterer tubes. Several
sealed glass test tubes contain a compound for which we want to determine the critical temperature. From one bottle to the other, the filling varies such that, as in the previous example, starting with the two-phase state as evidenced by the observation of
the liquid vapor interface, we may, by increasing the temperature, observe that some
of the test tubes fill with liquid, and others with vapor. If the filling corresponds
exactly to the critical density, the interface then remains at an almost invariant level
and, as we cross the critical temperature, disappears in situ giving rise to the critical
opalescence phenomenon. In this case, the change of pressure as a function of temperature merges with the vapor pressure curve up to the critical point. Beyond it follows the critical isochore, which is practically linear and for which the slope is equal to
that of the vapor pressure curve at the critical point.

3.2.3 Extensions to the Corresponding States Principle


3.2.3.1

Extension to Polar Compounds

Attempts have been made to expand the correlations of corresponding states to polar
compounds. We have cited above the inclusion of such compounds in the database used

82

3. Predicting Thermodynamic Properties of Pure Substances

to determine residual quantities as a function of the critical compressibility factor


(Section 3.2.1) as well as for the second virial coefficient.The first problem encountered is
to characterize such compounds.The parameters used up to this point (critical points, critical compressibility factor, or acentric factor) are themselves sensitive to polarity and
autoassociation by hydrogen bonding. Their values clearly deviate from the correlations
that may be put forth for the hydrocarbons as a function of molecular weight, for example.
The choice of dipolar moment turns out to be deceptive. Finally, the experimental data,
when they exist, are often limited to vapor pressure, density in the liquid phase, and possibly to the second virial coefficient. However, it is appropriate to cite the work of Stiel who
introduced a polarity factor, analogous to the acentric factor, starting from the vapor
pressure determined at a reduced temperature of 0.6:
X = log lo(Pu(T, = 0,6)/Pc) + 1 . 7 +~1.552

(3.33)

The method has been applied to the calculation of molar volumes of the saturated liquid phase [Halm and Stiel, 19701 and then of the compressibility factors for the vapor and
liquid phases [Stipp et al., 1973, Halm et al., 19851 via the equation:

z = z(0)+ wz(1) + xz(2)

(3.34)

where values Z(O),Z(),etc., are functions of the reduced variables.Finally,more recently an


extension to the Lee and Kesler method has been proposed that includes a third reference
compound,water [Wu and Stiel, 19851, for which the properties are represented using the
Keenan and Keyes [1969] equation of state.

3.2.3.2

Extension to Mixtures

Although we have not yet approached the subject of mixtures and the calculation of their
thermodynamic properties, we shall present here the so-called method of pseudocritical
variables (or properties).
The composition of a mixture is known by the mole fraction of each component.We will
consider the mixture as a pure substance whose critical properties (and possibly the acentric factor) are obtained from the same properties of pure substances using the rules of
empirical weighting.We can therefore calculate the reduced variables of the mixture using
the methods developed for pure substances,namely their residual quantities:compressibility factor, residual enthalpy, etc.
First, we shall consider the case of correlation of corresponding states of two parameters (critical temperature, and critical pressure). The most simple of the rules of weighting
[Kay, 19361 is linear:
Tc,m = CziTc,i and Pc,m = x z i P c , i
(3.35)
where zi denotes the mole fraction of component i.
We may also apply such a linear rule to the acentric factor.
The results obtained in this manner are generally acceptable, especially for the vapor
phase, taking into account the simplicity of the method. Of course, the vapor pressure of a
mixture cannot be calculated in this way since for a mixture, at a given temperature, the
limits of the two-phase range (bubble and dew point pressure) do not have the same value

3. Predicting Thermodynamic Properties of Pure Substances

83

and the liquid vapor equilibrium can be calculated only if we can measure the contribution
of each component to the Gibbs energy of the phases in equilibrium. The fugacity coefficient of the mixture that is determined by applying the method of pseudocritical variables
may therefore not be directly applied to the calculation of the equilibrium, and represents
only the residual property of the mixture.
The Lee and Kesler method can be applied to mixtures. The authors recommend the
following mixing rules:
(3.36)
(3.37)

am=&ai

(3.38)

where:

2,

C,I

Zc,i = 0.2905 - 0.085 mi

. = z.- RTcj
$1

and

(3.39)

Pc,i
Zc,m= 0.290 5 - 0.085 wm

(3.40)
(3.41)

3.2.4 Conclusion Concerning the Corresponding States Principle


The calculation of thermodynamic properties very often depends on the corresponding
states principle either explicitly, as when the Lee and Kesler method is applied, or implicitly when we use the equation of state for which parameters are correlated as a function of
reduced temperature and the acentric factor, for example.This law is applied first and foremost to pure substances but, in fact, it has been extended to mixtures. Utilized within reasonable limits, it is of inestimable value. However, it is advisable to be aware of extension
misuse. It may not be applied to polar compounds. Its extension to compounds of high
molecular weight is subject to caution and it is appropriate in this case to analyze the
molecular interactions in terms of interactions between the constituent groups of the molecule. Of course, in this case we must also take into account the size and shape of the molecule. We must also remember that we do not always make use of the necessary parameters (critical properties) either because they have not been measured, or because they are
located in a temperature zone where the molecule dissociates. It is therefore necessary to
predict them, for example, using the structure correlation properties. We shall provide
some examples. Such predictions do not generally improve the precision of the corresponding states principle. We must, however, whenever possible, validate its application using
available experimental values.

84

3.3

3. Predicting ThermodynamicProperties of Pure Substances

STRUCTURE PROPERTY CORRELATIONS

The observed change of the value of a property of components in a homologous series


(n paraffins,aromatic alkyls,primary alcohols, etc.) with the length of the chain is an experimental fact that naturally tends to break down the value of this property in contributions
that are attributed to the constituent groups of the molecule. Having established this relationship,we are now in a position to predict the values of the given property for new molecular structures made from the same groups. It is a forceful approach. Indeed, the number of
groups that we find in the area of organic chemistry is incomparably more restricted that
the number of distinct, chemical entities. Depending on the property in question, it is used
with more or less success.We shall offer some examples that relate to quantities already
encountered, but we shall later apply this method for calculating the deviation from ideality (Chapter 7) or for predicting thermochemical properties (Chapter 13).
It is necessary to first specify certain easily understood limits. In general, these methods
are poorly applied to the first members of homologous series.They are often considered a
special group. This restriction is not serious as long as it involves only properties of pure
substances and not mixtures. Indeed, the properties of these compounds are generally well
known. We may not, however, use them to develop laws that are particular to the homoiogous series to which they belong as individual members.
Furthermore, insofar as it is possible, we must include the proximity effect. The
constituent groups of a molecule are not isolated. A definition that is too narrow and
would neglect the influence exerted by proximity would cause gross errors. We know, for
example, that the hydroxyl groups of amdiols are not independent, and characteristic of
the alcohols, they are separated by numerous CH, groups. A fortiori, it is impossible to
separate a polysubstituted carbon in autonomous groups. Like Benson et al. [1969], therefore we must always define a group by the nature of the central atom and of those atoms
immediately surrounding it. In this way, for example, the methylene group of a linear
paraffin C-(C), (H), will be differentiated from an alkylbenzene [C-(Car)(C)(H)J,
where C, stands for the aromatic carbon. The disadvantage of such a description is obviously a considerable increase in the number of groups, and therefore the length of the
experimental database used to evaluate the contributions of each group. In addition, for
polyfunctional molecules, it turns out to be insufficient.The interaction between two functions may exert itself even if these two functions are separated by a methylene grouping.
We must also discriminate between the cidtrans isomers, and the ortho, meta, and para
configurations of the aromatic nucleus, etc.
Finally, the general characteristics of the molecular structure (volume, shape, cyclic
structures) are only taken into account indirectly. This way, extrapolation to long chain
hydrocarbons appears often imprecise, sometimes nonsensical, when confronted with
experimental data.
In addition, we must distinguish group contributions when they are applied to the ideal
gas properties or to real compounds. If molecular interactions do not exist, we may more
reasonably attribute to each group a contribution of its own. The entirety of these contributions will form a characteristicvector for the property in question. On the other hand, if
we must include the cohesion forces, we shall substitute group interactions for molecular
interactions and end up with a characteristicmatrix.

85

3. Predicting Thermodynamic Properties of Pure Substances

Such will be the case when we introduce the UNIFAC (see Chapter 7) method of calculating deviations from ideality in the liquid phase, for example. It should undoubtedly be
the case for correlations concerning critical points, the heat of vaporization,etc.
These questions were recently the subject of a synthetic investigation [Gani and
Constantinou, 19951 and new approaches that depend on a more precise investigation of
the geometry of the molecule were proposed [Constantinou and Gani, 19941.

3.3.1 Properties of the Ideal Gas


Thermodynamic statistics allow us to build the ideal gas partition function from the energies of translation, or rotation of all or part of the molecule, and the longitudinal and transversal vibrations of the interatomic bonds. The required parameters for applying the
obtained expressions are molecular mass, the moments of inertia of the molecule and of
the constituent groups, and the Raman or infrared absorption frequencies.Those relating
to molecular geometry (bond angles and lengths) are characteristic of atomic groupings.
As for the absorption frequencies, we know that, as a first approximation,they are related
to the nature of the bond in question and little sensitive to the environment.It is therefore
natural that the thermodynamic properties of the ideal gas, molar heat capacity, and the
entropy reported at absolute zero, may be broken down into group contributions.
For molecules or groups, the possible symmetry of certain structures may render several
distinct configurations indistinguishable.For example, such is the case for the three positions of the methyl group due to a 120rotation around the C-C bond, or even the twelve
positions of the methane molecule that are also obtained by a 120rotation around one of
the four C-H bonds. It should clearly be taken into account in the description of a property such as entropy that is related to the degree of order of a system. This symmetry is
expressed by the degree of external symmetry and the degree of global symmetry of
the molecule. The degree of external symmetry Q is equal to the number of indistinguishable positions that the molecule may assume by simple, fixed rotation on itself. Its evaluation is facilitated by the use of molecular models.Table 3.8 provides the values for some of
their structures.
Table 3.8
Amount of external symmetry for some compounds

Compound
Butene 1
Ethanol
Tertiary butanol
1,3 Butadiene
Propane
Orthoxylene
Metaxylene
Isobutane
Ammonia

I = I1
1
1
1
2
2
2
2
3
3

Compound
Methyl chloride
Trichloroacetonitrile
Paraxylene
Ethylene
Anthracene
1,3,5 Trimethylbenzene
Benzene
Neopentane
Carbon tetrachloride

3
3
4
4
4

6
12
12
12

86

3. Predicting Thermodynamic Properties of Pure Substances

Global symmetry,s, essentially differs from it by rotation of the methyl groups.We state:
(3.42)

s = o3n

where n is equal to the number of methyl groups in the molecule. Certain correlations
incorporate the symmetry of these methyl groups into their contribution, and specifically
their external symmetry only.
Finally, we are led to introduce certain corrective terms related to cidtrans isomerism,
or to the position (ortho, meta, para) of the substitutions on the aromatic nucleus.
We have:

N~A

C;

C ~

so=

2 NiAs;+x
i

N~ACG

(3.43)

NiAsP-Rlns

(3.44)

using A and A to designate the contributions and the corrective terms that relate to heat
capacity and entropy respectively. Ni is the number of groups of type i, and Nj the number
of corrections of type j .
Table 3.9 provides an example of such a correlation, limited to hydrocarbon structures
[Benson et al., 19691.They are also valid for groups containing double bonds, triple bonds,
for halogen and sulfur derivatives,the alcohols, ethers, ketones, amines, and the nitriles.

Table 3.9
Group contributions in the calculation of molar
heat capacity at constant pressure and of molar entropy.
Standard state:ideal gas, P = 1 bar, temperature 298.15 K [Benson et al., 19691
Aso
(ca1.K-)

6.19
5.50
4.54
4.37
3.24
2.67
6.19
5.84
4.88
1.12

30.41
9.42
-12.07
-35.10
11.53
-7.69
30.41
9.3
-12.2
-1.61

We may also cite the method proposed by Rihani and Doraiswamy [1965].The method
gives the group contribution in the form of a polynomial function of the temperature
(cg = a + bT + cT2 + dT3).Table 3.10 shows some examples.The group definitions are less
detailed than Bensons work, but because of this fact, the application of the method is simpler and the risk of error less.

a7

3. Predicting Thermodynamic Properties of Pure Substances

Table 3.10
Group contributions for the calculation of molar heat capacity at constant pressure.
Standard state: ideal gas, P" = 1 bar [Rihani and Doraiswamy, 19651. Units: calf mol-I
Aa

Ab-10'

0.608 7
0.394 5
-3.523 2
-5.830 7
-1.457 2
-1.388 3
0.121 9

2.143 3
2.136 3
3.415 8
4.454 1
1.914 7
1.515 9
1.217 0

~ ~ ~ 1 0 4 Ad.10'
-0.085 2
-0.119 7
-0.281 6
-0.420 8
-0.123 3
-0.106 9
-0.085 5

0.113 5
0.2596
0.8015
1.2630
0.298 5
0.265 9
0.212 2

(1) Aromatic carbon bound to an aromatic carbon and an aliphatic carbon (toluene).

(2) Aromatic carbon bound to two aromatic carbons (naphtalene).

EXAMPLE 3.3

Application of the Benson, and Rihani and


Dorsaiswamy methods to ortho ethyl toluene
We shall calculate the molar heat capacity and entropy of ortho ethyl toluene at
the standard state at 298.15 K. The values from the literature are, respectively,
37.94 cal .mol-l . K-l and 95.42 cal * mol-' * K-l.
First we shall use the Benson method.
We specify:
0 four aromatic carbons bound to two other aromatic carbons and a hydrogen:

4 * [Car-(Car)z(H)l;
0

two aromatic carbons bound to two aromatic carbons and a paraffin carbon:

2 * [Car-(Car>,(C>l;
0

a paraffin carbon bound to an aromatic carbon and 3 hydrogens:


1 * [C-(car)(W31;

a paraffin carbon bound to an aromatic carbon, a paraffin carbon, and two hydrogens:
1* [C- (Car)(C)(H)J;
a paraffin carbon bound to another paraffin carbon and three hydrogens:

1 * [C-(C)(HM;
In addition, there are two methyl groups which we will take into account in calculating the entropy term (symmetry term).
Finally, we will introduce the ortho interaction correction.
Therefore, we arrive at the following calculation (Table 3.11):

88

3. Predicting Thermodynamic Properties of Pure Substances

Table 3.11

Example of the applicationof the Benson method to the calculation of


heat capacity and entropy for orthoethyltoluene

11.53
-7.69
30.41
6.19

30.41
- 4.4

1.12

-1.61

The calculation of heat capacity and entropy shall be performed using Equations 3.43
and 3.44. From the preceding table, we obtain the following values:
cp" = 37.64 cal * mol-I * K-'
and so = 94.85 cal . mol-l . K-l
We again perform the same calculation using the Rihani and Doraiswamy method.
Table 3.12 below summarizes the calculations.
Table 3.12

Group contributionsin the calculation of


molar heat capacity at constant pressure and molar entropy.
Standard state: ideal gas, P" = 1bar [Rihani and Doraiswamy, 19651
Groups
CarH
Car (1)
CH3
CHZ

o-ethyltoluene

b . 102

4 0 4

-1.4572
-1.3883
0.6087
0.3945
-6.9935

1.914 7
1.515 9
2.143 3
2.136 3
17.113 5

-0.1233
-0.106 9
-0.085 2
-0.1197
-0.997 1

0.265 9
0.259 6
2.212 4

These results lead to 298.15 K, at cp" = 35.74 cal . mol-I . K-I.

3.3.2 Critical Coordinates


Knowing the critical properties is an absolute necessity each time we wish to apply one of
the correlations of corresponding states. This means that we are forced to fill in the gaps of
the literature by proposing predictive methods that, for the most part, themselves rely on
group contributions. Simmrock et al. [1986] lists fifty-six methods for the calculation of
critical temperature, fifty-five for critical pressure, and fifty-four for critical volume, and
determines the precision for most of them. The choice is considerably more difficult in
that, depending on the family (hydrocarbons, ketones, alcohols, etc.), one or the other of
these methods seems preferable.

89

3. Predicting Thermodynamic Properties of Pure Substances

We shall limit ourselves to the description of the method proposed by Lydersen [1955],
which, taking into account the precision shown by Joback [1984],is considered one of the
best. It may be applied to hydrocarbons, halogenated derivatives, alcohols, ketones, aldehydes, acids, esters, amines, nitriles, thiols, and mercaptans.
We use the following equations:

3 = 0.584 + 0.965
TC

NiATi-

(3.45)

z
+z

Pc (bar) = 0.113 + 0.0032n -

Ni APi

(3.46)

V c (cm3/mol)= 17.5

Ni Avi

(3.47)

where AT, AP,and Av stand for the increments attributed to the constituent groups for the
calculation of temperature, pressure, and critical volume, Tb the boiling point at atmospheric
pressure, and Ni the number of groups of type i in the molecule.Table 3.13 shows the increment values for the main hydrocarbon gr0ups.A complete listing is given by Reid et al. [1987].
Table 3.W
Contributions of hydrocarbon groups to critical properties [Lydersen,19551

AP
(bar)

AV
(cm3-mol-' )

0.0141
0.0189
0.0164
0.0067
0.0113
0.0129
0.0117

-0.001 2
0
0.0020
0.0043
-0.002 8
4.000 6
0.001 1

65
56
41
27
56
46
38

0.010 0
0.0122
0.004 2
0.0082
0.0143

0.0025
O.OO0 4
0.006 1
0.001 1
O.OO0 8

48
38
27
41
32

Groups
Aliphatic Groups

-CH3
-CHz>CH>C<
=CHz
=CH=C<

Cyclic Groups
-CHz>CH>C<
=CH-

=C<

We note that for the calculation of critical temperature, the normal boiling point must
be known. This is a disadvantage since this property is sensitive to structure (configuration
of the hydrocarbon skeleton, among other things) and is difficult to predict. The Joback
method contains group contributions for this property, but the results are of very variable
precision [Reid et al., 19871.
The method proposed by Ambrose [1978,1979]is of comparable precision, but is more
complicated to apply. It is explained in detail in the work by Reid et al. [1987] and, like
Lydersen,covers a large number of functions.
When using the group contribution correlations, we must be wary of the misuse of
extrapolations. The calculation of critical properties is an example. Tsonopoulos [1987]

90

3. Predicting Thermodynamic Properties of Pure Substances

demonstrated that the blind application to a paraffin containing 100 carbon atoms resulted
in a negative reduced temperature using the Lydersen method, and that the two methods
by Ambrose and Lydersen deviate from each other starting from q,,hydrocarbons.

3.3.3 Calculation of Molar Volume in the liquid Phase


We could cite many other examples of calculating the thermodynamic properties of pure
substances by group contribution.The breaking down into groups proposed by Benson has
been applied to the calculation of properties in the liquid phase by Domalski and Hearing
[1988,1990].
An especially simple method was proposed by Elbro et al. [1991] to obtain the molar
volume in the liquid phase with reasonable precision.

3.4

EXAMPLES OF THE RELATIONSHIPS


BETWEEN TH ERMODYNAMlC PROPERTIES

We have seen that applying the corresponding states principle requires that we know the
critical properties.This disadvantage is not small as these data are sometimes unavailable.
The value for the normal boiling temperature, vapor pressure values, and density finally
make up a more readily available database. A simple example will demonstrate that these
values may be combined using the equations that we have given to supply the missing critical data. More systematically,a group of empirical equations without theoretical basis but
with reasonable precision may be used to advantage.

3.4.1 Calculation of Critical Properties


from Vapor Pressure and Density Data
It often happens that we have at our disposal data such as vapor pressure or density, but
not the critical properties. In this case, we may assess these unknown values that, with
the use of the corresponding state correlations or the equations of state, concur with available data [see, for example, Rogalski et al., 1992; Vetere, 1987, 19891. We shall provide a
simple example of such a calculation.

EXAMPLE3.4

Calculation of the critical properties for n -butane


We have the measured vapor pressures and molar densities of n-butane at 0C and
at 100C.The results are listed below.

3. Predicting Thermodynamic Properties of Pure Substances

91

Table 3.14
Molar density and vapor pressure of n-butane

100

1.035
8.010

1.036
15.28

The correlations that seem best adapted to this calculation are the Rackett equation
(Section 3.2.1.2), and the simplified expression for vapor pressure as a function of
reduced temperature and the acentric factor (Eq. 3.16).
We write the Rackett equation:
(3.14)
in the form:

[ (

InvL*u=In-+
R T ~ I + I - -E r ] l n Z R a
PC
where application at two temperatures T, and T2yields:

The Rackett compressibilityfactor will be estimated using the equation:


ZRa= 0.29056 - 0.08775 w

in addition,we have:

v L > u = 1lpLP

1 $7

so that for the acentric factor we obtain the following expression:


0.29056-exp{[(l-+r-(1-$)
211 - 1 In

o=

0.087 75

To use the vapor pressure data, we use the equation:


PU
7
log,, - =-- (1 + 0) - -1
e
3
and so, at temperatures Tl and T2:

(:

P1"
7
log,, - = - - (1 + o) - - P;
3

:( ):

or:

(3.16)

92

3. Predicting ThermodynamicProperties of Pure Substances

In this way we have two independent expressions for the acentric factor as a function
of critical temperature. By making a series of assumptions about this temperature we
shall determine the value appropriate to each expression,as shown in Table 3.15 below.
Table 3.15
Calculation of the acentric factor as a function of assumptions
made about the critical temperature of n-butane
T
(K)

(Rackett)

400
410
420
424
425
426
430

-0.910 13
-0.40071
0.02181
0.17221
0.208 36
0.243 97
0.38120

0.276 36
0.24523
0.215 58
0.220 13
0.201 28
0.19846

We thus find T, = 424.8K7a value in line with the experimental value (425.18K). From
this value, we have the values for critical pressure, 37.34 bar (exp: 37.37 bar), and the
acentric factor: 0.202 (exp: 0.199).
We must note that we are dealing with two sets of different data: vapor pressure and
density. Vapor pressure data, although more plentiful, would have certainly determined the curve for vapor pressure, but not the final point of this curve, which is the
critical point.
The models used must be understood to be valid within the range under consideration. If the vapor pressure data had been at low temperature (below the normal boiling temperature), Equation 3.16 could have produced gross errors.
We may also use an equation of state that is able to simultaneously show the vapor
pressures and the volumetric properties in the liquid phase, and whose parameters
are a function of the critical properties and the acentric factor, such as the one proposed by Schmidt and Wenzel [1981]. The Lee and Kesler model discussed above
lends itself especially well to such an evaluation.However,we must point out that the
compoundsfor which the critical properties are unavailable are generally located outside of the normal range of application of equations of state and the corresponding
states principle.As they are in fact considered intermediary parameters, critical properties that are calculated in this fashion must be located within a temperature zone in
which actual calculations will not be carried out.

3.4.2 Calculation of the Heat of Vaporization: Watson Equation


It is desirable to also provide an example of the Watson equation [1931,1943].It allows for
the calculation of the heat of vaporization Ahu at any temperature provided that at any
particular temperature To,the value of this property is known.

93

3. Predicting Thermodynamic Properties of Pure Substances

It is written:

1- T,

Ah"

0.38

(3.48)

iq=(=)

The exponent 0.38 is absolutely empirical. With this method, the known value for heat
of vaporization serves to improve the prediction founded on the reduced temperature. It
was modified by Thek and Stiel[1966,1967].

3.4.3 Empirical Equations Developed from


the Normal Boiling Point and Density
A certain number of empirical equations were proposed for the calculation of thermodynamic properties starting with the normal boiling point and the density. Their accuracy is
mixed, their basis totally empirical, and their range of application limited to the hydrocarbons, but they have merit in that they require only data that are generally available. We
will mention those developed by Riazi and Daubert [1980,1987].

X=aTrpn

Example:

(3.49)

where Tb is the normal boiling temperature expressed in Kelvin, p is "specific gravity" at


60"F, meaning the density at 288.7 K relative to water at the same temperature, X the property in question, such as the molecular mass, critical properties, critical volume, molar volume in the liquid phase at 20C at atmospheric pressure wL,heat of vaporization Ah", and
the heat capacity of the ideal gas at 0,600, and 1200F(255,590, and 920 K). Of course, the
numerical values of the exponents m and n depend on the property in question. They are
provided in Table 3.16 along with the average deviations noted by the authors. This method
has been recently expanded to compounds of high molecular weight where the boiling
temperature at atmospheric pressure is unknown. After a comparative study of several
calculation methods for the critical properties and the acentric factor, Voulgaris et al.
[1991] recommended the Riazi and Daubert method.
Table 3.16
Parameters of the Riazi and Daubert correlation applied to the calculation of molar mass,
critical points, density, heat of vaporization, and heat capacity
Units

Property

M
=c

pc
vc
VL

P
Ah0
CP (0 OF)
cp (600 O F )

c p (1 200 O F )

g.mo1-'
K
bar
cm3 .mol-'
cm3 mol-'
g/cm3
J .mol-'
J .mol-' K-'
J .mol-' K-'
J .mol-' K-'
+

1.6607
19.0623
5.53031 lo7
1.7842
2.6594 loT4
0.983719
38.4502
8.1344
8.577 10-5
1.421

2.1962
0.58848
-2.3125
2.3829
2.1262
0.002016
1.1347
2.672 4
2.421 9
2.385 3

-1.0164
0.359 6
2.3201
-1.683
-1 368 8
1.0055
0.0214
-2.363
-1.943 6
-1.932

Deviation
(Yo)

2.6
1.3
3.1
2.8
2.8
2.8
1.4
3.3
3.5
4.2

94

tl

3. Predicting Thermodynamic Properties of Pure Substances

EXAMPLE 3.5

Riazi and Daubert method


We shall apply the Riazi and Daubert method to isobutane (Tb = 261.4 K, spgr
= 0.5653), benzene (Tb = 353.2 K, spgr = 0.8845), and methanol (Tb = 337.9 K, spgr
= 0.7978).
Table 3.17 below shows the results of the calculation. We note that while the results
are reasonably satisfactory for the hydrocarbons, the correlation may not be applied
to methanol, an autoassociated polar compound.
Table 3.17
Example of the application of the Riazi and Daubert correlation

Property

M
TC

pc
VC

vin

P
Aha
c, (0 O F )
C, (600 O F )
C, (1200 O F )

Isobutane

Benzene

Methanol

Ealculated Zxperimental Calculated Experimental Zalculated Zxperimental


60.4
410.8
37.84
268.3
106.5
0.560
21 OOO
90.3
186
249

58.1
408.1
36.4
259
104.1
0.558
21 400
84.5
167.8
219.8

14.2
576
53.3
258.7
87.5
0.880
29 850
70
161
215

78.1
562.2
48.98
258.9
89
0.879
30750
68.42
157.6
204.2

74.2
562
49
259
96.6
0.793
28 330
79.38
177.1
236.3

32
512.6
80.96
117.8
40.46
0.792
38290
41.59
66.42
85.94

This correlation is also applicable to petroleum cuts. In this case, the pseudocritical
points are obtained. As an example, we may consult the work of Yu and Eser [1995]
that compares several prediction methods for the critical points of jet fuel.

CONCLUSION
The physicochemical properties may in principle be evaluated using the molecular structure. It is a developing approach, but it still may not be substituted for more empirical
methods. These methods, correlations of corresponding states, and group contributions,
also implicitly depend on the concept of molecular interactions. They were developed
thanks to the acquisition of experimental data, which they combine in a more concise
form. This experimental basis is their strength, but it imposes limits.To apply them too far
beyond the range for which they were defined may lead to gross errors.

3. Predicting ThermodynamicProperties of Pure Substances

95

REFERENCES
Ambrose D (1978,1980) Correlation and prediction of vapour liquid critical properties. I. Critical
temperatures of organic compounds. NPL Rep. Chem., 92, National Physical Laboratory,
Teddington,Great Britain.
Ambrose D (1979) Correlation and prediction of vapour liquid critical properties. 11. Critical pressures and volumes of organic compounds. NPL Rep. Chem., 98, National Physical Laboratory,
Teddington,Great Britain.
Benson SW, Cruickshank FR, Golden DM, Haugen GR, ONeal HE, Rodgers AS, Shaw R, Walsh R
(1969) Additivity rules for the estimation of thermochemical properties. Chemical Reviews, 69,
279-324.
Constantinou L, Gani R (1994) AIChE J.,40,1697-1710.
Domalski ES, Hearing ED (1988) Estimation of the thermodynamic properties of hydrocarbons at
298.15 K. J. Phys. Chem. Re$ Data, 17,1637-1678.
Domalski ES, Hearing ED (1990) Heat capacities and entropies of organic compounds in the
condensed phase. Volume 11. J. Phys. Chem. Re$ Data, 19,881-1048.
Elbro HS, Fredenslund Aa, 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. 30,2516-2582.
Emschwiller G (1959) Chimie Physique. Presses Universitaires de France, Paris.
Gani R, Constantinou L (1995) Molecular structure based estimation of properties for process
design. Seventh Conference on fluid properties and chemical process design, Snowmass, Colorado,
June 1995.
Gubbins KE (1989) The future of thermodynamics. Chem. Eng. Prog., 85 (2), 38-49.
Halm RL, Stiel LI (1970) Saturated liquid and vapor densities for polar fluids. AZChE J. 16,3.
Halm et al. (1985) AZChE J., 31,1632.
Hougen OA, Watson KM, Ragatz RA (1959) Chemical Process Principles. Wiley & sons.
Joback KG (1984) M.S. thesis in chemical engineering, Massachussets Institute of Technology,
Cambridge, Mass.
Kay WB (1936) Density of hydrocarbon gases and vapors at high temperature and pressures. Ind.
Eng. Chem.,28,1014-1019.
Keenan JH, Keyes FG, Hill PG, Moore JG (1969) Steam Tables. Thermodynamic Properties of Water
Including Vapor, Liquid and Solid Phases. Wiley, 1969.
Lee BI, Kesler MG (1975) A generalized thermodynamic correlation based on three-parameter corresponding state. AIChE J.,21,510-527.
Lydersen AL (1955) Estimation of critical properties of organic compounds. Univ. Wisconsin Coll.
Eng., Eng. Stn., Rep., 3.
Meyer EF, Renner TA, Stec KS (1971) Cohesive energy in polar organic liquids. 11.The alkyl nitriles
and the 1-chloroalkanes.J. Phys. Chem. 75,642-648.
Pitzer KS, Curl Jr RF (1957) Empirical equation for the second virial coefficient. J. Am. Chem. SOC.,
79,2369.
Pitzer KS, Lipmann DZ, Curl RF,Jr., Huggins CM, Petersen DE (1955) Compressibility factor, vapor
pressure and entropy of vaporization.J. Am. Chem. SOC.,77,3433.
Rackett HG (1970) Equation of state for saturated 1iquids.J. Chem. Eng. Data, 15,514-517.

96

3. Predicting ThermodynamicProperties of Pure Substances

Rackett HG (1971) Calculation of bubble point volumes of hydrocarbon mixtures. J. Chem. Eng.
Data, 16,308-310.
Reid RC, Prausnitz JM, Poling BE (1987) The Properties of Gases and Liquids, fourth edition.
McGraw-Hill Book Co.
Riazi MR, Daubert TE (1980) Simplify property predictions. Hydrocarbon Processing, March, 115116.
Riazi MR, Daubert TE (1987) Characterization parameters for petroleum fractions. Ind. Eng. Chem.
Res., 26,755-759.
Riazi MR Daubert TE (1987) Improved characterization of wide boiling range undefined petroleum
fractions. Ind. Eng. Chem. Res., 26,629-632.
Riazi MR, Daubert TE (1980) Simplify property prediction. Hydrocarbon Processing, March, 115117.
Riedel L (1954) Eine neue universelle Dampfdruckformel. Chem. Eng. Tech.,26,83.
Rihani DN, Doraiswamy LK (1965) Estimation of heat capacity of organic compounds from group
contributions. I&EC Fundamentals,4,17-21.
Rogalski M, Mato FA, Neau E (1992) Estimation of hydrocarbon critical properties from vapour
pressures and liquid densities. Chem. Eng. Sci., 47,1925-1932.
Schmidt G, Wenzel H (1981) Estimation of critical data by equation of state. Canad. J. of Chem. Eng.,
59,527.
Silverberg PM, Wenzel A (1965) The variation of latent heat with temperature. J. Chem. Eng. Data,
10,363-366.
Simmrock KH, Janowsky R, Ohnsorge A (1986) Critical data of pure substances. Chemistry Data
series, Vol 11,Dechema, Frankfurt.
Spencer CF,Danner RP (1972) Improved equation for prediction of saturated liquid density.J. Chem.
Eng. Data, 17,236-241.
Spencer CF, Adler SB (1978) A critical review of equations for predicting saturated liquid density.
J. Chem. Eng. Data, 23,82-89.
Stipp GK, Bai SD, Stiel LI (1973) Compressibility factors of polar fluids in the gaseous and liquid
regions.AIChE J., 19,227.
Stipp GK, Bai SD, Stiel LI (1973) Compressibility factor of polar fluid in the gaseous and liquid
region.AIChE J., 19,1227-1233.
Thek RE, Stiel LI (1966) AIChE J., 12,599.
Thek RE, Stiel LI (1967) AZChE J.,13,626.
Thinh Tp, Duran JL, Ramalho RS (1971) Equations improve Cp predictions. Hydrocarbon
Processing, January, 98-104.
Tsonopoulos C (1974) An empirical correlation of second virial coefficients.AIChE J., 20,263-272.
Tsonopoulos C (1987) Critical constants of normal alkanes from methane to polyethylene. AIChE J.,
33,2080-2083.
Vetere A (1987) Methods for estimation of critical volumes. The Chemical EngineeringJ., 34,151-153.
Vetere A (1989) Estimation of critical pressures by the Rackett equation. Chem. Eng. Sci., 44,791795.
Voulgaris M, Stamatakis S, Magoulas K, Tassios D (1991) Prediction of physical properties for nonpolar compounds, petroleum and coal liquid fractions. Fluid Phase Equilibria, 64,73-106.
Watson KM (1931) Ind. Eng. Chem.,23,360.

3. Predicting Thermodynamic Properties of Pure Substances

97

Watson KM (1943) Ind. Eng. Chem.,35,398.


Wu GZA, Stiel LI (1985) A generalized equation of state for the thermodynamic properties of polar
fluids.AIChEJ.,31,1632-1644.
Yaws CL, Yang HC, Hopper JR, Cawley WA (1991) Equation for liquid density. Hydrocarbon
Processing, January, 103-106.

Yu J, Eser S (1995) Determination of critical properties ( T c , Pc) of some jet fuels. Ind. Eng. Chern.
Res., 34,404-409.

Equations of State

The equations of state express by means of a mathematical expression:


E( T,P,V,N) = 0
(4.1)
the relationship between the temperature and pressure, volume occupied, and the amount
of matter for a pure substance or mixture. This equation is more often reduced to one
mole:
e (T,P,v) = 0
(44
The most widely known example is the van der Waals equation [1873]:
a
p = -RT
-(4.3)
V - b V

This work uncovered a field that has been the subject of numerous studies that are still
ongoing, and into which we can provide but a glimpse. In 1983,R.C. Reid presented a statistic on the literature devoted to the topic and especially on the Redlich-Kwong equation
of state (Table 4.1), and concluded: Zt is a full-time job just to maintain a familiarity with
the new developments in this field!
Table 4.1
References from ChemicalAbstracts on equations of state [Reid, 19831

Period

Equations of State

Redlich-Kwong Equation

1982
1980-1981
1977-1979
1972-1976
1967-1971

450
885
1111
1406
927

39
30

There is no doubt whatsoever that this interest is rooted in the ability (at least in principle) of the equations of state to calculate all thermodynamic properties, in particular the
phase equilibria.
Using the equation of state, it is indeed possible, by application of Equations 1.44,1.45,
and 1.50 through 1.53, to calculate the effect of volume or pressure variation on the thermodynamic properties. Through integration (Eqs. 2.27 to 2.34), we can then calculate the

100

4. Equations of State

deviations from the ideal gas laws or residual quantities. For example, for enthalpy we
have:
h,,,( T,P) = h( T,P) - h#(T,P) =

or:

hres(T,P) = h( T,P) - h#(T,P) =

:\

lo'[

(T($

(al

v-T -

)"

dP

- P) dv + Pv - RT

(2.27)

(2.29)

If we also use the heat capacity in the standard state, the value of the thermodynamic
property may be calculated using:
T

h(T,P)=ho(To)+I cp"dT+(h-h#)
To

(4.4)

Conversely, an equation of state may be developed from an expression relating one of


the residual values to the conditions of temperature and volume (or pressure). The following expression for residual Helmholtz energy:
a-a'=-RTln(

--)-

V-b

a
V

(4.5)

yields the van der Waals equation by derivation with volume at constant temperature and
applying Equation 1.44:
dA=-SdT-PdV

or

da=-sdT-Pdv

(1.44)

This relationship between equations of state and expressions of residual properties


must be stressed. Indeed, statistical thermodynamic observations generally yield the value
of the residual quantity usually related to Helmholtz energy through the partition function, from which the equation of state results.
Furthermore, this relationship is the key to phase equilibria calculations using equations of state. At this point, it is appropriate to explore one of the objections that may be
raised when applying equations of state to the calculation of thermodynamic functions in
the liquid phase. If we state that there is an equation of state capable of simultaneously
describing both the liquid and vapor states for a pure substance,we see that in the pressure
volume diagram, a continuous isotherm trace like the one in Figure 4.1 can be drawn. On
this isotherm we can identify:
0 Curve A B related to the vapor phase
0 Point B representing this vapor at its dew point
0 Point C representing the liquid at its bubble point, and
0 Curve CD related to the liquid phase.
If the equation of state is well chosen, this line will fit the experimental data. However,
on the continuous line corresponding to the equation of state, we can also identify a curve
BM that we may relate to a metastable vapor state that occurs late compared to the
condensation phenomenon, and that is to a certain extent experimentallyreproducible.
We also see a curve N C related to a liquid that is also metastable. Curve MN on the
other hand does not correspond to any physical reality because along this curve the
mechanic stability condition dPldv < 0 is not met.

101

4. Equations of State

100

200

300

400

500

Molar volume (cm3. rno1-l)

figure 4.1 Calculation of a subcritical isotherm using an equation of state.

We can now have doubts concerning the validity of a residual function calculation in the
liquid phase that uses an equation of state since it will include an integration step over an
unstable zone for all thermodynamic values. For example, we have:

by applying Equation (2.30).


The vapor liquid equilibrium condition in particular is obtained by equation:

which can be read in Figure 4.1 using the equal areas contained between isotherm P(v)
and the straight ordinate Po.This condition is very important. We call it Maxwell's condition. It is a concrete expression of the relationship that exists between the equation of state
and the vapor liquid equilibrium.
In fact, the doubt surrounding the validity of an integration within an unstable range is
removed if we consider that the thermodynamic model has its origin in a Gibbs energy
model that we assume is valid for the two phases. Furthermore, in the absence of any other
hypothesis,the equation of state that results is obtained through a derivation that does not
include the unstable region:

( T)

v - RT
- = J(g-g#)

(4.7)
T

102

4. Equations of State

We can see that an equation of state may very well represent the volumetric behavior
within stable zones, but yield poor results for the vapor pressure if the representation of
the unstable zone is insufficient. On the other hand, it may allow for good calculations of
vapor pressure but yield poor values for volume (such is the case for most two parameter
cubic equations that we shall encounter later on).

Classificafionof Eguafionsof Sfafe


It would be good to be able to formulate a simple classification for the equations of state.
It is even necessary, in that an account should be methodical. For this reason, we shall discuss in sequence the equations of state derived from the virial development,those derived
from the van der Waals theory, etc., selecting the structure of the equation as our classification criterion. The qualities that characterize this or that equation are of a variable
nature, and it would be necessary to have each one of them correspond to a particular classification.
Because the expressions that are commonly used are imperfect and do not meet all
needs, it seems to us that the most important criterion is the range of application of the
equation in question:
0 What compounds, or what families
What properties
0 What range of pressure and temperature.
Hence, we shall, for example, attribute particular importance to the equations of state
devised to represent the liquid-vapor equilibria of non-polar compounds.
We shall also pay attention to the predictive character of the equation of state, most
often obtained by application of the corresponding states principle.
One criterion that may appear to be related to this predictive character is the number of
parameters and, of course, their physical significance.This criterion assumes a particular
importance when applied to mixtures because most often we must define a law of composition for each of these parameters.
The mathematical expression must be taken into account. We shall draw a distinction
between equations explicit for pressure or volume. We will consider its complexity, which,
upon resolution of the equation of state, may entail a lengthening of the computing time
that may be prohibitive. Above all, this complexity risks introducing serious incongruities
at the limits of the application range of the equation of state, or in extrapolation.

4.1

EQUATIONS OF STATE DERIVED FROM


THE VlRlAL DEVELOPMENT

The value of the compressibilityfactor for a fluid may be developed as a series around the
point where real fluid and ideal gas merge, which corresponds to a density of zero.

4. Equations of State

Thus, we have:

103

B C
z = 1 +- + 7+...
v

The coefficients B , C , etc. for pure substances are only a function of temperature and
are called the second, third, etc. virial coefficients. We have already encountered the second virial coefficient in our study of the properties of real fluids.
If the function Z( T,v) developed in this way is the unknown function exactly corresponding to the experiment, the virial coefficients will be regarded as defined physical
properties. They will be subject to experimental determinations,and compilations of them
exist. Example 4.1 gives an example of the treatment of such experimentalresults. We have
also provided a correlation of second virial coefficients within the scope of the corresponding states principle (Chapter 3, Section 3.2.2.1).
On the other hand, if the function Z(T,v) is an equation of state, the corresponding
development also allows us to define the expressions for the virial coefficients associated
with this equation. For example:

Of course, applying these equations allows us to find the experimental values of the virial coefficientsonly if the equation of state correctly represents the fluid. Examples 4.1 and
4.4 that follow illustrate these distinctions.
We can also define a development for pressure:
Z = 1 + B P + C P 2 + ...

(4.10)

whose parameters are related to the virial coefficients by equations,the first two of which
we shall consider:
B
B= (4.11)
RT
C - B2
C= R ~ T ~

(4.12)

These sequential developments are not looked upon as an equation of state since they
have an infinite number of terms whose values are unknown and whose calculation is
impossible. Yet their truncation has given rise to the virial equations of state that we
characterize according to the degree of truncation and by the nature of the development
from which they are derived,whether volume or pressure. Certain equations of state result
from a combination of truncation and empirical terms that assures them extensive validity.
The best known is the Benedict,Webb, and Rubin equation of state.

104

4. Equations of State

D EXAMPLE 4.1
Determination of the second and third virial coefficients
for ethane from experimental data
We shall consider a set of pressure,volume, temperature data for ethane, and using
these data, we shall determine the second and third virial coefficients at several temperatures.
To accomplish this, the development will be used as follows:

C
(Z-l)w=B+-+
2,

D
-+...
02

and for a given temperature, we will use the values of the product of (Z - l)w as a
function of the relation 1lw.The limit of this product for llw + 0 is equal to the second
virial coefficient B , and the slope at the origin of the curve is equal to the third virial
coefficient C. Table 4.2 shows the data for pressure and molar volume at 25C,and the
calculation intermediates.
Table 4.2
Determination of the second and third virial coefficients for ethane at 25C

P
(bar)
1
5
10
15
20
25
30
35
40
41.876

(cm3. mol-*)

1OOOIV

( Z - l)w

24 602
4 767
2 282
1449
1028
770
592.2
456.5
337.4
286.85

0.992 5
0.961 4
0.920 5
0.876 7
0.829 1
0.776 6
0.716 7
0.6445
0.5444
0.484 6

0.040 6
0.209 8
0.4382
0.6902
0.973 1
1.2987
1.6887
2.1908
2.9642
3.486 2

-185.7
-183.9
-181.4
-178.7
-175.6
-172.1
-167.8
-162.3
-153.7
-147.8

Using Figure 4.2, we can determine the values for B and C at 0,20, and 25C. The
results are listed in Table 4.3.
Table 4.3
Second and third virial Coefficients for ethane at 0,20, and 25C
(cm3.mol -)
273.15
293.15
298.15

-223
-193
-186

11105
10970

105

4. Equations of State

-140-

-1 60 z

h
7-

ru

-180-

1 ooo/v

Figure 4.2 Determinationof the second and third virial coefficient of


ethane from experimental data.

We note that despite the relatively extended pressure interval, the development of
the truncated virial after the third term is a good representation of the data, with the
exception of the critical isotherm for which we observe appreciable deviations.

4.1.1 Volume Virial Equation of State Truncated after the Second Term
The simplest and most applied of all truncations is written as:
B
Z=l+-

(4.13)

As can be expected in view of its simplicity,its application range is very limited. It takes
into account the vapor phase at low pressure only, and its precision decreases with increasing pressure. In general, we state that it may be used up to half the critical pressure.

This equation of state is explicit for pressure but of the second order for volume, and
only the root corresponding to the vapor phase is to be retained. Note that with an iterative calculation where pressure is unknown, it may happen that it has no real root. On a
diagram Z ( P ) analogous to those shown previously (Fig. 2.3 or 3 . Q it is represented by the
parabolic equation:

106

4. Equations of State

that passes through points P = 0, Z = 1 and P = 0, Z = 0. The former corresponds to the


ideal gas; the latter is extended well beyond the range of application for the equation. This
parabola is tangential to isotherm Z(P) at the P = 0 and Z = 1coordinates.
The values for the second virial coefficient that will be used when applying this equation
of state may come from experimental data specific to the component investigated.However,
most often they are predicted using a correlation such as the one that we introduced
(Chapter 3, Section 3.2.2.1) as a function of reduced coordinates and the acentric factor.
The expressions for residual functions at given temperature and pressure that correspond to this equation of state are simple. They are obtained via the previously described
procedure (Chapter 2, Section 2.4) as the equation is not explicit in volume. We find:
B-T= h - h # = RT
hresidual(T,P)

2,

dB
dT

(4.14)

(4.15)
(4.16)

4.1.2 Volume Virial Equation of State Truncated after the Third Term
The equation:

z = l +B2 -, +vCT

(4.17)

is little used. Indeed, the addition of a third term does not allow for a representation of the
liquid and vapor states with acceptable precision. In particular, we note that the application of the critical constraint (Eq. 2.4) yields a critical compressibility factor of 1/3, clearly
greater than the experimental values. Furthermore, the values of the third virial coefficient
are not well known. Finally, when applying this equation of state to mixtures, a compositional dependence law must be defined for this coefficient. Supposedly representative of
the interactions between three molecules, its expression involves ternary parameters
whose estimation is empirical only.
More advanced developments have, however, been applied to natural gas. They allow
for a very satisfactory predictive density calculation [Jaeschke et al., 1991a,1991bl.

4.1.3

Pressure Virial Equation of State Truncated after the Second Term

The simplest of all these equations is written as:


BP
Z=l+RT

(4.18)

107

4. Equations of State

or:

v=v#+B

(4.19)

It is explicit for both pressure and volume. Note that the use of the virial equation of
state with pressure truncated after the second term substitutes the isotherms of the compressibility factor with the bundle of lines that is tangential to them at points P = 0, and
2 = 1, and whose slope is equal to BIRT. It is therefore easy to qualitatively estimate its
validity range. It is temperature dependent as an examination of the Z(P) diagram shows
(Fig. 2.3 or 3.8). At low reduced temperatures, it is limited by the two-phase envelope, but
at higher reduced temperatures, it extends to very elevated pressures, the isotherms Z(P)
being practically linear. Of course, the equation can be applied to the vapor phase only.
The expression for residual function is especially simple:
(4.20)

(4.21)
Even though this form of the virial equation of state yields results that are slightly less
precise than those of expression (4.13), we prefer it for reasons of simplicity.

EXAMPLE 4.2

Calculation of an isentropic compression


We shall calculate temperature variation as a function of pressure during the compression of n-butane, which is supposed adiabatic and reversible, and therefore isentropic. The pressure virial equation of state truncated after the second term will be
used. The second virial coefficient will be calculated using the simplified expression:
B=a+-

P
T2

Heat capacity will be furnished by the equation:


c,=a+bT
The corresponding numerical data are:

a = 197.1 cm3 . mol-l

p = 8.31

a = 22.97 J . mol-I . K-

b = 24.98

O7 cm3 . mol-I . K2

J mol-I . K-
1

The initial conditions are: T = 278.15 K, P = 1bar, and the final pressure equals 6 bar.
By applying Equation 1.53 that expresses the variation of entropy with temperature
and pressure:
d S = C pd-T- ( g )
d P or d.s=cp--($-)
dT
dP
(1.53)
P
T
P
T

108

4. Equations of State

it is possible to derive the temperature as related to pressure along the isentropic:

(4.22)
The terms on the right hand side may be arrived at by applying the equation of state.
It is recommended to use the heat capacity at standard state cp" (ideal gas) and the
residual term for the calculation of the heat capacity. In any event, we cannot integrate the obtained differential equation.
It seems simpler to express the entropy using:
S(

T,P) = so( To,Po)+ [so(T,P) - so( To,Po)]


+ [s#( T,P) - so( T,P")] + [s( T,P) - s#( To,P)]

(4.23)

In this equation, we recognize successively:


An arbitrary "origin" value chosen for entropy, at standard state (ideal gas, pressure of 1bar, temperature To).
The inclusion of the temperature, thereby obtaining entropy as a function of temperature, always at standard state.
The calculation for the variation of entropy of the ideal gas between standard pressure and pressure P (this calculation would not have occurred with enthalpy).
The residual entropy term.
Applying the virial equation of state for pressure truncated after the second term, this
residual term is calculated as follows:

a(v - v # )
d(s-s#)~=-(

dB

7d)
P=- dT dP

dB
dT P
andhence' s-s#=- -

(4.24)

P dB
In - - - P
Po dT

(4.25)

and is therefore written as:


s( T,P) = so(To,Po)+

T
IT0

- dT - R

If we designate the initial conditions with T, and P,, during isentropic compression
the temperature and pressure are then linked by the expression:

+ITo$
T

so(To,P")

P dB
dT- R In - - - P
Po dT
= so(T0,P") +

I,

cp"

dT - R In

5
- ( $)
Po

p,

(4.26)

T = Ti

We verify that the pressures and temperatures To and Poas well as the entropy under
these conditions so (To,Po)are eliminated from this equation.To perform the calculations, and in particular to obtain the numerical value for the second member, we shall
use the value so (To,Po)= 312.12 J mol-' ' K-' at 298 K provided by the entropy tables.
Taking into account the expressions that were given for the second virial coefficient
and heat capacity at standard state, we obtain:

109

4. Equations of State

T
p
2P
so(TO,Po)+aln- +b(T-To)-Rln - + - P
TO
Pa T3
Tl
TO

p,
Po

2P

= so( To,Po)+ a In - + b ( Tl - To)- R In - + -

T:

Taking into account the numerical data, the initial value for entropy (right-hand side
of this equation) equals 307.08 J * mol-' K-'.
Table 4.4 below shows the change in temperature during compression as a result of
the preceding equation. For each temperature, we have added the value for vapor
pressure. It is appropriate to be sure that there is no condensation during the course
of compression as the preceding logic is specific to a homogeneous vapor phase.

P
(bar)
1.01325
2
3
4

T
(K)
283.15
300.4
310.7
318.1
323.9
328.7

5
6

PO
(bar)
1.49
2.61
3.54
4.35
5.06
5.72

We observe that between 324 and 329 K, the vapor pressure becomes less than the
calculated pressure. There is therefore condensation.
The risk of isentropic condensation may in some cases be estimated by applying a
very simplified logic. If we assume that compression is applied to a saturated vapor at
atmosphericpressure,in order to avoid condensation the isentropic slope must be less
than the slope of the vapor pressure curve P( T):
dPb
(%),ST

When applying the Clapeyron equation, if the vapor phase is comparable to an ideal
gas and the molar volume of the liquid phase is negligible, we have:
dPa
dT

Ahb P

-- - -

RT T

or

dlnPa
dln T

-=-

Aha
RT

(4.27)

Furthermore, for an ideal gas we have:


(1.69)

or in other words:

(;FT),
;
=

(4.28)

110

4. Equations of State

The condition may therefore be expressed as:


c;

Ah0

RT

-<-

The molar heat capacity for hydrocarbons increases steadily with molecular weight
while the adimensional enthalpy of vaporization AhVRT varies little (Trouton law,
Chapter 2, Table 2.7). From this we can conclude that the lighter hydrocarbons
(methane, ethane) do not undergo condensation during isentropic compression
begun at the normal boiling temperature compared to compounds of higher molecular weight. The previous, more precise calculation shows that the limit is found to be
approximately at n-butane.
We can finally compare the results of Table 4.4 with the ones we obtain by comparing
n-butane vapor to an ideal gas. As we have just seen, for an isentropic change we may
write:
R
T =p$!(
(4.29)
Applying this equation at a pressure of 6 bar, and neglecting the variation of heat
capacity with temperature, we find that T = 332 K.

0 EXAMPLE4.3
Calculation of fugacity in the liquid and vapor phase
Considering Equation 4.21:
(4.21)
by applying Equation 1.53 that expresses the variation of entropy with temperature
and allows us to obtain a simple expression for the fugacity of a pure substance in the
saturated vapor phase:
f v = P exp

(g)

and

='f

P'exp

BP'
(=)

(4.30)

We cannot apply the virial equation of state truncated after the second term to the
liquid phase. We shall therefore estimate the Poynting correction (Eq. 2.51):
fl=f"exp[

wL(P- P?
RT

(4.31)

and ultimately:
(4.32)
These calculations will be applied to the calculation of the liquid-vapor equilibrium
coefficients.

111

4. Equations of State

4.1.4 The Benedict, Webb, and Rubin Equation [I 9401


This equation of state has been generally used to allow the interpolation or smoothening
of experimental results that described the volumetric behavior of light hydrocarbons.If we
designate by p the "molar density", which is the opposite of the molar volume, p = l/v, it is
written as:

>) p2 + (bRT

- a)p3 aap6

CP3

+ T2 (1 + lip2) exp (- lip2)

(4.33)

In theory, this equation may be applied to vapor and liquid states, and to the calculation
of residual thermodynamic properties (enthalpy, entropy, fugacity coefficient).
As we see, it can be considered a development of the compressibility factor limited to
the second order as the additional term is empirical. At a given temperature, it has five
parameters, and in order to take into account their dependence on temperature, a total of
eight parameters. The values proposed for these parameters are generally specific to the
compound being examined.They may vary according to the range or the property that is
under investigation. It was also formulated using reduced coordinates so that the parameters depend only on the critical coordinates, or possibly on the acentric factor. Here we
show the form proposed by Starling [1971,1972]:
B o R T - A o -co
- + AD- T2 T3

bRT-u--

'T)

Eo
T4)'

( dy) p5+-(1+lip2)exp(-lip2)
T2

p2+a a+-

CP

(4.34)

where p, denotes the critical molar density and w the acentric factor:

(4.35)

We have also come across the formula used by Lee and Kesler (Eq. 3.25) to show the
properties of a simple fluid (methane) and of a reference fluid. We will have a chance to
cite more recent modifications regarding equations of state that are specific for certain
pure substances.
If the precision obtained is generally much greater than that obtained with equations
derived from the van der Wads theory, it must be maintained that the large number of

112

4. Equations of State

parameters and their lack of physical significanceundermine the definition of composition


rules and, because of this fact, their application to mixtures.

4.2

EQUATIONS OF STATE DERIVED FROM


THE VAN DER WAALS THEORY

We recall the equation of state proposed by van der Waals:

p = - RT
-v-b

a
v2

(4.3)

Firstly, we recognize a repulsion term that takes into account the volume of the molecules
using parameter b, or covolume,and secondly an attraction term, or internal pressure that
is dependent on the parameter a.
The values for these parameters are determined by applying critical constraints, and
at this time we can elaborate on the consequences of such a choice. There are three
constraints.The equation of state is satisfied at the critical point, as are the equations:

if),=(=),
a2P

=0

for T = T, and P = P,

Therefore, in the case of the van der Waals equation of state, we will have:
RT
a
p,= -- v,
v,-b
RT,

2a

2RT,

6a

=O

These three equations relate five quantities:the critical points T,, P,and v,,and the two
parameters of the equation of state a and b. We are therefore unable a priori to fix the
three critical coordinates. Generally, we choose to use the two parameters of the equation
of state and the critical volume (or the critical compressibility factor) as a function of the
critical temperature and the critical pressure that are more widely known. For this reason
and for the two parameter equations of state, the critical points correspondingto the equation of state will not coincide with the experimental values except for the values of pressure and temperature, with volume (or compressibilityfactor) being mostly overevaluated.
We end up with the equations:
R ~ T , ~
a=na(4.36)
P C

113

4. Equations of State

The numerical values for the adimensional parameters Ra and Rb,and the critical compressibility factor are:
27
64

Q = a

1
R - b- 8

and

3
Z= 8

(4.38)

If we take parameters a and b independent of temperature, all of the isotherms


obtained have the shape required, at least qualitatively, for describing liquid and vapor
states (the subcritical isotherm has a maximum and a minimum), or the supercritical
state (the isotherm is monotone). Furthermore, taking into account the two previous equations, we can express the van der Waals equation using reduced coordinates:
(4.39)

This equation describes the variation of the compressibility factor with reduced pressure and reduced temperature, and exhibits the corresponding states principle as a consequence of the van der Waals equation of state. This is how the law was introduced initially.
We need to emphasize the essential qualities for what was the work of a pioneer: the
representation for the different fluid states, and the prediction of the corresponding states
principle. We must also note the simplicity of the expression. At a given temperature and
pressure, the expression is third degree for volume and may be resolved easily without the
need for an iterative process. Additionally,the roots corresponding to the liquid and vapor
phases are easily identifiable.The equations of state derived from the van der Waals equation are often called cubic equations of state.
Yet, as van der Waals himself remarked,this equation of state is of unacceptable precision.
The compressibilityfactor that correspondsto it has a value (3/8) that is too high,and the calculation of molar volume in the liquid phase results in unacceptable systematic errors in
excess. In addition,if we determine the vapor pressure by applying the Maxwell condition of
the results are erroneous.
equality of areas contained between isotherms P(v) and isobar P,
These facts in part explain the considerableamount of work that has continued the studies of
van der Waals, and that has led to the equations of state that are today commonly applied.

4.2.1 The Soave-Redlich-Kwongand Peng-Robinson Equations of State


With the goal of better representing the volumetric properties of fluids, Redlich and
Kwong [1949] proposed an empirical modification to the van der Waals attraction term.
They also introduced a temperature dependence for the parameter a. The Redlich-Kwong
equation of state in its initial form is:
RT
p=-v-b

fTv(v+b)

(4.40)

where covolume b is related to the critical points by Equation (4.37), and the attraction
parameter is expressed as a function of these same coordinates by the expression:

114

4. Equations of State

(4.41)
For our purposes, we prefer to write the equation of state in the form:
RT
p=--v-b

a(T)
v(v+b)

(4.42)

(4.43)

with:

1
a(TJ = -

and, as with the van der Waals equation:

(4.44)

(4.37)
By applying the critical constraints, we determine the values of the parameters Lla and

a,,and of the critical compressibility factor. We find that:


1

Qa

9(2Ii3- 1)

= 0.42748

z = 1-

and

"

(4.45)

The results as they concern the calculation of density are definitely improved, especially
for the vapor phase, but remain mediocre for vapor pressure. The Redlich-Kwong equation of state, however, has proved one of the most useful. In particular, it embodies one of
the components of the method proposed by Chao and Seader [1961] for the calculation of
liquid vapor equilibria.
The two-parameter cubic equations of state are in fact unable to represent both volumetric behavior and vapor pressures with an acceptable precision [Abbott, 19891, and it is
appropriate to determine a priority. This was understood by Soave [1972] who applied the
Redlich-Kwong equation of state to the calculation of hydrocarbon vapor pressures.
Considering that the covolume retains the value corresponding to the critical point
(Eq. 4.37), it is possible to determine the value of the attraction parameter a at any temperature from the corresponding vapor pressure data (T,P"), by applying the equation of
state to the liquid and vapor phases and including the equilibrium condition (fugacity
equality). To fit the values thus obtained, Soave proposed applying the following expression:
(4.46)
a ( ~ ,=)[I + m (1 - f l ) 1 2
The parameter m is specific for the component in question but its value was correlated
as a function of the acentric factor:

m = M o + M I @ +M2m2
with:

Mo = 0.48, MI = 1.574,

M2 = - 0.176

(4.47)
(4.48)

115

4. Equations ofstate

For temperatures greater than the critical temperature, the previous equations are
admittedly applied without modification.
The value of Q, is, of course, the same as the value in the Redlich-Kwong equation of state
(0.42748.. .) so that the critical constraints are respected. The compressibility factor resulting
from the equation of state also remains untouched. Its value is equal to 1/3, clearly less than
the value that corresponds to the van der Waals equation, but still too high compared to
experimental values which, we recall, are located within the range of 0.25-0.29. We shall
therefore not be surprised that the Soave-Redlich-Kwong method results in systematic
excess deviation in the calculation of molar volumes, particularly for the liquid phase. For
hydrocarbons, this discrepancy is particularly noticeable when molecular weight increases.
Vapor pressures on the other hand, are pretty well represented as a result of the criterion that
was chosen for the definition of a (7). Between the boiling temperature at atmospheric pressure and the critical point, the deviations are 1 to 2% (Table 4.5). On the other hand, extrapolation towards low reduced temperatures is most often poor. The numerous modifications
that have been suggested since Soaves work aimed essentially at improving vapor pressure
calculation and at allowing an acceptable reproduction of molar volumes of the liquid phase.
The first of these calculations results from the work of Peng and Robinson [1976] who
modified the attraction term of the equation of state, which becomes:

RT
a
p = -v - b v2+2bv-b2

(4.49)

The method of parameter calculation is unchanged, and equations 4.37,4.43,4.46, and


4.47 are applied. It is understood that the numerical values of constants a,, f i b , M,, M I ,
and M2 are modified. They are shown in Table 4.6.
Due to the nature of the database, reproduction of vapor pressure is improved
(Table 4.5), but progress is primarily seen in the calculation of densities in the liquid phase.
Table 4.5
Average relative deviations (%) in the calculation of vapor pressures and densities
with the modified van der Waals (vdW*), the Soave-Redlich-Kwong (SRK),
and the Peng-Robinson (PR) equations of state [Rauzy, 19821

Component
Methane
Ethane
Propane
n-Butane
n-Pentane
n-Hexane
n-Heptane
n-Octane
n-Nonane
n-Decane

Densities
(T, = 0.7)

Vapor Pressures
(0.55 < T, < 1)
vdW

SRK

PR

vdW

SRK

PR

1.47
1.11
1.34
1.26
1.48
1.70
1.02
1.55
1.83
2.25

1.29
0.92
0.92
1.05
1.26
1.46
0.79
1.30
1.56
1.97

0.72
0.56
0.42
0.37
0.43
0.82
0.66
0.71
0.60
1.01

-34.5
-40.0
41.7
-44.0
-47.0
-50.9
-52.9
-55.7
-57.0
-55.9

-1.7
-5.3
-6.3
-7.8
-9.9
-13.8
-13.8
-15.6
-16.5
-15.4

10.2
1.5
6.1
4.8
3.0
0.6
-0.6
-2.2
-3.0
-2.1

* The van der Waals equation has been treated according to a method analogous to the one applied by
Soave (Eqs. 4.43 to 4.47).

116

4. Equations of State

Calculation of Thermodynamic Properties


By applying equations 2.29,2.31 to 2.33, and 2.40, it is possible to calculate the residual
properties, in other words the differences between the value of a thermodynamicproperty
for a fluid to which we apply one of the simple cubic equations of state, and the value corresponding to the ideal gas under the same conditions of temperature and pressure. As the
equation of state is specific for pressure but not volume, the method introduced in
Chapter 2, Section 2.4 is used. We shall do so for the enthalpy and the fugacity coefficient
with respect to the general cubic equation:
RT
p=-W-b

a
(V-brl)(v-br2)

(4.50)

which, depending on the value of parameters rl and r, represent the equations of state of
van der Waals (r, = r2 = 0), Soave-Redlich-Kwong (rl = 0, r2 = -l), or Peng-Robinson
(rl = -1 - lh,r2 = -1 + lh).
Calculation of the Residual Enthalpy
First, we will calculate the residual internal energy at the given temperature and volume:

u , , , ~ T,w)
~ ~ =~ U(
~ T,P,w)
(
-U#

T( g ) , - P ] dw

(2.28)

i.e., by using general formula (4.50):

( dd;)jI

u(T,w)- u#(T,w)= u - T -

yielding:

u( T,w) - u#(T,w) - 1

RT

a-T-

RT

dw
(w - br,) (w - br,)
da
dT

lJ(w,b,r,,r2)

(4.51)

(4.52)

where function U(w,b,rl,r2)depends on the equation of state considered (Table 4.6). Since
the internal energy of the ideal gas is independent of volume, (or pressure), its residual value
at a given temperature and pressure is the same.For the residual enthalpy we have therefore:
h (T,P) - h#(T,P) - 1
RT
RT

a-Tb

da
dT

U(~,b,rI,r,)+ 2 - 1

(4.53)

If we use a temperature that is less than the critical temperature and a pressure equal to
the vapor pressure (itself calculated using the equation of state), the application of this
equation to the saturated vapor phase (value of the volume corresponding to the largest
root of the equation of state), and then to the saturated liquid (value of the volume corresponding to the smallest root) gives us the residual enthalpy of the two phases. Their difference is the enthalpy of vaporization. In general, such an approach yields acceptable results
Note that the residual enthalpy expression uses the derivative of the attraction parameter a as related to temperature. If we were to calculate residual heat capacity, the second
derivative would be used as well. Comparison of the results of such calculations is a means

117

4. Equations of State

of validating the proposed a( T ) equations [Trebble and Bishnoi, 19871.The one proposed
by Soave (Eq. 4.46) yields satisfactory results (Fig. 4.3).

250

300

350
Temperature (K)

Figure 4.3 Heat capacity of ethane at constant pressure (P= 6 MPa);


calculation using the Soave-Redlich-Kwongequation of state (-)
and experimental data (A).

Fugacity Coefficient
First we will calculate the residual Helmholtz energy for a given temperature and volume:

1.e.:
yielding:

:I (

U(T,v)-U#(T,V)=

--

vT+?+(v-brl)(v-br,)

(4.54)

(4.55)

where the function U(w,b,rl,r2) is the same as the one used for the residual enthalpy
(Table 4.6).
To obtain the residual Gibbs energy at given temperature and pressure, we proceed as
discussed in Chapter 2, Section 2.4.2 and thus obtain:

118

4. Equations of State

that yields the expression for the fugacity coefficient:

(4.57)
in which Z is the compressibility factor and the function U is identical to the one used in
the residual enthalpy expression.
Table 4.6
Van der Waals, Soave-Redlich-Kwong,and Peng-Robinsonequations expressed by
the general equations 4.50,4.43,4.37,4.46,4.47,4.53,
and 4.57
Equation of state
Variable

t-GGGZ
0
0
27/64
118
0.500 0
1.5883
-0.175 7
318
b
-U

Soave-Redlich-Kwong

0
-1
1/9(21/3-1)
(21" - i y 3
0.48
1.574
-0.176
113
1 In( v-br,
r, - r 2
v-br,

-)

Peng-Robinson
-1 -Ih
-1iIh
0.45724
0.07780
0.37464
1.54226
-0.269 92
0.307
1
u-br,

-)

r1-r2 In( v -br,

In order to calculate vapor pressure, we must proceed using iterations:having formed a


pressure hypothesis, we solve the equation of state creating in this way two roots that correspond to the liquid and vapor states. For each root, we calculate the fugacity coefficient
by application of the previous expression 4.57. Fugacity, and therefore the fugacity coefficients that relate to each phase should be equivalent.
Table 4.6 above lists for the van der Waals, Soave-Redlich-Kwong,and Peng-Robinson
equations the values of the principal variables that occur in the equation of state
(Eq. 4.50), the expression for covolume b (Eq. 4.37), the attraction parameter a (Eqs. 4.43,
4.46, 4.47), and for the function U(w,b,rl,r2).To the van der Waals equation of state, we
applied the method recommended by Soave in order to calculate the attraction parameter
(Eqs. 4.43,4.46,4.47).

0 EXAMPLE4.4
Expressing the second virial coefficient relating to equations
of state as expressed in the general form 4.50
First, recall that the second virial coefficient may be defined by the equation:

4. Equations of State

119

To the equation of state:


RT
p=-u -b

a
(U

- br1) (U - br2)

we shall apply the following variable transformation:


b
q= U

(4.50)

(4.58)

by introducing the reduced density or packing fraction, a variable that expresses


the fraction of total volume occupied by the molecules.
The equation of state is therefore written as:
(4.59)
Furthermore, the second virial coefficient will be determined using the following
eauation:

(4.60)
Using the equation of state, we may write:

B=b- RT

and derive:

(4.61)

We note that this expression for the second virial coefficient is independent of parameters rI and rT
We may ask ourselves if in addition it approximates experimental values. We know
(Chapter 3, Section 3.2.2.1) that these values are correlated by the following equations with acceptable accuracy:

BPC = F(O)(T,) + o F ( l ) (T,)

(3.18)

with:

0.330 0.1385 0.0121 0.000607


F(O)(T,)= 0.1445 - -- -- -Tr
Tf
Tr3
T,8

(3.19)

and:

0.331 0.423 0.008


F()(Tr)= 0.0637 + -- -- Tf
T?
T,8

(3.20)

RTC

If, for example, by applying these equations we calculate the second virial coefficient
at the critical temperature (T, = l),we find:

BPC = -0.336707 - 0.0363 o


RTC

120

4. Equations of State

Application of the equation formulated from the equation of state, always at the critical temperature, on the other hand yields the following expression:

or even still, by using the general equations 4.37,4.43, or 4.46:

BP, = L!b - SZ,= 0.086 64 - 0.427 48 = -0.340 84


RT,

We can therefore determine the value of the acentric factor for which the second virial coefficient formulated from equations 3.18 to 3.20 is the same. We find o = 0.113.
Note that the second virial coefficient that relates to the cubic equations of state that
we have just examined is, at the critical temperature, independent of the acentric factor. The same is true for the critical compressibilityfactor.

CJ EXAMPLE 4.5
Application of the equation from
the Soave-Redlich-Kwongequation of state at low pressure
If we apply any one of the equations of state illustrated by Equation 4.50 at zero pressure, we observe that the equation becomes of the second order. In the case of a
Soave-Redlich-Kwongequation of state it may be written as:
(4.62)
This equation has real roots, as shown in Figure 4.4, if the reduced temperature is less
than a limit that we can easily estimate by using Equations (4.37,4.43,4.46,and 4.47).
We can question their physical significance.As an example, we use benzene at its boiling temperature under atmospheric pressure (353.25 K). The values of the variables a
and b obtained from these same equations,and from the values of the critical properties, and of the acentric factor (T, = 562.16 K, P, = 48.98 bar, o = 0.2108) are:
b = 82.679 cm3* mol-l,

m = 0.80428,

a = 1.3611

and -= 10.6872
bRT

The roots of the second order equation are equal to 105cm3.mol-l and 696 cm3 mol-l.
The highest corresponds to an unstable state, for which the derivative of the pressure
with volume is positive (0.052 bar *
as we can see from the calculation of this
derivative from the equation of state.The lowest on the other hand, is very close to the
value we would obtain by applying the equation of state at atmospheric pressure and in
the liquid phase (meaning saturated liquid, since we used boiling temperature at
atmospheric pressure). Indeed, when the root is at zero pressure,compressibility is very
low (0.025 cm3 * bar-').

121

4. Equations of State

200

400

600

Molar volume (cm3 . mol-1)

Figure 4.4 Roots for the Soave-Redlich-Kwongequation of state at


zero pressure. Application to benzene at 80C.

EXAMPLE 4.6

Application of the Soave method to the van der Waals equation of state
We shall take here a very detailed look at the process used by Soave to identlfy the
parameters for an equation of state.'Ihe numerical application relates to benzene at 25C.
Applying the van der Waals equation of state, we shall first determine the parameters
at the critical temperature. To do this, we use the critical conditions or, more simply,
write the equation of state, as a third order equation in volume:
a
ab
+ b)v2 + v - -= 0
(4.63)
v3-(
P
P
At the critical point, the equation has a triple root, and is therefore equivalent to the
equation:
(v - vc)3 = 0
We thus obtain three equations relating the critical coordinates to the parameters:
RTC (1+ 0,)
3vc = P C

ab

v3= pc

122

4. Equations of State

now we perform the following transformation of variables:

where a,,and Q,, are numbers with no dimension. The relationships between the
parameters and the critical coordinates become;
3Zc=1+ab
3z,2=aa z;=fi,,ab
The solutions for this system are:
1
z = 3an=27
64
8
' 8
At the critical temperature, and for benzene (T, = 562.1 K, P, = 49.04 bar), the values
for the variables are therefore;

a=-

a = 1.8817 J * m3 rnol-'

b = 1.1929 lo4 m3 . mol-'

Next we shall consider parameter b as independent of temperature and we shall calculate parameter a such that at 25C the experimental value of vapor pressure,
12.692 kPa (see Table 2.6) is exactly calculated by the equation of state.
If we designate
and
as the roots of the equation of state at equilibrium of the
liquid and vapor phases, and designate P as the vapor pressure, the following equations must be verified:
~

"

(4.64)
(4.65)
They represent the application of the equation of state to the liquid phase and to the
vapor phase. The fugacity coefficients of these two phases must also be equal. For the
van der Waals equation of state, the expression for this coefficient is (Table 4.6):
P ( v - b)
a b
lny,=-ln
+Z-l-(4.66)
RT
bRT v
and therefore we have:
P"(v L,' - b) +ZGC'-l-- a b
- In
RT
bRT v C a
~

P"(v vu - b )
a
b
+ Z W - 1 - -(4.67)
RT
bRT vv'
The three preceding equations allow for the determination of the three unknowns: a,
and vyu. To simplify the problem, we take into account the fact that pressure is
low and, therefore, the fugacity coefficient is very close to unity since the saturated
vapor may be compared to an ideal gas. This simplification allows us to restrict the
problem to the liquid phase and retain only the equations:
= - In

123

4. Equations of State

2,

(cm3 . mol-I)

(J.m3. mol-I)

In cp

150
140
130
135
136

1.816
2.345
3.91
2.875
2.743

2.875
1.39
-3.32
-0.161 5
0.231

Conclusions Concerning the Soave and Peng-Robinson Equations of State

By proposing a novel method for the determination of the attraction parameter of equations of state derived from the van der Waals theory, Soave allowed for the calculation of
liquid vapor equilibria in mixtures and apolar compounds,and petroleum fluids in particular. The application of such equations of state to a pure substance should be viewed only as
a prerequisite, and we will return to these methods when studying mixtures. In particular,
it is appropriate to note that if we are only dealing with the calculation of the vapor pressure for a pure substance (apolar), application of the preceding methods (Eq. 3.32 from
Lee and Kesler, for example) is simpler and more precise. If we also want to know the den-

124

4. Equations of State

sity at saturation, the Rackett method is preferred (Eq. 3.14). Finally,for a pure substance,
if we are interested in all thermodynamic properties, the Lee and Kesler method yields
better results. Whatever the case, the cubic equations of state that we have just discussed
provide only a mediocre representation of the critical region, as shown in Figure 4.5 that
allows for a comparison of experimental data with the calculation of the critical isotherm
achieved by using the Soave-Redlich-Kwongand Peng-Robinson equations of state.
It is therefore for liquid vapor equilibria of mixtures under pressure that the SoaveRedlich-Kwong and Peng-Robinson equations of state find their principal application,and
seem at the present time to be irreplaceable.We shall return (Chapter 8) to these methods
by defining mixing rules that link the values of the parameters with the composition and
upon which the practical value of an equation of state strongly depends.

4.2.2

Recent Developments of Cubic Equations of State

The generalized application of the Soave-Redlich-Kwongequation of state to calculations


in chemical engineering and to the exploitation of petroleum fluids (natural gas, crude
oils) has given rise to a large amount of work aimed at rectifying the shortcomings.
Limiting ourselves in this chapter to the formulation of these equations for pure substances, which is only a first but essential step, these flaws lie within predictions that are
sometimes inadequate for vapor pressures, and generally mediocre for volume. Without
trying to provide an exhaustive account of modifications that have been proposed, we will
present a few examples.
30

25

h
E

20

9 15
3

u)

ff 10
5

I
4

10

12

14

16

18

Molar density (mol . 1-1)

Figure 4.5 Calculation of the critical isotherm for ethane using


the Soave-Redlich-Kwong(-)
and the Peng-Robinson (- -)
equations of state. Comparison to experimental data (A).

4. Equations of State

4.2.2.1

125

Dependence of Attraction Parameter a on Temperature

It may be essential to precisely reproduce the vapor pressures for the components of a
mixture. Such is the case with C; (butane, isobutane,and butenes) hydrocarbon fractionation studies where relative volatilities are close to unity. In this case, the selection of an
appropriate a( T ) equation is more important than the choice of an improved mixing rule.
It is equally important to make use of precise expressions when studying and comparing
the mixing rules whose effects may be masked by an inadequacy of the a( T ) law.
If for such componentswe have access to accurate experimentalvapor pressure data, or
a satisfactory P( T ) expression, it is then possible to match an attraction parameter value
to each point on the experimentalvapor pressure curve. We may then correlate the results
obtained in the form of an a( T ) law that is more accurate than the one proposed by Soave,
but specific to each component.The main problems concern the extrapolation of this law
to temperatures that are higher than the critical temperature, or to low temperatures, and
its possible generalization for predicting its parameters as was done for Equation 4.47
using the acentric factor.
Extrapolation beyond the critical point requires particular attention. If we adopt two
distinct expressions according to the temperature range, one must be aware that the calculations for enthalpy will undergo discontinuity if the derivative daldT is not continuous. It
will be the same for heat capacity if the second derivative is not also continuous.
Furthermore, in the supercriticalregion we cannot rely on volumetric data that are poorly
represented by these equations. However, we must note that this difficulty occurs only in
cases of compounds for which the range of study includes the critical temperature, which
limits their number.
As for extrapolation to low temperatures, we are faced with a scarcity of experimental
data that allows for the estimation of the quality of the proposed correlation.
Among the numerous expressions that have been proposed, we quote some examples.
Mathias and Copeman [1983] apply the following equation to the Peng-Robinson equation of state:
+ c2(i - fl12 + c3(1- -)3]2
(4.68)
~ ( T =J [I + cl(i - -1
which is the same as the original equation if parameters c2 and cg equal zero. It is in this
way that it is applied for the supercritical region. The values of the three parameters are
specific for the substance being studied, and there is no attempt to generalize them. Of
course, this expression may be applied to either the Soave-Redlich-Kwongequation or the
van der Waals equation as long as the numerical values of the parameters differ.
While preserving the general expression for the attraction term:
R~T,~
(4.43)
a(T) = Q, -a(T,)
pc
Stryjek and Vera [1986] proposed an expression for the a parameter as a function of
reduced temperature:
(4.69)
a = mo+ ml(l + *)(0.7 - T,)
where the parameter mo is calculated from the acentric factor, and the parameter m lis
adjusted using the experimental data [see Proust and Vera, 19891. In the supercritical

126

4. Equations of State

domain, m, is considered zero. Its application to the calculation of vapor pressures for a
large number of polar and nonpolar compounds yields good results. It is, however, important that the values for the critical coordinates and the acentric factor are identical to those
used by the authors.
When the only modifications introduced to the equation of state relate to the modification of the law linking the attraction parameter to temperature, the consequences for the
calculation of the volume are minimal. Of course, it is not the same if the attraction term is
modified. The examples that we shall provide now have both a more profound modification, and often the suggestion of a new a( T) law.

4.2.2.2

Modifications of the Attraction Term

Applying the critical constraints to an equation of state having only two parameters allows
only to comply with two of the experimental critical coordinates, and the critical compressibility factor is determined by the form of the equation, generally an underestimation.
The use of a third parameter provides an additional flexibility if we wish to estimate liquid
vapor equilibrium and volume at the same time.
Schmidt and Wenzel [1980], Harmens and Knapp [1980], Heyen [1980], and Patel and
Teja [1982] suggest an equation of state that corroborates the general equation:
RT
a
p=-where u + w = 1
(4.70)
v - b v 2 + u b v +wb2
The fundamental inadequacy of equations of state to correctly represent the critical
zone and the saturated liquid at the same time makes it impossible to choose a third
parameter such that the three critical coordinates are exactly described. Although the
methods chosen by the authors for the calculation of the parameters differ, the results are
not fundamentally different. The most applied of these equations is undoubtedly that of
Patel and Teja, which has the form:
RT
a
p=-(4.71)
v - b v (V + b) + C ( V - b)

c,,

As we noted above, the critical compressibility factor, imposed on the equation of


state differs from the experimental value. It is related to the acentric factor by the equation:
6, = 0.329032 - 0.076799~+ 0.0211 9 4 7 ~ ~
(4.72)
This equation of state is similar to that of Redlich-Kwong for simple fluids (o= 0 ) and
to that of Peng-Robinson when the acentric factor is in the neighborhood of 0.3. It provides a useful compromise between the two equations.
We
Through the critical condition intermediary, the other parameters depend on
apply the usual equations:
R ~ T , ~
a ( T ) =aa
-a ( T , )
(4.43)

cc.

P
C

(4.37)

127

4. Equations of State

RTC
c=ac-

(4.73)

PC

is the smallest positive root of the equation:


(4.74)

a, and a, are calculated using the equations;


Qa= 3

c, + 3(1-2 cc)s2,+

+ 1 - 3 Cc

0,=1-3(,

(4.75)
(4.76)

The variation of the attraction parameter with temperature is monitored by an equation


that is identical to the one proposed by Soave:

~ ( T =J [I + m(1- V T J ] ~

(4.46)

where the parameter m is related to the acentric factor (for apolar substances):
m = 0.452413 + 1.3098201- 0.29593701~

(4.77)

The results obtained using this method are analogous to those produced by the method
proposed by Schmidt and Wenzel. They are improved, particularly for the calculation of
volumes when compared to the two parameter equations. The concept of translation will
allow us to investigate this improvement. It remains understood that the behavior in the
critical zone remains poor.
The equation of state proposed by Heyen distinguishes itself by the fact that covolume b varies with temperature.Thus it is possible to represent molar volumes for saturated
liquid with good precision. However,Trebble and Bishnoi [1987] have recently shown that
such a variation may cause incoherencesfor the calculation of heat capacity.
Volume Translation and the General Cubic Equation
We can in fact recover the various equations that we just cited using the volume translation concept, introduced by Martin [1979], and whose value was shown by PCneloux et al.
[1982].
If we suppose that we are using two equations of state e(T,P,v) = 0 and e(T,P,v)= 0
such that at a given temperature and for all values of pressure they provide the values v
and v for the volume, differing only by a constant value c(v = v - c), that (possibly) only
depends on temperature, their graphical representation of volume versus pressure
(Fig. 4.6) is made up of two curves, one translated from the other. If at a certain pressure P
the Maxwell condition (equality of areas located between curve P ( v ) and isobar P) is
respected for one, it is also respected for the other. These two equations of state therefore
yield the same results for the calculation of vapor pressure. Hence the calculation of volumes and phase equilibria are to a certain extent independent. This method is, of course,
valid for any equation of state, but is especially effective in affording otherwise lacking
flexibility to the methods derived from the van der Waals theory.

128

4. Equations of State

g 4
v

p!
3
0
cn

p!

100

200

300

400

500

Molar volume (cm3 . mol-1)

Figure 4.6 Translatedequations of state.

These equations are listed according to the general formula:


RT
a
p = -V- b
V[v+ yb]

(4.78)

v=v-c

(4.79)

b=b-c

(4.80)

The first equation is the generalized cubic equation that yields, via translation
(parameter c), the equation:

RT
a
p = -v - b (v + C)[V+ c + y(b + c)]

(4.81)

Like equations 4.50 and 4.70, this equation may be representative of all the cubic equations that we have previously encountered.We establish the following correspondences:
Redlich-KwongEquation: y= 1

and

2
Peng-Robinson Equation: y= - and
V2-1

c =0

(4.82)
c=-b(Ih-l)

All the equations of state stemming from the general expression 4.50 that, for example,
correspond to this last value of and for which parameter a is calculated according to the
equations recommended by Peng and Robinson, will yield the same results for the liquidvapor equilibrium calculation. Yet the calculation of molar volume will depend on the
value of c. This value may be adjusted for the substance in question. Indeed, we frequently
use data such as density at ordinary temperature or, for low molecular weight compounds,

4. Equations of State

129

the normal boiling temperature. We may then calculate parameter c such that the equation
of state returns exactly such data. It has also been proposed that c be correlated as a function of the Rackett factor.
Translation of the van der Waals, Redlich-Kwong, and Peng-Robinson equations has the
effect of evening out the considerable differences they produced in the calculation of
molar volume.
The translation term may vary with temperature, and we may be tempted to subject it to
a law such that liquid volume is restored along the saturation curve. It results in a rapid
variation of its value around the critical point that cannot be extrapolated within the critical region. In addition, neither the critical isotherm nor the isotherms in the critical zone
are really improved because the translation does not correct for the shortcomings of the
cubic equations within this region.
The differences between the thermodynamic properties of the liquid and vapor phases
in equilibrium are obviously not modified by translation. Therefore, the enthalpy of vaporization remains unchanged.

EXAMPLE 4.7

Application of the translation to the van der Waals equation


Returning to the previous example, we will determine the translation parameter for
benzene such that the molar volume in the liquid phase at 25C is represented exactly.
The experimental value is available in the literature: 88.3 cm3 .mol-l. We furthermore
calculated the value that corresponds to the van der Waals volume (135.4 cm3. mol-l)
where the a parameter is determined such that at the given temperature, the vapor
pressure is correctly described. The translation that must be performed is therefore
equal to the difference between these two values, or 135.4 - 88.3 = 47.1 cm3 mol-.
After application of equations 4.79 and 4.80, the newly derived equation of state may
be written as:
a
p = - -RT
(4.83)
v-b
(v+c)~
The value of parameter a is unchanged, but parameter b undergoes translation and,
taking into account the previously determined value (119.29 cm3 . mol-), becomes
equal to 72.19 cm3 .mol-.
We recognize the Clausius equation of state in the preceding equation.

f i e SBR Equation of State


Whatever the modifications to the van der Waals equation of state, it seems impossible to
satisfactorily represent the critical region. This shortcoming is to a certain extent common to any analytical representation of this region. It is particularly striking in the case of
equations that are as simple as the ones we have dealt with up until this point. We can try

130

4. Equations of State

to remedy this by abandoning the cubic structure of the equation of state. It is in this way
that Behar et aZ. [1985] proposed to develop the attraction term of the Redlich-Kwong
equation using density:
(4.84)
The four parameters of this equation vary with temperature. Their value has been correlated as a function of reduced temperature and the acentric factor. In order to better
represent the vapor pressure of compounds with high molar mass, their formulation has
been revised recently [Jullian et aZ., 19891.
The critical constraint has not been included in the determination of the parameters.
Nevertheless, because of the close attention paid to the volumetric behavior in the critical
region, the calculated critical points are reasonably close to the experimental data.
This equation marks undeniable progress in the calculation of volumes as a function of
pressure and temperature in the critical zone and for the liquid phase. It yields results analogous to those obtained when applying the Soave or Peng-Robinson methods for the estimation of vapor pressures. The fact that due to its structure its resolution demands an iterative calculation should not be considered a serious handicap. It has been applied to the
calculation of phase equilibria for petroleum mixtures (crude oil and natural gases).
However, it applies only to non-polar compounds.
We can compare this equation to the one proposed by Suzuki and Sue [1989] that has a
similar structure, and has been applied to alcohols, ammonia, and water on the condition
that specific parameters are used. Results have been generally acceptable.
It is not pertinent to list here the innumerable variations of the van der Waals equation
of state that have been proposed. They embody a fair amount of empiricism, and the
acceptable results that have been obtained are in fact due to a fortunate balance between
the inaccuracies contained within the expression for the repulsion term, and the various
forms that have been proposed for the attraction term.

4.2.2.3

Application of the Concept of Group Contribution

It is difficult to apply the equations that we have just described to hydrocarbons of high
molecular weight. Indeed, their parameters are calculated using critical points and these
points are not always known or cannot be determined experimentally due to heat decomposition. It is certainly possible to apply one of the numerous structure property equations
proposed for these parameters (Chapter 3), but we are aware of their faults, especially with
respect to extrapolation. Furthermore, the equation of state will be applied more often
within a range of temperatures that are much lower than the critical temperature, and the
application of the usual equations to the calculation of these parameters necessitates a
cumbersome detour. Finally, we note that the application of the corresponding states principle to high molecular weight structures is a poorly justified extension of it.
It has therefore been proposed that we apply the methods of group contribution to the
equation of state parameters. We recall the example [Coniglio, 1991; Coniglio et aZ., 19931
of the general cubic equation 4.78 where y is fixed at the value corresponding to the PengRobinson equation of state (see Eq. 4.82).

131

4. Equations of State

The parameters to be determined are therefore covolume b, the attraction parameter a,


and the translation term c.
We consider that the covolume b and the van der Waals volume Vware proportional:
(4.85)

where the van der Waals volume is calculated using:


3

in which N j represents the number of groups of type j , Vw,jis the contribution attributed to
them, S , , is the corrective term that corresponds to structure k, and Ik is the number of k
structures present. The groups in question correspond to the simplest hydrocarbon structures (Table 4.8), and in order to take into account the ring structures, we introduced the
term Sw,k.
Table 4.8
Group contributions and structural increments for the estimation of
van der Waals volume (Eq. 4.86) and translation volume (Eqs. 4.96 and 4.97)
v w

(cm3/mol)

(cm3/mol)

13.67
10.23
6.78
3.33

30.294 0
11.9370
-8.203 9
-28.045 8

8.06
5.54
4.74

16.7162
-2.898 6
7.1473

Naphthenic ring
Cyclopentyl or cyclohexyl
Free or condensed, trans
(trans-decaline)

-1.14

35.593 2

Naphthenic ring
Cyclopentyl or cyclohexyl
Free or condensed, cis (cis-decaline)

-2.50

37.376 8

Methyl ring
Condensed at benzene
(tetraline, indane, etc.)

-1.66

37.904 8

Groups

Wanes
CH3
CH2
CH
Quaternary C
Aromatic groups
CH
Substituted C
Condensed C ring
Structures
~

132

4. Equations of State

The calculation for the attraction parameter is based on the following equation:
(4.87)
where a (Tb)is the value that corresponds to boiling temperature at atmospheric pressure.
Functionsf,(rn) andf2(rn) are dependent on a form parameter rn:
(4.88)
(4.89)
with C, = 12.5295 and C , = 41.3891,x = 0.05, and y = 1.
The form parameter rn is calculated from term S, which is itself arrived at by group contributions:
rn = 0.341 90 + S + 0.18473 S 2
(4.90)
7

s=

NjMj+

zk6vnk

(4.91)

k=l

j= 1

where N j is the number of type j groups, Mi is the contribution attributed to them, 6mk is
the corrective term corresponding to structure k, and Zk the number of k structures
involved. The groups under examination are the same as those that relate to the calculation of the covolume, but the structural effects are different and take into account the
effects of chain length, substitution, non-planar aspect, and steric hindrance (Table 4.9).
For certain compounds (Table 4.10), a special correction is necessary and the structure correction terms are done away with. We then have:
rn = 0.34190 + S + 0.18473 S2+ Am
(4.92)
7

s= 2 N

~ M ~

(4.93)

j=1

The translation term c varies with temperature according to the law:


C ( T ) = C(Tb)[I + a ( 1 - Y )+ p(1- Y ) , ]
with:
Y = exp (1- TI Tb), a = 1.8746
T: - 0.2464, p = 2.2079 10" Tb- 1.2179

(4.94)
(4.95)

where c (Tb) is the translation at boiling temperature at atmospheric pressure. It can be


estimated either from the density data, or the equation:
C(Tb) = -0.236 1 Tb + 59.6102 m2 + s

(4.96)

where S is calculated by group contribution:


7

(4.97)
the groups and the structures are the same as those applied to the estimation of covolume.
The numerical values for the increments are shown in Table 4.8. In the special case of

133

4. Equations ofState

cyclopentane and cyclohexane, it is necessary to correct the c (Tb)expression by adding a


corrective term Ac equal to -2.498 and -3.9711, respectively.
The equation of state defined in this manner has been applied to the calculation of
vapor pressure, liquid phase density, heat of vaporization, and specific heat at constant
pressure in the liquid phase. Special attention is paid to the range of low reduced temperatures, and to high molecular weight compounds. The results are of a precision comparable
to experimental precision.
Table 4.9
Group contributions and structural increments for the estimation
of the form parameter m (Eqs. 4.90 and 4.91)
Groups

Alkanes
CH3
CHZ
CH
Quaternary C
Aromatic groups
CH
Substituted C
Condensed C ring
Structures

M
0.041 25
0.04303
0.028 02
0
0.02678
0.03667
0.00449
&?l

Normal alkanes(
Ik = N;0.5 for N, a 7

0.045 23

Substituted alkanes (2)(3)


Ik = Nqp + N,,, - 6 for N , =z8
Ik = 2 for N , a 9

0.007 02

m
P
free or condensed (cis- or trans-) cyclopentyl
in the naphthenes
free or condensed (cis-or trans-) cyclohexyl
in the naphthenes
Isopropyl or t-butyl bound to an aromatic
ring or naphthenic
poiyaro~ties(3)(4)
I, = 1for naphthalene and its derivatives
Ik = (-1) a 0.15 (Nc2 + 2 Nc3) for the other
compounds (phenanthrene, pyrene, etc.)

-0.04019
-0.06837
0.00746
0.018 12

(1) N , is the total number of carbon atoms.


(2) Substituted alkanes without quaternary carbons or contiguous double substitution.
(3). NGPis the number of carbons from the principal chain, Nsubs,the number of substi-

tutions
(4) a is the number of aromatic nuclei, Nc2 and Nc3 are the number of aromatic carbon
atoms common to 2 or 3 rings.

134

4. Equations of State

Compound
Cyclopentane
Cyclohexane
Isopropylcyclohexane
Benzene
Toluene

4.2.2.4

Am

-0.043 22
-0.051 43
-0.05022
0.02136
0.007 08

Equations of State for Rigid Spheres and Hard Chains

In a comparison of rigid sphere molecules where the repulsion energy is either infinite,
when the particles are in contact, or zero, and deriving inspiration from statistical thermodynamics and molecular dynamics, Carnahan and Starling [1969] propose an expression
for the compressibility factor as follows:

(4.98)
in which q is the reduced density:

V0

q= -

(4.99)

and v O is the actual volume of one mole of rigid spheres.


In uniting this expression for the repulsion term and the expressions proposed by van
der Waals or Redlich-Kwong for the attraction term [Carnahan and Starling, 19721, we
obtain equations of state that may be compared to those having the van der Waals repulsion term. Such a comparison was made by Marchand [1984], who demonstrated that, provided that the determination of the parameters for the two kinds of equations are done in
the same way (without inclusion of the critical constraint, and by adjusting to the vapor
pressure and volume data), the results were equivalent. Dohrn and Prausnitz [1990] also
proposed a development for the Redlich-Kwong attraction term that is analogous to the
formula embraced by BChar et al. [1985].
We may rightly conclude from this that it is necessary to better take into account the
molecular form, and expand a less empirical attraction term. Many authors have done this,
and we can cite the BACK equations [Chen and Kreglewski, 19771, PHCT or Perturbed
Hard Chain Theory [Beret and Prausnitz, 1975; Donohue and Prausnitz, 1978; Donohue
and Vilmachand, 19881, and COR or Chain of Rotators [Chien et al., 19831. Without getting into the details of the expressions proposed by the authors of these equations, we point
out that the repulsion term, derived from the Carnahan Starling term, has an additional
parameter. In general, it is adjustable and represents the length of the molecular chain or
the degrees of freedom of the molecular structure. The attraction term was inspired by the
work of Alder et al. [1972],and we find a volume parameter (already a part of the repulsion
term), and an energy parameter represented by a characteristic temperature. These three
parameters are unique to each component and may be substituted, to a certain extent, for
the critical coordinates and the acentric factor, which we do not need to know.

135

4. Equations of State

These equations of state have recently been the subject of a comparative study
[Solimando, 19911.We shall report (Table 4.11) the results that pertain to the calculations
of molar volume at saturation, vapor pressure, and the critical coordinates. These last values, as determined by the equation of state, differ from the experimental values since the
critical constraints are not taken into account. For purposes of comparison, this table also
lists the results obtained with the general cubic equation, expressed as PRc:
RT
a
p = -d - b v[v+ yb]

(4.78)

in which the value for parameter yis the one corresponding to the Peng-Robinson equation, and the translation calculated such that the liquid phase molar volume at a reduced
temperature of 0.7 is restored exactly. For this last equation, the critical constraints have
been taken into account.
Table 4.11
Comparative results of equations PRc, BACK, PHCT, and COR. Relative deviations (%)
of vapor pressure, molar volume in liquid phase at saturation, density in the supercritical
region and absolute deviation (K) from critical temperature [Solimando, 19911
Equation
PRc
BACK
PHCT
COR

6PU

6V
-

Pa
0.54
0.60
0.90
0.30

zIL,a

6P
P

6Tc

2.4
0.21
1.04
0.28

3.05
2.21
5.53
2.40

2.43
17.00
5.03

We observe that the equations confer, in comparison with the PRc cubic equation, a certain improvement, particularly with regard to the calculation of molar volumes in the liquid phase. This improvement is due in part to the abandonment of the critical constraint,
and the error on the critical temperature is far from being negligible. In the case of the
PHCT equation, it is unacceptable and leads to a degeneration of results within the critical
region. It is also appropriate to take into account the erratic behavior of the BACK equation at low reduced temperature that impedes the calculation of vapor pressure.
In this comparative study, the COR equation stands out as superior. Its expression is
given below:
Z = 1+(4~-2112)/(1-~)3+0,5~(a-1)[311+3a~-((a-1)~2]/(1-~)3
4

+ [1+ 0,5 c

(B,

+ B,T*/T + B,T/T*)]

mA,,,,,(T */T) ( V * / V ) ~ (4.100)


n=l m=l

with:

(4.101)

B, = 0.20095, B, = 0.019, B, = -0.0632, a = 1.078, and the universal coefficients A,,,,,are


provided in Table 4.12.

136

4. Equations of State

7
A,,

Table 4.12
coefficients for the COR equation [Chien et al., 19831
1

-9.042 14
-125.11
525.415
-859.803
634.635
-167.336

-1.125 17
548.709
-2566.20
4 471.80
-3 402.75
939.226

-0.809 958
-838.503
4398.77
-8 598.81
7 409.9085
-2365.34

1
-0.672 378
438.783
-2 482.01
5289.80
-5017.09
1784.58

To each component correspond three characteristicparameters, c, T* and v*, for which


Table 4.13 provides the values for some hydrocarbons. .
Table 4.W
Characteristic parameters for the COR equation [Chien et al., 19831

I &

Component
Methane
Ethane
Propane
n-Butane

I
I

4.3

152.19
226.48
263.53
293.52

21.07
30.56
41.37

1.899
3.206

SPECIFIC EQUATIONS OF STATE FOR CERTAIN PURE SUBSTANCES

For a certain number of compounds that are important in industry such as water, ammonia, and the light hydrocarbons, numerous and varied experimental data are available
(PVT measurements, vapor pressure, heat capacities,speed of sound). Relying on these
data which have been subject to rigorous prior evaluation, one has proposed equations of
state that are capable of reproducing them with excellent precision. We may therefore
consider the results of the calculation to be quasi-experimentaldata.
The equations of state are often specific if not to one component, at least to a small
group of compounds.They lend themselves poorly to the representation of mixtures, especially to the calculation of vapor-liquid equilibria. Their form is relatively complex and
they contain numerous parameters. We shall provide some examples.
Younglove and Ely 119871 apply an equation to the light hydrocarbons that is similar to
the Benedict,Webb, and Rubin equation where p is molar density, expressed in moles per
liter, P is pressure, and T is the temperature. The equation is written as:
(4.102)
1

4. Equations of State

137

The coefficients A , and B, are temperature dependent and the equation has a total of
32 parameters. It may be applied to methane, ethane, propane, n-butane, and isobutane.
The residual values relating to the thermodynamic properties (internal energy, enthalpy,
entropy, and heat capacities at constant volume and constant pressure) as well as the speed
of sound can thus be calculated.
By applying the Maxwell condition it is also possible to calculate the vapor pressure.
However, the properties of the liquid and vapor phases at equilibrium (vapor pressure and
molar density) are provided by auxiliary (and explicit) correlations as a function of temperature, and the result is practically identical.
The range of application for this formula is large: up to 200 MPa and 600 K for methane,
and 70 MPa and 600 K for ethane, for example. Precision is good: 0.2% for the calculation
of density, except in the immediate proximity of the critical point. The analysis of this precision as a function of the compound, the range, and the property is proposed by the
authors.
To better represent the critical region (flattening of the P ( v ) curve, and increase of heat
capacity at constant volume), Schmidt and Wagner [1985] apply an equation of state to
oxygen that has two exponential terms. In fact, it is a characteristic equation that relates
residual Helmholtz energy:
(4.103)
to reduced molar density and the inverse of reduced temperature:

6 = -P
PC

and

z = - Tc
T

(4.104)

It is written as:

The equation has also been applied by Ely [1985] to carbon dioxide, by Jacobsen et al.
[1986] to nitrogen, and by Friend et al. [1989] to methane. It is derived from a more general
formula in which only the significant terms have been retained.
The equation of state in the form P(T,p) is obtained through the derivative of
Helmholtz energy as it relates to molar density and, via the derivative relating to temperature, we obtain entropy, enthalpy, and the heat capacities (see Chapter 1,Section 1.6.2).For
example:
(4.106)
As in the case of the Younglove and Ely formula, auxiliary correlations are proposed for
the calculation of the saturated phase properties.
Haar and Gallagher [1978] applied a virial equation to ammonia that had 9 coefficients
and a total number of parameters (taking into account the influence of temperature) numbering 44. An analogous formula has been applied to water vapor [Haar et al., 1984;Kestin
et al., 1984;Dobbins et al., 19881.

138

4. Equations of State

It seems, however, that no analytic equation of state is applicable in a clearly satisfactory way to the immediate surroundings of the critical point. It is therefore necessary to
use scaling factors. The IUPAC proposed formulas for carbon dioxide, for example,
[Angus et al., 19761 which combine an analytic equation and an equation specific to the
critical region. Nevertheless, they have one disadvantage; in the region where the non-analytic term cannot be overlooked, the integrations necessary for the calculation of residual
values must be accomplished numerically.
It must be pointed out that because of their precision, these formulas may allow for an
improvement of the Lee and Kesler method. If, for example, we are interested in light
hydrocarbon mixtures, we may choose methane and n-butane as the reference compounds
and calculate their properties using the Younglove and Ely formula. In this instance, precision will be higher than that obtained with Equation 3.25. Finally, like Lee and Kesler, we
will interpolate as a function of the acentric factor.

4.4

THE TAlT EQUATION

The Tait equation [Dymond and Malhotre, 19881 is not strictly speaking an equation of
state. It is rather a specific expression of the compressibility of liquids whose application is
limited to temperatures that are clearly lower than the critical temperature. It is written as:

v ( e T ) = v ( O , T ) 1 - C l n 1+ B:TJ

(4.107)

where v (0,T) represents the molar volume at zero pressure (practically the same as the
molar volume at atmospheric pressure). Parameter C is virtually independent from the
nature of the component under examination:
C = 0.0894

(4.108)

Parameter B varies with temperature, and the expression:


B = B, exp (-Bl T)

(4.109)

is often used. Finally, the value of v (0,T) may be expressed as a function of the heat expansion coefficient a:
v(0,T) = v, exp ( a T )
(4.110)
Values B,, B,, vo ,and a are specific to the substance being studied.
The Tait equation may be applied to mixtures as long as it is understood that its parameters are then composition dependent.

4. Equations of State

139

REFERENCES
Abbott MM (1989) Thirteen ways of looking at the van der Waals equation. Chem. Eng. Progress,
February, 25-37.
Alder BJ,Young DA, Mark MA (1972) Studies in molecular physics. X. Corrections to the augmented
van der Waals theory for the square well fluids. J. Chem. Phys., 56,3013-3029.
Angus S, Armstrong B, de Reuck KM (1976) Carbon Dioxide Thermodynamic Tables of the Fluid
State. IUPAC commission on thermodynamics and thermochemistry. Pergamon Press, London.
Behar E, Simonet R, Rauzy E (1985) A new non-cubic equation of state. Fluid Phase Equilibria, 21,
237-255.
Benedict M, Webb GB, Rubin LC (1940) An empirical equation for thermodynamic properties of
light hydrocarbons and their mixtures. J. Chem. Phys., 8,334.
Beret S, Prausnitz JM (1975) Perturbed hard-chain theory: an equation of state for fluids containing
small or large molecules.AZChEJ.,21,1123-1132.
Carnahan NF, Starling KE (1969) Equation of state for non-attracting rigid spheres. J. Chem. Phys.,
51,635-636.
Carnahan NF, Starling KE (1972) Intermolecular repulsions and the equation of state for fluids.
AZChEJ., 18,1184-1189.
Carrier B, Rogalski M, Peneloux A (1988) Correlation and prediction of physical properties of hydrocarbons with the modified Peng-Robinson equation of state. 1.Low and medium vapor pressures.
Ind. Eng. Chem. Res., 27,1714-1721.
Chao KC, Seader JD (1961) A general correlation of vapor-liquid equilibria in hydrocarbon mixtures.
AZChE J., 7,598-605.
Chen SS, Kreglewski A (1977) Application of the augmented van der Waals theory. I. Pure Fluids.
Ber. Bunsenges. Phys. Chem.,81,1048-1052.
Chien CH, Greenkorn RA, Chao KC (1983) Chain of rotators equation of state. AIChE J., 29,560571.
Coniglio S (1991) Estimation des propriCtCs thermodynamiques des hydrocarbures lourds. Research
Report. Laboratoire de Chimie Physique, FacultC des Sciences de Luminy.
Coniglio S, Rauzy E, Berro Ch (1993) Representation and prediction of thermal properties of hydrocarbons. Fluid Phase Equilibria, 87,53-88.
Dobbins RA, Mohammed K, Sullivan DA (1988) Pressure and density series equations of state for
steam as derived from the Haar-Gallagher-Kell formulation. J. Phys. Chem. Data, 17 (l), 1-8.
Dohrn R, Prausnitz JM (1990) A simple perturbation term for the Carnahan Starling equation of
state. Fluid Phase Equilibria, 61,53-69.
Donohue MD, Prausnitz JM (1978) Perturbed hard-chain theory for fluid mixtures: thermodynamic
properties for mixtures in natural gas and petroleum technology.AZChE J., 24,849-860.
Donohue MD, Vilmachand P (1988) The perturbed hard-chain theory. Extensions and applications.
Fluid Phase Equilibria, 40,185-211.
Dymond JH, Malhotre R (1988) The Tait equation, 100 years on. Znt. J. Thermophysics, 9 (6), 941-951.
Ely JF (1985) An equation of state model for pure CO, and CO, rich mixtures. Gas Processors
Association. Proceedings of the 65th Annual Convention, 185-192.
Friend DG, Ely JF, Ingham H (1989) Thermophysical properties of methane.J. Phys. Chem. Re$ Data,
18,583-638.
Haar L, GallagherJS (1978) Thermodynamicproperties of amm0nia.J. Phys. Chem. Re$ Data, 7,635-792.

140

4. Equations of State

Haar L, Gallagher JS, Kell GS (1984) NBS/NRC Steam Tables, Thermodynamic and Transport
Properties and Computer Programs for Vapor and Liquid States of Water. Hemisphere Press, New
York.
Harmens A, Knapp H (1980) Three-parameter cubic equation of state for normal substances. Ind.
Eng. Chem., Fundam., 19,291-294.
Heyen G (1980) A cubic equation of state with extended range of application. In: Chemical
Engineering Thermodynamics,Newman SA (Ed.), p 175,Ann Arbor Science.
Jacobsen RT, Stewart RB (1986) Thermodynamic properties of nitrogen from the freezing line to
2000 K at pressures up to 1000 MPa. J. Phys. Chem. Re$ Data, 15,735-909.
Jaeschke M, Audibert S, van Canenghem P, Humphreys AE, Jansen van Rosmalen R, Pellel Q,
Schouten JA, Michels JPJ (1991a) Simplified GERG virial expansion for field use. SPE
Production Engineering,August, 350-355.
Jaeschke M, Audibert S, van Vanenghem P, Humphreys AE, Jansen van Rosmalen R, Pellel Q,
Schouten JA, Michels JPJ (1991b) Accurate prediction of compressibility factors by the GERG
virial expansion. SPE Production Engineering,August, 343-349.
Jullian S, Barreau A, Behar E,Vidal J (1989) Application of the SBR equation of state to high molecular weight hydrocarbons. Chem. Eng. Sci., 44,1001-1004.
Kestin JV, Sengers JV, Kamgar-Parsi B, Levelt Sengers JMH (1984) Thermophysical properties of
fluid H,O.J. Phys. Chem. Re$ Data, 13 (l), 175.
Marchand B (1984) h u d e comparte, duns la rtgion critique, des tquptions dttat fondtes sur la thtorie
des sphZres dures ou sur le tenne de rtpulsion de van der Waals. Editions Technip, Paris.
Martin JJ (1979) Cubic equations of state-Which? Ind, Eng. Chem. Fundam., 18,Sl-97.
Mathias PM, Copeman ThW (1983) Extension of the Peng-Robinson equation of state to complex
mixtures: evaluation of the various forms of the local composition concept. Fluid Phase Equilibria,
13,91-108.
Patel NC, Teja A (1982) A new cubic equation of state for fluids and fluid mixtures. Chem. Engng.
Sci., 37,463-473.
Peneloux A, Rauzy E, Freze R (1982) A consistent correction for Redlich-Kwong-Soave volumes.
Fluid Phase Equilibria, 8,7-23.
Peng D-Y, Robinson DB (1976) A new two-constant equation of state. Ind. Eng. Chem. Fundam., 15,
59-64.
Proust P, Vera JH (1989) PRSV :The Stryjek-Vera modification of the Peng-Robinson equation of
state. Parameters for other pure compounds of industrial interest. Can. J. Chem. Eng., 67,170-173.
Rauzy E (1982) Les mCthodes simples de calcul des Cquilibres liquide-vapeur sous pression. Thesis,
UniversitC dAix-Marseille 11.
Redlich 0, Kwong JNS (1949) On the thermodynamics of solutions. V. An equation of state.
Fugacities of gaseous solutions. Chem. Rev., 44,233-244.
Reid RC (1983) Retrospective comments on physical properties correlations.Fluid Phase Equilibria,
13,l-14.
Rogalski M, Carrier B, Solimando R, Peneloux A (1990) Correlation and prediction of physical properties of hydrocarbons with the modified Peng-Robinson equation of state. 2. Representation of
the vapor pressures and of the molar volumes. Ind. Eng. Chem. Res., 29,659-666.
Schmidt G, Wenzel H (1980) A modified van der Waals type equation of state. Chem. Engng. Sci., 35,
1503-1512.
Schmidt R, Wagner W (1985) A new form of the equation of state for pure substances and its application to oxygen. Fluid Phase Equilibria, 19,175-200.

4. Equations of State

141

Simnick JJ, Lm HM, Chao KC (1979) The BACK equation of state and phase equilibria in pure fluids
and mixtures. Equations of state in Chemical Engineering and Research. Advances in Chemistry
series, 182,209-234.
Soave G (1972) Equilibrium constants from a modified Redlich-Kwong equation of state. Chem.
Engng. Sci., 27,1197-1203.
Solimando R (1991) fiquations d'6tat susceptibles de representer les fluides pktroliers. Sc. D. thesis,
UniversitC d'Aix-Marseille.
Starling KE (1971) Thermo data refined for LPG. Hydrocarbon Processing, 3,101.
Starling KE, Han MS (1972) Thermo data refined for LPG. Hydrocarbon Processing, 5,101.
Starling KE, Han MS (1972) Thermo data refined for LPG. Part 14, Mixtures. Hydrocarbon
Processing, May, 129-132.
Stryjek R, Vera JH (1986) An improved Peng-Robinson equation of state for pure compounds and
mixtures. Can. J. Chem. Eng., 64,323-333.
Suzuki K, Sue H, Smith RL, Inomata H, Arai K, Saito S (1989) New equation of state based on a significant structure model. Fluid Phase Equilibria, 47,17-38.
Trebble MA, Bishnoi PR (1987) Accuracy and consistency of ten cubic equations of state for polar
and non-polar compounds. Fluid Phase Equilibria, 29,465-474.
Trebble MA, Bishnoi PR (1987) Development of a new four-parameter cubic equation of state. Fluid
Phase Equilibria, 35,l-18.
Trebble MA, Bishnoi PR (1988) Thermodynamic property prediction with the Trebble Bishnoi equation of state. Fluid Phase Equilibria, 39,111-128.
van der Waals JD (1873) The equation of state for gases and liquids. In: Nobel Lectures in Physics, 1,
254-265. Elsevier,Amsterdam (1967).
Younglove BA, Ely JF (1987) Thermophysical properties of fluids. 11. Methane, Ethane, Propane,
Isobutane, and normal Butane. J. Phys. Chem. Re$ Data, 16,577-797.

Characterization of Mixtures

In the majority of cases, thermodynamics is applied to processes involving mixtures. It is


therefore appropriate to examine the extent to which the methods applied to pure substances lend themselves to systems that are characterized by conditions of temperature,
pressure (or volume), and composition. In particular, we hope that properties of mixtures
may be inferred from the properties of their components with the aid of simple and predictive rules.
The first step in this endeavor consists of defining a convention that allows for strict characterization of the various attributes that must be known. More precisely, we need to know
the contribution of each component to a particular property (volume,enthalpy, etc.), and the
potential that determines exchanges of matter. Using a defined set of definitions and
mathematical equations, this characterization will provide a core structure into which we
may introduce the various thermodynamic models to be presented in the following chapters.

So as not to lose sight of the physical reality of this structure, we will do our best to
explain it in terms of molecular interaction concepts. As a matter of fact, the mixing
process is accompanied by a redistribution of attraction and repulsion forces that are
exerted between entities of different chemical natures.
This observation alone tells us that the simple and predictive compositional laws that
we desire are, in general, only an initial approximation.
We shall therefore point out the deviations to these elementary laws or excess values.
The application of equations of state to mixtures will be discussed later on (Chapter 8).
Composition Ranges and Molar Values
Let us assume that all the components of a mixture are known.The problems encountered
with complex or multicomponent mixtures such as natural gases, crude oils, and polymers will be addressed in subsequent chapters (see Chapters 11and 12).
A mixture is now defined by the nature of its components and the quantity (number of
moles N i ) of each one of them. From them we derive the total number of moles N,, and its
composition expressed as mole fractions. This composition is generally known as x i . If in
addition we wish to determine the physical state of a homogeneous phase, we reserve the
notation xi for the liquid phase, and use yi for the vapor phase.

N, = E N i

(5.1)

144

5. Characterization of Mixtures

Any extensive property (volume V , enthalpy H , etc.) of a homogeneous phase may be


divided by the number of moles within that phase, thus defining the corresponding molar
value (v, h, etc.).Throughout this chapter we shall often use volume in our examples as it is
the most concrete property, but established equations may be easily generalized to the
other extensive properties.

5.1

PARTIAL MOLAR VALUES IN THE HOMOGENEOUS PHASE

5.1.1 Definitions. Main Equations


If the mixture in question is divided into several phases, liquid and vapor, for example, calculation of its extensive properties is the result of the sum:

v =v " + VL
Therefore we shall examine the properties of a homogeneous phase.The question is the
following: can we determine these properties from the properties of the mixture components?
Later in this chapter, we will state that the components of the mixture are in the reference state if they are pure, at the same temperature, at the same pressure, and in the same
physical state as the mixture.Thisdefinition has the advantage of simplicity,but as we shall
see, may present certain difficulties upon calculation of the properties relative to this state.
It is, however, necessary if we wish to attempt to answer the preceding question.
If we designate by the notation v;, hi*, etc. molar volume and molar enthalpy for each
component in this state, for example, can we state?
V = C N i v ; , H = z N i h f , ...

(5.3)

We could assume that each component independently contributes to the property of the
whole. This is certainly false because of the molecular interactions that occur between
compounds of a different nature, and due to the irreversibility of the mixing process that
does not take into consideration the rules of simple additivity for values such as entropy or
Gibbs energy.
It is therefore necessary to more narrowly define these contributions: they are partial
molar values. Using volume as an example, we define the partial molar volume of component i in the mixture with the equation:

This corresponds to the variation of volume of the mixture produced by the addition of
component i, if all other conditions (temperature, pressure, amounts of other components)
are kept constant.

5. Characterization of Mixtures

145

We also note that any extensive value is a homogeneous function of the first degree as
related to the number of moles. If we multiply the quantity of each component by a
constant value, the extensive values are multiplied by this same factor. So:
V(ANl, AN2, ...) = AV(N,, N2, ...)
We may therefore apply the Euler theorem to the extensive values and write:
V = CNiV,

(5.5)

This equation, applicable to any extensive value, internal energy, enthalpy, entropy,
Gibbs energy, etc., resembles the similar equation (Eq. 5.3) introduced previously as a
hypothesis (not verified) for molar properties of pure components in the reference state.
Partial molar values make up the operative component contribution for mixture properties. To do this, they depend first and foremost on the nature of the component.
However, they represent also the response to the environment when this component is
added, and are therefore related to the nature of the other components and the composition of the mixture. Of course, they are a function of temperature and pressure.
If the mixture is reduced to a pure substance, the partial molar value has as its limit the
molar property of the compound in the reference state:
if xi + 1

then

Gi+ v:

(5.6)

If on the other hand, the mole fraction of the component approaches zero, the limit
value of the partial molar value is said to be at infinite dilution:
if xi + 0

then

Gi+ vlm

(5.7)

We observe that it is generally under these conditions that the partial molar values differ the most from the properties of the component in the pure state. This fact is easily
explained by stating that in a diluted medium, the molecular environment is completely
different from that of the pure substance.
Of course, the same equations that we derived from the thermodynamic properties also
exist for their partial molar values. For example:

5.1.2

Determination of Partial Molar Values

If the property being investigated is known through a model such as an equation of state,
for instance, the application of the definition equation allows for the calculation of partial
molar values relating to it. If this property is known from experimental data, prior modeling is necessary.

146

5. Characterization of Mixtures

CT EXAMPLE 5.1
Methane propane mixture: calculation of partial molar volumes
We shall illustrate this procedure by examining the partial molar properties of
methane (1) propane (2) binary mixtures. Table 5.1 shows the volumetric properties
of this mixture at 344.15 K at a pressure of 1.377 MPa as a function of the mole fraction of methane, designated yi since the mixture is gaseous.
Table 5.1
Methane (1) propane (2) system.
Variation of molar volume with composition
T = 344.15 K, P = 1.377 MPa
Yl

v experimental

(cm3.mol -l)

v calculated
(cm3.moI-')

0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1

1748
1791
1835
1873
1910
1941
1967
1992
2016
2 035
2 047

1748
1794
1836
1874
1 909
1941
1 969
1 994
2 015
2 033
2 047

In order to model the data, we will apply the virial equation of state for pressure, truncated after the second term. Its application to a mixture requires a second virial coefficient expression as a function of composition. This expression is quadratic:
RT
v = - + B(T,y)
(5.10)

(5.11)
or for a binary mixture:
B = BlJY 1" + 2B,,,YlY,

+ B2,2Y22

It must be noted that in principle, the virial equation of state thus truncated may be
applied only at lower pressures. Here the choice was made to allow a less cumbersome calculation, but the errors introduced into the results are minor and do not
change the significance of this example.
To determine parameters B,,,, B,,2, and B,,, we apply this equation of state to the two
components, methane and propane. Their molar volumes correspond to compositions
y, = 1 and y , = 1 respectively as well as to the equimolar mixture. We find:
Bl,l = -31 cm3 * mol-'

B,,, = -93.5 cm3 . mol-'

B,,, = -330 cm3 . mol-l

Table 5.1 shows that across the entire composition range, the experimental data are
well fitted (it is understood that the values that served to determine parameters are
not to be taken into account in this comparison).

147

5. Characterization of Mixtures

Having modeled the data and in order to apply the definition equation of partial
molar values, we shall express the volume occupied by the total number of moles Nt
as a function of the number of moles of each component:

RT
V =(Nl + N2) - + ( N l + N2)
P
+2B

(A)
+ N2

N1
N2 + B
2,2 Nl
Nl + N2 Nl + N2

~~

1,2

or:
We verify that the expression obtained is definitely of the first order with respect to
the number of moles Nl and N2 and we derive:

v1=

(E),,

RT
=

2(Bl,lNl + B1,2N2) - Bl,lN? + 2B1.2NlN2 + B2,&


W l + N2I2
Nl + N2

The expression is homogeneous, of the zero order with respect to the number of
moles, and may then be expressed as:

RT
A similar expression may be written for Z2, and more generally we have:
(5.12)
Using the previously obtained numerical values for the parameters, it is possible to
calculate the partial molar volumes for methane and propane in the mixture at
344.15 K and 1.377 MPa. Table 5.2 lists these values.
Table 5.2
Methane (1) propane (2) system.
Variation of partial molar volumes with composition
T = 344.15 K, P = 1.377 MPa

0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1

2221
2 188
2 158
2 132
2110
2091
2 075
2063
2 054
2 049
2047

1748
1750
1755
1764
1776
1792
1811
1833
1859
1889
1 922

148

5. Characterization of Mixtures

Figure 5.1 illustrates the experimental data and the calculated values for the molar
volume of the mixture, as well as partial molar volumes of the components.The figure
invites commentary.
First of all, we note deviations compared to the behavior of the ideal gas as the value
of v# = RT/P or 2078 cm3 . mol-I is represented on the figure. These deviations are
larger for propane (where the reduced temperature is equal to 0.93) than for methane
(T, = 1.76).They depend, of course, on composition.
The change of molar volume with composition is not linear.The mixture is accompanied
by a dilatation and its calculation by linear combination of the molar volumes of the
two componentsyields only a first approximation. On the other hand, the mixing law that
we have applied to the second virial coefficient B takes this phenomenon into account.
Later on we shall see that we are dealing with a non-idealmixture of real gases.
On Figure 5.1, we have illustrated a graphical construction method of partial molar
values for binary mixtures.Taking a mixture of composition y , (yl = 0.7 on the graph),
we draw the tangent to curve v(y)down to the y1 abscissa point. It cuts across the axes
representing the components of the mixture (yl = 1 for methane, and y1 = 0 for
propane) on points where the ordinate is equal to the partial molar volumes Zland G2
respectively.This construction is valid for any extensive property. It need not be considered a means of achieving the most precise values for partial molar values of
course, but it visually represents their change and we will have occasion to return to it.

17

0.25

0.5

I
I

0.75

Yl

Figure 5.1 Methane (1)propane (2) system.Variation of molar volume


and partial molar volumes with composition. T = 344.15 K,P = 1.377 m a .

5. Characterizationof Mixtures

149

It is therefore easy to state the differences between the partial molar values of the two
components GI and G2 and the molar volumes of the pure components q *and 21;.
Between a partial molar value and the pure substance property there is only similarity if the mixture reduces itself to a pure substance. The other components are then
said to be at infinite dilution. It is under these conditions that the differences
between partial molar values and the properties of pure substances will be generally
the greatest, as we have already stated. This is also visible from Table 5.2 and we note,
for example, that the partial molar volume of infinitely diluted methane in propane is
greater than the molar volume of the ideal gas. In certain cases, (liquid phase mixtures
of components with a very different molar volume, for example), it may happen that
one of the partial molar volumes has a negative value.
The operation that we have performed on the volumetric data may be applied to any
extensive value. The modeling steps, the application of the definition of partial molar
values, and their derivation remain always valid. We observe that the more or less important deviations from the linear mixing laws Wiu be translated by differences between
partial molar values and the properties of the pure components in the reference state.

In fact, if we have insisted on the definition and the calculation of the partial molar values, it is not because of the immediate practical importance for most of them. We are primarily interested in the molar volume of the mixture and its enthalpy, and not in the partial molar volumes of its components or the partial molar enthalpies. However, it is the
contribution of each component to the Gibbs energy of the mixture that we need to know
in order to determine the equilibria between phases or the chemical equilibria; that is to
say partial molar Gibbs energy.The relationship between this property and the other partial molar values, however, warrants the importance placed on these values.

5.2

CHEMICAL POTENTIAL

5.2.1 Definition
Partial molar Gibbs energy is generally referred to as chemicalpotential and symbolized
(5.13)
The total differential of Gibbs energy is therefore expressed as:
dG=VdP-SdT+

pidNi
i

or, by applying the Gibbs-Helmholtz equation:


G
V
H
d - = - dP- -d T +
RT RT
R T ~

Pi
dNi
i

RT

(5.14)

150

5. Characterization of Mixtures

5.2.2 Equilibrium Condition Between Phases


The partial molar Gibbs energy is the potential that regulates the exchanges between
phases as well as chemical transformations, just as temperature and pressure are the
potentials corresponding to thermal exchanges and volume variations.
For example, if we consider a mixture divided into two phases, liquid and vapor, the
transfer of matter between the two phases satisfies the material balances relative to each
component:
dNf + dNy = 0
and at the equilibrium condition:
(1.36)

dG,, = 0

Gibbs energy, an extensive variable, is expressed from the contributions of each phase:
G=G

+ G

and therefore

dG,, = dG$,

+ dGF,

=0

The elementary variations of the Gibbs energy of each phase are related to the chemical potentials of each component in the liquid and vapor phases:

and, taking into account the equation that expresses material balances:

Since the transfers are independent (as opposed to what occurs during a chemical reaction), we conclude from this that at equilibrium the chemical potential of each component
has the same value in each phase:
pf = py
(5.15)
This condition is of course generalized in the event that the system is divided into more
than two phases: liquid, liquid, vapor and solid, for example.Thus to the condition of minimization (at constant temperature and pressure) of Gibbs energy, correspond the preceding equalities, in a number equal to the number of components. There is, however, a loss
of information because these equalities express an extremum condition, and not a minimum condition such that false equilibria may correspond to it. We will return to this
point.
The previous equations impose constraints on the intensive properties of a system in
equilibrium (temperature, pressure, composition of each phase) on which the chemical
potentials are dependent. This may be seen in the rule of phases. If we say that n is the
number of components in a system, and qis the number of phases present, then the system
variance, or the number of intensive properties that may (and must) be given to determine
the state is equal to:
V=n+2-q
(5.16)

151

5. Characterization of Mixtures

5.2.3 Relationships between the Chemical Potential and


the Other Thermodynamic Properties
For the chemical potential, it is appropriate to present a number of relationships. Firstly,
for a pure substance, it is the same as molar Gibbs energy, which we will from now on note
as $.
Moreover, we shall apply to it Equation (5.3, established using volume and valid for
any extensive value:
G = ENipi
(5.17)

5.2.3.1 The Cibbs-Duhem Equation


We rewrite the previous equation in the form:

G =ENi Pi
RT

RT

and consider the elementary variation of function G/RT:


dG = E - dP.N i +
RT
i RT

zNid

and, accounting for Equation 5.14, we obtain:


Pi
V
zNid-=-dPi
RT RT

RT

H
- dT
RT2

(5.18)

It is called the generalized Gibbs-Helmholtzequation and is most often applied at constant temperature and pressure. Therefore for one mole of a mixture:

E x i dpi = 0
i

or

Exid Pi = 0
i
RT

at constant T and P

(5.19)

The chemical potentials are not independent. Their variations with composition are
related by the previous equation, the so-called Gibbs-Duhem equation. We shall most
often apply it to values derived from the chemical potentials to be discussed later on:
fugacity, activity, and activity coefficients. When these values are determined experimentally, the Gibbs-Duhem equation allows us to apply a test of coherence to them. In certain
cases, the equation also allows to complement incomplete data.

5.2.3.2

Dependence of Chemical Potential on Pressure and Temperature

The chemical potentials of the components of a mixture depend on pressure and temperature. We can easily formulate the following equations that correspond to those for a pure
substance (Eqs. 1.45 and 1.55):
(5.20)

152

5. Characterization of Mixtures

%)pNi,Nj

(5.21)

= -si

(5.22)
If, as we emphasized earlier, values such as volume and enthalpy are rather used, in practice, for the overall properties of the mixture and not of the components,we see that in fact,
their partial molar values determine the variations of the chemical potentials with temperature and pressure, and therefore have an important role in the evaluation of equilibria.

5.2.3.3

Relationshipsbetween the Chemical Potential


and the Other Thermodynamic Functions

Chemical potential may be introduced from other thermodynamic functions. It can be


shown that:
p . = (G)
-

=()

Ni ? ; p ~ ~Ni

=()Ni

~s,N~

=()

~ S N Ni
~

(5.23)
T,~N~

Demonstration of these equations is simple.For example,from expressions5.14 and 1.33:


G=A+PV

andthus

dG=dA+PdV+VdP

and

dG=VdP-SdT+ zpidNi
i

we obtain:
dA=-PdV-SdT+ c p i d N i

(5.24)

The chemical potential is defined as the partial molar Gibbs energy, but the equalities of
5.23 do no identify it in any way with internal energy,enthalpy, or partial molar Helmholtz
energy because only the first derivation is done at constant temperature and pressure, the
definition conditions for partial molar values.
We are especially interested in the last of these:
pi=

(G)

(5.25)

ZWj

that will be applied when the mixture is represented by an equation of state that is explicit
in pressure.

5.3

FUCACITY

5.3.1 Definition
Like the Gibbs energy, the chemical potential can be calculated only from an arbitrary origin. Furthermore, it approaches --oo if the pressure or concentration approaches zero.

5. Characterization of Mixtures

153

Finally, it is a relatively abstract value. It is customarily substituted with fugacity, defined


for the components of the mixture in the same way as for pure substances (Chapter 2,
Section 2.4.3, Equation 2.35). We shall first consider only the isothermal variations of the
chemical potential:
at constant temperature: RT d In = dpi
(5.26)
and so, between two states at the same temperature, the fugacity ratios will be substituted
by the chemical potential differences:
RTAlnh=Api

(5.27)

If the properties are brought to standard state (pure substance, ideal gas state, pressure
at 0.1 MPa), then:
f.
RTln 2 = p i - p p
(5.28)

fi"

where the value of fi" is equal to 0.1 MPa.


If we wish to compare both the pure component and the component in the mixture, we
have:
f. - p i - y r
(5.29)
RTln

f; -

Finally, we note that the equality of chemical potentials at equilibrium between phases
results in that of fueacities:
"
RTln fi"
- =p/-pf=O

ff

(5.30)

5.3.2 Dependence of Fugacity on Temperature,


Pressure, and Composition
From the definition of fugacity (Eq. 5.26) and the variation of chemical potential with
pressure (Eq. 5.20), we immediately arrive at the expression:
(5.31)
In order to express the dependence of fugacity on temperature, we will use expressions
5.22 and 5.28 and note that the value of f: is independent of temperature:
(5.32)
where hp is the enthalpy of component i in the standard state.
Finally, we may apply the Gibbs-Duhem equation to fugacity:

c "i(d lnfi)T,p

=0

(5.33)

However, at this point it must be pointed out that unlike what we had done for pure
substances, only the variations of fugacity are defined. Indeed, we have not examined the
values taken by the fugacities of a mixture of ideal gases.

154

5.4

5. Characterization of Mixtures

MIXING VALUES ACTIVITY

5.4.1 Definitions
It is appropriate to return to the comparison we made when introducing this chapter. We
compare the properties of the components of a mixture on the one hand when they are
pure, under the same conditions of temperature and pressure as the mixture and in the
same physical state, this state serving as our reference state, with, on the other hand, the
actual contribution of each component to the properties of the mixture. For example, take
the difference between the value of the volume or enthalpy before mixing:

EN,.;, xNihi*,...
and its actual value. The difference is called the volume of mixing (therefore having to do
with contraction or dilatation), and mixing enthalpy (or heat of mixing):

V ' = V - ~ N i v z * , H M = H - C N i h t *,...

(5.34)

Figure 5.2 represents the volume of mixing for the methane propane system calculated
from the data provided by Table 5.1:

50

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Yl

Figure 5.2 Volume of mixing for methane (1) propane (2) system.
T = 344.15 K, P = 1.377 MPa.

The mixing values are extensive properties, and their partial molar values can be
defined:
cM=2).--2)*
1
I
I ) hM=
I
h1. - h ?1
(5.35)

155

5. Characterization of Mixtures

Of course, we similarly define the Gibbs energy of mixing, with the Gibbs energies in
the reference state denoted as p::

G M= G - C N i p f

(5.36)

and we use the expression 5.17 to write this equation in the form:

G M= C N i ( p i - p f )

(5.37)

The difference between the chemical potentials in the mixture and in the reference
state may be expressed in terms of fugacity.We have:

G M= R T E N , In fi

ff

(5.38)

The ratio of fugacities in the mixture and in the reference state is called "activity":

fi
a.= I

ff

(5.39)

such that the Gibbs energy of mixing is also expressed by the equation:

G M= R T C N i In ai

(5.40)

and the partial molar Gibbs energies of mixing are expressed directly as a function of
activity:
(5.41)

5.4.2 Dependence of Activity on Temperature,


Pressure, and Composition
Activity is temperature, pressure, and composition dependent. It is easy to set up the following equations:
(5.42)
(5.43)

As for the variation of activity with composition, in general we must be content to apply
the Gibbs-Duhem equation:
(5.44)
xi(d In a J T p= 0

2
i

and remember that activity approaches 1if the mole fraction approaches 1.
If we consider a mixture whose components are, if not indiscernible or without molecular interaction, but at least similar from the point of view of chemical structure and molecular volume, such as mixtures of linear C&, paraffins, or mixtures of benzene, toluene,
xylenes, we may, at least to a first approximation, predict that the mixing process involves

156

5. Characterization of Mixtures

neither a large variation in volume nor appreciable heat. Since the intermolecular interaction energies are similar for the pure components, their redistribution upon mixing
causes no practical internal energy modification of this mixture. Effectively, this is what is
observed experimentally.
Under these conditions, and taking into account the preceding equations that describe
the variations of activity with pressure or temperature, we may also predict, at least initially, that this value is solely dependent upon composition.The ideal solution is thus introduced, if not defined.

5.5

THE IDEAL SOLUTION

Lewis [a9231 defined the ideal solution using the condition of proportionality between
fugacity in a mixture and the mole fraction:
f

at given T and P, -= const.


Xi

The fugacity of the mixture is reduced to that of the pure substance if the mole fraction
approaches unity. We therefore have:
(5.45)
f iid = f t?xi
It is of course the same to state that in an ideal solution, activity is equal to the mole
fraction:
.;d = xi
(5.46)
Under these conditions, as activity is independent of temperature and pressure, the volume of mixing and the enthalpy of mixing are zero:
2,M,id

=0

and

hM,id = 0

(5.47)

This is not true for entropy, Helmholtz energy, or Gibbs energy. Applying Equations 5.40 and 5.46, for the Gibbs energy we have:
G M , ~=

R T C NIn~xi

(5.48)

It is possible to derive therefrom the Helmholtz energy of mixing:


AM$

= G M k J - p V M M = G M , i d = R T ~ In
N xi~

(5.49)

as well as the entropy of mixing:


SM,id =

H M,id - G M,id
T

-REN~In xi

(5.50)

These terms convey the irreversibility of the mixing process that always involves an
increase in entropy and a decrease in Gibbs energy.
There is of course a particular case of a solution that is perfectly logical to view as ideal.
It is a mixture of ideal gases since the molecular interactions are zero within such a mixture. We are therefore in a position to specify the value for fugacity in an (ideal) mixture of

5. Characterization of Mixtures

157

ideal gases. The reference fugacity, of the pure substance that is, is therefore equal to the
pressure. Expressing the mole fraction as yi, we have:

fi" = Pyi

(5.51)

This equation allows us to finalize the definition of fugacity.

5.6

CALCULATION OF FUGACITY

To determine fugacity for a mixture component,we use the expressions:


RTd lnfi = dpi

(5.26)

and:

(5.20)

that both express the variations of the chemical potential with pressure. In addition,considering that the ideal gas laws apply to a real fluid when density approaches zero, we write:
lim ( f i ) p + o = PYi
(5.52)
often written as:
(5.53)
lim ( f i ) p + o = PYi
as it is understood that any fluid occupies an infinite volume as pressure approaches 0.
The calculation of the fugacity of a mixture is therefore based, according to these equations, on the knowledge af the partial molar volumes:
RTln

PYi

= fop(Zi-

F)

dP

(5.54)

The ratio:

q.= fi

(5.55)
PYi
is the fugacity coefficient of the component of a mixture.
Application of Equation 5.54 presents a dilemma if the model that we are using does
not allow for specific calculation of partial molar volumes as a function of temperature,
pressure, and composition. Such is often the case when we use an equation of state that is
applicable to the liquid and vapor phases. We shall address this problem in Chapter 8
(Section 8.3.2),which deals with equations of state applied to mixtures.

'

# EXAMPLE5.2
Methane propane mixture: calculation of fugacity in a mixture
To illustrate these definitions, we return to the methane-propane mixture that we
have already studied.We stated that the equation of state:
RT
(5.10)
v = - + B(?;y)
P

158

5. Characterization of Mixtures

may be applied to it, and we confirmed that the usual quadratic expression of the second virial coefficient:
(5.11)
represented the experimental data with acceptable accuracy. From it, we derived the
value of partial molar volume:
(5.12)
Taking into account the method for the calculation of the fugacity that we just presented (Eq. 5.54), we have:
RT In
PYi

= fop(Zi

F)

(5.56)

d P = ( 2 c BUyi - B ) P
i

and may calculate the values for fugacity of the two components as well as their
fugacity coefficients using the previously determined numerical values for coefficients Bi,j.Table 5.3 provides the results of this calculation.
The fugacities are to be compared with the partial pressures. The difference, which is
specified by the value of the fugacity coefficient,results from the fact that the system
is neither an ideal mixture, nor an ideal gas.
Finally we note that these same results yield the values of two specific values, the
fugacities of the two components in the reference state that coincide with the fugacities in mixture when the mole fraction of the component is equal to 1 (with the other
mole fraction equaling zero):
fi* = 1.357 MPa,

f;

= 1.175 MPa

Table 5.3
Methane (1) propane (2) system. T = 344.15 K , P = 1.377 MPa.
Fugacity coefficients and fugacity
Yl
~

0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1

1.071
1.054
1.039
1.026
1.015
1.006
0.998
0.993
0.988
0.986
0.985

0.853
0.854
0.856
0.860
0.865
0.871
0.879
0.888
0.900
0.913
0.928

0
0.145
0.286
0.424
0.559
0.693
0.825
0.957
1.089
1.222
1.357

~ ~ _ _ _ _ _

1.175
1.058
0.943
0.829
0.714
0.600
0.484
0.367
0.248
0.126
0

159

5. Characterization of Mixtures

5.7

EXCESS VALUES AND ACTIVITY COEFFICIENTS

5.7.1 Definitions
When studying pure substances, we emphasized the fact that the ideal gas state could
provide an acceptable approximation of the real fluid at low density, and most of all, a
calculation intermediate. Similarly, with mixtures, the concept of the ideal solution makes
up only one step in the calculation of the properties of a mixture. It would not be unreasonable to confine ourselves to this one step for a property such as volume. It is an approximation that could be risky for enthalpy. It is often unacceptable for the calculation of
phase equilibria, namely chemical potential or fugacity. The terms describing the difference between the properties of a real mixture and those of an ideal mixture are called, if
we are talking about extensive properties, excess values. As such, we have the definitions for excess volume, excess enthalpy, excess entropy, and excess Gibbs energy. We
state that:
VE=V-Vid

HE=H-Hid

SE=S-Sid

GE=G-Gid

(5.57)

For these extensive properties we have of course the same equations as for the thermodynamic properties from which they are derived, for example:
GE

= H E -TSE

G ~ = A ~ + P v ~
etc.
They have partial molar values, in particular, excess chemical potential:

(5.58)

(5.59)

(5.60)

In the case of volume, internal energy, and enthalpy, the excess value and the mixing
value coincide. Indeed, the mixing value of the ideal solution is zero. We can indiscriminately state:
V=XNiw;+VM
(5.61)

v = z N i w ; + VE

or:

(5.62)

The same is not true for entropy or Gibbs energy:


G =XNip;+GM

but:

+ z N i R T In xi + G

(5.63)

S = z N i s ; - z N i R lnxi + S E

(5.64)

G=

and similarly:

and

A = E N i @ ; - Pw,*) + z N i R T In xi + A

(5.65)

160

5. Characterization of Mixtures

5.7.2 Dependence of Excess Values on Temperature,


Pressure, and Composition
The excess values are of course dependent upon temperature, pressure, and composition.
For excess Gibbs energy, for example, we have:
(5.66)

(5.67)

(5.68)
Analogous equations may be set up for corresponding partial molar values. So:

(5.69)

(5.70)
As for the influence of composition, we may apply the Gibbs-Duhem equation:
xidp? = 0

at constant temperature and pressure

(5.71)

5.7.3 Activity Coefficients


For the Gibbs energy of mixing, we have defined the activities (Eq. 5.39). We shall do the
same for excess Gibbs energy and introduce the activity coefficients;
(5.72)
or:

f .1 = f

*yx.
1 1

(5.73)

These activity coefficients are related to the chemical potentials and to the excess Gibbs
energy by the following equations:

pi= pl*+ RT In xi + RT In yi

(5.74)
(5.75)

161

5. Characterization of Mixtures

GE=ENiRTlnx

(5.76)

Therefore, we may express the activity coefficients using an excess Gibbs energy model
(Eq. 5.75) and conversely, the excess Gibbs energy from the activity coefficients (Eq. 5.76).

5.7.4 Dependence of Activity Coefficients on Temperature,


Pressure, and Composition
The activity coefficients are related to the excess chemical potentials (Eq. 5.73). We therefore apply Equations 5.69 through 5.71 to them and define the influence of pressure and
temperature on this property:
(5.77)

(5.78)

and:

The Gibbs-Duhem equation is most commonly applied to the activity coefficients:


xid In yi= 0

at constant temperature and pressure

(5.79)

We must note here that, taking into account the low values of excess volumes, the deviations from ideality in the liquid phase are generally minor and may most often be neglected.

EXAMPLE5.3

Methane propane mixture: calculation of activity coefficients

By applying the definition equation, we can calculate the activity coefficients for
methane and propane in a mixture, as well as the excess Gibbs energy.The results of
these calculations are shown in Table 5.4.
The activities are to be compared to the mole fractions, and the deviations from the
ideal behavior are specified by the activity coefficient values and excess Gibbs
energy. Note that these deviations are small, which is normal, as the mixture is in the
gas state and has low density.
If we consider the nature of the model selected to represent the variation of molar
volume of the mixture with the composition, it is not surprising that the deviations
from ideality are symmetric.
i = 1,2
Y i c v i ) = Yi(1 -Yi),

162

5. Characterization of Mixtures

Table 5.4
Methane (1) propane (2) system: T = 344.15 K, P = 1.377 MPa.
Activity, Gibbs energy of mixing, activity coefficients,and excess Gibbs energy
a2

Y1

0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1

1
1.001
1.003
1.008
1.013
1.021
1.031
1.042
1.055
1.070
1.087

1.087
1.070
1.055
1.042
1.031
1.021
1.013
1.008
1.003
1.001
1

0
-909
-1 394
-1 698
-1 868
-1 923
-1 868
-1 698
-1 394
-909
0

1
0.901
0.803
0.705
0.608
0.511
0.412
0.313
0.211
0.107
0

0
0.107
0.211
0.313
0.412
0.511
0.608
0.705
0.803
0.901

0
22
38
50
58
60
58
50
38
22
0

Figure 5.3 shows the change, with the composition, of the excess Gibbs energy and
excess chemical potentials in the methane propane system. The latter correspond to
the excess Gibbs energy using the graphical method of drawing a tangent to the curve
gE ( x ) , which we have already used for partial molar volumes.

250 I

0.2

0.4

0.6

0.8

Y1

Figure 5.3 Methane (1) propane (2) mixture. Calculation of Gibbs


energy and chemical potentials. T = 344.15 K, P = 1.377 MPa.

5. Characterization of Mixtures

163

Relationships befween the Fugacity Coefficients


The fugacity coefficients provide a convenient indication of the difference that exists
between the chemical potentials of a component and the result of the calculation performed in the ideal gas state.
For a pure compound, we write;
R T l n p : = l o (vf-,)dP
RT

(2.39)

where v;is the partial molar volume of the pure component.


For a mixture component, the equation becomes;
RTln 2
f. =
PYi

lo -)

(Gi- RT d p

(5.54)

where Gi is the partial molar volume of component i in the mixture .


One might also define the fugacity coefficient of the mixture:
RTln cp, = RTln

$ = j o p ( v -F)d p

(5.80)

with v as the molar volume of the mixture. The fugacity coefficients defined in this way are
of course different, but related by a number of equations. In particular:
Pi
y= -

(5.81)

(5.82)
(5.83)

5.8

COMPARISON O F T W O METHODS FOR


CALCULATl NG FUG AClTY

A close examination of the preceding equations allows us to note that multiple approaches
exist for the same property. So in the case of chemical potential or fugacity, we may adopt
one or the other of the two methods that follow:
0 Apply to the mixture an equation of state that takes into account the influence of
composition using the mixing laws. This is the path chosen in the following calculations:
Equations 5.10,5.11 and Table 5.1
- mixture volume
Equation 5.12 and Table 5.2
- partial molar volumes
Equation 5.56 and Table 5.3
- fugacities in mixture
Equation 5.39 and Table 5.4
- activities

164

5. Characterization of Mixtures

- activity coefficients
- excess Gibbs energy

Equation 5.72 and Table 5.4


Equation 5.76 and Table 5.4
Apply to the components of a mixture an equation of state without considering what
the influence of composition will be on the variables of this equation. The method is
complemented by the use of an excess Gibbs energy model. In order to ensure coherence with the first approach, the equation selected for the equation of state will be:

* = RT
-+ BJT)

'D;

(4.19)

and for excess Gibbs energy:

(5.84)
By applying Equation 5.75, we would have derived the expressions for activity coefficients:

Therefore the progression encompasses the following steps and equations:


reference fugacities
Equation 2.39
Equation 5.84and Table 5.4
excess Gibbs energy
Equation 5.75 and Table 5.4
activity coefficients
Equation 5.72 and Table 5.4
activities
Equation 5.39 and Table 5.3
mixture fugacities
Equation 5.31 and Table 5.2
partial molar volumes
Equation 5.5 and Table 5.1
volume of mixture
Of course, since the coherence between the two approaches is ensured by the choice of
the excess Gibbs energy model, the results would have been identical even though they
were obtained in a different order.
When we look at liquid vapor equilibria, we will have the opportunity to apply the two
approaches to the calculation of fugacity. The first approach will essentially be concerned
with equilibria at high pressure in systems having components with high vapor pressure or
supercritical components.The second will be used at low pressure to calculate the hgacity
of components of a liquid mixture. Bringing the two methods together will finally allow for
expansion of the range of application of equations of state.
The following table gathers the expressions generally applied as a function of the phase
and the method applied.
-

Vapor Phase

Liquid Phase

First method

fi" = Py,(pi"

ff = Pyiqi"

Second method

fi" = pq;vyiyi"

fi" = pqpqy,yi"

5. Characterization of Mixtures

165

We note that the second method is never applied to the vapor phase. The approaches
will therefore be characterized by the calculation method applied to the liquid phase. Of
course,what we are stating here for the calculation of fugacity is equally valid for the other
thermodynamic properties.

5.9

ASYMMETRIC CONVENTION: THE HENRY CONSTANT

Throughout the previous discussion, the mixture properties were related to what is
conventionally called symmetric reference states. In these reference states, written as f;*
for fugacity,for example,each component was pure, and under the same conditions of temperature and pressure as the mixture, and in the same physical state.
It may happen that this state cannot be attained, and therefore we cannot arrive at the
properties experimentally.Such is the case for electrolyte solutions, for which we cannot
conceive ions in the pure state. It is also the case for solutions containing permanent
gases whose critical temperature is less than the temperature in question. A reference
state in which these permanent gases will be in the liquid state and pure, may, strictly
speaking, be possible, but not attainable.
For such systems, it is easy to distinguish solventsfrom solutes. For the solvent or
solvents,we retain the reference state that we have adopted up to this point, and we have:
(5.73)
For the solute or solutes, we most often choose the infinitely diluted state. Under these
conditions,we define the Henry constantas the reference fugacity using the equation:
(5.86)
At a finite concentration we use:

f;=%y;x,

(5.87)

This equation defines yi as the activity coefficient for the asymmetric convention.
As opposed to the reference fugacity defined in symmetric convention, the Henry
constant depends not only on the temperature and pressure, but also on the nature of the
solvent or solvents.
Bringing together Equations 5.73 and 5.87 allows us to relate the activity coefficients
defined in symmetric and asymmetric convention:
f .*L

y!=y
I

(5.88)

and express the Henry constant as a function of the reference fugacity and the activity
coefficient at infinite dilution in symmetric convention:

% = x,,,fI*L

(5.89)

166

5. Characterization of Mixtures

Note that we apply this formulation mainly for the solubility of gases in the liquid
phase.
We may also generalize this asymmetric convention in defining a reference state for the
solute, where it has a finite concentration. For example, we may use molality (number of
moles per kilogram of solvent) equal to unity. We come across this convention in the study
of electrolytes.

REFERENCE
Lewis GN, Randall M (1923) Thermodynamics and the Free Energy of Chemical Substances.
McGraw-Hill, New York.

Mixtures: Liquid-Vapor Equilibria

Having defined the principal thermodynamic properties at work in the characterization of


a mixture (partial molar quantities, activity coefficients, fugacity, etc.), it is reasonable to
first introduce the models that are commonly adopted for the prediction of these properties. These models are the models for excess Gibbs energy, and the mixing laws applied to
the equations of state. However, these models have, most often, been developed to represent phase equilibria. It is therefore appropriate to first describe the two-phase behavior of
mixtures. Priority is given here to liquid-vapor equilibria with the understanding that the
phenomena of liquid-liquid demixing, crystallization, and multi-phase equilibria will be
discussed further on.
Unlike pure substances, according to conditions of temperature and pressure a mixture
may divide itself into several phases with different compositions. Liquid-vapor equilibria
occur upon exploitation of a gas or petroleum reservoir. They allow for the separation of
mixture components via distillation. The latent heat accompanying the phase change is put
to good use during energy transfers. Familiarity with these equilibria is therefore a necessary condition for the simulation and optimization of chemical engineering operations.
This knowledge is based above all on experimental data. Predictive methods exist, but
they are supported by results of measurement for the determination of parameters that
are employed and govern their validity. We must emphasize this point. However, we will
not develop the experimental techniques that must often be adapted to each case and take
into account the nature of the mixture, and the ranges of pressure and temperature. For
further information, we may consult the works of Abbott [1986], Marsh [1989], Deiters
[1986], and Holste [1986].
Similarly, the databases,bibliographic or numeric, that collect the results of these measurements need to be known. The most important was begun by the University of
Dortmund [Gmehling et al., 1978-19841, but it is also worthwhile to cite that of Kehiaian
[1973-19931,which covers all the deviations from the ideal behavior, and provides data that
has been carefully monitored. Finally, the bibliographic database of Wichterle et al. [19731985,19931is close to exhaustive.
We will describe the phenomena of vaporization (or condensation), discuss the general
principles for the calculation of liquid-vapor equilibria as well as the principal problems
that arise, and take a look at the various forms that the equilibrium equation may assume
according to the nature of the mixture and the temperature and pressure conditions.

168

6.1

6. Mixtures: Liquid-Vapor Equilibria

DESCRIPTION OF THE VAPORIZATION OR


CONDENSATION PHENOMENA

The description of a mixture requires knowledge of pressure, temperature, and composition. In order to graphically illustrate the phase changes, we shall first consider binary mixtures evolving at constant pressure, and then at constant temperature. Figures 6.1 and 6.3,
and Table 6.1 refer to the propane (1) n-pentane (2) mixture. Equilibria have been calculated using the Soave-Redlich-Kwong method to be discussed in Chapter 8. Taking into
account the nature of the system, we can assume that the results of the calculation are very
close to reality.
Table 6.1
Liquid-vapor equilibrium of a propane (1) n-pentane (2) mixture;
at (p = 5 bar) isobars and (T = 300 K) isotherms
T=300K

P=5bar

O.OO0
0.050
0.100
0.150
0.200
0.250
0.300
0.350
0.400
0.450
0.500
0.550
0.600
0.650
0.700
0.750
0.800
0.850
0.900
0.950
1.OO0

%bubble

dew

(C)
92.018
81.909
72.650
64.302
56.849
50.223
44.333
39.086
34.392
30.174
26.363
22.902
19.743
16.844
14.173
11.699
9.399
7.251
5.236
3.338
1.543

(bar)
0.717
0.752
0.791
0.835
0.884
0.938
1.000
1.071
1.152
1.246
1.357
1.490
1.651
1.851
2.106
2.440
2.898
3.563
4.605
6.425
9.987

92.018
89.987
87.888
85.713
83.456
81.106
78.652
76.081
73.376
70.518
67.482
64.237
60.743
56.946
52.770
48.110
42.801
36.570
28.908
18.644
1.543

0.717
1.125
1.537
1.954
2.375
2.801
3.232
3.668
4.109
4.556
5.009
5.467
5.933
6.405
6.885
7.373
7.870
8.378
8.899
9.434
9.987

6.1.1 Isobaric liquid-Vapor Equilibrium Diagram


First, we shall consider the example of the propane (1) n-pentane (2) mixture at a pressure
of 5 bar.
Figure 6.1 indicates the compositions on the abscissa, namely the mole fraction of one
of the components with z1 the entire mixture, x1 the liquid phase, and y , the vapor phase.

169

6. Mixtures: Liquid-Vapor Equilibria

0.25

0.5

0.75

Mole fraction of propane

Figure 6.1 Isobaric liquid-vaporequilibrium diagram (p = 5 bar) of


the propane (1) n-pentane (2) system.

Since the mixture is binary, the sum of the mole fractions of both components must be one.
Hence, the mole fractions of the second component are readily derived from either z,, x,,
or y,. As a general rule, index 1will be attributed to the most volatile component,with the
lowest boiling point, or the highest vapor pressure.
On the ordinate, we have the temperature. Specifically on the abscissa axes 1and 0, we
have the boiling temperatures of the first and second component respectively, 1.5"C for
propane and 92C for n-pentane, at the pressure under consideration.
If we look at a mixture with overall composition z , (zl = 0.5 for example), this mixture
is a homogeneous liquid at low temperature. If, at constant pressure,we provide heat to the
system,we first cause the temperature of the liquid to rise, and reach the "bubble point" Tb
(at 26.4"C), where a first bubble of gas appears.At this point, the liquid phase evidently has
the same composition as the entire mixture Xi,bubble = zi,but the vapor phase appears at a
different composition Yi,bubble. The vapor phase is generally richer in volatile component.
In this example,y, = 0.915. We say that the system is at its bubble point.
If we continue adding heat, we generally observe that temperature increases, and that
the vapor phase simultaneously develops at the expense of the liquid phase. The system
splits itself into two phases whose compositionsx, and y, differ from that of the mixture as
a whole. These phases are related by the equilibrium conditions of equality of chemical
potentials, pressure, and temperature. In addition, the compositions of these phases, the

170

6. Mixtures: Liquid-Vapor Equilibria

number of moles in the liquid phase N L ,in the vapor phase N, and for each component in
the total system N , are related by material balances:
N , = NLxi + Ny,

(6.1)

For example, at 40C, x1 = 0.34,y , = 0.83, and the vaporized fraction is equal to:
N
N L + N

z1 - x1 - 0.5 - 0.34

y1 -XI

0.83 - 0.34

= 0.33

In general, vaporization of a mixture progresses up to the dew point.The mixture still


verifies the liquid-vapor equilibrium conditions, but the liquid phase disappears. The coorof the liquid phase,
dinates of this point are the dew temperature, the composition
and Yi,dew = zi of the vapor phase. In the case of the equimolar mixture of propane, with
n-pentane at a pressure of 5 bar, the dew temperature is 67.5Cand we have x1 = 0.13.
If we continue to apply heat to the system, it becomes homogeneous and we say that we
are dealing with a superheated vapor.
The description of the vaporization processes may be repeated for any composition of
this binary mixture. The bubble points describe the bubble curve and the dew points the
dew curve. These two curves that delimit the two-phase range meet on the axes representing the pure substances as the bubble and dew temperatures coincide (monovariant
system). Under the bubble curve is the range of the homogeneous liquid phase, termed
subcooled, and above the dew curve the zone of the superheated vapor phase. Inside
the lens, the system is in a partial vaporization state. It is neither at its bubble point nor at
its dew point, and it is separated into two phases: the liquid phase is at its bubble point, and
the vapor phase at its dew point.
Of course, the bubble and dew temperatures depend on the pressure. We shall examine
this dependence using isothermal equilibrium diagrams.
Ternary Systems
It is possible to show the composition of a ternary system using a triangular diagram. The
summits represent the components of the system, the three sides the binary systems
formed by two of the three components, and the interior of the triangle the actual ternary
mixtures. In Figure 6.2, perpendicular to the plane of the triangle, we have the temperature
bordering a prism with the three lens shaped areas of binary equilibrium drawn over its
surface at a given pressure. Inside the prism, we show the evolution of bubble or dew temperatures of the ternary mixture, thus defining the corresponding surfaces.
The intersection of the bubble and dew surfaces with an isotherm plane determines two
curves along which the points representing the mixtures at equilibrium correspond by
pairs.
A point situated between the two surfaces represents a two-phase system while the
points representing the two phases in equilibrium are found on the bubble and dew surfaces, and on the two curves corresponding to the temperature investigated. They are
aligned with the point that represents the entire mixture due to linear constraints imposed
by the material balances.

171

6. Mixtures: Liquid-Vapor Equilibria

+ plane
e=e,
B

Figure 6.2 Isobaric liquid-vapor equilibrium diagram of a ternary mixture.

6.1.2 Isothermal liquid-Vapor Equilibrium Diagrams


Evolution with Temperature
Critical Point and Retrograde Condensation
The graph of the isothermal liquid-vapor equilibrium diagram for a binary mixture is similar to the one that we just made for the isobar diagram.
Figure 6.3 relates to the propane n-pentane mixture at 300 K. The pressures are drawn
on the ordinates and on the axes representing the pure substances; we have the vapor pressures for the two components of the system, 10 bar for propane, and 0.7 bar for n-pentane.
At low density, the mixture is in the vapor state and pressure is low. A volume decrease
causes subsequently a pressure increase, the appearance of a first drop of liquid at P = 1.36
bar (we are now at the dew pressure), then the partial condensation of the mixture that
will split into two phases, liquid and vapor, finally the disappearance of the last bubble of
vapor at P = 5 bar (bubble pressure), and the compression of a homogeneous liquid phase.

1 72

6. Mixtures: Liquid-Vapor Equilibria

"
1
10

$ 8

e
p!
v)
3

g 6

0.25

0.5

0.75

Mole fraction of propane

Figure 6.3 Isothermal liquid-vapor equilibrium diagram (T = 300 K) for


the propane (1)n-pentane (2) system.

On the isothermal diagram, we find the equilibrium lens, but the positions of the liquid and
vapor areas and the bubble and dew curves are inverted in comparison to those observed
on the isobaric diagram.
It is interesting to examine the change of such a diagram with temperature, and
Figures 6.4 and 6.5 allow us to do just that. They pertain to the ethane benzene system and
the equilibria were calculated using the Soave-Redlich-Kwongmethod, as was done in the
previous example.For example, on this diagram we can follow the change of state of a mixture with overall composition z , = 0.7.
At low temperature, (10 and 25"C, Fig. 6.4) the equilibrium lenses have the shape
described previously. We note that the dew curve is for the most part close to either the
P = 0 or the y, = 1axes.This shape is due to the large volatility difference between the two
components. For the mixture investigated,the bubble (P = 23 and 31 bar respectively) and
dew pressures increase with temperature.
If the temperature is greater than the critical temperature of one of the components
(Fig 6.5),for example 90C, as the critical temperature of ethane is 305 K, the lens extends
only over a portion of the full composition range. Mixture z1 = 0.7 still has a dew point at
this temperature (4.7 bar) as well as a bubble point (P= 76.9 bar). The bubble and dew
curves no longer meet at the zl = 1 axis, but at a point corresponding to the maximum of

6. Mixtures: Liquid-Vapor Equilibria

173

40 -

z
e
h

P3 30 2
P
a
20 -

10 -

0-

1
0
0.25
0.5
0.75
1
Mole fraction of ethane

Figure 6.4 Isothermal liquid-vapor equilibrium diagram ( 8 = 10 and


25C) for the ethane (1)benzene (2) system.

pressure. At this point, the liquid and vapor phases at equilibrium are identical in composition and in their properties (density, enthalpy, entropy, etc.). This is the critical point of
liquid-vapor equilibrium at the given temperature.
At 155C,the critical point corresponds to the mixture z1 = 0.7. The physical change of
state of this mixture as pressure increases will show subsequently a homogeneous vapor
phase, a dew point (26.4 bar), and a two-phase mixture. Yet, we will not observe a bubble
point, and when the critical pressure (108 bar) is reached, the liquid-vapor interface disappears, and we observe the critical opalescence phenomenon.At high pressures, the mixture will once again be homogeneous.
At 175C,the composition of the critical point of ethane is less than that of the mixture
investigated. For this mixture, the increase in pressure is still evidenced by an initial dew
point (P = 44.5 bar), by separation into two phases, but one observes that the liquid phase
proportion goes through a maximum and it is through a second dew point where, at high
pressure (P = 106.7 bar), we leave the equilibrium lens to enter into the homogeneous
zone. The mixture has no bubble point.
Finally, at 210 and 235C,still for the z1 = 0.7 mixture, there is no longer either a dew or
bubble point, and the mixture is homogeneous regardless of pressure.

174

6. Mixtures: Liquid-Vapor Equilibria

15

12.5

rn
n

10

2
s

2 7.5

2.5

0
0

0.25
0.5
0.75
Mole fraction of ethane

Figure 6.5 Isothermal liquid-vapor equilibrium diagram ( 8 = 60,90,


155,175,210,and 235C) for the ethane (1) benzene (2) system.

It is possible to depict the change of the bubble and dew pressures as a function of temperature for a mixture of fixed composition. This is what is done on Figure 6.6 for the
example discussed here. We notice the meeting of the bubble and dew curves at the critical
point (the slope dPldT for both curves is identical in this point), the existence of a temperature extremum (cricondentherm) and a pressure extremum (cricondenbar, very close to
the critical point in the present example). Between the critical temperature and the cricondentherm, for example at 448 K (175C),the mixture being investigated has two dew pressures, and its evolution with decreasing pressures from the high pressure range where it is
homogeneous first shows a high dew point, then a liquid phase appears, a phenomenon
called retrograde condensation, then a maximum deposit of liquid phase, and finally the
low dew point. Bubble and dew points delimit the two-phase region inside which this
mixture splits into two phases, and outside which it is homogeneous.
In this same diagram, we have drawn the two-phase region of another mixture (zl = 0.2)
for which the cricondenbar (T = 533.9 K and P = 68.2 bar) is clearly distinct from the critical point (T = 540.6 K and P = 67.8 bar). In the pressure interval defined in this way, a temperature variation forms two bubble points.

6. Mixtures: Liquid-Vapor Equilibria

175

100

75
h

2
n

?!

; 50
v)

25

0
300

400

500

600

Temperature (K)

Figure 6.6 Ethane benzene system: two-phase envelope for two


distinct compositions (solid line: z1= 0.7;dotted line: z1 = 0.2)
and locus of critical points.

All of the critical points together determine the critical points locus. All the twophase envelopes fall within this line that extends from the critical point of the light component to the critical point of the heavy component. Beyond the range defined by the two
vapor pressure curves and the critical point locus, the system is homogeneous regardless of
its composition.
Depending on the composition of the mixture, the cricondenbar may be found on the
bubble or dew curve and we may observe two bubble or dew temperatures for pressures
between the critical pressure and the cricondenbar. In this last case, we are talking about
second type retrograde condensation.
The qualitative description that we just gave relates to a binary mixture. For a multicomponentcomplex mixture,the graph in the temperature pressure diagram remains valid
and we can draw the two-phase envelope of a mixture of fixed composition. Retrograde
condensation phenomena may be observed. Such is particularly the case for natural gases
where the liquid deposit caused by a decrease in pressure is observed during the course of
the exploitation of the reservoir (decrease in pressure due to the extraction of gas), in the
well (pressure drop during ascent), and in the separation installations at the well head.
It is important to note that the critical points locus may extend over a very large pressure range. For example, this is the case if the mixture contains hydrocarbons of very different molecular weights. So the critical pressure of the mixture that we were looking at
was 108 bar, and for the mixtures of methane with high molecular weight paraffins,such as
hexadecane for example, the pressure reaches several hundreds of bar. In particular, it is
clear that the critical coordinates of a mixture have nothing to do with its pseudocritical
coordinates. The first indicate a particular state of the mixture, often experimentally

176

6. Mixtures: Liquid-Vapor Equilibria

observable, and the second are merely parameters related to an extension of the corresponding states principle, and without physical reality.
Finally, we emphasize that the graph we have drawn corresponds to the simplest case.
According to the nature of the components and the appearance of liquid-liquid equilibria,
the two-phase envelope, and the critical points locus may have more complex forms. A
mixture may have no critical point, or may have several of them [see Rowlinson and
Swinton, 1982,p. 191,and Kreglewski, 1984,p. 1331.

6.1.3 Azeotropic Systems


Up to this point we have considered it normal that for a binary system the bubble and dew
temperatures at constant pressure, or the bubble and dew pressures at constant temperature are monotone functions of the composition. This is in no way a steadfast rule.
Suppose we have two components having the same vapor pressure at a given temperature.
Either, for example, the bubble curve will present an extremum or it will be isobaric. In this
last case, the bubble pressure would be independent of composition, and the system, at
least at this temperature, would behave like a pure substance from the point of view of
liquid-vapor equilibria. It is of course the first hypothesis that must be retained: the bubble curve has an extremum. As for the hypothesis that two components have the same
vapor pressure, it is in no way improbable, and examples of mixtures whose components
have close vapor pressures are plentiful. Such is the case for the cyclohexane benzene
system that has a maximum pressure azeotrope, or even the hexafluorobenzene benzene
system with two azeotropic compositions, one at maximum, and the other at minimum
pressure.
The appearance of an extremum of the bubble curve is not specific to such systems.
Deviations from ideality defined in the previous chapter may cause it as well. For example,
such is the case for heptane dimethylformamide mixtures rich in heptane, even though the
difference in boiling temperature is close to 50C. We can also cite the carbon dioxide
ethane system, where the components have vapor pressures differing by 50% at 10C.
In summary, this phenomenon may be caused by two types of behaviors: neighboring
volatility of mixture components, and/or deviations from ideality.
We show that (Section 6.3.1) in this case, if the bubble curve has an extremum, so has
the dew curve, at the same composition. At the extremum, bubble and dew pressures have
the same value, and liquid and vapor phases have the same composition. The liquid-vapor
equilibrium is therefore not selective; we say that there is an azeotrope. Figure 6.7 illustrates this phenomenon. We note that for the compositions in component 1 (the most
volatile), less than the azeotropic composition, the vapor phase is richer in this component
than the liquid phase, in agreement with what was observed in the normal case. On the
other hand, beyond the azeotropic composition, the opposite is true and selectivity is
reversed. Distillation of such a mixture that, in this case, has a bubble and dew pressure
maximum, will yield the azeotropic mixture at the top of the column, and not the most
volatile component. At the bottom of the column, we have one or the other of the two
components, according to the composition of the original mixture. The azeotropic phenomenon therefore constitutes a limit to the distillation separation processes.

177

6. Mixtures: Liquid-Vapor Equilibria

70

65

g 0.9

6
60
P3

/
/
/

v)
3

d
F

0.8

0.5

50

0.7

0.6

55

45

1
0.0

'

'

40

0.25

0.5

0.75

0.25

0.5

0.75

'

Mole fraction of hexane

Mole fraction of hexane

Figure 6.7 Azeotropic liquid-vapor equilibrium diagram of the hexane (1)


acetone (2) system at 50C (left,A) or at atmospheric pressure (right, B).

A mixture having the azeotrope composition therefore vaporizes at a given temperature, like a pure substance. Pressure remains constant during vaporization. However, if we
examine this same diagram at another temperature, we will observe that the azeotropic
composition has changed.This change differentiates the azeotropic mixture from the pure
substance, and allows us to circumvent the previously discussed separation limit. The two
components of a system having an azeotrope may be separated via two distillations working at different pressures. Figure 6.8 [Rowlinson and Swinton, 1982,page 1081 collects the
data obtained by several authors, and shows the change in water composition of the
azeotrope as a function of temperature in the water ethanol system.Table 6.2 [Otsuki and
Williams, 19531 specifies these data for pressures higher than atmosphericpressure.
Table 6.2
Change in composition (mole fraction)
of the ethanol water azeotrope with pressure

Pressure

Temperature
(ethanol)

0.1
0.344
0.689
1.378
2.068

351.5
385.75
408.85
437.35
455.75

0.894
0.882
0.874
0.862
0.852

178

6. Mixtures: Liquid-Vapor Equilibria

0.15

460
Temperature (K)

300

350

Figure 6.8 Change in the composition of the water ethanol azeotrope as a


function of temperature [according to Rowlinson and Swinton, 1982,page 1081.

T = 283.15 K

T = 263.15K

0.25

0.50

0.75

Mole fraction of ethane

Figure 6.9 Isothermal liquid-vapor equilibrium diagrams for


the ethane (1) carbon dioxide (2) system.

6. Mixtures: Liquid-Vapor Equilibria

179

It may occur that the azeotrope exists only within a limited pressure range. In other
cases, the carbon dioxide ethane mixture for example, it persists up to the critical zone.
Figure 6.9 shows the isothermal diagram that forms two separate lenses when the temperature approaches the critical temperatures (both close to 30C) of the components.
The extremum of the bubble and dew points characterizingan azeotrope is most often
a maximum, and corresponds to the positivedeviations from ideality from the point of
view of the Gibbs energy (gE > 0, activity coefficients greater than unity). This is what is
observed for hydrocarbon mixtures, or mixtures of hydrocarbons with polar solvents. The
negative deviations from ideality can cause the appearance of azeotropes with a pressure minimum, which are often attributed to complexation phenomena (by hydrogen
bonding or hydration) in the liquid phase.
The correspondence between isothermal and isobaric azeotropic diagrams of a given
system is simple. For the azeotropic composition, a maximum of pressure corresponds to a
minimum of temperature. Figure 6.7 demonstrates this agreement. On the isothermal diagram as well as on the isobaric diagram, the azeotrope is observed at 5OoCand at atmospheric pressure.

6.2

THE LIQUID-VAPOR EQUILIBRIUM CONDITION


THE EQUILIBRIUM COEFFICIENT

We have already shown that the equilibrium condition between two phases is expressed by
the equality of chemical potentials or fugacities.We have:

p/=p,+

or

f~=f~,i=I,n

(6.2)

According to the nature of the mixture and the temperature and pressure conditions,
this condition may assume various forms that result from the fugacity expressions and its
simplificationsthat we have already come across.
If the temperature is less than the critical temperature of all the components, the vapor
pressure is defined for each one of them. Liquid phase fugacity may be calculated (see
Chapter 2, Sections 2.4.3. and 2.4.4) by using:
fi

- p qi l p q x , yi

with Pi representing vapor pressure, qicthe fugacity coefficient of the substance at saturation, the Poynting correction,xi the mole fraction, and xL the activity coefficient in the
liquid phase. For the vapor phase we have:

where P is pressure, y i the mole fraction,and pithe fugacity coefficient in the vapor phase.
The equilibrium condition may be written using the equilibrium coefficient,the ratio
of the mole fractions in the vapor and liquid phases:
Yi
K.= Xi

180

6. Mixtures: Liquid-Vapor Equilibria

therefore:
In this expression the first factor is the Pialpratio, the most important part of the liquid
vapor equilibrium coefficient.We specifically observe that at first approximation this coefficient varies with temperature like the vapor pressure, and is inversely proportional to
pressure.
The second part contains the fugacity coefficient at saturation, qi",the Poynting correction 9, and the fugacity coefficient in the vapor mixture. Its importance depends on the
temperature which governs vapor pressure and therefore the fugacity coefficient at saturation. The Poynting correction and the fugacity coefficient of the vapor mixture are temperature and pressure dependent. Finally,we have the activity coefficient.
However, we note that if all vapor pressures are low, so will be the total pressure. Under
these conditions the equilibrium coefficient is expressed by the equation:
K . =y L. =-Pi" yL
at low pressure
I X i
P '
It is understood that the "low pressure" restriction applies as much to total pressure as to
the vapor pressures.This expression could not be applied,for example,to the calculation of
the solubility of ethane in benzene at atmospheric pressure, since the vapor pressure of
ethane has several MPa.
Remaining within the framework of these last restrictive hypotheses, we see that the
previous equation allows us to express the bubble pressure as a function of the composition of the liquid phase:

P=

Pi"Xiyi"

(6.6)

So the shape of the bubble curve is directly related to deviations from ideality.
= l),we can state:
If, finally,we can neglect these deviations

(xL

Pi"
ideal solutions at low pressure: Ki= 3 = xi
P
and:

P=

PYX,

(6.8)

This particular case relates to Raoult's law, according to which "the partial pressure in
the vapor phase is proportional to the mole fraction in the liquid phase". We have seen
that the general Equation (6.4) splits the equilibrium coefficient into three factors. The
first is the Pi"/ P ratio, relating to ideal solutions at low pressure, the second,where we find
the fugacity coefficients and the Poynting correction, intervenes if pressure reaches a few
atmospheres (according to the desired precision, etc.), and the third is the activity coefficient in liquid phase, which may be very important (see Chapter 7).
We can also envision the case of ideal solutions under pressure. This hypothesis can be
valid as long as we do not approach the critical region. We note that the product that figures in the numerator of general Equation 6.4 is the fugacity of the pure substance in the
liquid phase:

6. Mixtures: Liquid-Vapor Equilibria

181

and the denominator is equal to the fugacity in the vapor phase, also for a pure substance:

So that for ideal solutions,the equilibrium coefficient becomes:


&,ideal

fl?"
fi*"

=-

As we see, the hypothesis of ideal behavior results in the fact that this coefficient is only
pressure and temperature dependent. It is independent of composition. Note, however,
that as a general rule, one of the reference fugacities that figure into the above equation
relates to a hypothetical state: indeed, if, at pressure P and temperature T, a mixture is in
the liquid-vapor equilibrium state, its "light" components are vapor in the pure state, and
its "heavy" components are liquids.

0 EXAMPLE6.1
Calculation of the ideal solution equilibrium Coefficient,
where the vapor phase is represented by the virial equation of
state for pressure truncated after the second term
If Bi is the second virial coefficient,v k is the molar volume in the liquid phase, and Piu
is the vapor pressure, the fugacity in the vapor phase is, using Equation (4.21),equal to:

Fugacity in the liquid phase is calculated by accounting for the Poynting correction:

such that the equilibrium coefficient is expressed by the equation:

Ki= Pi" exp (vF- B i )( P - Pi")]


P
RT

(6.10)

Once again, we find as the main term the ratio of vapor pressure to total pressure. It is
corrected by a multiplier whose effect is to bring the equilibrium coefficientscloser to
unity. Indeed, since the second virial coefficient is negative (at least under the conditions where this equation is applicable),we see that:
Pi"
ifP,O> P
then
1< K i < P
and

ifP,?<P

then

Pi" < K i < l


P

182

6. Mixtures: Liquid-Vapor Equilibria

This result is easily understood if we consider that the deviations from the laws of the
ideal gas, which are shown here by the virial equation of state, have a tendency to
approach the properties of the two phases. The selectivity of the liquid-vapor equilibrium, which is expressed by the ratio of the equilibrium coefficients of the substances
making up the mixture, is therefore diminished.
Note that the preceding equation allows a very simple empirical representation of the
variation of the equilibrium coefficients with pressure:
In (PK,) = ai(T ) P + pi(T )

(6.11)

Because of its simplicity,we will apply this method to the propane n-pentane mixture
at 300 K studied in the beginning of this chapter (Table 6.1, Fig. 6.3) even though the
vapor pressure of propane at this temperature (10 bar) makes the use of the virial
equation of state truncated after the second term a bit excessive for evaluating fugacity at saturation.
To calculate the vapor pressures, the second virial coefficients, and the molar volumes
in liquid phase, we shall apply the Lee and Kesler (Eq. 3.24) correlation, the
Tsonopoulos (Eqs. 3.18 to 3.20), and the Rackett method (Eq. 3.14).
At 300 K and a pressure of 3.232 bar, we obtain the results listed below (Table 6.3):
Table 6.3
Calculation of the ideal equilibrium coefficients of propane
and n-pentane at 300 K and 3.232 bar
Component
Propane
n-Pentane

Po
(bar)

B
(cm3. mol-l)

(cm3.mol-')

Ki

10
0.725

-390
-1 203

90
116

2.72
0.256

UL

We can derive the composition of the phases at equilibrium by resolving:


x1+x2=1
Yl+Y2=1
Y 1 = KlXl
Y2 = 4 x 2

The final result is:


x1 =

1-K2
~

K1- K2

and therefore we have the following numerical values:


x1 = 0.302

and

y, = 0.822

that we can compare to those provided by Table 6.1, x1 = 0.3 and y, = 0.83, for the
same conditions of pressure and temperature. As different calculation methods have
been applied, we estimate that agreement is excellent.

6. Mixtures: Liquid-Vapor Equilibria

183

It is important to once again note that in this ideality hypothesis, the equilibrium coefficients depend solely on the nature of the component being examined, and the temperature and pressure. The calculation of liquid-vapor equilibria is greatly facilitated.
We may attempt to apply the corresponding states principle to these coefficients, as
did Hougen and Watson [1959, p. 9411. However, such correlations include the prediction of the vapor pressures, a major factor of the equilibrium coefficient, that can most
often be obtained with better precision by applying a vapor pressure equation that is
specific for the component in question.

In any event, Equation 6.4, which we consider general, is valid only as long as the vapor
pressure is defined, as we pointed out when the equation was introduced. In addition, if the
temperature is only slightly lower than the critical temperature of the light components,
the calculation of their fugacity in the liquid phase utilizes a Poynting correction for a compressible fluid, which is not negligible if pressure is low. The result is only an approximation.
It is therefore preferable to evaluate the elements of the equilibrium condition by applying
one and the same model to the two phases, liquid and vapor, namely an equation of state.

0;
K.= 0:

(6.12)

This expression is of course also valid in the specific cases that we have considered up to
this point.
In summary, we have two paths for expressing the liquid-vapor equilibrium coefficient.
The first applies different models to the liquid and vapor phases. For the liquid phase, we
must calculate the vapor pressure of each component, the fugacity at saturation (using an
equation of state in the vapor phase for pure components), the Poynting correction (using
density data in liquid phase or an equation of state for pure substances in liquid phase),
and finally, the activity coefficients. For the vapor phase, we must calculate the fugacity
coefficients of each component in the vapor mixture by applying an equation of state to
this mixture. We call this path a heterogeneous method since, from the point of view of
its calculation, it mimics the heterogeneity of the mixture. This path has a number of
advantages. It lends itself to the simplifications that we pointed out, its expression breaks
the equilibrium coefficient down into significant elements, and it may be adapted to the
case that is investigated. The main choice lies in the expression and the calculation of the
activity coefficients in the liquid phase. We shall see (Chapter 7) that the models proposed
to this effect are numerous, and that some are predictive and take into account the molecular structure of the solution. One of the main characteristics of this calculation method is
its flexibility. It is preferred for the calculation of liquid-vapor equilibria at low pressure.
However, the application of different models to the two phases in equilibrium does not
allow for the representation of continuity of the physical states when we approach the critical point. Then, heterogeneity becomes incoherence.
The second path, which applies the same model to the two phases present does not run
into this obstacle. Based on an equation of state, it leads not only to the liquid-vapor equilibria calculations, but also to the calculation of other properties of the mixture: density,
heat capacities, enthalpy, entropy, etc. at least as long as the equation of state is accurate for
all of these properties. It may appear to be more general since it may be applied as well to

184

6. Mixtures: Liquid-Vapor Equilibria

low as to high pressure. In fact, it is less general because the equations of state that we use
(with the exception of the virial equation truncated after the second term, which is only
valid at moderate pressure and for the vapor phase) and the mixing rules that are associated with them (see Chapter 8) generally apply to apolar mixtures only.

6.3

DEPENDENCE OF THE EQUILIBRIUM CONDITIONS


ON TEMPERATURE, PRESSURE, AND COMPOSITION

The change in the equilibrium coefficient with conditions of temperature, pressure, and
composition is, in general, complex and is investigated by applying the equations that we
have just established [see Rowlinson and Swinton, 1982,p. 1041.We shall examine two specific cases.

6.3.1

Dependence of Bubble Pressure on Composition

We will limit ourselves to a binary system. For a homogeneous phase, there are three independent variables, temperature, pressure, and mole fraction of one of the components, x1
for example in the liquid phase, or y 1 in the vapor phase. At constant temperature, the variation of the chemical potentials of the components is therefore written as:
dpv=i?rdP+
and:

()
dy,
3Yl

i = 1,2 in the vapor phase

(H)
3x1 ZP

i = 1,2 in the liquid phase

ZP

dpf=Z,&dP+

kl

However, at liquid-vapor equilibrium, the equilibrium conditions:


pil/-p,&=o

i=1,2

allow for only one independent variable, for example xl. For any variation of xl, since the
equilibrium condition is respected, we have:
d(p/-pf) =0

i = 1,2

or:
The two preceding equations (i = 1and i = 2) are multiplied by y i , and we take the sum:

(3

[ y , ( V y - Z , L ) +y2(ZT-Z;j2L)] -

185

6. Mixtures: Liquid-Vapor Equilibria

In this equation, we note that:

by applying the Gibbs Duhem equation, and that:


ylv~+y2v;=vv

We therefore obtain:

(6.13)
Furthermore, it is possible to express the first derivatives of the chemical potentials as a
function of the second derivative of the Gibbs energy. Indeed, we have:
g = X l P l + (1 - X d k

and therefore, by taking into account the Gibbs Duhem equation:

These last two equations may be resolved in terms of pl and k :

which when derived with respect to x1yield:

such that in the expression (6.13), the numerator is simplified:

t )

Y,( 3x1

7;P + Y2(

EP

= (Y1X2-Y,X1)

$)Ep=

(Yl

a2g
(a),p

-4

to yield the equation:


(6.14)
This equation is applicable to any point on the bubble curve except the critical point, for
which the numerator and the denominator are simultaneously zero as the two phases in
equilibrium are identical.
Specifically,the equation allows us to illustrate the specificity of the azeotrope that we have
already pointed out, known as the Gibbs-Konovalow law. It states that any extremum of the
bubble curve causes identity of the compositions in the liquid and vapor phase, and vice versa.

186

6. Mixtures: Liquid-Vapor Equilibria

6.3.2 Dependence of Bubble Pressure on Temperature


Clapeyron Equation Applied to a Mixture
Let us consider a mixture with n components, of fixed composition, at its bubble point. As
previously, we observe that any simultaneous change of temperature preserving equilibrium is evidenced by the equations:

Multiplying each of these differential equations by yi and taking the sum, we obtain:
n

Ic Yidp.Y= C YidpL
i= 1

i=l

According to the generalized Gibbs Duhem equation (Eq. 5.18), we can state:
n

C,
i=l

py
vv
h"
yid = - dP- - d T
RT RT
R T ~
n

or:
Moreover, the composition of the liquid phase remaining constant, we have:
p;
.u;
d - = -dPRT RT

h,&

- dT

R T ~

and ultimately we have the expression:


n

dP,

(6.15)

dT

It may be interpreted or established in the following manner: at liquid-vapor equilibrium, condensation of a mole of vapor of composition yi in a large quantity of liquid phase
with composition xi takes place at constant composition of the liquid phase. This condensation is reversible and is accompanied by zero variation in Gibbs energy:
AGTp= 0

This equation is preserved if the temperature varies, always remaining at the bubble
point:
dW

, P >

=0

therefore, we have:
AVdP - AS dT = 0

6. Mixtures: Liquid-Vapor Equilibria

187

Since the Gibbs energy variation is zero, we can replace entropy variation with the ratio
AH/T and obtain the equation:
-dP,
=-

dT

AH
TAV

(6.16)

which by its form is similar to the Clapeyron equation. However, here the A operator represents the difference between the properties of a mole of mixture with composition yi in
the vapor phase of the same composition on the one hand, and in a liquid phase with composition xion the other hand, as specified in Equation 6.15.

6.3.3 CoherenceTests
The preceding general equations show that the liquid-vapor equilibrium data, temperature, pressure, and compositions of the two phases, are not independent. They must also
satisfy the coherence tests that are the consequences of the Gibbs-Duhem equation.
Generally,these tests may be applied only to binary systems and on the condition of using
a large number of isothermal measurements for which the precision has been estimated.
The oldest of these is based on the variation of the activity coefficients with composition. If we have:
-g E- - Q = x1 In % + x2 In

RT

*/2

we see that by taking into account the Gibbs-Duhem Equation (5.18) and the bond
between the mole fractions of a binary mixture:
x1 + x2 = 1

then:
This equation can be integrated between the limits x1 = 0 and x1 = 1, and accounting for
the fact that Q(0) = Q(1) = 0, we see that:
(6.17)
This condition may be verified graphically by using the experimentalvalues of the activity coefficients (the knowledge of which supposes a complete set of data: temperature,
pressure, and the compositions of the two phases). The area determined by the axis of the
compositions and the In %/% curve must have an algebraic value of zero. This often used
criterion must be abandoned in favor of treating the data with a thermodynamic model
without bias, such as the Redlich-Kister equation (see Chapter 7, Section 7.2).
If the data are not isothermal, the generalized Gibbs-Duhem law must be applied, and
the heats of mixing must be known [see Vidal, 1974,p. 4311.
It has also been proposed [Boissonas, 19391 that, the Gibbs-Duhem equation be
applied to the calculation of the bubble pressures by stepwise graphical integration. The

188

6. Mixtures: Liquid-Vapor Equilibria

results are compared to the experimental values. This method is applied with great
difficulty and is hardly ever used.

6.3.4 Stability and Critical Point Conditions


The equilibrium condition between phases, expressed by the identity of the chemical
potentials or fugacities,does not contain all the information supplied by the second law
of thermodynamics.We used only the Gibbs energy extremum condition at constant temperature and pressure, and not the minimum condition. It is therefore possible to acquire
additional information concerning the stability of a state.
In order to be stable, the concavity of the surface representing the variation (isotherm
and isobar) of Gibbs energy with composition G(x,, x2,xg,x4, etc.) must be oriented
toward the positive free enthalpies.
Let us consider a supposedlyhomogeneous system,characterized by the number of moles
of each component N i , and Gibbs energy G,and break it down into two subsystemsThefirst
is made up of a small perturbation of the initial system, and for each component contains
Ni - ANi moles. Its Gibbs energy will be calculated by a development of the second order:

Similarly,for the subsystem made up of ANi moles:

The global variation will thus be:

which must be positive whatever the ANi variations. The quadratic formula whose variables are the second derivatives of the Gibbs energy with respect to the number of moles
must therefore be positive definite.
For a binary system, we shall derive the simple equation relating to the Gibbs energy of
mixing:
aZgM

->0

ax 1

or

a2gM

-> O

ax;

(6.18)

Figure 6.10 illustrates this condition. It relates to the hexane (1) methanol (2) system at
298.15 K, and the excess Gibbs energy has been calculated using the NRTL model (see
Chapter 7) on the assumption that the system is homogeneous and liquid. To obtain the
Gibbs energy of mixing, we applied equations:
G M= G - xNipL;

(5.36)

189

6. Mixtures: Liquid-Vapor Equilibria

0.02
0

-0.02

-0.04

-0.06

-0.08
-0.1
I

-0.12

0.2

0.4

0.6

0.8

Mole fraction of hexane

Figure 6.10 Variation of the Gibbs energy of mixing of a n-hexane


methanol system in liquid phase as a function of the composition.
8 = 25C. Stability and liquid-liquid immiscibility.

G = xNip;

+ xNiRTlnxi + G E

(5.63)

g M= gE + x x i R T l n x i

or for one mole:

Note that the curve for the Gibbs energy of mixing has two inflection points, termed A
and B that correspond to the equation:

Between these two inflection points, the concavity condition is violated, and a mixture
illustrated by point M , for example,is unstable.
If we studied the shape of curve gM(x) with temperature, we could observe that at a
higher temperature the two inflection points approach each other and end up merging.We
are now at the critical point of the mixture that is characterized by the expressions:
azgM

-=0
ax2

and

a3g~

-= 0 at the critical point

(6.19)

ax3

In fact, the stability condition presented in this way is necessary but insufficient.
Consider line CD tangential to curve g" (x) at two points. Its intersectionswith axes x1 = 0

190

6. Mixtures: Liquid-Vapor Equilibria

and x2 = 1 correspond, according to the graphical depiction shown previously (Chapter 5,


Section 5.1.2), to the partial molar quantities for the Gibbs energy of mixing of components 2 and 1respectively.The partial molar quantities are themselves related to the activities or fugacities:
(5.41)

with

a,=

fi
-

' fi*

(5.39)

As the two points C and D have the same tangent at curve gM (x),we find that the
(activities) or (fugacities) of the two components are the same at C and D:

or:
and that the two mixtures are in equilibrium.
If we now juxtapose the two mixtures C and D in any proportion, since the Gibbs
energy is an extensive property, its value for the system thus obtained will change within
the segment of line CD. Any point of this segment represents a heterogeneous system
formed by the juxtaposition of mixtures C and D , which are in equilibrium. We note that
the values for the Gibbs energy relative to such points are less than those corresponding to
a mixture of the same global composition but considered homogeneous, and for which the
representation would be on the arc of curve CD. Such is the case with the pairs of points M
and M',or N and N'. The homogeneous mixture whose composition corresponds to the
point M abscissa is unstable not only because it does not satisfy the concavity condition,
but also because its separation into two phases (C and D ) in suitable proportions causes a
decrease in Gibbs energy.
It is the same for point N to which the heterogeneous system N' corresponds, but we
note that for this mixture N , the condition of concavity is respected. It is therefore necessary, but insufficient.
Later on (Chapter 8), we will apply this method to the calculation of liquid-vapor equilibria under pressure.
The short analysis that has been presented here was developed specifically by Baker
[l978,1982],Heidemann[1980], and Michelsen [1982a and b].These authors specifically proposed methods that simplify the calculations and apply the stability test to the Helmholtz
energy whose calculation using equations of state is easier than that of the Gibbs energy.

6.4

LIQUID-VAPOR EQUILIBRIUM PROBLEMS

First, we shall list the quantities that characterize a mixture divided into two phases, liquid
and vapor. Then we shall do the same for the relationships between these quantities in
order to determine the number of degrees of freedom for a system.This method, similar to

6. Mixtures: Liquid-Vapor Equilibria

191

the one we follow to establish the rule of phases from a balance of the intensive variables,
will be extended to the extensive variables.
Let us take a mixture containing n components. Since their quantity (number of moles
Ni of each one of the components) is known, we can find the global composition z i .
In order to characterize the mixture and calculate its properties,we must know the pressure, temperature, quantity of each of the phases present (number of moles N L , N"), and
their compositions (xi,y i ) ,in total, 2N + 4 quantities.From these quantities,the models that
we are using allow us to verify that the equilibrium conditions are respected (calculation of
chemical potentials or fugacities), to determine the volume, and all the thermodynamic
properties (enthalpy, entropy, etc.).
We use a material balance for each component (Eq. 6.1):
Ni = NLxi + Nvyi

(6.1)

or n equations. The equilibrium conditions involve chemical potentials or fugacities:

or equilibrium coefficients:

or again n equations. Finally the equations for mole fractions:


xxj=l

and

Cyi=1

for a total of 2n + 2 equations.


In order for the problem to be solvable,we need two quantities to be determined, either
from those we have listed, or from among the data that depend on them (total volume,
enthalpy,etc.).
So we can state that the total variance of a system is equal to two, and this result can
be expanded to a multiphase equilibrium. There is no contradiction whatsoever with
the phase rule because ( 1 ) we have considered that the total quantity of each component
is known, and (2) by including the quantity of each phase from among the unknowns,
we have expanded the variance to the extensive quantities. For example, we know
that for water at atmospheric pressure, the liquid-vapor equilibrium temperature is
fixed (intensive variance equal to one), and equal to 100C, but we cannot specify the
volume occupied,which, according to the extent of vaporization, may vary from 20 cm3to
30 liters.
Of course, in the choice of two data for the liquid-vapor equilibrium problem, intensive
variance must be respected. For a pure substance,from between these two data, only one
intensive variable can be selected, and the same is true for a binary mixture at the threephase liquid-liquid-vapor equilibrium.
The fact that the total variance is reduced to two and is independent of the number of
components will allow us to briefly introduce the principal problems of liquid-vapor equilibrium as a function of the nature of the available data.

192

6. Mixtures: Liquid-Vapor Equilibria

6.4.1 At Given Temperature (or Pressure) and Vaporized Fraction


Since the total number of moles is known, the quantities in each phase N L and NV are
derived directly from the vaporized fraction. The other unknowns are the pressure (or
temperature), and the composition of each of the phases.
Two important, specific cases must be mentioned. If the vaporized fraction approaches
zero or one, then the composition of the disappearing phase approaches a defined limit
that can be determined using the equilibrium equations, while the composition of the
remaining phase becomes identical to that of the overall mixture. These limits are nothing
but the bubble and dew points of the mixture.
Finally,note that the problem can have no solution when the mixture contains one or more
supercritical components, or have several solutions in the case of retrograde condensation.

6.4.2 At Given Temperature and Pressure


The unknowns are the N L and N quantities of each phase and their compositions. As
before, we rely on the material balances and the equilibrium equations to solve the problem. Of course, we must note that the problem only has a physical solution if the data are
found within the interior of the two-phase envelope. We can say that a problem is indeterminate if the overall composition, temperature, and pressure of the mixture coincide
exactly with the azeotropic coordinates.

6.4.3 Case Where One of the Data is a Thermodynamic Property


Problems for which the data are the temperature or the pressure and, additionally, a property of the system, may be perfectly conceivable in practice.
This case has been illustrated for the calculation of the pressure in a bottle of known
volume, containing a known quantity of liquefied gas of known composition, at a given
temperature.
Of course, we must also mention the case of expansion accompanying the partial liquefaction of the mixture. Depending on the problem, this expansion will be considered isenthalpic or isentropic and the data for the problem are the final pressure and the value of
enthalpy or entropy.

6.5

CALCULATION ALGORITHMS

We will not go into a detailed description of the algorithms that have been proposed to
solve the liquid-vapor equilibrium problems. However, some examples will allow us to
specify the principal difficulties.

6. Mixtures: Liquid-Vapor Equilibria

193

In general, the equilibrium coefficients are a function of temperature, pressure, the


nature of all the components, and the compositions of the liquid and vapor phases. Several
of these variables are among the unknowns, such that the solution to the problem is iterative. The discussion that we provide below can undergo many modifications aimed at speeding up the convergence of the resolution process. We can also mention some simplifications.
First, let us look at the case where these equilibrium coefficients can be calculated from
Equation 6.4:

6.5.1 Calculation of the Bubble Point


The composition of the liquid phase is identical to that of the mixture:
x 1. = z 1.

The equilibrium equations:


K.=2
xi

and the condition:


CYi = 1
are combined to give us the equation known as the bubble point relationship:

2 Kixi = 1

(6.20)

which must be validated.

6.5.1.1

Calculation of the Bubble Pressure

Since the temperature is known, we may a priuri calculate the vapor pressures and the activity coefficients that figure into the expression for the equilibrium coefficient.The iterative
calculation will start by taking the vapor phase as an ideal gas mixture (qy = 1)and forming
a preliminary hypothesis for pressure, P = 1bar for example, or applying Equation 6.6:

P=

pyxiy;

(6.6)

We therefore have all variables available to calculate the equilibrium coefficient, the
compositions of the vapor phase, and the s u m x K i x i . This sum does not fulfill condition
6.20 such that these products, K i x i ,cannot be assimilated to the composition of the vapor
phase, They will, however, yield a preliminary evaluation after normalization:
(6.21)

194

6. Mixtures: Liquid-Vapor Equilibria

The initialization accomplished, the iterative cycle will be continued by formulating a


new hypothesis for pressure based on the equation:
(6.22)
where n is the current iteration.
By making use of the available data, the fugacity coefficients in the vapor phase may be
calculated.
Each iterative step will in turn apply the preceding Equations 6.21 and 6.22 and the calculation of the fugacity coefficients 'piv, until condition 6.20 is met. Convergence is generally rapid.

6.5.1.2

Calculation of the Bubble Temperature

In this case, we do not make use of the value of the activity coefficients since they are temperature dependent. We can make two temperature hypotheses corresponding respectively to the boiling temperatures of the most volatile component, reduced by 20C (in
anticipation of azeotrope risk), and of the least volatile, increased by 2OOC. For each
hypothesis, the sum Kixiwill be calculated in an intermediate iterative cycle focused on
the compositions in the vapor phase, as the equilibrium coefficients are calculated from
Equation 6.4. This cycle will be initiated by taking the vapor phase as an ideal gas mixture
(cp" = l),followed by normalization of the Kixi products. It quickly ends with the stabilization of the sum Ki xi. The values thus obtained for the two hypotheses of temperature
are usually that one is greater than 1,and the other is less than 1,and new hypotheses for
temperature are created by dichotomy. For each new hypothesis,the intermediate iterative
cycle allowing the stabilization of the sum Kixi will be repeated until condition 6.20 is
met.

c.

c.

6.5.2

Calculation of the Dew Point

The composition of the vapor phase is identical to that of the mixture:


y I. = x 1.
The equilibrium equations:
K.=

3
xi

and the condition:


E X i

=1

are combined to give us the equation known as the dew point relationship:
(6.23)
which must be validated.

6. Mixtures: Liquid-Vapor Equilibria

6.5.2.1

195

Calculation of the Dew Pressure

We do not have the composition of the liquid phase and we therefore admit that, to begin
the calculation, it is an ideal mixture. An initial pressure hypothesis allows for the calculation of the equilibrium coefficients (Eq. 6.4), and the y i / K i ratios. These may not be
considered equal to the unknown composition in the liquid phase since the Equation 6.23
is not respected, and, even if it had been, they were calculated with the ideal mixture
hypothesis. Their normalization:
Yi
-

x.= Ki

(6.24)

22
i Kj

provides an evaluation of the liquid composition for the next iteration, permitting the calculation of the activity coefficients.A new pressure hypothesis is provided by the equation:

Pn=

Pn-1

c (2)
Ki
i

(6.25)

n-1

The Equations 6.24 and 6.25 as well as the calculation of the activity coefficients are
applied to each subsequent iteration. They normally converge rapidly.

6.5.2.2

Calculation of the Dew Temperature

This calculation will be performed in a manner analogous to the calculation of bubble temperature. Initialization will involve two levels of pressure, obtained from the vapor pressures of the extreme components of the mixture. For each pressure hypothesis, an intermediate iterative cycle, itself initialized by conceding the ideality of the liquid phase, will
allow the stabilization of the yi /Ki ratios and the calculation of their sum. Through
dichotomy we obtain the temperature at which condition 6.23 is validated.

6.5.3

Partial Vaporization

As we have seen, the two imposed data may be the pressure and the temperature, the pressure and the vaporized fraction, or the temperature and the vaporized fraction, while the
unknowns are, respectively,the vaporized fraction, the temperature or the pressure, as well
as, of course, the composition of the phases in equilibrium.
The combination of material balances (Eq. 6.1) and the equilibrium conditions (Eq. 6.3)
allows us to express the mole fractions as a function of these equilibrium coefficients and
the numbers of moles in the vapor and liquid phase:
x.=

Ki Ni
Ni
and y . =
N ~ + N ~ K ~
NL+NVKi

(6.26)

196

6. Mixtures: Liquid-Vapor Equilibria

To comply with the constraint equations, combined in the form:


2 X i

-2 y i= 0

we therefore write:

By dividing the numerator and the denominator of this equation by E N i and using the
vaporized fraction defined as:

we obtain the following equation that must be verified:

z+
1

(Ki - 1) zi
=O
( K i - 1) FV

(6.27)

Proposed by Rachford and Rice [1952], this resolution method applies in principle,
whether the system is ideal or not. If we need to account for all the elements of Equation 6.4,
we may proceed by dichotomy around the unknown variable, the vaporized fraction, temperature, or pressure. For each step of the iterative process, the equilibrium coefficients will
be evaluated by an intermediate cycle with the goal of stabilizing their value, and the sum:

2 1 +(Ki( ~- 1~ ) zi- F~1 )


will be calculated until condition 6.27 is validated.

6.5.4

Application to Ideal Solutions

In the case of ideal solutions, the equilibrium coefficient of any component of the mixture
is dependent only upon the nature of the component in question, the temperature, and the
pressure. It is independent of the nature of the other components, and of the compositions
of the two phases in equilibrium. It can be expressed as a function of the fugacities in the
reference states (pure substance, same temperature and same pressure as the mixture, in
liquid and vapor states):

If, in addition, pressure is low, and if the vapor pressures of all of the components are
low, we then write:

and:

197

6. Mixtures: 1iquid-Vapor Equilibria

When the temperature appears among the data, the solution to the problem of bubble
or dew pressure is immediate. If not, iterations are performed to improve temperature
until Equation 6.8 is validated.
In the more general case where, while retaining the ideality hypothesis, we must take
into account the deviations of the vapor phase from the ideal gas laws, as well as the
Poynting correction, all the equilibrium problems will have an iterative solution. We shall
illustrate this point with an example.

EXAMPLE 6.2

Calculation of bubble and dew pressures in an ideal mixture


Once again we shall consider the case of the propane n-pentane system (Section 6.1,
Table 6.1, Fig. 6.1 and 6.3, Example 6.1), and for an equimolar mixture, we will calculate the bubble and dew pressures at 300 K. The equilibrium coefficients will be calculated by the method used previously (Section 6.2, Example 6.1).
The pressure (bubble or dew) will be initialized at 1 bar, and the iterations will be
conducted by application of Equations 6.22 or 6.25.
Tables 6.4 and 6.5 below summarize the calculations.
Table 6.4
Calculation of the bubble pressure of an equimolar propane n-pentane mixture at 300 K
P=lbar

Component
Propane
n-Pentane
CKi .xi

P = 4.58 bar

P = 4.97 bar
K

K-x

8.42
0.736

4.21
0.368
4.579

1.97
0.194

.X

0.985
0.097
1.082

1.834
0.183

P = 4.998 bar
K

*X

0.917
0.0915
1.0085

K-x

1.82 0.91
0.182 0.091
1.0009

Table 6.5
Calculation of the dew pressure of an equimolar propane n-pentane mixture at 300 K

Component
Propane
n-Pentane
CYi/Ki

P = l bar

P=1.354bar

y/K

8.42
0.736

0.0593
0.6793
0.7386

6.263
0.554

P=1.378bar

I y / K I K
0.0798 6.156
0.545
0.903
0.9824

Y/K
0.0812
0.9176
0.998 8

P=1.38bar
K
6.148
0.544

y/K
0.0813
0.9186
0.99992

These results are compared to those shown earlier on (Table 6.1): Pbubble = 5 bar and
Pdew
= 1.36 bar.
We can estimate the agreement to be very satisfactory even though the thermodynamic models that were applied are different.
Of course, the calculation mode would have been identical for a mixture containing
more than two components.

198

6. Mixtures: Liquid-Vapor Equilibria

EXAMPLE 6.3

Partial vaporization at given temperatures and pressures


The unknowns are the vaporized fraction and the composition of phases in equilibrium. The equilibrium coefficients are known because of the ideality hypothesis, and
will remain invariant throughout the iterations,such that in Equation 6.7:
(6.27)

only the vaporized fraction is unknown. This equation is reduced to the resolution of
an expression of degree n - 1 for F,, where n is the number of components. We will
determine its root, which should be between 0 and 1. If there is no root that corresponds to this partial vaporization condition,it then follows that the mixture is in the
subcooled state (F < 0),or is a superheated vapor (Fv> 1). In the example below we
give the calculation intermediates for several vaporized fraction values.
We wish to fractionate a mixture in one theoretical stage such that the vapor pressure
of the liquid phase obtained does not exceed 3.5 bar at 300 K. The composition of the
mixture and the ideal equilibrium coefficients (calculated by applying Equation 6.10
according to the example used in Section 6.2, Example 6.1) are provided in Table 6.6.
Table 6.6
Composition of a propane n-pentane mixture
undergoing partial vaporization, and the equilibrium
coefficients of the components at 300 K and 3.5 bar

Component
Propane
n-Butane
Isobutane
n-Pentane
Isopentane

z
0.2
0.3

0.4
0.05
0.05

K
2.525

0.7708
1.066
0.240 1
0.3140

At the end of a partial vaporization where the temperature and pressure are 300 K
and 3.5 bar, we obtain a vapor phase that will be at its dew point and a liquid phase at
its bubble point, and whose vapor pressure or bubble pressure will therefore be in
line with the imposed specification. Of course, the mixture itself will be neither at its
dew point nor its bubble point.
We shall apply Equations 6.26 and determine its value as a function of the hypothesis
made for the vaporized fraction:

- 1)zi
c 1 +(Ki(Ki1) FV
Table 6.7 below lists the results of these calculations.

199

6. Mixtures: liquid-Vapor Equilibria

Table 6.7
Calculation of partial vaporization of a propane n-pentane mixture
at the given temperature and pressure (300 K and 3.5 bar)
1

FV

Component
Propane
n-Butane
Isobutane
n-Pentane
Isopentane

0.2
0.3
0.4
0.05
0.05
1

EX-EY

0.505
0.2312
0.4263
0.0120
0.0157
1.19
-0.19

x
0.0792
0.3892
0.3754
0.2083
0.1592
1.2113

0.5117

0.473 7

0.2
0.3
0.4
0.05
0.05
1

0.116 1
0.3366
0.387 9
0.078 1
0.074 1
0.993

0.2932
0.2594
0.413 4
0.01875
0.023 3
1.008

0.1123
0.339 9
0.387 0
0.081 8
0.077 0
0.998

Y
0.2837
0.262 0
0.4124
0.0196
0.0242
1.002

and finally:
0.524

FV

Component

Propane
n-Butane
Isobutane
n-Pentane
Isopentane
L

0.1112
0.3409
0.386 7
0.083 1
0.078 1
1

0.2807
0.262 8
0.412 1
0.0199
0.024 5

I
0

We note that the Equations 6.26 yield xi = zi for the value FV = 0, and yi = zi for the
value FV = 1, independent of the values of the equilibrium coefficients.Yet in these two
cases, the 6.27 condition is not met and we observe that the sum that should cancel out
changes sign in the interval from 0-1. We can also verify that the pressure lies between
the dew pressure (2.82 bar) and the bubble pressure (4.25 bar) of this mixture, at 300K.
We must not forget that for a binary mixture the solution to the problem is immediate, and is expressed by the simple equations:
(6.28)

(6.29)
These equations are always valid but lead to the solution immediately if the mixture
is ideal.
The problem of partial vaporization at given vaporized fraction and pressure (or temperature) is handled in the same way using successive hypotheses of the unknown
(temperature or pressure). However, the equilibrium coefficients should be calculated again for each new value of the variable on which they are dependent.

200

6. Mixtures: Liquid-Vapor Equilibria

If between the two data given for the calculation of the liquid-vapor equilibrium there
is a quantity such as volume, entropy, or enthalpy that does not figure into the equations
for material balance and equilibrium balance, the resolution process may then require a
double iteration, even in the case of ideal mixtures. We shall illustrate this in the next
example. It is similar to the example in Chapter 3 (Example 3.2).

EXAMPLE 6.4

Calculation of equilibrium at given temperature and volume


200 grams of a propane (10% molar), n-butane (40% molar), isobutane (40% molar),
n-pentane (5% molar), and isopentane (5% molar) mixture are contained within a
bottle having a volume equal to one liter. We wish to know the pressure, the relative
amount, and the composition of each of the liquid and vapor phases at 300 K.
To solve the problem, we need to have access to a calculation method for equilibrium
coefficients, and for volumes. It is not necessary that these two methods are absolutely
coherent. In this case, with the hypothesis of ideal mixture, the equilibrium coefficients will be calculated according to the method used in the preceding examples
(Example 6.1, Equation 6.10). Proceeding in this fashion, we will determine the second virial coefficients Bi for each component, as well as their molar volumes in the
liquid phase vf. Using the terms v; and vE for the molar volumes of the liquid and
vapor phases, we can write in accordance with the ideality hypothesis:

V = N~ V;

therefore:

+N ~ V ;

V = E N i [(l- F V ) v k + F

V;

We shall perform the calculations for one mole of mixture. The volume specification
is then:
v = 290.62 cm3 .mol-'
A prior calculation allows us to know the bubble and dew points of the mixture, and
the corresponding molar volumes:
bubble point
dew point

P = 3.56 bar
P = 2.60 bar

v = 103.9 cm3 . mol-I


v = 8892 cm3 . mol-I

Since the volume imposed is between these two limits, the mixture is in the two-phase
state. To specify this state, we shall proceed using successive hypotheses for pressure.
For each hypothesis we will solve the problem of partial vaporization at the given
temperature and pressure as in the preceding mathematical example, and for each of
these intermediate calculations, since we have the vaporized fraction, we will calculate the molar volume of each phase, as well as that of the mixture. The successive
pressure hypotheses are performed by dichotomy.

201

6. Mixtures: Liquid-Vapor Equilibria

The table below shows the state of the mixture and its overall molar volume for each
of the pressure hypotheses after resolution of partial vaporization at the given temperature and pressure.
Table 6.8
Calculation of partial vaporization of a mixture at given temperature and volume

P = 3.54 bar. F"= 0.0174

P = 3.51 bar. F"= 0.0418

Ki

xi
2.519
0.0940
0.769
0.404
1.0631 0.399
0.051 6
0.239
0.051 5
0.313
371

Propane
n-Butane
Isobutane
n-Pentane
Isopentane

214

urn

Yi
0.2368
0.311
0.424
0.0124
0.016 1

IP

= 3.52 bar. F"= 0.0293

Yi

2.509
0.766
1.059
0.239
0.312

0.0958
0.403
0.399
0.0511
0.0510
290.6

0.2403
0.309
0.423
0.0122
0.015 9

6.5.5 Non-Ideal Solutions


6.5.5.1

Non-Ideal Solutions at low Pressure

Around atmospheric pressure, the vapor phase can be fairly well assumed to be an ideal
gas mixture. However, the liquid phase may show large deviations from ideality.The equilibrium coefficients may therefore be expressed by the equation:

They are therefore independent of the vapor phase composition.The calculation of the
bubble pressure requires no iteration:

P=

c
i

PYXi yi"

(6.6)

A study of the activity coefficient models will allow us to apply these simplified equations.

6.5.5.2 General Case


If the precision of the data so permits, or if the pressure level does not allow for the
hypothesis of the ideal gas applied to the vapor phase, we must take into account the
fugacity coefficients and the Poynting correction. The equilibrium coefficient calculation
uses the general Equation 6.4

The example below illustrates the iterative process that we have just summarized
(Section 6.5.2.1).

2 02

6. Mixtures: Liquid-Vapor Equilibria

# EXAMPLE6.5
Calculation of the dew pressure in an acetone (y, = 0.5982)
water (y2 = 0.4018) mixture at 323.15 K
The unknowns are the composition of the liquid phase and the pressure. In Equation 6.4 the vapor pressure and the fugacity coefficient at saturation are fixed by the
data. They will be calculated by application of the Antoine Equation, and the virial
equation of state for pressure truncated after the second term.The second virial coefficient B will be calculated by the Tsonopoulos method (Chapter 3, Section 3.2.2.1).
The results (Table 6.9) are as follows:

Component
Acetone
Water

@Pa)

P?

(cm3/mole)

Bi

81.332
12.326

-1 363
-718

0.959
0.997

An initial pressure hypothesis is made based on the ideality of the liquid phase and by
assuming that the vapor phase is an ideal gas:

We find P = 25.03 kPa, x1 = 0.1841, x, = 0.8159.


Furthermore, we must calculate the Poynting correction and the activity coefficients
as a function of these values that change during the process of iteration.
The Poynting correction will be calculated using Equation 2.41, with the molar volume
in liquid phase being estimated by the Rackett method (Chapter 3, Section 3.2.1.2).
The activity coefficients will be obtained using the NRTL model (Chapter 7,
Section 7.6.2), and the values for the variables of this model will be taken from the
data bank established by Gmehling and Onken [1978]:

C,, = 2716 J . mol-l, C,, = 4359 J mol-l, a,,= 0.4927

Table 6.10 below summarizes the calculations.As specified, for each pressure hypothesis, the values for the compositions in the liquid phase are obtained by an intermediate iteration cycle that normalizes the yi/Ki ratios, without change in pressure, allowing for stabilization of the compositions. After stabilization, pressure is adjusted by
application of Equation 6.25:
(6.25)

6. Mixtures: Liquid-Vapor Equilibria

203

Although the initial pressure and composition hypothesis is a result of the law of ideal
solutions, we find that the compositions in the liquid phase stabilize after four iterations. For subsequent pressure hypotheses, this stabilization is obtained after the second iteration. Finally, three pressure iterations are sufficient to arrive at the final
result.
This result can be compared to the experimental data: 30 kPa, xi = 0.029, a comparison
that validates the NRTL method.
Table 6.10
Iterative calculation of the dew pressure of
an acetone (yl = 0.5982) water (yz= 0.4018) mixture at 323.15 K

6.5.6 General Calculation Method of liquid Vapor Equilibria


The preceding substitution methods may be easily transferred to the case where the equilibrium coefficients are expressed by Equation 6.12:
'PL
K.= 'PY

(6.12)

Regardless of the expression chosen, they generally lead to a solution with reasonable
calculation time when the mixture is under moderate pressure, and specificallywhen it is far

204

6. Mixtures: Liquid-Vapor Equilibria

from its critical point (whose coordinates are generally unknown). Yet we have observed
and it can be demonstrated, that convergence of this procedure in some cases is very slow,
and other methods have been developed specificallyfor the homogeneousmodels.A similar equation of state simultaneously describes the liquid and vapor phases. However, it is
also applicable if different models are attributed to the two phases in equilibrium.
The first of these simultaneously solves all of the equations imposed on the system. It
was proposed by Asselineau et al. [1979] and Michelsen [1980].
We have seen that the equilibrium state was determined as soon as the values of two of
the variables (temperature, pressure, quantity of each of these phases) or thermodynamic
properties (volume, enthalpy, etc.) were fixed.We shall denote these values T*, P*,V*, and
H*.If n is the number of components, we have also seen that the material balances, the
equilibrium equations, and the constraint equations yield 2n + 2 equations for the 2n + 4
variables. According to the type of problem, two of the equations such as T - T* = 0,
P - P* = 0, NV - Nv*= 0, V - V*= 0, H - H* = 0, etc., complete the system. This is not linear
as the products of the number of moles times the compositions figure into the material balances, and as the equilibrium equations most often involve the mixture fugacity coefficients or the activity coefficients,and the expression for these quantities as a function of
temperature, pressure, and composition, may be complex.The same is true for the volume
or enthalpy expression. We therefore apply the Newton method for the resolution of this
2n + 4 system, which requires prior determination of the expressions for the Jacobian variables of this system, namely the derivatives of each one of the equations that we want to
solve with respect to the variables (?: Nv,NL, xi,yi), expressions that are dependent on
the models applied to the mixture (activity coefficients,equations of state, etc.). We note,
however, that from one problem to the next, only the last two equations vary. Their nature
is related to the type of problem we want, such that the method has a degree of appreciable generality.It was introduced by Asselineau et al. [1979]. After resolving the problem, a
simple extension allows for the determination of the derivatives of the equilibrium conditions relative to any one of the variables. Of course, this kind of system may be simplified
in specific cases of bubble point, dew point, or partial vaporization at given temperature
and pressure.

r:

A second set of methods is based on the minimization of the Gibbs energy and prior
study of the stability of the system using the method called tangent plane.
Proposed by Michelsen [1982a, 1982b, 19851, it was the subject of many developments
[Nghiem, 1985,Aganval et al., 19851, and we shall introduce some of its principles with an
example for the calculation of liquid-vaporequilibria using an equation of state (Chapter 8).

6.6

SOLUBILITY OF CASES IN LIQUIDS

To finish up, we consider the specific case of systems made up of permanent gases,components whose critical temperature is below the range of temperature under examination,
and solvents,components whose vapor pressure is low within this temperature range.
We wish to determine the solubility of the gases in the liquid phase as a function of temperature and pressure.

6. Mixtures: Liquid-Vapor Equilibria

205

Of course, we may apply the methods based on an equation of state to a system defined
in this way (see Chapter 8). However, knowledge of the Henry constants (see Chapter 5,
Section 5.9) often provides a simple solution to the problem at hand.
First we shall consider the case of a binary system: a permanent gas (component 1)and
a solvent (component 2).
The Henry constant of the gas in the solvent is defined by the equation:
(5.86)
It depends on the identity of the gas, of the solvent, and on the temperature.
At moderate pressure (a few dozen bar), the solubility of the gas in the liquid phase is
weak, such that we may write:

fi" = 31112 x1

(6.30)

While still at moderate pressure, the fugacity of the gas in the vapor phase is close to its
partial pressure:
fi" = PY1
(6.31)
The equilibrium condition therefore leads to the following expression for solubility of
the gas in the liquid phase:
PY1
XI = (6.32)
%I2

So, as a first approximation,the solubility of a gas in a liquid solvent is proportional to


its partial pressure. This is the Henry law statement, and the range of composition and
pressure for which this law is reasonably complied with is often designated as the range of
application of the Henry law. This range varies according to the nature of the gas. If it is at
a temperature that is clearly higher than its critical temperature, the approximation represented by Equation 6.31 is well verified. Such is the case for solutions of nitrogen and
methane for example, and for pressures up to 100 bar.
The previous equations show that the Henry constant plays the role of a corrected
vapor pressure, especially in order to take into account the non-ideality (in the symmetric
convention sense) of the mixture.
We can apply these equations to solutions diluted with a "subcritical" component, and
express the equilibrium condition either with the equation:

or by using the Henry constant:


%I2
Kl = p

(6.33)

and thereby more easily relate the Henry constant to vapor pressure, activity coefficient at
infinite dilution, fugacity coefficients at saturation, and the Poynting correction.

206

6. Mixtures: Liquid-Vapor Equilibria

If the system in question has several gas solutes, we can say that as long as their solubility remains low, their interactions are negligible, and Equation 6.32 is applicable to each
one of them.
On the other hand, if the solvent is a mixture, the Henry constants for each of the
solutes depend on the composition of the solvent.

REFERENCES
Abbott MM (1986). Low pressure phase equilibria:measurement of VLE. Fluid Phase Equilibria, 29,
193-207.
Agarwal RK, Li Y-K, Nghiem LX, Coombe DA (1991) Multiphase multicomponent isenthalpic flash
calculation.J. Canadian Petroleum Technology, 30, No. 3,69-75.
Asselineau L, Bogdanic G, Vidal J (1979) A versatile algorithm for calculating vapour-liquid equilibria. Fluid Phase Equilibria, 3,273-290.
Baker LE, Luks KD (1978) Critical point and saturation pressure calculations for multicomponent
systems.Paper presented at the 53 rd SPE meeting, Houston, October 1978, ref. SPE 7478.
Baker LE, Pierce AC, Luks KD (1982) Gibbs energy analysis of phase equilibria. SPE Journal,
October, 731-741.
Boissonas (1939) Helv. Chim. Acta, 22,541.
Deiters UK, Schneider GM (1986) High pressure phase equilibria: experimental methods. Fluid
Phase Equilibria, 29,145-160.
Gmehling J, Onken U (1978) Vapor-liquid equilibrium data collection. Chemistry Data Series,
Dechema.
Heidemann RA, Khalil AM (1980) The calculation of critical points. AZChE J., 26,769-779.
Holste JC, Hall KR, Eubank PT, Marsh KN (1986) High pressure PVT measurements. Fluid Phase
Equilibria, 29,161-176.
Hougen OA, Watson KM, Ragatz RA (1959) Chemical Process Principles. Wiley & sons, Ed.
Kehiaian H International Data Series, Ser. A. Thermodynamic properties of non-reacting binary systems of organic substances.The Texas A&M University.
Kreglewski A (1984) Equilibrium Properties of Fluids and Fluid Mixtures. Texas A&M University
Press, College Station, Texas.
Marsh KN (1989) New methods for vapor liquid equilibria measurements. Fluid Phase Equilibria, 52,
169-184.
Michelsen ML (1980) Calculation of phase envelopes and critical points for multicomponent mixtures. Fluid Phase Equilibria, 4,l-10.
Michelsen ML (1982a) The isothermal flash problem. Part I. Stability. Fluid Phase Equilibria, 9,l-19.
Michelsen ML (1982b) The isothermal flash problem. Part 11. Phase split calculation. Fluid Phase
Equilibria, 9,21-40.
Michelsen ML (1985) Simplified flash calculations for cubic equations of state. Znd. Eng. Chem. Proc.
Des. Dev., 25,184-188.
Nghiem LX, Li Y-K, Heidemann RA (1985) Application of the tangent plane criterion to saturation
pressure and temperature computations. Fluid Phase Equilibria 21,30-60.

6. Mixtures: Liquid-Vapor Equilibria

207

Otsuki H, Williams FC (1953) Effect of pressure on vapor-liquid equilibria for the system ethyl alcohol water. Chem. Eng. Progress; Symposium Series, 49 (6), 55-67.
Rachford HH Jr, Rice JD (1952) J. Petrol. Technol.,4, (l), 19; (2), 3.
Renon H, Asselineau L, Cohen G, Raimbault C (1971) Calcul sur ordinateur des Cquilibres Ziquidevapeur et liquide Ziquide. Editions Technip,Paris.
Rowlinson JS, Swinton FL (1982) Liquid and Liquid Mixtures. 31d edition, Buttenvorth, London.
Vidal J (1974) Thermodynamique. Les mCthodes appliqutes au raffinage et au gCnie chimique. Editions Technip,Paris.
Wichterle I, Linek J, Hala E (1973-85) Vapor-Liquid Equilibrium Bibliography. Elsevier,Amsterdam.
Wichterle I, Linek J, Wagner Z, Kehiaian HV (1993) Vapor-Liquid Equilibrium Bibliographic Data
Base. Eldata, SARL,Montreuil, France.

Deviations from Ideality


in the Liquid Phase

In the preceding chapters we have shown that we can estimate the liquid-vapor equilibria
at low pressure and the properties of a mixture in the liquid phase by using the following
steps:
0 Calculate the properties of each component in the reference state.
0 Calculate the properties of the mixture in the ideal state.
0 Calculate the excess quantities.
The first of these steps involves the application of the methods presented in Chapters 2
and 3. However, we may run into some difficulties. A liquid mixture may contain components that, if pure, would be in the vapor state or possibly in the supercritical state. In particular, such is the case if it is in equilibrium with a vapor phase (meaning at its bubble
point). Under these conditions and for the lightest of its components, the reference state is
not stable and corresponds to a superheated liquid. To estimate the properties of such a
state requires simplifyinghypotheses.These hypotheses are plausible as long as we are at a
temperature significantly below the critical temperature of the component in question.We
consider the molar volume equal to the molar volume at saturation. The same is true for
the enthalpy.Finally,for fugacity we will use the Poynting correction even if the pressure is
less than the vapor pressure. Near the critical point, this procedure becomes questionable
and, at a temperature higher than the critical temperature, it may not be applied because
the properties at saturation do not exist.
This estimation process for liquid phase properties is therefore only viable if the temperature is dearly lower than the critical temperature of all the components. We w i l l use
this hypothesis.

The second step involves the mixing properties that characterize the ideal solution.
The volume of mixing of the ideal mixture is zero as is the heat capacity of mixing
or the enthalpy of mixing. This is not true for the entropy or the Gibbs energy of mixing.
The last step involves experimental data, correlation, or even prediction of excess quantities on the basis of models. The goal of this chapter is to introduce some concrete
examples of deviation from ideality, primarily emphasizing the effects that these deviations have in terms of liquid-vapor equilibrium.As we are already familiar with the essen-

21 0

7. Deviations from Ideality in the Liquid Phase

tial role of activity coefficients,we shall introduce the models that are most often applied
to their correlation and prediction.

7.1

EXCESS QUANTITIES

7.1.1 Excess Volume, Excess Heat Capacity


These quantities are often neglected in calculations.It is true that the excess volume (the
dilatation or contraction upon mixing) is generally very low. Figure 7.1 [Berro, 19861 shows
the variation of this quantity with the composition for the ethanol n-heptane mixture. Note
that the maximum value of excess quantity is 0.5 cm3 . mol-l, or around 0.5% in relative
value. The precision with which the properties in the reference state and the ideal solution
are evaluated is usually much lower. Therefore, the excess volume may rightly be neg1ected.A review of the experimental data was done by Battino [1971] and more recently by
Handa and Benson [1979].

0.5

Xi

Figure 7.1 Mixing volume of the ethanol (1) n-heptane (2) system.
[Berro, 19861.

7. Deviations from Ideality in the Liquid Phase

21 1

By applying Equation 5.78, we can conclude that the activity coefficients under
these conditions are practically independent of pressure. Indeed, we shall see that the
models relating to excess Gibbs energy retain only composition and temperature as variable.
As for excess heat capacity that determines the influence of temperature on the heat of
mixing, it is also commonly neglected. Figure 7.2 [Costas and Patterson, 19851 for the benzene n-decane mixture shows that it may, however, reach -4 J K-l. mol-l, a value that we
can compare to the benzene heat capacity at the same temperature, 137 J/mol. In fact, we
have access to many experimental measurements concerning this property, but the models
that are applied to the deviation from ideality cannot represent excess Gibbs energy, heat
of mixing, and excess heat capacity all at the same time.

0.5

XI

Figure 7.2 Excess molar heat capacities of the benzene (1)


n-decane (2) system [Costas and Patterson,19851.

However, this is a quantity that seems particularly sensitive to the structure of a solution
and to the molecular interactions that occur upon mixing. The example of the 1,4 dioxan
cyclohexane system is given in Figure 7.3 [Trejo et al., 19911. The excess quantity is low in
relative value, but the shape of its variation with the composition reveals the complexity of
the system and represents a challenge to any model that is not based on a real understanding of the phenomena at work.

21 2

7. Deviations from Ideality in the Liquid Phase

4.2

UI

4.6

OQ
-0.8

-1.0

II

0.5

Figure 7.3 Excess molar heat capacities for the 1-Cdioxan (1)
cyclohexane (2) system [Trejo et al., 19911.

7.1.2 Heat of Mixing


It is very clear that this quantity is part of enthalpic balances. It also enters into the calculation of separation processes by both enthalpic balance at each stage and by the fact that
partial excess molar enthalpies determine the variation of the activity coefficients with
temperature, as Equation 5.77 shows:
(5.77)
For example, it has been shown that omitting this quantity from the calculation of the
cyclohexanolcyclohexanone mixture distillation yields a product of unacceptable purity or
a production that is diminished by 30% [Zudkevitch,19781.
The order of magnitude and the sign of the heat of mixing are variable. For binary
hydrocarbon systems, the influence of structure has been the subject of a detailed study
[see Abdoul et al., 1991,for example]. Generally moderate and positive (Fig. 7.4), the quantity is very sensitive to temperature when the mixture components have very different
chain lengths (Fig. 7.5). Furthermore, it may reach several kilojoules in the case of specific
interactions. It may also happen that a system, seemingly ideal from the point of view of
Gibbs energy, has a high heat of mixing. Such is the case with water dimethylformamide
mixtures. As with heat capacity,but generally to a lesser degree, its variation with the com-

213

7. Deviations from Ideality in the Liquid Phase

1 200

900

a.

600

300

0.5 x (benzene)

(J . mol-1)

90

b.

60

30

0.5 x (n-hexane)

Figure 7.4 Excess enthalpies for hydrocarbon mixtures


a. benzene + n-heptane:(0);
+ n-undecane: (0);
+ n-pentadecane ( t ).
b. n-hexane + n-octane: (A); + n-decane:(0);
n-dodecane (0);
+ n-hexadecane( *).

21 4

7. Deviations from Ideality in the Liquid Phase

hE

(J . mol-1)

0.5

x (hexane)

Figure 7.5 Variation of excess enthalpy of the n-hexane n-hexadecane


system with temperature.
298.16 ( *);313.15 (0);
324.15 (0);333.15 (0);
349.15 (A).

position may be the sign of complex molecular interactions. This is the case with the water
ethanol system [Fig. 7.6, Larkin et al., 19751 for which deviations from the ideal behavior
may be interpreted by the rupture and recombination of hydrogen bonds.
The model for heat of mixing is generally paired with that of excess Gibbs energy as
soon as the model applied to this second property takes into account the influence of temperature. However, it must be emphasized that the change of excess Gibbs energy with
composition is generally simpler than that of heat of mixing, and priority is most often
given to the calculation of activity coefficients. Hence, the models are often less accurate
when applied to the heat of mixing. So, in the case of the above-mentioned water ethanol
mixture, excess Gibbs energy is moderate and its variation with composition leaves nothing to be predicted from the shape of the hE(x)curves.

7.1.3

Excess Gibbs Energy and Activity Coefficients

Generally, the models involve excess Gibbs energy, but the practical impact of deviations
from ideality on phase equilibria is evaluated in terms of activity coefficients.We have seen
that at low pressure the liquid-vapor equilibrium coefficient, ratio of the mole fractions in
both phases, is calculated using Equation 6.5:

21 5

7. Deviations from Ideality in the Liquid Phase

600

400

200
7
h

-I
! o
3

UI

.E

-200

-400

-600

-800
0

0.2

0.4

0.6

0.8

x (ethanol)

Figure 7.6 Variation of the excess enthalpy for the ethanol (1)
water (2) system with temperature [Larkin et al., 19751.

y. = Pi yL
K.= 2
I X i
P I

at low pressure

(6.5)

where Piarepresents the vapor pressure, P the pressure, and the activity coefficientin the
liquid phase. It is therefore in terms of the activity coefficients that we shall provide a few
examples. The activity coefficients are composition dependent, and as we have already
mentioned, it is in a dilute medium that the behavior of a component usually differs the
most from its properties in an ideal solution. The activity coefficients at infinite dilution
therefore represent, in a way, a scale of non-ideality. Table 7.1 concerns hydrocarbon
mixtures. It shows that mixtures containing paraffins and aromatics are unwaveringly non
ideal. In a dilute solution of benzene, the volatility of heptane is increased by about 70%,
such that despite the difference in boiling temperature (80C for benzene, 100C for heptane), the separation by simple distillation is impossible,as predicted by the pinch zone
seen in the liquid-vapor equilibrium diagram of this mixture (Fig. 7.7). In any case, such
deviations cannot be neglected. However, if the molar volumes are not appreciably different, hydrocarbon mixtures of the same family may be considered ideal.

21 6

7. Deviations from Ideality in the Liquid Phase

Table 7.1
Activity coefficients at infinite dilution in hydrocarbon mixtures
Components
1
Hexane
Heptane
Heptane
Hexane
Heptane
Cyclohexane
Cvclohexane

2
Benzene
Benzene
Benzene
Toluene
Toluene
Benzene
Toluene

("C)
1.7
1.7
1.6
1.8
1.4
1.5
1.35

1.4
1.35
1.3
1.6
1.3
1.4
1.35

Mole fraction (benzene)

1 .o

0.75
h

a
c
a

0.5
v

0.254

0 I
0.0

0.25

0.5

I
I
0.75

x (benzene)

Figure 7.7 Liquid-vapor equilibrium diagram of the benzene (l),


n-heptane system at atmospheric pressure.

21 7

7. Deviations from Ideality in the Liquid Phase

The mixtures containing both apolar compounds (hydrocarbons) and polar compounds
show deviations from ideality of an entirely different order of magnitude. As an example,
Table 7.2 lists activity coefficients at infinite dilution for heptane and benzene in some
polar compounds.

Solvent
Acetone
Methanol
Ethanol
Dimethylformamide
Dimethylsulfoxide
Ethylene glycol

(C)

Yr

(C)

y;

40
40
40
25
25
25

6.4
34
13
21
121
750

31
30
45
25
25
25

1.6
7.2
5.1
1.4
3.5
32

We observe that in a dilute medium, the volatility of a hydrocarbon is considerably


modified and that the polar compound acts selectively as the activity coefficient is very
sensitive to the nature of the hydrocarbon. As such, in dimethylformamide the relative
volatility of heptane and of benzene is multiplied by the activity coefficients ratio, 21A.4.
In an extractive distillation performed in the presence of this solvent, heptane will be eliminated at the head of the column although its boiling temperature is 20C lower than that
of benzene.
The elevated values for the activity coefficients generally give rise to partial miscibility
in the liquid phase. This can be observed for heptane in methanol, dimethylformamide,
dimethylsulfoxide, and ethylene glycol.The aromatic hydrocarbon miscibility in these solvents is, if not total, at least higher, and the liquid liquid extraction purification processes
rely on this selectivity (see Chapter 9).
The examples of deviation from the ideal mixture that we have just discussed yield
positive deviations. The activity coefficients are greater than one and the excess Gibbs
energy is positive. If it exists, the azeotrope phenomenon is observed by a pressure maximum (or a temperature minimum). Such is most frequently the case, at least for non electrolyte mixtures. We shall see, however, that the molecular volume differences cause negative deviations that may be very large in the case of polymer solutions.
At the molecular level, the mixture of two polar compounds is accompanied by the rupture of dipolar interactions between identical molecules and the reestablishment of such
interactions between different molecules. It is difficult to predict the resulting sign and
magnitude of their deviations from ideality.

218

7.2

7. Deviations from Ideality in the Liquid Phase

CORRELATION OF LIQUID VAPOR


EQUILIBRIA AT LOW PRESSURE
COHERENCETEST

We return to the calculation of liquid-vapor equilibria in order to introduce a method for


the determination of activity coefficients that is applicable to incomplete binary data.
This method uses a perfectly flexible excess Gibbs energy expression. It is capable of
representing the variation of this property and the related activity coefficients with the
composition, regardless of the complexity of this variation. Due to this fact, it is not based
on any particular conception of the structure and properties of the solution, and we may
not expect prediction of higher order systems, such as ternary systems, for example.
The following example will help us grasp its principle.

EXAMPLE 7.1

Correlation of liquid-vapor equilibrium data for


the acetonitrile ( 1 ) toluene (2)system
For the acetonitrile toluene system we have access to the experimental data shown in
Table 7.3.
Table 7.3
Vapor pressures of acetonitrile (1)
toluene (2) mixtures at 318 K
X1

0
0.027
0.040 5
0.098
0.213
0.301
0.4795

9.88
12.11
13.06
16.197
20.226
22.379
25.276

0.596
0.735
0.866
0.934
0.973 5
1

26.423
27.448
28.110
28.142
27.963
27.751

Note that the data are at low pressure, and that the vapor pressures of the two system
components whose values are given in the same table (27.751 kPa for acetonitrile
and 9.88 kPa for toluene) are themselves low. We therefore observe that at first
approximation, the vapor phase is a mixture of ideal gases, and we neglect the
Poynting correction and apply the equilibrium equation in the form expressed in
Equation 6.5:
y.
K . = =

Pi
- yL
P

at low pressure

7. Deviations from Ideality in the Liquid Phase

219

Yet the system is not ideal, as evidenced by the polarity difference of its constituents
and the existence of an azeotrope (maximum of the curve P(x)).
On the other hand, these data are incomplete:we only have a series of bubble points
(temperature, pressure, composition of the liquid phase), but we ignore the corresponding compositions of the vapor phase that would allow for a direct estimation of the
activity coefficientsby application of the preceding Equation 6.5. This type of data is frequently encountered when the staticmethod for experimentaldetermination is used.
To correlate these values, we will assume a model for the non-ideality of the mixture:

g E = CoRTxlx,

(7.1)

It has but one parameter, C,, that must be determined, and of course, it will be necessary to verify the validity of the model. To this excess Gibbs energy model, using
Equation 5.75:
(5.75)

we can relate the following expressions for the activity coefficients:


In

= Cox, and

In fi = Cox:

(7.2)

that we incorporate into Equation 6.6 that gives the value of the bubble pressure:

P=

cPiXiyi
i

to yield:

Pea, = P y x , exp (Cox;)+ P& exp (Cox,)

For a given value of parameter C,, and for each experimental point (T,xl, x,), we can
therefore compare the experimental value of the bubble pressure Pexp,provided in
Table 7.3, with the calculated value Peal, which is derived from the preceding equation, and evaluate the average quadratic deviation:

where the sum is extended to all nexpexperimental determinations. This deviation


depends only on parameter C, and the best value of this parameter will be determined
by minimizing it. In the case shown here, we find that C, = 1.2. Of course,this procedure
does not validate the model at all. For this purpose, it is necessary to have a closer look
at the correlation that was obtained.Figure 7.8 andTable 7.4 demonstrate that the data
are adequately represented and, taking into account the hypotheses that were made
(the vapor phase behaves as a mixture of ideal gases), we may accept the result as being
satisfactory.We may therefore complete the data by calculating the activity coefficients and the composition of the vapor phase on the basis of the optimized C,value:

p1
Y , = -x1 exp
Pcal

pz
Y , = -x2 exp (cox:>
Pcal

220

7. Deviations from Ideality in the Liquid Phase

The values we obtain are listed in Table 7.4 and represented in Figure 7.8.
Table 7.4
Liquid-vapor equilibrium of acetonitrile(1) toluene (2) mixtures at 318 K

Yl

9.88
11.96
12.89
16.24
20.64
22.71
25.20
26.23
27.20
27.81
27.91
27.86
27.75

0.213
0.4795
0.596
0.735
0.866
0.934
0.973 S

30

1
1
1
1.01
1.06
1.11
1.32
1.53
1.91
2.46
2.85
3.12
3.32

3.32
3.11
3.02
2.65
2.1
1.8
1.38
1.22
1.09
1.02
1.01
1
1

0
0.195
0.263
0.447
0.602
0.661
0.731
0.767
0.816
0.883
0.933
0.971
1

25

';ii 20

I /

!?
3

!?
a

/
/
/

15

10

1
0

0.25

0.5

0.75

x1, Y1

Figure 7.8 Liquid-vapor equilibrium diagram for the acetonitrile (1)


toluene (2) system. T = 318 K.

7. Deviations from Ideality in the Liquid Phase

22 1

Estimation of Experimental Precision


Coherence Test
In the preceding example we should, in fact, have compared the calculated deviations,
after optimization, to the experimental uncertainties. These bear on temperature ( 6 0 ,the
compositions (&), and pressure (6P).Calculating the pressure for the values of the composition and temperature taken as true, we may relate the relative uncertainties at these
quantities to the pressure by defining a resulting uncertainty W.

and the function that we minimize is defined by:

where nexpand nparrespectively denote the number of experimental points (except the
pure substances) and the number of parameters (equal to one in the previous example).
This average quadratic deviation o,, according to its definition, must be close to unity in
order for the correlation to be deemed satisfactory. If such is not the case, we may first
blame the model, and consider it unable to represent the system in question.
In practice, we apply the expression proposed by Redlich-Kister [1948] for the excess
Gibbs energy:

g = RTx,x2[Co+ C,(x, - x2) + C2(x,- x2)2 + C3(x,- x2)3 + ...]

(7.6)

At first glance, this expression is perfectly flexible. It will be successivelytruncated after


the first, then the second, then the third term, each time optimizing the parameters (minimization relating to the C, variables, then C,, C, ,etc.). The improvement can be continually monitored by evaluating the decrease in the average quadratic deviation defined by
Equation 7.5. When this improvement is no longer substantial,we can say that the residual
deviation is indeed the result of experimentaluncertainty. If it differs significantly from the
value of one, we can say that these uncertainties were underestimated, or, more rarely,
overestimated.
Of course, in the case where we proceed with such an analysis,we must account for the
imperfection of the vapor phase and the Poynting correction, and express the equilibrium
coefficientsin the general form 6.4:

the fugacity coefficients being calculated,at low pressure, using the virial equation of state
truncated after the second term. The virial coefficients will be preferably estimated from
experimental data, or, if none are available,from a predictive correlation such as that from
Tsonopoulos (see Chapter 5). In this case, the calculation of the binary coefficient B1,2may
introduce a bias into the data correlation.

222

7. Deviations from Ideality in the Liquid Phase

Finally,we note that this method may only be applied when the number of experimental points is large compared to the number of adjusted parameters. Other excess Gibbs
energy expressions have been proposed, for example,based on the use of orthogonal polynomials [Christiansen and Fredenslund, 1975; Klaus and van Ness, 19671.
When we use complete data (that is to say that we have available temperature, pressure, and the composition of the two phases), we may then determine their coherence. For
this we define three objective functions:

SQy =

nexp - npar

It is understood that the resulting uncertainty for the compositions in the vapor phase
Ay is estimated in the same way as the one relative to pressure AP. The definition for
the two last objective functions comes back to subdividing the data into two incomplete
sub-sets.Through minimization of each one of these three objective functions we determine the parameters of the model. The results must be the same within the uncertainty
interval of these parameters. Table 7.5 shows this approach [Neau, 19791 applied to a
methanol n-propanol system [Berro et al., 19751. In this case, the Redlich-Kister equation
is rewritten as:
rn

and we effectively observe the coherence of the parameters. We also note that the results
from incomplete TPx or complete TPxy data are equivalent as far as precision obtained
for the parameters. It is not the same for Txy data, which contain less information.
Table 7.5
Correlation of the equilibrium data of the methanol n-propanol system [Berro, 19751.
Data processing and coherence measurement
Data 5 p e
Objective Function

TPxy
SQpy

Cl

0.064 7 0.001 7
0.008 1 0.0006
0.0029 i 0.0007

c2
c
3

*
*

TPx
SQP
0.063 9
0.008 2
0.0032

* 0.001 3
* 0.0007

* 0.000 7

TXY
SQY
0.0645 f 0.0034
0.006 3 0.001 6
0.003 6 0.001 6

*
*

7. Deviations from Ideality in the Liquid Phase

223

As we have already noted (see Chapter 6, Section 6.3.3),the coherence tests may use only
data at low pressure. Under these conditions only can the vapor phase be strictly represented by a virial equation of state truncated after the second term. In addition, these tests
are applicable only to binary mixtures for which we have numerous data that covers the
entire composition interval, and whose uncertainty has been estimated. These items were
studied by E. Neau [1979],PCneloux et al. [1975,1976,1990],and Neau and PCneloux [1981].

7.3

INFLUENCE O F VARYING MOLAR VOLUME:


THE COMBINATORIAL TERM

Up until this point, we attributed the deviations from ideality to molecular interactions,
and it is just this matter that will intervene in the models that are to be discussed further
on, and that are commonly applied.
However, we must emphasize that in a mixture containing compounds of very different
molar volumes, even if these compounds are of a similar chemical nature, we may observe
non-negligible deviations from ideality. Such is the case for solutions of monomers within
its polymer, or for example, polyethylene in paraffinic hydrocarbons.
According to Flory [1942]and Huggins [1941,1942],these deviations from ideality are of
entropic origin, and correspond to the possibilities of distribution of the polymer molecules
in a tridimensional lattice.The term combinatorial refers to these types of deviations.
If we consider a binary system composed of a high molecular weight compound (the
polymer),that can be represented by a chain containing p segments,with each segment
occupying one node of the lattice (Fig. 7.9), as well as a low molecular weight compound
whose molecule occupies only one site, a mixture containing Nl molecules of this second
compound and N2 molecules of the polymer will occupy a lattice of Nl + pN2 sites.
We will assess the various possible arrangements. To do this, we first look at the first
segment of the first polymer molecule:there are N , + pN2 possibilities. For the second segment, we must account for the number of neighboring sites z, which we call the coordination number of the lattice. The third segment may occupy only z - 1 sites, one of the bordering sites being occupied by the second segment, and it is the same for subsequent segments up to the last site of the first polymer molecule. This first molecule therefore has:
P, = ( N , + p N 2 ) z(z - l)p-2distinct configurations
If we take the kthpolymer molecule, we only have (Nl + pN2)- ( k - 1)p vacant sites, and
we will consider that the number of configurations for the placement of each segment has
been reduced in proportion with the remaining sites.We therefore have:

or in other words:
Pk =

(4
+p(N2)
+ pN2Y-l
+

z(z - l ) P - 2 distinct configurations

224

7. Deviations from Ideality in the Liquid Phase

For the Nz polymer molecules, the total number of possible configurations Q12,is equal
to the product of the terms so determined, divided by N,!, as the N, polymer molecules are
indistinguishable.The molecules of the low molecular weight component that occupy only
one lattice site are placed on the remaining sites without introducing new configurations
since they are themselves indistinguishable. Therefore, in total, for the mixture under
examination we have:

We must now compare these configurationsto those evaluated for the pure substances.
with N , = 0
Q,,, is equal to one, and Q2,zis determined using the same equation as for
of course.
From these numbers of configurations and according to the results of statisticalthermodynamics,we can calculate the entropy of mixing:
S

= k(ln Q1,2 - In Q

where k is the Boltzmann constant.We replace


the Stirling approximation:

~ -JIn Q2,J
Ql,l, Q2,2with their values and apply

lnN!=NlnN-N
to arrive at the equation:
SM
-=

-(NI In Nl N1PN2 +Nzln Nl


+

+ PN2

Figure 7.9 Schematic representation of the Flory reticular model.

7. Deviations from ldeality in the Liquid Phase

225

As each site has the same volume, the ratio of molecular volumes v; is equal to p , and
we obtain the volumetric fractions:

We thus arrive at the expression for the entropy of mixing for one mole:
SM

- = -(xl In Qil

+ x2 In a2)

(7.7)

This expression differs from the entropy of mixing of the ideal solution, which is written
as:
= -(xl In x1 + x2 In x 2 )

S L l

R
such that excess entropy is equal to:

where:

v=xlvl +x2v2

To this entropy term, which we shall generalize for a mixture of n components, corresponds a component of excess Gibbs energy:

This term is always negative. By applying Equation 5.75:

(5.75)
we get the following expression for the activity coefficients (which are always less than 1):
Qii

Qi

v;

Xi

2,

In *I= In - + 1- - =In
Xi

v.
+1-2
v

(7.10)

In particular, at infinite dilution in a binary mixture:


(7.11)
since the ratio v;/v;

is equal t o p , we ultimately have:


1
1
lny,"=ln-+l-P
P

and

lny;=lnp+l-p

226

7. Deviations from Ideality in the Liquid Phase

As an example, table 7.6 shows the values that these activity coefficients may take as a
function of the molar volume ratio:

rl

2
5
10
100
lo00

0.82
0.44
0.245
2.7 lo-*
2.7 10-3

Y2m
0.74
0.09
1.2 10-3
10-40
10-430

It goes without saying that we can only take these results as a qualitative indication of
the large deviations from ideality caused by the differences in molar volume. The hypotheses on which the model is based (analogous lattice for a crystalline system, linear molecules, etc.) are debatable.
In order to account for molecular form, it has been proposed to characterize them both
by their volume and by their external surface. If we have linear molecules, these two properties remain proportional when the chain length increases, for example, in the n-paraffins.
It is not the same, if due to cyclization or branching the molecules have a more compact
form, or even globular. If we designate the external surface of the molecule of component
i as qi, and the corresponding surface fraction as 0,:
(7.12)

the model proposed by Stavermann [I9501leads to the equation:


(7.13)

where we also recognize the term proposed by Flory. The Stavermann model was utilized
in the UNIQUAC and UNIFAC methods, which we shall introduce later.
To illustrate the deviations from ideality for polymer solutions, we must also take into
account the fact that the proportion of free volume (meaning volume not occupied by
the molecule itself) is generally smaller for a polymer than for a solvent (see Chapter 11).
In any event, we may not overlook the role played by molecular interactions and the
enthalpic component. A simple means of accounting for them consists of expressing the
enthalpy of mixing using the equation:

hE = R T x ( x , + p ~ , ) @ , @ ,

7. Deviations from Ideality in the Liquid Phase

227

where x, called the interchange parameter, is empirical, and must be determined from
experimental data. In this case, for a binary mixture the excess Gibbs energy is:
(7.14)
In conclusion, the term entropic or combinatorial that we have just introduced is
only one component of non-ideality.

Molecular interactions play a role in the energy balance accompanying the mixing
process. They exert a poorly understood but undeniable influence on the surroundings at a
short distance from the interaction focal point (molecule or group). The models that we
shall now consider are focused on their contribution.

7.4

THE CONCEPT OF LOCAL COMPOSITION

The notion of local composition was introduced by G.M. Wilson [1964].It is the source of
the models deemed best for the correlation and prediction of deviation from ideality. They
include the Wilson equation, the NRTL model, and UNIQUAC.
If we look at a binary mixture and first examine the system formed by the components
before mixing, and then after mixing (Fig. 7.10), we note that:
0 Before mixing, each molecule is surrounded by similar molecules, and exhibits with
each one of them an interaction energy E,,, for the first component and ~ 2 for
, ~the
second. The numbers of closest neighbors, z , and z2, are called the coordination numbers;
0 After mixing, z ; molecules are distributed around each molecule of component 1 and
the composition of this environment depends of course on the overall composition,
but need not be identical to it. The molecules of a polar component may have the tendency to group due to the effect of orientation forces, and in this way exclude the molecules from an apolar component from their neighboring area. This tendency may
continue until demixing.We denote the mole fractions of components 1and 2, x2,, and
x1,,, respectively, around a molecule of component 1.The description of a center of
attraction formed by a molecule of component 2 will be similar and will introduce the
,
x ~ ,As
~ .for the molecular interactions in the mixture, they are
quantities z ; , x , , ~ and
of three types according to the nature of the centers of attraction: ql, ~ 5 when
, ~ the
centers are the same, and E ~ =, ~ ~ 2 if, they
~ are different. Of course, the values introduced here correspond to averages, and should only be considered as the parameters
of a model.
On these bases, we shall establish the energy balance of the mixture with n, molecules
of component 1with n2 molecules of component 2.
Beginning with pure component 1, in order to extract one molecule, it is necessary to
furnish an energy equal to z , ~ , , ,and
, its introduction into the mixture translates into an
interaction of z ~ ( x , , , +
~ ~x2,1%,1)
,~
or a variation equal to z , ~ , -, ~z; ( X ~ , ~ E +
~ , x, ~ , , % , ~ ) .
Similarly, for a molecule of component 2, the transfer is accompanied by a variation in

228

7. Deviations from Ideality in the Liquid Phase

For one mole of the mixture the total energy


energy equal to ~ ~- z; 5
( x ~,, ~ +
~
E ~ , ~
variation, using 9(for Avogadros number, is therefore written:
1

LW=

2 b , [ z l & 1 , 1 - Z ~ ( ~ 1 , 1 & 1 , +1 x 2 , 1 5 , 1 ) 1

+ n,[Z25,2-Z;(X1.2&1,2

+x2,25,2)11

where the term 1/2 is introduced since otherwise each interaction is counted twice.
Such an expression contains too many inaccessible quantities to be truly useful. We
must introduce approximations and propose a way to evaluate local compositions. We also
have to move from energy balance to excess Gibbs energy. Each of these steps will establish an empirical model whose value will depend on the physical meaning and on its predictive power.
Firstly, we will admit that the degrees of coordination are not modified by the operation
of mixing. Furthermore, around a given molecule, the sum of the local compositions is
equal to 1:
and:

Figure 7.10 Concept of local compositions.


XI

= XZ = 0.5;

XI,] =

2
-

9X2J

=-

TX1.2

1
65 ,x2,2 = 6.

such that for one mole of mixture, the expression for the excess internal energy, or internal
energy of mixing, becomes:
1

2 9([ Z l X l X 2 , 1 ( 5 , 1 - E1.1)

U E = U M = --

+ Z2X2X1,2(&1,2

-5,211

(7.15)

Note that only the differences between the molecular interactions, 9,1- q l ,q2- ~ 2 , ~
appear in this expression.

7. Deviations from Ideality in the Liquid Phase

229

7.4.1 The Lattice Model


In order to further simplify this internal energy of mixing, we concede that the components, whether in the pure state or in the mixture, have identical molecular volumes. Each
molecule occupies one segment in the ordered lattice as the one pictured in Figure 7.9
since this lattice is unchanged upon mixing.
It is also acceptable that the local compositions are identical to the global compositions.
Equation 7.15 may be written as:

To get excess Helmholtz energy, we apply the Gibbs-Helmholtz equation:

aand thus

aE
RT
=

uE

d(

f )+ const.

TO

If we admit that .qj interactions (and therefore the internal energy of mixing) are not
temperature dependent, we obtain:
aE
RT

uE
+
RT

- = - const.

The integration constant is obtained by considering the limit as T + -. The excess


Helmholtz energy is then equal to excess entropy, which is zero due to the hypothesis
established for the local mole fractions (the mixture is random). Therefore:

Accounting for the lattice hypothesis that is identical for the pure substance as well as
for the mixture, the excess volume is null, and we end up with the same expression for
excess Gibbs energy:

The term A12,related to the molecular interactions,may be determined from the liquidvapor equilibrium experimental data, as we have seen previously (Section 7.2,Example 7.1).

7.4.2 The Quasi Chemical Model


Guggenheim [1952] introduced the quasi chemical model, and its main principles have
been presented by Prausnitz et al. [1986]. This model may be considered a special case of
the models for local mole fractions [see Panayiotou and Vera, 1980; Vera, 19861. The

230

7. Deviations from Ideality in the Liquid Phase

establishment of molecular interactions upon mixing is expressed as a reversible chemical


transformation:
[1-11 + [2-21 + 2 [l-21
If nY is the number of neighboring ij in the mixture containing n1 molecules of component 1,and n2molecules of component 2, then the equilibrium condition can be expressed by:

:,z'

=K

n1,1n2,2
The equilibrium constant K is related to the energies of interaction E ~by
, ~the equation:

K1,2= 4 exp 21,2

-2;

- %,2

Considering the material balances acting on each component, we may calculate the values for the neighboring nkjand the local mole fractions.
This model as well as several of its variants that have been published, are undoubtedly
less empirical than the Wilson, NRTL, or UNIQUAC models. However, it introduces certain practical disadvantages that have limited its use. Firstly, it has only one parameter (the
equilibrium constant) per binary, which detracts from its flexibility. Secondly, its application to a mixture with n components necessitates solving a system of equations of the second order, which is less desirable from the point of view of efficient calculation. It has been
applied in group contribution interpretations by Kehiaian [1983,1988], and by Panayiotou
and Vera [1980] and High and Danner [1990] to describe polymer solutions, as we shall see
later on (Chapter 11).

7.4.3 General Remarks


The most commonly applied models, regular solutions, the Wilson model, NRTL, UNIQUAC, and UNIFAC may be generated from the local composition concept, such as
expressed in Equation 7.15:
(7.15)
These concepts will be discussed later. For now, we will mention that this equation uses
two parameters related to the differences between the energy interactions q1,E ~ ,and
~,
~ 2 , The
~ . ratios of the local mole fractions in these models will be expressed with these differences. For example, for a binary system (see Section 7.6.3), we have:
X2,l

x2

- - z2,1
X1,l
x1

where

22,1= exp

(-

It has been emphasized [Flemr, 1976; McDermott and Ashton, 19771 that such a definition does not respect the material balance imposed on the local mole fractions:
XlX1,lf

x2x1,2 = x1

and

XlX2,1+

x2x2,2= x2

7. Deviations from Ideality in the Liquid Phase

231

Taking into account the fact that the sum of the local mole fractions is equal to 1:

+ x2,1= 1

x1,2+ x2,2= 1

and

we derive the following equation:


XlX2.1

= x2x1,2

and one condition relating to parameters 21,2and


1
-1

- X1

71.2

1
--

x2

22,l

that is not fulfilled by the preceding definition equation for these parameters. On the other
hand, the quasi-chemical model does comply with the condition [Panayiotou and Vera,
19801.
We need to also mention that the concepts we have presented do not include free volume, which in the case of a lattice model, is depicted by the difference between the volume
of each lattice segment and the true volume of the molecules.We shall return to this point
in Chapter 11,which is dedicated to polymer solutions.

7.5

REGULAR SOLUTIONS

The theory of regular solutions was introduced by Hildebrand [1924, 19701, and then
Scatchard [1931]. It would be more natural and historically accurate to introduce this theory
as being an outgrowth of the van der Waals equation of state, of the mixing rules relating to
it (see Chapter 8), and of the work of van Laar [1910,1913].However, it may also be developed from the concept of local composition [Moelwyn-Hughes,1961,p. 7741 by calling upon
the following hypotheses:the local compositionswill be taken as the volumetric fractions:
x2,1 = @2, x1,1

= @I,

x1,2

= @I,

x2,2 = @2,

and the expression 7.15 for the excess internal energy becomes:

Inspired by the expression for the dispersion energy and therefore limiting the application of this model to mixtures of apolar compounds (see Chapter 3, Introduction) we have:

to obtain:

232

7. Deviations from Ideality in the Liquid Phase

but the terms between the square brackets may be expressed as a function of the residual
energy of the liquid (changed sign):

and the ratio of these residual energies and the volume illustrates what we have termed
"the cohesive energy density of the liquid".
Note that the law of composition defined in this way is especially simple, and may be
summarized in the following expression:

The solubility parameter of a compound will be defined by the equation:


(7.16)
and we shall therefore express the excess internal energy in the form:
U E =v@1@2(61
- S2)2

(7.17)

To get the excess Gibbs energy, we state that the excess volume is zero. Excess internal
energy and excess enthalpy (heat of mixing) are therefore identical. Finally, we neglect
excess entropy such that we get:

gE=v@1@2(61-62)2

(7.18)

Note that this model can predict only positive deviations from ideality.
This expression may be extended to a mixture of n components in the form:
(7.19)
The corresponding expressions for activity coefficients for a binary mixture are:
In

v;
x=(6, - ~ 5 ~@;
2)~
RT

and

In y2 = - (6,- 62)2@;
RT

(7.20)

and for n components:


V .*

h y i = 2 (6-6 )2
RT '

(7.21)

where 6
, is the average solubility parameter of the mixture, calculated by weighting the
solubility parameters of the components with their volumetric fractions:

sm=~@i~

(7.22)

We recognize the striking feature of this model; it requires only easily accessible quantities
such as the molar volumes of pure substances in the liquid state, w:, and the solubilityparametersnese are defined by Equation 7.16 and may be calculated from the heat of vaporization:

233

7. Deviations from rdeality in the Liquid Phase

(7.23)
Their values are available in the literature and in the data banks. For example,below we
give the values from Chao and Seader [1961], who applied the theory of regular solutions
to the calculation of liquid-vapor equilibria.
Table 7.7
Molar volumes and solubility parameters [Chao and Seader,19611
Compound
Ethane
Propane
n-Butane
n-Pentane
n-Hexane
n-Heptane
Ethylene

V*

(cm3.mo~-1)
68
84
101.4
116.1
131.6
147.5
61

12.38
13.09
13.77
14.36
14.87
15.2
12.44

Propene
Butene 1
Pentene 1
Cyclohexane
Benzene
Toluene
p-Xylene

79
95.3
110.4
108.7
89.4
106.8
124

13.15
13.83
14.42
16.77
18.74
18.25
17.94

We note that these authors took these parameters to be independent of temperature,


and they included light hydrocarbons in the range of application for the proposed method.
The examination of the values of the solubility parameters is sufficient for predicting
the importance of the deviations from ideality. So we can state that the values are not very
different from one n-alkane to the next. It is the same for paraffin to olefin. On the other
hand, the solubility parameters of the aromatics are clearly higher. On the basis of this
model, we conclude that aromatic and paraffin hydrocarbon mixtures are not ideal.
Besides, we know this from experimental values of the activity coefficients at infinite dilution (Table 7.1).

EXAMPLE7.2

Calculation of the solubility of ethane in benzene and


n-Heptane using the regular solutions model
The solubility of gases in the liquid phase depends closely on the deviations from ideality of the liquid solution. We will look at the case of ethane (component 1)at 25C
and at atmospheric pressure dissolved in either benzene (component 2) or n-heptane
(component 3).
The calculation of the equilibrium coefficients will rely on the application of the general Equation 6.4:

234

7. Deviations from Ideality in the Liquid Phase

Of course, in applying this equation, we must take into account the values for the
vapor pressures of ethane, benzene, and heptane. These values have already been
mentioned: for ethane Py = 4.1876 MPa (Table 2.2), for benzene Pz = 12.692 kPa
(Example 4.6), and finally for heptane, application of the Antoine equation
(Table 2.5) yields a value of P; = 6.09 kPa. Furthermore, the total pressure is equal to
1atmosphere.From these values, we can conclude that the fugacity coefficients at saturation for benzene and n-heptane are very close to unity (since the saturated vapor
may be easily considered as an ideal gas). The same is true for the Poynting corrections relative to these two components, as the total pressure differs little from the
vapor pressure and for the fugacity coefficients in the vapor mixture, as the total pressure is low. We therefore rewrite the preceding equation in the form:

In the case of ethane, we must, however, calculate the fugacity coefficient at saturation and the Poynting correction.This was done in Example 2.3 to find cpp= 0.689.The
Poynting correction will be calculated using the molar volume data of the pure substance in the saturated liquid phase, vLra = 95 cm3 . mol-' (Example 2.3) according to
equation:

4 = exp

v,"*(P - Pi",
95*10-6(101325 -4187600)
= exp
= 0.855
RT
8.3145-298.15

Therefore, ultimately:
K, =

4 187600
*0.689*0.855~
= 24.34./,,
101325

12692
6 090
K2= -*/2 = 0 . 1 2 5 6 ~ ~ K3 = -y3 = 0.060 1%
101325
101325
The calculation for the solubility of ethane in benzene will naturally result from the
application of Equation 6.28:
xl=

1-K2
~

K1- K2

(6.28)

However, we note that the equilibrium coefficients depend on the composition


through the activity coefficients, and are therefore a matter of iterative calculation.
This calculation is begun by assuming the solution to be ideal, which provides us with
an initial value for the composition in the liquid phase. It then continues by taking
into account the activity coefficients until the composition stabilizes. Since the

235

7. Deviations from Ideality in the Liquid Phase

mixture is made up of hydrocarbons,we apply the theory of regular solutionsin order


to calculate the activity coefficients:
In

=v; (6, - 6z)20,2

RT

and

In y2 = 4 (6, - 62)2@7
RT

(7.20)

and use the values of the solubility parameters and molar volumes suggested by Chao
and Seader (Table 7.7) to yield:

and:

In % =

68
(12.38 - l8.74)@; = 1.1090;
8.3145 *298.15

In y2 =

89.4
(12.38 - 18.74)207= 1.45907
8.3145 ~298.15

Table 7.8 below shows the results for each iteration.

Xl

0.036 1
0.012 6
0.012 2
0.012 1

@l

0.027 7
0.009 6
0.009 3

*/1
1
2.853
2.967
2.97

rz
1
1.001
1.OOo 1
1.OOo 1

Kl
24.34
69.44
72.23
72.29

K2
0.125 6
0.125 7
0.125 6
0.125 6

The solubility of ethane in benzene at atmospheric pressure and 25C is thus close to
1.2% (in mole fractions).
If we repeat the calculation with n-heptane instead of benzene as the solvent,the
activity coefficient value for ethane is considerably modified:
68
In /1=
(12.38 - 15.2)0; = 0.218 0 ;
8.3145 e298.15
and the solubility of ethane, calculated as above, is close to 2.9%.
Note that due to the very different volatilities of the components (ethanehenzene or
ethaneln-heptane), we could have begun the calculation at infinite dilution. In this
case,taking benzene as the solvent,the first iteration would have yielded a result close to:
1- 0.125 6
x.=
= 0.0118

mvl YL

and a second iteration would have been practically useless.


We also note that the molar volume value for ethane that we used for the calculation
of the Poynting correction is not the same as the one applied to the calculation of the
activity coefficients. In the first case, we took the value corresponding to saturated
ethane at 25C. In the second case, we used the value recommended by Chao and
Seader,which corresponds more to the properties of ethane at its boiling temperature
at atmosphericpressure.

236

7. Deviations from Ideality in the Liquid Phase

This application demonstrates the importance of the deviations from ideality in the solvent power of a compound. Solubility is high when the deviations are low (and possibly
negative).This influence is apparent in liquid-liquid solubility,solid-liquid solubility, as
well as in gas-liquid solubility.
For the theory of regular solutiom, we need to retain the following essential items:
Range of application: mixtures of apolar compounds (especially hydrocarbons)
0 A generally satisfactory precision in terms of the d d a t i o n of activity coefficients
0 Predictive nature.
Undoubtedly, due to this predictive nature but also because the entropy term was neglected, we cannot expect a perfect estimation of liquid-vapor equilibria, and as for the
key pairs of a process, it is appropriate to go back to the experimental data. The example
illustrated in Figure 7.11 [Jose et al., 19921 for the benzene n-tetradecane system clearly
300
gE

T=283.15

gE T=293.15
gE

T=303.15

+ g E T=308.15

200

gE T=313.15

+ gE

T=323.15

o gE T=333.15

r
h

100
0,

gE

gE T=353.15

gE T=363.15

T=343.15

Q gE T=373.15

-100

-4
0

0.2

0.6

0.4

0.8

X1

Figure 7.11 Excess Gibbs energy of the benzene (1)n-tetradecane (2)


system [Jose et al., 19921.

237

7. Deviations from Ideality in the Liquid Phase

shows that the theory of regular solutions cannot claim to represent the variation of excess
Gibbs energy of such mixtures as a function of composition and temperature.
Numerous modifications have been proposed to improve its accuracy.The most simple
consists of coupling the Flory theory (Section 7.3) and the Scatchard-Hildebrandtheory
within the excess Gibbs energy expression by stating:
(7.24)
As shown in the example from Figure 7.12, the calculation of the activity coefficients
approaches experimental data. We must, however, point out that application of the Flory
theory to mixtures containing aromatic hydrocarbons is difficult to defend.
One modification,perhaps more empirical, consists of stating the expression for excess
Gibbs energy as:

(7.25)
where the binary parameter ki,jis determined from experimental data. Note that the law of
variation with composition remains identical to the one predicted by Equation 7.19, with

v)

.-(u
0

2?
.>

/
0

1.5

Mole fraction of n-hexane

Figure 7.12 Calculation of the activity coefficients for the n-hexane (1)
benzene (2) system at 55C. Regular solutions (- - - - -); Flory
entropy term + regular solutions:(
>.

238

7. Deviations from Ideality in the Liquid Phase

only the amplitude of the deviations from ideality adjusted to the experimental data using
the k , parameter.
As we stated at the beginning of this section,the regular solution method may be developed from the van der Waals equation of state. It also results from the van Laar [1920,
19131 equations that describe the variation of activity coefficients with composition using:

In

A1,2
Al2x1)

(7.26)

and

1+A2,lXZ

Parameters Alz and A2,,are determined from experimental data. However, if we state:

we come back to the expression 7.19, which is predictivein the sense that the parameters
are calculated from the properties of the pure substances.

7.6

EMPIRICAL MODELS BASED ON


THE CONCEPT OF LOCAL COMPOSITION

Whether formulated from the van der Waals equation of state or from the concept of local
composition, the regular solution theory distinguishesitself by its predictive capacity. Only
the properties that relate to the pure components are needed for its application. This is not
true for the methods that we will discuss now. They are characterized by the expressions of
excess Gibbs energy containingbinary parameters whose values must be determined from
experimental data, coming essentially from binary liquid-vapor equilibria. The process
therefore appears similar to the one we used in Section 7.2 and would limit itself to a single correlation of data if these methods did not lead to the calculation of high order, ternary, or n component system equilibria once the parameters are identified. Such is the
character of the Wilson, NRTL, and UNIQUAC methods that play an essential role in the
practical calculations of liquid-vaporequilibria.

7.6.1 The Wilson Equation


The model proposed by Wilson is based both on the concept of local composition and the
Flory theory. It is based on the assumption that Equation 7.9, to which the development of
this last theory leads, remains valid in the presence of molecular interactions with the
condition that the volumetric fraction Oi is replaced by the local volumetric fraction of
component i surrounding a molecule i, {i,i:
gE=

R T ~ X , tiI, i ~
Xi

239

7. Deviations from Ideality in the Liquid Phase

These local volumetric fractions are calculated from the molar volumes of the components w;, and the intermolecular interaction energy,

.;xi exp

t..=
41

(-2)

[w;xjexp

(-$)I

;=l

We state:

such that the excess Gibbs energy expression becomes:


n

g E = -RT

xi In

(7.27)

i=l

and the activity coefficients are expressed by the equation:


(7.28)

Parameters Ai,jor
are determined using experimental data (note that the Ai,; parameters must always be positive).The Wilson equation has been applied to a great number of
polar or non polar systems. It allows for their correlation in a generally very satisfactory
way. The prediction of liquid-vapor equilibria of high order systems is generally considered
weak. We can show that, simply due to its mathematical form, the Wilson equation may
never be used to illustrate liquid-liquiddemixing.

7.6.2 The NRTL Equation


The NRTL (Non Random Two Liquids) equation proposed by Renon and Prausnitz
[1968] is based on the expression for the internal energy of mixing as a function of local
compositions discussed above:
1
uE = uM = -2

A! tZlXlX2,1(%,1 - E1,J + Z2X2X1,2(%2

- %,2)1

Given that the degrees of coordination z1 and z2 are equal, and stating:

cj,i= --21 %Z(&. Id. - &.C I.)


it can be generalized to any mixture:

(7.15)

240

7. Deviations from Ideality in the Liquid Phase

As was done by Wilson, the local compositions were calculated using the Cij parameters. A third parameter called non-randomness parameter, which is in fact empirical, is
introduced to yield:
xj exp
x . .=
I,

(- aj,2)

n
k=l

where aj, = ai,j .


It is inserted into the expression for excess internal energy. As in the theory of regular
solutions, we neglect the excess entropy and excess volume terms and thereby obtain the
value of excess Gibbs energy:

(7.29)

The activity coefficients are expressed by the equation:

(7.30)

(7.31)

where:

For every binary system, the NRTL equation therefore has three parameters that must
= %,l. In addition, it has been suggested
be adjusted to experimental data C1,2,C2,1and q,2
to account for the variations of these parameters with temperature by stating for instance:

c I.d. = CI!d ? ) + c1 9 1! ? ( T - T , )

and

a],.I . = a ]!> I? ) + aI d! ? ( T - T , )

(7.32)

We then have six parameters. Deriving the 8IRTratio with respect to temperature and
applying the Gibbs-Helmholtz equation, we obtain the expression for excess enthalpy
(heat of mixing), and the simultaneouscorrelation of liquid-vapor equilibria and the heats
of mixing is possible (Fig. 7.13).
In fact, the number and the character of the parameters that we may determine are mainly
dependent on the number and the character of the data we use, and the amplitude of the deviations from ideality,as shown in Table 7.9 [Renon et al.,1971,p. 301.We may account for parameter variations with temperature only if the liquid-vapor equilibrium data extend over a wide
temperature range, or if we can associate them with heat of mixing data. Even if we limit ourselves to three parameters, the application to moderately non-ideal or weakly polar systems
will show that these parameters are correlated. It is therefore suggested to fix parameter a,
since the standardvalues are equal to 0.2 for weakly polar systems, and 0.3 for polar systems.
It has also been suggested that this same parameter be fixed at -1 [Marina and Tassios, 19731.

7. Deviations from Ideality in the Liquid Phase

241

a.

b.

Figure 7.W Use of the NRTL model to the calculation of (a) liquid
vapor equilibrium at atmospheric pressure, and (b) heat of mixing
for the acetone (1)water (2) systems [Renon et al., 19711.

The NRTL equation is also applicable to the calculation of mutual solubilities in the
liquid phase. As shown in Table 7.10, it allows for a generally very good representation of
liquid-vapor equilibria (bubble pressures and composition of the vapor phase), heats of
mixing, mutual solubilities and activity coefficients at infinite dilution for a large number
of polar or non-polar binary systems.Table 7.11 summarizes a series of trials relative to the

2 42

7. Deviations from Ideality in the Liquid Phase

Table 7.9
Determination of the NRTL.model parameters as a function
of the nature of the data and the degree of non-ideality [Renon et al., 19711
Available Data
Order of
magnitude
of g E

g:,

Weak
< 0.30 RT

gE,hM
over a wide range
of temperatures

gE,hM
at a single
temperature

gE
at a single
temperature

6 parameters

4 parameters

2 parameters

(benzene-n-heptane)

a$2= 0.2 a:,?= 0


(water-acetic acid)

c;,,= CT2 = a;,z = 0

a$,z= 0.2
(benzenedimethylsulfoxide)

Strong
g",,, > 0.30 RT

6 parameters

6 parameters

(acetone-water)

(ethanol-cyclohexane)

3 parameters

c ; ,=~c T , ~ =

Table 7.10
Use of the NRTL model to the calculation of liquid-vapor equilibrium pressures,
the heats of mixing, the liquid-liquid equilibria, and the activity coefficients at infinite dilution
Systems

5 cal Quantity Deviations


pc

Character

Hydrocarbons
and hydrocarbons
or carbon
tetrachloride
Alcohols
and hydrocarbons
or carbon
tetrachloride
Polar
and hydrocarbons
or carbon
tetrachloride
Alcohols and alcohols
Alcohols and polars
Polars and polars
Alcohols and water
Polars and water

loo P"

0.5

0.6

1.5

1.3

1.5
3
3
1
1.5
1.5

1.5
2
1
0.5
2
1.5

(n-hexane-ethanol)

Table 7.11
Application of the NRTL model to the correlation and prediction of phase equilibria, heats of mixing,
and activity coefficients at infinite dilution for the acetone (1)n-hexane (2) system

Average Quadratic Deviations


between Calculated and Measured Quantities (0)

Data Used
T ("C)
P(atm)

loo (+YE")

5
(caUmol)

(Y")Ex
Number
of points
Trial 1

1.5

0.9

0.8

1.2

Trial 2

1.5

0.9

1.4

2.0

Trial 3

Trial 4
Trial 5
Trial 6
Trial 7

1.8

E
27.2

1.1

0.8

0.8

0.9

0.5

1.5

1.5

3.0

0.5

0.9

1.2

40.7

13.7

9.0

0.5

1
0.8
I
0.9
I
3.5
I
1
7
.
7
1
1.0

+
133.8

h,

t
i

244

7. Deviations from Ideality in the Liquid Phase

determination of the six parameters of the acetone, hexane system. We use the liquidvapor isothermal equilibrium data between -20 and 55C,as well as activity coefficients at
infinite dilution for hexane in acetone at atmospheric pressure, the mutual solubilities in
the liquid phase between -90 and -4OC, and finally the heats of mixing.
From one trial to the next, we have used a portion of this database for the parameter
calculation, and we can then examine to what extent the data not used in the adjustment
are predicted. In particular, we observe that from the data for heat of mixing and the
isothermal liquid-vapor equilibrium data at 20C only, the full data set is predicted satisfactorily (Trial 2). On the other hand (Trial 7), using the mutual solubilities only is not
appropriate to predict either liquid-vaporequilibria,or the heats of mixing.
It is then possible to calculate these same properties for higher order mixtures, with n
components. However, we must emphasize the fact that the simultaneous correlation of
the liquid-vapor and the liquid-liquid equilibria is more delicate. The prediction of miscibility gaps for complex systems most often requires that ternary data be available. A
detailed study of the possibilities of this method and the calculation programs that relate
to it have been published [Renon et al., 19711.
The local environment could have been defined in terms of volume fractions, as proposed by Bruin and Prausnitz [1971].This approach has the advantage of containingas a
special case, the regular solution model if the a parameter is null. Under these conditions,
for a complex mixture the parameters that relate to the binary hydrocarbons may be calculated a priori.

7.6.3 The UNIQUAC Model


The UNIQUAC, Universal Quasi Chemical, model [Abrams and Prausnitz, 1975;
Maurer and Prausnitz, 19781 is itself also based on the concept of local composition.
However, it expresses the energy balance of the mixing operation by taking into account
the external surfaces of the molecules.The molecule of component i is broken down into ri
segments, and its external surface is described by a parameter qi. Using z to denote the
coordination number, the molecule of component i is in contact with zqi segments belonging to neighboring molecules. If we take component i as being pure, in order to remove a
molecule from its environment,we must furnish energy equal to z qi qi.In the mixture, it
is surrounded by zqi segments in proportion to )i.Its condensation will be accompanied
by an energy transfer equal to Cjz qi 8j,iE ~ , ~ .
The equation (analogous to expression 7.15) expressing the internal energy of mixing is
therefore:
n

We state:

7. Deviations from Ideality in the Liquid Phase

245

and with 6, as the global surface fractions:

we determine the local environment with the equation:

where:
resulting in the expression:

In opposition to the approach used for the NRTL model, and this is essential, we apply
the Gibbs-Helmholtz equation in order to calculate the excess Helmholtz energy:

setting the integration limit Toapproaching infinity. At high temperature, it is assumed that
the interaction term is zero such that the integration constant is obtained from the combinatorial entropy expression of Stavermann-Guggenheim (Eq. 7.13):
(7.13)
As the integration result relates to Helmholtz energy, it is assumed equivalent to Gibbs
energy since the excess volume is low, and it is designated using the term residual excess
Gibbs energy:
n

(7.33)

so finally we have:
gE -- g Ecombinatorial -tgresidual
E

(7.34)

246

7. Deviations from Ideality in the Liquid Phase

The activity coefficients are calculated by applying the following equations:


In

= In &ombinatonal + In %,residual

1. = z (ri - qi) - (ri- 1)


2

where:

and

z = 10

(7.35)

(7.36)

(7.37)

The UNIQUAC model therefore deals with two types of parameters. The first are representative of the volume (ri) and surface (qi) of each component and were calculated
from the volumes and the surface of the molecules proposed by Bondi [1964,1968]. The
, ~ T,~)are binary parameters and must be calculated by correlation with
second ( A u ~or
experimental data. It has been proposed to account for their dependence on temperature
using:
(7.38)
As with the NRTL equation, a detailed study of the UNIQUAC model and the calculation programs necessary for its use have been published [Prausnitz et al.,19801.

7.6.4 The Wilson, NRTL, UNIQUAC Models

Conclusion
The three models that we have just introduced are derived from the concept of local composition. They have been established based on different hypotheses concerning the relationships between internal energy of mixing and excess Gibbs energy (we may speak of
enthalpic models (NRTL) or entropic models (Wilson, UNIQUAC)), and from different expressions for the immediate environment of a molecule. The equations (symbols
included), differ, but they may also be compared on the basis of the results obtained and
the application range.
In contrast to the model of regular solutions, they are applicable to mixtures of polar or
non-polar compounds.Their extension to ionic solutions has even been proposed.
They are not totally predictive because the binary parameters (supposedly representative of different molecular interactions) must be determined from experimental data
relating to the binary systems in question. Their value is that they allow predicting
the properties of more complex equilibrium systems, in other words, multicomponent
systems.

7. Deviations from Ideality in the Liquid Phase

247

If we stick to the calculation of liquid-vapor equilibria and the Wilson equation, the
NRTL and the UNIQUAC models yield comparable results, and are generally accurate.
The calculation of heats of mixing by application of the Gibbs-Helmholtz equation is
generally possible only if we take into account parameter variation with temperature,
which doubles their number.
As for liquid-liquid equilibria, we retain that because of its mathematical structure, the
Wilson equation does not represent demixing. On the other hand, the NRTL and UNIQUAC equations may be applied to the correlation and the prediction of these phenomena.
Parameter determination, whose principle was explained in Section 7.2, must adhere to
the rules of common sense; the amount and the nature of the available data must be
respected. For example, we can account for the variation of parameters with temperature
only if we use either equilibrium data within a wide temperature range, or equilibrium
data plus heat of mixing data.
Similarly,if we wish to apply the NRTL equation or the UNIQUAC equation to the calculation of liquid-liquid equilibria, we must not be satisfied with parameters determined
from liquid-vapor equilibrium data. The latter, within the full miscibility region, may be
well correlated or predicted, and the solubility boundaries represented in a qualitative
fashion only because of the extreme sensitivity of this type of equilibrium to activity coefficient values. The experimental database must therefore include binary liquid-liquid equilibrium data and, preferably, ternary data, for there to be any hope of an accurate prediction of the n component mixture.
Endowed with three parameters, the NRTL model is, in principle, able to be applied to
a very large spectrum of deviations from ideality. This is also a weakness because their
simultaneous determination requires more data.
Finally, we shall be cautious for the risk of parameter intepcorrelation. For a given
binary data base, it may happen that an entire set of parameter pairs represents the experimental base equally well. In a way, too many parameters are dealt with. Application of the
models just discussed to mixtures of non-polar compounds, and especially hydrocarbons, is
an example of such a phenomenon. To eliminate this lack of a unique solution, we may be
led to insert a relationship between the parameters, their equality, for example, in the case
of the UNIQUAC [Soave, 19921 model, or to enrich the database with new measurements that are more accurate, or of a different character. For example, an inter-correlation
of parameters relating to some binary systems can be eliminated by using ternary data
related to the same systems.
In any event, application of these models to a complex mixture involves a considerable
amount of correlation. Let us not forget that a system made up of ten components, which
is nothing exceptional in practice, contains 45 binary systems. However, there are databases that, for a large number of systems, specify the values for the parameters of the models we have just introduced in addition to the experimental data. The best known and the
most abundant in data, is the one started by the University of Dortmund [Gmehling J. and
Onken U., 19781.An example of its contents is given in Table 7.12.
Yet it happens that no experimental data relating to the studied system is available. If
this system is one of the key pairs of a separation, we must then go back to the lab.
Otherwise, we can apply a group contribution method.

248

7.7

7. Deviations from Ideality in the Liquid Phase

GROUP CONTRIBUTION METHODS

Even if we exclude ionic solutions, the number of compounds that may occur in a liquid
phase mixture is such that we cannot hope to one day dispose of all the numerical values
for the parameters operating within the models that we just examined. However, we can
note that the molecular interactions on which these models are based lend themselves to a
more refined, and above all, simplified analysis. If we disregard the simplest compounds,
these interactions are in fact the result of interactions that exist between the constituent
groups of the molecules. Therefore, in a mixture of n alkanes, we distinguish between the
CH,-CH,, CH,-CH,,
and CH,-CH, etc. interactions. This analysis is considerably simplified. In effect, the number of atomic groups is much lower than the number of individual chemicals, and we can draw a parallel between the group structures, the compounds,
and the solutions on the one hand, and letters, words, and sentences on the other.
However, there are a number of difficulties:
The first, as we have already pointed out (Chapter 3, Section 3.3), relates to the definition
of groups. Can we say that the terminal methyl group of a paraffin is the same as the one
present in ethanol, acetone, or even toluene? Do the three methyl groups in isopentane have
identical properties? Each group should be defined not only according to its nature, but also
by the nature of its immediate environment within the molecule containing it. We shall thus
distinguish between the CH,(CH,), CH,(CH), CH,(CH,OH), and CH,(C,,) groups in the
previous example. This has been done in some cases (see Chapter 3, Section 3.3.1, Benson
method), but a multiplication of distinct groups results from it, and therefore, parameters
representing their interactions.As these parameters must be determined from experimental
measurements, once again we find ourselves confronted with the problem of availability of
data. The method will undoubtedly prove to be more precise, but less predictive.
Similarly,two contiguous groups (in a)exert an undeniable influence on each other. For
example, we cannot describe the properties of ethylene glycol using two methylene groups
and two hydroxyl groups. This proximity effect blurs once the hydrocarbon chain separating the two functional groups lengthens. It has been more closely studied by Kehiaian [1983].
Some structures, and in particular the first terms in homologue series, evade the group
composition. Sometimes they involve compounds of considerable practical importance.
For example, we can cite methanol, solvents such as NN-dimethylformamide, and the chlorofluorocarbons of methane and ethane, etc. Such components themselves form a group
whose structural properties (volume, surface) are sometimes of a greater order of magnitude than those of other groups.
Despite these difficulties, for binary or higher order systems the group contributions
methods are commonly applied to the prediction of excess quantities with good reason
when we are not seeking the same level of precision afforded by models that rely directly
on the correlation of experimental data.
Special mention must be made of the DISQUAC (DISpersive QUAsi Chemical) model
introduced by Kehiaian [1983,1988].The primary goal is not to provide a predictive calculation method of complex equilibria, but rather to elucidate and evaluate the contribution
of each structure to the various excess quantities (Gibbs energy, heat of mixing, heat
capacity). Emphasis is directed at the irregularities and problems that we have mentioned,
and their interpretation.

Table 7.12
Liquid-vaporequilibrium database [Gmehling et al., 1978-19841

(1) ACETONE

C 3H60

CONST ANT S :
HARGULLS
VAN LAAR
WILSON
NRTL
UNIOUAC

A12

A2 1

962.8113

I . 6352
I . 6399
341.4214

699.6135
-42.4060

599.8541
485.5658

1.5061
1.5055

EXPERIHENTAL DATA
X1
Y I
P MH HG
0.0
0.0651
0.1592
590-20 0.2549
0.3478
611. 3 0
0.k429
632.60
639.60
0.5210
633.80
0.5901
0.6202
631. 1 0
631.00
0.1168
0.7923
621.80
623.30
0.8022
603.40
0.8692
0.9288
583.20
0.9658
543.30
1.0000
505.00
339.40
444.60
545.80

0.0
0.2828
0.4442

0-5163
0.5560
0.5866
0.6068
0.6258
0.6339
0.6662
0-1034
0.1292
0.1583
0.8255
0.9003
1.0000

MEAN DEVIATION:
MAX.

DEVIATION:

ALPHA12

3. LO
10.11
3.32

0.0

-0.0032

2.88
0.04

0.0240

8.65
0.26
3.05
-1.42

6.82
-4.81
-1.38

0.0010
-0.0075
-0.0044
0.0

4. 21

0.0086

-2.33

10.11

5'

-0.0104
-0.0 I63
-0.0111
-0- 01 3 3
-0.0096
-0.0049
-0.0034
0.0011
0.0041

4.90
1.17

8
=

0.4817

MARGUL E S
D I F F P O IFF Y l

4.21

x
1
-

0.0240

V A N LAAR
OIFF P O l F F Y l

4.21
3.25
10.42

3.63
5- I2
7.29
8.10
0.28
3.06
-1.45
2.79
-0.01
-2.52
6.58

0.0
-0.0030

-0.OlOl
-0.01 61
-0.0111
-0.01 35
-0.0100

-0.0053

-5.00
-1.38

-0.0039
0.0001
0.0038
0.0238
0.0010
-0.0012
-0.0041
0.0

4.30

0.0086

10.42

0.0230

WILSON
O l F F P O l F F I1

4.2 1
-3.48
9.40
1.95
10.83
12-06

12.30
3.12
5.69
0.19
4.26

1-22

-3-11
3.82
-7.90
-1.38
b.

I4

12.30

0.0
-0.0122
-0.0065

-0.0063
-0.0018

-0.0074
-0.0083
-0.0071
-0.0079
-0.0069
-0.0037
0.0165
-0.0025
-0.0052
0.0001
0.0

mrL
DIFF P OIFF Y l
4.21

-2.44
9.34

7. I2
9.82
11-16
11.55
2.48
5.09

0.24
3.81
0.80
3.22
4.30
-7.26
-1.38

0.0
-0.0109
-0.0014
-0.0080
-0.0094
-0.0084
-0.0081
-0.00 7 5
-0.0073
-0.0056
-0.0021

0.0182
-0.001 4
-0.0054
-0.0008
0.0

UNIOUAC
OlF F P OI FF Y l

4-21
2-36
10.16
4.01
5.78
1.82
9.07
0.55

3.30
-1.25
2.88
-0.00
-2.11
6-11
-5.38
-1.38

0.0
-0 0042
-0.0098
-0.01*9
-0.0165
-0.0128
-0.0098
-0.0057
-0.0044
-0.0002
0.0029
0.0230
0.OOOT
-0.0069
-0.0035
0.0

0.0071

5.62

0.0012

6.39

0.0082

0.0165

11.55

0.0182

10.16

0.0230

N
P
rD

250

7. Deviations from Ideality in the Liquid Phase

Proposed with the intent of a more widespread application, the ASOG and UNIFAC
methods both adhere to the first and third propositions and hypotheses alluded to by
G.M. Wilson [1962]:
The partial molar excess Gibbs energy:
(5.75)
is the sum of two contributions,the first, referred to as combinatorial,corresponds to
the differences of size and form of the components of the mixture, and the second,
called residual, corresponds to the interactions between groups. We therefore have:

In % = In
0

%,combinatorial-tIn %,residual

(7.39)

The residual term is estimated by substituting the notion of group solution for the
notion of chemical compound mixture. For example,in a mixture containing one mole of
hexane and one mole of acetone,hexane supplies 2 moles of CH, groups and 4 moles of
CH, groups, and acetone 2 moles of CH, groups and one mole of the C=O group.The
molar composition of the group solutionis therefore 4/9 for the CH, groups,4/9 for the
CH,groups and 1/9 for the C=O group. Of course, the components themselves will be
considered solutions of special groups since the composition of hexane is 2/6 for CH,
and 4/6 for the CH, groups, and acetone is 2/3 for CH, ,and 1/3 for the C= 0 groups.
These group solutions are not ideal and each group k is characterized by its mole fraction Xk and its activity coefficient r k . The residual activity coefficients themselves,1;.,
are related to them by the expression:

In

%,residual =

c
k

vk,i

(In rk - In

rk,i)

(7.40)

where vk,i denotes the number of groups k in component i, r k , i the activity coefficient
of group k in the mixture of groups that make up pure compound i, and r k the activity coefficient for the same group in the actual mixture.
The difference between the ASOG and UNIFAC methods results from the models that
have been selected to express the combinatorial and residual terms, as well as the definition of groups. They were the subject of a comparison by Gupte and Daubert [1986] that
does not attribute a decisive superiority to either of them. We shall more fully develop the
more familiar UNIFAC model.

7.7.1 The ASOC Method


Introduced by Wilson [1962] and Derr and Deal [1969], this method has been developed
by Kojima and Tochigi [1979], and then by Tochigi et al. [1981].
We apply the general expression for activity coefficients 7.39 that has just been introduced, and Equation 7.40 for the calculation of the residual term.
The combinatorial term is expressed starting with the Flory theory where vi designates
the number of atoms contained within component i, with the exception of the hydrogen
atoms. We have:

7. Deviations from Ideality in the Liquid Phase

In

%ombinatonal = In

Vi

"i
+1- C ?xi
i

C5.j
i

For the residual term, the activity coefficients of groups


Wilson equation:

25 1

(7.41)

rkare calculated using the


(7.42)

m=l

where X , is the fraction of each group 1 in the mixture:

C Xjv1.j
(7.43)

and q i is the number of groups 1 in component i.


The interaction parameters between groups ak,[are temperature dependent according to:
In aS1 = mS1+

nk
7
1

(7.44)

where mk,[and nk,lwere obtained by liquid-vapor equilibrium data regression.

7.7.2 The UNIFAC Method


The UNIFAC method was proposed by Fredenslund et al. [1975].It flows directly from the
propositions defined by Wilson (Eqs. 7.39 and 7.40) and the UNIQUAC method. The combinatorial term is calculated from the Stavermann-Guggenheim expression:
(7.13)
from which we obtain the expression for the activity coefficients:

I.= - (r i - q i ) - (ri- 1)

where

t = 10

(7.36)

In order to calculate the volume and surface fractions of the components we need to
know the corresponding molecular parameters ri and q i ,these being calculated from the
volume and surface parameters from each group R, and Qk using simple additivity rules:
(7.45)

252

7. Deviations from Ideality in the Liquid Phase

The residual term is calculated from:


(7.40)
we must calculate the activity coefficients r k for the groups.The UNIQUAC method yields
the expression:
(7.46)

where the surface fractions for groups @k,are derived from their mole fraction x
expressions:

by the

(7.47)

and:

(7.48)

The interaction parameters between groups (k;,ldepend on the temperature:


(7.49)
Of course the preceding expressions 7.46 to 7.48 are applied equally to the calculations
of r k and rk,i.
R o categories of groups are defined. The main groups currently number around 50,
and correspond to one type of interaction, and therefore to a particular pair of the aL1
parameters. Interaction parameters can be found between the CH, and the CH,CO
groups, for example. Within some of these main groups, subgroups are defined that differ
by the values of the Bondi Rk and Qk parameter$.For example, we have subgroups CH,,
CH,, CH, and C for the main CH, group, and CH,CO and CH,CO for the main
CH,CO group. There are a total of 108 subgroups. In the appendix of this chapter,
Tables A3.1 and A3.2 (Appendix 3), taken from the work of Hansen et al. [1991] specify the
nature of some of the groups and subgroups,and provide the parameters Rk,Qk,and ak,,.
Since it was initially defined, this method has been regularly refined [Skjold-Jorgensen
et al., 1979;Gmehling et al., 1982;Weidlich et al., 1983;Tiegset al., 1987;Hansen et al., 19911,
and the matrix of interaction parameters determined by correlation of experimental data
has been substantially expanded, as shown in Figure 7.14.
Application to the calculation for liquid-liquid equilibria is possible (see Chapter 9), but
a specially adapted interaction parameter matrix [Magnussen et al., 19811 must then be
used.

0
.
*

5:
%o
p00
$000

!$oooo
~ o o o o o q
~ 0 0 0 0 0 0
~ 0 0 0 0 0 0 0
~ 0 0 0 0 0
00 0

~0.000.000

go.0.00.000.00
.0.0.000000000

g000.0.00000000
$00.0.00000000000
~0.00000.000000000

0'"

~ 0 0 0 0 0 0 0 0 . 0 0 0 0 0 0 0 0 0

~000.000.0.00.00.000

~oOOOOO.000.00000000og
~00000.000.0000.000.00
g000.0000.000.000000000

0 %

~0.0000000000

~..000000000

~ 0 0 0 0 . 0 0 0 0 0 ~

7. Deviations from rdeality in the Liquid Phase

~0000..00000000000000000

0..0.0..0..0.0.00000.000

~.....0000.0..0.0.0.0.0000.0000~

~........000.000.000.000000000

~....00000..0.0.000.00000.000

~..00.0.0.00.0.000.00.000.00

0
.
0
.
.
0
.
.
.
.
.
0
.
0
.
0
0
0
.
.
0
.

$j..00000.0000000.00.oo.ooo$j

g..00....00.0.0.000.0000*

~....
~.......

~..0...00.00000.000..00000000000*
$.....0..0.000.0000000.0*000000000
"0000....000000000.00..00000000000

$ 0 0
0 0
00..000000000000.00.~.
~...0.0.00..000.00000000
.00..

~
.
.
.
0
.
.
.
.
.
.
.
.
.
0
.
.
.
.
.
.
0
.
0
.
.
.
0
.
.
0
0
0
0
0
0
0
0
~.0000...00...000.0000000.0.0.000000000
~..0...0........0.0....0...0~0.00000.000

$...000000....00000000*000..0.0.000000000$

m...0.000..0.000..00*00.000000000.00.00000.
C
.
.
.
.
0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0
.
.
0
0
0
.
.
0
0
.
0
.
.
0
.
0
.
.
0
.
0

0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0
.
.
0
.
.
.
0
.
.
.
.
.
0
.
0
0
0
.
0
0
0
.
.

~..O....O..O...............

"...................

m..........o..o..o.o.
.

N . . . . . . . . . . . . . . . . . . . ~ O ~ O O O ~ O O O O O ~ ~ ~ ~ ~ ~ ~ ~ O O ~ O ~ ~ ~ ~ ~
.

253

254

7. Deviations from Ideality in the Liquid Phase

In addition to these extensions, many variations of this method have been proposed
[Fredenslund and Rasmussen, 19851. In particular, Larsen et al. [1987], Weidlich and
Gmehling [1987], and Gmehling et al. [1993] have modified the combinatorial term and
specified parameter variation with temperature by taking into account the heats of mixing
in the database used for their determination.
Application to polymer solutions is also possible (see Chapter 11).
More recently, it has been demonstrated that this method may be incorporated into
equations of state for the definition of mixing rules. We shall return to this topic in
Chapter 8.

U EXAMPLE 7.3
Calculation of the activity coefficients in
the 3-pentanone n-heptane system
In an effort to explain the UNIFAC method, we shall provide a numerical example.
We shall calculate the activity coefficient of 3-pentanone in an n-heptane solution at
353.15 K. The mole fraction of 3-pentanone is equal to 0.056.
Each of the two components will be first described by its constituent subgroups to
obtain the values for the
coefficients:

3-Pentanone
n-Heptane

and so we have v,,, = 2, v,,, = 1, v , , =~1, v,,, = 2, v,,, = 5 , v , , ~


= 0.
We then look up the values for the volume ( R k )and surface ( Q k )parameters for each
group:
0.901 1
0.848

0.6744
0.540

1.4457
1.18

So, R, = 0.9011, R, = 0.6744, .....Q, = 0.848, etc.


We create the matrix for interaction parameters between groups in the same fashion.
Note that the CH, and CH, groups are two subgroups derived from the same main
CH, group. The interaction parameters with the CH,CO group therefore have the
same value, and their reciprocal interaction parameters are zero.

CH,
CH2
CH2C0

26.76

26.76

476.4
0

7. Deviations from Ideality in the Liquid Phase

255

Calculation of the combinatorial term

The volume and surface parameters of the components are calculated by applying
equation 7.45:
rl=

2 vk , l Rk --2.0.9011

+1*0.6744+1-1.4457=3.9223

We find:
rl = 3.9223 q1 = 3.416
r2 = 5.1742 q2 = 4.396.
We now derive the volumetric and surface fractions of the components:

0.056.3.9223
xlrl
= 0.043 033
0.056.3.922
3 + 0.944 .5.1742
xlrl + x2r2
0.056 * 3.416
el = x141 -0.056.3.416
+ 0.944.4.396 = 0.044066
xlql + x242
Q1 =

as well as the li terms defined by Equation 7.36:

10
2

1 - - (rl - q l ) - (rl - 1) = - (3.9223 - 3.416) - (3.9223 - 1) = 4.3908

-2

we find that l2 = -0.2832, and finally:

= In

0.043033
0.056

10

+3.4161n
2

0.044066
0.040 033
- 0.3908 +
0.043 033
0.056
(0.056*0.3908+ 0.944.0.2832) = -0.027

Calculation of the residual term

Using Equation 7.49, we shall find the interaction terms Ymnas follows:
v1.3 = exp

( li:t )
--

= 0.2595

The table below summarizes the results that relate to this term:

CH,
CH2
CH2C0

0.927

0.927

0.259 5
1

First, we must calculate the activity coefficients for each of the three groups in the
pure component 1,l-k,l, k = 1,3. The mole fractions of these groups are obtained by
Equation 7.48, applied to the component in question only:

256

7. Deviations from Ideality in the Liquid Phase

X1,l

v1.1
vl,l + '2,l

= 0.5,

+ '3.1

X2.1

= 0.25,

X3.1

= 0.25

and we derive the surface parameters:


0
'9'

0.848 * 0.5
0.848.0.5
+
0.540-0.25
+ 1.180.0.25 = 0.496 5
- Q1Xl,,+ Q2X2,,+ Q3X3,,
-

Q1xi.i

02,1

= 0.158 1,

@3,1

= 0.3454

Equation 7.46 remains to be applied to obtain the values for r k , l . For example, for 1-1,1:

@1%,3

@2y2,3

'3y3.3

which is:
1-In (0.4965-1 + 0.1581.1 + 0.345 4-0.927)
0.496 5 * 1
0.496 5 . l + 0.158 1* 1 + 0.345 4 * 0.927
In 1-1,1= 0.848

0.158 1* 1
0.496 5.1 + 0.158 1.l + 0.345 4 * 0.927
0.345 4 * 0.259 5
0.496 5.0.2595 + 0.1581.0.2595 + 0.345 4.1

and therefore:
In 1-1,1= 0.848 [ l + 0.025 5 - 0.5093 - 0.1622 - 0.17301 = 0.1527
In the same way, we find that r2,1
= 0.0972 and In r3,1
= 0.4368.
A similar calculation is then applied to the mixture itself.We determine the mole fractions for each group in the mixture (Eq. 7.48). For the first group:
VI,IXl + v1,2x2

x1=
'I,lX1

+ v1,2x2 + '2,Ix1

+ v2,2x2

'3,IX1

'3,2"2

2.0.056 + 2.0.944
2 * 0.056 + 2.0.944 + 1.0.056 + 5 * 0.944 + 1 .0.056 + 0.0.944

= 0.297 4

and similarly for X2= 0.69906 and X3= 0.0082.


The surface fractions that correspond to these compositions (Eq. 7.47):
0, = 0.3907

0, = 0.594 1

0, = 0.015 22

257

7. Deviations from Ideality in the Liquid Phase

once again allow for the application of Equation 7.46 to calculate the activity coefficients for each group in the mixture. For example:

In I-, = 0.848

1 - In (0.3907.1 + 0.594 1 1+ 0.015 22.0.927)


0.390 7 * 1
0.390 7 .1 + 0.594 1* 1+ 0.015 22 * 0.927
-

0.594 1. l
0.390 7 * 1+ 0.594 1. l + 0.015 22 * 0.927

0.015 22 * 0.259 5
0.390 7.0.259 5 + 0.594 1-0.2595 + 0.015 2 * 1 ,

yielding:
In I-, = 0.848 [l + 0.000 1111- 0.391 1- 0.594 7 - 0.014591 = 0.000552
We also find In I-, = 0.000351 and In I-, = 1.5769.
Finally, by applying Equation 7.40:

In

%,residual = vl,l

(In

- In 1,l)

+ v2,1

(In r2 - In

r2,1)

v3,1

(In

I-3 - In I-3,1)

= 2(0.00055 - 0.1527) + l(0.000351- 0.097 2) + l(1.5769 - 0.4368) = 0.738 9

The combinatorial term must be taken into account (Eq. 7.39), ultimately yielding:
In
and therefore:

x = In ~,,combinatoria,+ In x,,residual = -0.027 + 0.7389 = 0.711 9


fi = 2.037

7.7.3 Group Contribution Method


Conclusion
We have already emphasized the value of the group contributions method. It compensates
for the lack of experimental data, which means that methods that rely directly upon such
data should be the methods of choice, when possible. However, it happens frequently that
we find ourselves in a mixed situation: given a mixture containing three or more components, we have access to data for some of the binary components, but not for all of them. A
priori, it seems that applying one of the methods necessitating prior determination of the
parameters from experimental data (Wilson equation, NRTL method, or UNIQUAC) cannot be undertaken since for some of the binary components, such a determination is
impossible. Is it necessary to apply the UNIFAC method to the system and in this way lose
the information from experimental data? A composite solution is possible. We choose a
model: Wilson equation, NRTL, or UNIFAC. From the available experimental data, we
determine the parameters of this model. The parameters relating to other binary components may be calculated by adjusting the activity coefficients predicted by the UNIFAC
(or ASOG) model.
Regardless of the model, we must not neglect the proximity effect.

258

7. Deviations from Ideality in the Liquid Phase

It also seems that the UNIFAC or ASOG methods such as they have been presented,
find their application limited to systems whose components do not have too much of a difference in molar volume, despite the fact that this difference is accounted for by a combinatorial term. Their application to polymer solutions has been done, but at the expense of
large modifications. We shall discuss these modifications later on (Chapter 11).

7.8 ASSOCIATED SOLUTIONS


The molecular interactions may give rise to the formation of associated compounds, as is
the case for systems having compounds prone to hydrogen bonding. There are numerous
examples. We can mention acetic acid that, even in vapor phase and at atmospheric pressure, mainly exists as a dimer, as seen by the density value. One mole of acetic acid, or 60 g,
at its boiling point at atmospheric pressure (118C), occupies a volume in the saturated
vapor phase that is close to 16 literdmole, which is around half of what we would calculate
for an ideal gas. These combinations may cause polymerization, as in the case of alcohols
that have the general formula:
@-OH),

n=l,w

Along with these auto-associationswe know that complexes may form due to transassociation,between two different alcohols, for example. They form a copolymermixture. Other examples exist between two compounds where neither compound undergoes
auto-associationitself. One acts as a hydrogen bond donor, and the other as an acceptor.
The example most often cited is that of chloroform and acetone mixtures. These two compounds react according to the stoichiometry:

c1

I
I

2 C1-C-H

CHC13 + CH3-CO-CH3

c1

... O=C

,CH3
\

CH3

It is appropriate to account for this phenomenon in the analysis of thermodynamic


properties of such mixtures. An initial approach uses the combined ideal solutions
hypothesis,which we illustrate below.

Combined ideal solutions: the chloroform ( I ) acetone (2) system


We denote the quantities (number of moles) of chloroform and acetone placed in solution
N l , and N2,respectively. If we ignore the complexation reaction mentioned previously, we
can calculate the composition (xl,x2) of the mixture:

N,
In fact, the formation of a complex with 1/1 stoichiometry is going to modify this composition. If we designate the extent of the reaction 5, the quantities of chloroform,acetone,

7. Deviations from Ideality in the Liquid Phase

259

and of the complex are, respectively, equal to Nl - 5, N 2 - 5, and


sponding compositions are:

5, and the actual corre-

x; =

Nl

N1- 5
N2-5
5
+ N2 - 5' x' -- N l + N2 - 5' xj = N , + N 2 - 5

where subscript 3 is for the complex.


If ratio u is:

and if we consider the ternary mixture chloroform (1) acetone (2) complex (3) ideal, then
the reaction equilibrium is expressed by the equation:

--

x;x;

u(1- u )

(XI - u ) (x, - u )

=K

where K is the chemical equilibrium constant. The value for u can be obtained:

The liquid-vapor equilibrium of this supposedly ideal ternary mixture is expressed by


the equations:
Y, = p
p1" XI, y , = p
pz"x,,
f

P=Pyx;+P;x;

given that the complex has a negligible vapor pressure compared to that of chloroform and
acetone, and that the reaction has progressed little.
If we had ignored the complex formation, we would have expressed the liquid-vapor
equilibrium on the basis of detectable mole fractions, yet taking into account the deviations from ideality:
p1"
Y, = P "/1,,

P?
P

Y2 = - xx2,

p = p y "/1q
+p;

EX2

and putting the two expressions together, gives the activity coefficients:

%=-,

X1

y2=-

xs
x2

The experimental data from Goral et al. [1985] enable us to determine the value of the
equilibrium constant by minimizing the discrepancies between the experimental and calculated values for the pressure. We find that K = 1.25 and Table 7.13 as well as Figure 7.15
allow us to verify that the model is reasonably applied.
In this case, we note that the deviations from ideality are negative. They relate to the
fact that the actual mole fractions of acetone and chloroform are less than the observable
mole fractions.

2 60

7 . Deviations from Ideality in the Liquid Phase

Table 7.W
Correlation of liquid-vapor equilibria of the chloroform (1) acetone (2)
system at 313.15 K using the associated ideal solutions model

o.Ooo0
0.0303
0.083 4
0.143 1
0.2082
0.287 8
0.361 6
0.4270
0.497 8
0.5628
0.601 8
0.638 1
0.685 5
0.7305
0.803 9
0.8602
1.Oooo

56.5700
55.6300
53.8800
51.9300
49.8200
47.2700
45.0800
43.4200
42.0400
41.2900
41.0900
41.0700
41.3000
41.9300
43.2500
44.5400
48.0900

56.5700
55.4970
53.593 1
51.4435
49.1425
46.5090
44.4002
42.9336
41.8867
41.4729
41.4746
41.6333
42.0455
42.6207
43.855 7
44.9790
48.0900

Ylexp

Y I cal

Y2

o.Ooo0
0.0140
0.0407
0.075 0
0.1189
0.1844
0.2595
0.3385
0.436 5
0.5340
0.593 7
0.648 4
0.7169
0.7767
0.860 5
0.9120
1.Oooo

o.Ooo0
0.012 1
0.0366
0.070 2
0.1157
0.1865
0.2686
0.354 0
0.456 2
0.5536
0.611 1
0.662 6
0.725 9
0.780 8
0.8584
0.908 1
1.0000

0.4444
0.459 8
0.488 8
0.524 7
0.568 1
0.626 8
0.685 8
0.740 1
0.7982
0.848 3
0.875 7
0.899 0
0.925 9
0.947 3
0.973 8
0.9874
1.Oooo

1.Oooo
0.999 5
0.995 8
0.986 7
0.970 1
0.939 1
0.899 2
0.855 7
0.801 8
0.7486
0.716 1
0.686 1
0.647 7
0.6128
0.5597
0.522 7
0.4444

o.Ooo0
0.0139
0.0408
0.0751
0.1183
0.1804
0.2480
0.3160
0.3974
0.477 4
0.527 0
0.573 7
0.6347
0.6920
0.7828
0.8494
1.Oooo

1.Oooo
0.9692
0.9127
0.8455
0.7682
0.6688
0.574 1
0.4903
0.4026
0.327 3
0.285 2
0.2483
0.2037
0.165 1
0.1098
0.0731
o.Ooo0

60

50
h

g!

I2

40

30
0.0

0.2

0.4

0.6

0.8

Mole fraction of chloroform

Figure 7.15 Application of the associatedideal solutions model to the


chloroform (1) acetone (2) system.
Curves: calculated values; points: experimental values [Goral et al., 19851.

7. Deviations from Ideality in the Liquid Phase

261

With alcohol hydrocarbon mixtures, one of the components in the pure state is polymerized by auto-association, and the measurement of its vapor pressure is in fact for a
polymer mixture. According to traditional laws of equilibrium displacement, the addition
of an inert solvent, the hydrocarbon, will cause the polymerization reaction to regress,
release the more volatile forms that are less combined, and thus be responsible for positive
deviations from ideality, corresponding to what is observed.
General Equations for Associated Solutions
The model for associated ideal solutions is very much simplified. We know that the hydrocarbon mixtures are not ideal, and we do not know whether it is caused by an association
process. We are therefore led to suppose that in an associated solution, the actual chemical
entities themselves make up a non-ideal solution. This fact results in generally complex
models.
There are, however, some general equations that we shall illustrate using the case of
solutions with an auto-associative compound (alcohol, for example) in an inert solvent.
They lend themselves to more complex cases.
Let us denote the auto-associative compound A, the solvent B, and state the stoichiometry for auto-association:

i=2,w

iA,+Ai

(7.50)

The apparent number of moles of component A, N(A) and the (real) number of moles
of component B, N(B) are known on the basis of the mass of the components and the
molar mass of the monomer for A, and of solvent B. On the other hand, we ignore the
number of moles for each of the species A i , (A,). The material balance, however,
imposes the relationship:
m

N(A)=

C. (Ai)

(7.51)

Furthermore, the tiextents of these reactions are such that the Gibbs energy of the system G at the given temperature and pressure, is minimal. For each of them we have:
dN(Ai) = d t i
m

dN(A,) = -C i d t i
2

The elementary variation of Gibbs energy corresponding to the progression variation


d t i is therefore (as a function of the number of moles and actual chemical potentials):
m

C1

( d ~ ) T p = P(Ai) dN(Ai) =

C b(Ai) - ~ P ( A ~ )dIt i
2

and the minimum condition imposes:

(7.52)

2 62

7. Deviations from Ideality in the Liquid Phase

If we now displace the equilibrium by the addition of A or B:


m

( d ~ ) , p= C. p(Ai) d ~ ( ~+iCL(B)
)
d(~)
1

which, by application of Equation 7.52 becomes:


m

(dG),,

ip(Al) dN(A,)

+ p(B) dN(B)

1
m

= p(AI)

i dN(A,)

+ p(B) dN(B)

however, by applying the material balance Equation 7.50


m

dN(A) =

i dN(A,)
1

and of course:

dN(B) = dN(B)

that we substitute in the preceding expression for dGT,Pthat yields:


( W , , =AJ

dN(A) + P V ) dN(B)

This last reaction defines the apparent chemical potentials for A and B:
P(A) = A,)

and

P(B)= B )

(7.53)

where:

So the apparent chemical potential of the associated compound is equal to the actual
chemical potential of the monomer.
The preceding equations are independent of the hypotheses that may be made for:
0 Specifying the auto-association process by limiting its extension, for example, or by
introducing equations between the equilibrium constants of the reactions. For example, we often state that for reactions of form:

i = 1,03
the equilibrium constant is independent of the degree of polymerization i.
0 Calculating the molar volume of the polymer. We often apply an additivity rule.
0 Expressing the non-ideality of the real solution. In principle, any model may be
applied. It is, however, logical to account for a combinatorial term (see this chapter,
Section 7.3, and Flory, 1944) since the solution contains chemical species characterized by a long chain length and an energy (or residual) term. Examples are provided in the work of Renon [1967a, 1967bl.
The final models are often complex, have too many parameters that are often inter-correlated, and which, finally, do not appear to offer a decisive advantage from the point of
view of chemical engineering when compared to models that we have previously discussed.
Note that this concept of combined solutions has also been applied to equations of state.
Ai+Al

7. Deviations from Ideality in the Liquid Phase

7.9

2 63

IONIC SOLUTIONS

Deviations from ideality for ionic solutions result in a large part from long distance interactions between anions and cations, solvation of ions, and the formation of ion pairs. To
this is of course added the interactions due to attraction forces exerted between the solvent molecules at average distance, and repulsion at short distance, as we have already
encountered.
These deviations from ideality are evidenced in electrochemistry, but also when determining phase equilibria: the precipitation of salts from a saturated solution, splitting of two
liquid phases, and liquid-vapor equilibria. For example, an azeotrope may be substantially
displaced or even eliminated by the addition of salts.
Ion activity cannot be estimated individually due to the electroneutrality condition that
prevents a solution from containing a single species and, because of their absence in the
vapor phase that does not allow us to obtain the liquid-vapor equilibrium coefficients for
each one of them.
We shall neither discuss nor summarize the thermodynamics of ionic solutions in this
study. The reader is invited to consult the works of Lewis et al. [1961, pages 298 to 3721,
Robinson and Stokes [1965], the Cruz Journal [1977], and the articles by Pitzer [1973,
19791, Chen [1979], and Renon [1979).
Instead, we shall summarize the example proposed by Sander et al. [1986] and Macedo
et al. [1990] for the calculation of liquid-vapor equilibria of electrolyte solutions. It uses an
extension of the UNIQUAC model already introduced in this chapter.
Only the solvents are involved in the liquid-vapor equilibrium. Therefore, this model
limits itself to the calculation of activity coefficients of the solvent(s), yet takes into
account the changes caused by the presence of ions. The model states the liquid-vapor
equilibrium condition using the traditional formula:

fi'=

fiL

p y i q : = p ; q ; q x i yi"
Since the model was developed from liquid-vapor equilibrium data at low pressure, this
condition is reduced to:

Pyi = p;xi yi"


and in these equations y; represents the activity coefficient of the solvent or solvents,
defined from the usual reference state (pure substance). Three contributions stand out in
the activity coefficient expression:
(7.54)
The first term is from the Debye and Huckel [1923] theory, which is explained in the
present case by Cardoso and O'Connel[1987].The second term is the combinatorial term
from the UNIQUAC model, which has not been modified. The third, or residual term,
retains the form defined by the UNIQUAC model, but the calculation of the interaction
parameters accounts for the presence of ions.

2 64

7. Deviations from Ideality in the Liquid Phase

The Debye and Huckel term is expressed by the equation:

In

where:

b - H=

2 AMidi
b3di

[l+bfi- l + b f i
a

A = 1.327757 .los

(ET)~"

and

- 2 In (1 + b lh)]

(7.55)

b = 6.359696

Mi being the molar mass of component i, di the density (expressed in kg/cm3),d, the solvent density (or the solvent mixture), E the dielectric constant, I the ionic force, and T the
temperature.
The expression for the residual term remains unchanged:
(7.37)
L

k=l

where the expression for the interaction energy as a function of temperature is:

(7.56)
In the presence of electrolytes, the ai,, parameters will be a function of the composition.
If subscript i is a solvent, and the subscriptsj or k are ions, we have:

(7.57)

The parameters of this model (volumes and surface of the ions, the interaction terms
a;,i,$ k , i ) were calculated for as much as experimental data existed for the solutions where
the solvent was: water, methanol, ethanol, n-propanol, acetone, and n-butanol, and the
alkaline cations, alkaline earth salts, nickel, mercury, the halide anions, acetate, and nitrate.
All in all, the results were excellent.
We might imagine a model based on group contributions. However, its development
would be hindered by the lack of experimental data in sufficient quantity.The same would
be true for its extension to liquid-liquid equilibria and the precipitation of salts at saturation.

7. Deviations from Ideality in the Liquid Phase

265

REFERENCES
Abdoul W, Rauzy E, PCnelouxA (1991) Group contribution equation of state for correlating and predicting thermodynamic properties of weakly polar and non-associating mixtures. Binary and multicomponents. Fluid Phase Equilibria, 68,47-102.
Abrams DS, Prausnitz JM (1975) Statistical thermodynamics of liquid mixtures: a new expression
for the excess Gibbs energy of partly or completely miscible systems. AIChE J., 21,116-128.
Battino R (1971) Volume changes on mixing for binary mixtures of liquids. Chem. Rev., 71 (l), 5-44.
Benson GC (1986) Excess volume of the Ethanol Heptane system. International Data Series;Selected
Data on Mixtures, A series, p. 264. H. Kehiaian, Ed.; Thermodynamic Research Center, Texas
A&M University.
Berro C, Deyrieux A, PCneloux A (1975) J. Chim. Phys., 10,1118.
Bondi A (1964) van der Waals volume and radii. The Journal of Physical Chemistry, 68 (3), 441-451.
Bondi (1968) Physical Properties of Molecular Crystals, Liquids and Glasses. Wiley, New York.
Bruin S, Prausnitz JM (1971) One parameter equation for excess Gibbs energy of strongly non-ideal
mixtures. Ind. Eng. Chem. Process Des. Dev., 10,562-572.
Cardoso MJE, OConnel JP (1987) Activity coefficients in mixed solvent electrolytes solutions.Fluid
Phase Equilibria, 33,315-326.
Chao KC, Seader JD (1961) A general correlation of vapor-liquid equilibria in hydrocarbon mixtures.
AIChE J.,7,598-605.
Chen C-C, Britt HI, Boston JF, Evans LB (1979) Two new activity coefficient models for the vaporliquid equilibrium of electrolyte systems. In: Thermodynamics of Aqueous Systems with Industrial
Applications. S. A. Newman, Ed., ACS Symposium Series, 1980, p. 61.
Christiansen W,Fredenslund Aa (1975) Thermodynamic constitency using orthogonal collocation or
computation of equilibrium vapor compositions at high pressures. AIChE J., 21,49-56.
Costas M, Patterson D (1985) Excess heat capacity of the Benzene Decane system. International
Data Series; Selected Data on Mixtures, A series, p. 214. H. Kehiaian, Ed.; Thermodynamic
Research Center, Texas A&M University.
Cruz JL (1977) Revue bibliographique des reprksentations analytiques de coefficients dactivitk
ioniques moyens dans les solutions binaires electrolytiques. Rev. Inst. Franq. du Pktrole, 32,393.
Debye P, Huckel E (1923) Theory of electrolytes I. Freezing point lowering and related phenomena.
Physik Z., 24,185.
Derr EL, Deal CH (1969) Analytical solution of groups. Correlation of activity coefficients through
structure group parameters. Inst. Chem. Eng. Symp. Ser., 32,3,40.
Flemr V (1976) A note on excess Gibbs energy equations based on local composition concept.
Collect. Czech. Chem. Commun.,41,3347.
Flory PJ (1942) Thermodynamics of high polymer so1utions.J. Chem. Physics, 10,51-61.

Flory PJ (1944) Thermodynamics of heterogeneous polymers and their solutions.J. Chem. Phys., 12,
425-438.
Flory PJ (1970) Thermodynamics of polymer solutions. Discussions Faraday SOC.,49,7-29.
Fredenslund Aa, Jones RL, Prausnitz JM (1975) AIChE J. 21,1086.
Fredenslund Aa, Gmehling J, Rasmussen P (1977) Vapor-Liquid Equilibrium Using UNIFAC.
Elsevier, New York.
Fredenslund Aa, Rasmussen P (1985) From UNIFAC to SUPERFAC and back? Fluid Phase
Equilibria, 24,115-150.

2 66

7. Deviations from Ideality in the Liquid Phase

Gmehling J, Onken U (1978) Vapor-liquid equilibrium data collection. Chemistry Data Series,
Dechema.
Gmehling J, Rasmussen P, Fredenslund Aa (1982) Vapor-liquid equilibria by UNIFAC group contribution. Revision and extension 2. Ind, Eng. Chem. Process Des. Dev., 21,118-127.
Gmehling J, Li, Jidding, Schiller M (1993) A modified UNIFAC model. 11. Present parameter matrix
and results for different thermodynamic properties. Ind. Eng. Chem. Res., 32,178-193.
Goral M, Kolasinska G, Oracz P, Warycha S (1985) Vapor-liquid equilibria in the ternary system
methanol-chloroform-acetoneat 313.15 and 323.15 K. Fluid Phase Equilibria, 23,89-116.
Guggenheim EA (1952) Mixtures. Oxford University Press.
Handa YP, Benson GC (1979) Volume changes on mixing two liquids: a review of the experimental
techniques and the literature data. Fluid Phase Equilibria, 3,185-249.
Hansen HK, Rasmussen F, Fredenslund Aa, Schiller M, Gmehling J (1991) Vapor-liquid equilibria by
UNIFAC group contribution. Revision and extension 5. Ind. Eng. Chem. Res., 30,2352-2355.
High MS, Danner RP (1990) Application of the group contributions lattice-fluid EOS to polymer
solutions.AIChE J.,36,1625-1632.
Hildebrand JH, Scott RL (1924,1936, 1950) The Solubility of Nonelectrolytes. Reinhold Publishing
Corp., New York.
Hildebrand JH, Prausnitz JM, Scott RL (1970) Regular and Related Solutions. Van NostrandReinhold, Princeton.
Huggins ML (1941) J. Chem. Phys. 9,440.
Huggins ML (1942) Ann. N. Y. Acad. Sci. 43,l.
Jose J, Blonde1 A, Mokbel I, Kasehgari H, Viton C (1992) Mesure de la masse molaire moyenne de
coupes pktrolikres et des tensions de vapeur dhydrocarbures lourds. 1992 PIRSEM Report.
Kehiaian HV (1983) Group contribution methods for liquid mixtures: a critical review. Fluid Phase
Equilibria, 13,243-252.
Kehiaian HV, Marongiu B (1988) A comparative study of thermodynamic properties and molecular
interactions in mono and polychloroalkane + n-alkane or + cyclohexane mixtures. Fluid Phase
Equilibria, 40,23-78.
Klaus RL, van Ness HC (1967) Chem. Eng. Prog. Symp. Ser., 63 (81), 88.
Kojima K, Tochigi K (1979) Prediction of Vapor Liquid Equilibria by the ASOG Method. Kodonsha
Elsevier, Tokyo.
Larkin JA (1975) Excess enthalpy of the Water Ethanol system.J. Chem Thermodynamics,7,137-148.
Larsen BL, Rasmussen P, Fredenslund Aa (1987) A modified UNIFAC group contribution model
for prediction of phase equilibria and heat of mixing. Ind. Eng. Chem. Res., 26,2274-2286.
Lewis GN, Randall M, Pitzer KS, Brewer L (1961) Thermodynamics.p. 332-348. McGraw-Hill.
McDermott C, Ashton N (1977) Note on the definition of local composition. Fluid Phase Equilibria,
1,33-35.
Macedo E, Skovborg P, Rasmussen P (1990) Calculation of phase equilibria for solutions of strong
electrolytes in solvent/water mixtures. Chem. Eng. Science, 45,875-882.
Magnussen Th, Rasmussen P, Fredenslund Aa (1981) UNIFAC parameter table for prediction of
liquid-liquid equilibria.Ind. Eng. Chem. Process Des. Dev., 20,331-339.
Marina JM, Tassios DP (1973) Effective local compositions in phase equilibrium correlations. Ind.
Eng. Chem. Process Des. Dev. 12,67-71.
Marsh KN (1989) New methods for vapor-liquid equilibria measurements. Fluid Phase Equilibria, 52,
169-184.

7. Deviations from Ideality in the Liquid Phase

267

Maurer G, Prausnitz JM (1978) On the derivation and extension of the UNIQUAC equation. Fluid
Phase Equilibria, 2,91-99.
Moelwyn-Hughes (1961) Physical Chemistry, 2nd edition. Pergamon Press.
Neau E (1979) Contribution au traitement des donnkes et i la prevision des Cquilibres liquide-vapeur
des solutions moleculaires.Sc. D. thesis, Universite dAix-Marseille11.
Neau E, PCneloux A (1981) Estimation of model parameters. Comparison of methods based on the
maximum likelihood principle. Fluid Phase Equilibria, 7,l-19.
Panayiotou C, Vera JH (1980) The quasichemical approach for non-randomness in liquid mixtures.
Expressions for local surfaces and local compositions with an application to polymer solutions.
Fluid Phase Equilibria, 5,55-80.
Peneloux A, Deyrieux R, Neau E (1975a) Reduction des donnCes exPCrimentales et theorie de linformati0n.J. Chim. Phys., 10,1101-1107.
PCneloux A, Deyrieux R, Neau E (1975b) RCduction des donnkes sur les Cquilibres liquide-vapeur
isothermes. Critkres de prCcision et de coherence. Analyse de linformation. J. Chim. Phys., 10,
1107-1117.
PCneloux A, Deyrieux R, Canals L, Neau E (1976) The maximum likelihood test and the estimation
of experimental inaccuracies.Application to data reduction for liquid-vapor equilibrium.J. Chim.
Phys., 7,8,708-716.
PCneloux A, Neau E, Gramajo A (1990) Variance analysis fifteen years and now. Fluid Phase
Equilibria, 56,l-16.
Pitzer KS (1973) Thermodynamics of electrolytes I: Theoretical basis and general equations.
J. Phys. Chem. 77,268.
Pitzer KS (1979) Thermodynamics of aqueous electrolytes at various temperatures, pressures and
compositions. In: Thermodynamics of Aqueous Systems with Industrial Applications.
S. A. Newman, Ed., ACS Symposium series, 1980,p. 451.
Prausnitz JM, Anderson JM, Grens EA, Eckert CA, Hsieh R, OConnel JP (1980) Computer
Calculations for Multicomponent Vapor-Liquid and Liquid-Liquid Equilibria. Prentice-Hall,Inc.,
Englewood Cliffs, New Jersey.
Prausnitz JM, Lichtenthaler RN, de Azavedo EG (1986) Molecular Thermodynamics of Fluid Phase
Equilibria. Prentice-Hall,Inc., Englewood Cliffs, California.
Redlich 0,Kister AT (1948) Znd. Eng. Chem., 345-348.
Renon H, Prausnitz JM (1967) On the thermodynamics of alcohol-hydrocarbons solutions. Chem.
Eng. Sci., 22,299-307 and (errata) 22,1891.
Renon H, Prausnitz JM (1968) Ind. Eng. Chem. Process Des. Dev. 7,210.
Renon H, Asselineau L, Cohen G, Raimbault C1 (1971) Calcul sur ordinateur des kquilibres liquidevapeur et liquide-liquide. Application d la distillation des mtlanges non idkaux et a lextractionpar
solvants. Editions Technip, Paris.
Renon H (1979) Representation of NH3-H,S-Hz0,
NH3-S0,-Hz0
and NH3-C0,-H20
vapor-liquid equilibria. In: Thermodynamics of Aqueous Systems with Industrial Applications.
S. A. Newman, Ed., ACS Symposium series, 1980,p. 173.
Robinson RA, Stokes RH (1965) Electrolyte Solutions, 2nd edition, Butterworth, London.
Scatchard G (1931) Chem. Rev. 8,321.
Sander B, Fredenslund Aa, Rasmussen P (1986) Calculation of vapor-liquid equilibria in mixed solventlsalt systems using an extended UNIQUAC equation. Chem. Eng. Sci., 41,1171-1183.
Soave G (1992) Application of the UNIQUAC equation to regular solutions. Fluid Phase Equilibria,
77,133-137.

2 68

7. Deviations from Ideality in the Liquid Phase

Skjold-Jorgensen S, Kolbe B, Gmehling J, Rasmussen P (1979) Vapor-liquid equilibria by UNIFAC


group contribution. Revision and extension. Ind. Eng. Chem., Process Des. Dev., 18,714-722.
Stavermann AJ (1950) The entropy of polymer solutions. Rec. Trav. Chim. the Netherlands, 69,163.
Tiegs D, Gmehling J, Rasmussen P, Fredenslund (1987) Vapor-liquid equilibria by UNIFAC group
contribution. Revision and extension 4. Ind. Eng. Chem. Rex, 26,159.
Tochigi K, Lu BCY, Ochi K, Kojima K (1981) On the temperature dependance of ASOG parameters
for VLE calculations.AIChE J., 27,1022-1024.
Trejo LM, Costas M, Patterson D (1991) International Data Series, A series, Selected Data on
Mixtures, p. 50.
van Laar JJ (1910) Sechs Vortrage uber das Thermodynamische Potential. 2. Physik. Chem. 72,723
and 83,599.
Vera JH (1986) On local compositions concept. Fluid Phase Equilibria, 26,313-316.
Weidlich U, Macedo E, Gmehling J, Rasmussen P (1979) Vapor-liquid equilibria by UNIFAC group
contribution. Revision and extension 3. Ind. Eng. Chem., Process Des. Dev., 22,676-678.
Weidlich U, Gmehling J (1987) A modified UNIFAC model. I. Prediction of VLE, hE,and y". Ind. Eng.
Chem. Res., 26,1372-1381.
Wilson GM, Deal CH (1962) Activity coefficients and molecular structure. Ind. Eng. Chem. Fundam.,
1,20-23.
Wilson GM (1964) Vapor liquid equilibrium. XI. A new expression for the free energy of mixing.
J. Am. Chem. SOC,86,127-130.
Zudkevitch D (1978) Impact of thermodynamic and fluid properties on design and economics of
separation processes, Encyclopedia of Chemical Processing and Design. J. McKetta.

Application of Equations of State


to Mixtures
Calculation of Liquid-Vapor
Equilibria Under Pressure

At least in principle, equations of state permit the simultaneous calculation of the vapor
pressures and residual properties (density, enthalpy, entropy, heat capacity) of liquid and
vapor phases, when applied to pure substances. These calculations are performed in a
coherent fashion and the results that are obtained validate the general equations that link
the temperature, pressure, and the thermodynamic functions.They therefore represent a
very promising pathway, but of course only on the condition that the range of application
for these equations is properly specified and respected: which compounds? which conditions? which properties?
In this chapter, we emphasize the calculation of liquid-vapor equilibria under pressure.
Indeed, the difficulties of this topic are at the source of development for the equations of
state, as pointed out earlier. The calculation of these equilibria is also the criterion most
often applied to evaluate new approaches.
If we look at mixtures, the compositionvariable must be taken into account in the formulation of equations of state. These are written in the form:
E ( T , P , v N l , N 2 , N 3...)
, =0

(8.1)

which shows temperature, pressure, volume, and the number of moles of each component,
or for one mole of mixture:
e(T,P,:v,zl,z2,z3
,...) = 0

(8.2)

In this last equation, ziis the mole fraction of component i, usually represented by xi if
the mixture is homogeneous and liquid, and by yiif it is in the vapor state.
In fact, these expressions must also contain a list of the parameters for the equation of
state in question, the virial coefficients,covolume, attraction parameter, etc. already seen
for pure substances.It is generally at the parameter level that we see the influence of composition according to the mixing laws, which we shall define case by case.
We may also ask ourselves what the parameters of an equation of state really are. For
those derived from the van der Waals theory (Soave-Redlich-Kwong [1972] or Peng-

2 70

8. Application of Equations of State to Mixtures

Robinson [1976] equation of state), apparently they are two in number, a and b. Since they
are most often calculated from the critical points and the acentric factor, we can also say
that these last quantities are the actual parameters, and that it is appropriate to apply the
mixing laws to them. We shall see that this has been done.
Most often the mixing laws respect the structure of the equation of state, such that we
implicitly form the hypothesis that a mixture of given composition behaves like an artificial pure substance; we therefore have a one fluid model. This is a hypothesis that nothing substantiates,other than its simplicity and the results obtained.
The fact that the application of equations of state to mixtures and the calculation of
liquid-vapor equilibria are practically inseparable,leads us to carefully specify the calculation method for partial molar quantities, in particular chemical potentials and fugacity
coefficients. Certainly, the definitions and the equations developed in Chapter 5 apply, but
in practice we need to account for the structure of the equation of state and the mixing
laws that we choose. For the principal examples that will be mentioned here, we need to
make this point clear.
Finally,a fundamental question must be asked. For a homogeneous phase, the extensive
properties and their partial molar quantities can be, at least in principle, expressed and calculated using two distinct paths.The first one we described and applied (Chapters 5 and 7)
uses reference states, mixing quantity, and excess quantity. The second uses equations of
state and mixing laws. Insofar as the models pertaining to each of these paths both address
the problem and have been selected, we end up with the same result. Consequently, an
implicit relationship exists between these two paths. Defining this relationship has led to
new methods that, in design, combine the equations of state and excess quantities,and as
such the qualities of each of these paths.

8.1

EXTENSIONS OF THE CORRESPONDINGSTATES PRINCIPLE

These extensionshave already been mentioned (Chapter 3, Section 3.2.3.2) and it is within
the framework of the Lee and Kesler [1975] method that they are most often utilized. Yet
they may be generalized for many other methods, as the examples that follow will show. In
particular, they are used in equations of state that apply the corresponding states principle
for the calculation of the parameters of pure substances.An example of this generalization
is provided by the work of YC [1990] who, for the calculation of the speed of sound, combined, on the one hand, the equations of state of Soave-Redlich-Kwong, of PengRobinson, of Simonet-Behar-Rauzy, of Benedict, Webb and Rubin as used by Lee and
Kesler, with, on the other hand, the mixing laws that we shall discuss. As an example, the
equation of state from Soave-Redlich-Kwongthat we introduced in Chapter 4 using:
(4.42)

(4.43)

8. Application of Equations of State to Mixtures

271

(4.37)

a(T)= [ l + m(1m =M,

(4.46)

+ MI w + M 2 0 2

(4.47)

may be written using the compressibilityfactor 2

For a mixture, if we therefore define calculation rules for what we shall call pseudocritical points, and an acentric factor Tc,,, Pc,,, and oc,,, we may apply the equation of
state and calculate the compressibility factor as a function of the reduced coordinates
defined as:
T
P
T=and P r = (8-3)
T,m
Pcm

8.1.1 Calculation Rules for Pseudocritical Points


The definition of pseudocritical points, introduced by Kay [1936] in its most simple form:

q,,,= XziTGi

and

Pc,, =

cziP,,,

(3.35)

has had many variations. The best known is the one proposed by Lee and Kesler, mentioned earlier (Chapter 3, Section 3.2.2.3):

(3.36)
(3.37)
0,

xziwi

(3.38)

where:

(3.39)
Zc,i= 0.2905 - 0.085 mi
VC,,

and
zizj

i j

Zc,m
= 0.2905 - 0.085 om

(3.40)
(3.41)

2 72

8. Application of Equations of State to Mixtures

It was modified by Plocker et al. [1978] for the purpose of better calculating the liquidvapor equilibria. It therefore has a binary parameter ki,jthat may be adjusted to experimental data. Equation 3.36 is replaced with:

where:

(8.5)

In works related to the calculation of density in the liquid phase, Spencer and Danner
[1973], Hankinson and Thomson [1979], and finally Teja [1980], proposed the following
rules.
Spencer and Danner rules:

Hankinson and Thomson rules:

c,m
= (0.291 - 0.08 0 , )

RTqm
-

(8.10)

vc,m

with the linear Equation 3.38 being applied to the calculation of 0,.
The Teja mixing law has two adjustable parameters
and E , ~which,
,
in the absence of
binary experimental data, we shall take to be equal to 1.It is written:
(8.11)
(8.12)

2 73

8. Application of Equations of State to Mixtures

ZC, zizjvc,i,jTc,i,j
Tcm =

(8.14)
"c,m

(8.15)
Finally, Pedersen et al. [1984], in order to predict the viscosities of complex mixtures,
applied the following rules:

Indisputably,these rules are the product of much empiricism.


Figures 8.la and 8.lb show the change of pseudocritical pressure as a function of pseudocritical temperature for the methane n-octane and the ethane benzene systems respectively. Evidently,differences appear as a function of the method that has been applied,with
the exception of those from Hankinson and Thomson on the one hand, and Teja on the
~ both taken
other hand that lead to identical results if the binary parameters gi,j, and K , are
to equal 1.The differences are relatively small in the case of the second system. However,
one should account for the composition,and Table 8.1 specifies the pseudocritical values of
the equimolar mixtures for each system.

Methane, n-octane

Tcm

Pc,m

6)

(MPa)

(K)

(MPa)

379.70
504.74
425.02
362.90
410.23
410.23
429.48

3.54
3.73
3.58
3.06
3.45
3.45
3.50

433.75
469.55
442.31
426.99
439.54
439.54
443.4

4.89
5.29
5.08
4.90
5.05
5.05
5.08

e' m

Kay
Spencer et al.
Lee et al.
Plocker et al.
Hankinson et al.
Teja
Pedersen et al.

Ethane, benzene

'em

2 74

8. Application of Equations of State to Mixtures

5.5

.. .
*---

8
0

8
8

4.5
h

E
g

3
fn

!!

3.5

2.5
I

200

300

400

500

Temperature (K)

Figure 8.la Pseudocritical coordinatesof methane, n-octane mixtures.


Kay Method: -;
Spencer and Danner: - - - -;
Lee and Kesler: - - -; Plocker et al.: .-.-.-;
Hankinson and Thomson: .-.-.--; Teja et al.: ------;
Perdersen et al.: ....-....-.

Of course, it is important not to confuse the pseudocritical coordinates, which are the
parameters for a model only, with the true critical coordinates. The true critical point
(Chapter 6, Section 6.1.2) relates to conditions of temperature and pressure that characterize a real, experimentally observable physical phenomenon in which the two phases in
liquid-vapor equilibrium become indistinguishable.They may differ considerably from the
pseudocritical coordinates. For example, in the ethane benzene mixture, the critical pressure reaches 10.5 MPa while the pseudocritical pressure does not go above 5.35 MPa. For
hydrocarbon mixtures with very different molecular weights, critical pressure may reach
several hundred bar.

275

8. Application of Equations of State to Mixtures

5.4

5.3

I
I

--*
8

\
\
\

5.2

\
\
\

I
I

a!z

I
I
I

I
I

5.1
ln
ln

e!

4.9

4.8
300

400

500

600

Figure 8.lb Pseudocritical coordinates of ethane benzene mixtures.


Spencer and Danner:
- - -;
Kay Method: -;
Lee and Kesler: - - -; Plocker et al.: .-.-.-;
Hankinson and Thomson: .-.-.-;
Teja et al.:-..-..-;
Perdersen et al.: ....-....-.

--

8.1.2 Calculation of Thermodynamic Properties


and Fugacity Coefficients in a Mixture
The calculation of the properties for a mixture truly presents only one problem: is the system homogeneous? If it is, we may calculate the molar volume by application of the equation of state, and the other thermodynamic properties as if we were dealing with a pure
substance. If it is not, we take the sum of the properties of each phase, for example, for the
calculation of volume:
V = NLvL+ Nvvv
The answer to this question assumes that we know how to solve the phase equilibrium
problem. If we use the same model for this, then we must calculate the fugacities of each
component.

2 76

8. Application of Equations of State to Mixtures

We will illustrate the principle of such a calculation using the simplest rule, from
Kay (3.35):
T , m = Cxi Te,i

and

pC,m =

CxiPc,i

(3.35)

Application of the Lee and Kesler method or of an equation of state such as SoaveRedlich-Kwongallows us, for the homogeneous phase with composition zi, to calculate the
compressibility factor 2, the residual terms for enthalpy and entropy, and the fugacity
coefficient of the mixture that we shall call qm.To do this, we apply exactly the same equations as for a pure substance with the understanding that the reduced coordinates are
defined in relation to the pseudocritical coordinates by Equation 8.3.
In this way, note that we may calculate:
0 the fugacity coefficient of the mixture qm;
0 the fugacity coefficients for the components of the mixture in the pure state, .
9
;
These latter are not to be confused with the fugacity coefficients for the same components in the mixture, cpi.
To obtain these values, we note that:

G-G#
= Nt In qm
RT

or:

where Nt is the total number of moles.


Derivation with respect to the number of moles of component i yields:
lnqi= p.- p! =lnqm+Nt(-) q m
(8.18)
RT
aNi I;cNj
Since the fugacity coefficient qmis a function of reduced coordinates, and possibly the
acentric factor, we have:

However, we have:

T r = -,T

T$m

Taking into account the Kay rule:

and:

therefore

aTr aNi

T aT,m
T t m aNi

2 77

8. Application of Equations of State to Mixtures

Furthermore, by application of the Gibbs-Helmholtz equation we can write:

therefore:

A series of analogous calculations yields:


(8.20)
Finally, applying a linear mixing rule to the acentric factor gives us:

And ultimately:

+(?)

(mi - 0,)

(8.21)

Zf3"

If we apply the Lee and Kesler method that uses the simple fluid properties (o= 0) and
the properties for a reference fluid with acentric factor w"', for any residual term, and
especially for the fugacity coefficient of the mixture, we state:
In q,

= (In q)(O)

+ om

(In q) ( r ) - (In q) (O)

and the preceding equation becomes:

+
8.2

(In q)( r ) - (In q) (O)


(CB~
- w,)
a(')

(8.22)

VlRlAL EQUATIONS OF STATE TRUNCATED


AFTERTHE SECOND TERM

At this point, we shall return only very briefly to these equations. An example of their
application to the calculation of vapor phase properties was discussed in Chapter 5. Recall:
B
(4.13)
Z=l+2,

2 78

8. Application of Equations of State to Mixtures

for the equation derived from development in density, and:

BP
Z=l+RT

(4.18)

for the equation corresponding to the development in pressure.


In either case, the second virial coefficient B is related to the compositions by a quadratic mixing rule, that can be demonstrated by statistical mechanics:

(5.11)
which is, for a binary mixture:

B = B,,,Y? + 2B,,2Y,Y2 + B2,2Y;


The residual properties are calculated with the equations established in Chapter 4. For
the fugacity coefficients in a mixture, and for the development in pressure truncated after
the second term, we have:
RTln q i = ( 2 E B , y i - B ) P
i

(5.56)

As we have already emphasized, these expressions should be applied only to the vapor
phase at low pressure. This application is common when calculating liquid-vapor equilibria
with the equilibrium coefficient expressed as in Equation 6.4, and it is normal to account
for the corrections introduced when we make use of high precision experimental data,
even if the pressure is close to atmospheric pressure.
The main problem is knowing the second virial coefficients, especially their binary
terms Bi,i.We rarely use experimental data for these terms [Dymond and Smith, 19801.
We have already discussed and applied the Tsonopolous correlation (Chapter 3,
Section 3.2.2.1) to pure substances. It may be extended to the calculation of the binary
terms by defining the binary parameters Tc,i,j,Pc,i,j,and
This has been done. We apply
the following equations (among which we recognize some that have been previously
mentioned):
Tc,i,j =

kij)

(8.23)
(8.24)

where:

and:

(8.25)

and finally:

(8.26)

8. Application of Equations of State to Mixtures

8.3

2 79

EQUATIONS OF STATE DERIVED FROM


THE VAN DER WAALS THEORY

The range of application for the virial equation of state is narrow.


It is certainly possible to calculate the liquid-vapor equilibria under pressure by
applying one of the extensions of the corresponding states principle mentioned above,
especially the Lee and Kesler method [1975] as modified for this purpose by Plocker
et al. [1978].
However, it is the equations of state derived from the van der Waals theory that have
provided the most satisfactory solution. It is the most commonly applied solution to this
problem. Here we shall look at the equations using the general expression:
RT
p=-v -b

a
(v- br,) (v - br2)

(4.50)

in which the values for parameters rl and r2 are characteristic of the equation in question,
as seen in Table 4.6.

8.3.1 The Classical Mixing Rules


In general, we combine these equations of state with the so-called classical mixing rules:
a =CC.ai,jzizj
i

(8.27)

b = xbizi

(8.28)

where:

aV = V

G ( 1 - kU)

where kj,, = ki,j

(8.29)

In the preceding expressions the terms ai,i and bi are the parameters for the pure substances.The calculation of the alj binary terms makes use of parameter ki,j.,called the interaction parameter, determined from experimental data for phase equilibrium. We shall
come back to this term.
As before, the composition of the mixture is designated by zi.When the system is separated into two phases, the Equations 4.50, and 8.27 to 8.29 are applied to each phase. It is
understood that the corresponding composition,xi for the liquid phase, and yi for the vapor
phase (see Section 8.5) are at work. At a given temperature and pressure, the equations of
state may be solved, and the proper root assigned to the phase in question. The thermodynamic properties are then calculated for one and the other phase by applying the equations from Table 4.6.
It is important to point out the relationships that exist between the classical mixing rules
and the regular solutions theory. They make the following calculation possible.

2 80

8. Application of Equations of State to Mixtures

EXAMPLE 8.1

Classical mixing rule and internal energy of mixing


Using the very simple case from the van der Waals equation, given that the parameters a and b are temperature independent, we calculate the variation of internal
energy accompanying the mixing of two components, z1moles for component 1,and
z2 moles for component 2, in order to form one mole of mixture. The mixture and the
pure components are at the same temperature and under conditions such that the
reduced densities are the same:
b
b
b
77=1=1=211

'u2

'u

To calculate the residual energy we use equation:

Residual internal energy (parameters a and b being temperature independent) is

The variation of internal energy is therefore equal to:

where the subscripts are for the pure components. The term:
u # - [Z,U?

+ Z2U?]

is zero because the mixture of ideal gases occurs without internal energy variation.
What remains is:

Taking the mixing laws (8.27) through (8.29) into account, and taking the interaction
parameter kij to be zero, this equation may be restated (after some mathematical
manipulation!) as:

Note that the terms:

..

Z;b;
-

b
are analogous to the volume fractions, and that the ratios:
ai
bi

bi

8. Application of Equations of State to Mixtures

281

are analogous to energies per unit of volume, such that the preceding equation may
be compared to the equation pertaining to regular solutions (Scatchard-Hildebrand
theory, see Chapter 7, Section 7.5):
u =

v q aj2(6, - 6,) 2

(7.17)

where 4 is the solubility parameter, or the square root of cohesive energy density. In
this way the classical mixing laws proposed by van der Waals and the model for regular solutions fundamentallyconform to the same theory of molecular interactions;the
ranges of application are the same: mixtures of apolar compounds.
Mixing rule applied to the translationparameter

We have seen [PBneloux,19821 that any equation of state may be corrected using translation. For this, we have said that the result of the equation of state before correction was an
intermediate calculation, a fictitious volume, designated v,and that the volume was
obtained according to equation:
v=v-c
(4.79)
The value of the translation, c, is independent of pressure. Under these conditions, for
pure substances the calculation of the liquid-vapor equilibrium, in other words the vapor
pressure, is not modified.
For this parameter, we also need to define a mixing law. We shall do this in such as way
that, for mixtures as for pure substances,the translation does not modify the calculation of
phase equilibria. For this we have:
c =ccizi
(8.30)
i

Indeed, if we calculate the fugacity coefficients from the classical equation:


RTln 2
f. =

lo

(Zi - --)dP
RT

PZi

(5.54)

we note that the partial volumes are themselves translated from ci, which does not
depend on composition:
- 0. = v!- c .
I
I
1
such that the fugacities,before and after translation, are related by:

If the liquid-vapor equilibrium condition is verified before translation:

fl!L(ze4
=fi(Te:y)
then it will also be verified after translation:

fC(ze4=fY(zer)
Let us point out once again that correction by translation and this definition for a mixing rule relating to translation parameters are not exclusive to equations of state derived
from the van der Waals theory, but apply to any equation of state.

2 82

8. Application of Equations of State to Mixtures

8.3.2 Calculation of Chemical Potentials and Fugacity Coefficients


Phase equilibrium assumes equality of temperature, pressure, and chemical potentials (or
fugacities). If we wish to evaluate chemical potentials using the definition equation:
(5.13)
we end up with the equation:
R T l nf.
L = [ o ( < - pRT
)dP
PZi

(5.54)

Yet, when we apply the same equation of state both to the liquid phase and the vapor
phase, it cannot be explicit in volume, and we cannot easily calculate the partial molar volumes. We therefore prefer to start with the equation:
(5.23)
where A is Helmholtz energy.We also know (see Equation 1.44a) that:

We therefore obtain:

The same goes for an ideal gas:

and:
However, in the latter residual term, the real fluid and the ideal gas are considered to be
under the same conditions of temperature and volume. To obtain the fugacity coefficient,
it is necessary to consider the residual term that corresponds to the same conditions of
temperature and pressure, and therefore add:

2 a3

8. Application of Equations of State to Mixtures

Thus we obtain the definitive equation:


R T l n r p , = ~ m vaNi
[ - (T,N~
~ )+ y ] d V - R T l n Z

(8.31)

This equation is applied to the calculation of fugacity coefficients each time that the
equation of state is explicit in pressure, and among others, to those illustrated by Equation
(4.50). It is therefore necessary to express the total number of moles N,, and the total volume V in the equation. Equation 4.50 therefore will be written as:
N:a

N,RT
p = -V- N,b

(8.32)

(V- N,br,) (V- N,br2)

Equations 8.31 and 8.32 are independent from the selected mixing rule for the equation
of state parameters, and are also to be applied when we shall define the mixing laws from
the excess quantities (Section 8.4). We detail below the application in the case of classical
mixing rules.
The mixing rules (Eqs. 8.27 and 8.28) yield the expressions:

N,b=zbiNi
1

which we substitute into Equation 8.32. We may then calculate the partial derivative of the
pressure as it relates to the number of moles and integrate according to Equation 8.31.
The result is:
lnqi=-ln

P ( v - b)
RT
~

bi
+ -(Z-1)+
b

:)

a
( 2i
zyzj v - br,
(8.33)
bRT(r, - r2)
v - br,

8.3.3 Application Range and Results


At least in principle, the van der Waals equation of state is applied only to non-polar compounds, and there is even more of a reason when the classical mixing rules that we have
defined are applied to it. This leaves us with an enormous range, especially petroleum fluids. Whether in the exploitation of reservoirs or in the refining industry, the Soave-RedlichKwong or the Peng-Robinson equations of state are commonly used. They have been the
subject of numerous modifications,which we briefly discussed in Chapter 4, but as long as
we limit ourselves to non-polar compounds,the classical mixing laws are the most used.
At this level, the main problem is the determination of the interaction parameter kV.We
shall return to this point.
The results are generally good. Figure 8.2 (identical to Figure 6.9) shows the equilibrium
diagram for the ethane carbon dioxide system. For this calculation, the value of the interaction parameter is equal to 0.13 and independent of temperature. We observe that the
experimental values are well correlated, as are the azeotropic behavior and the area close

284

8. Application of Equations of State to Mixtures

6
T = 293.15 K
T = 283.15 K

n
2

ln

p!

\\

T = 263.15K

2
I

0.25

0.5
0.75
Mole fraction of ethane

Figure 8.2 Liquid-vapor equilibrium diagram of the ethane (1)


carbon dioxide (2) system.
Calculated curves, experimental points.

to the critical region. Similarly,Figure 8.3 [Huron et al., 19771shows the critical points locus
for carbon dioxide, hydrogen sulfide, and paraffin hydrocarbon mixtures.
It can be tempting to extend this method to other compound families.Asselineau ef al.
[1978] showed that the Soave-Redlich-Kwong method could be applied to mixtures of
chlorofluoro-compounds, after determination of the interaction parameter of course.
Figure 8.4 shows that this is the case for the CF,, CHF, mixture. Since this result, many
other systems from the same family have been described in this way. Goral et al. [1981]
have even applied this method to more polar systems and, according to them, only mixtures containing oxygenated compounds had unacceptable results.
We believe caution is called for in this area. A priori, the classical mixing rules should be
applied to non-polar systems only, and either after determining the binary interaction
parameter, or after having verified that this parameter may be assumed to be zero.
In any event, a recommendation is in order. The liquid-vapor equilibria of a mixture will
not properly correlate or be predicted if the vapor pressures of the components are not
themselves correctly calculated.We have seen that application of the Soave-Redlich-Kwong
or Peng-Robinson equations of state in their original form yields an average error on the

285

15

a"

Eg! 10
3
(D
fn

g!

-1 00

+loo

+200

+300

Temperature (OC)

-a"

15

E
f 10fn
(D

g!

5-

-25

50
Temperature IOC)

100

Figure 8.3 Critical points loci for carbon dioxide, paraffins C,-C,, (top)
and hydrogen sulfide, paraffins C,-C, (bottom).
Calculated curves, experimental points [Huron et aL, 19771.

286

8. Application of Equations of State to Mixtures

0.25

0.50

0.75

Mole fraction of CF4

Figure 8.4 Liquid-vapor equilibrium diagram of the carbon tetrafluoride,

carbon trifluoride system.


Calculated curves, experimentalpoints [Asselineau et al., 19781.

order of 1to 2% within the range of temperature between the boiling temperature at atmospheric pressure and the critical temperature. These results worsen considerably at lower
temperatures, and we must pay attention to this point for the treatment of systems containing compounds with low vapor pressure or, of course, compounds other than hydrocarbons.

8.3.4

The Binary Interaction Parameter

In theory, the interaction parameter k,, within the calculation for the binary term abi:
ai,j= %(I-

ki,,)

(8.29)

must be determined from the experimental liquid-vapor equilibrium data. It may be neglected in the case of paraffin mixtures. Elsewhere in the literature we find numerous pub-

8. Application of Equations of State to Mixtures

287

lications in which laboratory results are correlated using the Soave-Redlich-Kwong or


Peng-Robinson equations of state, thus determining the optimal values for this parameter.
It is particularly appropriate to mention the works of Grabowski and Daubert [1978a,b, c],
Moysan et al. [1986], and above all the Knapp et al. compilation [1982].
Yet there are still problems.
In the first place, take the case of the methane paraffin mixtures. Figure 8.5 shows the
equilibria lenses of the methane hexadecane system calculated using the Peng-Robinson
equation, as well as the experimental values published by Glaser et al. [1985]. An initial
calculation was done with a zero value for the interaction parameter, and yielded a critical
pressure value slightly higher than 400 bar. It is clear that the experimental values for bubble pressure are continuously underestimated. A second calculation was done with
ki,j= 0.06. The experimental values are pretty well represented, with the exception of the
critical zone because the critical pressure was probably overestimated. It therefore seems
that we cannot neglect the interaction parameter to represent binary liquid-vapor equilibria of binary mixtures formed from methane and paraffins.

80

60

5 40
e!

3
u)

e!

20

A''
I

0.5
0.75
Molaire fraction of methane

0.25

Figure 8 5 Liquid-vapor equilibrium diagram for the methane (1)


hexadecane (2) system.
Calculation using the Peng-Robinson method.
T = 300 K. - - - - - :kbl. . = 0;
: ki,j= 0.06.

2 88

8. Application of Equations of State to Mixtures

This result deserves to be tempered however. In effect, the preceding calculation is the
result of the application of the Peng-Robinson equation as it was proposed initially. Under
these conditions, we should note that the parameters of the pure substances are relatively
uncertain. Methane is found well above its critical temperature and the application of the
equations proposed by Soave (expressions 4.43,4.46, and 4.47), and taken up by Peng and
Robinson (numerical values from Table 4.5), warrants caution. The same is true for hexadecane, for which the reduced temperature is on the other hand very low. The vapor pressure of this compound is poorly predicted by these expressions. For this reason, some
authors [Pedersen et al., 19891 have preferred to take the interaction parameter to be zero
for all hydrocarbon binary systems.Yet, it is to be observed that this proposition is applied
within the framework of a study dedicated to reservoir fluids, complex mixtures by virtue
of the number of components, and accompanied by a special treatment for the heavy fractions. Additionally, for such mixtures the number of components is very high, and the simplification afforded by this hypothesis allows for an appreciable reduction in calculation
time [Hendricks, 19881. Table 8.2 gives the interaction parameters recommended by these
authors for binary systems containing nitrogen, carbon dioxide, and hydrogen sulfide.
Table 8.2
Soave-Redlich-Kwong equation of state;
interaction parameters between hydrocarbons and nitrogen,
carbon dioxide, and hydrogen sulfide [Pedersen eta/.,19891
Component

NZ

NZ

0
0
0
0.02
0.06
0.08
0.08
0.08
0.08
0.08
0.08

COZ
HZS

c,
CZ
c
3

i-C,
n-C,
i-C,
n-C,
n-C,

0
0
0.12
0.12
0.15
0.15
0.15
0.15
0.15
0.15
0.15

0
0.12
0
0.08
0.07
0.07
0.06
0.06
0.06
0.06
0.05

The hydrogen containing binary systems, frequently encountered in the refining industry, also pose a difficult problem that has been the subject of several studies [Grabovski and
Daubert, 1978c;Moysan et al., 19831.Firstly, we cannot satisfactorily describe the entirety of
the equilibrium diagram. If we limit ourselves to pressures below 300 bar, it is then possible
to obtain valid results, but the values for the interaction parameter are especially high. For
example, k , = 0.742 for the application of the Soave-Redlich-Kwongequation to the hydrogen n-butane binary system at 120C.These parameters are only slightly sensitive to the
nature of the hydrocarbon component, but they vary with temperature. Finally, and most
importantly, in opposition to what is generally observed, the results are only slightly sensitive to the value of the ki,jparameter. All these observations are pretty well explained if we
take into account the very high value for the reduced temperature of hydrogen (T, = 33 K).
The application of Equations 4.43,4.46, and 4.47 proposed by Soave yield a very low attrac-

8. Application of Equations of State to Mixtures

289

tion parameter a value for hydrogen. It is therefore the same for the binary parameter ai,j
obtained by applying mixing rule 8.29 if the values of k , are not themselves especially high.
Another, more general problem concerns the variation of the interaction parameter k ,
with temperature. This variation is often neglected. In any event, it generally remains moderate. Yet, it needs to be emphasized that it may have an importance that is not negligible
for the calculation of mixing enthalpy.
The values for the interaction parameter are most often determined by minimization of
the differences between experimental data for bubble pressure and the results of the calculation. Since the data relative to composition of the vapor phase (dew points) are more
rare, they are not generally taken into consideration. If we are interested in light-heavy
binary systems such as the methane hexadecane system, the content of heavy compound
in the vapor phase is very low (Fig. 8.5), and the relative error for this value could be considerable, and cannot be controlled without experimental data. This fact must be remembered when we seek to predict the retrograde dew pressures of natural gases. The sensitivity of this value to the content of heavy component is very high, and often the calculation
is only performed thanks to a special adjustment of the interaction parameters, or even of
the nature of the heavy components. We must therefore emphasize the value of predictive
methods specific to this problem [Pedersen et al., 19891.

EXAMPLE 8.2

Relationship between the Henry constant and the binary interaction parameter
It has also been suggested to determine the interaction parameter from the Henry
constant, meaning from gas solubilities in the hydrocarbon solvents. This constant is
defined by equation 5.86:
(5.86)

If x1 + 0, and if the solvent is a pure substance x2 + 1, under the Henry constant


measurement conditions, the pressure approaches vapor pressure P; of the solvent.
Equation 8.33 becomes:

a22

2-

v; - b2r,
(1 - k1,J - %) In
b2
u;- b2r2

and it is understood that the values for P; and u; validate the equation of state and
the liquid-vapor equilibrium equation for the pure solvent.
We see that the Henry constant is explicitly related to the interaction parameter, as
well as to the properties of the solvent at saturation. This method is excellent for calculating the portion of the equilibrium diagram that corresponds to the diluted zones,
but it is not generally possible to predict the diagram as a whole, and the critical zone
in particular, in this way.

290

8. Application of Equations of State to Mixtures

8.3.5 Alternatives on the Classical Mixing Rules


8.3.5.1

Dependence of the Attraction Parameter on Composition

As we have pointed out, the classical mixing rule relating to the attraction parameter:
a =CCaUzizj
i

(8.27)

ai,j= d K ( 1- ku)

where kj,i = ki,j

(8.29)

may not be applied to polar mixtures called asymmetric. Several modifications have
been proposed that involve a variation of the interaction parameter with composition.
Panagiotopoulos and Reid [1986] suggest a linear variation:
I,]
l,]
Id
51 I
k..=k..+(k..-k..)z.

wherek;.=O

(8.34)

in which we point out that in a system with n components, it is not invariant if we interchange compounds i and j.
Adachi and Sugie [1986] avoid this deficiency by stating:
k41. . = k..
+ 1.41.(z 1. - 2.)
41
I

where kI!d. = k..


1.. = -1.41., k..
= 1.. = 0
Vh
41
51

(8.35)

Stryjek and Vera [1986] represent the variation of ki,jas:


41 = k!.z.
v 1 + k!.z.
1.1 I
k..

(8.36)

For a binary system, these last two equations are equivalent.


Finally, Schwartzentruber and Renon [19891introduce three parameters per binary:
(8.37)

As Michelsen and Kistenmacher [1990a] have shown, these mixing rules have a fundamental flaw. For example, if we apply the Adachi and Sugie model and we consider a ternary mixture having two identical components (subscripts 2 and 3) so that:
a2= a3
k;,3 = 4 3 = 0, k;,3 = k;,2,4,3 = 11,2
the results of the calculation of parameter a relative to the mixture, and therefore any calculation using the equation of state, will depend on the relative proportions of compounds
2 and 3. Hence, we cannot apply these models to complex mixtures (petroleum fluids) that
often have very similar compounds.
These models pose another problem, i.e. dilution. By applying Equations 8.27,8.29,
and 8.34 to 8.37, we notice that the terms depending on zi or zj become negligible if the
number of compounds increases substantially, which again, is the case with petroleum mixtures.
In summary, we cannot advocate the application of these mixing rules despite the
indisputable flexibility they bring to binary data correlation. However, we note that alter-

8. Application of Equations of State to Mixtures

291

natives have been proposed [Mathias, 19921that enable us to circumvent the deficiencies
just highlighted. Yet they remain entirely empirical.

8.3.5.2

Application of a Quadratic Mixing Rule to the Covolume

Application of a linear mixing law to the covolume:


b =2bizi

(8.28)

is suggested by the fact that the experimental measurements of density in liquid phase at
high pressure show that excess volume is low, and decreases (in absolute value) as pressure
increases.
Nevertheless, it has been proposed to apply a quadratic mixing rule to the covolume b.
It is similar to the one being used for the attraction parameter:
b =CCbi,jzizj
i

(8.38)

If the binary term is calculated using the arithmetic mean of parameters b,, and bj,j,this
rule is identical to the linear combination 8.28. More generally:
(8.39)
introduces a second interaction parameter li,j,and confers added flexibility to the model.
In practice, it ends up that in this way we can acceptably correlate the liquid-vapor equilibria of systems with polar compounds, such as the acetone water system. However, we
believe that caution is called for when applying this rule and, in general, it should be
avoided. If the system in question contains only apolar components, we often run into the
intercorrelation phenomenon of the two parameters ki,jand li,j,since several pairs of values plot the equilibrium diagram in a more or less equivalent fashion. Their values are
therefore indeterminate. In the presence of polar compounds, this intercorrelation is
minor or does not exist, but sometimes we wind up with a high value for parameter li,jthat
leads to an unrealistic value for excess volume in the dense phase. Indeed, for a dense
phase, at high pressure, for example, or even for a liquid phase far from the critical zone,
when the vapor pressure of a system does not exceed a few bars, the densities are influenced directly by the value of the covolume. Again, such is the case in the acetone water
system for which the optimal values for parameters ki,jand li,j result in an excess volume
close to 20%.
Yet, sometimes we have access to the excess volume values. Under these conditions,the
application of a quadratic mixing rule to the covolume is justified, and may even prove
necessary to simultaneously correlate the equilibrium data with the density data.
Finally, we note that in applications involving the theory of rigid spheres, Equation 8.38
is frequently applied using:

which is the equation that corresponds to molecular diameter additivity.

292

8. Application of Equations of State to Mixtures

8.3.6 Calculation of the Thermodynamic Properties of the Mixture


As we have pointed out, the mixing rules that are applied to the equations of state derived
from the theory of van der Waals have been tested on the calculation of liquid-vapor equilibria, and it is from the equilibrium data that we determined the optimal values for the
interaction parameters that operate within these rules. It is often the case that we also calculate the other thermodynamicproperties of a mixture: density, enthalpy, or specific heat.
Such is particularly the case when we are dealing with the experimental data that relate to
a reservoir fluid: volume deposit during retrograde condensation,for example.
We have made it clear (Chapter 4) that for pure substances, the calculation of densities
by this type of equations of state yields acceptable results, but not very precise. The same is
true for mixtures, and the results are weakly sensitive to the interaction parameter values
from the classical mixing rule. Note, however, that if we calculate properties at saturation,
since they (bubble temperature or pressure, for example) are strongly influenced by the
value of parameter ki,j,the same will be true, indirectly,for the calculated density.
We have only a small amount of data concerning the other properties, such as enthalpy
or specific heat. In general, the results appear to be acceptable.
Remember that we can make use of a composite method. We can calculate the phase
equilibria using the Soave-Redlich-Kwongor Peng-Robinson method, and then the properties of the phases in equilibrium by another method, for example, the Lee and Kesler
method. Even if thermodynamic coherence of the results is no longer assured, their precision is generally improved.

8.4

MIXING RULES AND EXCESS FUNCTIONS

There are two kinds of methods for the calculation of liquid-vapor equilibria.
The first ones apply to each phase a separate model. The fugacities in the vapor phase
are calculated using an equation of state, for example, the ideal gas equation or the virial
equation truncated after the second term. The fugacities in the liquid phase are calculated
from a reference state, the pure substance under the same conditions of temperature, pressure, and physical state, and the laws of ideal solutions are corrected using an excess Gibbs
energy model or an activity coefficient model. There are a number of these models, and
they may be selected in terms of the properties of the solution (polarity, hydrogen bond
association, etc.). The model is sometimes predictive, such as the UNIFAC model. The
method proposed by Chao and Seader [1961] is an example of this. We shall refer to these
models as heterogeneous because the different models used for the two phases are a
reflection of the heterogeneity of the system.They are commonly applied at low pressure,
and have the advantage of flexibility, but are unable to explain the continuity within the
critical zone that exists between the vapor state and the liquid state. Within this region,
heterogeneity becomes incoherence.
On the other hand, the models we call homogeneoususe the same model, most often
an equation of state, for the two phases present. Continuity in proximity to the critical
point is respected.The Soave-Redlich-Kwongequations of state together with the classical

293

8. Application of Equations ofState to Mixtures

mixing rules are the simplest example. Yet we have seen that the range of application is
limited to apolar systems.
Is it possible to combine the advantages of the two methods, both heterogeneous and
homogeneous?
Since we arrive at the same results using different pathways, it is probable that each
implicitly contains the essential concepts of the other. We shall attempt to explain this relationship using the equations of state derived from the van der Waals theory. However, we
note that, in general, the principles may be applied to other equations of state.

8.4.1 Calculation of Excess Quantities Using Equations of State:


The Problem of Reference States
Let us take a binary mixture, heptane benzene, for example, at room temperature and at
atmospheric pressure. Let us designate the reference state (pure substances, same conditions of temperature, pressure, and physical state) by the exponent *. The flagged quantities are for the components;those that are not flagged refer to the mixture. By applying the
Soave-Redlich-Kwong method, it is possible to calculate the molar volumes in the liquid
phase for each of the pure components 2,; and 2,;. The classical mixing rules may be
applied (a zero value for the interaction parameter is acceptable), and we can then calculate the parameters, solve the equation of state, and obtain molar volume v for a mixture of
composition zl, z2.
Taking the difference, we obtain the excess volume:
2,E = 2, - <2,;z,

+4 z 2 )

Yet we note that this equation implies, as we have just seen, the simultaneous application of the equation of state (which is not explicit in volume) to the pure substances,and to
the mixture.
Similarly,we may calculate excess Gibbs energy. To do this, note that the fugacity coefficient is related to residual Gibbs energy by the equation:

f =RTlncp
g(T,P)-g#(T,P)=RTln P

(2.38)

for the pure substances as well as for the mixture. Therefore, we may state:
g E = RT(1n (p,- cziIn cp:)

(5.83)

it being understood that these fugacity coefficients are calculated using the equation:
lncp=-In

P ( v - b)
RT

a
+ 2- 1 + U(v,b,r1,r2)
bRT

(4.57)

If we use the Soave-Redlich-Kwongequation, for each of the pure substances we have:


RT ai,i
p=vr-bi v:(v;+bi)

294

8. Application of Equations of State to Mixtures

and, for the mixture:

In qm=-In

P(vm - b m )
RT

am
+ Zm- 1+ In

b,RT

Vm + b m
~

Vm

Since the temperature, pressure, and composition are given, the preceding equations
and the choice of a mixing rule for parameters a and b allow for the calculation of excess
Gibbs energy. Conversely, we know that the excess Gibbs energy data obtained by application of a model such as UNIFAC, may allow for the calculation of the attraction parameter a of the mixture, and consequently, constitute a mixing law.
In fact, there are difficulties if we apply this method to mixtures containing compounds
with very different volatilities [Lermite and Vidal, 19881,which is the general case with the
Soave-Redlich-Kwong or Peng-Robinson equations of state where the predilection range
is the calculation of the liquid-vapor equilibria at high pressure. This problem is related to
the different physical states to which the equation of state may be applied. We shall explain
this problem using an example.

EXAMPLE 8.3

Characterizing the roots of the Soave-Redlich-Kwong equation of state


Table 8.3 shows some examples of the resolution of the Soave-Redlich-Kwong equation of state applied to the ethane (1) propane (2) mixture at 40C and 25 bar. The
detailed results are provided in Appendix 4.
The first column of Table 8.3 lists the mole fraction of ethane, columns 2,3, and 4 have
the molar volume values, and columns 5 , 6 , and 7 show the fugacities for each component corresponding to the molar volumes. As a first step, we shall analyze the values
provided for molar volume, and we note that they have been arranged into three different columns.The reason for this is to characterize the roots of the equation of state.
According to the parameter values, specifically the temperature for a pure substance,
the temperature and composition for a mixture, and the pressure, a cubic equation
of state may have one or three roots. If there are three roots, their characterization is
not a problem.The smallest root is attributed to the liquid phase, and the largest to the
vapor phase. As for the intermediate value, it has no physical significance (it is found
within the zone of mechanical instability since compressibility is positive). This is what
occurs for the compositions z1 = 0.34 or 0.54. The intermediate root was not used and
the liquid and vapor roots listed in two separate columns. However, it may happen
that the equation of state has but a single root, even if isotherm P(v) shows a minimum and a maximum when the pressure is either greater than the maximum pressure
(the root for the liquid state), or less than the minimum pressure (vapor state). This

295

8. Application of Equations of State to Mixtures

also happens, at any pressure, if the isotherm is monotone (for a pure substance the
temperature is then higher than the critical temperature). In order to characterize a
unique root, we calculate the pseudocritical volume of the mixture, meaning the
value for which, at a certain temperature, the first and second derivatives of pressure
in relation to volume are zero. This volume is related to parameter b by the equation:
v c--

z,

b Q b
and for the Soave-Redlich-Kwong equation of state, using the values from Table 4.6:

3
Table 8.3
Ethane (1) pro ane (2) system at 313.15 K and 25 bar. Molar volume
(cm3.mol- ) and fugacities (bar) as a function of composition

Fugacities

Molar Volume

Z1

102.81

0
11.867

0.18

103.22

6.282
9.753

0.34

105.2

644.1

11.604
7.913

7.464
11.074

0.54

113.78

731.04

17.75
5.691

11.56
7.869

0.64

763.52

13.60
6.218

0.88

821.59

18.14
2.471

0.90

830.70

18.98
1.772

851.94

21.07
0

For this value of volume at the temperature in question (here 40C),we then calculate
the derivative of pressure with respect to volume. If it is positive, the curve P(v) has
two extrema and a comparison of the root with the pseudocritical volume v,, allows
for its characterization. It is liquid if it is less than v,, vapor if greater. If the derivative is negative, isotherm P(v) is monotone. Supercritical for a pure substance, it
may, however, correspond to liquid or vapor states in the case of a mixture. Such a
root therefore may not be easily characterized as long as the liquid-vapor equilibrium
calculation has not been done.

296

8. Application of Equations of State to Mixtures

Referring to Table 8.3, we see that for a composition z1 = 0.18, the equation of state
has only one liquid root with a value of 103.22 cm3/mol,for which we have the corresponding fugacities of 6.282 bar for ethane, and 9.753 bar for propane. For a composition z1 = 0.34, there is one liquid root (105.2 cm3/mol) for which we have the
corresponding fugacities for ethane (11.604 bar) and propane (7.913 bar), and one
vapor root (644.1 cm3/mol) and the associated fugacities for this root of 7.464 bar
for ethane, and 11.074 bar for propane. We also note that in this case, the values for
molar volume or supercriticalfugacities (zl = 0.9 or 1) are continuous with those
corresponding to the vapor state.The table in Appendix 4 is the same as Table 8.3,but
more detailed in terms of composition.
Figure 8.6 illustrates what we have just discussed.It relates to ethane propane mixtures
at 40C. Several curves graphing the change in pressure as a function of molar volume
are shown for several concentrations. Suppose we wished to calculate an excess value
for a mixture containing 34% ethane, at a pressure of 25 bar. We observe that for this
mixture the equation of state has three roots, the smallest corresponding to the liquid
state. For propane, there is a single root, which the graph unambiguously attributes to
the liquid state. However, for ethane, since the temperature (40C) is greater than the
critical temperature, there is only one root whose value is characteristic of a gas under
pressure.Therefore we cannot define coherent reference states for the components of
this system, and if we apply the preceding process to the calculation of excess volume,
we arrive at a very negative value which is characteristic not of deviations from ideality, but very simply of the differences in physical state between the mixture and the
pure components.The same is true for calculating the excess Gibbs energy.
This problem must be solved if we seek to replace the mixing laws with models using
excess functions.

-5

200

600
Molar volume (1x13. mol-1)
400

800

Figure 8.6 Isotherms P ( v ) for the ethane (1) propane (2) mixtures at
313 K. Mole fractions of ethane:0;0.34; 1.

1000

8. Application of Equations of State to Mixtures

297

EXAMPLE 8.4

Analysis of the ideality of the ethane ( 1 ) propane (2)system.


Dependence of fugacities with composition at 3 13 K and at 25 bar
We look at the ideality of this ethane propane mixture. The ideality condition
mentioned by Lewis relies on the proportionality of fugacity and mole fraction. If
we show the dependence of the fugacities for ethane and propane (Table 8.3 and
Appendix 4) in the liquid or vapor phase with composition, we observe that this
proportionality condition is more or less validated, as shown in Figure 8.7. The
equations:
f y = fI?Vyi

or

f / = fI?Lxi

may be applied. However, the reference fugacities may be obtained only from the
equation of state for ethane in vapor (or "supercritical") phase, f;" = 21.07 bar, and
propane in the liquid phase,fitL = 11.87 bar.The other reference states are hypothetical and the corresponding values are obtained by extrapolation of the f ( z ) lines:
fi*L = 32.6 bar, and f;" = 16.7 bar.

0.25

0.50

0.75

Mole fraction of ethane

Figure 8.7 Examination of the ethane (1) propane (2) system ideality.
Variation of fugacities with composition at 313 K and 25 bar.
Vapor phase: - - - - - - A ; Liquid phase:
+;
Linear extrapolations: - - - - - - - - - - -.

298

8. Application of Equations of State to Mixtures

Yet, having verified the quasi-ideality of the mixture and calculated the reference
fugacities, we may calculate the liquid-vapor equilibrium at 40C at a pressure of
25 bar using the equations:

fYY, = f;%
We find that x1 = 0.34 and y1 = 0.53.
We must emphasize that this approximate calculation assumes that ideality has been
verified,which is not generally the case. Later on (Section 8.5), we introduce general
calculation methods for the liquid-vapor equilibrium.
We also need to specify that the ideality mentioned here is not directly related to the
fact that the interaction parameter is zero.

8.4.2 Mixing Rules Derived from


Excess Cibbs Energy at Infinite Pressure
A review of Figure 8.6 shows that at high pressure the behavior of a subcritical compound
such as propane, or of a supercritical compound such as ethane, and binary mixtures of
both components, tends toward that of a dense fluid, in opposition to what is observed at
25 bar. For this reason, it has been proposed [Vidal, 19781 that the reference state be
defined as the fluid approaching infinite pressure.
Under these conditions,the molar volume approaches a well-defined limit, covolume b.
The excess volume approaches zero if we apply the linear mixing law to the covolume. As
for excess Gibbs energy, such as it results from Equation 5.83,it approaches a limit whose
expression is very simple:

(8.40)
Parameter A is characteristic of the applied equation of state. For the Soave-RedlichKwong equation, it is equal to In 2.
If we apply the classical mixing rule, we find:
(8.41)

where:

(8.42)

and:

(8.43)

8. Application of Equations of State to Mixtures

2 99

The excess Gibbs energy expression at infinite pressure that results is analogous to the
expression from the theory of regular solutions.Indeed, we note that the ratios:

represent volumetric fractions at infinite pressure, and that the terms:

are, by their nature and numerical values, comparable to the solubility parameters. We
have already established this relationship for the energy of mixing at constant density.
We return to expression 8.40. It is very important to note that this equation is independent of the mixing rule applied to parameter a.
It therefore allows us to define new mixing rules by presenting it in the form:

where g : is an expression for excess Gibbs energy. This expression can be chosen as a
function of the problem, and the nature of the mixture, as we do when we apply a heterogeneous method such as the Wilson equation, the NRTL models, UNIQUAC, or
UNIFAC. Huron and Vidal [1979] have proposed a modification of the NRTL model
where, within the expression for local compositions, the mole fractions are replaced by
volume fractions:

(8.45)

This modification has an important advantage. We have seen that the classical mixing
rule is applied to a large number of systems. It is therefore desirable to select a mixing rule
which, for such systems, can be reduced to the classical rule thanks to a specific choice of
parameters. Such is the case for this modified NRTL rule if we have:

a..=O
>I

c..
=g .. - g . .
Id
61
61

and:
where:
g..=I

(bi+ bj)
2

3 00

0.4

0.3

0.2

0.1
0.25

0.50

0.75

0.25

Y1

0.50

0.75

Xi

y1

Xi

yt

P(MPa)

2.5

2
= 523 K

1.5
0.25

0.50

0.75

X1

y1

0.25

0.50

0.75

Figure 8.8 Correlation of liquid-vapor equilibria for the acetone (1)


water (2) system by application of the classical mixing rule (dotted
curves) and Equations 8.28,8.44,and 8.45 (solid curves, experimental points) [Huron et af.,19791.

The mixing law defined by Equations 8.44 and 8.45 has been applied with good results
to many systems containing polar components. Figure 8.8 shows that in the case of an acetone water system the classical mixing rule yields very poor results, with the prediction of
an immiscibility zone that, in fact, does not exist, while the new mixing rule allows us to
correlate the experimental data up to and within the critical zone, especially by following
the change in the azeotropic composition.The prediction of ternary liquid-vapor equilibria
from binary data is generally possible, as shown in Table 8.4.

301

8. Application of Equations of State to Mixtures

Table 8.4
Liquid-vapor equilibria of the acetone methanol water system.
Average quadratic deviations for the calculation of bubble temperatures.
Experimental data [Griswold, 1949,19521
T(K)

1 Acetone Methanol

Acetone Water

Methanol Water

0.7
0.85
0.9
1.5

0.9
1
0.95
0.7

328-333
323-343

473
523

Ternary Data
.05
.3

.25

Analogous results may be obtained if the excess Gibbs energy at infinite pressure is
expressed by the unmodified NRTL model, or by the UNIQUAC model. The correlation
of binary data and study of their dispersion is also possible using a Redlich-Kister model.
It is appropriate, however, to point out the flaws in this method:
0 We can question the physical significance of a reference state at infinite pressure
since the properties of a fluid can never be measured in this state.
0 The mixing rules thus defined, as stated by the authors, may not verify the quadratic
condition that statistical mechanics imposes on the second virial coefficients. We
know that if we apply the Soave-Redlich-Kwongor Peng-Robinson equation of state,
the second virial coefficient is expressed by the equation:

such that any mixing rule should respect the condition:


a
b -- = CCZit,Bv
RT

where:

This condition is verified by the classical mixing law. However, it is not verified if the
excess Gibbs energy at infinite pressure is expressed using the NRTL, UNIQUAC, or
UNIFAC model.
In fact, it is at low pressure that this quadratic rule is important, and we have never
observed practical consequences due to this deficiency.
We can express excess Gibbs energy at infinite pressure using one of the known and
proved models. However, the parameters of these models must be determined and we
cannot use the values determined from excess Gibbs energy at low pressure such as
they have been assembled in some databases [Gmehling et ul., 19801. Indeed, the values for excess Gibbs energy at infinite pressure and at low pressure are not the same.
From a practical point of view, this is the methods main disadvantage.

3 02

8. Application of Equations of State to Mixtures

8.4.3 Mixing Rules and Excess Functions at Constant Packing Fraction


In light of the remarkable results that they provide, and also because their expression can
be regarded as complex, we shall develop these methods proposed by PCneloux et al.
[1989] in detail.

8.4.3.1

Formulation of Equations of State Derived from


the van der Waals Theory in Terms of Packing Fraction

In general, for given conditions of temperature and pressure, the definition for mixture
quantities is somewhat arbitrary,and its evaluation or interpretation cannot be reduced to
a single balance of intermolecular cohesion energies.This is particularly marked if the stable physical state is not the same for the mixture and all its components, permanent gas,
liquid solvent, or liquid solution, for example. A major element concerning the mixing
quantity may be the variation of energy for the change of physical state, and we could adequately speak of a heat of dissolution. In any formalism concerning mixing quantities,it
is therefore important to carefully select the states of reference such that the model pertains to an unambiguous property. In general, the definition of temperature and pressure is
not sufficient to address this criterion, as we have seen. At infinite pressure, any fluid,
whether supercritical or not, tends toward a dense state, and this limit has been proposed
as a reference. To examine the mixture and its components in the same state of molar volume is unacceptable.The value 40 cm3/molis acceptable for methane at its boiling temperature at atmospheric pressure, but unacceptable for a high molecular weight hydrocarbon.
It is therefore appropriate to take into account the nature of the compound, or more
specifically,its actual or molecular volume. Quantities such as the van der Waals volume, or
even the critical volume, may be proposed for that purpose.
All of the equations of state derived from the van der Waals theory contain a repulsion term that depends on the volume and on a parameter related to the actual volume of
the molecules, covolume b. It is in relation to this parameter that it seems appropriate to
define the reference volume when we study these models.
By definition,in this study we state that packing fraction (reduced density, density) is
the ratio of the covolume to the molar volume:
b
17= -

(8.46)

2)

When the packing fraction is the same for the mixture and its components, they have an
equal proportion of free volume, relative to their actual volume.
The equations of state that we are considering are found in an especially simple form as
a function of this variable. So, if we state:
a

a=bRT
the van der Waals equation becomes:

(8.47)

8. Application of Equations of State to Mixtures

3 03

and the more general Equation 4.50

This observation may be generalized to all the equations of state derived from the van
der Waals theory, which can be written in the form:

z = -1-71
-

(8.49)

a71v471)

For the van der Waals, Soave-Redlich-Kwong,and Peng-Robinson equations, a is independent of packing fraction, and depends only on temperature and, for a mixture, composition. The y function itself depends only on the packing fraction and the chosen equation
of state (via the intermediary of constants rl, r2).

8.4.3.2 Calculation of the Helmholtz Energy A


For an equation of state that is applicable to liquid and vapor states (and explicit for pressure), the corresponding characteristic function is Helmholtz energy, and not Gibbs
energy.
If N is the number of moles, and A and V the correspondingHelmholtz energy and volume, we know that, at a constant temperature and number of moles:

Accounting for the definition of packing fraction, we have:


b h b
q=-=v
v

dV - dq

and

v 7 1

at constant number of moles, such that:


NZRT

dA,= -d71
71

For the ideal gas (denoted by superscript #), the compressibilityfactor is equal to 1,and:
NRT

dA#,=-dq
17

The variation of residual Helmholtz energy with packing fraction is therefore equal to:
d(A -A#), =

NRT(2 - 1)

If 71 + 0, then A -A# + 0, and we derive:

71

d71

3 04

8. Application of Equations of State to Mixtures

Applied to an equation of state such as the one represented by Equation 8.49, this
equation becomes:
(8.50)
A -A# = -NRT[ln (1 - 17) + a5(17)]
(8.51)

where:

Note that in the equation, A - A# must be read as: A(?: Ni,17) - A#(I; Ni,
9).The real
fluid and the ideal gas are under the same conditions of temperature, composition, and
packing fraction. However, they are not under the same conditions of pressure.
So for general Equation 4.50, we shall write:

8.4.3.3

Application to a Mixture and its Components

As in the case of mixing rules created from excess Gibbs energy at infinite pressure, in
order to define the equation of state for a mixture when those of its components are
known, we may proceed in two ways. The first usually is based on the application of mixing
laws relative to the parameters, and generally assumes that the mixture and its components
comply with the same equation of state. We can then calculate the mixing quantities,
specifically the Helmholtz energy of mixing. The second is based on the definition of the
Helmholtz energy of mixing.
Classical approach: the mixing laws are defined

Let us assume that the equation of state for the mixture and its components is defined, that
is to say that parameters bi, b, a,,a as well as functions (or ti),
and y (or 5) are defined.
Such is the case with the usual application of equations SRK or PR, since the parameters
of the pure substances are calculated from the critical points, and the mixture parameters
are calculated using the mixing laws, linear for b, quadratic for a, thereby defining
a = aibRT. In addition, functions y.and Yare identical.
Let us consider the mixture and its components in the same state of packing fraction 17:
-bl- --=b2

u1

212

... - -1 = 7

(8.53)

2,

Therefore, according to Equation 8.50, and by denoting molar Helmholtz energy as a


(this symbol is not to be confused with a, which is the attraction parameter for the equation of state):
a,=a~-RT1n(1-q)-RTai~,(q)

305

8. Application of Equations of State to Mixtures

Therefore, we can calculate the Helmholtz energy of mixing at constant temperature


and packing fraction, defined by:

(w,17) = ~ ( T , x17), -Cziai (T,17)

aM
We find

u ~ ( T , x , =~ () u # - ~ z ~ u-; RT(~{-&cx~{~)
)

The first term corresponds to the mixture in the ideal gas state. However, neither the
pressure nor the molar volume are constant:
b.

v.= 2

and

v=

c 1 7

17

In order to calculate this term, we use a reference state defined by the value for molar
volume, v o ,at the same temperature, of course. If a; stands for the Helmholtz energy of
the components, and uo for that of the mixture, both in the ideal gas state, in this reference
state:
V.
bi
= up- RT In a (?;q ) = up- RT In

a#(T,q) = ao- RTln-

and:

v17

V0
V

therefore:

= ao - RTln-

VOV

V0

= (ao-zziap) - RT (In b - X z i In bi)

a#- &a:

Furthermore, we know that for an ideal mixture of ideal gases achieved at constant volume (or pressure):
ao -xziap = RT&

In zi

such that we obtain:


aM(T,z,q)= RT

Xzi In zi+&In

b.

(a{-z~,a~{~)

(8.54)

We note that this expression in broken down into three terms. The first represents the
ideal mixture, the second has been introduced for the calculation of the Helmholtz energy
of mixing at constant packing fraction and is identical to the expression for the Flory combinatorial term, and the third is sometimes characterized as residual. It refers to the
attraction terms that exist in the equation of state:
ftt

-E

(TZ,17) = - ~ ~ ( a 5Ziaiti)

(8.55)

This term may therefore be calculated within the scope of the initial hypothesis, the
equation of state defined for the mixture and its components.

306

8. Application of Equations of State to Mixtures

Definition of the equation using the attraction term

Let us look at the case where for each component the equation of state (terms b, ,a , , 5,)
is defined. In addition, let us adopt a linear mixing law for the covolume:
b=

C bizi

(8.28)

However, the values for a and 6 relating to the mixture (or rather their product) are not
de6ned
We shall define the expression for the function uzt as a function of temperature, packing fraction, and composition.The product & will be derived by applying Equation 8.55 in
the form:

(8.56)
as well as the product a y b y application of Equation (8.51):

The function u z must satisfy certain conditions. It is analogous to an excess function,


and approaches 0 if the mixture reduces to a pure substance (xl + l),or if the packing
fraction approaches zero.
It depends on temperature, packing fraction, and composition. It is possible to have
these last two variables intervene simultaneously by defining "DDLC", or density dependent local compositions. However, from a practical point of view, we prefer to dissociate the
effects of composition and packing fraction by stating:

azt(rz,17)= @ ( T , Z ) - ! m

(8.58)

F(q) is a function that depends only on packing fraction, and approaches zero at the
same time as this variable. GE(T,z ) depends only on temperature and composition, and
cancels itself out for xi = 1(pure substance).It seems reasonable to express dependence of
the attraction term with composition, using an excess Gibbs energy model.
The preceding equations show that the terms a5 and a y are related. It is necessary to
point out an important special case,since it in fact constitutes the rule: for all components,we
choose the same equation of state (the same functions or ti,regardless of i), and for this
equation of state the function 5 depends on packing fraction only. In this case, and with no
other information,it is logical to apply this same function to the expression F(q).By stating:
(8.59)
Equation 8.56 becomes:

8. Application of Equations of State to Mixtures

307

that is to say, after eliminating function 5:

or:

(8.60)

and we can easily notice, in this case, the identity that exists with the use of excess Gibbs
energy at infinite pressure (Eq. 8.44).

8.4.3.4

Results: Abdoul Group Contributions Method

Of course, we may apply this method to the correlation of binary data for polar systems
[Lermite and Vidal, 19881, using the UNIQUAC model, for example, for representing the
GE(TJ) function.The results are analogous to those obtained using excess Gibbs energy at
infinite pressure and the modified NRTL model. It is more interesting to look at the
method of group contributions proposed by Abdoul et al. [1991].
The choice of an excess function model at work in Equation 8.60 and using group
contributions is not necessarily a simple one. At first glance, the UNIFAC [Fredenslund
et al., 19771and the Lermite and Vidal[1992] or ASOG [Kojima and Tochigi,19791models
come to mind. Yet, none of these models reduces to the classical mixing law, whose validity
is well known in the case of apolar systems. Furthermore, for such systems the two interaction parameters between groups are strongly correlated, and due to this fact, poorly determined. The importance given here to apolar systems is in no way exaggerated. They provide the groundwork for our databases and the practical application of calculation methods for phase equilibria at high pressure.
Since the classical mixing rule is conveniently applied to such systems, it is appropriate
to analyze it in terms of excess functions. Equations 8.41 to 8.43 are absolutely equivalent
to the classical mixing law [Vidal, 1978;PCneloux, 19891.Adjusting the experimental data
or predicting the binary parameters ki,jor Ei,jis therefore the same,but the Ei,jparameters
are directly related to excess Helmholtz energy. It is therefore this parameter that we shall
express by group contributions. Abdoul et al. [1991] applies a modification of the
Guggenheim reticular model to it, and as first approximation states:
(8.61)
where:

(8.62)

where ai,jis the proportion of groups k in component i. The interactions between groups,
A , , , in principle depend on temperature, and their values are determined from a large
experimental database of liquid-vapor equilibria at low pressure, high pressure, and heats
of mixing. The components considered include nitrogen, methane, ethane, carbon dioxide,
and hydrogen sulfide,each of which make up a distinct group, and the paraffin family, represented by three groups, plus cyclanes and aromatics, for a total of 17 groups. However, it

8. Application of Equations of State to Mixtures

308

was observed that it is necessary to account for a chain length effect with the addition of
a second empirical term. Ultimately we have:
a
a..
1
(8.63)
= -- b x x @ i q E i ,+
j E,
b
bi
2
i
j

-XZ,*

where E2 is determined from the principal hydrocarbon chain lengths:

The results are of remarkable precision. Figure 8.9 illustrates the distribution of errors
at bubble pressure as a function of the number of systems in question.

:i
I

50

High
pressures

Low
pressures

10

dP/P(%)

dP/P(%) 10

Figure 8.9 Application of the Abdoul method to the prediction of


liquid-vapor equilibria for hydrocarbon systems under pressure.
Distribution of the errors concerning bubble pressure.

Essentially, this model contains two hypotheses. First, the validity of the classical mixing
rule corrected by the chain length term, the importance of which is generally low. Then,
most importantly, the possibility of applying a group contributions method to the intermolecular interaction parameter Ei,i.
We can attempt to extend this method to systems other than hydrocarbon mixtures provided, however, that for such systems, the classical mixing rule is applicable.
This has been done [Fransson,19931for mixtures containing chlorofluoro-hydrocarbons.
It is known [Asselineauet al., 19781that the Soave method combined with the classical mixing rule allows us to calculate the liquid-vapor equilibria of such mixtures.This shows that
for each system it is possible to determine an interaction parameter value kY or Ei,j such
that the representation of the equilibria is acceptable.This Ei,jparameter is predicted by
the expression 8.62, and we added the groups below to those defined by Abdoul et al.:
CCl,F,, CClF,, CHClF,, CHF,, CClF,, CCl,F, CHF2,CF,, CF,

8. Application of Equations of State to Mixtures

309

To conclude this section, it seems to us that the Abdoul method is the best current
approach for the prediction of liquid-vapor equilibria for non-polar mixtures.
We also note that nothing prevents application of different equations of state to the various components of the system,better adapted and more specific than the Soave-RedlichKwong or Peng-Robinson equations of state. To illustrate the solubility of heavy hydrocarbons in supercritical solvents,Barna et al. [1994] have ascribed a very specific equation of
state to supercritical solvents, such as those introduced in Chapter 4, Section 4.3, and a
generalized cubic equation of state to heavy hydrocarbons.
We have just introduced the definition for mixing rules using excess Gibbs energy at
infinite pressure, or by using Helmholtz energy at constant packing fraction. These methods have a deficiency in practice, which we have already pointed out. The numerical values
for parameters in the excess function expression must be determined by experimental data
regression. Nevertheless, at least for equilibria at low pressure, there are large databases
that provide such numerical values for the Wilson equation, and the NRTL and
UNIQUAC models. They cannot be used because excess Gibbs energy at infinite pressure,
or Helmholtz energy at constant packing fraction, do not have the same value as excess
Gibbs energy at low pressuce. The same is true for the UNIFAC model.
The two methods that we shall now discuss seem to provide a solution to this problem.

8.4.4 The "MHVZ" Method


The initial answers to this problem are the results of observations by Mollerup 119861,
Heidemann and Kokal[1990], and finally Michelsen [199Ob].We shall look at these observations.
The linear mixing law is applied to the covolume:
b=

bizi

(8.28)

We have seen (Chapter 4, Example 4.5) that an equation of state such as the SoaveRedlich-Kwong equation may have two roots at zero pressure. For this, the temperature
must be less than a value for which the parameter:
a
bRT

a =-

(8.47)

is higher than the well defined limit:


a>3+2fi
We shall designate the smallest of these roots (in molar volume) as wo . It corresponds to
the liquid state, its compressibility is very weak, and it provides an excellent approximation
of the value of molar volume in liquid phase at a pressure close to atmospheric pressure.
If we say, that uo is the ratio:
(8.65)

310

8. Application of Equations of State to Mixtures

it is related to the value of a by the equation:


1
u0(a)= - [ ( a - 1 ) - ~ a 2 - 6 a + 1 ]
2

(8.66)

The fugacity coefficient applied to the Soave-Redlich-Kwongequation of state may be


written as:
-b
v+b
bf =-In v+Z-1-aln In b
RT
b
and at zero pressure:
bfo
In - =-ln(uo-1)-1-aln
RT

uo + 1

-=Q(u,(a),a)=q(a)

(8.67)

UO

for the mixture, and:

u +1
In bf0,i = - l n ( ~ ~ , ~ - l ) - l - a ~ l n
(8.68)
= Q(uO(ai),ai)=q(ai)
RT
%,i
for each of its components.
The function q(a)(or q(ai))is strictly the result of Equation 8.66, but this equation can
be applied only if there exists a root at zero pressure.
However, at first approximation and within the range:
10<a<13
it may be expressed by the linear equation:

q(a)= 40 + q1a

and

q(ai) = 40 + q1ai

(8.69)

where q1= -0.593 in the case of the Soave-Redlich-Kwongequation.


If we now calculate excess Gibbs energy at zero pressure:

goE =lnfo-Ezi1nh,,
RT

(8.70)

we obtain:
(8.71)
and, taking into account the approximation of function q(a):
(8.72)

or:

(8.73)

This equation is close to the expression for excess Gibbs energy at infinite pressure
(8.40) or Helmholtz energy of mixing at constant packing fraction (8.54).

8. Application of Equations of State to Mixtures

31 1

If a mixing rule has been previously defined,then the value of g f is defined.Conversely,


if we have the value of gf, then the value of parameter a as it relates to the mixture is
known:
(8.74)

or:

(8.75)

It is important to note that, unlike excess Gibbs energy at infinite pressure or excess
Helmholtz energy at constant packing fraction, excess Gibbs energy at zero pressure
is very close to the excess Gibbs energy of a dense phase (liquid mixture) at pressure
approaching atmosphericpressure. Indeed, if we say that v E is the excess volume, we arrive
at:

Jd

.-D

gE-gf=

vEdP

and this term is absolutely negligible as long as the phase in question is a dense phase, and
pressure P remains low.
We may therefore calculate go using the same models (NRTL, UNIQUAC, UNIFAC)
and using the same numerical values for the parameters as for gE. It is therefore possible to
apply expression 8.74 or 8.75 at the same time using the model (mathematical expression)
and parameters (numerical values) published in current databases and based on the application of heterogeneous methods. The method therefore provides a solution to the problem pointed out above. The results obtained are good [Michelsen,1990bl.
Yet it has been observed [Dahl, 1990al that a quadratic expression of the q(a)function
is preferable:
(8.76)
44 = 40 + q1a + 92a2
where q1 = -0.487, and q2 = -0.0047 in the case of the Soave-Redlich-Kwongequation of
state. Under these conditions, after substitution into Equation 8.71, we obtain:
(8.77)
and the calculation of parameter a relative to the mixture is performed by resolving a
second degree equation, as all other quantities are known. This second version of the
Michelsen method is known as the MHV2 method.
The results are excellent.As an example,Figure 8.10 shows the liquid-vapor equilibrium
diagram of the acetone water system. We applied the MHV2 method and thus combined
the Soave-Redlich-Kwongequation of state and the UNIQUAC model. The parameters
were taken from the Gmehling et al. [1980] databases, and determined by modification of
a heterogeneous model (the vapor phase is represented by the virial equation of state
truncated after the second term, and the liquid phase using activity coefficients).Thissame
graph uses the experimental data from Chaudry et al. [1980]. We can state that they are
perfectly represented, and in this way the excess function has been faithfully incorporated
into the equation of state model.

312

8. Application of Equations of State to Mixtures

80

1
I

0.25

0.5

0.75

Mole fraction of acetone

Figure 8.10 Prediction of the equilibrium diagram for the acetone (1)
n-hexane (2) system using the MHV2 method. T = 323.15 K.

We may have difficulties if some of the components are found at the limits of the value
range for a within which we applied a second degree polynomial expression to the q(a)
expression. It may therefore be preferable to use another q(a)expression, such as the one
suggested by Soave [1992].
Of course, if we are talking about liquid-vapor equilibria under pressure, the system
probably contains some light components (methane, ethane, carbon dioxide, etc.). It is still
possible to apply an excess Gibbs energy model, but the parameters must be determined
for binary systems containing such components, because this could not have been accomplished by a heterogeneous method. Similarly,with equilibria under pressure, temperature can be significantly higher than the range for which the parameters in the database
have been published. We are now extrapolating,and the results may be worse.
It is of course very usual to express excess Gibbs energy using a group contributions
model. Dahl et al. [1990a, b, c] have done just that using the UNIFAC model modified by
Larsen [1987]. Again, in this case the results are reliable, even at high temperature, as
shown in Figure 8.11.This figure shows the acetone water equilibrium diagram up to the
critical zone. We may compare the results to those of Figure 8.8, which were obtained using
excess Gibbs energy at infinite pressure. The interaction parameters of the light compounds and atomic groups were determined, and gas solubilities were predicted for
systems composed of hydrogen, carbon monoxide, methane, carbon dioxide, hydrogen
sulfide, and water. The MHV2 model was also applied to the liquid-liquid equilibria pre-

8. Application of Equations of State to Mixtures

31 3

6-

***** EXPER

MHV2

4-

T = 473 K

m
c

s!

n
2-

"

T = 423 K

0.5
Mole fraction of acetone

Figure 8.11 Prediction of liquid-vapor equilibria for the acetone (1)


water (2) system using the MHV2 method [Dahl et al., 1990al.

diction in the critical zone [Dahl et aZ., 1992a1, and ultimately extended to solutions
containing electrolytes [Dahl et aZ., 1992bl.
In summary, the MHV2 method is based on a dense reference state, but at zero pressure, which allows for the use of previously defined models (expressionsand parameters),
and on the more specific application of the UNIFAC model that provides to it the advantages of the group contributions methods.

8.4.5 The Wong and Sandler Method [I 9921


The method proposed by Wong ef al. [1992 a and b], at least in principle, allows us to solve
two problems that we have already discussed, namely those concerning the rules of excess
Gibbs energy derived at infinite pressure. Like the MHV2 method, it may use the models
(expressions and parameter values) that already exist. In addition, it retains the quadratic
variation of the second virial coefficient. This method is based primarily on the expression
for excess Helmholtz energy at given pressure and temperature.
The residual Helmholtz energy is expressed by the equation:
a -a# = -RT In

P(v-b)
a
v+b
--lnRT
b
v

(8.78)

314

8. Application of Equations of State to Mixtures

Applying this expression to the mixture and its components at the same pressure and temperature,we derive from it the excess Helmholtz energy at the given pressure and temperature:

P ( v - b)
RT - Ci z i In P (vl!- bi)
aE=-RT[ln RT
8
v+b
-In-b
v

a.

c z i 2 In
i

bi

1-

vI!+ bi

(8.79)

which differs from excess Gibbs energy only by the term PvE:

aE = gE - pvE
At infinite pressure, the limit of excess Helmholtz energy is equal to:

;;)

zi 3 In2

a:=-(;-?

(8.80)

or:

(8.81)

Furthermore, if B is the second virial coefficient of the mixture related to the parameters of the equation of state by equation:
a
B
a
B=b-or - = I - (8.82)
RT
b
bRT
we derive:

(8.83)

The second virial coefficient is calculated using a quadratic mixing rule:

B=

CCBvzizj
i

(5.11)

where:

(8.84)
8..

B4.1 . = b4 1. . - 1
RT

and (see Chapter 4):

(4.61)

such that Equation 8.83 becomes:

(1 - li,j)zizj
18
851
\----J

If we select an excess Helmholtz energy model at infinite pressure, once the parameters
of the expression and the li,jparameter are known, parameter b for the mixture is obtained
by application of equation 8.85, and parameter a is obtained by Equation 8.81.

31 5

8. Application of Equations of State to Mixtures

However, it has been observed that if we calculate the excess Helmholtz energy for a
dense phase using Equation 8.79 at increasing pressures, this property varied little with
pressure. By denoting the excess Helmholtz energy as a: for the dense phase close to
atmospheric pressure, we can state:
a: = a;

The excess Helmholtz energy aE, is calculated using any available model (expression and
parameters), such as those discussed in the previously mentioned database from Gmehling
et al. [1980]. These databases must provide the value for g:, and for the li,i parameter such
that the excess Gibbs energy, at low pressure and for the dense phase calculated by the
equation of state, is as close as possible to the excess Gibbs energy calculated by the model.
This procedure is completely predictive when we study a system for which the parameters of an excess Gibbs energy model are available. We may also apply the UNIFAC model
[Orbey et al., 19931.
The authors have applied the van Laar and the NRTL models and obtained good
results, even when extrapolating the calculations beyond the range of temperature for
which the parameters of the model were determined. Of course, in the presence of light
components,as for the MHV2 model, all model parameters must be determined since they
are not found in the literature.
Table 8.5 shows some of the results obtained by using this model.
Table 8.5
Prediction of liquid-vapor equilibria using the Wong
and Sandler method [Hernandez Garduza, 19931

System
Methanol-benzene
Aceton-water
Acetone-methanol
Water-methanol
Water-ethanol
Water-propanol

T (K)
313-493
313-523
373-473
373-523
423-623
423-573

SP (%)
2.45
1.84
1.65
1.59
2.22
3.13

SY
0.02
0.08
0.021
0.012
0.012
0.023

Note that the mixing rule relating to the covolume is no longer linear. Due to this fact, it
may yield unrealistic values for excess volume in the dense phase, and profoundly alter the
influence of pressure on phase equilibria.

8.4.6 Advantages and Disadvantages of Mixing Laws


Derived from Models and Excess Functions
The relationship established between the heterogeneous and homogeneous methods has
proved advantageous. The calculation of phase equilibria under pressure using the equations of state may thus be extended to mixtures containing polar components.The group
contributions methods may be applied to the equations of state, where they reinforce their
predictive ability.

316

8. Application of Equations of State to Mixtures

We have discussed four alternatives. We must make a special case for the Abdoul
method that concerns only mixtures of non-polar compounds, but which, for these systems,
seems to us to be the most precise.
It would be interesting to make an extensive comparison of the MHV2 method and the
method proposed by Wong and Sandler since they are both characterized by the possibility of using the models and parameters that have already been determined from equilibrium data at low pressure. These methods have been investigated by Knudsen [1992],
Hernandez Garduza [1993], and Huang and Sandler [1993]. The MHV2 method is more
developed, and combined with the UNIFAC model, has undeniable merits.
This does not mean that all the problems have been resolved. We would like to mention
two problems that do not, in our estimation, currently have an acceptable solution.
The first has to do with mixtures containing molecules of very different sizes. The
results deteriorate as this difference increases. For example, such is the case with methane
paraffin systems. Here we have a range similar to polymer solutions. The difficulties are
possibly due to an erroneous evaluation of the combinatorial term and the free volume
term (see Chapter 11) in the understanding of excess quantities or the mixing rules.
The second example is three-phase equilibria, and in particular, equilibria between
methanol and the light hydrocarbons, methane, ethane, or propane. Figure 8.12 shows the
equilibrium diagrams for systems made up of methane and either ethane or ethylene
[Ma and Kohn, 1964; Zeck and Knapp, 19861.We notice that, within a temperature range
close to the critical temperature of the hydrocarbon, there is a three-phase equilibrium
zone. In this instance, no method can truly account for the experimental data in a satisfactory fashion. We must add that this problem is an important one from an application point
of view since methanol is used as a solvent for the treatment of natural gases by serving as
an anti-hydrate additive. The problem is further complicated by the presence of water.

8.5

CALCULATION OF LIQUID-VAPOR EQUILIBRIA

Liquid-vapor equilibria under pressure are most generally calculated using equations of
state. The parameters involved in the mixing rules are determined using experimental
equilibrium data. In the preceding discussion, we considered this problem resolved.
First, we go back to the calculation principles and the equations related to the model,
which must be simultaneously verified using the equations of state derived from the theory
of van der Waals as an example. The general principles, however, may be easily applied to
other equations of state.
The calculation procedure is iterative. At each step, we use values for temperature T,
pressure P, and compositions for phases in equilibrium x i , and yi.These values may either
be given in the problem definition, they may be the result of the previous iteration or of
the initialization procedure.
From the temperature we can calculate the parameters for the components of mixture
ai,i,and bi, as well as the binary terms ai,jby application of Equation 8.29:

aY= d G ( 1 - ki,J

where kj,i= k,

(8.29)

Y
a0

-7

m
Q)

N
II

I-

Y'

21

ml
'I

I
I

Y;

I-

(D

Y
II

c-

8. Application of Equations of State to Mixtures

*C,_.C.-.-.-.

f
0

0:

'0

31 7

318

8. Application of Equations of State to Mixtures

The mixing laws are then applied to the liquid phase of composition xi,for example:
aL =CC auxixj
i

(8.27)

(8.28)

b = zbixi
i

if we are dealing with the classical mixing laws. We then solve the equation of state:

RT
w -b

aL
(w - b Lrl) ( w L - b Lr2)

p=--

(4.50)

If this equation has three roots, we consider only the smallest one. If we find only one
root, we must check that it corresponds to a dense phase.
Finally,we calculate the fugacities:

In qf = -In

P ( v L - b L ) bi
+ -(ZL-l)
bL
RT

aL
+ bLRT(rl
- r2)

w L - bLrl
wL-bLr2

(8.33)

The preceding steps are repeated for the calculation of fugacity coefficients in the vapor
phase. We apply the mixing laws with compositions yi:

(8.27)

av =ZCai,jYiYj
i

"

b = zbiyi

and solve for the equation of state:


RT
p=-vV-bV

(8.28)

a"
(wV-bVr,) (wV-bVr2)

(4.50)

If we find three roots, we choose the largest one. If there is only one root, we check that
it corresponds to a vapor phase.
The expression for the fugacity coefficients is then:

h q,!'

= -In

P ( v V - b V ) bi
+ -( Z V - 1)
b"
RT

a"
bVRT(rl- r2)

('y:" );:

"

iIn w
vv-bvrl
- bvr2
(8.33)

The equilibrium condition is ultimately written as:


q;xi = qyyi

(8.86)

It seems that the sequence of operations is simple and that it is fundamentally based on
initialization of missing quantities, and on the process that perpetuates iteration until
condition 8.86 is verified.

319

8. Application of Equations of State to Mixtures

In fact, the calculation of liquid-vapor equilibria under pressure presents certain difficulties due to:
0 the proximity of the critical zone;
0 the presence of the retrograde condensation region;
0 the rather small differences between the properties of the liquid phase and the properties of the vapor phase;
0 the possible continuity between liquid-liquidequilibria and liquid-vaporequilibria.
Therefore,it happens occasionally that a calculation yields the trivialsolution,according to which the two phases in equilibrium are identical, and the system is mistakenly
located outside of the two-phase coexistence zone. This problem is easy to understand
when the resolution for the equation of state for either one or both of the phases has only
one root.
The usual methods, the so-called substitution methods, currently applied in the calculation of liquid-vapor equilibria at low pressure,may be used in many cases. Each single step
is rapid, but we observe that their convergence may be very slow.
The calculation algorithms have been the subject of several studies. A review of these
studies was done by Heidemann [1983] and, more recently, by Michelsen [1992] and
PCneloux [1992].We shall not go into the details of these studies but shall limit ourselves to
stating the principle of the most commonly used methods.

8.5.1 Newton Method


If a system is known by the quantity of each of its components, the whole of its properties
may be determined if we know the temperature, pressure, the quantity (number of moles)
of each of the phases present, and the composition (mole fractions) of these phases. For a
two-phase state, there are therefore 2N + 4 quantities. For each component,we have equations for material balance and equilibrium conditions. The sum of the mole fractions in
each phase is equal to one. We hence have 2N + 2 equations. The problem is therefore
defined as soon as we have two definite quantities selected from among those we have discussed, or from among the thermodynamic properties (total volume, enthalpy, etc.). For
example,if these given quantities are temperature T *and the total volume V*:
T = T*
and:

+VN

~ v* ~

where vL and vv are the values for molar volume in liquid and vapor phase, calculated
from the equation of state.
These two constraint equations complement the problem, made up of a system of
2N + 4 equations (non-linear) and 2N + 4 unknowns. Asselineau et al. [1979] proposed
applying the Newton method to solve it. The Jacobian matrix must therefore be known.
This matrix is made up of the partial derivatives of each of the functions defined by the
equations that we have just listed with respect to the variables,and in particular, the fugacities with respect to temperature, pressure, and composition.From one type of problem to
the next, and according to the problem data, only the final two lines of the system change.
It is obvious that the system thus formed may be simplified in many cases.

320

8. Application of Equations of State to Mixtures

It is well known that the Newton method is characterized by fast convergence speed,
but also by a particular weakness relative to initialization. Most often, for this initialization
we apply the equilibrium coefficients expression proposed by Wilson:
In Ki= In

P .
+ 5.373 (1 + mi)
P

(8.87)

In general, the calculation converges in an acceptable manner, as shown in Figure 8.13,


which, for a mixture resembling a natural gas, provides the number of iterations as a function of the problem.

100
Temperature ( O C )

200

Figure 8.13 Calculation of liquid-vapor equilibria for a natural gas.


: vaporized iso-fraction lines;
-: isenthalpic lines;
-.-.-.-.: isentropic lines;
+ : calculations at given temperature and pressure;
A : calculations at given temperature and vaporized fraction;
:calculations at given pressure and vaporized fraction.

-----

8. Application of Equations of State to Mixtures

321

8.5.2 Tangent Plane Method


It is important to mention that the equality of fugacities corresponds to an extremum of
the function (Gibbs energy) that we wish to minimize, and it may occur that the calculation
yields a false solution. More recent and effective algorithms [Michelsen, 1982a, b] have
been proposed that are based on the so-called tangent plane principle, which we shall
illustrate using an example.

EXAMPLE 8.5

Calculation of the liquid-vapor equilibrium of the ethane, propane system


At 40C, the binary ethane propane mixture was calculated using the Soave-RedlichKwong method. Table 8.3 and Appendix 4 provide the values for the roots of the
equation of state and fugacitiesfor each one of the components at a pressure of 25 bar
and as a function of composition.We see (Example 8.3) that depending on the value
of the composition, the equation of state has one or two roots, listed in this table in
three separate columns referring to the liquid,vapor,or undetermined states.
However, it is only a solution to the equation of state, not a solution to the calculation
of liquid-vapor equilibrium.This calculation requires the use of fugacities, and we can
verify that for a liquid phase compositionx, = 0.34 and vapor phase y, = 0.53 (approximately), the equality condition of the fugacities is verified. These values correspond
to the simplified solution that was obtained (Example 8.4) making use of the fact that
the mixture was practically ideal.
We shall calculate (Table 8.6) the variation of Gibbs energy accompanyingthe mixing of
pure ethane and pure propane. It is expressed as a function of the chemical potentials:

gM = 1 [(
Q= ZlY, + Z z k ) - ( Z l K + Z Z M l
RT

or of the fugacities:

RT

Q = z1 In

fl

+ z2 In fz

fl

f2

(8.88)
(8.89)

where fi is the fugacity of component i in the mixture, and fi* the fugacity of this same
component in the pure state, under the same conditions of temperature and pressure
provided by Table 8.3, for the extreme values of the composition:
f ; = 21.07 bar

fl= 11.87 bar

For example, for a mole fraction of ethane of 0.35, we find:


11.927
7.8
Q L = 0.35 In -+ 0.65 In -= -0.472
21.075
11.87
and:

7.67
21.075

10.92
11.87

Qv= 0.35 In -+ 0.65 In -= -0.408

where the exponents L and V correspond to the liquidand vaporroots respectively.

322

8. Application of Equations of State to Mixtures

Table 8.6

Adimensional Gibbs energy of mixing (Q(x) = P / R T )


for the ethane,propane system at 40C and 25 bar,
calculated using the Soave-Redlich-Kwongequation
Q(X)

Z1

0
0.04
0.08
0.12
0.16
0.20
0.24
0.28
0.32
0.36
0.40
0.44
0.48
0.52
0.56
0.60
0.64
0.68
0.72
0.76
0.80
0.84
0.88
0.92
0.96
1

-0.147
-0.237
-0.305
-0.357
-0.398
-0.429
-0.451
-0.466
-0.473
-0.474
-0.469
-0.458
-0.441
-0.419
-0.391

$M/RT

-0.210
-0.280
-0.333
-0.378
-0.417
-0.449
-0.475
-0.495
-0.509
-0.517
-0.518
-0.513
-0.501
-0.482
-0.456
-0.420
-0.375
-0.318
-0.246
-0.152
0

If we now draw (Fig. 8.14) the Q(z,) diagram made up of two distinct curves, we may
say that up to approximately z , = 0.42, the liquid phase is more stable since its Gibbs
energy is lower. Beyond that point, the vapor phase is more stable. Obviously, that
does not solve the problem of the liquid-vapor equilibrium calculation,since we know
that at equilibrium the two phases have different compositions. However, we observe
that it is possible to draw a tangent common to the two curves, where the points of
contact are located at around 34 to 53%.Any point located on the segment made by
this tangent and limited by these compositions represents the Gibbs energy of a heterogeneous system made up of a phase containing 34% ethane, and a phase containing 53% ethane, in proportions that can be determined by the rule of levers. This system corresponds well to a minimization of Gibbs energy. Ultimately we see that:
0 if z1 < 0.34, the system is homogeneous: the stable state corresponds to the liquid
root of the equation of state;
0 between z1 = 0.34 and z1 = 0.53 the stable state is two-phase; the two phases have
compositions of x1 = 0.34 and y, = 0.53.Their proportions depend on the global composition of mixture z,;

323

8. Application of Equations of State to Mixtures

-0.25

k
2

-0.5
I

I
I

Y
-0.75

0.4
0.6
0.8
Mole fraction of ethane

0.2

Figure 8.14 Gibbs energy of mixing calculated by the Soave-RedlichKwong method.


Calculation of the liquid-vapor equilibria using the tangent plane method.

if z1> 0.53, the system is homogeneous: the stable state corresponds to the vapor
root of the equation of state.
This method may be generalized to a mixture containing more than two components.
It is the best way to resolve liquid-vaporequilibria.

Stability of a mixture: the tangent plane criterion


The preceding calculation will help us to understand the method proposed by Michelsen
[1982a] to verify the stability of a mixture. At given temperature and pressure, the Gibbs
energy defines a surface g ( z ) as a function of compositions. Given a mixture of composition zi,there exists a tangent plane Po to this surface.There may exist, for other compositions, other tangent planes that are parallel to Po. If one of them is located at a height
lower than Po with respect to the scale of Gibbs Energies, then the mixture of composition
ziis unstable. If there exists a merger with Po, then the mixture that corresponds to it is in
equilibrium with the mixture of composition zi.
It may be desirable to precede any equilibrium calculation with a stability test. This
avoids useless calculation if the mixture is stable, and above all, provides an excellent initialization for the equilibrium calculation if the mixture is found to be unstable.

324

8. Application of Equations of State to Mixtures

CONCLUSION
Current works dedicated to the equations of state and mixing laws attest both to their
practical importance and the subjects theoretical value, as well as to the shortcomings of
the methods most often employed.
The calculations used to evaluate the potential of an oil or natural gas deposit and to
improve exploitation, apply the equations of state derived from the van der Waals theory.
The same is true for refining processes, and it has even been proposed that the PengRobinson equation of state be used to determine methanol synthesis equilibria (see
Chapter 13).
The representation of all thermodynamic properties, phase equilibria, densities, heat
capacities, etc., using a single model is a seductive thought, and presents a challenge for the
inventor of such models.
However, as pointed out by Prausnitz, the interaction parameter, or the corrective
term necessary for the mixing laws, is evidence of our lack of understanding. The extensions sanctioned with the introduction of the concept of excess function fall far short of
solving all the problems.
The prediction of liquid-liquid or three-phase equilibria pushes the limits of our models.
The calculation of these equilibria also requires particularly robust algorithms. No matter
how refined these algorithms are today, we must be aware that there is risk of error in such
cases. It is no easy task to combine swiftness with safety. However, these two qualities are
sought after when we study a complex operation such as the exploitation of a reservoir, a
polyphasic chemical reactor, or if we are interested in dynamic simulation of chemical
engineering operations.

REFERENCES
Abdoul W, Rauzy E, Peneloux A (1991) Group contribution equation of state for correlating and predicting thermodynamic properties of weakly polar and non-associating mixtures. Binary and multicomponents. Fluid Phase Equilibria, 68,47-102.
Adachi Y, Sugie H (1986) A new mixing rule - Modified conventional mixing rule. Fluid Phase
Equilibria 28,103-118.
Asselineau L, Bogdanic G, Vidal J (1978) Calculation of thermodynamic properties and vapor-liquid
equilibria of refrigerants. Chem. Eng. Sci., 33,1269-1276.
Asselineau L, Bogdanic G, Vidal J (1979) A versatile algorithm for calculating vapour-liquid equilibria. Fluid Phase Equilibria, 3,273-290.
Bama LR, Rauzy E, Berro Ch, Blanchard JM (1994) An excess function-equation of state model for
solubility of hydrocarbon solids in supercriticalcarbon dioxide. Fluid Phase Equilibria, 1,191-208.
Chao KC, Seader JD (1961) A general correlation of vapor-liquid equilibria in hydrocarbon mixtures.
AZChE J., 7,598-605.
Chaudry MM, van Ness HC, Abbott MM (1980) J. Chem. Eng. Data, 25,254-257.
Dahl S, Michelsen ML (1990a) High pressure vapor liquid equilibrium with a UNIFAC based equation of state. AIChEJ.,36,1829-1836.

8. Application of Equations of State to Mixtures

325

Dahl S, Fredenslund Aa, Rasmussen P (1990b) The MHV2-UNIFAC model: an extended group
contribution model for prediction of gas solubility and vapor-liquid equilibria at low and high
pressures. Sep 9008 report, Institut for kemiteknik, DTH, Lyngby, Denmark.
Dahl S, Fredenslund Aa, Rasmussen P (19%) The MHV2 model: a UNIFAC based model for prediction of gas solubility and vapor-liquid equilibria at low and high pressures. Sep 9023 report,
Institut for kemiteknik, DTH, Lyngby, Denmark.
Dahl S, Dunalewicz A, Fredenslund Aa, Rasmussen P (1992a) The MHV2 model prediction of phase
equilibria at sub- and supercritical conditions.J. Supercritical Fluids, 5,42-47.
Dahl S, Macedo E (1992b) The MHV2 model a UNIFAC based equation of state model for vaporliquid and liquid-liquid equilibria of mixtures with strong electrolytes. Ind. Eng. Chem. Rex, 31,
1195-1201.
Dymond JH, Smith EB (1980) The Virial Coefficients of Pure Gases and Mixtures. Clarendon Press,
Oxford.
Fransson E, Vamling L, Vidal J (1993) The Abdoul-Rauzy-Penelouxgroup-contribution equation of
state extended to CFC-containing mixtures Chem. Eng. Sci., 48,1753-1759.
Fredenslund Aa, Gmehling J, Rasmussen P (1977) Vapor-Liquid Equilibrium Using UNIFAC.
Elsevier, New York.
Glaser M, Peters CJ,van der Kooi HJ, Lichtenthaler RW (1985) Phase equilibria of the system
methane + n-hexadecane.J. Chem. Thermodynamics, 17,803-815.
Gmehling J, Onken U, Arlt W, Grenzheuser P, Weidlich U, Kolbe B, Rarey-Nies J (1980-1984) Vaporliquid equilibrium data collection. Dechema Data Series, Frankfurt.
Goral M, Maczynski A, Schmidt G, Wenzel G (1981) Vapor-liquid equilibrium calculations in binary
systems of hydrocarbons + related compounds not containing oxygen. Comparison between
methods using equations of state and activity coefficients.Ind. Eng. Chem. Fund., 20,267-277.
Grabowski MS, Daubert (1978a) A modified Soave equation of state for phase equilibrium calculations. 1.Hydrocarbons systems. Ind. Eng. Chem. Process Des. Dev., 17,443-448.
Grabowski MS, Daubert (1978b) A modified Soave equation of state for phase equilibrium calculations. 2. Systems containing CO,, H,S, N,, and CO. Ind. Eng. Chem. Process Des. Dev., 17,448-453.
Grabowski MS, Daubert (1978~)A modified Soave equation of state for phase equilibrium calculations 3. Systems containing Hydrogen. Ind. Eng. Chem. Process Des. Dev., 18,300-306.
Griswold J, Buford CB (1949) Vapour-liquid equilibria of acetone-methanol-water.Znd. Eng. Chem.,
41,2347-33.
Griswold J, Wong SY (1952) Phase equilibria of the acetone-methanol-watersystem from 100C into
the critical region. Chem. Eng. Prog., Symp. Ser.,48, No. 3,18-34.
Hankinson RW, Thomson GH (1979) A new correlation for saturated densities of liquids and their
mixtures. AIChE J.,25,653-662.
Heidemann RA (1983) Computation of high pressure phase equilibria.Fluid Phase Equilibria,14,55-78.
Heidemann RA, Kokal SL (1990) Combined excess free energy models and equations of state. Fluid
Phase Equilibria, 56,17-37.
Hendricks EM (1988) Reduction theorem for phase equilibrium problems. Ind. Eng. Chem. Res., 27,
1728-1732.
Hernandez Garduza (1993) Etude des modkles thermodynamiques susceptibles de representer les
melanges contenant essentiellement des hydrocarbures, de l'eau et des alcools. Sc.D. Thesis,
Universite de Provence Aix-Marseille I.
Huang H, Sandler S (1993) Prediction of vapor-liquid equilibria at high pressures using activity coefficient parameters obtained from low pressure data: a comparaison of two equation of state
mixing rules. Ind. Eng. Chem. Res., 32,1498-1503.

326

8. Application of Equations of State to Mixtures

Huron MJ, Dufour GN, Vidal J (1977) Vapor-liquid equilibrium and critical locus curve calculations
with the Soave-Redlich-Kwong equation for hydrocarbon systems with carbon dioxide and hydrogen sulphide.Fluid Phase Equilibria, 1,247-265.
Huron MJ,Vidal J (1979) New mixing rules in simple equations of state for representing vapor-liquid
equilibria of strongly non-ideal mixtures. Fluid Phase Equilibria, 3,255-271.
Kay W B (1964) Density of hydrocarbon gases and vapors at high temperature and pressures. Znd.
Eng. Chem., 28,1014-1019.
Knapp H, Doring R, Oellrich R, Plocker U, Prausnitz JM, Langhorst R, Zeck S (1982) Vapor-liquid
equilibria for mixtures of low boiling substances. Dechema Data Series, Dechema, Frankfurt,
Germany.
Knudsen K (1992) Phase equilibria and transport of multiphase systems. Sc.D. Thesis, Institut for
Kemiteknik. Danmarks Tekniske Hojskole.
Kojima K, Tochigi K (1979) Prediction of Vapor-Liquid Equilibria by ASOG Method. Kodonsha
Elsevier,Tokyo.
Larsen BL, Rasmussen P, Fredenslund Aa (1987) A modified UNIFAC group contribution model for
prediction of phase equilibria and heats of mixing. Znd. Eng. Chem. Rex, 26,2274.
Lee BI, Kesler MG (1975) A generalized thermodynamic correlation based on three-parameter corresponding state. AZChE J., 21,510-527.
Lermite Ch, Vidal J (1988) High pressure polar compounds phase equilibria calculation: mixing rules
and excess properties. Fluid Phase Equilibria, 42,l-19.
Lermite Ch, Vidal J (1992) A group contribution equation of state for polar and non-polar compounds. Fluid Phase Equilibria, 72,111-130.
Ma YH, Kohn JP (1964) Multiphase and volumetric equilibria of the ethane methanol system at temperatures between - 40 and 100C.J. Chem. Eng. Data, 9,3-8.
Mathias PM, Klotz HC, Prausnitz JM (1992) Equations-of-state mixing rules for multicomponent
mixtures. The problem of invariance.Fluid Phase Equilibria, Fluid Phase Equilibria, 67,31-44.
Michelsen ML (1982a) The isothermal flash problem. Part I. Stability. Fluid Phase Equilibria, 9,l-19.
Michelsen ML (1982b) The isothermal flash problem. Part 11. Phase split calculation. Fluid Phase
Equilibria, 9,21-40.
Michelsen M, Kistenmacher H (1990a) On composition dependent interaction coefficients. Fluid
Phase Equilibria, 60,47-58.
Michelsen ML (1990b) A modified Huron-Vidal mixing rule for cubic equations of state. Fluid Phase
Equilibria, 60,213-219.
Michelsen ML (1992) Phase equilibrium calculations.What is easy and what is difficult? Proceedings
of the ESCAPE1 Conference,paper S19, Comput. Chem. Eng.
Mollerup J (1986) A note on the derivation of mixing rules from excess Gibbs energy models. Fluid
Phase Equilibria, 25,323-327.
Moysan JM, Huron MJ, Paradowski H,Vidal J (1983) Prediction of the solubility of hydrogen in hydrocarbon solvents through cubic equations of state. Chem. Eng. Sci., 38,1085-1092.
Moysan JM, Paradowski H, Vidal J (1986) Prediction of phase behaviour of gas containing systems
with cubic equations of state. Chem. Eng. Sci., 41,2069-2074.
Orbey H, Sandler SI, Wong DSH (1993) Accurate equation of state predictions at high temperatures
and pressures using the existing UNIFAC model. Fluid Phase Equilibria, 85,41-54.
Panagiotopoulos A, Reid M (1986) New mixing rules for cubic equations of state for highly polar
asymetric systems. In: Equations of state - Theories and applications, ed. by Chao K and
Robinson R, Am. Chem. SOC.Symp. Ser., 300,571-582.

8. Application of Equations of State to Mixtures

32 7

Pedersen KS, Fredenslund Aa, Christensen PL, Thomassen P (1984) Viscosity of crude oils. Chem.
Eng. Sci., 39,1011-1016.
Pedersen KS, Fredenslund Aa, Thomassen P (1989) Properties of Oils and Natural Gases. Gulf
Publishing Company, Houston, Texas.
PBneloux A, Rauzy E, Freze R (1982) A consistent correction fot Redlich-Kwong-Soavevolumes.
Fluid Phase Equilibria, 8,7-23.
PBneloux A, Abdoul W, Rauzy E (1989) Excess functions and equations of state. Fluid Phase
Equilibria, 47,115-132.
PBneloux A (1992) Utilisation des Bquations d'6tat pour le calcul des propriBtBs des fluides mono- et
polyphasiques. Entropie, 172/173,73-82.
Peng D-Y, Robinson DB (1976) A new two-constant equation of state. Znd. Eng. Chem. Fundam., 15,
59-64.
Plocker U, Knapp H, Prausnitz JM (1978) Calculation of high-pressure vapor-liquid equilibria from a
corresponding- states correlation with emphasis on assymetric mixtures. Znd. Eng. Chem. Process
Des. Dev., 17,324-332.
SchwartzenrubberJ, Renon H (1989) Development of a new cubic equation of state for phase equilibrium calculations.Fluid Phase Equilibria, 67,99-110.
Soave G (1972) Equilibrium constants from a modified Redlich-Kwong equation of state. Chem.
Engin. Sci., 27,1197-1203.
Soave G (1992) A new expression of q(a) for the modified Huron-Vidal method. Fluid Phase
Equilibria, 72,325-327.
Spencer CF, Danner RP (1972) Improved equation for prediction of saturated liquid density.1 Chem.
Eng. Data, 17,236-241.
Stryjek R, Vera JH (1986) An improved Peng-Robinson equation of state for pure compounds and
mixtures. The Canadian J. Chem. Engin., 64,323-333.
Teja A (1980) A corresponding states equation for saturated liquid densities.AZChE J., 26,337-345.
Vidal J (1978) Mixing rules and excess properties in cubic equations of state. Chem. Engin. Sci., 33,
787-791.
Wong DSH, Sandler SI (1992a) A theoretically correct new mixing rule for cubic equations of state.
AZChE J., 38,671-680.
Wong DSH, Orbey H, Sandler SI (1992b) An equation of state mixing rule using available activity
coefficient model parameters and which allows extrapolation over large ranges of temperature
and pressure. Ind. Eng. Chem. Res., 31,2033-2039.
Ye S (1990) Mesure et exploitation de la vitesse d'ultrasons dans les liquides en pression; application
I? des fluides complexes d'origine pBtrol2re. Sc.D. Thesis, UniversitC de Pau.
Zeck S, Knapp H (1986) Vapor-liquid and vapor-liquid-liquid phase equilibria of binary and ternary
systems of nitrogen, ethene and methanol experiment and evaluation. Fluid Phase Equilibria, 26,
37-58.

Liquid-Liquid and
Liquid-Liquid-Vapor Equilibria

Under ordinary conditions of temperature and pressure, water and hydrocarbons are
practically immiscible. The compounds that are strongly auto-associated by hydrogen
bonding such as ethylene glycol or methanol, and the polar compounds acetonitrile and
dimethylformamide,for example, are only partially miscible with the saturated hydrocarbons. More generally, partial miscibility in liquid phase is a commonly observed phenomenon. This fact may be explained by the existence of local compositions that exclude from
the intermolecular attraction center any compounds with a structure that is foreign to this
center.
Like the liquid-vapor equilibria, the liquid-liquid equilibria are selective, and used in
separation processes. Their prediction is tricky, however, because it is based on familiarity
with, and an exact representation of deviations from ideality. The vapor pressure of the
components is not at work in these phase separations.
First we shall give some examples of such equilibria using binary and ternary mixtures
in order to describe the variety of circumstances they create.
Liquid-vapor and liquid-liquid equilibria may occur simultaneously,according to conditions of temperature and pressure. This is what happens during distillation in the presence
of water vapor (stripping). Heteroazeotropic distillation is also based on these threephase equilibria. The result is an infinite variety of circumstances, only the most simple of
which will be discussed here.
Of course, there exist liquid-liquid equilibria critical points, and their representation
within the temperature/pressure range leads to diagrams that are comparable to those
from Chapter 6 (Section 6.1.2, Fig. 6.6),but that are in fact much more complex because
the liquid-liquid and liquid-vapor critical points loci are combined, which sometimes
reveals the continuity that exists between these two types of phase equilibria.We shall discuss them only briefly.
The calculation methods for liquid-liquid equilibria are based, as we have stated, on
familiarity with deviations from ideality. We do not have the space for a lengthy analysis,
merely to point out their increased complexity.

330

9.1

9. Liquid-Liquid and Liquid-Liquid-VaporEquilibria

LIQUID-LIQUID EQUILIBRIA AND DEVIATIONS FROM IDEALITY

Lack of miscibility is related to deviations from ideality within the mixture, as we have
mentioned using examples of activity coefficients at infinite dilution (Chapter 7,
Section 7.1.3,Table 7.2). Elevated values for this term may result in partial solubility.This
relationship may be explained by examination of the diagram (Fig. 9.1) illustrating the
change of Gibbs energy of mixing in the liquid phase with composition, at constant temperature and pressure, when calculated by a non-ideality model, NRTL or UNIQUAC, for
example, (Chapter 8, Section 8.6.2).The curve is drawn for all compositions,not taking into
account possible demixing.
We have seen that the diffusional stability condition (see Chapter 6, Section 6.3.4) is
expressed for a binary mixture by the equation:
(6.18)
This condition is not respected by the mixtures whose composition is found between the
abscissa of points C and D. Such mixtures are unstable, and cannot exist in a homogeneous
liquid phase. If at constant pressure we represented the change of compositions xc and xD

500 1

....
Figure 9.1 Variation of Gibbs energy of the hexane methanol system
at 25C and atmospheric pressure. Calculation by the NRTL model.
Graphical representation of demixing boundaries.

9. Liquid-Liquid and Liquid-Liquid-Vapor Equilibria

33 1

as a function of temperature, we would obtain the spinodal curve, whose calculation is


(relatively) easy from the expression for excess Gibbs energy.
The necessary diffusional stability condition is, however, insufficient.Indeed we observe
that there is a line tangent to curve $(x) at two points, and that on this line, any point P of
the segment between the two points of contact can represent a heterogeneous system
made up of two phases of composition xc and xD in appropriate proportions. The Gibbs
energy of this heterogeneous system is less than that of the assumed homogeneous system
of the same composition, whose representative point M is located on the ?(x) curve. It is
therefore more stable.
The solubility limits therefore correspond to the abscissa of points A and B. Determination of the corresponding activities (or the partial molar Gibbs energies of mixing) using
the graphical tangent method (see Chapter 5, Section 5.1.2) on the $(x) curve, confirms
the fact that they are identical in the two phases since they are equal to the ordinates of the
intersectionsof the tangent with the axes x1 = 0 and x1 = 1.
A similar analysis has been done in the study of liquid-vapor equilibria under pressure
(Chapter 8, Section 8.5.2, Fig. 8.14).
The mixtures whose composition lies between xA and xc, and xD and xE are called
metastable. At constant pressure, the change of compositions xA and xE with temperature
is given by the binodalcurve,the locus of equilibrium points.The spinodal curve is inside
the binodal curve.
We have already pointed out that deviations from ideality in dense liquid phase (that is
to say at temperatures clearly lower than the critical temperature of the components) are
not very sensitive to pressure (Chapter 5, Section 5.7.4).The same is therefore true for liquid-liquid equilibria, and the examples that follow involve equilibria measured at atmospheric pressure. However, this is an approximation, and large variations in pressure may
extend or narrow the demixing range. The spinodal and binodal curves correspond to surfaces within the pressure, temperature, and composition space.
In any event, it is important to remember that a decrease in pressure may cause the
appearance of a vapor phase, and therefore a liquid-liquid-vapor equilibrium.

9.2

GENERAL DESCRIPTION OF LIQUID-LIQUID EQUILIBRIA

9.2.1 Binary Systems


At constant pressure, the liquid-liquid equilibria are generally sensitive to temperature, as
shown in Figure 9.2 for the methanol heptane system [Tagliaviniand Arich, 19581.This figure shows the binodal curve, in other words, the solubility limits. A mixture whose representative point M is within the binodal curve separates into two liquid phases L and L
(points A and B) according to proportions that may be measured by the rule of levers:

332

9. Liquid-Liquid and Liquid-Liquid-Vapor Equilibria

ase L

phase

0.25

0.5

0.75

Mole fraction of methanol

Figure 9.2 Equilibrium diagram for the methanoln-heptane system.


[Tagliavini and Arich, 19581.

This proportion will change with temperature and we see that in this example,the mutual
solubility of heptane and methanol increases with temperature.This is related to a decrease
in deviations from ideality (positive enthalpy of mixing, endothermic mixture), and constitutes by far the most common case.An increase in temperature is going to cause phase L to
disappear. Beyond a given temperature the system with compositionxM is homogeneous.
The two arcs of the binodal curve meet at C. At this point, the two phases in equilibrium
are identical: we have a critical point. It is called the upper critical solution temperature
(UCST). Table 9.1 provides the upper critical solution temperature for some hydrocarbon,
solvent pairs.
It may occur, depending on the value of pressure, that the liquid-vapor equilibrium (or
thermal decomposition) interferes with the liquid-liquid equilibrium. The diagram is then
truncated. This case will be studied later on. Frequently the diagram is also truncated due
to the absence of experimental data.
More rarely, it may happen that an increase in temperature extends immiscibility, and
there exists a lower critical solution temperature (LCST) on the diagram for the binodal
curve. Such is the case for the dipropylamine water mixture [Horbson R.W., 1941,Fig. 9.31.
Such a phenomenon is practically the rule for polymer solutions.
Finally, a liquid-liquid equilibrium diagram may be closed and have, as in the case of
the tetrahydrofurane water system, two critical points, as shown in Figure 9.4.

333

9. Liquid-L iquid and Liquid-Liquid- Vapor Equilibria

Table 9.1
Critical solution temperatures (C) of
polar hydrocarbonlsolvent systems [Francis, 19441

n-Hexane

Methanol
Ethanol
Acetone
Acetonitrile
Ammonia
Aniline
Ethylene glycol

n-Heptane 1-Heptene

Benzene

12

26
180

50

25

0.25

0.5
0.75
Mole fraction of dipropylamine

Figure 9.3 Equilibrium diagram for the dipropylamine water system.


[Horbson, 19411.

An extensive review of the literature [Sorensen ef al., 19791 gives us the relative frequency of each type of diagram:
0 Diagram with an upper critical solution temperature: 41%
0 Truncated diagram: 53%
Diagram with a lower critical solution temperature: 2% (not including polymers)
Closed diagram: 4%.

334

9. Liquid-Liquid and Liquid-Liquid-Vapor Equilibria

100

1275
5:

0
I

0.25

0.5

0.75

Mole fraction of tetrahydrofurane

Figure 9.4 Equilibrium diagram for the tetrahydrofurane water system.


[Mathous et al., 19721.

9.2.2 Ternary Systems


We often illustrate the compositions of ternary systems using an equilateral triangular diagram each peak of which represents one of the components, and each side one of the
binary systems. The composition corresponding to a point within the triangle is obtained
for each component by drawing a line parallel to the side opposite the peak representing
it, and taking the intersection of this parallel with either one of the other two sides.
In this way we can illustrate the liquid-liquid equilibria at given temperature and pressure, and the appearance of the diagram will depend primarily on the reciprocal solubilities of the components, taken two at a time.
So for the n-heptane, benzene, methanol system at 280 K, we know that the binary
n-heptane methanol system has a miscibility gap. If we add benzene to a heterogeneous
n-heptane methanol mixture, it acts as a third solvent and, in sufficient quantity, favors
the formation of a homogeneous mixture. Figure 9.5 shows the curve for the limits of the
heterogeneous range. It is referred to as a closed diagram. On this curve, the points in
liquid-liquid equilibrium correspond in pairs shown by the equilibrium lines. The curve
also shows the ternary critical point, where the liquid-liquid equilibrium line reduces to a
single point.
If we increase temperature, most often the immiscibility range becomes smaller. It disappears when we reach the upper critical solution temperature of the immiscible binary
system.

335

9. L iquid-Liquid and Liquid-L iquid-Vapor Equilibria

n-Heptane

Methanol

Figure 9.5 Liquid-liquid equilibrium diagram for the n-heptane,benzene,


methanol system at 280 and 306 K [Wittrig, 19771.
:two-phase equilibrium boundary;- - - - - : equilibrium lines.

If the ternary system contains two partially miscible binary systems, we obtain an
open diagram such as the one in Figure 9.6 for the n-hexane, n-heptane, methanol system. If we increase the temperature, the extent of the immiscibility range diminishes. If we
surpass the lowest of the critical solution temperatures, (in the n-hexane methanol system), the diagram then becomes closed. At this temperature, the ternary critical point
intersects with the critical point of the binary system, which becomes miscible in all proportions.
The two types of diagrams that we have just described are the most common. There are
many other, relatively minor possibilities. They have been listed by Sorensen et al. [1979].
It may happen that a three-phase equilibrium exists, such as the one shown in Figure 9.7
for the methyl 2-butene 2, acetonitrile, water system at 20C.The system therefore has two
degrees of freedom, and at given pressure and temperature, the compositions of the three
phases in equilibrium are independent of the global composition of the heterogeneous
mixture. Outside of the triangle made up of points A, B, and C, for these three phases in
equilibrium, there are three distinct two-phase equilibrium zones, one of which extends all
the way to the side representing the methyl 2-butene 2, water binary system, which is
practically immiscible in any proportions. The two other binary systems are completely
miscible.

336

n-Heptane

Methanol

Figure 9.6 Liquid-liquid equilibrium diagram of the n-hexane,n-heptane,


methanol system at 306 K [Wittrig, 19771.
: two-phase equilibrium boundaries;- - - - -: equilibrium lines.
*Acetonitrile

T = 293.15 K

Methyl 2-Butene 2

Water

Figure 9.7 Liquid-liquid equilibrium diagram for the methyl 2-butene 2,


acetonitrile,water system at 293.15 K [Mikitenko, 19671.
:two-phase equilibrium boundaries;- - - - - :equilibrium lines;
-.-.-.- :three-phase equilibrium lines.

337

9. Liquid-Liquid and Liquid-Liquid-Vapor Equilibria

Finally we can cite the case of water, dimethylsulfoxide, tetrahydrofurane mixtures


shown in Figure 9.8. Although at 20C the three binary systems are completely miscible,
there exists an islet of immiscibility inside the ternary diagram, and the boundary of the
two-phase region has two critical points. We recall that the diagram for the water, tetrahydrofurane (Fig. 9.4) system had a lower critical solution temperature of around 75C.
Above this temperature, the ternary diagram becomes closed,similar to the one shown
in Figure 9.5.

Water

THF

Figure 9.8 Liquid-liquid equilibrium diagram for the water,


dimethylsulfoxide,tetrahydrofurane system at 293.15 K.
----- :equilibriumlines; A: critical points [Wolski, 19701.

9.3

SELECTIVITY OF THE LIQUID-LIQUID EQUILIBRIUM

The selectivity of liquid-vapor equilibria has its origin both in the vapor pressure differences of the components as well as in the deviations from ideality of the mixture that may
increase or attenuate volatility differences.The liquid-liquid equilibria depend only on the
deviationsfrom ideality.These deviations are, however, selective.We have seen (Chapter 7,
Table 7.2) that the activity coefficients at infinite dilution for hydrocarbons in polar solvents, which are generally high, depend very much on the chemical nature of the hydrocar-

338

9. Liquid-L iquid and Liquid-Liquid-Vapor Equilibria

bon. We have also just seen that the paraffins are only partially soluble in methanol, yet
there is complete miscibility of the benzene methanol pair.
This selectivity is used to our advantage in separation processes using liquid-liquid
extraction.We will demonstrate this point using an example.
Table 9.2 and Figure 9.9 show the liquid-liquid equilibrium data for the n-hexane, benzene, dimethylsulfoxidesystem at 293.15 K.
Table 9.2
Liquid-liquid equilibria for n-hexane (l), benzene (2), dimethylsulfoxide (3)
at 293.15 K. Compositions in mole fractions [Waksmundski et aL,19641
41)

x (2)

0.915
0.842
0.787
0.721
0.664
0.597
0.536
0.456
0.377
0.309

0.083
0.151
0.202
0.267
0.320
0.382
0.435
0.490
0.519
0.526

0.002
0.007
0.011
0.012
0.016
0.021
0.029
0.054
0.104
0.165

0.035
0.036
0.044
0.048
0.054
0.064

0.081
0.096
0.125
0.150

0.040
0.080
0.132
0.184
0.226
0.276
0.333
0.385
0.433
0.460

0.925
0.883
0.824
0.767
0.721
0.660
0.586
0.519
0.442
0.390

Benzene

A
n-Hexane

DMSO

Figure 9.9 Liquid-liquid equilibrium diagram for the n-hexane, benzene


dimethylsulfoxide system at 293.15 K [Waksmundski et al., 19641.

339

9. Liquid-Liquid and Liquid-Liquid-VaporEquilibria

These data show that if we add 100 moles of dimethylsulfoxideto a mixture containing
68 moles of hexane and 32 moles of benzene, we obtain a heterogeneous mixture and the
compositions of the two liquid phases (commonly referred to as extract and raffinate) correspond to the compositions found in the third line of Table 9.2. We note that the proportions of hexane and benzene are clearly different in these two phases. Table 9.3 summarizes the properties (quantities and compositions) of the two phases.

Input
n-Hexane
Benzene
Dimethylsulfoxide
Total quantity

68
32
100
200

Raffinate
62.7
16.2
0.9
79.8

0.787
0.202
0.011

Extract
5.3
15.8
99.1
120.2

0.044

0.132
0.824

Separation selectivityis generally defined by the ratio:


0.132
x2
0.202
=
= ___ = 11.6
0.044
Xi
0.787
-

&,I

x;

As we shall see, it is related to the activity coefficients ratios.

9.4

LlQUl D-LIQU I D-VAPOR EQUILI BRlA

The appearance of the liquid-liquid-vapor equilibrium diagram depends on both the


extent of the immiscibility range (and therefore the deviations from ideality), and on the
relative volatilities of the components (ratio of their vapor pressures). The binary isobaric
diagrams 9.10a and 9.10b illustrate the simplest cases.
On Figure 9.10a, we first note that the coexistence curve for the two liquid phases L and
Lis truncated, similar to the one shown in Figure 9.2. The dotted portion can be described
at higher pressure only.
Startingwith a mixture in the vapor state (point M ) , we lower the temperature.When we
reach the dew temperature (point P ) , a liquid phase L appears (point Q),which has a different composition than the vapor. If we continue to decrease temperature, the points representing the liquid and vapor phases describe the PH and QA curves respectively.When
the composition of the liquid phase L and the temperature are such that the solubility limit
is reached (point A ) , a second liquid phase Lappears, represented by point B, and the
vapor phase is represented by point H. The system is now monovariant and, while the pres-

340

9. Liquid-Liquid and Liquid-Liquid-Vapor Equilibria

sure remains steady, condensation occurs at constant temperature and at constant compositions of the three phases in equilibrium. Their proportions vary, however, and the point
representative of the global composition of the two liquid phases will evolve from point A
in the direction of B as the formation of phase L advancesThe process continues until this
global composition is identical to that of the initial mixture. Condensation is then complete.

x, x, y

a.

x, x, y

b.

Figure 9.10 Liquid-liquid-vapor isobaric equilibrium diagrams.

If the composition of the initial mixture x, had been lower than that of the ~ a t u r a t e d ~ ~
mixture x,, condensation would have remained in a two-phase state throughout.
Finally, if it intersects with the composition of the mixture represented by point H,
condensation would occur immediately in a three-phase state until all vapor is gone. The
proportions of the two liquid phases would remain constant, and be determined using the
ratio of segments A H and HB.
Conversely, we see that vaporization of a mixture divided into two liquid phases will
always begin with a monovariant step. The global composition of the two liquid phases
evolves toward points A or B according to whether x, is lower, or higher than x H .
Point H , the intersection of the two arcs of the dew curve, is called the heteroazeotrope.
Figure 9.1Ob is characteristic of a mixture with two compounds of very different volatility. Due to this difference in volatility, there is no heteroazeotrope. Depending on the composition of the initial mixture, we may observe several types of condensation:
xM < x,: condensation occurs in a two-phase state.
X, < x, < xB:condensation begins in a two-phase state, and ends in a three-phase state.
The points representative of the two liquid phases and the vapor phase are A , B, and H.
xB < x, < xH:condensation begins in a two-phase state, continues via a three-phase
(and invariant) step, and ends in a two-phase state.
xH < xM: condensation occurs in a two-phase state.

9. Liquid-L iquid and Liquid-Liquid- Vapor Equilibria

341

Figure 9.11 [de Swaan Arons, 19951 summarizes some of the possibilities according to
the values of the relative volatility (on the abscissa) and non-ideality (on the ordinate).
The equilibrium diagrams are isothermal diagrams,in opposition to the ones we have discussed here, and can be commented in the same way. We note the possibility of observing
an azeotrope outside of the demixing zone. We can also consult the work of Rowlinson and
Swinton [1982] on the evolution of liquid-liquid-vapor diagrams and the azeotrope as a
function of pressure and temperature conditions.

Figure 9.11 Change in liquid-vapor and liquid-liquid-vapor equilibrium diagrams


as a function of the relative volatility (PTlPi')and the degree of non-ideality (gE).
(Fluid Phase Equilibria, 104, J. de Swaan Arons. The systematic study of phase
behaviour and the emerging coherence of phenomena, 97-118,O 1995. With kind
permission of Elsevier Science-NL, Sara Burgerhartstraat 25,1055 KV Amsterdam,
The Netherlands).

342

9. Liquid-Liquid and Liquid-Liquid-VaporEquilibria

The shape of the liquid-liquid-vapor equilibrium diagram changes with temperature.


The three-phase equilibrium pressure increases and the heteroazeotrope leaves the
demixing range to yield a diagram that most often has a homoazeotrope within the twophase zone.
We have noted that pressure had little influence on liquid-liquid equilibria. However, at
high pressures (several hundreds of bars) we may observe this influence that principally
depends on excess volume. At a given temperature, the demixing range may be extended
or diminished before finally disappearing at a liquid-liquid critical point. In this case, the
critical points locus will have both a liquid-vapor line, and a liquid-liquidline that terminate at the critical solution temperature determined at low pressure. These two lines
may be connected. With the help of a calculation based on the van der Waals equation of
state, it is possible to predict several types of critical loci [Scott and van Konyenbourg,
1970;van Konyenbourg and Scott, 1980;Rowlinson and Swinton, 1982, Chapter 61.
However, very few systems have been perfectly characterized experimentally,especially
where liquid-liquid and liquid-vapor equilibria approach the critical zone. We can cite
the data relating to the ethane methanol and ethylene methanol mixtures shown in
Figures 9.12 [Ma K.H. and Kohn J.P., 1964;Zeck S. and Knapp H., 19861 and 9.13 [Deiters,
19931.First of all we note the influence of the chemical nature of the hydrocarbon.The ethylene methanol system has a lower solution temperature while no liquid-liquid critical
point has been determined for the ethane methanol system. For the latter system, the
phase equilibria change with temperature according to the steps (qualitative) illustrated in
Figure 9.12. We note the probable presence of two critical points at a temperature that is
slightly greater than the critical temperature of ethane, then the merging of the liquidvapor and liquid-liquid equilibria at a temperature slightly greater than the critical temperature of ethane [Ma and Kohn, 19641.
The methane n-hexane system (or greater homologues), like the ethylene methanol system, has a lower critical solution temperature and an immiscibility zone in the liquid phase
in a temperature range that is close to the critical temperature of methane [Chen et al.,
1976, Lin et al., 19771. An analogous phenomenon is observed for systems made up of
ethane and long chain hydrocarbons, such as n-nonadecane. The explanation is certainly
not found in the differences in polarity or molecular interactions, but in free volume, as we
shall see with polymer solutions.
It is very clear that the combination of various binary equilibrium possibilities, with or
without immiscibility, and with or without azeotrope, leads to a great variety of ternary liquid-liquid-vapor equilibrium diagram types. We shall limit ourselves to a very simple
example of a system, the ternary mixture A , B, C, illustrated by Figure 9.14. The liquid-liquid equilibria are shown in an open diagram analogous to those previously shown
(Fig. 9.5 or 9.9), and the critical point Q is found on the curve delimiting the immiscibility
range.Al1 immiscible ternary mixtures made up of the union of two liquid phases (points L
and L)in equilibrium,and whose representative point M is therefore located on the equilibrium line LL, are themselves in equilibrium with a similar vapor phase (point V).
On Figure 9.14 we have drawn the locus of points that represent the composition of the
vapors with all the heterogeneous mixtures. This locus depends on the non-ideality of the
system and the vapor pressures of the components. In this case it is limited by the heteroazeotrope P corresponding to the binary mixture AB, and R, the point in equilibrium
with the critical mixture in the liquid phase Q.

9. Liquid-Liquid and Liquid-L iquid- Vapor Equilibria

.9 E
o

343

344

9. Liquid-Liquid and Liquid-L iquid- Vapor Equilibria

P
P

Figure 9.W Liquid-liquid-vaporequilibria and loci of critical points:


ethane methanol system.

9. Liquid-Liquid and Liquid-Liquid-VaporEquilibria

345

Figure 9.14 Liquid-liquid-vaporequilibrium diagram for a ternary mixture.


:phase composition lines; - - - - -: liquid-liquid equilibrium lines;
-. -.-.- :liquid-vapor equilibrium lines.

9.5

CALCULATION METHODS

The calculation of liquid-liquid or liquid-liquid-vapor equilibria is based on the general


rule of equalities of chemical potential or fugacities (see Chapter 5, Section 5.2.2):
or

y:=y,&'

fF=fF'

i = l,n

(9-3)

for the equilibrium between liquid phases L and L', and:


y 1F. = y 1F.' = y ,

or

fF =fF'=f/

i =l,n

(9.4)

for the equilibrium between two liquid phases L and L', and a vapor phase V.
If we apply a calculation method for fugacities derived from an expression for activity
coefficients:
= fl?" yi*"x'
(5.73)

fi"

"

we note that the reference fugacity &* is the same for the two liquid phases, and that the
liquid-liquid equilibrium condition is reduced to the equality of the activities:

yLx.
1
1 = yf'xf

or

a: = a:'

(9.5)

346

9. Liquid-Liquid and Liquid-Liquid-VaporEquilibria

We see that the liquid-liquid equilibrium depends only on the deviations from ideality.
And thus, for the partition coefficient it is:

If, as in the example given for the n-hexane (l), benzene (2), dimethylsulfoxide or
DMSO (3) system,we are interested in the selectivity of the solvent,defined by the ratio of
partition coefficients of benzene and hexane, we see that selectivity is also expressed as a
function of activity coefficients:

In the special case of strongly immiscible binary systems, n-hexane and DMSO,
for example, the activities of n-hexane (denoted below by subscript 1) and dimethylsulfoxide are close to one in the two phases, and if we say that L is the phase rich in DMSO,
we can write:

So the solubility (also called solvent power) of hydrocarbons in polar solvents varies
with the inverse of the activity coefficients at infinite dilution. As for the selectivity of a
solvent with respect to two hydrocarbons, at first approximation it may be calculated by
the ratio of the activity coefficients at infinite dilution of these hydrocarbons in the
solvent:
A
a211

I1

y2L

(9.9)

However, we note that this approximation does not include the non-ideality of the
n-hexane benzene mixture.
It is not necessary to engage in a lengthy review of the models applied to the calculation
of liquid-liquid or liquid-liquid-vapor equilibria. These models (most often the calculation
of excess Gibbs energy, and equations of state) have already been described (Chapters 7
and 8). Yet we must emphasize the extreme sensitivity of the calculation of excess Gibbs
energy or activity coefficients values. This sensitivity is illustrated in Table 9.4 ,and
Figure 9.15 that list the values for excess Gibbs energy and Gibbs energy of mixing calculated by two models for the propane methanol system. We note that these two models
yield very similar values for excess Gibbs energy (upper portions of the diagram).Yet after
calculating the Gibbs energy of mixing,by adding Gibbs energy of an ideal mixture, we get
two curves (lower part of the diagram). One of these curves (solid lines) lends itself to the
construction illustrated in Figure 9.1 that, after drawing a tangent to curve e ( x ) at two
points, allows for the determination of the solubility limits. The dotted lines characterize a
mixture that is miscible in all proportions.

347

9. Liquid-L iquid and Liquid-Liquid-Vapor Equilibria

Table 9.4
Excess Gibbs energy and Gibbs energy of mixing for the propane (1)
methanol (2)system calculated by two models. T = 313.15 K,P = 13 bar
First Model

gEIRT

0
0.3850
0.5675
0.5741
0.4003
0

0
0.2
0.4
0.6
0.8
1

Second Model

I
I

gMIRT
0

-0.1154
-0.1055
-0.098 9
-0.1001
-0

gEIRT
0
0.3685
0.5475
0.5738
0.4354
0

gMIRT
0

-0.1319
-0.1255
-0.0992
-0.065 0
0

1 ,

0.8

-0.2

1
0

0.25

0.5

0.75

Mole fraction of propane

Figure 9.15 Calculation of excess Gibbs energy (upper portion of the


diagram) and Gibbs energy of mixing (lower portion) of the propane (1)
methanol (2) system using two models.

If the same model is often applied to the representation of liquid-vapor and liquid-liquid equilibria,it often happens that we use two different sets of interaction parameters to
perform the calculation.An example of this is provided by the UNIFAC model (Chapter 7,
Section 7.7.2) for which a specific set of interaction parameters between groups of liquidliquid equilibria has been suggested [Magnussen et al., 19811.The problem is the same, if
not more complex,when the applied model is derived from an equation of state.

348

9.6

9. Liquid-Liquid and Liquid-Liquid-VaporEquilibria

WATER, HYDROCARBON SYSTEMS

Mixtures containing water and hydrocarbons are of considerable importance. In petroleum reservoirs, hydrocarbons are in contact with an aqueous liquid phase that is rich in
salts. During refining, we often perform distillation in the presence of water vapor (stripping) for the distillation of heavy cuts.
It is well known that the miscibility of water in hydrocarbons (and vice versa) is weak,
and we often lack experimental data, or the accuracy of such data is questionable.
This is especially true for data relative to the solubility of water in hydrocarbons.These
data have been compiled by Daubert and Danner [1985] and empirical correlations have
been proposed. Equation 9.10 is an example. It allows us to calculate the solubility in mole
fractions (with approximately 30% precision), of water in a single hydrocarbon, or in a
mixture of hydrocarbons characterized by the mass ratio W,/W, of hydrogen to carbon:
(9.10)
The pressure is the three-phase equilibrium pressure, and is in fact (see the section that
follows) very close to the sum of the vapor pressures of water and of the hydrocarbon phase.
Data for the solubility of hydrocarbons in water is more available, in particular because
of their importance in the area of pollution. Several correlations have been made by Yaws
[1990,1991a,1991bl.We shall come back to this point (Section 9.6.2).

9.6.1 Total Immiscibility Hypothesis:


Calculation of the Three-Phase Equilibrium
Within a wide range of temperature and pressure, the mutual solubility of water and
hydrocarbons is very low, and it often occurs that the problem at hand is such that it can be
neglected. The calculation of phase equilibria is therefore relatively simple. In an effort to
simplify the matter, we shall introduce the principle using low pressure, such that the vapor
phase may be approximated by a mixture of ideal gases.

Binary systems
Due to the hypothesis of total immiscibility in liquid phase, the equilibrium diagram is similar to the one in Figure 9.10, but the curves representing the limit of liquid-liquid solubility (abscissa point x and x) and the bubble curve (curve AQ, for example) are merged with
abscissa lines 0 and 1, which relate to the hydrocarbon and water.
Let us say that water is indicated by subscript 1, the aqueous liquid phase by L, the
hydrocarbon by subscript 2, the hydrocarbon liquid phase by L, and the vapor phase by V.
The equilibrium conditions are then:
L-V equilibrium
P y , = PT
(9.11)
L-V equilibrium

Py2 = P:

9. Liquid-Liquid and Liquid-Liquid-Vapor Equilibria

349

They allow us to calculate the bubble pressure at the given temperature of the heterogeneous liquid mixture:
(9.12)
bubble =
I-k 7
To calculate the bubble temperature at a given pressure, we find the temperature of the
vapor pressures such that the preceding equation is satisfied.
The composition of the vapor at the heteroazeotrope is found using equation:

(9.13)

y1=

The calculation of the dew point must take into account two possibilities.The first drop
of liquid phase to settle is made up of either water or hydrocarbon.The two Equations 9.11
are therefore simultaneously tested. At given pressure, we thus obtain two values for temperatures Tl (L-V equilibrium) and T2(L-V equilibrium).If Tl > T2,the dew point corresponds to the appearance of an aqueous phase. If not, it corresponds to the appearance
of a hydrocarbon phase. Finally, if by coincidence Tl = T2,we are at the heteroazeotrope.
The reasoning is similar for determining the dew pressure. The criteria are reversed. We
note that on the isothermal equilibrium diagram the two arcs of the dew curve are hyperbolic.
Systems made up of water and a mixture of hydrocarbons

The equilibrium conditions are:


L

and:

+ V equilibrium PyI= Py

+ V equilibrium Pyi= PPxf y:


n

where:

zx(=l
2

(9.14)

i = 2,n

(9.15)

and

xy,=1
1

There are three possibilities:


Equilibrium between an aqueous liquid phase ( L ) and the vapor phase: Equation
9.14 must be verified.
0 Equilibrium between a hydrocarbon liquid phase (L) and the vapor phase: Equations 9.15 must be verified.
0 Three-phase equilibrium: Equations 9.14 and 9.15 must both be verified.
If, for example, we follow a condensation process, generally we see two subsequent
steps. During the first step, there is a two-phase L-V or L-V equilibrium according to
what is initially verified (by decrease in temperature at constant pressure, or increase in
pressure at constant temperature), either 9.14 or 9.15. If water condenses first, the relative
proportions of the hydrocarbons in the vapor phase remain unchanged, but their total
concentration increases. This remains true until we are able to simultaneously verify
Equations 9.14 and 9.15. We then come to a second step with three-phase condensation.
The compositions of the vapor phase, the liquid hydrocarbon phase, and the proportions of
the three phases will vary, all at the same time. The example below explains the calculation
of this last step.

350

9. Liquid-Liquid and Liquid-Liquid-Vapor Equilibria

EXAMPLE 9.1

Water (I), benzene (2), toluene (3) equilibria at atmospheric pressure

(x

The hydrocarbon liquid phase may be considered ideal = 1,i = 2,3). By application
of Equations 9.14 and 9.15, we obtain the compositions of the hydrocarbon liquid
phase and the vapor phase during three-phase equilibrium:
,

x2 =

P - Pp- P;
PF- P ;

p1"
y , = -,y2
P

PF
P

and
I

= -x2

and

(9.16)

x i = I-x;
Y3 =

P;

p x3

(9.17)

The liquid-liquid-vapor equilibrium temperature fluctuates between 69 and 84"C,values that correspond to the water, benzene, and toluene binary heteroazeotropes
respectively.Table 9.5 shows the results of the calculation, and the triangular diagram
in Figure 9.16 shows the locus of the vapor phases in three-phase equilibrium as well
as some isothermal equilibrium lines. The composition temperature binary diagrams
are represented on the three sides of the triangle.

Figure 9.16 Liquid-liquid-vaporequilibrium diagram of the water,


benzene, toluene system at atmospheric pressure.

351

9. Liquid-L iquid and L iquid-Liquid- Vapor Equilibria

Table 9.5

Water (l), benzene (2), toluene (3) three-phaseequilibria at atmospheric pressure

69
70
72
74
75
76
78
80
82
84

Yl

Y2

Y3

1
0.927
0.783
0.66
0.56
0.5
0.375
0.241
0.14
0.028

0
0.073
0.217
0.34
0.44
0.5
0.625
0.759
0.86
0.972

0.3
0.308
0.33
0.36
0.38
0.394
0.427
0.465
0.5
0.545

0.7
0.673
0.607
0.535
0.478
0.441
0.351
0.244
0.15
0.051

0
0.019
0.063
0.105
0.142
0.165
0.222
0.291
0.35
0.404

9.6.2 Application of Equations of State to the Calculation of


Phase Equilibria for Water Hydrocarbon Systems
In deposits and during production, natural gases are found to be in contact with an aqueous phase, often saline. Later on (Chapter lo), we shall see that hydrate formation may
result from such contact.This compound,however, only exists in the solid phase.
We shall consider only systems made up of hydrocarbons, most often of low molar mass,
and water, possibly salt laden. We use data pertaining to the behavior of water hydrocarbon mixtures under high pressure, especially within the critical zone [for example, see
Brunner, 19901.The solubilities of light hydrocarbonsin water are also available (Solubility
Data Series, 1982,1986,1987).
In order to use the same methods as those generally applied in the calculation of phase
equilibria in petroleum reservoirs, we usually seek to adapt the cubic equations of state to
water hydrocarbon systems. We know, however, that due to the polarity of water, neither
the equation of state itself, nor the mixing rules are applicable. We shall provide three
examples where these methods have been successfully adapted.
One modification of the Peng-Robinson equation (see Chapters 4 and 8) has been suggested by Soreide and Whitson [1992]. In order to faithfully reproduce the vapor pressure
of water, the function 47)is modified. If we apply the classical mixing rule (Eqs. 8.27 and
8.29) to parameter a, we cannot simultaneously reproduce the solubility of hydrocarbons
in water and water in hydrocarbons. Two distinct values are therefore proposed for the
interaction parameter kti.The parameter relating to the hydrocarbon phase and the vapor
phase is temperature independent. The parameter that relates to the aqueous phase
depends on the temperature and salt content by a correlation that is a function of the acentric factor of the hydrocarbon whose solubility we wish to calculate. The results are good
[de Hemptinne, 19941, and the greatest deviations are found at high temperature.
Nevertheless,it is appropriate to note that they are very sensitive to the value of the acentric factor, and that only sodium chloride solutions have been studied. Finally, it is suggested that the application of this method to heavier hydrocarbons be verified.

352

9. Liquid-Liquid and Liquid-Liquid-Vapor Equilibria

In order to avoid the disadvantage associated with the use of two different interaction
parameters,Darridon [1992] has introduced a coupling function that assures the continuity
of the representation, and a group contribution method has been applied to the calculation
of the parameters. Saline solutions have not been studied.
We must emphasize that the problem of representing water hydrocarbon equilibria is
often aggravated by the presence of a third component such as methanol. This third solvent is used for two purposes: as a hydrate formation inhibitor, and as a selective solvent
for the absorption of acid gas (CO,, H,S), found in natural gas. The representation of
phase equilibria for such systems poses a problem that the previously cited methods cannot solve. Application of the method using the excess functions at constant packing fraction (Chapter 8, Section 8.4.3) has been tried [Hocq, 19941.The definition for the excess
function takes into account a chemical term that stands for the auto-association of
methanol. The method has been applied to a large number of ternary, quaternary, and
higher order systems, but we observe large deviations in the concentration of the minority
compounds (methanol and water in hydrocarbon phase, hydrocarbons in the aqueous
phase). Furthermore, due to the lack of experimental data, the method cannot be applied
to mixtures simultaneously containing water, methanol, and light hydrocarbons.

CONCLUSION
It must be admitted that the problem of calculating liquid-liquid-vapor equilibria has not
yet received a satisfactory solution. We have pointed out the difficulty.We must be wary of
any method that is not founded upon experimental data. Any simulation with units using
such equilibria most often requires laboratory measurements, and a certain degree of
experience at the time of data correlation.There is still much to be done in this area.

REFERENCES
Brunner E (1990) Fluid mixtures at high pressures. IX. Phase separation and critical phenomena in 23
(n-alcanes + water) mixtures. J. Chem. Therm., 22,335-353.
Chen RJJ, Chappelear PS, Kobayashi R (1976) J. Chem. Eng. Data, 21,213.
Darridon JL (1992) Mesure et representation des Cquilibres de phases sous pression de melanges
deau, de paraffines et de dioxyde de carbone. Thesis, Universite de Pau.
Daubert Th E, Danner RP (1985) Technical Data Book - Petroleum Refining, chapter 9. American
Petroleum Institute, metric edition.
Deiters UK, Bluma M, Hintziger G, Kraska Th (1995) Phanomenologie und Berechnung von Fluid Phasengleichgevichten.Course at the University of Koln: Unpublished works.
de Hemptinne J-Ch (1994) Comparaison de differents moddes de calcul pour la solubilite des gaz
(CH,, CO,, H,S) dans leau en presence de sels. ZFP Report No. 41439.
de Swaan Arons (1995) The systematic study of phase behaviour and the emerging coherence of phenomena. Fluid Phase Equilibria, 104,97-118.

9. Liquid-Liquidand Liquid-Liquid-VaporEquilibria

353

Francis AW (1944) Solvent selectivity for hydrocarbons measured by critical solution temperature.
Znd. Eng. Chem.,36,764-771 et 1096-1104.
Hocq (1994) h d e experimentale et modelisation thermodynamique des melanges mkthanol, eau,
hydrocarbures. Sc.D. Thesis, UniversitC dAix-Marseille 111.
Horbson RW, Hartman RJ, Kanning EW (1941) J. Am. Chem. SOC.63,2094.
Lin YN, Chen RJJ, Chappelear PS, Kobayashi R (1977) J. Chem. Eng. Data, 22,402.
Ma YH, Kohn JP (1964) Multiphase and volumetric equilibria of the ethane methanol system at temperatures between -40 and 100C.J. Chem. Eng. Data, 9,3-8.
Magnussen T, Rasmussen P, Fredenslund Aa, Znd. Eng. Chem. Proc. Des. Dev., 20,331.
Matous J, Novak JP, Sobr J, Pick J (1972) Collect. Czech. Chem. Comm.,37,2653.
Mikitenko P (1967) Unpublished works.
Rowlinson JS, Swinton FL (1982) Liquids and liquid mixtures. Butterworth, London.
Scott RL, van Konyenbourg PH (1970) Discuss. Faraday. SOC.,49,87.
Solubility Data Series (1982) Ethane. Vol. 9. Pergamon Press.
Solubility Data Series (1986) Propane, n-butane, isobutane. Vol. 24. Pergamon Press.
Solubility Data Series (1987) Mkthane. Vol. 27-28. Pergamon Press.
Soreide I, Whitson CH (1992) Peng-Robinson predictions for hydrocarbons, CO, and H,S with pure
water and NaCl brine. Fluid Phase Equilibria, 77,217-240.
Sorensen JM, Magnussen Th, Rasmussen P, Fredenslund Aa (1979) Liquid-liquid equilibrium data:
their retrieval, correlation and prediction. Part I: Retrieval. Fluid Phase Equilibria, 2,297-309.
Tagliavini G,Arich G (1958) Ric. Sci. 28,1902.
van Konyenbourg PH, Scott RL (1980) Phil. Trans.,A298,495.
Vorobeva AT, Karapetyants M Kh (1967) Zh. Fiz. Khim., 41,1144.
Waksmundski A, Stelmach K, Wolski T (1964) Przem. Chem. 43,194-197,445-449,548-550.
Wittrig TS (1977) Ph. D. Thesis, Illinois.
WolskiT (1970) Rocz. Chem.,44,2237.
Yaws CL, Yang H-C, Hopper JR, Hansen KC (1990) Hydrocarbons solubility data. Chem. Eng.,
97 (4), 177-182.
Yaws CL, Yang H-C (1991a) Henrys law constants for 362 organic compounds in water. Chem. Eng.,
98 (ll), 179-185.
Yaws CL, Pan X (1991b) New correlation calculates reliable paraffin solubilities. Oil & Gas Journal,
89 (4), 79-81.
Zeck S, Knapp H (1986) Vapor-liquid and vapor-liquid-liquid phase equilibria of binary and ternary
systems of nitrogen, ethene and methanol: experiment and evaluation. Fluid Phase Equilibria, 26,
37-58.

10

Fluid-Solid Equilibria
Crystallization Hydrates

The study of liquid-solid equilibria, or crystallization, is an enormous field of study. Even


with a sketchy introduction, we would stray far from the petroleum applications that are
the subject of this study.While emphasizing the specifics of such phase changes and establishing the form according to which the equilibrium condition is most often expressed, we
shall limit ourselves to the discussion of two examples. On the one hand, the deposit of
paraffins from a crude oil or gas-oil hydrocarbon liquid phase [see Pion, 19941, and on the
other, solid phase deposits of high molar mass compounds from a natural gas [see Ungerer
et al., 19951.
There is another type of equilibrium involving natural gases and a solid phase: hydrate
formation. These are clathrate compounds formed by one or more light hydrocarbon
molecules trapped within a crystalline structure made up of water molecules and stabilized by the presence of the hydrocarbon. In natural gas reservoirs, or within the transport
pipes for this gas, the presence of water and the conditions of temperature and pressure
are favorable for the formation of these compounds. The practical consequences (pipeline
obstruction) of these formations can be disastrous. We shall conclude this chapter with a
brief description of these compounds and the calculation method for such equilibria.
We have repeatedly emphasized the continuity that exists between the liquid and vapor
phases. An initial analysis shows that the case is very different for the solid phase. In particular, the primary physicochemical characteristic of a solid phase is its crystalline structure. This structure can vary considerably even with compounds of similar chemical structure. Furthermore, depending on the temperature range, any given compound may exist in
various structures or allotropic forms. Table 10.1 shows the structures for linear paraffins.
We note the differences that exist depending on the parity of the number of carbon atoms,
and depending on temperature.
Due to this fact, the formation of a homogeneous mixture in the solid state is far from
being the rule. The presence of a solid mixture requires not only a favorable energy balance (decrease in Gibbs energy), but also geometric compatibility: crystalline structure
and grid parameters.
Furthermore, the nucleation phenomena have considerable importance during the crystallization process. It is common to observe states outside the equilibrium state resulting
from a lag in phase change. This occurs within a particularly large temperature range that
may reach several dozens of degrees. These lags occur both for liquid-solid transitions as

356

10. Fluid-Solid Equilibria. Crystallization. Hydrates

for transitions between two crystalline forms.The results often show up as discrepancies in
the available experimental data, as well as in the thermodynamic equilibrium and the
actual behavior.
Finally, it would be inadequate to approach the subject of liquid-solid equilibria without
taking into account the interface phenomena that is not limited to the nucleation process.
Table 10.1
Crystalline structures of the n-paraffins CnH2n+2
[Asbach et al., 19701
Temperatures Slightly Lower
Than the Melting Temperature

11 <o d
n d< 39

Even
n < 22
n = 24,26
26 < n < 32
34 < n < 42
n>44

10.1

Low Temperatures

Orthorhombic or hexagonal

Orthorhombic

Triclinic
Monoclinic or hexagonal
Monoclinic or hexagonal
Orthorhombic or hexagonal
Orthorhombic

Triclinic
Triclinic or monoclinic
Monoclinic
Orthorhombic
Orthorhombic

LIQUID-SOLID EQUILIBRIUM DIAGRAM

The simplest diagrams correspond to the case of two totally immiscible substances in the
solid state. Figure 10.1 shows the case for the metaxylene (melting temperature ef=-48C)
paraxylene (melting temperature Of = 13.2"C) mixture. If we chill a mixture containing
65% metaxylene, at -27C a solid phase appears that is made up of pure paraxylene. By
decreasing the temperature, this phase accumulates and the liquid mixture becomes more
concentrated in metaxylene. On the temperature composition diagram, its representative
points describe the arc of a curve. If the initial mixture had had more than 87.8% metaxylene, the solid phase deposited would have been composed of pure metaxylene, crystallization would have decreased the concentration of metaxylene in the liquid phase, and the
representative point for this phase would describe a second arc. These two arcs intersect at
one point where a liquid phase composed of 87.8% metaxylene, and two solid phases made
up of the two, pure isomers coexist. The point is called the "eutectic" point, and since variance is equal to 1 at given pressure, one atmosphere in this instance, the temperature is
fixed and equal to -52.5"C. Mixtures of the n-paraffins often yield such diagrams, n-decane
+ n-dodecane, n-C30 + n-C,,, for example. We note that the description for such a phase
change is completely analogous to the description that we would have made for the condensation of a binary mixture whose two compounds have a negligible mutual miscibility
in the liquid phase, water and hydrocarbon, for example.
We have seen that depending on temperature, a compound may exist in various crystalline forms. On the crystallization diagram, existence of these forms is observed as a

357

10. Fluid-Solid Equilibria. Crystallization. Hydrates

-20

0.2

0.4

0.6

0.8

Mole fraction of metaxylene

Figure 10.1 Liquid-solid equilibrium diagram for the metaxylene


paraxylene mixture.

Mole fraction of A

I
1

Figure 10.2 Liquid-solid equilibrium diagram for a mixture where one of


the components exists in two crystalline forms, a and p.

break in the slope of the curve representing the liquid phase composition change as a
function of temperature, such as shown in Figure 10.2. This disruption in the slope is
related to the enthalpy of transition between the two crystalline forms.The break is found
in the crystallization diagrams for the paraffins, but is not very marked, and is located very

358

10. Fluid-Solid Equilibria. Crystallization.Hydrates

close to the melting point. We can cite the n - q , + n-C3,,n-C,, + n-C,, , n-C,, + cyclohexane systems.Table 10.2 [Pion, 19941 lists the melting and transition temperatures of the
components.
Table 10.2
Temperatures and enthalpies of melting
and transition for the n-paraffins
(Units:K and J mol-')
+

Paraffin

n-C,

317.26
334.60
338.70
354.65

48 125
64 668
68 438
137 580

315.85
330.80
334.50
353.15

29 286
35 241
37 489

There may be partial miscibility,and in this case, the crystallization diagram is similar to
the one shown in Figure 10.3. In this figure, we recognize the eutectic point, the threephase equilibrium line, and the coexistence ranges of the liquid phase with solid phase S,,
with solid phase S,, and of the two solid phases S, and S,. Remaining in the field of n-paraffin mixtures, we cite the n-C,, + n-Ci3 or the n-C,, + n-C,, mixtures.
For the case of total miscibility, the crystallization diagram may have the very simple
shape of a liquid-vapor equilibrium diagram, such as in Figure 6.1. We may also observe
the appearance of a temperature minimum, as in the azeotropic diagram, and at this point,
the curves for the liquid and solid phases are tangential. Solid solutions covering the entire
composition range do not seem to exist for the n-paraffin mixtures, unless parity is

Mole fraction of A

Figure 103 Liquid-solid equilibrium diagram with partial miscibilityin


solid phase.

359

10. Fluid-Solid Equilibria. Crystallization. Hydrates

the same and the number of carbon atoms differs little. This can be observed for the n-C,,
+ n-c,?,,n-C, + n-C24,and the n-C,, + n-c,?, systems.
Finally, we ought to mention the crystallization of partially miscible systems in the liquid phase. They are prone to split into two liquid phases L, and L,, shown on Figure 10.4,
in which the two compounds are additionally assumed to be totally immiscible in the solid
phase. If we call them A and B, and the two phases that they make up in the crystalline
state S, and S,, depending on the initial mixture composition, we observe the solid-liquid
two-phase equilibrium S,, L,, or S,, L,, or the liquid-liquid equilibrium L,, L,, or finally
S, L,. The system has two eutectic points, E and D .
There are many other possibilities,in particular due to the formation of compounds of
defined stoichiometry, with or without congruent melting points, but the diagrams that we
have briefly discussed seem to be the most common in the field that interests us.

Mole fraction of A

I
1

Figure 10.4 Crystallization diagram of a partially miscible mixture in


liquid phase.

10.2

CALCULATION OF CRYSTALLIZATION EQUILIBRIA

10.2.1 General Equations


If only compounds with low vapor pressure are involved in the mixture, the pressure has
practically no influence on the liquid-solid equilibrium. This influence can be evaluated
applying the Poynting corrections to one or the other phase, and after calculation, turns
out to be minor for pressure variations in the dozens of bars. We shall therefore neglect
this variable in the following discussion. We shall introduce it only for the purpose of
studying the crystallization of heavy compounds from a natural gas.

3 60

10. Fluid-Solid Equilibria. Crystallization. Hydrates

First, we shall establish the equilibrium equation for the simplest case, where the crystallized phase is made up of a pure substance (Fig. 10.1).
If we say that subscript i is the component that crystallizes,fL*s
and fi*" are the fugacity
values in the pure state, solid and liquid respectively, and xi and yf the mole fraction and
the activity coefficientin liquid phase, we can first state the equality of the fugacities of this
component in the two phases:

fi" = ft?"Xi yi"

(10.1)

Of course, for this component in the pure state the liquid state is unstable as soon as the
temperature is lower than the melting temperature. It is characterized as a sub-cooled liquid state.
The ratio fL?/fi" is related to the Gibbs energy difference at temperature T for the
change from solid to liquid state Ag;(S + L) for the pure substance, such that we may
rewrite the previous equation as:
(10.2)
This Gibbs energy difference is not known, but may be calculated as a function of its
value at the melting temperature Tmof the component,Ag'T, (S + L). It is zero at this temperature, since there is equilibrium between the two states, solid and liquid, of the pure
substance:

where:
In order to calculate the derivative of the Gibbs energy with respect to temperature, we
apply the Gibbs-Helmholtz equation, and thus obtain:

For the pure substance, A h x S + L) is the enthalpy of melting at temperature T. It


depends on the temperature through the heat capacities of the two phases whose difference is written as Ac*,(S -+ L). We assume that this term is independent of temperature
and equal to its value at the melting temperature T,,,.We hence obtain:
Ah*,(S + L ) = Ah*,m(S -+ L ) + Ac; ( S + L ) - ( T - T m )

(10.5)

361

10. Fluid-Solid Equilibria. Crystallization. Hydrates

and we may write Equation 10.4 as:


Ag>(S + L)
RT
-

(S + L)
RTm

T
($
-l)+

AcL (S + L)
R

Tm
T

- + 1-

5)
(10.6)
T

Ultimately, the equilibrium Equation 10.2 is written as:

In terms of the compound that crystallizes, it uses the heat of melting and the difference
of heat capacities.The latter is often neglected, and we then have:
(10.8)
This form is used when determining the molecular weight of a solute by cryoscopy. We
prepare a dilute solution of this solute of known composition by weight. The preceding
equation is applied to the solvent for which we measure the melting point depression. In
this equation, the enthalpic term is calculated either from literature data, or from a prior
calibration. We admit that because the solute is diluted, the activity coefficient for the solvent is equal to 1, such that we can calculate the molar composition xi. By comparing it
with composition by weight, we derive the molar mass of the solute.
Of course, if the solid phase presents various crystalline forms (Fig. 10.2), we need to
break the calculation down into several steps, separated by the transition temperatures. We
account for the continuity of the Gibbs energy or fugacity of the solid. If S, and S2 are the
two crystalline forms in equilibrium at the transition temperature T,, then:

(4 + S2) = 0
and:

or

&>,(Sl+ L) =

jy =

(S, + L )

(10.9)
(10.10)

On the other hand, the enthalpy of the solid phase is discontinuous, and its stepwise
evolution will show the transition enthalpies.
In general, it is necessary to take into account the deviations from ideality in the liquid
phase, shown here by the activity coefficient
They depend on temperature and composition, and may be calculated using models that we have already introduced (Chapter 7).
However, we note that for a given binary mixture, the crystallization equilibria and the liquid-vapor equilibria generally do not involve a comparable temperature range. Hence, the
use of activity coefficients obtained from the regression of liquid-vapor equilibrium measurements in liquid-solid equilibrium calculations most often involves an extrapolation of
temperature. In this way, the binary methanol water system has been studied many times
above 50C, but until recently [Jose, 19941,we did not have access to data that provided an
accurate evaluation of activity coefficients within the crystallization zone.
The influence of the non-ideality model in the liquid phase can be great. Figure 10.5
allows us to compare the experimental values of the crystallization temperatures for benzene in n-undecane mixtures with the values calculated using three hypotheses: ideal solution, UNIFAC model, and the method proposed by Abdoul et al. [1991], already men-

xL.

3 62

10. Fluid-Solid Equilibria. Crystallization.Hydrates

270

xg 260
h

-E

al

250

F
240

230

:0

0.2

0.4

0.6

0.8

Mole fraction of mundecane

Figure 10.5 Calculation of crystallization equilibrium for the n-undecane


benzene system [Zhou,19911.
: ideal solution;
0:experimental values;
: Abdoul method.
. -: UNIFAC method

-----

tioned during the discussion of the calculation of liquid-vapor equilibria under pressure
(Chapter 8, Section 8.4.3.).
When there is a solid solution, the equilibrium equation is written as:

fi*sxsy s = fi*Lxk y;

(10.11)

where fi*, x i , and are, respectively, the fugacities of the component in the pure state, the
mole fractions, and the activity coefficients. Exponents L and S specify the physical state,
liquid or solid. The ratio fi*L/fi*s may be calculated using heat of melting and the differences in heat capacity as above, and the equilibrium Equation 10.7 becomes:

The activity coefficients in the liquid phase will be calculated using one of the previously discussed models. However, there are no data for deviationsfrom ideality in the solid
phase. We know only that miscibility is generally limited,and often very low. The proposed
models for the calculations of these deviations in the liquid phase probably require serious
modifications.Characteristics such as the nature of the crystalline lattice and its geometric
parameters may have a significant influence. Nevertheless, it has been proposed to use the
concept of regular solutions for the solid phase in order to predict the crystallization equilibria of the paraffins [Won, 19891.

10. Fluid-Solid Equilibria. Crystallization.Hydrates

363

10.2.2 Paraffin Crystallization


The problem of solid hydrocarbon deposit from crude oil inside wells or transport pipes
during the exploitation of some fields, or from petroleum fractions such as gas-oils, has
obvious practical implications. The development of a model capable of predicting conditions of temperature and pressure at which crystallization begins, or the rate of crystallization under given conditions,is less obvious, but recently has been the subject of many studies [see, for example, Coutinho, 1995; Pion, 1994;Won, 19891. In particular, we must deal
with a number of unknowns:
0 The components are poorly identified and too numerous to be able to be accounted
for individually.It is necessary to group them (see Chapter 12).For these components,
we do not necessarily have elementary data such as temperature and enthalpy of
melting or transition, for example.
0 These deposits may be classified into two categories [Pion, 1994;Gilby, 19831. On the
one hand, macrocrystallinewaxes,mainly formed from normal paraffins,and on the
other, microcrystalline waxes, containing appreciable proportions of branched
paraffins and alkylcyclopentanesor cyclohexanes.
0 We ignore the nature of the deposit formed: pure crystalline phases or several solid
solutions with limited solubility characteristics.
0 As we have just seen, our non-ideality models in the crystalline phase do not have a
theoretical basis.
0 Finally, we have little experimental data concerning the conditions that lead to these
deposits.
Coutinho [1995] presents a sketch (probably simplified) of phase diagrams for paraffin
mixtures with similar chain lengths showing the differences due to chain parity. It also
shows the existence of solid solutions within the hexagonal crystalline zone, also called
rotator as well as, at low temperatures, the equilibria between the mono- and triclinic
and orthorhombic phases. This sketch is shown in Figure 10.6.
The model proposed by Won [1989] assumes the formation of a single solid phase made
of a homogeneous mixture containing all crystallized components. In Equation 10.7, the
differences in heat capacities are neglected, and the enthalpic term is the sum of the heat
of melting and the heats of transition.This hypothesis is correct if the temperature or temperatures of transition are close to the melting temperature, which is the case as shown by
the data in Table 10.2.Finally,the regular solutions model (see Chapter 7) is applied to the
two phases.The solubility parameters in the solid phase are calculated for each component
from the sum of the heat of melting of the solid, and the heat of vaporization of the liquid.
In general, it seems that this model produces a large overestimation of the initial crystallization temperatures.
Hansen [1987] believes that the solid phase is an ideal mixture. However, he takes into
account the solid-liquid interfacial energy term in the expression for the Gibbs energy of
this phase. The liquid phase is represented with the Flory-Huggins model (see Chapter 7,
Equation 7.4). The results obtained for North Sea crude oils appear good, but it must be
mentioned that the model was established from these very paraffin oils.
The recent study by Coutinho et al. [1994] leads to the inclusion of a combinatorial
term, a free volume term, and an energy interaction term in the non-ideality model applied

3 64

10. Fluid-Solid Equilibria. Crystallization. Hydrates

Liquid

Liquid

Orthorhombic

0.5

0.5

Liquid

Rotator II
p!

0
.-c
.0

Orthorhombic

2
I-

I
0.5

Figure 10.6 Phase diagrams of linear paraffin mixtures.


a two even chains;b. two odd chains;c. odd and even chains.

to the liquid phase, analogous to the methods applied to polymer solutions (see
Chapter 11).The hexagonal and mono- or triclinic phases are represented by a symmetric
excess Gibbs energy model for mole fraction [Coutinhoet al., 1995al.However, the Wilson
equation (see Chapter 7) is applied to the orthorhombic phases and the parameters are
determined from the enthalpies of sublimation [Coutinhoet al. 1995bl.The theoreticaljustification for these choices remains weak, but this method allows for a good prediction of
experimental data relating to ternary mixtures [Coutinho and Stenby, 19951.Yet these are
rare, and often imprecise.

10. Fluid-Solid Equilibria. Crystallization. Hydrates

365

10.2.3 Fluid-Solid Phase Transition at High Pressure


Some recently discovered natural gas deposits are characterized by:
0 Conditions of particularly high temperature and pressure.
0 A fluid composition that, next to high proportions of methane (generally greater than
60% molar) also contain heavy hydrocarbons with up to 40 carbon atoms.
These features are reflected in the phase diagram that in some cases may show phase
transitions between the single-phase fluid, two-phase liquid + vapor, and the threephase liquid + vapor + solid zones, as shown in Figure 10.7. In this diagram we also note
that the value of the critical temperature (approximately 75C) indicates that the fluid is a
gas condensate within the reservoir,and a light oil in the treatment plant and transport pipes.
Furthermore,at the transport conditions,we can observe the presence of a solid phase.
Under these conditions, we cannot establish a calculation method for deviation from
ideality in liquid phase as we have done previously. It is better to represent the fluid phase
using an equation of state. This is what was done by Soave [1979] for fluid-solid equilibria
in carbon dioxide,light hydrocarbon mixtures.
A similar model was extended by Ungerer et al. [1995] to the fluid deposits mentioned
above, and verified using experimental data relating to synthetic mixtures whose composition is close to that of real fluids.The crystallizable components were phenanthrene or hexan-paraffin). Phase transitions analogous to the ones shown in Figure 10.7
triacontane (C,
for a natural gas were indeed observed, especially the formation of a solid phase at temperatures 10 to 30C below the melting temperature of the pure crystallizable component.
The fugacity of the component that crystallizes, in the pure state as a subcooled liquid
and in a fluid mixture f i L is calculated using the Peng-Robinson equation of state.The
interaction parameter between methane and the crystallizable component are two-phase
fitted on the envelope (dew and bubble curves).The calculation of the fugacity of the same
component in the solid state is taken from Equation 10.7 (Section 10.2.1),but this time the
influence of pressure on the crystallization equilibrium must be taken into account. The
data necessary for the calculation are the melting temperature T, and the transition temperatures Tkk = 2, n at reference pressure po,usually equal to atmospheric pressure; the
corresponding variations in enthalpy Ahm and Aht,k; the differences in heat capacity
between the solid phases and the pure substance in a subcooled liquid state Ace,, and the
difference in molar volume Av, between the liquid and solid states. We end up with the
following expression for the Gibbs energy difference (that is to say, the fugacities ratio)
between the two states:pure substance in crystallized solid state in the form of allotrope n,
and pure substance in subcooled liquid state, both under the same conditions of temperature T and pressure P. In this expression, subscript i is the component that crystallizes, to
which the values characterizing melting (T, ,Ah,, Av,) and the transitions (Tkk = 2, n ,
Ahk,t,Ace k ) are assigned

366

10. Fluid-Solid Equilibria. Crystallization. Hydrates

50

100

150

200

Temperature (C)

Figure 10.7 Phase diagram of a condensate gas [Ungerer et al., 19951.

If the transition temperatures are close to the melting temperature, and if we neglect
the heat capacity differences,this equation becomes:
n

The equilibrium equation is identical to the one that we have already applied:
(10.2)
except that the activity coefficient in liquid phase is calculated from the equation of state:
SL

(10.15)

3 67

10. Fluid-Solid Equilibria. Crystallization. Hydrates

The model established in this manner effectively allows us to quantitatively plot the
entire phase diagram. It is appropriate to note that only one adjustable parameter has
been used: the interaction parameter of the equation of state. It has been fitted on the liquid-vapor equilibrium data. The calculation of phase transitions corresponding to crystallization is therefore purely predictive.

10.3

HYDRATES

10.3.1 Generalities
Certain gases, especially the low molecular weight hydrocarbons (methane, ethane,
propane, isobutane), either pure or in mixture, may create hydrate crystals in the presence
of water under given conditions of temperature and pressure.The crystalline framework is
made up of water molecules, which is stabilized by the hydrocarbon molecules that are
found within the cavities of this framework. We are talking about clathrate compounds,
within the larger family of the clathrates. It is especially important in practice to be familiar with the stability conditions of such structures. Under reservoir conditions,large quantities of natural gas may be found in the hydrate state. Under transport conditions,the formation of these solid phases may cause pipeline plugging.
In this study we shall consider this question only very briefly. A very complete study has
been proposed by E. D. Sloan [1990].
Basically,we can distinguish two crystalline structures,I and 11,which differ in their geometric parameters, and whose viability depends on the hydrocarbon that stabilizes them.
Each one of them has two types of cavities, large or small. Table 10.3 shows the general
characteristics of the unit cell of type I or 11, and Table 10.4 shows the different possible
combinations as a function of the nature of the component responsible for structure stability.
Another structure, called structure IT,has recently been observed by Ripmeester ef al.
[1987]. It is made up of three types of cavities MI, M2,and M3,according to the stoichiometry [3M,, 2M2, M 3 , 34 H20]. However, this structure appears only in the simultaneous
presence of a light hydrocarbon (methane) and a hydrocarbon such as methylcyclohexane,
for example.
Table 10.3
General characteristics of type I and I1 cells

I
Number of water molecules
Number of small cavities
Number of large cavities
Diameter of small cavities (A)

Structure1

7.95
8.6

Structure11

7.82
9.46

368

10. Fluid-Solid Equilibria. Crystallization. Hydrates

Table 10.4

Possible content of cavities


Structure I
Cavity
Small
Large
Carbon dioxide
Nitrogen
Hydrogen sulfide
Methane

%a--p&T

+
+
+
+

Ethane
Propane

n-Butane

Isobutane

StructureI1
Cavity

It is surprising to see a structure consisting of water stabilized by the presence of hydrocarbon molecules when we know that in the liquid phase, water and hydrocarbons are
practically immiscible, and do not appear to have any mutual affinity whatsoever. This
point has been debated by BChar et al. [1994]. It seems that in fact, in diluted aqueous solution, the presence of hydrocarbons causes the formation of water molecule aggregates,
linked by hydrogen bonds, therefore having an ordered structure with decreased entropy.
These aggregates thus make up the hydrate precursors.

10.3.2

Phase Diagrams

Figure 10.8 [BChar et al., 19941 shows the phase diagram for an ethane water mixture.
Along the dotted line we see the vapor pressure curve for ethane (the curve for water is
not shown), as well as several zones:
0 Zone G-L,,where ethane in gaseous phase ( G ) and water in liquid phase (L,)
coexist.
0 Zone L,-L,, separated from the above zone by the ethane vapor pressure curve,
where two liquid phases coexist: one (L,) made up of water, and the other (L,) from
ethane.
0 Zone H-G,also two-phase, representing the simultaneous presence of the gaseous
hydrocarbon ( G )and hydrate (H).
0 Zone H-L,, separated from the H-G phase by the vapor pressure curve for
ethane. Two phases are present here: the hydrate phase (H), and the liquid ethane
phase (L,).
0 And finally, below WC,for pressures below the vapor pressure of ethane, the G-Z
zone, relating to gaseous ethane and ice.
It is understood that the L, and L, phases are not made up of water or ethane exclusively,but the mutual solubility of these compounds in liquid phase is low.

10. Fluid-Solid Equilibria. Crystallization. Hydrates

3 69

10

Figure 10.8 Phase equilibria for ethane,water, hydrate.


-.-.-.*

*-

.ethane vapor pressure curve;

.three-phase equilibria (see text);

----- :water crystallisation.

The line (solid) separating the existence ranges of the hydrate (zones H-G or H-L,) and
the ranges where water is liquid (zones G-L, and 154,) represents monovariant threephase states H-L,-L, or H-G-L,.Point Q2located at the intersection of this line and the
ethane vapor pressure curve is invariant,since four phases are in equilibrium:the aqueous
liquid phase, the liquid and vapor phases made up of ethane and the hydrate phase. The
same is true for point Q, that represents the coexistence of ice, liquid water, gaseous
ethane, and the hydrate.
As we can see, the risk of hydrate formation appears at around 15Cif the pressure is on
the order of a few dozen bars. For methane, the temperature of hydrate formation varies
from 0C at 13 bar to 20C at 125 bar.
In the case of a natural gas or oil, the diagram is shown by Figure 10.9.The liquid-vapor
equilibrium curve (or vapor pressure) of the hydrocarbon is split into dew and bubble
curves that merge at the critical point K (see Chapter 6, Fig. 6.6). Inside this envelope
are located the three-phase ranges relative to the coexistence of the two liquid and
gaseous hydrocarbon phases with either liquid water (G + L , + L,) or the hydrate
( H + G + L,).

3 70

10. Fluid-Solid Equilibria. Crystallization. Hydrates

c
h

I
I

ti+ L,

Figure 10.9 Phase equilibria for natural gas (or crude oil), water
hydrates.
----- :two-phase envelope of the hydrocarbon mixture.
-:
three-phaseequilibria.

10.3.3 Calculation of Hydrate Formation Equilibria


The methods applied to the calculation of the equilibria of hydrate formation are derived
for the most part from the theory proposed by van der Waals and Platteeuw [1959] on the
basis of the following hypotheses:
0 The hydrate is considered a solid solution of the compound dissolved within the
unstable crystalline frame.
0 This crystalline frame is not distorted by the presence of the compound stabilizing it.
0 One cavity may not contain more than a single molecule of solute.
0 Insertion of the solute is considered as a single adsorption, and described by the
Langmuir theory.
In this study, we follow the practical development proposed by Parrish and Prausnitz
[19721.
If pf is the chemical potential of water in the hydrate devoid of any occupation,Nk the
number of cavities of type k, and Xi,k the fraction of cavity of type k occupied by the
hydrocarbon of rank i, then the chemical potential of water in hydrate p,,, is expressed by
the equation:
(10.16)

10. Fluid-Solid Equilibria. Crystallization. Hydrates

371

At equilibrium and depending on temperature, this chemical potential is equal to that


of ice or the liquid aqueous phase. In the latter case, we can also express it as a function of
the chemical potential of pure water in the liquid state p;L, the mole fraction in the aqueous phase xw, and its activity coefficient yw:

pw = p i L + RT In ywxw

(10.17)

The equilibrium condition is therefore expressed by the equation:


(10.18)
or:

, u ~ - & ~ = - R Nkln
T ~ 1-cX.
k

+RTlnywxw

(10.19)

,*)

It is necessary to calculate the values for the cavity fractions of type k occupied by the
hydrocarbon of rank i, x i , $ ' We do this using the Langmuir equation, which is generally
applied to adsorption processes:
(10.20)
where fi is the fugacity of the light component of rank j in liquid or vapor phase, and c j , k is
the Langmuir parameter characteristicof the adsorption of this compound within the cavity of type k.
We must also know the difference A& between the chemical potentials of water in the
empty hydrate p;, and pure water in the liquid state p;L:
AP; = P w*H -dL

(10.21)

It depends only on pressure P and temperature T, but is related to reference states


where at least one, the one relating to the empty hydrate, is unstable. In order to calculate
it, we use the experimental dissociation data of a reference hydrate, that formed by
methane, for example. If we say that To is a reference temperature (273.15 K), and Poand
Prefare the dissociation pressures of the reference hydrate at temperatures To and T, then
we may state:

and:

(10.23)

The values Ah; and Av: are at work in these equations, and respectively represent the
differences between the enthalpies and the molar volumes of water in the empty hydrate,
and in the pure liquid state.We consider that these values are independent of temperature
and pressure. The variation of the dissociation pressure of the reference hydrate with temperature dPre,ldT is expressed from an empirical correlation of the experimental data of
dissociation Pref(
T).

3 72

10. Fluid-Solid Equilibria. Crystallization. Hydrates

The Langmuir constants were obtained by Parrish and Prausnitz from the Kihara
potential (see Chapter 3) whose parameters were calculated from experimental data for
dissociation of hydrates. The same is true for the values Ah,* and Av:.
Many variations of this method have been proposed. In particular, we may quote the
works of Ng and Robinson [1976,1977,1980] as well as those of Munck et al. [1988].
One of the finer points concernsthe estimation of the activity of water in expression 10.19.
The aqueous phase contains low quantities of hydrocarbons, which may be neglected.
However, we must account for the solubility of carbon dioxide in water, and the fact that
reservoir water contains large quantities of dissolved salts. Munck et al. [1988] therefore proposes applying the UNIQUAC method to the calculation of the activity coefficient of water,
yw,such as it has been modified by Sander and later Macedo (see Chapter 7, Section 7.9).
The formation of hydrates may be inhibited by the addition of compounds such as
methanol or ethylene glycol to the aqueous phase. Of course, they must be taken into
account in the application of Equation 10.19. In this case, we may apply one of the models
discussed previously (Chapter 7) to the calculation of the activity coefficient of water, but
we must remember that the numerical values for the parameters of these models were
generally determined by liquid-vapor correlation in a temperature range that is very different from that of hydrate formation.

REFERENCES
Abdoul W, Rauzy E, Peneloux A (1991) Group contribution equation of state for correlating and predicting thermodynamic properties of weakly polar and non associating mixtures. Binary and multicomponents. Fluid Phase Equilibria, 68,47-102.
Anderson FE, Prausnitz JM (1986) Inhibition of gas hydrates by methanol. AZChEJ.,32,1321-1333.
Asbach GI, Kilian HG, Colab X (1970) Zur Thermodynamik von Mischsystem aus n-Paraffinen. Ber.
Bunsenges. Phys Chem.74 (8/9), 812-823.
BBhar E, Delion AS, H e m JM, Sugier A, Thomas M (1994) Le problkme des hydrates dans le
contexte de la production et du transport polyphasiques des petroles bruts et des gaz naturels.
Rev. Znst. Frang. du pttrole, 49 (3), 241-263.
Coutinho JAP (1995) Phase equilibria in petroleum fluids: multiphase regions and wax formation.
Thesis, Technical University of Denmark, Department of Chemical Engineering.
Coutinho JAP,Andersen SI, Stenby EH (1995) Evaluation of activity coefficient models in prediction
of alkane solid-liquid equilibria. Fluid Phase Equilibria, 103,23-29.
Coutinho JAP,Andersen SI, Stenby EH (1995a) Solid-liquid equilibria of n-alkanes using a modified
delta lattice parameter model. 7th Fluid phase properties and equilibria for chemical process
design, Snowmass, Colorado,June 18-23,1995.
Coutinho JAP, Knudsen K, Andersen SI, Stenby EH (1995b) Predictive local composition models for
solid-liquid and solid-solid equilibrium in n-alkanes. Wilson equation for binary systems. AZChE J.
Coutinho JAP, Stenby EH (196) Predictive local composition models for solid-liquid and solid-solid
equilibrium in n-alkanes. Wilson equation for multicomponent systems. Z&EC Research 35 (3),
918-925.
Fredenslund Aa, Gmehling J, Rasmussen P (1977) Vapor-liquid equilibrium using UNZFAC.Elsevier,
New-York, 1977.

10. Fluid-Solid Equilibria. Crystallization.Hydrates

3 73

Gilby GW (1983) The use of ethylene-vinyl acetate copolymers as flow improvers in waxy crude oil.
Chemical in the oil industry, Special publication, proceedings.
Hansen AB, Fredenslund Aa, Pedersen KS, Ronningsen HP (1987) A thermodynamic model for predicting wax formation in crude oils. AIChE J.,34,1937.
Hildebrand JH, Prausnitz JM, Scott RL (1970) Regular and Related Solutions. Van NostrandReinhold, Princeton.
Jose J (1994) Unpublished work.
Larsen BL, Rasmussen P, Fredenslund Aa (1987) A modified UNIFAC group contribution model for
prediction of phase equilibria and heat of mixing. Znd. Eng. Chem. Res., 26,2274-2286.
Munck J, Skjold Jorgensen S (1988) Computation of the formation of the gas hydrates. Chem. Eng.
Sci., 43,2661-2672.
Ng H-J, Robinson DB (1976) The measurement and prediction of hydrate formation in liquid hydrocarbon-water systems. Znd. Eng. Fundam., 15,293-298.
Ng H-J, Robinson DB (1977) The prediction of hydrate formation in condensed systems. AIChE 1,
23,477-482.

Ng H-J, Robinson DB (1980) A method for predicting the equilibrium gas phase water content in gashydrate equilibrium. Ind. Eng. Chem. Fundam., 19,33-36.
Parrish WR, Prausnitz JM (1972) Dissociation pressures of gas hydrates formed by gas mixtures. Ind.
Eng. Chem., Proc. Des. Dev. 11,26-34.
Pion S (1994) Modklisation thermodynarnique compositionnelle de la cristallisation des paraffines i
partir des effluents de production. IFP Report 41426.
Ripmeester JA,Tse JS, Ratcliffe CI, Powell BM (1987) A new clathrate hydrate structure. Nature 325,
135-136.
Sloan ED (1990) Clathrate Hydrates of Natural Gas. Marcel Dekker, New York.
Ungerer Ph, Faissat B, Leibovici C, Zhou H, Behar E, Moracchini G, Courcy JP (1995) High pressurehigh temperature reservoir fluids: investigation of synthetic condensate gases containing a solid
hydrocarbon. Fluid Phase Equilibria, 111,287-311.
van der Waals, JH, Platteeuw, JC (1959) Clathrate solutions. Adv. Chem. Phys., 2,1-57.
Won KW (1989) Thermodynamic calculation of cloud point temperatures and wax phase compositions of refined hydrocarbon mixtures. Fluid Phase Equilibria, 53,377-396.
Zhou H (1991) ModClisation des Cquilibres solide-liquide des systkmes binaires normale-paraffine,
hydrocarbure.IFP Report 38705.

11

Polymer Solutions and Alloys

It would be presumptuous to attempt to introduce the calculation methods for the thermodynamic properties of polymers, their solutions, and their mixtures in just a few pages.
Many studies have been dedicated to the subject, and in particular we may refer to the
work of van Krevelen [1990]. We address this topic because of its importance in certain
petroleum applications such as transport via flexible pipes. We shall limit ourselves to the
calculationof liquid-vapor equilibria for polymer solutions and the solubilitiesof polymer
mixtures. These calculations present an ideal example for the application of the methods
that we have seen until now. In particular, we shall apply the concepts of activity coefficients, group contributions, and equations of state.
A polymer is characterized by its chemical nature, first and foremost by its structural
pattern, A,whose repetition makes up the polymer chain A,,.If we have a copolymer, several patterns, A and B, for example, occur in random, alternating, or sequential fashion to
produce, respectively, the structures:

-A-A-B-B-B-A-B-B-A-A-A-A-B-A-A-B-A-B-A-B-A-B-A-B-A-B-A-B-A-A-A-A-A-A-A-B-B-B-B-B-B-BThe proportions of these patterns and their alternation are mainly determined by the
mode of synthesis.The same is true for branching, or the possible bridging between chains.
Let us note that the formation of chain structures may cause steric hindrances not found
in the primary patterns.
The nature of the terminal groups may greatly influence the chemical properties of a
polymer, but for the physical properties, this influence is generally weak due to dilution
of these groups within a chain of high molecular weight.
Most often a polymer is a mixture of components of similar structures,but chain lengths,
and therefore molar mass (Mi),varying within a wide range. We say that it is polydispersed.
Distribution (number of moles Ni, or mass wi,of these components as a function of molar
mass) may be determined by gel permeation (or exclusion chromatography).It is rare that we
proceed to such fractionation.Most often we simply use average molar mass by number:

E NiMi
(11.1)

376

11. Polymer Solutions and Alloys

determined by tonometry (analysis of terminal groups), and/or the average molar mass
by weight:

C. W

C. N~M,?

~ M ~

(11.2)
i

obtained by light diffusion or sedimentation rate measurements. Molar mass and dispersity, that is to say the range over which the degree of polymerization is extended, has a
great influence on some of the physical properties of polymers and their solutions, especially on solubility in the liquid phase and the formation of a homogeneous mixtures.
Figure 11.1shows the solubility limits for polystyrene in acetone. A polymer with molar
mass equal to 4800 is soluble in any proportions between (approximately) -50C and
190C. This temperature range narrows as molar mass of the polymer increases and
reaches 10300, for example. For these two polymers,we observe two critical solution temperatures.The lower critical solution temperature is higher than the upper critical solution
temperature. Finally, if the molar mass is around 19800, there is no longer a temperature
range in which solubility is total, but solubility reaches a maximum at around 60C.
Polymers may exist in the solid, or crystalline,state, in the liquid state, either molten or
in solution in a solvent, and finally in the vitreous state.The latter state is characterized by
a random dispersion of molecules like in the liquid phase, but by a higher packing density
(or small free volume), such that molecular movement and fluidity are greatly reduced.
The change from liquid to vitreous state is similar to the so-called second order phase transitions in which there is no discontinuity of properties such as density, Helmholtz energy or

10 300

140

J.
0 7

19 800

10 300

-507

4800
0.1
0.2
Polymer volume fraction

0.3

Figure 111 Influence of molar mass on polystyrene solubility in acetone


[Slow et al., 19721.

377

11. Polymer Solutions and Alloys

enthalpy, but of the derivative of these values with respect to temperature. However, we
must note that vitreous transition is observed throughout an appreciable temperature
range, which may be attributed to the polymer polydispersity. Furthermore, this temperature range depends on the chilling rate. As such, the vitreous state actually does not qualify as an equilibrium state. The vitreous transition temperature can be calculated by a
group contribution method [van Krevelen, 1990,p. 130-1501.It must be added that vitreous
and crystalline states may coexist in the form of a heterogeneous mixture, and that a polymer chain may be part of both the vitreous phase and crystalline spherulites. In a semicrystalline polymer, the crystallized fraction depends on the thermal history of the polymer, especially the crystallization kinetics.
In the following discussion, we consider only liquid or vitreous states. Their properties
may be represented using equations of state [Rodgers, 19931.The calculation of density as
a function of temperature and pressure is often performed using the Tait equation [see
Chapter 4, Section 4.4, and Danner and High, 19921.We may also proceed using additive
group contributions [Elbro et al., 19911.

11.1

POLYMERS IN SOLUTION

The examples that we provide logically extend the preceding chapters. By no means do
they cover the main points of the studies dedicated to polymer solutions. Many such studies exist due to the fundamental and practical interest in this field. As always, the use of
experimental data to support any calculation method is indispensable.We must mention in
this context, the database established by Hao et al. [1991].
We must point out that the liquid-vapor equilibria calculation is in fact based on the
prediction of the activity of the solvent, since the polymer does not pass into the vapor
phase (except in the case of supercritical extraction,which we will not discuss here). We
can show that the results (by calculation or experimentation) are little influenced by the
value of molar mass and its distribution. The same is not true for the calculation of solubilities in the liquid phase (liquid-liquid equilibria) that require an exact calculation of the
activities of both the solvent and the polymer.

11.1.1 The Flory-Huggins Model


From the calculation of different distribution possibilities of molecules in a lattice grid (see
Chapter 7, Section 7.3, Fig. 7.9) Hory [1942] and Huggins [1941,1942]showed that large
differences in size between molecules that make up a mixture determined an entropy
of mixing that was markedly different from that of the ideal solutions.The result is a socalled combinatorialcomponent in the expression for excess Gibbs energy and activity
coefficients:
E
g combinatorial = - Ts&binatorial

RTCxiIn

@.
Xi

*
= RTCxi
In

2).
2)

(7.9)

3 78

11. Polymer Solutions and Alloys

@.
@.
vr
= In 2 + 1- 2
x,wmbinatorial
= In Xi + 1- 2
Xi
V:

In

(7.10)

These terms are constantly negative, but are only one element in the calculation of deviations from ideality. By derivation with respect to temperature it appears that the resulting
enthalpy of mixing is zero. Such solutions are called athermal. In fact the interaction
energy between the components of the mixture is also accounted for using an enthalpic
term. The so-called Flory-Hugginsmodel is often applied:

g E = x1 In 0 1 + x2 In @2 + x(xl + p ~ ~ ) @ ~ @ ~
RT

X1

(7.14)

x2

in which p is the ratio of molar volumes of the polymer and the solvent,and parameter xis
related to the molecular interaction energies by the inter-exchange parameter w12:
(11.3)
In practice, this x parameter is determined from experimental data and assumed to be
composition independent. Application of this model has quickly revealed its deficiencies.
It can be approached using the theory of regular solutions (see Chapter 7, Section 7.5)
by stating:
V
x= 1
(6,-6,)2
RT

(11.4)

where vl is the molar volume of the solvent, and 6, and S,are the solubility parameters of
the solvent and the polymer. We know that this theory only applies to non-polar solutions,
but it has been extended by Hansen [1967] to polymer solutions containingpolar groups. It
may serve as a qualitative guide in the study of solvents that are compatible with a given
polymer. This author breaks the cohesion energy down into three terms corresponding to
the forces of dispersion,dipole-dipole interaction forces, and hydrogen bonds:

and thus replaces Equation 11.4 by:


(11.5)
The value of 6, is assumed to be equal to that of the hydrocarbon of the same structure
(homomorphic),and the values for SP and 6, are determined using solubility data.These
data have been tabulated [see van Krevelen, 1990, part VII, Table IV, for example]. In a
three-dimensionaldiagram, the compatibility of a polymer and a solvent can be estimated
by examining the proximity of their representative points. A simplified version of this
method [Bagley et al., 19711 limits representation to two parameters, 6, and 6h, where 6, is
defined by the equation:

In the 6h, 6, space, Figure 11.2 elucidates the grouping of polystyrene solvents around
the representative point of this polymer.

3 79

11. Polymer Solutions and Alloys

20

+
15

+
+ ++
++ *
++ +++

10

.
5

10

15

6V

Figure 11.2 Polystyrene solubility: application of the Bagley method


[1971] [van Krevelen, 1990,p. 2071.
:solvents; + : non-solvents.

11.1.2 The Influence of Free Volume


From the geometric data (bond angles and lengths),we can calculate the actual volume of
the molecules, also called the van der Waals volume, vuw. The values obtained are much
lower than those resulting from the density of the dense fluid phase because, even in liquid
or vitreous phase, a major fraction (from 30 to 50%) of the volume is free and available for
molecular motion. This fraction depends on the compound, and Figure 11.3 [Elbro et al.,
19911shows that the ordinary solvents,except water, on the one hand, and the polymers on
the other hand, are clearly distinguishable on the basis of this criterion.The polymers are
indisputably more compact.
Placing a polymer in solution therefore causes a large variation in free volume, that is to
say, in molecular interaction distances. It corresponds to an expansion of the polymer and
condensation of the solvent [Patterson, 19691. This phenomenon must be taken into
account if we wish to arrive at a realistic theory of solutions and account for certain exper-

3 80

11. Polymer Solutions and Alloys

0.6

1.4

1.8

Specific volume (cm3 9-1)

Figure 113 Proportion of free volume in solvents (0)and polymers (A)

[Elbroe t d . , 19901.

imental data. For polymer solutions, the general rule is that a rise in temperature favors
demixing, causing a lower critical solution temperature that is higher than the upper critical solution temperature (Fig. 11.1).We therefore attribute demixing at low temperature to
energy effects, while demixing at high temperature is interpreted in terms of free volume
variation. Similarly, it has been observed (Fig. 11.4) that an increase in pressure brought
about an increase in solubility.This may be interpreted by considering that the differences
between the free volume fractions in the solvent and the polymer are reduced by the
increase in pressure, since the solvent is more compressible.
Credit for the first works to account for this parameter is due to Prigogine [1957] and
Flory [1964,1965,1970].The equation of state proposed by Flory is written as:
(11.6)

11. Polymer Solutions and Alloys

381

200 -

J P(MPa)
150 -

%
E
F

100\

I
I

500

\
10

20

40

30

Polymer mass fraction

Figure 114 Influence of pressure on polystyrene solubility in acetone


[Zeman and Patterson, 19721.

In this equation the values for pressure, volume, and temperature are "reduced" using
the characteristicvalues Y, v*, and c:
(11.7)
The parameter c was introduced by Prigogine to account for the number of degrees of
freedom of the molecule, but in fact, its physical significance must be regarded with great
caution.
Note that this equation of state only concerns the liquid phase. Indeed, the compressibility factor does not approach 1 when density approaches zero.
Oishi and Prausnitz [1978] accounted for three elements that determine deviation from
ideality in a solution by writing the activity coefficient in the form:

In

x = In y c + In y,""

+ In y,P

(11.8)

where In xc is the combinatorialterm, derived from the Flory combinatorialterm as modis the free volume term
ified by Stavermann [1950] (see Chapter 7, Equation 7.13), In

xE"

3 82

11. Polymer Solutions and Alloys

xR

derived from the Flory equation of state, and finally In is a residual term representing
the interaction energies, identical to the one used in the UNIFAC model with regard to
expression and the numerical values for the parameters.Application of this model requires
knowledge of the molar volumes of the components and of the solution. This is because
the equation of state is applied for expressing the activity coefficient, and not for calculating molar volumes.
We shall provide three examples from among the recently proposed models. They take
into account each element of the non-ideality and further incorporate the notion of group
contributions. The first two are narrowly derived from the lattice model introduced
by Flory. The concept of free volume corresponds to an expansion of the lattice grid
(see Chapter 7, Section 7.3, Fig. 7.9). The third model is also derived from the lattice form,
where the grid elements, at constant volume, are occupied by either the segments of
the polymer (or solvent) chain, or by holes whose proportion varies with that of free
volume.

11.1.3 The Entropic-FV Model


This model [Elbro et al., 19901combines the combinatorial term and the free volume term
into a single expression.We define a free volume fraction:
(11.9)

as an expression in which vvwis the van der Waals volume, whose numerical values are
furnished by Bondi [1964,1968]and 2) the molar volume.
The activity coefficient is then:
(11.10)
The residual term is expressed either using the UNIQUAC model (see Chapter 7), and
the interaction parameters are then optimized from experimental data, or from the
UNIFAC model [Kontogeorgiset al., 19931.This method may be classified among the socalled activity coefficient methods because the values for the molar volumes of the components, the solvent, and the polymers must be known, as in the Oishi and Prausnitz
model.They may be calculated by the group contributions method proposed by Elbro et al.
[1991] for this purpose. The results obtained by applying the UNIFAC version of
the model are generally good. Figure 11.5 shows these results and allows for a comparison
of the calculated values and the experimental values for the activity coefficients in
weight (ai= ai/wi) at infinite dilution for solvents in polymers. The greatest deviations
involve chloride derivatives for which the contributions of the UNIFAC model may prove
unreliable.

383

11. Polymer Solutions and Alloys

O/

0
0

2.5

-1

0
0

Figure 115 Prediction of activity coefficients at infinite


dilution for solvents in polymers using the entropic-FV
method [Bogdanic,1995bl.

11.1.4 The GC-Flory Model


This model [Bogdanicet aZ., 1994,1995alhas the same structure as the Oishi and Prausnitz
model. The activity coefficient is therefore expressed as a sum of three terms:
In

= In y c

+ In y F V +

In y r

(11.8)

However, it is the original Flory expression that is applied to the calculation of the combinatorial term:

(7.10)
In addition, the free volume and attraction (or residual) terms are obtained from an
equation of state analogous to the one proposed by Flory, but modified such that the limit
at low density is respected.

3 84

11. Polymer Solutions and Alloys

In this equation of state, there is an attraction term Eattr:


1

p = -RT
--W

Zj

+C

Eatt,

(11.11)

G3-1

The resulting expression for the free volume activity coefficient is as follows:
In y T v = 3(1 + Ci)
In

q13

-1

(G1l3-1)

- Ciln

2).

f
w

(11.12)

Linear mixing laws are applied to parameters Ci and w:.


The attraction term Eattr
is inspired from the UNIFAC model, such that the residual
component becomes:

(11.14)

Application of the model requires that we know parameters qi, w 1*, Ci,and 5,;.The first
three are obtained by additive group contributions: if we say that vi is the number of
groups of type n in component i, and that 9
(
, and Qn are the volume and surface parameters of these groups and identical to the ones used in the UNIFAC method, then we have:
W: = 1.448-15.17

V; Q

(11.15)

(11.16)
n

The Ci parameters are temperature dependent:


(11.17)

The interaction energies between molecules i and j are calculated from interactions
between the groups rn and n, weighted by the surface fractions Om, 6,:
(11.18)

The molar volumes in liquid phase for each component and for the mixture are
obtained from the equation of state, in opposition to the models above. We may therefore

385

11. Polymer Solutions and Alloys

state that the GC-Flory model is an equation of state model, even if it is applied using the
activity coefficient expression. It is necessary, however, to mention that if the vapor pressure of the solvents is needed, it must be obtained by a different correlation (the Antoine
equation, for example).
Parameters C , , , C,,, C,S E,,,,, and A,,, are determined from an experimental database relating to compounds of low molar mass: thermal expansion in the liquid phase and
heats of vaporization for C , , , C,,, C i ,and liquid-vapor equilibria for em,,, and A&,,,, .
The GC-Flory model has been applied to the calculation of solvent activities in
homopolymer and copolymer solutions. The activity coefficients are calculated by regarding the polymer as a mixture of pseudocomponents (see Chapter 12). It has been shown
that the influence of the number of pseudocomponents, as well as the polydispersity of the
polymer with respect to the molar masses and the chemical composition on the calculation
of liquid-vapor equilibria, may be neglected. The results are good, as evidenced by the
comparison of calculated and experimental activity coefficients at infinite dilution for solvents in polymers (Fig. 11.6a) or copolymers (Fig. 11.6b).

-1

4
3 1
8%

-c

2l
1

0
0

3
'n

"&

Figure 1l.b Prediction of activity coefficients at infinite


dilution for solvents in polymers using the GC-Flory method
[Bogdanic et nl., 19941.

3 a6

11. Polymer Solutions and Alloys

/.. ..

83

c:
K
-

,6'

Figure 116b Prediction of activity coefficients at infinite


dilution for solvents in copolymers using the GC-Flory
method [Bogdanic et al., 1995al.

11.1.5 The GCLF Equation of State


(Group Contribution Lattice Fluid)
The equation of state applied by High and Danner [1990] to polymer solutions is derived
from the works of Sanchez and Lacombe [1978], from the Guggenheim lattice model
[1952],and from the model developed by Panaiyotou and Vera [1980,1982].Its conception
differs fundamentally from that of the preceding models, that are inspired by the van der
Waals theory. We shall introduce this method in more detail.
The components of the system are situated on a lattice with a degree of coordination z
( z = 10). Each cell of the lattice of invariant volume v,, (equal to 9.75 cm2 .mol-') is either
occupied by a segment belonging to one of the components, or empty, and therefore a gap.
The proportion of gaps depends on the total volume of the system. Each component is
characterized by a volume vt?and an interaction energy q i .
First we shall take a look at the model as applied to a pure substance, denoted by subscript i.

11. Polymer Solutions and Alloys

387

The number of segments ri occupied by the component is equal to the ratio of its characteristic volume vI: to the volume of one cell v h :
*
V .
r. = 2
(11.19)
vh

The two terminal segments are in contact with 2 (z - 1) neighboring cells and, taking
into account the chemical bonds, the other segments with (ri - 2) (z - 2) cells for a total of
(z - 2) ri + 2 cells, such that the surface parameter qi attributed to component i is defined
by the equation:
(11.20)
zqi = (z - 2)ri + 2
When temperature, pressure, and the number of moles of the component Ni are given,
the number of holes Nh is unknown.The total volume of the lattice is equal to:
(11.21)
or, for one mole of component:
(11.22)
The surface fraction (or possible contacts) 0, attributed to the component is equal to:
(11.23)
Finally we define the reduced coordinates using the equations:
(11.24)
The equation of state is therefore written as:

(11.25)
In order to apply this equation of state to a mixture, using the linear mixing laws,we first
define the surface parameters q , and the volume parameters, r and v*:

r = x x i r i ,v * = C x i v I ! ,q = &qi

(11.26)

and the surface fraction attributed to the components:


0=

Zei

(11.27)

The interaction energy between the two different components E& is calculated using the
geometric average, possibly corrected by an interaction parameter ki,j:
E. . =
51

1.1

(1 - k41. .)

(11.28)

388

11. Polymer Solutions and Alloys

The equilibrium constant of the quasi-chemical model is derived from this approach
[see Chapter 7, Section 4.2, and Guggenheim, 19521:

(11.29)
and rl,l:

and the local order parameters


r1,z =

1+

vl

2
- 48,e2(1-

(11.30)

K)

e, q 1+ e2r1,,
=1

where:
In these last equations,the

5,

(11.31)

parameter is calculated without considering the holes:

(11.32)
The energy parameter of the mixture is then obtained by the equation:

The reduced coordinates are then:

(11.34)
and we can solve the equation of state, which is identical to the one applied to pure substances:

(11.35)
The activity of one of the components of the mixture is calculated by applying this equation of state to the pure component, then to the mixture, under the same conditions of temperature T and pressure P
*

V.

lnai= -= l n a i + l n 4 + q i h
RT
V

+qi(

=)+(

- 8

y)lnrci (11.36)

The concept of group contributions has been applied to the calculation of the characteristic parameters v f and q j using equations:
(11.37)

11. Polymer Solutions and Alloys

and:

Ei,j,T=

c
k

@k,i@l,j

3 89

(11.38)

where vk,i and @k,i are, respectively, the number of groups k and the surface fraction of
group k in component i:
@k,i

L
@kiQk

(11.39)

@,iQ,

In this equation, Q k stands for the surface parameter of group k, and its value is the one
used in the UNIFAC method (see Chapter 7, Section 6).
The parameters of this model are temperature dependent.They have been determined at
300 and 400 K, either for components,v r , E ~q ~j ,or
, for groups, RtT,and E ~ ,from
~ , an experimental database made up of vapor pressures and densities of low molar mass compounds.
This method has been applied to the calculation of liquid-vapor equilibria for systems
of low molar mass compounds, as well as to polymer solutions.The results are generally
good. The molar volumes are calculated from the equation of state. The same may be true
for the vapor pressures, but in fact this value is obtained by applying a different equation
(the Antoine equation, for example).
However, for the moment, the GC-LF-EOSmodel remains limited to a relatively limited number of groups.

11.1.6 Extension to Liquid-Liquid Equilibria


In principle, the methods we have just described may be applied to the calculation of liquid-liquid equilibria.This has been done for the GC-Flory model [Saraiva et al., 19951,for
polymer solutions, and polymer-polymer mixtures. The various types of diagrams can be
depicted. However, prediction is only semi-quantitative due to the sensitivity of these
equilibria to the value and distribution of the molar masses, which are generally not well
known. Furthermore, we know that even for low molar mass compounds that are completely identified, the UNIFAC method may be applied to liquid-liquid equilibria only if
we use a special matrix of interaction parameters.

11.2

POLYMER MIXTURES

It is now common to formulate polymer mixtures (blends) with the purpose of manufacturing a material that combines the features, often complementary, of its components.
The miscible mixtures are made up of a homogeneous phase down to the molecular level.
However, very often the blends are heterogeneous, even if they have, macroscopically,a
homogeneous appearance. That does not mean that they are without practical interest.
Their properties depend on the nature of the dispersed phase and the continuous phase, as
well as on the surface tension (adhesion). Blends have been the subject of many studies.
For example,we may consult articles by Paul and Barlow [1980] and Paul [1986].

390

11. Polymer Solutions and Alloys

Miscibility is often determined by thermometric analysis.The appearance of two vitreous transition temperatures attests to the presence of two phases. A single transition is
interpreted as perfect miscibility.
We shall see that it is sometimes possible to establish the conditions of miscibility by
application of the Flory-Huggins theory. However, remember that the molar mass values
of the polymers (or copolymers) exert an undeniable influence on the behavior of these
mixtures [see Broseta and Fredickson, 1990, for example].
We can show that if we apply the Flory-Huggins theory to a mixture of two polymers
with degrees of polymerization p1 and p 2 respectively, and whose inter-exchange parameter is denoted by xm,the critical solution point then corresponds to the equation:
1
(11.40)
and the critical composition, expressed in volume fractions, is equal to:
@l,crit

=
6

(11.41)

6 2
+ 6 2

For high degrees of polymerization, we see that the value of xcritmis almost zero. The
miscibility condition is therefore:
Xm<

This inequality may be interpreted as an indication of affinity, necessary between the components of both polymers.
In the case of a mixture of two random copolymers A1,Bx, C1,Dy, the application of
the so-called mean field theory has been proposed [Kambour et al., 1983; ten Brinke
et al., 19831 and thus to break the xmparameter down into:
Xm

= (1 -x)(l

-Y)XA,C +

(1 - ~ I Y X A ,+Dx ( l - Y ) X ~ C
+xYxB,D - x ( l -x)xA,B - Y ( ~ - Y ) x ~ ,

(11.42)

The miscibility condition is found in the form of a space within the x , y plane delimited
by the taper of equation:
xA,Bx2 + x c , D y 2 + ( x A , c - xA,D

- xec + X,&Y

+ (xB,c-xA,B-xA,C)x

+ ( x A , D - x A , c - x c D ) ~ + (xA,C-xcritm)

=O

(11.43)

This method has been applied (among others) to mixtures of copolymers poly(phenylene oxide, sulfonated phenylene oxide) or poly(pheny1ene oxide) partially sulfonated
with copolymers poly(orthochlorostyrene,parafluorostyrene), by Vukovic et al. [1995].
Miscibility was determined by observation of the vitreous transition point. Prior studies
[Vukovic et al., 19951 of mixtures of poly(pheny1ene oxide) partially sulfonated with
poly(orthoch1orostyrene)or poly(parafluor0styrene) have allowed for the determination
of five of the xi,j parameters. The sixth parameter, relating to the parafluorostyrene,
orthochlorostyrene interaction has been fitted to the experimental data. These data, and
the result of the calculation, are depicted in Figure 11.7.The black circles represent immiscible mixtures, and the white circles are miscible mixtures.We see that orthochlorostyrene

391

11. Polymer Solutions and Alloys

is immiscible with poly(pheny1ene oxide) regardless of the degree of sulfonation, that


poly(pheny1ene oxide) is immiscible with all poly(orthochlorostyrene,parafluorostyrene)
polymers, including the homopolymers. However, a composition range exists for which
these two copolymers are miscible, and this range is limited by the ellipse of Equation 11.43whose parameters are listed inTable 11.1.

70

E 604,

E
8

a
v)

50-

a,

40 0

....
. .
. .
0

. .

--

.ac

.
0

00

0
0

00'

....d

Figure 117 Miscibility of copolymers poly(pheny1ene oxide, sulfonated


phenylene oxide) with copolymers poly(orthochlorostyrene, parafluorostyrene) [Vukovic et aL, 19951.
:immiscible mixtures; o :miscible mixtures.

Interaction
PPO, pFSt
PPO, OClSt
SPPO, pFSt
SPPO, OClSt
PPO, SPPO
oClSt, pFSt

Xi i
0.037 (a)
0.02 (a)
0.093 (a)
0.28 (a)
0.318 (a)
0.083 (b)

(a) Vukovic et al., 1993.(b) Vukovic et al., 1995.

392

11. Polymer Solutions and Alloys

CONCLUSION
Despite its considerable practical importance and fundamental value, the thermodynamics
of polymers, their solutions, and their mixtures is in its infancy. In this field, the states of
matter are poorly defined. The chain structure (or polymer lattice) is taken into account
only in a very empirical manner, while the properties of block polymers or alternating
polymers differ greatly. Nevertheless, the most important physicochemical phenomena
have been elucidated (combmatorial term, free volume term, molecular interactions,ranges
of molecular weight variation). The application of group contribution methods leads to predictive methods that are of a satisfactory precision for the calculation of liquid-vapor equilibria. Significant progress can be expected for the calculation of liquid-liquidequilibria.

REFERENCES
Saraiva A, Bogdanic G, Fredenslund Aa (1995) Revision of the Group-contribution Flory equation of
state for phase equilibria calculations in mixtures with polymers. 2. Prediction of liquid-liquid
equilibria for polymer solutions. Znd. Eng. Chem., 34,1835-1941.
Bagley EB, Nelson TP, Scigliano JM (1971) J. Paint. Technol.,43,35.
Bogdanic G, Fredenslund Aa (1994) Revision of the Group-contribution Flory equation of state for
phase equilibria calculations in mixtures with polymers. 1.Prediction of vapor-liquid equilibria for
polymers solutions. Znd. Eng. Chem. Res., 33,1331-1340.
Bogdanic G, Fredenslund Aa (1995a) Prediction of vapor-liquid equilibria for mixtures with copolymers. Znd. Eng. Chem. Res., 34,324-331.
Bogdanic G, Fredenslund Aa (1995b) Application of the group-contributions methods to polymer
solution thermodynamics. 1.Vapor-liquid equilibria. Kem. i Znd. 44 (lo), 415-427.
Bondi A (1964) van der Waals volume and radii. J. Phys. Chem., 68 (3), 441-451.
Bondi (1968) Physical Properties of Molecular Crystals, Liquids and Glasses. Wiley, New York.
Broseta D, Fredickson GH (1990) Phase equilibria of copolymerhomopolymer ternary blends: molecular weight effect. J. Chem. Phys. 93 (4), 2927-2938.
Danner R, High MS (1992) Polymer solution handbook. Design Institute for Physical property data,
AIChE. NSRDS and AIChE, The Pennsylvania State University,University Park.
Elbro HS, Fredenslund Aa, Rasmussen P (1990) A new simple equation for the prediction of solvent
activities in polymer solutions. Macromolecules,23,4707-4714.
Elbro HS, Fredenslund Aa, Rasmussen P (1991) Group contribution method for the prediction of
liquid densities as a function of temperature for solvents,oligomers and polymers. Znd. Eng. Chem.
Rex, 30,2576-2582.
Flory PJ (1942) Thermodynamics of high polymer solutions.J. Chem. Physics, 10,51-61.
Flory PJ, Orwell RA,Vrij A (1964) Statistical thermodynamics of chain molecules liquids. I. An equation of state for normal paraffin hydrocarbons. J. Am. Chem. SOC.,86,3507-3514.
Flory PJ, Orwell RA,Vrij A (1964) Statistical thermodynamics of chain molecules liquids. 11. Liquid
mixtures of normal paraffin hydrocarbons. J. Am. Chem. SOC.,86,3515-3520.
Flory PJ (1965) Statistical thermodynamics of liquid mixtures.J. Amer. Chem. Soc., 87 (9), 560-565.
Flory PJ (1970) Thermodynamics of polymer solutions. Discussions Faraday SOC.,49,7-29.

11. Polymer Solutions and Alloys

393

Guggenheim EA (1952) Mixtures. Clarendon press, Oxford.


Hansen (1967) J. Paint. Technol.,39,104 et 511.
Hao W, Elbro HS, Alessi P (1991) Polymer solution data collection. Dechema Chemistry Datu Series;
Dechema, Frankfurt.
High MS, Danner RP (1990) Application of the group contributions lattice-fluid EOS to polymer
solutions.AIChE J., 36,1625-1632.
Huggins ML (1941) J. Chem. Phys. 9,440.
Huggins ML (1942) Ann. N. YAcud. Sci. 43,l.
Kambour RP, Bendler JT, Bopp RC (1983) Phase behavior of polystyrene, poly(2,6-dimethyl-l,4phenylene oxide), and their brominated derivatives. Macromolecules 16,753-757.
Kontogeorgis GM, Fredenslund Aa, Tassios D (1993) Simple activity coefficient model for the prediction of solvent activity in polymer solutions. Ind. Eng. Chem. Res., 32,362-372.
Oishi T, Prausnitz JM (1978) Estimation of solvent activities in polymer solutions using a groupcontribution method. Ind. Eng. Chem. Process Des. Dev., 17,333.
Panayiotou C, Vera JH (1980) The quasi-chemical approach for nonrandomness in liquid mixtures.
Expressions for local surfaces and local compositions with an application to polymer solutions.
Fluid Phase Equilibria, 555-80.
Panayiotou C, Vera JH (1982) Statistical thermodynamics of r-fluids and their mixtures. Polymer J.,
14,681-694.
Patterson D (1969) Review. Free volume and polymer solubility. A qualitative view. Macromolecules
2,672.
Paul DR, Barlow JW (1980) Polymer blends (or alloys). J. Macromol. Sci.-Rev. Macromol. Chem.,
C18(1), 109-168.
Paul DR (1986) Polymer blends: phase behavior and property relationships. Advances in Chemistry
Series, 211; D.R. Paul and L.H. Sperling, Eds.
Prigogine I (1957) The Molecular Theory of Solutions. North Holland, Amsterdam.
Rodgers PA (1993) Pressure-volume-temperature relationships for polymeric liquids: a review of
equations of state and their characteristic parameters for 56 polymers. J. Appl. Polym. Sci., 48 (6),
1161-1180.
Sanchez IC, Lacombe RH (1978) Statistical thermodynamics of polymer solutions. Macromolecules,
11(6), 1145-1156.
Slow KS, Delmas G, Patterson D (1972) Cloud-point curves in polymer solutions with adjacent upper
and lower critical solution temperatures. Macromolecules, 5 (l), 29-34.
Staverman AJ (1950) The entropy of polymer solutions. Rec. Trav. Chim.,the Netherlands, 69,163.
ten Brinke G,Karasz FE, MacKnight WJ (1983) Phase behaviour in polymer blends: poly (2,6-dimethyl-1,4-phenylene oxide) and halogen-substituded styrene copolymers. Macro-molecules, 16,
1827-1832.
van Krevelen DW (1990) Properties of Polymers. Their Correlation with Chemical Structure; their
Numerical Estimation and Prediction from Additive Group Contributions. Third edition. Elsevier,
Amsterdam.
Vukovic R, Zuanic M, Bogdanic G, Kurusevic V, Karasz FE, MacKnight WJ (1993) Polymer, 34,1449.
Vukovic R, Bogdanic G, Erceg A, Fles D, Karasz FE, MacKnight WJ (1995) Copolymer blends of
phenylsulphonylated poly(2,6-dimethyl-l,4-phenyleneoxide and poly(p-fluorostyrene-co-p(o)chlorostyrene). ThermochimicaActa 2398,l-10.
Zeman L, Patterson D (1972) Pressure effect in polymer solution phase equilibria. 11. Systems showing upper and lower critical solution temperatures. J. Phys. Chem., 76 (8), 1214-1219.

12

Multicomponent Mixtures

The calculation methods for the thermodynamic properties that we have introduced are
based on the knowledge of the composition of the system being investigated, namely a list
of components, and the quantity (number of moles) for each. In addition, it is necessary
that these components have a limited number, otherwise the calculation (and memory)
limitations of our computer systems will be exceeded.
In fact, petroleum fluids, crude oil, natural gas, and even petroleum fractions (gasoline,
kerosene,etc.) do not conform to this condition.As evidence of the number of components
present in a crude oil, we need look no further than the number of identified hydrocarbons
and of badly identified cuts in a straight run gasoline gas chromatography analysis
(Table 12.1),or to consult the detailed analysis for this gasoline (see Appendix 5 ) . In a natural gas, the analysis unambiguously identifies the light components nitrogen, carbon
dioxide,methane, and C+C4 hydrocarbons.Yet we know that heavier components are
also present with carbon atoms reaching several dozens. They are poorly identified, but
even in low proportion, we cannot neglect them. They are functionally responsible for the
retrograde condensationphenomenon and for the very high values of the critical pressure.

Table 12.1
Analysis of a straight-run gasoline by gas chromatography;identified compounds
and chromatographic cuts sorted by number of carbon atoms and chemical family.
See Appendix 5 for a more detailed analysis.

396

12. Multicomponent Mixtures

Due to its practical importance and fundamental value, the problem of petroleum fluid
characterization has been the subject of several studies. The topic is too broad for the
scope of this text. We refer the reader to works by Rojey et al. [1994], Gravier [1986], and
Pedersen et al. [1989],among others.
The polymers are another example. They are actually mixtures of componentsA,,A,, ...
A,, etc. that correspond to n degrees of polymerization, most often varying across wide
ranges. We refer to them as polydispersed. The distribution of these mixtures exerts an
undeniable influence on their solubility in ordinary solvents or their mutual solubility (see
Chapter 11).
How do we solve this problem? Two methods are proposed.
The first is in common use, and simplifies the composition of the real mixture by lumping
similar, individual chemicals into a single pseudocomponent according to their physicochemical properties. Such simplificationis sometimes extreme so as to leave only four or five
pseudocomponents.The calculation methods discussed in previous chapters are then applied.
The second method, less often used, substitutes a continuous distribution function for
the elementary list of components (with their proportions). Its use implies a revision of the
calculation methods that we have presented. It is particularly well adapted to polymers for
which this distribution is generally consistent,and determined by the polymerization reaction mechanism.
We shall examine the concept and the salient features of these two approaches. We also
note that each one of these approaches has many variations, as a witness to their shortcomings.

12.1

PSEUDOCOMPONENTS

12.1.1 Complex Mixture Analysis


As an example,we use a petroleum fluid, crude oil, or natural gas. We shall first analyze the
fluid.
In order to characterize a crude oil, it is W e d . The column and the reflux ratio are chosen such that the separation of the components of the mixture are as good as possible,if not
perfect [Roussel and Boulet, 1994a;Paradowski,19941.By graphing temperature as a function
of distilled amount, we obtain the TBP(True Boiling Point) curve. In case the fluid contains
a limited number of components,the graph would show a step function,each step corresponding to the boiling temperature of the successivecomponentsthat are collected at the head of
the column. For a multicomponentmixture, we obtain a continuous curve (fig. 12.1). It
allows us to define the fractions that will be considered as pseudocomponents. For each fraction, density, and possibly molecular weight, are determined. These fractions correspond to
lumps of a large number of components according to their boiling temperature.
However, very often more detailed analyses are used.
The gas chromatography technique achieves component separation according to
decreasing volatility. Calibration allows for the comparison of retention time and boiling

397

12. Multicomponent Mixtures

Temperature

Fraction 1

I
I

I
I
I

I
I
I

1 - 1

Components
defined
as pure
substances I

I
I
I
I

I
I

Distilled

mass or volume

Figure 12.1 TBP distillation and crude oil partitioning into fractions
defined by their average boiling temperature [Paradowski,19941.

temperature, and this analysis method is often called simulated TBP.A large number of
components are thus separated, identified, and quantified [Durand et al., 1989; Roussel
and Boulet, 1994bl.We may then group them.
Beyond a certain point the components can no longer be separated. We may then turn
to gel permeation chromatography methods that separate components according to
molecular volume. In order to specify the chemical nature (saturated hydrocarbons, aromatics) we use liquid chromatography. Nuclear magnetic resonance is used to establish
molecular structure.
In this way an abundant amount of analytic data is available,overabundant in fact compared to our calculation possibilities.
The most common analytic methods express the proportions of the components (or
pseudocomponents) in mass fraction. The organization of the thermodynamic methods
requires conversion to mole fractions, and therefore molar mass information. For identified components, this creates no problems. For cuts defined by average boiling temperature and density, we can apply correlations that use these data, such as the ones proposed

398

12. MulticomponentMixtures

by Riazi and discussed in Chapter 3, Section 3.4.2 [see also Paradowski, 19941.Finally, for
the condensed natural gas fractions or crude oils we may proceed with direct measurement
by cryoscopy or tonometry. From these data and the molar masses of the fractions defined
by analysis of the fluid, we can calculate the molar mass of the last fraction using a mass
balance [Durand et al., 19891.This last value hence has a very large uncertainty.This uncertainty must be taken into account in the characterization of the heavy fractions, that has
therefore an appreciable degree of freedom.

12.1.2 Lumping
Lumping may result directly from the TBP curve as shown in Figure 12.1, or taken from
the chromatographic analysis.There are many ways in which to proceed.
Figures 12.2a, b, and c [Ruffier-Meray,1990, Ruffier-Meray et al., 19901 show various
ways of lumping the hydrocarbons eluted by gas chromatography between n-nonane and
n-decane. This elution time interval corresponds, approximately, to the range of boiling
temperatures.
The roughest lump (Fig. 12.2a) combines all of these hydrocarbons into a single pseudocomponent.
However, Figure 12.2b shows that lumping according to the number of carbon atoms
may differ from the previous 1ump.Thisdifference is due to the fact that ramifications may

Figure l2A Chromatographic analysis of a hydrocarbon fraction of


n-nonane to n-decane.Elution by order of decreasing volatility.

399

12. Multicomponent Mixtures

Figure l2.2b Hydrocarbon fraction from n-nonane to n-decane.


Partitioning according to number of carbon atoms.

n-c,
Paraffins
Naphtenes
Aromatic rings

oc,

]clo

-I

F
Figure l2& Hydrocarbon fraction from n-nonane to n-decane.
Partitioning of C,, hydrocarbons according to chemical family.

400

12. Multicornponent Mixtures

affect the boiling temperature such that in the given range of elution time hydrocarbons
with both 9 and 10 carbon atoms appear.
Finally, we can take the chemical family into account (paraffins, naphthenes, aromatic
rings), as shown in Figure 12.2~.
The number of pseudocomponents is found to be considerably increased.
The dynamic clouds method
A lumping algorithm has been proposed by Monte1 and Gouel[1984]. It is based on the
method of dynamic clouds. We shall introduce the iterative steps of this method.
(a) The number of desired pseudocomponents is fixed. We divide the mixture into that
number of classes in a relatively arbitrary fashion. For each class we define a center, the
most abundant component, for example.
(b) The distancesbetween each component of the mixture and the centers are calculated based on metrics defined by certain thermodynamic properties (critical points,
boiling temperature, parameter values for an equation of state, etc.).
(c) Each component is then tied into the center closest to it. New classes, equal in number to the old classes but not necessarily having the same components, are as such defined.
(d) For each class we define a new center by weighting the thermodynamic properties of the components belonging to this class.
(e) The calculation is repeated starting from (b) until no component changes class.The
centers then define the pseudocomponents.

12.1.3 Thermodynamic Properties of Pseudocomponents


The application of the calculation methods for thermodynamic properties assumes that a
certain number of properties are known, such as the critical point, the acentric factor, the
boiling temperature at atmospheric pressure, or even the parameters of the equation of
state to be applied. With the exception of the lightest components of a petroleum fluid,
these methods are applied to pseudocomponents defined by a lumping technique. It is
therefore required to determine these properties by weighting the characteristic values of
the real components identified and lumped within the pseudocomponent. For the critical
properties, for example,we apply the equations defined relative to pseudocritical properties
(see Chapter 8, Section 8.11), most often limiting ourselves to linear weighting [Kay, 19641.
When the calculation of thermodynamic properties is based on the application of an
equation of state derived from the van der Waals theory (most often the case), a logical
method has been proposed by Carrier et al. [1989],Neau et al. [1993], and Leibovici [1993].
The pseudocomponent is defined such that the calculation results are unchanged for the
group when we have a homogeneous phase. The calculation obviously depends on the mixing law selected. If we apply the equations defined in Chapter 4 to the calculation of pure
substance and pseudocomponent parameters:

R~T:
a(T) = Q, pc

(4.43)

12. Multicomponent Mixtures

401

(4.37)

where a (T,) is a function of the reduced temperature, such as the one proposed by Soave
(Eq. 4.46), for example.
If we then apply the classic mixing laws:
a=

aUzizj
i

b=

(8.27)

cbizi

(8.28)

ai,j= V G ( 1 - ki,j)

with kJ.1. .= kY. .

(8.29)

then, if we say that TGmand PGmare the pseudocritical coordinates of the pseudocomponent we may state for any temperature [Leibovici,19931:

If the temperature is equal to the pseudocritical temperature of the pseudocomponent,


function a, is equal to one, and we obtain:

Similarly, the mixing law relating to covolume yields:


(12.3)

In the preceding equations, the sums concern the components lumped within the
pseudocomponent,and zi denotes the mole fraction of component i in the pseudocomponent (not in the mixture).The a,functions depend on the component not by their form, but
through their parameters (for example,the acentric factor, if we apply Equation 4.46).
The pseudocritical coordinates may thus be obtained by solving Equations 12.2 and
12.3. The value of function a, at any temperature T may then itself be obtained at any
temperature by application of Equation 12.1. If we wish to apply Equation 4.46 to the
pseudocomponent,then an acentric factor wmmay be attributed to the pseudocomponent
from a list of values for a, that covers a given temperature range.
The interaction parameters for the pseudocomponents are determined by application
of the same principle (invariance of the properties for the single-phase system).
For cuts whose chemical composition is unknown, we can apply correlations based on
the few available properties (density,average boiling temperature) such as those described
in Chapter 3, Section 3.4 [see also Paradowski, 19941. Alternatively, we may identify those
cuts to one or more defined hydrocarbons, whose properties are known or may be calculated using group contributions (Chapter 3, Section 3.3.2), since the latter are defined by
analysis,specifically nuclear magnetic resonance.

402

12.1.4

12. MulticomponentMixtures

Representing the Heavy Fraction of Natural Gases

Compared to the detailed analysis, lumping results is an undeniable loss of information.


However, it is necessary due to the large number of components that is unacceptable with
our calculation tools. Furthermore, as advanced as our analysis methods may be, the heaviest fractions of petroleum fluids remain poorly identified. As a result, we are left with
some degrees of freedom that may be put to our advantage by adjusting the calculation of
phase equilibria for a petroleum fluid to fit the experimental data. This flexibility is very
useful for a natural gas where retrograde condensation is particularly sensitive to the
nature of the heaviest components despite their low concentrations.
We may therefore envision the representation of this heavy fraction using a pseudocomponent whose critical properties (T,,P,) and acentric factor w are adjusted on the
curve representing the liquid volume deposit as a function of pressure. However, we risk
ending up with an incoherent T,, P,, o triplet considering the usual relationships
between these values and molecular weight.
It has been proposed [Nabec, 1989; Barreau et al., 19911 (and it seems infinitely more
reasonable) to consider the heavy fraction as a mixture containing a paraffin hydrocarbon,
and one or two alkylaromatic hydrocarbons whose properties must remain compatible
with the density of this fraction, and whose properties are calculated using the correlations
established by Robinson and Peng [1978]. We shall use the example of a natural gas whose
composition is provided by Table 12.2.
Figure 12.3 shows two adjustment steps: in 12.3a, the heavy fraction is combined with a
linear paraffin hydrocarbon in C,,, C,,, or C,,: in 12.3b,the optimal solution is defined by
representing this fraction by a mixture of n-tridecane, and two alkylbenzenes with the formulas C19H32and %,Ha, and mole fractions equal to 0.03707,0.00433 and 0.00025 respectively.We observe that we thus obtain an excellent correlation of the retrograde condensation data. However, it may appear that density data require resorting to polynuclear aromatic pseudocomponents.
Table 12.2
Composition of the natural gas whose retrograde condensation curve is
represented by Figure 12.3. Molar mass of heavy fraction CI1+:193g/mole
Component
Nitrogen
Carbon Dioxide
Methane
Ethane
Propane
n-Butane
n-Pentane
n-Hexane
n-Heptane
n-Octane
n-Nonane
n-Decane
c11+

Mass Fraction

Mole Fraction

0.108 17
0.550 46
0.131 40
0.028 924
0.026 235
0.034 095
0.024 471
0.029 229
0.004 178 6
0.005 398 7
0.007 488 0
0.008 307
0.041 642

0.138 90
0.449 90
0.294 60
0.034 6
0.021 4
0.021 1
0.012 2
0.012 2
0.001 5
0.001 7
0.002 1
0.002 1
0.007 7

12. Multicomponent Mixtures

403

a.

Pressure (MPa)

b.

1.
A
\

I
5

10
Pressure (MPa)

e\

20

Figure 12.3 Correlation of the retrograde condensation of a natural


gas (curves are calculated, points are experimental).Representation
of the heavy fraction by: a paraffin pseudocomponent (graph (a),
first step), or three pseudocomponents,one paraffin, and two alkylaromatics (graph (b), final step).

404

12.2
12.2.1

12. MulticomponentMixtures

CONTINUOUS THERMODYNAMICS
Definition

If the chemical components of a complex mixture are continuously distributed as a function of a property Z,such as boiling temperature, the number of carbon atoms, or molecular weight, we may then define a distribution function F(Z) such that the product of F(Z) AZ
represents the mole fraction (or mass fraction) of the species whose property Z is between
Z - AZ/2 and Z + AZ/2. Figure 12.4 thus illustrates the discontinuous representation by a
finite number of pseudocomponentson the one hand, and on the other the continuous representation by the distribution curve.
Function F(Z) must be selected such that it best represents the reality such as it is known
by analysis of the complex mixture. In this way it may be obtained by derivation of the
TBPcurve. It is also desirable that the function lends itself to mathematical treatment.
We shall give two examples. Of course, the function must respect the relationship:

jI F ( Z ) dz= 1
which is substituted for the more classical equation for mole fractions:

Figure 12.4 Representationof a complex mixture by a distribution


function or by a finite number of pseudocomponents.

(12.4)

12. Multicomponent Mixtures

405

This distributionfunctionwill be used within the expressionfor any extensiveproperty.For


example,for the volume of a single-phasesystem,if N, is the total number of moles,we have:
V =N, 21(T,8z1,zz,...z,)
for a discreet mixture, and:
V =Nt V ( T , e F ( Z ) )

(12.5)

for a continuous mixture, where 21 is a function of temperature, pressure, and the distribution variable Z,and, most often, of the distribution function F. 21 is therefore an equation of
state explicit in volume.
If we take the mixture to be ideal (see Chapter 5, Section 5.5), then the equations:

v =N,cziv;(T,P)
and:

F ( I ) v* (T,P,Z)dZ

(12.6)

correspond to each other, as it is understood that v; ( T P )is the molar volume of pure component i in the reference state, and v*( TRZ) is the function that links the molar volume to
the distribution variable.
This last example shows that application of the concept of continuous thermodynamics
requires as essential conditions: selection of a distribution variable Z and its associated
function F(Z) on the one hand, and on the other, the expression of what we have called a
characteristic function (equation of state, Helmholtz energy expression, Gibbs energy,
etc.), which depends on two variables ( T and T and V) as with simple mixtures, but also
on the distribution variable Z and the distribution function that replaces composition.

r:

12.2.2 Chemical Potential, Fugacity Coefficient, and


Equilibrium Condition Between Phases
Applying the distribution functions to the calculation of thermodynamic properties has
been rigorously reviewed by Salacuse and Stell [1982]. Here we shall provide only a few
essential equations.
Like volume, the Gibbs energy of a homogeneous phase is expressed by an equation
such as:
G = N, G(TW(I))
(12.7)
and the chemical potential is defined by the equation:
(12.8)
We may apply a model based on the definition of excess quantities (see Chapter 5), and
in this case we state that:

p [ T W , W ) l =p * ( T W ) + RTln F ( I ) + RTln y[TW,F(I)l

(12.9)

406

12. MulticomponentMixtures

where the activity coefficients yare most often expressed by derivation of an excess Gibbs
energy model:
(12.10)
If we apply an equation of state explicit for pressure, then the fugacity coefficient is
expressed by the equation [Cottermann et al., 19851:

which is similar to Equation 8.31.


The liquid-vapor equilibrium condition is expressed as before by the equality of chemical potentials:

or, in terms of fugacities:

P F W qL[I;ez,FL(I)] = P F V ( 0 qV[I;eZ,FV(Z)l

(12.13)

These equations link the distribution functions of the two phases, as the first of the following examples shows,borrowed from Ratzsch and Kehlen [1983].

12.2.3 Application Examples


12.2.3.1 liquid-Vapor Equilibrium in an Ideal Solution
We have a liquid phase (defined by its distribution function), assumed ideal, characterized
by a Gaussian distribution as a function of the boiling temperature at atmospheric pressure, denoted by To:
(12.14)
The maximum of this distribution has an abscissa T:, and its standard deviation is oL.
We shall determine the distribution function of the vapor phase at equilibrium as well as
the pressure, at given temperature T:
If it were a discreet mixture, the equilibrium equation at low pressure would be:
Yi - P p
xi P

If Pa (I; To) is the vapor pressure of the fluid element whose boiling temperature at
atmospheric pressure is To,we may write:

12. Multicomponent Mixtures

407

(12.15)
and:

P = / TO P(T,To)FL(To)dTo

therefore:

(12.16)

(12.17)

In order to have an expression for the function Pa (?: To),we can apply the ClausiusClapeyron equation:
(2.10)
in the form:

[(

31

(T, To) = Po exp A 1--

where Po is the atmospheric pressure and we arrive at the following result [Ratzsch and
Kehlen, 19831:
(12.18)

where:

(12.19)

As for the bubble pressure, it is expressed by the equation:

(12.20)
It is important to note that the distribution functions of the two phases in equilibrium
are the same, and their parameters are related by simple equations.
On the other hand, if we look at a mixture with distribution function F(Z) in a state of
partial vaporization, the material balance that must be respected between any element
characterized by the distribution variable interval dZ is:
N,F(Z) dZ = N L F L(Z)dZ + NvFv(Z) dZ
N ~ N'=
+ N,

(12.21)

It appears that functions Fv(Z) and FL(Z)do not have the same form as F(Z). It seems we
can get around this problem by adopting orthonormal function developmentsas the distribution function,with the number of terms depending on the complexity of the distribution
and the desired precision [Halpin and Quicke, 19901.

408
12.2.3.2

12. MulticomponentMixtures

Excess Gibbs Energy of a Polymer Solution


in SemicontinuousThermodynamics

We have another example using polymer solutionsin a solvent [Hillmann,19851.Given the


molar mass of the solvent and the average molar mass of the polymer (in number), we can
calculate the solvent mole fraction x , and the polymer mole fraction X = 1- x , from the
mass ratios.The method applied is called semicontinuous.The polymer is represented by
a distribution function and the variable is the degree of polymerization (or molar mass) for
example, and the solvent remains a distinct,defined chemical species.
The model applied to the calculation of excess Gibbs energy is that of Flory-Huggins
(Chapter 7, Section 7.3). If @standsfor mole fractions:
gE
@l
@2
- = x1 In - + x2 In - + x(xl
RT
X1
x2

(7.14)

we have an equation that in this case may be written as:

gE
RT

- =xlln

Wl

-+ X
W

1
2

+ - w {@*

jI A(l,Z)@(Z)dZ +

A(Z,Z)@(Z)@(Z)dZ dZ
I

(12.22)

where A(1,Z) and A(Z,Z) represent the interaction parameters between the solvent and
the polymer with degree of polymerization Z,and two polymers with degrees of polymerization Z and Z,respectively.
The molar volume of the solution w depends on the distribution variable. It is calculated
according to:

w = xlw; + X

F(Z)w*(Z) dZ

(12.23)

I
1

What needs to be given is the distribution function and the relationships between the
degrees of polymerization Z on the one hand, and the molar volumes w*(Z) and the interaction parameters A(1,Z) and A(Z,Z) on the other. We may then calculate the activity coefficients using equations:
(5.75)
for the solvent and, for the polymer:
(12.10)
This approach lends itself particularly well to the study of the influence of the degree of
dispersion on polymer solubility.
It is quite clear that the semicontinuous method may be generalized to cases where
several components are distinct,as in the following example.

12. Multicornponent Mixtures

409

Retrograde Condensation of a Natural Gas


Application of the Soave-Redlich-Kwong
Equation of State

12.2.3.3

Semicontinuous thermodynamics has been applied by Cottermann and Prausnitz [1985]


to calculations of retrograde condensation of natural gases. The lightest components
(C,C,)
are identified,and the C,, fraction is represented by the Schultz [1940] distribution function F(M),where M is proportional to molar mass:

F(M)=

(1 - p

- 1

exp

PuNw

(-7)

(12.24)

In this equation, Tis the gamma function, and yis the distribution origin. The average
M , and the variance oare given by expressions:

Mo=ap+y

02=ap2

and

IIIyo points may be

emphasized in this example.


First, the calculation of retrograde condensation runs into the problem we have already
pointed out. The material balances (see Equation 12.21) impose a relationship between
any material element of the global mixture on the one hand and the same element in the
liquid and vapor phases on the other. As a result, these two phases may not have the same
distribution function as the global mixture. The solution that has been proposed
[Cottermann and Prausnitz, 19951, called quadrature, results in defining pseudocomponents. According to the authors, this definition is less arbitrary and embodies an appreciable economy of calculation time.
Secondly, since the thermodynamic model is the Soave-Redlich-Kwong equation of
state (see Chapter 4, Section 4.2.1), we need to define a calculation method for the parameters a (M,M), and b (M) that corresponds to the material element characterized by value
M of the distribution variable. The following assumptions are made:
b(M) = b,

+ blM

G(KZj=

+ al(T)M

(12.25)
(12.26)

where b, and b, are constants, and a, and a1 are polynomial functions of temperature.
As for the binary terms (see Chapter 8, Section 8.3.1), a(i,M) and a(M,M), the usual
geometric average is applied.The interaction parameter k ( i , M )is independent of M and M.
If we denote as X the mole fraction of the continuous fraction (X = 1- xi),then the
1
classic mixing laws are written as follows:
(12.27)
1

a=

ccxixiai,j+ 2X
i

xi
i

F(M)a (i, M) dM
I

+X 2

I,,

F(M)F(M)a(M, M) dM dM

(12.28)

41 0

12. MulticomponentMixtures

Figure 12.5 shows the results of the calculations.The analysis provides a discreet representation of the global mixture, for which the distribution function is substituted (solid
line). The calculation of partial vaporization generates the composition of the phases in
equilibrium, and we note that the distribution of the liquid phase is indisputably different
from that of the global mixture, and more complex.

30

20

10

30
c

0
.-c
S

g
20
+
v)

10

60

100

150

200

250

300

Molar mass

Figure 125 Retrograde condensationof a natural gas;distribution curves


for the vapor phase and the liquid phase obtained by retrograde condensation (T = 367 K,P = 13.9 MPa) [Cottermann and Prausnitz, 19851.

12. Multicomponent Mixtures

41 1

CONCLUSION
Whether we prefer to illustrate a complex mixture using a finite number of pseudocomponents or by a distribution function, a very large number of alternatives are available.
First we have the analytic procedure: gas chromatography can be more or less detailed,
and be applied within variable limits, with the heavy fraction starting at Cll+ for some,
and up to %,+ for others.
Then, for a discreetrepresentation, we need to fix the number of pseudocomponents,
which is not the same depending on whether our aim is to simulate the exploitation of a
natural gas reservoir, the operation of surface separators, or laboratory measurements.
Finally, the lumping method and the calculation method for pseudocomponent properties
vary according to the authors, as we have seen.
With continuous thermodynamics,it is required to select the distribution variable, and
we observe that in order to better respect physical reality, we may be led to define two
variables: boiling temperature and aromaticity, for example, in a petroleum fluid. This
choice is necessary for copolymers whose properties depend on the distribution of molar
mass and of chemical composition.The distribution function must then be defined, and its
parameters determined as a function of available data: TBP distillation curve, chromatographic analysis of petroleum fluids, average molar mass in number, in weight, or even of a
higher order for a polymer, etc.
In many cases, the only analytic data available is insufficient for satisfactory prediction
of all thermodynamic properties,and experimentalphase equilibrium data must contribute
to the model. Such is the case for light oils, characterized by a high vaporized fraction as
soon as the pressure drops below the bubble pressure and that of condensate gases.
These numerous alternatives can be found in the literature. For examples, we may consult the works mentioned at the beginning of this chapter. Furthermore, the problems
resulting from the mixture complexity in the case of petroleum fluids are not limited to the
calculation of thermodynamic properties [Montel,19931.

REFERENCES
Barreau A, Braunschweig B, Emami E, Behar E (1991) A knowledge based system for the automation
of thermodynamic models adjustment. Paper presented at CAIPEP (Conference on Artificial
Intelligence in Petroleum Exploration and Production) 91, College Station,Texas.
Carrier B (1989) Modelisation des coupes lourdes des fluides petroliers. Sc.D.Thesis, Marseille.
Carrier B, Rogalski M, PBneloux A (1989) Choice of pseudocomponents for flash calculations of
petroleum fluids in: C: fraction characterization, Chorn L.G. and Mansoori G.A. Taylor and
Francis, Eds. New York, 123-136.
Cotterman R, Bender R, Prausnitz JM (1985a) Phase equilibria for mixtures containing very many
components. Development and application of continuous thermodynamics for chemical process
design. Ind. Eng. Chem. Process Des. Dev.,24,194-203.
Cotterman RL,Prausnitz JM (1985b) Flash calculations for continuous or semi continuous mixtures
using an equation of state.Ind. Eng. Chem. Process Design & Development,24,434-443.

41 2

12. MulticomponentMixtures

Durand JP, Fafet A, Barreau A (1989) Direct and automatic GC analysis for molecular weight
determination and distribution in crude oils and condensates up to C& J. High Resolution
Chromatography,12 (4), 230-233.
Gravier JF(1986) Cows de production. 2. Propriktks des fluides de gisement. Editions Technip, Paris.
Halpin TPJ, Quicke N (1990) A new method for continuous thermodynamics applied in an equation
of state. SPE Reservoir Engineering 5 (4), 617-622.
Hillmann BTh (1985) Continuous thermodynamics of fluid mixtures. Ph. D. Thesis, Georgia Institute
of Technology. Ed. University microfilms international, Ann Arbor, Michigan.
Kay W B (1964) Density of hydrocarbon gases and vapors at high temperature and pressures. Znd.
Eng. Chem.,28,1014-1019.
Leibovici CF (1993) A consistent procedure for the estimation of properties associated to lumped
systems. Fluid Phase Equilibria,84,l-8.
Montel F, Gouel PL (1984) SOC.Pet. Engineers, 13119, 59th Annual Technical Conference and
Exhibition,Houston, Texas.
Montel F (1993) Phase equilibria needs for petroleum exploration and production industry. Fluid
Phase Equilibria,84,343-367.
Nabec R (1989) Contribution au dkveloppement dun systkme expert de sklection et dajustement de
mod6les thermodynamiques appliques aux fluides de gisement. Mkmoire industriel ESIEA (&ole
supkrieure dinformatique,dklectronique et dautomatique), Paris.
Neau E, Jaubert J-N, Rogalski M (1993) Characterization of heavy oils. Znd. Eng. Chem. Res.,32,11961203.
Paradowski H (1994) Mkthodes de calcul des propriktks physiques des hydrocarbures. In: Pktrole
brut, produits pktroliers, schkmas de fabrication.Wauquier JP, Ed., Editions Technip,Paris.
Pedersen KS, Fredenslund Aa, Thomassen P (1989) Properties of Oils and Natural Gases. Gulf
Publishing Company, Houston, Texas.
Ratzsch MT, Kehlen H (1983) Continuous thermodynamics of complex mixtures. Fluid Phase
Equilibria,14,225-234.
Robinson DB, Peng DY (1978) The characterization of heptanes and heavier fractions for the GPA
Peng-Robinson programs. GPA (Gas ProductorsAssociation) Research Report RR28.
Rojey A, Dur,and B, Jaffret C, Jullian S, Valais M (1994) Le gaz naturel. Production, traitement,
transport. Editions Technip, Paris.
Roussel J-C1, Boulet R (1994a) Composition des pktroles bruts et des produits pktroliers. In: Le
Rafinage du Pktrole. Vol. 1. Pktrole brut, produits pktroliers, schkmas de fabrication.Wauquier JP,
Ed., fiditions Technip, Paris.
Roussel J-C1, Boulet R (1994b) Fractionnement et analyse Blkmentaire des pktroles bruts et des
coupes pktrolikres. In: Le Rafi,nage du Pktrole. Vol. 1. Pktrole brut, produits pktroliers, schkmas de
fabrication.Wauquier JP, Ed., Editions Technip, Paris.
Ruffier-Meray V (1990) Reduction optimale des donn6es analytiques nkcessaires B la caracterisation
thermodynamique des gaz naturels. Sc. D. Thesis CNAM, Paris.
Ruffier-Meray V, Barreau A, Bkhar E (1990) Optimal Reduction of the Analytical Data Necessaryfor
the Thermodynamic Characterizationof Natural Gases.
Salacuse JJ, Stell G (1982) Polydisperse systems: statistical thermodynamics, with applications to
several models including hard and permeable spheres. J. Chem. Phys. 77 (7), 3714-3725.
Schultz GVZ (1940) Phys. Chem., B47,155.

13

Chemical Reactions

We shall touch only briefly upon the details of chemical transformations. Our discussion
takes a more condensed look at the chapters dedicated to this subject in an earlier work
[Vidal, 19741. Otherwise, the fundamental principles are presented in most general texts.
Chemical reactions are represented by a more or less complex stoichiometry that generally has little to do with the fundamental steps involved in the mechanism of the reaction. They may be the result of a combination of several reactions. For example, hydrocracking of 5 moles of n-octane forms:
propane
2 moles
n-butane
3 moles
isobutane
3 moles
n-pentane
0.5 mole
isopentane 1.5 mole
The overall process may be represented by:

5 n-CSH1, + 5 H2 + 2 C,H, + 3 n-C,Hl, + 3i-C4H,O + 0,5n-C5H,2 + 1,5n-C5H,,


which is the result of a number of simple reactions, each one affected by an appropriate
weighting factor:

Component Reaction

Weighting

2
3
3
1.5

The purpose is to represent the stoichiometry using the simple expression:

vlA, + v2A2+ ... + v; A ; + vi A; + ...

(13.la)

with the stoichiometric coefficients v, ,v2 etc. attributed to the reagentsA,, A, etc. of the
left-hand side of the expression, and v;, vi etc. attributed to the products A;, A; etc. in
the right-hand side of the expression.

414

13. Chemical Reactions

We also use the more condensed formula:

EviAi= 0

(13.lb)

where the stoichiometric coefficients of the products are positive, and those of the
reagents are negative.
For the remainder of this chapter we shall use both conventions side by side as each has
its merits. The corresponding equations will be notated using letters a or b.
The stoichiometry respects the material balance relative to the elements, and allows us
to establish the balance for the products and reagents by taking into account the extent of
the reaction. We may calculate the variation in the number of moles resulting from the
reaction:
A v = [v; + V; + ...I - [vI + V, + ...I
(13.2a)
or more simply by using the convention applied in Equation 13.lb:

A v = Evi

(13.2b)

As we shall see, it also provides the enthalpic balance, that is to say the heat of reaction.
However, this calculation is only meaningful as long as the enthalpy of each component is
expressed according to a reference state that leaves the reaction unaffected, namely the elements. It is also in terms of this origin that we must evaluate the Gibbs energy of the system,
and find its minimum value in order to know the composition of the system at equilibrium.
The values that permit us to calculate the enthalpy and the Gibbs energy of a compound
in relation to the elements are the standard enthalpy of formation and the standard Gibbs
energy of formation.They are the fundamental thermochemical data.
In order to determine the heat of reaction and proceed to the thermal balance of a reactor or a reactive section, we must also take into account the physical state of the system,
depending on the temperature and pressure conditions. Both the heat capacities at constant pressure and the residual quantities that we have already discussed are part of this
calculation. Similarly,it is possible to calculate the Gibbs energy of the mixture under reaction conditions and the chemical potentials for each component using fugacity coefficients
or activities.
These calculations make use of the principles and methods presented in the previous
chapters.

13.1

THERMOCHEMICAL DATA

13.1.1 Standard Enthalpy of Formation


Standard Gibbs Energy of Formation
The following data are available for a large number of compounds in the standard state
and at temperature T:
0 standard enthalpy of formation (Ah;)T

13. Chemical Reactions

41 5

standard Gibbs energy of formation (Ag;),


standard entropy S;
0 standard heat capacity (c;),;
by standard state is meant the pure substance in the ideal gas state (except for carbon in its
graphite structure, and sulfur in its rhombic structure), and at a pressure equal to 0.1 MPa.
The temperature is often equal to 298.15 K. Appendix 1lists such data.
Standard enthalpy (or Gibbs energy) of formation is equal to the variation of enthalpy
(or Gibbs energy) that results from the formation of a compound from its elements, where
the compounds and elements are at the same temperature T, and in the standard state. For
elements, the standard state is the stable molecular form (diatomic hydrogen, etc.). Their
standard enthalpies of formation are zero.
Standard entropy is calculated by application of the third law of thermodynamics. It is
zero for any pure, solid, crystalline substance at absolute zero.
llvo comments are in order regarding these values:
0 They are not independent. Using the values for standard entropies,SF,we can calculate the standard variation of entropy resulting from the formation reaction, (As;),.
For a compound with a basic formula of CVIHV2OV3NV4,
for example,we have:
0
0

(As;),= G(cv1Hv20v3N"4)

The standard enthalpy of formation, the standard Gibbs energy of formation, and the
standard entropy of formation are related by the equation:

For the standard enthalpies or Gibbs energies of formation,if we use only the data for
a single temperature To,the values at another temperature T may be calculated using
the variation of heat capacity that is related to the formation reaction. They are calculated using equations that are analogous to those to be presented later for any reaction. However, it is preferable to calculate the standard enthalpy for the reaction in
question at this temperature To,and then calculate its variation as a function of temperature (see Section 13.1.4.2).
The values of the heat capacities and standard entropy can be calculated using ideal
gas statistical thermodynamics. This, however, is not the case for the standard
enthalpy of formation. Its value is obtained from experimental measurements, most
often using heats of combustion. For example, for propane we may state:
(u;)298.15(c3H8)

= 3. (Ls.h,0)298.15
(c>+

(u%%3.15
(H2) - (u,")298.15(C3H8)

(13.5)

where Ah," are the combustion enthalpies that are equal to -393.51 kJ.mo1-l for
carbon, -285.83 kJ mol-I for hydrogen, and -2,219.9 kJ amo1-l for propane. Taking
into account the uncertainty (even when low in relative value) of the calorimetric
measurements for heats of combustion, we see that the standard enthalpies of

41 6

13. Chemical Reactions

formation themselves have a significant uncertainty (approximately 15 kJ*g-l for


the hydrocarbons).
The same is therefore true for the Gibbs energies of formation, which are related to
the enthalpies of formation by Equation 13.4. Finally, this uncertainty influences the
constant for the law of mass action (see Section 13.3.2, Eq. 13.29). An error of 600 J in
Gibbs energy of reaction results in a relative error of approximately 20% for this constant.
This situation is, however, sometimes improved thanks to other heats of reaction measurements, as well as direct determinations of the equilibrium constant and its dependence on
temperature.

13.1.2 Application of Group Contribution Methods


With regard to molar heat capacity at constant pressure, and the molar entropy of the ideal
gas, we introduced (Chapter 3, Section 3.3.1) prediction methods using group contributions. The very nature of these properties that are a direct consequence from the internal
kinetic energies of the molecule provides the justification for these methods. These energies include translation and rotation motions of all or part of the molecule,and interatomic
bond vibrations.
As for heats of reaction, some (approximate) constants are well known:
0 paraffin dehydrogenation:125 kJ .mol-
0 paraffin cracking: 75 kJ * mol-
0 paraffin hydrocracking:-50 kJ mol-l
0 naphthene aromatisation:210 kJ-mol-l
n-paraffin isomerization:-8 kJ.mol-l.
Similarly,observation of regularity in homologue series has, in a first stage, allowed the
introduction of the concept of heat of bonding [Pauling, 19601. This concept was later
developed using group contributions and, in the same manner as the data provided in
Chapter 3, Section 3.3.1, we find below some of the values proposed by Benson el al. [1969]
for the calculation of standard enthalpies of formation (Table 13.1), and by Van Krevelen
and Chermin [1951] then Chermin [1961] for the calculations of standard Gibbs energies
of formation (Table 13.2).They are found by using the expressions:
Ah;=

2 NiA(Ahfi) + 2i NjA(Ahij)

(13.6)

in which the i subscripts refer to the constituent groups of the molecules and the j subscripts to the corrections for structure, and:
Agf= A

+ BT + RT In Q

(13.7)

where parameters A and B are obtained by group contributions and orepresents the number of external symmetry of the molecule (see Chapter 3, Section 3.3.1,Table 3.8).
Use of these tables shall be made clear by applying them to orthoethyltoluene.

41 7

13. Chemical Reactions

Table l3.1
Group contributions for the calculation of standard enthalpy of formation.
The standard state is: ideal gas, P = 0.1 MPa, temperature 298.15 K [Benson,19691

Groups

A [Ahf(kcal)]
- 10.08

-4.95
-1.90
0.5
3.3
5.51
-10.08
-4.86
-0.98
0.57

Table 13.2
Group contributions for the calculation of standard Gibbs energy of formation.
The standard state is: ideal gas, Po= 0.1 MPa [Chermin, 19611.Units: cal . mol-
Groups
CH3
CHZ
CH
C
Branches:
3 contiguous CH groups
contiguous CH and C groups
C P
C;p
Substitutions:
ortho
meta
para
3 contiguous CH groups
contiguous CH and C groups

300-600 K

600-1 500 K

-10833 + 21.76 T
-5283 + 24.43 T
-756 + 29.42 T
3 060 + 36.36 T

-12 393 + 24.36 T


-5 913 + 25.48 T
-756 + 29.42 T
3 840 + 35.06 T

2 312
1625

2312
1625

3 100 + 6.1 T
5 280 + 9.94 T

2 536 + 7.04 T
5 634 + 9.35 T

955 + 0.55 T
352 - 0.57 T
-183 + 1.05 T
2 312
1625

1687 - 0.67 T
574 - 0.94 T
615 - 0.28 T
2 312
1625

(1)Aromatic carbon bound to an aromatic carbon and an aliphatic carbon (toluene).

EXAMPLE 13.1

Calculation of the enthalpy of formation and Gibbs energy


of formation of orthoethyltoluene
First we shall use the Benson method [1969] by looking at the group decomposition
from Chapter 3, Section 3.3.1, Example 3.3 for the calculation of molar heat capacity
and entropy. Table 13.3summarizes the calculations.

41 8

13. Chemical Reactions

Groups

A [Ah; (kcal)]

3.3
5.51
-10.08
-4.86
-10.08
0.57

2
1
1

1
1

For standard entropy at 298.15 K, we find s; = 94.85 cal mo1-l.K-l or 397 J.mo1-'.
We may now calculate the entropy of formation by applying Equation 13.3and using
the values for the entropy of carbon (5.69 J.mo1-') and hydrogen (130.6 Jamol-')
from Appendix 1:

= 397 - [9.5.69 + 6430.61 = -438 J.mol-'-K-'

We can then calculate (Eq. 13.4) the Gibbs energy of formation at 298.15 K:
(Ag;), = (Ah;), - T(As;)~= -962 + 298,15-438 = 129627 J-mol-'
If we now apply the van Krevelen and Chermin method, we obtain the expression for
Gibbs energy of formation as a function of temperature. The corresponding contributions are shown in Table 13.4:

Groups
CH3
CH2
CarH
Car
Substitutions:
Ortho

300-600 K

-10833 + 21.76 T
-5283 + 24.43 T

2
1

3 100 + 6.1 T
5 280 + 9.94 T

955 + 0.55 T

for a total Ag; = -3034 + 112.78 T i n cal.mol-' or -12694 + 471.9 T i n J-mol-' and,
therefore, at 298.15 K: 128 kJ.mol-'.
This result, as well as the one obtained using the Benson method (129.6 kJ-mol-') are
close to the value given in the databases: 131.8 kJ-mol-' .

13. Chemical Reactions

41 9

The discrepancy may seem important. However, it is must be noted that when we calculate enthalpy or Gibbs energy of a reaction (see next section), most of the reagent
and product constituent groups will be unchanged, such that there will be error compensation. Thus, the equilibrium between the three forms of ethyltoluene are calculated by a simple difference in the contributions in the ortho, meta, and para substitutions. The resulting precision is excellent.

13.1.3 Coherent Enthalpy Data


Up until now, we had considered that enthalpy (or internal energy) of a pure substance
could be related to an arbitrary origin. As a matter of fact, as long as the system undergoing transformations preserves its components, the values attributed to this origin are eliminated when we proceed to the application of the first law for calculating an enthalpy (or
internal energy) difference.
This is no longer true if a chemical reaction is involved. However, we can use coherent
enthalpy data by using a temperature To (298.15 K, for example) as an origin and, for the
value of enthalpy, the standard enthalpy of formation at this temperature. Therefore, for a
pure substance, Equation 2.54 is written:

h*(T, P ) = (Ah;)ro +

c: d T + (h*- h#),,

(13.8)

TO

and, if we limit ourselves to standard enthalpies:

(13.9)
This is the same as choosing the elements at temperature To as the origin.The values for
enthalpy calculated in this way may be applied to any energy balance, whether the transformation involves a chemical reaction or not.
For Gibbs energy, we state:
g;l = h;l - T s ; ~= (Ah;)To +j:c;dT-Ts;l

(13.10)

where entropy is referenced to absolute zero, as required by the third principle.

13.1.4

Standard Enthalpy and Gibbs Energy of Reaction

13.1.4.1 Definition and Calculation from Standard Enthalpies


and Cibbs Energies of Formation
Standard enthalpy (or Gibbs energy) of a reaction at temperature T, AH: (or AG:) is
equal to the enthalpy (or Gibbs energy) change resulting from the total transformation of
the reagents into products according to the specified stoichiometry.The reagents and prod-

42 0

13. Chemical Reactions

ucts are in their standard state, that is to say pure, in the ideal gas state (with the exception
of carbon and sulfur) at a pressure of 0.1 MPa and at temperature T. Neither residual
properties nor mixing quantities have anything at all to do with these values.
As we shall see, these values form the basis of energy balance and equilibrium calculation. They are rather intermediate calculation values than thermochemicaldata.
They are obtained very simply from the standard enthalpies (or Gibbs energies) of formation. Very succinctly,standard enthalpy (or Gibbs energy) of reaction is equal to the difference between standard enthalpy (or Gibbs energy) of formation of the products, and
the standard enthalpy (or Gibbs energy) of formation of the reagents. More precisely, if
the reaction is described by the stoichiometry in 13.la, then the enthalpy of reaction is calculated according to the equation:

AH;= [ v ; ( A ~ ; ) ~ ( +Av;(Ah&(A;)
;)
+ ...I
- [ V l ( q T ( A l )+ v 2 ( q M 4 2 ) + .-I

(13.11a)

Similarly,for the Gibbs energy of reaction:


A%=

[v;(&;)T(A;) + v;(&;)T(A;) + ***I


-[vl(Ag;)T(Al)+ v2(&;),(-42) + -*I

(13.12a)

or if we prefer to apply the formula in 13.lb:


MOT
= Cvi<a;)AAi>

and:

AG =

vi (Ag;) T (Ail

(13.11b)
(13.12b)

Finally, if we make use of coherent enthalpy or Gibbs energy data (Eqs. 13.9 and
13.10), the preceding expressions are easily transposed. All we need to do is replace the
enthalpies (or Gibbs energies) of formation with the standard, coherent enthalpies (or
Gibbs energies).For example, Equations 13.11b and 13.12b become:

AH; = c v i h $ ( A i )

(13.13b)

Xvig;(Ai)

(13.14b)

AG ;
=

13.1.4.2

Dependence of Enthalpy of Reaction and

Cibbs Energy of Reaction on Temperature


We apply the general equation:
AG;=AH;-TAS;

(13.15)

and separately calculate enthalpy variation, and then the standard entropy of reaction with
temperature.
This last value is calculated from either standard entropies SOT, or entropies of formation
using the equations:

AS.,= [v;s;(A;)+ v;s;(A;) + ...I - [V1S$(Al)+ v ~ s ; ( A+~...I


)
= [v;(As;)),(A;)
+ v;(As;)~(A;)
+ ...I

-[Vl(AS;)T(Al)+ v2(As;)T(A2)+ -.I

(13.16a)

42 1

13. Chemical Reactions

or:

AS; =

V,S;(AJ =

CV~(AS?)T
(AJ

(13.16b)

according to whether we apply stoichiometry 13.la or 13.lb.


Let us say that AC; is the difference in heat capacity at constant pressure between the
reagents and the products, calculated using the given stoichiometry:
ACF = [v;c;(A;)

+ v;c;(A;) + ...I

or:

- [vlc;(Al) + vzc;(Az) + ...I

(13.17a)
(13.17b)

ACF = ZvicF(Ai)

Under these conditions,we can write:


T

ACFdT

AH;=AH%+I

(13.18)

TO

and:

(13.19)

We end up with the equation:


AG;l= AHF0 - T ASF0 +

ACF dT - T

dT

(13.20)

TO

TO

which has no approximation. Its application depends specifically on the expression that
we chose to represent the heat capacities and their dependence on temperature (see
Chapter 1, Section 1.7.2).
In many cases, the AC; term is low. Enthalpy and entropy of reaction vary little with
temperature, and the Gibbs energy of reaction may be expressed by the equation:
AG;=AH;o-

TASF0

(13.21)

We see that we may thus, as an initial approximation, calculate the standard Gibbs
as a function of temperature, if known, and at any specific temperenergy of reaction AG ;
ature To, its value AG F0, as well as the value for enthalpy of reaction AH; Similarly, a
0
linear expression for Gibbs energy of reaction, such as the one we end up with using van
Krevelen and Chermin group contributions (Eq. 13.7, Table 13.2), implicitly contains the
standard enthalpy of reaction data.

13.2
13.2.1

HEAT OF REACTION AND ENERGY BALANCE


Heat of Reaction

The term heat of reaction deserves to be clarified if we wish to give it a physical meaning. Let us consider a chemical process described by one of the stoichiometries:
vlAl + vZA,+ ... + v;A;

+ v;A; + ...

(13.la)

422

13. Chemical Reactions

or:

cviAi= 0

(13.lb)

The elementary changes in number of moles (Ni) of the reagents and products are
related by the equations:
dNi
dNi
---dN1 - ... =--- - ... - ... = ... = d{
(13.22a)
Vl

v;

Vf

dii
=dc

or more simply:

(13.22b)

Vi

where { is the extent of reaction.


If the reaction takes place in a homogeneous phase, the enthalpy change of the reacting
medium is expressed, at constant temperature and pressure, as a function of the partial
molar enthalpies, and of the extent of reaction, taking into account the previous equations:

dH = chidNi = [(vih;+ ... + vfhi+ ...) - (vlhl + ... + vihi + ...)] d{

(13.23a)

d H = x h i dNi = [xvihi]d{

(13.23b)

or:

These equations may be written in the more global form:


dH=AHdc

(13.24)

where AH represents the partial derivative of enthalpy at constant temperature and pressure with respect to the extent of reaction, or the enthalpy change resulting from the reaction according to its stoichiometry (extent equal to one), under the reacting medium conditions (temperature, pressure, composition). It is good to define heat of reaction in this way

AH = ( v ;h1 + ... + ~ f h+i ...) - (v1h1+ ...+ vi hi + ...)

(13.25a)

or (according to the convention chosen for stoichiometry):

AH= Z V i h i

(13.25b)

This expression thus contains a contribution resulting from the state of the medium
(residual terms, enthalpy of mixing), as well as a chemical contribution.Both contributions
must be taken into account in the calculation using Equations 13.25a or 13.25b.
If the enthalpies of the components are coherent in the pure state (Eq. 13.8), and
accounting for the partial molar enthalpies of mixing h,?, we may then state:

hi(T,P ) = (Ah&o

+ ck. dT + (h: -

= h5i + (h: -

TO

+ h,?

+ h,? = hYi + h,?

(13.26)

After substitution into Equations 13.25a or 13.25b, and taking Equation 13.13b into
account,we obtain:
AH = AH; + {[v;(h;- hi) + ...I - [v1(h;- h#,)+ ...I)

+ [ ( v ; h y +...)-( v l h p + ...)I

(13.27a)

13. Chemical Reactions

or:

42 3

(13.27b)

This expression shows the respective roles of the chemical term AH;,the residual
terms, and the partial molar enthalpies of mixing in the heat of reaction, such as defined by
Equation 13.24.
We may also combine the chemical term with the residual terms by stating:

AH=AH*,+[(v;hy
or:
where:
or:

+ ...)-( V J y + ...)I

AH = AH*,
+

viE?

(13.28a)
(13.28b)

AH*,=AH;+ ([v;(h;-h;#)+...I- [v,(h; -h#,)+ ...I}

(13.29a)

AH*,=AH;+ x v i ( h * - h # )

(13.29b)

In these last equations,AH*,


is the enthalpy change resulting from the complete transformation of the reagents into products in accordance with the stoichiometry.The reagents
and products are pure, under the same conditions of temperature and pressure as the
reacting medium, and in the same physical state. If, for example, the reaction is studied in
the liquid phase, the residual enthalpies are then close to the enthalpies of condensation.
provides a good approxThe contribution of the heats of mixing is often neglected. AH*,
imation for the heat of reaction as it takes place in the reacting medium.
If the reacting medium is heterogeneous, but essentially each of the reagents or products is found and remains within one of the constituent phases, it is more convenient to
select reference states that are close to reality. For example,consider a reaction that occurs
in a two-phase liquid-vapor environment,such as the hydrogenation of benzene at atmospheric pressure and at ordinary temperature. We consider that both in the reacting
medium as in the reference state, hydrogen is in the vapor phase and benzene and cyclohexane are in the liquid phase. The same is true for hydration of propylene in isopropyl
the reference states for each compound must be specialcohol. When we calculate AH*,,
fied. Similarly,heat of combustion data (thermal power) makes sense only if we specify the
states for the hydrocarbon and for the combustion water (liquid or gaseous).
If some reagents and products undergo a change of phase, then to any variation in the
extent of reaction d{, corresponds an elementary variation in phase equilibrium and in the
components distribution. Equations 13.23 must then be replaced by an expression that
takes this equilibrium into account. For a system in the liquid-vapor equilibrium state, for
example, if NL and iVv are the number of moles in liquid and vapor phase, and hL and hV
are the molar enthalpies of these phases, at constant temperature and pressure we can
write:
(13.30)
.Each of the terms in this equation must be calculated using the methods for calculating
phase equilibria and the properties of homogeneous phases discussed in previous chapters.

424

13.2.2

13. Chemical Reactions

Energy Balance of a Reactor or a Reacting Section

In practice, it often occurs that the reactor feed (reagents,recycle, etc.) is broken down into
several flows of matter characterized by different temperature and pressure conditions,
and by different physical states. We may also include a separation vessel, a unit of product
fractionation, within the energy balance. In any event, for an open system at steady-state,
we apply Equation 1.16
H2-H,=Q+W

or

AH=Q+W

(1.16)

to express the first law. As the W term is generally zero, energy balance is reduced to the
calculation of the difference between the enthalpy of the effluent H2 and the enthalpy of
the feed H,. For either one we take the sum of the enthalpies for the various flows of matter. If these enthalpies are calculated relative to the elements,that is to say that we take the
enthalpies of formation into account, by applying Equations 13.8 or 13.9, there is no need
to account for the enthalpies of reaction, which are implicit within the coherent data.
I t on the other hand, as is often the case, the physical(temperature and pressure variation, physical state, enthalpies of mixing) and chemical (enthalpy of reaction) terms are
separated in the calculation, we then substitute to the actual transformation an open cycle
that:
0 Brings each of the flows of matter that make up the feed to a reference state selected
as a function of the available thermochemical data, such as the standard state at temperature To. If we say that H , is the enthalpy of the feed calculated with respect to
this reference state, the enthalpy change is equal to -Hl. It incorporates sensitive
enthalpies (resulting from temperature variations), latent heats for phase changes,
heats of mixing, etc.;
0 Transforms, in this state, the reagents into products, taking into account the stoichiometry and the extent of reaction.At this point, the reaction enthalpy calculated in
the assumed reference state comes into play, or AH:o for an extent of reaction equal
to one (and if we choose the standard state as the reference state);
Brings the products to the effluent conditions.We say that H2 is the enthalpy of the
effluent calculated with respect to the reference state. Like H,, it incorporates sensitive enthalpies (temperature variations), latent heats for phase changes, heats of mixing, etc.
Throughout the entire cycle, the total enthalpy change (for an extent of reaction equal
to one, and if we choose the standard state as the reference state) is equal to:
AH = -HI

+ AH;^ + H~

(13.31)

Since the calculation respects both the start and the end of the real process, we thus
obtain the energy balance.
Often the reactor or the reacting section is adiabatic. Enthalpy variation is therefore
zero, and the unknown is, for example, the state, the composition, and the temperature of
the effluent.Flame temperatures are calculated in this fashion. Similarly,in catalytic distillation an isenthalpic balance is established around each stage, including the heat and
extent of reaction. For these two examples and in a very general way, we must take chemical equilibria into account in order to calculate the extent of reaction.

42 5

13. Chemical Reactions

13.3

CHEMICAL EQUILIBRIA

The extent of a reaction may be limited by kinetic factors. We may therefore eliminate
some transformations from among all those that a system may undergo, because without a
catalyst or initiator, in effect these transformations do not take place. Except in this case,
we shall not discuss the kinetic limitation any more than we did the mass transfer kinetics during the study of phase changes.
On the other hand, we must consider that, a priori,any reaction is in equilibrium, and
therefore evaluate the limits of its extent imposed by thermodynamics.To accomplish this,
we account for the stoichiometry and apply the second law in the form:
dGTP= 0

(1.36)

13.3.1 The Equilibrium Condition


Let us consider a reaction described by the general stoichiometry:
v1A1 + v2A,+ ... + v;A;+ viA,+

...

(13.la)

The elementary variations in the amounts of reagents and products (Ni, number of
moles) are related by the equations:

--dN1 - ... - dNi


Vl
vi

dN;
dN;
- ... - -- ... - -=
v;

V;

... = d t

(13.22a)

where 5 is the extent of reaction.


The variation in Gibbs energy of the reacting medium at constant temperature and
pressure is expressed as a function of chemical potentials and of the extent:
dG =&

dNi = [(v; p; + ... + vlp;

+ ...) - (vlpl + ... + vipi + ...)I

d t (13.32a)

The equation may be written in the more global form:


dG = AG d5
where:

AG=(v;&+

...+ v ; , u ~ +...)-( ~1/.~1+


,..+ ~ i p i +...)

or:

AG =

vipi

(13.33)
(13.34a)
(13.34b)

The evolution of the reacting medium occurs only in the sense of a decrease in Gibbs
energy. The sign of dg is therefore identical to that for AG. At equilibrium, Gibbs energy is
minimal for any transformation performed at constant temperature and pressure, and
therefore:
AG=O

and

(13.35)

42 6

13. Chemical Reactions

13.3.2 The law of Mass Action


The value of AG depends on both the physical conditions (temperature, pressure, composition) and the variation of Gibbs energy resulting from the chemical transformation.
These two factors appear if we express the chemical potential relative to its standard
state:
RTln

f;l = p i - p r
I?

(5.28)

and the equilibrium condition yields:

The right hand side of the equation is the change of Gibbs energy caused by the reaction being performed under standard conditions,with a minus sign. It is the standard Gibbs
energy of reaction AG;, and the equation is most often expressed as:

(-)

f{

v i ...

(5)
?..

f;

(13.37)

...

(-)

f;

or:

...
v:

...

f;

= K

...
where:

(%Ti..
(13.38)

...

RTln K = -AG;

(13.39)

The values of the reagent and product fugacities in the reacting mixture fl!,fi,depend on
composition, pressure, and physical state conditions. Their standard state values, $O, &,
depend on the choice made for that state. Generally this state is the ideal gas state at pressure Po= 0.1 MPa. They therefore have the value:$O = & = PO = 0.1 MPa.
We make note of a special case where one of the phases within the system is dense and
contains only one component (most often a solid phase). Under these conditions, taking
the standard state for this component to be the same solid state at a pressure of 0.1 MPa,
as pressure has only negligible influence on fugacity in this phase, the fugacities and &
are equal, and their ratio is dropped in the first part of Equation 13.38.
The right hand side of the equilibrium condition involves the constant KO, which
itself depends on the reaction considered, its stoichiometric coefficients through Equations 13.39 and 13.12, and on the choice of the standard state.

42 7

13. Chemical Reactions

To calculate the fugacities within the mixture, and according to the physical state, the
pressure, and non-ideality,we can apply any of the various formulas and the models associated with them: activity coefficients, fugacity coefficients, equations of state, etc.We shall
not review these alternatives as they were introduced in Chapters 5,7, and 8. We may
choose from a number of approximations,for example:
0 In the vapor phase, we can assume that the system behaves as an ideal gas, thereby
simplifying the law of mass action such that the fugacity ratios are replaced by the
partial pressures of the components.
0 For a mixture in the liquid phase, assume ideality. The equilibrium may then be
expressed in terms of mole fractions and the equilibrium constant K O is replaced by a
constant K*. The latter is calculated from the change of Gibbs energy AG* using an
equation similar to expression 13.39, but where the reference states for the reagents
and products are selected as the pure substances in the liquid phase.
0 One may wish to express the law of mass action, using mole fractions, or using the
concentrations in moles per liter, etc.
Various expressions result from these approximations.We have mentioned them previously [Vidal, 19741. We shall present a few examples.

EXAMPLE 13.2

Synthesis of methyl tertiary butyl ether


Methyl tertiary butyl ether (MTBE), used as an anti-knock agent in gasoline, is
obtained from the reaction of methanol and isobutene,in liquid phase, and in contact
with acidic resins.
CH3
CH3OH + CHZ=C,

,CH3

CH,

I
I

t
,CH3-O-C-CH3
CH3

The equilibrium constant K O may be calculated as a function of the standard enthalpies


of formation,which are listed in Appendix 1. If we wish to perform the calculations at
25"C, for example,and determine the equilibrium in terms of mole fractions in the liquid phase, we must take the non-ideality of this phase into account, which is evident
due to the large differences in polarity of the components.For methanol and MTBE
whose vapor pressures are low under these conditions,we can state:
fi = ppyixi
where Pp,*I;. ,and xiare, respectively, the vapor pressures, activity coefficients,and the
mole fractions of these two components.
Since the vapor pressure of isobutene is higher, it is desirable to account for the fugacity coefficient at saturation:
f." = p ?(pay&
1
1 1 1
and the Poynting correction is neglected.

42 8

13. Chemical Reactions

Under these conditions, and assigning subscripts 1,2, and 3 to isobutene, methanol,
and MTBE respectively, the law of mass action may be written in the form:

where P" denotes standard state pressure (0.1 MPa).


Constant K* depends on temperature only. It corresponds to the Gibbs energy of
reaction when we use the pure components in liquid phase as the reference state. The
left hand side of the equation contains the activities. Finally, if we isolate the mole
fractions, we find:
-x3
- - K * -Yl Y2
x1 x2

Y3

but the ratio on the left hand side depends on the temperature (like K*) and compositionsvia the activity coefficients.These must therefore be determined by correlation
of experimental liquid-vapor equilibrium data and application of one of the models
described previously (NRTL or UNIQUAC, for example, Chapter 7, Section 7.6.2 and
7.6.3), or predicted by a method such as UNIFAC (Chapter 7, Section 7.7.2). This has
been done by several authors [Delion et al., 1986;Oost et al., 19951for this or related
systems.

EXAMPLE 13.3

Hydration ofpropylene
The hydration of propylene yields isopropanol.Under ordinary conditions of temperature and pressure, water and alcohol are essentially in liquid phase, and propylene in
vapor phase. The reacting medium will naturally be characterized by the mole fractions of the first components in the liquid phase and by the partial pressure of propylene. The fugacities will be calculated using the equation:

for water and alcohol, and for propylene, accounting for pressure, we write:

where P is the total pressure. Subscripts 1,2, and 3 are for water, propylene, and isopropanol, and the law of mass action is written in the form:

where, if we include the mole fractions in liquid and vapor phase:

13. Chemical Reactions

42 9

This calculation assumes that the liquid-vapor equilibrium is established, and we note
that in this case it is irrelevant whether the fugacity is calculated in either of the two
phases, as they are equal.

D EXAMPLE 13.4
Synthesis of methanol
Methanol is obtained from the reaction of a mixture containing hydrogen, carbon
monoxide, and carbon dioxide. The conversions have been calculated by Chang et al.
at temperatures ranging from 200 to 30O0C,and at pressures from 5 to 35 MPa. Under
these conditions, the mixture (in vapor state) cannot be considered as an ideal gas,
and the fugacities are calculated from the fugacity coefficients:

such that the law of mass action related to the reaction:

CO + 2H,

+ CH,OH

where the subscripts 1,2, and 3 denote carbon monoxide, hydrogen, and methanol, is
written:

or:
The mole fraction ratio in the first part therefore depends on temperature, pressure,
and the composition itself via the fugacity coefficients.
These are calculated by application of the Soave-Redlich-Kwongor Peng-Robinson
equation of state (see Chapters 4 and 8 ) to the mixture of H,, CO, CO,, CH,OH, and
H,O. Here we have an extension of the range of application for these equations, but
we have no model that applies to polar mixtures under pressure. The calculation
above assumes that there is no liquid phase formation. If a liquid dropout is observed,
the liquid-vapor equilibrium must be taken into account.

13.3.3 The laws of Equilibrium Displacement


The extent of a reaction may be modified by varying the conditions of temperature, pressure, or by the addition of a compound (reagent, product, or inert component).The equilibrium condition (13.35) remains valid such that we can state:
dAG = 0

(13.40)

430

13. Chemical Reactions

regardless of the variations of temperature dT, pressure dP, extent of reaction de, or
external material input dN,. We therefore state:
aAG
aAG
aAG
aAG
dAG= -d T + -d P + - d t + x - d N i e = O
ap
aT
ae
Nie

(F)
>o

where

(13.41)
(13.35)

TP

We already know that:

and:
which may be written:
aAG
-aT

AH
T

since AG = 0. Substitution into Equation 13.41yields:


AH
aAG
aAG
dAG=--dT+AVdP+-dt+C-dNie=O
T
at
aNie

(13.42)

And for a closed system (dN, = 0), we obtain:

AH

($)E=
rn

=O

(13.43)

This equation may be compared to the Clapeyron equation (Chapter 2, Section 2.2.2.1)
relative to the liquid-vapor equilibrium of a pure substance, or to the equation expressing
the variation of bubble pressure with temperature in a mixture (Chapter 6, Section 6.16).
Similarly:
(13.44)

and:

(13.45)

In the preceding equations, the quantity AV is defined in a manner analogous to AH


(Eqs. 13.25a or 13.25b) or to AG (Eqs. 13.34a or 13.34b).
We know that (aAG/dQTPis positive (Eq. 13.35) and therefore, the two Equations 13.44
and 13.45, which determine the influence of temperature and pressure on the extent of
reaction, can be summarized by the law of Le Chatelier: Any variation of one of the independent parameters tends to produce a variation in the state of equilibrium in the direction that produces a change in the opposite direction of the parameter considered.

43 1

13. Chemical Reactions

We may also specify the influence of temperature on the equilibrium constant by applying the Gibbs-Helmholtz equation to the Gibbs energy of reaction. We obtain:
/alnKo\

AHo

(13.46)

This equation is often applied assuming that the standard enthalpy of reaction is independent of temperature. It then makes it possible to determine the heat of reaction from
equilibrium measurements performed at two temperatures.
The influence of the addition of a reagent or product cannot be specified [Vidal, 19741
except in very simple cases (ideal gas, ideal solution).We cannot state that the reaction is
going to change in such a way as to consume the added compound, even if this is most
often the case. Finally,if we have an inert compound (diluent), the addition is favorable to
the reaction in the direction where the reaction increases the number of moles produced.

13.4

CALCULATION OF SIMULTANEOUS CHEMICAL EQUILIBRIA

Often several independent chemical transformations occur simultaneouslywithin the reaction medium. A very simple example is provided by the isomerization of paraffins. For npentane, isopentane,and neopentane,we see two independent stoichiometries,for example:
n-pentane + isopentane
and
n-pentane + neopentane
If we express the equilibrium condition in the vapor phase, assumed to be ideal, assigning subscripts 1,2, and 3 to n-pentane, isopentane, and neopentane respectively, and the
subscripts I and I1 to the two reactions described,then:

Yz

Y3

Yl

Yl

- =K, and
taking into account the general relationship:
Y1+

YZ + Y 3 = 1

and we can show that if we consider a paraffinic hydrocarbon with n isomers, and if
is the Gibbs energy of formation of the isomer of order i, then the equilibrium in the vapor
phase is described by:

432

13. Chemical Reactions

This particular case is easy to deal with because on the one hand we have assumed that
the reaction medium was an ideal gas, and on the other hand because the particularly simple stoichiometry of the isomerization reactions leads to resolution of a linear system.The
opposite is generally true, and this problem has been the subject of many investigations.Of
note on this topic is the work of Smith and Missen [1982].
We shall limit ourselves to a few elements concerning one of the methods for solving
such problems.
Consider a system made up of n components where the initial quantities Ni,oare known.
Among the various possible chemical reactions, some may be eliminated for thermodynamic (very small equilibrium constants) or kinetic (lack of appropriate catalyst, for
example) reasons.We retain only p stoichiometries,identified by coefficients v ~ which
, ~ , are
positive for the products and negative for the reagents, with i denoting the component
ranging from 1to n, and j denoting the reaction, ranging from 1top.
Furthermore, if is the extent of the reaction of order j , then the material balances are
expressed as:
D

(13.47b)
It provides the sum of the number of moles as well as the mole fractions of each component. Since temperature and pressure are fixed, we can calculate the chemical potentials
and the Gibbs energy of the mixture using an equation of state or an activity coefficients
model

f.
pi = p; + RT In 2

and

fi

G=
i=l

where the chemical potentials are related to a coherent origin:

1,

p& = g & = h $ . - Ts& = (Ah;)To,i+


92

cki dT-

Ts;.

(13.10)

We are thus led to find the minimum of a function (Gibbs energy, G) as a function of the

p variables 5 j . If we state that the partial derivative of the Gibbs energy with respect to
each extent of reaction is zero, we find the p mass action laws applied to the stoichiometries involved. We thus find the solution after solving a system of p (non-linear) equations
and p unknowns. We may, and it appears preferable, apply a constrained minimization

algorithm (the numbers of moles must remain positive).


Note that the stoichiometries that characterize the evolution of the system can be
replaced by an equivalent system obtained from the linear combination of some of them.
Such a substitution may turn out to be beneficial from the point of view of numerical accuracy.
The most classic and useful of the algorithms, however, disregard the observed
stoichiometries. Given a list of components characterized by their general formula
(C,H,O,N,, for example) they automatically map out all stoichiometriescapable of relating these components, retain those that form independent entities whose number is determined from the rank number of the matrix containing the stoichiometries and optimized

13. Chemical Reactions

43 3

from the point of view of numerical precision, and minimize Gibbs energy as a function of
the extent of the retained reactions. Initially,they were developed to solve the problem of
flame temperature calculation. Indeed in this case, all reactions that may be imagined
based on the individual chemical natures present (CO,, H,O, CO, H,, 0,, H, 0,OH) must
be retained, due to thermal activation. However, this method does not properly address
catalytic transformations that favor a more restricted number of reactions, which is
described by experimentation.It may therefore lead to erroneous solutions. For example,
a calculation for metathesis reactions, such as:
butene-1 + butene-2 + propylene + pentene-2
2 butene-1 -+ ethylene + hexene-3
ethylene + butene-2 + 2 propylene
characterized by the invariance in the number of moles, will be entirely false if the algorithm that disregards the nature of the components takes the actual reactions and substitutes them with a group such as:
butene-1 + 2 ethylene
butene-1 + butene-2
3 butene-1 + 4 propylene
3 butene-1 + 2 hexene-3
5 butene-1 -+ 4 pentene-2
Great caution must therefore be observed when evaluating the validity of processes
applying the so-called method of automatic scripting of chemical reactions, as elegant as
the formula may be. By excluding certain processes due to the mechanism of activation, we
calculate an equilibrium that we may describe as metastable.However, if we do not take
into account these internal constraints, we define the system incorrectly,as we emphasized
in the introduction to the first chapter.
One additional difficulty comes from the possible distribution of the system into several
phases and the possible variation in the number of phases as a function of operating conditions.As we have stated, the Gibbs energy of a system, which must be minimized, is calculated by taking into account the specific properties of each phase, using appropriate
models.

REFERENCES
Benson SW, Cruickshank FR,Golden DM, Haugen GR, ONeal HE, Rodgers AS, Shaw R, Walsh R
(1969) Additivity rules for the estimation of thermochemical properties. Chemical Reviews 69,
279-324.
Chang T, Rousseau RW, Ferrell JK (1983) Use of the modificationof the Redlich-Kwong equation of
state for phase equilibrium calculations. Systems containing Methanol. Znd. Eng. Chem. Proc.Des.
Dev. 22,462-468.
Chermin HAG (1959) Petrol. Refiner 38, (12) 117.
Chermin HAG (1961) Thermo data for petrochemicals. Parts 26 to 32. Petrol. Refiner 40, (2), 145; (3),
181; (4), 127; (5), 234; (6), 179; (9), 261; (lo), 145.

434

13. Chemical Reactions

Delion A,Torck B, Hellin M (1986) Equilibrium constant for the liquid phase hydration of isobutylene over ion-exchange resin. Znd. Eng. Chem. Proc.Des.Dev. 25 (4), 889-893.
Oost C, Sundmacher K, Hoffmann U (1995) Synthesis of tertiary amyl methyl ether (TAME): equilibrium of the multiple reactions. Chem. Eng. Technol. 18 (2) 110-117.
Pauling L (1960) The Nature of Chemical Bond. Cornell University Press, Ithaca, New York.
Pedley JB, Naylor RD, Kirby SP (1977) ThermochemicalData of Organic Compounds. Chapman and
Hall, London.
Rihani DN, Doraiswamy LK (1965) Estimation of heat capacity of organic compounds from group
contributions. Z&EC Fundamentals,4,17-21.
Smith WR, Missen RW (1982) Chemical Reaction Analysis: Theory and Algorithms.Wiley, New York.
Stull DR, Westrum EF, Sinke GC (1987) The Chemical Thermodynamics of Organic Compounds.
R.E. Krieger publishing company, Malabar, Florida.
Van Krevelen DW, Chermin HAG (1951) Estimation of the free enthalpy (Gibbs free energy) of formation of organic compounds from group contributions. Chem. Eng. Sci. 1 (2), 66-80.
Vidal J (1974) Thermodynamique.Mkthodes appliqukes au raffinage et au gknie chimique, Vol. 2,
Chapters 16 to 19. fiditions Technip, Paris.

A Grozdana, tant aimte


Paris 1976, Zagreb 1997

<< Les gens ont des Btoiles qui ne sont pas les mikes.
Pour les uns qui voyagent, les ttoiles sont des guides,
pour dautres qui sont savants elles sont des problkmes.
Mais toutes ces Btoiles-lil se taisent.

Toi tu auras des Btoiles cornme personne nen a,.. B


Le Petit Prince, A. de Saint-ExupBry

All men have stars,


but they are not the same things for different people.
For some, who are travelers, the stars are guides.
For others, who are scholars, they are problems.
But all these stars are silent.
You-you alone-wille have stars as no one else has them ...
The Little Prince, A. de Saint-ExupBry

[Translationfrom:
Antoine de Saint-Exuptry,Le Perit Prince.
Translated from the French by Katherine Woods,
Harcourt Brace Jovanovich, Inc., New York, 19431

Symbols

Roman Alphabet
:molar Helmholtz energy
: equation of state parameter
: activity of component i (Chapter 5)
: Helmholtz energy
:equation of state parameter
:second virial coefficient
:translation parameter in an equation of
state
C :binary parameter in an excess Gibbs
energy model (Chapter 7)
E :total energy (Chapter 1)
f :fugacity
:molar Gibbs energy
g
G :Gibbs energy
h :molar enthalpy
h :Planck's constant
H :enthalpy
k :interaction parameter (Chapter 8)
:Boltzmann constant
K :equilibrium constant
M :molar mass
n
:number of components
N :number of moles
P :pressure
F :vapor pressure
4
:Poynting correction
qi :surface parameter of component i
Q :heat quantity
Qk
:surface parameter of group k
ri :volume parameter of component i
R :ideal gas constant
R, :volume parameter of group k
:molar entropy
S
:entropy
S
T :absolute temperature
:molar internal energy
U
U :internal energy
:molar volume
11
V :volume
Wi :mass fraction
W :work
:mole fraction of component i in the liquid
Xi
phase
yi :mole fraction of component i in the vapor
phase

z :coordination number
zi :mole fraction of component i
2

:compressibility factor

Greek Alphabet
y
: activity coefficient
6 : solubility parameter
7
: molar, packing fraction
8
:temperature
0, : surface fraction of component i
p
: chemical potential
v
:stoichiometric coefficient
4 :extent of reaction
p
: molar density
0 :fugacity coefficient
@ :volume fraction
o :acentric factor

Subscripts
c
: critical (critical temperature, T,, etc.)
cal :calculated
exp :experimental
:formation (enthalpy of formation, Ah;,
f
etc.)
i, j :component
k, I :group
P : at constant pressure
T : at constant temperature
V :at constant volume
SuperscriptslExponents
c : at saturation
* :reference state
:standard state
#
:ideal gas state
id :ideal solution
E :excess property
L :liquid state
M :mixing property
V :vapor state
O

Operators
exp :exponential: exp (x) = ex
In :natural logarithm
log,, :base 10 logarithm

Database

Many numerical databases exist for thermodynamic properties of pure substances and
mixtures.The references for this appendix list some of them.
For a number of components,we have compiled the data that seem to us to be the most
needed for the application of the methods introduced throughout this text.
First we have molar mass, and properties in the standard state (ideal gas at a pressure
equal to 0.1 MPa, with the exception of carbon as graphite), at 298.15 K. We list heat capacity at constant pressure, entropy, enthalpy of formation, and Gibbs energy of formation.
We then have the critical coordinates, critical temperatures, critical pressure and the
critical compressibilityfactor, as well as the acentric factor. These properties are required
for the correlations using corresponding states (Chapter 3) and, frequently,for the calculation of parameters for the equations of state (Chapter 4).
The values related to phase changes, melting temperature, enthalpy of melting, boiling
at atmospheric pressure, and, at this temperature, enthalpy of vaporization and density in
the liquid phase, complete this database. Note that we have not given the parameters for a
vapor pressure equation.At low pressure, we may apply the Antoine equation (Table 2.5),
and for non-polar components, up to the critical point, we may apply the Lee and Kesler
correlation (Chapter 3, Equation 3.32). In order to calculate density for a saturated liquid,
we use the Rackett method (Chapter 3, Equation 3.13 or 3.14). Finally, the variation of
heat of vaporization with temperature may be calculated utilizing the Watson method
(Chapter 2, Equation 2.22).
We note that for carbon dioxide, the boiling temperature, the heat of vaporization, and
density in the liquid state are given for a pressure of 0.5135 MPa (triple point).

43 6

Appendix 1. Database

Symbols and Units


molar mass (g/mol)
heat capacity in the ideal gas state at 298.15 K (J * mo1-l. K-l)
entropy in the ideal gas state at a pressure of 0.1 MPa, at 298.15 K (J . mol-' .K-')
enthalpy of formation in the ideal gas state at 298.15 K (J-mol-')
Gibbs energy of formation in the ideal gas state at a pressure of 0.1 MPa,
at 298.15 K (J.mo1-')
critical temperature (K)
critical pressure (MPa)
critical compressibility factor
acentric factor
melting temperature at atmospheric pressure (K)
enthalpy of melting at atmospheric pressure (J * mol-l)
boiling temperature at atmospheric pressure (K)
density in the liquid state, at boiling temperature at atmospheric pressure, and at
this pressure (g/cm3)
enthalpy of vaporization at boiling temperature at atmospheric pressure (J -mol-')
Abbreviations
co2

H2S
n-CN
cyclo-C,
me-cyclo-C,
~ycl0-C~
me-cyclo-C,
MTBE

carbon dioxide
hydrogen sulfide
normal alkane C,H2N +*
cyclopentane
methylcyclopentane
cyclohexane
methylcyclohexane
methyl tertiary butyl ether

Component

C (graphite)

Hydrogen

Nitrogen

12.01

2.01
28.8
130.6
0
0
33.2
1.31
0.031
-0.215

28
29.1
191.5
0
0
126.1
3.39
0.292

8.64
5.69
0
0

Water

18
33.6
188.7
-241 820
-228 600
647.1
22.06
0.229
0.345
273.1
6000
373.1
0.973
40 800

44.01
37.2
213.7
-39350
-39 440
304.2
7.38
0.274
0.228
216.6
9 020
216.6
1.18
15300

63.1
720
77.4
0.808
5 543

32
29.3
205
0
0
154.6
5.04
0.288
0.022
54.4
444
90.2
1.142
6 825

Ammonia

H2S

0.040

17
35.5
192.7
4 5 900
-16400
405.6
11.28
0.242
0.252
195.4
5660
239.7
0.681
23 330

34.1
34.2
205.6
-20 630
-33 440
373.5
8.96
0.284
0.083
187.7
2 380
212.8
0.943
18890

(*) For carbon dioxide,the boiling temperature,heat of vaporization, and the density in the liquid state are
given for a pressure of 0.5135 MPa (triple point).

Methane

16
35.7
186.3
-74 850
-50 820
190.6
4.60
0.288
0.011
90.7
940.
111.7
0.424
8 165

1Ethane
30.1
53.1
229.1
-83 850
-31 950
305.4
4.88
0.284
0.099
90.3
2 860
184.6
0.545
14690

Propane

n-Butane

44.1
74.5
270.2
-104 680
-24 400
369.8
4.25
0.280
0.152
85.5
3525
231.1
0.583
18800

58.1
98.3
309.9
-125 650
-16560
425.2
3.80
0.274
0.199
134.9
4 660
272.7
0.602
22 435

43 8

Appendix 1. Database

Component
n-Pentane

n-Hexane

n-Heptane

n-Octane

72.2
119.8
349.5
-146710
-8 770
469.6
3.37
0.269
0.249
143.4
8390
309.2
0.610
25 990

86.2
141.8
388.7
-166 940
-80
507.4
3.01
0.264
0.305
177.8
13080
341.9
0.615
29110

100.2
164.4
428
-187 650
8 150
540.3
2.74
0.263
0.351
182.6
14055
371.6
0.615
31730

114.2
187
467.2
-208 820
15920
568.8
2.49
0.259
0.396
216.4
20 740
398.8
0.612
34 765

128.3
209.4
506.4
-228 870
24 730
595.7
2.31
0.255
0.438
219.6
15470
424
0.608
37 150

142.3
232.1
545.7
-249530
32970
618.5
2.12
0.249
0.484
243.5
28715
447.3
0.604
40020

156.3
257
585
-270 290
41 630
638.8
1.97
0.243
0.536
247.6
22 180
469.1
0.598
42 520

170.3
279.8
624.2
-290870
50 040
658.2
1.82
0.242
0.573
263.6
36 840
489.5
0.593
44 340

184.4
301.9
663.4
- 311500
58450
675.8
1.72
0.236
0.619
267.8
28 500
508.6
0.582
45 795

198.4
324.7
702.7
-332040
66 820
692.4
1.62
0.237
0.662
279
45 070
526.7
0.595
49 030

212.4
347.3
741.9
-352750
75 230
706.8
1.52
0.228
0.705
283.1
34590
543.8
0.572
49 570

226.4
370.1
778.3
-373 340
83 680
720.6
1.42
0.220
0.747
291.3
53360
560
0.565
52 520

n-C,

439

Appendix 1. Database

Component

Ahu (J.mo1-')

Isobutane

Isopentane

58.1
96.5
295.4
-134180
-20 760
408.1
3.65
0.282
0.177
113.5
4 540
261.4
0.595
21 400

72.2
118.3
343.6
-152 970
-13 300
460.4
3.38
0.270
0.227
113.3
5 150
301
0.613
24 800

cyclo-c,

Me-cyclo-C,

Cyclo-c,

Me-cyclo-C,

70.1
83.3
292.9
-77 030
38 870
511.8
4.50
0.273
0.194
179.3
610
322.4
0.725
27 260

84.2
109.3
339.9
-106700
35 770
532.8
3.78
0.272
0.230
130.7
6930
345
0.699
29 340

84.2
107.1
298.2
-123 140
31760
553.5
4.07
0.273
0.212
279.7
2 740
353.9
0.721
29 890

98.2
135.1
343.3
-154770
27 280
572.2
3.47
0.269
0.235
146.6
6 750
374.1
0.6%
31 820

cis-Decaline

trans-Decaline

Tetraline

138.3
166.7
377.7
-169240
85 520
702.3
3.24
0.267
0.294
230.2
9 490
469
0.758
39 800

138.3
167.6
374.5
-182 170
73 550
687.1
2.84
0.238
0.254
242.8
14410
460.5
0.737
38 620

132.2
152.2
369.6
26610
167100
720.2
3.62
0.267
0.328
237.4
12450
480.8
0.820
42 350

440

Appendix 1. Database

I
I

Component
Benzene

Toluene

Ethylbenzene

o-Xylene

78.1
82.3
269.2
82 927
129658
562.2
4.90
0.271
0.211
278.7
9 865
353.2
0.816
30750

92.1
104.7
319.7
49 999
122290
591.8
4.11
0.264
0.264
178.2
6 635
383.8
0.780
33 600

106.2
127.1
360.5
29 790
130574
617.2
3.61
0.263
0.304
178.2
9 185
409.4
0.761
35 920

106.2
132.4
352.8
18995
122077
630.4
3.73
0.263
0.313
248
13600
417.6
0.770
37 OOO

m-Xylene

p-Xylene

Naphtalene

106.2
125.7
357.7
17240
118850
617
3.54
0.259
0.326
225.3
11570
412.3
0.757
36 330

106.2
125.3
352.1
18030
121270
616.3
3.51
0.260
0.326
286.4
17115
411.5
0.753
35 820

128.2
132.7
333.1
150580
224 100
748.4
4.05
0.269
0.302
353.4
18980
491.1
0.862
43 420

Ethylene

Propylene

28.1
42.9
219.5
52280
68 125
282.4
5.03
0.277
0.085
104
3350
169.5
0.570
13460

42.1
64.7
266.6
19710
62 140
365.6
4.67
0.289
0.140
87.9
3000
225.5
0.610
18460

Butene-1

Butene-2-cis

56.1
85.7
307.8
-540
70240
419.6
4.02
0.276
0.187
87.8
3850
266.9
0.626
22 440

56.1
79
300.8
-6 990
65 855
435.6
4.21
0.272
0.203
134.3
7310
276.9
0.642
23430

Appendix 1. Database

441

~-

Component
Butene-2-trans

Isobutene

56.1
87.8
296.5
-11 170
62 970
428.6
4.10
0.274
0.218
167.6
9 760
274
0.627
22 900

56.1
89.1
293.6
-16900
58 075
417.9
4
0.275
0.189
132.8
5 930
266.3
0.627
22 210

Acetone

2-Butanone

Methanol

Ethanol

58.1
75.6
295.3
-217 150
-152720
508.2
4.70
0.233
0.306
178.4
5 690
329.4
0.748
29 790

72.1
103.2
338.1
-238 360
-146060
535.5
4.15
0.249
0.324
186.5
8440
352.8
0.739
31220

32
44
239.7
-200 670
-162 420
512.6
8.10
0.224
0.566
175.5
3 200
337.9
0.749
35 140

46.1
65.4
282.6
-234 430
-167 900
516.3
6.38
0.248
0.637
159.1
4 930
351.4
0.734
39400

Diethylether

MTBE

Tetrahydrofurane

74,l
116.7
341
-252 130
-121 750
466.7
3.64
0.262
0.285
156.9
7 190
307.6
0.697
27 090

88.1
127.9
353
-292 880
-125 440
497.1
3.43
0.273
0.267
164.6
7 600
328.4
0.702
28 170

72.1
77.4
297.3
-184 180
-79 680
540.1
5.19
0.259
0.225
164.7
8540
339.1
0.835
30 090

Acetonitrile
41.1
52.1
243.5
74 040
91 820
545.5
4.83
0.184
0.338
229.3
8910
354.8
0.717
30300

442

Appendix 1. Database

REFERENCES
Arlt W, Macedo MEA, Rasmussen P, Sorensen JM (1979-1987) Liquid-liquid equilibrium data collection. Chemistry Data Series,Vol. V. Dechema, Frankfurt.
Christensen C, Gmehling J, Rasmussen P, Weidlich U, Holderbaum T (1984-1991) Heats of mixing
data collection. Chemistry Data Series,Vol. 111.Dechema, Frankfurt.
Christensen JJ, Hanks RW, Izatt RM (1982) Handbook of heats of mixing. John Wiley & sons, New
York.
Daubert TE, Danner FW (1985) Data compilation tables of properties of pure compounds. Design
Institute for Physical Property Data; American Institute of Chemical Engineers (DIPPR
/AIChE).
Engels H (1990) Phase equilibria and phase diagrams of electrolytes. Chemistry Data Series,Vol. XI.
Dechema, Frankfurt.
Figurski G Vapor-liquid equilibrium data for electrolyte solutions. Chemistry Data Series,Vol. XIII.
Dechema, Frankfurt.
Gmehling J, Menke J, Fischer K, Krafczyk J (1995) Azeotropic data. Verlag Chemie, Weinheim,
Germany.
Gmehling J, Onken U, Arlt W, Grenzheuser P, Weidlich U, Kolbe B, Rarey JR (1978-1994) Vaporliquid equilibrium data collection. Chemistry Data Series,Vol. I. Dechema, Frankfurt.
Gmehling J,Tiegs D, Medina A, Soares M, Bastos J, Alessi P, Kikic I, Schiller M, Menke J (1986-1994)
Activity coefficients at infinite dilution. Chemistry Data Series,Vol. IX. Dechema, Frankfurt.
Kehiaian HV, Marsh K (Eds.) (1973-1995) Znternational data series. Selected data on mixtures.
Thermodynamics Research Center, The Texas A&M University, College Station, Texas.
Knapp H, Doring R, Oellrich L, Plocker U, Prausnitz JM, Langhorst R, Zeck S (1982-1989) Vaporliquid equilibria for mixtures of low boiling substances. Chemistry Data Series,Vol. VI. Dechema,
Frankfurt.
Knapp H, Teller M, Langhorst R (1987) Solid-liquid equilibrium data collection. Chemistry Data
Series,Vol. VIII. Dechema, Frankfurt.
Reid RC, Prausnitz JM, Sherwood ThK (1977) The properties of gases and liquids. 3rd edition.
McGraw-Hill,New York.
Reid RC, Prausnitz JM, Poling BE (1987) The properties of gases and liquids. 4th edition. McGrawHill, New York.
Simmrock KH, Janowsky R, Ohnsorge A (1986) Critical data of pure substances. Chemistry Dam
Series,Vol. 11.Dechema, Frankfurt.
Wen Hao, Elbro HS, Alessi P (1992) Polymer solution data collection. Chemistry Dnta Series,
Vol. XIV. Dechema, Frankfurt.
Wichterle I, Linek J, Wagner 2,Kehiaian H V (1993) Vapor-liquidequilibrium bibliographic database.
ELDATA SARL, Paris.

Appendix

l e e and Kesler Method


Compressibility Factor, Residual Terms for
Enthalpy, Entropy, and Heat Capacity at
Constant Pressure, and Fugacity Coefficient

The following tables allow the calculation of the compressibility factor, some residual
properties, and the fugacity coefficient as a function of the reduced coordinates and the
acentric factor w,by application of the Lee and Kesler method discussed in Chapter 3,
Section 3.2.2.3.
The tables provide the values for the following quantities:
Compressibilityfactor Z(O)of the simple fluid: Table A2.1.
Corrective term A Z Table A2.2.
Adimensional residual enthalpy of the simple fluid (with opposite sign):

(hi:)@),
Table A2.3.

- -

Corrective term (with opposite sign): -A

(hi:),
Table A2.4.
-

Adimensional residual entropy of the simple fluid (with opposite sign):

-( $)@),
Table A 2 5

Corrective term (with opposite sign): -A

( is#),
- Table A2.6.

Logarithm (base 10) of the fugacity coefficient of the simple fluid (with opposite sign):
-(log,,

);

(0)

,Table A2.7.

Corrective term (with opposite sign) -A

loglo -

, Table A2.8.

444
0

Appendix 2. L e e and Kesler Method

Adimensional residual heat capacity at constant pressure of simple fluids:

(cpRcT)

,Table A2.9.

Corrective term: A

R')

,Table A2.10.

Application of these tables to a fluid with acentric factor m is accomplished using the
following expressions:
Z=Z(O)+mAZ

(3.31)

R
loglo

f = (log,,
P

;)"+'

mA(log10

);

Appendix 2. Lee and Kesler Method

445

Table A2.1
Compressibility factor for the simple fluid

- --- ---

0.0100
0.0200
0.0300
0.04oO
0.0500
0.06oO
0.0700
0.0800
0.09oO
0.1000
0.1100
0.1200
0.1300
0.1400
0.1500
0.1600
0.1700
0.1800
0.1900
0.2000
0.2100
0.2200
0.2300
0.2400
0.2500
0.2563
0.2563
0.2600
0.2700
0.2800
0.2900
0.3000
0.3100
0.3200
0.3300
0.3400
0.3500
0.3600
0.3700
0.3800
0.3900

0.8

0.9

0.94

0.98

1.02

1.04

1.06

1.1

1.2

1.5

0.9935
0.9870
0.9804
0.9737
0.9669
0.9601
0.9532
0.9462
0.9391
0.9319
0.9246
0.9172
0.9097
0.9021
0.8944
0.8866
0.8786
0.8705
0.8623
0.8539
0.8453
0.8366
0.8277
0.8186
0.8093
0.8034
0.0425
0.0431
0.0448
0.0464

0.9954
0.9908
0.9862
0.9815
0.9768
0.9721
0.9673
0.9625
0.9577
0.9528
0.9478
0.9429
0.9379
0.9328
0.9277
0.9226
0.9174
0.9121
0.9068
0.9015
0.8961
0.8906
0.8851
0.8795
0.8739
0.8703
0.8703
0.8682
0.8624
0.8566
0.8507
0.8447
0.8387
0.8325
0.8263
0.8200
0.8136
0.8071
0.8005
0.7938
0.7869

0.9960
0.9919
0.9879
0.9838
0.9797
0.9755
0.9714
0.9672
0.9630
0.9587
0.9544
0.9501
0.9458
0.9414
0.9370
0.9326
0.9281
0.9236
0.9191
0.9145
0.9099
0.9052
0.9005
0.8958
0.8910
0.8880
0.8880
0.8862
0.8814
0.8765
0.8715
0.8665
0.8615
0.8564
0.8512
0.8460
0.8407
0.8354
0.8300
0.8246
0.8191

0.9965
0.9929
0.9893
0.9857
0.9821
0.9785
0.9748
0.9711
0.9674
0.9637
0.9600
0.9562
0.9524
0.9486
0.9448
0.9409
0.9370
0.9331
0.9292
0.9253
0.9213
0.9173
0.9132
0.9092
0.9051
0.9025
0.9025
0.9010
0.8968
0.8926
0.8884
0.8842
0.8799
0.8756
0.8713
0.8669
0.8625
0.8580
0.8535
0.8490
0.8444

0.9967
0.9933
0.9899
0.9866
0.9832
0.9797
0.9763
0.9729
0.9694
0.9659
0.9624
0.9589
0.9553
0.9518
0.9482
0.9446
0.9410
0.9373
0.9337
0.9300
0.9263
0.9225
0.9188
0.9150
0.9112
0.9088
0.9088
0.9074
0.9035
0.8996
0.8957
0.8918
0.8879
0.8839
0.8799
0.8758
0.8717
0.8676
0.8635
0.8593
0.8551

0.9969
0.9937
0.9905
0.9873
0.9842
0.9809
0.9777
0.9745
0.9712
0.9679
0.9646
0.9613
0.9580
0.9547
0.9513
0.9480
0.9446
0.9412
0.9377
0.9343
0.9308
0.9273
0.9238
0.9203
0.9168
0.9146
0.9146
0.9132
0.9096
0.9060
0.9024
0.8988
0.8951
0.8914
0.8877
0.8839
0.8802
0.8764
0.8726
0.8687
0.8649

0.9970
0.9941
0.9911
0.9881
0.9851
0.9820
0.9790
0.9760
0.9729

0.9972
0.9944
0.9916
0.9888
0.9859
0.9831
0.9802
0.9773
0.9745
0.9716

0.9975
0.9950
0.9925
0.9900
0.9874
0.9849
0.9824
0.9798
0.9772
0.9747
0.9721
0.9695
0.9669
0.9643
0.9617
0.9591
0.9564
0.9538
0.9511
0.9485
0.9458
0.9431
0.9404
0.9377
0.9350
0.9333
0.9333
0.9323
0.9295
0.9268
0.9240
0.9213
0.9185
0.9157
0.9129
0.9101
0.9073
0.9044
0.9016
0.8987
0.8958

0.9981
0.9962
0.9943
0.9924
0.9904
0.9885
0.9866
0.9846
0.9827
0.9808
0.9788
0.9769
0.9749
0.9730
0.9710
0.9690
0.9671
0.9651
0.9631
0.9611
0.9591
0.9572
0.9552
0.9532
0.9511
0.9499
0.9499
0.9491
0.9471
0.9451
0.9431
0.9410
0.9390
0.9370
0.9349
0.9329
0.9308
0.9288
0.9267
0.9246
0.9226

0.9991
0.9982
0.9973
0.9964
0.9954
0.9945
0.9936
0.9927
0.9918
0.9909
0.9900
0.9891
0.9882
0.9872
0.9863
0.9854
0.9845
0.9836
0.9827
0.9818
0.9809
0.9800
0.9791
0.9781
0.9772
0.9767
0.9767
0.9763
0.9754
0.9745
0.9736
0.9727
0.9718
0.9709
0.9700
0.9691
0.9682
0.9673
0.9663
0.9654
0.9645

0.0480

0.0497
0.0513
0.0530
0.0546
0.0562
0.0579
0.0595
0.0612
0.0628
0.0644

0.9698
0.9667
0.9636
0.9605
0.9574
0.9542
0.9511
0.9479
0.9447
0.9415
0.9383
0.9350
0.9318
0.9285
0.9252
0.9219
0.9198
0.9198
0.9186
0.9152
0.9119
0.9085
0.9051
0.9017
0.8983
0.8948
0.8914
0.8879
0.8844
0.8808
0.8773
0.8737

0.9686
0.9657
0.9628
0.9599
0.9569
0.9539
0.9510
0.9480
0.9450
0.9419
0.9389
0.9359
0.9328
0.9297
0.9266
0.9247
0.9247
0.9235
0.9204
0.9173
0.9141
0.9110
0.9078
0.9046
0.9014
0.8981
0.8949
0.8916
0.8883
0.8850
0.8817

446

Appendix 2. L e e and Kesler Method

Table A2.1 (cont.)


Compressibility factor for the simple fluid

0.8

0.9

0.94

0.98

1.02

1.04

1.06

1.1

1.2

1.5

0.8701
0.8665
0.8628
0.8592
0.8555
0.8517
0.8480
0.8442
0.8404
0.8366
0.8328
0.8289
0.8250
0.8210
0.8181
0.8181
0.8171
0.8131
0.8091
0.8050

0.8784
0.8750
0.8717
0.8683
0.8649
0.8614
0.8580
0.8545
0.8510
0.8475

0.8930
0.8901
0.8872
0.8842
0.8813
0.8783
0.8754
0.8724
0.8694
0.8664
0.8634
0.8603
0.8573
0.8542
0.8519
0.8519
0.8512
0.8481
0.8449
0.8418
0.8387
0.8355
0.8323
0.8292
0.8259
0.8227
0.8195
0.8162
0.8129
0.8096
0.8063
0.8030
0.8008
0.8008
0.7996
0.7%2
0.7928
0.7894
0.7860
0.7825
0.7790

0.9205
0.9184

0.9636
0.9627
0.9618
0.9609

0.4oOo
0.4100
0.4200
0.4300
0.4400
0.4500
0.4600
0.4700
0.4800
0.4900
0.5000

0.5100
0.5200
0.5300
0.5374
0.5374
0.5400
0.5500
0.5600
0.5700
0.5800
0.5900

0.6oOo
0.6100
0.6200
0.6300
0.6400
0.6500
0.6600
0.6700
0.6800
0.6900
0.6963
0.6963
0.7000
0.7100
0.7200
0.7300
0.7400
0.7500
0.7600

0.0661
0.0677
0.0693
0.0709
0.0726
0.0742
0.0758
0.0775
0.0791
0.0807
0.0823
0.0840
0.0856
0.0872
0.0884
0.0884
0.0888

0.0904
0.0921
0.0937
0.0953
0.0969
0.0985
0.1002
0.1018
0.1034
0.1050
0.1066
0.1082
0.1098
0.1114
0.1130
0.1141
0.1141
0.1147
0.1163
0.1179
0.1195
0.1211
0.1227
0.1243

0.7800
0.7729
0.7657
0.7584
0.7508
0.7432
0.7353
0.7273
0.7191
0.7106
0.7019
0.6930
0.6837
0.6742
0.6668
0.0906
0.0910
0.0926
0.0942
0.0958
0.0974
0.0990
0.1006
0.1022
0.1037
0.1053
0.1069
0.1085
0.1101
0.1117
0.1133
0.1148
0.1158
0.1158
0.1164
0.1180
0.1196
0.1211
0.1227
0.1243
0.1258

0.8135
0.8079
0.8021
0.7963
0.7904
0.7845
0.7784
0.7723
0.7660
0.7597
0.7532
0.7466
0.7400
0.7331
0.7280
0.7280
0.7262
0.7191
0.7118
0.7044
0.6968
0.6890
0.6810
0.6728
0.6643
0.6556
0.6465

0.6371
0.6274
0.6171
0.6064
0.5951
0.5876
0.1222
0.1227
0.1243
0.1258
0.1273
0.1289
0.1304
0.1319

0.8398
0.8351
0.8304
0.8257
0.8209
0.8160
0.8111
0.8062
0.8012
0.7961
0.7910
0.7858
0.7805
0.7752
0.7712
0.7712
0.7699
0.7644
0.7589
0.7533
0.7476
0.7418
0.7360
0.7300
0.7240
0.7179
0.7116
0.7052
0.6988
0.6921
0.6854
0.6785
0.6740
0.6740
0.6714
0.6642
0.6568
0.6492
0.6414
0.6333
0.6250

0.8509
0.8466
0.8423
0.8380
0.8336
0.8292
0.8248
0.8203
0.8157
0.8111
0.8065

0.8018
0.7971
0.7923
0.7888
0.7888
0.7875
0.7827
0.7777
0.7727
0.7677
0.7626
0.7574
0.7522
0.7469
0.7415
0.7361
0.7306
0.7250
0.7193
0.7135
0.7077
0.7039
0.7039
0.7017
0.6956
0.6895
0.6832
0.6768
0.6702
0.6636

0.8610
0.8570
0.8531
0.8491
0.8451
0.8411
0.8370
0.8329
0.8287
0.8246
0.8204
0.8161
0.8118
0.8075
0.8043
0.8043
0.8031
0.7988
0.7943
0.7898
0.7853
0.7807
0.7761
0.7715
0.7668
0.7620
0.7572
0.7523
0.7474
0.7424
0.7373
0.7322
0.7290
0.7290
0.7271
0.7218
0.7165
0.7111
0.7056
0.7001
0.6945

0.8009
0.7968
0.7926
0.7884
0.7842
0.7799
0.7756
0.7712
0.7668
0.7624
0.7579
0.7534
0.7505
0.7505
0.7488
0.7442
0.7395
0.7348
0.7300
0.7252
0.7203

0.8440

0.8404
0.8368
0.8332
0.8305
0.8305
0.8296
0.8259
0.8223
0.8186
0.8148
0.8111
0.8073
0.8035
0.7996
0.7958
0.7919
0.7879
0.7840
0.7800
0.7760
0.7719
0.7693
0.7693
0.7678
0.7637
0.7595
0.7553
0.7511
0.7468
0.7425

0.9163
0.9142
0.9121
0.9100
0.9079
0.9058
0.9037
0.9016
0.8994
0.8973
0.8952
0.8930
0.8914
0.8914
0.8909
0.8887
0.8866
0.8844
0.8822
0.8800
0.8779
0.8757
0.8735
0.8713
0.8691
0.8669
0.8646
0.8624
0.8602
0.8580
0.8565
0.8565
0.8557
0.8535
0.8512
0.8490
0.8467
0.8444
0.8422

0.9600
0.9591
0.9582
0.9573
0.9564
0.9555
0.9546
0.9537
0.9528
0.9519
0.9512
0.9512
0.9510
0.9501
0.9492
0.9483
0.9474
0.9465
0.9456
0.9447
0.9438
0.9429
0.9420
0.9412
0.9403
0.9394
0.9385
0.9376
0.9370
0.9370
0.9367
0.9358
0.9349
0.9340
0.9331
0.9323
0.9314

447

Appendix 2. Lee and Kesler Method

Table A2.1 (cont.)


Compressibility factor for the simple fluid

--0.8

0.9

0.94

0.98

1.02

1.04

0.7700
0.7800
0.7900
0.8OOO
0.8100
0.8200
0.8300
0.8400
0.8500
0.8600
0.8700
0.8800
0.8890
0.8890
0.8900
0.9OoO

0.9100
0.9200
0.9300
0.9400
0.9500
0.9600
0.9700
0.9800
0.9820
0.9840
0.9860
0.9880
0.9900
0.9910
0.9920
0.9930
0.9940
0.9950
0.9960
0.9970
0.9980
0.9990
1.oooO
1.0010
1.0020

0.1259
0.1275
0.1291
0.1307
0.1323
0.1339
0.1355
0.1371
0.1387
0.1403
0.1419
0.1435
0.1449
0.1449
0.1451
0.1467
0.1483
0.1499
0.1514
0.1530
0.1546
0.1562
0.1578
0.1594
0.1597
0.1600
0.1603
0.1607
0.1610
0.1611
0.1613
0.1615
0.1616
0.1618
0.1619
0.1621
0.1623
0.1624
0.1626
0.1627
0.1629

0.1274
0.1290
0.1305
0.1321
0.1337
0.1352
0.1368
0.1383
0.1399
0.1414
0.1430
0.1445
0.1459
0.1459
0.1461
0.1476
0.1492
0.1507
0.1523
0.1538
0.1553
0.1569
0.1.584
0.1600
0.1603
0.1606
0.1609
0.1612
0.1615
0.1616
0.1618
0.1620
0.1621
0.1623
0.1624
0.1626
0.1627
0.1629
0.1630
0.1632
0.1633

0.1334
0.1349
0.1365
0.1380
0.1395
0.1410
0.1425
0.1441
0.1456
0.1471
0.1486
0.1501
0.1515
0.1515
0.1516
0.1531
0.1546
0.1562
0.1577
0.1592
0.1607
0.1622
0.1637
0.1652
0.1655
0.1658
0.1661
0.1664
0.1667
0.1668
0.1670
0.1671
0.1673
0.1674
0.1676
0.1677
0.1679
0.1680
0.1682
0.1683
0.1685

0.6165
0.6076
0.5983
0.5887
0.5786
0.5680
0.5568
0.5449
0.5321
0.5181
0.5026
0.4849
0.4663
0.1742
0.1743
0.1748
0.1755
0.1762
0.1771
0.1780
0.1790
0.1800
0.1811
0.1821
0.1824
0.1826
0.1828
0.1830
0.1833
0.1834
0.1835
0.1836
0.1837
0.1838
0.1839
0.1840
0.1842
0.1843
0.1844
0.1845
0.1846

0.6567
0.6498
0.6426
0.6353
0.6278
0.6201
0.6121
0.6039
0.5954
0.5867
0.5775
0.5680
0.5592
0.5592
0.5581
0.5477
0.5367
0.5250
0.5124
0.4989
0.4840
0.4673
0.4481
0.4248
0.4194
0.4136
0.4074
0.4006

0.3931
0.3890
0.3847
0.3800
0.3749
0.3693
0.3629
0.3555
0.3464
0.3338
0.2918
0.2575
0.2508

0.6888
0.6829
0.6770
0.6710
0.6649
0.6587
0.6523
0.6458
0.6392
0.6325
0.6256
0.6185
0.6120
0.6120
0.6113
0.6038
0.5962
0.5883
0.5802
0.5719
0.5633
0.5543
0.5450
0.5353
0.5333
0.5313
0.5293
0.5273
0.5252
0.5242
0.5231
0.5221
0.5210
0.5200
0.5189
0.5178
0.5168
0.5157
0.5146
0.5135
0.5124

0.7154
0.7104
0.7053
0.7002
0.6950
0.6897
0.6844
0.6790
0.6735
0.6679
0.6622
0.6565
0.6513
0.6513
0.6506
0.6447
0.6386
0.6325
0.6262
0.6198
0.6133
0.6067
0.5999
0.5929
0.5915
0.5901
0.5887
0.5873
0.5858
0.5851
0.5844
0.5837
0.5830
0.5822
0.5815
0.5808
0.5801
0.5793
0.5786
0.5779
0.5771

1.06

1.1

1.2

1.5

0.7382
0.7338
0.7293
0.7249
0.7203
0.7158
0.7111
0.7065
0.7017
0.6970
0.6921
0.6872
0.6828
0.6828
0.6823
0.6773
0.6722
0.6671
0.6619
0.6566
0.6513
0.6459
0.6404
0.6349
0.6337
0.6326
0.6315
0.6304
0.6292
0.6287
0.6281
0.6275
0.6270
0.6264
0.6258
0.6252
0.6247
0.6241
0.6235
0.6229
0.6224

0.7755
0.7720
0.7685
0.7649
0.7613
0.7577
0.7540
0.7504
0.7467
0.7429
0.7392
0.7354
0.7320
0.7320
0.7316
0.7278
0.7240
0.7201
0.7162
0.7122
0.7083
0.7043
0.7002
0.6962
0.6954
0.6945
0.6937
0.6929
0.6921
0.6917
0.6913
0.6909
0.6904
0.6900
0.6896
0.6892
0.6888
0.6884
0.6880
0.6876
0.6871

0.8399
0.8376
0.8353
0.8330
0.8307
0.8284
0.8261
0.8237
0.8214
0.8191
0.8167
0.8144
0.8123
0.8123
0.8120
0.8097
0.8073
0.8049
0.8026
0.8002
0.7978
0.7954
0.7930
0.7906
0.7901
0.78%
0.7891
0.7887
0.7882
0.7879
0.7877
0.7875
0.7872
0.7870
0.7867
0.7865
0.7862
0.7860
0.7858
0.7855
0.7853

0.9305
0.9296
0.9287
0.9278
0.9269
0.9261
0.9252
0.9243
0.9234
0.9225
0.9217
0.9208
0.9200
0.9200
0.9199
0.9190
0.9182
0.9173
0.9164
0.9155
0.9147
0.9138
0.9129
0.9121
0.9119
0.9117
0.9116
0.9114
0.9112
0.9111
0.9110
0.9109
0.9109
0.9108
0.9107
0.9106
0.9105
0.9104
0.9103
0.9103
0.9102

----

------- ----

448

Appendix 2. Lee and Kesler Method

Table A2.1 (cont. and end)


Compressibility factor for the simple fluid

0.8

0.9

0.94

0.98

- -1.0030
1.oO40
1.0050
LOO60
1.0070
1.0080
1.0090
1.0100
1.0120
1.0140
1.0160
1.0180
1.0200
1.0300
1.0400
1.0500
1.0600
1.0700
1.0800
LO900
1.1000
1.1100
1.1200
1.1300
1.1400
1.1500
1.1600
1.1700
1.1800
1.1900
1.2000

0.1630
0.1632
0.1634
0.1635
0.1637
0.1638
0.1640
0.1642
0.1645
0.1648
0.1651
0.1654
0.1657
0.1673
0.1689
0.1705
0.1721
0.1737
0.1752
0.1768
0.1784
0.1800
0.1816
0.1831
0.1847
0.1863
0.1879
0.1894
0.1910
0.1926
0.1942

0.1635
0.1636
0.1638
0.1640
0.1641
0.1643
0.1644
0.1646
0.1649
0.1652
0.1655
0.1658
0.1661
0.1676
0.1692
0.1707
0.1722
0.1737
0.1753
0.1768
0.1783
0.1798
0.1813
0.1829
0.1844
0.1859
0.1874
0.1889
0.1904
0.1920
0.1935

0.1686
0.1688
0.1689
0.1691
0.1692
0.1694
0.1695
0.1697
0.1700
0.1703
0.1706
0.1709
0.1712
0.1727
0.1741
0.1756
0.1771
0.1786
0.1801
0.1816
0.1831
0.1846
0.1861
0.1875
0.1890
0.1905
0.1920
0.1935
0.1949
0.1964
0.1979

1.02

1.04

1.06

0.1847
0.1848
0.1850
0.1851
0.1852
0.1853
0.1854
0.1855
0.1858
0.1860
0.1863
0.1865
0.1867
0.1879
0.1891
0.1904
0.1916
0.1928
0.1941
0.1954
0.1966
0.1979
0.1992
0.2005
0.2018
0.2031
0.2044
0.2057
0.2070
0.2084
0.2097

0.2465
0.2434
0.2408
0.2388
0.2370
0.2355
0.2341
0.2329
0.2309
0.2292
0.2277
0.2265
0.2254
0.2216
0.2193
0.2179
0.2171
0.2166
0.2165
0.2165
0.2167
0.2171
0.2176
0.2181
0.2188
0.2195
0.2202
0.2210
0.2219
0.2228
0.2237

0.5113
0.5102
0.5091
0.5080
0.5069
0.5057
0.5046

0.5034
0.5011
0.4988
0.4964
0.4940
0.4916
0.4791
0.4657
0.4513
0.4356
0.4187
0.4002
0.3803
0.3597
0.3395
0.3214
0.3063
0.2944
0.2852
0.2781
0.2727
0.2686
0.2654
0.2629

0.5764
0.5756
0.5749
0.5741
0.5734
0.5727
0.5719
0.5711
0.5696
0.5681
0.5666
0.5651
0.5635
0.5557
0.5477
0.5394
0.5309
0.5222
0.5131
0.5038
0.4942
0.4843
0.4741
0.4636
0.4529
0.4419
0.4307
0.4194
0.4081
0.3970
0.3863

0.6218
0.6212
0.6206
0.6200
0.6195
0.6189
0.6183
0.6177
0.6165
0.6153
0.6142
0.6130
0.6118
0.6058
0.5997
0.5935
0.5872
0.5809
0.5744
0.5677
0.5610
0.5542
0.5472
0.5401
0.5329
0.5256
0.5182
0.5107
0.5030
0.4953
0.4875

1.2

1.1

- ---0.6867
0.6863
0.6859
0.6855
0.6851
0.6846
0.6842
0.6838
0.6830
0.6821
0.6813
0.6805
0.6796
0.6754
0.6712
0.6669
0.6626
0.6582
0.6538
0.6494
0.6449
0.6404
0.6359
0.6313
0.6267
0.6221
0.6174
0.6127
0.6080
0.6032
0.5984

1.5

0.7850
0.7848
0.7845
0.7843
0.7841
0.7838
0.7836
0.7833
0.7828
0.7824
0.7819
0.7814
0.7809
0.7785
0.7760
0.7736
0.7711
0.7687
0.7662
0.7638
0.7613
0.7588
0.7563
0.7538
0.7513
0.7489
0.7464
0.7438
0.7413
0.7388
0.7363

0.9101
0.9100
0.9099
0.9098
0.9097
0.9097
0.9096
0.9095
0.9093
0.9091
0.9090
0.9088
0.9086
0.9078
0.9069
0.9060
0.9052
0.9043
0.9035
0.9026
0.9018
0.9009
0.9001
0.8992
0.8984
0.8975
0.8967
0.8958
0.8950
0.8941
0.8933

----

449

Appendix 2. Lee and Kesler Method

Table A2.2
Compressibility factor: corrective term
- - - -- 0.8

0.9

0.94

0.98

-- -- -

1.02

1.04

1.a6

1.1

1.2

1.5

0.0100

-0.0044

-0.0019 -0.0013

4.0009 -0.0007

-0.0005

4.OoO4

4.oO02

D.0000

O.OOO4

D.0008

0.0200

-0.0088

-0.0039 -0.0027

4.0018

-0.0014

-0.0010 4.0007

-0.0005

0.0000

0.0008

0.0015

0.0300

4.0134

4.0059

-0.0041

4.0026

-0.0021 -0.0015

4.0011

-0.0007

0.0000

0.0011

0.0023

0.0400

4.0180

-0.0079 -0.0054

4.0035

-0.0028 -0.0021 4.0014

-0.0009

0.0000

0.0015

D.0031

0.0500

4.0228

-0.0099

-0.0068 -0.0044 -0.0034 -0.0026

4.0018

-0.0011

0.0000

0.0019

0.0039

0.O600

-0.0277

4.0120

-0.0082

4.0053 -0.0041 -0.0031 4.0022

-0.0013

D.0000

0.0023

0.0046

0.0700

4.0327

4.0141

-0.0096

4.0062 -0.0048

-0.0036

-0.0025

4.0016

D.OOOO

0.0027

0.0054

0.0800
0.0900

4.0379

-0.0162 -0.0111

-0.0071 -0.0055

-0.0041 -0.0029

4.0018

D.OOO1

0.0031

0.0062

-0.0432

-0.0183

-0.0125

4.0081 -0.0062

-0.0046

-0.0032

4.0020

0.0001

0.0035

0.0070

0.1000

-0.0487

4.0205

-0.0139

-0.0090

-0.0069 -0.0051 -0.0036

4.0022

0.0001

0.0040

0.0078

0.1100

-0.0543

4.0227 -0.0154

-0.0099

-0.0076

-0.0057

4.0039

4.0024

D.OOO2

0.0044

0.0086

0.1200

4.0601

-0.0249 -0.0169

4.0108

-0.0083

-0.0062

-0.0043

4.0026

0.0002

0.0048

0.0094

0.1300

4.0662

4.0272

-0.0184

4,0117

-0.0090

-0.0067

4.0046

4.0028

0.0003

0.0052

0.0102

0.1400

-0.0724

4.0295

-0.0199

-0.0127

-0.0097

-0.0072

-0.0049

4.0030

0.0003

0.0057

0.0110

0.1500

-0.0789

4.0319

-0.0214

4.0136 -0.0104

-0.0077

-0.0053

4.0032

O.ooo4

0.0061

0.0118

0.1600

-0.0857

4.0342 -0.0230

4.0146 -0.0112

-0.0082

-0.0056

4.0033

O.ooo4

0.0066

0.0126

0.1700

-0.0927

4.0367 -0.0245

-0.0155

-0.0119

-0.0087

-0.0059

4.0035

0.0005

0.0070

0.0134

0.1800

-0.1001

4.0391 -0.0261 -0.0165

-0.0126

-0.0092

-0.0063

-0.0037

O.OOO6

0.0075

0.0142

0.1900

-0.1079

4.0416 -0.0277

-0.0174

-0.0133

-0.0097

-0.0066 4.0039

O.ooo6

0.0079

0.0150

0.2000

-0.1160

4.0442 -0.0293

-0,0184 -0.0140

-0.0102

-0.0069

4.ma

0.0007

0.0084

0.0158

0.2100

-0.1246

4.0468

-0.0309

-0.0193 -0.0147

-0.0107

-0.0072

4.0042

0.0008

0.0089

0.0166

0.2200

-0.1337

4.0494

-0.0326

-0.0203

-0.0154

-0.0112

-0.0075

4.0043

0.0009

0.0093

0.0174

0.2300

-0.1434

-0.0521 -0.0342

-0.0213 -0.0162

-0.0117

-0.0079

-0.0045

0.0010

0.0098

0.0182

0.2400

-0.1538

4.0549

-0.0359

-0.0223 -0.0169

-0.0122

-0.0082

-0.0046

0.0011

0.0103

0.0190

0.2500

-0.1650

4.0577 -0.0376

-0.0233

-0.0176

-0.0127

-0.0085

-0.0048

0.0012

0.0108

0.0198

0.2563

-0.1725

4.0595

-0.0387

-0.0239 -0.0180

-0.0130

-0.0087

4.0049

0.0013

0.0111

0.0204

0.2563

-0.0177

-0.0595

-0.0387

-0,0239 -0.0180

-0.0130

-0.0087

-0.0049

0.0013

0.0111

0.0204

0.2600

-0.0179

4.0606 -0.0394

-0.0243 -0.0183

-0.0132

-0.0088

-0.0049

0.0013

0.0113

0.0207

0.2700

-0.0186

-0.0636 -0.0412

-0.0253 -0.0190

-0.0137

-0.0091 -0.0051

0.0015

0.0118

0.0215

0.2800

-0.0193

4.0666

-0.0429

-0.0263 -0.0198

-0.0142

-0.0094

-0.0052

0.0016

0.0123

0.0223

0.2900

-0.0199

4.0698

-0.0448

-0.0147

-0.0096

-0.0053

0.0017

0.0129

0.0231

0.3000

-0.0206

-0.073C

-0.0273 -0.0205
-0.0466 -0.0283 -0.0212

-0.0151 -0.0099

-0.0054

0.0019

0.0134

0.0240

0.3100

-0.0213

-0.0763 -0.0485

-0.0293 -0.0219

-0.0156

-0.0102

-0.0055

0.0020

0.0139

0.0248

0.3200

-0.0219

-0.0797

-0.0304 -0.0227

-0.0161 -0.0105

-0.00%

0.0022

0.0145

0.0256

0.3300

-0.0226

-0.0832 -0.0524

-0.0314

-0.0234

-0.0166

-0.0107

-0.005i

0.0024

0.0150

0.0265

0.3400

-0.0233

-0.0866

-0.0543

-0.0325

-0.0241

-0.0171

-0.0110

-0.005E

0.0025

0.0156

0.0273

0.3500

-0.0239

-0.090f

-0.0564 -0.0335

-0.0245

-0.0175

-0.0113

-0.0055

0.0027

0.0161

0.0281

0.3600

-0.0246

-0.0942

-0.0584

-0.0346

-0.023

-0.018C

-0.0115

-0.OoM

0.0029

0.0167

0.0290

0.3700

-0.0252

-0.0985

-0.0605

-0.0357

-0.026? -0.0185

-0.0118

-0.m 0.0031

0.0173

0.0298

0.3800

-0.0259

-0.102i

-0.0627

-0.0368

-0.0271

-0.0185

-0.0120

-0.0061

0.0033

0.0178

0.0307

0.3900

-0.0266 -0.107Z

-0.0649

-0.0379

-0.0271

-0.0194

-0.0122

-0.006;

0.0035

0.0315
- -

-0.0504

0.0184

450

Appendix 2. Lee and Kesler Method

Table A2.2 (cont.)


Compressibility factor: corrective term

--

0.4000
0.4100
0.4200
0.4300
0.4400
0.4500
0.4600
0.4700
0.4800
0.4900
0.5000
0.5100
0.5200
0.5300
0.5374
0.5374
0.5400
0.5500
0.5600
0.5700
0.5800
0.5900
0.6oOo
0.6100
0.6200
0.6300

0.6400
0.6500
0.6600
0.6700
0.6800
0.6900
0.6963
0.6963
0.7000
0.7100
0.7200
0.7300
0.7400
0.7500
0.7600

--

---

0.8

0.9

0.94

0.98

1.02

1.04

1.06

-0.0272
-0.0279
-0.0285
-0.0292
-0.0298
-0.0305
-0.0311
-0.0318
-0.0324
-0.0331
-0.0337
-0.0344
-0.0350
-0.0356
-0.0361
-0.0361
-0.0363
-0.0369
-0.0376
-0.0382
-0.0388
-0.0395
-0.0401
-0.0407
-0.0414
-0.0420
-0.0426
-0.0433
-0.0439
-0.0445
-0.0452
-0.0458
-0.0462
-0.0462
-0.0464
-0.0470
-0.0477
-0.0483
-0.0489
-0.0495
-0.0502

-0.1118
-0.1166
-0.1217
-0.1271
-0.1328
-0.1389
-0.1454
-0.1525
-0.1601
-0.1685
-0.1777
-0.1880
-0.1997
-0.2133
-0.2248
-0.0361
-0.0363
-0.0368
-0.0374
-0.0380
-0.0385
-0.0391
-0.0396
-0.0402
-0.0407
-0.0413
-0.0418
-0.0424
-0.0429
-0.0435

-0.0671
-0.0694
-0.0718
-0.0742
-0.0767
-0.0793
-0.0819
-0.0847
-0.0875
-0,0905
-0.0935
-0.0967
-0.1001
-0.1035
-0.1062
-0.1062
-0.1072
-0.1111
-0.1152
-0.1195
-0.1242
-0.1292
-0.1346
-0.1405
-0.1471
-0.1543
-0.1625
-0.1718
-0.1828
-0.1959
-0.2124
-0.2341
-0.2526
-0.0483
-0.0485
-0.0489
-0.0493
-0.0496
-0.0500
-0.0504
-0.0508

-0.0390
-0.0401
-0.0412
-0.0424
-0.0435
-0.0447
-0.0458
-0.0470
-0.0482
-0.0494
-0.0507
-0.0519
-0.0532
-0.0545
-0.0555
-0.0555
-0.0558
-0.0571
-0.0585
-0.0598
-0.0612
-0.0627
-0.0641
-0.0656
-0.0671
-0.0687
-0.0703
-0.0720
-0.0737
-0.0755
-0.0773
-0.0792
-0.0805
-0.0805
-0.0812
-0.0833
-0.0855
-0.0878
-0.0902
-0.0929
-0.0957

-0.0285
-0.0293
-0.03CQ
-0.0308
-0.0315
-0.0323
-0.0330
-0.0337
-0.0345
-0.0352
-0.0360
-0.0367
-0.0375
-0.0382
-0.0388
-0.0388
-0.0390
-0.0397
-0.0405
-0.0413
-0.042a
-0.0428
-0.0435
-0.0443
-0,0450
-0.0458
-0.0466
-0.0473
-0.0481
-0.0488
-0.0496
-0.0504
-0.0508
-0.0508
-0.0511
-0.0519
-0.0526
-0.0534
-0.0542
-0.0549
-0.0557

-0.0198
-0.0203
-0.0207
-0.0212
-0.0216
-0.0220
-0.0224
-0.0229
-0.0233
-0.0237
-0.0241
-0.0245
-0.0249
-0.0253
-0.0255
-0.0255
-0.0256
-0.0260
-0.0264
-0.0267
-0.0270
-0.0274
-0.0277
-0.0280
-0.0283
-0.0286
-0.0288
-0.0291
-0.0293
-0.0295
-0.0297
-0.0299
-0.0300
-0.0300
-0.0301
-0.0302
-0.0304
-0.0304
-0.0305
-0.0306
-0.0306

-0.0125
-0.0127
-0.0129
-0.0131
-0.0133
-0.0135
-0.0137
-0.0138
-0.0140
-0.0142
-0.0143
-0.0144
-0,0146
-0.0147
-0.0147
-0.0147
-0.0148
-0.0149
-0.0149
-0.0150
-0.0151
-0.0151
-0.0151
-0.0151
-0.0151
-0.0151
-0.0150
-0.0150
-0.0149
-0.0148
-0.0146
-0.0145
-0,0144
4.0144
-0.0143
-0.0141
-0.0139
-0.0136
-0.0133
-0.0130
-0.0127

-0.0062
-0.0062
-0.0063
-0.0063
-0.0063
-0.0063
-0.0063
-0.0063
-0.0062
-0.0062
-0.0061
-0.0061

-0.0440

-0.0445
-0.0449
-0.0449
-0.0451
-0.0456
-0.0461
-0.0467
-0.0472
-0.0477
-0.0482

--

---

1.1

0.0038
0.0040
0.0042
0.0045
0.0048
0.0050
0.0053
0.0056
0.0059
0.0062
0.0066
0.0069
-0.o060 0.0073
-0.0059 0.0076
-0.0058 0.0079
-0.0058 0.0079
-0.0058 0.0080
-0.0057 0.0084
-0.0056 0.0088
-0.0054 0.0092
-0.0053 0.0096
-0.0051 0.0101
-0.0049 0.0106
-0.0047 0.0110
-0.0045 0.0115
-0.0042 0.0120
-0.oO40 0.0126
-0.0037 0.0131
-0.0034 0.0137
-0.0031 0.0142
-0.0027 0.0148
-0.0023 0.0154
-0.0021 0.0158
-0.0021 0.0158
-0.0020 0.0161
-0.0015 0.0167
-0.0011 0.0174
-0.0006 0.0181
-0.OOO1 0.0188
O.OOO4 0.0196
O.OOO9 0.0203

1.2

1.5

0.0190
0.0196
0.0202
0.0209
0.0215
0.0221
0.0228
0.0234
0.0241
0.0247
0.0254
0.0261
0.0268
0.0275
0.0280
0.0280
0.0282
0.0289
0.0296
0.0303
0.0311
0.0318
0.0326
0.0334
0.0341
0.0349
0.0357
0.0365
0.0373
0.0382
0.0390
0.0399
0.0404
0.0404
0.0407
0.0416
0.0425
0.0434
0.0443
0.0452
0.0461

0.0323
0.0332
0.0340
0.0349
0.0358
0.0366
0.0375
0.0383
0.0392
0.0401
0.0409
0.0418
0.0427
0.0435
0.0442
0.0442
0.0444
0.0453
0.0462
0.0470
0.0479
0.0488
0.0497
0.0506
0.0515
0.0524
0.0532
0.0541
0.0550
0.0559
0.0568
0.0577
0.0583
0.0583
0.0586
0.0595

0.0604
0.0613
0.0623
0.0632
0.0641

45 1

Appendix 2. Lee and Kesler Method

Table A2.2 (cont.)


Compressibility factor: corrective term

0.8

0.9

0.94

0.98

1.02

1.04

1.06

1.1

1.2

1.5

----0.7700
0.7800
0.7900
0.8OoO

0.8100
0.8200
0.8300
0.8400

0.8500
0.8600
0.8700
0.8800
0.8890
0.8890
0.8900
0.9000
0.9100
0.9200
0.9300
0.9400
0.9500
0.9600
0.9700
0.9800
0.9810
0.9820
0.9830
0.9840
0.9850
0.9860
0.9870
0.9880
0.9890
0.9900
0.9910
0.9920
0.9930
0.9940
0.9950
0.9960
0.9970

-0.0508
-0.0514
-0.0520
-0.0526
-0.0532
-0.0539
-0.0545
-0.0551
-0.0557
-0.0563
-0.0569
-0.0575
-0.0581
-0.0581
-0.0582
-0.0588
-0.0594

-0.0488
-0.0493
-0.0498
-0.0503
-0.0508
-0.0514
-0.0519
-0.0524
-0.0529
-0.0534
-0.0539
-0.0544
-0.0549
-0.0549
-0.0549
-0.0554
-0.0560
O
.
m-0.0565
-0.0606 -0.0570
-0.0612 -0.0575
-0.0618 -0.0580
-0.0624 -0.0585
-0.0630 -0.0590
-0.0636 -0.0595
-0.0637 -0.0595
-0.0637 -0.0596
-0.0638 -0.0596
-0.0638 -0.0597
-0.0639 -0.0597
-0,0640 -0.0598
-0,0640 -0.0598
-0.0641 -0.0599
-0,0641 -0.0599
-0.0642 -0.0600
-0.0643 -0.0600
-0.0643 -0.0601
-0.0644 -0.0601
-0.0644 -0.0602
-0,0645 -0.0602
-0.0646 -0.0602
-0.0646 -0.0603

-0.0512
-0.0516
-0.0520
-0.0524
-0.0528
-0.0532
-0.0536
-0.0540
-0.0544
-0.0547
-0.0551
-0.0555
-0.0559
-0.0559
-0.0559
-0.0563
-0.0567
-0.0571
-0.0575
-0.0579
-0.0583
-0.0587
-0.0591
-0.0595
-0.0596
-0.05%
-0.0597
-0.0597
-0.0597
-0.0598
-0.0598
-0.0599
-0.0599
-0.0599

-0.0988
-0.1021
-0.1058
-0.1100
-0.1147
-0.1203
-0.1270
-0.1352
-0.1461
-0.1614
-0.1866
-0.2529
-0.3192
-0.0705
-0.0703
-0.0690
-0.0679
-0.0670
-0.0663
-0.0658
-0.0653
4.0650
-0.0647
-0.0644
-0.0644
-0.0644
-0.0644
-0.0644
-0.0643
-0.0643
-0.0643
-0.0643
-0.0643
-0.0643
O
.
m-0.0642
-0.0600 -0.0642
- O . m -0.0642
-0.0601 -0.0642
-0.0601 -0.0642
-0.0602 -0.0642
-0.0602 -0.0642

-0.0565 -0.0306 -0.0123


-0.0572 -0.0305 -0.0119
-0.0580 -0.0304 -0.0114
-0.0588 -0.0303 -0.0109
-0.0595 -0.0301 -0.0103
-0.0603 -0.0299 -0.0097
-0.0611 -0.0296 -0.0091
-0.0618 -0.0292 -0.0084
-0.0626 -0.0288 -0.0076
-0.0634 -0.0284 -0.0068
-0.0642 -0.0278 -0.0059
-0.0649 -0.0272 -0.0050
-0.0656 -0.0265 -0.0041
-0.0656 -0.0265 -0.0041
-0.0657 -0.0264 -0.OO40
-0.0665 -0.0256 -0.0029
-0.0673 -0.0246 -0.0017
-0.0682 -0.0235 -0.ooo4
-0.0690 -0.0222 0.0010
-0.0699 -0.0208 0.0025
-0.0708 -0.0191 0.0041
-0.0719 -0.0172 0.0059
-0.0730 -0.0150 0.0078
-0.0745 -0.0125 0.0098
-0.0747 -0.0122 0.0100
-0.0749 -0.0119 0.0102
-0.0751 -0.0116 0.0105
-0.0753 -0.0114 0.0107
-0.0755 -0.0111 0.0109
-0.0758 -0.0108 0.0111
-0.0760 -0.0105 0.0113
-0.0763 -0.0102 0.0116
-0.0766 -0.0099 0.0118
-0.0769 -0.0096 0.0120
-0.0772 -0.0092 0.0122
-0.0776 -0.0089 0.0125
-0.0780 -0.0086 0.0127
-0.0785 -0.0083 0.0129
-0.0791 -0.0079 0.0132
-0.0798 -0.0076 0.0134
-0.0806 -0.0072 0.0137

0.0015
0.0021
0.0028
0.0035
0.0042
0.0050
0.0058
0.0066
0.0075
0.0085
0.0095
0.0105
0.0115
0.0115
0.0116
0.0128
0.0140
0.0153
0.0167
0.0181
0.0196
0.0212
0.0229
0.0247
0.0248
0.0250
0.0252
0.0254
0.0256
0.0258
0.0260
0.0261
0.0263
0.0265
0.0267
0.0269
0.0271
0.0273
0.0275
0.0277
0.0279

0.0211
0.0219
0.0228
0.0236
0.0245
0.0254
0.0264
0.0273
0.0283
0.0294
0.0304
0.0315
0.0325
0.0325
0.0327
0.0338
0.0350
0.0363
0.0375
0.0388
0.0402
0.0416
0.0430
0.0445
0.0446
0.0448

0.0449
0.0451
0.0452
0.0454
0.0455
0.0457
0.0458
0.0460
0.0461
0.0463
0.0465
0.0466
0.0468
0.0469
0.0471

0.0470
0.0480
0.0489
0.0499
0.0509
0.0519
0.0529
0.0539
0.0549
0.0560
0.0570
0.0581
0.0590
0.0590
0.0592
0.0603
0.0614
0.0625
0.0636
0.0647
0.0659
0.0671
0.0683
0.0694
0.0696
0.0697
0.0698
0.0699
0.0700
0.0702
0.0703
0.0704
0.0705
0.0707
0.0708
0.0709
0.0710
0.0711
0.0713
0.0714
0.0715

0.0650
0.0659
0.0668
0.0677
0.0687
0.0696
0.0705
0.0714
0.0724
0.0733
0.0742
0.0751
0.0760
0.0760
0.0761
0.0770
0.0779
0.0789
0.0798
0.0808
0.0817
0.0826
0.0836
0.0845
0.0846
0.0847
0.0848
0.0849
0.0850
0.0851
0.0852
0.0853
0.0854
0.0855
0.0856
0.0856
0.0857
0.0858
0.0859
0.0860
0.0861

------

452

Appendix 2. Lee and Kesler Method

Table A2.2 (cont. and end)


Compressibility factor: corrective term

0.8

0.9

-0.0647
-0.0647
-0.0648
-0.0649
-0.0649
-0.0650
-0.0650
-0.0651
-0.0652
-0.0652
-0.0653
-0.0653
-0.0654
-0.0655
-0.0655
-0.0656
-0.0656
-0.0657
-0.0658
-0.0658
-0.0659
-0.0659
-0.0660

-0.0603
-0.0604
-0.0604
-0.0605
-0.0605

0.9980
0.9990
1.m
1.0010
1.0020
1.0030
1.0040
1.0050
1.0060
1.0070
1.0080
1.oO90
1.0100
1.0110
1.0120
1.0130
1.0140
1.0150
1.0160
1.0170
1.0180
1.0190
1.0200
1.0300
LO400
1.0500
1.06oO
1.0700
1.0800
1.0900
1.1000
1.1100
1.1200
1.1300
1.1400
1.1500
1.1600
1.1700
1.1800
1.1900
1.2000

-0.0666
4.0672
-0.0678
-0.0684

-0.0690
-0.0696
-0.0702
-0.0708
-0.0714
-0.0720
4.0726
-0.0731
-0.0737
-0.0743
-0.0749
-0.0755
-0.0761
-0.0767

-0.0606

-0.0606
-0.0607
-0.0607
-0.0608
-0.0608
-0.0609
-0.0609
-0.0610
-0.0610
-0.0611
-0.0611
-0.0612
-0.0612
-0.0613
-0.0613
-0.0614
-0.0614
-0.0619
-0.0624
-0.0629
-0.0634
4.0639
-0.0644
-0.0649
-0.0653
-0.0658
-0.0663
-0.0668
-0.0673
-0.0677
-0.0682
-0.0687
-0.0692
-0.0697
-0.0701

0.94

0.98

-0.0602 -0.0641 -0.0817


-0.0603 -0.0641 -0.0833
-0.0603 -0.0641 -0.0879
-0.0604 -0.0641 -0.0807
-0.0604 -0.0641 -0.0783
-0.0604 -0.0641 -0.0767
-0.0605 -0.0641 -0.0754
-0.0605 -0.0641 -0.0743
-0.0606 -0.0641 -0.0734
-0.0606 -0.0640 -0.0726
-0.0606 -0.0640 -0.0719
-0.0607 -0.0640 -0.0713
-0.0607 -0.0640 -0.0707
-0.0608 -0.0640 -0.0702
-0.0608 -0.0640 -0.0697
-0.0608 -0.0640 -0.0693
-0.0609 -0.0640 -0.0689
-0.0609 -0.0640 -0.0685
-0.0610 -0.0640 -0.0682
-0.0610 -0.0640 -0.0678
-0.0610 -0.0640 -0.0675
-0.061 1 -0.0640 -0.0672
-0.061 1 -0.0640 -0.0670
-0.0615 -0.0639 -0.0649
-0.0619 -0.0639 -0.0636
-0.0623 -0.0640 -0.0626
-0.0627 -0.0640 -0.0620
-0.0631 -0.0641 -0.0615
-0.0635 -0.0642 -0.0611
-0.0639 -0.0643 -0.0608
-0.0643 -0.0644 -0.0607
-0.0647 -0.0645 -0.0605
-0.0651 -0.0646 -0.0605
-0.0655 -0.0648 -0.0604
-0.0659 -0.0650 -0.0604
-0.0663 -0.0651 -0.0604
-0.0667 -0.0653 -0.0605
-0.0671 -0.0655 -0.0606
-0.0675 -0.0657 -0.0607
-0.0679 -0.0659 -0.0608
-0.0683 -0.0661 -0.0609

1.02

1.04

4.0069 0.0139
-0.0065 0.0142
4.0062 0.0144
-0.0058 0.0147
4.0054 0.0149
4.0050 0.0152
4.0047 0.0154
4.0043 0.0157
4.0039 0.0159
4.0035 0.0162
4.0030 0.0165
4.0026 0.0167
-0.0022 0.0170
4.0018 0.0173
-0.0013 0.0175
4.OOO9 0.0178
4.oO04 0.0181
0.0184
O.oo00
0.0005 0.0186
0.0010 0.0189
0.0015 0.0192
0.0020 0.0195
0.0025 0.0198
0.0080 0.0228
0.0146 0.0261
0.0226 0.0297
0.0322 0.0336
0.0438 0.0378
0.0576 0.0423
0.0733 0.0473
0.0895 0.0527
0.1031 0.0585
0.1106 0.0648
0.1103 0.0714
0.1031 0.0785
0.0912 0.0859
0.0769 0.0934
0.0620 0.1009
0.0476 0.1082
0.0344 0.1148
0.0227 0.1205

1.06
0.0281
0.0283
0.0285
0.0287
0.0289
0.0291
0.0293
0.0295
0.0297
0.0299
0.0301
0.0303
0.0305
0.0307
0.0310
0.0312
0.0314
0.0316
0.0318
0.0320
0.0323
0.0325
0.0327
0.0350
0.0374
0.0399
0.0426
0.0454
0.0484

0.0515
0.0548
0.0582
0.0619
0.0656
0.0696
0.0738
0.0781
0.0825
0.0872
0.0919
0.0968

1.1

0.0472
0.0474
0.0476
0.0477
0.0479
0.0480
0.0482
0.0484
0.0485

0.0487
0.0488
0.0490
0.0492
0.0493
0.0495
0.0497
0.0498
0.0500
0.0501
0.0503
0.0505
0.0506
0.0508
0.0525
0.0543
0.0561
0.0579
0.0598
0.0618
0.0638
0.0659
0.0680
0.0702
0.0724
0.0747
0.0771
0.0795
0.0820
0.0845
0.0871
0.0897

1.2
0.0716
0.0718
0.0719
0.0720
0.0721
0.0722
0.0724
0.0725
0.0726
0.0727
0.0729
0.0730
0.0731
0.0732
0.0734
0.0735
0.0736
0.0737
0.0739
0.0740
0.0741
0.0742
0.0744
0.0756
0.0769
0.0782
0.0795
0.0808
0.0821
0.0835
0.0848
0.0862
0.0876
0.0890
0.0904
0.0918
0.0933
0.0947
0.0962
0.0976
0.0991

1.5

0.0862
0.0863
0.0864
0.0865
0.0866
0.0867
0.0868
0.0869
0.0870
0.0871
0.0872
0.0873
0.0874
0.0874
0.0875
0.0876
0.0877
0.0878
0.0879
0.0880
0.0881
0.0882
0.0883
0.0892
0.0902
0.0911
0.0921
0.0931
0.0940
0.0950
0.0959
0.0969
0.0978
0.0988
0.0997
0.1007
0.1017
0.1026
0.1036
0.1045
0.1055

453

Appendix 2. Lee and Kesler Method

Table A23
Adimensional residual enthalpy of the simple fluid (with opposite sign)

0.8

0.0100
0.0100
0.0200
0.0300
0.0400
0.0500
0.06oO
0.0700
0.0800
0.09oO
0.1000
0.1100
0.1200
0.1300
0.1400
0.1500
0.1600
0.1700
0.1800
0.1900
0.2000
0.2100
0.2200
0.2300
0.2400
0.2500
0.2563
0.2563
0.2600
0.2700
0.2800
0.2900
0.3000
0.3100
0.3200
0.3300
0.3400
0.3500
0.3600
0.3700
0.3800

0.0152
0.0152
0.0305
0.0460
0.0617
0.0776
0.0937
0.1100
0.1264
0.1432
0.1601
0.1773
0.1947
0.2124
0.2304
0.2486
0.2672
0.2860
0.3053
0.3248
0.3448
0.3651
0.3859
0.4071
0.4288
0.4511
0.4653
4.5092
4.5091
4.5090
4.5089
4.5087
4.5086
4.5085
4.5083
4.5082
4.5081
4.5079
4.5078
4.5076
4.5075

0.9

0.94

0.98

1.02

0.0122
0.0122
0.0244
0.0368
0.0492
0.0618
0.0744
0.0872
0.1001
0.1130
0.1261
0.1393
0.1527
0.1661
0.1797
0.1934
0.2073
0.2212
0.2354
0.2496
0.2641
0.2787
0.2934
0.3083
0.3234
0.3387
0.3484
0.3484
0.3541
0.3698
0.3857
0.4017
0.4180
0.4346
0.4513
0.4683
0.4856
0.5032
0.5210
0.5391
0.5576

0.0112
0.0112
0.0225
0.0339
0.0454
0.0569
0.0685
0.0802
0.0920
0.1038
0.1158
0.1278
0.1400
0.1522
0.1646
0.1770
0.1895
0.2022
0.2149
0.2278
0.2407
0.2538
0.2670
0.2803
0.2938
0.3074
0.3160
0.3160
0.3211
0.3349
0.3489
0.3631
0.3774
0.3918
0.4065
0.4212
0.4362
0.4513
0.4667
0.4822
0.4979

0.0104
0.0104
0.0209
0.0314
0.0419
0.0526
0.0633
0.0741
0.0849
0.0958
0.1068
0.1178
0.1290
0.1402
0.1514
0.1628
0.1742
0.1857
0.1973
0.2090
0.2208
0.2326
0.2445
0.2566
0.2687
0.2809
0.2886
0.2886
0.2932
0.3057
0.3182
0.3308
0.3435
0.3564
0.3694
0.3824
0.3956
0.4090
0.4224
0.4360
0.4497

1.04

1.06

1.1

1.2

1.5

0 . m

0.m
0.0084
0.0169
0.0253
0.0339
0.0424
0.0510
0.05%
0.0683
0.0770
0.0857
0.0945
0.1033
0.1121
0.1210
0.1299
0.1388
0.1478
0.1569
0.1659
0.1751
0.1842
0.1934
0.2027
0.2120
0.2213
0.2272
0.2272
0.2307
0.2401
0.2495
0.2591
0.2686
0.2782
0.2879
0.2976
0.3073
0.3172
0.3270
0.3369
0.3469

0.0072
0.0072
0.0144
0.0216
0.0289
0.0361
0.0434
0.0507
0.0581
0.0654
0.0728
0.0802
0.0876
0.0951
0.1025
0.1100
0.1175
0.1251
0.1326
0.1402
0.1478
0.1555
0.1631
0.1708
0.1785
0.1862
0.1911
0.1911
0.1940
0.2018
0.20%
0.2174
0.2253
0.2331
0.2411
0.2490
0.2570
0.2650
0.2730
0.2810
0.2891

0.0048
0.0048
0.0096
0.0144
0.0192
0.0240
0.0288
0.0336
0.0385
0.0433
0.0481
0.0530
0.0578
0.0627
0.0675
0.0724
0.0772
0.0821
0.0870
0.0919
0.0968
0.1017
0.1066
0.1115
0.1164
0.1213
0.1244
0.1244
0.1262
0.1311
0.1361
0.1410
0.1460
0.1509
0.1559
0.1608
0.1658
0.1707
0.1757
0.1807
0.1857

0.0100
0.0100
0.0201
0.0302
0.0404
0.0506
0.0609
0.0713
0.0817
0.0922
0.1027
0.1133
0.1240
0.1347
0.1455
0.1564
0.1673
0.1783
0.1894
0.2006
0.2118
0.2231
0.2345
0.2460
0.2575
0.2692
0.2765
0.2765
0.2809
0.2927
0.3046
0.3166
0.3286
0.3408
0.3531
0.3654
0.3779
0.3905
0.4032
0.4160
0.4289

0.0097
0.0097
0.0194
0.0291
0.0389
0.0488
0.0587
0.0687
0.0787
0.0888
0.0989
0.1091
0.1193
0.1296
0.1400
0.1504
0.1609
0.1714
0.1821
0.1927
0.2035
0.2143
0.2252
0.2361
0.2471
0.2582
0.2652
0.2652
0.2694
0.2806
0.2920
0.3034
0.3148
0.3264
0.3380
0.3498
0.3616
0.3735
0.3855
0.3976
0.4098

0.0093
0.0093
0.0187
0.0281
0.0376
0.0471
0.0566
0.0662
0.0759
0.0856
0.0953
0.1051
0.1149
0.1248
0.1348
0.1448
0.1549
0.1650
0.1752
0.1854
0.1957
0.2060
0.2164
0.2269
0.2374
0.2480
0.2547
0.2547
0.2587
0.2694
0.2802
0.2911
0.3020
0.3130
0.3241
0.3353
0.3465
0.3578
0.3692
0.3807
0.3922

0 . m

0.0180
0.0271
0.0363
0.0454
0.0546
0.0639
0.0732
0.0825
0.0919
0.1013
0.1108
0.1203
0.1299
0.1395
0.1492
0.1589
0.1687
0.1785
0.1884
0.1983
0.2083
0.2183
0.2284
0.2385
0.2449
0.2449
0.2487
0.2590
0.2693
0.2797
0.2901
0.3006
0.3112
0.3218
0.3325
0.3433
0.3541
0.3650
0.3760

454

Appendix 2. Lee and Kesler Method

Table A23 (cont.)


Adimensional residual enthalpy of the simple fluid (with opposite sign)

0.3900
0.4000
0.4100
0.4200
0.4300
0.4400
0.4500
0.4600
0.4700
0.4800
0.4900
0.5000
0.5100
0.5200
0.5300
0.5374
0.5374
0.5400
0.5500
0.5600
0.5700
0.5800
0.5900

4.5073
4.5071
4.5070
4.5068
4.5067
4.5065
4.5063
4.5062
4.5060
4.5058
4.5056
4.5054
4.5053
4.5051
4.5049
4.5047
4.5047
4.5047
4.5045
4.5043
4.5041
4.5039
4.5037
4.5035
4.5033
4.5031
4.5029
4.5027
4.5025
4.5022
4.5020
4.5018
4.5016
4.5014
4.5014
4.5013
4.5011
4.5009
4.5007
4.5004
4.5002
4.5000
4.4997
4.4995
4.4992
-

0.6OOo
0.6100
0.6200
0.6300
0.6400
0.6500
0.6600
0.6700
0.6800
0.6900
0.6963
0.6963
0.7000
0.7100
0.7200
0.7300
0.7400
0.7500
0.7600
0.7700
0.7800
0.7900

0.8

--0.94

0.9

0.98

1.02

1.04

D.5764
D.5956
D.6151
D.6350
D.6554
D.6762
D.6975
D.7194
D.7418
0.7648
D.7885
0.8129
0.8380
0.8641
0.8912
D.9119
4.0662
4.0666
4.0679
4.0691
4.0704
4.0716
4.0728
4.0739
4.0751
4.0762
4.0773
4.0784
4.0795
4.0805
4.0816
4.0826
4.0836
4.0842
4.0842
4.0846
4.0856
4.0865
4.0874
4.0884
4.0893
4.0902
4.0910
4.0919
4.0927
-

0.5138
0.5300
0.5463
0.5629
0.5798
0.5969
0.6143
0.6319
0.6499
0.6682
0.6868
0.7057
0.7251
0.7448
0.7649
0.7801
0.7801
0.7855
0.8065
0.8280
0.8501
0.8728
0.8961
0.9201
0.9448
0.9704
0.9968
1.0243
1.0529
1.0827
1.1140
1.1470
1.1820
1.2053
3.8154
3.8167
3.8202
3.8236
3.8269
3.8302
3.8333
3.8364
3.8394
3.8423
3.8451
~

1.06

1.1

1.2

1.5

- ---

0.4636
0.4776
0.4918
0.5061
0.5205
0.5352
0.5500
0.5649
0.5801
0.5954
0.6110
0.6267
0.6427
0.6589
0.6752
0.6876
0.6876
0.6919
0.7088
0.7259
0.7433
0.7610
0.7790
0.7973
0.8159
0.8349
0.8543
0.8740
0.8941
0.9147
0.9357
0.9572
0.9793
0.9936
0.9936
1.0019
1.0251
1.0490
1.0736
1.0990
1.1252
1.1524
1.1805
1.2099
1.2406
-

0.4419
0.4551
D.4683
0.4817
0.4953
0.5089
0.5227
0.5367
0.5508
0.5650
0.5794
0.5939
0.6087
0.6236
0.6386
0.6499
0.6499
0.6539
0.6693
0.6850
0.7008
0.7169
0.7332
0.7497
0.7664
0.7834
0.8007
0.8182
0.8360
0.8542
0.8726
0.8914
0.9105
0.9228
0.9228
0.9299
0.9498
0.9701
0.9908
1.0120
1.0336
1.0558
1.0785
1.1019
1.1259
-

3.4221
D.4345
D.4470
0.4596
D.4723
0.4851
D.4980
D.5111
D.5242
D.5375
D.5510
0.5645
0.5782
0.5920
0.6060
0.6164
0.6164
0.6201
0.6343
0.6488
0.6633
0.6781
0.6930
0.7081
0.7234
0.7388
0.7545
0.7703
0.7864
0.8027
0.8192
0.8360
0.8529
0.8639
0.8639
0.8702
0.8877
0.9055
0.9236
0.9420
0.96M
0.9798
0.9992
1.0190
1.0392
-

0.4038
0.4156
0.4274
0.4393
0.4512
0.4633
0.4755
0.4878
0.5001
0.5126
0.5252
0.5379
0.5507
0.5636
0.5766
0.5863
0.5863
0.5897
0.6030
0.6164
0.6299
0.6435
0.6573
0.6713
0.6853
0.6995
0.7139
0.7284
0.7431
0.7579
0.7730
0.7882
0.8035
0.8134
0.8134
0.8191
0.8349
0.8508
0.8670
0.8834

0.9OoO
0.9169
0.9340
0.9513
0.9689
-

0.3870
0.3981
0.4093
0.4206
0.4319
0.4433
0.4548
0.4664
0.4781
0.4898
0.5017
0.5136
0.5256
0.5377
0.5499
0.5591
0.5591
0.5623
0.5747
0.5872
0.5998
0.6125
0.6253
0.6383
0.6513
0.6645
0.6778
0.6912
0.7047
0.7184
0.7322
0.7462
0.7602
0.7693
0.7693
0.7745
0.7889
0.8034
0.8181
0.8329
0.8480
0.8631
0.8785
0.8941
0.9098
-

D.3569
D.3670
D.3771
D.3873
D.3975
D.4078
D.4182
D.4286
0.4391
0.4496
0.4602
D.4709
0.4816
0.4924
0.5033
0.5114
0.5114
0.5142
0.5252
0.5363
0.5474
0.5586
0.5699
0.5813
0.5927
0.6043
0.6159
0.6276
0.6393
0.6512
0.6631
0.6752
0.6873
0.6950
0.6950
0.6995
0.7118
0.7242
0.7367
0.7493
0.7620
0.7748
0.7877
0.8007
0.8139
-

0.2972
0.3053
0.3135
0.3217
0.3299
0.3382
0.3464
0.3547
0.3631
0.3715
0.3799
0.3883
0.3967
0.4052
0.4138
0.4201
0.4201
0.4223
0.4309
0.4395
0.4482
0.4569
0.4656
0.4744
0.4832
0.4920
0.5009
0.5098
0.5187
0.5277
0.5367
0.5457
0.5548
0.5606
0.5606
0.5639
0.5731
0.5823
0.5915
0.6008
0.6101
0.6195
0.6289
0.6383
0.6478
-

0.1907
0.1957
0.2007
0.2057
0.2107
0.2157
0.2207
0.2258
0.2308
0.2358
0.2409
0.2459
0.2510
0.2560
0.2611
0.2649
0.2649
0.2662
0.2712
0.2763
0.2814
0.2865
0.2916
0.2967
0.3018
0.3069
0.3120
0.3172
0.3223
0.3274
0.3326
0.3377
0.3428
0.3461
0.3461
0.3480
0.3531
0.3583
0.3635
0.3686
0.3738
0.3790
0.3842
0.3894
0.3946
-

455

Appendix 2. Lee and Kesler Method

Table A23 (cont. and end)


Adimensional residual enthalpy of the simple fluid (with opposite sign)

0.8000
0.8100
0.8200
0.8300

0.8400
0.8500
0.8600
0.8700
0.8800
0.8890
0.8890
0.8900
0.9OoO

0.9100
0.9200
0.9300
0.9400
0.9500
0.9600
0.9700
0.9800
0.9900
1.oooO
1.0100
1.0200
1.0300
1.0400
1.0500
1.0600
1.0700
1.0800
1.0900
1.1000
1.1100
1.1200
1.1300
1.1400
1.1500
1.1600
1.1700
1.1800
1.1900
1.2000
1.2000

0.8

0.9

0.94

0.98

1.02

1.04

1.06

1.1

1.2

1.5

4.4990
4.4987
4.4985
4.4982
4.4980
4.4977
4.4975
4.4972
4.4970
4.4967
4.4967
4.4967
4.4964
4.4962
4.4959
4.4957
4.4954
4.4951
4.4948
4.4946
4.4943
4.4940
4.4937
4.4934
4.4932
4.4929
4.4926
4.4923
4.4920
4.4917
4.4914
4.4911
4.4908
4.4905
4.4902
4.4899
4.4896
4.4893
4.4890
4.4887
4.4884
4.4881
4.4878
4.4878

4.0936
4.0944
4.0952
4.0960
4.0968
4.0975
4.0983
4.0990
4.0998
4.1004
4.1004
4.1005
4.1012
4.1019
4.1026
4.1032
4.1039
4.1046
4.1052
4.1058
4.1064
4.1070
4.1076
4.1082
4.1088
4.1094
4.1099
4.1105
4.1110
4.1115
4.1121
4.1126
4.1131
4.1136
4.1140
4.1145
4.1150
4.1155
4.1159
4.1163
4.1168
4.1172
4.1176
4.1176

3.8479
3.8506
3.8533
3.8559
3.8584
3.8609
3.8633
3.8656
3.8680
3.8700
3.8700
3.8702
3.8724
3.8746
3.8768
3.8788
3.8809
3.8829
3.8849
3.8868
3.8887
3.8906
3.8924
3.8942
3.8959
3.8977
3.8993
3.9010
3.9026
3.9042
3.9058
3.9074
3.9089
3.9104
3.9119
3.9133
3.9147
3.9161
3.9175
3.9189
3.9202
3.9215
3.9228
3.9228

1.2727
1.3065
1.3424
1.3805
1.4215
1.4659
1.5148
1.5697
1.6331
1.7014
3.4154
3.4173
3.4344
3.4496
3.4635
3.4761
3.4878
3.4986
3.5088
3.5183
3.5272
3.5357
3.5438
3.5514
3.5587
3.5656
3.5723
3.5787
3.5849
3.5908
3.5965
3.6019
3.6072
3.6124
3.6173
3.6221
3.6268
3.6313
3.6357
3.6399
3.6441
3.6481
3.6520
3.6520

1.1505
1.1760
1.2023
1.2294
1.2576
1.2869
1.3174
1.3493
1.3827
1.4142
1.4142
1.4180
1.4553
1.4950
1.5376
1.5838
1.6344
1.6908
1.7551
1.8310
1.9260
2.0615
2.5738
3.0081
3.0924
3.1464
3.1868
3.2193
3.2466
3.2702
3.2909
3.3095
3.3263
3.3417
3.3558
3.3690
3.3812
3.3926
3.4034
3.4136
3.4232
3.4324
3.4410
3.4410

1.0597
1.0808
1.1023
1.1242
1.1467
1.1698
1.1935
1.2178
1.2428
1.2658
1.2658
1.2685
1.2951
1.3225
1.3509
1.3803
1.4109
1.4428
1.4760
1.5109
1.5476
1.5863
1.6273
1.6711
1.7180
1.7688
1.8240
1.8848
1.9524
2.0282
2.1140
2.2110
2.3182
2.4309
2.5409
2.6413
2.7287
2.8031
2.8660
2.9193
2.9650
3.0045
3.0391
3.0391

0.9868
1.0050
1.0235
1.0423
1.0614
1.0809
1.1007
1.1209
1.1415
1.1603
1.1603
1.1625
1.1839
1.2058
1.2282
1.2512
1.2746
1.2987
1.3234
1.3487
1.3748
1.4016
1.4292
1.4578
1.4872
1.5177
1.5493
1.5820
1.6161
1.6516
1.6886
1.7272
1.7677
1.8101
1.8545
1.9012
1.9502
2.0015
2.0550
2.1106
2.1679
2.2263
2.2853
2.2853

0.9258
0.9419
0.9582
0.9748
0.9916
1.0086
1.0259
1.0434
1.0612
1.0773
1.0773
1.0792
1.0975
1.1161
1.1351
1.1543
1.1738
1.1937
1.2140
1.2346
1.2556
1.2770
1.2989
1.3212
1.3439
1.3671
1.3909
1.4152
1.4400
1.4655
1.4916
1.5183
1.5457
1.5738
1.6027
1.6323
1.6628
1.6941
1.7263
1.7594
1.7933
1.8282
1.8640
1.8640

0.8271
0.8404
0.8539
0.8675
0.8812
0.8951
0.9090
0.9231
0.9374
0.9502
0.9502
0.9517
0.9663
0.9809
0.9957
1.0107
1.0258
1.0411
1.0565
1.0721
1.0879
1.1038
1.1199
1.1363
1.1528
1.1694
1.1863
1.2034
1.2207
1.2382
1.2559
1.2739
1.2920
1.3104
1.3291
1.3480
1.3671
1.3865
1.4061
1.4260
1.4462
1.4666
1.4873
1.4873

0.6573
0.6668
0.6764
0.6861
0.6957
0.7055
0.7152
0.7250
0.7349
0.7438
0.7438
0.7448
0.7547
0.7647
0.7748
0.7849
0.7950
0.8052
0.8154
0.8256
0.8360
0.8463
0.8567
0.8672
0.8777
0.8883
0.8989
0.9095
0.9202
0.9310
0.9418
0.9527
0.9636
0.9746
0.9856
0.9967
1.0078
1.0190
1.0302
1.0415
1.0528
1.0642
1.0756
1.0756

0.3998
0.4050
0.4102
0.4154
0.4206
0.4259
0.4311
0.4363
0.4416
0.4463
0.4463
0.4468
0.4521
0.4573
0.4626
0.4678
0.4731
0.4784
0.4836
0.4889
0.4942
0.4995
0.5048
0.5101
0.5154
0.5207
0.5260
0.5313
0.5366
0.5419
0.5473
0.5526
0.5579
0.5633
0.5686
0.5739
0.5793
0.5846
0.5900
0.5954
0.6007
0.6061
0.6115
0.6115

456

Appendix 2. Lee and Kesler Method

Table A2.4
Adimensional residual enthalpy (with opposite sign): corrective term

0.8

0.9

0.94

0.98

1.02

1.04

1.06

1.1

1.2

1.5

----0.0100
0.0100
0.0200
0.0300
0.04oO
0.0500
0.06oO
0.0700

0.0800
0.09oO
0.1000
0.1100
0.1200
0.1300
0.1400
0.1500
0.1600
0.1700
0.1800
0.1900
0.2000
0.2100
0.2200
0.2300
0.2400
0.2500
0.2563
0.2563
0.2600
0.2700
0.2800
0.2900
0.3000
0.3100
0.3200
0.3300

0.3400
0.3500
0.3600
0.3700
0.3800
0.3900
0.4OoO

0.4100

0.0212
0.0212
0.0429
0.0649
0.0874
0.1104
0.1339
0.1579
0.1825
0.2077
0.2336
0.2601
0.2874
0.3155
0.3445
0.3744
0.4053
0.4374
0.4707
0.5053
0.5415
0.5794
0.6192
0.6612
0.7057
0.7530
0.7844
5.2628
5.2629
5.2631
5.2633
5.2635
5.2637
5.2639
5.2641
5.2643
5.2645
5.2647
5.2649
5.2651
5.2653
5.2656
5.2658
5.2660

0.0137
0.0137
0.0275
0.0414
0.0556
0.0698
0.0843
0.0989
0.1137
0.1286
0.1438
0.1591
0.1747
0.1905
0.2065
0.2227
0.2391
0.2558
0.2728
0.2900
0.3075
0.3253
0.3434
0.3619
0.3807
0.3998
0.4121
0.4121
0.4194
0.4393
0.4597
0.4805
0.5018
0.5237
0.5460
0.5690
0.5926
0.6169
0.6419
0.6677
0.6944
0.7221
0.7507
0.7806

0.0116
0.0116
0.0232
0.0349

0.0468
0.0587
0.0708
0.0830
0.0952
0.1076
0.1201
0.1327
0.1455
0.1584
0.1714
0.1845
0.1978
0.2112
0.2248
0.2385
0.2524
0.2665
0.2807
0.2951
0.3097
0.3245
0.3339
0.3339
0.3395
0.3546
0.3700
0.3856
0.4015
0.4176
0.4339
0.4505
0.4674
0.4846
0.5021
0.5199
0.5380
0.5566
0.5754
0.5947

0.0098
0.0098
0.0197
0.0296
0.0396
0.0496
0.0597
0.0699
0.0802
0.0905
0.1009
0.1114
0.1219
0.1325
0.1432
0.1540
0.1649
0.1758
0.1869
0.1980
0.2092
0.2205
0.2319
0.2434
0.2550
0.2667
0.2741
0.2741
0.2785
0.2904
0.3024
0.3146
0.3268
0.3392
0.3517
0.3643
0.3771
0.3900
0.4030
0.4162
0.4295
0.4430
0.4567
0.4705

0.0090
0.0090
0.0181
0.0272
0.0364
0.0457
0.0549
0.0643
0.0737
0.0831
0.0926
0.1022
0.1118
0.1215
0.1312
0.1410
0.1509
0.1608
0.1708
0.1809
0.1910
0.2012
0.2114
0.2218
0.2322
0.2427
0.2493
0.2493
0.2532
0.2639
0.2746
0.2854
0.2963
0.3M2
0.3183
0.3294
0.3407
0.3520
0.3634
0.3749
0.3866
0.3983
0.4101
0.4221

0.0083
0.0083
0.0167
0.0251
0.0336
0.0420
0.0506
0.0591
0.0677
0.0764
0.0851
0.0938
0.1026
0.1115
0.1203
0.1293
0.1382
0.1472
0.1563
0.1654
0.1746
0.1838
0.1931
0.2024
0.2118
0.2212
0.2271
0.2271
0.2307
0.2402
0.2498
0.2595
0.2692
0.2789
0.2888
0.2987
0.3086
0.3186
0.3287
0.3389
0.3491
0.3594
0.3698
0.3802

0.0077
0.0077
0.0154
0.0231
0.0309
0.0387
0.0466
0.0544
0.0623
0.0703
0.0782
0.0862
0.0943
0.1023
0.1104
0.1186
0.1268
0.1350
0.1432
0.1515
0.1598
0.1682
0.1765
0.1850
0.1934
0.2019
0.2073
0.2073
0.2105
0.2191
0.2277
0.2363
0.2450
0.2538
0.2626
0.2714
0.2803
0.2892
0.2981
0.3071
0.3162
0.3253
0.3344
0.3436

0.0071
0.0071
0.0142
0.0213
0.0285
0.0357
0.0429
0.0501
0.0574
0.0647
0.0720
0.0793
0.0867
0.0940
0.1015
0.1089
0.1163
0.1238
0.1313
0.1389
0.1464
0.1540
0.1616
0.1692
0.1769
0.1846
0.1894
0.1894
0.1923
0.2001
0.2078
0.2156
0.2234
0.2313
0.2392
0.2471
0.2550
0.2630
0.2710
0.2790
0.2870
0.2951
0.3032
0.3113

0.0060

O.Oo40

O.Oo60

0.0040

0.0121
0.0182
0.0243
0.0304
0.0365
0.0426
0.0487
0.0548
0.0610
0.0672
0.0733
0.0795
0.0857
0.0920
0.0982
0.1044
0.1107
0.1169
0.1232
0.1295
0.1358
0.1421
0.1484
0.1547
0.1587
0.1587
0.1610
0.1674
0.1737
0.1801
0.1865
0.1928
0.1992
0.2056
0.2120
0.2185
0.2249
0.2313
0.2378
0.2442
0.2507
0.2571

0.0081
0.0121
0.0162
0.0202
0.0243
0.0283
0.0323
0.0363
0.0404

0.0010
0.0010
0.0019
0.0029
0.0038
0.0048
0.0057

0.0066

0.0076
0.0085
0.0094
0.0444 0.0103
0.0484 0.0111
0.0524 0.0120
0.0564 0.0129
0.0604 0.0137
0.0644 0.0146
0.0684 0.0154
0.0724 0.0162
0.0764 0.0171
0.0804 0.0179
0.0843 0.0187
0.0883 0.0195
0.0923 0.0203
0.0962 0.0210
0.1002 0.0218
0.1026 0.0223
0.1026 0.0223
0.1041 0.0225
0.1080 0.0233
0.1120 0.0240
0.1159 0.0248
0.1198 0.0255
0.1237 0.0262
0.1276 0.0269
0.1315 0.0276
0.1354 0.0282
0.1392 0.0289
0.1431 0.02%
0.1469 0.0302
0.1508 0.0309
0.1546 0.0315
0.1584 0.0321
0.1622 0.0327

-----------

45 7

Appendix 2. Lee and Kesler Method

Table A2.4 (cont.)


Adimensional residual enthalpy (with opposite sign): correctiveterm

x
0.4200
0.4300
0.4400
0.4500
0.4600
0.4700
0.4800
0.4900
0.5000
0.5100
0.5200
0.5300
0.5374
0.5374
0.5400
0.5500
0.5600
0.5700
0.5800
0.5900
0.6oOo

0.6100
0.6200
0.6300
0.6400
0.6500
0.6600

0.6700
0.6800
0.6900
0.6963
0.6963
0.7000
0.7100
0.7200
0.7300
0.7400
0.7500
0.7600
0.7700
0.7800
0.7900
0.8000
0.8100

0.8

0.9

0.94

5.2663
5.2665
5.2667
5.2670
5.2672
5.2675
5.2677
5.2680
5.2683
5.2685
5.2688
5.2690
5.2692
5.2692
5.2693
5.2696
5.2699
5.2702
5.2704
5.2707
5.2710
5.2713
5.2716
5.2719
5.2722
5.2725
5.2728
5.2731
5.2734
5.2737
5.2739
5.2739
5.2740
5.2744
5.2747
5.2750
5.2753
5.2756
5.2760
5.2763
5.2766
5.2770
5.2773
5.2776

0.8117
0.8442
0.8783
0.9143
0.9523
0.9926
1.0358
1.0822
1.1326
1.1877
1.2489
1.3178
1.3755
4.2584
4.2582
4.2575
4.2568
4.2562
4.2555
4.2550
4.2544
4.2539
4.2534
4.2529
4.2524
4.2520
4.2516
4.2512
4.2509
4.2505
4.2503
4.2503
4.2502
4.2500
4.2497
4.2495
4.2492
4.2490
4.2488
4.2487
4.2485
4.2484
4.2483
4.2482

0.6144
0.6346
0.6553
0.6764
0.6981
0.7205
0.7434
0.7671
0.7915
0.8167
0.8428
0.8699
0.8908
0.8908
0.8981
0.9275
0.9583
0.9906
1.0246
1.0607
1.0990
1.1400
1.1842
1.2322
1.2851
1.3440
1.4107
1.4883
1.5814
1.6995
1.7959
3.8587
3.8578
3.8553
3.8529
3.8507
3.8486
3.8467
3.8448
3.8431
3.8415
3.8400
3.8386
3.8373

0.98

0.4845
0.4986
0.5130
0.5275
0.5422
0.5572
0.5724
0.5878
0.6034
0.6193
0.6354
0.6518
0.6642
0.6642
0.6686
0.6856
0.7029
0.7206
0.7386
0.7570
0.7758
0.7950
0.8147
0.8349
0.8556
0.8769
0.8987
0.9213
0.9445
0.9686
0.9843
0.9843
0.9935
1.0194
1.0464
1.0745
1.1040
1.1351
1.1679
1.2027
1.2400
1.2802
1.3239
1.3720

1.02

1.04

1.06

1.1

-0.4341
0.4463
0.4586
0.4710
0.4835
0.4962
0.5090
0.5219
0.5350
0.5482
0.5616
0.5752
0.5853
0.5853
0.5888
0.6027
0.6168
0.6310
0.6454
0.6600
0.6748
0.6898
0.7051
0.7205
0.7363
0.7522
0.7685
0.7850
0.8018
0.8189
0.8299
0.8299
0.8363
0.8541
0.8722
0.8907
0.9097
0.9290
0.9489
0.9692
0.9901
1.0115
1.0336
1.0563

--- --

0.3907
0.4013
0.4119
0.4226
0.4335
0.4443
0.4553
0.4664
0.4775
0.4887
0.5001
0.5115
0.5200
0.5200
0.5230
0.5346
0.5463
0.5581
0.5699
0.5819
0.5940
0.6063
0.6186
0.6310
0.6436
0.6562
0.6690
0.6820
0.6950
0.7082
0.7166
0.7166
0.7215
0.7349
0.7485
0.7622
0.7761
0.7901
0.8043
0.8186
0.8331
0.8478
0.8626
0.8776

0.3528
0.3621
0.3714
0.3808
0.3902
0.3997
0.4092
0.4187
0.4284
0.4380
0.4477
0.4575
0.4648
0.4648
0.4673
0.4772
0.4871
0.4971
0.5071
0.5172
0.5273
0.5375
0.5478
0.5581
0.5684
0.5788
0.5893
0.5998
0.6104
0.6210
0.6278
0.6278
0.6317
0.6424
0.6532
0.6640

0.6749
0.6858
0.6968
0.7078
0.7188
0.7299
0.7410
0.7521

0.3195
0.3277
0.3359
0.3441
0.3524
0.3607
0.3690
0.3774
0.3858
0.3942
0.4026
0.4111
0.4174
0.4174
0.4195
0.4281
0.4366
0.4452
0.4537
0.4624
0.4710
0.4797
0.4883
0.4971
0.5058
0.5145
0.5233
0.5321
0.5409
0.5497
0.5553
0.5553
0.5585
0.5673
0.5762
0.5850
0.5939
0.6028
0.6116
0.6205
0.6294
0.6382
0.6470
0.6559

1.2

1.5

-0.2636
0.2701
0.2766
0.2830
0.2895
0.2960
0.3025
0.3090
0.3155
0.3220
0.3285
0.3350
0.3399
0.3399
0.3415
0.3480
0.3546
0.3611
0.3676
0.3740
0.3805
0.3870
0.3935
0.4OoO

0.4064
0.4129
0.4193
0.4258
0.4322
0.4386
0.4426
0.4426
0.4449
0.4513
0.4577
0.4640
0.4703
0.4765
0.4828
0.4890
0.4952
0.5013
0.5075
0.5135

0.1660
0.1698
0.1735
0.1773
0.1810
0.1848
0.1885
0.1922
0.1958
0.1995
0.2032
0.2068
0.2095
0.2095
0.2104
0.2140
0.2176
0.2212
0.2247
0.2282
0.2318
0.2352
0.2387
0.2422
0.2456
0.2490
0.2524
0.2557
0.2591
0.2624
0.2645
0.2645
0.2657
0.2689
0.2721
0.2753
0.2785
0.2817
0.2848
0.2879
0.2909
0.2940
0.2970
0.2999

0.0333
0.0339
0.0345
0.0350
0.0356
0.0362
0.0367
0.0372
0.0377
0.0382
0.0387
0.0392
0.03%
0.0396
0.0397
0.0402
0.0406
0.0410
0.0415
0.0419
0.0423
0.0427
0.0431
0.0435
0.0438
0.0442
0.0445
0.0448
0.0452
0.0455
0.0457
0.0457
0.0458
0.0461
0.0463
0.0466
0.0468

0.0471
0.0473
0.0475
0.0477
0.0479
0.0481
0.0483

458

Appendix 2. L e e and Kesler Method

Table A2.4 (cont. and end)


Adimensional residual enthalpy (with opposite sign): corrective term

0.8

0.9

0.94

0.98

1.02

1.04

1.06

0.8200
0.8300

0.8400
0.8500
0.8600
0.8700
0.8800
0.8890
0.8890
0.8900
0.9OoO

0.9100
0.9200
0.9300
0.9400
0.9500
0.9600
0.9700
0.9800
0.9900
1.oooO
1.0100
1.0200
1.0300
1.04oO
1.0500
1.06oO
1.0700
1.0800
1.09oO
1.1000
1.1100
1.1200
1.1300
1.1400
1.1500
1.1600
1.1700
1.1800
1.1900
1.2000
1.2000

5.2780
5.2783
5.2787
5.2790
5.2794
5.2797
5.2801
5.2804
5.2804
5.2804
5.2808
5.2812
5.2815
5.2819
5.2823
5.2826
5.2830
5.2834
5.2838
5.2841
5.2845
5.2849
5.2853
5.2857
5.2860
5.2864
5.2868
5.2872
5.2876
5.2880
5.2884
5.2888
5.2892
5.2896
5.2900
5.2904
5.2908
5.2912
5.2916
5.2921
5.2925
5.2925

4.2481
4.2481
4.2480
4.2480
4.2480
4.2480
4.2480
4.2480
4.2480
4.2480
4.2481
4.2482
4.2482
4.2483
4.2484
4.2485
4.2486
4.2488
4.2489
4.2491
4.2492
4.2494
4.2496
4.2498
4.2500
4.2502
4.2505
4.2507
4.2509
4.2512
4.2515
4.2517
4.2520
4.2523
4.2526
4.2529
4.2532
4.2536
4.2539
4.2542
4.2546
4.2546

1.1

1.2

1.5

0.3028
0.3057
0.3086
0.3114
0.3142
0.3170
0.3197
0.3221
0.3221
0.3223
0.3250
0.3276
0.3301
0.3326
0.3351
0.3375
0.3399
0.3422
0.3445
0.3467
0.3489
0.3511
0.3532
0.3552
0.3572
0.3591
0.3610
0.3628
0.3646
0.3663
0.3679
0.3695
0.3711
0.3725
0.3740
0.3753
0.3766
0.3778
0.3790
0.3801
0.3811
0.3811

0.0484

-3.8361
3.8349
3.8338
3.8328
3.8319
3.8311
3.8303
3.8296
3.8296
3.8295
3.8289
3.8282
3.8277
3.8271
3.8267
3.8263
3.8259
3.8255
3.8252
3.8250
3.8248
3.8246
3.8244
3.8243
3.8242
3.8242
3.8242
3.8242
3.8242
3.8243
3.8244
3.8245
3.8246
3.8248
3.8250
3.8252
3.8254
3.8256
3.8259
3.8262
3.8265
3.8265

1.4259
1.4873
1.5595
1.6479
1.7638
1.9370
2.3388
3.4153
3.4134
3.3970
3.3840
3.3733
3.3645
3.3571
3.3510
3.3458
3.3415
3.3378
3.3347
3.3321
3.3299
3.3281
3.3266
3.3254
3.3244
3.3237
3.3231
3.3227
3.3225
3.3224
3.3225
3.3227
3.3229
3.3233
3.3237
3.3243
3.3248
3.3255
3.3262
3.3270
3.3270

1.0798
1.1041
1.1293
1.1555
1.1828
1.2114
1.2414
1.2696
1.2696
1.2730
1.3065
1.3423
1.3808
1.4227
1.4688
1.5204
1.5798
1.6509
1.7418
1.8763
2.3825
2.7434
2.7802
2.8037
2.8216
2.8363
2.8489
2.8601
2.8702
2.8794
2.8879
2.8958
2.9033
2.9103
2.9170
2.9234
2.9295
2.9353
2.9409
2.9463
2.9516
2.9516

0.8927
0.9080
0.9235
0.9391
0.9549
0.9709
0.9870
1.0015
1.0015
1.0032
1.0196
1.0361
1.0527
1.0693
1.0859
1.1024
1.1188
1.1350
1.1507
1.1659
1.1801
1.1931
1.2044
1.2132
1.2185
1.2189
1.2125
1.1968
1.1693
1.1284
1.0777
1.0299
1.0046
1.0165
1.0686
1.1536
1.2605
1.3783
1.4984
1.6146
1.7231
1.7231

0.7633
0.7745
0.7857
0.7969
0.8081
0.8193
0.8305
0.8404
0.8404
0.8416
0.8527
0.8637
0.8746
0.8855
0.8962
0.9067
0.9170
0.9272
0.9370
0.9465
0.9556
0.9643
0.9725
0.9800
0.9868
0.9927
0.9977
1.0015
1.0039
1.0049
1.0041
1.0013
0.9963
0.9890
0.9792
0.9670
0.9525
0.9364
0.9194
0.9030
0.8885
0.8885

0.6647
0.6734
0.6822
0.6909
0.6995
0.7081
0.7167
0.7243
0.7243
0.7252
0.7336
0.7419
0.7501
0.7583
0.7663
0.7741
0.7818
0.7894
0.7967
0.8039
0.8108
0.8174
0.8238
0.8299
0.8356
0.8409
0.8458
0.8503
0.8543
0.8577
0.8605
0.8627
0.8642
0.8649
0.8648
0.8639
0.8621
0.8593
0.8556
0.8510
0.8454
0.8454

0.5196
0.5255
0.5315
0.5374
0.5432
0.5490
0.5547
0.5598
0.5598
0.5604
0.5659
0.5715
0.5769
0.5822
0.5875
0.5927
0.5978
0.6027
0.6076
0.6123
0.6170
0.6215
0.6259
0.6301
0.6342
0.6381
0.6418
0.6454
0.6489
0.6521
0.6551
0.6579
0.6605
0.6629
0.6651
0.6670
0.6686
0.6700
0.6711
0.6720
0.6726
0.6726

0.0486
0.0487
0.0488
0.0490
0.0491
0.0491
0.0492
0.0492
0.0492
0.0493
0.0493
0.0494
0.0494
0.0494
0.0494
0.0494
0.0494
0.0494
0.0494
0.0493
0.0492
0.0492
0.0491
0.0490
0.0489
0.0487
0.0486
0.0485
0.0483
0.0481
0.0480
0.0478
0.0476
0.0473
0.0471
0.0469
0.0466
0.0464
0.0461
0.0458
0.0458

459

Appendix 2. lee and Kesler Method

Table A2.S
Adimensional residual entropy of the simple fluid (with opposite sign)

0.8

0.9

0.94

0.98

0.0100
0.0100
0.0200
0.0300
0.04oO
0.0500
0.06oO
0.0700
0.0800
0.0900
0.1000
0.1100
0.1200
0.1300
0.1400
0.1500
0.1600
0.1700
0.1800
0.1900
0.2000
0.2100
0.2200
0.2300
0.2400
0.2500
0.2563
0.2563
0.2600
0.2700
0.2800
0.2900
0.3000
0.3100
0.3200
0.3300
0.3400
0.3500
0.3600
0.3700
0.3800
0.3900
0.4000
0.4100

1.0125
1.0125
1.0252
1.0381
1.0511
1.0644
1.0778
1.0915
1.1054
3.1196
3.1339
3.1486
3.1635
3.1787
3.1942
3.2100
3.2261
D.2426
D.2594
0.2767
D.2943
D.3124
D.3310
D.3501
0.3697
D.3899
0.4029
5.4497
5.4358
5.3996
5.3647
5.3311
5.2987
5.2674
5.2371
5.2078
5.1795
5.1520
5.1253
5.0993
5.0741
5.0496
5.0257
5.0025

0.0090
0.0090
0.0180
0.0271
0.0364
0.0457
0.0551
0.0646
0.0742
0.0840
0.0938
0.1037
0.1138
0.1240
0.1343
0.1447
0.1552
0.1659
0.1767
0.1877
0.1988
0.2101
0.2215
0.2331
0.2449
0.2568
0.2644
0.2644
0.2689
0.2813
0.2938
0.3065
0.3194
0.3326
0.3460
0.3597
0.3736
0.3878
0.4023
0.4171
0.4322
0.4476
0.4634
0.47%

D.0079
D.0079
D.0159
0.0240
D.0321
0.0404
0.0486
0.0570
0.0654
0.0740
0.0825
0.0912
0.1000
0.1088
0.1178
0.1268
0.1359
0.1452
0.1545
0.1639
0.1734
0.1831
0.1928
0.2027
0.2127
0.2228
0.2292
0.2292
0.2330
0.2433
0.2538
0.2645
0.2752
0.2861
0.2972
0.3084
0.3198
0.3314
0.3431
0.3551
0.3672
0.3795
0.3920
0.4048

1.02

1.04

1.06

1.1

1.2

1.5

0.0063
0.0063
0.0127
0.0191
0.0256
0.0321
0.0386
0.0452
0.0518
0.0585
0.0653
0.0720
0.0789
0.0857
0.0927
0.0996
0.1067
0.1138
0.1209
0.1281
0.1353
0.1426
0.1500
0.1574
0.1649
0.1725
0.1772
0.1772
0.1801
0.1878
0.1955
0.2033
0.2112
0.2192
0.2272
0.2353
0.2435
0.2517
0.2601
0.2685
0.2770
0.2856
0.2943
0.3030

I.Oo60
l.Oo60
1.0120
1.0181
1.0242
1.0304
1.0366
1.0428
1.0491
1.0554
1.0618
1.0682
3.0746
3.0811
3.0876
3.0942
3.1008
3.1075
3.1142
D.1210
D.1278
0.1346
0.1415
0.1485
D.1555
D.1626
0.1671
0.1671
0.1697
0.1769
0.1841
0.1914
0.1988
0.2062
0.2137
0.2212
0.2288
0.2365
0.2443
0.2521
0.2599
0.2679
0.2759
0.2840

0.0057
0.0057
0.0114
0.0172
0.0230
0.0288
0.0347
0.0406
0.0465
0.0525
0.0585
0.0646
0.0707
0.0768
0.0830
0.0892
0.0954
0.1017
0.1080
0.1144
0.1208
0.1273
0.1338
0.1403
0.1469
0.1535
0.1577
0.1577
0.1602
0.1670
0.1737
0.1806
0.1875
0.1944
0.2014
0.2084
0.2155
0.2227
0.2299
0.2371
0.2445
0.2519
0.2593
0.2668

3.0052
3.0052
D.0103
D.0155
D.0208
D.0260
D.0313
D.0367
D.0420
D.0474
0.0528
D.0582
0.0637
0.0692
0.0747
0.0802
0.0858
0.0914
0.0971
0.1028
0.1085
0.1142
0.1200
0.1258
0.1317
0.1376
0.1413
0.1413
0.1435
0.1494
0.1554
0.1615
0.1675
0.1736
0.1798
0.1860
0.1922
0.1985
0.2048
0.2111
0.2175
0.2240
0.2304
0.2370

0.0041
0.0041
0.0082
0.0123
0.0164
0.0206
0.0247
0.0289
0.0331
0.0373
0.0415
0.0458
0.0500
0.0543
0.0586
0.0629
0.0672
0.0716
0.0759
0.0803
0.0847
0.0891
0.0936
0.0980
0.1025
0.1070
0.1098
0.1098
0.1115
0.1160
0.1206
0.1251
0.1297
0.1343
0.1389
0.1436
0.1483
0.1529
0.1576
0.1624
0.1671
0.1719
0.1767
0.1815

D.0023
D.0023

0.0071
0.0071
0.0142
0.0214
0.0286
0.0359
0.0432
0.0506
0.0581
0.0656
0.0732
0.0808
0.0885
0.0963
0.1041
0.1120
0.1200
0.1280
0.1362
0.1444
0.1526
0.1610
0.1694
0.1779
0.1865
0.1952
0.2007
0.2007
0.2039
0.2128
0.2218
0.2308
0.2399
0.2492
0.2585
0.2680
0.2776
0.2872
0.2970
0.3069
0.3170
0.3271
0.3374
0.3479

0.0067
0.0067
0.0134
0.0202
0.0270
0.0339
0.0408
0.0478
0.0548
0.0619
0.0690
0.0762
0.0835
0.0908
0.0981
0.1056
0.1130
0.1206
0.1282
0.1359
0.1436
0.1514
0.1593
0.1672
0.1752
0.1833
0.1884
0.1884
0.1915
0.1997
0.2080
0.2164
0.2249
0.2334
0.2421
0.2508
0.2597
0.2686
0.2776
0.2867
0.2959
0.3052
0.3147
0.3242

----- -

0.0046

D.0069
0.0091
0.0114
0.0137
0.0160
0.0183
0.0206
0.0230
0.0253
0.0276
0.0299
0.0322
0.0346
0.0369
0.0392
0.0416
0.0439
0.0463
0.0486
0.0510
0.0533
0.0557
0.0581
0.0596
0.0596
0.0604
0.0628
0.0652
0.0676
0.0700
0.0724
0.0747
0.0771
0.0795
0.0819
0.0843
0.0868
0.0892
0.0916
0.0940
0.0964

460

Appendix 2. Lee and Kesler Method


Table A25 (cont.)
Adimensional residual entropy of the simple fluid (with opposite sign)

--

--

0.4200
0.4300
0.4400
0.4500
0.4600

0.4700
0.4800
0.4900
0.5000
0.5100
0.5200
0.5300

0.5374
0.5374
0.5400
0.5500
0.5600
0.5700
0.5800
0.5900
0.6OoO

0.6100
0.6200
0.6300
0.6400
0.6500
0.6600
0.6700
0.6800
0.6900
0.6963
0.6963
0.7000
0.7100
0.7200
0.7300
0.7400
0.7500
0.7600
0.7700
0.7800
0.7900
0.8000
0.8100

0.8

0.9

0.94

0.98

1.02

1.04

1.06

1.1

1.2

1.5

4.9798
4.9578
4.9362
4.9152
4.8946
4.8745
4.8549
4.8357
4.8169
4.7986
4.7806
4.7629
4.7500
4.7500
4.7456
4.7287
4.7121
4.6958
4.6798
4.6641
4.6486
4.6335
4.6186
4.6040
4.5896
4.5755
4.5616
4.5479
4.5345
4.5212
4.5129
4.5129
4.5082
4.4954
4.4827
4.4703
4.4580
4.4460
4.4340
4.4223
4.4107
4.3993
4.3881
4.3770

0.4962
0.5132
0.5307
0.5487
0.5672
0.5864
0.6061
0.6265
0.6477
0.6697
0.6926
0.7165
0.7351
4.2360
4.2321
4.2168
4.2019
4.1872
4.1729
4.1588
4.1450
4.1314
4.1180
4.1050
4.0921
4.0795
4.0670
4.0548
4.0428
4.0310
4.0236
4.0236
4.0193
4.0079
3.9966
3.9855
3.9746
3.9639
3.9533
3.9428
3.9325
3.9224
3.9124
3.9025

0.4177
0.4309
0.4444
0.4581
0.4721
0.4864
0.5009
0.5159
0.5311
0.5467
0.5627
0.5791
0.5915
0.5915
0.5959
0.6132
0.6310
0.6493
0.6682
0.6877
0.7080
0.7289
0.7507
0.7734
0.7971
0.8220
0.8482
0.8758
0.9051
0.9365
0.9576
3.7318
3.7286
3.7199
3.7113
3.7028
3.6944
3.6860
3.6778
3.6696
3.6616
3.6536
3.6457
3.6379

0.3584
0.3692
0.3800
0.3910
0.4022
0.4136
0.4251
0.4368
0.4487
0.4608
0.4731
0.4856
0.4950
0.4950
0.4983
0.5113
0.5245
0.5379
0.5516
0.5656

0.3339
0.3436
0.3535
0.3635
0.3737
0.3839
0.3943
0.4049
0.4156
0.4264
0.4374
0.4486
0.4570
0.4570
0.4599
0.4714
0.4831
0.4950
0.5070
0.5193
0.5318
0.5445
0.5574
0.5706
0.5840
0.5977
0.6116
0.6259
0.6405
0.6553
0.6650
0.6650
0.6706
0.6862
0.7021
0.7185
0.7353
0.7526
0.7704
0.7887
0.8075
0.8270
0.8472
0.8680

0.3119
0.3209
0.3299
0.3391
0.3483
0.3577
0.3672
0.3767
0.3864
0.3962
0.4062
0.4162
0.4238
0.4238
0.4264
0.4368
0.4472
0.4578
0.4686
0.4795
0.4906
0.5018
0.5132
0.5248
0.5366
0.5485
0.56M
0.5730
0.5856
0.5984
0.6066
0.6066
0.6114
0.6247
0.6382
0.6520
0.6661
0.6804
0.6951
0.7101
0.7254
0.7411
0.7572
0.7737

0.2922
0.3004
0.3088
0.3172
0.3257
0.3343
0.3429
0.3517
0.3606
0.3695
0.3786
0.3877
0.3946
0.3946
0.3970
0.4063
0.4158
0.4254
0.4351
0.4449
0.4548
0.4649
0.4751
0.4854
0.4959
0.5065
0.5172
0.5281
0.5392
0.5504
0.5576
0.5576
0.5618
0.5733
0.5851
0.5970
0.6091
0.6215
0.6340
0.6467
0.6597
0.6729
0.6864
0.7001

0.2744
0.2820
0.2897
0.2975
0.3053
0.3133
0.3212
0.3293
0.3374
0.3457
0.3540
0.3623
0.3686
0.3686
0.3708
0.3793
0.3880
0.3967
0.4055
0.4144
0.4234
0.4325
0.4417
0.4510
0.4604
0.4699
0.4796
0.4893
0.4992
0.5092
0.5156
0.5156
0.5193
0.5295
0.5399
0.5504
0.5611
0.5719
0.5828
0.5939
0.6052
0.6166
0.6282

0.2435
0.2502
0.2568
0.2635
0.2703
0.2771
0.2840
0.2909
0.2979
0.3049
0.3120
0.3191
0.3244
0.3244
0.3263
0.3335
0.3408
0.3482
0.3556
0.3631
0.3706
0.3782
0.3859
0.3936
0.4014
0.4093
0.4173
0.4253
0.4334
0.4415
0.4468
0.4468
0.4498
0.4581
0.4665
0.4750
0.4836
0.4922
0.5009
0.5098
0.5187
0.5277
0.5368
0.5460

0.1863
0.1912
0.1961
0.2010
0.2059
0.2108
0.2158
0.2208
0.2258
0.2309
0.2359
0.2410
0.2448
0.2448
0.2461
0.2513
0.2564
0.2616
0.2668
0.2721
0.2773
0.2826
0.2879
0.2933
0.2987
0.3041
0.3095
0.3149
0.3204
0.3259
0.3294
0.3294
0.3315
0.3370
0.3426
0.3483
0.3539
0.3596
0.3653
0.3711
0.3769
0.3827
0.3885
0.3944

0.0989
0.1013
0.1037
0.1062
0.1086
0.1111
0.1135
0.1160
0.1184
0.1209
0.1234
0.1258
0.1277
0.1277
0.1283
0.1308
0.1333
0.1357
0.1382
0.1407
0.1432
0.1457
0.1482
0.1507
0.1532
0.1557
0.1583
0.1608
0.1633
0.1658
0.1674
0.1674
0.1683
0.1709
0.1734
0.1760
0.1785
0.1811
0.1836
0.1862
0.1887
0.1913
0.1938
0.1964

0.5799

0.5945
0.6094
0.6247
0.6403
0.6564
0.6728
0.6897
0.7070
0.7249
0.7365
0.7365
0.7433
0.7623
0.7819
0.8022
0.8233
0.8452
0.8680
0.8918
0.9167
0.9430
0.9706
1.oooO

0.6400

46 1

Appendix 2. Lee and Kesler Method

Table A2.5 (cont. and end)


Adimensional residual entropy of the simple fluid (with opposite sign)

----0.8

0.9

0.94

0.98

1.02

1.04

1.06

1.1

1.2

1.5

0.8897
0.9122
0.9357
0.9602
0.9860
1.0130
1.0416
1.0686
1.0686
1.0719
1.1042
1.1388
1.1764
1.2173
1.2627
1.3137
1.3725
1.4427
1.5320
1.6615
2.1674
2.5942
2.6709
2.7174
2.7502
2.7752
2.7951
2.8113
2.8247
2.8361
2.8458
2.8540
2.8612
2.8674
2.8727
2.8773
2.8814
2.8848
2.8878
2.8904
2.8926
2.8926

0.7906
0.8080
0.8259
0.8443
0.8632
0.8827
0.9029
0.9216
0.9216
0.9238
0.9455
0.9679
0.9913
1.0157
1.0411
1.0678
1.0958
1.1253
1.1565
1.1897
1.2251
1.2631
1.3042
1.3489
1.3980
1.4524
1.5134
1.5823
1.6610
1.7504
1.8498
1.9544
2.0563
2.1486
2.2281
2.2948
2.3502
2.3963
2.4349
2.4674
2.4952
2.4952

0.7141
0.7284
0.7430
0.7578
0.7730
0.7886
0.8045
0.8191
0.8191
0.8208
0.8375
0.8546
0.8721
0.8902
0.9087
0.9278
0.9475
0.9677
0.9886
1.0103
1.0326
1.0558
1.0799
1.1049
1.1309
1.1580
1.1864
1.2160
1.2471
1.2798
1.3141
1.3502
1.3883
1.4285
1.4708
1.5153
1.5619
1.6104
1.6605
1.7117
1.7632
1.7632

0.6520
0.6641
0.6765
0.6890
0.7018
0.7148
0.7280
0.7401
0.7401
0.7415
0.7552
0.7691
0.7834
0.7979
0.8127
0.8278
0.8432
0.8590
0.8751
0.8915
0.9084
0.9256
0.9433
0.9614
0.9800
0.9991
1.0186
1.0387
1.0594
1.0806
1.1025
1.1250
1.1482
1.1722
1.1968
1.2222
1.2485
1.2755
1.3034
1.3321
1.3616
1.3616

0.5553
0.5647
0.5742
0.5838
0.5935
0.6033
0.6133
0.6223
0.6223
0.6233
0.6335
0.6438
0.6542
0.6648
0.6754
0.6863
0.6972
0.7083
0.7196
0.7309
0.7425
0.7542
0.7661
0.7781
0.7903
0.8026
0.8152
0.8279
0.8408
0.8539
0.8672
0.8807
0.8944
0.9083
0.9224
0.9368
0.9513
0.9661
0.9811
0.9964
1.0118
1.0118

0.4003
0.4062
0.4122
0.4182
0.4242
0.4303
0.4364
0.4419
0.4419
0.4426
0.4487
0.4549
0.4612
0.4675
0.4738
0.4801
0.4865
0.4929
0.4994
0.5059
0.5124
0.5190
0.5256
0.5323
0.5390
0.5457
0.5525
0.5593
0.5661
0.5730
0.5799
0.5869
0.5939
0.6009
0.6080
0.6152
0.6223
0.6295
0.6368
0.6441
0.6514
0.6514

0.1990
0.2016
0.2041
0.2067
0.2093
0.2119
0.2145
0.2168
0.2168
0.2171
0.2197
0.2223
0.2249
0.2275
0.2301
0.2327
0.2353
0.2380
0.2406
0.2432
0.2458
0.2485
0.2511
0.2538
0.2564
0.2591
0.2617
0.2644
0.2670
0.2697
0.2723
0.2750
0.2777
0.2803
0.2830
0.2857
0.2884
0.2910
0.2937
0.2964
0.2991
0.2991

---0.8200
0.8300
0.8400
0.8500
0.8600
0.8700
0.8800
0.8890
0.8890
0.8900

0.9oOo
0.9100
0.9200
0.9300
0.9400
0.9500
0.9600
0.9700
0.9800
0.9900
1.oooO
1.0100
1.0200
1.0300
1.0400
1.0500
1.0600
1.0700
1.0800
1.0900
1.1000
1.1100
1.1200
1.1300
1.1400
1.1500
1.1600
1.1700
1.1800
1.1900
1.2000
1.2000

4.3660
4.3552
4.3446
4.3341
4.3237
4.3134
4.3033
4.2943
4.2943
4.2933
4.2834
4.2737
4.2641
4.2545
4.2451
4.2358
4.2267
4.2176
4.2086
4.1997
4.1910
4.1823
4.1737
4.1652
4.1568
4.1485
4.1403
4.1322
4.1241
4.1161
4.1083
4.1005
4.0927
4.0851
4.0775
4.0700
4.0626
4.0553
4.0480
4.0408
4.0336
4.0336

3.8928
3.8832
3.8737
3.8644
3.8552
3.8461
3.8371
3.8292
3.8292
3.8283
3.8195
3.8109
3.8023
3.7939
3.7856
3.7774
3.7692
3.7612
3.7533
3.7454
3.7377
3.7300
3.7224
3.7149
3.7075
3.7002
3.6929
3.6857
3.6786
3.6716
3.6646
3.6578
3.6510
3.6442
3.6375
3.6309
3.6244
3.6179
3.6115
3.6052
3.5989
3.5989

3.6301
3.6225
3.6149
3.6074
3.6000
3.5927
3.5854
3.5790
3.5790
3.5783
3.5711
3.5641
3.5572
3.5503
3.5434
3.5367
3.5300
3.5234
3.5168
3.5103
3.5039
3.4975
3.4912
3.4850
3.4788
3.4727
3.4666
3.4606
3.4546
3.4487
3.4429
3.4371
3.4314
3.4257
3.4200
3.4144
3.4089
3.4034
3.3980
3.3926
3.3872
3.3872

1.0313
1.0649
1.1014
1.1412
1.1856
1.2359
1.2949
1.3592
3.1073
3.1083
3.1165
3.1229
3.1280
3.1321
3.1352
3.1376
3.1393
3.1405
3.1413
3.1416
3.1416
3.1413
3.1407
3.1399
3.1389
3.1376
3.1362
3.1347
3.1330
3.1312
3.1292
3.1272
3.1251
3.1228
3.1206
3.1182
3.1158
3.1133
3.1108
3.1082
3.1056
3.1056

---

---

--

462

Appendix 2. Lee and Kesler Method

Table A2.6
Adimensional residual entropy (with opposite sign): corrective term

0.8
0.0100
0.0100
0.0200
0.0300
0.O400

0.0500
0.0600
0.0700
0.0800
0.0900
0.1000
0.1100
0.1200
0.1300
0.1400
0.1500
0.1600
0.1700
0.1800
0.1900
0.2000
0.2100
0.2200
0.2300
0.2400
0.2500
0.2563
0.2563
0.2600
0.2700
0.2800
0.2900
0.3000
0.3100
0.3200
0.3300
0.3400
0.3500
0.3600
0.3700
0.3800
0.3900
0.4OoO
0.4100

0.0222
0.0222
0.0449
0.0680
0.0917
0.1159
0.1406
0.1660
0.1920
0.2188
0.2463
0.2745
0.3037
0.3337
0.3648
0.3970
0.4303
0.4650
0.5011
0.5389
0.5784
0.6199
0.6636
0.7099
0.7592
0.8119
0.8470
5.4805
5.4803
5.4799
5.4794
5.4789
5.4785
5.4781
5.4776
5.4772
5.4768
5.4764
5.4759
5.4755
5.4751
5.4747
5.4743
5.4739

0.9
0.0133
0.0133
0.0267
0.0402
0.0539
0.0678
0.0819
0.0961
0.1105
0.1251
0.1399
0.1549
0.1701
0.1855
0.2012
0.2171
0.2332
0.2496
0.2663
0.2833
0.3005
0.3181
0.3360
0.3543
0.3729
0.3918
0.4040
0.4040
0.4112
0.4310
0.4513
0.4721
0.4933
0.5151
0.5375
0.5605
0.5842
0.6086
0.6338
0.6599
0.6868
0.7148
0.7439
0.7742

0.94
0.0109
0.0109
0.0220
0.0331

0.0444
0.0557
0.0672
0.0788
0.0904
0.1022
0.1141
0.1262
0.1383
0.1506
0.1630
0.1756
0.1883
0.2011
0.2141
0.2273
0.2406
0.2541
0.2678
0.2816
0.2956
0.3098
0.3189
0.3189
0.3243
0.3389
0.3537
0.3688
0.3841
0.3997
0.4155
0.4316
0.4480
0.4646
0.4816
0.4989
0.5166
0.5346
0.5530
0.5719

0.98
0.0091
0.0091
0.0183
0.0275
0.0368
0.0462
0.0557
0.0652
0.0747
0.0844
0.0941
0.1039
0.1137
0.1237
0.1337
0.1438
0.1540
0.1642
0.1746
0.1850
0.1955
0.2062
0.2169
0.2277
0.2386
0.2496
0.2565
0.2565
0.2607
0.2719
0.2833
0.2947
0.3063
0.3179
0.3297
0.3417
0.3537
0.3659
0.3783
0.3908
0.4034
0.4162
0.4291
0.4423

1
0.0083
0.0083
0.0167
0.0252
0.0337
0.0422
0.0508
0.0595
0.0682
0.0769
0.0857
0.0946
0.1035
0.1125
0.1215
0.1306
0.1398
0.1490
0.1583
0.1677
0.1771
0.1866
0.1962
0.2058
0.2155
0.2253
0.2315
0.2315
0.2351
0.2451
0.2551
0.2652
0.2754
0.2856
0.2960
0.3064
0.3169
0.3275
0.3383
0.3491
0.3600
0.3710
0.3821
0.3933

1.02
0.0076
0.0076
0.0153
0.0231
0.0308
0.0386
0.0465
0.0544
0.0623
0.0703
0.0783
0.0863
0.0944
0.1026
0.1108
0.1190
0.1273
0.1356
0.1440
0.1524
0.1609
0.1694
0.1780
0.1866
0.1953
0.2040
0.2095
0.2095
0.2128
0.2217
0.2306
0.2395
0.2485
0.2576
0.2667
0.2759
0.2852
0.2945
0.3039
0.3134
0.3229
0.3325
0.3421
0.3519

1.04

1.06

1.1

1.2

1.5

---

--

0.0070
0.0070
0.0141
0.0212
0.0283
0.0354
0.0426
0.0498
0.0570
0.0643
0.0716
0.0790
0.0863
0.0937
0.1012
0.1086
0.1161
0.1237
0.1313
0.1389
0.1465
0.1542
0.1619
0.1697
0.1775
0.1853
0.1903
0.1903
0.1932
0.201 1
0.2091
0.2171
0.2251
0.2332
0.2413
0.2495
0.2577
0.2659
0.2742
0.2825
0.2909
0.2993
0.3078
0.3163

0.0037
0.0037
0.0075
0.0112
0.0150
0.0187
0.0225
0.0262
0.0300
0.0337
0.0375
0.0412
0.0450
0.0487
0.0525
0.0562
0.0599
0.0637
0.0674
0.0711
0.0749
0.0786
0.0823
0.0861
0.0898
0.0935
0.0958
0.0958
0.0972
0.1009
0.1047
0.1084
0.1121
0.1158
0.1195
0.1232
0.1268
0.1305
0.1342
0.1379
0.1415
0.1452
0.1488
0.1525

0.0065
0.0065
0.0129
0.0195
0.0260
0.0325
0.0391
0.0457
0.0523
0.0590
0.0657
0.0724
0.0791
0.0858
0.0926
0.0994
0.1062
0.1131
0.1199
0.1268
0.1338
0.1407
0.1477
0.1547
0.1618
0.1688
0.1733
0.1733
0.1759
0.1830
0.1902
0.1973
0.2045
0.2118
0.2190
0.2263
0.2336
0.2410
0.2483
0.2557
0.2631
0.2706
0.2781
0.2856

0.0055
0.0055
0.0110
0.0165
0.0220
0.0276
0.0331
0.0387
0.0443
0.0499
0.0555
0.0611
0.0667
0.0724
0.0780
0.0837
0.0894
0.0951
0.1008
0.1065
0.1123
0.1180
0.1238
0.1295
0.1353
0.1411
0.1447
0.1447
0.1469
0.1527
0.1586
0.1644
0.1703
0.1761
0.1820
0.1879
0.1938
0.1997
0.2056
0.2116
0.2175
0.2234
0.2294
0.2354

---------

0.0014
0.0014
0.0028
0.0042
0.0056
0.0070
0.0084
0.0098
0.0112
0.0126
0.0140
0.0154
0.0167
0.0181
0.0194
0.0208
0.0221
0.0235
0.0248
0.0262
0.0275
0.0288
0.0301
0.0314
0.0328
0.0341
0.0349
0.0349
0.0354
0.0366
0.0379
0.0392
0.0405
0.0418
0.0430
0.0443
0.0455
0.0468
0.0480
0.0493
0.0505
0.0517
0.0530
0.0542

--

463

Appendix 2. Lee and Kesler Method

Table A2.6 (cont.)


Adimensional residual entropy (with opposite sign): correctiveterm

------

0.4200
0.4300
0.4400
0.4500
0.4600
0.4700
0.4800
0.4900
0.5000
0.5100
0.5200
0.5300
0.5374
0.5374
0.5400
0.5500
0.5600

0.5700
0.5800
0.5900
0.6oOo

0.6100
0.6200
0.6300
0.6400
0.6500
0.6600
0.6700
0.6800
0.6900
0.6963
0.6963
0.7000
0.7100
0.7200
0.7300
0.7400
0.7500
0.7600
0.7700
0.7800
0.7900
0.8000
0.8100

0.8

0.9

0.94

0.98

1.02

1.04

1.06

1.1

1.2

1.5

5.4736
5.4732
5.4728
5.4724
5.4721
5.4717
5.4713
5.4710
5.4706
5.4703
5.4700
5.4696
5.4694
5.4694
5.4693
5.4690
5.4686
5.4683
5.4680
5.4677
5.4674
5.4671
5.4668
5.4665
5.4662
5.4659
5.4656
5.4653
5.4650
5.4648
5.4646
5.4646
5.4645
5.4642
5.4640
5.4637
5.4635
5.4632
5.4630
5.4627
5.4625
5.4622
5.4620
5.4618

0.8059
0.8392
0.8741
0.9110
0.9501
0.9917
1.0364
1.0846
1.1370
1.1947
1.2589
1.3315
1.3926
4.2778
4.2774
4.2760
4.2745
4.2731
4.2718
4.2705
4.2692
4.2679
4.2667
4.2655
4.2644
4.2632
4.2621
4.2611
4.2600
4.2590
4.2584
4.2584
4.2581
4.2571
4.2562
4.2553
4.2544
4.2535
4.2527
4.2519
4.2511
4.2503
4.2495
4.2488

0.5911
0.6109
0.6311
0.6519
0.6732
0.6952
0.7178
0.7411
0.7652
0.7902
0.8160
0.8429
0.8636
0.8636
0.8710
0.9003
0.9310
0.9632
0.9973
1.0335
1.0720
1.1133
1.1580
1.2068
1.2605
1.3205
1.3889
1.4685
1.5646
1.6869
1.7873
3.8326
3.8313
3.8279
3.8247
3.8217
3.8188
3.8161
3.8134
3.8110
3.8086
3.8063
3.8042
3.8021

0.4556
0.4690
0.4827
0.4965
0.5106
0.5248
0.5393
0.5540
0.5689
0.5841
0.5996
0.6153
0.6272
0.6272
0.6313
0.6477
0.6643
0.6813
0.6986
0.7163
0.7345
0.7530
0.7720
0.7915
0.8116
0.8322
0.8534
0.8753
0.8979
0.9213
0.9366
0.9366
0.9455
0.9708
0.9971
1.0247
1.0535
1.0840
1.1162
1.1505
1.1873
1.2269
1.2702
1.3179

0.4047
0.4161
0.4277
0.4394
0.4512
0.4632
0.4753
0.4875
0.4998
0.5123
0.5250
0.5378
0.5474
0.5474
0.5508
0.5639
0.5773
0.5908
0.6044
0.6183
0.6324
0.6467
0.6612
0.6760
0.6910
0.7062
0.7217
0.7375
0.7536
0.7699
0.7805
0.7805
0.7866
0.8037
0.8211
0.8389
0.8571
0.8757
0.8948
0.9144
0.9345
0.9552
0.9766
0.9986

0.3617
0.3716
0.3815
0.3915
0.4016
0.4118
0.4221
0.4325
0.4429
0.4534
0.4640
0.4747
0.4828
0.4828
0.4855
0.4964
0.5074
0.5185
0.5297
0.5410
0.5524
0.5639
0.5756
0.5873
0.5991
0.6111
0.6232
0.6354
0.6478
0.6603
0.6683
0.6683
0.6729
0.6856
0.6985
0.7116
0.7248
0.7381
0.7516
0.7652
0.7791
0.7930
0.8072
0.8215

0.3249
0.3335
0.3422
0.3509
0.3596
0.3684
0.3773
0.3862
0.3952
0.4042
0.4132
0.4223
0.4291
0.4291
0.4315
0.4407
0.4500
0.4593
0.4687
0.4781
0.4876
0.4972
0.5068
0.5165
0.5262
0.5359
0.5458
0.5557
0.5656
0.5756
0.5820
0.5820
0.5857
0.5958
0.6059
0.6162
0.6264
0.6368
0.6471
0.6575
0.6680
0.6785
0.6891
0.6996

0.2932
0.3007
0.3083
0.3160
0.3236
0.3313
0.3390
0.3468
0.3546
0.3624
0.3702
0.3781
0.3840
0.3840
0.3860
0.3939
0.4019
0.4099
0.4179
0.4259
0.4340
0.4421
0.4502
0.4583
0.4665
0.4747
0.4829
0.4911
0.4994
0.5077
0.5129
0.5129
0.5160
0.5243
0.5326
0.5409
0.5493
0.5577
0.5660
0.5744
0.5828
0.5912
0.5995
0.6079

0.2414
0.2473
0.2533
0.2593
0.2654
0.2714
0.2774
0.2834
0.2895
0.2955
0.3016
0.3076
0.3121
0.3121
0.3137
0.3198
0.3258
0.3319
0.3380
0.3441
0.3501
0.3562
0.3623
0.3683
0.3744
0.3805
0.3865
0.3926
0.3986
0.4047
0.4085
0.4085
0.4107
0.4167
0.4227
0.4287
0.4347
0.4406
0.4466
0.4525
0.4584
0.4643
0.4701
0.4760

0.1561
0.1598
0.1634
0.1670
0.1706
0.1742
0.1778
0.1814
0.1850
0.1885
0.1921
0.1956
0.1983
0.1983
0.1992
0.2027
0.2062
0.2097
0.2132
0.2167
0.2201
0.2236
0.2270
0.2305
0.2339
0.2373
0.2406
0.2440
0.2474
0.2507
0.2528
0.2528
0.2540
0.2573
0.2606
0.2639
0.2671
0.2703
0.2735
0.2767
0.2799
0.2830
0.2861
0.2892

0.0554
0.0566
0.0578
0.0590
0.0602
0.0613
0.0625
0.0637
0.0648
0.0660
0.0671
0.0683
0.0691
0.0691
0.0694
0.0706
0.0717
0.0728
0.0739
0.0750
0.0761
0.0772
0.0783
0.0794
0.0805
0.0815
0.0826
0.0836
0.0847
0.0857
0.0864
0.0864
0.0868
0.0878
0.0888
0.0898
0.0908
0.0918
0.0928
0.0938
0.0948
0.0958
0.0967
0.0977

464

Appendix 2. Lee and Kesler Method

Table A2.6 (cont. and end)


Adimensional residual entropy (with opposite sign): corrective term

0.8

0.8200
0.8300
0.8400
0.8500
0.8600
0.8700
0.8800
0.8890
0.8890
0.8900
0.9000
0.9100
0.9200
0.9300
0.9400
0.9500
0.9600
0.9700
0.9800
0.9900
1.oooo
1.0100
1.0200
1.0300
1.o400
1.0500
1,0600
1.a700
1.0800
1.0900
1.1000
1.1100
1.1200
1.1300
1.1400
1.1500
1.1600
1.1700
1.1800
1.1900
1.2000
1.2000

5.4615
5.4613
5.4611
5.4609
5.4606
5.4604
5.4602
5.4600
5.4600
5.4600
5.4598
5.4596
5.4594
5.4592
5.4590
5.4588
5.4587
5.4585
5.4583
5.4581
5.4579
5.4578
5.4576
5.4574
5.4573
5.4571
5.4569
5.4568
5.4566
5.4565
5.4563
5.4562
5.4561
5.4559
5.4558
5.4556
5.4555
5.4554
5.4553
5.4551
5.4550
5.4550

0.9

0.94

0.98

1.02

1.04

1.06

1.1

1.2

1.5

3.8002
3.7983
3.7965
3.7948
3.7932
3.7916
3.7901
3.7889
3.7889
3.7887
3.7874
3.7861
3.7849
3.7837
3.7826
3.7815
3.7805
3.7795
3.7786
3.7777
3.7769
3.7761
3.7754
3.7746
3.7740
3.7733
3.7727
3.7721
3.7716
3.7710
3.7705
3.7701
3.7696
3.7692
3.7688
3.7685
3.7682
3.7678
3.7675
3.7673
3.7670
3.7670

1.3714
1.4326
1.5047
1.5932
1.7097
1.8844
2.2920

1.0213
1.0449
1.0694
1.0948
1.1214
1.1492
1.1785
1.2060
1.2060
1.2094
1.2422
1.2772
1.3150
1.3561
1.4014
1.4524
1.5110
1.5813
1.6714
1.8052
2.3106
2.6707
2.7069
2.7297
2.7470
2.7611
2.7732
2.7837
2.7932
2.8019
2.8098
2.8172
2.8242
2.8307
2.8368
2.8426
2.8482
2.8535
2.8586
2.8635
2.8683
2.8683

0.8360
0.8506
0.8654
0.8804
0.8956
0.9109
0.9264
0.9403
0.9403
0.9420
0.9578
0.9736
0.9896
1.0057
1.0217
1.0377
1.0536
1.0693
1.0846
1.0993
1.1132
1.1259
1.1369
1.1456
1.1509
1.1515
1.1455
1.1305
1.1039
1.0644
1.0155
0.9695
0.9456
0.9584
1.0103
1.0946
1.2001
1.3161
1.4343
1.5487
1.6553
1.6553

0.7103
0.7209
0.7316
0.7422
0.7529
0.7636
0.7743
0.7838
0.7838
0.7849
0.7956
0.8061
0.8166
0.8271
0.8374
0.8475
0.8575
0.8673
0.8769
0.8861
0.8951
0.9036
0.9116
0.9190
0.9258
0.9318
0.9368
0.9408
0.9435
0.9449
0.9445
0.9424
0.9381
0.9317
0.9229
0.9119
0.8988
0.8841
0.8687
0.8538
0.8408
0.8408

0.6163
0.6246
0.6329
0.6412
0.6495
0.6577
0.6659
0.6732
0.6732
0.6740
0.6821
0.6901
0.6980
0.7059
0.7136
0.7212
0.7287
0.7360
0.7432
0.7502
0.7570
0.7636
0.7699
0.7759
0.7817
0.7871
0.7921
0.7968

0.4817
0.4875
0.4932
0.4989
0.5046
0.5101
0.5157
0.5206
0.5206
0.5212
0.5266
0.5320
0.5374
0.5426
0.5478
0.5529
0.5580
0.5630
0.5678
0.5726
0.5773
0.5819
0.5863
0.5907
0.5949
0.5990
0.6030
0.6068
0.6104
0.6139
0.6173
0.6205
0.6235
0.6262
0.6288
0.6312
0.6334
0.6354
0.6371
0.6386
0.6399
0.6399

0.2923
0.2953
0.2984
0.3013
0.3043
0.3073
0.3102
0.3128
0.3128
0.3131
0.3159
0.3188
0.3216
0.3243
0.3271
0.3298
0.3325
0.3351
0.3377
0.3403
0.3428
0.3453
0.3478
0.3502
0.3526
0.3550
0.3573
0.3596
0.3618
0.3640
0.3661
0.3682
0.3703
0.3723
0.3743
0.3762
0.3781
0.3799
0.3817
0.3834
0.3851
0.3851

0.0987
0.0996
0.1005
0.1015
0.1024
0.1033
0.1042
0.1050
0.1050
0.1051
0.1060
0.1069
0.1078
0.1087
0.1096
0.1104
0.1113
0.1121
0.1130
0.1138
0.1146
0.1155
0.1163
0.1171
0.1179
0.1187
0.1195
0.1203
0.1210
0.1218
0.1225
0.1233
0.1240
0.1248
0.1255
0.1262
0.1270
0.1277
0.1284
0.1291
0.1297
0.1297

4.2481
4.2474
4.2467
4.2461
4.2454
4.2448
4.2442
4.2437
4.2437
4.2436
4.2431
4.2425
4.2420
4.2415
4.2410
4.2405
4.2400
4.2395
4.2391
4.2387
4.2382
4.2378
4.2374
4.2371
4.2367
4.2363
4.2360
4.2357
4.2353
4.2350
4.2347
4.2344
4.2342
4.2339
4.2336
4.2334
4.2332
4.2329
4.2327
4.2325
4.2323
4.2323

3.3568
3.3548
3.3373
3.3232
3.3116
3.3019
3.2937
3.2867
3.2808
3.2756
3.2712
3.2674
3.2641
3.2612
3.2588
3.2566
3.2547
3.2532
3.2518
3.2506
3.2496
3.2488
3.2482
3.2476
3.2472
3.2469
3.2467
3.2466
3.2466
3.2466
3.2467
3.2469
3.2472
3.2472

-----

0.8009
0.8046
0.8078
0.8103
0.8123
0.8135
0.8141
0.8138
0.8127
0.8108
0.8081
0.8044
0.8000
0.8000

---- --

465

Appendix 2. Lee and Kesler Method

Table A2.7
Logarithm (base 10) of the fugacity coefficient of the simple fluid (with opposite sign)
--- ---

0.8

0.9

0.94

0.98

1.02

1.04

1.06

1.1

1.2

1.5

0.0100
0.0100

0.0028
0.0028

0.0020
0.0020

0.0017
0.0017

0.0015
0.0015

0.0014
0.0014

0.0014

0.0013

0.001 1
0.001 1

O.ooo8
0.0008

O.OOO4
O.OOO4

0.0056
0.0085

0.0040
0.0060

0.0035
0.0052

0.0031
0.0046

0.0029

0.0013
0.0026

0.0012
0.0012

0.0200
0.0300

0.0014
0.0027

0.0113
0.0142
0.0170

0.0080
0.0100
0.0120

0.0070
0.0088

0.0062
0.0077

0.0058
0.0073

0.0039
0.0052

0.0024
0.0036
0.0049

0.0022
0.0033
0.0043

0.0017
0.0025

0.0400
0.0500

0.0041
0.0055
0.0068

0.0008
0.0012
0.0016

0.0087
0.0102

0.0082
0.0096

0.0054
0.0065

0.0140
0.0160

0.0093
0.0108
0.0124

0.0061
0.0073

0.0199
0.0229

0.0105
0.0123
0.0141

0.0065
0.0078

0.0900
0.1000

0.0258
0.0287

0.0181
0.0201

0.0159
0.0177

0.0140
0.0156

0.0117
0.0131
0.0146

0.0110
0.0124
0.0138

0.0091
0.0104
0.0117

0.0085
0.0098
0.0110

0.0076
0.0087
0.0098

0.1100
0.1200

0.0317
0.0347

0.0222
0.0243

0.0194
0.0212

0.0171
0.0187

0.0161
0.0176

0.0152
0.0166

0.0130
0.0143
0.0156

0.0122
0.0135

0.1300
0.1400
0.1500

0.0377
0.0407
0.0438

0.0263
0.0284
0.0305

0.0231
0.0249
0.0267

0.0203
0.0219
0.0235

0.0191
0.0206
0.0221

0.0180
0.0194
0.0208

0.1600
0.1700

0.0468
0.0499

0.0326
0.0347

0.0285
0.0304

0.0251
0.0267

0.0236
0.0251

0.0222
0.0236

0.0196
0.0209
0.0222

0.1800
0.1900

0.0530
0.0562

0.0322
0.0340

0.0283
0.0299

0.0593
0.0625
0.0657

0.0359
0.0378
0.0396

0.0279
0.0293
0.0307

0.2300
0.2400
0.2500

0.0690

0.0315
0.0332
0.0348
0.0364

0.0266
0.0281
0.0296

0.0250
0.0264

0.2000
0.2100
0.2200

0.0368
0.0389
0.0411

0.0600
0.0700
0.0800

0.0432
0.0454

0.0028
0.0032

0.0109
0.0120

0.0075
0.0083
0.0091

0.0036
0.0040
0.0044

0.0147
0.0160
0.0172

0.0131
0.0142
0.0153

0.0100
0.0108
0.0117

0.0048
0.0051
0.0055

0.0184
0.0197

0.0164
0.0175

0.0125
0.0133

0.0059
0.0063

0.0236
0.0249

0.0209
0.0222
0.0235

0.0187
0.0198
0.0209

0.0142
0.0150
0.0159

0.0067
0.0071
0.0075

0.0262
0.0276

0.0247
0.0260

0.0220
0.0231

0.0167
0.0175

0.0079
0.0083

0.0289
0.0303
0.0316
0.0330

0.0272
0.0285
0.0298
0.0311

0.0242
0.0254
0.0265

0.0184
0.0192
0.0201

0.0087
0.0091
0.0095

0.0169
0.0182

0.0533
0.0541

0.0465
0.0465
0.0472

0.0408
0.0408
0.0414

0.0383
0.0383
0.0388

0.0359
0.0359
0.0365

0.0338
0.0338
0.0343

0.0319
0.0319
0.0323

0.0276
0.0283
0.0283
0.0288

0.0209
0.0215
0.0215
0.0218

0.0099
0.0101
0.0101
0.0103

0.1028
0.1179

0.0563
0.0585

0.0491
0.0510

0.0430
0.0447

0.0404
0.0419

0.0379
0.0394

0.0357
0.0371

0.0336
0.0349

0.0299
0.0310

0.0226
0.0235

0.1324
0.1464
0.1599
0.1730
0.1856
0.1979
0.2097

0.0607
0.0630
0.0652
0.0675
0.0698
0.0721
0.0744
0.0767

0.0529
0.0548
0.0568
0.0587
0.0607
0.0626
0.0646
0.0666

0.0790
0.0814
0.0838
0.0861
0.0885

0.0686
0.0706
0.0726
0.0746
0.0766

0.0464
0.0480
0.0497
0.0514
0.0531
0.0548
0.0565
0.0582
0.0599

0.0435
0.0451
0.0466
0.0482
0.0498
0.0514
0.0530
0.0545
0.0561
0.0577
0.0594
0.0610
0.0626

0.0409
0.0423
0.0438
0.0453
0.0467
0.0482
0.0497
0.0512
0.0527
0.0542
0.0557
0.0572
0.0587

0.0384
0.0398
0.0412
0.0425
0.0439
0.0453
0.0467
0.0481
0.0495
0.0509
0.0523
0.0537
0.0551

0.0362
0.0375
0.0388
0.0400
0.0413
0.0426
0.0439
0.0452
0.0466
0.0479
0.0492
0.0505
0.0518

0.0322
0.0333
0.0344
0.0356
0.0367
0.0379
0.0390
0.0402
0.0413
0.0425
0.0436

0.0243
0.0252
0.0260
0.0269
0.0278
0.0286
0.0295
0.0303
0.0312
0.0320
0.0329
0.0338
0.0346

0.0107
0.0111
0.0115
0.0119
0.0123
0.0127
0.0131
0.0135
0.0138
0.0142
0.0146
0.0150
0.0154
0.0158
0.0162

0.2700
0.2800
0.2900
0.3000
0.3100
0.3200
0.3300
0.3400
0.3500
0.3600
0.3700
0.3800
0.3900
0.4100

0.0058
0.0066

0.0357
0.0373

0.0811
0.0871

0.4OoO

0.0020
0.0024

0.0381
0.0397

0.0722
0.0755
0.0776

0.2563
0.2563
0.2600

0.0312
0.0327
0.0342

0.0033
0.0041
0.0050

0.0322
0.0336
0.0350

0.0475
0.0497
0.0519
0.0533

0.2213
0.2324
0.2433
0.2539
0.2641
0.2741

0.0415
0.0434
0.0453

0.0044

0.0616
0.0634
0.0651
0.0669

0.0448
0.0460

466

Appendix 2. L e e and Kesler Method

Table A2.7 (cont.)


Logarithm (base 10) of the fugacity coefficient of the simple fluid (with opposite sign)

0.8

0.9

0.94

0.98

1.02

1.04

1.06

1.1

1.2

1.5

0.0909
0.0934
0.0958
0.0983
0.1008
0.1033
0.1058
0.1084
0.1110
0.1136
0.1162
0.1188
0.1208
0.1225
0.1244
0.1316
0.1387
0.1456
0.1525
0.1592
0.1657
0.1722
0.1785
0.1847
0.1909
0.1969
0.2028
0.2086
0.2143
0.2199
0.2234
0.2234
0.2254
0.2309
0.2362
0.2415
0.2467
0.2518
0.2568
0.2618
0.2667
0.2715
0.2762
0.2809

0.0787
0.0807
0.0828
0.0849
0.0869
0.0890
0.0912
0.0933
0.0954
0.0976
0.0997
0.1019
0.1035
0.1035
0.1041
0.1063
0.1085
0.1108
0.1130
0.1153
0.1176
0.1199
0.1223
0.1246
0.1270
0.1294
0.1319
0.1343
0.1368
0.1394
0.1410
0.1420
0.1440
0.1494
0.1548
0.1600
0.1651
0.1702
0.1752
0.1801
0.1850
0.1898
0.1945
0.1991

0.0686
0.0704
0.0721
0.0739
0.0757
0.0775
0.0793
0.0811
0.0829
0.0847
0.0865
0.0884
0.0897
0.0897
0.0902
0.0921
0.0939
0.0958
0.0977
0.0996
0.1015
0.1034
0.1053
0.1073
0.1092
0.1112
0.1132
0.1151
0.1171
0.1192
0.1205
0.1205
0.1212
0.1232
0.1253
0.1274
0.1295
0.1316
0.1337
0.1359
0.1381
0.1402
0.1425
0.1447

0.0642
0.0659
0.0675
0.0691
0.0708
0.0725
0.0741
0.0758
0.0775
0.0791
0.0808
0.0825
0.0838
0.0838
0.0842
0.0860
0.0877
0.0894
0.0911
0.0929
0.0946
0.0964
0.0982
0.0999
0.1017
0.1035
0.1053
0.1071
0.1090
0.1108
0.1120
0.1120
0.1126
0.1145
0.1164
0.1182
0.1201
0.1220
0.1240
0.1259
0.1278
0.1298
0.1318
0.1337

0.0602
0.0617
0.0633

0.0565
0.0580
0.0594
0.0608
0.0622
0.0637
0.0651
0.0666
0.0680
0.0695
0.0709
0.0724
0.0735
0.0735
0.0739
0.0753
0.0768
0.0783
0.0798
0.0813
0.0828
0.0843
0.0858
0.0873
0.0888
0.0904
0.0919
0.0934
0.0950
0.0965
0.0975
0.0975
0.0981
0.0996
0.1012
0.1028
0.1044
0.1059
0.1075
0.1091
0.1107
0.1124
0.1140
0.1156

0.0532
0.0545
0.0558
0.0571
0.0585
0.0598
0.0612
0.0625
0.0639
0.0652
0.0666
0.0680
0.0690
0.0690
0.0693
0.0707
0.0721
0.0735
0.0748
0.0762
0.0776
0.0790
0.0804
0.0818
0.0832
0.0847
0.0861
0.0875
0.0889
0.0904
0.0913
0.0913
0.0918
0.0932
0.0947
0.0961
0.0976
0.0991
0.1005
0.1020
0.1035
0.1050
0.1065
0.1080

0.0471
0.0483
0.0495
0.0506
0.0518
0.0530
0.0542
0.0554
0.0565
0.0577
0.0589
0.0601
0.0610
0.0610
0.0613
0.0625
0.0637
0.0649
0.0661
0.0673
0.0685
0.0698
0.0710
0.0722
0.0734
0.0747
0.0759
0.0771
0.0783
0.0796
0.0804
0.0804
0.0808
0.0821
0.0833
0.0846
0.0858
0.0871
0.0883
0.0896
0.0909
0.0921
0.0934
0.0947

0.0355
0.0364
0.0372
0.0381
0.0390
0.0398
0.0407
0.0416
0.0425
0.0433
0.0442
0.0451
0.0457
0.0457
0.0460
0.0468
0.0477
0.0486
0.0495
0.0504
0.0512
0.0521
0.0530
0.0539
0.0548
0.0557
0.0566
0.0575
0.0584
0.0592
0.0598
0.0598
0.0601
0.0610
0.0619
0.0628
0.0637
0.0646
0.0655
0.0664
0.0673
0.0682
0.0691
0.0701

0.0166
0.0170
0.0174
0.0178
0.0182
0.0186
0.0190
0.0194
0.0198
0.0202
0.0206
0.0210
0.0212
0.0212
0.0213
0.0217
0.0221
0.0225
0.0229
0.0233
0.0237
0.0241
0.0245
0.0249
0.0253
0.0257
0.0261
0.0265
0.0269
0.0272
0.0275
0.0275
0.0276
0.0280
0.0284
0.0288
0.0292
0.0296
0.0300
0.0304
0.0308
0.0312
0.0316
0.0320

0.4200
0.4300
0.4400
0.4500
0.4600
0.4700
0.4800
0.4900
0.5000
0.5100
0.5200
0.5300
0.5374
0.5374
0.5400
0.5500
0.5600
0.5700
0.5800
0.5900

0.6OoO
0.6100
0.6200
0.6300

0.6400
0.6500
0.6600
0.6700
0.6800
0.6900
0.6963
0.6963
0.7000
0.7100
0.7200
0.7300
0.7400
0.7500
0.7600
0.7700
0.7800
0.7900
0.8000
0.8100

0.2839
0.2934
0.3027
0.3117
0.3205
0.3292
0.3376
0.3458
0.3539
0.3618
0.3695
0.3770
0.3826
0.3826
0.3845
0.3917
0.3988
0.4058
0.4126
0.4193
0.4259
0.4324
0.4387
0.4450
0.4511
0.4571
0.4630
0.4689
0.4746
0.4802
0.4837
0.4837
0.4858
0.4912
0.4966
0.5018
0.5070
0.5122
0.5172
0.5222
0.5271
0.5319
0.5366
0.5413

0.0648
0.0663
0.0679
0.0694
0.0710
0.0725
0.0741
0.0757
0.0772
0.0784
0.0784
0.0788
0.0804
0.0820
0.0836
0.0852
0.0868
0.0884
0.0901
0.0917
0.0933
0.0950
0.0966
0.0983
0.0999
0.1016
0.1033
0.1044
0.1044
0.1050
0.1067
0.1084
0.1101
0.1118
0.1136
0.1153
0.1170
0.1188
0.1206
0.1224
0.1242

467

Appendix 2. Lee and Kesler Method

Table A2.7 (cont. and end)


Logarithm (base 10) of the fugacity coefficient of the simple fluid (with opposite sign)

0.8200
0.8300
0.8400

0.8500
0.8600
0.8700
0.8800
0.8890
0.8890
0.8900
0.9000
0.9100
0.9200
0.9300
0.9400
0.9500
0.9600
0.9700
0.9800
0.9900
1.oooo
1.0100
1.0200
1.0300
1.04oO
1.0500
1.0600
1.0700
1.0800
1.0900
1.1000
1.1100
1.1200
1.1300
1.1400
1.1500
1.1600
1.1700
1.1800
1.1900
1.2000
1.2000

0.8

0.9

0.94

0.98

1.02

1.04

1.06

1.1

1.2

1.5

0.5459

0.2855
0.2901
0.2946
0.2990
0.3033
0.3077
0.3119
0.3157
0.3157
0.3161
0.3202
0.3243
0.3284
0.3323
0.3363
0.3402
0.3440
0.3478
0.3515
0.3552
0.3589
0.3625
0.3661
0.3696
0.3731
0.3765
0.3800
0.3833
0.3867
0.3900
0.3932
0.3965
0.3996
0.4028
0.4059
0.4090
0.4121
0.4151
0.4181
0.4211
0.4240
0.4240

0.2037
0.2082
0.2127
0.2171
0.2214
0.2257
0.2299
0.2337
0.2337
0.2341
0.2382
0.2423
0.2463
0.2502
0.2541
0.2580
0.2618
0.2656
0.2693
0.2730
0.2766
0.2802
0.2838
0.2873
0.2907
0.2942
0.2976
0.3009
0.3042
0.3075
0.3107
0.3139
0.3171
0.3203
0.3234
0.3264
0.3295
0.3325
0.3355
0.3384
0.3413
0.3413

0.1470
0.1493
0.1516
0.1540
0.1564
0.1589
0.1614
0.1637
0.1641
0.164s
0.1685
0.1725
0.1764
0.1802
0.1841
0.1878
0.1916
0.1952
0.1989
0.2025
0.2061
0.2096
0.2131
0.2165
0.2199
0.2233
0.2266
0.2299
0.2332
0.2364
0.2396
0.2427
0.2459
0.2489
0.2520
0.2550
0.2580
0.2610
0.2639
0.2668
0.2697
0.2697

0.1357
0.1378
0.1398
0.1419
0.1439
0.1460
0.1482
0.1501
0.1501
0.1503
0.1525
0.1547
0.1569
0.1592
0.1615
0.1638
0.1662
0.1686
0.1711
0.1737
0.1765
0.1798
0.1831
0.1864
0.1896
0.1929
0.1961
0.1993
0.2025
0.2056
0.2087
0.2118
0.2148
0.2178
0.2208
0.2238
0.2267
0.2296
0.2325
0.2354
0.2382
0.2382

0.1260
0.1278
0.1296
0.1314
0.1333
0.1351
0.1370
0.1387
0.1387
0.1389
0.1408
0.1427
0.1447
0.1466
0.1486
0.1506
0.1526
0.1546
0.1567
0.1587
0.1608
0.1629
0.1651
0.1673
0.1695
0.1717
0.1740
0.1764
0.1787
0.1812
0.1837
0.1862
0.1889
0.1915
0.1942
0.1969
0.1996
0.2023
0.2050
0.2077
0.2103
0.2103

0.1173
0.1189
0.1206
0.1222
0.1239
0.1256
0.1273
0.1288
0.1288
0.1290
0.1307
0.1324
0.1341
0.1359
0.1376
0.1394
0.1412
0.1429
0.1447
0.1465
0.1484
0.1502
0.1521
0.1539
0.1558
0.1577
0.1596
0.1615
0.1635
0.1655
0.1675
0.1695
0.1715
0.1735
0.1756
0.1777
0.1798
0.1820
0.1841
0.1863
0.1885
0.1885

0.1095
0.1110
0.112s
0.1140
0.1155
0.1171
0.1186
0.1200
0.1200
0.1201
0.1217
0.1233
0.1248
0.1264
0.1280
0.1296
0.1312
0.1328
0.1344
0.1360
0.1377
0.1393
0.1409
0.1426
0.1443
0.1459
0.1476
0.1493
0.1510
0.1527
0.1545
0.1562
0.1.580
0.1597
0.1615
0.1633
0.1651
0.1669
0.1687
0.1705
0.1724
0.1724

0.0960
0.0973
0.0986
0.0999
0.1011
0.1024
0.1038
0.1049
0.1049
0.1051
0.1064
0.1077
0.1090
0.1103
0.1117
0.1130
0.1143
0.1157
0.1170
0.1184
0.1197
0.1211
0.1224
0.1238
0.1252
0.1265
0.1279
0.1293
0.1307
0.1321
0.1335
0.1349
0.1363
0.1377
0.1391
0.1406
0.1420
0.1434
0.1449
0.1463
0.1478
0.1478

0.0710
0.0719
0.0728
0.0737
0.0746
0.0755
0.0764
0.0773
0.0773
0.0774
0.0783
0.0792
0.0801
0.0810
0.0820
0.0829
0.0838
0.0847
0.0857
0.0866
0.0875
0.0884
0.0894
0.0903
0.0912
0.0922
0.0931
0.0941
0.0950
0.0959
0.0969
0.0978
0.0988
0.0997
0.1007
0.1016
0.1026
0.1035
0.1045
0.1054
0.1064
0.1064

0.0323
0.0327
0.0331
0.0335
0.0339
0.0343
0.0347
0.0350
0.0350
0.0351
0.0355
0.0359
0.0363
0.0366
0.0370
0.0374
0.0378
0.0382
0.0386
0.0390
0.0394
0.0398
0.0402
0.0405
0.0409
0.0413
0.0417
0.0421
0.0425
0.0429
0.0433
0.0437
0.0440

0.5505
0.5550
0.5594
0.5638
0.5681
0.5724
0.5761
0.5761
0.5766
0.5807
0.5848
0.5888
0.5928
0.5968
0.6006
0.6045

0.6083
0.6120
0.6157
0.6194
0.6230
0.6266
0.6301
0.6336
0.6371
0.6405
0.6438
0.6472
0.6505
0.6537
0.6570
0.6601
0.6633
0.6664
0.6695
0.6726
0.6756
0.6786
0.6816
0.6845
0.6845

0.0444

0.0448
0.0452
0.0456
0.0460

0.0464
0.0467
0.0471
0.0471

468

Appendix 2. Lee and Kesler Method

Table A2.8
Logarithm (base 10) of the fugacity coefficient (with opposite sign): corrective term

- - - - - - - - -- -

0.0100
0.0100
0.0200
0.0300
0.04oO
0.0500
0.0600
0.0700
0.0800
0.09oO

0.1000
0.1100
0.1200
0.1300
0.1400
0.1500
0.1600
0.1700
0.1800
0.1900
0.2000
0.2100
0.2200
0.2300
0.2400
0.2500
0.2563
0.2563
0.2600
0.2700
0.2800
0.2900
0.3000
0.3100
0.3200
0.3300
0.3400
0.3500
0.3600
0.3700
0.3800
0.3900
0.4OoO
0.4100

0.8

0.9

0.94

0.98

1.02

1.04

1.06

0.0019
0.0019
0.0038
0.0057
0.0077
0.0096
0.0116
0.0136
0.0157
0.0177
0.0198
0.0220
0.0241
0.0263
0.0286
0.0308
0.0331
0.0355
0.0379
0.0403
0.0428
0.0454
0.0480
0.0506
0.0534
0.0562
0.0580
0.4769
0.4770
0.4773
0.4776
0.4779
0.4782
0.4785
0.4788
0.4791
0.4794
0.4797
0.4800
0.4803
0.4806
0.4809
0.4812
0.4814

0.0008
0.0008
0.0017
0.0025
0.0034
0.0043
0.0051

0.0006
0.0006
0.0012
0.0018
0.0023
0.0029
0.0035
0.0041
0.0047
0.0053
0.0059
0.0065
0.0071
0.0078
0.0084
0.0090
0.0096
0.0102
0.0109
0.0115
0.0121
0.0128
0.0134
0.0141
0.0147
0.0154
0.0158
0.0158
0.0160
0.0167
0.0173
0.0180
0.0187
0.0194
0.0200
0.0207
0.0214
0.0221
0.0228
0.0235
0.0242
0.0250
0.0257
0.0264

O.ooo4
O.OOO4

0.0003
0.0003
O.OOO6

0.0002
0.0002

O.OOO9

0.0007
0 . m
0.0011
0.0013
0.0016
0.0018
0.0020
0.0022
0.0025
0.0027
0.0029
0.0031
0.0034
0.0036
0.0038
0.0040
0.0042
0.0045
0.0047
0.0049
0.0051
0.0053
0.0056
0.0057
0.0057
0.0058

0.0002
0.0002
0.0003
0.0005
0.0006
0.0008

0.0001
0.0001
0.0002
0.0003

O.Oo0

O.OOO6

0.0011
0.0013
0.0014
0.0016
0.0017
0.0019
0.0020
0.0022
0.0023
0.0025
0.0026
0.0028
0.0029
0.0031
0.0032
0.0034
0.0035
0.0037
0.0038
0.0039
0.0039
0.0040
0.0041
0.0043
0.0044
0.0046
0.0047
0.0049
0.0050
0.0051
0.0053
0.0054
0.0056
0.0057
0.0058

0.0007
0.0008
0.0009
0.0010
0.0011
0.0012
0.0013
0.0014
0.0014
0.0015
0.0016
0.0017
0.0018
0.0019
0.0020
0.0021
0.0021
0.0022
0.0023
0.0024
0.0024
0.0024
0.0025
0.0026
0.0026
0.0027
0.0028
0.0029
0.0030
0.0030
0.0031
0.0032
0.0032
0.0033
0.0034
0.0035
0.0035

O.Oo60

0.0069
0.0077
0.0086
0.0095
0.0104
0.0113
0.0122
0.0132
0.0141
0.0150
0.0160
0.0169
0.0179
0.0188
0.0198
0.0208
0.0218
0.0228
0.0234
0.0234
0.0238
0.0248
0.0258
0.0269
0.0279
0.0290
0.0301
0.0311
0.0322
0.0334
0.0345
0.0356
0.0368
0.0380
0.0392
0.0404

0.0008
0.0011
0.0015
0.0019
0.0023
0.0027
0.0031
0.0035
0.0039
0.0042
0.0046
0.0050
0.0054
0.0058
0.0062

0.0066
0.0070
0.0074
0.0078
0.0082
0.0086
0.0090
0.0094
0.0098
0.0101
0.0101
0.0102
0.0106
0.0110
0.0114
0.0118
0.0122
0.0127
0.0131
0.0135
0.0139
0.0143
0.0147
0.0152
0.0156
0.0160
0.0164

0.0012
0.0015
0.0018
0.0021
0.0024
0.0027
0.0030
0.0033
0.0036
0.0039
0.0042
0.0045
0.0048
0.0051
0.0054
0.0057
O.Oo60

0.0063

0.0066
0.0069
0.0072
0.0075
0.0077
0.0077
0.0078
0.0082
0.0085

0.0088
0.0091
0.0094
0.0097
0.0100
0.0103
0.0106

0.0109
0.0112
0.0115
0.0119
0.0122
0.0125

O.ooo4

O.Oo60

0.0062
0.0064
0.0067
0.0069
0.0071
0.0073
0.0075
0.0078
0.0080
0.0082

0.0084
0.0086
0.0088
0.0091

------

O.Oo60

0.0061

O.ooo4
0.0005

1.1

1.2

O.oo00 4.0002
O.oo00 -0.ooo2
O.oo00 4.0003
O.oo00 -0.0005
O.oo00 4.0007
O.oo00 4.0008
O.oo00 4.0010
O.oo00
O.oo00
O.oo00
O.oo00

O.oo00
O.oo00
O.oo00
O.oo00
-0.0001
4.0001
-0.0001
4.0001
4.0001
4.0001
4.0001
-0.0001
-0.0002
4.0002
-0.0002
4.0002
-0.0002
4.0002
-0.0003
-0.0003
-0.0003
-0.0003

4.0012
-0.0013
4.0015
-0.0017
4.0018
-0.0020
4.0022
4.0024
-0.0025
4.0027
-0.0029
4.0031
-0.0033
4.0034
4.0036
-0.0038
4.0040
4.0042
-0.0044

-0.0045
-0.0045
4.0045
-0.0047
-0.0049
-0.0051
4.0053
-0.oO04 -0.0055
-0.oO04 -0.0057
-0.ooo4 4.0059
4.oO04 4.0061
4.0005 4.0063
-0.0005 -0.0065
-0.0005 -0.0067
-0.ooo6 4.0069
-O.m -0.0071
-0.0007 4.0073
4.0007 -0.0075

1.5
4.0003
-0.0003
4.0007
-0.0010
4.0013
4.0017
4.0020
-0.0023
-0.0027
4.0030
-0.0034
4.0037
4.0040
-0.0044
4.0047
-0.0051
-0.0054
-0.0057
-0.0061
4.0064
-0.0068
4.0071
-0.0075
-0.0078
4.0081
-0.0085
-0.0087
4.0087
4.0088
-0.0092
-0.0095

-0.0099
-0.0102
-0.0106
-0.0109
-0.0113
4.0116
-0.0120
-0.0123
-0.0127
4.0130
-0.0134
4.0137
-0.0141

- - -- -

469

Appendix 2. Lee and Kesler Method

Table A2.8 (cont.)


Logarithm (base 10) of the fugacity coefficient (with opposite sign): corrective term
---0.8

0.9

0.94

0.98

--1

1.02

1.04

- -- 1.06

1.1

1.2

0.4200
0.4300
0.4400
0.4500
0.4600
0.4700
0.4800
0.4900
0.5000
0.5100

0.4817
0.4820
0.4823
0.4826
0.4829
0.4832
0.4835
0.4838
0.4841
0.4844

0.5200

3.4047 0.0559 0.0350

0.5300
0.5374
0.5374
0.5400
0.5500
0.5600
0.5700
0.5800
0.5900
0.6OOO
0.6100
0.6200
0.6300
0.6400
0.6500
0.6600
0.6700
0.6800
0.6900
0.6963
0.6963
0.7000
0.7100
0.7200
0.7300
0.7400
0.7500
0.7600
0.7700
0.7800
0.7900
0.8000
0.8100

0.4850
0.4852
0.4852
0.4853
0.4856
0.4858
0.4861
0.4864
0.4867
0.4870
0.4873
0.4876
0.4879
0.4882
0.4885
0.4887
0.4890
0.4893
0.4896
0.4898
0.4898
0.4899
0.4902
0.4905
0.4908
0.4911
0.4913
0.4916
0.4919
0.4922
0.4925
0.4928
0.4931

0.0417
0.0429
0.0442
0.0456
0.0469
0.0483
0.0497
0.0512
0.0527
0.0543
0.0576
0.0595
0.1971
0.1971
0.1974
0.1977
0.1980
0.1983
0.1986
0.1989
0.1992
0.1994
0.1997
0.2000
0.2003
0.2006
0.2009
0.2011
0.2014
0.2016
0.2016
0.2017
0.2020
0.2023
0.2025
0.2028
0.2031
0.2034
0.2036
0.2039
0.2042
0.2045
0.2047

0.0272
0.0279
0.0287
0.0294
0.0302
0.0310
0.0317
0.0325
0.0334
0.0342
0.0358
0.0365
0.0365
0.0367
0.0376
0.0385
0.0394
0.0403
0.0412
0.0422
0.0432
0.0442
0.0452
0.0463
0.0474
0.0486
0.0498
0.0512
0.0526
0.0535
0.1183
0.1184
0.1187
0.1190
0.1193
0.1196
0.1199
0.1202
0.1205
0.1208
0.1211
0.1214
0.1217

1.5

0.0168
0.0173
0.0177
0.0181
0.0186
0.0190
0.0194
0.0199
0.0203
0.0208

0.0128
0.0131
0.0134
0.0137
0.0140
0.0143
0.0146
0.0150
0.0153
0.0156
3.0212 0.0159
0.0216 0.0162
0.0220 0.0164
0.0220 0.0164
0.0221 0.0165
0.0225 0.0168
0.0230 0.0172
0.0234 0.0175
0.0239 0.0178
0.0244 0.0181
0.0248 0.0184
0.0253 0.0187
0.0258 0.0190
0.0262 0.0194
0.0267 0.0197
0.0272 0.0200
0.0277 0.0203
0.0282 0.0206
0.0286 0.0209
0.0291 0.0213
0.0295 0.0215
0.0295 0.0215
0.0296 0.0216
0.0302 0.0219
0.0307 0.0222
0.0312 0.0225
0.0317 0.0228
0.0322 0.0232
0.0328 0.0235
0.0333 0.0238
0.0339 0.0241
0.0345 0.0244
0.0351 0.0248
0.0357 0.0251

0.0093
0.0095
0.0097
0.0099
0.0101
0.0103
0.0105
0.0108
0.0110
0.0112

0.0062
0.0064
0.0065
0.0066
0.0068
0.0069
0.0070
0.0071
0.0073
0.0074

0.0114 3.0075
0.0116
0.0117
0.0117
0.0118
0.0120
0.0122
0.0124
0.0126
0.0128
0.0130
0.0132
0.0134
0.0136
0.0138
0.0140
0.0142
0.0144
0.0146
0.0148
0.0149
0.0149
0.0150
0.0151
0.0153
0.0155
0.0157
0.0159
0.0160
0.0162
0.0164
0.0166
0.0167
0.0169

0.0076
0.0077
0.0077
0.0077
0.0079
0.0080
0.0081
0.0082
0.0083
0.0084
0.0085
0.0086
0.0088
0.0089
0.0090
0.0091
0.0092
0.0092
0.0093
0.0094
0.0094
0.0094
0.0095
0.0096
0.0097
0.0098
0.0098
0.0099
0.0100
0.0101
0.0101
0.0102
0.0102

0.0036
0.0036
0.0037
0.0038
0.0038
0.0039
0.0039
0.0040
0.0041
0.0041
1.0042
0.0042
0.0042
0.0042
0.0043
0.0043
0.0043
0.0044
0.0044
0.0045
0.0045
0.0045
0.0046
0.0046
0.0046
0.0047
0.0047
0.0047
0.0047
0.0047
0.0047
0.0047
0.0047
0.0048
0.0048
0.0048
0.0048
0.0048
0.0048
0.0048
0.0048
0.0047
0.0047
0.0047

-0.0007
4.0008
4.0008
-0.oO09
4.oO09
-0.0010
4.0010
4.0011
-0.0012
4.0012

-0.0077
4.0079
-0.0082
-0.0084
4.0086
-0.0088
4.0090
-0.0092
-0.0094
4.0097

-0,0144
4.0148
-0.0151
-0.0155
4.0158
-0.0162
-0.0165
-0.0169
-0.0172
4.0176

0.0013 4.0096 4.0179


4.0013
4.0014
4.0014
4.0014
4.0015
-0.0015
4.0016
-0.0017
4.0017
4.0018
-0.0019
4.0020
-0.0021
-0.0021
-0.0022
4.0023
-0.0024
4.0025
4.0026
-0.0027
4.0027
-0.0027
4.0028
-0.0029
-0.0030
-0.0031
-0.0032
-0.0033
-0.0035
-0.0036
-0.0037
-0.0038
-0.oO40

4.0101
-0.0103
4.0103
4.0103
-0.0106
-0.0108
-0.01 10
-0.01 13
-0.0115
-0.0117
-0.0120
-0.0122
-0.0124
-0.0127
-0.0129
4.0132
-0.0134
4.0137
4.0139
-0.0141
4.0141
-0.0142
4.0144
4.0147
-0.0149
-0.0152
-0.0155
-0.0157
-0.0160
-0.0163
-0.0165
-0,0168
-0.0171

4.0183
-0.0186
-0.0186
4.0187
4.0190
-0.0194
4.0197
-0.0201
-0.0205
4.0208
-0.0212
4.0215
-0.0219
-0.0223
-0.0226
4.0230
-0.0233
-0.0237
4.0241
-0.0243
4.0243
-0.0244
4.0248
-0.0252
-0.0255
-0.0259
-0.0263
-0.0266
-0.0270
-0.0274
-0.0277
-0.0281
4.0285

470

Appendix 2. Lee and Kesler Method

Table A2.8 (cont. and end)


Logarithm (base 10) of the fugacity coefficient (with opposite sign): corrective term

x
0.8200
0.8300
0.8400
0.8500
0.8600
0.8700
0.8800
0.8890
0.8890
0.8900
0.9000
0.9100
0.9200
0.9300
0.9400
0.9500
0.9600
0.9700
0.9800
0.9900
1.m
1.0100
1.0200
1.0300
1.0400
1.0500
1.0600
1.0700
1.0800
1.0900
1.1000
1.1100
1.1200
1.1300
1.1400
1.1500

1.1600
1.1700
1.1800
1.1900
1.2000
1.2000

--------

--

0.8

0.9

0.94

0.98

1.02

1.04

1.06

1.1

1.2

1.5

0.4933
0.4936
0.4939
0.4942
0.4945
0.4948
0.4950
0.4953
0.4953
0.4953
0.4956
0.4959
0.4962
0.4965
0.4967
0.4970
0.4973
0.4976
0.4979
0.4982
0.4984
0.4987
0.4990
0.4993
0.4996
0.4998
0.5001
0.5004
0.5007
0.5010
0.5012
0.5015
0.5018
0.5021
0.5024
0.5026
0.5029
0.5032
0.5035
0.5038
0.5040
0.5040

0.2050
0.2053
0.2056
0.2058
0.2061
0.2064
0.2066
0.2069
0.2069
0.2069
0.2072
0.2074
0.2077
0.2080
0.2082
0.2085
0.2088
0.2090
0.2093
0.2096
0.2098
0.2101
0.2103
0.2106
0.2109
0.2111
0.2114
0.2116
0.2119
0.2122
0.2124
0.2127
0.2129
0.2132
0.2134
0.2137
0.2140
0.2142
0.2145
0.2147
0.2150
0.2150

0.1219
0.1222
0.1225
0.1228
0.1231
0.1233
0.1236
0.1238
0.1238
0.1239
0.1241
0.1244
0.1247
0.1250
0.1252
0.1255
0.1258
0.1260
0.1263
0.1266
0.1268
0.1271
0.1273
0.1276
0.1279
0.1281
0.1284
0.1286
0.1289
0.1291
0.1294
0.1296
0.1299
0.1302
0.1304
0.1307
0.1309
0.1312
0.1314
0.1316
0.1319
0.1319

0.0363
0.0369
0.0376
0.0383
0.0391
0.0400
0.0410

0.0254
0.0257
0.0260
0.0264
0.0267
0.0270
0.0273
0.0276
0.0276
0.0276
0.0280
0.0283
0.0286
0.0289
0.0292
0.0296
0.0299
0.0302
0.0305
0.0309
0.0312
0.03 15
0.0318
0.0321
0.0324
0.0327
0.0329
0.0332
0.0334
0.0337
0.0339
0.0341
0.0344
0.0346
0.0348
0.0351
0.0353
0.0355
0.0357
0.0360
0.0362
0.0362

0.0170
0.0172
0.0174
0.0175
0.0176
0.0178
0.0179
0.0180
0.0180
0.0181
0.0182
0.0183
0.0184
0.0185
0.0186
0.0187
0.0188
0.0189
0.0189
0.0190
0.0190
0.0190
0.0190
0.0190
0.0190
0.0189
0.0188
0.0186
0.0184
0.0182
0.0178
0.0175
0.0170
0.0166
0.0162
0.0158
0.0155
0.0153
0.0151
0.0149
0.0148
0.0148

0.0103
0.0103
0.0104
0.0104
0.0105
0.0105
0.0105
0.0105
0.0105
0.0105
0.0106
0.0106
0.0106
0.0106
0.0106
0.0106
0.0105
0.0105
0.0105
0.0104
0.0104
0.0103
0.0102
0.0101
0.0100
0.0099
0.0098
0.0096
0.0095
0.0093
0.0091
0.0089
0.0086
0.0084
0.0081
0.0078
0.0074
0.0071
0.0067
0.0063
0.0058
0.0058

0.0047
0.0046
0.0046
0.0046
0.0045
0.0045
0.0044
0.0044
0.0044
0.0044
0.0043
0.0043
0.0042
0.0041
0.0040
0.0040
0.0039
0.0038
0.0037
0.0035
0.0034
0.0033
0.0032
0.0030
0.0029
0.0027
0.0025
0.0024
0.0022
0.0020
0.0018
0.0015
0.0013
0.0011
0.0008
0.0005
0.0002
4.0001
4.m
4.0007
4.0010
4.0010

4.0041
-0.0042
4.0044
-0.0045
4.0047
4.0048
4.0050
-0.0051
4.0051
4.0051
4.0053
-0.0054
-0.0056
4.0058
-0.0060
4.0061
4.0063
4.0065
4.0067
4.0069
-0.0071
4.0073
4.0075
4.0078
4.0080
-0.0082
4.0084
4.0087
-0.0089
4.0092
4.0094
4.0097
4.0100
4.0103
4.0105
4.0108
-0.01 11
-0.0114
-0.0117
4.0120
4.0124
-0.0124

4.0173
-0.0176
4.0179
-0.0182
4.0184
4.0187
-0.0190
-0.0193
4.0193
-0.0193
4.0196
-0.0199
4.0202
-0.0205
-0.0208
4.021 1
4.0214
-0.0217
4.0220
4.0223
-0.0226
4.0229
-0.0232
4.0236
4.0239
-0.0242
4.0245
-0.0248
4.0252
4.0255
4.0258
-0.0262
-0.0265
-0.0269
-0.0272
-0.0275
-0.0279
-0.0282
-0.0286
-0.0290
-0.0293
-0.0293

4.0288
4.0292
4.0296
-0.0299
4.0303
4.0307
-0.0310
-0.0314
4.0314
-0.0314
4.0318
-0.0322
4.0325
-0.0329
4.0333
4.0336
-0.0340
-0.0344
4.0348
-0.0351
-0.0355
4.0359
-0.0363
4.0366
-0.0370
-0.0374
4.0378
-0.0382
4.0385
4.0389
-0.0393
-0.0397
4.0400
-0.0404
-0.0408
-0.0412
4.0416
4.0419
4.0423
4.0427
4.0431
4.0431

0.0557
0.0557
0.0560
0.0564
0.0567
0.0570
0.0573
0.0576
0.0579
0.0582
0.0585
0.0588
0.0591
0.0593
0.0596
0.0599
0.0601
0.0604
0.0607
0.0609
0.0612
0.0614
0.0617
0.0620
0.0622
0.0625
0.0627
0.0630
0.0632
0.0634
0.0637
0.0639
0.0642
0.0642

471

Appendix 2. lee and Kesler Method

Table A2.9
Adimensional residual heat capacity at constant pressure for the simple fluid

x
0.0100
0.0100
0.0200
0.0300
0.0400
0.0500
0.O600

0.0700
0.0800
0.0900
0.1000
0.1100
0.1200
0.1300
0.1400
0.1500
0.1600
0.1700
0.1800
0.1900
0.2000
0.2100
0.2200
0.2300
0.2400
0.2500
0.2563
0.2563
0.2600
0.2700
0.2800
0.2900
0.3000
0.3100
0.3200
0.3300
0.3400
0.3500
0.3600
0.3700
0.3800
0.3900
0.4000
0.4100

0.8

0.9

0.94

0.98

1.02

1.04

1.06

1.1

1.2

1.5

----- 0.0358
0.0358
0.0728
0.1108
0.1500
0.1906
0.2324
0.2757
0.3206
0.3671
0.4154
0.4655
0.5177
0.5721
0.6289
0.6882
0.7504
0.8155
0.8840
0.9561
1.0323
1.1128
1.1983
1.2892
1.3863
1.4902
1.5593
3.7362
3.7337
3.7272
3.7208
3.7143
3.7080
3.7016
3.6953
3.6891
3.6829
3.6768
3.6707
3.6646
3.6586
3.6526
3.6467
3.6408

0.0251
0.0251
0.0507
0.0769
0.1036
0.1310
0.1589
0.1875
0.2168
0.2467
0.2774
0.3089
0.3411
0.3742
0.4082
0.4431
0.4790
0.5159
0.5538
0.5929
0.6332
0.6748
0.7177
0.7620
0.8079
0.8553
0.8859
0.8859
0.9044
0.9553
1.0082
1.0631
1.1202
1.1796
1.2417
1.3064
1.3741
1.4450
1.5194
1.5975
1.6797
1.7664
1.8581
1.9552

0.0221
0.0221
0.0445
0.0674
0.0907
0.1145
0.1387
0.1634
0.1886
0.2143
0.2406
0.2674
0.2948
0.3228
0.3514
0.3807
0.4107
0.4414
0.4728
0.5050
0.5381
0.5719
0.6067
0.6424
0.6791
0.7168
0.7410
0.7410
0.7556
0.7955
0.8366
0.8790
0.9227
0.9678
1.0144
1.0625
1.1123
1.1639
1.2174
1.2728
1.3304
1.3903
1.4526
1.5175

0.0195
0.0195
0.0393
0.0595
0.0799
0.1008
0.1219
0.1435
0.1654
0.1877
0.2104
0.2336
0.2571
0.2811
0.3056
0.3306
0.3560
0.3820
0.4085
0.4355
0.4631
0.4913
0.5202
0.5496
0.5798
0.6106
0.6303
0.6303
0.6422
0.6745
0.7076
0.7415
0.7763
0.8120
0.8486
0.8861
0.9247
0.9644
1.0052
1.0472
1.0905
1.1350
1.1809
1.2283

0.0184
0.0184
0.0370
0.0560
0.0752

0.0948
0.1146
0.1348
0.1553
0.1762
0.1974
0.2189
0.2409
0.2632
0.2859
0.3091
0.3326
0.3566
0.3811
0.4060
0.4314
0.4574
0.4838
0.5108
0.5383
0.5664
0.5844
0.5844
0.5952
0.6245
0.6545
0.6852
0.7165
0.7486
0.7815
0.8151
0.8496
0.8849
0.9212
0.9583
0.9965
1.0356
1.0759
1.1172

0.0173
0.0173
0.0349
0.0528
0.0709
0.0893
0.1079
0.1269
0.1461
0.1656
0.1854
0.2056
0.2261
0.2469
0.2681
0.2896
0.3115
0.3337
0.3564
0.3794
0.4029
0.4268
0.4512
0.4760
0.5012
0.5270
0.5434
0.5434
0.5533
0.5801
0.6074
0.6353
0.6638
0.6928
0.7225
0.7529
0.7839
0.8156
0.8480
0.8812
0.9152
0.9500
0.9856
1.0221

0.0164
0.0164
0.0330
0.0498
0.0669
0.0842
0.1018
0.1196
0.1376
0.1560
0.1745
0.1934
0.2126
0.2320
0.2518
0.2718
0.2922
0.3129
0.3340
0.3554
0.3771
0.3992
0.4217
0.4446
0.4679
0.4916
0.5067
0.5067
0.5158
0.5403
0.5653
0.5908
0.6168
0.6433
0.6703
0.6978
0.7259
0.7546
0.7838
0.8137
0.8442
0.8753
0.9072
0.9397

0.0155
0.0155
0.0312
0.0471
0.0632
0.0796
0.0961
0.1129
0.1299
0.1471
0.1645
0.1823
0.2002
0.2184
0.2369
0.2557
0.2747
0.2940
0.3136
0.3335
0.3537
0.3743
0.3951
0.4163
0.4379
0.4597
0.4737
0.4737
0.4820
0.5046
0.5276
0.5510
0.5749
0.5991
0.6238
0.6489
0.6745
0.7005
0.7271
0.7541
0.7817
0.8098
0.8385
0.8677

0.0139
0.0139
0.0280
0.0423
0.0567
0.0713
0.0860
0.1010
0.1161
0.1314
0.1469
0.1626
0.1784
0.1945
0.2108
0.2273
0.2440
0.2609
0.2780
0.2954
0.3130
0.3308
0.3489
0.3673
0.3858
0.4047
0.4167
0.4167
0.4238
0.4432
0.4629
0.4829
0.5031
0.5237
0.5446
0.5658
0.5873
0.6092
0.6314
0.6540
0.6769
0.7002
0.7239
0.7479

---------

0.0109
0.0109
0.0218
0.0329
0.0440
0.0553
0.0667
0.0781
0.0897
0.1013
0.1131
0.1250
0.1370
0.1491
0.1613
0.1736
0.1861
0.1987
0.2113
0.2242
0.2371
0.2502
0.2634
0.2767
0.2901
0.3037
0.3123
0.3123
0.3175
0.3313
0.3453
0.3595
0.3738
0.3882
0.4028
0.4176
0.4325
0.4476
0.4628
0.4782
0.4938
0.5095
0.5254
0.5415

0.0058
0.0058
0.0117
0.0176
0.0235
0.0295
0.0354
0.0414
0.0475
0.0535
0.0596
0.0657
0.0718
0.0780
0.0842
0.0904
0.0966
0.1029
0.1091
0.1155
0.1218
0.1282
0.1346
0.1410
0.1474
0.1539
0.1580
0.1580
0.1604
0.1670
0.1735
0.1801
0.1867
0.1934
0.2001
0.2068
0.2135
0.2203
0.2271
0.2339
0.2408
0.2477
0.2546
0.2615

- -

472

Appendix 2. lee and Kesler Method

Table A2.9 (cont.)


Adimensional residual heat capacity at constant pressure for the simple fluid

0.8

0.9

0.94

0.98

---- 0.4200
0.4300
0.4400
0.4500
0.4600
0.4700
0.4800
0.4900
0.5000
0.5100
0.5200
0.5300
0.5374
0.5374
0.5400
0.5500
0.5600
0.5700
0.5800
0.5900
0.6OoO
0.6100
0.6200
0.6300
0.6400
0.6500
0.6600
0.6700
0.6800
0.6900
0.6963
0.6963
0.7000
0.7100
0.7200
0.7300
0.7400
0.7500
0.7600
0.7700
0.7800
0.7900
0.8000
0.8100

3.6349
3.6291
3.6233
3.6176
3.6119
3.6062
3.6006
3.5950
3.5895
3.5840
3.5785
3.5731
3.5691
3.5691
3.5677
3.5624
3.5570
3.5517
3.5465
3.5413
3.5361
3.5309
3.5258
3.5207
3.5157
3.5106
3.5057
3.5007
3.4958
3.4909
3.4878
3.4878
3.4860
3.4812
3.4764
3.4716
3.4669
3.4621
3.4575
3.4528
3.4482
3.4436
3.4390
3.4345

2.0583
2.1680
2.2852
2.4107
2.5455
2.6910
2.8486
3.0201
3.2076
3.4140
3.6424
3.8974
4.1070
5.9122
5.9019
5.8625
5.8239
5.7861
5.7492
5.7130
5.6775
5.6428
5.6088
5.5755
5.5428
5.5108
5.4793
5.4485
5.4183
5.3886
5.3700
5.3700
5.3594
5.3308
5.3027
5.2751
5.2479
5.2213
5.1950
5.1693
5.1440
5.1190
5.0945
5.0704

1.5852
1.6559
1.7298
1.8073
1.8885
1.9739
2.0637
2.1585
2.2586
2.3645
2.4770
2.5966
2.6906
2.6906
2.7242
2.8607
3.0071
3.1648
3.3352
3.5202
3.7218
3.9427
4.1862
4.4564
4.7583
5.0985
5.4859
5.9320
6.4531
7.0722
7.5305
8.9520
8.8980
8.7552
8.6193
8.4896
8.3657
8.2473
8.1338
8.0251
7.9207
7.8205
7.7240
7.6312

1.2772
1.3277
1.3799
1.4339
1.4898
1.5477
1.6078
1.6702
1.7351
1.8025
1.8727
1.9459
2.0023
2.0023
2.0223
2.1021
2.1857
2.2732
2.3650
2.4615
2.5630
2.6701
2.7832
2.9029
3.0299
3.1648
3.3086
3.4622
3.6268
3.8037
3.9229
3.9229
3.9944
4.2007
4.4250
4.6697
4.9380
5.2338
5.5621
5.9288
6.3417
6.8108
7.3496
7.9762

1.1598
1.2035
1.2486
1.2950
1.3428
1.3922
1.4431
1.4957
1.5501
1.6063
1.6645
1.7248
1.7710
1.7710
1.7873
1.8522
1.9195
1.9896
2.0625
2.1384
2.2175
2.3002
2.3866
2.4770
2.5718
2.6712
2.7757
2.8857
3.0017
3.1241
3.2055
3.2055
3.2538
3.3912
3.5372
3.6927
3.8586
4.0363
4.2270
4.4323
4.6541
4.8945
5.1563
5.4424

1.02

1.04

1.06

1.1

1.2

1.5

---

---

1.0596
1.0980
1.1374
1.1778
1.2194
1.2621
1.3059
1.3511
1.3976
1.4454
1.4947
1.5455
1.5843
1.5843
1.5979
1.6521
1.7079
1.7657
1.8255
1.8873
1.9513
2.0177
2.0866
2.1582
2.2325
2.3098
2.3904
2.4743
2.5619
2.6533
2.7135
2.7135
2.7490
2.8491
2.9541
3.0643
3.1802
3.3022
3.4308
3.5666
3.7103
3.8626
4.0244
4.1966

0.7724
0.7973
0.8227
0.8484
0.8746
0.9013
0.9285
0.9561
0.9843
1.0130
1.0422
1.0720
1.0944
1.0944
1.1023
1.1332
1.1648
1.1969
1.2298
1.2632
1.2974
1.3323
1.3679
1.4043
1.4414
1.4794
1.5182
1.5578
1.5984
1.6399
1.6667
1.6667
1.6823
1.7258
1.7702
1.8158
1.8624
1.9102
1.9592
2.0094
2.0609
2.1138
2.1680
2.2236

0.9730
1.0071
1.0419
1.0776
1.1141
1.1515
1.1899
1.2292
1.2695
1.3109
1.3533
1.3970
1.4301
1.4301
1.4418
1.4878
1.5352
1.5839
1.6341
1.6858
1.7391
1.7940
1.8506
1.9091
1.9696
2.0320
2.0966
2.1635
2.2328
2.3046
2.3515
2.3515
2.3790
2.4563
2.5367
2.6202
2.7072
2.7977
2.8922
2.9908
3.0938
3.2015
3.3143
3.4326

0.8976
0.9280
0.9591
0.9909
1.0233
1.0564
1.0903
1.1250
1.1604
1.1966
1.2337
1.2717
1.3005
1.3005
1.3106
1.3504
1.3913
1.4331
1.4761
1.5201
1.5654
1.6118
1.6595
1.7086
1.7590
1.8109
1.8642
1.9192
1.9758
2.0342
2.0722
2.0722
2.0944
2.1565
2.2206
2.2869
2.3553
2.4261
2.4994
2.5753
2.6539
2.7354
2.8200
2.9079

0.5577
0.5742
0.5908
0.6076
0.6246
0.6418
0.6592
0.6768
0.6947
0.7127
0.7309
0.7494
0.7632
0.7632
0.7680
0.7869
0.8060
0.8254
0.8450
0.8648
0.8849
0.9052
0.9258
0.9466
0.9677
0.9891
1.0107
1.0327
1.0549
1.0774
1.0918
1.0918
1.1002
1.1232
1.1466
1.1703
1.1944
1.2187
1.2434
1.2684
1.2937
1.3194
1.3454
1.3718

0.2685
0.2755
0.2825
0.2896
0.2967
0.3038
0.3110
0.3182
0.3254
0.3327
0.3399
0.3472
0.3527
0.3527
0.3546
0.3620
0.3694
0.3768
0.3843
0.3918
0.3993
0.4069
0.4145
0.4221
0.4298
0.4375
0.4452
0.4530
0.4608
0.4686
0.4736
0.4736
0.4765
0.4844
0.4923
0.5003
0.5083
0.5163
0.5243
0.5324
0.5406
0.5487
0.5569
0.5651

Appendix 2. Lee and Kesler Method

473

Table A29 (cont. and end)


Adimensional residual heat capacity at constant pressure for the simple fluid

----0.8

0.8200
0.8300

3.4299
3.4254

0.8400
0.8500
0.8600
0.8700
0.8800
0.8890
0.8890
0.8900
0.9000
0.9100
0.9200
0.9300
0.9400
0.9500
0.9600
0.9700
0.9800
0.9900
1.m
1.0100
1.0200
1.0300
1.0400

0.9

0.94

0.98

1.02

1.04

1.06

1.1

1.2

1.5

4.3802
4.5766
4.7871

3.5567
3.6872

2.9992
3.0941
3.1929

2.2808

1.3986
1.4257
1.4532

0.5734
0.5817

----5.0467
5.0234

7.5417
7.4554

3.4210

5.0004

3.4165
3.4121

4.9778

7.3722
7.2918

3.4078
3.4034
3.3995
3.3995
3.3991
3.3948
3.3905
3.3862
3.3820

4.9555
4.9336
4.9120
4.8929
4.8929
4.8907
4.8698
4.8491
4.8288

3.3778
3.3736

4.8087
4.7890
4.7695

3.3694
3.3653

4.7503
4.7314

3.3612
3.3571
3.3530
3.3490

4.7127
4.6943
4.6761

3.3450
3.3410

4.6582
4.6405
4.6230
4.6058
4.5888

1.0500
1.0600

3.3370
3.3331
3.3291

1.0700
1.0800

3.3252
3.3213

4.5721
4.5555
4.5392

1.0900
1.1000
1.1100

3.3175
3.3136

4.5230
4.5071

1.1200
1.1300
1.1400
1.1500

3.3098
3.3060
3.3022
3.2985
3.2947

1.1600
1.1700
1.1800
1.1900
1.2000
1.2000

3.2910
3.2873
3.2836
3.2800
3.2763
3.2763

4.4914
4.4758
4.4605
4.4453
4.4303
4.4155
4.4009
4.3864
4.3721
4.3580
4.3580

7.2141
7.1389

8.7158
9.6051
10.6990
12.0852
13.9125

7.0661
7.0029

16.4598
10.3277
16.1838

7.0029
6.9956

14.6866
14.4265

6.9273
6.861 1
6.7968
6.7343
6.6737
6.6147
6.5574
6.5016
6.4474
6.3945
6.3430
6.2928
6.2438
6.1961
6.1495
6.1040
6.0595
6.0162
5.9737
5.9323
5.8917

5.7567
6.1038
6.4895
6.9209
7.4071
7.9598
8.5945
9.2491
9.2491

9.3318
12.2714 10.2003
10.5728 11.2404
L9.1915
L8.0413

12.5113

5.0133
5.2571
5.5208
5.8070
6.0848
6.0848
6.1186
6.4594
6.8337
7.2468

3.8246
3.9693
4.1221
4.2835
4.4545
4.6163
4.6163
4.6358
4.8285
5.0335
5.2522

14.1040 7.7050
L7.0652 16.1661 8.2164
L6.2240 18.9543 8.7906
L5.4898 22.9626 9.4400
L4.8421 29.2864 10.1804
L4.2654 40.9899 11.0319
L3.7481 71.7092 12.0210
13.2808
13.1830
L2.8560 00.0552 14.5652

5.4859
5.7363
6.0050
6.2941

L2.4678 61.7228
L2.1113 46.6040
I1.7826 38.2432
11.4782 32.8485
L1.1955 29.0398

8.6166
9.1382
9.7127

5.8521

L0.9319 26.1883
10.6855 23.9611
10.4546 22.1665
10.2376 20.6848
10.0331 19.4375

5.8133
5.7753
5.7381
5.7017
5.6661
5.6311
5.5969
5.5633
5.5304
5.5304

9.8402
9.6577
9.4848
9.3206
9.1645
9.0159
8.8742
8.7388
8.6094
8.6094

18.3708
17.4463
16.6362
15.9195
15.2802
14.7057
14.1859
13.7135
13.2817
13.2817

16.2328
18.2767
20.8249
24.0584
28.2264
33.6421
40.5863
48.9447
57.4702
63.5602
55.1655
62.6774
57.8639
52.2864
46.8690
42.0230
37.8559
34.3319
31.3642
31.3642

6.6060
6.9434
7.3093
7.7072
8.1414

L0.3474
L 1.05 10

L1.8331
L2.7047
L3.6779

3.2958
3.4031
3.5150
3.6319
3.7411
3.7411
3.7541
3.8820
4.0159
4.1562
4.3035
4.4583
4.6210
4.7924
4.9729
5.1635
5.3648
5.5778
5.8033
6.0424
6.2962
6.5660
6.8530
7.1588
7.4849
7.8330

L4.7656
L5.9807

8.2048
8.6023
9.0274

L7.3343
18.8334
l0.4769
12.2497
l4.1165
L6.0161
l7.8610
l9.5460
30.9664
30.9664

9.4822
9.9685
10.4882
11.0431
11.6342
12.2622
12.9269
13.6269
14.3591
14.3591

2.3395
2.3998
2.4618
2.5256
2.5912
2.6586
2.7207
2.7207
2.7281
2.7996
2.8732
2.9491
3.0273
3.1080
3.1912
3.2770

1.4811
1.5094
1.5381
1.5672
1.5936
1.5936
1.5967
1.6267
1.6570
1.6878
1.7190
1.7507
1.7829

3.3656
3.4571

1.8155
1.8486
1.8821

3.5516
3.6492
3.7501

1.9162
1.9508
1.9858

3.8545
3.9624
4.0740

2.0214
2.0575
2.0941

4.1895
4.3091
4.4329
4.5611

2.1313
2.1690
2.2073

4.6939
4.8315
4.9741
5.1217
5.2748
5.4333
5.5976
5.7677
5.9439
6.1264
6.3152
6.5104
6.5104

2.2461
2.2855
2.3255
2.3660
2.4072
2.4489
2.4913
2.5342
2.5778
2.6220
2.6668
2.7123
2.7584
2.7584

0.5900
0.5984
0.6068
0.6152
0.6237
0.6313
0.6313
0.6322
0.6407
0.6493
0.6578
0.6665
0.6751
0.6838
0.6926
0.7013
0.7101
0.7190
0.7278
0.7367
0.7457
0.7546
0.7636
0.7727
0.7817
0.7908
0.8000
0.8091
0.8183
0.8276
0.8368
0.8461
0.8554
0.8648
0.8742
0.8836
0.8931
0.9026
0.9121
0.9121

-----------

474

Appendix 2. Lee and Kesler Method

Table A2.10
Adimensional residual heat capacity at constant pressure: corrective term

-- -

0.0100
0.0100
0.0200
0.0300
0.0400
0.0500
0.0600
0.0700
0.0800
0.0900
0.1000
0.1100
0.1200
0.1300
0.1400
0.1500
0.1600
0.1700
0.1800
0.1900
0.2000
0.2100
0.2200
0.2300
0.2400
0.2500
0.2563
0.2563
0.2600
0.2700
0.2800
0.2900
0.3000
0.3100
0.3200
0.3300
0.3400
0.3500
0.3600
0.3700
0.3800
0.3900
0.4000
0.4100

--

0.8

0.9

0.94

0.98

1.02

1.04

1.06

1.1

1.2

1.5

0.0972
0.0972
0.1986
0.3046
0.4154
0.5316
0.6536
0.7820
0.9174
1.0605
1.2121
1.3732
1.5449
1.7283
1.9251
2.1369
2.3658
2.6143
2.8854
3.1828
3.5109
3.8753
4.2832
4.7438
5.2688
5.8744
6.3046
10.4357
10.4371
10.4409
10.4446
10.4482
10.4517
10.4552
10.4586
10.4620
10.4653
10.4685
10.4717
10.4748
10.4779
10.4808
10.4838
10.4867

0.0581
0.0581
0.1177
0.1789
0.2417
0.3063
0.3726
0.4409
0.5112
0.5836
0.6582
0.7352
0.8148
0.8969
0.9820
1.0700
1.1612
1.2558
1.3541
1.4563
1.5626
1.6734
1.7891
1.9099
2.0364
2.1690
2.2554
2.2554
2.3081
2.4544
2.6086
2.7713
2.9434
3.1259
3.3199
3.5265
3.7473
3.9839
4.2383
4.5127
4.8100
5.1332
5.4865
5.8744

0.0481
0.0481
0.0971
0.1472
0.1984
0.2508
0.3043
0.3591
0.4151
0.4725
0.5313
0.5915
0.6533
0.7167
0.7817
0.8485
0.9172
0.9878
1.0604
1.1352
1.2123
1.2917
1.3737
1.4583
1.5457
1.6361
1.6944
1.6944
1.7297
1.8266
1.9271
2.0315
2.1398
2.2526
2.3700
2.4923
2.6201
2.7536
2.8933
3.0398
3.1936
3.3553
3.5256
3.7054

D.0400
D.0400
0.0808
0.1222
0.1644
D.2073
0.2510
0.2955
0.3409
0.3871
0.4342
D.4823
D.5313
0.5813
0.6323
0.6844
0.7376
0.7919
0.8475
0.9043
0.9624
1.0218
1.0826
1.1449
1.2087
1.2741
1.3160
1.3160
1.3412
1.4100
1.4806
1.5531
1.6275
1.7041
1.7829
1.8639
1.9474
2.0334
2.1220
2.2135
2.3079
2.4055
2.5064
2.6108

0.0366
0.0366
0.0739
0.1117
0.1501
0.1891
0.2288
0.2691
0.3101
0.3518
0.3942
0.4374
0.4813
0.5260
0.5715
0.6179
0.6651
0.7133
0.7623
0.8123
0.8634
0.9154
0.9685
1.0227
1.0780
1.1346
1.1707
1.1707
1.1923
1.2514
1.3117
1.3735
1.4366
1.5013
1.5675
1.6353
1.7048
1.7761
1.8492
1.9242
2.0012
2.0802
2.1615
2.2451

0.0336
0.0336
0.0677
0.1022
0.1373
0.1728
0.2089
0.2456
0.2827
0.3205
0.3588
0.3977
0.4372
0.4774
0.5182
0.5596
0.6018
0.6446
0.6882
0.7325
0.7776
0.8235
0.8702
0.9177
0.9661
1.0154
1.0468
1.0468
1.0656
1.1167
1.1689
1.2220
1.2762
1.3315
1.3879
1.4455
1.5042
1.5642
1.6255
1.6881
1.7521
1.8175
1.8844
1.9529

0.0309
0.0309
0.0621
0.0938
0.1258
0.1583
0.1912
0.2245
0.2583
0.2926
0.3273
0.3625
0.3982
0.4343
0.4711
0.5083
0.5461
0.5844
0.6233
0.6628
0.7028
0.7435
0.7848
0.8268
0.8694
0.9127
0.9402
0.9402
0.9567
1.0014
1.0468
1.0930
1.1400
1.1878
1.2364
1.2858
1.3361
1.3873
1.4394
1.4924
1.5464
1.6014
1.6575
1.7146

3.0284
3.0284
3.0571
0.0861
D.1155
0.1452
0.1753
0.2057
D.2365
0.2677
D.2992
D.3311
0.3635
0.3962
0.4293
0.4629
0.4969
0.5313
0.5662
0.6015
0.6373
0.6736
0.7103
0.7476
0.7853
0.8236
0.8479
0.8479
0.8624
0.9018
0.9417
0.9822
1.0232
1.0649
1.1071
1.1500
1.1935
1.2376
1.2825
1.3279
1.3741
1.4210
1.4686
1.5170

0.0241
0.0241
0.0485
0.0730
0.0978
0.1228
0.1481
0.1736
0.1993
0.2253
0.2516
0.2781
0.3048
0.3318
0.3591
0.3866
0.4144
0.4425
0.4709
0.4995
0.5285
0.5577
0.5872
0.6170
0.6472
0.6776
0.6968
0.6968
0.7083
0.7394
0.7708
0.8025
0.8345
0.8669
0.8996
0.9326
0.9660
0.9997
1.0338
1.0682
1.1030
1.1382
1.1738
1.2097

0.0164
0.0164
0.0329
0.0495
0.0662
0.0830
0.0999
0.1168
0.1338
0.1509
0.1681
0.1854
0.2027
0.2202
0.2377
0.2553
0.2730
0.2908
0.3087
0.3266
0.3446
0.3627
0.3809
0.3992
0.4176
0.4360
0.4476
0.4476
0.4545
0.4731
0.4918
0.5105
0.5294
0.5483
0.5673
0.5863
0.6054
0.6246
0.6439
0.6633
0.6827
0.7022
0.7217
0.7413

0.0061
0.0061
0.0122
0.0183
0.0244
0.0304
0.0365
0.0426
0.0486
0.0546
0.0607
0.0667
0.0727
0.0787
0.0847
0.0906
0.0966
0.1026
0.1085
0.1144
0.1203
0.1262
0.1321
0.1380
0.1438
0.1497
0.1533
0.1533
0.1555
0.1613
0.1671
0.1729
0.1787
0.1844
0.1901
0.1958
0.2015
0.2072
0.2129
0.2185
0.2241
0.2297
0.2353
0.2408

Appendix 2. Lee and Kesler Method

475

Table A2.10 (cont.)


Adimensional residual heat capacity at constant pressure: corrective term

0.8

0.4200
0.4300
0.4400
0.4500
0.4600
0.4700
0.4800
0.4900
0.5OOO
0.5100
0.5200
0.5300
0.5374
0.5374
0.5400
0.5500
0.5600
0.5700
0.5800
0.5900
0.6oOo
0.6100
0.6200
0.6300
0.6400
0.6500
0.6600
0.6700
0.6800
0.6900
0.6963
0.6963
0.7oM)
0.7100
0.7200
0.7300
0.7400
0.7500
0.7600
0.7700
0.7800
0.7900
0.8000
0.8100

10.4895
10.4922
10.4949
10.4976
10.5002
10.5028
10.5053
10.5077
10.5101
10.5124
10.5147
10.5170
10.5186
10.5186
10.5192
10.5214
10.5235
10.5255
10.5276
10.5295
10.5315
10.5334
10.5352
10.5370
10.5388
10.5405
10.5422
10.5439
10.5455
10.5470
10.5480
10.5480
10.5486
10.5501
10.55 15
10.5529
10.5543
10.5557
10.5570
10.5.583
10.5595
10.5607
10.5619
10.5631

0.9

0.94

0.98

1.02

1.04

1.06

1.1

1.2

6.3028
6.7789
7.3119
7.9131
8.5977
9.3852
10.3024
11.3859
!2.6881
.4.2860
16.2991
18.9223
I46
9.5985
9.6078
9.6429
9.6767
9.7091
9.7402
9.7701
9.7987
9.8263
9.8528
9.8782
9.9026
9.9261
9.9486
9.9702
9.9911
10.0110
10.0233
10.0233
10.0302
L0.0487
L0.0664
L0.0834
10.0998
L0.1155
L0.1306
10.1450
L0.1590
L0.1723
L0.1851
L0.1974

3.8955
4.0970
4.3110
4.5389
4.7821
5.0426
5.3222
5.6236
5.9495
6.3032
6.6890
7.1116
7.4529
7.4529
7.5770
8.0926
8.6675
9.3133
10.0448
10.8813
11.8485
12.9812
14.3280
15.9588
17.9782
20.5501
23.9475
28.6635
35.6971
47.4659
60.3894
9.0749
9.1351
9.2904
9.4337
9.5660
9.6882
9.8013
9.9060
10.0030
10.0929
10.1764
10.2539
10.3258

2.7189
2.8310
2.9473
3.0680
3.1935
3.3241
3.4602
3.6020
3.7502
3.9050
4.0671
4.2371
4.3687
4.3687
4.4155
4.6031
4.8006
5.0091
5.2295
5.4629
5.7107
5.9744
6.2556
6.5564
6.8790
7.2261
7.6008
8.0070
8.4488
8.9318
9.2622
9.2622
9.4622
10.0481
10.6991
11.4275
12.2489
13.1835
14.2577
15.5072
16.9811
18.749C
20.9128
23.6288

2.3310
2.4195
2.5107
2.6046
2.7014
2.8013
2.9044
3.0110
3.1212
3.2352
3.3533
3.4756
3.5694
3.5694
3.6025
3.7342
3.8711
4.0134
4.1616
4.3160
4.4771
4.6454
4.8214
5.0057
5.1989
5.4018
5.6151
5.8398
6.0769
6.3275
6.4939
6.4939
6.5928
6.8744
7.1738
7.4929
7.8339
8.1993
8.5919
9.0151
9.4729
9.9699
10.5117
L1.1050

2.0229
2.0946
2.1681
2.2433
2.3204
2.3994
2.4805
2.5637
2.6490
2.7367
2.8268
2.9194
2.9899
2.9899
3.0146
3.1126
3.2134
3.3173
3.4242
3.5345
3.6483
3.7657
3.8870
4.0123
4.1418
4.2759
4.4146
4.5584
4.7074
4.8619
4.9630
4.9630
5.0224
5.1892
5.3625
5.5429
5.7307
5.9265
6.1307
6.3440
6.5668
6.7998
7.0437
7.2992

1.7728
1.8321
1.8925
1.9542
2.0171
2.0813
2.1468
2.2136
2.2818
2.3515
2.4227
2.4954
2.5504
2.5504
2.5697
2.6456
2.7232
2.8026
2.8838
2.9669
3.0519
3.1389
3.2280
3.3192
3.4127
3.5084
3.6065
3.7071
3.8102
3.9159
3.9844
3.9844
4.0244
4.1356
4.2497
4.3668
4.4870
4.6104
4.7370
4.8670
5.oO04
5.1373
5.2778
5.4220

1.5661
1.6160
1.6666
1.7181
1.7705
1.8236
1.8777
1.9326
1.9884
2.0452
2.1030
2.1617
2.2059
2.2059
2.2214
2.2821
2.3438
2.4067
2.4706
2.5356
2.6017
2.6691
2.7376
2.8073
2.8782
2.9504
3.0238
3.0986
3.1746
3.2520
3.3018
3.3018
3.3307
3.4109
3.4924
3.5752
3.6595
3.7452
3.8323
3.9208
4.0107
4.1020
4.1946
4.2886

1.2460
1.2827
1.3197
1.3572
1.3950
1.4333
1.4719
1.5110
1.5504
1.5903
1.6306
1.6713
1.7018
1.7018
1.7124
1.7539
1.7958
1.8382
1.8810
1.9242
1.9678
2.0118
2.0563
2.1011
2.1464
2.1921
2.2381
2.2846
2.3315
2.3787
2.4088
2.4088
2.4263
2.4743
2.5226
2.5712
2.6201
2.6694
2.7189
2.7686
2.8186
2.8688
2.9192
2.9697

--- -----

1.5

0.7610
0.7807
0.8005
0.8204
0.8403
0.8603
0.8803
0.9003
0.9204
0.9406
0.9608
0.9810
0.9960
0.9960
1.0013
1.0216
1.0419
1.0622
1.0826
1.1030
1.1234
1.1438
1.1642
1.1846
1.2051
1.2255
1.2459
1.2662
1.2866
1.3069
1.3198
1.3198
1.3272
1.3475
1.3677
1.3878
1.4079
1.4279
1.4479
1.4677
1.4875
1.5072
1.5268
1.5462

0.2464
0.2519
0.2574
0.2628
0.2683
0.2737
0.2791
0.2845
0.2898
0.2951
0.3004
0.3057
0.3096
0.3096
0.3109
0.3162
0.3214
0.3265
0.3317
0.3368
0.3418
0.3469
0.3519
0.3569
0.3619
0.3668
0.3717
0.3766
0.3814
0.3862
0.3892
0.3892
0.3910
0.3957
0.4004
0.4051
0.4097
0.4143
0.4189
0.4234
0.4279
0.4324
0.4368
0.4412

476

Appendix 2. Lee and Kesler Method

Table A2.10 (cont. and end)


Adimensional residual heat capacity at constant pressure: corrective term

x
0.8200
0.8300
0.8400
0.8500
0.8600
0.8700
0.8800
0.8890
0.8890
0.8900
0.9000
0.9100
0.9200
0.9300
0.9400
0.9500
0.9600
0.9700
0.9800
0.9900
1.m
1.0100
1.0200
1.0300
LO400
1.0500
1.0600
1.0700
1.0800
1.0900
1.1000
1.1100
1.1200
1.1300
1.1400
1.1500
1.1600
1.1700
1.1800
1.1900

1.2000
1.2000

---0.8

0.9

0.94

0.98

10.5642
10.5653
10.5663
10.5673
10.5683
10.5693
10.5702
10.5710
10.5710
10.5711
10.5720
10.5728
10.5737
10.5744
10.5752
10.5759
10.5767
10.5773
10.5780
10.5786
10.5792
10.5798
10.5804
10.5809
10.5814
10.5819
10.5823
10.5828
10.5832
10.5836
10.5839
10.5843
10.5846
10.5849
10.5852
10.5854
10.5857
10.5859
10.5861
10.5863
10.5864
10.5864

10.2092
10.2205
10.2313
10.2417
10.2517
10.2612
10.2704
10.2782
10.2782
10.2791
10.2875
10.2955
10.3032
10.3105
10.3175
10.3241
10.3305
10.3366
10.3423
10.3478
10.3531
10.3580
10.3627
10.3672
10.3714
10.3754
10.3792
10.3827
10.3861
10.3892
10.3922
10.3949
10.3975
10.3999
10.4021
10.4042

10.3927
10.4549
10.5127
10.5663
10.6162
10.6627
10.7058
10.7418
10.7418
10.7459
10.7832
10.8178
10.8499
10.8798
10.9074
10.9330
10.9568
10.9787
10.9990
11.0177
11.0349
11.0507
11.0652
11.0785
11.0906
11.1016
11.1115
11.1205
11.1286
11.1358
11.1422
11.1478
11.1527
11.1568
11.1604
11.1633
11.1656
11.1674
11.1686
11.1693
11.1696
11.1696

27.1484
31.9049
38.7150
49.3291
68.3307
113.2381
154.8589

11.7579
12.4805
13.2851
14.1873
15.2070
16.3697
17.7096
19.0968
19.0968
19.2724
21.1217
23.3481
26.0855
29.5417
34.0567
40.2304
49.2340
63.7135
91.2664
166.947t

10.4061

10.4078
10.4094
10.4108
10.4121
10.4121

2.8896
3.291 1
6.4393
8.6710
10.2995
11.5136
12.4332
13.1377
13.6819
14.1041
14.4329
14.6874
14.8839
15.0342
15.1473
15.2302
15.2885
15.3267
15.3483
15.3563
15.3531
15.3401
15.3197
15.2929
15.2607
15.2241
15.1838
15.1405
15.0948
15.0470
14.9976
14.9469
14.8953
14.8953

I71.543t
104.4691
79.1339
65.4147
56.6792
50.5637
46.0085
42.4670
39.6228
37.2808
35.3134
33.6319
32.1791
30.9071
29.7827
28.7803
27.8802
27.0673
26.3275
25.6517
25.6517

1.02

1.04

7.5672 5.5698
7.8484 5.7215
8.1437 5.8769
8.4542 6.0360
8.7806 6.1989
9.1242 6.3654
9.4859 6.5354
9.8259 6.6904
9.8259 6.6904
9.8666 6.7087
10.2674 6.8850
10.6889 7.0640
11.1317 7.2450
11.5956 7.4276
12.0801 7.6107
12.5831 7.7934
13.1009 7.9743
13.6271 8.1515
14.1506 8.3230
14.6538 8.4859
15.1084 8.6370
15.4697 8.7718
15.6665 8.8852
15.5865 8.9706
15.0526 9.0198
13.7896 9.0235
11.3905 8.9693
7.3271 8.8436
1.1753 8.6293
-6.5234 8.3078
-12.5215 7.8588
-10.520f 7.2624
3.7588 6.5020
27.6759 5.5706
54.6618 4.4799
79.0211 3.2730
97.7070 2.0388
109.947~ 0.9237
116.203~ 0.1306
117.4211 -0.1048
114.7512 0.4372
114.751: 0.4372

1.06

1.1

1.2

1.5

4.3838
4.4803
4.5779
4.6766
4.7763
4.8767
4.9780
5.0691
5.0691
5.0797
5.1818
5.2840
5.3859
5.4874
5.5879
5.6870
5.7843
5.8791
5.9707
6.0584
6.1412
6.2180
6.2877
6.3490
6.4002
6.4396
6.4652
6.4750
6.4664
6.4368
6.3833
6.3029
6.1925
6.0491
5.8697
5.6518
5.3940
5.0958
4.7590
4.3872
3.9882
3.9882

3.0203
3.0710
3.1216
3.1723
3.2228
3.2732
3.3234
3.3681
3.3681
3.3733
3.4229
3.4720
3.5206
3.5686
3.6159
3.6623
3.7077
3.7521
3.7952
3.8370
3.8772
3.9157
3.9522
3.9867
4.0188
4.0485
4.0753
4.0991
4.1196
4.1365
4.1496
4.1585
4.1630
4.1627
4.1574
4.1467
4.1303
4.1079
4.0794
4.0443
4.0027
4.0027

1.5656
1.5848
1.6038
1.6228
1.6415
1.6601
1.6785
1.6948
1.6948
1.6967
1.7147
1.7325
1.7501
1.7674
1.7845
1.8013
1.8178
1.8341
1.8500
1.8657
1.8810
1.8959
1.9105
1.9248
1.9386
1.9520
1.9650
1.9776
1.9898
2.0014
2.0126
2.0233
2.0335
2.0432
2.0523
2.0609
2.0689
2.0763
2.0832
2.0894
2.0950
2.0950

0.4455
0.4498
0.4541
0.4583
0.4625
0.4667
0.4708
0.4744
0.4744
0.4748
0.4789
0.4829
0.4868
0.4907
0.4946
0.4984
0.5022
0.5060
0.5097
0.5133
0.5169
0.5205
0.5240
0.5275
0.5309
0.5343
0.5377
0.5410
0.5442
0.5474

- -

0.5506
0.5537
0.5568
0.5598
0.5628
0.5657
0.5686
0.5715
0.5743
0.5770
0.5797
0.5797

Appendix

Surface, Volume, and


Interaction Parameters Applied
in the UNIFAC Method

The UNIFAC method [Fredenslund, 19751 is introduced in Chapter 7, Section 7.7.2.


The following tables A3.1 and A3.2 provide the values for the parameters of volume R,,
surface Q , for some groups, as well as the interaction parameters a,,, between these
groups, for application of the UNIFAC method.

Table A3.1
Volume ( R k )and surface ( Q k ) parameters applied in the UNIFAC method
(Eq. 7.45)
-

lCH,

2C=C

CH,
CH,
CH
C
CH,=CH
CH =CH
CH, =C
CH =C

c=c
3ACH

ACH
AC

No
1
2
3
4
5
6
7
8
9
10
11

Rk

Qk

Example

Decomposition

0.901 1
0.6744
0.4469
0.2195

0.848
0.540
0.228
0

Hexane

2 CH, and 4 CH,

2-Methyl-propane
Neopentane

3 CH, and 1CH


4CH,andlC

1.3454
1.1167
1.1173
0.8886
0.6605

1.176
0.867
0.988
0.676
0.485

Hexene-1
Hexene-2
2-Methyl-1-butene
2-Methyl-2-butene
2,3-Dimethylbutene

1CH,, 3 CH,, 1CH, =CH


2CH,, 2 CH,, 1CH =CH
2CH,, 1CH,, 1CH, =C
3 CH,, 1CH =C
4CH3,1C=C

0.5313
0.3652

0.400
0.120

Naphtalene
Styrene

8 ACH, 2 AC
1CH, =CH, 5 ACH, 1AC

478

Appendix 3. Surface, Volume, and Interaction ParametersApplied in the UNIFAC Method

Table A3.1 (cont.)


Volume ( R k )and surface ( Q k ) parameters applied in the UNIFAC method
(Eq. 7.45)
Decomposition

Example
4ACCH,
I

SOH

ACCH,
ACCH,
ACCH

1 OH

12 1.2663
13 1.0396
14 0.8121
I

I 15 1

1.0000

0.968
0.660
0.348

Toluene
Ethylbenzene
Cumene

5 ACH, 1ACCH,
lCH,, SACH, lACCH,
2 CH,, 5AC, 1ACCH

1.200

2-Propanol

2 CH,, 1CH, 1OH

6CH,OH

CH,OH

16

1.4311

1.432

Methanol

1CH,OH

7H,O

H,O

17

0.9200

1.400

Water

1H,O

Phenol

5 ACH, 1ACOH

Butanone
3-Pentanone

1CH,, 1CH,, 1CH,CO


2 CH,, 1CH,, 1CH,CO

10CHO

CHO

21

0.9980

0.948

Propionic aldehyde

1 1CH,, 1CH,, 1CHO

Table A3.2
Interaction parameters uk,,applied in the UNIFAC method (Eq. 7.49)

I
1CH,
2 c=c
3 ACH
4 ACCH,
5 OH
6 CH,OH
7 H,O
8 ACOH
9 CH,CO
10CHO

1CH,
2 c=c
3 ACH
4 ACCH,
5 OH
6 CH,OH
7 H,O
8ACOH
9 CH,CO
10CHO

lCH,

2C=C

3 ACH

0
-35.36
-11.12
-69.7
156.4
16.51
300
275.8
26.76
505.7

86.02
0
3.446
-113.6
457
-12.52
496.1
217.5
42.92
56.3

61.13
38.81
0
-146.8
89.6
-50
362.3
25.34
140.1
23.39

76.50
74.15
167
0
25.82
-44.5
377.6
244.2
365.8
106

I 6CH,OH I

7H,O

8ACOH

9CH,CO

1333
526.1
1329
884.9
-259.7
-101.7
324.5
0
-133.1
-155.6

476.4
182.6
25.77
-52.1
84
23.39
-195.4
-356.1
0
128

697.2
787.6
637.4
603.3
-137.1
0
289.6
-265.2
108.7
-340.2

1318
270.6
903.8
5 695
353.5
-181
0
401.8
472.5
480.8

986.5
524.1
636.1
803.2
0
249.1
- 229.1
- 451.6
164.5
529

10CHO
677
448.8
347.3
586.8
-203.6
306.4
-116
-27 1
-37.36
0

Appendix

Properties of
the Ethane (1) Propane (2) System
at 45C and at 2.5 MPa
as a Function of Composition

This appendix lists the values for molar volumes of the ethane (1) propane (2) mixture at
45C and at 2.5 MPa as a function of composition, as well as the fugacities of these two
components in the mixture. This table is similar to Table 8.3 (Chapter 8, Example 8.3), but
uses a more detailed compositional scale (by steps of 2%).
It is used in Examples 8.3,8.4, and 8.5:
Column 1

mole fraction of ethane


mole fraction of propane
molar volumes (cm3.m o P )
fugacities of ethane (Pa)
fugacities of propane (Pa)

Columns 2,3,4
Columns 5,6,7

Table Ad1
Ethane (1)propane (2) mixture
Application of the Soave-Redlich-Kwong equation of state: molar volumes and fugacities

1
Volume

Composition

I1 L
Undefined

1
2
1
2
1
2
1
2
1
2

o.oO0
1.oO0
0.020
0.980
0.040
0.960
0.060
0.940
0.080
0.920

102.81
102.79
102.79

102.83

Liquid

0.000000E+00
0.1 18672E+07
0.711426E+05
0.116301E+07
0.141964E+06
0.113935E+07
0.212459E+06
0.111575E+07
0.282619E+06
0.109220E+07

Fugacites
Vapor

Undefined

480

Appendix 4. Properties of the Ethane ( 1 ) Propane (2) System

Table A4.1 (cont.)


Ethane (1)propane (2) mixture
Application of the Soave-Redlich-Kwong equation of state: molar volumes and fugacities
P = 2.5 MPa

:omposition
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2

0.100
0.900
0.120
0.880
0.140
0.860
0.160
0.840
0.180
0.820
0.200
0.800
0.220
0.780
0.240
0.760
0.260
0.740
0.280
0.720
0.300
0.700
0.320
0.680
0.340
0.660
0.360
0.640
0.380
0.620
0.400
0.600
0.420
0.580
0.440
0.560
0.460
0.540
0.480
0.520
0.500
0.500
0.520
0.480
0.540
0.460

T = 313.15 K

Volume
Liquid

Vapor

102.87
102.93
103.00
103.10
103.22
103.36

519.96

103.52

550.75

103.71

572.64

103.94

590.55

104.19

606.05

104.49

619.89

104.82

632.48

105.20

644.10

105.63

654.93

106.12

665.10

106.67

674.71

107.30

683.84

108.02

692.53

108.85

700.85

109.80

708.83

110.91

716.50

112.22

723.90

113.78

731.04

Fuaacites
Undefined

Liquid

Vapor

0.352440E+06
0.106871E+07
0.421912E+06
0.104527E+07
0.491028E+06
0.102189E+07
0.559780E+06
0.998563E+06
0.628159E+06
0.975294E+06
0.696154E+06
0.952083E+06
0.763757E+06
0.928931E+06
0.830954E+06
0.905839E+06
0.897734E+06
0.882809E+06
0.964083E+06
0.859841E+06
0.102998E+07
0.836938Et06
0.109542Et07
0.814101E+06
0.116037Et07
0.791334E+06
0.122482E+07
0.768640Et06
0.128873E+07
0.746022E+06
0.135208E+07
0.723487E+06
0.141482E+07
0.701040E+06
0.147692E+07
0.678690Et06
0.153831E+07
0.656450E+06
0.159892Et07
0.634337Et06
0.165866E+07
0.612373E+06
0.171739E+07
0.590595E+06
0.177491E+07
0.569057E+06

0.463363E+06
0.131774E+07
0.502020E+06
0.128997E+07
0.542174E+06
0.126067E+07
0.582802E+06
0.123067E+07
0.623625E+06
0.120025E+07
0.664533E+06
0.1 16951E+07
0.705473E+06
0.113852E+07
0.746422E+06
0.110733E+07
0.787365E+06
0.107594E+07
0.828298E+06
0.104439E+07
0.869219E+06
0.101269E+07
0.910129E+06
0.980838E+06
0.951030Et06
0.948847Et06
0.991925E+06
0.916719E+06
0.103282E+07
0.884460E+06
0.107371Et07
0.852070E+06
0.111461E+07
0.819551E+06
0.1 15552E+07
0.786905E+06

Undefined

481

Appendix 4. Properties of the Ethane ( I ) Propane (2) System

Table A4.1 (cont. and end)


Ethane (1) propane (2) mixture
Application of the Soave-Redlich-Kwong equation of state: molar volumes and fugacities

P = 2.5 MPa:
:ompition
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2

0.560
0.440
0.580
0.420
0.600
0.400
0.620
0.380
0.640
0.360
0.660
0.340
0.680
0.320
0.700
0.300
0.720
0.280
0.740
0.260
0.760
0.240
0.780
0.220
0.800
0.200
0.820
0.180
0.840
0.160
0.860
0.140
0.880
0.120
0.900
0.100
0.920
0.080
0.940
0.060
0.960
0.040
0.980
0.020
1.OOo
o.oO0

T = 313.15 K

Volume
Liquid

Vapor

115.70

737.94

118.15

744.62

121.47

751.11

126.69

757.40
763.52
769.47
775.26
780.91
786.42
791.80
797.04
802.17
807.19
812.09
816.89
821.59

Undefined

Liquid

Undefined

Appendix

Detailed Analysis of
a Straight-Run Gasoline Cut

Using gas chromatography, this appendix takes a detailed look at a gasoline straight-run
cut, the summary of which may be found in Table 12.1. Since the gasoline has undergone
no chemical transformation, it contains no ethylenic hydrocarbons, and the aromatic compounds correspond to those present in the crude oil from which it is extracted by simple
distillation.
Sulfur-containing compounds are not listed.

Table A5.1
Composition expressed in mass %
of a straight-run gasoline.
Gas chromatograph analysis
(the detailed analysis separates 260 components;
below, some have been lumped)

n-Alcanes
Component
Propane
n-Butane
n-Pentane
n-Hexane
n-Heptane
n-Octane
n-Octane
n-Nonane
n -D ecane
n-Undecane
n-Dodecane

W %

0.18
1.93
4.07
5.34
5.52
5.1
5.1
4.53
4
2.77
0.84

484

Appendix 5. Detailed Analysis of a Straight-Run Gasoline Cut

Isoalkanes
Component
Isobutane
Isopentane
2,2-Dimethylbutane
2,3-Dimethylbutane
2-Methylpentane
3-Methylpentane
2,2-Dimethylpentane
2,4-Dimethylpentane
3,3-Dimethylpentane
2-Methylhexane
2,3-Dimethylpentane
3-Methylhexane
3-Ethylpentane
2,2-Dimethylhexane
2,5-Dimethylhexane
2,4-Dimethylhexane
3,3-Dimethylhexane
2,3,4-Trimethylpentane
2,3-Dimethylhexane
2-Methyl-3-ethylpentane
2-Methylheptane
4-Methylheptane
3,4-Dimethylhexane
3-Methylheptane
3-Eth ylhexane
2,3,5-Trimethylhexane
2,2-Dimethylheptane
2,4-Dimethylheptane
4,CDimethylheptane
2,6-Dimethylheptane
2,5-Dimethylheptane
3,5-Dimethylheptane
3,3,4-Trimethylhexane
2,3-Dimethylheptane

Component

W %

0.27
2.03
0.04
0.28
2.21
1.62
0.03
0.14
0.03
1.51
0.52
1.51
0.16
0.10
0.21
0.32
0.05
0.08
0.32
0.24
1.86
0.63
0.06
1.42
0.56
0.05
0.04
0.28
0.02
0.55
0.45
0.11
0.21
0.57

4-Ethylpentane
4-Methyloctane
2-Methyloctane
3-Ethylpentane
3-Methyloctane
2,2-Dimethyloctane
4,4-Dimethyloctane
3,5-Dimethyloctane
2,7-Dimethyloctane
2,6-Dimethyloctane
3,3-Dimethyloctane
3,6-Dimethyloctane
Methyl-ethylheptane
4-Ethyloctane
5-Methylnonane
4-Methylnonane
2-Methylnonane
3-Ethyloctane
3-Methylnonane
ClO-Isoparaffin
5-Methyldecane
4-Methyldecane
2-Methyldecane
3-Methyldecane
C, ,-Isoparaffin
2,4-Dimethyldecane
2,6-Dimethyldecane
2,5-Dimethyldecane
5-Methylundecane
4-Methylundecane
2-Methylundecane
3-Methylundecane
C,,-Isoparaffin
2,6-Dimethylundecane

W%

0.12
0.77
0.83
0.20
1.14
0.20
0.03
0.56
0.12
0.82
0.10
0.19
0.59
0.13
0.27
0.81
0.71
0.05
0.78
0.34
0.42
0.51
0.50
0.57
0.26
0.15
0.24
0.12
0.16
0.25
0.24
0.26
0.13
0.09

~~

Cycloalkanes and Alkylcycloalkanes


.~
Component
Cyclopentane
1-Methylcyclopentane
Cyclohexane
Dimethylcyclopentane
Methylcyclohexane
Ethylcyclopentane
Trimethylcyclopentane
Dimethylcyclohexane
Methylethylcyclopentane
Ethy lcyclohexane
Tetramethy lcyclopentane
Trimethylcyclohexane

w %
0.30
1.01
0.78
1.06
1.54
0.31
0.39
0.59
0.62
0.81
0.06
0.53

Component

Dimethylethylcyclopentane
MethylethyIcyclohexane
n-Butylcyclopentane
Me thylpropylcyclohexane
C,- Alkylcycloalkanes
Methylpropylcyclohexane
n-Butylcyclohexane
Cis and trans decalines
Clo-Alkylcycloalkanes
C, ,-Alkylcycloalkanes
n-Hexylcyclohexane
C,2-Alkylcycloalkanes

W %

0.31
0.88
0.26
0.68
1.17
0.68
0.37
0.15
1.09
2.02
0.01
0.64

485

Appendix 5. Detailed Analysis of a Straight-Run Gasoline Cut

Aromatics and Alkyls Aromatics


Component
Benzene
Toluene
Ethylbenzene
Metaxylene
Paraxylene
Orthoxylene
Isopropylbenzene
n-Propylbenzene
1-Methyl-3-ethylbenzene
1-Methyl-4-ethylbenzene
1,3,5-Trimethylbenzene
1-Methyl-2-ethylbenzene
1,2,4-Trimethylbenzene
1,2,3-Trimethylbenzene

1-Methyl-4-isopropylbenzene
1-Methyl-3-isopropylbenzene
Indane
1-Methyl-2-isopropylbenzene
1-3-Diethylbenzene

W %

0.5
1.82
0.98
1.14
0.48
0.99
0.15
0.37
0.86
0.56
0.31
0.62
1.12
0.45
0.20
0.18
0.24
0.25
0.22

Component

1-Methyl-3-n-propylbenzene
1-Methyl-4-n-propylbenzene
1,3-Dimethyl-5-ethylbenzene
1-Methyl-2-n-propylbenzene
1,4-Dimethyl-2-ethylbenzene
1,3-Dimethyl-4-ethylbenzene
1-Methylindane

1,2-Dimethyl-4-ethylbenzene
1,3-DimethyI-2-ethylbenzene
1,2-DimethyI-3-ethylbenzene
1,2,3,5-Tetramethylbenzene
1,2,4,5-Tetramethylbenzene
5-Methylindane
4-Methylindane
1,2,3,4-Tetramethylbenzene
Naphtalene
C,,-Alkyls aromatiques
C, ,-Alkyls aromatiques
Cl2-Alkylsaromatiques

w%
0.32
0.56
0.28
0.57
0.30
0.39
0.09
0.25
0.10
0.41
0.20
0.26
0.21
0.38
0.30
0.06
0.66
0.81
0.01

Appendix

Units

We have almost systematically used the international system of units, as well as some other
commonly used units that are directly derived from them. However, sometimes thermodynamic properties are expressed in Anglo-Saxon units. Below we list some of the more common units with their conversion factors.
International System

Property
Mass
Material quantity
Temperature
Pressure

Unit
kilogram
mole
kelvin
pascal
kilopascal
megapascal
cubic meters
joule

Volume
Energy

Symbol
kg
mol
K
Pa
kPa = lo3 Pa
MPa = lo6Pa
m3
J

Other Commonly Used Units and Anglo-Saxon Units


Property

Unit (Symbol)

Mass

pound (lb)

Temperature

degrees centigrade ("C)


or degrees Celsius
degrees Fahrenheit (OF)
Rankine (R)

To Convert to:

Calculation

kilogram
kelvin

m(kg) = 0.4536.m(lb)

kelvin
kelvin

T(K) = @(OF) + 459.67/1.8


T(K) = T(R)/1.8

bar
atmosphere
millimeters of mercury
(mmHg)
pounds per square inch (psi)

pascal
pascal

P(Pa) = lo5P (bar)


P(Pa) = 101325.P(atm)

pascal
pascal

P(Pa) = 133.3.P (mmHg)


P(Pa) = 6895.P (psi)

Volume

cubic centimeter (cm3)


cubic foot (ft3)

cubic meter
cubic meter

V(m3) = 10-6.V (cm3)


V(m3) = 0.0283. V (ft3)

Energy

thermochemical calorie (cal)


BritishThermal Unit (Btu)

joule
joule

E(J) = 4.184.E (cal)


E(J) = 1055.E (Btu)

Pressure

T(K) = q 0 c ) + 273.15

INDEX

Index Terms

Links

A
Acentric factor

68

correlations using corresponding states


and acentric factor
numerical data

71
435

relationship with the critical


compressibility factor
Activity

69
154

155

dependence on temperature, pressure, and


composition

155

in an ideal solution

156

Activity coefficients

160

calculation by
ASOG method

250

Flory theory

223

377

NRTL model

239

246

regular solution theory

231

UNIFAC method

251

UNIQUAC model

244

246

Wilson equation

238

246

dependence on temperature, pressure, and


composition

161

effect on liquid-vapor equilibrium

214

in ionic solutions

263

in polymer solutions

223

377

This page has been reformatted by Knovel to provide easier navigation.

382

Index Terms

Links

Activity coefficients (Cont.)


in the chemical equilibrium expression

427

in the solid phase

362

Antoine Equation

41

ASOG method

250

Associated solutions

258

Azeotrope

176

B
Benedict, Webb, and Rubin equation of state

75

Boiling temperature (numerical data)

435

Bubble point

169

calculation

193

dependence on composition

184

dependence on temperature

186

111

C
Carnahan and Starling

134

Characteristic functions

22

Chemical equilibria

425

calculation

427

simultaneous

431

the equilibrium condition

425

Chemical potential

149

dependence on composition

151

dependence on pressure and temperature

151

Chemical reactions

413

Clausius-Clapeyron

38

Coherence test

431

187

160

218

This page has been reformatted by Knovel to provide easier navigation.

Index Terms

Links

Combinatorial term

223

Compressibility factor
calculation using the corresponding states

30
76

443

Constant (Henry)

165

205

Continuous thermodynamics

404

Coordinates (pseudocritical)

82

Corresponding states

64

289

271

application to the calculation of residual


values
correlations based on the acentric factor

65
71

correlations based on the critical


compressibility factor

69

extension to mixtures

82

270

Lee and Kesler method

75

443

Cricondenbar

174

Cricondentherm

174

Critical compressibility factor

68

correlations using corresponding states and


the critical compressibility factor
numerical data
relationship with the acentric factor
Critical coordinates

69
435
69
33

calculation by group contributions

88

calculation from other properties

90

numerical data

93

435

Critical point

33

34

171

mixture

171

175

332

pure substance
Critical solution temperature

33
332

390

This page has been reformatted by Knovel to provide easier navigation.

Index Terms

Links

D
Dew point
calculation
Distribution function

169
194
404

E
Enthalpy
dependence on temperature and pressure

6
19

excess

159

of formation

414

numerical data

435

of melting (numerical data)

435

of mixing

154

of reaction

419

of vaporization
numerical data
residual
calculation by equation of state
Entropy
dependence on temperature and pressure

44

92

50

51

443

106

116

435

9
19
159

numerical data

435

calculation from an equation of state


Equations of state
application to

212

46

excess

residual

44

53
106

443

99

269

99

mixtures

269

phase equilibrium of mixtures

316

phase equilibrium of pure substances

100

121

This page has been reformatted by Knovel to provide easier navigation.

Index Terms

Links

Equations of state (Cont.)


thermodynamic property calculations

99

106

116

292
Benedict, Webb, and Rubin

111

Carnahan-Starling

134

cubic

112

127

279

279

290

292

106

mixing rules

298
Peng-Robinson

115

Redlich-Kwong

113

rigid spheres

134

Soave-Redlich-Kwong

113

specific

136

Tait equation

138

van derWaals

112

virial

102

105

146

277

11

425

150

153

179

Equilibrium condition
Equilibrium condition between phases

345
Excess values
calculation using equation of state

159
293

dependence on temperature, pressure , and


composition

160

heat capacity

211

volume

210

F
First law

3
8

This page has been reformatted by Knovel to provide easier navigation.

Index Terms

Links

Flory theory

223

Fluid-solid equilibria

355

calculation

359

description

356

Free volume
Fugacity

378

365

379
53

calculation from activity coefficients

160

calculation from equations of state

106

calculation using corresponding states

443

152

117

dependence on temperature, pressure and


composition
in mixtures

153
153

calculation

157

in an ideal solution

156

of a pure substance

53

calculation

54

282

G
Gibbs energy
dependence on temperature and pressure
excess

13
19
159

at infinite pressure

298

at zero pressure

309

relationship with activity coefficients

160

relationship with fugacity coefficients


and the excess Gibbs energy
relationship with mixing rules
of formation
numerical data
of mixing

163
293
414
435
155

This page has been reformatted by Knovel to provide easier navigation.

282

Index Terms

Links

Gibbs energy (Cont.)


of reaction
residual
calculation from equations of state
Gibbs-Duhem equation

419
51
106

118

151

153

161

187

Gibbs-Helmholtz equation

20

Group contributions

84

248

activity coefficient

248

257

critical coordinates

88

155

application to the calculation of

enthalpy of formation

416

equations of state parameters

130

properties of the ideal gas

385

85

H
Heat capacity

ideal gas
calculation

25

numerical data

435

85

residual
calculation from corresponding states

443

Heat of mixing

154

Heat of reaction

421

Heat of vaporization
dependence on temperature

44

46

44

92

numerical data

435

Helmholtz energy

12

dependence on temperature and volume

212

19

This page has been reformatted by Knovel to provide easier navigation.

386

Index Terms

Links

Helmholtz energy (Cont.)


excess
at constant packing fraction

303

at infinite pressure

314

residual
calculation from equations of state

117

Henry Constant

165

Heteroazeotrope

340

Hydrates

367

calculation of formation equilibria

370

formation conditions

367

205

289

I
Ideal gas

23

Ideal gas state

23

Interaction parameter

279

24

286

290
Internal energy
dependence on temperature and volume
Ionic solutions

3
19
263

L
Lattice model
Lee and Kesler method

229
75

443

Liquid phase (properties in the)


mixtures
pure substances

164
54

This page has been reformatted by Knovel to provide easier navigation.

289

Index Terms

Links

Liquid-liquid equilibria

329

calculation

345

description

331

in polymer mixtures

389

selectivity

337

Liquid-liquid-vapor equilibria

339

calculation

345

description

339

Liquid-vapor equilibria

167

calculation

190

condition

179

correlation

218

282

dependence on temperature, pressure, and


composition

184

Liquid-vapor equilibrium coefficient

179

Local composition

227

Lumping

398

183

M
Melting
enthalpy

435

temperature

435

MHV2 method

309

Mixing rules
alternatives

290

classical

279

derived from excess Gibbs energy

298

309

derived from excess Helmholtz energy

304

314

Mixing values

154

Mixtures (multicomponent)

395

This page has been reformatted by Knovel to provide easier navigation.

316

Index Terms
Molecular simulation
Multicomponent mixtures

Links
63
395

N
NRTL model
Numerical databases

239
26

435

P
Paraffins (crystallization)

363

Partial molar values

144

Peng-Robinson equation of state

115

Petroleum fluids

395

Polymers

375

activity coefficients calculation

377

blends

389

equations of state

383

Flory theory

223

group contributions

386

using continuous thermodynamics

408

Poynting correction

56

Pseudocomponents

396

Pseudocritical coordinates

145

382

384

385

386

82

271

229

388

Q
Quasi chemical model

R
Rachford-Rice
Rackett equation

196
70

This page has been reformatted by Knovel to provide easier navigation.

Index Terms

Links

Raoult (law of)

180

Redlich-Kwong equation of state

113

Reduced variables

65

Reference state

144

Regular solutions

231

Residual properties

49

293

53

65

443
calculation from equations of state

106

calculation using corresponding states

76

enthalpy

50

Gibbs energy

51

Retrograde condensation

171

Rigid spheres (equation of state for)

134

116

S
Scatchard-Hildebrand
Second law

237
8

Soave-Redlich-Kwong equations of state

113

Solubility of gases in liquids

204

Solubility parameter

232

Stability condition

188

323

330

26

58

415

Standard state

435
State (reference)

144

293

State (standard)

26

58

435
Stavermann

226

This page has been reformatted by Knovel to provide easier navigation.

415

Index Terms

Links

T
Tait equation

138

Thermochemical data

414

Thermodynamics (continuous)

404

435

U
UNIFAC method

251

UNIQUAC model

244

Units

487

477

V
van der Waals equation of state
Vapor pressure

112
31

33

calculation using corresponding states

74

77

Clapeyron equation

36

correlations

37

41

Virial coefficient

30

72

103

102

105

106

146

277

127

281

Virial equations of state

Volume translation

W
Water, hydrocarbon systems

348

Wilson equation

238

This page has been reformatted by Knovel to provide easier navigation.

34

También podría gustarte