You are on page 1of 254

TR diss

2168

CRITICAL PHENOMENA
IN BINARY FLUID MIXTURES

Classification of Phase Equilibria


with the Simplified-Perturbed-Hard-Chain Theory
Cover: p,T-Projection of a binary mixture showing a closed liquid-liquid
critical curve and a high-pressure critical curve

Copyright © 1992, by A. van Pelt, Rotterdam, the Netherlands

All rights reserved. No part of this publication may be reproduced, stored in a


retrieval system or transmitted, in any form or by any means, electronic,
mechanical, photocopying, recording, or otherwise, without the written prior
permission of the author.

Printed in the Netherlands


CRITICAL PHENOMENA
IN BINARY FLUID MIXTURES

Classification of Phase Equilibria


with the Simplified-Perturbed-Hard-Chain Theory

KRITISCHE VERSCHIJNSELEN
IN BINAIRE FLUÏDE MENGSELS

Klassifikatie van fasenevenwichten


met de Simplified-Perturbed-Hard-Chain Theory

PROEFSCHRIFT

ter verkrijging van de graad van doctor aan de Technische Universiteit Delft op
gezag van de Rector Magnificus Prof. drs. P.A. Schenck,
in het openbaar te verdedigen ten overstaan van een commissie aangewezen
door het College van Decanen
op donderdag 17 december 1992 te 19.00 uur, door

Antonius van Pelt

Chemisch Technoloog
Geboren te Rotterdam
Dit proefschrift is goedgekeurd door de promotor
Prof. dr. ir. J. de Swaan Arons

en de toegevoegd promotor
Dr. ir. C.J. Peters

De promotiecommissie bestaat uit:

Prof. P.A. Schenck , Rector Magnificus


Prof. J. de Swaan Arons , promotor
Dr. C.J. Peters , toegevoegd promotor
Prof. P.H.E. Meijer , The Catholic University of Washington , V.S.
Prof. M.D. Donohue , Johns Hopkins University , V.S.
Prof. G.M. Schneider , Ruhr Universität Bochum , Duitsland
Dr. U.K. Deiters , Ruhr Universität Bochum , Duitsland
Prof. A.H.M. Levelt , Katholieke Universiteit Nijmegen , Nederland
De natuur gedoogt dat gij haar bespiedt,

niet dat gij haar ontraadselt.

Pythagoras
DANKWOORD / ACKNOWLEDGEMENT / DANKSAGUNG

Op deze plaats wil ik iedereen bedanken die heeft bijgedragen aan het totstandkomen van
dit proefschrift. In de eerste plaats zijn dat natuurlijk mijn promotor Prof.dr.ir. Jakob de Swaan
Arons en mijn begeleider Dr.ir. Cor Peters. Zij hebben het mij mogelijk gemaakt dit promotie-
onderzoek in de sektie "Toegepaste thermodynamika en fasenleer" te verrichten, het manuscript
van dit proefschrift gelezen en dit bekommentarieerd. Dr.ir. Theo de Loos bedank ik voor zijn
suggestie berekeningen rond het Van Laar punt uit te voeren, zie paragraaf 8.4. Verder bedank ik
alle, niet met name genoemde, (ex-)leden van onze sektie voor hun collegialiteit.
Prof. Paul Meijer (Catholic University of America, Washington DC.) dank ik voor vele
interessante discussies, suggesties en adviezen.
Herrn Priv. Doz. Dr. Ulrich Deiters bin ich dankbar für seine Einladung fünf Monate an
der Ruhr-Universität Bochum, Deutschland, zu verbringen, für die Bereitstellung der Programme
zur Berechnung globaler Phasendiagramme und für die Einweisung in ihre Benutzung.
Selbstverständlich danke ich auch den anderen Mitarbeitern des "UKD-Teams": Andreas Bolz,
Thomas Kraska, Gereon Hintzen und Martin Bluma.
Uiteraard bedank ik ir. Jaap Kleimeer en Kees Ravesteyn van de systeemgroep voor het
operationeel houden van de computers en installeren van software.
Dr. Tony Wells is gratefully acknowledged for reading the manuscript of this PhD-thesis
and correcting my (ab)use of the english language.
Tenslotte ben ik dank verschuldigd aan de stichting Scheikundig Onderzoek in
Nederland (SON), de tweede-geldstroom organisatie voor de chemie die deel uitmaakt van de
Nederlandse Organisatie voor Wetenschappelijk Onderzoek (NWO), voor de financiële
ondersteuning van dit projekt (projektnummer: 700-343-024).

En 'last but not least' bedank ik Henriëtte 'Piep' Sla, die, door haar voortdurende
aanmoediging, er mede voor gezorgd heeft dat dit proefschrift en het hieraan voorafgegane werk
tot een goed einde is gebracht.

Rotterdam, 18 oktober 1992,

Ton van Pelt


i

Contents
1. INTRODUCTION 1

1.1 The role of thermodynamics in process development . . . . . . . . . . . . . . . . . . . . . . . . . . . 1


1.2 Supercritical fluid extraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Relevance of this work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4 Outline of this thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

2. GENERAL DISCUSSION OF FLUID PHASE EQUILIBRIA 13

2.1 Classification of fluid phase behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13


2.2 Type I phase behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.3 Type II phase behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.4 Type III phase behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.5 Type IV phase behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.6 Type V phase behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.7 Type VI phase behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

3. THE SIMPLIFIED-PERTURBED-HARD-CHAIN THEORY: DEVELOPMENT 27

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.1.1 The Van der Waals type equations of state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.1.2 The Benedict-Webb-Rubin type equations of state . . . . . . . . . . . . . . . . . . . . . . . . 28
3.1.3 Reference fluid equations of state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.1.4 Augmented hard body equations of state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.2 The Perturbed Hard Chain Theory and its family members . . . . . . . . . . . . . . . . . . . . . 31
3.3 The partition function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.4 The free volume . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.5 The mean potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.6 Mixing rules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
ii Contents

4. THE SIMPLIFIED-PERTURBED-HARD-CHAIN THEORY: PERFORMANCE 56

4.1 Pure component parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56


4.1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
4.1.2 The optimization procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
4.1.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.2 The temperature dependence of the attractive term . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
4.3 Virial coefficients from the SPHCT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
4.4 The isochoric heat capacity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.5 Binary flash calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
4.6 Calculation of Henry coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
4.7 Calculation of critical curves in binary mixtures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
4.7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
4.7.2 The Hicks-Young algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
4.7.3 Stability tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
4.7.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
4.7.5 Discussion and conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

5. TRANSITION STATES IN THE GLOBAL PHASE DIAGRAM 113

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113


5.2 Thermodynamic conditions for the various transition states . . . . . . . . . . . . . . . . . . . . 116
5.2.1 The tricritical state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
5.2.2 The double critical endpoint . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
5.2.3 The mathematical double point . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . 126
5.2.4 The azeotropic boundary curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
5.2.5 The azeotropic critical endpoint . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
5.2.6 The zero temperature endpoint . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141

6. COMPUTATIONAL TECHNIQUES 143

6.1 Evaluation of the equations and unknowns . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143


6.2 Optimization methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
6.2.1 Gradient methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
6.2.2 Method of steepest descent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
6.2.3 Newton-Raphson method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
6.2.4 Gauss method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
6.2.5 The Marquardt method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
6.3 Calculation of the tricritical curve . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
6.4 The use of MAPLE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
6.4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
6.4.2 Derivation of expressions for Gnx with Maple . . . . . . . . . . . . . . . . . . . . . . . . . . 155
6.4.3 Derivation of expressions for AiVjx with Maple . . . . . . . . . . . . . . . . . . . . . . . . . . 158
Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
Contents iii

7. THE GLOBAL PHASE DIAGRAM FROM THE SPHCT EQUATION 163

7.1 The global phase behaviour of equal-sized and equally flexible molecules . . . . . . . 163
7.2 The global phase behaviour of molecules with equal flexibilities but slightly
different sizes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
7.3 The global phase behaviour of molecules with equal flexibilities but large size
differences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
7.4 The global phase behaviour of molecules with equal sizes but unequal flexibilities 183
7.5 The global phase behaviour of molecules with unequal flexibilities and sizes . . . . . . 186
Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190

8. SPECIAL PHENOMENA FROM THE GLOBAL PHASE DIAGRAM 191

8.1 The type VI/VII-region in the global phase diagram . . . . . . . . . . . . . . . . . . . . . . . . . . 191


8.2 Critical curves in the binary mixture ethane(1)+n-butane(2) . . . . . . . . . . . . . . . . . . . . 198
8.3 Calculation of type VIII phase behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
8.3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
8.3.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
8.3.3 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
8.4 The Van Laar point in the T,x-plane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
8.4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
8.4.2 The neighbourhood of the Van Laar point . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
8.4.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
8.4.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226

9. CONCLUSIONS AND PROSPECT 227

SUMMARY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230

SAMENVATTING . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233

APPENDIX . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237

CURRICULUM VITAE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240


1

CHAPTER 1

INTRODUCTION

This thesis is mainly concerned with the prediction of phase equilibria in the critical
region and critical phenomena in binary fluid mixtures with a semi-theoretical equation of state.
With our investigations we hope to contribute to the understanding of these phenomena and
relate them to the structure of the individual molecules. The predictive capabilities, as far as
critical phenomena are concerned, of the used semi-theoretical model are explored in a general
way, e.g., we did not focus on systems of practical importance but we investigated all possible
variations in critical phase behaviour. However, in the various chapters, it will be shown that the
connection between the model and the behaviour of real systems has not been lost. In section 1.1
it is explained why theoretical investigations of thermodynamic models are important. In section
1.2, a separation process from the chemical industry that might benefit from our research is
introduced. Section 1.3 gives a general overview of previous investigations that have been made
in this field of research. This chapter ends with a summary of the scope and structure of the
thesis in section 1.4.

1.1 The role of thermodynamics in process development

Phase equilibrium thermodynamics is one of the most important fundamental sciences


for process development in the chemical industry. The results achieved by thermodynamics in
the last decades have led to a profound change in working methods for the development of all
processes in which phase equilibria play an important role. Whereas in the past the development
of these processes was dominated by experimental methods, e.g., laboratory experiments, the
working methods have changed significantly, becoming model-supported calculation methods
[1]. This holds in particular for unit operations like distillation and absorption. With the aid of
process simulators that contain physical property models, chemical processes can be designed
and/or optimized with a computer. We want to emphasize that for these computations one is
restricted to systems with thermodynamic properties that can accurately be described by a model.
This does not mean that basic data have become superfluous. On the contrary, investigations on
physical property models have shown that thermodynamic properties of many mixtures, like
phase equilibria, especially if the mixtures contain polar components or electrolytes, are hard to
predict from a model alone. In this respect basic data still are indispensible.
2 Chapter 1

In general, the use of physical property models in process development can lead to a
considerable reduction of costs with respect to process development that is based on experiments
only. However, it may be understood that this strategy depends on the availability of physical
property models. In this work, a physical property model, the Simplified-Perturbed-Hard-Chain
Theory (SPHCT), is thoroughly investigated with respect to the prediction of critical phenomena
in binary fluid mixtures. On the one hand, predictions of thermodynamic properties for pure
components and binary mixtures from the SPHCT are compared with experimental data. On the
other hand, a general overview is presented of the critical phenomena that can be obtained from
the SPHCT. This part is the most important aspect of this work. This general overview is
extremely useful for obtaining insight into the complicated world of phase behaviour.
The investigations that have been done are purely theoretical. It is not difficult to find a
practical example, to which the knowledge of critical phenomena is applied. Supercritical fluid
extraction is a good illustration of a practical application in the chemical industry where
knowledge about the critical region of mixtures is necessary. By no means do we want to focus
on the theoretical or practical implications of the process development of supercritical fluid
extraction. The next section only serves to show a link between our theoretical investigations and
one possible practical application. Our work might be seen as a theoretical support of the
description of all processes in which fluid phase equilibria play an important role.

1.2 Supercritical Fluid Extraction

Supercritical fluid extraction processes use the unique dissolving properties of solvents
near their critical point [2-4]. In the Propane Deasphalting Process lube oils are refined with
near-critical propane. Propane is also used in the Solexol Process for the purification and
separation of vegetable and fish oils. The Rose Process uses near-critical butane or pentane for
the extraction of high molecular weight components from oil residues. Supercritical carbon
dioxide at 160 bar and 70°C is used for the extraction of nicotine from tobacco [2-4].
A simplified flowsheet of a supercritical extraction plant is shown in Fig.1.1. Gaseous
solvent from the separator (5) and fresh solvent is condensed by a cooling machine (6) and led to
a solvent buffer vessel (7). The pressure of the liquified solvent is increased to the extraction
pressure by a pump (8). At this pressure, the solvent is heated to the desired temperature (9).
Thereafter, the solvent enters the extraction column (3) at the bottom. At the same time the
heated feed stream enters the extraction column at the top of the extraction column (3). After the
countercurrent extraction, the loaded solvent leaves the extraction column at the top. The
pressure of the loaded solvent is usually reduced (4) below the critical pressure of the solvent to
ensure a good separation of the solvent and the extract in the separator (5). The raffinate that
leaves the extraction column at the bottom can contain up to 50 wt.% solvent. The pressure of
this stream is reduced (10) to separate the solvent (12) from the raffinate in a separator (11).
Introduction 3

Figure 1.1: Simplified flowsheet of a supercritical extraction plant.

This solvent stream (12) may be recycled. To compensate for the losses of solvent in the raffinate
that may still occur, fresh solvent is added before the solvent recycle stream is condensed again.

The increased solubility phenomenon of the solute in the supercritical solvent in the
extraction column forms the basis of supercritical fluid processing. The enhanced solvating
power of a supercritical solvent is illustrated in Fig.1.2. In this figure the solubility of solid
naphthalene, yi, in supercritical carbon dioxide is shown as a function of the reduced pressure at
various temperatures [5,6]. At the saturated vapour pressure of solid napthalene for the three
reduced temperatures, i.e., pr≈0, the mole fraction of naphthalene in the vapour phase must be 1.
Increasing the pressure above the saturated vapour pressure of naphthalene leads to a sharp
decrease of the solubility. It is shown in Fig.1.2 that if the pressure is raised above the critical
pressure of carbon dioxide (the solvent), the mole fraction of naphthalene (the solute) increases
4 Chapter 1

strongly.

Figure 1.2: Solubility of naphthalene in Figure 1.3: Isotherms for pure CO2,
supercritical CO2 at various Tc=304 K, pc=7.38 MPa,
temperatures. Experimental ρc=0.01064 mole.cm-3. The
data were taken from [6]. reduced temperatures are: a:
0.79, b: 0.89, c: 0.99, d:
1.05, e: 1.20, f: 2.01.
Experimental data were
taken from [7].

This large solubility enhancement in the vicinity of the critical pressure of carbon dioxide
is mainly caused by the rapid increase in solvent density, see Fig.1.3. For reduced temperatures
(T/Tc) ranging from 1.0 to 1.2 small changes in the reduced pressure (p/pc) can change the
reduced density (ρ/ρc) of a solvent from about 0.1, a gas-like density, to about 2.5, a liquid-like
density. By adjusting the pressure and/or temperature, the properties of a supercritical solvent
can be changed from a gas, with little solvating power, to a liquid, with good solvating power
[8]. Besides the density of the solvent, there are more thermodynamic properties that influence
the solubility of a solute in a solvent, like the saturated vapour pressure of the solute and the
chemical structure of the solute and the solvent. The effect of the vapour pressure is visible at
high reduced pressures (pr>3) in Fig.1.2. The increase in solubility of naphthalene with
temperature at a specified reduced pressure, can be explained with the increased saturated vapour
pressure at higher temperatures. The effect of the chemical structure on the solubility can easily
be explained with the 'like-dissolves-like' rule. In general it can be stated that the solubility in
non-polar solvents, such as ethane and propane, will be smaller with decreasing vapour pressures
and increasing polarity of the solute. This solubility behaviour may change when more polar
solvents are used. The solubility effect of chemical structure differences can be identified by
defining an enhancement factor, Ei, as a measure for the affinity of a solvent for a solute:
Introduction 5

actual mole fraction solute i y .p


Ei = = i sat (1.1)
ideal mole fraction solute i xi . pi

The factor Ei represents the enhancement of the actual mole fraction of the solute, yi, over the
value xi.pisat/p, that can be calculated according to Raoult's law, if gas and liquid phase non-
idealities and pressure effects on the liquid or solid phase are neglected. If pure supercritical
solvents are used the choice is usually limited to non-polar solvents such as carbon dioxide,
ethane and propane. Higher solubilities can sometimes be obtained by using mixed solvents. For
more polar compounds a higher solubility can be obtained by adding small amounts of polar
entrainers such as methanol, ethanol, acetone or water to the non-polar solvent.
It is obvious that knowledge of the phase equilibria that occur in the supercritical fluid
extraction process is necessary for the design of this process. Basic data obtained from
experimental research are of the highest importance to acquire this knowledge. However, it can
be useful to have some knowledge of the densities and heat capacities of the various streams,
compositions of coexisting phases, such as raffinate and loaded solvent, etc., before one even
starts to do experimental research. These rough predictions obtained from a physical property
model can then be used to direct this experimental research, or if they are accurate enough, to
support process development. It should be stressed again that the example of the supercritical
extraction process only shows a link between our theoretical investigations on critical
phenomena and a practical application.

1.3 Relevance of this work

In the previous section an example was discussed of the application of supercritical


solvents in chemical engineering. It is obvious that for the design of such a process a
comprehensive knowledge of the thermodynamic properties of the solvent and the solute at near-
critical conditions of the solvent is necessary. It also has been mentioned in section 1.1, that
model-based calculations become more and more important for the design of chemical processes,
even though experimental data are still necessary. Equations of state provide chemical engineers
with useful tools for modelling the phase equilibria. However, more research has to be done to
improve the accuracy of the equations of state currently used by many chemical engineers in
process simulators.
In this work a semi-theoretical equation of state was tested in the critical region of binary
fluid mixtures. However, the emphasis of this work is not on the applicability of this equation of
state to the development of a particular process, but on its significance for a general and
systematic investigation of phase behaviour and critical curves in binary fluid mixtures. Research
in this field of science is not new but started more than two decades ago. Van Konynenburg and
Scott [9,10] showed that almost all known types of fluid phase equilibria - vapour-liquid, liquid-
liquid and gas-gas - can be generated, at least qualitatively using the Van der Waals equation of
state (VDW equation) and the Van der Waals mixing rules. They also suggested a classification
6 Chapter 1

of binary phase diagrams. Their classification is based on the absence or presence of three-phase
curves, and if present, on the way critical curves are connected to these three-phase curves. They
distinguished six different classes of binary fluid phase behaviour. However, due to the
limitations of the equation of state they were able to generate only five of these classes with the
VDW equation of state. Because the classification of Van Konynenburg and Scott [9,10] forms
the basis of our work, this classification will be discussed in the next chapter.
Furman et al. [11] made similar studies of lattice-gas mixtures. The resultant lattice-gas
equation of state was extended by Furman and Griffiths [12] to the Van der Waals lattice-gas
equation of state and these investigators discovered new types of phase behaviour. Van
Konynenburg and Scott completed their earlier work with the findings of Furman and Griffiths
[12] in an extensive review on fluid phase behaviour [13]. Furman and coworkers [11,12]
introduced another classification procedure. Their classification does not distinguish between the
various types of phase behaviour in a pressure-temperature-projection (p,T-projection) of the
critical curves in a binary system, but distinguishes between the various phenomena within one
mixture. They called a homogeneous phase: A. Examples are liquid (L), solid (S) and vapour (V).
A two-phase equilibrium is called A2, like L+V, S+L, S+V and L1+L2. When both coexisting
phases become identical, the critical state is called: B, e.g. L=V. Equilibrium between three
phases is called A3. Examples are L1+L2+V, S+L+V and S1+S2+L. When, in case of a two-phase
equilibrium, both phases have the same composition, the mixture is said to be azeotropic.
Furman and coworkers called this state: A2az. A critical endpoint, i.e., a critical phase that is in
equilibrium with an alternate (noncritical) phase, is called BA, e.g., L1=L2+V, and a critical
azeotropic point is called Baz. A quadruple point, a point in the p,T-plane where four phases are
B

in equilibrium, is denoted by A4, e.g., S1+S2+L+V. Even phenomena that are rare in binary
mixtures can be classified according to the classification of Furman and coworkers. A tricritical
point (unsymmetric) is denoted by C, e.g., L1=L2=V. A tricritical point is the point where two
critical endpoints on the same three-phase curve coincide. When a critical azeotropic phase (Baz) B

is in equilibrium with an another, noncritical, phase this phase is denoted by BazA. A tricritical
B

phase that is in coexistence with an alternate noncritical phase, is denoted by CA. Although this
classification gives a detailed insight into the phase behaviour of a certain binary mixture, it does
not give a general overview of possible phase behaviour in binary mixtures.
Clancy et al. [14] focussed on the influence of polar forces on the phase behaviour. They
used an equation of state that was derived from perturbation theory. However, they did not
include an investigation of all possible phase diagrams.
Mazur and coworkers [15,16] and Boshkov and coworkers [17,18] used the Ree equation
of state for Lennard-Jones molecules. They did extensive research on the global phase diagrams
for binary mixtures that are composed of equal-sized molecules. New classes were added by
them to the classification scheme of Van Konynenburg and Scott [9,10].

Meijer and coworkers [19-27] investigated the global phase behaviour of the symmetric
lattice-gas [28], the Van der Waals lattice-gas and the Tompa-model [29]. The latter model is a
variant of the lattice gas model for compressible binary mixtures in which the possibility of size
Introduction 7

difference between the components is included. They also used the classification of Van
Konynenburg and Scott [9,10] to distinguish between the various types of fluid phase behaviour.
Most of their research focussed on mathematical double points and on the phase behaviour
around the Van Laar point in the global phase diagram.
Systematic investigations of global phase behaviour were undertaken recently by Deiters
and Pegg [30] and Kraska and Deiters [31]. Deiters and Pegg [30] investigated all possible phase
behaviour that can be predicted with the Redlich-Kwong equation of state (RK equation). Their
investigations were based on the classification according to Van Konynenburg and Scott. Kraska
and Deiters [31] did similar calculations on the Carnahan-Starling-Redlich-Kwong equation of
state (CSRK equation). A new nomenclature for phase diagram classes of binary fluid mixtures
has recently been proposed by Bolz [32]. His classification scheme is a combination of the
classification of Van Konynenburg and Scott and the classification according to Furman and
coworkers. The new nomenclature primarly describes the topology and connectivity of critical
curves. Moreover, information about the critical curves is contained in this nomenclature: the
basic idea of Bolz was that the construction of a p,T-projection of the critical curves of a binary
system should be possible from the name.
The critical point of a pure component is always the starting point of either a critical
curve or a sequence of critical curves connected by three-phase curves. One starts with the
critical curve originating at the critical point of the component with the higher critical
temperature. All segments of the critical curve, that are connected by three-phase lines, are taken
into account. The target of this critical curve is indicated with a superscript. Possible targets are:
C for critical curves going upwards to a compact state at infinitely high pressures;
P for critical curves going to the other pure critical point;
Z for critical curves going to an endpoint, from where a three-phase curve goes to absolute
zero;
Q for curves going to an endpoint, from where a three-phase curve goes to a quadruple
point (four fluid states in equilibrium).

Subsequently, the critical curve originating at the critical point with the lower critical
temperature is described, in the same way as mentioned above. Then, additional critical curves
are classified as follows:
l for critical curves coming from infinitely high pressures and going to an endpoint;
n for critical curves with two endpoints;
u for critical curves coming from and going to infinitely high pressures, going through a
pressure minimum; these curves have no endpoints;
o for critical curves that form completely stable closed loops.

Finally, additional information can be added directly behind the critical curve to which it
refers. The following abbreviations should be used:
H for heteroazeotropic behaviour;
8 Chapter 1

A for azeotropic behaviour;


Q for a quadruple point;
M for maxima, the number of maxima may be indicated by a superscript.

Several common examples of the new nomenclature are: 1P, 1Pl, 1C1Z. These examples
correspond to types I, II and III phase behaviour according to the classification of Van
Konynenburg and Scott [9,10]. Rare phase behaviour, like the phase behaviour of binary systems
from the shield region, have longer names like: 1PAnlQ. An advantage of this nomenclature is
the fact that it is very descriptive.

It is obvious that a lot of work has already been done in the field of classification of
phase behaviour. However, most of the work that has been done sofar is based on two-parameter
equations of state that are more or less empirical. Initially we intend to approach the field of
classification of phase behaviour armed with a semi-theoretical equation of state, that has
primarily been designed to account for the nature of chainlike molecules. The equation of state is
chosen in such a way that new types of binary phase behaviour are to be expected, although this
cannot be ensured in advance. The equation of state should correspond to the objectives of our
research group at the University of Technology in Delft, one of which is the study of phase
behaviour of chain-like molecules. It is also hoped that our research will compliment the efforts
made previously by other investigators.
Following the trend of investigations on the VDW equation of state towards more
advanced equations of state like the RK and CSRK equation, it is logical to explore the ability of
the Simplified-Perturbed-Hard-Chain Theory to predict and describe critical curves and phase
behaviour in binary fluid mixtures. The equation of state that results from this theory, the SPHCT
equation, satisfies all three conditions that are mentioned above.

1.4 Outline of this thesis

It is impossible to cover all subjects that are related to the calculation of high pressure
phase equilibria in one PhD-thesis and so it is necessary to restrict oneself to a couple of subareas
of this field of research. It is not our objective to make improvements on the way the various
types of phase behaviour in binary systems are classified. We will stay close to the classification
system of Van Konynenburg and Scott [9,10,13]. We have chosen not to emphasize
investigations of types of phase behaviour that are unlikely to be found experimentally, like the
types of phase behaviour found in the so-called shield-region [11,12]. We also will not pay much
attention to possible numerical optimization procedures, that would be less time-consuming or
mathematically more elegant. However, we will emphasize the global phase behaviour that can
be predicted from the SPHCT equation and special attention is paid to particular critical
phenomena that can be calculated from this equation like high-pressure critical curves and liquid-
liquid immiscibility loops. Liquid-liquid immiscibility loops cannot be calculated from the
Introduction 9

VDW, RK and CSRK equations. High-pressure critical curves have been calculated from the RK
and CSRK equation but this type of critical curve has not gained much attention in this respect.
This chapter has given a short introduction on the subject. The role critical phenomena
can play in the chemical industry has been explained by means of an example: supercritical fluid
extraction. The various types of critical curves in binary fluid mixtures and the classification of
these critical curves according to Van Konynenburg are discussed in chapter 2. An extensive
derivation of the SPHCT equation of state that has been used to calculate critical curves in binary
mixtures, is given in chapter 3.
In chapter 4 we try to justify the choice of the SPHCT equation of state for this
investigation. In order to do so, a number of experimentally determined thermodynamic
properties such as saturated vapour pressures of pure components, temperature dependence of
the attractive term of the equation of state, second and third virial coefficients, the molar
isochoric heat capacity, binary isothermal phase diagrams, Henry coefficients and critical curves
in binary systems, have been compared with results calculated from the SPHCT equation of
state.
Chapter 5 deals with the global phase diagram. All phenomena represented in a global
phase diagram (tricritical points, Van Laar points, shield region, etc.), including the mathematical
conditions to calculate these phenomena from an equation of state, are discussed.
In chapter 6, the calculation procedures for the different curves in the global phase
diagram are discussed. The numerical algorithm for solving a set of non-linear equations, the
Marquardt-algorithm, is discussed. Most of the mathematical conditions have been elaborated
with the symbolic algebra program MAPLE. We will digress on the use of this symbolic
computation program in this chapter.
Chapter 7 describes the global phase diagrams that have been calculated from the
SPHCT equation of state. It turns out that from this equation of state ordinary types of phase
behaviour, as well as some additional interesting features, like closed-loop liquid-liquid
immiscibility, can be calculated. Another exceptional phenomenon observed is the appearance of
two tricritical points on the same critical curve, with the possibility that these tricritical points
coincide. This gives rise to new areas in the global phase diagram, and additionally, to new types
of binary phase behaviour. A comparison is made with global phase diagrams predicted from
other equations of state, like the VDW and RK equation of state.
In chapter 8, some peculiarities in the calculated global phase diagrams are discussed,
such as the already mentioned closed loop liquid-liquid immiscibility, the Van Laar point and
high-pressure immiscibility. Finally, in chapter 9 an outlook on the possible development of the
subject of global phase diagrams is given.
10 Chapter 1

NOMENCLATURE

E enhancement factor
L liquid phase
p pressure
S solid phase
T temperature
V vapour phase
x mole fraction of a component in the liquid phase
y mole fraction of a component in the vapour phase

Greek letters

ρ density

Superscripts and subscripts

az azeotropic
c critical property
i component index
r reduced property
sat saturated property

REFERENCES

1. S. Zeck. Thermodynamics in process development in the chemical industry. Importance,


benefits, current state and future development. Fluid Phase Equilibria 70, (1991), 125-140.
2. M.A. McHugh and V.J. Krukonis. Supercritical fluid extraction: principles and practice.
Butterworths. Stoneham. (1986).
3. M.E. Paulaitis, V.J. Krukonis, R.T. Kurnik and R.C. Reid. Supercritical fluid extraction.
Rev. Chem. Eng. 1, (1983), 179-250.
4. D.F. Williams. Extraction with supercritical gases. Chem. Eng. Sci. 36, (1981), 1769-1788.
5. Y.V. Tsekhanskaya, M.B. Iomtev and E.V. Mushkina. Solubility of naphthalene in ethylene
and carbon dioxide under pressure. Russ. J. Phys. Chem. 38, (1964), 1173-1176.
6. M.A. McHugh and M.E. Paulaitis. Solid solubilities of naphthalene and biphenyl in
supercritical carbon dioxide. J. Chem. Eng. Data 25, (1980), 326-329.
7. IUPAC, International Thermodynamic Tables of the Fluid State. Vol. 3: Carbon dioxide.
Butterworths, London, (1977).
8. S.R. Allada. Solubility parameters of supercritical fluids. Ind. Eng. Chem. Process Des. Dev.
23, (1984), 344-348.
Introduction 11

9. P.H. van Konynenburg. Critical lines and phase equilibria in binary mixtures. PhD-thesis
University of California Los Angeles. (1968).
10. R.L. Scott and P.H. van Konynenburg. 2. Static properties of solutions: van der Waals and
related models for hydrocarbon mixtures. Discuss. Faraday Soc. 49, (1970), 87-97.
11. D. Furman, S. Dattagupta and R.B. Griffiths. Global phase diagram for a three-component
model. Phys. Rev. B, 15, (1977), 441-464.
12. D. Furman and R.B. Griffiths. Global phase diagram for a Van der Waals model of a binary
mixture. Phys. Rev. A, 17, (1978), 1139-1148.
13. P.H. van Konynenburg and R.L. Scott. Critical lines and phase equilibria in binary van der
Waals mixtures. Phil. Trans. 298A, (1980), 495-540.
14. P. Clancy, K.E. Gubbins and C.G. Gray. Thermodynamics of polar liquid mixtures. Discuss.
Faraday Soc. 66, (1978), 116-129.
15. V.A. Mazur, L.Z. Boshkov and V.B. Fedorov. Phase behaviour in two component Lennard-
Jones systems. Dokl. Akad. Nauk SSSR. 282, (1985), 137-140.
16. V.A. Mazur, L.Z. Boshkov and V.G. Murakhovsky. Global phase behaviour of binary
mixtures of Lennard-Jones molecules. Physics Letters 104, (1984), 415-418.
17. L.Z. Boshkov and V.A. Mazur. Phase equilibria and critical lines of binary mixtures of
Lennard-Jones molecules. Russ. J. Phys. Chem. 60, (1986), 16-19.
18. L.Z. Boshkov. On the description of phase diagrams of two component mixtures with a
closed domain of demixing based on the one-fluid model of an equation of state. Dokl.
Akad. Nauk. SSSR. 294, (1987), 901-905
19. P.H.E. Meijer and M. Napiorkowski. The three-state lattice gas as model for binary gas-
liquid systems. J. Chem. Phys. 86, (1987), 5771-5777.
20. P.H.E. Meijer. Binary gas-liquid systems classification. Soc. Fr. de Chimie, Int. Symp. on
Supercritical Fluids, Oct. 17-19, Nice. M. Perrut ed., Vol. 1, (1988), 239-244.
21. P.H.E. Meijer. The influence of the chain length of long molecules on the equation of state in
binary gas liquid mixtures. J. Stat. Phys. 53, (1988), 543-548.
22. P.H.E. Meijer. Study of the critical line and its double point in the intermediate model.
Physica A 152, (1988), 359-364.
23. P.H.E. Meijer, M. Keskin and I.L. Pegg. Critical lines for a generalized three state binary
gas-liquid lattice model. J. Chem. Phys. 88, (1988), 1976-1982. Erratum: J. Chem. Phys. 90,
(1989), 3408.
24. P.H.E. Meijer. The van der Waals equation of state around the van Laar point. J. Chem.
Phys. 90, (1989), 448-456.
25. P.H.E. Meijer, I.L. Pegg, J. Aronson and M. Keskin. The critical lines of the van der Waals
equation for binary mixtures around the van Laar point. Fluid Phase Equilibria 58, (1990),
65-80.
26. P.H.E. Meijer and I.L. Pegg. Structure of the critical lines for the lattice gas model. Physica
A 174, (1991), 391-405.
27. M. Keskin, M. Gençaslan and P.H.E.Meijer. Evaluation and comparison of critical lines for
various models of gas-liquid binary systems. J. Stat. Phys. 66, (1992), 885-896.
28. H. Tompa. Phase relationships in polymer solutions. Trans. Faraday Soc. 45, (1949), 1142-
1152.
29. J.A. Schouten, C.A. ten Seldam and N.J. Trappeniers. The two-component lattice-gas model.
Physica 73, (1974), 556-572.
12 Chapter 1

30. U.K. Deiters and I.L. Pegg. Systematic investigation of the phase behavior in binary fluid
mixtures. I. Calculations based on the Redlich-Kwong equation of state. J. Chem. Phys. 90,
(1989), 6632-6641.
31. Th. Kraska and U.K. Deiters. Systematic investigation of the phase behavior in binary fluid
mixtures. II. Calculations based on the Carnahan-Starling-Redlich-Kwong equation of state.
J. Chem. Phys. 96, (1992), 539-547.
32. A. Bolz. Vergleichende Untersuchung globaler Phasendiagramme. PhD-thesis, Ruhr-
Universität Bochum, Germany, (1992).
13

CHAPTER 2

GENERAL DISCUSSION OF FLUID PHASE EQUILIBRIA

Van Konynenburg has shown that the Van der Waals equation of state when applied to
binary mixtures, predicts qualitatively most types of fluid phase equilibria that have been
observed experimentally. In this chapter these experimentally determined phase diagrams are
discussed, starting with an introduction in section 2.1. Special attention is paid to fluid phase
equilibria and critical phenomena. Van Konynenburg established a classification of the different
types of fluid phase behaviour in binary mixtures in which he distinguished six main types: type I
to VI, see sections 2.2-2.7.

2.1 Classification of fluid phase behaviour

In everyday life three states of matter exist: solid, liquid and vapour. When a pure liquid
is heated, it usually transforms into a vapour. The temperature at which this happens depends on
the applied pressure. At its boiling temperature, the liquid phase is said to be in equilibrium with
a vapour phase. The graphical representation in the pressure-temperature-plane (p,T-plane) of
these phases and their mutual coexistence is called a phase diagram. The phase diagrams,
describing all kinds of phase equilibria between vapour, liquid and/or solids in a binary system,
are called complete phase diagrams. Because of the great variety of phenomena, not all complete
phase diagrams can be discussed in this chapter. Valyasko [1] focussed on complete phase
diagrams showing all kinds of critical phenomena. Complete phase diagrams with phenomena
like polymorphism, azeotropy, formation of solid solutions and new compounds were not
discussed by Valyashko [1]. According to his classification, complete phase diagrams fall into
two groups, as far as critical phenomena are concerned:

Group I: Complete phase diagrams in which critical phases are not in equilibrium with a
solid phase

Group II: Complete phase diagrams in which critical phases occur that are in equilibrium
with a solid phase

It should be noted that critical phenomena play an important role in this classification and we
will continue with a more precise definition of critical phenomena. The critical point of a fluid is
the point in pressure-temperature-composition-space (p,T,x-space), at which the coexisting liquid
14 Chapter 2

and vapour phases become indistinguishable. All physical properties, such as density, refractive
index, etc., of the two coexisting phases will become identical. For a pure component this occurs
at a fixed temperature and pressure. For a binary system critical points may exist over a range of
pressures, temperatures and compositions, resulting in a critical curve.
Van Konynenburg and coworkers [2-4] extensively discussed a number of binary phase
diagrams that later turned out to belong to the group I phase diagrams according to the
classification of Valyashko [1]. In the studies of Van Konynenburg and coworkers [2-4] the solid
phase equilibria were not taken into account. Fig. 2.1 shows the classification of binary phase
diagrams suggested by Van Konynenburg and coworkers [2-4]. Their classification is based on
the presence or absence of three-phase lines and, when present, on the way critical curves are
connected to them. In type I mixtures, only vapour-liquid separation occurs, whereas in type II to
type VI mixtures liquid-liquid immiscibility also occurs in some regions of the phase diagram.
For type II and type VI mixtures, the vapour-liquid and liquid-liquid critical curves are quite
distinct. In type III, type IV and type V mixtures, the critical curve is separated into two or more
parts. These branches of the critical curve can no longer be clearly identified as being vapour-
liquid or liquid-liquid. A detailed discussion of the experimentally observed phase diagrams can
be found in several reviews, e.g., in Rowlinson and Swinton [5]. We will restrict ourselves here
to a short overview of the various types of experimentally determined phase diagrams by briefly
discussing these types according to the classification of Van Konynenburg and Scott. Van
Konynenburg [2] introduced the concept "class" as well as "type". Class 1 mixtures (types I and
II) possess a continuous critical curve from one pure fluid critical point to the other, whereas
class 2 mixtures (types III, IV and V) do not. Strictly speaking, Van Konynenburg and Scott [2-
4] assign type VI to a third class, because type VI mixtures show a closed liquid-liquid
immiscibility loop. Moreover, those authors did not use the term "type VI", but this term has
been used by subsequent authors [5] and is now in standard usage.
Each type of phase behaviour will be illustrated with one or more examples of binary
mixtures that show this phase behaviour. A detailed overview of experimentally determined
critical data for binary systems can be found in [6,7].

2.2 Type I phase behaviour.

Type I mixtures have a continuous vapour-liquid critical curve without any liquid-liquid
immiscibility. Often, in real mixtures, crystallization of one of the components masks a low
temperature liquid-liquid phase separation. Type I behaviour occurs in binary systems with two
substances that are chemically similar and/or have critical properties that are comparable.
Typical examples are ethane + 2-methylpropane [8], CO2 + O2 [9,10] and benzene +
cyclohexane [11,12]. Mixtures of substances belonging to a homologous series deviate from
simple type I behaviour only when their size difference, and thus the values of their critical
properties, exceeds a certain ratio.
Fluid phase equilibria 15

Figure 2.1: The six main types of binary fluid phase behaviour, according
to the classification of Van Konynenburg and Scott [2-4].

For example, in the n-alkane series, with methane as one component a change from type I to type
V occurs first with n-hexane [13], whereas, with ethane as the lighter component, n-octadecane is
the first to exhibit partial miscibility in the liquid state [14]. With propane, this behaviour occurs
when the second component is n-C30 [15].
Type I mixtures can be conveniently distinguished first by considering the shape of the
continuous critical curve that connects the critical points of the two pure components. In
addition, the presence or absence of azeotropy should be noted. Fig.2.2 represents the various
shapes of the p,T-projection of the critical curve. In this diagram the pure vapour pressures are
represented as solid curves and critical curves are shown as dotted curves. The pure component
critical points are shown as open circles.
16 Chapter 2

Mixtures corresponding to curve c, where the


critical curve is almost linear in both the p,T-
and the T,x-planes, are usually formed from
substances with very similar critical properties,
e.g., n-decane + benzene [16]. Mixtures
corresponding to curve b where the critical
curve is convex upwards and frequently
exhibits a maximum in both the p,T- and p,x-
planes are extremely common among type I
systems and occur whenever there are mode-
rately large differences between the critical
temperatures or volumes of the pure
components, e.g., propane + n-octane [17].
Mixtures whose critical curve conforms to the
shape of curve a are extremely rare. In this
case there is partial immiscibility at tem-
peratures higher than the critical temperatures
of both components. By common defintion the Figure 2.2: Six possible types of continuous
substances should thus be considered to be in critical curves for type I mix-
tures. The solid phases are not
the gaseous state. Phase behaviour like this, is shown.
an example of so-called 'gas-gas immiscibility
of the third kind'. A binary system that exhibits this phase behaviour is cycloheptane +
tetraethylsilane [18]. Type I mixtures, whose critical loci are everywhere concave upwards,
frequently show a pressure minimum as in curve d. A typical example is propane + hydrogen
sulphide [19]. The critical curve e extends through a temperature minimum in the p,T-
projection. This phase behaviour is observed for several mixtures and is usually associated
with the occurrence of a maximum-pressure azeotrope extending up to the critical curve. A
typical example is acetone + n-pentane [20]. The final type of critical locus, shown in Fig.2.2,
is critical curve f. The critical curve has a temperature minimum and is concave upwards.
This type of phase behaviour is found in mixtures that are composed of an n-alkanol and a n-
alkane [21]. The calculational work of Van Konynenburg and coworkers [2-4] have shown
that if a binary mixture of equal-sized molecules with a continuous vapour-liquid critical
curve exhibits maximum-pressure azeotropy, there is always a liquid-liquid critical curve at
relatively low temperatures. This means that the binary system belongs to type II. In other
words: computationally a type I mixture with maximum-pressure azeotropy does not exist for
equal-sized molecules. This is schematically shown in Fig.2.3.
It is shown in Fig.2.3 that the azeotropic curve is tangent to the p,T-projection of the
vapour-liquid critical curve. Consequently, the azeotropic curve does not meet the critical curve
at the temperature minimum of the critical curve. This type of azeotropy is commonly called
"positive", "maximum-pressure" or "minimum-boiling" azeotropy. The latter term refers to an
isobar on a T,x-cross-section along which the azeotrope is the two-phase point of minimum
Fluid phase equilibria 17

temperature. The term maximum-pressure azeotropy is preferred by us and will be used in this
work. An example of a mixture that forms a maximum-pressure azeotrope is the system CO2 +
ethane [22].

Figure 2.3: Type II phase behaviour with a Figure 2.4: Type I phase behaviour with a
maximum-pressure azeotrope. minimum-pressure azeotrope.

"Negative", "minimum-pressure" or "maximum-boiling" azeotropy is also possible. Here,


in contrast to Fig.2.3, the critical curve displays a maximum in the critical temperature. The
minimum-pressure azeotrope in a p,T-projection lies to the right of both pure-fluid vapour
pressure curves, see Fig.2.4. If the bubblepoint temperature of a mixture is higher than that of
either pure fluid, it follows that the cross (1-2) interaction is more strongly attractive than either
the (1-1) or (2-2) interaction. This suggests the idea that the unlike molecules will undergo a
chemical reaction. According to this concept the mixture is composed of three components,
instead of two. A binary system that shows minimum-pressure azeotropy is dimethyl ether + SO2
[23]. Another possibility, occurring in the binary ethanol + benzene [24], is that an azeotrope
exists at lower pressures; however, as the pressure is increased it terminates on a pure fluid
vapour pressure curve and does not reach the critical curve. In almost all binary systems showing
minimum-pressure azeotropy, the required interaction between the unlike molecules is lost at
temperatures that correspond to the vapour-liquid critical region.
The vapour pressure curves of pure components sometimes intersect in the p,T-plane.
The point of intersection is called a Bancroft point. Binary mixtures with a Bancroft point
usually show maximum-pressure azeotropy. However, systems that contain a Bancroft point can
show a special form of azeotropy: double azeotropy, see Fig.2.5. An example of a binary system
that exhibits double azeotropy is benzene + hexafluorobenzene [25-27]. More recently double
azeotropy was found in the system 2,6-dimethylpyridine + 4-methylpyridine [27]. Both
azeotropic curves end at an azeotropic endpoint. This is illustrated in Fig.2.5.
18 Chapter 2

Figure 2.5: Type I mixture with a Bancroft Figure 2.6: Type I phase behaviour, including
point (B) and showing double high-pressure immiscibility. This
azeotropy. Both azeotropic type of phase behaviour is often
curves end at an endpoint. considered to be a special form of
type VI phase behaviour.
Fig.2.6 illustrates the phase diagram for water + 3-methylpyridine [28]. Both components
are fully miscible in the liquid and vapour phase at low pressures, which suggests that this
mixture belongs to the type I mixtures. At higher pressures, i.e., above 1400 bar, liquid-liquid
phase separation occurs. This phenomenon is called 'high-pressure immiscibility'. Because this
type of phase behaviour shows close resemblance with special forms of type VI phase behaviour,
see section 2.7, it is often considered as a special form of type VI phase behaviour.

2.3 Type II phase behaviour

As with type I mixtures, type II mixtures have a continuous vapour-liquid critical curve.
In addition there is a three-phase line, L1L2V in the p,T-projection between the vapour pressure
curves, ending at an upper critical endpoint (UCEP). From this UCEP a second critical curve of
liquid-liquid nature goes to infinite pressures where the fluid mixture approaches closest packing.
The liquid-liquid critical curve can also end at an endpoint S+(L1=L2). Actually the liquid-liquid
critical curve can be extended to pressures lower than the pressure of the UCEP. In this case the
extended part of the critical curve is metastable until it reaches the point of thermodynamic
instability. Generally the liquid-liquid critical curve has one of the three shapes in the p,T-plane,
that are shown in Fig.2.7. In this figure the dash-dotted curve represents the three-phase line
L1L2V. The three possible shapes of the liquid-liquid critical curve can be described as follows:
Fluid phase equilibria 19

a the liquid-liquid critical curve has a


negative slope in the p,T-projection,
(∂ p / ∂ T)c < 0.
b the liquid-liquid critical curve has a
positive slope in the p,T-projection,
(∂ p / ∂ T)c > 0.
c the critical curve has a negative slope in the
UCEP and changes via a temperature
minimum to a liquid-liquid critical curve with
a positive slope at higher pressures.

An example of a type II system is carbon


dioxide + n-octane [29]. It is possible that a
type II system exhibits maximum- or
minimum-pressure azeotropy. Van Konynen-
burg and Scott denoted type II phase
behaviour with maximum-pressure azeotropy
as type II-A, see Fig.2.3. Type II phase
Figure 2.7: Three possible types of critical behaviour with minimum-pressure azeotropy
curves in type II mixtures. could not be calculated from the Van der
Waals equation of state for equal-sized
molecules. In the binary system water + phenol [30] positive azeotropy occurs as well as liquid-
liquid immiscibility. Minimum-pressure azeotropy together with liquid-liquid immiscibility is
extremely rare. The occurrence of this behaviour at low vapour pressures implies that the unlike
molecules weakly attract each other in a certain composition range, i.e., the composition range
where the demixing occurs. On the other hand they form a negative azeotrope outside this
composition range, which implies strong attractive forces between the unlike molecules.
According to [5], there is strong evidence that the system acetic acid + triethylamine exhibits this
kind of phase behaviour. This type of phase behaviour is schematically shown in Fig.2.8.
If the three-phase line is above the vapour pressure curve of the most volatile component,
a heterogeneous azeotrope or heteroazeotrope is formed. This type of phase behaviour, type II-
HA, is illustrated in Fig.2.9. Type II-HA is probably the most common form of type II phase
behaviour and is found if two liquids of similar vapour pressure are only partially miscible.
Heteroazeotropy occurs in the mixtures methanol + n-hexane and methanol + n-heptane [21,31].
20 Chapter 2

Figure 2.8: Binary mixture showing type II Figure 2.9: Binary mixture showing type II-
phase behaviour, including a HA phase behaviour.
minimum-pressure azeotrope.

2.4 Type III phase behaviour

Type III mixtures have two distinct critical curves, one starting at the critical point of the
component with the higher critical temperature that goes to infinite pressures. The other critical
curve starts at the critical point of the component with the lower critical temperature and meets a
three-phase line L1L2V at an UCEP. Four possible kinds of type III behaviour are shown in
Fig.2.10 in p,T-projections.
A typical example of a system that has a pressure maximum and a pressure minimum in
its critical curve that goes to infinite pressures is the system ethane + methanol [32]. This
corresponds to curve a in Fig.2.10. This type of phase behaviour is also called type IIIm, where
the subscript m refers to the pressure-minimum. A binary mixture that corresponds to curve b in
Fig.2.10 is ethane + nitromethane [33]. The curves c and d in Fig.2.10 show different kinds of
gas-gas immiscibility. In gas-gas immiscibility of the first kind the critical curve starting at the
critical point of the component with the higher critical temperature moves to higher pressures
and temperatures on increasing the mole fraction of the other component, see curve d in Fig.2.10.
A typical example of a binary system that shows gas-gas immiscibility of the first kind is He +
Xe [34]. In gas-gas immiscibility of the second kind the critical curve starting at the critical point
of the component with the higher critical temperature moves to higher pressures but initially
passes through a temperature minimum on increasing the mole fraction of the other component,
see curve c in Fig.2.10. Binary mixtures that display gas-gas immiscibility of the second kind are
NH3 with CH4, Ar or N2. These binary systems are discussed by Schouten [35].
Fluid phase equilibria 21

Figure 2.10: Four possibilities for type III Figure 2.11: Type III phase behaviour with
mixtures. heteroazeotropy.

Binary mixtures have been investigated where the three-phase line L1L2V lies everywhere
above the vapour pressure curve of the more volatile component. This phenomenon is called
heteroazeotropy and this type of phase behaviour, type III-H, is illustrated in Fig.2.11. The
binary systems H2O + lower n-alkane up to n-C24H50 [36] are known to exhibit type III-H beha-
viour. Binary mixtures that are composed of H2O + n-alkane with a carbon number that is higher
than 24, still show type III-H phase behaviour [36] although the connectivity of the critical
curves differs from the one that occurs at lower carbon numbers (i.e., n < 24).

2.5 Type IV phase behaviour

In type IV phase behaviour three distinct critical curves exist, as is shown in Fig.2.12.
One of them is a liquid-liquid critical curve that starts at the UCEP of a three-phase line L1L2V
and rapidly goes to infinite pressures. The second critical curve starts at the critical point of the
component with the lower critical temperature and ends in an UCEP of a second three-phase line
L1L2V. The third critical curve starts at the critical point of the component with the higher critical
temperature and ends at a Lower Critical End Point (LCEP) of the second three-phase line L1L2V.
The third critical curve continuously changes its character. Near the critical point of the pure
component with the higher critical temperature the critical curve has a vapour-liquid character,
whereas near the LCEP the critical curve has a liquid-liquid character. Type IV phase behaviour
is known to occur in the binary systems methane + 1-hexene [37] and CO2 + n-tridecane [38].
22 Chapter 2

Figure 2.12: Type IV phase behaviour. Figure 2.13: Type V phase behaviour.

2.6 Type V phase behaviour

Type V mixtures, see Fig.2.13, have two distinct critical curves and correspond to type
IV mixtures without the steep liquid-liquid critical curve. However, in mixtures that are known
to belong to the type V mixtures the distinct liquid-liquid critical curve may be obscured by the
appearance of the solid phase. Mixtures of n-alkanes with large size differences display type V
phase behaviour. While the system methane + n-pentane still displays type I phase behaviour,
type V behaviour occurs in the system methane + n-hexane [13], whereas, with ethane as the
lighter component, n-octadecane is the first to exhibit partial miscibility in the liquid state [14].
With propane, this behaviour occurs when the second component is n-C30 [15]. A well-studied
example of a mixture showing type V behaviour is the previously mentioned system methane +
n-hexane [13]. In this system immiscibility is found: at temperatures below the LCEP (182.46 K,
[39,40]) methane and n-hexane are fully miscible in the liquid state.

2.7 Type VI phase behaviour

Binary mixtures showing type VI phase behaviour have a continuous vapour-liquid


critical curve between the critical points of the pure components. In addition, type VI mixtures
have a closed liquid-liquid critical curve that starts at a LCEP of a three-phase line L1L2V and
ends at an UCEP on the same three-phase line. The liquid-liquid critical curve can be extended
below both endpoints in the p,T-projection. Each of the metastable parts of the critical curve ends
Fluid phase equilibria 23

at a point where local instability begins. These two points are connected by an unstable critical
curve that runs more or less parallell to the three-phase line L1L2V. Together the stable,
metastable and unstable parts of the liquid-liquid critical curve form one closed critical curve.
Type VI phase behaviour is schematically shown in diagram a of Fig.2.14. The maximum on this
liquid-liquid critical curve is called a 'hypercritical point'. Type VI phase behaviour is found in
mixtures where strong intermolecular bonding, such as hydrogen-bonding, can occur. A common
example of type VI phase behaviour can be found in the system H2O + 2-butoxyethanol [41].
Schneider [41] has written an extensive review on type VI phase behaviour.

Figure 2.14: Four possibilities of type VI phase behaviour. Diagram c can also be seen as a
special form of type I phase behaviour.
24 Chapter 2

In addition to the relatively low-pressure liquid-liquid immiscibility, a quite distinct


liquid-liquid critical curve can appear at higher pressures, see diagram b of Fig.2.14. This type of
phase behaviour occurs in the binary system heavy water + 2-methylpyridine [41]. Note the
resemblance with type I phase behaviour with a high pressure immiscibility region, shown in
diagram c of Fig.2.14. If the high-pressure critical curve interferes with the low-pressure critical
curve, a 'tube of immiscibility' appears in the p,T,x-space. An example of this type of phase
behaviour can be found in the binary mixture water + 3-methylpyridine [27], see diagram d in
Fig.2.14.
The various types of phase behaviour considered in this section and the previous sections
do not cover all experimentally determined phase diagrams. However, most of the qualitatively
different types of phase behaviour have been discussed. Van Konynenburg [2] evaluated the
various types of phase equilibria that could be calculated with the Van der Waals equation of
state in a systematic way. In the following chapters, we will also perform a systematic
investigation of fluid phase behaviour and critical curves in binary mixtures. However, we have
chosen a model that has been developed primarly to account for the nature of chainlike
molecules: the Simplified-Perturbed-Hard-Chain-Theory (SPHCT). The SPHCT and the
equation of state that results from it, will be thoroughly discussed in the next chapter.

NOMENCLATURE

L liquid phase
LCEP Lower Critical EndPoint
p pressure
S solid
T temperature
UCEP Upper Critical Endpoint
V vapour
x mole fraction of the heavier component

REFERENCES

1. V.M. Valyashko. Complete phase diagrams of binary systems with different volatility
components. Z. Phys. Chemie 267, (1986), 481-493.
2. P.H. van Konynenburg. Critical lines and phase equilibria in binary mixtures. PhD-thesis
University of California Los Angeles. (1968).
3. R.L. Scott and P.H. van Konynenburg. 2. Static properties of solutions: van der Waals and
related models for hydrocarbon mixtures. Discuss. Faraday Soc. 49, (1970), 87-97.
4. P.H. van Konynenburg and R.L. Scott. Critical lines and phase equilibria in binary van der
Waals mixtures. Phil. Trans. 298A, (1980), 495-540.
5. J.S. Rowlinson and F.L. Swinton. Liquids and Liquid Mixtures Butterworths Monographs in
Chemistry, London, Chap. 6, (1982).
Fluid phase equilibria 25

6. C.P. Hicks and C.L. Young. The gas-liquid critical properties of binary mixtures. Chem.
Rev. 75, (1975), 119-175.
7. R.J. Sadus. High pressure phase behaviour of multicomponent fluid mixtures. Elsevier,
Amsterdam, (1992).
8. Th.W. de Loos, H.J. van der Kooi and P.L. Ott. Vapor-liquid critical curve of the system
ethane + 2-methylpropane. J. Chem. Eng. Data 31, (1986), 166-168.
9. H.S. Booth and J.M. Carter. The critical constants of carbon dioxide-oxygen mixtures. J.
Phys. Chem. 34, (1930), 2801-2825.
10. G.H. Zenner and L.I. Dana. Liquid-vapor equilibrium compositions of carbon dioxide-
oxygen-nitrogen mixtures. Chem. Eng. Progr., Symp. Ser., No.44, 59, (1963), 36-41.
11. E.J. Partington, J.S. Rowlinson and J.F. Weston. The gas-liquid critical temperatures of
binary mixtures. Trans. Faraday Soc. 56, (1960), 479-485.
12. C.P. Hicks and C.L. Young. Critical temperatures of mixtures of quasi-spherical molecules.
Trans. Faraday Soc. 67, (1971), 1605-1611.
13. A.J. Davenport and J.S. Rowlinson. The solubility of hydrocarbons in liquid methane. Trans.
Faraday Soc. 59, (1963), 78-84.
14. J. Specovius, M.A. Leiva, R.L. Scott and C.M. Knobler. Tricritical phenomena in "quasi-
binary" mixtures of hydrocarbons. 2. Binary ethane systems. J. Phys.Chem. 35, (1981),
2313-2316.
15. C.J. Peters, H.J. van der Kooi, J.L. de Roo, J. de Swaan Arons, J.S. Gallagher and J.M.H.
Levelt Sengers. The search for tricriticality in binary mixtures of near-critical propane and
normal paraffins. Fluid Phase Equilibria 51, (1989), 339-351.
16. S.C. Pak and W.B. Kay. The critical properties of binary hydrocarbon systems. Ind. Eng.
Chem. Fundam. 11, (1972), 255-267.
17. A. Kreglewski and W.B. Kay. The critical constants of conformal mixtures. J. Phys. Chem.
73, (1969), 3359-3366.
18. C.P. Hicks and C.L. Young. Critical temperatures of mixtures of quasi-spherical molecules.
Trans. Faraday Soc. 67, (1971), 1598-1604.
19. W.B. Kay and G.M. Ramboseck. Liquid-vapor equilibrium realtions in binary systems.
Propane-hydrogen sulfide systems. Ind. Chem. Chem. 45, (1953), 221-226.
20. W.B. Kay. P,T,x Diagrams in the critical region. Acetone-n-alkane systems. J. Phys. Chem.
68, (1964), 827-831.
21. Th.W. de Loos, W. Poot and J. de Swaan Arons. Vapour-liquid equilibria and critical
phenomena in methanol + n-alkane systems. Fluid Phase Equilibria 42, (1988), 209-227.
22. M.R. Moldover and J.S. Gallagher. Critical points of mixtures: an analogy with pure fluids.
AIChE J. 24, (1978), 267-278.
23. J.R. Noles, J.A. Zollweg and W.B. Streett. Vapor-liquid equilibrium in the associating
system dimethyl ether-sulfur dioxide, presented at the AIChE 1989 Annual Meeting, San
Francisco, November 9, 1989.
24. J.M. Skaates and W.B. Kay. The phase relations of binary systems that form azeotropes. N-
alkyl-alcohol-benzene systems: methanol through n-butanol. Chem. Eng. Sci. 19, (1964),
431-444.
25. W.J. Gaw and F.L. Swinton. Occurence of double azeotrope in the binary system
hexafluorobenzene + benzene. Nature 212, (1966), 283-284.
26. W.J. Gaw and F.L. Swinton. Thermodynamic properties of binary mixtures containing
hexafluorobenzene. Trans. Faraday Soc. 64, (1968), 2023-2034.
27. A.P. Schmidt and H. Schwarz. Phase equilibrium measurements in the binary systems of
2,6-dimethylpyridine and 2-, 3- or 4-methylpyridine and of hexafluorobenzene and benzene.
Annual meeting AIChE, Washington D.C., paper 24t, Nov. 27 - Dec. 2, 1988.
26 Chapter 2

28. G.M. Schneider. Druckeinfluß auf die Entmischung flüssiger Systeme. II. Löslichkeit von
H2O und D2O in Methylpyridinen und Methylpiperidinen. Z. Phys. Chem. NF. 39, (1963),
187-197.
29. G.M. Schneider. Druckeinfluß auf die Entmischung flüssiger Systeme. IV. Entmischung
flüssiger n-Alkan-CO2-Systeme bis -60°C und 1500 bar. Messungen zum Problem der sog.
"Entmischung in der Gasphase". Ber. Bunsenges. Phys. Chem. 70, (1966), 10-16.
30. K. Roth, G.M. Schneider and E.U. Franck. Phasengleichgewichte flüssig-flüssig und flüssig-
fest in den Systemen Cyclohexan-Methanol und Phenol-Wasser bis 6000 bar. Ber. Bunsen-
ges. Phys. Chem. 70, (1966), 5-10.
31. A. Zawissa. High-pressure liquid-vapour equilibria, critical state, and p(Vm,T,x) to 448.15 K
and 4.053 MPa for {x C6H14 + (1-x) CH3OH}. J. Chem. Thermodynamics 17, (1985), 941-
947.
32. E. Brunner. Fluid mixtures at high pressures. II. Phase separation and critical phenomena of
(ethane + an n-alkanol) and of (ethene + methanol) and of (propane + methanol). J. Chem.
Thermodynamics 17, (1985), 871-885.
33. Z. Alwani and G.M. Schneider. Fluid mixtures at high pressure. Phase separation and critical
phenomena in binary mixtures of a polar component with supercritical carbon dioxide,
ethane and ethene up to 1000 bar. Ber. Bunsenges. Phys. Chem. 80, (1976), 1310-1315.
34. J. de Swaan Arons and G.A.M. Diepen. Gas-gas equilibria. J. Chem. Phys. 44, (1966), 2322-
2330.
35. J.A. Schouten. Phase equilibria in binary systems at very high-pressures. Physics Reports
(Review Section of Physics Letters) 172, (1989), 33-92.
36. E. Brunner. Fluid mixtures at high pressures. IX. Phase separation and critical phenomena in
23 (n-alkane + water) mixtures. J. Chem. Thermodynamics 22, (1990), 335-353.
37. A.J. Davenport, J.S. Rowlinson and G. Saville. Solution of three hydrocarbons in liquid
methane. Trans. Faraday Soc. 62, (1966), 322-327.
38. J. van der Steen, Th.W. de Loos and J. de Swaan Arons. The volumetric analysis and
prediction of liquid-liquid-vapor equilibria in certain carbon dioxide + n-alkane systems.
Fluid Phase Equilibria 51, (1989), 353-367.
39. R.J.J. Chen, P.S. Chappelaer and R. Kobayashi. Dewpoint loci for methane-n-hexane and
methane-n-heptane binary systems. J. Chem. Engng. Data 21, (1976), 213-219.
40. Y.-N. Lin, R.J.J. Chen, P.S. Chappelaer and R. Kobayashi. Vapor-liquid equilibrium of the
methane-n-hexane system at low temperatures. J. Chem. Engng. Data 22, (1977), 402-408.
41. G.M. Schneider. Druckeinfluß auf die Entmischung flüssiger Systeme. I. Geschlossene
Mischungslücken bis 5000 bar. Z. Phys. Chem. NF. 37, (1963), 333-352.
27

CHAPTER 3

THE SIMPLIFIED PERTURBED HARD CHAIN THEORY: DEVELOPMENT

The choice of the equation of state is of major importance to our investigations. The
equation of state resulting from the Simplified-Perturbed-Hard-Chain-Theory (SPHCT) was
chosen for the classification of critical curves in binary fluid mixtures. The SPHCT equation of
state is a member of a family of equations of state that started with the Perturbed-Hard-Chain
equation of state (PHCT). We will start this chapter with a general overview of equations of state
in section 3.1. In section 3.2 the augmented hard sphere equations of state are discussed. Special
attention is paid to the origin and development of the PHCT. The derivation of the SPHCT
partition function is discussed in section 3.3. A more detailed analysis of the repulsive part, the
attractive part and the mixing rules of the SPHCT is given in the sections 3.4, 3.5 and 3.6.

3.1 Introduction

To date, no investigator has reported a perfectly general equation of state that can be
applied to a wide variety of mixtures, whose components differ largely in polarity and size.
Instead, specialized equations have been developed that can be applied to limited ranges of
mixtures, e.g., natural gases. The main problem in the development of what might be called the
perfect equation of state can be found in our limited knowledge of the interactions of chemically
dissimilar molecules. Nevertheless equation of state methods for the calculation of phase
equilibria in non-polar and polar mixtures have developed rapidly during the last twenty years.
An extensive overview of equations of state has been recently given by Anderko [1]. Generally
speaking, equations may be classified as being a member of one of four possible families [2]:

i Van der Waals family


ii Benedict-Webb-Rubin family
iii Reference fluid equations of state
iv Augmented hard body equations of state

A short introduction to these four classes of equations of state is given in the following
sections.
28 Chapter 3

3.1.1 The Van der Waals type equations of state

Equations of state within this family are probably the most widely used for engineering
calculations. This popularity arises from the simplicity of the equations which enables them to be
expressed in terms of dimensionless variables. Generally these equations can be written as a
polynomial of degree 3 in the compressibility ratio. Normally they are presented as:

RT a ( T, ω )
p= - (3.1)
V - b f ( V, T, b,... )

where f represents a quadratic function of the molar volume, V, the temperature, T, the covolume
parameter, b. The f-function sometimes contains additional parameters. The most important
variants of Eq.(3.1) are summarized in Table 3.1. These equations are typically written for pure
fluids and then extended to mixtures. The members of this family most successful in correlating
phase equilibria, (Soave-Redlich-Kwong, Peng-Robinson), adjust the temperature dependence of
the a-parameter so as to provide good agreement with pure component vapour pressures.

Table 3.1: Summary of common Van der Waals type equations of state.

Equation f(V,T,b,...) Ref.

Van der Waals V2 [3]

Redlich-Kwong T 1/2.V(V+b) [4]

Soave V(V+b) [5]

Peng-Robinson V(V+b) - b(V-b) [6]

Fuller V(V+cb) [7]

Schmidt-Wenzel V 2 + ubV + wB2 [8]

Harmens-Knapp V 2 + Vcb - (c-1)b2 [9]

Patel-Teja V(V+b) + c(V-b) [10]

3.1.2 The Benedict-Webb-Rubin type equations of state

The second class of equations of state that are commonly used in engineering phase
equilibria are those of the Benedict-Webb-Rubin (BWR) family. The original BWR equation
[11] contains eight adjustable constants and is given by:
Development SPHCT 29

C0
p = ρ R T + ( B0 R T - A0 - 2
) ρ2 +( b R T - a ) ρ3 +
T (3.2)

ρ3
a α ρ6 +( c ) ( 1 + γ ρ 2 ) e-γ ρ
2

2
T

Later, extra terms were added in the factors in front of the ρ2-, ρ3- and ρ6-terms. In order to use
Eq.(3.2) for mixtures, mixing rules are used. In general they are given by:
r
⎛ n


1
w = ⎜⎜ xi w ⎟⎟
ir (3.3)
⎝ i=1 ⎠

Values of r for each coefficient are given in Table 3.2.


Table 3.2: The exponents of the
It should be noted that, unlike the mixing rules for the
BWR mixing rules.
Van der Waals equation, these mixing rules have no
theoretical basis. This type of equation offers improved
Constant (w) r
thermodynamic property correlations but does not offer
any better results for the correlation of mixture phase A0 2
behaviour. The primary problem is the large number of
B0
B 1
coefficients for every component in the mixture and the
lack of any theoretical guidelines. Another drawback of C0 2
this type of equation is the large number of volume a 3
roots that can be obtained from this equation. This is
due to the high degree of the molar volume in these b 3
type of equations of state. c 3

α 3

γ 2

3.1.3 The reference fluid equations of state

Complex equations of state have been developed to represent highly precise (and
accurate) p,V,T-data and thermodynamic property data for a certain fluid, without regard to
mathematical simplicity or generalization to other fluids. Examples of this class of equations is
the nonanalytic equation derived by Goodwin [12] and a modified BWR equation [13]. In order
to apply these type of equations to other pure fluids or mixtures, corresponding states principles
must be used. Corresponding states principles are based on the assumption that for non-polar or
slightly polar substances the compressibility ratio, Z, and other thermodynamic properties can be
expressed as a linear function of the acentric factor, ω, at a specified reduced temperature, Tr
(=T/Tc), and pressure, pr (=p/pc):
30 Chapter 3

Z = Z ( 0 ) ( p r , T r ) + ω Z (1) ( p r , T r ) (3.4)

Z(0) represents the compressibility ratio of a "simple" fluid, e.g., argon, and Z(1) is the deviation of
the compressibility ratio of the real fluid from Z(0). Lee and Kesler [13] have tabulated Z(0) and
Z(1) for 0.3<Tr<4.0 and 0<pr<10.0. They used a modified BWR equation of state, see section
3.1.2, to obtain values from p,V,T-data, saturated vapour pressure data and the critical points of
Ar, Kr and CH4 for the simple fluid, Z(0)(pr,Tr), and the same properties of n-octane for the
reference fluid, Z(1)(pr,Tr). For Ar, Kr and CH4, Z=Z(0) and for n-octane Z=Z(ref). For the reference
fluid it follows that:

Z
( ref )
= Z ( 0 ) + ω ref Z ( 1 ) (3.5)

or:
( ref )
- Z( 0 )
Z
(1)
=Z (3.6)
ω ref
For a pure fluid it follows that:

ω
Z = Z ( 0 ) + [ Z ( ref ) - Z ( 0 )] (3.7)
ω ref
Both the compressibility ratio of the simple fluid, Z(0), and the reference fluid, Z(1), can be
calculated from a reduced and modified BWR equation of state [13]. If pc, Tc and ω of a
substance are known, this method can be used to calculate the molar volume (or compressibility
ratio) at a specified temperature and pressure. In order to extend this method to mixtures mixing
and combining rules must be introduced for the pseudo critical temperature and volume, Tc and
Vc. Usually they are chosen in accordance with the Van der Waals one-fluid theory, although
other choices are possible.

3.1.4 Augmented hard body equations of state

The fourth category of equations of state is similar in some respects to the Van der Waals
family, but is set apart because of the abandonment of the Van der Waals repulsion term
RT/(V-b). These equations start with theoretically based hard body equations of state and add
terms to account for the effect of molecular attraction. The hard body terms are the Wertheim-
Thiele equation [14,15], the Carnahan-Starling equation of state for hard spheres [16,17] and the
equations of Gibbons [18], Boublik and coworkers [19-21] or Nezbeda and Leland [22] for rigid
non-spherical bodies. Examples of this class of equations are the Perturbed Hard Chain Theory
[23-26] and the augmented Van der Waals theory [27-30]. These two models have been applied
Development SPHCT 31

to polar/non-polar systems with reasonable success. Generally speaking, this family of equations
of state is in a developmental stage, as will be shown in the next section. Because of their
mathematical complexity, these equations have not yet found widespread industrial use. We
believe, however, that these kinds of equations provide us with the most promising route to the
calculation of phase equilibria in polar and non-polar systems.

3.2 The Perturbed Hard Chain Theory and its family members

Using simple models for a liquid, and recognizing that some rotational and vibrational
motions depend on density, Prigogine [31] and Flory [32] developed analytical equations of state
for polymeric molecules. Unfortunately, the equations developed by Flory and Prigogine did not
reduce to the ideal gas equation, and hence were limited to calculations of liquid phase
properties. The Perturbed Hard Chain Theory (PHCT) developed by Beret and Prausnitz [23,24]
and Donohue and Prausnitz [25,26] is applicable for both liquid and vapour phase properties of
fluids ranging in molecular size from methane to large hydrocarbons and polymers, see Fig.3.1.
The PHCT has been applied succesfully to calculate thermodynamic properties of numerous and
varied systems of industrial interest.
Beret and Prausnitz [23,24], who
proposed an approximation for the density-
dependent rotational and vibrational
motions, applied the PHCT to pure fluids
including polymers. Donohue and Prausnitz
[25,26] changed the constants in the
attractive term of the PHCT to fit more
accurately the properties of the alkanes.
More important, they also extended the
PHCT to mixtures by using mixing rules
derived from perturbation theory and
showed that the mixture distribution
coefficients and heat of mixing for several
non-polar fluid mixtures can be calculated
accurately with low values of a binary
interaction coefficient (usually less than
Figure 3.1: PHCT provides an equation of
0.02). A binary interaction coefficient is an state for fluids containing simple
adjustable parameter that can be used to or complex molecules, covering
give an optimal description of experi- all fluid densities
mentally determined phase behaviour.
Kaul et al. [33] calculated Henry coefficients using the PHCT with similar low values of binary
interaction coefficients. They extended the PHCT to calculate the second virial coefficient of
both pure fluids and mixtures. Liu and Prausnitz [34] showed that the PHCT can also be used to
32 Chapter 3

accurately calculate the solubilities of gases in liquid polymers when the light component is
supercritical. Liu and Prausnitz [35,36] also applied the PHCT for phase equilibrium calculations
in polymeric systems (polymer-solvent, polymer-polymer, polymer-polymer-solvent) taking into
account the molecular weight distribution of polymeric molecules. Ohzono et al. [37] correlated
the pure component parameters reported by Donohue and Prausnitz [26] and Kaul et al. [33]
succesfully with the group volumes of Bondi [38].
The PHCT equation of state is succesful but mathematically complex. Consequently,
computer calculations, especially for mixtures are time consuming and costly. Since the PHCT
has already proved to be valuable in calculating properties of various types of systems, it is
useful to consider modifications of the theory which simplify the equation of state. Such
simplifications would make the PHCT equation of state more useful to the engineering
community. Gmehling et al. [39] truncated the perturbation expansion in the attractive term of
the PHCT equation of state after the second order term. The Truncated Perturbed Hard Chain
equation of state (TPHCT) has ten universal constants instead of the twenty-one universal
constants in the four expansion terms of the PHCT. The results for pure fluids were still accurate,
but are not available yet for non-polar mixtures. Wilhelm and Prausnitz [40] used the TPHCT to
calculate properties of narrow-boiling fractions of crude oil. The vapour pressures and saturated
liquid volumes were calculated with reasonable accuracy.
Kim et al. [41] replaced the attractive term of the PHCT equation of state with a
theoretical but simple expression based on the local composition model of Lee et al. [42]. The
resulting equation of state, the Simplified Perturbed Hard Chain equation of state, (SPHCT),
reproduces both experimental vapour pressures and liquid-density data for a number of fluids
with good accuracy. Calculations for mixtures require relatively simple mixing rules. The
SPHCT equation of state was improved by Vimalchand et al. [43] through implementation of an
approximation for the hard-sphere fluid structure.
Vimalchand and Donohue [44] and Morris et al. [45] modified the PHCT by replacing
the square-well potential with a soft-core Lennard-Jones potential energy function. In their
Perturbed Soft Chain Theory (PSCT) the molecular attractions are accounted for by a
perturbation expansion. This softening of the repulsive part of the intermolecular potential is
important for a correct description of the thermodynamic properties of fluids at high
temperatures. In the PSCT small values of the binary interaction coefficient are sufficient to give
accurate fits for mixtures of non-polar fluids.
All equations of state that were previously mentioned do not take into account the
increase in attractions due to dipolar and quadrupolar forces. Fairly accurate multipolar mixture
calculations, without a binary interaction coefficient, were obtained using the Perturbed
Anisotropic Chain Theory (PACT) by Vimalchand and Donohue [44,46]. In the PACT
anisotropic multipolar interactions were explicitly combined with the PHCT equation of state by
an additional perturbation expansion. Cotterman et al. [47,48] developed an equation of state
based on PHCT ideas which also is applicable to multipolar fluids.
Development SPHCT 33

The Chain of Rotators Theory (COR), derived by Chien et al. [49], is also related to the
PHCT. In this equation of state, a chain molecule is not composed of spherical segments but is
seen as a dumbbell-shaped molecule. Chien et al. [49] obtained an accurate correlation of pure
fluid properties using the COR Theory. However calculations for mixtures generally required
large values of binary interaction coefficients. The Deiters equation of state [50-52] is similar to
the PHCT and the COR. This equation of state contains corrections for non-spherical molecular
shape, softness of the repulsive potential and three-body effects. The Deiters equations of state
calculates critical compressibilty ratios that lie in the range between 0.23 and 0.292, which is in
agreement with values for real systems.
Ikonomou and Donohue [53] extended the PACT to obtain an equation of state which
takes into account the existence of hydrogen bonding. This new theory, the Associated Perturbed
Anisotropic Chain Theory (APACT) is a generalization of the PACT to include hydrogen
bonding compounds and mixtures. Although the APACT gives good fits for hydrogen bonding
systems it is lengthy and cumbersome. This equation of state has been further extended to cover
Lewis acid-base mixtures [54-56]. Ikonomou and Donohue [57], starting from the SPHCT,
derived a simpler equation of state for hydrogen-bonding systems called COMPACT. Even
though significant improvement was achieved in the calculation of phase equilibria for several
hydrogen bonding systems, mixtures that display highly complex phase behaviour, like n-butane
+ water, require a more complicated equation of state, such as the APACT. Ciocca et al. [58]
made Prigogine's c-parameter in the SPHCT, see section 3.3, density-dependent, which led to the
Augmented Simplified Perturbed Hard Chain Theory (ASPHCT). The ASPHCT was later
extended to polar fluids through the addition of a perturbation term which takes into account the
polarity of the molecules. This new equation of state of Ciocca and Nagata [59] was called the
Simplified Perturbed Anisotropic Chain Theory (SPACT).
Jin et al. [60] have reformulated the non-polar interactions in terms of group-group
interactions. This led to the Group Perturbed Soft Chain Theory (GPSCT). Using the group
parameters and the GPSCT equation of state, calculated results show good agreement with
experimental data.
Lu and Sheng [61] used the PHCT with quantum corrections to calculate high pressure,
low temperature vapour-liquid equilibria of systems containing hydrogen. Accurate correlation
of phase equilibria for many hydrogen containing systems was also obtained.
The "next-generation" of equations of state to be derived from the PHCT, will probably
be a combination of the ideas in the PHCT and those of the Generalized Flory-Dimer equation of
state (GFD), developed recently by Hall and coworkers [62-64]. The major focus of the research
on the GFD equation of state has been the development of a new equation for the heavier
constituents of natural gas (alkanes) which rigorously takes into account the chain-like nature
and flexibility of these molecules. Basically, the GFD-equation of state can be rearranged into a
form that is similar to the PHCT [65]. The GFD was further simplified to the so-called
Simplified Generalized Flory-Dimer (SGFD) equation of state by the introduction of a simple
density dependence for some of the pure component parameters.
34 Chapter 3

In summary, the PHCT and the theories derived from it [66,67], have been succesfully
used in a variety of applications which include highly non-ideal phase behaviour due to strongly
polar and hydrogen bonding fluids, water - hydrocarbon systems, crude oil fractions, hydrogen
solubility in coal-derived liquid, solubility of light gases and volatile solutes in polymers, low
temperature quantum effects in systems containing hydrogen, solid solubility in supercritical
fluids and entrainer effects. The most simple member of the family of equations of state based on
the PHCT, is the SPHCT equation of state. For our investigations the SPHCT equation of state
will be used for a number of reasons:

- The SPHCT equation has a statistical mechanical background.


- The SPHCT equation is suitable for chainlike molecules.
- The SPHCT equation has a relatively simple mathematical form.
- The SPHCT equation correlates binary phase behaviour reasonably well.
- The SPHCT equation has received considerable attention from various research groups
[68-70].

3.3 The partition function

In this section we will develop the SPHCT from statistical mechanics. A general equation
for the canonical partition function of a system with N molecules can be found in many
textbooks, e.g., Findenegg [71]:

- Η ( N,V)
1 3r 3r 3 r 3 r (3.8)
Z ( N,V, T ) = ∫ .. ∫ e kT
d r 1 .. d r N d p1 .. d p N
N! h3N

where Z represents the canonical partition function of the system, also called the phase integral.
According to the official notation [72] we may call the canonical partition function either Q or Z,
and we may call the configuration integral, a property we will meet later in this section, either Z
or Q. To avoid confusion about the nomenclature, it is stressed here that we have adopted the
symbol Z for canonical partition function (German: Zustandssumme) and Q for the configuration
integral. In Eq.(3.8) d31 = dxi.dyi.dzi is a differential volume element around the position ri of the
molecule i, d3i = dpx,i.dpy,i.dpz,i is a differential change of the impulse coordinates. The
integration is over 3N position coordinates and 3N impulse coordinates. The total energy of the N
molecules depends on the position and impulse vectors. The total energy function is called the
Hamiltonian. The Hamiltonian, H, is composed of the kinetic energy of the N molecules and the
potential energy of the interactions of the N molecules, the so called configurational energy UN:

1 N r r r
Η= ∑
2m i = 1
( p 2x,i + p 2y,i + p 2z,i ) + U N ( r 1 , r 2 , .. , r N ) (3.9)
Development SPHCT 35

The integration over the impulses in Eq.(3.9) can be performed directly:


3N r r r
1 ⎡ ⎤
+∞ -p 2 -U N ( r 1, r 2 , .. ,r N )
r r
N! h3N ⎣ -∫∞
Z ( N,V, T ) = dp ⎥ ∫ ∫
. .. e .d 3 r 1 .. d 3 r N
2mkT kT
⎢ e (3.10)
⎦ V V

As will be shown later [73] there is an extra contribution to the kinetic energy for chainlike
molecules by rotations and vibrations. For the moment we will continue with an expression for
the partition function of N spherical molecules. Elaboration of Eq.(3.10) leads to:
3
N
⎛ 2 π m k T ⎞2 QN
Z ( N,V, T ) = ⎜ 2
⎟ . Q N = 3N (3.11)
⎝ h ⎠ Λ

with:
h
Λ= 1 (3.12)
( 2 π m k T )2

r r r
UN - ( r 1, r 2 , .. ,r N )
1 3r 3r

N! V∫ V∫
Q N ( N,V, T ) = . .. e kT
d r 1 .. d r N (3.13)

The integral QN includes all possible configurations of the system of N molecules and is called
the configuration integral. The property Λ is the De Broglie wavelength.
The link between the canonical partition function of a system of N molecules and the
macroscopic Helmholtz energy is given by the following equation [71,73]:

A = - k T ln Z N (3.14)

in which A represents the molar Helmholtz energy. All pressure explicit equations of state can be
derived when one has knowledge about the canonical partition function of the system of N
molecules:

⎛∂ A⎞ ⎛ ∂ ln Z N ⎞
p ( N,V, T ) = - ⎜ ⎟ =kT ⎜ ⎟ (3.15)
⎝ ∂ V ⎠T ⎝ ∂ V ⎠T

where p represents the pressure.


It is shown by Eqs.(3.14) and (3.15) that an expression for ZN forms the key to any
equation of state. We will now derive an expression for ZN that will lead us to the SPHCT. We
will briefly digress on intermolecular interactions. The intermolecular interactions (Van der
Waals interactions) are based on induced dipole-induced dipole or disperse interactions (London
interactions). For polar molecules the dipole-dipole interactions (Keesom interactions) and the
interactions between dipoles and induced dipoles (Debye interactions) play an important role.
36 Chapter 3

The interaction energy between two point molecules i and j depends only on the distance
between those molecules: rij = │ri-rj│. This interaction energy is called the pair potential u(rij).
Directive forces between molecules are not taken into account. Van der Waals [3] proposed in
his thesis two approximations that have generally been accepted.

i The pair potential u(rij) is separated into attractive and repulsive contributions. The attractive
potential [uattr(rij) < 0] depends approximately on the sixth power of the distance rij. At very
small distances the repulsive potential [urep(rij)>0] dominates because of the penetration of
the electron shells of the different molecules.
ii For the attractive interactions the Mean Field Approximation (MFA) is assumed. The MFA
states that the attractive interaction ui between a molecule i and all other molecules of the
system can be replaced by a mean interaction φ. This mean interaction is proportional with
the density N/V, but independent of the exact positions of the molecules.

The intermolecular interactions are assumed to be pairwise additive. This means that the
interaction energy of a system of three molecules i, j and k is equal to the sum of the three pair
interactions. Deiters [50-52] added in his equation of state an approximation to include three-
body interactions as well. These three-body interactions are very important for the behaviour of
fluids at high densities. In the SPHCT three-body interactions are ignored. For a system of N
molecules the pairwise additivity assumption together with the MFA leads to:

r r r 1 N N 1
U N ( r 1 , r 2 , .., r N ) = u ( r 12 ) + u ( r 13 ) + ..+ u ( r N -1 , N ) = ∑ ∑ u ( r ij ) = N φ (3.16)
2 i=1 j=1 2
j≠i

where φ is the mean potential that all molecules j exert on a central molecule i. The factor 1/2 is
required in Eq.(3.16) to avoid counting each interaction twice, e.g., u(r12) ≡ u(r21). The
configuration integral in Eq.(3.13) can now be written as follows:
⎛ -N φ ⎞
1 ⎜⎝ 2 k T ⎟⎠ r r
QN = e . ∫ .. ∫ d 3 r 1 .. d 3 r N (3.17)
N! V V

The integral in Eq.(3.17) can be replaced by the free volume to the N-th power, which leads to
the generalized Van der Waals partition function:
N -Nφ
1 V f 2kT
Z ( N,V, T) = 3N
. .e (3.18)
Λ N!

Prigogine [31] took the energetic effects of vibrations and rotations of the individual molecules
into account as well. The N molecules cannot be seen as spherical molecules anymore, but rather
as chainlike molecules. Prigogine [31] proposed the following partition function:
Development SPHCT 37

N N
N ⎛ V f ⎞
-Nφ
1 ⎛V ⎞
⎜ 3 ⎟ . ( qr .qv ) . ⎜⎜ ⎟⎟ . e
2kT
Z ( N,V, T) = (3.19)
N!⎝Λ ⎠ ⎝ V ⎠

An artificial separation is proposed in the vibrational and rotational partition function: a part of
the rotations and vibrations is internal while the rest is external. The internal vibrations and
rotations are supposed to be a function of temperature only, the external vibrations and rotations
are a function of both temperature and density. It is assumed [31] that the partition function of
the rotations and vibrations can be factorized according to Eq.(3.20):

( q r . q v ) = ( q r . q v )int . ( q r . q v )ext (3.20)

Donohue and Prausnitz [25,26] assumed the contribution of the external rotation and vibration
partition function to be one-third of the translational partition function:
( 3c-3 )
⎡ V f 2ckT ⎤

3
( q r .qv )ext = ⎢ .e ⎥ (3.21)
⎣ V ⎦

The partition function according to Donohue and Prausnitz [25,26] is:


N Nc
1 ⎛ V ⎞ ⎡V f ⎤ N
Z ( N,V, T ) = .⎜ ⎟ .( q r . qv )int (3.22)
N ! ⎝ Λ3 ⎠ ⎢⎣ V ⎥⎦

where there are 3c degrees of freedom for each molecule that contributes to the configuration
integral (that is c=1 for spherical molecules considered previously, and c>1 for chainlike
molecules).
We have now obtained an expression for the canonical partition function for a system of N
(nonspherical) molecules. However, from this expression we cannot yet derive an equation of
state as we need information about the density dependence of Vf and φ. For the temperature and
density dependence of φ we will analyse the spatial structure of the system of N molecules. The
average interaction energy φ that a certain molecule i encounters, is a function of the positions of
all the other N-1 molecules in the system. Radial distribution functions contain information about
the positions of all other molecules. The only radial distribution function that is under
consideration here is the pair distribution function g(rij). This g(rij) is a correction factor to
calculate the local density from the bulk density (N/V) at a specified distance rij from a central
molecule i. Often this g(rij) is seen as the probability of the presence of a molecule j in a spherical
shell with infinitesimal thickness dr and volume 4πr2dr around a central molecule i. Strictly
speaking, this is not correct because radial distribution function can have values higher than 1
and probabilities cannot.
Except at low densities, the pair distribution function strongly depends on density and
weakly on temperature [73]. It is the temperature and density dependence of the pair distribution
function that should give us the temperature and density dependence of the mean potential. If the
38 Chapter 3

summation over all molecules in Eq.(3.16) is replaced by the integral of u(r)g(r) over all possible
distances one obtains the following expression for the total interaction energy:
2 ∞
N N
UN= . ∫ g(r, ,T) u(r) 4π r 2 dr (3.23)
2V 0 V

According to Sandler [73], the mean potential φ is related to the total interaction energy by the
following equation:

1 ⎡ ⎤
T T ∞
-2kT U N dT = - N k T N
φ= ∫
N T = ∞ kT 2 ∫ 2 ⎢ ∫
V T = ∞ kT ⎣ r = 0
u(r) . g(r, ,T ) 4π r 2 .dr ⎥ dT
V ⎦
(3.24)

In summary the problem of finding an expression for the canonical partition function ZN of a
system with N molecules as in Eq.(3.22) can be split up into two parts. We need to find
expressions in density and temperature for:
i the free volume Vf
ii the mean potential function φ

3.4 The free volume

To derive an accurate expression for the free volume it is more convenient to use the pair
correlation functions than the pair distribution function. In analogy with the radial distribution
functions, the correlation functions can be defined for pairs, triplets etc. From now on only pair
correlation functions and pair distribution functions are considered. The pair correlation
functions can be divided into the total correlation function h(r12) and the direct correlation
function c(r12). Like the pair distribution function, both correlation functions depend on density
and temperature. In this section we take into account only the dependence of the pair distribution
functions and pair correlation functions on the separation distance. The notation for the
distribution function will be g(r) instead of g(r,N/V,T). The total correlation function is defined
as:

h ( r 12 ) = g ( r 12 ) - 1 (3.25)

The total correlation function is a scaled (or shifted) pair distribution function. Schematic
drawings of g(r), h(r) and c(r) are shown in Figs.(3.2)-(3.4). The so-called direct correlation
function was first introduced by Ornstein and Zernike [74]. According to them, the total
correlation between two molecules 1 and 2 can be separated into two parts.
Development SPHCT 39

Figure 3.2: A schematic drawing of the Figure 3.3: A schematic drawing of the
pair distribution function g total correlation function h
as a function of the reduced as a function of the reduced
distance r/σ. distance r/σ.

i A direct effect of molecule 1 on molecule


2, characterized by c(r12). This direct
effect is only noticeable on short distances
from molecule 1. The direct effect has the
same range as the intermolecular
potential.
ii An indirect effect of molecule 1 on 2 via
molecules 3, 4, .... This indirect effect has
a longer range than the direct effect.

The short range effect of c(r12) is clearly shown in


Fig.(3.4). The indirect effect is in principle the
sum over all contributions of all other molecules.
The direct correlation function is related to the
total correlation function and is defined by the
Ornstein-Zernike relation [74]: Figure 3.4: A schematic drawing of the
direct correlation function c as
a function of the reduced dis-
tance r/σ.
h ( r 12 ) = c ( r 12 ) + ρ ∫ c ( r 13 ).h ( r 32 ) dr 3 (3.26)
40 Chapter 3

The first term on the right side of Eq.(3.26) is the direct part and the second term is the indirect
part of the total correlation function h(r12). The total correlation under the integral sign can be
replaced by all possible direct correlation functions
h (r12 ) = c (r12 ) + ρ ∫ c (r13 ).c (r32 ) dr3 + ρ 2 ∫ ∫ c (r13 ).c (r34 ).c (r42 ) dr3 dr4 + ..... (3.27)

The second term on the right hand side of


Eq.(3.26) is the indirect correlation between
molecule 1 and molecule 2 via molecule 3. A
graphical representation of the Ornstein-Zernike
relation is shown in Fig.3.5. In this figure (a)
represents the direct effect of molecule 1 on
molecule 2 and (b) and (c) represent the indirect
effect. Part of the indirect effect is caused by
molecule 3, see Fig.3.5(b). This indirect effect is
given by c(r13).c(r32) averaged over the position of
molecule 3. A second contribution to the indirect
effect is caused by molecules 3 and 4, averaged
over the positions of molecules 3 and 4, see
Fig.3.5(c). Eq.(3.27) cannot be solved because
both c(r) and h(r) are not known. For c(r) a
number of expressions have been developed that
close the Ornstein-Zernike relation. These expres-
sions are called 'closures' and can be divided into Figure 3.5: Direct (a) and indirect (b,c)
two groups [75]: the theoretical closures and the effects of molecule 1 on
semi-empirical closures. The Yvon-Green-Born molecule 2.
(YGB) solution and the Kirkwood solutions belong to the first group. The Percus-Yevick (PY)
Approximation, the Hypernetted Chain (HNC) Approximation and the Mean Spherical
Approximation (MSA) belong to the second group.

NkT 1 N 2 ⎛∂u⎞
2 ∫
p= - r⎜ ⎟ g(r) 4 π r 2 dr (3.28)
V 6 V r = 0 ⎝ ∂r ⎠

To test the consistency of these approximations, two expressions can be derived that relate
the pair distribution function with (macroscopic) thermodynamical properties, see Baxter [75]:

NkT N
κ T = 1 + ∫ (g(r) - 1) 4π r 2 dr (3.29)
V V r=0

where κT is the isothermal compressibility, κT =(1/ρ).(∂ ρ/∂ p)T. Eq.(3.29) differs from Eq.(3.28)
in that it is not restricted to systems with pairwise-additive intermolecular potentials. For a
system of hard spheres the pair-wise additivity assumption is strictly valid.
Development SPHCT 41

If g(r) is known exactly, both equations must give identical results for a system of hard spheres.
However, because g(r) is known only approximately, different answers may be obtained from
Eq.(3.28) and Eq.(3.29). The 'compressibility pressure' obtained by integrating Eq.(3.29) with
respect to r may differ from the 'virial pressure' given by Eq.(3.28).
The pair distribution function is exactly known at zero density [76]:
-u(r) (3.30)
lim g(r) = e kT
N
→0
V

For higher densities Percus and Yevick [77] assumed:

⎛ -u(r) ⎞ ⎛ u(r) ⎞ (3.31)


c(r) = ⎜⎜ e kT - 1 ⎟⎟.( h(r) + 1 ).⎜⎜ e kT ⎟

⎝ ⎠ ⎝ ⎠

The PY-approximation relates the radial distribution function g(r) to the direct correlation
function c(r). With h(r)=g(r)-1 this leads to:

⎛ u(r)

c (r)= ⎜⎜ 1 - e kT ⎟⎟.g(r) (3.32)
⎝ ⎠
The Ornstein-Zernike relation, Eq.(3.26), can now be solved analytically for a system of hard
spheres, i.e., molecules with the following potential function: u(r)=∞ if r<σ and u(r)=0 if r≥σ.
However, it is not possible to obtain a solution for a more realistic intermolecular potential
function. The solution of Percus and Yevick [77] is valid only for a system of molecules that
either do not have an attractive potential or are in the situation where the attractive potential
plays no role, i.e., at infinite temperatures. The results from the pressure and compressibility
equations are respectively:

⎛ p ⎞ 1+ 2 η + 3η 2
⎜⎜ ⎟⎟ = (3.33)
⎝ ρ k T ⎠p ( 1 - η )2

⎛ p ⎞ 1 +η +η 2
⎜⎜ ⎟⎟ = (3.34)
⎝ ρ k T ⎠c ( 1 - η )
3

where η is the reduced density, i.e., the ratio of the true volume of the N molecules over the total
volume V: η=Nπσ3/6V. Because Eqs.(3.33) and (3.34) are almost the same but not identical the
PY-closure is thermodynamically inconsistent. An empirical but accurate equation for a system
composed of hard spheres, was obtained by Carnahan and Starling [16]. Their basic idea was to
develop a rounded Padé approximation to describe the virial coefficients of the dimensionless
virial equation of state. It turned out that their hard-sphere equation of state equals one-third of
Eq.(3.33) added to two-thirds of Eq.(3.34).
42 Chapter 3

p 1 +η +η 2 - η 3
=
ρ kT ( 1 - η )3 (3.35)

The Carnahan-Starling equation [16] describes


interactions between hard spheres that are dominated by
short range repulsive forces, see Fig.3.6. Both equations
suggest that the pressure becomes infinite only as η
approaches 1. This is a physically impossible high
density, since η cannot exceed the reduced density of a
hexagonal or cubic close-packed array of spheres,
namely τ=π√2/6=0.7405. In case of a non-regular
closest packing the reduced density is about 0.65 [71].
Real fluids at normal pressures have reduced densities
of about 0.5 [71]. The close-packed density limit can be
introduced in η by Eq.(3.36):

π N σ3
2.
πσ3 N 6 2 = (3.3.6) Figure 3.6: Schematic representation of
η = =
6 V V the hard-sphere potential
τ .V * function.
V
Eq.(3.36) shows that at closest packing η equals τ instead of 1. Another physically unrealistic
phenomenon is: if η > 0.8 then g(r) can be negative at some values of r [77]. Despite these
theoretical shortcomings the Carnahan-Starling expression is much better than the cubic
RT/(V-b)-term [77] and the Carnahan-Starling expression is used as the repulsive term in the
PHCT and SPHCT equations of state. The free volume term from the Carnahan-Starling
equation is given by [73]:

N -η ( 4 -3 η )
⎛ Vf⎞ ( 1 - η )2
⎜⎜ ⎟⎟ = e (3.37)
⎝ V ⎠

Because the molecules are not treated as spheres in the SPHCT but as chains, we do not define
the closest packed volume V * as is done in Eq.(3.36) but as V *=Nsσ3/√2, where σ represents the
diameter of a segment and s represents the number of segments per molecule.
The only remaining unknown in the canonical partition function for the system of N
molecules is the mean potential. An expression for the mean potential will be derived in the next
section.
Development SPHCT 43

3.5 The mean potential

The original derivation of the expression for the mean potential is given in [42,73]. We will
investigate the case in which the molecules interact with the square-well potential, see Fig.3.7:

⎧ ∞ for r < σ
⎪ (3.38)
u ( r )= ⎨ - ε for σ < r < R σ
⎪ 0 for R σ < r

To proceed, Eq.(3.23) is rewritten to:

2 ∞
N . g(r, N
2 V ∫0
UN = ,T) u(r) 4π r 2 dr =
V

2
N
= N ∫ (-ε ) . g ( r, V ,T ) 4π r 2 dr =
2V R′
(3.39)

- . N c ( ρ ,T )
2

In the second form of the Eq.(3.39) it should be Figure 3.7: Schematic representation of
noted that only the attractive well contributes to the the square-well potential
mean potential. The region σ<r<Rσ in which the function.
square-well is important, that is where u(r)=-ε, has
been denoted by R'. One should also recognize that the number of molecules within the range of
the potential of a central molecule, denoted as Nc(ρ,T) is given by:

N
N c ( ρ ,T )=
V ∫ g ( r, ρ , T) 4π r
R′
2
dr (3.40)

where Nc is an occupation number. Clearly Nc(ρ,T) is a function of temperature and density


because of the dependence of g(r,ρ,T) on these properties. Combining Eqs.(3.24) and (3.39) leads
to [42,73]:

N c ( ρ ,T )
T
φ=kT ε
T =∞
∫ k T2
dT (3.41)

To obtain a closed form of the partition function for the square-well fluid an expression for the
occupation number as a function of temperature and density is needed. Therefore we consider a
lattice in which each site has a coordination number Zm. A binary mixture of molecules will be
distributed on this lattice. One of the species will later be replaced by empty sites, so that the
44 Chapter 3

composition dependence of the binary mixture will become the density dependence of the pure
fluid. We now assume that, because of the attractive interactions, the two molecular species 1
and 2 are not randomly distributed about a central molecule, but rather follow the local composi-
tion model:

N 21 N 2 ε 21 - ε 11 N 2 - ε
= .e 2kT = . e2 k T (3.42)
N 11 N 1 N1

where Nji is the number of molecules j on lattice sites within the coordination sphere of a central
molecule i, and εij is the square-well depth for an i-j interaction. Also, anticipating that species 2
will be vacancies, we have set ε21=0 and ε=ε11. The factor 2 in the argument of the exponential
function in Eq.(3.42) is arbitrary. In a later version of the SPHCT [43], Eq.(3.42) has been
replaced by a more complicated expression. Now using the lattice requirement, that:

N 11 + N 21 = Z m
(3.43)

yields:

Z m N1
N 11 = ε (3.44)
-
N1+ N 2 .e 2 kT

where Zm is the total lattice-site coordination number for a molecule 1. To proceed we now
recognize that species 2 represents lattice vacancies so that N1 and N2 are related to the fluid
volume by:
*
N1 V
= (3.45)
N1+ N 2 V

where V *=Nσ3/√2 is the close-packed volume. Elimination of N2 from Eq.(3.44) by Eq.(3.45)


and observing that N11 now is the coordination number Nc, gives:
ε
*
Z m V e2 k T
N 11 = N c = ε
(3.46)
V + V * ⎛⎜ e 2 k T - 1 ⎞⎟
⎝ ⎠

For the R=1.5 square-well fluid, the lattice coordination number Zm=18 [42,73]. Substituting
Eq.(3.46) in Eq.(3.41) gives:
Development SPHCT 45

T ε
*
Z m V e2 k T
φ=kT ε ∫ 2 ⎡ *⎛
ε
⎞ ⎤
dT (3.47)
T =∞ k T ⎢V + V ⎜ e 2 k T - 1⎟⎥
⎣ ⎝ ⎠⎦

To solve the integral in Eq.(3.47), we define a new variable Ψ:


ε
Ψ= (3.48)
2kT

Now dT can be transformed to dΨ:

2 k T2
dT = .d Ψ (3.49)

Subsitution of Eqs.(3.48) and (3.49) into Eq.(3.47) leads to the following transformed integral:
Ψ Ψ
e
φ = - 2 k T ZmV . ∫ (V - V )+ V dΨ
*
* * Ψ
(3.50)
Ψ=0 e

Using the Standard Integral Tables of Weast [78], standard integral no. 530, integration leads to:
Ψ
⎡1 ⎤
φ = - 2 k T Z m V ⎢ * . ln ( V - V * + V * . eΨ ) ⎥
*
(3.51)
⎣V ⎦Ψ = 0

Elaboration of the integration limits and back transformation from Ψ to T gives:


⎡ ⎤
⎢ V ⎥ (3.52)
φ = 2 k T Z m ln ⎢ ε ⎥
⎢V + V * ⎛⎜ e 2 k T - 1⎞⎟ ⎥
⎣ ⎝ ⎠⎦
Eq.(3.52) gives us the temperature and density dependence of the mean interaction potential.
This expression is the one we were looking for and the one that will be used to complete our
expression of the canonical partition function ZN, see Eq.(3.22). Before we do this we need to
adjust Eq.(3.52) to segmental molecules. The kinetic energy of a chain-like molecule is ckT
instead of kT and the intermolecular attraction ε has to be replaced by εq in which ε is a
characteristic energy per segment that has an external surface area and q is the external surface
area per molecule. The partition function now becomes:

N
1 ⎛ V ⎞ ⎡ - η ( 4 - 3 η2 ) ⎤ N c - N c Z m ⎡⎢ ln V * ⎤
(3.53)
Z ( N,V, T ) = ⎜ ⎟ . e ( 1 -η ) ⎥ . e ⎥
N ! ⎝ Λ 3 ⎠ ⎢⎣
⎣ V +V Y ⎦

46 Chapter 3

with η=τV*/V and Y=exp(εq/2ckT)-1. With Eq.(3.14) this leads to:


⎛V ⎞
A = - k T ln ( Z N ) = ln ( N ! ) - N k T ln ⎜ 3 ⎟
⎝Λ ⎠
(3.54)

η ( 4 - 3η ) ⎛ V ⎞
+N kT c + N k T c Z m ln ⎜ ⎟
( 1 -η ) 2 *
⎝ V +V Y ⎠

With Eq.(3.15) the SPHCT equation of state now becomes:

R T c R T 4η - 2η2 R T Z m c V* Y
p= + - (3.55)
V V ( 1 - η )3 V V + V * Y

where N=NA, Avogadro's number and R=NAk, the gas constant.

3.6 Mixing rules

In the previous sections the SPHCT was derived for pure fluids. In this section the
SPHCT will be extended to fluid mixtures. The most widely used method for extending
equations of state to mixtures is the 'quadratic' or 'Van der Waals' mixing rules, first proposed by
Van der Waals in 1890 [79]. According to these mixing rules it is assumed that the properties of
a fluid mixture are the same as those of a hypothetical pure fluid at the same pressure and
temperature, but having the characteristic constants appropriately averaged over the composition.
The averaging functions or mixing rules are quadratic in the mole fraction:
n n
w = ∑ ∑ xi x j wij (3.56)
i=1 j=1

where wij represents a constant of the equation of state. For pure component i the value of w
reduces to the pure component parameter wii. In case i is not equal to j, wij is determined by an
appropriate combining rule, with or without binary interaction coefficients. For the Van der
Waals equation of state the quadratic mixing rules are in accordance with the 'one-fluid'
approach. The one-fluid mixing rules describe how the microscopic properties ε and σ of a fluid
mixture can be calculated from the ε and σ of the pure components. For a fluid mixture with a
square-well potential function between the i and j molecules, it can be stated that:

u ij (r) = ε ij . f ( r / σ ij ) (3.57)

where εij represents the intermolecular interaction energy and σij is the interaction distance.
Development SPHCT 47

The one-fluid mixing rules [80-82] for ε and σ are:


n n
< σ 3 > = ∑ ∑ xi x j σ ij3 (3.58)
i=1 j=1

n n
< ε σ 3 > = ∑ ∑ xi x j ε ij σ ij3 (3.59)
i=1 j=1

These equations suggest a quadratic composition dependence of the co-volume and energetic
interaction parameters respectively. However, by setting σij=(σii+σjj)/2 in Eq.(3.58) the co-
volume is represented by a linear average. Two or more fluid concepts have received less
attention and are mostly used for equations of state based on corresponding states principles.
In order to model complex phase behaviour of highly nonideal mixtures, mixing rules
other than the previously mentioned Van der Waals mixing rules are required. Several authors
[83-86] have proposed various forms of the Van der Waals mixing rules that use composition
dependent binary interaction coefficients. Such mixing rules are not generally applicable and
may fail for simple mixtures [87]. Moreover, they might lead to inconsistencies at the low
density limit where the second virial coefficient should be a quadratic function of the
composition. Sometimes these new mixing rules lead to inconsistencies when a component is
split into two or more identical fractions [88]. One of the most promising developments,
especially for mixtures that are composed of polar molecules, may be that proposed by Wong
and Sandler [89,90]. They developed a new mixing rule based on the Van der Waals mixing
rules for cubic equations of state by equating the Helmholtz energy at infinite pressure from an
equation of state to that from an activity coefficient model. Consequently, their new mixing rule
produces the correct low density and high density limits without density dependent binary
interaction coefficients [89,90].
In the case of the SPHCT equation of state and its subsequent modifications the one-fluid
approach is not used. The mixing rules for the SPHCT equation of state are based on a
combination of results of the perturbation theory for a mixture of square-well molecules with
mixing rules derived from lattice theory. For the flexibility parameter and the covolume a linear
mixing rule is used. In case of the covolume this linear mixing rule is applied to the segments of
each molecule.
n
< c > = ∑ xi ci (3.60)
i=1

n n 3
< V > = ∑ xi V = ∑
* * N A si σ ii
i xi (3.61)
i=1 i=1 2

The terms in angular brackets, <...>, represent mixture properties. For the attractive parameter in
the original PHCT, mixing rules are used that are similar to the one-fluid mixing rule but are now
adjusted to segmental molecules. Donohue [25] based his ideas on the 'lattice requirement':
48 Chapter 3

n n
< ε q s σ 3 > = ∑ ∑ xi x j ε ij qi s j σ ij3 (3.62)
i=1 j=1

The standard method for introducing a binary interaction coefficient into the mixing rule is to
assume a corrected geometric mean rule for the energy parameter:

ε ij = ε ii ε jj .( 1 - k ij ) (3.63)

A mixing rule was developed for the SPHCT equation of state by Kim et al. [41] that is similar to
that of the PHCT. One of the most striking aspects of the mixing rule for the attractive parameter,
<cV*Y>, is that the segmental interactions that are mixed, are weighted by the flexibility and
covolume parameters of the corresponding molecules:
n
⎡ ⎛ ε ij qi ⎞ ⎤
< cV * Y > = ∑ xi x j ci V *ji ⎢exp⎜⎜ ⎟⎟ - 1⎥ (3.64)
i, j=1 ⎣ ⎝ 2 ci k T ⎠ ⎦

The mixing rule in Eq.(3.64) is based on the 'mean density' approach. Finally, we give the
multicomponent expression of the SPHCT equation of state:
⎡ ⎤
⎢ ⎥
R T < c > R T ⎡4η - 2η ⎤ R T ⎢ Z m < c V* Y > ⎥
2
(3.65)
p= + ⎢ ⎥ -
⎣ (1 - η ) ⎦ V ⎢V + ⎛⎜ < c V Y > ⎞⎟ ⎥
3 *
V V
⎢ ⎥
⎣ ⎝ <c > ⎠⎦

where η=τ.V*/V. The geometrical constant τ and the maximum coordination number Zm have the
following values [42,73]:

τ = 0.7405
Zm = 18.0

The pure component parameters Y, V* and T* are defined by Eqs.(3.66), (3.67) and (3.68):

⎛ * ⎞
Y = exp⎜ T ⎟ - 1 (3.66)
⎝ 2T ⎠

* sσ3
V = NA (3.67)
2

* εq
T = (3.68)
ck
Development SPHCT 49

The applied mixing rules were proposed by Donohue and Prausnitz [25,26] and Kim et al. [41]:
n
< c > = ∑ xi ci (3.69)
i=1

n n 3
< V * > = ∑ xi V *i= ∑ xi N A si σ ii (3.70)
i=1 i=1 2

n
⎡ ⎛ ε ij qi ⎞ ⎤
< cV * Y > = ∑ xi x j ci V *ji ⎢exp⎜⎜ ⎟⎟ - 1⎥ (3.71)
i, j=1 ⎣ ⎝ 2 ci k T ⎠ ⎦

with:

s j σ ji
3
*
V = NA
ji (3.72)
2

and with cross terms given by:

σ ii + σ jj
σ ij = (3.73)
2

ε ij = ε ii .ε jj . (1 - k ij ) (3.74)

where kij represents the binary interaction coefficient.

In the next chapter the accuracy of the SPHCT equation of state will be tested. It is
obvious that it is impossible to give a systematic survey of the quality of all possible
thermodynamic properties that can be calculated from the SPHCT equation of state. Therefore
we have restricted ourselves to a limited number of thermodynamic properties.

NOMENCLATURE

A molar Helmholtz energy H Hamiltonian


b covolume parameter of the Van der h Planck's constant
Waals type equations of state. h total correlation function
c 1/3 of the total degrees of freedom of h12 total pair correlation function
a molecule k Boltzmann constant
c direct correlation function kij binary interaction coefficient
c12 direct pair correlation function mi mass of molecule i
g radial distribution function N number of particles
g12 pair distribution function NA Avogadro's constant
50 Chapter 3

Nc occupation number Greek symbols


p pressure
pi impulse of molecule i ∂ partial derivative
Q configuration integral ε intermolecular potential energy per
q number of external segments per unit area
molecule η reduced density
R ratio of square-well width to the hard κT isothermal compressibility
core diameter of a segment Λ de Broglie-wavelength
R gas constant ρ density
ri position of molecule i σ hard core diameter of a segment
rij distance between molecule i and j τ geometrical constant
s number of segments per molecule Φ mean potential
T absolute temperature ψ transformed temperature
T* characteristic temperature ω acentric factor
UN total interaction energy, also called
configuration energy
uij intermolecular pair potential
V molar volume
V* covolume parameter of the SPHCT
equation of state
Vf free volume
Z compressibility ratio
Zm close-packed coordination number
ZN canonical partition function

Subscripts

attr attractive
exp experimental
ext external
i component i
IG ideal gas
int internal
r rotational
r reduced property
rep repulsive
seg segment
t translational
v vibrational
Development SPHCT 51

REFERENCES

1. A. Anderko. Equation of state methods for the modelling of phase equilibria. Fluid Phase
Equilibria 61, (1990), 145-225.
2. T.W. Leland. Proc. of the 2nd Intern. Conf. on Phase Equilibria and Fluid Properties in the
Chemical Industry, West-Berlin, (1980).
3. J.D. van der Waals. Over de continuiteit van den gas- en vloeistoftoestand. PhD-thesis,
Leiden, (1873).
4. O. Redlich and J.N.S. Kwong. On the thermodynamics of solutions. V. Chem. Rev. 44,
(1949), 233-244.
5. G. Soave. Equilibrium constants from a modified Redlich-Kwong equation of state. Chem.
Engng Sci. 27, (1949), 1197-1203.
6. D.Y. Peng and D.B. Robinson. A new two-constant equation of state. Ind. Eng. Chem.
Fundam. 15, (1976), 59-64.
7. G.G. Fuller. A modified Redlich-Kwong-Soave equation of state capable of representing the
liquid state. Ind. Eng. Chem. Fundam. 15, (1976), 254-257.
8. G. Schmidt and H. Wenzel. A modified van der Waals equation of state. Chem. Eng. Sci.
35, (1980), 1503-1512.
9. A. Harmens and H. Knapp. Three parameter cubic equation of state for normal substances.
Ind. Eng. Chem. Fundam. 19, (1980), 291-294.
10. N.C. Patel and A.S. Teja. A new cubic equation of state for fluids and fluid mixtures. Chem.
Eng. Sci. 37, (1982), 463-473.
11. M. Benedict, G.R. Webb and L.C. Rubin. An empirical equation for thermodynamic
properties of light hydrocarbons and their mixtures. I. Methane, ethane, propane and butane.
J. Chem. Phys. 8, (1940), 334-335.
12. R.D. Goodwin. Thermophysical properties of methane from 90 to 500 K at pressures to 700
bar. NBS Technical Note 653.
13. B.I. Lee and M.G. Kesler. A generalized thermodynamic correlation based on three-
parameter corresponding states. AIChE J. 21, (1975), 510-527.
14. M.S. Wertheim. Exact solution of the Percus-Yevick integral equation for hard spheres.
Phys. Rev. Lett. 10, (1963), 321-323.
15. E. Thiele. Equation of state for hard spheres. J. Chem. Phys. 39, (1963), 474-479.
16. N.F. Carnahan and K.E. Starling. Equation of state for non-attracting rigid spheres. J. Chem.
Phys. 51, (1969), 635-636.
17. N.F. Carnahan and K.E. Starling. Thermodynamic properties of a rigid-sphere fluid. J.
Chem. Phys. 53, (1970), 600-603.
18. R.M. Gibbons. The scaled particle theory for particles of arbitrary. Mol. Phys. 17, (1969),
81-86.
19. T. Boublik. Hard sphere equation of state. J. Chem. Phys. 53, (1970), 471-472.
20. T. Boublik and I. Nezbeda. Equation of state for hard dumbbells. Chem. Phys. Lett. 46,
(1977), 315.
21. T. Boublik. Equation of state for hard convex body fluids. Mol. Phys. 42, (1981), 209-216.
22. I. Nezbeda and T.W. Leland. Conformal theory of hard non-spherical molecule fluids. J.
Chem. Soc., Faraday Trans. II 75, (1979), 193-200.
23. S. Beret and J.M. Prausnitz. Perturbed hard chain theory: an equation of state for fluids
containing small or large molecules. AIChE J. 21, (1975), 1123-1132.
24. S. Beret and J.M. Prausnitz. A generalized Van der Waals equation for polymers and other
fluids. Macromolecules 8, (1975), 878-882.
52 Chapter 3

25. M.D. Donohue and J.M. Prausnitz. Statiscal thermodynamics of solutions in natural gas and
petroleum refining. Report RR-26, Gas Processors Assn. (1977).
26. M.D. Donohue and J.M. Prausnitz. Perturbed hard chain theory for fluid mixtures in natural
gas and petroleum technology. AIChE J. 24, (1978), 849-860.
27. A. Kreglewski, R.C. Wilhoit and B.J. Zwolinski. Thermodynamic properties of conformal
mixtures calculated from the hard-sphere equation of state. J. Phys. Chem. 77, (1973), 2212-
2217.
28. A. Kreglewski and R.C. Wilhoit. Thermodynamic properties of systems with specific
interactions calculated from the hard-sphere equation of state. I. Binary systems with one
inert component. J. Phys. Chem. 78, (1974), 1961-1967.
29. A. Kreglewski and R.C. Wilhoit. Thermodynamic properties of systems with specific
interactions calculated from the hard-sphere equation of state. II. Binary systems of electron
donors and acceptors. J. Phys. Chem. 79, (1975), 449-454.
30. A. Kreglewski and S.S. Chen. Applications of the augmented van der Waals theory of
fluids. II. Phase equilibria in mixtures. J. Chim. Phys. 75, (1978), 347-352.
31. I. Prigogine. The molecular theory of solutions. Amsterdam, (1957).
32. P.J. Flory. Statistical thermodynamics of liquid mixtures. J. Am. Chem. Soc. 87, (1965),
1833-1838.
33. B.K. Kaul, M.D. Donohue and J.M. Prausnitz. Henry's constants and second virial
coefficients from Perturbed-Hard-Chain Theory. Fluid Phase Equilibria 4, (1980), 171-184.
34. D.D. Liu and J.M. Prausnitz. Thermodynamics of gas solubilities in molten polymers. J.
Appl. Polymer Sci. 24, (1979), 725-733.
35. D.D. Liu and J.M. Prausnitz. Molecular thermodynamics of polymer compatibility: effect of
contact agility. Macromolecules 12, (1979), 454-458.
36. D.D. Liu and J.M. Prausnitz. Calculation of phase equilibria for mixtures of ethylene and
low-density polyethylene at high pressures. Ind. Eng. Chem. Prod. Des. Dev. 19, (1980),
205-211.
37. M. Ohzone, Y. Iwai and Y. Arai. Correlation of solubilities of hydrocarbon gases and
vapors in molten polymers using Perturbed-Hard-Chain Theory. J. Chem. Eng. Japan 17,
(1984), 550-553.
38. A. Bondi. Physical properties of molecular crystals, liquids and glasses. John Wiley, New
York, (1968).
39. J. Gmehling, D.D. Liu and J.M. Prausnitz. High pressure vapor-liquid equilibria for
mixtures containing one or more polar components. Application of an equation of state
which includes dimerization equilibria. Chem. Eng. Sci. 34, (1979), 951-958.
40. A. Wilhelm and J.M. Prausnitz. Vapour pressures and saturated liquid volumes for heavy
fossil fuel fractions from a perturbed hard chain equation of state. Fuel 64, (1985), 501-508.
41. C-H. Kim, P. Vimalchand, M.D. Donohue and S.I. Sandler. Local composition model for
chainlike molecules: a new simplified version of the perturbed hard chain theory. AIChE J.
32, (1986), 1726-1734.
42. K.H. Lee, M. Lombardo and S.I. Sandler. The generalized Van der Waals partition function.
II: Application to the square-well fluid. Fluid Phase Equilibria 21, (1985), 177-196.
43. P. Vimalchand, A. Thomas, I.G. Economou and M.D. Donohue. Effect of hard-sphere
structure on pure-component equation of state calculations. Fluid Phase Equilibria 73,
(1992), 39-55.
44. P. Vimalchand and M.D. Donohue. Thermodynamics of quadrupolar molecules: The
Perturbed-Anisotropic-Chain Theory. Ind. Eng. Chem. Fundam. 24, (1985), 246-257.
Development SPHCT 53

45. W.O. Morris, P. Vimalchand and M.D. Donohue. The Perturbed-Soft-Chain Theory: An
equation of state based on the Lennard-Jones potential. Fluid Phase Equilibria 32, (1987),
103-115.
46. P. Vimalchand, I.C. Celmins and M.D. Donohue. VLE- Calculations for mixtures
containing multipolar compounds using the Perturbed-Anisotropic-Chain Theory. AIChE J.
32, (1986), 1735-1738.
47. R.L. Cotterman, B.J. Schwarz and J.M. Prausnitz. Molecular thermodynamics for fluids at
low and high densities. Part I: Pure fluids containing small or large molecules. AIChE J. 32,
(1986), 1787-1798.
48. R.L. Cotterman and J.M. Prausnitz. Molecular thermodynamics for fluids at low and high
densities. Part II: Phase equilibria for mixtures containing components with large
differences in molecular size or potential energy. AIChE J. 32, (1986), 1799-1812.
49. C.H. Chien, R.A. Greenkorn and K.C. Chao. Chain-of-rotators equation of state. AIChE J.
29, (1983), 560-571.
50. U.K. Deiters. A new semi-empirical equation of state for fluids - I. Derivation. Chem. Eng.
Sci. 36, (1981), 1139-1145.
51. U.K. Deiters. A new semi-empirical equation of state for fluids - II. Application to pure
substances. Chem. Eng. Sci. 36, (1981), 1147-1151.
52. U.K. Deiters. A new semi-empirical equation of state for fluids - III. Application to phase
equilibria in binary mixtures. Chem. Eng. Sci. 37, (1982), 855-861.
53. G.D. Ikonomou and M.D. Donohue. Thermodynamics of hydrogen-bonded molecules: The
Associated-Perturbed-Anisotropic-Chain Theory. AIChE J. 32, (1986), 1716-1725.
54. M.D. Donohue, P. Vimalchand and G.D. Ikonomou. Phase equilibrium predictions for polar
and hydrogen bonding mixtures. Fluid Phase Equilibria 30, (1986), 307-314.
55. I.G. Economou, G.D. Ikonomou, P. Vimalchand and M.D. Donohue. Thermodynamics of
Lewis acid-base mixtures. AIChE J. 36, (1990), 1851-1864.
56. J.R. Elliott Jr., S.J. Suresh and M.D. Donohue. A simple equation of state for non-spherical
and associating molecules. Ind. End. Chem. Res. 29, (1990), 1476-1485.
57. G.D. Ikonomou and M.D. Donohue. COMPACT: A simple equation of state for associated
molecles. Fluid Phase Equilibria 33, (1987), 61-90.
58. G. Ciocca, I. Nagata and V. Brandani. Density dependence of the external degrees of
freedom: application of a simplified version of the Perturbed Hard Chain Theory. Fluid
Phase Equilibria 41, (1987), 59-80.
59. G. Ciocca and I. Nagata. The simplified perturbed anisotropic chain theory: a new equation
of state for polar fluids. Fluid Phase Equilibria 58, (1990), 265-282.
60. G. Jin, J.M. Walsh and M.D. Donohue. A group-contribution correlation for predicting
thermodynamic properties with the Perturbed-Soft-Chain-Theory. Fluid Phase Equilibria
31, (1986), 123-146.
61. H.C. Lu and W. Sheng. High pressure low temperature vapour-liquid equilibria of hydrogen
containing systems. Fluid Phase Equilibria 23, (1985), 1-13.
62. R. Dickman and C.K. Hall. Equations of state for chain molecules: continuous space analog
of Flory theory. J. Chem. Phys. 85, (1986), 3023-3026.
63. K.G. Honnell and C.K. Hall. A new equation of state for athermal chains. J. Chem. Phys.
90, (1989), 1841-1855.
64. C.K. Hall, M.A. Denlinger and K.G. Honnell. Generalized Flory theories for predicting
properties of fluids containing long chain molecules. Fluid Phase Equilibria 53, (1989), 151-
158.
65. M.D. Donohue and C.K. Hall. A thermodynamic model for gas processing. I. Equation of
state for pure, chain fluids. To be published.
54 Chapter 3

66. M.D. Donohue and P. Vimalchand. The perturbed-hard-chain theory. Extensions and
applications. Fluid Phase Equilibria 40, (1988), 185-211.
67. P. Vimalchand and M.D. Donohue. Comparison of equations of state for chain molecules. J.
Phys. Chem. 93, (1989), 4355-4360.
68. C.J. Peters, J.L. de Roo, and J. de Swaan Arons. Measurements and calculations of phase
equilibria in binary mixtures of propane + tetratriacontane. Fluid Phase Equilibria 72,
(1992), 251-266.
69. Gasem, K.A.M., and Robinson, R.L. Jr. Evaluation of the Simplified Perturbed Hard Chain
Theory (SPHCT) for prediction of phase behaviour of n-parrafins and mixtures of n-
parrafins with ethane. Fluid Phase Equilibria 58, (1990), 13-23.
70. L. Ponce-Ramirez, C. Lira-Galeana and C. Tapia-Medina. Application of the SPHCT model
to the prediction of phase equilibria in CO2-hydrocarbon systems. Fluid Phase Equilibria 70,
(1992), 1-18.
71. G.H. Findenegg. Statistische Thermodynamik. Dr. Dietrich Steinkopff Verlag, Darmstadt,
Germany, (1985).
72. I. Mills, T. Cvitas, N. Kallay, K. Homann and K. Kuchitsu. International Union of Pure and
Applied Chemistry: Quantities, Units and Symbols in Physical Chemistry. Blackwell
Scientific Publications, Oxford, (1988).
73. S.I. Sandler. The generalized Van Der Waals partition function. I. Basic theory. Fluid Phase
Equilibria 19, (1985), 233 - 257.
74. L.S. Ornstein and F. Zernike. Accidental deviations of density and opalescence at the
critical point of a single substance. Proc.Acad.Sci. 17, (1914), 793-806.
75. R.J. Baxter in E. Eyring, D. Henderson and W. Jost. Physical chemistry VIII. A: An
advanced treatise; the liquid state. Academic Press, New York. (1971).
76. T.M. Reed and K.E. Gubbins. Applied statistical mechanics. McGraw-Hill, New York.
(1973), chap. 8.
77. J.S. Rowlinson and F.L. Swinton. Liquids and Liquid Mixtures Butterworths Monographs
in Chemistry, London, Chap. 7, (1982).
78. C.W. Weast. CRC Handbook of Chemistry and Physics. 62nd Ed., Boca Raton, (1982).
79. J.D. van der Waals. Molekulartheorie eines Körpers, der aus zwei verschiedenen Stoffen
besteht. Separat-Abdruck aus "Zeitschrift für physikalische Chemie" V2. Wilhelm
Engelmann, Leipzig, Germany, (1890), 133-173.
80. T.W. Leland, P.S. Chappelear and B.W. Gamson. Prediction of vapor-liquid equilibria from
the corresponding states principle. AIChE J. 8, (1962), 482-489.
81. T.W. Leland, J.S. Rowlinson and G.A. Sather. Statistical thermodynamics of mixtures of
molecules of different sizes. Trans. Faraday Soc. 64, (1967), 1447-1460.
82. R.C. Reid and T.W. Leland. Pseudo-critical constants. AIChE J. 11, (1965), 228-237.
83. A.Z. Panagiotopoulos and R.C. Reid. New mixing rule for cubic equations of state for
highly polar, asymmetric systems. ACS Symp. Ser. 300, (1986), 571.
84. Y. Adachi and H. Sugie. A new mixing rule - modified conventional mixing rule. Fluid
Phase Equilibria 28, (1986), 103-118.
85. R. Sandoval, G. Wilczek-Vera and J.H. Vera. Prediction of ternary vapor-liquid equilibria
with the PRSV equation of state. Fluid Phase Equilibria 52, (1989), 119-129.
86. J. Schwarzentruber, H. Renon and S. Wantanasiri. Development of a new cubic equation of
state for phase equilibrium calculations. Fluid Phase Equilibria 52, (1989), 127-134.
87. S.K. Shibata and S.I. Sandler. Critical evaluation of equation of state mixing rules for the
prediction of high pressure phase equilibria. Ind. Eng. Chem. Res. 28, (1989), 1893-1898.
Development SPHCT 55

88. M.L. Michelsen and H. Kistenmacher. On composition dependent interaction coefficients.


Fluid Phase Equilibria 58, (1990), 229-230.
89. D.S.H. Wong and S.I. Sandler. A theoretically correct mixing rule for cubic equations of
state. AIChE J. 38, (1992), 671-680.
90. D.S.H. Wong and S.I. Sandler. An equation of state mixing rule for nonideal mixtures using
available activity coefficient model parameters and which allows extrapolation over large
ranges of temperature and pressure. Ind. Eng. Chem. 31, (1992), 2033.
56

CHAPTER 4

THE SIMPLIFIED-PERTURBED-HARD-CHAIN THEORY: PERFORMANCE

In chapter 3 the SPHCT equation of state was derived. In this chapter the performance of
the SPHCT equation will be tested for several thermodynamic properties. In section 4.1 the pure
component parameters will be obtained for more than 100 fluids. The validity of the attractive
term will be examined in section 4.2. The virial coefficients that can be calculated from the
SPHCT equation will be discussed in section 4.3. In section 4.4 the calculations of a property
that diverges at the critical point will be considered, i.e., the isochoric heat capacity. In sections
4.5 to 4.7 the equation is applied to binary mixtures. We start with the calculation of isothermal
phase envelopes in section 4.5. The Henry coefficients that have been calculated from the
SPHCT equation will be discussed in section 4.6 and finally, in section 4.7, attention will be
given to the calculated critical curves in binary hydrocarbon mixtures.

4.1 Pure component parameters

4.1.1 Introduction

In this section the pure component parameters of the SPHCT equation of state will be
optimized so that calculations of saturated vapour pressures near the critical point can be made.
Because of the fundamental importance of the critical point it is necessary for any equation of
state to reproduce the critical point of a pure component properly. Kim et al. [1] presented pure-
component parameters that were optimized from saturated vapour pressure and liquid density
data. Gasem and Robinson [2] extended the availability of pure component parameters to higher
normal alkanes, up to C64H130. Similarly, Ponce et al. [3] optimized pure component parameters
for 70 pure fluids encountered in the oil industry (paraffins, olefins, naphtenes and aromatics). A
drawback of all three collections of parameters is the poor reproduction of the critical point, e.g.,
for ethane the overprediction of the critical temperature is about 10 K. For higher alkanes and
polar molecules, like H2O, this overprediction becomes much more pronounced. We (re)fitted
the pure component parameters of more than 100 components, including some polar fluids, so
that the critical point in the pressure-temperature- or p,T-plane is matched exactly.
Performance SPHCT 57

4.1.2 The optimization procedure

The critical point of a pure substance is given by two conditions, e.g., [4]:
⎛∂ p⎞
⎜ ⎟ =0 (4.1)
⎝ ∂ V ⎠T c

⎛ ∂2 p ⎞
⎜⎜ 2 ⎟⎟ = 0 (4.2)
⎝ ∂V ⎠T c

where V represents the molar volume. In addition the equation of state should yield the correct
value of the critical pressure:

p ( V c ,T c ) = p c (4.3)

Fixing the calculated pc, Tc and Vc at the


experimental values does not leave any degree
of freedom for fitting the parameters c, V* and
T* from saturated vapour pressure data. Then
three conditions with three unknowns are to be
satisfied; i.e., mathematically the system is
completely determined. Parameters obtained in
this way, overpredict saturated vapour pressures
and cause physically unrealistic values, e.g.,
0.05<c<0.10 was found for the normal alkane
series up to decane.
Acceptable calculation of saturated
vapour pressures near the critical point requires
Figure 4.1: Relative error of the calculated
the critical volume, as calculated by the SPHCT
vapour pressure curve for n-
equation of state, to be treated as an empirical
hexane.
parameter. Dohrn and Prausnitz [5] showed that
the calculated value of the critical volume (or the critical compressibility factor Zc) is not an
important indicator for the overall performance of an equation of state. The parameters c, V* and
T* in the SPHCT equation of state were calculated from critical data (e.g., pc and Tc) and
additionally fitted to saturated vapour pressure data. As a result the calculated vapour pressure
curve ends exactly at the critical point, see Fig.4.1.
The following trial and error procedure was adopted in obtaining c, V* and T*. Initially Vc
was set equal to 0.3464*R.Tc/pc, where R represents the gas constant. Using Eqs.(4.1)-(4.3),
values of c, V* and T* were calculated with the Marquardt nonlinear least-squares method.
58 Chapter 4

The Marquardt algorithm is extensively described in chapter 6. Using these parameters the
Absolute Average Deviation, AAD, along the vapour pressure curve was evaluated in the
temperature range 0.75 < T/Tc < 1.

sat sat
N pi,exp - pi,calc
∑i=1
sat
pi,exp
AAD = (4.4)
N

The value of Vc was then changed by a small


amount, ΔV, and new values of c, V*, T* and
AAD were obtained by solving Eqs.(4.1)-(4.4).
This procedure was repeated until a minimum
occurred in the value of the AAD. When
necessary the stepsize in Vc was adjusted.
Strictly speaking, this often used optimization
method is only correct when raw experimental
data are used. In the literature many saturated
vapour pressure data are often smoothed by a
multi-parameter equation, for example the
IUPAC-data series [6] and the API-tables [7].
To what extent the degree of data smoothing
effects the values of the parameters of the Figure 4.2: Schematic representation of
SPHCT equation is not known. the square-well potential
Kim et al. [1] used 36.0 as a value for function.
the maximum coordination number Zm, appearing in the attractive term of the SPHCT equation.
This high coordination number was chosen so that c equals 1.0 for methane. However, in theory
the maximum coordination number depends on R, the ratio of the width to the hard core diameter
σ, of the applied square-well potential in the model. This is shown in Fig.4.2. Sherwood and
Prausnitz [8] optimized values of σ and R from second virial coefficient data. They found an
average value of R approximately equal to 1.5 (1.60 for methane, 1.36 for n-pentane). On
geometrical grounds Lee et al. [9] and Lee and Chao [10] concluded that for R=1.5, Zm should be
18.0. Therefore Zm = 18.0 was used in our parameter optimization.

4.1.3 Results

Optimum values of c, V* and T* for many pure fluids, including polar substances, are
given in Table 4.1. For six classes of pure components the calculated vapour pressure curves are
shown in Fig.4.3. Fig.4.4 shows the variation of the new SPHCT equation of state parameters
with carbon number for the lower normal alkanes. When experimental data are scarce, Fig.4.4
Performance SPHCT 59

may be used to estimate the required parameters. It was shown by Peters et al. [11], that good
agreement was obtained between experimental data and calculated results of the system propane
+ tetratriacontane using extrapolated parameters for tetratriacontane. The linear behaviour that is
shown in Fig.4.4 suggests that these parameters relate to molecular properties and are not merely
adjustable constants.
The SPHCT equation of state contains seven parameters (T*, V*, c, ε, q, s, σ) for each
pure component, but only five of these are independent. Three of these parameters, i.e., c, V* and
T*, are necessary to calculate pure component properties. These were determined according to
the method described above. The other two parameters can be determined by correlating
parameters for many similar fluids. The potential square-well depth, ε or ε/k, was determined
from the slope of c.T* (=εq/k) versus n. The segmental volume, V*seg, was determined from the
slope of V* versus n, see Fig.4.4. With Eq.(3.67) the segmental diameter, σ, can be calculated. In
both cases, the normal alkanes from methane up to decane, were used for obtaining values of ε/k
and σ.
For a CH2-segment the volume is 6.7306 cm3.mole-1 and the diameter σ = 2.510 Å. This
order of magnitude agrees with experimental bond lengths in normal alkanes, for C-C: 1.541 Å
and for C-H: 1.073 Å, [12]. The potential energy per unit external surface area is ε/k = 125.8 K.
For methane this leads to a characteristic temperature T* = 138.661 K. This value is comparable
with the one that Sherwood and Prausnitz [8] calculated for methane (142.5 K) from second
virial coefficient data.
The assumed linear behaviour of the hard core volume and the energy density with the
carbon number (or segment number) leads to a general equation relating the parameters q and s.

ε .q
∂( )
k ε ∂q
= . = C1 (4.5)
∂n k ∂n

∂(s.Vseg
*
) ∂s
= Vseg
*
. = C2 (4.6)
∂n ∂n
60 Chapter 4

Figure 4.3: The calculated vapour pressure curve of six classes of compounds; a: normal
alkanes, b: cycloalkanes, c: 1-alkenes, d: alkyl-benzenes, e: 1-alkanols, f: alkyl-
amines. The value of n near each curve represents the carbon number of the
component under consideration.
Performance SPHCT 61

Figure 4.4: Pure component parameters as a function of carbon number for the normal
alkanes.
62 Chapter 4

Combining both equations leads to:

dq
= C3 ⇒ q = C3 .s + C4 (4.7)
ds

Fitting q as a function of s leads to: C3 = 1


and C4 = -1 for the normal alkanes, see Fig.
4.5. With this relation between s and q, the
previously mentioned separation of (sσ) and
(εq/k) can be done in the following way:
The number of segments s can be deter-
mined from s = V*/Vseg*. Eq.(4.7) can be
used to calculate q and with Eq.(3.68) ε/k
can be calculated. Donohue and Prausnitz
[13,14] applied the relation s = V*/Vseg* to
all kinds of molecules. For Vseg* in this
relation these investigators used the volume
of one CH2-segment, even if the molecule
did not contain any CH2-segments. They
concluded that the segment size is quite Figure 4.5: Surface area of the n-alkanes as a
arbitrary and does not influence the calcu- function of the segment number for
lated results. In this way the SPHCT equa- the normal alkanes.
tion of state can also be applied to molecules that do not belong to a homologous series, for
instance H2O, CO2 and NH3. In preliminary binary mixture calculations of polar systems,
Eq.(4.7) was used. Strictly speaking this is not correct: the relation between q and s depends on
the structure of the molecule. However, Eq.(4.7) with the values (C3,C4) = (1,-1), gave quite
satisfactory results for polar mixtures as well.

Table 4.1.Pure component parameters for the SPHCT equation of state.

NORMAL c V* T* AAD N lit.


3
ALKANES - cm /mole K % -
METHANE 1.2943 17.292 138.661 1.17 25 6
ETHANE 1.6755 23.834 204.225 1.32 16 7
PROPANE 1.9904 31.067 234.618 1.52 20 7
N-BUTANE 2.2118 38.404 261.571 1.49 23 7
N-PENTANE 2.5349 45.324 278.172 1.73 25 7
N-HEXANE 2.8412 52.427 291.409 2.39 26 7
N-HEPTANE 3.0833 59.511 303.757 1.21 28 7
Performance SPHCT 63

N-OCTANE 3.4305 66.001 311.449 2.12 30 7


N-NONANE 3.8043 71.749 317.050 1.80 31 7
N-DECANE 4.1717 76.187 322.819 1.99 31 7

CYCLO- c V* T* AAD N lit.


ALKANES - cm3/mole K % -
C-PROPANE 1.8531 26.183 258.131 1.30 11 7
C-BUTANE 2.1624 31.905 284.848 1.22 12 7
C-PENTANE 2.1700 39.250 316.486 1.51 14 7
C-HEXANE 2.2728 46.104 337.905 1.24 15 7
C-HEPTANE 2.4892 51.587 359.673 1.78 17 7
C-OCTANE 2.4617 59.697 386.388 0.76 17 7
ALKENES c V* T* AAD N lit.
3
- cm /mole K % -
ETHENE 1.6326 21.381 190.545 1.28 41 6
PROPENE 1.8989 28.463 235.210 1.43 20 6
1-BUTENE 2.1922 35.914 258.080 1.45 19 17
1-PENTENE 2.4228 43.629 278.703 1.74 25 7
1-HEXENE 2.6737 51.137 294.230 1.69 27 7
1-HEPTENE 2.9923 57.807 304.512 1.94 28 7
1-OCTENE 3.2969 64.627 313.357 2.00 29 7
1-NONENE 3.5867 71.161 321.330 2.09 31 7
1-DECENE 3.7483 81.254 330.549 1.76 32 7
1-UNDECENE 4.0465 87.990 335.828 1.48 33 7
1-DODECENE 4.2800 96.303 341.572 4.50 34 7
1-TRIDECENE 4.8152 99.293 341.672 1.67 32 7
1-TETRADECENE 4.9504 110.109 347.787 1.51 33 7
1-PENTADECENE 5.1933 119.479 351.632 1.13 32 7
1-HEXADECENE 5.5133 126.266 354.010 1.23 35 7
1-HEPTADECENE 5.7707 135.520 356.910 2.18 37 7
1-OCTADECENE 6.2163 142.455 357.304 1.35 32 7
1-NONADECENE 6.5447 150.150 358.957 0.98 30 7
ALKYNES c V* T* AAD N lit.
- cm3/mole K % -
ETHYNE 2.1944 17.007 190.630 1.52 9 17
64 Chapter 4

ALKYL- c V* T* AAD N lit.


BENZENES - cm3/mole K % -
BENZENE 2.3071 38.472 342.039 1.75 27 17
TOLUENE 2.5769 46.596 348.919 1.84 31 7
ETHYL-BENZENE 2.8312 53.290 354.867 2.03 32 7
N-PROPYL-BENZENE 3.0697 60.188 359.402 1.55 34 7
M-XYLENE 2.9833 53.474 349.981 2.00 32 7
O-XYLENE 2.8891 52.180 360.499 1.99 33 7
P-XYLENE 2.9752 53.615 349.762 2.07 32 7

MONO-CHLORO- c V* T* AAD N lit.


3
ALKANES - cm /mole K % -
CH3Cl 2.0839 22.039 262.213 1.51 22 17
C2H5Cl 2.1835 30.323 284.146 1.46 13 18

POLY-FLUORO- c V* T* AAD N lit.


3
ALKANES - cm /mole K % -
CF4 2.0953 21.262 142.161 1.57 21 19
C2F6 2.5991 30.921 172.291 1.94 16 19

ALKANES WITH c V* T* AAD N lit.


3
DIFF HALOGENS - cm /mole K % -
CBrF3 2.1037 29.991 212.317 1.25 26 19
CHClF2 2.3260 24.953 223.962 1.49 34 19
CHCl2F 2.1786 30.013 279.031 1.92 27 7
CH2ClF 2.3636 27.424 258.421 5.16 22 7
CClF3 2.0582 27.499 189.711 1.24 19 7
CCl2F2 2.3543 31.332 232.808 0.15 59 19
CCl3F 2.2164 36.611 289.693 1.43 43 19
C2HClF4 2.4631 35.018 238.689 0.62 21 7
C2H2Cl2F2 2.7090 34.438 288.616 2.27 26 7
C2H3ClF2 2.5355 31.079 242.568 1.87 22 7
C2ClF5 2.5362 36.392 209.090 1.53 18 7
C2Cl2F4 2.5464 41.698 247.764 2.18 39 19
C2Cl3F3 2.6719 45.151 284.495 0.79 10 19
Performance SPHCT 65

MONOHYDRIC c V* T* AAD N lit.


ALIPHATIC ALCO- - cm3/mole K % -
HOLS
METHANOL 4.6811 16.085 261.170 2.31 26 7
ETHANOL 5.1075 20.354 256.941 1.97 27 7
1-PROPANOL 4.6238 26.489 274.209 1.20 28 7
1-BUTANOL 4.3448 33.358 291.647 0.85 29 7
1-PENTANOL 4.3255 39.502 304.954 1.36 30 7
1-HEXANOL 4.4406 45.284 315.144 2.77 32 7
1-HEPTANOL 4.6058 51.287 323.079 3.49 33 7
1-OCTANOL 4.6572 58.004 332.796 2.73 33 7
1-NONANOL 4.9806 63.769 336.276 3.51 35 7
1-DECANOL 5.2538 69.932 340.437 3.36 35 7

ALIPH. MONO- c V* T* AAD N lit.


CARBOXYLIC ACIDS - cm3/mole K % -
ETHANOIC ACID 3.9047 28.136 315.564 2.75 31 7
PROPANOIC ACID 4.2942 30.866 317.842 1.84 31 7
BUTANOIC ACID 5.4759 29.170 309.425 2.68 32 7

PRIMARY c V* T* AAD N lit.


ALKYLAMINES - cm3/mole K % -
METHANAMINE 2.6174 18.603 252.455 1.20 23 7
ETHANAMINE 2.5910 26.135 268.662 1.41 24 7
1-PROPANAMINE 2.7870 32.563 287.694 3.25 26 7
1-BUTANAMINE 2.9222 40.099 300.398 1.49 27 7

AROMATIC c V* T* AAD N lit.


AMINES - cm3/mole K % -
ANILINE 3.3740 37.713 384.337 2.15 36 7
M-TOLUIDINE 3.5399 47.415 385.281 1.39 37 7
O-TOLUIDINE 3.7996 51.918 370.786 3.08 36 7
P-TOLUIDINE 3.6122 44.862 381.514 3.97 36 7

NITRILES c V* T* AAD N lit.


3
- cm /mole K % -
C2N2 2.5903 21.427 236.359 1.68 11 18
66 Chapter 4
SILICON c V* T* AAD N lit.
3
COMPOUNDS - cm /mole K % -
SiF4 4.5537 13.753 139.239 2.94 8 18
SiH4 1.7784 20.871 176.991 2.95 8 18

INORGANIC c V* T* AAD N lit.


COMPOUNDS - cm3/mole K % -
CO 1.4483 15.230 93.142 1.41 13 17
CS2 1.9898 25.597 346.517 1.80 15 18
GeCl4 2.5937 46.061 323.794 1.55 15 18
HBr 1.6833 15.985 242.462 1.68 10 18
HCl 1.8183 14.493 211.645 1.98 11 18
HI 1.5060 20.125 293.090 1.70 13 18
H2O 3.1627 8.725 361.705 2.12 89 15
H2S 1.6471 15.890 251.122 1.05 11 18
H2Se 1.4449 18.527 288.328 1.30 12 18
NH3 2.5269 11.695 240.247 1.25 11 16
N2H4 2.7023 14.112 380.196 0.84 17 18
NO 5.0708 6.688 90.681 2.69 6 18
N2O 1.9268 15.398 198.359 1.15 9 18
PH3 1.5571 19.451 221.851 1.75 10 18
SnCl4 2.5465 51.139 350.079 1.57 16 18
SO2 2.5879 17.634 253.579 1.67 26 17
SO3 3.3277 16.967 271.145 2.90 13 18

ELEMENTS c V* T* AAD N lit.


3
- cm /mole K % -
Ar 1.2319 13.064 111.661 1.21 39 6
H2 0.4616 14.463 36.450 0.89 46 20
N2 1.4072 15.031 89.278 1.14 33 6
O2 1.3271 12.606 111.659 1.18 15 17
Xe 1.1520 21.314 219.565 0.97 5 18

POLYFUNCTIO- c V* T* AAD N lit.


NAL COMPOUNDS - cm3/mole K % -
CO2 2.3205 13.872 184.626 1.41 13 6
COS 1.7144 23.039 251.038 1.78 11 18
HCN 3.6376 25.319 246.524 3.21 13 18
Performance SPHCT 67

It is shown in Table 4.1 that for most components the vapour pressure curve in the
temperature range 0.75 < T/Tc < 1.00 can be described with an absolute average deviation of
1-3 %. Unfortunately, the calculated critical compressibility factor, Zc, is too high, varying
between 0.34 and 0.35 for most components. This means that the critical molar volume is badly
calculated. Near the critical point of a pure fluid errors in calculated fluid densities are often
20 % and more. Because our investigations are directed towards the calculation of critical curves
in binary fluid mixtures we omitted density data from our parameter optimizations. Additionally,
critical density data for binary mixtures are scarce. When they are available, the experimental
uncertainties are usually greater than in density data at low or moderate pressures. A possible
way to improve density calculations from an equation of state is the implementation of a volume
translation in accordance with Peneloux et al. [21]. However, the bad density calculation in the
critical region is not only due to an artefact in the structure of the SPHCT equation. We will
briefly digress in the following paragraph on a fundamental limitation in all "so-called" classical
equations of state. A classical equation of state is an equation in which p is an explicit,
continuous function of V and T.

Even if we had used the most accurate density data in our optimization procedure to
obtain pure component parameters, we would not have achieved a perfect description of the
phase boundary in the p,T- and p,V-plane in the neighbourhood of the critical point. This is due
to the fact that the SPHCT equation is a classical equation of state. A critical point can be
calculated from any classical equation of state, like the SPHCT equation, but in general the
critical point and the phase boundary do not simultaneously agree with experimental data. When
we look at the coexistence curve in the T,ρ-plane, where the density ρ is defined as ρ=1/V, we
see that the coexistence curve is not parabolic around the critical point, as calculated with the
SPHCT equation. The true coexistence curve is much flatter, almost cubic instead of parabolic, at
the top. Consequently, optimizing the parameters of the SPHCT equation of state to the
coexistence boundary at low pressure will cause the calculated critical point in general to be too
high in temperature and pressure. Optimization of the pure component parameters to reproduce
the experimental critical point will be at the expense of a correct phase boundary description at
low pressures, especially in T,ρ-plane. The density calculation along the coexistence curve near
the critical point cannot simply be improved by introducing a new parameter into the equation of
state nor by taking experimental density data into account.

It should be noted further from Table 4.1 that for most non-polar molecules the trends in
the homologous series are as expected: increasing chain lengths lead to increasing values for c,
V* and T*. Most striking in Table 4.1 is the fact that for a completely spherical molecule, e.g.,
methane or argon, the c- parameter is larger than 1.0. This is a result of the fact that methane is
assumed to be built up from 2.6 CH2-segments. During the parameter optimization a certain
flexibility is assigned to these 2.6 CH2-segments. When the value of c is fixed to 1.0, the AAD
will be about 5.1 %, instead of the minimum value of 1.3 % when c = 1.2943. The difference
between segment number and carbon number for the normal alkanes at low carbon numbers
68 Chapter 4

leads to the vertical intercepts of the fitted lines of εq/k and V* versus n, at n = 0. However, as the
chain becomes longer the number of segments approaches the carbon number of the normal
alkanes.

4.2 The temperature dependence of the attractive term.

The attractive term of any equation of state should yield the correct limiting behaviour at
high temperatures. For many equations of state, e.g., the Soave-Redlich-Kwong [22] and the
Peng-Robinson [23] equation, the temperature dependence of the attractive term is based purely
on empirical grounds. The SPHCT equation has a peculiar temperature dependence in the
attractive term. The attractive term has an exponential temperature function in both the
numerator and the denominator. Because of this peculiar temperature dependence, the limiting
behaviour at high temperatures of the attractive term of the SPHCT equation is investigated in
this section and compared with experimental data and other equations of state.
The attractive term of the SPHCT equation of state is given by the following equation:
⎡ ⎤
⎢ * ⎥
attr R T ⎢ Z m . < cV Y > ⎥
p =- (4.8)
V ⎢ ⎛ < cV * Y > ⎞ ⎥
⎢V + ⎜⎜ ⎟⎟ ⎥
⎣ ⎝ < c > ⎠⎦

with:
⎛ T* ⎞
⎜⎜ ⎟⎟
⎝ 2T ⎠
Y =e -1 (4.9)

For a pure substance the attractive term can be factorized in a temperature independent and a
temperature dependent term:
⎡ ⎛⎜⎜ T * ⎞⎟⎟ ⎤
T ⎢ e ⎝ ⎠ - 1⎥
2T

R c Zm ⎢ ⎥
attr
p =- . ⎣ ⎦ (4.10)
⎛ T* ⎞
V ⎜⎜ ⎟
2T ⎟
Vr + e ⎝ ⎠ - 1

where Vr represents the reduced volume V/V*. The temperature dependent term can be made
dimensionless and is called Ψ u. The superscript u stands for 'unnormalized'.
⎡ ⎛⎜⎜⎝ 2 T1 r ⎞⎟⎟⎠ ⎤
Tr ⎢ e - 1⎥
u

⎣ ⎥⎦
Ψ = ⎛ 1 ⎞
⎜ ⎟
(4.11)
⎜ ⎟
⎝ 2T r ⎠
Vr + e -1
Performance SPHCT 69

where Tr represents the reduced temperature T/T*. It should be noted that Ψ u is also density-
dependent. The reduced volume varies from Vr=1 at the closest-packed density to Vr=∞ at zero
density. The temperature function Ψ u can be normalized to 1 at infinite temperatures by
multiplication with a factor 2Vr:
⎡ ⎛⎜⎜⎝ 2 T1 r ⎞⎟⎟⎠ ⎤
2 Vr Tr ⎢ e - 1⎥

⎣ ⎥⎦
Ψ = Ψ u . 2V = r ⎛ 1 ⎞
(4.12)
⎜ ⎟
⎜ 2T ⎟
⎝ r⎠
Vr + e -1

The attractive term of the SPHCT equation of state can now be written as:
attr R cT * * Z m
p =- 2
.V . . Ψ(Tr ,Vr ) (4.13)
V 2

It is important to test the limiting behaviour at high temperatures of the Ψ-function of the
SPHCT equation of state by comparing calculated results with experimental data. However, the
Ψ-function can not be generated from experiments. This problem can be solved by producing the
'experimental' data with a computer. To do so, the equations to produce these data have to be
exact. The pressure equation is an exact relation if three-body and more-body forces are ignored.
This equation has been mentioned when an expression was derived for the repulsive term for the
SPHCT equation of state in chapter 3:

N k T 1 N2 ⎛d u⎞
2 ∫
p= - r⎜ ⎟ g(r) 4π r dr
2
(4.14)
V 6 V r=0 ⎝ d r ⎠

To use Eq.(4.14), we need to select an expression for the intermolecular potential function, u(r),
and the pair distribution function, g(r). A reasonably realistic potential function is the Lennard-
Jones potential function:
⎛ σ 12 σ 6 ⎞
u ( r ) = 4 ε ⎜⎜ 12 - 6 ⎟⎟ (4.15)
⎝r r ⎠

At zero density the radial distribution function is exactly known:


⎛ u( r )⎞
⎜- ⎟
⎝ kT ⎠
g ( r )= e (4.16)

Thus at zero density all functions that are necessary to determine the pressure are known,
although the Lennard-Jones potential function is an approximation. From Eqs.(4.14)-(4.16), the
following expression is obtained for Lennard-Jones molecules at zero density:
70 Chapter 4

⎡ 4 ⎛ ⎛ 1 ⎞12 ⎛ 1 ⎞6 ⎞ ⎤
Ra 8 ⎡ 12 ⎛ 1 ⎞12 6 ⎛ 1 ⎞6 ⎤ ⎢⎢⎣ - Tr ⎜⎜⎝ ⎜⎝ r* ⎟⎠
∞ -⎜ * ⎟
⎝r ⎠




p = - 2 b. π 2 . ∫ ⎢- * ⎜ * ⎟ + * ⎜ * ⎟ ⎥ . e
attr ⎠ ⎦
. r* 3 dr* (4.17)
V 3 ⎢ r ⎝r ⎠
r* = 1 ⎣ r ⎝ r ⎠ ⎦⎥

In Eq.(4.17) the following properties are introduced: a=ε/k, b=Nσ3/√2, r*=r/σ, R=N.k and
Tr=T/a=k.T/ε. The lower limit of the integral has been set to r*=1, because only at r*≥1 does the
intermolecular potential contributes to the attractive pressure. The integral was solved
numerically. In case of infinitely high temperatures, the integral approaches the value 2/3. To
compare the results for Lennard-Jones molecules with those obtained for the SPHCT equation,
we have to normalize the integral. If the value of the integral is normalized to 1 at infinitely high
temperatures, Eq.(4.17) becomes:
R a 16
p
attr,LJ
= - 2 . b. π 2 . Ψ LJ (4.18)
V 9

with:
⎡ 4 ⎛ ⎛ 1 ⎞12 ⎛ 1 ⎞6 ⎞ ⎤
3 ⎡ 12 ⎛ 1 ⎞
∞ 12
6⎛1⎞
6
⎤ ⎢⎢⎣ - Tr ⎜⎜⎝ ⎜⎝ r* ⎟⎠
-⎜ * ⎟
⎝r ⎠




Ψ = ∫ ⎢- * ⎜ * ⎟ + * ⎜ * ⎟
⎠ ⎦
LJ
⎥ .e . r* 3 dr* (4.19)
r* = 1
2 ⎢⎣ r ⎝ r ⎠ r ⎝r ⎠ ⎥⎦

This integral was solved numerically for various values of the reduced temperature, Tr. The
upper limit of r*=∞ was replaced by r*=50. Values for r* higher than 50 did not significantly
contribute to the value of the integral. Values of Ψ LJ are given in Table 4.2.

Table 4.2 :Calculated values for Ψ LJ as a function of the reduced temperature at zero density.

Tr Ψ LJ Tr Ψ LJ Tr Ψ LJ Tr Ψ LJ

0.1 367.569 0.6 1.607 1.2 1.237 20.0 1.012

0.2 8.367 0.7 1.483 1.5 1.181 50.0 1.005

0.3 3.246 0.8 1.400 2.0 1.130 100.0 1.002

0.4 2.214 0.9 1.342 5.0 1.048 200.0 1.001

0.5 1.814 1.0 1.298 10.0 1.023 500.0 1.000

The similarity of the temperature dependence of the SPHCT equation of state with the
data for Lennard-Jones molecules is striking, see Fig.4.6. This similarity becomes even more
apparent if the numerical constants of the attractive terms of both the SPHCT and the Lennard-
Jones equation are evaluated: Zm/2 = 9 and 16/9 π√2 = 7.8985
Performance SPHCT 71

This is another support for the choice Zm=18


instead of the value Zm=36, that was used by
Kim et al. [1]. The results for the Ψ-function of
the SPHCT equation are represented by the solid
curves in Fig.4.6 and the data for the Lennard-
Jones molecules at zero density by open circles.
Each curve is marked by a number that
represents the reduced density, i.e., ρr=V*/V. The
curves were calculated from the SPHCT
equation by choosing c=q=1, i.e., the molecules
are assumed to be spherical. We have also
chosen εSW=εLJ. This assumption is roughly
valid for methane (εSW/k =142.5 K, εLJ/k =
148.9 K, [8]), although εSW can be considerably
smaller or larger than εLJ, e.g., for argon
Figure 4.6 : The Ψ-function of the εSW/k=93.3 K, εLJ/k=117.7 K, [8] and for
SPHCT equation as a tetrafluoromethane εSW/k=191.1 K, εLJ/k= 151.5
function of the reduced K, [8]. Because we do not try to correlate
temperature at various experimental data but show some general trends,
densities. the numerical differences between εSW and εLJ
are of minor concern at this moment. The data
for the Lennard-Jones molecules, represented by the open circles, were taken from table 4.2. It is
shown in Fig.4.6 that the Ψ-function of the SPHCT equation at zero density is in excellent agree-
ment with the data obtained for Lennard-Jones molecules at zero density. Unfortunately, no
comparison can be made for the Ψ-function of the SPHCT equation with Ψ LJ-data at higher
densities in this way. This is because a simple exact expression for g(r) is not available for higher
densities. Moreover, at higher densities three-body and more-body interactions become
important. The maximum of the Ψ-function of the SPHCT equation of state at low values of the
reduced temperature cannot be explained. There is no physical reason to justify this maximum.
However, it is not expected that this maximum plays an important role when phase equilibrium
calculations are made. The maximum only appears at very low reduced temperatures at which
the fluid phase is usually obscured by a solid phase.
It is interesting to compare the Ψ-function of the SPHCT equation of state with the Ψ-
functions of other equations of state. For various equations of state the Ψ-functions are given
below. In Eqs.(4.20)-(4.23) VDW represents the Van der Waals equation, RK the Redlich-Kwong
equation, PR the Peng-Robinson equation and D represents the Deiters equation of state.
72 Chapter 4

VDW
Ψ =1 (4.20)

RK 1
Ψ = (4.21)
Tr

[
Ψ = 1+ κ ( 1 -
PR
Tr ) ]
2
(4.22)

⎡ ⎛


⎟ ⎤
λ ⎢ ⎜ y ( Vr ) ⎟

Tr + ⎜ λ ⎟
D Vr .⎢ e ⎜⎜ Tr + ⎟⎟
⎝ Vr ⎠ ⎥
Ψ = ⎢ - 1⎥ (4.23)
y ( Vr ) ⎢ ⎥
⎢ ⎥
⎣ ⎦
In Fig.4.7 all Ψ-functions are shown as a function of the reduced temperature at zero density. At
zero density the Ψ-function of the Deiters equation is identical to the Ψ-function of the SPHCT
equation of state. In Fig.4.8, a zoomed out view of Fig.4.7, the limiting behaviour of the
attractive term at high temperatures is clearly demonstrated for each equation of state. From
Figs.(4.7) and (4.8) the following conclusions can be drawn:
- For the VDW-equation the Ψ-function is independent of the reduced temperature. Only at
infinitely high temperatures is the value of the Ψ-function correct: Ψ VDW(Tr=∞)=1.

- In case of the RK-equation the Ψ-function has a vertical asymptote at T=0 K. At all other
temperatures the Ψ-function is monotonically decreasing but approaches the wrong horizontal
asymptote at infinitely high temperatures: Ψ RK(Tr=∞)=0. This means that at high reduced
temperatures the attractive pressure vanishes.

- The Ψ-function of the PR-equation has no asymptotes at T=0 K nor at infinitely high
temperatures. At low reduced temperatures it is a decreasing function. For spherical
molecules, i.e., ω=0, the Ψ-function has a minimum at (Tr,Ψ PR)=(13.46,0) after which it starts
increasing at higher reduced temperatures.

- The Ψ-function of the Deiters equation of state (D), [25,26], is, besides the Ψ-function of the
SPHCT equation, the only Ψ-function that is density-dependent. If the Ψ D-function in
Eq.(4.23) is simplified by setting λ=0 and y(Vr)=0.5, which is allowed at zero density, then the
resulting expression for Ψ D is identical to the expression of the Ψ-function of the SPHCT
equation at zero density. The curve that is drawn in Fig.4.7 for the SPHCT equation is also
valid for the Deiters equation. The Ψ D-function is monotonically decreasing and approaches
the right horizontal asymptote at infinitely high temperatures, Ψ D(Tr=∞)=1. At higher
densities the low temperature vertical asymptote is dependent on the density for this equation.
The Ψ D-function does not display the low-temperature maximum at higher densities which
was observed for the SPHCT equation.
Performance SPHCT 73

Figure 4.7: The Ψ-functions of various Figure 4.8: A zoomed out view of Fig.4.7.
equations of state as a func- The limiting behaviour of the
tion of the reduced tempera- various equations at zero den-
ture at zero density. sity and high temperatures is
clearly demonstrated.

The most important conclusion that can be drawn is that the SPHCT and the Deiters equation of
state are the only equations of state examined that possess an attractive term that fullfills the
condition of a non-zero horizontal asymptote at infinitely high temperatures. It is a surprise to
find that the equations that are normally used in chemical engineering, such as the PR- and the
RK-equation, do not fullfill this high temperature boundary condition at zero density. It is
dangerous to use those equations for extrapolation over a large temperature range, especially if
molecules with low characteristic temperatures, e.g., H2 and N2, are involved.

4.3 Virial coefficients from the SPHCT

In the previous section the high temperature limit of the SPHCT equation of state was
tested. In this section we will investigate the low density limit of the SPHCT equation. The
equation of state for an ideal gas is p=ρkT, with ρ=N/V, the particle density. Any equation of
state for real gases can be expanded in a series of ρ:

p
Z= = 1 + B ( T ).ρ + C ( T ). ρ 2 + ..... (4.24)
ρkT

B(T) is called the second virial coefficient, C(T) the third virial coefficient, etc. For a given gas
the virial coefficients are functions of the temperature only. It has not only a simple mathematical
74 Chapter 4

form but it also has a statistical mechanical background [27]. The methods of statistical
mechanics allow derivation of the virial equation and provide physical significance to the virial
coefficients [27]. Thus, for the expansion in ρ, the term B.ρ arises on account of interactions
between pairs of molecules, the C.ρ2 term on account of three-body interactions, etc. Since two-
body interactions are many times more common than three-body interactions, and three-body
interactions are many times more numerous than four-body interactions, etc., the contribution to
the compressibility factor Z of the successively higher-ordered terms falls off rapidly.
The virial equation, given by Eq.(4.24), is an infinite series. For engineering purposes,
use of the virial equation of state is practical if only two or three terms are required instead of the
infinite series. This is realistic for gases and vapours at low to moderate pressures. The virial
equation truncated to two terms represents the p,V,T-behaviour of most vapours at subcritical
temperatures up to a pressure of about 15 bar. At higher temperatures the two-term virial
equation is appropriate for gases over an increasing pressure range as the temperature increases.
For pressures up to 50 bar the virial equation truncated to three terms usually provides acceptable
results [28]. Since virial coefficients beyond the third are rarely known, virial equations of more
than three terms are rarely used.
If an equation of state, like the SPHCT equation of state, is at hand there is no need to use
a virial equation for the calculation of p,V,T-behaviour. However, every equation of state should
reduce in its low pressure limit to the virial equation. This makes it interesting to look at the
values of the virial coefficients as calculated from the SPHCT equation of state. The second and
third virial coefficient can be calculated from any equation of state according to the following
equations:

( Z -1 )
B = lim (4.25)
ρ →0 ρ

( Z -1 )

ρ
C = lim (4.26)
ρ →0 ∂ρ

For the SPHCT equation of state, application of Eqs.(4.25) and (4.26) leads to:

B = 4 < c > τ < V * > - Z m < cV * Y > (4.27)

10 (< c > τ < V * >) + Z m < cV * Y >2


2

C= (4.28)
<c>
Performance SPHCT 75

Figure 4.9 : Second virial coefficients as a Figure 4.10 : Third virial coefficients as a
function of temperature for function of temperature for
methane, ethane and propane. methane and ethane. For an
For an explanation of symbols, explanation of symbols, see
see text. text.

Eq.(4.27) shows that the composition-dependence of the second virial coefficient is quadratic.
This is in agreement with theory, see section 3.6. For methane, ethane and propane the calculated
values of the second and third virial coefficients are shown as a function of temperature in
Fig.4.9 and 4.10 and compared to experimental data. The calculations made with the SPHCT
equation are represented by the solid curves in these figures. The number that marks each curve
represents the carbon number of the alkane. The dashed curves are calculations that are made
with the Peng-Robinson equation. The experimental data, represented by the open circles, were
taken from Dymond and Smith [29]. It can be concluded from Fig.4.9 and Fig.4.10 that the
agreement with experimental data is good for the three alkanes and slightly better than the
calculations with the well-known Peng-Robinson equation of state. However, the maximum in
the third virial coefficient is not calculated by either of the equations of state.

4.4 The isochoric heat capacity

The final pure component thermodynamic property that will be calculated with the
SPHCT equation of state in this section is the molar isochoric heat capacity (or molar heat
capacity at constant volume), Cv. We chose this thermodynamic property for a special reason, i.e.
the isochoric heat capacity is one of thermodynamic properties of a pure fluid that diverges at the
76 Chapter 4

critical point [4]. In this section not only will the values of Cv calculated from the SPHCT
equation be compared to experimental data, but also the behaviour of the calculated Cv around its
point of divergence will be considered. The isochoric heat capacity can be defined as the quantity
of heat necessary to increase the temperature of one mole of a substance by one degree. The Cv
for a monoatomic substance, e.g., argon, can be calculated from any equation of state from the
following equation [4]:

3 ⎛∂2 p⎞
C v = R - T ∫ ⎜⎜ ⎟ dV
∂ T 2 ⎟⎠V
(4.29)
2 V ⎝

When Eq.(4.29) is applied to the SPHCT equation of state, we obtain the following expression
for Cv:
2
3 c (Vr - 1)(Y + 1) ⎛ 1 ⎞
Cv = R + R Z m ⎜ ⎟⎟ (4.30)
2 (Vr + Y )2 ⎜⎝ 2 Tr ⎠

where Vr=V/V* and Tr=T/T*. It is stressed that Eq.(4.30) is valid only for pure monoatomic
fluids.

Figure 4.11: The isochoric heat capacity as Figure 4.12: Calculated results of the iso-
a function of the density at choric heat capacity as a func-
various temperatures. For an tion of the density for various
explanation of symbols, see temperatures. For an explana-
text. tion of symbols, see text.

In Fig.4.11 the calculated values of Cv for argon (solid curves) at supercritical


temperatures are compared with experimental data (open circles, boxes and triangles) as a
function of density, ρ. The experimental data were taken from [6]. The black dots represent the
Performance SPHCT 77

experimentally determined values of Cv along the vapour pressure curve of argon. The left
branch represents the saturated vapour and the right branch the saturated liquid. At the critical
density there is a vertical asymptote. The dotted curve in Fig.4.11 represents the vapour pressure
curve that is calculated from the SPHCT equation. It is obvious that the calculated isotherms and
the vapour pressure curve are a long way from the experimental data. It is also shown that the
calculated curves do not show the divergence in Cv at the critical density: the value of Cv at the
critical point is 16.50 J.mole-1.K-1. In Fig.4.12 the calculated isotherms (solid) and the vapour
pressure curve (dotted) are shown over a larger density range. The experimental data have been
left off this figure. It is shown that the limiting value of Cv at closest packing is identical to the
value at zero density: 3R/2. There is no theoretical justification for the limiting value of 3R/2 at
the closest-packed volume. Before we briefly digress on the divergence problem, we can
conclude that the calculation of some thermodynamic properties, like Cv's, is beyond the
possibilities of the SPHCT equation of state.
According to Eq.(4.30) the molar isochoric heat capacity is a continuous function of
density and temperature. However, it is discussed by Rowlinson and Swinton [4] that in reality at
the critical point Cv should approach infinity. This qualitative error in the description of this
property not only holds for the SPHCT equation of state, but appears in every classical equation
of state. So, it should be noted that with a classical equation of state, like the SPHCT equation,
the molar isochoric heat capacity is not only badly described in the critical region, but the
description turns out to be fundamentally wrong. It is another aspect of the same problem that
was already mentioned in section 4.1, when the calculated densities along the coexistence curve
near the critical point were discussed. These fundamental limitations of analytic or classical
equations of state led to the development of the so-called nonclassical equations. Eqs.(4.31) and
(4.32) are "simple scaling" equations and are thus examples of non-classical equations of state
[30]:
p p
= c [ 1 + C 3 (-t )2-α + C 4 t + C 5 t 2 + C 6 t 3 + ... ] (4.31)
RT RTc

ρ = ρ c [ 1 + _ C 1 (-t )β + C 2 t ] (4.32)

where t=(T-Tc)/Tc and where α is the critical exponent characterizing the divergence of the molar
isochoric heat capacity on the one-phase critical isochore:

C v = C v,0 (-t ) (4.33)

The critical exponent β characterizes the curvature of the phase boundary in the T,ρ-plane.
Calculated from an analytic equation of state α=0 so that Cv does not diverge, which is in
contradiction to experiment. Careful experiments on fluids extremely close to the critical point
have shown that α=0.110±0.001 and β=0.325±0.001, in contrast to the classical values, 0 and 0.5
respectively. However, this subject is outside the scope of this thesis and it will not be dealt with.
A recent overview of non-classical equations of state can be found in Bruno and Ely [30].
78 Chapter 4

4.5 Binary flash calculations

In the previous section several pure component properties were discussed that can be
predicted with the SPHCT equation of state. In the following sections some properties of binary
mixtures are discussed. Probably the most important property that can be described with the
SPHCT equation is the equilibrium state between two (or more) coexisting phases. If a closed
system containing several components existing in many phases is left for a sufficiently long time,
it will reach a state where the macroscopic properties of the phases will not change in time. Of
course on microscopic scale changes will continue to occur. Such a state is called equilibrium.
The criteria for equilibrium between the phases can be found in numerous textbooks on
thermodynamics, e.g., [31]:

1. The temperatures of the coexisting phases are equal.


2. The pressures of the coexisting phases are the same.
3. A component must have the same chemical potential, μi, in each coexisting phase.

In the case of a vapour and a liquid phase the criteria are:


L V
T =T (4.34)

L V
p =p (4.35)

μ iL = μVi i = 1, 2, ..., n (4.36)


The first two criteria, which represent thermal and mechanical equilibria, seem almost trivial and
often they are not stated. Introducing fugacities, the third condition for each component i
becomes:

L V
f i ( p, T, xi ) = f i ( p, T, yi ) (4.37)

Two equations define the fugacity of component i, see [31]: dμi=RTdlnfi and fi/yip→1 if p→0. If
the pressure approaches zero then the vapour phase behaves according to the ideal gas law.
Integration of the first equation from p=p0 to p leads to:

⎛ f ⎞ ⎛ f ⎞
μ i ( p, T, x ) = μ i ( p0 , T, x ) + R T ln ⎜⎜ i ⎟⎟ = μ i ( p0 , T ) + R T ln ⎜⎜ i ⎟⎟ (4.38)
⎝ x i p0 ⎠ ⎝ p0 ⎠

where p0 is a pressure that is very close to zero. The subscript 0 refers to the ideal gas state. In
Eq.(4.38) xi is removed from the logarithmic term because μi(p0,T,x)=μi(p0,T)+RTlnxi. The latter
equality is valid for an ideal gas. It is proven by Eq.(4.38) that Eq.(4.37) is equivalent to
Eq.(4.36). The fugacity of component i in each phase can be written as [31]:

f i ( p, T, xi ) = xi φ i p
L L
(4.39)
Performance SPHCT 79

f i ( p, T, yi ) = yi φ i p
V V
(4.40)

The fugacity coefficients φiL and φiV can be calculated from any equation of state with the
following exact thermodynamic relationships [31,32]:
⎡⎛∂ p⎞

RT ⎤
R T ln φ = ∫ ⎜
L
i
⎢ ⎜ ⎟
⎟ - ⎥ dVt - R T ln Z L (4.41)
Vi L ⎢
∂n Vt ⎥
⎣ ⎝ i ⎠V t ,T,n j ⎦

∞ ⎡⎛∂ p⎞ RT ⎤
R T ln φ = ∫ ⎢ ⎜⎜ ⎟⎟ ⎥ d Vt - R T ln Z V
V
i - (4.42)
⎢ ⎝ ∂ ni ⎠V t ,T,n j Vt ⎥
ViV ⎣ ⎦

The following expression results for the calculation of the fugacity coefficient of component i in
a mixture of n components according to the SPHCT equation of state [1]:

4 η - 3 η 2 < ci > V *i 4 η - 2 η 2
ln φ i = ci +
( 1 - η )2 < V * > ( 1 - η )3

⎡ c c ⎛ < cV *Y > ⎞⎤ (4.43)


- < c > Z m ⎢ 2 - i + i ln ⎜⎜ 1 + ⎟⎟ ⎥
⎣ <c> <c> ⎝ < c >V ⎠⎦


[ ] ⎤
n

⎢ V ( c i - 2 < c > ) + ∑ x j ci V ji Yij + c j Vij Y ji - 2 < cV Y > ⎥


* * *

- < c > Zm ⎢
j=1 ⎥ - ln Z
⎢ < c > V + < cV *Y > ⎥
⎢ ⎥
⎣ ⎦
The reduced density η was already defined in chapter 3 as: η=τ.<V *>/V and the property Yij in
the fourth term of Eq.(4.43) is equal to:

⎛ ε ij qi ⎞
Y ij = exp ⎜⎜ ⎟⎟ - 1 (4.44)
⎝ 2 ci k T ⎠
With this equation for the fugacity coefficients we are able to perform phase equilibrium
calculations. With Eqs.(4.43) and (4.44), the distribution coefficients between the vapour and the
liquid phase for each component are calculated:

yi φi
Ki = = (4.45)
xi φ Vi

Subsequently the calculation of the vapour and liquid phase compositions is made according to
the following scheme. The calculated distribution coefficients are used in an equilibrium stage
80 Chapter 4

flash calculation at a fixed pressure and temperature [31-33]. In this calculation a total feed of F
moles with an overall composition zi in the two-phase region is considered. The material balance
equations over the equilibrium stage are:

F = L +V (4.46)

zi . F = xi . L + yi .V (4.47)

Solving these equations for xi and yi by introducing α=V/F and the distribution coefficient Ki
gives:
zi
xi = (4.48)
( Ki - 1 )α + 1

zi . K i
yi = (4.49)
( K i - 1 ).α + 1

Since both sets of mole fractions must sum up to unity, the difference F must equal zero:
n n n
( K i - 1 ).z i
F ( α , z i ) = ∑ y i - ∑ xi = ∑ (4.50)
i=1 i=1 i = 1 ( K i - 1 ).α + 1

The benefit of this objective function is apparent from its derivative with respect to the vapour-
to-feed ratio α:

⎛∂ F ⎞ n
( K i - 1 )2 .zi
⎜ ⎟= - ∑ (4.51)
⎝ ∂ α ⎠ i = 1 [ 1 + ( K i - 1 ).α ]2
Since ∂ F(α)/∂ α is always negative a (step-limited) Newton-Raphson method can be used to
determine the value of α for which F(α)=0 at fixed distribution coefficients Ki and overall
compositions zi. Bubble point and dew point calculations [33] provide an initial estimate for α.
The split is estimated as:

T - T BubblePoint
α= (4.52)
T DewPoint - T BubblePoint

Fig.4.13 shows the flow diagram for a typical isothermal flash calculation. Five to ten iterations
are usually sufficient for a wide-boiling mixture. Considerably fewer iterations are required for a
close boiling mixture.
Three binary mixtures were selected from literature to test the validity of the SPHCT
equation of state with respect to the prediction of isothermal phase diagrams. Data of the two
binary mixtures nitrogen(1) + argon(2) and nitrogen(1) + methane (2) were taken from Miller et
al. [34]. The pure component parameters for the SPHCT equation of state were taken from Table
4.1. For both systems the binary interaction coefficients, k12, were set to 0.0. Fig.4.14 and
Fig.4.15 show the results for the systems nitrogen(1) + argon (2) and nitrogen(1) + methane (2)
Performance SPHCT 81

respectively. In both figures the experimental data


are represented by the open circles and the calcu-
lated results are represented by the solid lines. For
the system nitrogen(1) + argon(2) good agreement
was obtained between the experimental data and
the calculated results.
For the system nitrogen(1) + methane(2)
the agreement is worse, especially when the
mixture is rich in methane. The description can be
improved by introduction of a binary interaction
parameter. However, in this case, the errors in the
calculated results are caused by a bad prediction
of the saturated vapour pressure of pure methane
at the specified temperature. This is due to the fact
that the pure component parameters, c, V* and T*
for methane were optimized on the temperature Figure 4.13: Flow diagram for isothermal
range 143-190.5 K. flash calculation.

Figure 4.14: p,x,y-Section of nitrogen(1) + Figure 4.15: p,x,y-Section of nitrogen(1) +


argon(2). Experimental data methane(2). Experimental
were taken from [32]. data were taken from [32].

It can be seen from Fig.4.15 that both the dewpoint and the bubblepoint curve will fall into place
if the saturated vapour pressure of methane is moved a little upwards. This demonstrates the
importance of an equation of state giving an accurate description of the pure component vapour
82 Chapter 4

pressure curves as a function of temperature.


A similar calculation was carried out for
the binary mixture methane(1) + ethane(2) at
four different temperatures: 158.15 K, 172.04 K,
186.11 K and 199.92 K, see Fig.4.16. The pure
component parameters for the calculations of
this binary mixture were taken from table 4.1.
The binary interaction coefficient was set to 0.0
for each temperature. The solid circles in
Fig.4.16 represent the experimental bubble point
data, while the open circles represent the
experimental dewpoint data. The experimental
data of this binary system are reported by
Wichterle et al. [35]. The agreement between the
calculated results and the experimental data is
good. It was already mentioned that the SPHCT
equation of state was developed for mixtures
with large size differences. An extensive study
of the description of isothermal phase diagrams Figure 4.16: p,x,y-Section of methane(1) +
with the SPHCT equation of state was ethane(2). Experimental data
performed by Rijkers [28] for the systems were taken from [33].
methane(1) + n-decane(2), + n-dodecane(2), + n-hexadecane(2) and methane(1) + benzene(2).
Rijkers [28, p.96] concluded that "The SPHCT equation performs best in describing binary l+g
data when its parameters for methane are tuned on vapor pressure data. This especially holds for
the restricted pressure range [0.1 ≤ p ≤ 10 MPa]."

4.6 Calculation of Henry coefficients

Flash calculations in combination with an equation of state provide a powerful method


for determining the vapour and liquid compositions in coexisting phases. Another way to
perform phase equilibrium calculations is via the activity coefficient or γ,φ-method. According to
the γ,φ-method, the vapour phase is described with an equation of state and the liquid phase with
an excess model. One of the applications based on this concept is the already mentioned
calculation of gas solubilities in liquids. This method can only be used for the calculation of
isothermal phase envelopes at temperatures below the critical temperatures of both components.
The equilibrium relation now becomes [31]:
Performance SPHCT 83

p
Vi sat

,L
RT
p isat
xi γ i p i
sat
θ ( p ,T) e
i
sat sat
i dp = yi θ i (p,T, yi ) p (4.53)

From this equilibrium relation a number of factors can be ignored under certain conditions. If the
vapour phase can be described with the ideal gas law then: θi(p,T,y)=1 and θisat(pisat,T)=1. The
exponential factor in Eq.(4.53), also called the Poynting-correction, is near 1, especially at low to
moderate pressures. If both components are chemically more or less similar and non-polar, like
propane and n-butane, the activity coefficients are near 1 over the whole concentration range.
Eq.(4.53) now reduces to Raoult's law. This simplified equilibrium relation is the simplest way to
perform phase equilibrium calculations, including the calculation of the solubility of gases in
liquids:

xi . pisat = yi . p i = 1, 2,..., n (4.54)

Another way to estimate the solubility, xi, of a gas in a liquid is Henry's law [31]. Henry's law has
different forms dependent on the pressure and composition ranges where the gas solubility is to
be calculated. A recent overview on the use of Henry's law is given by Carroll [36]. The
mathematical expression of the most used form, the strict Henry's law, is:

xi . H ij = yi . p i = 1, 2,..., n (4.55)

This expression is good when gas solubilities less than 1 mole percent are to be calculated at
pressures below 2 bar. The Henry constant, Hij, is not a pure component property. In binary
mixtures, the Henry coefficient H21 (component 2 is infinitely diluted in component 1) is defined
as:

⎛ f2 ⎞
H 2,1 = ⎜ ⎟ (4.56)
lim ⎜ ⎟
xi → 0, p → pisat x2
⎝ ⎠

In Eq.(4.56), component 2 represents the gaseous component or solute, and component 1 is the
liquid solvent. From Eq.(4.55) it can be seen that:
i the solubility of a gas, x2 in Eq.(4.55), is proportional to the partial pressure of this gas.
ii the solubility of a gas is inversely proportional to the value of the Henry coefficient.

To use Eq.(4.55) for calculation of the gas solubility in a binary system the value of H21 is
needed. Henry coefficients can be obtained, according to Eq.(4.56), by extrapolation of
experimental data of f2/x2 as a function of x2 to x2=0. Another way to obtain Henry coefficients,
is by using an equation of state [37]:
84 Chapter 4

f 2L x2 φ2L p
lim H 21 = lim = lim = φ2L . p1sat (4.57)
x2 →0, p → p1sat x2 →0, p → p1sat x2 x2 →0, p → p1sat x2

The fugacity coefficient of the solute (2) in the liquid phase is calculated with the equation of
state at the saturated molar liquid volume of the solvent (1). The saturated vapour pressure is also
calculated with the equation of state. Because the Henry coefficient is defined along the vapour
pressure curve of the solvent, the pressure is not kept constant if the Henry coefficient is
calculated as a function of the temperature.

Figure 4.17: Henry coefficients of methane Figure 4.18: Henry coefficients of ethane
in five solvents as a function of in six different solvents as a
temperature function of temperature.

With the SPHCT equation of state the Henry coefficients of methane, ethane, propane,
nitrogen and carbon dioxide in various solvents have been calculated, see Fig.4.17-4.21. The
calculated results are represented by the solid lines in Fig.4.17-4.21, and the open symbols
represent the experimental data. The experimental data were taken from [38]. These gases and
solvents are commonly encountered in natural gas processing. The binary interaction coefficients
were obtained by optimization on experimental Henry coefficient data. The following objective
function has been used to minimize the absolute average deviation, AAD:
calc
1 N
| H 21,i - H 21,i |
exp

AAD =
N

i=1 H 21
exp
(4.58)
,i

The optimized values for the binary interaction coefficients are given in table 4.3.
Performance SPHCT 85

Figure 4.19: Henry coefficients of propane in Figure 4.20: Henry coefficients of nitrogen
three solvents as a function of in three different solvents as a
temperature. function of temperature.

Figure 4.21: Henry coefficients of carbon


dioxide in three different solvents
as a function of temperature.
86 Chapter 4

Table 4.3: Optimized binary interaction coefficients for the description of Henry coefficients for
components encountered in natural gas processing. The components are
abbreviated as: n-Cn for the normal alkanes, c-Cn for the cycloalkanes and BNZ for
benzene.

System 1000.k12 System 1000.k12

C1 + C2 -7.458 C3 + n-C4 5.779

C1 + C2 5.887 C3 + n-C5 2.013

C1 + n-C5 25.146 C3 + BNZ -3.194

C1 + H2S 20.539

C1 + CO2 -8.346 N2 + C1 22.444

N2 + C2 5.042

C2 + C3 -13.480 N2 + n-C4 17.881

C2 + n-C4 -0.285

C2 + n-C5 -0.453 CO2 + C3 107.886

C2 + n-C7 18.713 CO2 + n-C4 110.002

C2 + c-C6 3.622 CO2 + H2S 63.449

C2 + BNZ 11.055

The agreement between the calculated Henry coefficients and the experimental data is
reasonable. At low temperatures, the trend that Henry coefficients increase with increasing
temperatures, is correctly predicted. In general, the gas solubility in a liquid decreases at
increasing temperatures. Because the Henry coefficient is inversely proportional to the gas
solubility xi, the value of the Henry coefficient should increase with increasing temperature.
However, it is seen from Figs.(4.17)-(4.21) that Henry coefficients can decrease at higher
temperatures. According to Gmehling and Kolbe [37] it is possible that the partial molar enthalpy
of the gaseous solute is higher in the liquid phase than it is in the gas phase. This causes the
solubility of the gas in the liquid to increase if the temperature is raised. We should bear in mind
that the prediction of Henry coefficients is an extremely stringent test of the theory. Errors in the
mixing rules are exponentially magnified in the value of the Henry coefficient [13].
Performance SPHCT 87

4.7 Calculation of critical curves in binary mixtures

The final test for the SPHCT equation of state concerns its ability to describe critical
curves. We will restrict ourselves here to binary hydrocarbon mixtures. It has already been
shown in chapter 2 that the critical curves and the way they are connected to three-phase lines,
determine the type of phase behaviour, according to the classification of Van Konynenburg [39].
For this reason, we take a particular interest in the quality of the calculated critical curves and to
this topic more attention will be given than to the previous tests.

4.7.1 Introduction

The thermodynamic conditions of critical points in a binary mixture in terms of the molar
Gibbs energy, G, can be found in many textbooks on thermodynamics e.g., [4, chap.6]

⎛ ∂ 2G ⎞ ⎛ ∂ 3G ⎞
⎜⎜ ⎟ =0 ,
2 ⎟
⎜⎜ ⎟ =0
3 ⎟ (4.59)
⎝ ∂ x ⎠ p,T ⎝ ∂ x ⎠ p,T

The composition variable, x, represents the mole fraction of the heavier component, i.e.
component 2. In the following the abbreviated notation for multiple differentiations of [4] is
used:

⎛ ∂ iG ⎞
Gix ⎜⎜ i ⎟⎟
= (4.60)
⎝ ∂ x ⎠ p,T

According to this notation, the critical conditions can be written as: G2x=0 and G3x=0. In
Eq.(4.59) the partial derivatives of G are used to define the criteria for the critical state in a
binary mixture. However, G has pressure, temperature and composition (p, T and x) as
independent variables. Because most equations of state contain the molar volume V, T and x as
independent variables, a Helmholtz energy formulation of the critical conditions is to be
preferred. The molar Helmholtz energy, A, also has V, T and x as independent variables. The
molar Gibbs energy is related to the molar Helmholtz energy, in accordance with the following
fundamental equation: dG=dA+d(pV). The critical conditions can now be written down in terms
of molar Helmholtz energy derivatives. To perform the transformations from the molar Gibbs
energy criteria to the molar Helmholtz energy criteria the Jacobian Transformation procedure is
used [40].

⎛ AVx2 ⎞
G2 x = A2 x - ⎜⎜ ⎟⎟ (4.61)
⎝ A2V ⎠

2 3
⎛A ⎞ ⎛A ⎞ ⎛ AVx ⎞
G3 x = A3 x - 3 AV 2 x ⎜⎜ Vx ⎟⎟ + 3 A2Vx ⎜⎜ Vx ⎟⎟ - A3V ⎜⎜ ⎟⎟ (4.62)
⎝ A2V ⎠ ⎝ A2V ⎠ ⎝ A2V ⎠
88 Chapter 4

To determine the critical points we therefore need an expression for A in terms of V, T and x. The
molar Helmholtz energy can be obtained from any pressure explicit equation of state via
Eq.(4.63):

A ( V, T, x ) = ( 1 - x ) A*1 ( V 0 , T ) + x A*2 ( V 0 , T ) +
V
(4.63)
R T [ ( 1 - x ) ln( 1 - x ) + x ln( x)] - ∫ p dV
V0

where A1*(V0 ,T, x) and A2*(V0 ,T, x) are the pure component molar Helmholtz energies at the
specified conditions V0, T and x in the ideal gas state. The integration of the pressure over the
volume is done from V0, the reference volume at p=1 bar in the ideal gas state, to V, the volume
at the desired pressure in the real gas state. The real gas state in our investigation is described by
the SPHCT equation of state. Because the SPHCT equation of state reduces to the ideal gas law
at infinite molar volume, the integral in Eq.(4.63) can be separated in accordance with Eq.(4.64):

V V ∞ V
- ∫ p dV = - ∫ p SPHCT
dV -∫p id
dV -∫p id
dV (4.64)
V0 ∞ V V0

By performing the differentiations of A with respect to x, the reference volume, V0, will disappear
from the molar Helmholtz derivatives. Accordingly, the reference volume will not influence the
actual value of these molar Helmholtz derivatives. The partial derivatives of the Helmholtz
energy with respect to volume and mole fraction were easily obtained with the computer algebra
program MAPLE. This extremely useful tool will be discussed in section 6.4. The expressions
for all molar Helmholtz derivatives according to the SPHCT equation of state, can be found in
Appendix A.

Both equations in Eq.(4.59) are highly non-linear. Consequently they cannot be solved in
a straightforward manner for temperature, volume and composition. An iterative technique is
required in which one of the variables (V, T or x) is fixed and the other two are varied to obtain
values at which Eqs.(4.61) and (4.62) are satisfied. This will be discussed in the next section.

4.7.2 The Hicks-Young algorithm

Simultaneous solution of equations in N dimensions is much more difficult than finding


roots in the one dimensional case. The principle difference between one and many dimensions is
that in one dimension it is possible to bracket or 'trap' a root between bracketing values. A root is
bracketed in the interval [a,b] if f(a) and f(b) have opposite signs. If this is the case and if the
one-dimensional function is continuous, then at least one root must lie in that interval. Finally the
root can be located to any desired accuracy within these bracketing bounds with a suitable
Performance SPHCT 89

iterative method, e.g., the secant method or the bisection method. If a function depends on only
one variable, i.e., the one-dimensional case, this 'bracket-iterate' idea is highly recommended by
Press et al. [41]. Rowlinson and Swinton [4, chap.8] recommended the Hicks-Young algorithm
for the calculation of critical points in binary fluid mixtures. In essence, the calculation of critical
points in a binary fluid mixture comes down to finding a root in two dimensions. Hicks and
Young [42] proposed an algorithm for the two-dimensional case that is an extension of the
'bracket-iterate' concept for the one-dimensional case. We modified their algorithm by a
simplification that reduces the number of iterations, necessary to calculate a critical point in a
binary mixture.
The two conditions that define a critical
point in a binary fluid mixture are given by
Eq.(4.59). If a value for the composition x is
chosen this system of equations can be solved.
Hicks and Young [42] considered a specified
rectangle in the volume-temperature plane. This
rectangle is called the 'search area'. Both
functions G2x and G3x have zero contour curves.
These zero contour curves divide the V,T-plane
into regions where the functions are positive and
negative, see Fig.4.22. The solution(s) of our set
of equations that we seek are those points (if any
are present!) which are common to the zero
contour curves of both G2x and G3x. To find all
common points we must map out the full zero
contour curves of both functions. Hicks and Figure 4.22: Zero contour curves of G2x
Young [42] proposed a method to trace one of (solid) and G3x (dashed) in
the V,T-plane.
the zero contour curves, G2x=0, in the specified
search area from its entry point to its exit point. At each subsequent step the sign of the other
function, G3x, is evaluated. A change of sign of the function value of G3x along the zero contour
curve of G2x indicates a critical point. They split the problem into three parts:

(i) locating the entry and exit point(s) of the 'G2x zero contour curve' on the boundary of the
search area
(ii) tracking the 'G2x zero contour curve' in the search area
(iii) more precise location of the solution

These three items will be elaborated in the next paragraphs.


90 Chapter 4

The location of the entry points

Entry points of the G2x zero contour curve on the border of the search area in the V,T-
plane are located by stepping around the boundary of the search area with suitably small steps
ΔV and ΔT. The sign of the function value of G2x is evaluated at each step. A change of sign in
the value of G2x locates a root on the boundary of the search area. Subsequent, repeated bisection
of the interval will locate the position of this root (entry point) to any desired accuracy. This 'root
trapping' procedure on the boundary of the search area has the disadvantage that roots in G2x that
are too close to each other (less than the initial stepsize) may both be missed altogether. During
the 'tracking' procedure the zero contour curve of G2x will be traced from its entry point to its exit
point. A list is kept of all the entry points. To avoid tracing the same curve twice entry points are
checked against this list before they are used.

The tracking procedure

Hicks and Young [42] proposed to


track each curve by taking a series of small
steps along the curve. We propose a method
that is simpler and reduces the number of
iterations drastically. Our method shows some
similarities with the Simplex minimization
algorithm. Around an entry point a small 'two-
dimensional' box or rectangle is constructed,
see Fig.4.23. At each corner of the box,
numbered 1 to 4, the value of G2x is calculated.
The box has sides of length ΔV and ΔT. The
sign of the value of G2x is compared at 1 and
2, 2 and 3, 3 and 4, 4 and 1. If the sign
changes between any pair then the zero
contour curve of G2x passes between them an
odd number of times, i.e., there is at least one
local exit point between the pair. In general Figure 4.23: Schematic drawing of the
the size of the rectangle should be small overturning box in the search
enough so that there is only one entry side and area and the more precise
one exit side. Once the entry and exit side are location of a critical point.
determined the sign of the value of G3x is
evaluated at each corner of this box. If the sign
of G3x is the same at each corner it can be concluded that the zero contour curve of G3x does not
pass through the box. Now we construct a new box of the same size against the exit side of the
present box. The present box and the new box have two corners in common and the function
values of G2x and G3x can be stored. The exit side of the present box will be the entry side of the
Performance SPHCT 91

new box. With this procedure the construction of the new box only requires the function
evaluations of G2x and G3x at the two new corners. By comparing the signs of the values of G2x
the exit side of the new box is determined. Again a change of sign of the values of G3x at the
corners of the new box indicates the presence of the zero contour curve of G3x in the new box. If
there is no change of sign in the values of G3x, a third box is constructed next to the exit side of
the second box. This procedure is repeated until the tumbling box leaves the search area. The
entry point and exit point will be removed from the list of entry points that was made by stepping
around the boundary of the search area. In this way all entry points are traced to their exits. This
global two-dimensional root bracketing is simple and fast. For each new box the signs of the
values of G2x and G3x have to be evaluated at the two new corners, i.e., four function evaluations
have to be made per step. If a change of sign of the value of G3x is detected on the corners of a
box it is certain that the zero contour curve of G3x passes through the box. It may be possible that
the zero contour curves of G2x and G3x intersect in the box. Consequently, a closer inspection of
this box is necessary.

Convergence to the root's value

If the zero contour curve of G3x passes through the box the entry points and exit points of
the zero contour curves of G2x and G3x on boundary of the box are located precisely. Both zero
contour curves in the box are approximated by linear interpolations between their respective
entry and exit points. If these two curves intersect it is certain that an intersection of both zero
contour curves occurs in the box. If not, both contour curves are more or less parallel inside the
box and a new box is constructed adjacent to the exit side of the present box. If a crossing of both
contour curves is to be expected the coordinates of the present box and corresponding
information about the entry side and exit side of the zero contour curve of G2x is stored. A
smaller box is constructed around the entry point of the zero contour curve of G2x in the present
box. The lengths of the sides of this new box are one-tenth of the original side lengths. This is
schematically shown in Fig.4.23. With this smaller box the same procedure is repeated as is
described in the previous section, i.e., the entry side and exit side of the zero contour curve of
this smaller box are determined by evaluating the signs of G2x at the corners of this smaller box.
Thereafter the signs of G3x are evaluated at the corners of this smaller box. If there is no change
of sign in G3x at the corners of this smaller box, a new (small) box is constructed next to the exit
side of the present box. If the zero contour curve of G3x passes through the smaller box, then the
box sides are reduced again by a factor ten. This procedure is repeated until the intersection point
of both contour curves is trapped in a box that has sides that are smaller than the desired accuracy
in temperature and volume. Linear interpolation between the entry and exit point of the zero
contour curves of G2x and G3x on the boundary of this final box, gives the values of the volume
and temperature of the critical point at this specified composition. The equation of state gives the
corresponding pressure. The original box is restored with the original side-lengths ΔV and ΔT
and a new box is constructed next to the present one. The zero contour curve of G2x is then traced
92 Chapter 4

until a new critical point at the same composition is detected or until the tumbling box leaves the
search area.

The Hicks-Young algorithm with the modification that we proposed is simple and
elegant. It performed well for the calculation of critical points from the SPHCT equation.
However, mathematically more sophisticated algorithms can be used. Especially if the
elimination of the temperature from the critical conditions is feasible, other solution methods are
recommended, as is the case for the Van der Waals and the Redlich-Kwong equation [43]. These
convenient algebraic manipulations shorten the computation time significantly. In case of the
SPHCT equation of state the elimination of temperature from the critical conditions is not
possible because of the exponential temperature dependence of the numerator and the
denominator in the attractive term of the equation of state.
We want to make the following comments on the (modified) Hicks-Young algorithm. If
the zero contour curve of G2x is a closed curve that is situated completely inside the search area,
critical points may be missed. However, this problem might be overcome by redefining the
search area. A disadvantage of the (modified) Hicks-Young algorithm is its relatively slow
convergence because numerical differentiation of critical conditions, Eq.(4.61) and (4.62), with
respect to the iteration variables, V and T, is not used. However, the lack of use of numerical
derivatives of Eq.(4.61) and (4.62) causes the algorithm to be numerically stable. This algorithm
offers two other advantages. The first advantage is the possibility to build in other equations of
state into the program because no equation of state specific manipulations have been performed.
The second, more important, advantage is the fact that initial estimates for the critical points are
not needed: this makes the Hicks-Young algoritm especially suitable for scanning large V,T-
areas at a specified composition for possible solutions of the critical conditions.

4.7.3 Stability tests

All critical points that have been calculated should be subjected to three stability tests.
These stability tests distinguish between thermodynamically stable, metastable and unstable
critical points. The stability test can be subdivided in local stability tests (i.e., the tests on
mechanical and material instability) and a global stability test.

The mechanical stability test

If (∂ p/∂ V)T,x>0, i.e., A2v<0, then the critical phase is said to be 'mechanically unstable'.
A (small) density fluctuation will cause the critical phase to split into a liquid and a vapour
phase. Mechanical instability can only occur in azeotropic mixtures, if the liquid and the
coexisting vapour phase, both with the azeotropic composition, become critical [4].
Performance SPHCT 93

The material stability test

If G4x<0 the critical phase is unstable with respect to a separation into two phases with
different compositions, i.e., the critical phase is 'materially unstable'. If the first non-zero
derivative of the molar Gibbs energy with respect to the composition at constant temperature and
pressure is an even derivative and is positive, the critical phase will be stable under any small
fluctuation in the local compositions in the critical phase. We will discuss the behaviour of the
derivatives of the molar Gibbs energy with respect to the mole fraction in chapter 5. Both the
mechanical and material stability test are extensively discussed by Rowlinson and Swinton [4,
chap.6].

The global stability test

The third stability test takes the global stability of a critical phase into account. If a
critical phase passed both local stability tests, it might still be possible that it is
thermodynamically metastable. This happens if a phase separation is possible that would lower
the molar Gibbs energy of the system at the same pressure and temperature. To ensure that a
critical phase is locally stable only the signs of two derivatives (A2v, G4x) have to be evaluated at
the critical pressure, temperature and composition.
The global stability criterion requires a full investigation of the molar Gibbs energy plane
at the pressure and temperature of the critical phase. Deiters and Pegg [43] presented the
following mathematical formulation. If a critical phase c is to be stable with respect to an
auxiliary phase a, then ΔG has to be positive, see Fig.4.24. The difference ΔG is defined as
follows:

Δ G = Ga (xa ) - Gc (xa ) (4.65)

If ΔG is negative, as is the case in Fig.4.24(b), a mixture with the critical composition would split
into two phases because the molar Gibbs energy of the two coexisting phase would be lower than
that of the critical phase. The dotted lines in Fig.4.24(a) and (b) that are tangent to the points a
and c, have the following dependence on the composition:

G a ( x) p ,T = ( μ 2a - μ1a ) x + μ1a (4.66)

G c ( x ) p ,T = ( μ 2c - μ1c ) x + μ1c (4.67)


where μiF represents the chemical potential of component i in phase F. Substitution of Eqs.(4.66)
and (4.67) into Eq.(4.65) and recognizing that ΔG >0 for the critical phase to be stable, leads to
the following inequality:

( μ 2a - μ 2c ) . xa + ( μ1a - μ1c ) . ( 1 - xa ) > 0 (4.68)


94 Chapter 4

Figure 4.24: Diagram A shows a locally and globally stable critical point, whereas the
critical point in diagram B is locally stable but globally unstable. The latter
critical point is said to be metastable.

Eq.(4.68) should hold for every auxiliary composition in the range <0,1>, i.e., for every point a
along the solid curve in Fig.4.24. In case the left-hand side of Eq.(4.68) equals zero and the
auxiliary composition does not equal the composition of the critical phase, we have reached the
boundary between stability and metastability along the critical curve. Such a critical point is
called a critical endpoint. Once a critical point was calculated and had passed the local stability
tests, we tested all auxiliary phases in the composition range from xa=0.02 to xa=0.98 with step
size Δxa=0.02. If necessary the stepsize in the auxiliary phase composition was reduced.

4.7.4 Results

For several binary hydrocarbon mixtures the experimentally determined critical points
are compared with critical curves that have been calculated with the SPHCT equation of state.
The pure component parameters c, V * and T * were taken from Table 4.1. There are several ways
to obtain values for the binary interaction coefficient, k12. They can be estimated from theory
(McCoubrey and Hudson, [44]) or optimized from experimental data.
McCoubrey and Hudson [44] derived from the Margenau-London [45] theory of
dispersion forces the following expression for ε12:
⎡ 2 ( I 1 I 2 ) ⎤ ⎡ 26 σ 13 σ 32 ⎤ (4.69)
ε 12 = ⎢ ⎥⎢ 6⎥
(ε1ε2 )
⎣⎢ I 1 + I 2 ⎦⎥ ⎣ ( σ 1 + σ 2 ) ⎦
Performance SPHCT 95

In their investigations McCoubrey and Hudson [44] used the Lennard-Jones 12:6-potential for
the intermolecular potential between spherical molecules. Recognizing that the ionization
potentials of most non-polar hydrocarbons hardly differ [12], the first factor on the right-hand
side of Eq.(4.69) becomes unity. Using Eq.(3.74) and Eq.(4.69) leads to the following expression
for k12:

6 σ 11 3
σ 322
k 12 = 1 - 2 (4.70)
[ σ 113 + σ 322 ]6
In the SPHCT all molecules are split up in a specified number of segments. Each segment has the
volume of a CH2-group, 6.73 cm3.mole-1, see section 4.1. Accordingly, all segments have equal
diameters (2.51 A). The intermolecular interaction between two unlike molecules is calculated
by summing up the intersegmental interactions between the segments of the unlike molecules. If
Eq.(4.70) is applied to the SPHCT equation of state, the binary interaction coefficient is zero for
all intersegmental interactions, because all segments have the same segmental diameter.
However, it turns out that, especially when the size difference between the unlike molecules is
large, a binary interaction coefficient is needed for an accurate description of the vapour-liquid
critical curve in binary hydrocarbon mixtures.
There are several thermodynamic properties that can be determined experimentally, from
which the binary interaction coefficient can be optimized. Cabrerizo et al. [46] obtained k12-
values for the Prigogine-Flory-Patterson model from excess enthalpy, excess volume and excess
Gibbs energy data. Rijkers [28] calculated k12-values for the Peng-Robinson equation of state
(PR) from virial cross coefficients (B12). Hicks and Young [47] optimized binary interaction
B

coefficients from critical temperature and critical composition data. However, the most common
procedure for obtaining k12-values is from data regression on isothermal phase envelopes. A very
large number of k12-values for the PR equation of state determined in this way, is given by
Gmehling and Onken [48]. Because our aim is to test the performance of the SPHCT equation of
state in the critical region we optimized k12-values on critical data. The objective function that
was minimized is given by Eq.(4.71):
2
1 N ⎛ p calc exp
- p c,i ⎞
AAD = ∑ ⎜ c,i ⎟ + ( xcalc
c,i - x c,i )
exp 2
(4.71)
N ⎜ exp
p c,i ⎟
i=1 ⎝ ⎠

where AAD represents the Absolute Average Deviation of the calculated critical pressure from
the experimentally determined critical pressure at a certain temperature. Here pc,iexp is the
experimental critical pressure and pc,icalc the calculated critical pressure. By using Eq.(4.71) it is
assumed that the experimental critical temperatures do not contain any errors, which is of course
not correct. However, it can be assumed that the relative error in the measured critical
temperature is smaller than the relative errors in the critical compositions and critical pressures.
Most experimentally determined critical temperatures are given in 5 digits while most critical
pressures and critical compositions are given in 3 to 4 digits.
96 Chapter 4

Critical density data for binary systems are


scarce and are therefore omitted from our
optimization procedure. To perform the
optimization of k12 the Hicks-Young algorithm
had to be extended. The Hicks-Young algorithm
as described in the previous section, gives a
critical volume and a critical temperature at a
specified critical composition. For optimizing
binary interaction coefficients, the program was
nested in a iterative loop that converged to the
specified (experimental) critical temperature.
As an example the dependence of the
AAD on the values of the binary interaction
coefficients for n-alkane systems including n-
butane, is shown in Fig.4.25. In this figure the
solid curve (1) represents the system n-butane +
n-pentane, the 'dash-dot-dot-dotted' curve (2) Figure 4.25: Diagram showing the rela-
represents n-butane + n-hexane, the dashed tionship between the AAD
and the binary interaction
curve (3) represents n-butane + n-heptane, the coefficient.
dash-dotted curve (4) represents n-butane + n-
octane and the dotted curve (5) represents the system n-butane + n-decane. The SPHCT equation
of state was tested for a large number of non-polar binary hydrocarbon systems. Experimental
data are tabulated by Hicks and Young [47]. The optimal k12-values for n-alkane + n-alkane
systems are given in Table 4.4, for n-alkane + cycloalkane systems are given in Table 4.5, for n-
alkane + aromatics systems are given in Table 4.6 and for other hydrocarbon systems are given
in Table 4.7. For each class of binary systems an example is shown in Fig.(4.26-4.29).
Performance SPHCT 97

Table 4.4: Optimized k12-values for the SPHCT equation of state for n-alkane + n-alkane
mixtures. If a binary system is marked by an asterisk (*), experimental data were
available over a limited composition range. The column, marked with AAD, shows
the absolute average deviation with respect to the critical pressure and the critical
composition, defined by Eq.(4.71). AADx represents the absolute average deviation
in the calculated critical composition and AADp represents the absolute average
deviation in the critical pressure. If the column AADp contains a '-', experimental
critical pressures were not available. The fifth column contains the number of
datapoints that was used for the regression. The literature indication in the last
column of this table refers to the dataset of Hicks and Young [47], that was used
for optimization.

System 103.k12 AAD AADx AADp N lit.


% % %
C1 + C2 23.5 1.54 0.87 1.23 8 b
C1 + C3 4.5 3.39 1.20 3.04 12 b,c
C1 + n-C4 -27.0 9.62 3.93 8.13 10 a,b
C1 + n-C5 -50.4 6.01 4.26 3.82 5 a
C1 + n-C6* -135.4 13.0 11.33 5.61 3 a,b
C1 + n-C7 -68.1 9.35 5.21 6.45 8 a
C1 + n-C8* -128.5 14.64 14.29 2.94 2 a
C1 + n-C9* -101.0 6.59 6.36 1.69 4 a
C1 + n-C10* -104.2 5.37 4.34 2.98 7 a

C2 + C3 -2.2 2.98 1.02 2.71 6 a


C2 + n-C4 36.7 1.11 0.30 1.01 8 a,b
C2 + n-C5 54.8 2.78 0.89 2.56 6 a,b
C2 + n-C6* 52.4 2.87 1.69 2.26 3 a
C2 + n-C7 18.0 5.33 2.29 4.51 15 a,b
C2 + n-C10* -50.0 13.32 3.48 12.27 6 a

C3 + n-C4 5.6 0.32 0.17 0.27 6 c


C3 + n-C5 22.0 0.84 0.63 0.42 5 b
C3 + n-C6 25.5 0.61 0.27 0.54 6 a
C3 + n-C7 33.9 2.76 1.42 2.29 5 a
98 Chapter 4

System 103.k12 AAD AADx AADp N lit.


% % %
C3 + n-C8 29.9 4.46 2.22 3.52 7 a
C3 + n-C10* 14.8 6.03 1.68 5.77 5 a

n-C4 + n-C5 3.1 0.27 0.17 0.20 9 a


n-C4 + n-C6 8.1 0.51 0.23 0.40 9 a
n-C4 + n-C7 24.1 1.17 0.55 0.91 14 a
n-C4 + n-C8 34.3 1.23 0.82 0.90 9 a
n-C4 + n-C10* 41.4 1.91 0.64 1.78 3 a

n-C5 + n-C6 1.2 0.50 0.50 - 5 a


n-C5 + n-C7 8.3 1.31 0.25 1.28 3 a
n-C5 + n-C8* 20.5 0.49 0.49 - 6 a
n-C5 + n-C9 5.7 5.78 3.18 4.10 9 b
n-C5 + n-C10* 12.9 1.54 1.54 - 7 a

n-C6 + n-C7 -0.8 0.88 0.25 0.83 9 a


n-C6 + n-C8 0.5 0.99 0.31 0.91 9 a
n-C6 + n-C10 17.3 1.82 0.86 1.44 9 a

n-C7 + n-C8 0.0 1.02 0.19 1.00 9 a


Performance SPHCT 99

Table 4.5: Optimized k12-values for the SPHCT equation of state for n-alkane + cycloalkane
mixtures. If a binary system is marked by an asterisk (*), experimental data were
available over a limited composition range. If the column AADp contains a '-',
experimental critical pressures were not available. The fifth column contains the
number of datapoints that was used for the regression. The literature indication in
the last column of this table refers to the dataset of Hicks and Young [47], that was
used for optimization.

System 103.k12 AAD AADx AADp N lit.

% % %

cC5 + C5 -4.9 0.60 0.34 0.41 9 a

cC5 + C7 0.6 1.16 0.78 0.78 9 a

cC5 + C8 5.0 1.13 0.54 0.94 9 a

cC5 + C9 8.7 1.26 0.76 0.94 9 a

cC6 + C1* -60.4 10.42 9.69 3.48 6 a

cC6 + C2 1.0 6.44 4.50 3.93 8 a

cC6 + C5 -12.9 0.60 0.60 - 5 a

cC6 + C6 -5.1 0.77 0.46 0.49 9 b

cC6 + C7 -3.2 1.38 1.16 0.47 9 a

cC6 + C8 -1.7 2.24 1.71 1.23 9 a

cC6 + C9 0.7 1.39 0.99 0.89 9 a

cC6 + C10 4.2 0.52 0.36 0.27 9 a


100 Chapter 4

Table 4.6: Optimized k12-values for the SPHCT equation of state for n-alkane + aromatic
mixtures. The used abbreviations for the aromatics are: BNZ=benzene,
TOL=toluene and C2BNZ=ethylbenzene. If a binary system is marked by an
asterisk (*), experimental data were available over a limited composition range.
The column, marked with AAD, shows the absolute average deviation with respect
to the critical pressure and the critical composition, defined by Eq.(4.71). AADx
represents the absolute average deviation in the calculated critical composition
and AADp represents the absolute average deviation in the critical pressure. If the
column AADp contains a '-', experimental critical pressures were not available.
The fifth column contains the number of datapoints that was used for the
regression. The literature indication in the last column of this table refers to the
dataset of Hicks and Young [47], that was used for optimization.

System 103.k12 AAD AADx AADp N lit.

% % %

C2 + BNZ 29.1 20.26 17.91 8.68 6 a

C3 + BNZ 15.6 5.56 3.21 4.49 4 a

nC5 + BNZ 17.0 3.64 3.64 - 5 a

nC6 + BNZ 10.1 0.52 0.22 0.45 9 b

nC7 + BNZ 11.5 0.66 0.37 0.44 9 a

nC8 + BNZ 12.9 1.38 0.76 1.03 5 a

nC9 + BNZ 14.9 1.46 0.99 0.98 9 a

nC10 + BNZ 22.3 0.44 0.33 0.23 9 b

nC5 + TOL 58.0 1.50 1.50 - 4 a

nC6 + TOL 6.8 5.33 1.09 5.14 3 b

nC5 +C2BNZ 15.2 0.70 0.32 0.57 9 a

nC6 +C2BNZ 10.1 0.64 0.32 0.52 9 a

nC7 +C2BNZ 6.7 0.83 0.43 0.69 9 a

nC8 +C2BNZ 6.1 0.92 0.44 0.78 9 a


Performance SPHCT 101

Table 4.7: Optimized k12-values for the SPHCT equation of state for other hydrocarbon +
hydrocarbon mixtures. The used abbreviations for the aromatics are:
BNZ=benzene, TOL=toluene, oXYL=o-xylene and C2BNZ=ethylbenzene. If a
binary system is marked by an asterisk (*), experimental data were available over
a limited composition range. The column, marked with AAD, shows the absolute
average deviation with respect to the critical pressure and the critical composition,
defined by Eq.(4.71). AADx represents the absolute average deviation in the
calculated critical composition and AADp represents the absolute average
deviation in the critical pressure. If the column AADp contains a '-', experimental
critical pressures were not available. The fifth column contains the number of
datapoints that was used for the regression. The literature indication in the last
column of this table refers to the dataset of Hicks and Young (1975), that was used
for optimization.

System 103.k12 AAD AADx AADp N lit.


% % %
cC5 + cC6 -2.6 0.60 0.38 0.42 9 a
cC5 + cC7 18.5 4.40 4.40 - 6 a
cC5 + cC8 6.4 1.55 1.55 - 7 a
cC6 + cC7 8.7 1.16 1.16 - 6 a
cC6 + cC8 10.8 0.88 0.88 - 9 a

BNZ + cC5 10.0 0.56 0.54 0.14 9 a


TOL + cC6 9.3 0.40 0.40 - 4 a

BNZ + TOL -1.9 1.20 0.92 0.58 9 c


BNZ + oXYL 5.8 0.66 0.42 0.48 9 a
BNZ +C2BNZ 2.7 0.35 0.17 0.30 9 a
TOL + oXYL 0.1 0.46 0.38 0.21 9 a
TOL +C2BNZ -2.3 0.93 0.92 0.13 9 a
oXYL+C2BNZ -0.1 1.85 1.83 1.59 9 a

For the n-alkane + n-alkane systems a series of calculated critical curves is shown with n-butane
as one of the components, see Figs.4.26(a) and (b). For the n-alkane + cycloalkane systems a
series is shown in Figs.4.28(a) and (b) with cyclohexane as one of the components. In
Figs.4.29(a) and (b) the calculated critical curves are compared with experimental data for the n-
alkane systems with benzene. Finally, in Figs.4.30(a) and (b) critical curves of systems of
102 Chapter 4

benzene and other aromatics are shown. For the selected groups of systems the results are discus-
sed in more detail. If the size ratio between the unlike molecules is not too large, the description
of the critical curves with the SPHCT equation of state is good. For systems with large size
differences, like C2H6 + n-C7H16, C2H6 + n-C10H22, C3H8 + n-C8H18 and C3H8 + C10H22 it is not
possible to obtain a good description of the critical curve in the p,T-plane and the T,x-plane with
one value for k12: different values for k12 should be used for a description of the vapour-liquid
critical curve in the p,T-plane and the T,x-plane. In general for systems with large size
differences the critical pressures are calculated too high. In binary systems of n-butane, n-
pentane, etc. with the higher n-alkanes the description of the critical curves is good.

Van Konynenburg [39] classified binary fluid mixtures according to the appearance of
their critical curves in a p,T-projection. Type I mixtures show a continuous vapour-liquid critical
curve and the absence of liquid-liquid immiscibility, see also chapter 2. Most of the calculated
phase diagrams in the binary hydrocarbon mixtures that are under consideration in this chapter,
show type I phase behaviour. However, it is stressed here that calculated phase diagrams that we
call type I, show type II phase behaviour if a liquid-liquid critical curve exists at temperatures
below the lower temperature boundary of our search area. It is rather inefficient to start the
search for critical points at T=0K, if the vapour-liquid critical curve is situated above T=300K.
For most hydrocarbon mixtures the search area for the calculation of critical points ranged from
100K below the critical temperature of the most volatile component to 100K above the critical
temperature of the heavy component. Moreover, it will be shown in chapter 8 that the calculated
composition range of the liquid-liquid critical curve that forms the difference between type I and
II phase behaviour, can be extremely small, e.g., Δx=0.005. For this reason it easily can be
overlooked during the calculations. For the same reason a calculated phase diagram that is said to
be type V, may show type IV phase behaviour. Because we are interested in the description of
the vapour-liquid critical curve, the possible existence of a liquid-liquid critical curve at very low
temperatures is not of importance to us at this time. We shall therefore speak of type I or V phase
behaviour, although the considered binary mixture may show type II or IV phase behaviour.

n-alkane + n-alkane systems

Mixtures of substances belonging to a particular homologuous series, like the n-alkanes,


deviate from simple type I behaviour only when their size difference, and thus the values of their
critical properties, exceeds a certain ratio. Davenport and Rowlinson [49] pointed out that with
methane as one component a change from type I to type V occurs first with n-hexane, whereas
with ethane as the lighter component, n-octadecane is the first to show partial miscibility
(Specovius et al., [50]). With propane this behaviour occurs when the second component is n-C30
(Peters et al., [51]). With our optimized binary interaction coefficients these transitions from type
I to type V phase behaviour are calculated to occur at lower solute carbon numbers. In binary
mixtures with methane as a solvent, type V behaviour is predicted when the solute is n-butane,
Performance SPHCT 103

see Figs.4.27(a) and (b). At the Upper Critical EndPoint (UCEP) a vapour-liquid critical phase is
in equilibrium with a second liquid phase (L1=V+L2). The UCEP is located at (p,T,x) = (50.652
bar, 195.046 K, 4.854.10-3). The coexisting liquid phase has composition x=0.10664. At the
Lower Critical EndPoint (LCEP) a liquid-liquid critical phase is in equilibrium with a vapour
phase (L1=L2+V). The LCEP is located at (p,T,x) = (38.466 bar, 188.493 K, 4.3683.10-2). The
coexisting vapour phase has composition x=9.0394.10-4. The critical curve from the critical point
of pure methane to the UCEP is stable and from the UCEP to the solid upward triangle
metastable. Between the solid upward triangle and the solid downward triangle the critical curve
is unstable. From the solid downward triangle to the LCEP the critical curve is metastable again
and from the LCEP to higher temperatures the critical curve is stable again, see Figs.4.27(a) and
(b). Between the critical endpoints a three-phase line L1L2V is situated. This curve has not been
calculated and is not shown in Figs.4.27(a) and 4.27(b).

Figure 4.26(a): p,T-Projection of the critical Figure 4.26(b):T,x-Projection of the cal-


curves in the systems n-buta- culated critical curves in
ne(1) + n-alkanes(2). The open the binary systems n-buta-
symbols represent experimental ne(1) + n-alkanes(2). For
binary critical points and the an explanation of
solid symbols the experimental symbols, see Fig.4.26(a).
critical points of the pure
components. The numbers
represent the carbon numbers
of the pure n-alkanes.
104 Chapter 4

Figure 4.27(a): p,T-Projection of the calcu- Figure 4.27(b): T,x-Projection of the calcu-
lated critical curve in the lated critical curve in the
binary system methane(1) + binary system methane(1) +
n-butane(2). The open circle n-butane(2). For an explan-
represents the critical point of ation of symbols, see Fig.4.27
pure methane, the open (a). Parts of the critical curve
triangles the critical end- that are marked with s are
points and the solid triangles stable, m metastable and u
the boundary points between unstable.
metastability and instability.

For binary mixtures with ethane as the solvent, type V behaviour is predicted when n-
decane is the solute. As there were no critical data available for the binary systems C2H6 + C8H18
and C2H6 + C9H20 we could not optimize binary interaction coefficients for these systems, but it
might be possible that type V phase behaviour would be predicted for these mixtures as well. Of
course, the SPHCT equation of state is capable of calculating type V phase behaviour in the
correct binary mixtures when the binary interaction coefficients are optimized on critical data in
the region where this immiscibility phenomenon occurs. For binary systems with propane and
the higher n-alkanes as solvent only type I phase behaviour was predicted.

n-alkane + cycloalkane systems

In this group of binary mixtures, n-alkanes with cylcopentane and cyclohexane were
considered. Again, for binary mixtures with large size differences such as methane or ethane
with cyclohexane, the relative errors are large (10.4 % and 6.4 %).
Performance SPHCT 105

Figure 4.28(a): p,T-Projection of the cal- Figure 4.28(b): T,x-Projection of the cal-
culated critical curves in the culated critical curves in the
binary systems n-alkanes(1) binary systems n-alkanes(1)
+ cyclohexane(2). The open + cyclohexane(2). For an
symbols represent experimen- explanation of symbols, see
tal critical points of the b- Fig.4.28(a). of the pure
inary mixture, the solid sym- components. The numbers
bols represent the experi- near the pure critical points
mental critical points of the represent the carbon number
pure components. The num- of the n-alkanes.
bers near the pure critical
points represent the carbon
number of the n-alkanes.

For all systems under consideration, type I phase behaviour was predicted with the SPHCT
equation of state, except for the system methane + cyclohexane. For this system, type V phase
behaviour was calculated. Whether this is correct or not cannot be judged with existing
experimental data. In analogy with the n-alkane mixtures, demixing, and so type V phase
behaviour, might be predicted at too low size differences. So it might well be possible that in
reality the system methane + cyclohexane also exhibits type I phase behaviour. For the systems
of cyclopentane or cyclohexane with the n-alkanes from n-butane up to n-decane the description
of the critical curve in the p,T-plane and the T,x-plane is excellent with relative errors less than
1% in most systems.

n-alkane + aromatics systems


To show the effect of physical dissimilarity on phase behaviour in binary systems this
106 Chapter 4

group of systems can be compared with the n-alkane + cycloalkane systems. The deviations in
the description of the vapour-liquid critical curve of the n-alkanes + cyclohexane are comparable
with the deviations on the n-alkane + benzene systems. However, the k12-values of the latter
group of systems are larger. In general, the same conclusions can be drawn for this category of
binary systems as for the n-alkane + n-alkane and n-alkane + cycloalkane systems. If the size
differences between the unlike molecules are small, only a low value for the binary interaction
coefficient (k12<0.01) is needed to give an excellent description of the binary vapour-liquid
critical curve.
The vapour-liquid critical curves in these systems under consideration are all continuous.
All binary systems that were investigated in this category exhibit type I or II phase behaviour
according to the classification of Van Konynenburg [39]. In the system ethane + benzene critical
points were detected with liquid-like densities at relatively low temperatures. Because no
experimental data of a liquid-liquid critical curve in this system were available, this critical curve
was discarded from the optimization.

Figure 4.29(a): p,T-Projection of the calculated Figure 4.29(b): T,x-Projection of the


critical curves in the binary calculated critical cur-
ves in the binary systems
systems n-alkanes(1) + benzene
n-alkanes(1) + benzene
(2). The open symbols represent (2). For an explanation
experimental critical points of of symbols, see Fig.4.29
the binary mixtures, the solid (a).
symbols represent the experi-
mental critical points of the pure
components. The numbers near
the pure critical points represent
the carbon number of the pure
n-alkanes.
Performance SPHCT 107

hydrocarbon + other hydrocarbon systems

The description of the vapour-liquid critical curves in the p,T-plane and the T,x-plane of
this group of systems is excellent. It should be noticed that the binary systems under
consideration all have similar-sized molecules.

Figure 4.30(a): p,T-Projection of the calculated Figure 4.30(b): T,x-Projection of the calcu-
critical curves in the binary lated critical curves in the
systems that are composed of binary systems that are com-
aromatics. The open symbols posed of aromatics. For an ex-
represent experimental critical planation of symbols, see Fig.
points of the binary mixture, the 4.30(a).
solid symbols represent the
experimental critical points of
the pure components.

4.7.5 Discussion and conclusions

It can be concluded from Tables 4.4-4.7 and Figs.4.26-4.30 that the SPHCT equation of
state is capable of giving an accurate description of vapour-liquid critical curves in binary
hydrocarbon systems. For binary mixtures of molecules that are approximately equally sized,
such as n-butane or cyclopentane and the n-alkanes from n-pentane up to n-decane, the
description is excellent, both in the p,T-plane and in the T,x-plane. If the size difference between
the unlike molecules is large, e.g., V1*/V2*>2, a value for the binary interaction coefficient
different from k12 = 0, is needed. For most systems the description in the T,x-plane is better than
108 Chapter 4

the description of the critical curve in the p,T-plane.


The k12-values for many binary hydrocarbon systems were optimized from binary critical
data. An objective function was used that is analogous to objective functions used for
optimization of binary interaction coefficients from isothermal phase envelopes [33]. Strictly
speaking the objective function in Eq.(4.71) is not correct. With this objective function the sum
of the vertical distances from all experimental critical points to the calculated critical curve in the
p,T- and T,x-plane is minimized. In this way it is assumed that the experimentally determined
critical temperatures contain no errors, which is of course not correct. A better objective function
for optimizing k12-values is the minimum of the sum of the perpendicular distances from all
experimental points to the calculated critical curve in p,T,x-space. However, using the
'perpendicular distance' criterion requires a far greater numerical effort. Our objective function
gives at least a good indication of the optimal k12-values.

In general the following concluding remarks can be made about the description of critical
curves in binary systems with an equation of state. Because critical phenomena are non-classical
and the SPHCT equation of state is a classical equation, it cannot be expected that the SPHCT
equation of state with optimized k12-values will give a perfect description on both critical points
and binodal curves near a binary critical point. This is shown by the fact that experimentally
determined isothermal phase envelopes and p,x-cross sections from the p,T,x-space of a binary
system are always flatter near the binary critical point than the calculated binodals. This problem
occurs with every classical equation of state, see also section 4.1.

The binary interaction coefficient is related to the deviation of the depth of the square-
well potential of unlike molecules in a binary mixture from the geometric mean of the square-
well potentials in the pure fluids, Eq.(3.74). However, by optimizing values for k12, other
inadequacies of the equation of state are taken into account as well. Optimization of the k12-
values also is handicapped by the fact that for many binary systems critical data are available
only over a limited concentration range. This can have a profound influence on the value of the
binary interaction coefficient. For some binary mixtures the values of the binary interaction
coefficient were optimized on T,x critical data only because critical pressures were not available.
Nevertheless, the value of the binary interaction coefficient gives some information on the
intermolecular interaction between unlike molecules and the effect of size difference on these
unlike interactions.

The McCoubrey-Hudson theory, applied to the SPHCT equation, suggests that binary
interaction coefficients are zero even for interactions between short and long molecules. It is
proven here, that although the optimal values of k12 are small for most binary systems, non-zero
values for k12 are necessary for a good description of critical curves in binary hydrocarbon
mixtures.
Performance SPHCT 109

NOMENCLATURE

A molar Helmholtz energy data point i


Ai* molar Helmholtz energy of pure Ti* characteristic temperature of com-
component i ponent i
AAD absolute average deviation Tr reduced temperature, Tr=T/T*
a characteristic energy for Lennard- u intermolecular potential
Jones molecules, a=ε/k V molar volume
B second virial coefficient V0 molar reference volume at p=1 atm in
b closest-packed volume for Lennard- the ideal gas state
Jones molecules, b=NA.σ3/_2 Vi* closest-packed volume of pure
C third virial coefficient component i
CV molar isochoric heat capacity Vr reduced molar volume, Vr=V/V*
c one-third of the total degrees of x mole fraction of the heavier com-
freedom per molecule ponent, i.e., component 2.
F feed stream Z compressibility factor, Z=pV/RT
fi fugacity of component i Zm maximum coordination number
G molar Gibbs energy zi mole fraction of component i in the
g pair distribution function feed
I ionization potential
k Boltzmann constant Greek symbols
k12 binary interaction coefficient
NA Avogadro's number α phase split ratio
p pressure α critical exponent for Cv
pattr attractive term of an equation of state β Critical exponent for the density
pc,icalc calculated critical pressure of data ∂ partial derivative
point i Δ differential quantity
pc,iexp experimental critical pressure of data ε intermolecular potential energy per
point i unit external surface area
pi saturated vapour pressure of com- μi chemical potential of component i
ponent i ρ density, ρ=N/V
R gas constant ρr reduced density, ρr=V*/V
R ratio of the width of the square-well σ hard core diameter of a segment
potential and the hard core diameter of τ geometrical constant
a segment Ψ temperature function of an equation of
r intermolecular distance state
r* reduced distance ω acentric factor
q external surface area of a molecule
s number of segments per molecule Subscripts and superscripts
T absolute temperature
Tc,iexp experimental critical temperature of c critical property
110 Chapter 4

D Deiters
LJ Lennard-Jones
PR Peng-Robinson
RK Redlich-Kwong
SW square-well
u unnormalized property
VDW Van der Waals

REFERENCES

1. C-H. Kim, P. Vimalchand, M.D. Donohue and S.I. Sandler. Local composition model for
chainlike molecules: a new simplified version of the perturbed hard chain theory. AIChE J.
32, (1986), 1726-1734.
2. K.A.M. Gasem and R.L. Robinson, Jr. Evaluation of the Simplified Perturbed Hard Chain
Theory (SPHCT) for prediction of phase behaviour of n-parrafins and mixtures of n-
parrafins with ethane. Fluid Phase Equilibria 58, (1990), 13-23.
3. L. Ponce-Ramirez, C. Lira-Galeana and C. Tapia-Medina. Application of the SPHCT
model to the prediction of phase equilibria in CO2-hydrocarbon systems. Fluid Phase
Equilibria 70, (1992), 1-18.
4. J.S. Rowlinson and F.L. Swinton. Liquids and Liquid Mixtures Butterworths Monographs
in Chemistry, London, (1982), Chap.3.
5. R. Dohrn and J.M. Prausnitz. A simple perturbation term for the Carnahan-Starling
equation of state. Fluid Phase Equilibria, 61, (1990), 53-69.
6. IUPAC, International Thermodynamic Tables of the Fluid State. Butterworths, London.
Vol. 1: Argon, (1971), Vol. 2: Ethylene, (1972), Vol. 3: Carbon dioxide, (1977), Vol. 5:
Methane, (1978), Vol. 6: Nitrogen, (1977), Vol. 7: Propylene, (1980).
7. American Petroleum Institute. Research Report 44, Selected Values of Physical and
Thermodynamic Properties of Hydrocarbons and Related Compounds, Carnegie Press,
(1988).
8. A.E. Sherwood and J.M Prausnitz. Intermolecular potential functions and the second and
third virial coefficients. J. Chem. Phys. 41, (1964), 429-437.
9. K.H. Lee, M. Lombardo and S.I. Sandler. The generalized Van Der Waals partition
function. II. Application to the square-well fluid. Fluid Phase Equilibria 21, (1985), 177-
196.
10. R.J. Lee and K.C. Chao. Coordination number and thermodynamics of square-well fluid
mixtures. Mol. Phys. 61, (1987), 1431-1442
11. C.J. Peters, J.L. de Roo and J. de Swaan Arons. Measurements and calculations of phase
equilibria in binary mixtures of propane + tetratriacontane. Fluid Phase Equilibria 72,
(1992), 251-266.
12. C.W. Weast. CRC Handbook of Chemistry and Physics, 62nd Ed., Boca Raton, (1982).
13. M.D. Donohue and J.M. Prausnitz. Statistical thermodynamics of solutions in natural gas
and petroleum refining. Report RR-26, Gas Processors Assn., (1977).
14. M.D. Donohue and J.M. Prausnitz. Perturbed Hard Chain Theory for fluid mixtures:
Thermodynamic properties for mixtures in natural gas and petroleum technology. AIChE J.
24, (1978), 849 - 860.
15. ASME Steam tables in [12], E16-E21.
Performance SPHCT 111

16. F. Din. Thermodynamic functions of gases. Volume 1 : Ammonia, Carbon Dioxide and
Carbon Monoxide, Butterworths Scientific Publications, London, (1956).
17. L.N. Canjar and F.S. Manning. Thermodynamic properties and reduced correlations for
gases, Gulf Publishing Company, Houston, (1967).
18. International Critical Tables of Numerical Data, Physics, Chemistry and Technology. Vol.
III, McGraw & Hill, New York, (1928).
19. ASHRAE. Thermodynamic properties of refrigerants, New York, (1980).
20. J.V. de Palma and G. Thodos. Vapor-pressure behavior of parahydrogen. J. Chem. Engng.
Data 11, (1966), 31-37.
21. A. Peneloux, E. Rauzy and R. Freze. A consistent correction for Redlich-Kwong-Soave
volumes. Fluid Phase Equilibria 8, (1982), 7-23.
22. G. Soave. Equilibrium constants from a modified Redlich-Kwong equation of state. Chem.
Engng Sci. 27 (1949), 1197-1203.
23. D.Y. Peng and D.B. Robinson. A new two-constant equation of state. Ind. Eng. Chem.
Fundam. 15 (1976), 59-64.
24. T.M. Reed and K.E. Gubbins. Applied statistical mechanics. McGraw-Hill, New York.
(1973), chap. 8.
25. U.K. Deiters. A new semiempirical equation of state for fluids - I. Derivation. Chem. Eng.
Sci. 36, (1981), 1139-1145.
26. U.K. Deiters. Entwicklung einer semiempirischen Zustandsgleichung für fluide Stoffe und
Berechnung von Fluid-Phasengleichgewichten in binären Mischungen bei hohen Drücken.
PhD-thesis Ruhr-Universität Bochum, West-Germany, (1979).
27. G.H. Findenegg. Statistische Thermodynamik. Dr. Dietrich Steinkopff Verlag, Darmstadt,
Germany, (1985), chap.8.
28. M.P.W.M. Rijkers. Retrograde condensation of lean natural gas. PhD-Thesis. University of
Technology Delft, (1991).
29. J.H. Dymond and E.B. Smith. The virial coefficients of pure gases and mixtures. Clarendon
Press, Oxford, (1980).
30. T.J. Bruno and J.F. Ely. Supercritical Fluid Technology. Reviews in modern theory and
application. CRC Press, Boca Raton, (1991).
31. J.M. Smith and H.C. van Ness. Introduction to chemical engineering thermodynamics.
McGraw-Hill, Inc., New York. 4th. Ed., (1987), chap. 10-14.
32. R.C. Reid, J.M. Prausnitz and B.E. Poling. The properties of gases and liquids. 4th Ed.
McGraw-Hill. New York. (1987).
33. T.F. Anderson and J.M. Prausnitz. Computational methods for high-pressure phase
equilibria and other fluid phase properties using a partition function. 2. Mixtures. Ind. Eng.
Chem. Process Des. Dev. 19, (1980), 9-14.
34. R.C. Miller, A.J. Kidnay and M.J. Hiza. Liquid-vapor equilibria at 112.00 K for systems
containing nitrogen, argon, and methane. AIChE J. 19, (1973), 145-151.
35. I. Wichterle and R. Kobayashi. Vapor-liquid equilibrium of methane-ethane system at low
temperatures and high pressures. J. Chem. Engng. Data 17, (1972), 9-12.
36. J.J. Carroll. What is Henry's law? Chem. Eng. Prog., September 1991, 48-52.
37. J. Gmehling and B. Kolbe. Thermodynamik. Georg Thieme Verlag Stuttgart, New York,
(1988), chap. 4.
38. J.M. Prausnitz and P.L. Chueh. Computer calculations for high-pressure vapor-liquid
equilibria. Prentice-Hall, Inc. Englewood Cliffs, New Jersey, (1968), appendix D.
39. P.H. van Konynenburg. Critical lines and phase equilibria in binary mixtures. PhD-thesis
University of California Los Angeles. (1968).
40. Shaw. The derivation of thermodynamical relations for a simple system. Phil. Trans. Roy.
Soc. London A 234, (1935), 299-328.
112 Chapter 4

41. W.H. Press, B.P. Flannery, S.A. Teukolsky, and W.T. Vetterling. Numerical recipes.
Cambridge University Press, New York, (1986), Chap. 9.
42. C.P. Hicks and C.L. Young. Theoretical prediction of phase behaviour at high temperatures
and pressures for non-polar mixtures. J. Chem. Soc., Faraday Trans. 2, 73, (1977), 597-612.
43. U.K. Deiters and I.L. Pegg. Systematic investigation of the phase behavior in binary fluid
mixtures. 1. Calculations based on the Redlich-Kwong equation of state. J. Chem. Phys. 90,
(1989), 6632-6641.
44. J.C. McCoubrey and G.H. Hudson. Intermolecular forces between unlike molecules. Trans.
Faraday Soc. 56, (1960), 761-766.
45. H. Margenau. Van der Waals forces. Rev. Mod.Phys. 11, (1939), 1-35.
46. U. Cabrerizo, R.G. Rubio, C. Menduina and J.A. Renuncio. Thermodynamics of
hydrocarbon systems using an Ising-like model. J. Phys. Chem. 90, (1986), 889-894.
47. C.P. Hicks and C.L. Young. The gas-liquid critical properties of binary mixtures. Chem.
Rev. 2, 75, (1975), 119-175.
48. J. Gmehling and U. Onken. Vapour-Liquid Equilibrium Data Collection Dechema Data
Series. Frankfurt, West-Germany, (1982).
49. A.J. Davenport and J.S. Rowlinson. The solubility of hydrocarbons in liquid methane.
Trans. Faraday Soc. 59, (1963), 78-84.
50. J. Specovius, M.A. Leiva, R.L. Scott and C.M. Knobler. Tricritical phenomena in "quasi-
binary" mixtures of hydrocarbons. 2. Binary ethane systems. J. Phys.Chem. 35, (1981),
2313-2316.
51. C.J. Peters, H.J. van der Kooi, J.L. de Roo, J. de Swaan Arons, J.S. Gallagher and J.M.H.
Levelt Sengers. The search for tricriticality in binary mixtures of near-critical propane and
normal paraffins. Fluid Phase Equilibria 51, (1989), 339-351
113

CHAPTER 5
TRANSITION STATES IN THE GLOBAL PHASE DIAGRAM

In chapter 2 it was discussed which types of fluid phase behaviour can occur in binary
systems. The main goal of this chapter is to demonstrate how continuous transformations can
take place between these types of phase behaviour. After the introduction in section 5.1, we will
discuss the various types of transition states between the different types of phase behaviour in
section 5.2.

5.1 Introduction

In Fig.5.1 the evolution is shown of the phase behaviour of binary mixtures in the p,T-
plane due to increasing dissimilarities between the unlike molecules with respect to size and
interaction. In practice it is not only a size or interaction effect that causes the change in phase
behaviour, but a combination of both effects. For this reason, the dominating factor, causing a
particular transition in phase behaviour, is surrounded by the double quotes in Fig.5.1.
In chapter 7 it will be shown how these various types of phase behaviour can be located
on a map. The aim of this mapping is to facilitate the discussion of binary fluid phase behaviour.
Van Konynenburg [1] was the first investigator to elaborate this idea. He called this map the
"master diagram", later the term "global phase diagram" was introduced [2]. The boundaries
between the regions of the various types of phase behaviour will be considered in more detail in
the following sections. It should be stressed here that the global phase diagrams are based on
calculations using an equation of state. Real binary mixtures form a point in a specified two-
dimensional global phase diagram. It is possible that this point is situated in the wrong region:
i.e., it is possible that the equation of state predicts the wrong type of phase behaviour for a
practical mixture. In this chapter we will not worry about that: the transition states between the
various types of phase behaviour are under consideration and the agreement with real mixtures is
neglected.
In a way similar to Van Konynenburg and Scott [1,3,4], we define four dimensionless
parameter ratios. This reduction of the number of degrees of freedom allows us to make global
phase diagram calculations.

ζ = ε 11 ε 22
-
(5.1)
ε 11 + ε 22
114 Chapter 5

Figure 5.1: The evolution of binary fluid phase behaviour shown in p,T-projections.

*
- *
ξ = V*1 V 2*
(5.2)
V 1 +V 2

-
χ = c1 c 2 (5.3)
c1 + c 2

λ = ε 11 ε 22 ε 12
+ -2 (5.4)
ε 11 + ε 22

It seems that ζ, ξ and χ depend on two variables, see Eqs.(5.1), (5.3) and (5.4). This is not
true, e.g., ζ only depends on the ratio of ε11/ε22, i.e., ζ=(θ-1)/(θ+1) where θ=ε11/ε22. In a similar
way ξ depends on the ratio V1*/V2* and χ depends on the ratio c1/c2. The variable λ depends on
ε11/ε22 and ε12/ε22. The way in which ζ, λ, ξ and χ are defined depend on the equation of state used
for the global phase diagram calculations, in our case the SPHCT equation of state. For the
Redlich-Kwong and Carnahan-Starling-Redlich-Kwong they are slightly different defined [5,6].
Transition states 115

Although this chapter gives a general introduction to the global phase diagram, it is obvious that
sometimes we cannot avoid referring to the equation of state that is used. Global phase diagrams
can be constructed using any two of the four dimensionless ratios, see Eqs.(5.1)-(5.4). Van
Konynenburg and Scott [1,3,4] chose ζ for the horizontal axes and λ for the vertical axes. We
also have chosen to discuss ζ,λ-global phase diagrams, keeping the other two dimensionless
parameters at a constant value. However, other two-dimensional global phase diagrams could
just as well be used, e.g, ζ,ξ-diagrams [3]. ζ is a measure for the relative difference in the
potential (square-) well depth of both components. λ is a measure for the relative difference in 1-
2 interactions with respect to the 1-1 and 2-2 interactions. ξ and χ are measures for the relative
size and flexibility differences of both components. The parameters ζ, ξ and χ only depend on the
pure component parameters of both components. The λ-parameter also depends on the mixture
parameter ε12 and thus k12. ζ, ξ and χ can have values between -1 and 1, λ can have values
between minus infinity and 1. However, extremely negative values for λ are not realistic
physically.

It will be shown in chapter 6 that the SPHCT parameters of one of the components needs
to be defined to reduce the degrees of freedom for global phase diagram calculations. In our
investigations component 2 is used as a reference component. The pure component parameters of
the reference component are:

c2 = 1.0000
V2* = 20.000 cm3.mole-1
T2* = 239.998 K

These pure component parameter values of the SPHCT equation of state are fictitious. However,
their values are chosen so that they are physically realistic.

The limits of the dimensionless parameters and their physical meaning are listed below.
ξ→ 1 V1* → ∞ ζ→ 1 ε11 → ∞
* *
ξ = 0 V1 = V2 ζ= 0 ε11 = ε22
*
ξ → -1 V1 → 0 ζ → -1 ε11 = 0

χ→ 1 c1 → ∞ λ→ 1 ε12 = 0, k12 = 1
χ = 0 c1 = c2 λ= 0 ε12 = (ε11.ε22)0.5, k12 = 0
χ → -1 c1 → 0 λ → -∞ ε12 = ∞, k12 = -∞

The value of the global phase diagram parameter λ is related to the binary interaction coefficient,
k12, according to the following relation:

λ =1- 1 - ζ 2 .( 1 - k 12 ) (5.5)
116 Chapter 5

5.2 Thermodynamic conditions for the various transition states

In this section we will discuss the transition states in the global phase diagram. If the
thermodynamic conditions of a transition state are known, then we can derive expressions from
the equation of state under consideration, that fulfill these conditions. Subsequently, we are able
to calculate the curves that represent the boundaries between the various types of phase
behaviour in the global phase diagram. In the following subsections the transition states will be
summarized, and when possible, illustrated by experimentally determined examples.

5.2.1 The tricritical state

Because the tricritical curves play an important role in the global phase diagram, extra
attention will be paid to this multicritical phenomenon. We will digress on the mathematical
conditions for equilibria between two or three phases and the mathematical conditions for critical
and multicritical states, according to the work of Michelsen [7-10]. We recall, see chapter 4, that
an ordinary critical point is a thermodynamic state at which two coexisting phases become
identical. Similarly, the thermodynamic state at which three phases become identical
simultaneously is a tricritical point (TCP). This can be extended to critical points of even higher
order, such as tetracritical and pentacritical points, where four and five phases become identical
respectively. The constraint on the minimum number of components in a system required for
higher order critical phenomena to occur can be obtained from an analysis of the phase rule [11].
Gibbs phase rule may be written as follows:

f = n - r +2 -φ (5.6)

where f represents the number of degrees of freedom, n the number of components, r the number
of phases and φ is the number of extra conditions. For a normal critical point in a binary mixture,
n=2, r=2 and φ=1. Similarly, at a k-th order critical point, φ=(k-1). Thus the phase rule for a k-th
order critical point may be written:

f = n - r + 2 - ( k - 1 )= n - r + 3 - k (5.7)

Since f ≥ 0 and r ≥ k, the minimum number of components required for an k-th order critical
point is given by:

n≥2k -3 (5.8)

For an ordinary critical point (k=2), n≥1; for a tricritical point (k=3), n≥3 and for a tetracritical
point (k=4), n≥5.
TCPs have been located experimentally in three component systems. In ternary systems,
the TCP is a thermodynamic state with zero degrees of freedom: it exists only at an unique
Transition states 117

pressure, temperature and composition (p, T and x). In quaternary systems, one variable must be
specified in order to define the tricritical point. Since the 1960's a number of ternary (and
quaternary) systems are known that exhibit tricritical phenomena. They have been reviewed by a
number of authors [11-13]. Some ternary hydrocarbon systems with experimentally located
TCPs are shown in Table 5.1.
Eq.(5.8) shows that it is highly
unlikely to have a TCP in a binary
Table 5.1: Ternary systems with experimen-
mixture. However, the Gibbs phase rule is
tally located tricritical points.
based on the assumption that in a binary
mixture the components and thus their System Ref.
properties, e.g., molecular size, flexibility, CH4 + C2H6 + n-C8H18 [14]
and the interaction between the unlike
molecules cannot be seen as variables. CH4 + C3H8 + n-C8H18 [14]
This is, of course, true if we try to CH4 + n-C4H10 + n-C8H18 [14]
determine a TCP in a binary mixture
CH4 + n-C5H12 + n-C8H18 [15]
experimentally. With an equation of state
on the other hand, one is not restricted to CH4 + CO2 + n-C8H18 [14]
the fixed properties of existing
CH4 + n-C4H10 + N2 [16]
components nor to the real interaction
between the unlike molecules. If one of CH4 + n-C5H12 + N2 [16]
the parameters of one of the pure
components in a binary mixture or the interaction between the unlike molecules in a binary
mixture is seen as a variable, we obtain an extra degree of freedom. This extra degree of freedom
allows us to calculate a TCP in a binary system without any difficulty.
Such an approach has also been realized by experimental investigators in a class of
systems termed 'quasi-binary'. The quasi-binary approximation assumes that two components of
a ternary system are similar enough that they can be treated as a single pseudo-component. To
illustrate this we imagine the pseudo-n-alkane C18.5H39. This pseudo-n-alkane should be
composed of two n-alkanes with about the same carbon number, e.g., n-C18H38 and n-C19H40
rather than e.g., a mixture of n-C16H34 and n-C21H44. The ratio of the mole fractions of these two
components is then more or less constant in all phases present. Creek et al. [17,18] used the
quasi-binary approximation in their study of the systems methane + n-pentane + 2,3-
dimethylbutane and methane + 2,2-dimethylbutane + 2,3-dimethyl-butane. Experimental
measurements on both systems in the three-phase region confirm that the relative mole fraction
in all phases is constant within experimental error. A list of quasi-binary systems that show a
TCP is shown in table 5.2.
118 Chapter 5

Table 5.2: Quasi-binary mixtures with experimentally located tricritical points.

System Ref.

C2H6 + n-C16H34 + n-C20H42 [19,20]

CH4 + n-C5H12 + 2,3-(CH3)2C4H8 [17,18]

CH4 + 2,2-(CH3)2C4H8 + 2,3-(CH3)2C4H8 [17,18]

C2H6 + n-C17H36 + n-C18H38 [11]

The thermodynamic criteria for a TCP may be expressed in a number of ways. In this
section the development of the criteria from Michelsen [7-10] is reproduced and illustrated with
quantitative examples. We start by considering the local stability of a test phase with
composition x0. It is assumed that the molar Gibbs energy, G(p,T,x), can be expanded in a Taylor
series in the composition variable, x, at constant pressure and temperature, around the test phase
0. This Taylor series expansion has the following form:

⎛∂G⎞ 1 ⎛ ∂ 2G ⎞ 1 ⎛ ∂ 3G ⎞
G ( x ) = G ( x0 ) + ⎜ ⎟ .( x - x ) + ⎜
⎜ ⎟
2 ⎟
.( x - x ) 2
+ ⎜⎜ 3 ⎟⎟ .( x - x0 ) 3 + ...... (5.9)
⎝ ∂ x ⎠x0 2 ⎝ ∂ x ⎠x0 6 ⎝ ∂ x ⎠x0
0 0

The tangent plane to this curve in the test point has the following equation:

⎛∂G⎞
G TP ( x) = G ( x0 ) + ⎜ ⎟ .( x - x0 ) (5.10)
⎝ ∂ x ⎠x0

In the case of a binary mixture at constant pressure and temperature there is, strictly speaking, a
tangent line at the test point and not a tangent plane. However, we adopted the term 'tangent
plane' to stay close to the nomenclature of Michelsen [7-10]. When we subtract the equation for
the tangent plane, Eq.(5.10), from the Taylor series expansion, Eq.(5.9), we obtain the 'tangent
plane distance' function, F(x):

1 ⎛ ∂ 2G ⎞ 1 ⎛ ∂ 3G ⎞
F ( x ) = ⎜⎜ ⎟ .( x − x ) 2
+ ⎜ ⎟ .( x − x0 ) 3 + ......
2 ⎝ ∂ x 2 ⎟⎠ x0 6 ⎜⎝ ∂ x 3 ⎟⎠ x0
0 (5.11)

For convenience we adopt the abbreviations introduced by Michelsen [7-10]: b= ½ . (∂ 2G/∂ x2)x0
c = 1/6 . (∂ 3G/∂ x3)x0 , d = 1/24 . (∂ 4G/∂ x4)x0 , e = 1/120 . (∂ 5G/∂ x5)x0 and f = 1/720 . (∂ 6G/∂ x6)x0 .
The polynomial for F up to the sixth order term now becomes:
Transition states 119

F ( x ) = b.( x − x0 ) 2 + c.( x − x0 ) 3 + d.( x − x0 ) 4 + e.( x − x0 ) 5 + f.( x − x0 ) 6 + (5.12)

(
+ O ( x − x0 ) 7 )
The polynomial series may be truncated after a term is reached that is large with respect to zero.
If b is large relative to zero, the stability of the test phase is determined by the sign of b. If b is
positive, then F(x) will be positive for some x near x0 and the test phase will be locally stable. If b
is zero, stability is determined by the higher-order coefficients in the expansion. A nonzero value
of c with b=0 implies that the mixture is intrinsically unstable, since F(x) becomes negative at a
composition that is a little higher than x0 (if c is negative) or negative at a composition that is a
little smaller than x0 (if c is positive). If an even order term coefficient is zero, then, for the test
phase to be locally stable, the next uneven order term coefficient should be zero as well. Local
stability of the test phase requires the first non-zero coefficient in Eq.(5.12) to be even and
positive.

Any composition x, with x ≠ x0, say x = x1, for which:

⎛ ∂ F ( x) ⎞
F ( x1 ) = 0 and ⎜ ⎟ =0 (5.13)
⎝ ∂ x ⎠x= x1

is valid, is the composition of a phase in equilibrium with the test phase. The equations in
Eq.(5.13) are identical to μ10=μ11 and μ20=μ21. In the latter expression the subscript refers to the
component and the superscript refers to the phase. An example is shown in Fig.5.2 and Fig.5.3.
In Fig.5.2 the dotted line represents the horizontal axis (G=0), the dashed line is the line that is
tangent to the test phase with composition x0 and the solid curve is called the G-curve. The G-
curve can be calculated from an equation of state. Because this section is intended to illustrate
the thermodynamic conditions for phase equilibria and critical phenomena, the following simple
expression is used to describe the G-curve: G=1000.x4-1600.x3+940.x2-239.x+22.3.
In Fig.5.3 the dotted line represents the horizontal axis (F=0) and the solid line is the
tangent plane function. The two coexisting phases 0 and 1 in Figs.5.2 and 5.3 have been found in
the following way. We started with a composition of the test phase at e.g., x0=0.9 and calculated
the polynomial coefficient b for this point on the G-curve to make sure that the test phase was
stable. The plane that is tangent to the G-curve at the composition of the test phase did not touch
the G-curve at a composition different from x=x0. The same was done for x0=0.8, 0.7, 0.6 and
0.5. At x0=0.5 it turned out that another phase existed, i.e., the phase with composition x=x1, that
belonged both to the G-curve as the tangent plane, i.e., the tangent plane at x0 had become a
double tangent plane.

The coefficients of the polynomial that approximates G as a function of x, see Eq.(5.12),


at the composition of the test phase with composition x0=0.5 in Fig.5.2 and 5.3 are: b=40, c=400
and d=1000. In this example the higher order terms are zero. The composition of the coexisting
phase is x1=0.3. The advantage in using the tangent plane distance function, F, can be seen from
120 Chapter 5

Fig.5.3. Geometrically F(x) is the vertical distance from the molar Gibbs energy surface at
composition x to the tangent plane (the plane that is tangent to the molar Gibbs energy surface at
the test phase with composition x0) at composition x. The sign of b at the test phase with
composition x0=0.5, i.e., b=+40, shows that the test phase with this composition is stable. If the
composition of the test phase is chosen to be x0=0.4, i.e., at the local maximum in Fig.5.3, the
polynomial coefficients of the curve in Fig.5.3 would have been different: b=-20, c=0 and
d=1000. The negative value of b and consequently of (∂ 2G/∂x2), in this case (b=-20), shows that
the test phase with composition x0=0.4 is unstable.

Figure 5.2: G(x) versus x in case of a Figure 5.3: F(x) versus x in case of a
two-phase equilibrium. two-phase equilibrium.

In order to describe a phase equilibrium between the test phase and an alternate phase
(x=x0 and x=x1) the expansion around x0 must be carried out to the fourth power in x and
factorized as follows:

F ( x ) ≅ d.( x − x0 ) 2 .( x − x1 ) 2 (5.14)

This ensures that the conditions for equilibrium with the test phase, F=0 and ∂F/∂x=0, are
satisfied for the alternate phase. Eq.(5.14) is a polynomial approximation of the real curve that is
calculated from an equation of state. If it is assumed that it is strictly valid, then the polynomial
coefficients b and c that represent the higher derivatives around the test point can be expressed as
functions of the difference between the compositions of the coexisting phases, i.e., d: b=d(x1-x0)2
and c=-2d(x1-x0). By definition the critical point is the state where two equilibrium phases
become identical. The alternate phase must then have the same composition as the test phase,
that is: x1=x0. The Taylor series expansion around the test point x0 then becomes:
Transition states 121

F ( x ) ≅ d.( x − x0 ) 4 (5.15)

The conditions for a critical point may now be written as b = c = 0, or in terms of the molar
Gibbs energy:

⎛ ∂ 2G ⎞ ⎛ ∂ 3G ⎞
⎜⎜ ⎟
2 ⎟
= 0 and ⎜⎜ ⎟ =0
3 ⎟
(5.16)
⎝ ∂ x ⎠ p,T ⎝ ∂ x ⎠ p,T

For the critical phase to be stable it is also necessary to have d > 0, or in terms of the molar Gibbs
energy:

⎛ ∂ 4G ⎞
⎜⎜ ⎟ >0
4 ⎟
(5.17)
⎝ ∂ x ⎠ p,T

The same procedure can be applied for the derivation of the equations for a three-phase
equilibrium and a tricritical point. In order to describe phase equilibrium between the test phase
and two alternate phases (x=x0, x=x1 and x=x2) the expansion must be carried out to the sixth
power in x and factorized as follows:

F ( x ) ≅ f.( x - x0 )2 .( x - x1 )2 .( x - x2 )2 (5.18)

An example is given in Fig.(5.4) and Fig.(5.5).

Figure 5.4: G(x) versus x in case of a Figure 5.5: F(x) versus x in case of a
three-phase equilibrium. three-phase equilibrium.

The G-curve in Fig.5.4 has been calculated from the following simple equation: G=10000.x6-
32000.x5+41400.x4-27680.x3+10081.x2-1895.x+143.8. The coefficients in Eq.(5.12) around the
122 Chapter 5

test phase with composition x0=0.5 are: b=36, c=120, d=-1100, e=-2000 and f=10000. The
compositions of the coexisting phases are x1=0.3 and x2=0.8. The factorization, as in Eq.(5.18),
ensures that the conditions for equilibrium with the test phase, F=0 and ∂F/∂x=0, are satisfied for
the alternate phases. Again, if the higher order terms can be neglected, the polynomial
coefficients can be expressed as a function of the difference between the compositions of the
coexisting phases with the test phase composition and coefficient f. By definition the tricritical
point is the state where three equilibrium phases become identical. The two alternate phases must
then have the same composition as the test phase, that is x1 = x2 = x0. The Taylor series expansion
around the test point x0 then becomes:

F ( x ) ≅ f .( x - x0 )6 (5.19)

The conditions for a critical point may now be written as b = c = d = e = 0, or in terms of the
molar Gibbs energy:

⎛ ∂ 2G ⎞ ⎛ ∂ 3G ⎞ ⎛ ∂ 4G ⎞ ⎛ ∂ 5G ⎞
⎜⎜ 2 ⎟⎟ = 0 , ⎜⎜ 3 ⎟⎟ = 0 , ⎜⎜ ⎟ =0 ,
4 ⎟
⎜⎜ 5 ⎟⎟ = 0 (5.20)
⎝ ∂ x ⎠ p,T ⎝ ∂ x ⎠ p,T ⎝ ∂ x ⎠ p,T ⎝ ∂ x ⎠ p,T

For the critical phase to be stable it is also necessary to have f > 0, or in terms of the molar Gibbs
energy:

⎛ ∂ 6G ⎞
⎜⎜ 6 ⎟⎟ > 0 (5.21)
⎝ ∂ x ⎠ p,T

The criteria for a TCP are thus described by four equations in the thermodynamic variables,
Eq.(5.20), with Eq.(5.21) as a criterium for its local stability. The tricritical transition state is
schematically shown in Fig.5.6 as the transition state between type I and type V phase behaviour.

In Fig.5.6 the solid curves that are marked with 1 and 2 represent the vapour pressure curves.
The open circles represent the pure component critical points and the critical endpoints (U,L).
We recall that a critical endpoint is a point where a critical phase coexists with an alternate, non-
critical phase. The dotted curves in Fig.5.6 represent the critical curves, the dashed-dotted curves
the L1L2V three-phase line and the TCP is shown as a solid circle.
At the TCP a vapour-liquid critical curve of type I phase behaviour is about to be
interrupted. Between the critical endpoints a L1L2V-three-phase line has just appeared. If we
approach the TCP from type V phase behaviour, then the TCP is formed when an UCEP and a
LCEP of the same three-phase line coincide. The transition between type II and type IV phase
behaviour also takes place via a critical curve that contains a TCP.
The conditions for a tetracritical point are derived in an analogous way. In the shorthand
notation of Rowlinson and Swinton [21], where (∂ iG/∂xi)p,T is represented by Gix, they can be
written as:
Transition states 123

Figure 5.6: The tricritical point in the p,T-plane. It is shown that a p,T-projection with a
tricritical point forms the transition state between type I and type V phase
behaviour.

Gix = 0 with i = 2, 3, ...,7 (5.22)

For the tetracritical phase to be locally stable, the eighth derivative of the molar Gibbs energy
with respect to the composition at constant p and T should be positive, i.e., G8x>0. In order to
determine the global stability of the tricritical and tetracritical phase the test procedure described
in section 4.7 should be followed.

5.2.2 The double critical endpoint.

A second transition state between two types of phase behaviour is the double critical
endpoint (DCEP). A binary mixture that contains a DCEP shows the phase behaviour that is
intermediate between type III and type IV phase behaviour. This is shown in Fig.5.7.

In Fig.5.7 the solid curves that are marked with 1 and 2 represent the vapour pressure curves.
The open circles represent the pure component critical points and the critical endpoints. The
dotted curves represent the critical curves. The dashed-dotted curves represent the L1L2V three-
phase lines and the DCEP is shown as a solid circle. The double critical endpoint in Fig.5.7 is the
point where the vapour-liquid critical curve in the p,T-projection of a type III phase diagram
touches the three-phase line L1L2V. If the pressure minimum in the vapour-liquid critical curve
emerging from the critical point of component 2 is below the three-phase line, then the three-
phase line is interrupted, see the diagram on the right in Fig.5.7.
124 Chapter 5

Figure 5.7: The DCEP in the p,T-plane. It is shown that a p,T-projection with a DCEP forms
the transition state between type III and type IV phase behaviour.

The low temperature branch of the three-phase line now contains an UCEP, whereas the high
temperature branch of the three-phase line contains a LCEP. This leads to type IV phase
behaviour. If we approach the DCEP from type IV phase behaviour then the DCEP is formed
when an UCEP and a LCEP of two different three-phase lines merge.
The DCEP is a global transition state, i.e., nothing peculiar happens in the p,T-projection
of system showing a DCEP as far as topology of the critical curves is concerned. The DCEP is
the point on the critical curve, that emerges from the critical point of component 2, where a
critical phase (L1=L2) coexists with a vapour phase. Like the TCP, the DCEP can only be located
experimentally in quasi-binary mixtures. Experimental research by Brunner [22] of mixtures of
n-alkanes(1) + ammonia(2) shows the existence of a DCEP, although Brunner [22] calls the
DCEP a "hypercritical" point. In the pseudo-binary mixture n-C18.5H39(1) + NH3(2), the p,T-
projection of the critical curve lies only 0.05 MPa above the L1L2V three-phase line and is almost
parallel to it. The pseudo-component n-C18.5H39 is composed of n-octadecane and n-nonadecane.

The derivation of the mathematical equations that define a DCEP follows from
straightforward thermodynamic relations. One of the conditions is the tangent slope condition.
The tangent slope condition states that the slope of the critical curve is equal to the slope of the
L1L2V three-phase line at the DCEP. We start the derivation of an expression for the slope of the
three-phase line in the p,T-projection from the following fundamental equation:

d G = - S d T +V d p + Gx d x (5.23)

The chemical potentials can be calculated from:


μ 1 = G - x G x and μ 2 = G + (1 − x) G x (5.24)
Transition states 125

The differentials of the chemical potential can be determinated by differentiating both


expressions in Eq.(5.24) with respect to the composition x:

d μ 1 = d G - G x .d x - x.d G x (5.25)

d μ 2 = d G - G x .d x + ( 1 - x ).d G x

Eq.(5.23) can be inserted into both equations of Eq.(5.25) to eliminate dG. The analogous
expression for dGx, dGx= -Sx dT + Vx dp + G2x dx, is used to eliminate dGx from both equations
in Eq.(5.25). The remaining expressions for dμ1 and dμ2 are:

d μ 1 = (- S + x. S x ).d T + ( V - x.V x ).dp - x.G 2x .dx (5.26)

d μ 2 = (- S - ( 1 - x ). S x ).d T + ( V + ( 1 - x ).V x ).dp + ( 1 - x ).G 2x .dx (5.27)

At a critical endpoint two phases; a critical and an alternate phase; are in equilibrium. The
alternate phase will be denoted by the superscript a and the critical phase by the superscript c.
Deiters [5] has postulated that:

d μ ic = d μ ia with i = 1 , 2 (5.28)

Insertion of Eq.(5.26) and Eq.(5.27) applied to the critical and non-critical phase, into Eq.(5.28)
yields a set of two linear equations, one equation per component. When these equations are
subtracted from another we obtain (with G2xc=0):

c c a a a a
S x .d T - V x .dp = S x .d T - V x .d p - G 2x .d x (5.29)

With Eq.(5.29) the variables Sxa, Vxa and G2xa are eliminated from the elaborated Eq.(5.28). The
remaining terms can be rearranged into:

d p S c - S a - ( x c - x a ). S cx
= (5.30)
d T V c - V a - ( xc - x a ).V cx

Eq.(5.30) is the slope of the three-phase line at a critical endpoint in a p,T-projection. For a
DCEP this slope is the same as the slope of the critical curve. The slope of a critical curve is
derived more easily. When the fundamental equation, Eq.(5.23) is differentiated twice with
respect to the composition variable x, we obtain the following expression:
(5.31)
d G 2x = - S 2x .d T + V 2x .d p + G 3x .d x
Along the critical curve dG2x=0 and G3x=0. This leads to the following expression for the slope
of the critical curve in a p,T-projection:
126 Chapter 5

⎛ d p ⎞ S c2x
⎜ ⎟= c (5.32)
⎝ d T ⎠c V 2x

When the slopes of the three-phase line and the critical curve are equated, we obtain the so-called
'slope criterion':
c c a c a c
S 2x S - S - ( x - x ). S x
c
= c a c a c
(5.33)
V 2x V - V - ( x - x ).V x

Additional equations are G2x=G3x=0 and μia=μic with i=1,2. The remaining problem is to find
expressions for the derivatives Vjx and Sjx, that are defined at constant pressure and temperature.
In analogy to the derivation of the critical conditions, these partial derivatives also can be
expressed in terms of partial Helmholtz energy derivatives with the Jacobian transformation
method of Shaw [23]:

AVx
V x= - (5.34)
A2V

1
V 2x = - .( AV2x + 2. A2Vx .V x + A3V .V 2x ) (5.35)
A2V

S x = - ( ATx + ATV .V x ) (5.36)

S 2x = - ( AT2x + 2. ATVx .V x + AT2V .V x + ATV .V 2x )


2
(5.37)

5.2.3 The mathematical double point

A third transition state between two types of phase behaviour is given by the
mathematical double point (MDP). A MDP occurs if two critical curves intersect in the p,T,x-
space of a binary mixture. The MDP is often seen as being unstable by definition [24]. However,
in this section it will be proven that two types of MDPs exist: a stable and an unstable one. In the
global phase diagram the MDPs can lead to a boundary between two different types of phase
behaviour. Because not much attention so far has been given to the MDP we will extensively
derive the mathematical conditions that lead to a MDP.
We will start by considering a MDP that is unstable. We refer to this type of MDP as a
MDP of the first kind. By definition, at a MDP, two critical curves intersect or the critical curve
is said to be self-crossing. An example of a binary mixture that shows a MDP is given in Fig.5.8.
It is shown in this figure that this p,T-projection does not form a transition state between two
types of phase behaviour, i.e., type IV phase behaviour is shown in this figure. In Fig.5.8 the
solid curves that are marked with 1 and 2 represent the vapour pressure curves.
Transition states 127

The open circles represent the pure


component critical points and the critical
endpoints. The dotted curves represent the
stable parts of the critical curves. The dashed-
dotted curves represent the L1L2V three-phase
lines. The metastable and unstable critical
curves are shown as solid lines and the MDP as
a solid circle.
The mathematical conditions for the
MDP in reduced density coordinates were
derived by Meijer [24]. He also derived the
necessary conditions for a MDP in the equation
of state variables V, T and x [25]. The fact that
two critical curves intersect in the V,T,x-space
leads to the observation that at this intersection
point there is no unique direction of the critical
curve, in either the V,x-plane nor in the T,x-
plane. This observation was used by Meijer
Figure 5.8: Schematic drawing of a p,T-
[25] to derive the conditions for a MDP in the
projection that contains a
independent variables V,T and x. To obtain an mathematical double point.
expression for (∂V/∂x) at a certain point along
the critical curve, we make a Taylor series expansion around this critical point. The derivation of
the MDP conditions can also be made by studying (∂T/∂x) along the critical curve. The choice
between V and T is quite arbitrary. To avoid confusion about the paths of differentiation we
define GII=G2x and GIII=G3x at constant pressure and temperature. The Taylor-series along the
critical curve, truncated after the first term around this critical point is given by Eqs.(5.38) and
(5.39).

⎛ ∂ G II ⎞ ⎛ ∂ G II ⎞ ⎛ ∂ G II ⎞
dG = 0 = ⎜⎜ ⎟⎟ .dT + ⎜⎜ ⎟⎟ .dV + ⎜⎜ ⎟⎟ .dx
II
(5.38)
⎝ ∂T ⎠V,x ⎝ ∂V ⎠T,x ⎝ ∂x ⎠V,T

⎛ ∂ G III ⎞ ⎛ ∂ G III ⎞ ⎛ ∂ G III ⎞


dG =III
0 = ⎜
⎜ ∂T ⎟ ⎟ .dT + ⎜
⎜ ∂V ⎟ ⎟ .dV + ⎜⎜ ⎟⎟ .dx (5.39)
⎝ ⎠V,x ⎝ ⎠T,x ⎝ ∂x ⎠V,T

Elimination of the dT-terms in Eqs.(5.38) and (5.39) leads to:

⎛ ∂ G II ⎞ ⎛ ∂ G III ⎞ ⎛ ∂ G III ⎞ ⎛ ∂ G II ⎞
⎜ ⎟⎜ ⎟-⎜ ⎟⎜ ⎟ (5.40)
⎛ ∂V ⎞ ⎝ ∂T ⎠ ⎝ ∂x ⎠ ⎝ ∂T ⎠ ⎝ ∂x ⎠
⎜ ⎟ =
⎝ ∂ x ⎠G II=G III=0 ⎛ ∂ G II ⎞ ⎛ ∂ G III ⎞ ⎛ ∂ G III ⎞ ⎛ ∂ G II ⎞
⎜ ⎟⎜ ⎟-⎜ ⎟⎜ ⎟
⎝ ∂V ⎠ ⎝ ∂T ⎠ ⎝ ∂V ⎠ ⎝ ∂T ⎠
128 Chapter 5

If a critical curve is self-crossing at a certain point, the critical curve has no unique direction at
this point. This means that along the critical curve the derivative (∂V/∂x) cannot exist at a MDP.
Because (∂V/∂x) must be undefined at a MDP both the numerator and the denominator of
Eq.(5.40) must be zero. This leads to the following two conditions to be fulfilled at a MDP
(besides the normal critical conditions: G2x=0 and G3x=0):
II III III II
GT . G x - GT . G x = 0 (5.41)

II III III II
GV . GT - GV . GT = 0 (5.42)

In Eqs.(5.41) and (5.42) again the shorthand notation of Rowlinson and Swinton [21] is used.
Deiters [26] has established a relationship between the conditions of a MDP and G4x. When
Eq.(5.41) is multiplied by GVII and Eq.(5.42) is multiplied by GxII and the resulting equations are
added, the terms containing GTIII cancel out. In the remaining equation the factor GTII can be
eliminated:
II III II III
GV .G x - G x .GV = 0 (5.43)

According to the Jacobian transformation method [23], it can be proven that:


III III
G 4x = G x + V x .GV (5.44)

When GxIII=G4x-Vx.GVIII, obtained from Eq.(5.44), is substituted into Eq.(5.43) and the remaining
terms are rearranged, we obtain:
II III II II
G 4x .GV - GV .( G x + V x .GV ) = 0 (5.45)

According to the Jacobian transformation method of Shaw [23], the expression within the
brackets in Eq.(5.45) equals G3x, which equals zero along the critical curve. This simplifies
Eq.(5.45) significantly:
II
G 4x .GV = 0 (5.46)

With Eq.(5.46) we have shown that two types of MDPs exist. We propose the following
nomenclature: MDPs where G4x=0 will be referred to as MDPs of the first kind and MDPs
where GVII=0, will be referred to as MDPs of the second kind. It is stressed here that both types
of MDPs can be calculated by Eqs.(5.41) and (5.42). MDPs that have been calculated from an
equation of state so far are MDPs of the first kind [23,27-34].
Transition states 129

It can be understood from section 5.2.1, that as long as G5x≠0, a MDP of the first kind is
unstable. A schematic example of a MDP of the first kind in p,T-plane was already given in
Fig.5.8. It was proven by Deiters [35] that if G4x=0 along a critical curve, a cusp occurs in p,T-
plane. This is exactly what happens in the p,T-plane in the event of a MDP of the first kind. At a
MDP of the first kind two cusps appear in p,T-plane that touch each other, see Fig.5.8. If G5x=0
at a MDP of the first kind then the MDP becomes tricritical. The tricritical MDP is called the
"Van Laar point" [24]. The local stability of the Van Laar point is dependent on the sign of G6x.

In Fig.5.9 the p,T-projection is shown of the


critical curves of a binary system with a Van
Laar point. At the Van Laar point four branches
of two critical curves and a three-phase line
coincide. One of the four branches is unstable.
The Van Laar point is represented in Fig.5.9 by
a black dot. All critical curves are represented by
dotted lines in this figure. In section 8.4, we will
discuss the Van Laar point and the way the
critical curves behave near this point in the T,x-
plane.
A MDP of the second kind, where
II
GV =0 holds, can be stable. As G4x is not equal
to zero at this type of MDP it cannot be
recognized in the p,T-plane by the two cusps.
Because G4x can be positive in a MDP of the
second kind, it can be stable and thus form a
transition state between two types of phase Figure 5.9: Schematic p,T-projection of the
behaviour. A schematic drawing of a MDP of critical curves of a binary
the second type is shown in Fig.5.10. mixture with a Van Laar point.

MDPs of the second kind have been determined experimentally in a number of systems.
An extensive overview on experimental MDPs of the second kind has been given by Schneider
[36], although Schneider did not use the term MDP. However, MDPs of the second kind have
not been calculated with an equation of state previously. We will show in chapter 8 that with the
SPHCT equation it is possible to calculate phase diagrams that show a MDP of the second kind.

In the previous equations the molar Gibbs energy was used to define the criteria for the
MDP. However, to calculate MDPs with the SPHCT equation we need expressions in terms of
partial molar Helmholtz energy derivatives. For the GII-derivatives the analogous expressions in
terms of molar Helmholtz energy derivatives are:
130 Chapter 5

Figure 5.10: The MDP of the second kind in the p,T-plane. It is shown that a p,T-projection
with a MDP of the second kind forms the transition state between two different
forms of type VI phase behaviour.

⎛ AVx ⎞
G x = A3x - AV2x ⎜⎜ ⎟⎟ - AVx Q x
II
(5.47)
⎝ A2V ⎠

⎛ AVx ⎞
GV = AV2x - A2Vx ⎜⎜ ⎟⎟ - AVx QV
II
(5.48)
⎝ A2V ⎠

⎛ AVx ⎞
GT = AT2x - ATVx ⎜⎜ ⎟⎟ - AVx QT
II
(5.49)
⎝ A2V ⎠

with:
⎛A ⎞ ⎛A ⎞ ⎛ A ⎞
Q x = ⎜⎜ V2x ⎟⎟ - ⎜⎜ 2Vx ⎟⎟ . ⎜⎜ Vx ⎟⎟ (5.50)
⎝ A2V ⎠ ⎝ A2V ⎠ ⎝ A2V ⎠

⎛A ⎞ ⎛A ⎞ ⎛ A ⎞
QV = ⎜⎜ 2Vx ⎟⎟ - ⎜⎜ 3V ⎟⎟ . ⎜⎜ Vx ⎟⎟ (5.51)
⎝ A2V ⎠ ⎝ A2V ⎠ ⎝ A2V ⎠
Transition states 131

⎛A ⎞ ⎛A ⎞ ⎛ A ⎞
QT = ⎜⎜ TVx ⎟⎟ - ⎜⎜ T2V ⎟⎟ . ⎜⎜ Vx ⎟⎟ (5.52)
⎝ A2V ⎠ ⎝ A2V ⎠ ⎝ A2V ⎠

The partial derivatives of GIII are in terms of molar Helmholtz energy derivatives:

2 3
⎛ AVx ⎞ ⎛A ⎞ ⎛A ⎞
G = A4x - 3 AV3x ⎜⎜ ⎟⎟ + 3 A2V2x ⎜⎜ Vx ⎟⎟ - A3Vx ⎜⎜ Vx ⎟⎟ + P.Q x
III
x (5.53)
⎝ A2V ⎠ ⎝ A2V ⎠ ⎝ A2V ⎠

2 3
⎛ AVx ⎞ ⎛A ⎞ ⎛A ⎞
G = AV3x - 3 A2V2x ⎜⎜ ⎟⎟ + 3 A3Vx ⎜⎜ Vx ⎟⎟ - A4V ⎜⎜ Vx ⎟⎟ + P.QV
III
V (5.54)
⎝ A2V ⎠ ⎝ A2V ⎠ ⎝ A2V ⎠

2 3
⎛ AVx ⎞ ⎛A ⎞ ⎛A ⎞
G = AT3x - 3 ATV2x ⎜⎜ ⎟⎟ + 3 AT2Vx ⎜⎜ Vx ⎟⎟ - AT3V ⎜⎜ Vx ⎟⎟ + P.QT
III
T (5.55)
⎝ A2V ⎠ ⎝ A2V ⎠ ⎝ A2V ⎠

with:

2
⎛ ⎞ ⎛ ⎞
P = - 3 AV2x + 6 A2Vx ⎜⎜ AVx ⎟⎟ - A3V ⎜⎜ AVx ⎟⎟ (5.56)
⎝ A2V ⎠ ⎝ A2V ⎠

Besides the critical conditions, G2x=0 and G3x=0, the conditions in Eqs.(5.41) and (5.42)
have to be fulfilled. These conditions are derived from the fact that at a MDP the critical curve
has no unique direction in the V,x-plane. The same holds for the direction in the T,x-plane. This
will be proven as follows. Instead of the elimination of the dT-terms from Eqs.(5.38) and (5.39)
we can just as well eliminate the dV-terms. We then obtain an expression for ∂T/∂x along the
critical curve. Because there is no unique direction of the critical curve in the T,x-plane we set
the numerator and denominator of this expression equal to zero. This gives:

II III III II
GV . G x - GV . G x = 0 (5.57)

II III III II
GT . GV - GT . GV = 0 (5.58)

We must prove that Eqs.(5.57) and (5.58) are identical with Eqs.(5.41) and (5.42). It is obvious
that Eq.(5.42) is identical to Eq.(5.58) and it was also proven that Eq.(5.43) which is identical to
Eq.(5.57), can be obtained from Eqs.(5.41) and (5.42). In other words: it is proven that a MDP in
the V,x-plane automatically implies a MDP in the T,x-plane.
132 Chapter 5

A final point we want to make about MDPs is the determination of the two directions of
the critical curves at a MDP. To determine these directions in the V,x-plane, the series
expansions in Eqs.(5.38) and (5.39) should be truncated after the quadratic terms in V and x.
Elimination of the dT-terms then leads to the following equation:

(G III
T G 2V - G 2V GT ) dV + 2 ( GT GVx - GVx GT ) dV dx + ( GT G 2x - G 2x GT ) dx = 0
II III II 2 III II III II III III II 2
(5.59)

The factor 1/2 of the Taylor series expansion has been eliminated from Eq.(5.59). We introduce a
shorthand notation so that Eq.(5.59) can be written as:
2 2
M 2V dV + 2 M Vx dV dx + M 2x dx = 0 (5.60)

where:

M iVjx = ( GT G iVjx - G iVjx GT )


III II III II
(5.61)

If Eq.(5.60) is divided by dx2 a quadratic expression for (dV/dx)c is obtained:

2
⎛dV ⎞ ⎛dV ⎞
M 2V ⎜ ⎟ +2 MV x ⎜ ⎟+ M 2x=0 (5.62)
⎝dx ⎠ ⎝dx ⎠

Solutions for (dV/dx)c that can be derived from this quadratic expression are:

2
⎛dV ⎞ M Vx ± M Vx - M 2V M 2x
⎜ ⎟ = - (5.63)
⎝ d x ⎠c M 2V M 2V

5.2.4 The azeotropic boundary curves

The fourth type of curve that plays an important role in the global phase diagram is the
azeotropic boundary curve. According to the Van der Waals equation of state (VDW), the
azeotropic composition of a mixture is independent of the temperature, if the unlike molecules
are equal-sized [1]. Thus for a mixture of equal-sized molecules there is either azeotropy from
T=0 K up to the critical curve or there is no azeotropy at all. The transition state between these
two possibilities is formed when the azeotropic curve coincides with the vapour pressure curve
of one of the components in the p,T-plane. According to the VDW equation the azeotropic
Transition states 133

composition is temperature dependent if the mixture is composed of molecules of different size.


This means that azeotropy might occur at T = 0 K but might disappear at a certain temperature
when the azeotropic curve runs into one of the vapour pressure curves. Van Konynenburg [1]
concluded that there is not a unique boundary curve between phase diagrams with and without
azeotropy in global phase diagram, if the molecules differ in size (ξ ≠ 0). In this case there is a
transition region. Van Konynenburg [1] calculated the boundaries of this transition region: one is
formed when the azeotropic line runs into the critical point of one of the pure components, the
"high-temperature" azeotropy limit, and the other boundary is formed when the azeotropic line
runs into the vapour pressure curve of one of the pure components at T = 0 K, the "low-
temperature" azeotropy limit.

For the SPHCT equation the situation is slightly different. If ξ = 0 and χ = 0, see
Eqs.(5.1)-(5.4), high-temperature and low-temperature azeotropy may exist, because the
azeotropic composition is temperature dependent in this case. The low-temperature azeotropy
boundary cannot be evaluated numerically with the SPHCT equation of state because of the
temperature-dependence of the attractive term of the equation of state. The conditions for the
high-temperature azeotropic line in the global phase diagram are mathematically simple:

μ 1L = μV1 at x = 0 or x = 1 (5.64)

μ 2L = μV2 at x = 0 or x = 1 (5.65)

x = y = 0 or x = y = 1 (5.66)

For the case ξ = χ = 0 we can obtain an analytical expression for the azeotropic lines in the
global phase diagram. From Eqs.(5.64) and (5.65) it follows that AxL = AyV [1]. In the case when
ξ = χ = 0, the expression for Ax can be simplified to:

Z m . < cV Y ′ >
*
Ax = - c R T . (5.67)
c V + < cV * Y >

where <cV *Y'> = ∂ <cV *Y>/∂ x. The condition AxL = AyV leads to:

< cV * Y ′ >L < cV * Y ′ >V


L * L
= V * V
(5.68)
cV + < cV Y > cV + < cV Y >
134 Chapter 5

Because <cV *Y'> and <cV *Y> only depend on temperature and composition, they are equal in
the vapour and the liquid phase in the case of azeotropy. Eq.(5.68) holds if either VL=VV or <cV
*
Y'>=0. Along the azeotropic curve in a p,T-diagram V L≠V V (except when the azeotropic line
meets the critical curve). So, a general condition for azeotropy in binary mixtures with ξ=χ=0 is
<cV *Y'>=0. Elaboration of the latter expression leads to:

T11* T12* *
T22
-( 1- x ) e 2T
+ ( 1 - 2x ) e 2T
+xe 2T
=0 (5.69)

For x=1 Eq.(5.69) leads to λ=ζ and for x=0 Eq.(5.69) leads to λ=-ζ. The azeotropic composition
is temperature dependent according to the following equation:

T11* T12*
(5.70)
az e 2T
-e 2T
x =
2 T11* T12* *
T22
e 2T
-2e 2T
+e 2T

For mixtures composed of molecules that differ in size and/or flexibility, the conditions of the
boundary curves for high-temperature azeotropy in the global phase diagram, Eqs.(5.64)-(5.66),
are evaluated numerically.

5.2.5 The azeotropic critical endpoint

The transition state between type III-A and type III-H phase behaviour is formed by a
p,T-projection of a binary mixture that contains an azeotropic critical endpoint (ACEP). The
nomenclature may be confusing because various other names are used for this type of phase
behaviour. According to Van Konynenburg [1] there are two possibilities for type III behaviour.
If the L1L2V three-phase line is situated between the vapour pressure curves of both components
in the p,T-projection, the mixture shows type III phase behaviour, and if the three-phase line is
above both vapour pressure curves the binary mixture shows type III-HA phase behaviour. The
"HA" refers to heteroazeotropy. Boshkov [37] has used the term type "III-H" to denote type III
with heteroazeotropy. We will also use the term type "III-H" instead of type "III-HA". The
ACEP is more or less similar to the high-temperature azeotropic transition state in the p,T-plane.
The ACEP is formed if a binary mixture that exhibits type III-H phase behaviour, also shows
homoazeotropy. An extra condition for an ACEP is that the azeotropic curve must meet the
L1L2V-three-phase line in its endpoint [37]. At lower temperatures the azeotropic curve is located
above the L1L2V-three-phase line. A schematic drawing of a p,T-projection of a binary system
with an ACEP is given is Fig.5.11.
Transition states 135

Figure 5.11: The azeotropic critical endpoint in p,T-space. It is shown that a p,T-projection
with an ACEP forms the transition state between type III-H and type III-A phase
behaviour.

In Fig.5.11, the solid curves that are marked with 1 and 2 represent the vapour pressure curves.
The open circles that are connected to these solid curves represent the pure component critical
points. The dotted curves represent the critical curves. The dashed-dotted curves represent the
L1L2V three-phase lines. The open circles connected to the dashed dotted curve represent the
upper critical endpoint. The ACEP is shown as a solid circle in the figure in the middle. The
mathematical conditions are given by Deiters and Pegg [5]:

A2V = AVx = A3V = 0 (5.71)

μ ic = μ ia with i = 1, 2 (5.72)

5.2.6 The zero temperature endpoint

The transition state between type I and type II phase behaviour is formed by the zero
temperature endpoint (ZTEP) according to the Van der Waals equation [1]. A ZTEP is obtained
if in a type II mixture the critical endpoint on the L1L2V-three-phase line is located at T=0 K. At a
ZTEP the critical molar volume approaches the covolume, with p and T approaching zero. A
ZTEP has never been observed experimentally as, obviously, it will be obscured by one or more
solid phases.
The conditions that are mentioned above have a number of interesting implications for
the SPHCT equation that will be studied in this section. We will start by considering the
136 Chapter 5

consequences of the critical molar volume approaching the closest packed volume in the Van der
Waals (VWD) equation of state. Van Konynenburg [1] also derived conditions for the ZTEP.
Because his derivation was placed in the context of excess properties, we will derive the
conditions directly from the molar Gibbs energy. The VDW equation is given by:

RT a
p= - (5.73)
V -b V2

with the following mixing rules:

n n n n
a = ∑ ∑ xi x j aij and b = ∑ ∑ xi x j bij (5.74)
i=1 j=1 i=1 j=1

If we only consider the condition that the critical molar volume approaches the covolume, then
the critical conditions G2x=G3x=0 simplify to:

mix attr
G 2x = A2x + A2x = 0 (5.75)

mix attr
G 3x = A3x + A3x = 0 (5.76)

where the superscript mix denotes ideal mixing term (i.e., the RT{xln(x)+(1-x)ln(1-x)}-term). The
simplifications given above are strictly valid for the VDW equation of state. It can be shown that
the repulsive terms of the molar Helmholtz energy cancel out in the equations for G2x and G3x if
the molar volume approaches the covolume. For the VDW equation the A3xattr-term in Eq.(5.76)
is equal to zero. This leads to the following expressions:
2
∂ a (5.77)
RT ∂ 2
G 2x = - x =0
( 1- x ) x b

R T ( - 1+ 2 x )
G 3x = 2 2
=0 (5.78)
x ( x-1 )

This set of equations has two solutions that are recognized readily. It can be seen from Eq.(5.78)
that the A3xmix-term is zero for xc=0.5, independent of the value of the critical temperature. The
critical temperature can subsequently be calculated from Eq.(5.77). This gives the following
solution for Tc:
Transition states 137

1 ∂2a
Tc= (5.79)
4 b R ∂ x2

At closest-packing of the molecules in a binary mixture, i.e., when V = b, there exists one critical
point with xc = 0.5 and Tc specified by Eq.(5.79). The pressure at this socalled "jamming point"
[24] is infinite. To illustrate this peculiar phenomenon we calculated one example with the VDW
equation. For the binary system methane(1)+ethane(2) with k12 = 0, the temperature of the
jamming point is 80 K. This point can be seen as the limiting case of the type II liquid-liquid
immiscibility critical curve at infinite pressures. In general, the jamming point also can belong to
the liquid-liquid critical curve of a type III mixture. This depends solely on whether the liquid-
liquid critical curve is connected to an UCEP or a pure component critical point at lower
pressures. According to the VDW equation the phase behaviour of the mixture of methane(1) +
ethane(2) belongs to type II.
The second solution of Eqs.(5.77) and (5.78) leads to the ZTEP. If T = 0 K, Eq.(5.78) is
satisfied for all possible values for xc. If T = 0 K, Eq(5.77) is only satisfied if (∂ 2a/∂ x2) = 0. For
the ZTEP there is no unique pressure because both the numerator and the denominator of the
RT/(V-b)-term in Eq.(5.72) become zero. In the parameters of the global phase diagram, the
elaborated condition for a ZTEP, i.e., (∂ 2a/∂ x2) = 0, leads to the following expression: λ = 0. At
the ZTEP the critical pressure is undefined and one can say that the liquid-liquid critical curve
merges with the T = 0-axis.

For the SPHCT equation of state the situation is different. It was already stated in section
3.4 that if the molar volume approaches the closest packed volume according to the SPHCT
equation of state, then the value of the pressure is still finite. This is a limitation of the Carnahan-
Starling expression that is used for the repulsive term. However, in practical situations this
"maximum pressure" limit plays no role. To give an impression, we have calculated the
maximum attainable pressure for methane at 298.15 K. With the pure component parameters
from table 4.1, the pressure at closest packing is 1.9 105 bar. In the p,T,x-plane there is a "ceiling"
according to the SPHCT equation of state above which no phase equilibrium calculations can be
made. This plane starts at (p,T) = (0,0) and rises very steeply. In engineering calculations this
plane is not even approached because for most substances at ambient temperatures, the plane lies
at about 104 bar or even higher.
We will continue by looking for solutions from the SPHCT equation that are similar to
the two solutions that were found for the Van der Waals equation at closest-packing. We start
with the jamming point solution and proceed with the ZTEP-solution. Because the value of the
repulsive term in the pressure equation at the jamming point (i.e., the situation in V,T,x-space
when the critical molar volume approaches the closest-packed volume of the mixture) is still
finite, we cannot neglect the other terms with respect to the repulsive terms in the critical
conditions. Consequently the simplications made in Eqs.(5.75) and (5.76) are not valid for the
SPHCT equation of state. The only way to approximate the jamming point(s) is by calculation of
138 Chapter 5

critical curves at densities near the closest-packing density. It turns out that solving the critical
conditions for values of Vc that are close to the closest-packed volume show that a high- pressure
critical curve can exist. This gives rise to two jamming points.
In Figs.5.12 and 5.13 the high-pressure critical curves of four binary mixtures are shown
in the p,T- and V,x-plane (solid curves), including the closest-packing limit (dotted line). The
fictitious pure component parameters are:
c1 = 2.000 c2 = 2.000
* 3
V1 = 27.00 cm .mole -1
V2*= 35.00 cm3.mole-1
T1*= 205.00 K T2*= 260.00 K

Figure 5.12: p,T-projection of calculated Figure 5.13: V,x-projection of calculated


high-pressure critical curves high-pressure critical curves
for various values of k12. for various values of k12.
Although these SPHCT parameters are fictitious, their values are realistic physically. The critical
pressures and temperatures are 42.7 bar and 323.6 K for component 1 and 41.8 bar and 410.4 K
for component 2. In Figs.5.12 and 5.13 k12 = 0.0425 for curve 1, k12 = 0.0420 for curve 2,
k12 = 0.0410 for curve 3 and k12 = 0.0400 for curve 4. The closest-packing limit in the V,x-plane is
exactly calculated, in the p,T-plane the closest packing limit is approximated by
pCPL ≈ <c>.RT/<V*>.106.74. This approximation only takes the repulsive contribution to the
closest-packing limit into account at a fixed composition of xc = 0.30. It is highly probable that
even if binary mixtures could be found that have these pure component parameters and this
binary interaction parameter, the high pressure critical curve would be obscured by the presence
of solid phases.
Mathematically the origin of the high-pressure critical curve is to be found in the
composition dependence of the attractive term. Because the attractive term of the Van der Waals
equation of state is quadratic in terms of composition, the A3xattr-term vanishes, giving rise to
Transition states 139

only one jamming point at xc = 0.5. For other equations of state, like the Redlich-Kwong and the
SPHCT equation, the atrractive term is more complicated and the A3xattr-term does not vanish in
the expression for G3 = 0. As a consequence more than one jamming point can exist at different
compositions and temperatures. It should be emphasized that, due to the structure of the
Carnahan-Starling expression, the pressure of a jamming point, predicted from the SPHCT
equation, is still finite.
As shown above, two solutions exist for the Van der Waals equation at closest packing.
The first solution led to the jamming point and the second solution led to the ZTEP. At T = 0 K a
solution can only be obtained from the VDW equation if the molar volume approaches the
closest-packed volume. In this case the pressure is undefined. According to the SPHCT equation
the pressure is not undefined at T = 0 K, as is the case with the VDW equation, but the pressure
equals zero regardless of the value of the molar volume or the composition. The boundary case
of a ZTEP, calculated from the SPHCT equation is therefore different from the one obtained for
the Van der Waals equation, as is shown in Fig.5.14. The only point of the ZTEP curve in the
global phase diagram that can be solved analytically, is in case ξ = χ = 0, where a ZTEP occurs at
(ζ, λ) = (0,0).

Figure 5.14: (a) Approaching the ZTEP with a Van der Waals type equation of state.
(b) Approaching the ZTEP with the SPHCT equation of state.

In Fig.5.14 the solid curves represent the vapour pressure curves, the open circles the critical
points and the critical endpoint. The dotted curves represent the critical curves. The solid circle is
the point where the liquid-liquid critical curve becomes unstable, if this point is approached from
high pressures. Fig.5.14 shows how the liquid-liquid critical curve shifts towards the T=0 axis.
140 Chapter 5

We can cause this shift of the liquid-liquid critical curve by lowering the value of the binary
interaction coefficient k12. Fig.5.14(a) shows that the liquid-liquid critical curve will eventually
merge with the line T = 0K for a Van der Waals type of equation. In this limiting case of type II
phase behaviour a ZTEP is formed. For a Carnahan-Starling type of equation, such as the
SPHCT equation, the liquid-liquid critical curve coincides with the origin (p,T) = (0, 0). This is
shown in Fig.5.14(b).

NOMENCLATURE

A molar Helmholtz energy ξ dimensionless parameter ratio


b polynomial coefficient χ dimensionless parameter ratio
c polynomial coefficient φ number of extra conditions
ci one-third of the external degrees of
freedom of molecule i Subscripts and superscripts
d polynomial coefficient
e polynomial coefficient 0 test phase
F tangent plane distance 1 coexisting phase 1
f number of degrees of freedom 2 coexisting phase 2
f polynomial coefficient a alternate phase
G molar Gibbs energy c critical phase
k order of a critical point I component i
k12 binary interaction coefficient
n number of components
p pressure
R gas constant
r number of phases
S molar entropy
T (absolute) temperature
Ti* characteristic temperature
V molar volume
Vi* closest-packed molar volume of
component i
x composition, i.e., the mole fraction of
the heavier component

Greek symbols

ζ dimensionless parameter ratio


λ dimensionless parameter ratio
μi chemical potential of component i
Transition states 141

REFERENCES

1. P.H. van Konynenburg. Critical lines and phase equilibria in binary mixtures. PhD-thesis
University of California Los Angeles. (1968).
2. D. Furman, S. Dattagupta and R.B. Griffiths. Global phase diagram for a three-component
model. Phys. Rev. B, 15, (1977), 441-464.
3. R.L. Scott and P.H. van Konynenburg. 2. Static properties of solutions: van der Waals and
related models for hydrocarbon mixtures. Discuss. Faraday Soc. 49, (1970), 87-97.
4. P.H. van Konynenburg and R.L. Scott. Critical lines and phase equilibria in binary van der
Waals mixtures. Phil. Trans. 298A, (1980), 495-540.
5. U.K. Deiters and I.L. Pegg. Systematic investigation of the phase behavior in binary fluid
mixtures. I. Calculations based on the Redlich-Kwong equation of state. J. Chem. Phys. 90,
(1989), 6632-6641.
6. Th. Kraska and U.K. Deiters. Systematic investigation of the phase behavior in binary fluid
mixtures. II. Calculations based on the Carnahan-Starling-Redlich-Kwong equation of state.
J. Chem. Phys. 96, (1992), 539-547.
7. M.L. Michelsen. The isothermal flash problem. Part I. Stability. Fluid Phase Equilibria 9,
(1982), 1-19.
8. M.L. Michelsen. The isothermal flash problem. Part II. Phase-split calculation. Fluid Phase
Equilibria 9, (1982), 21-40.
9. M.L. Michelsen. Calculation of critical points and phase boundaries in the critical region.
Fluid Phase Equilibria 16, (1984), 57-76.
10. M.L. Michelsen and R.A. Heidemann. Calculation of tri-critical points. Fluid Phase
Equilibria 39, (1988), 53-74.
11. C.M. Knobler and R.L. Scott. Multicritical points in fluid mixtures: experimental studies.
In: C. Domb and J.L. Lebowitz (Editors), Phase transitions and critical phenomena.
Academic Press. Vol.9, chap.2, 163-231.
12. R.L. Scott. Multicritical phenomena in fluid mixtures. In J.V. Sengers (Editor), Proc. of the
8th. Symp. on Thermophysical Properties. Vol.1: Thermophysical properties of fluids.
ASMF, 397-404.
13. K.D. Luks. The occurence of multiphase equilibria behaviour. Fluid Phase Equilibria 29,
(1986), 209-224.
14. J.P. Kohn and K.D. Luks. Liquid-liquid-vapor equilibria in cryogenic LNG mixtures.
Research report RR-49. Gas Processors Association, Tulsa, OK, (1981).
15. J.P. Kohn, R.C. Merrill and K.D. Luks. Liquid-liquid-vapor equilibria in cryogenic LNG
mixtures-Phase II. Research report RR-60. Gas Processors Association, Tulsa, OK, (1982).
16.J.P. Kohn, R.C. Merrill and K.D. Luks. Liquid-liquid-vapor equilibria in cryogenic LNG
mixtures. Phase III- Nitrogen rich systems. Research report RR-67. Gas Processors Associa-
tion, Tulsa, OK, (1983).
17. J.L. Creek, C.M. Knobler and R.L. Scott. Tricritical points in ternary mixtures of
hydrocarbons. J. Chem. Phys. 67, (1977), 366-368.
18. J.L. Creek, C.M. Knobler and R.L. Scott. Tricritical phenomena in "quasi-binary" mixtures
of hydrocarbons. 1. Methane systems. J. Chem. Phys. 74, (1981), 3489-3499.
19. J.R. Wagner Jr., D.S. McCaffrey Jr. and J.P Kohn. Partial miscibility phenomena in the
ternary system ethane-n-hexadecane-n-eicosane. J. Chem. Eng. Data 13, (1968), 22-24.
20. G.D. Efremova and A.V. Shvarts. Liquid-liquid-gas equilibrium for the ethane-n-eicosane-
n-hexadecane system. Russ. J. Phys. Chem. 44, (1970), 470.
142 Chapter 5

21. J.S. Rowlinson and F.L. Swinton. Liquids and Liquid Mixtures Butterworths Monographs
in Chemistry, London, (1982).
22. E. Brunner. Fluid mixtures at high pressures. VI. Phase separation and critical phenomena
in 18 (n-alkane + ammonia) and 4 (n-alkane + methanol) mixtures. J. Chem. Thermodyna-
mics 20, (1988), 273-297.
23. Shaw. The derivation of thermodynamical relations for a simple system. Phil. Trans. Roy.
Soc. London A 234, (1935), 299-328.
24. P.H.E. Meijer, M. Keskin and I.L. Pegg. Critical lines for a generalized three state binary
gas-liquid lattice model. J. Chem. Phys. 88, (1988), 1976-1982. Erratum: J. Chem. Phys. 90,
(1989), 3408.
25. P.H.E. Meijer. Private communications. Delft, The Netherlands, (1989).
26. U.K. Deiters. Private communications. Bochum, Germany, (1991).
27. P.H.E. Meijer and M. Napiorkowski. The three-state lattice gas as model for binary gas-
liquid systems. J. Chem. Phys. 86, (1987), 5771-5777.
28. P.H.E. Meijer. Binary gas-liquid systems classification. Soc. Fr. de Chimie, Int. Symp. on
Supercritical Fluids, Oct. 17-19, Nice. M. Perrut ed., Vol. 1, (1988), 239-244.
29. P.H.E. Meijer. The influence of the chain length of long molecules on the equation of state
in binary gas liquid mixtures. J. Stat. Phys. 53, (1988), 543-548.
30. P.H.E. Meijer. Study of the critical line and its double point in the intermediate model.
Physica A 152, (1988), 359-364.
31. P.H.E. Meijer. The van der Waals equation of state around the van Laar point. J. Chem.
Phys. 90, (1989), 448-456.
32. P.H.E. Meijer, I.L. Pegg, J. Aronson and M. Keskin. The critical lines of the van der Waals
equation for binary mixtures around the van Laar point. Fluid Phase Equilibria 58, (1990),
65-80.
33. P.H.E. Meijer and I.L. Pegg. Structure of the critical lines for the lattice gas model. Physica
A 174, (1991), 391-405.
34. M. Keskin, M. Gençaslan and P.H.E.Meijer. Evaluation and comparison of critical lines for
various models of gas-liquid binary systems. J. Stat. Phys. 66, (1992), 885-896.
35. U.K. Deiters. Phänomenologie und Berechnung von Phasendiagrammen. Bochum,
Germany, (1992).
36. G.M. Schneider. Phasengleichgewichte in flüssigen Systemen bei hohen Drucken. Ber.
Bunsenges. Phys. Chem. 70, (1966), 497-612.
37. L.Z. Boshkov and V.A. Mazur. Phase equilibria and critical lines of binary mixtures of
Lennard-Jones molecules. Russ. J. Phys. Chem. 60, (1986), 16-19.
143

CHAPTER 6
COMPUTATIONAL TECHNIQUES

In chapter 5 it was discussed which peculiar critical phenomena can occur in a binary
fluid mixture, e.g. a tricritical point, a Van Laar point, etc. In the global phase diagram, most of
these peculiar critical phenomena are represented by curves. In this section it will be explained
how these curves are numerically evaluated. In section 6.1 an inventory is made of the equations
and unknowns to calculate boundary curves in the global phase diagram. In section 6.2 various
algorithms for solving a non-linear set of equations are considered. A detailed discussion of the
calculation of the tricritical curve in the global phase diagram is made in section 6.3. Finally, in
section 6.4 we will explain how the symbolic algebra program Maple was used to derive
mathematical expressions from thermodynamic equations.

6.1 Evaluation of the equations and unknowns

We start with the summing up of the unknowns and the equations for the calculation of
one point on the tricritical curve in the global phase diagram, see Table 6.1.

Table 6.1 :List of unknowns and equations to be solved for the calculation of a tricritical point.

Unknowns Number Equations Number

c1 , V1* , T1* 3 Choose c2 , V2* , T2* 3

c2 , V2* , T2* 3 Equation of state 1

k12 1 Definitions ζ , λ , ξ , χ 4

ptc , Ttc , xtc , Vtc 4 Tricritical conditions 4

ζ,λ,ξ,χ 4

Total 15 Total 12

In this chapter we will focus on the calculation of the tricritical curve in the global phase
diagram. From numerical point of view it does not matter whether we calculate the tricritical
curve or any other boundary curve in the global phase diagram. As we shall see it all comes
144 Chapter 6

down to solving a set of non-linear equations. It is shown in Table 6.1 that in order to calculate a
tricritical point, we have three degrees of freedom. It is obvious that we have to define three
unknowns for the system to be determined. These unknowns have to be chosen with care. In our
calculations, three of the four dimensionless parameter ratios (ζ, λ, ξ, χ) were given a fixed value,
e.g., ζ, ξ, χ or λ, ξ, χ.
For the calculation of the tricritical curve in the global phase diagram the following
procedure, proposed by Deiters [1] is used:

Iteration steps

1. Initialize, a. Choose c2 , V2* , T2*


b. Choose ξ , χ; i.e., c1 and V1* are determined with Eqs.(5.3)-(5.4)
c. Choose ζ; i.e. T1* is determined with Eq.(5.1)
2. Choose starting values Vtc0 , Ttc0 , xtc0 and k120.
3. Solve the tricritical conditions (G2x = G3x = G4x = G5x = 0). This gives us: Vtc , Ttc , xtc , k12
and thus λ.
4. Compute ptc from the equation of state.
5. Perform local and global stability tests on the calculated tricritical point.
6. Write all variables to an output device (printer, file).
7. Choose a new value for ζ (i.e., T1* is given a new value) and go to step 2.

If, instead of ζ, a value of λ was chosen in step 1c, step 3 would have given us the value of T1*
(and thus ζ) instead of k12 (and thus λ). Of course it is also possible to use ξ or χ as the variables
to be scanned or calculated. This general iteration scheme is also valid for the calculation of the
other curves in the global phase diagram. The only difference from the calculation of the
tricritical curve is then made in the mathematical conditions to be fulfilled in step 3. There are no
differences in the number or types of the variables.
The most important step in the iteration scheme is step 3, where four non-linear equations
have to be solved. This will be discussed in the following section.

6.2 Optimization methods

In general the problem of solving a set of non-linear equations can be transformed into a
minimization problem, in which the number of equations equals the number of unknowns. To
explain this, we define a vector of k functions, f(x), where f(x) = [f1(x), f2(x), ..., fk(x)]T and x =
[x1, x2, ..., xk]T. In this chapter, the notation according to the IUPAC-convention [2] is used: i.e.,
vectors are printed in bold face italic type (non-capital letters), matrices are printed in bold face
italic type (capital letters). The superscript T denotes a transpose of the corresponding vector or
the transpose of the corresponding matrix. In our example, the calculation of a tricritical point,
k = 4, f1 = G2x, f2 = G3x, f3 = G4x and f4 = G5x, x1 = xtc, x2 = Vtc, x3 = Ttc and x4 = k12.
Computational techniques 145

Solving f(x) = 0 gives us the same solution vector x as the minimization of the objective
function F(x), where

F ( x )= [ f 1 ( x ) ]2 + [ f 2 ( x ) ]2 + .....+ [ f k ( x ) ]2 (6.1)

To solve simultaneously a set of


non-linear equations, we used the
Marquardt algorithm [3]. In this
section the basic principles of the
Marquardt algorithm are
explained without going into
detail. A more detailed discussion
on non-linear optimization
methods can be found in [4].
Because the Marquardt
minimization method is a
combination of different
minimization methods, we will
start with a short consideration of
these minimization methods.
They are iterative and follow the
general scheme depicted in
Fig.6.1.
In this approach we try to find a
sequence x1, x2, . . . , xN of
vectors in the parameter space
such that xN minimizes F(x)
Figure 6.1: Flow diagram for iterative optimization approximately. Starting with
methods. some initial vector x1, each of the
following elements in the
sequence is based on the preceding one in that we add a vector Δxn, called a step or correction
vector, to xn in order to determine xn+1. Thus
x n +1 = x n + Δ x n (6.2)

Although it might not be the shortest way to the minimum of F(x), we usually require
F(xn+1) < F(xn). A step that meets this condition is called acceptable. Ideally, the procedure
should terminate when no further reduction of the objective function can be obtained. For
practical purposes the iteration stops, for example, if for a prespecified small εj > 0 :

( xn+1 − xn )T .( x n+1 − xn ) < ε 1 (6.3)


146 Chapter 6

F ( x n ) - F ( xn+1 ) < ε 2 (6.4)

[ ΔF ( xn )]T .ΔF ( xn ) < ε3 (6.5)

n = nmax with nmax = a prespecified upper bound for the


(6.6)
number of iterations

Usually a combination of these criteria is used. The various minimization methods are
distinguished by the way they choose their correction vector Δ x. A number of the minimization
methods is discussed in the following subsections.

6.2.1 Gradient methods

Given a point xn in the k-dimensional parameter space, we search for a direction vector δ
that is directed downhill, a direction in which the objective function F(x) decreases. However, we
have to be careful not to proceed too far in this direction since this may lead beyond the
minimum and, consequently, uphill again. On the other hand, a too small step will be inefficient.
In other words, we have to choose an appropriate step size t and a step direction vector δ such
that:

F ( xn + t δ ) < F ( xn ) (6.7)

If δ is a downhill direction vector, a small step in that direction will always decrease the
objective function. Thus we are looking for a vector δ such that F(xn+tδ) is a decreasing function
of t for t sufficiently close to zero. Consequently, for our direction vector δ, it holds:
T
d F ( xn + t δ ) ⎡ ∂ F ( x ) ⎤ ⎡ d ( xn + t δ )⎤
⎥ = [ Δ F ( xn ) ] . δ < 0
T
|t = 0 = ⎢ ⎥ .⎢ (6.8)
dt ⎣ ∂ x ⎦ xn ⎣ dt ⎦t = 0

It can be proven [4] that the direction of the (locally) steepest descent is given by:

δ = - Pn .Δ F ( xn ) (6.9)

where Pn is any positive definite matrix. It is obvious that [∇F( xn )]T. Pn.∇F( xn ) > 0 for all
vectors ∇F(x) ≠ 0. Therefore, [∇F(xn)]T.δ = -[∇F(x)]T.Pn.∇F(xn) < 0 if ∇F(x) ≠ 0. As the
general form of an iteration we obtain from Eq.(6.2):
(6.10)
xn +1 = xn - t n .Pn .Δ F (xn )

where tn is the step size in the n-th iteration.


Computational techniques 147

Clearly, there are many downhill directions on a hypersurface, at least as many as there
are positive definite matrices. Since our limited horizon in most cases does not allow an a priori
optimal choice, many different ways to the minimum are proposed in the literature. In some
algorithms the step size is determined simultaneously with the direction, whereas other methods
specify the direction only and allow various possibilities for the step size selection. These can
range from using the first trial that fulfills Eq.(6.3) to an optimization of the step size in a given
direction. For a description of some possible procedures see [5]. Determination of the optimal
step size in each iteration is likely to decrease the number of steps to the minimum but increases
the computational cost for each iteration. It was found [5,6] that for a number of test problems it
did not pay to use the most expensive procedures for the step size selection and according to
Dennis [7] efforts to improve the gradient methods focus on modifications of the direction rather
than the step size.
Usually the gradient methods are distinguished by the way the step direction is chosen.
Since almost all methods use the basic formula Eq.(6.10) it is the direction matrix Pn that defines
this step direction. In the following subsections we will discuss possible choices, some of which
are summarized below. In this summary the following notation is used:

F(x) represents the objective function to be minimized. This objective function is composed of
the product of the vectors [f(x)]T and f(x).
J(x) represents the Jacobian matrix (matrix of first derivatives) of the vector f(x).
H(x) represents the Hessian matrix (matrix of second derivatives) of the objective function
F(x).

Steepest descent:
Pn = I k (6.11)

where Ik represents the identity matrix.

Newton-Raphson:
-1
⎡∂ 2 F (x ) ⎤
Pn = ⎢ ⎥ = [ H(x)]-1 (6.12)
⎣ ∂x
2

Gauss:

[
Pn = 2. [J ( xn )]T . J ( xn ) ]
-1
(6.13)
148 Chapter 6

Marquardt:

[
Pn = 2. [J ( xn )]T . J ( xn ) + λM Qn ]
-1
(6.14)

where λM is the Marquardt-parameter.

Quadratic hill climbing:


-1
⎡ ∂ 2 F ( x) ⎤
|xn + λM I k ⎥ = [ H ( xn ) + λM I k ]
-1
Pn = ⎢ (6.15)
⎣ ∂x
2

6.2.2 Method of Steepest Descent

It can be shown that the initially steepest descent is obtained if we choose:


P n = Ik (6.16)

in all iterations, where k is the dimension of the parameter space as before. Although this method
is very simple, its use cannot be recommended in most cases, since it may converge very slowly
if the minimum lies in a long and narrow valley. The objective function is then said to be ill-
conditioned. It is clear that using the same direction matrix Pn in each iteration does not allow a
flexible adjustment to different shapes of the objective function surface. However the steepest
descent method can be valuable if combined with other algorithms.

6.2.3 Newton-Raphson Method

The Newton-Raphson method, or simply the Newton algorithm, uses the inverse of the
Hessian matrix to specify the step direction in each iteration, that is:
-1
⎡∂ 2 F ( x ) ⎤
|x n ⎥ = [ H ( x n ) ]
-1
Pn = ⎢ (6.17)
⎣ ∂x
2

In order to see why this direction matrix is chosen, we approximate F(x) at xn by its Taylor series
expansion up to the quadratic terms:

1
F ( x ) ≅ F ( xn ) + Δ F ( xn )T .( x - xn ) + ( x - xn )T .H ( xn ).( x - xn ) (6.18)
2

If this expansion is differentiated with respect to x, we obtain:

Δ F ( x ) ≅ Δ F ( xn ) + H ( xn ).( x - xn ) (6.19)
Computational techniques 149

The conditions for a minimum are ∇F(x)=0, which implies that:

xn+1 = xn - [ H (xn ) ] .Δ F ( xn )
-1
(6.20)

Hence, if F(x) is quadratic, i.e., we have an exact equality in Eq.(6.20), we reach the minimum in
one step of length one. By introducing the step size tn, we obtain a step-limited Newton-Raphson
method. In general, the Hessian may not be positive definite outside a small neighbourhood of
the minimum and thus iterations of the form in Eq.(6.20) may lead to a maximum or a saddle
point. Consequently, a step according to Eq.(6.20) may not be acceptable. Other disadvantages
are that the first and second derivatives must be determined analytically, which places a heavy
burden on the user of this algorithm. Whereas the analytic or numerical determination of the
elements of the gradient is acceptable in most cases, analytic or numerical computation of
higher-order partial derivatives is considered to be a possible source of error. The gradient
methods described in the following subsection can be considered as efforts to overcome the
disadvantage that analytic determination of second order partial derivatives is necessary, without
giving up the advantage of fast local convergence.

6.2.4 Gauss Method

The Gauss method, sometimes called the Gauss-Newton method, is based on the
possibility to approximate the Hessian if the objective function is of a special form: the product
of a vector with its transpose. When a system of non-linear equations is to be solved, the
objective function is nearly always defined as the sum of squares of the function vectors, see
section 6.1:

F ( x) = [ f ( x) ] . f ( x)
T
(6.21)

where again f(x)=[f1(x), f2(x), ..., fk(x)]T. The gradient vector of F(x) is obtained by differentiating
the objective function F(x) with respect to the parameter vector x:

Δ F ( x) = 2 [ J ( x)] . f ( x)
T
(6.22)

where J(x) is the Jacobian matrix of f(x). The Hessian matrix of F(x) is obtained by
differentiating the gradient vector of F(x) with respect to the parameter vector x:

H ( x) = 2 [J ( x )] .J ( x) + 2 R ( x )
T
(6.23)
150 Chapter 6

The matrix R(x) contains the following elements:

⎛ k 2
∂ fj
2
∂ fj ⎞
⎜ k

⎜ ∑ f j ∂ x2 ... ∑ f j
∂ x1 .∂ xk ⎟
⎜ j=1 1 j=1

⎜ ⎟
R ( x) =⎜ ... ... ... ⎟ (6.24)
⎜ ⎟
⎜ k 2 2 ⎟
⎜ ∂ fj k ∂ fj ⎟
⎜⎜ ∑ j ∂ x .∂ x
f ... ∑ fj
∂ xk2 ⎟⎟
j=1 1 k j=1
⎝ ⎠
Since all elements of R(x) are products of the functions fj with a second derivative of fj, all
elements are exactly zero at the minimum. Close to the minimum, the matrix R(x) hardly
contributes to the Hessian matrix and can be neglected. The product of the Jacobian matrix with
its transpose times two is used as an approximation for the Hessian matrix. Consequently,

[
Pn = 2 [J ( xn )]T . J ( xn ) ]
-1
(6.25)

The Gauss method is also applicable to objective functions differing from Eq.(6.21). The
simplicity of this algorithm and its good local convergence properties [7] make it an attractive
method where applicable. However, the direction matrix Pn might not be positive definite if xn is
not close to the solution vector xn*, which minimizes F(x) and, if the elements of the R(x) matrix
are large, the convergence can be slow.

6.2.5 The Marquardt Method

The Marquardt algorithm [3], sometimes referred to as the Marquardt-Levenberg


method, can be used to modify procedures that do not guarantee a positive definite direction
matrix Pn. This algorithm utilizes the fact that

Pn + λM .Qn > 0 (6.26)

if Q is a positive definite matrix and the scalar λM is sufficiently large. A possible choice for Q is
the identity matrix Ik. Typically, this method is used in combination with the Gauss algorithm
and [J(x)]T.J(x) is modified rather than its inverse. This is justified by the fact that the inverse of
a positive definite symmetric matrix is also positive definite. Thus the new direction matrix is
given by:

[
Pn = 2. [J ( x)]T . J ( xn ) + λM Qn ]
-1
(6.27)

where Ik can be used as Qn. For a λM close to zero, this method is equivalent to the Gauss
Computational techniques 151

algorithm, Eq.(6.24). On the other hand, for increasing λM, the steepest descent method is
approached, Eq.(6.16). Since the Gauss algorithm performs very well in the neighbourhood of
the minimum, we start with a small λM and decrease λM > 0 in each iteration unless this results in
an unacceptable step. For the typical step tn always takes the value one. Similarly, we could
modify the Hessian matrix of the objective function F(x) and use:

Pn = [ H ( xn ) + λM I n ]-1 (6.28)

as the direction matrix. This algorithm is usually referred to as the quadratic hill-climbing
method. An improvement may be obtained for both the Marquardt method and the quadratic hill-
climbing method by using a matrix Q that is different from Ik. Marquardt's method appears to
perform very well in practice even if the initial parameter vector x1 is not close to the minimum
of the objective function. Indeed, it has become the standard algorithm for solving a set of non-
linear equations.
Several attempts have been made to economize Marquardt's method. The most obvious
improvement is to keep the Jacobian J(x) fixed for several iterations. This is justified by the fact
that in the vicinity of the minimum, the matrix J(x) changes very little. This strategy, however,
slows down the convergence, which is at best first order if J(x) is kept constant. Another
alternative is to use different approximations to J(x) and update the inverse of J(x) from step to
step; techniques have been provided by [8-10]. A comprehensive description of these schemes
can be found in [11].

6.3 Calculation of the tricritical curve

It was shown in the previous sections that a Marquardt routine was used to solve a set of
nonlinear equations. In this section we will present a step by step description of all numerical and
mathematical actions that are performed to calculate a global phase diagram. It is by no means
our objective to give a detailed discussion on the programs that have been written for our
research. However, we feel forced to give a more detailed description of the way the curves in
the global phase diagram are calculated, because these calculations form the basis of our
research.
We restrict ourselves to an elaboration of the calculation of the tricritical curve in the
global phase diagram. The procedure described below was proposed by Deiters [1]. It has
already been mentioned that in principle the calculation procedure is the same whatever type of
curve is to be calculated in the global phase diagram. In our example ξ and χ are kept constant
and ζ is given a specified value. The value of λ is subsequently calculated.
We start by making a series of calculations of critical curves in p,T-space for binary type
V mixtures. How these calculations are done is extensively described in section 4.7. By changing
the value of k12 in each p,T-diagram, we cause the L1L2V three-phase line in the p,T-plane to
shrink to zero length. The values that are obtained for Vc, Tc, xc and k12, when the three-phase line
152 Chapter 6

disappears, are used as starting values for the program that calculates the tricritical points
directly. In this program, the following steps are made:

1. We start in the main program. The reference component is defined by setting c2 = 1.0,
V2* = 20.0 cm3.mole-1 and T2* = 239.998 K.

2. The values of the two parameter ratios, ξ and χ are assigned the value they had in the p,T-
calculations, e.g., ξ = 0.0 and χ = 0.0.

3. With the values for ξ and χ and the initial value for k12 we calculate c1, V1*, T1*, ζ and λ. The
vector of unknowns, x, contains three elements: λ, xc and Vc. The fourth variable, Tc, is
intrinsically calculated as will be shown below. By using three variables, we also need
three equations: G2x is made zero by definition, as we will explain later.

4. Before the Marquardt-routine is called, the variable vector is transformed. The objective
of this transformation is to avoid that the variables λ, xc and Vc take physically unrealistic
values during the optimization procedure. The transformation functions are:

λ
λt = (6.29)
1 -|λ |

t 2 xc - 1
xc = (6.30)
1 - | 2 xc - 1 |

Vct = ln (Vc ) (6.31)

The variable vector x has the elements xT = (λt, xct, Vct).

5. In the main program the Marquardt-subroutine is called. The Marquardt-subroutine will


evaluate G3x, G4x and G5x at the specified (transformed) vector as is described in steps 6-9.
With these values the Jacobian elements are calculated by numerical differentiation in the
Marquardt-subroutine. The subroutine that contains the equations for G3x, G4x and G5x will
be referred to as the function-subroutine.

6. Initially the back transformation of the variable vector is carried out in the function-
subroutine:

λt
λ= (6.32)
1 +| λ t |
Computational techniques 153

1 + xct (6.33)
xc =
(
2 1 + | xct | )
t
Vc = eVc (6.34)

It is obvious from Eqs.(6.32)-(6.34) that the actual values of λ will always lie in the range
<-1,1>, the value of xc will always be in the range <0,1> and Vc will always be positive.

7. From λ the value of k12 is evaluated.

8. Subsequently Tc is calculated from the condition G2x = 0 in the function-subroutine. We


assume a high value for Tc, e.g., 500 K, and calculate G2x from Vc, Tc, xc and k12. In general
this value will be strongly positive. Then Tc is lowered by ΔT, e.g., 1 K, and G2x is
evaluated again. This is repeated until G2x becomes negative. Within this last interval a
one-dimensional Newton-Raphson search [12] is performed to determine the value of Tc
accurately. It should be noted that although the composition does not change in the search
for Tc, the mixing rule calculation for <cV *Y> has to be performed at each step because
<cV *Y> depends on temperature. Through the calculation of Tc one of the conditions for a
tricritical point, G2x=0, is fullfilled by definition.

9. Next G3x, G4x and G5x are calculated in the function-subroutine. Because these functions
have values that strongly differ in magnitude (G5x can be 103 to 105 times as much as G4x,
which in turn can be 103 to 105 times as much as G3x away from the tricritical point), they
can interfere with the minimization procedure. Thus they need to be scaled without
changing the value of their actual roots. We have chosen a way in which G3x, G4x and G5x
scale themselves on their actual values. All expressions for Gnx are of the following form:

a + b - c + d - e + f + g - .....
G nx = (6.35)
z

where the terms a, b, c, .. etc. represent partial derivatives of the molar Helmholtz energy.

s ( a + b - c + d - e + f + g - ..... ) / z
G nx = (6.36)
( | a | +| b | +| c |+ | d |+ | e |+ | f | + | g |+ ..... ) / | z |

According to Eq.(6.36) the function vector f is calculated: f T = (G3xs, G4xs, G5xs).

10. A return is made to the Marquardt-subroutine. As described in steps 6 to 9, the Marquardt-


subroutine calls the function subroutine a number of times to calculate the elements of the
Jacobian matrix. In section 6.2, it is described how a new parameter vector is determined
from the Jacobian matrix. With this new parameter vector, steps 6 to 9 are repeated until
154 Chapter 6

the norm of the function vector and the norm of the gradient vector are smaller than a
prespecified value, e.g., 2.10-5.

11. If the accuracy criteria are fullfilled a return is made to the main program. A back
transformation is then performed on the transformed solution vector. Subsequently, the
local and global stability tests are performed with the solution vector, see section 4.7.

12. A new value of ζ is chosen. Starting values for λ, xc and Vc are made by linear or quadratic
extrapolation from the previous one or two points of the tricritical curve, that have already
been calculated and the iteration scheme is re-entered at step 4.

This iteration scheme gives good insight into the way tricritical curves in the global
phase diagram are calculated. In the next chapter we will give the results that have been obtained
for global diagrams.

6.4 The use of Maple

6.4.1 Introduction

In the previous sections we have extensively described how we have solved the
necessary equations in order to calculate a global phase diagram. We did not describe the way
how these equations are programmed using the Fortran 77 program language. The listings of the
programs would not only take dozens of pages but they are also not interesting for the reader.
However, we will briefly discuss the use of Maple. The software package Maple used for the
derivation of expressions for Gnx and AiVjx, is a mathematical manipulation language [13,14]. The
type of computation provided by Maple is known by various names such as 'algebraic
manipulation', 'symbolic computation' or 'computer algebra'. A basic feature of such a language
is the ability to, explicitly or implicitly, leave the elements of a computation unevaluated. A
corresponding feature is the ability to perform 'simplification' of expressions involving
unevaluated elements. Maple is designed for the computation one learns in high school algebra
and university mathematics - from fractions, factoring of polynomials, and solution of equations,
through calculus (limits, derivatives, exact solutions to indefinite and definite integrals,
differential equations and power series) to matrix algebra and beyond. Maple can also be used
for numerically solving equations, although we did not use it in this way.
It is beyond the scope of this thesis to describe all the mathematic calculations that can be
performed with Maple. We will restrict ourselves to an extensive description of the procedure we
followed to obtain expressions for Gnx and AiVjx.
Computational techniques 155

6.4.2 Derivation of expressions for Gnx with Maple

In this section we will describe how expressions are obtained for Gnx as a function of
AiVjx with Maple. We have chosen to keep the expressions for Gnx and AixjV separate. By doing
this we can later replace the AixjV-expressions that will be derived from the SPHCT equation of
state by AiVjx-expressions from other equations of state. The expressions for Gnx are independent
of the equation of state that is used. As an example we will derive an expression for G3x. A
Maple session is started on a VAX-2000 workstation by typing maple. Maple will print the
Maple logo and a "prompt" (>) on screen, to show that it is waiting for commands. Next we
define G2x as a function of A2x, AVx and A2V:

> g2x := a2x(v,x) - (avx(v,x))^2/a2v(v,x);

The variables between brackets of the partial derivatives, e.g., v and x in a2x(v,x), indicate that
a2x is some function of v and x, without giving an explicit expression for a2x on these variables.
According to the Jacobian transformation method of Shaw [15] we need to define the following
matrix for the calculation of G3x:

⎛ A2V AVx ⎞
⎜ ⎟
⎜ ⎟
⎜ ∂ G2 x ∂ G2 x ⎟
⎜ ⎟
⎝ ∂V ∂x ⎠

Maple has pre-programmed knowledge of common mathematical operations, like differentiation.


The matrix elements M21 (dg2xdv) and M22 (dg2xdx) are evaluated by Maple through the
following commands:

> dg2xdv := diff(g2x,v);


> dg2xdx := diff(g2x,x);

The matrix, g3xmat, can now be declared in Maple and expressions can be assigned to the
elements:

> g3xmat := array(1..2,1..2);


> g3xmat[1,1] := a2v(v,x);
> g3xmat[1,2] := avx(v,x);
> g3xmat[2,1] := dg2xdv ;
> g3xmat[2,2] := dg2xdx ;

The expression for G3x is now simply calculated by evaluating the determinant of this matrix and
dividing it by A2V. Because we need the determinant-option of Maple, we will load the LINALG-
package into memory:
156 Chapter 6

> with(linalg,det);

The basic Maple program, referred to as the kernel, can do many simple mathematical
calculations involving numbers and formulae. When more complicated calculations are wanted,
like the evaluation of a determinant, Maple can draw on a library of mathematical procedures.
These are kept in files on the disk storage of the workstation. Now Maple is ready to evaluate an
expression for G3x:

> g3x := det(g3xmat)/a2v(v,x);

Maple cannot derive an explicit expression for the partial derivatives, e.g., (∂A2x/∂V)T,x, because
we have not given an explicit expression for A2x as a function of V and x. This expression
depends on the particular equation of state that is used. The expression for G3x contains terms
like diff(a2x(v,x),v). We can replace this particular expression by AV2x. In Maple these
substitutions are done by:

> g3x := subs(diff(a2x(v,x),v)=av2x, g3x) ;


> g3x := subs(diff(avx(v,x),v)=a2vx, g3x) ;
> g3x := subs(diff(a2v(v,x),v)=a3v , g3x) ;

> g3x := subs(diff(a2x(v,x),x)=a3x , g3x) ;


> g3x := subs(diff(avx(v,x),x)=av2x, g3x) ;
> g3x := subs(diff(a2v(v,x),x)=a2vx, g3x) ;

Maple is designed to leave most expressions the way we type them in, or the way they are
created during a computation. There are exceptions to this philosophy, i.e., certain
transformations are automatically performed, e.g., turning 0*x into 0, or turning x*2 and x+x into
2*x. However, Maple's automatic simplification does not always put an expression into simplest
terms, e.g., it does not factor or expand polynomial expressions, nor does it remove greatest
common divisors from rational functions. The Maple function simplify will apply all these
standard simplifications to the required expression. There may be situations where the factorized
result is more clear than the expanded result, such as (x+1)100. Maple can also simplify an
expression while trying to avoid expansions as much as possible, using the function normal. For
G3x both Maple functions give the same result and the final expression for G3x can be simplified
by:

> g3x := simplify(g3x);


Computational techniques 157

The expression for G3x that Maple has derived is:

a3x . a2v 3 - 3 avx . av2x . a2v 2 + 3 avx 2 . a2vx . a2v - avx 3 . a3v
g3x :=
a 2v 3

We can continue this Maple session with the derivation of an expression for G4x from this
expression, in a similar way to how the expression of G3x was derived from G2x. However, we
will skip the description of the derivation of the higher order derivatives and continue with a
description of how the expressions are converted to Fortran 77 commands. The expression that is
given in the above equation is an 'integer expression' and cannot be used directly in a Fortran 77
program. We convert this expression to a 'double precision floating point expression' by:

> Digits := 12;


> precision := double;
> g3x := convert(g3x, float);

Maple can print the resulting expression using syntax of the Fortran 77 programming language.
Before using this command for the first time in a Maple session we must load it from the library.
If the expression is longer than one line, Fortran continuation conventions are used. We will also
load the 'optimize-package' from the library. The optimize-function reduces the number of
arithmetic operations necessary to compute a Fortran 77 expression, through common
subexpressions and binary powering:

> readlib(fortran);
> readlib(optimize);

Both options can be combined to write an optimized Fortran 77 version of the expression for G3x
into the file G3X.FOR:

> fortran(g3x, filename=`G3X.FOR`, optimized);

The expression for G3x can also be stored on disk in the Maple-format for later use:

> save g3x, `g3x.map`;

In a subsequent Maple session, e.g., to derive an expression for G4x, we do not have to type an
expression for G3x as we did here for G2x, but we can read it from disk by:

> read `g3x.map`;


158 Chapter 6

The Maple session can now be closed by:


> quit;

The file G3X.FOR now contains an optimized Fortran 77 expression for G3x. Before compiling
and linking this file to the program that will numerically evaluate G3x, we have to add the Fortran
77 subroutine-header and declare the variables. This is done via a text editor. It seems trivial to
use a powerful symbolic computation program like Maple for the derivation of such a simple
expression as G3x. The numerator of G3x only contains 4 terms. However, the higher order
derivatives of G with respect to x contain a huge number of terms in the numerator, causing the
time needed to derive the equations to increase considerably. The numerator of G8x contains 339
terms! As well as getting rid of a lot of tedious algebra, the chances of making errors during the
derivation or typing in the equations is far lower than if the derivation had been carried out by
hand. The derivation of an expression for G8x would take days, probably weeks, if it would be
done by hand, and again weeks to check it. Moreover, when all Maple commands are stored in a
file, a socalled 'script file', they can easily be checked. All commands to derive equations for Gnx
with n=2..8 were written into a script file. The time spent on the complete derivation of Gnx on a
VAX-2000 workstation is given below:

Time consumed (s):

derivation simplification
G3x 3.50 0.14
G4x 7.02 0.20
G5x 22.01 0.28
G6x 85.65 0.56
G7x 298.57 1.21
G8x 1043.26 4.32

The use of Maple gives us a tremendous time profit. The time for optimizing, 'fortranizing' and
storing the files on disk for the expression for Gnx should be added to the times given above. At
maximum, for G8x, this is 254.15 s. Nevertheless, Maple has turned out to be a valuable tool in
the derivation of these equations.

6.4.3 Derivation of expressions for AiVjx with Maple

Because the first principles of Maple have been explained in the previous section, we can
briefly discuss the means by which expressions are derived for the partial Helmholtz energy
derivatives AiVjx. We start with defining an expression for A(V,T,x) according to the SPHCT
equation, see also Eq.(4.63):
Computational techniques 159

> ap(v,x) := (1-x)*a1 + x*a2;


> ai(v,x) := -r*t*ln(v/v0);
> am(v,x) := r*t*(x*ln(x)+(1-x)*ln(1-x));
> ar(v,x) := tau*vst(x)*(-3*tau*vst(x)+4*v)*c(x)*r*t/v-tau*vst(x))^2;
> aa(v,x) := r*t*zm*c(x)*ln(v*c(x)/(v*c(x)+a(x)));
> a(v,x) := ap(v,x)+ai(v,x)+am(v,x)+ar(v,x)+aa(v,x);

The Maple function a(v,x) now contains the expression for A(V,T,x) according to the SPHCT
equation of state. We foresee that we will have to make similar substitutions to those in the
previous section. Maple offers us the possibility to write procedures ourselves to shorten the
maple session. The following Maple procedure makes a number of substitutions into the
argument f. This procedure is called RW (rewrite):

> # Define procedure REWRITE


> rw := proc(f);
> local f2,f3,f4,f5,f6,f7,f8;
> f2 := f;
> f3 := subs(diff(c(x), x) = dcdx(x), f2);
> f4 := subs(diff(vst(x), x) = dvstdx(x), f3);
> f5 := subs(diff(a(x), x) = dadx(x), f4);
> f6 := subs(diff(dcdx(x), x) = 0, f5);
> f7 := subs(diff(dvstdx(x), x) = 0, f6);
> f8 := subs(diff(dadx(x), x) = d2adx2(x), f7);
> subs(diff(d2adx2(x),x) = 0, f8);
> end;
> # End of the procedure REWRITE

This procedure is called like any other Maple function. An expression for Ax is obtained by the
following Maple commands:

> ap1x(v,x) := rw(normal(diff(ap(v,x),x)));


> ai1x(v,x) := rw(normal(diff(ai(v,x),x)));
> am1x(v,x) := rw(normal(diff(am(v,x),x)));
> ar1x(v,x) := rw(normal(diff(ar(v,x),x)));
> aa1x(v,x) := rw(normal(diff(aa(v,x),x)));

An expression for a1x(v,x) can now be obtained by:

> a1x(v,x) := ap1x(v,x)+ai1x(v,x)+am1x(v,x)+ar1x(v,x)+aa1x(v,x);


160 Chapter 6

By separately differentiating the various terms we avoid the possibility of Maple mixing up the
different terms. This would lead to an expression for a1x(v,x) that is difficult to survey. Next
order derivatives are derived from the expressions above, e.g., for AVx:

> ap1v1x(v,x) := rw(normal(diff(ap1x(v,x),v)));


> ai1v1x(v,x) := rw(normal(diff(ai1x(v,x),v)));
> am1v1x(v,x) := rw(normal(diff(am1x(v,x),v)));
> ar1v1x(v,x) := rw(normal(diff(ar1x(v,x),v)));
> aa1v1x(v,x) := rw(normal(diff(aa1x(v,x),v)));
> a1v1x(v,x) := ap1v1x(v,x)+ai1v1x(v,x)+am1v1x(v,x)+ar1v1x(v,x)+aa1v1x(v,x);

In this manner expressions up to the eight order derivatives of A(V,T,x) are easily obtained.
Before the expressions are printed in Fortran 77 language, final substitutions have to be made,
i.e, the dependencies of c on x, vst on x, etc., should be removed from the final expressions for
AiVjx. This is done by the procedure RW1:

> # Define procedure REWRITE1


> rw1 := proc(f);
> local f2,f3,f4,f5,f6,f7,f8;
> f2 := f;
> f3 := subs(c(x) = c , f2);
> f4 := subs(vst(x) = vst , f3);
> f5 := subs(a(x) = a , f4);
> f6 := subs(dcdx(x) = dcdx , f5);
> f7 := subs(dvstdx(x)= dvstdx , f6);
> f8 := subs(dadx(x) = dadx , f7);
> subs(d2adx2(x) = d2adx2, f8);
> end;
> # End of the procedure REWRITE1

Maple's function factor will perform a complete factorization of polynomials in one or more
variables with respect to the coefficient field of rational numbers. Both functions are used for the
final simplification of AiVjx:
> ap1x(v,x) := factor(rw1(ap1x(v,x)));
> ai1x(v,x) := factor(rw1(ai1x(v,x)));
> am1x(v,x) := factor(rw1(am1x(v,x)));
> ar1x(v,x) := factor(rw1(ar1x(v,x)));
> aa1x(v,x) := factor(rw1(aa1x(v,x)));

The final expression for Ax is obtained by:


> a1x := ap1x(v,x)+ai1x(v,x)+am1x(v,x)+ar1x(v,x)+aa1x(v,x);
Computational techniques 161

All other derivatives of A(V,T,x) are obtained in the same way. The final operations to 'fortranize'
these expressions are similar to the ones discussed in section 6.4.2:

> precision := double;


> Digits := 12;
> a1x := convert(a1x, float);
> fortran(a1x, filename=`A1X.FOR`, optimized);

The Maple session is again closed by:


> quit;

The file A1X.FOR now contains an optimized Fortran 77 expression for Ax. Before compiling
and linking this file to the Fortran program that numerically evaluates Ax, we have to add the
Fortran 77 subroutine-header and declare the variables. This is again done via a text editor. If all
Maple-commands are written in a script-file the derivation of expressions for AiVjx (2 ≤ i+j ≤ 8)
only takes a number of minutes. By only changing the definitions of Arep and Aattr (ar(v,x) and
aa(v,x) in the Maple commands above), we can use the same script-file for the derivation of AiVjx
from other equations of state.

NOMENCLATURE

AiVjx partial derivative of the molar R matrix with residuals


Helmholtz energy with respect to V (i t step length
times) and x (j times) Tc critical pressure
ci one-third of the external degrees of Ti* characteristic temperature of compo-
freedom of component i, also called nent i
the flexibility parameter Vc critical volume
fi mathematical function to be minimi- Vi* covolume of component i
zed x parameter vector
f k-dimensional function vector xc critical composition
F objective function
Gnx partial derivative of the molar Gibbs ∇ Nabla operator, denoting a gradient
energy with respect to x (n times) vector
H Hessian matrix
I identity matrix Greek symbols
J Jacobian matrix
k dimensionality δ direction vector
k12 binary interaction coefficient ε accuracy limit
pc critical pressure ζ dimensionless parameter ratio
P semi-positive definite matrix λ dimensionless parameter ratio
Q positive definite matrix λM Marquardt-parameter
162 Chapter 6

ξ dimensionless parameter ratio


χ dimensionless parameter ratio

Superscripts and subscripts

s scaled
t transformed
T transpose
REFERENCES

1. U.K. Deiters. Private communications, Ruhr-Universität Bochum, Germany, (1990-


1991).
2. I. Mills, T. Cvitas, N. Kallay, K. Homann and K. Kuchitsu. International Union of Pure
and Applied Chemistry: Quantities, Units and Symbols in Physical Chemistry. Blackwell
Scientific Publications, Oxford, (1988).
3. D.W. Marquardt. An algorithm for least-squares estimation of nonlinear parameters. J.
Soc. Indust. Appl. Math. 11(2), (1963), 431-441.
4. S.L.S. Jacoby, J.S. Kowalik and J.T. Pizzo. Iterative methods for nonlinear optimization
problems. Prentice Hall, Inc., Englewood Cliffs, New Jersey, (1972).
5. P.D. Flanagan, P.A. Vitale and J. Mendelsohn. A numerical investigation of several one-
dimensional search procedures in nonlinear regression problems. Technometrics 11,
(1969), 265-284.
6. Y. Bard. Comparison of gradient methods for the solution of nonlinear parameter
estimation problems. SIAM J. Numer. Anal. 7, (1970), 157-186.
7. J.E. Dennis, Jr. Some computational techniques for the nonlinear least squares problem,
in G.D. Bryne and C.A. Hall, eds., Numerical solutions of systems of nonlinear algebraic
equations, Academic, New York, (1973), 157-183.
8. J.G.P. Barnes. An algorithm for solving non-linear equations based on the secant method.
Comput. J. 8, (1965), 66-72.
9. C.G. Broyden. A class of methods for solving nonlinear simultaneous equations. Math.
Comp. 19, (1965), 577-593.
10. M.J.D. Powell. A method for minimizing a sum of squares of non-linear functions
without calculating derivatives. Comput. J. 7, (1965), 303-307.
11. C.G. Broyden. Quasi-Newton methods and their application to function minimization.
Math. Comp. 21, (1967), 368-381.
12. W.H. Press, B.P. Flannery, S.A. Teukolsky, and W.T. Vetterling. Numerical recipes.
Cambridge University Press, New York, (1986).
13. B.W. Char, K.O. Geddes, G.H. Gonnet, M.B. Monagan and S.M. Watt. Maple reference
manual. 5th Ed. Watcom publications limited. Waterloo, Ontario, Canada, (1988).
14. B.W. Char, K.O. Geddes, G.H. Gonnet, M.B. Monagan and S.M. Watt. Maple, first
leaves: a tutorial introduction to Maple. Watcom publications limited. Waterloo, Ontario,
Canada, (1988).
15. Shaw. The derivation of thermodynamical relations for a simple system. Phil. Trans. Roy.
Soc. London A 234, (1935), 299-328.
163

CHAPTER 7
THE GLOBAL PHASE DIAGRAM FROM THE SPHCT EQUATION

In this chapter we discuss the results of our global phase diagram calculations. Section
7.1 deals with the global phase diagram of the SPHCT equation for molecules with the same size
and the same flexibility. In section 7.2, the global phase diagrams are discussed for molecules
with small size differences and equal flexibilities. The global phase diagrams for molecules with
large size differences and equal flexibilities are discussed in section 7.3. In section 7.4 we will
discuss the global phase diagrams for molecules that have the same size but non-equal
flexibilities and finally, in section 7.5 we will discuss the global phase diagrams for molecules
with non-equal sizes and non-equal flexibilities.

7.1 The global phase behaviour of equal-sized and equally flexible molecules

In this chapter the results of the global phase diagram calculations with the SPHCT
equation of state, are discussed. As already mentioned in the previous chapter, one of the
components needs to be defined, in order that the global phase diagram calculations can be
performed. For the readers' convenience, we recall the values of the (fictitious) pure component
parameters of the reference component (i.e., component 2):

c2 = 1.0000
V2* = 20.000 cm3.mole-1
T2* = 239.998 K

The values are chosen so that as far as their critical properties are concerned (Tc,2 = 300.84 K,
pc,2 = 68.41 bar and Vc, 2 = 127.25 cm3.mole-1), they are physically realistic. A molecule of the
reference component can be seen as a n-alkane molecule with 3.0 CH2-segments. Because we
choose c2 = 1, the molecule is not flexible but very stiff. This causes the critical pressure to be
higher from what one might expect for a n-alkane. In one of the following sections the influence
of the flexibility on the global phase behaviour is studied. For mixtures composed of molecules
of equal size (ξ = 0.0) and of equal flexibility (χ = 0.0), the global phase diagram in ζ,λ-space is
symmetric with respect to the λ-axis. The global phase diagram calculated from the SPHCT
equation of state for ξ = χ = 0, is shown in Fig.7.1. This figure shows three tricritical curves, that
have also been found with the Van der Waals equation (VDW) [1,2], the Redlich-Kwong
equation (RK) [3] and the Carnahan-Starling-Redlich-Kwong equation (CSRK) [4].
164 Chapter 7

Figure 7.1: A global phase diagram of the SPHCT equation for molecules with equal
sizes, ξ=0.0, and equal flexibilities, χ=0.0. The solid curves represent the
stable tricritical curves, the dotted curves the metastable parts of the tricritical
curves, the dash-dotted curves the azeotropic boundaries and the thin solid
curves the azeotropic critical endpoint curves. The open boxes on the curved
tricritical lines represent the Van Laar points and the open circle represents
the termination point of the symmetric tricritical line.
The global phase diagram of the SPHCT equation 165

Two of the three tricritical lines are curved and, as mentioned above, symmetric with
respect to the λ-axis. They enclose the region of type I and type II phase behaviour. The third
tricritical curve merges with the λ-axis from λ = 1 to λ = 0.24417. The tricritical curves in Fig.7.1
have been calculated very accurately: for the drawing of the curved tricritical lines we calculated
at least 200 tricritical points per curve and for the vertical tricritical curve we calculated
approximately 100 tricritical points. To draw the tricritical curves we calculated in total at least
500 tricritical points per global phase diagram. The same care has been taken for the drawing of
the tricritical curves in the global phase diagrams where ξ ≠ 0. The curved tricritical lines end
towards lower values of λ at about (ζ, λ) = (-1, -3.4) and (ζ, λ) =(1, -3.4). The tricritical mole
fractions approach the values xc → 1 and xc → 0 respectively. The exact value of λ at ζ = -1 or
ζ = 1 is hard to evaluate numerically because the tricritical mole fractions become very close to
either 0 or 1, e.g., at λ = -1.5, the tricritical mole fractions are 0.007 and 0.993 respectively.
Towards positive λ-values the curved tricritical lines intersect and finally become metastable.
The curved tricritical lines intersect twice. The upper intersection point at (ζ, λ) = (0.0, 0.33347)
has also been found for the VDW-, RK- and CSRK-equation of state [1-4]. This intersection
point is not a real intersection point but a consequence from the projection onto a
two-dimensional plane, i.e., the tricritical properties, pc, Tc, Vc and xc, are different at the various
tricritical curves. Below the termination point of the vertical tricritical curve the tricritical
pressure becomes negative. Coming from high λ-values, the vertical tricritical curve must
become metastable just above the termination point. Because the exact point where the vertical
tricritical curve becomes metastable is very hard to calculate, we used the point where the
vertical tricritical curve starts to have a negative pressure, as termination point of this tricritical
curve. In reality, the point where the metastability of this tricritical curve begins is so close to the
point where the tricritical pressure is negative, that they would be indistinguishable in Fig.7.1.
To demonstrate that the intersection
point of the three tricritical curves at (ζ, λ) =
(0.0, 0.33347) is a consequence of the
projection we calculated the tricritical mole
fractions along the tricritical curves. This is
illustrated in Fig.7.2., in which λ is shown as a
function of the tricritical mole fraction along the
tricritical curves (solid lines) in the global phase
diagram with ξ = χ = 0.0. The dotted line in
Fig.7.2. represents the intersection point of the
three tricritical curves in the ζ, λ-projection. A
binary mixture that is composed of molecules
with the pure component parameters and the
binary interaction parameter, that correspond
with the intersection point in the global phase
diagram (ζ, λ) = (0.0, 0.33347) does not have a Figure 7.2: The value of λ as a function
critical state of a higher order than 3, but three of the mole fraction along the
tricritical curves at ξ=0.0.
166 Chapter 7

tricritical points. The shield region is situated around this intersection point. The shield region
that has been discovered by Furman et al. [5,6], contains five types of phase behaviour. These
five types of phase behaviour all show four-phase equilibria. Because practical mixtures that
correspond to the shield region are not very likely to be found [7], we will not pay any attention
to the shield region and the corresponding types of phase behaviour. For a detailed overview one
is referred to [3,5-11].
The curved tricritical lines have been calculated to values of λ as low as λ = -1.5. The
lower intersection point was found at (ζ, λ) = (0.0, -1.15775). For the CSRK equation this
intersection point occurred at (ζ, λ) = (0.0, -0.9) [12]. Both the upper and lower intersection point
of the curved tricritical lines occur at lower λ values than calculated from the CSRK-equation,
λUSPHCT = 0.33347, λLSPHCT = -1.15775 and λUCSRK = 0.29, λLCSRK = -0.9. In the vertical direction
the enclosed type I and type II region is larger for the SPHCT equation than it is for the CSRK
equation. This also holds for the horizontal stretching of the enclosed region [11]:

Equation ζ of the intersection


point at λ=0
VDW 0.56
RK 0.40
SPHCT 0.40
CSRK 0.30

We can conclude that the type I and type II region is larger than the corresponding region of the
CSRK equation. Because the type I/II region has the shape of a kite, we will refer to this region
as the "kite-shaped" region in the following sections.
The curve of the double critical endpoints in the global phase diagram could only be
calculated with the SPHCT equation with extreme numerical effort: the starting values of the
double critical endpoint, e.g., λ, xc and Vc, sometimes had to be within 0.01% of the final
solution. The final solution was then used for the calculation of the next point of the double
critical endpoint curve. In almost every case, the program then crashed because the values of the
parameters at the double critical endpoint were not accurate enough to be used as starting values
for the calculation of the next double critical endpoint. We believe that the extreme effort needed
to calculate the double critical endpoint curve in the global phase diagram is not worthwhile. The
importance of the double critical endpoint curve is not great as far as the topology of the critical
curves of a binary mixture in p, V, T, x-space is concerned, i.e., the double critical endpoint is
nothing more than a kind of global stability test, see section 5.2.2. We performed calculations in
p,T-space that confirmed the location of the double critical endpoint curve in the global phase
diagram as reported by other investigators [1-4,11]. The results of these calculations are shown
in the next chapter. Suffice to say, the double critical endpoint curve is almost tangent to and
extremely close to the upper parts of the curved tricritical lines in the global phase diagram. The
double critical endpoint curve is so close to the upper parts of the curved tricritical lines in the
global phase diagram that both lines are almost indistinguishable.
The global phase diagram of the SPHCT equation 167

Calculations show that for some equations the tricritical curve and the double
critical endpoint curve can intersect in the global phase diagrams. For the VDW- and the
RK-equation the double critical endpoint curve and the tricritical curve intersect at a point, the
so-called "Van Laar point" [13], thus forming the regions for type IV and IV* phase behaviour
[2,3]. The Van Laar point is a point of symmetrical tricritical behaviour [3], though there is no
obvious symmetrical relationship between the interaction parameters. For the VDW-equation the
Van Laar point lies on the geometric mean curve [2,6]. The geometrical mean curve in the global
phase diagram is defined as the curve where k12 = 0. Meijer [14,15] has used this simplification
to study the critical behaviour around the Van Laar point analytically. According to Kraska and
Deiters [4,12], the double critical endpoint curve of the CSRK equation becomes metastable
above λ ≈ 0.04, i.e., before the Van Laar point is reached; hence the Van Laar point and
consequently type IV* phase behaviour cannot be calculated with the CSRK equation. For the
three-component lattice gas [5,6] the double critical endpoint curve and the tricritical curve
merge and consequently the classes IV and IV* disappear. For the SPHCT equation the Van Laar
points have been found for equal-sized and equally flexible molecules at the following
coordinates: (ζ, λ) = (-0.19760, 0.19131) and (ζ, λ) = (0.19760, 0.19131). The characteristic
temperature of component 1 is T1* = 358.205 K. The coordinates of the right Van Laar point in
p,V,T,x-space are: pc = 93.63 bar
Tc = 338.78 K
Vc = 84.27 cm3.mole-1
xc = 0.5307
Extensive research of the neighbourhood of the Van Laar point has been done for the SPHCT
equation. The results are given in section 8.4.
The mathematical double point curve of the first kind lies between the upper parts of the
curved tricritical lines and the double critical endpoint curves in the global phase diagram. This
mathematical double point curve also goes through the Van Laar point.
As was shown in section 5.2.4, the boundary lines for azeotropy at ξ = χ = 0.0 lie at the
diagonals of the ζ, λ-global phase diagram. Above both azeotropic boundary lines in the global
phase diagram we find binary mixtures that exhibit maximum-pressure azeotropy. Below both
azeotropic boundary lines in the global phase diagram, we will find binary mixtures that exhibit
minimum-pressure azeotropy. To the left and to the right of both azeotropic boundary lines, the
binary mixtures that correspond to these parts of the global phase diagram do not exhibit
azeotropy. The region where double azeotropy occurs is situated closely to the intersection point
of the azeotropic boundary lines. This region is extensively investigated for the VDW equation
by Van Konynenburg [1].
The boundary curve between type III-A and III-H, that was calculated for the first time
by Mazur et al. [9], has qualitatively the same shape for the SPHCT equation as it has for the
VDW-, RK- and CSRK-equation [3,4,11]. The right-hand azeotropic critical endpoint curve in
the global phase diagram approaches the azeotropic boundary line asymptotically towards
(ζ,λ) = (1,1) where xc → 1. Already at ζ = 0.50000, the λ-value of the azeotropic critical endpoint
curve is: λ = 0.52003 and the critical mole fraction xc = 0.99998. Towards the left the azeotropic
168 Chapter 7

critical endpoint curve approaches one of the curved tricritical lines in the global phase diagram.
According to Mazur et al. [9] the azeotropic critical endpoint curve ends to the left in the shield
region.
The geometric mean curve is part of a
circle going through the following three points
in the global phase diagram: (ζ, λ)=(-1, 1),
(ζ, λ) = (0, 0) and (ζ, λ) = (1, 1). This is readily
seen from Eq.(5.5). Binary mixtures that
correspond to a point on this curve have a
binary interaction coefficient k12 = 0.0. Ordinary
binary mixtures have binary interaction
coefficients between k12 = -0.10 and k12 = 0.20.
The region where the ordinary binary mixtures
can be found is shown in Fig.7.3 between the
dash-dotted lines. To get an impression of the
location of the kite-shaped type I/II region with
respect to the region where the pratical mixtures
are to be found (-0.10 ≤ k12 ≤ 0.20), we have
Figure 7.3: Three lines of constant values
also drawn the tricritical curves for mixtures of
of k12. The tricritical curves in
equally sized molecules in Fig.7.3. The case ξ=0.0 have also been
geometrical mean curve does not depend on ξ drawn to show the location of
nor on the value of χ, and is situated at exactly the type I/II region.
the same place in all of the following global phase diagrams.
Sofar, there are no qualitative differences found between the global phase diagram
calculated with the SPHCT equation, and the ones calculated with the VDW- and RK-equation.
The global phase diagram calculated with the CSRK-equation, slightly differs from the ones that
are calculated with the VDW, RK and SPHCT equation, because there is no Van Laar point and
thus type IV* cannot be found. At this point we want to introduce the work of Boshkov [16]. He
was the first researcher who was able to calculate type VI phase behaviour with a
pressure-explicit equation of state, i.e., a mixture with a liquid-liquid immiscibility loop.
Boshkov [16] was able to locate type VI in the global phase diagram and suggested the scheme
as shown in Fig.7.4. The region of type VI and VII is that small, that it is not visible in Fig.7.1. It
is located in the global phase diagram, where the tricritical curve is about to intersect the ζ-axis.
In Fig.7.4 the solid curve represents the tricritical curve and the dashed curve represents the
double critical endpoint curve. Although the type VI/VII region is very small, like the shield
region, it is of more practical importance, because type VI has been determined experimentally,
see section 2.7. As mentioned previously, we cannot calculate the double critical endpoint curve
so we were not able to confirm the scheme of Boshkov [16]. However, it turned out that the
SPHCT can also be used to calculate liquid-liquid immiscibility loops and we have performed
many calculations in the type VI/VII region. Our findings are in agreement with the scheme that
Boshkov presented. We will discuss the type VI/VII region in section 8.1.
The global phase diagram of the SPHCT equation 169

Figure 7.4: The neighbourhood of type VI phase behaviour in the global phase diagram.

7.2 The global phase behaviour of molecules with equal flexibilities but slightly
different sizes

In this section we will discuss the results of the global phase diagram calculations for
similar-sized molecules (0.0 < ξ ≤ 0.3). The corresponding global phase diagrams are shown in
Figs.7.5, 7.6 and 7.7.
170 Chapter 7

Figure 7.5: A global phase diagram of the Figure 7.6: A global phase diagram of the
SPHCT equation of state for SPHCT equation for ξ = 0.2
ξ = 0.1 and χ = 0.0. For an and χ = 0.0. For an explan-
explanation of symbols, see ation of symbols, see Fig.7.1.
Fig.7.1.

Although some results for molecules with large differences in size (0.4 ≤ ξ ≤ 1) have been taken
into account here, a more detailed discussion of the global phase diagrams for these molecules
will be given in the next section. The choice between molecules with slightly different sizes and
molecules with large size differences is quite arbitrarily and only serves to order the various
global phase diagrams. If the value of ξ is changed then the unlike molecules of a binary mixture
no longer have the same size. In case of ξ = 0.0, the reference component and the variable
component are composed of 3.0 CH2-segments. A CH2-segment has a molar volume of 6.73
cm3.mole-1. If ξ = 0.1, the molecules of the variable component are composed of 3.6
CH2-segments, if ξ = 0.2, the molecules of component 1 have 4.5 CH2-segments and finally if
ξ = 0.3 then the molecules of component 1 each contain 5.5 CH2-segments. Of course the
reference component has 3.0 CH2-segments at every value of ξ. Because the c-parameter has a
value of 1 for both components, the molecules cannot be seen as (pseudo-)n-alkanes, but should
be considered as stiff rods. In section 7.4 we will discuss the influence of the flexibility ratio χ on
the global phase behaviour of binary mixtures.
Global phase diagrams where ξ ≠ 0 are not symmetric with respect to the λ-axis. The
tricritical curves shift to the left of the global phase diagram if the value of ξ is increased.
Moreover, the tricritical curves also turn slightly clockwise around the origin. This is shown in
Fig.7.8. The data for this figure were taken from table 7.1. In this figure the open circles
represent the intersection points of the curved tricritical lines as a function of ξ. The solid circles
are extrapolated intersection points.
The global phase diagram of the SPHCT equation 171

Figure 7.7: A global phase diagram of the Figure 7.8: The movement of the kite-
SPHCT equation for ξ = 0.3 shaped type I/II region as a
and χ = 0.0. For an explana- function of ξ. For an explana-
tion of symbols, see Fig.7.1. tion of symbols, see text.

These extrapolations are based on a correlation with the intersection points of the azeotropical
boundary curves. The solid lines between two circles only serve to show the intersection points
that belong to the same value of ξ. The corresponding value of ξ is written near the lower
intersection points. From the extrapolation of the dotted lines, it seems highly probable that at
ξ = 1, both intersection points will coincide at (ζ, λ) = (-1, 1). It is shown in Fig.7.8 that the
kite-shaped region in which types I and II are situated, not only shifts to the left of the global
phase diagram and turns, but also that this enclosed region becomes smaller at higher values of ξ.
The extrapolated curves through the lower and upper intersection point in Fig.7.8, i.e., the dashed
curves, show that in the limiting case of ξ = 1, there will be no domain of type I and II left. Both
dashed curves will intersect at (ζ, λ) = (-1, 1), thus there are no binary mixtures to be found at
ξ = 1 where both components are fully miscible in the liquid state. If the size difference between
the unlike molecules is infinite, only type III or type Vh phase behaviour will occur in a binary
mixture. Whether type III or type Vh will occur depends on the ratio of the ε-values of both
components and the value of the binary interaction coefficient. The coordinates of the
intersection points are given in Table 7.1. Phase equilibrium calculations of molecules with
extreme size differences have been performed by Jackson et al. [17]. The coordinates of the
termination point of the "vertical" tricritical curve as a function of ξ at χ = 0.0 are given in Table
7.2.
172 Chapter 7

Table 7.1: This table gives the coordinates of the intersection points of the curved tricritical
lines as a function of ξ. LIP is an abbreviation for lower intersection point and
UIP means upper intersection point. The point marked with an * are extrapolated
points.

ξ LIP ζ LIP λ UIP ζ UIP λ

0.0 0.00000 -1.15775 0.00000 0.33347

0.1 -0.16522 -1.11509 -0.12790 0.33908

0.2 -0.31411 -1.00537 -0.24351 0.35302

0.3 -0.44657 -0.85005 -0.34911 0.37270

0.4 -0.56322 -0.66377 -0.44642 0.39717

0.5 -0.66509 -0.45604 -0.53696 0.42669

0.6 -0.75347 -0.23176 -0.62195 0.46294

0.7* -0.82928* 0.01018* -0.70797* 0.50821*

0.8* -0.89398* 0.26344* -0.79604* 0.57561*

p,T-Calculations have shown that the


Table 7.2: The coordinates of the
double critical endpoint curves are close to the
termination point of the
curved tricritical lines if the value of ξ is
vertical tricritical curve as a
increased. The problems that we had in the
function of ξ at χ = 0.0.
numerical evaluation of the double critical
endpoint curve also occured for the
ξ ζ λ
CSRK-equation at non-zero values of ξ [12].
Kraska [12] showed, by means of 0.0 0.00000 0.24417
p,T-calculations, that the double critical
0.1 -0.12708 0.25068
endpoint curves are close to the corresponding
curved tricritical lines for the CSRK equation 0.2 -0.22943 0.26662
as well. Both Van Laar points shift towards the 0.3 -0.32828 0.28932
left in the global phase diagram if ξ is
increased. The coordinates in the global phase 0.4 -0.41904 0.31790
diagram of the Van Laar points are given in 0.5 -0.50310 0.35318
Table 7.3. The Van Laar points are shown in
0.6 -0.58168 0.39702
Figs. 7.5, 7.6 and 7.7 as open boxes.
The global phase diagram of the SPHCT equation 173

Table 7.3: Global phase diagram coordinates of the Van Laar point of the SPHCT equation as
a function of ξ at χ = 0.0.

Van Laar point LEFT Van Laar point RIGHT

ξ ζ λ ζ λ

0.0 -0.19760 0.19131 0.19760 0.19131

0.1 -0.42948 0.13809 -0.45770 0.59473

0.2 -0.57316 0.15264 - -

0.3 -0.68326 0.19432 - -

In Fig 7.5 it is shown that the asymptotical limits for the azeotropic critical endpoint
curve at ξ = 0.1 are the same as in the case of ξ = 0.0: the azeotropic critical endpoint curve at the
righthand side of the global phase diagram approaches the azeotropical boundary curve at high
values of ζ and it approaches the right tricritical curve at low values of ζ. The azeotropic critical
endpoint curve at the lefthand side of the global phase diagram at ξ = 0.1 and χ = 0.0 behaves in a
similar way, see Fig.7.5.
The azeotropic boundary lines start at (ζ, λ) = (-1, 1) and (ζ, λ) = (1, 1) and go towards
lower values of λ for all values of ξ considered in this work. Although no analytical expression
could be derived for the azeotropical boundary curves, they appear to be straight lines. The
intersection point of the azeotropic boundary lines shifts, just as the intersection points of the
curved tricritical lines do, towards the left and upwards in the global phase diagram at higher
values of ξ. This is shown in Fig.7.9. In this figure the open triangles represent the points where
both azeotropic boundary lines intersect as a function of ξ. The dashed curves represent the
points where the curved tricritical lines intersect as a function of ξ. The extrapolation of the curve
going through the intersection points of the azeotropic boundary curves, i.e., the solid curve,
shows that at ξ = 1, the intersection of the azeotropic boundary curves occurs at (ζ, λ) = (-1, 1).
For ξ = 1, the azeotropic boundary curves are: λ = 1 and ζ = -1. In this case there are no mixtures
to be found with a phase behaviour that shows azeotropy. The data that have been used in Fig.7.9
have been taken from Table 7.4.
Even at a relatively small size ratio, ξ = 0.1, qualitative departures are observed in the
phase diagrams of the types I and V. In the phase diagrams of these types an additional critical
curve has been observed at extremely high pressures. This critical curve is a boundary of a
high-pressure liquid-liquid immiscibility, that extends over a wide temperature range, often
beyond the critical temperature of the less-volatile component. Phase diagram types with this
high-pressure immiscibility have been marked with a superscript "h", in accordance with the
nomenclature introduced by Kraska and Deiters [4]. The high-pressure critical curve is discussed
in section 5.2.4. This phenomenon was also observed for the RK- and the CSRK-equation, see
[4].
174 Chapter 7

Table 7.4: The coordinates of the intersec-


tion point of both azeotropical
boundary curves as a function
of ξ.

ξ ζ λ

0.0 0.00000 0.00000

0.1 -0.16678 0.01452

0.2 -0.31686 0.05164

0.3 -0.44981 0.10425

0.4 -0.56602 0.16775

0.5 -0.66651 0.23958

0.6 -0.75273 0.31922

0.7 -0.82642 0.40889 Figure 7.9: The course of the intersection


point of the azeotropic boun-
0.8 -0.88972 0.51632
dary curves as a function of ξ.

There is one final point we want to make about the global phase diagrams of
similar-sized molecules. The upper intersection point of the three tricritical curves in the
ζ,λ-global phase diagram is an exact intersection point. However, at higher values of ξ this
intersection point grows into an "intersection triangle", that enlarges with increasing values of ξ.
According to Kraska [12], who was the first to discover this phenomenon, the intersection
triangle can only be due to numerical inaccuracies. For reasons of symmetry there can only exist
an intersection point instead of an intersection triangle in the shield-region. Kraska [12] proposed
this triangle to be caused by numerical errors because of the extremely steep slopes of the
derivatives of the molar Gibbs energy with respect to the composition. Although we scaled our
equations for the derivatives of G, see section 6.4, and used analytical expressions for the
derivatives Gnx (n = 2,..,5), we also found the intersection triangle inside the shield region with
the SPHCT equation. In our investigations the intersection triangle turned out to be much
smaller. The height and width of the intersection triangle are shown in Table 7.5 and in Figs.7.10
and 7.11. On the one hand it may seem strange that if the intersection triangle is caused by
numerical inaccuracies, the height and width of the intersection triangle are smooth functions of
ξ. On the other hand, if the appearance of the intersection triangle is assumed to be caused by
numerical inaccuracies, we may explain the fact that we find smaller intersection triangles by the
use of analytical expressions for the derivatives Gnx instead of numerical expression, like Kraska
[12] did. In that case, the size of the intersection triangle gives us an idea of the accuracy of the
calculated tricritical curves. Of course, even if analytical expressions for the derivatives Gnx are
The global phase diagram of the SPHCT equation 175

used, round-off errors will occur. Because the intersection triangle is irrelevant to practical
mixtures we do not attempt to go into detail on it and no further efforts has been made to solve
this problem.

Table 7.5: The width Δ ζ and the height Δ λ of the intersection triangle in the shield region.

SPHCT CSRK [12]


ξ Δζ Δλ Δζ Δλ
0.0 0.00000 0.00000 - -
0.1 0.00038 0.00052 - -
0.2 0.00078 0.00101 0.003 0.004
0.3 0.00118 0.00154 0.006 0.01
0.4 0.00165 0.00214 0.013 0.02
0.5 0.00212 0.00278 - -
0.6 0.00275 0.00365 - -

We can conclude that all curves more or less follow the shifts of the curved tricritical
lines in the global phase diagram as a function of ξ. A good impression of the global phase
behaviour can thus be obtained by studying the shifts of the tricritical curves in the global phase
diagram. The shifting and turning of the tricritical curves in the global phase diagram at
increasing values of ξ has also been found for the VDW-, RK- and the CSRK-equation.

Figure 7.10: Schematic drawing of the intersec- Figure 7.11: The height and width of the
tion triangle, formed by the tricriti- intersection triangle as a
cal curves in the shield-region. function of ξ.
176 Chapter 7

For the VDW-equation the shift to the left has been observed previously [1], but the
turning of the kite-shaped region in the global phase diagram was not recognized by Van
Konynenburg [1]. Bolz [11] performed many global phase diagram calculations with the VDW
equation and observed the clockwise-turning of the kite-shaped region. The domains of the
different phase diagram types for mixtures composed of molecules with slightly different sizes,
are arranged in the same way as was the case for ξ = 0.0. The most important conclusion we can
draw is that the kite-shaped region that encloses the type I and II phase behaviour regions
becomes smaller at increasing values of ξ, and consequently the region where type III and V
occur, i.e., the regions that contain phase diagrams with an immiscibility area, become larger.

7.3 The global phase behaviour of molecules with equal flexibilities but large size
differences

In this section we will discuss the results of the global phase diagram calculations for
ξ > 0.3 and χ = 0.0. To obtain an idea about the number of CH2-segments of the molecules of the
variable component at these size ratios, their values are given below:

ξ 0.4 0.5 0.6

comp. 1. 7.0 9.0 12.0


comp. 2. 3.0 3.0 3.0

If the value for ξ is increased from ξ = 0 to 0.3, the angle in the left curved tricritical line be-
comes sharper, see Figs.7.5-7.7. Eventually a tetracritical point appears on the curved tricritical
line on the lefthand side of the global phase diagram. The coordinates of the tetracritical point in
the global phase diagram were determined by solving G2x = G3x = G4x = G5x = G6x = G7x = 0. For
χ = 0.0 the following coordinates were found:

ξ = 0.31390
ζ = -0.69626
λ = 0.20262
χ = 0.00000

The resulting residual value of the objective function F, see section 6.3, is F=3.1 10-8. The
coordinates of the tetracritical point in p,V,T,x-space are:
p = 26.08 bar
T = 150.24 K
x = 0.28698
V = 92.36 cm3.mole-1
The global phase diagram of the SPHCT equation 177

The p,T-projection of the critical curves of the mixture with the tetracritical point mentioned
above, is shown in Fig.7.12. At the tetracritical point G8x = 1.43 1010 J/mole, i.e., the tetracritical
point is locally stable. According to the global stability test, see section 4.7, the tetracritical point
is also globally stable. Although the residual value of the objective function is very low, care has
to be taken with the values given above. With slightly different starting values, the following
tetracritical coordinates were determined:

ξ = 0.31444 p = 25.92 bar G8x= 2.84 1010 J/mole


ζ = -0.69680 T = 149.93 K F= 5.7 10-8
λ = 0.20294 x = 0.26727
χ = 0.00000 V = 95.91 cm3.mole-1

The second solution is worse because the


residual value of the objective function
(F = 5.7 10-8) is larger than the one of the first
solution (F = 3.1 10-8). The coordinates show
that the global phase diagram parameters of the
tetracritical point in both solutions are almost the
same, the largest difference to be found is in ξ,
i.e., < 0.2 %. The tetracritical composition and
volume in both solutions are rather different,
6.9 % and 3.8 % respectively. Unfortunately we
were not able to obtain a solution with a lower
residual value of the objective function than the
first solution. For this reason we take the first
solution as the closest approximation of the
tetracritical point. A p,T-projection of the binary
mixture mixture that has this tetracritical point is
Figure 7.12: p,T-Projection of a binary shown in Fig.7.12. The parameters of the pure
system with a tetracritical components and the binary interaction
point. coefficient are:

c1 = 1.0000 c2 = 1.0000
T1* = 102.241 K T2* = 239.998 K
V1* = 38.30 cm3.mole-1 V2* = 20.00 cm3.mole-1

p1c = 15.22 bar p2c = 68.41 bar


T1c = 128.163 K T2c = 300.848 K
V1c = 243.67 cm3.mole-1 V2c = 127.24 cm3.mole-1
Z1c = 0.348 Z2c = 0.348
k12 = -0.110883
178 Chapter 7

In Fig.7.12 the solid curves represent the vapour pressure curves of the pure component and the
dotted curve the critical curve. The open circles are the critical points of the pure components
and the open box is the tetracritical point. The high-pressure immiscibility critical curve is
outside the scale of the diagram. A pressure minimum is observed along the high-pressure
immiscibility curve at (p, T) = (17.4 103 bar, 137.2 K). It is obvious from Fig.7.12 that this binary
mixture shows type I or II phase behaviour. Above T = 10 K no liquid-liquid immiscibilty at low
pressures was found for this mixture, which makes type I phase behaviour very likely.
At higher values of ξ the curved tricritical line that contained the tetracritical point is
separated into two curved tricritical lines. The Van Laar points were not found for ξ > 0.3 and
χ = 0.0. It is not clear whether the Van Laar points still exist after the bifurcation of the tricritical
curve in the global phase diagram. The global phase diagrams for ξ = 0.4, 0.5 and 0.6 are shown
in Figs.7.13, 7.14 and 7.15, respectively.

Figure 7.13: A global phase diagram of Figure 7.14: A global phase diagram of
the SPHCT equation for the SPHCT equation for
ξ = 0.4 and χ = 0.0. For an ξ = 0.5 and χ = 0.0. For an
explanation of symbols, see explanation of symbols, see
Fig. 7.1. Fig. 7.1.

It is shown in Figs.7.13, 7.14 and 7.15 that both branches of the left curved tricritical line enclose
a region in the global phase diagram that has never been obeserved before. Calculations have
been made to discover the topology of the critical curves for mixtures that are situated in this
region for ξ = 0.6 and χ = 0.0. A magnification of this region is shown in Fig.7.16.
The global phase diagram of the SPHCT equation 179

Figure 7.15: A global phase diagram of Figure 7.16: Enlargement of the global
the SPHCT equation for phase diagram of the SPHCT
ξ = 0.6 and χ = 0.0. For an equation for ξ = 0.6.
explanation of symbols, see
Fig. 7.1.

In Fig 7.16 the points a, b, c, d and e mark the binary mixtures where p,T-calculations
have been made. The results of these p,T-calculations are shown in Figs. 7.17-7.21. In Fig.7.22
we propose the scheme for the transitions of the various types of phase behaviour in this region
of the global phase diagram. We will now briefly describe the various transitions and the new
types of phase behaviour. In the Figs.7.17-7.21 the critical curves are shown as dotted curves,
irrespective of their kind of stability. The three-phase lines have not been shown in these figures.
At point a in Fig 7.16, the corresponding binary mixture shows type Vh phase behaviour,
see Fig.7.17 and 7.22. In Fig.7.17, a high-pressure immiscibility critical curve exists, although
this critical curve is not visible on this scale. The pressure minimum of this critical curve is
located at (p, T) = (2104.25 bar, 196.91 K). In Fig.7.17 the following parameters are used:
ξ = 0.60, ζ = -0.90 and λ = 0.37232 (k12 = -0.44). We have to increase the value of λ to 0.37668
(k12 = -0.43) to move towards point b in Fig 7.16. The p,T-projection of the critical curves of this
binary mixture is shown in Fig.7.18. Between the points a and b in Fig.7.16 an isolated critical
point is formed (at negative pressures), that forms a closed critical curve when the value of k12 is
increased to -0.43. We will call the type of phase behaviour that is shown in Fig.7.18, type VIIh,
see also Fig 7.22. In Fig.7.18 the high-pressure critical curve is not visible. The pressure
minimum of this critical curve is located at (p, T) = (1737.42 bar, 190.57 K). The pressure
minimum in b is 366.83 bar lower than the pressure minimum in point a in Fig.7.16.
180 Chapter 7

Figure 7.17: p,T-projection at ξ = 0.60, Figure 7.18: p,T-projection at ξ = 0.60,


ζ = -0.90 and λ = 0.37232 ζ = -0.90 and λ = 0.37668
(k12= -0.44). See point a in (k12= -0.43). See point b in
Fig.7.16. Fig.7.16.

Figure 7.19: p,T-projection at ξ = 0.60, Figure 7.20: p,T-projection at ξ = 0.60,


ζ = -0.90 and λ = 0.38103 ζ = -0.90 and λ = 0.39847
(k12= -0.42). See point c in (k12= -0.38). See point d in
Fig.7.16. Fig.7.16.
The global phase diagram of the SPHCT equation 181

If λ (or k12) is further increased to 0.38103 (k12 = -0.42), two phase behaviour
transformations occur. At λ = 0.38022 (k12 = -0.42187) a tricritical point appears in the
vapour-liquid critical curve just above the temperature maximum of the closed critical curve. In
Fig.7.16 we have passed the "horizontal" tricritical curve between point b and c. We will call the
type of phase behaviour just above the horizontal tricritical curve in Fig.7.16, type VII2h, see
Fig.7.22. In type VII2h phase behaviour the
vapour-liquid critical curve is interrupted twice
by a three-phase line. Between λ = 0.38022 and
λ = 0.38103 the closed critical curve merges
with the vapour-liquid critical curve. The
double critical endpoint forms the transition
state between type VII2h and type V2h phase
behaviour. Because the region between the
horizontal tricritical curve and the double
critical endpoint curve in Fig.7.16 is extremely
small, it is difficult to find a binary mixture that
exhibits type VII2h phase behaviour. At
λ = 0.38103 we are at point c in Fig.7.16. A
p,T-projection of the critical curves of the
binary mixture that corresponds to point c is
shown in Fig.7.19. This type of phase
Figure 7.21: p,T-projection at ξ = 0.60,
behaviour seems similar in p,T-space to the one
ζ = -0.87 and λ = 0.49216
that is found below the lower intersection point
(k12= -0.03). See point e in
of the curved tricritical lines in the global phase
Fig.7.16
diagram, i.e., Vsym. In type Vsym the demixing of
the liquid phase, that normally occurs near one of the vapour pressure curves in the p,T-
projection, occurs near both saturated vapour pressure curves.
The T,x- and V,x-projections of the critical curves of these binary mixtures show a certain
symmetry, hence type Vsym. However, we propose to call the new type of phase behaviour type
X, because the origin of the second three-phase line stems from a closed critical curve, see also
Fig.7.22. In giving names to the new types of phase behaviour we tried to stay as close as
possible to the original nomenclature of Van Konynenburg and Scott [1,2]. Sometimes, as in this
case, it would be confusing to add a super- or subscript to one of the existing types (I to VII) and
for this reason we added a fully new number to this nomenclature. The order of numbering is of
course arbitrary. The reason for calling this type of phase behaviour type X is the following: type
VIII and IX that will be described later in this section are similar to type VII and we preferred to
give these types of phase behaviour a number that is close to VII. The high-pressure critical
curve is not shown in Fig.7.19, it has a pressure minimum in p,T-space at the following
coordinates: (p,T) = ( 1358.32 bar, 184.00 K). The pressure minimum of the high-pressure
critical curve in point c in Fig.7.16 is 379.10 bar lower than the pressure minimum in point b in
Fig.7.16. If we only look at the topology of the critical curves in Figs.7.18 and 7.19, the
182 Chapter 7

transition from type VIIh to type X is via a mathematical double point of the first kind, see
section 5.2.3.
At k12 = -0.38, see point d in Fig.7.16, the high-pressure critical curve has exchanged
branches with the vapour-liquid critical curve at relatively low pressures. This is shown in
Fig.7.20 and Fig.7.22. A new type of phase behaviour has formed and we propose to call it type
IX. The transition from type X to type IX has taken place via a mathematical double point of the
second kind, see section 5.2.3. In p,T-space the high-pressure critical curve bounces on the
maximum of the vapour-liquid critical curve at lower pressures. It was already mentioned that
the pressure minimum in the high-pressure critical curve was declining when we moved from
point a to b to c in Fig. 7.16. Further increasing the value of k12 leads to the disappearance of one
of the three-phase lines in the vapour-liquid critical curve. This happens if the "vertical"
tricritical curve in Fig.7.16 is reached. We had to increase the value of ζ some to show this
transition, see Fig.7.16. To move from point d to point e we not only increased the value of k12 to
-0.03 but we also increased the value of ζ from ζ = -0.90 to ζ = -0.87. We propose to call the new
type of phase behaviour in and around point e in Fig.7.16, type VIII. The p,T-projection of the
critical curves of the mixture that corresponds to point e in Fig.7.16 is shown in Fig.7.21 and
Fig.7.22. The various types of phase behaviour are summarized in Fig.7.22.

Figure 7.22: New types of phase behaviour in the region sketched in Fig.7.16, see point a, b,.., e
The global phase diagram of the SPHCT equation 183

It is obvious that the nomenclature according to Van Konynenburg and coworkers [1,2] does not
suffice anymore. Although it is still possible to give names to the various types of phase
behaviour, the nomenclature is difficult to survey. Bolz [11] solved this problem by introducing a
new nomenclature that was described in chapter 1. According to this nomenclature the names of
the various types of phase behaviour are:

Vh ↔ VIIh ↔ VII2h ↔ X ↔ IX ↔ VIII [1,2]

2Pu ↔ 2Pnu ↔ 3Pnu ↔ 3Pu ↔ 1C3C ↔ 1C2C [11]

One of the advantages of the nomenclature according to Bolz [11] is that complicated phase
diagrams are easily given a name if we look at the p,T-projections of the phase diagrams. On the
other hand, we can easily reconstruct the p,T-projection of a binary mixture from its name
according to Bolz' classification. The classification of Van Konynenburg [1] did not have the
intention to describe the complicated phase diagrams, as shown in Figs. 7.17-7.21. The new
names were chosen by us by adding the super- and subscripts h or 2 to already existing names. If
this was not possible or if a conflict arose, we just added a new roman number to the already
existing types. Because complicated phase diagrams can today be calculated from equations of
state, the new classification according to Bolz [11] deserves closer study and attention.
We also performed global phase diagram calculations for molecules with extreme size
ratios, e.g., ξ = 0.7, 0.8 and 0.9. Because no new types of phase behaviour were found and the
global phase diagrams could only be calculated partially, we will not discuss those results.
We want to conclude this section with the remark that, although the new types of phase
behaviour can be calculated from the SPHCT equation of state, some types are not likely to
occur in practice. The molecules of component 1 are composed of 12 CH2-segments while the
flexibility parameter of this component is c1 = 1. The physical meaning of c1 = 1 is that molecules
of component 1 have no rotational and vibrational degrees of freedom, i.e., the molecules cannot
rotate or vibrate and act like stiff rods. It is not very likely that molecules that are composed of
12 CH2-segments act like stiff rods, e.g., compare n-decane: c = 4.1717. However, molecules
such as trans-decalin are composed of CH2-segments and may show more stiffness than n-
alkanes. In the next chapter we will show that type VIII behaviour can be calculated with
parameters of practical components, although the binary interaction coefficient will still be
fictitious.

7.4 The global phase behaviour of molecules with equal sizes but unequal flexibilities.

In this section we will briefly discuss some global phase diagrams for molecules with
unequal flexibilities and equal sizes. Spectacular new types of phase behaviour were found for
molecules with different sizes in section 7.3, but not for molecules with unequal flexibilities. We
could not locate a tetracritical point along one of the curved tricritical lines in the global phase
184 Chapter 7

diagram. For this reason, we will not discuss these global phase diagrams as extensively as we
did in the previous section. In general we can say that, from a numerical point of view, the global
phase diagrams for ξ = 0 and χ > 0 were a lot harder to calculate than global phase diagrams for
ξ > 0 and χ = 0. Of course, the conditions used to calculate the curves in the global phase diagram
remained the same. Above χ = 0.40 the global phase diagrams could only be partially calculated.
These results are not shown in this study. The global phase diagrams for χ = 0.10, 0.20, 0.30 and
0.40 are shown in Figs.7.23-7.26. It is shown in these figures that the kite-shaped type I/II region
shifts to the right of the global phase diagram and turns anti-clockwise. The kite-shaped region
becomes smaller at increasing values of χ, as was the case for ξ > 0. The termination point of the
vertical tricritical line moves to lower values of λ if the value of χ is increased. The termination
point reaches a minimum value of λ = 0.230 at χ ≈ 0.4. If the value of χ is further increased, the
termination point moves upwards again in the global phase diagram. The coordinates of the
termination point of the vertical tricritical curve are given in Table 7.6.

F igure 7.23: A global phase diagram of the Figure 7.24: A global phase diagram of the
SPHCT equation for ξ = 0.0 SPHCT equation for ξ = 0.0
and χ = 0.1. For an explana- and χ = 0.2. For an explana-
tion of symbols, see Fig.7.1. tion of symbols, see Fig.7.1.

The Van Laar point on the left tricritical curve moves very rapidly upwards and to the
right in the global phase diagram if the value of χ is increased. The Van Laar point on the right
tricritical curve moves downwards and to the right in the global phase diagram if the value of χ is
increased. If χ ≥ 0.3, the Van Laar point shifts upwards towards higher values of λ. The
coordinates of the Van Laar points are given in Table 7.7. We suggest that additional global
phase diagram calculations should be made to give a comprehensive review on the effect of χ on
the global phase diagram.
The global phase diagram of the SPHCT equation 185

Table 7.6: The coordinates of the ter-


mination point of the vertical
tricritical curve as a function
of χ at ξ = 0.0.

χ ζ λ

0.0 0.00000 0.24417

0.1 0.06056 0.23849

0.2 0.12463 0.23426

0.3 0.19296 0.23122

0.4 0.26654 0.22950

0.5 0.34683 0.22976

0.6 0.43588 0.23377

Figure 7.25: A global phase diagram of the Figure 7.26: A global phase diagram of the
SPHCT equation for ξ = 0.0 SPHCT equation for ξ = 0.0
and χ = 0.3. For an explana- and χ = 0.4. For an explana-
tion of symbols, see Fig.7.1. tion of symbols, see Fig.7.1.
186 Chapter 7

Table 7.7: Global phase diagram coordinates of the Van Laar point of the SPHCT equation
as a function of χ at ξ = 0.0.

Van Laar point LEFT Van Laar point RIGHT

χ ζ λ ζ λ

0.0 -0.19760 0.19131 0.19760 0.19131

0.1 0.34007 0.52751 0.34527 0.12423

0.2 0.77580 0.82501 0.43741 0.10240

0.3 - - 0.51061 0.10009

0.4 - - 0.57578 0.11094

7.5 The global phase behaviour of molecules with unequal flexibilities and sizes.

In this section we will briefly discuss some global phase diagrams for molecules with
unequal flexibilities and unequal sizes. Significant new types of phase behaviour were not found
for these mixtures. It eventuated that the global phase diagrams for ξ = χ > 0 were more difficult
to calculate than global phase diagrams with χ = 0.0. Above χ = 0.60 the global phase diagrams
could only be partially calculated. However, we expect that with more numerical effort, e.g.,
using better starting values and using smaller step sizes, the tricritical curves of these global
phase diagrams could also be calculated. The results for ξ = χ > 0.60 are not shown in this
chapter. The global phase diagrams for ξ = χ = 0.10, 0.20, 0.30, 0.40, 0.50 and 0.60 are shown in
Figs.7.27-7.32. It is shown in these figures that the kite-shaped type I/II region shifts to the left of
the global phase diagram and turns clockwise. The effect of ξ, i.e., pulling the kite-shaped region
to the left in the global phase diagram, is stronger than the effect of χ, i.e., pulling the kite-shaped
region to the right in the global phase diagram. The kite-shaped region becomes smaller at
increasing values of ξ and χ. The termination point of the vertical tricritical curve shifts towards
lower λ-values and to the left in the global phase diagram if the values of ξ and χ are increased. It
cannot be predicted whether it will go upwards again at even higher values of ξ and χ. The Van
Laar point on the left tricritical curve shifts towards the left and downwards in the global phase
diagram if the values of ξ and χ are increased. If ξ = χ ≥ 0.6, it moves towards higher λ-values.
The Van Laar point on the right tricritical curve moves relatively rapid upwards along the
tricritical curve in the global phase diagram. Again the effect of increasing ξ seems to dominate
the effect of increasing χ. The coordinates of the termination point of the vertical tricritical curve
and the Van Laar points are given in Table 7.8 and 7.9, respectively.
The global phase diagram of the SPHCT equation 187

Figure 7.27: A global phase diagram of the Figure 7.28: A global phase diagram of the
SPHCT equation for ξ = 0.1 SPHCT equation for ξ = 0.2
and χ = 0.1. For an explana- and χ = 0.2. For an explana-
tion of symbols, see Fig.7.1. tion of symbols, see Fig.7.1.

Figure 7.29: A global phase diagram of the Figure 7.30: A global phase diagram of the
SPHCT equation for ξ = 0.3 SPHCT equation for ξ = 0.4
and χ = 0.3. For an explana- and χ = 0.4. For an explana-
tion of symbols, see Fig.7.1. tion of symbols, see Fig.7.1.
188 Chapter 7

Figure 7.31: A global phase diagram of the Figure 7.32: A global phase diagram of the
SPHCT equation for ξ = 0.5 SPHCT equation for ξ = 0.6
and χ = 0.5. For an explana- and χ = 0.6. For an explana-
tion of symbols, see Fig.7.1. tion of symbols, see Fig.7.1.

For mixtures with ξ > 0.0 and χ > 0.0 we did


Table 7.8: The coordinates of the
not find a tetracritical point. However, it was
termination point of the
not possible to calculate the left tricritical curve
vertical tricritical curve as a
near the corner in the curve. The tetracritical
function of ξ and χ with ξ = χ.
point appeared exactly in this corner, for
ξ = 0.31390 and χ = 0.0. In that case it was also
ξ=χ ζ λ
very difficult to calculate the tricritical curve
around the tetracritical point. This is probably 0.0 0.00000 0.24417
due to the fact that in the neighbourhood of a 0.1 -0.06123 0.23993
tetracritical point, the values of the higher
derivatives, i.e., G6x and G7x approach zero. If 0.2 -0.11053 0.23831
ξ = χ ≥ 0.6, convergence problems in the 0.3 -0.15137 0.23744
calculation of the tricritical curve in this region
0.4 -0.18595 0.23652
occurred again. This may be caused by G6x and
G7x approaching zero, indicating that a 0.5 -0.21580 0.23489
tetracritical point exists for ξ and χ > 0.0. We 0.6 -0.24197 0.23228
suggest that additional global phase diagram
0.7 -0.26522 0.22814
calculations should be made to give a
comprehensive review on the effects of both ξ and χ on the global phase diagram.
The global phase diagram of the SPHCT equation 189

Table 7.9: Global phase diagram coordinates of the Van Laar point of the SPHCT equation
as a function of χ and ξ with ξ = χ.

Van Laar point LEFT Van Laar point RIGHT


ξ=χ ζ λ ζ λ
0.0 -0.19760 0.19131 0.19760 0.19131
0.1 -0.27939 0.18039 0.08270 0.21423
0.2 -0.35527 0.16759 -0.03684 0.25905
0.3 -0.40640 0.16732 -0.17109 0.33011
0.4 -0.45900 0.16177 -0.34908 0.45080
0.5 -0.51299 0.15142 -0.68063 0.71958
0.6 -0.55128 0.14965 - -
0.7 -0.57462 0.15748 - -

NOMENCLATURE

ci flexibility parameter, i.e., one-third of the external degrees of freedom of one molecule of
component i
F objective function
G molar Gibbs energy
Gnx n-th derivative of G with respect to x
k12 binary interaction coefficient
p pressure
T temperature
Ti* characteristic temperature of component i
V molar volume
Vi* closest-packed volume of component i
x mole fraction of component 2
Z compressibility factor

Greek symbols

ζ dimensionless parameter ratio, see Eq.(5.1)


λ dimensionless parameter ratio, see Eq.(5.2)
ξ dimensionless parameter ratio, see Eq.(5.3)
χ dimensionless parameter ratio, see Eq.(5.4)

Subscripts and superscripts


c critical
190 Chapter 7

REFERENCES

1. P.H. van Konynenburg. Critical lines and phase equilibria in binary mixtures. PhD-thesis
University of California Los Angeles. (1968).
2. P.H. van Konynenburg and R.L. Scott. Critical lines and phase equilibria in binary van der
Waals mixtures. Phil. Trans. 298A, (1980), 495-540.
3. U.K. Deiters and I.L. Pegg. Systematic investigation of the phase behavior in binary fluid
mixtures. I. Calculations based on the Redlich-Kwong equation of state. J. Chem. Phys. 90,
(1989), 6632-6641.
4. Th. Kraska and U.K. Deiters. Systematic investigation of the phase behavior in binary fluid
mixtures. II. Calculations based on the Carnahan-Starling-Redlich-Kwong equation of state.
J. Chem. Phys. 96, (1992), 539-547.
5. D. Furman, S. Dattagupta and R.B. Griffiths. Global phase diagram for a three-component
model. Phys. Rev. B 15, (1977), 441-464.
6. D. Furman and R.B. Griffiths. Global phase diagram for a Van der Waals model of a binary
mixture. Phys. Rev. A 17, (1978), 1139-1148.
7. I. Ch. Wei and R.L. Scott. In search of the Griffiths shield region. J. Stat. Phys. 52, (1988),
1315-1324.
8. V.A. Mazur, L.Z. Boshkov and V.B. Fedorov. Phase behaviour in two component
Lennard-Jones systems. Dokl. Akad. Nauk SSSR. 282, (1985), 137-140.
9. V.A. Mazur, L.Z. Boshkov and V.G. Murakhovsky. Global phase behaviour of binary
mixtures of Lennard-Jones molecules. Physics Letters 104, (1984), 415-418.
10. L.Z. Boshkov and V.A. Mazur. Phase equilibria and critical lines of binary mixtures of
Lennard-Jones molecules. Russ. J. Phys. Chem. 60, (1986), 16-19.
11. A. Bolz. Vergleichende Untersuchung globaler Phasendiagramme. PhD-thesis,
Ruhr-Universität Bochum, Germany, (1992).
12. Th. Kraska. Systematische Untersuchung des globalen Phasenverhaltens binärer fluider
Mischungen ausgehend von der Carnahan-Starling-Redlich-Kwong-Zustandsgleichung.
Diploma thesis, Ruhr-Universität Bochum, Germany, (1989).
13. P.H.E. Meijer, M. Keskin and I.L. Pegg. Critical lines for a generalized three state binary
gas-liquid lattice model. J. Chem. Phys. 88, (1988), 1976-1982. Erratum: J. Chem. Phys. 90,
(1989), 3408.
14. P.H.E. Meijer. The van der Waals equation of state around the van Laar point. J. Chem.
Phys. 90, (1989), 448-456.
15. P.H.E. Meijer, I.L. Pegg, J. Aronson and M. Keskin. The critical lines of the van der Waals
equation for binary mixtures around the van Laar point. Fluid Phase Equilibria 58, (1990),
65-80.
16. L.Z. Boshkov. On the description of phase diagrams of two component mixtures with a
closed domain of demixing based on the one-fluid model of an equation of state. Dokl.
Akad. Nauk. SSSR. 294, (1987), 901-905
17. G. Jackson, J.S. Rowlinson and C.A. Leng. Phase equilibria in model mixtures of spherical
molecules of different sizes. J. Chem. Soc., Faraday Trans. 1., 82, (1986), 3461-3473.
191

CHAPTER 8
SPECIAL PHENOMENA FROM THE GLOBAL PHASE DIAGRAM

In this chapter we will highlight four particular phenomena that occur in the global
phase diagram calculated from the SPHCT equation. In section 8.1 we will pay attention to the
type VI/VII-region. The course of the closed critical curve and the high-pressure critical curve
are studied in section 8.2. In section 8.3 we will describe the appearance of type VIII phase
behaviour and finally, in section 8.4 we will go into detail on the Van Laar point. In this chapter
the relation between real existing components and the particular phenomena under
consideration is emphasized. We also show the projections of the critical curves in a number of
different ways, in order to widen our knowledge of phase diagrams.

8.1 The type VI/VII-region in the global phase diagram

As already mentioned in section 7.1, it is possible to calculate type VI and type VII phase
behaviour from the SPHCT equation of state. The boundaries of the type VI/VII-region in the
global phase diagram are formed by the double critical endpoint curve. Unfortunately we were
not able to calculate this curve in the global phase diagram, see section 7.1. To confirm the
findings of Boshkov [1], we made a grid over the type VI/VII-region in the global phase diagram
for ξ = χ = 0.0 and calculated the phase diagrams in the grid points. This situation is shown in
Fig.8.1. In this figure the solid curves represent the mathematical double point curves, (the
MDPs are of the first kind), the dotted curve represents the tricritical curve, the open circles are
located on the spots where we have performed phase diagram calculations. It is impossible to
show all calculated phase diagrams. Eight phase diagrams were selected that give a good insight
into the topology of the critical curves in this region. These points are shown in Fig.8.2 by black
dots. The global phase diagram parameters of all black dots in Fig.8.2 are given in Table 8.1. The
other symbols have the same meaning as in Fig.8.1.
In the previous chapter we used p,T-diagrams to study the connectivity of the critical
curves. In this section we show T,x-projections of the critical curves of the binary systems that
are under investigation. One of the advantages of using T,x-projections of the critical curves is
the fact that the limit of stability (G4x = 0) is recognized by an extremum in the critical curve:
(∂ T/∂ x)c = 0. To prove this, we start with the following fundamental equation that was already
used in section 5.2.2:

d G = - S d T +V d p + Gx d x (8.1)
192 Chapter 8

Figure 8.1: Part of the global phase Figure 8.2: Magnification of Fig.8.1.
diagram of the SPHCT The black dots represent
equation for ξ = 0.0 and points of which the p,T-pro-
χ = 0.0. The open circles re- jection of the critical curves
present the points were p,T- are discussed.
calculations have been
made.
If Eq.(8.1) is differentiated three times with
Table 8.1: Coordinates of the black dots in
respect to the composition, we obtain:
the global phase diagram of
Fig. 8.2 where ζ = 0.364. d G3x = - S 3x d T + V 3x d p + G4x d x (8.2)

point λ k12 Along the critical curve it holds that


a 0.03135 -0.040 dG3x = 0, which leads to:
b 0.03600 -0.035 c
V 3x
c
G 4x (8.3)
d T = c d p+ c d x
c 0.05090 -0.019 S 3x S 3x
d 0.05184 -0.018 Insertion of dp = (S2xc/V2xc)dT, see section
e 0.05277 -0.017 5.2.2, into Eq.(8.3) and rearrangement of the
f 0.05370 -0.016 terms, gives:
g 0.05929 -0.010 ⎛ dT⎞ c c
V 2x G 4x (8.4)
⎜ ⎟= c c c c
h 0.06022 -0.009 ⎝ d x ⎠c S 3x V 2x - V 3x S 2x

From Eq.(8.4) it is evident that if G4x = 0 then (dT/dx)c = 0. Another advantage of the use of T,x-
projections of the critical curves is the absence of the cusps in the critical curves, that appear in
p,T-projections. For the p,T-calculations we have used the following pure component parameters:
Special phenomena 193

c1 = 1.000 c2 = 1.000
V1*= 20.00 cm3.mole-1 V2*= 20.00 cm3.mole-1
T1*= 514.71 K T2*= 240.00 K

We will start our investigations at point a in Fig.8.2. The T,x-projection of the critical curve that
corresponds to point a of Fig.8.2 is shown in Fig.8.3. In this figure the solid curves represent
locally stable critical curves. In Fig.8.3 we focus on the composition range [0.7, 1.0] because the
interesting phenomena occur in this composition range. The phase behaviour at point a belongs
to type I or II. We will see that the liquid-liquid critical curve at relatively low temperatures,
which is characteristic for type II phase behaviour, is very difficult to detect. While moving from
point a to point b in Fig.8.2, a closed critical curve forms. This behaviour is shown in Fig.8.4. In
this figure the dotted curves represent locally unstable critical curves. The origin of this closed
critical curve lies in the appearance of an isolated critical point between points a and b in Fig.8.2.

Figure 8.3: T,x-projection of the critical Figure 8.4: T,x-projection of the critical
curves of the binary mixture curves of the binary mixture
corresponding to point a in corresponding to point b in
Fig.8.2. Fig.8.2.

The right half of the closed critical curve is unstable (G4x < 0) and the left half is (locally) stable.
At the maximum and minimum of the closed critical curve, G4x equals zero which is in
agreement with Eq.(8.4). At point b no liquid-liquid critical curve was found. It is most likely
that the system shows type I phase behaviour, although type II phase behaviour is possible.
Moving upwards in the global phase diagram to point c in Fig.8.2, the closed critical curve
becomes larger. At point c the type II liquid-liquid critical curve has been found at relatively low
temperatures. This is shown in Fig.8.5. The concentration range of this critical curve is very
small, i.e. Δ x < 0.01. The magnification displayed in Fig.8.5 shows the liquid-liquid critical
194 Chapter 8

curve. The right half of the liquid-liquid critical curve is unstable (G4x < 0). The vapour-liquid
critical curve starts to show a dip near the closed critical curve. At λ = 0.051682 (k12 = -0.018167)
the tricritical curve in the global phase diagram is passed, see Fig.8.2. Just above the tricritical
curve, see point d in Fig.8.2, an S-shaped loop in the vapour-liquid critical curve can be
observed. The phase diagram corresponding to point d in Fig.8.2, is shown in Fig.8.6. Moving
further upwards to point e in Fig.8.2, the S-shaped loop becomes more pronounced, see Fig.8.7.

Figure 8.5: T,x-projection of the critical Figure 8.6: T,x-projection of the critical
curves of the binary mixture curves of the binary mixture
corresponding to point c in corresponding to point d in
Fig.8.2. Fig.8.2.

The closed critical curve almost touches the vapour-liquid critical curve. At low temperatures the
distance between the closed critical curve and the liquid-liquid critical curve also becomes
smaller. This is shown in Fig.8.7. At λ = 0.052192 (k12 = -0.016845) the closed critical curve and
the vapour-liquid critical curve merge. In Fig.8.2 we are exactly on one of the mathematical
double point curves, between point e and point f. Just above this mathematical double point
curve, see point f in Fig.8.2, there is a gap in the closed critical curve. This is shown in Fig.8.8.
When point g is reached in Fig.8.2 we are almost on the "horizontal" mathematical
double point curve. The closed critical curve, or what is left of it, almost touches the liquid-liquid
critical curve at low temperatures. This is shown in Fig.8.9. At λ = 0.059584 (k12 = -0.00968), we
have exactly reached the horizontal mathematical double point curve, see Fig.8.2. Just above this
mathematical double point curve, point h in Fig.8.2, the collision of both critical curves has
occurred. This is shown in Fig.8.10.
Special phenomena 195

This result is in agreement with the findings of Boshkov [1], although it is not a definite
confirmation, due to our inability to calculate the double critical endpoint curves in Figs.8.1 and
8.2. The transitions between the various types of phase behaviour and the intermediate phase
diagrams are summarized in Fig.8.11.

Figure 8.7: T,x-projection of the critical cur- Figure 8.8: T,x-projection of the critical cur-
ves of the binary mixture cor- ves of the binary mixture cor-
responding to point e in Fig.8.2. responding to point f in Fig.8.2.

Figure 8.9: T,x-projection of the critical cur- Figure 8.10: T,x-projection of the critical cur-
ves of the binary mixture cor- ves of the binary mixture cor-
responding to point g in Fig.8.2. responding to point h in Fig.8.2.
196 Chapter 8

Figure 8.11: Summary of the various phase diagrams and intermediate transition states. All
diagrams have been calculated from: y2-x3+8x2-20x-a=0. In this figure y may
represent the mole fraction and x the temperature. The equation above has no
physical meaning whatsoever and is only used for illustration [2].

The closed critical curve can also interfere with the high-pressure critical curve in
p,V,T,x-space, consequently leading to a mathematical double point of the second kind. This is
not shown in the global phase diagram of Fig.8.2, because high-pressure immiscibility does not
occur if ξ = 0. To show the T,x-projection of the critical curves of a binary mixture with a
mathematical double point of the second kind, we performed phase diagram calculations for
ξ = -0.2581 and χ = 0.0. The results of these calculations are shown in Figs.8.12, 8.13 and 8.14.
The pure component parameters of the components are:
c1 = 2.000 c2 = 2.000
* 3
V1 = 27.00 cm .mole -1
V2*= 35.00 cm3.mole-1
T1*= 205.00 K T2*= 260.00 K

The p,T-projections of the high-pressure critical curves have already been shown in section 5.2.6.
In Figs.8.12-8.14, the solid curves represent the stable parts of the critical curves and the dotted
curves the unstable parts. In Fig.8.12 k12 = 0.04510. There is a separate closed critical curve and a
separate high-pressure critical curve. Both curves seem to intersect twice but this is only a
consequence of the projection on the T,x-plane.
Special phenomena 197

Figure 8.13: T,x-projection of the critical


Figure 8.12: T,x-projection of the closed criti-
cal curve and the high-pressure curves, just after a mathematical
critical curve. double point has occurred.

At k12 = 0.04515, both critical curves have


merged. This is shown in Fig.8.13. Further
increasing the value of k12 to 0.04520 leads to a
widening of the gap in the closed critical curve
in the T,x-projection, see Fig.8.14. The
corresponding p,T-projection of the critical
curves of the system, shown in Fig.8.14, has a
tube of immiscibility, see section 5.2.3.
In the next section we will discuss a
binary mixture that might exhibit the type
VI/VII-phenomena. The word "might" is
appropriate because although the pure
component parameters are those of real
components, the binary interaction coefficient is
not. Even if the binary interaction coefficient
would be realistic, it would probably be Figure 8.14: T,x-projection of the critical
impossible to determine the closed immiscibility curves after the mathematical
region because in practice as it would most double point has occurred.
likely be obscured by solid phases.
198 Chapter 8

8.2 Critical curves in the binary mixture ethane(1)+n-butane(2)

It has already been shown that the mixture ethane(1) + n-butane(2) exhibits type I phase
behaviour, see section 4.7. The value of the binary interaction coefficient k12 that gives an
optimal description of the vapour-liquid critical curve is k12 = 0.0367, see Table 4.4. If, however,
the value of k12 is changed, a lot of interesting phenomena in this binary mixture are encountered.
It is stressed here that the diagrams that are obtained if k12 is varied, do no longer represent the
practical mixture ethane(1) + n-butane(2). We will discuss the critical phenomena that have been
calculated in this section and illustrate them with y1,y2-density diagrams. This representation
helps to understand the connectivity of the critical curves as discussed by Meijer et al.[3,4]. For
this purpose, the unstable parts of the critical curves are also shown. The origin of the y1,y2-
density diagrams stems from ternary lattice-gas mixtures. In lattice-gas theories of binary fluids,
volume variations are modelled by introducing vacant sites ("holes") as a third species. There are
now three lattice site fractions:
*
xi V i ,
yi = i = 1, 2
V
(8.5)
*
V
y0 = 1 - y1 - y 2 = 1 -
V

We used a similar approach in section 3.5, where a pure fluid was considered to be a binary
mixture composed of molecules and holes. The y1,y2-diagrams that we will use, have three pure
component points. We used the following conventions: the lower left corner of the density
triangle is the point where every lattice site contains a molecule of component 1. The lower right
corner of the density triangle is the point where every lattice site contains a molecule of
component 2 and the upper corner of the density triangle is the point where the lattice is
completely empty (or completely filled with holes). The left side of the triangle or "ethane-side",
thus contains mixtures of component 1 with holes and the right side of the triangle or "n-butane-
side" contains mixtures of component 2 with holes. The horizontal side of the triangle contains
binary mixtures that are composed of molecules of component 1 and 2 without holes! This
horizontal side is the equivalent of the pressure-ceiling in a p,T-diagram according to the SPHCT
equation. It should be noted that we used triangular coordinates for our y1,y2-density diagrams.
The triangular coordinates are normally used for the graphical representation of the Gibbs
triangle. The y1,y2-diagrams of Meijer et al.[3,4] and Kraska and Deiters [5] are slightly
differently defined and consequently look a little different. Their y1,y2-density triangles contains
an angle of 90°. In the following projections the (meta)stable critical curves are shown as solid
curves and the unstable critical curves as dotted curves.
For all mixtures, the pure component parameters of ethane and n-butane were used. The
pure component parameters were taken from Table 4.1. We start our investigations with
k12 = 0.0450. This binary mixture shows type Ih (1Pu) phase behaviour. This is shown in
Figs.8.15(a) and 8.15(b). It was already mentioned in chapter 1 and 7 that the nomenclature
Special phenomena 199

according to Bolz [6] deserves closer attention. We think that this nomenclature is at least a good
supplement to the nomenclature of Van Konynenburg and Scott [7,8], but the latter one might
even become superfluous. For this reason we we will give the names of each type of phase
behaviour according to both nomenclatures, the one of Bolz between brackets, e.g. Ih (1Pu).
In Fig.8.15(a) the vapour-liquid critical curve is situated in the upper part of y1,y2-density
triangle. It is striking to see, how relatively few lattice sites are occupied along the vapour-liquid
critical curve. The high-pressure critical curve is situated in the lower part of the y1,y2-density
triangle, close to the horizontal axis. The corresponding p,T-projection of the critical curves of
this binary mixture is shown in Fig.8.15(b). To get both critical curves on the same scale in
Fig.8.15(b), a logarithmic pressure axis was used. The pressure minimum along the high-
pressure critical curve is situated at about (p,T)=(20.0 103 bar, 90.5 K).

Figure 8.15(a): y1,y2-Density triangle for Figure 8.15(b): p,T-Projection for


C2H6 (1)+ n-C4H10 (2) with C2H6 (1) + n-C4H10 (2)
k12= 0.0450. with k12 = 0.0450.

Increasing the value of k12 to 0.0460 leads to the formation of a closed critical curve, thus
leading to type VIh phase behaviour. This is shown in Figs.8.16(a) and 8.16(b). This closed
critical curve originates from an isolated critical point. This isolated critical point appears at
k12 = 0.045406. The coordinates of the isolated critical point in p,V,T,x-space are:
picp = -1120.62 bar xicp = 0.1898
Ticp = 85.45 K Vicp = 53.93 cm3.mole-1

The dotted curves in Figs.8.16(a) and 8.16(b) show that a part of the closed critical curve is
unstable, as already discussed in section 8.1. It is also shown in Fig.8.16(a) that the high-pressure
critical curve shifts upwards in the triangle, i.e., to lower pressures in the p,T-plane. The inset in
200 Chapter 8

Fig.8.16(b) shows a magnification of the closed critical curve in the p,T-plane. The high-pressure
critical curve is not shown in Fig.8.16(b). This curve has a pressure minimum at about ( p,T ) =
(7.4 103 bar, 87.1K).

Figure 8.16(a): y1,y2-Density triangle for Figure 8.16(b): p,T-Projection for


C2H6 (1)+ n-C4H10 (2) with C2H6 (1) + n-C4H10 (2)
k12= 0.0460. with k12 = 0.0460.

Figure 8.17(a): y1,y2-Density triangle for Figure 8.17(b): p,T-Projection for


C2H6 (1)+ n-C4H10 (2) with C2H6 (1) + n-C4H10 (2)
k12= 0.0480. with k12 = 0.0480.
Special phenomena 201

At k12 = 0.0480 the high-pressure critical curve has merged with the closed critical curve. This is
shown in Figs.8.17(a) and 8.17(b). In the p,T-projection of the critical curves of this binary
mixture, a "tube of immiscibility" has formed, see also section 2.7 and Fig.2.14(d). Inside this
tube, two liquid phases coexist, outside this tube there is a homogeneous liquid phase. This type
of phase behaviour has never been calculated before from a pressure-explicit equation of state.
Now that this type of phase behaviour can be calculated from the SPHCT equation, we must
identify it according to the classification of Van Konynenburg and Scott [7,8] to distinguish it
from the other type(s) VI phase behaviour, see Fig.2.14.(a)-(d). In chapter 7 we have given a new
number to all types of phase behaviour in which the high-pressure critical curve interferes with
the critical curves at (relatively) low pressures. We therefore propose to call this type of phase
behaviour type XI (1Pll). The transition state between type VIh (1Pnu) and type XI (1Pll) phase
behaviour is formed by a mathematical double point of the second kind. The mathematical
double point in the binary mixture with the pure component parameters of ethane and n-butane is
found at k12 = 0.047284. The coordinates of the mathematical double point in p,V,T,x-space are:

pMDP = -818.30 bar xMDP = 0.220


TMDP = 80.31 K VMDP = 45.74 cm3.mole-1

We recall that for a mathematical double point of the second kind: G4x ≠ 0. To demonstrate this,
the value of G4x at this mathematical double point was calculated, resulting in: G4x,MDP = 9222.63
J.mole-1. If the value of k12 is increased to 0.0500, again a closed critical curve is present. This is
shown in Figs.8.18(a) and 8.18(b).

Figure 8.18(a): y1,y2-Density triangle for Figure 8.18(b): p,T-Projection for


C2H6 (1)+ n-C4H10 (2) with C2H6 (1) + n-C4H10 (2)
k12= 0.0500. with k12 = 0.0500.
202 Chapter 8

Because this closed critical curve is situated wholly at negative pressures, there is no
need to talk about a new type of phase behaviour. For experimentalists, the second closed critical
curve is not an important phenomenon. Even if the value of k12 is further increased, this curve
will not obtain a positive critical pressure. It should be mentioned that the curve is not unstable
because along this closed critical curve G4x is always positive. This also means that this closed
critical curve shows no cusps in the p,T-plane, as the first closed critical curve did. In the p,T-
projection of the critical curves, the second closed critical curve looks like an "egg" that rests on
the unstable part of the first closed critical curve, see Fig.8.18(b).
To study the development of this second closed critical curve we performed calulations
for this binary mixture with k12 = 0.0520. The results are shown in Figs.8.19(a) and 8.19(b). In
Fig.8.19(a), it is shown that the remaining part of the high-pressure critical curve and the first
closed critical curve enclose a larger region in the y1,y2-density triangle. The second closed
critical curve rapidly grows towards the upper corner of the y1,y2-density triangle while it widens
at the bottom. In the p,T-projection of the critical curves, the pressure maximum of the second
closed critical curve shifts upwards and slightly to the left, see Fig.8.19(b). Just before
k12 = 0.0540 the second closed critical curve merges with the "tube of immiscibility". This
transition takes place via a mathematical double point of the first kind. This leaves only two
distinct critical curves in this binary mixture. We will refer to these critical curves as the vapour-
liquid critical curve and the "alternate" critical curve. At positive pressures, no differences can be
observed in the connectivity of the critical curves in the p,T-projection.

Figure 8.19(a): y1,y2-Density triangle for Figure 8.19(b): p,T-Projection for


C2H6 (1)+ n-C4H10 (2) with C2H6 (1) + n-C4H10 (2)
k12= 0.0520. with k12 = 0.0520.
Special phenomena 203

Figure 8.20(a): y1,y2-Density triangle for Figure 8.20(b): p,T-Projection for


C2H6 (1)+ n-C4H10 (2) with C2H6 (1) + n-C4H10 (2)
k12= 0.0700. with k12 = 0.0700.

To demonstrate this clearly, calculations at k12 = 0.0700 were performed. The results of
these calculations are shown in Figs.8.20(a) and 8.20(b). The top of the second closed critical
curve almost touches the upper corner of the y1,y2-density triangle. In the p,T-projection, the
pressure maximum of the second closed critical curve shifts towards the origin of the p,T-plane,
see Fig.8.20(b). The alternate critical curve contains two parts of instability (G4x < 0). However,
the value of G4x is not continuous anymore along the alternate critical curve as should be
expected. This can be demonstrated with the transition on the alternate critical curve from
instability to (meta)stability that is closest to the ethane-side of the y1,y2-density triangle. Coming
from the upper corner of the y1,y2-density triangle and moving towards the lower left corner of
the triangle, the critical curve starts to exhibit stability. Near (y0, y1, y2) = (0.5, 0.5, 0.0) in
Fig.8.20(a) a transition to instability occurs along the critical curve. The value of G4x rises
asymptotically to "plus-infinity" when the transition point is approached. Just after the transition
point the sign of G4x turns to "minus-infinity" and its value increases (but still remains negative)
when we move away from the transition point. There is a discontinuity at the transition point in
G4x. All other properties, like pressure, temperature, composition and volume, are continuous
along the transition point. For this reason we decided that there exists only one alternate critical
curve in Figs.8.20(a) and 8.20(b), although G4x can be discontinuous along this critical curve.
At k12 = 0.10000 we observe the following trends, see Figs.8.21(a) and 8.21(b): the left
"leg" of the alternate critical curve shifts towards the "ethane-side" of the y1,y2-density triangle,
see Fig.8.21(a). The maximum in the alternate critical curve close to the upper corner of the
y1,y2-density triangle can no longer be distinguished from the upper corner itself, although there
204 Chapter 8

is no intersection. Unfortunately, no solutions were found when we tried to track the low-
temperature liquid-liquid critical curve to positive pressures, see Fig.8.22(b). The other
maximum in the alternate critical curve, where the transition from (meta)stability to instability
occurs, shifts upwards in the triangle and a collision with the vapour-liquid critical curve can be
expected at higher values of k12.

Figure 8.21(a): y1,y2-Density triangle for Figure 8.21(b): p,T-Projection for


C2H6 (1)+ n-C4H10 (2) with C2H6 (1) + n-C4H10 (2)
k12= 0.1000. with k12 = 0.1000.
Just before this collision will take place, a dip forms in the vapour-liquid critical curve. The
situation for k12 = 0.17200 is shown in Figs.8.22(a) and 8.22(b). The connectivity of the vapour-
liquid critical curve and the alternate critical curve will change via a mathematical double point
of the first kind.
It was extensively demonstrated in section 8.1 that a tricritical point has to appear on the
critical curve that is completely stable near the point of collision. In this case the tricritical point
will appear on the vapour-liquid critical curve in order that an unstable region on this critical
curve be created. The vapour-liquid critical curve and the alternate critical curve will form a
mathematical double point of the first kind at the point where both critical curves become
unstable (G4x = 0). After the mathematical double point the phase behaviour is type III. For
k12 = 0.17400 the situation is shown in Figs.8.23(a) and 8.23(b). The left "leg" of the alternate
critical curve seems to have merged with the ethane-side of the y1,y2-density triangle. This is of
course not possible because it would imply that a range of critical temperatures and pressures
would exist for pure ethane. This critical curve was traced to the point where it should bend to
the lower left corner of the y1,y2-density triangle. At this point the critical temperature along the
alternate critical curve is below 2K. Unfortunately, at this moment we were not able to trace it
any further and consequently we cannot explain what exactly happens with this critical curve.
Special phenomena 205

Figure 8.21(a): y1,y2-Density triangle for Figure 8.21(b): p,T-Projection for


C2H6 (1)+ n-C4H10 (2) with C2H6 (1) + n-C4H10 (2)
k12= 0.1000. with k12 = 0.1000.

Figure 8.23(a): y1,y2-Density triangle for Figure 8.23(b): p,T-Projection for


C2H6 (1)+ n-C4H10 (2) with C2H6 (1) + n-C4H10 (2)
k12= 0.1740. with k12 = 0.1740.
206 Chapter 8

8.3 Calculation of type VIII phase behaviour

8.3.1 Introduction

In section 8.2, we used the pure component parameters of ethane and n-butane for the
calculation of peculiar phase diagrams, including type XI (1Pll). In this section the SPHCT
equation of state is used for binary mixtures with pure component parameters representing
ammonia and a member of the homologous series CnF2n+2. The calculated phase behaviour,
including type VIII (1C2C), will probably not give an accurate description of the phase behaviour
of the mentioned binary systems, as explained in the next section. The only reason for choosing
these pure components, is to show that type VIII phase behaviour can be calculated with
physically realistic values for the pure fluids.

8.3.2 Results

The SPHCT equation of state parameters of NH3, CF4 and C2F6 were taken from Table
4.1. For the 'imaginary' carbon numbers between 1 and 2 of the CnF2n+2 series, the functions were
used:

c = 0.50774 . n + 1.5870 (8.6)

*
V = 9.6407 . n + 11.623 (8.7)

c .T * = 150.42 . n + 147.41 (8.8)

where n represents the carbon number. With Eq.(8.6)-(8.8) interpolations can be made between
the pure component parameters of CF4 and C2F6. In all calculations the value the binary
interaction coefficient, k12, was set equal to zero. In general, this assumption for k12 is often used
for the prediction of phase behaviour if both components of the binary system are chemically and
physically rather similar, see section 4.7. In this case a positive value for k12 is more realistic. For
NH3, the ε-value that was optimized on saturated vapour pressure data, contains a contribution
for the hydrogen bond formation. For the calculation of ε12, the contribution of hydrogen bond
formation can be ignored as a NH3-molecule and a CnF2n+2-molecule cannot form a hydrogen
bond. To correct this, the value of ε for NH3 should be split according to Eq.(8.9).

ε NH = ε disperse
3 NH 3
+ ε hydrogen
NH
bond
3
(8.9)

For the calculation of ε12, Eq.(8.10) should be used.

ε 12 = εC F
n 2n+2
. ε disperse
NH 3 (8.10)
Special phenomena 207

Because this separation into a disperse part and a hydrogen bonding part is hard to make another
kind of correction is needed. In practice, this is done by assigning a positive value to k12.
However, in this investigation the value for k12 was set to zero. This means that the predicted
phase behaviour will not give an accurate representation of the physical reality. While the aim of
our research is to explore the various types of phase behaviour, this deviance from the physical
reality is of minor importance.

Figure 8.24: Calculated p,T-projection of Figure 8.25: P,T-Projection of the calcu-


the critical curves of the lated high-pressure critical
binary system C2F6 (1) + curve of the binary mixture
NH3 (2) with k12 = 0.0. CnF2n+2 (1) + NH3 (2), with
n = 1.2780 and k12 = 0.0.
The investigations were started with the parameters of C2F6 and NH3, e.g., the carbon number n
of the perfluoro-alkane series has the value 2. Calculation of the critical curve with these pure
component parameters and with k12 = 0.0 leads to a continuous vapour-liquid critical curve in the
binary system, see Fig.8.24. Whether this phase diagram belongs to type I or type II behaviour in
the classification of Van Konynenburg [7,8], is hard to establish. From a numerical point of
view, it is difficult to calculate possible liquid-liquid critical curves close to 0 K, because of the
reciprocal temperature term in the exponential function of the attractive parameter in the SPHCT
equation.
If the value of n is decreased to n = 1.2780 this continuous vapour-liquid critical curve
changes quantitatively, i.e. the binary mixture still shows type Ih (1Pu) or IIh (1Pul) phase
behaviour. However, in the p,T-projection of the critical curves of this binary system an isolated
critical point appears at T = 226.35 K and p ≈ -124 bar. In addition to this isolated critical point at
n = 1.2780, a third critical phenomenon was discovered in this binary system: a high-pressure
critical curve that is situated completely above p = 10,000 bar, see also Fig.8.25. Although this
208 Chapter 8

high-pressure critical curve was not found for n = 2, it might be possible that even at higher
pressures or above the "pressure-ceiling" such a critical phenomenon occurs, see section 5.2.6.
The composition of this high-pressure critical curve remains constant at xNH3 ≈ 0.7. If the
decrease of the carbon number is continued, a closed critical curve will develop from the
previous mentioned isolated critical point, thus leading to type VIh (1Pnu) phase behaviour. The
maximum of this closed critical curve in the p,T-projection is called a hypercritical point. This
has also been discussed in sections 8.1 and 8.2.
As the value of the carbon number is decreased from n = 1.2780 to n = 1.15 a number of
observations can be made from Figs.8.26(a)-(d). The enclosed area of the closed critical curve is
enlarged. This can be seen from the T,x- and V,x-projections of the critical curves of this binary
mixture. In the p,T-projections it is shown that the closed critical curve is moving towards higher
pressures, while the high-pressure critical curve is moving towards lower pressures. At about
n = 1.11, a tricritical point appears on the continuous vapour-liquid critical curve, thus leading to
a transition from type VIh to type VIIh phase behaviour (1Pnu→2Pnu). The connectivity of the
critical curves does not change because of the appearance of the tricritical point. The critical
curve, on which the tricritical point has appeared, now contains a part that is locally unstable, see
section 8.1.

Figure 8.26(a): p,T-Projection of the calcu- Figure 8.26(b): p,T-Projection of the calcu-
lated high-pressure critical lated critical curves of the
curve of the binary system binary system CnF2n+2 (1) +
CnF2n+2 (1) + NH3 (2), with NH3 (2), with k12 = 0.0.
k12 = 0.0.
Special phenomena 209

Figure 8.26(c): T,x-Projection of the calcu- Figure 8.26(d): V,x-Projection of the calcu-
lated critical curves of the lated critical curves of the
binary system CnF2n+2 (1) + binary system CnF2n+2 (1) +
NH3 (2), with k12 = 0.0. NH3 (2), with k12 = 0.0.

Between carbon number values of n = 1.11 and n = 1.10, the vapour-liquid critical curve
and the closed critical curve merge to form one critical curve. This can be seen from the T,x-
projection of the critical curves of the binary mixture with n = 1.10, see Fig.8.27. The
connectivity of the critical curves has changed via a mathematical double point of the first kind.
Because this mathematical double point is unstable no changes will occur in the type of phase
behaviour according to the classification of Van Konynenburg and Scott [7,8]. After the
mathematical double point has formed, i.e., at a slightly lower value of n, see section 8.1, a
transition takes place from type VIIh to type Vh behaviour (2Pnu→2Pu). It is similar to the transi-
tion between type VII2h and type X (3Pnu→3Pu), that was discussed in section 7.3. The transition
state is formed by a double critical endpoint.
Further decreasing the value of the carbon number to n = 1.08, leads to an enlargement of
the gap in the 'newly formed' vapour-liquid critical curve. In the p,T-projection of the critical
curves it is shown that the minimum of the high pressure critical curve and the maximum of the
'new' vapour-liquid critical curve approach each other. This can be seen from Fig.8.28(a) and (b).
The collision of the critical curves takes place at a carbon number of n ≈ 1.07. In this case, unlike
the previous case, the critical curves do not merge but they exchange part of their branches. In
the p,T-diagram this can be seen as two branches of a hyperbole touching each other. At the
transition state they become identical with their asymptotes, and after the transition state two new
hyperbolic branches appear. We have called this transition state a mathematical double point of
the second kind, see section 5.2.3. The remarkable T,x-projection of the critical curves after the
transition at n = 1.07 is shown in Fig.8.29, where n has the value 1.05, see also section 8.1.
210 Chapter 8

Figure 8.27: T,x-Projection of the calcula-


ted critical curves of the
binary system CnF2n+2(1) +
NH3(2), with k12=0.0.

Figure 8.28(a): p,T-Projection of the high- Figure 8.28(b): p,T-Projection of the calcul-
pressure critical curve of the ated critical curves of the
binary system CnF2n+2(1) + binary system CnF2n+2(1) +
NH3(2) with k12=0.0. NH3(2) with k12=0.0.
Special phenomena 211

At a carbon number n=1.0, i.e., we now


have used the parameters of CF4 and NH3, the
topology of the critical curves of type VIII phase
behaviour is very well demonstrated, see
Fig.8.30(a) and (b). This type of phase
behaviour has never been observed
experimentally nor calculated with a pressure
explicit equation of state, although Schneider [9]
foresaw its possible existence. The three-phase
curve of this binary system has been calculated
with a computer program that was written by
Gallagher et al.[10]. The critical endpoints of
the three phase curve calculated with this
program, were found to be exactly on the critical
curves that were calculated with our own
software. It is shown in Fig.8.30(b) that the two
Figure 8.29: T,x-Projection of the calcula-
critical curves that go to high pressures diverge.
ted critical curves of the
One goes towards lower temperatures while the
binary system CnF2n+2 (1) +
other one is a 'gas-gas'-immiscibility critical
NH3 (2), with k12 = 0.0.
curve of the second kind. When the value of the
carbon number is decreased below n=1.0, no qualitative changes in phase behaviour take
place.

Figure 8.30(a): p,T-Projection of the cal- Figure 8.30(b): A zoomed-out view of


culated critical curves of the Fig. 8.30 (a).
binary system CF4(1) +
NH3(2) with k12=0.0.
212 Chapter 8

8.3.3 Discussion

The calculated transitions from type Ih (or IIh) via type VIh, type VIIh and type Vh to type
VIII (1Pu (1Pul) → 1Pnu → 2Pnu → 2Pu → 1C2C) are smooth and logical. It is not possible to
locate the previously mentioned mixtures in one global phase diagram because a change in the
carbon number n causes changes in ζ, λ, ξ and χ at the same time. To get an idea of where these
mixtures are approximately situated we calculated the global phase diagram parameters for
n = 1.15:

ζ = -0.72
λ = 0.31
ξ = 0.32
χ = -0.09

This is approximately the same region where we located type VIII behaviour in section 7.3.

The question arises as to why type VIh phase behaviour can be predicted with the SPHCT
and the Ree equation of state [1] but not with other equations of state, like the Van der Waals
[7,8], the Redlich-Kwong [11] and the Carnahan-Starling-Redlich-Kwong equation of state
[5,12]. For the SPHCT equation of state, the answer probably lies in the temperature dependence
of the attractive term. Unlike most other equations of state, the SPHCT equation of state has an
exponential temperature dependence. This, combined with a suitable choice for the values of c, ε
and σ, may possibly lead to type VIh phase behaviour. Moreover, the mixing rules in the SPHCT
equation also differ from the other equations mentioned above. This will have an influence on
the phase behaviour that is calculated from the SPHCT equation. Further research has to be done
to pin-point the mathematical origin of the closed critical curve phenomenon.

Experimentally, type VI behaviour occurs in systems with one polar and one less polar
compound, e.g., water-2-butoxyethanol [9]. The explanation for this behaviour, from the physical
point of view, is the possible formation of hydrogen bridges, an effect which is not explicitly
taken into account in the formulation of the SPHCT equation of state. However, when optimizing
pure component parameters from saturated vapour pressure data, very high values for ε are found
for NH3. This high value for ε may be the numerical translation of the physical phenomenon of
hydrogen bridge formation. By setting k12=0.0, which is unrealistic for these binary systems, the
value for ε12 contains a contribution for hydrogen bridge formation. Although this phenomenon
probably occurs in the binary system CnF2n+2-NH3 between the NH3-molecules, it will certainly
not occur between a NH3-molecule and a CnF2n+2-molecule. Although no critical data were
available for CnF2n+2-NH3-systems, type III phase behaviour will be most likely.
Special phenomena 213

8.4 The Van Laar point in the T,x-plane

8.4.1 Introduction

In this section we will focus on the T,x-projections of the critical curves of binary
mixtures in the neighbourhood of the Van Laar point in the global phase diagram. The Van Laar
point is defined by Meijer et al.[3] as the point in the global phase diagram where the tricritical
curve (TCP-curve), the double critical endpoint curve (DCEP-curve) and the mathematical
double point curve (MDP-curve) intersect. In this section we will only encounter mathematical
double points of the first kind, see section 5.2.3. From a topological point of view the Van Laar
point is important. The various types of phase behaviour II, III, IV and IV* (1Pl, 1C1Z, 2Pl and
2C1Z) are all situated around the Van Laar point in the global phase diagram, see Fig.8.31. It is
remarkable that none of the studies done so far shows the dependence of the critical temperature
on the composition of the critical curves around the Van Laar point. If one wants to locate the
Van Laar point or type IV* (2C1Z) experimentally, it is necessary to have knowledge about the
T,x-behaviour of the critical curves.

8.4.2 The neighbourhood of the Van Laar point

The global phase diagram around the Van Laar point (VL) is shown in Fig.8.31. TCP
represents the tricritical curve, MDP the mathematical double point curve (of the first kind),
DCEP is the double critical endpoint curve and VL is the van Laar point. The label gm on the
dashed curve stands for geometrical mean condition. ζ and λ are defined by Eqs.(5.1) and (5.2). It
should be noted that in real global phase diagrams the type IV* region has no substantial width.
The size ratio ξ and the flexibility ratio χ are arbitrarily fixed at a value of 0.1. All calculations
are local rather than global. Hence one cannot distinguish between stable and metastable
intervals: in the phase diagrams only the limit of stability is determined, i.e., G4x = 0.
Consequently we did not calculate (double) critical endpoints. Fig.8.32(a) shows the calculated
TCP-curve and the MDP-curve, see also Fig.8.31. Both curves were calculated with the SPHCT
equation of state. Because these curves are close together they seem to coincide. In fact, detailed
calculations close to the Van Laar point show that they intersect in this case. Fig.8.32(b) shows a
schematic drawing of the situation around the Van Laar point. The marked points in the global
phase diagram indicate points at which the critical curves have been calculated with the SPHCT
equation. The Van Laar point and a number of special points were located in the global phase
diagram. These points are marked in Fig.8.32(b). The values of ξ and χ are constant in each point
of the global phase diagram. We recall that the parameters of the reference component, i.e.
component 2, are chosen. By giving ξ and χ a certain value, - 0.1 in our case -, the values of c1
and V1* are defined as well. The equation of state parameters that vary in the global phase
diagram are T1* and k12.
214 Chapter 8

Figure 8.31: The regions II, III, IV and IV* in


the global phase diagram
around the van Laar point. This
schematic diagram is not drawn
to scale.

Figure 8.32: (a) Part of the global phase dia- Figure 8.32: (b) Schematic drawing of the
gram, calculated from the global phase diagram around
SPHCT equation of state. For the Van Laar point.
an explanation of symbols, see
Fig.8.31.
Special phenomena 215

The pure component parameters at the Van Laar point are:


c1 = 1.2222, c2 = 1.0000
V1 = 24.444 cm .mole , V2* = 20.000 cm3.mole-1
* 3 -1

T1* = 309.392 K, T2* = 239.998 K

These values of the pure component parameters of the SPHCT equation are fictitious. However,
their values are chosen so that they are physically realistic. To obtain a Van Laar point in the
above mentioned binary mixture we used k12 = 0.211532. In the Van der Waals case, according
to Meijer et al.[13,14], the value for k12 is exactly 0.0 at the Van Laar point. Our value of k12 is
rather high for a real binary system, although not impossible. At point 2 of the global phase
diagram the binary mixture shows type II (1Pl) phase behaviour. The pure component parameters
of both components at point 2 of the global phase diagram are the same as they are at point 1
since ζ has not changed. However, the value of λ decreases if we move from point 1 to point 2.
This means that the value of k12 in the binary system belonging to point 2 is lower than the value
of k12 at point 1. Vertical changes in the global phase diagram are always accompanied with
changes in the k12-value of the binary system while both pure component parameters of this
binary system remain the same.
From point 2 in the global phase diagram we move to point 3 in Fig.8.32(b). Point 3 is
situated on the TCP-curve in the global phase diagram. With this move we not only change the
value λ, but we also change the value of ζ. The binary mixture belonging to point 3 in the global
phase diagram has pure component parameters that are different from the ones at points 1 and 2.
In our case the equation of state parameters of component 2 are fixed. According to Eq.(5.1) the
value of ε11 must change. This change in the value of ε11 leads to T1* = 342.634 K. The values of
c1 and V1* are the same at every point of the global phase diagram since the values of ξ and χ are
fixed. Horizontal changes in the global phase diagram are always accompanied by a change in
one of the pure component parameters in the binary mixture. During this horizontal move in the
global phase diagram the value of k12 in the (changing) binary mixture will change as well. From
point 3 we move vertically in the global phase diagram to point 4 in Fig.8.32(b). As for all other
vertical changes, only the value of k12 in the binary mixture will change during this move. The
binary mixture in point 4 in the global phase diagram shows type IV (2Pl) behaviour. From point
4 we move vertically in the global phase diagram to point 5 in Fig.8.32(b). Point 5 is situated on
the MDP-curve. Because a MDP of the first kind in a binary system can never be stable (except
in the Van Laar point) the binary mixture will also show type IV (2Pl) phase behaviour.
If we move vertically upwards in the global phase diagram from the Van Laar point to
point 6 we have entered the type III (1C1Z) region. By moving to the left in the global phase
diagram we have to change the value of ε11, and as a consequence the value of T1*. At point 7 in
the global phase diagram we meet the TCP-curve again. At point 7, T1* has the value
T1*=290.354 K. Of course the value of k12 in the binary mixture belonging to point 7 in the global
phase diagram differs from the one in point 6. Decreasing the value of λ in the global phase
diagram, and accordingly the value of k12 in the binary mixture, brings us to point 8 in the global
phase diagram. The binary mixture that corresponds with point 8 in the global phase diagram
216 Chapter 8

shows type IV* (2C1Z) phase behaviour. By a further reduction in the value of λ we reach point 9
in the global phase diagram, see Fig.8.32(b). Point 9 is situated on the MDP-curve. Because of
the unstable character of the MDP of the first kind, the binary system corresponding to point 9 in
the global phase diagram shows type IV* (2C1Z) phase behaviour.
The coordinates of the points 1 to 9 are given in Table 8.2. The p,V,T,x-coordinates of the
special critical phenomena that occur in the binary systems belonging to these points are
summarized in Table 8.3.

Table 8.2: Coordinates of the global phase diagram variables of the special
critical phenomena at the marked points in Fig.8.32(b). The size
and flexibility ratios, ξ and χ, are both 0.1

No. Type ζ λ k12

1 VL 0.082699 0.214233 0.211532


2 II 0.082699 0.2141 0.2114
3 TCP 0.13312 0.17395846548 0.16654061625
4 IV 0.13312 0.17395850 0.16654065
5 MDP 0.13312 0.1739585242 0.1665406755
6 III 0.082699 0.2144 0.2117
7 TCP 0.05109 0.2389628769 0.2379677033
8 IV* 0.05109 0.2389628756 0.2379677020
9 MDP 0.05109 0.23896287499 0.23796770135

Table 8.3: Coordinates in p,V,T,x-space of the special critical phenomena at the


marked points in Fig.8.32(b).

No. Type p T x V
bar K - cm .mole-1
3

1 VL 80.287 323.109 0.5566 98.86


3 TCP 90.18 340.16 0.61464 92.64
5 MDP 90.17 340.149 0.61324 92.37
7 TCP 74.58 311.15 0.52336 102.57
9 MDP 74.58 311.148 0.52375 102.66
Special phenomena 217

8.4.3 Results

The T,x-projection of the critical


curves of the binary system with a Van Laar
point, point 1 in Fig.8.32(b), is shown in
Fig.8.33. In this T,x-projection, four critical
curves can be seen. These four critical curves
come together in the Van Laar point, marked
by an open box. In Fig.8.33 and also in the
following figures, the solid curves represent
the (meta)stable critical curves, the dotted
curves represent the locally unstable critical
curve. The open box represents the van Laar
point. A tricritical point is represented by an
open circle, a mathematical double point of
the first kind by a solid box and points where
the stability of the critical curves change from
unstable to metastable by an open triangle.
A temperature-maximum is represented by a
triangle up and a temperature-minimum by a Figure 8.33: T,x-Projection of the critical
triangle down. In Fig.8.33, one can also say curves of the binary system at
that two critical curves are present. In this point 1 in the global phase
respect both critical curves consist of two diagram of Fig.8.32(b). The
branches. One branch originates at the critical solid curves represent the
point of the pure component with the highest (meta)stable critical curves, the
critical temperature, i.e., component 1 at x=0. dotted curve the locally unstable
This branch of the critical curve is marked critical curve. The open box
with the letter a, see Fig.8.33 and is locally represents the van Laar point.
stable, i.e., G4x≥0 for every critical point on
this branch. The second branch originates in the critical point of the pure component with the
lowest critical temperature, i.e., component 2 at x=1. This branch is marked with the letter b, see
Fig.8.33 and is also locally stable. The third branch, marked with the letter c, comes from a point
at extremely high pressures. This point is often referred to as the jamming point. This branch is
locally stable as well. The fourth branch, marked with the letter d, is locally unstable, i.e., G4x≤0
for every critical point on this branch. Branch d is expected to originate at the point (p,T)=(0,0),
see also section 8.2. By moving around in the global phase diagram, as discussed above, the Van
Laar point in the binary mixture will break open and the connectivity of these branches of the
critical curves changes accordingly.
218 Chapter 8

Figure 8.34: T,x-Projection of the critical Figure 8.35: T,x-Projection of the critical
curves of the binary system at curves of the binary system at
point 2 in the global phase point 3 in the global phase
diagram of Fig.8.32(b). diagram of Fig.8.32(b).

Figure 8.36: (a) T,x-Projection of the type Figure 8.36: (b) Enlargement of the re-
IV critical curves of the gion around the S-shaped
binary system at point 4 in loop in one of the critical
the global phase diagram of curves in Fig.8.36(a).
Fig.8.32(b).
Special phenomena 219

By definition, changing the value of k12 in a binary mixture leads to a move along a
vertical line in the global phase diagram. By decreasing the value of k12 we move away from the
Van Laar point in the type II (1Pl) region, see point 2 in Fig.8.32(b) and Fig.8.34. As can be seen
from Fig.8.34 a continuous vapour-liquid critical curve between the critical points of the pure
components remains. The branches a and b are now connected. Additionally a liquid-liquid
critical curve exists at lower temperatures. This critical curve stems from the branches c and d.
Coming from extremely high pressures this critical curve is stable. This part of the critical curve
is again marked c. On this part of the critical curve we meet the point where metastability starts.
This point is a point of global instability and thus not visible in Fig.8.34, see also section 4.7. The
(metastable) critical curve then reaches a point of instability. This point is marked in Fig.8.34 by
an open triangle. The unstable part of the critical locus, branch d, is shown by the dotted curve.
By changing the characteristic temperature of component 1 to T1* = 342.634 K we move
along a horizontal line to the right in the global phase diagram. If we assume a suitable value for
k12, we create a TCP in our system, point 3 in Fig.8.32(b). The resulting T,x-projection is shown
in Fig.8.35. Basically this is a type II (1Pl) diagram with a TCP on the continuous vapour-liquid
critical locus. This TCP is a horizontal point of inflection on the critical curve in the T,x-
projection and it is marked by an open circle in Fig.8.35. Further increasing the value of k12 leads
to type IV (2Pl) phase behaviour (point 4 in Fig.8.32(b)), see Figs.8.36(a) and 8.36(b). The well-
known S-shaped loop appears in the critical curve. Again the unstable parts of the critical loci are
represented by dotted curves.

Figure 8.37(a): T,x-Projection of the critical Figure 8.37(b): Magnification of the region
curves of the binary system at around the mathematical
point 5 in the global phase dia- double point in Fig.8.37(a).
gram of Fig.8.32(b).
220 Chapter 8

Increasing the value of k12 a little brings us to a MDP of the first kind, point 5 in
Fig.8.32(b), see Figs.8.37(a) and 8.37(b). At this point in the global phase diagram, a normal type
IV (2Pl) phase diagram is found experimentally. The MDP of the first kind is always a transition
state between metastability and instability. So one can say it is at least metastable and thus it
cannot be found experimentally. The only exception is when the MDP coincides with the Van
Laar point. In that case it is the limit of stability and metastabilty at the same time, see also
section 5.2.3. From the T,x-projection of the critical curves, see Figs.8.37(a) and 8.37(b), it can
be seen that the minimum of the continuous vapour-liquid critical curve a-b coincides with the
maximum of the critical curve c-d that comes from infinite pressures and is expected to go to
( p,T ) = ( 0,0 ).
A further increase of the value of k12 leads to type III (1C1Z) phase behaviour. In the
MDP the connectivity of the critical branches will be changed. Branch a will be connected with
branch c, and branch b will be connected with branch d.

Figure 8.38: T,x-Projection of the type III Figure 8.39: T,x-Projection of the critical
critical curves of the binary curves of the binary system at
system at point 6 in the global point 7 in the global phase
phase diagram of Fig.8.32(b). diagram of Fig.8.32(b).

Turning back to the Van Laar point in the global phase diagram, point 1 in Fig.8.32(b),
an increase in the value of k12 leads to type III (1C1Z) behaviour, point 6 in Fig.8.32(b), see
Fig.8.38. The connectivity of the critical curves with respect to type II (1Pl) has changed. Coming
from infinite high pressures, the stable critical curve a-c ends in the critical point of the
component with the highest critical temperature. The critical curve that comes from the critical
point of the component with the lowest critical temperature, branch b, now undergoes the
changes from stable via metastable to unstable, i.e., branch d.
Special phenomena 221

By changing the characteristic temperature of component 1 to T1* = 290.354 K we move


to the left in the global phase diagram. At a suitable value of k12 we generate a TCP in the binary
system again (point 7 in Fig.8.32(b)). Basically we have type III (1C1Z) phase behaviour with a
TCP located on the a-c critical curve that comes from infinite pressures, see Fig.8.39. Again the
TCP is a horizontal point of inflection on the critical curve. If we stay close to the Van Laar
point, the TCP-curve and the MDP-curve are almost point-symmetric with respect to the Van
Laar point in the global phase diagram. So, by reducing the value of k12 (point 8 in Fig.8.32(b))
an S-shaped loop will develop in the liquid-liquid critical curve, see Figs.8.40(a) and 8.40(b).

Figure 8.40(a): T,x-Projection of the type IV* Figure 8.40(b): Enlargement of the region
critical curves of the binary around the S-shaped loop in
system at point 8 in the global one of the critical curves in
phase diagram of Fig. 8.32(b). Fig.8.40(a)

This type of phase behaviour, which only exists in a very narrow region of the global phase dia-
gram, is called type IV* (2C1Z) behaviour. On a further decrease of the value of k12 the minimum
of the S-shaped loop in the critical curve coincides with the maximum of the second critical
curve. We have reached the MDP-curve in the global phase diagram, point 9 in Fig.8.32(b). At
this MDP the connectivity of the critical branches will again be changed, see Figs.8.41(a) and
8.41(b). A further reduction of the value of k12 leads to type II (1Pl) phase behaviour. In this case
branch a is connected with branch b and branch c is connected with branch d, as discussed earlier
in this section. With this MDP we also reach the end of our investigations around the Van Laar
point in the global phase diagram. Based on all the calculations of critical curves in binary
sytems we are able to make a transition scheme. Fig.8.42 shows schematically the T,x-
projections of the critical curves and three-phase lines L1L2V of the various types of phase
behaviour and their continuous transition states. These types of phase behaviour occur in binary
systems that are located around the Van Laar point in the global phase diagram.
222 Chapter 8

Figure 8.41(a): T,x-Projection of the critical Figure 8.41(b): Enlargement of the region
curves of the binary system at around the mathematical
point 9 in the global phase dia- double point in Fig. 8.41(a).
gram of Fig.8.32(b).

Figure 8.42: Scheme of the four different types of phase behaviour in binary systems and their
transition states in the neighbourhood of the van Laar point. All diagrams are
T,x-projections of the critical curves and three-phase lines. The solid curves
represent the stable parts of the critical curves, the dotted curves represent the
locally unstable parts of the critical curves, the dashed curves represent the
three-phase lines.
1. system with a van Laar point (VL).
2. type II system.
3. type III system.
4. type IV system.
5. type IV* system.
6, 7. system with a tricritical point (T).
8, 9. system with a mathematical double point (MDP).
10, 11. system with a double critical endpoint (DCEP).
Special phenomena 223
224 Chapter 8

8.4.4 Conclusions

It is shown that the SPHCT equation of state describes qualitatively the same phase
behaviour around the Van Laar point as the Van der Waals [7,8] and the Redlich-Kwong
equations of state [11]. A continuous transition is possible from a binary system with a Van Laar
point into binary systems that show type II, III, IV and IV* (1Pl, 1C1Z, 2Pl and 2C1Z) phase
behaviour.
The region in the global phase diagram in which type IV* (2C1Z) phase behaviour is
located is found to be extremely small. To move around in this region in the neighbourhood of
the Van Laar point we need the k12-value to be known to at least 9 digits. The T,x-projections of
the critical curves of the binary mixtures show that the temperature range of the three phase loop
in the critical curve is very narrow (at maximum 0.01 K). The temperature range of the three-
phase line is expected to be even smaller. This makes an experimental verification of type IV*
almost impossible. Changing the pure component parameters of the two components does not
seem to offer a solution. We either move towards the Van Laar point or away from the Van Laar
point along the horizontal axes in the global phase diagram. If we move towards the Van Laar
point, the IV* (2C1Z) region becomes smaller and disappears at the Van Laar point. In the T,x-
projection the temperature range of the high temperature part of the three-phase line will shrink
and disappear at the Van Laar point. If we move away from the Van Laar point we will have
lower ζ-values and higher λ-values. In practice this means that the components become more
similar. In the limiting case (ζ = 0) both components have the same potential well depth. At the
same time the value of k12 needs to increase in accordance with the increasing λ-value. This
increase in k12 is opposite from what one expects when the components similarity increases.
With real molecules the value of k12 exceeds 0.20 only in exceptional cases.
The above discussion of the experimental verification of type IV* (2C1Z) behaviour has
three weak points. The first point is that this study did not take into account the double critical
endpoint phenomenon. The DCEP-curve is the lower boundary of the type IV* (2C1Z) region to
the left of the Van Laar point in the global phase diagram. In analogy with the studies on the Van
der Waals [7,8] and Redlich-Kwong equations of state [11] it is to be expected that this DCEP-
curve is situated extremely close to the TCP-curve in the global phase diagram.
The second point is that this study only covers the fixed size and flexibility ratios
ξ = χ = 0.1. Preliminary studies at different size and flexibility ratios show that the IV* (2C1Z)
region in the global phase diagram will not widen significantly. However, by changing the size
and flexibility ratios of the different components, the Van Laar point and the IV* (2C1Z) region
can be moved towards more physically realistic regions in the global phase diagram, i.e. binary
systems with smaller k12-values. With the RK and CSRK equation larger IV* regions have been
found in other parts of the global phase diagram.
The third point concerns the reliability of the SPHCT equation of state. It might be
possible that the 'perfect' equation of state that matches all experimental data points ever
measured, would show a larger type IV* (2C1Z) region. We think that the SPHCT cannot be
identified as this 'perfect' equation of state.
Special phenomena 225

NOMENCLATURE

ci flexibility parameter, i.e., one-third of the external degrees of freedom of one molecule of
component i
G molar Gibbs energy
Gnx n-th derivative of G with respect to x
k12 binary interaction coefficient
p pressure
S molar entropy
Snx n-th derivative of S with respect to x
T temperature
Ti* characteristic temperature of component i
V molar volume
Vnx n-th derivative of V with respect to x
Vi* closest-packed volume of component i
x mole fraction of component 2
yi reduced density fraction of component i

Greek symbols

Δ differential
ε potential well-depth
ζ dimensionless parameter ratio, see Eq.(5.1)
λ dimensionless parameter ratio, see Eq.(5.2)
ξ dimensionless parameter ratio, see Eq.(5.3)
χ dimensionless parameter ratio, see Eq.(5.4)

Subscripts and superscripts

c critical
icp isolated critical point
226 Chapter 8

REFERENCES

1. L.Z. Boshkov. On the description of phase diagrams of two component mixtures with a
closed domain of demixing based on the one-fluid model of an equation of state. Dokl.
Akad. Nauk. SSSR. 294, (1987), 901-905
2. P.H.E. Meijer. Private communications. Delft, (1992).
3. P.H.E. Meijer and M. Napiorkowski. The three-state lattice gas as model for binary gas-
liquid systems. J. Chem. Phys. 86, (1987), 5771-5777.
4. P.H.E. Meijer, M. Keskin and I.L. Pegg. Critical lines for a generalized three state binary
gas-liquid lattice model. J. Chem. Phys. 88, (1988), 1976-1982. Erratum: J. Chem. Phys. 90,
(1989), 3408.
5. Th. Kraska and U.K. Deiters. Systematic investigation of the phase behavior in binary fluid
mixtures. II. Calculations based on the Carnahan-Starling-Redlich-Kwong equation of state.
J. Chem. Phys. 96, (1992), 539-547.
6. A. Bolz. Vergleichende Untersuchung globaler Phasendiagramme. PhD-thesis, Ruhr-
Universität Bochum, Germany, (1992).
7. P.H. van Konynenburg and R.L. Scott. Critical lines and phase equilibria in binary van der
Waals mixtures. Phil. Trans. 298A, (1980), 495-540.
8. P.H. van Konynenburg. Critical lines and phase equilibria in binary mixtures. PhD-thesis
University of California Los Angeles. (1968).
9. G.M. Schneider. Phasengleichgewichte in flüssigen Systemen bei hohen Drucken. Ber.
Bunsenges. Phys. Chem. 70, (1966), 497-612.
10. J.S. Gallagher, C.J. Peters and J.M.H. Levelt-Sengers. Private communications.
11. U.K. Deiters and I.L. Pegg. Systematic investigation of the phase behavior in binary fluid
mixtures. I. Calculations based on the Redlich-Kwong equation of state. J. Chem. Phys. 90,
(1989), 6632-6641.
12. Th. Kraska. Systematische Untersuchung des globalen Phasenverhaltens binärer fluider
Mischungen ausgehend von der Carnahan-Starling-Redlich-Kwong-Zustandsgleichung.
Diploma thesis, Ruhr-Universität Bochum, Germany, (1989).
13. P.H.E. Meijer. The van der Waals equation of state around the van Laar point. J. Chem.
Phys. 90, (1989), 448-456.
14. P.H.E. Meijer, I.L. Pegg, J. Aronson and M. Keskin. The critical lines of the van der Waals
equation for binary mixtures around the van Laar point. Fluid Phase Equilibria 58, (1990),
65-80.
227

CHAPTER 9
CONCLUSIONS AND PROSPECT

The objective of this research was to investigate the global phase diagrams that can be
calculated from an advanced equation of state suitable for chainlike molecules. By doing this, we
hoped to find new types of fluid phase behaviour. We have chosen the equation that has been
derived from the Simplified Perturbed Hard Chain Theory: the SPHCT equation. A study of
global phase diagrams and the related study of critical curves in binary mixtures only makes
sense if this equation of state can be used in practice. Thus another question that must be
resolved is: Can this equation be used to correlate experimental data?
In chapter 4 we have demonstrated that this is the case. A number of thermodynamic
properties has been calculated from the SPHCT equation of state and compared with
experimental data. It was shown that this equation fulfills a number of thermodynamic conditions
and gives at least a reasonable description of certain thermodynamic properties. However, we
must add that testing the equation was not the main objective of our research.
An equation of state is very suitable for the description of critical phenomena. The
conditions for these critical phenomena can be applied to an equation of state which leads to a
number of equations to be solved for a number of variables, e.g, volume, composition. Extra
attention in this respect has been paid to mathematical double points in chapter 5. For these
points, the conditions had never been derived before in the natural variables of a pressure-explicit
equation of state, i.e., volume, temperature and composition. We also have shown that
mathematical double points can be stable and form the transition state between two different
types of phase behaviour.
Unlike other studies of global phase behaviour, we have used analytical expressions for
the derivatives of the Helmholtz energy. The use of the symbolic algebra package MAPLE saved
a tremendous amount of tedious algebra and reduced the chance of errors being made during the
derivation of mathematical expressions.
From the calculated global phase diagrams it can be concluded that the SPHCT equation
behaves differently to other equations of state that have so far been investigated. This is due
mainly to two phenomena that can occur in the pressure-temperature projection of binary
mixtures. These are the possibilities of calculating liquid-liquid immiscibility loops and high-
pressure critical curves. The critical curve that encloses the liquid-liquid immiscibility loops is
also referred to as a closed critical curve. In the T,x-plane this critical curve is closed, although
part of this closed critical curve is unstable. Liquid-liquid immiscibility loops were earlier
calculated by Boshkov, although his findings were never confirmed by other investigators. We
were not able to confirm his results, although our investigations strongly support his conclusions
228 Chapter 9

about the location of the type VI/VII-region (1Pn/2Pn-region) in the global phase diagram, see
chapter 8. Our investigations confirmed the existence of type VII (2Pn) phase behaviour. In
chapter 7 it is shown that the interference of the high-pressure critical curve with the closed
critical and vapour-liquid critical curve has led to completely new types of phase behaviour, type
VIII-XI (1C2C-1Pll). Some of these types were foreseen or measured, but up till now they have
never been calculated with an equation of state. Also in the global phase diagrams deviations
were observed from global phase diagrams that have been calculated from other equations of
state, like the existence of a tetracritical point. Although a tetracritical point was also found for
the Carnahan-Starling-Redlich-Kwong equation (CSRK equation), we have encountered a new
region in the global phase diagram where a new type of phase behaviour occurs, type IX (1C3C).
As far as the nomenclature of fluid phase diagrams is concerned, new intelligent proposals have
been made. Whereas the current nomenclature of Van Konynenburg and Scott is quite arbitrary,
the proposed nomenclature of Bolz is very descriptive and simple. In our opinion, this new
nomenclature certainly is a good candidate to replace the nomenclature of Van Konynenburg and
Scott.
In addition to the discovery of the new types of phase behaviour, we have also
thoroughly investigated the region around the Van Laar point and we confirm the existence of
type IV* phase behaviour. According to the results obtained from the SPHCT equation, it will be
extremely difficult to make an experimental verification of this type IV* (2C1Z), see chapter 8.
This research has answered a number of questions, but also raised many others. One of
the most annoying things encountered, was the inability to calculate the double critical endpoint
curves in the global phase diagram from the SPHCT equation. This is the result of numerical
problems. Although we believe that the location of the tricritical curves in the global phase
diagram is more important, it would be interesting to study the behaviour of the double critical
endpoint curves, especially in the region of the tetracritical point. In the neighbourhood of the
tetracritical point, the tricritical curves are extremely hard to calculate. The details on this
transition state in the global phase diagram have not been as yet clarified. A final point we want
to bring up is the location of the curve in the global phase diagrams that contains the
mathematical double points of the second kind. Although we obtained some results, we do not as
yet know how this curve is situated in the global phase diagram. Therefore these results have
been omitted from this thesis.
To complete this particular investigation, further research on the global phase diagram
should be done to answer the above questions. Of course, many subjects can be added to this,
like the phase behaviour in the shield-region. The recommendations listed so far all refer to the
SPHCT equation of state. It would be interesting to compare our results with those that will be
obtained from other theoretical equations of state, e.g., the APACT equation that has been
derived to take polar components into account. With the results from the global phase diagram,
we do not only study the phase behaviour as calculated from a particular equation of state, but
we can also study the influence that mixing rules have on fluid phase behaviour. In this respect, it
may be interesting to incorporate new, promising mixing rules, e.g., the Wong-Sandler mixing
rule, in equations of state and study the deviations from other global phase diagrams.
Conclusions and prospect 229

In conclusion, we think that the research on the global phase behaviour from equations of
state will lead to a better understanding of the overall capabilities of equations of state.
Moreover, it will provide us with a deeper insight in the consequences of intermolecular
behaviour on the fluid phase behaviour of binary systems.
230

Summary
This thesis is mainly concerned with the global phase behaviour that is calculated with
the Simplified Perturbed Hard Chain equation. The results of our investigations have extended
our knowledge of the systematics of phase equilibria in binary fluid mixtures. Our investigations
give insight into the consequences of an assumed model for the intermolecular interactions on
the phase behaviour of a binary system. This leads to a deepening of our theoretical knowledge.
The knowledge that is obtained about the possibilities of an equation of state can be used for the
description of experimental data of thermodynamic properties. Moreover, for the development of
equipment that is used in the chemical process industry, the use of simulation software, the so-
called 'flowsheeting programs' is becoming more and more important. In addition to models for
the description of the kinetics of chemical reactions, models for the design of particular
equipment and economic models, thermodynamic models, among them the equations of state,
form a very important part of these flowsheeting programs. A profound knowledge of the phase
behaviour that is predicted with an equation of state, is indispensable for the correct use of this
simulation software.
In chapter 1 we entered into the details of the relation between the theory of equations of
state and their application in flowsheeting programs. Supercritical fluid extraction was discussed
as an example of a unit operation from chemical industry. This unit operation, which has recently
received a lot of attention in the field of the production and/or purification of fine chemicals and
food stuffs, is based mainly on the special thermodynamic properties of a near-critical phase. In
our research we specifically investigated the possibilities of an equation of state in the critical
region. This chapter was closed with an inventory of the theoretical research that has been done
in this area.
In chapter 2 an introduction was given to the various types of binary fluid phase
behaviour as they are currently known. This has been done according to the classification of Van
Konynenburg and Scott. We tried to give at least 1 example of experimentally determined phase
behaviour for each type.
In chapter 3 an introduction was given to equations of state in general and an extensive
derivation of the equation of state that results from the Simplified Perturbed Hard Chain Theory
was also provided. We called this equation the SPHCT equation. The SPHCT equation
distinguishes itself from many other equations of state by having a sound theoretical background.
For this reason, the repulsive and attractive term were extensively discussed. This chapter was
closed by a short description of the mixing rules that are used.
In chapter 4 the quality of the SPHCT-equation is discussed. It eventuated that although
this equation of state has a sound theoretical background, this does not automatically guarantee
correct correlations of thermodynamic properties. It is not possible, and it was also not the goal
Summary 231

of this PhD-thesis, to test all the 'ins and outs' of the SPHCT equation. We restricted ourselves to
the calculation of a number of thermodynamic properties, of which the behaviour in the critical
region sometimes received extra attention.
It turned out that it is possible to have the vapour pressure curve, that is calculated from
the SPHCT equation, end exactly in the critical point in the p,T-plane. The deviation from
experimental vapour pressure data in the reduced temperature range [0.75-1] is less than 3% for
99 of the 108 pure fluids that have been investigated. Unfortunately, the saturated liquid and
vapour volume are less well described. At the calculated critical point, the compressibility ratio
of most pure substances lies between 0.34 and 0.35, which corresponds to a deviation of about
20% in the critical volume. It must be said that the Peng-Robinson equation better describes the
vapour pressure curve in the p,T-plane for small and/or polar molecules. For chainlike molecules
with a carbon number of approximately 10 or more, the SPHCT-equation often performs better.
From this it can be concluded that the ideas that have been built into the SPHCT about chain
behaviour are indeed effective.
The temperature dependence of the attractive term of the SPHCT equation deviates from
most other equations of state currently in use, like the Peng-Robinson and the Soave-Redlich-
Kwong equation of state (the PR- and SRK-equation). However, it is striking that the attractive
terms of the PR- and SRK-equation do not obey the condition that at zero density and high
temperatures, the attractive term must become constant, while the SPHCT-equation obeys this
condition. Another condition that an equation of state must fulfill is a correct description of the
virial coefficients as a function of temperature. For methane, ethane and propane, the SPHCT-
equation indeed describes the temperature dependence of the second virial coefficient correctly.
The maximum in the third virial coefficient for methane and ethane is not correctly described,
neither by the SPHCT-equation nor the PR-equation. The last pure component property
investigated with the SPHCT equation is the molar isochoric heat capacity, because of the
diverging behaviour of this property near the critical point. As was expected from every classical
equation of state, like the SPHCT equation, this divergence in the critical region is not predicted.
It should be concluded that the SPHCT equation is in error here, not only quantitatively but also
qualitatively.
The SPHCT equation has also been tested for binary mixtures. For mixtures consisting of
simple molecules, like methane, ethane, argon and nitrogen, binary phase equilibria are well
described. The Henry coefficients, that have been calculated from the SPHCT equation for
substances that are commonly encountered in natural gas processing are also in good agreement
with experimental data. The final property of the SPHCT equation that has been tested is the
description of critical curves in binary fluid mixtures. For most hydrocarbon mixtures the
description of the vapour-liquid critical curves is good as long as the unlike molecules do not
differ too much in size.
In chapter 5 the mathematical conditions are described for the various curves and points
that occur in the global phase diagram, including the tricritical curves, the tetracritical point, the
double critical endpoint curves, the Van Laar point, the azeotropical boundary curves, the
azeotropical endpoint curves and the mathematical double point curves. For the calculation of a
232 Summary

mathematical double point the conditions have been derived in volume, temperature and
composition coordinates. For a pressure-explicit equation of state, like the SPHCT equation, this
has never been done before in this form. Moreover, we have shown that there are two kinds of
mathematical double points, stable and unstable mathematical double points. This chapter is
closed with a short consideration on the 'pressure-ceiling' that exists in the p,T-plane according to
the SPHCT equation. This pressure ceiling is physically not realistic and a consequence of a
shortcoming in the Carnahan-Starling expression, that has been used in the repulsive term of the
SPHCT equation.
In chapter 6 numerical algorithms are discussed that have been used for the calculation of
the global phase diagrams. The Marquardt algorithm, used to solve sets of nonlinear equations, is
discussed extensively. This chapter is closed with a short description of the application of the
symbolic computation program Maple. With Maple we have derived expressions for the
derivatives of the molar Gibbs energy with respect to the composition and of the molar
Helmholtz energy with respect to the molar volume and the composition. The latter derivatives
depend on the equation of state that has been used, the first do not.
In chapter 7 results are given of the calculations of global phase diagrams. It turns out
that the SPHCT equation in this respect can behave differently than other equations of state. The
appearance of the tetracritical point and consequently the new types of phase behaviour are
extensively discussed. They have been numbered according to the classification of Van
Konynenburg and Scott and it is proposed to call these types of phase behaviour type VIII, IX, X
and XI. Moreover we calculated type VI and VII phase behaviour and a new version of type VII
that we propose to call type VII2h.
In chapter 8, some interesting phenomena previously described in chapter 7 are more
closely investigated. Amongst these are the possibility to calculate closed critical curves from the
SPHCT equation and the socalled high-pressure critical curves from the SPHCT equation. In
particular the interference of these critical curves gives rise to some interesting phase diagrams.
We tried to find practical binary mixtures that exhibit the interference of the closed critical curve
with the high-pressure critical curve. It turned out to be possible to find pure substances that obey
the demands on the pure component parameters, like ethane+n-butane and ammonia+a short
perfluoroalkane. Unfortunately the chances of finding the special critical curves that have been
predicted in these mixtures are low. The real binary mixtures, that are composed of these
components, do not have the requested binary interaction coefficients. Another possibility is that
the phenomena occur in a temperature range where solid phases are stable.
The Van Laar point is a special point in the global phase diagram. At this point four
branches of two critical curves come together with a three phase line. The regions in the global
phase diagram, where the types II, III, IV en IV* occur, are arranged around the Van Laar point.
These types of phase behaviour are extensively discussed in this chapter.
Finally, chapter 9 gives general conclusions and an opinion on work to be undertaken in
the future.
233

Samenvatting
Dit proefschrift behandelt voornamelijk het globale fasengedrag van binaire systemen,
zoals dat met de Simplified Perturbed Hard Chain toestandsvergelijking is berekend. Hierdoor
hebben wij onze kennis omtrent de systematiek van fasenevenwichten in binaire systemen
verdiept. Dit onderzoek geeft een inzicht in de konsequenties van een model op molekulaire
schaal op het binaire fluïde fasengedrag, hetgeen leidt tot een verdieping van onze theoretische
kennis, maar er is meer. Immers, de kennis, die verkregen wordt omtrent de mogelijkheden van
een toestandsvergelijking, kan worden gebruikt bij de beschrijving van experimentele data van
thermodynamische grootheden. Daarnaast wordt bij de ontwikkeling van apparatuur die in de
chemische procesindustrie wordt toegepast, meer en meer gebruik gemaakt van simulatiepro-
grammatuur, de zgn. 'flowsheeting programs'. Naast o.a. modellen ter beschrijving van de
kinetiek van chemische reakties, modellen voor de dimensionering van specifieke apparatuur en
kostenmodellen, vormen thermodynamische modellen, waaronder toestandsvergelijkingen, een
zeer belangrijk onderdeel van deze flowsheeting programs. Een grondige kennis van het
fasengedrag dat door een toestandsvergelijking wordt voorspeld, is onontbeerlijk voor een juiste
bediening van deze simulatieprogrammatuur.
In hoofdstuk 1 wordt nader ingegaan op de relatie tussen de theorie van de
toestandsvergelijkingen en hun toepassing in flowsheeting programs. Als voorbeeld van een 'unit
operation' wordt superkritische extraktie aangehaald. Dit proces, dat de laatste jaren meer en
meer naamsbekendheid heeft gekregen voor de produktie en/of zuivering van o.a.
fijnchemicaliën en voedingsmiddelen, berust grotendeels op de bijzondere thermodynamische
eigenschappen van een nabij-kritische fase. In dit onderzoek wordt specifiek gekeken naar de
mogelijkheden van een toestandsvergelijking in het kritisch gebied. Dit hoofdstuk wordt
afgesloten met een inventarisatie van theoretisch onderzoek, dat reeds op dit gebied is verricht.
In hoofdstuk 2 wordt een inleiding gegeven over de verschillende typen fluïde
fasengedrag zoals deze tot nu toe bekend zijn. Dit is gedaan aan de hand van de klassifikatie van
Van Konynenburg en Scott. Er is gepoogd van ieder type binair fluïde fasengedrag minstens 1
experimenteel voorbeeld te geven.
In hoofdstuk 3 wordt een inleiding gegeven over toestandsvergelijkingen in het
algemeen, alsmede een uitvoerige afleiding van de toestandsvergelijking die voortkomt uit de
Simplified Perturbed Hard Chain Theory. Deze toestandsvergelijking wordt verder afgekort tot
de SPHCT-vergelijking. De SPHCT-vergelijking onderscheidt zich van vele andere
toestandsvergelijkingen door een robuuste theoretische achtergrond. De repulsieve en attraktieve
term komen dan ook uitvoerig aan bod. Dit hoofdstuk wordt afgesloten met een korte
verhandeling over de gebruikte mengregels.
234 Samenvatting

In hoofdstuk 4 wordt de kwaliteit van de SPHCT-vergelijking behandeld. Het feit dat een
toestandsvergelijking een robuuste theoretische achtergrond heeft, blijkt toch geen automatische
garantie te geven omtrent de kwaliteit van de voorspellingen die ermee gedaan worden. Het is
niet mogelijk en ook niet het doel van dit proefschrift, alle 'ins and outs' van de SPHCT-
vergelijking te testen. Wij hebben ons dan ook beperkt tot de berekening van enige
thermodynamische grootheden, waarbij soms de nadruk is gelegd op hun gedrag in het kritisch
gebied.
De met de SPHCT-vergelijking berekende dampspanningskurve van een zuivere stof
eindigt in het p,T-vlak exakt in het kritisch punt. De afwijking van experimentele damp-
spanningen in het gereduceerde temperatuurinterval [0.75-1] is dan voor 99 van de 108
onderzochte stoffen minder dan 3%. Helaas worden de verzadigde vloeistof- en dampvolumina
minder goed beschreven. In het berekende kritisch punt is de kompressibiliteitsfaktor voor de
meeste stoffen gelegen tussen 0.34 en 0.35, wat korrespondeert met een afwijking van circa 20%
in het kritische volume. Het moet worden gezegd dat voor kleine en/of polaire molekulen de
Peng-Robinson-vergelijking betere resultaten geeft voor de dampspanningen en volumina in het
kritisch gebied. Voor ketenmolekulen met een koolstofgetal van circa 10 of hoger voldoet de
SPHCT-vergelijking vaak beter, waaruit gekonkludeerd kan worden dat de ingebouwde ideeën
omtrent ketengedrag wel degelijk effektief zijn.
De temperatuurafhankelijkheid van de attraktieve term van de SPHCT-vergelijking wijkt
af van die van de meest gangbare toestandsvergelijkingen, zoals de Peng-Robinson- en de
Soave-Redlich-Kwong-vergelijking (PR- en SRK-vergelijking). Het is opvallend dat de
attraktieve term van de PR- en SRK-vergelijking niet voldoen aan de eis dat deze konstant moet
worden bij lage dichtheid en hoge temperaturen, terwijl de SPHCT-vergelijking wel voldoet aan
deze eis. Een andere eis, waaraan een toestandsvergelijking moet voldoen, is een juiste
beschrijving van de viriaalkoëfficienten als funktie van de temperatuur. Voor methaan, ethaan en
propaan blijkt de SPHCT-vergelijking, net zoals de PR-vergelijking, de temperatuurafhanke-
lijkheid van de tweede viriaalkoëfficient inderdaad juist te beschrijven. Het maximum in de
derde viriaalkoëfficient voor methaan en ethaan wordt noch door de SPHCT- noch door de PR-
vergelijking, correct beschreven. De laatste zuivere stof eigenschap, die onderzocht is met de
SPHCT-vergelijking, is de isochore warmtekapaciteit vanwege het divergerend gedrag dat deze
grootheid rond het kritisch punt vertoont. Zoals van iedere klassieke vergelijking, als de SPHCT-
vergelijking, verwacht mag worden, wordt deze divergentie in het kritisch gebied niet voorspeld.
Er mag dan ook gekonkludeerd worden dat de SPHCT-vergelijking niet alleen kwantitatief maar
ook kwalitatief niet voldoet voor de beschrijving van de isochore warmtekapaciteit in het kritisch
gebied.
De SPHCT-vergelijking is ook getest voor binaire mengsels. Voor eenvoudige apolaire
molekulen als methaan, ethaan, argon en stikstof worden binaire fasenevenwichten goed
beschreven. Ook de Henrykoëfficienten die met de SPHCT-vergelijking berekend zijn voor
stoffen van belang in de gaswinning, komen goed overeen met experimentele data. De laatste
eigenschap van de SPHCT-vergelijking die wij hebben getest, is de beschrijving van kritische
lijnen in binaire fluïde mengsels. Voor de meeste mengsels van koolwaterstoffen is de
Samenvatting 235

beschrijving van de vloeistof-damp kritische lijnen ronduit goed te noemen zolang de molekulen
onderling niet te veel verschillen in grootte.
In hoofdstuk 5 worden de mathematische kondities beschreven van de verschillende
lijnen en punten die voorkomen in een globaal fasendiagram, waaronder de trikritische kurven,
het tetrakritisch punt, de dubbele kritische eindpuntkurven, het Van Laar punt, de azeotropische
kurven, de azeotropische kritische eindpuntkurven en de mathematische dubbelpuntkurven. Voor
de berekening van een mathematisch dubbelpunt zijn de kondities afgeleid in volume-,
temperatuur- en samenstellingskoördinaten. Voor een drukexpliciete toestandsvergelijking, zoals
de SPHCT-vergelijking, was dit nog niet eerder in deze vorm gedaan. Bovendien hebben wij
aangetoond dat er twee soorten mathematische dubbelpunten zijn, stabiele en instabiele. Dit
hoofdstuk wordt afgesloten met een korte beschouwing van het 'drukplafond' dat bestaat in het
p,T-vlak volgens de SPHCT-vergelijking. Dit drukplafond is fysisch niet reëel en is het gevolg
van een beperking in de Carnahan-Starling uitdrukking, die gebruikt is in de repulsieve term van
de SPHCT-vergelijking.
In hoofdstuk 6 wordt ingegaan op de numerieke methoden, die gebruikt zijn bij de
berekening van globale fasendiagrammen. Het Marquardt-algoritme, dat gebruikt is voor het
oplossen van stelsels niet-lineaire vergelijkingen, wordt daarbij uitvoerig behandeld. Dit
hoofdstuk wordt afgesloten met een korte beschrijving van het gebruik van het computer algebra
programma Maple. Met Maple zijn de uitdrukkingen afgeleid voor de partiële afgeleiden van de
molaire Gibbsenergie naar de samenstelling en van de molaire Helmholtzenergie naar het
molaire volume en de temperatuur. De laatstgenoemde afgeleiden zijn afhankelijk van de
gebruikte toestandsvergelijking, de eerste niet.
In hoofdstuk 7 zijn de resultaten weergegeven van de globale fasendiagramberekeningen.
Het blijkt dat de SPHCT zich in dit opzicht anders kan gedragen dan andere
toestandsvergelijkingen. Met name het ontstaan van een tetrakritisch punt en de daarbij
behorende nieuwe typen fasengedrag krijgen ruim de aandacht. Zij zijn volgens de klassifikatie
van Van Konynenburg en Scott genummerd en er is voorgesteld om deze typen, type VIII, IX, X
en XI te noemen. Daarbij is ook een nieuwe variant van type VII gevonden, die wij type VII2h
hebben genoemd.
In hoofdstuk 8 worden enkele bijzondere verschijnselen, die in hoofdstuk 7 beschreven
zijn, nader toegelicht. Hieronder valt de mogelijkheid om met de SPHCT-vergelijking gesloten
kritische kurven en de zogenaamde hoge-druk kritische kurven te berekenen. Vooral de
interferentie van deze typen kritische kurven levert interessante fasendiagrammen op. Gepoogd
is reëele systemen te vinden, die deze interferentie van de hoge-druk kritische kurve en de
gesloten vloeistof-vloeistof kritische kurve zouden kunnen vertonen. Het bleek mogelijk om
zuivere stoffen te vinden die voldoen aan de eisen, zoals ethaan+n-butaan en ammonia+een kort
perfluoroalkaan. Helaas is de kans klein, dat de genoemde typen fasengedrag daadwerkelijk
gevonden zullen worden in deze binaire stelsels. De reëele binaire systemen, bestaande uit de
genoemde zuivere stoffen, vertonen niet de vereiste binaire interaktiekoëfficienten.
Een andere mogelijkheid is dat de verschijnselen zich afspelen in een temperatuurgebied, waar
vaste fasen stabiel zijn.
236 Samenvatting

Een bijzonder punt in een globaal fasendiagram is het Van Laar punt. Volgens
berekeningen komen in dit punt vier takken (waarvan één instabiel) van twee kritische lijnen
tezamen met een driefasenlijn. De gebieden in het globale fasendiagram waar de typen II, III, IV
en IV* voorkomen, zijn gegroepeerd rond het Van Laar punt. Deze typen fasengedrag worden in
dit hoofdstuk uitvoerig besproken.
In hoofdstuk 9 tenslotte, worden algemene konklusies gegeven alsmede voorstellen
gedaan omtrent toekomstig werk.
237

APPENDIX A
PARTIAL HELMHOLTZ DERIVATIVES OF THE SPHCT EQUATION OF STATE

The molar Helmholtz energy according to the SPHCT equation of state consists of the
following five contributions:
1. a contribution from the pure components
2. a contribution from the ideal mixing term
3. a contribution from the ideal gas part
4. a contribution from the repulsive part
5. a contribution from the attractive part

A ( V, T, x ) = x1 A1* ( V0 , T ) + x2 A2* ( V0 , T ) + A mix ( T, x ) + (A1)

+ Aig ( V, T ) + A r ( V, T, x ) + A a ( V, T, x )

The partial derivatives of the molar Helmholtz energy are built up from the derivatives of each of
the separate terms. All derivatives AiVjx consist of four terms because the contribution of the pure
components of the molar Helmholtz energy, the first two terms of Eq.(A1), will disappear after
one differentiation with respect to the molar volume. All partial derivatives have the following
structure:
mix ig
AiVjx ( V, T, x ) = AiVjx ( T, x ) + AiVjx ( V, T ) +
(A2)
r a
+ AiVjx ( V, T, x ) + AiVjx ( V, T, x )

If i or j equals 1 the index in the partial derivative is omitted. For the critical conditions in a
binary fluid mixture the partial derivatives of the different contributions are given below:

The ideal mixing term.


RT
A2mix
x = (A3)
x (1- x )

- RT ( 1 - 2 x )
A3mix
x = (A4)
x2 (1 - x )2

A2mix mix mix mix mix (A5)


V = AVx = A3V = A2Vx = AV 2 x = 0
238 Appendix A

The ideal gas term.

RT
A2igV = (A6)
V2
2RT
A3igV = - (A7)
V3
AVxig = A2igx = A2igVx = AVig2 x = A3igx = 0 (A8)

The repulsive term.

In the expressions for the partial derivatives of the repulsive part of the molar Helmholtz energy
the following shorthand notation is used:

Vijk = τ i + j < V * > i < V *' > j V k with


(A9)
∂ <V >*
∂c
< V *' > = and c' =
∂x ∂x

In Eq.(A-9) the i, j and k are used as powers in Vijk and should not only be seen as indices.

r c V 100 ( V 100 - 4 V 001 )


A2 V = - 2 R T (A10)
( V 100 - V 001 )4

r c ( 2 V 111 - V 210 + 2 V 012 ) + c′ ( - 3 V 201 + V 300 + 2 V 102 )


AV x = - 2 R T (A11)
( V 100 - V 001 )4

r c ( 5 V 011 - 2 V 110 ) + c′ ( - 6 V 101 + 2 V 200 + 4 V 002 )


A2 x = 2 R T V 011 (A12)
( V 100 - V 001 )4

r c V 101
A3 V = 24 R T (A13)
( V 100 - V 001 )5

r c ( 10 V 111 - 2 V 210 + 4 V 012 ) + c′ (-5 V 201 + V 300 + 4 V 102 )


A2 V x = - 2 R T (A14)
( V 100 - V 001 )5

r c ( 5 V 012 + 2 V 111 - V 210 ) + c′ ( - 3 V 201 + V 300 + 2 V 003 )


AV 2 x = 4 R T V 010 (A15)
( V 100 - V 001 )5

r c ( 6 V 011 - 2 V 110 ) + c′ ( 5 V 002 - 7 V 101 + 2 V 200 )


A3 x = - 6 R T V 021 (A16)
( V 100 - V 001 )5
Appendix A 239

The attractive term.

In the expressions for the partial derivatives of the attractive part of the molar Helmholtz energy,
the following shorthand notation is used:

∂a ∂ 2a ∂c
a = < cV *Y > , a ′ = , a ′′ = 2 , c′ = (A17)
∂x ∂x ∂x
The number of primes after a variable, denotes the number of times this variable is differentiated
with respect to the composition x.

a a c( 2 cV +a )
A2V = - R T Z m 2 2
(A18)
V ( cV + a )

a ( a 2 c′ + a ′ c 2 V )
AVx = R T Z m (A19)
V ( c V + a )2

a - a′′ c 2 ( c V + a ) + ( a c′ - a′ c )2
A = RT Zm
2x (A20)
c ( c V + a )2

a c a ( 3 c V ( c V + a ) + a2 )
A = 2 RT Zm
3V 3 3 (A21)
V ( cV +a )

a c′ ( 3 c V + a ) + 2 a′ c V
2 3 2
a
A 2Vx = - RT Zm 2 3
(A22)
V ( cV +a )

a - a′′ c 2 ( c V + a ) + 2 ( a c′ - a′c )2
AV2x = - R T Z m (A23)
( c V + a )3

A3x = - R T Z m ( a c′ - a′ c ) .
a

(A24)
3 a′′ c2 ( c V + a ) + a c′ 2 ( 3 c V + a ) + a′ c c′ ( - 3 c V + a ) - 2 a′ 2 c2
2 2
c ( cV +a )
240

Curriculum Vitae
1. Personal data

Name Antonius (Ton) van Pelt


Date of birth 08 September 1964
Place of birth Rotterdam, The Netherlands
Marital status Married since 01 September 1992

2. Education

1970-1976 Primary school "De Kerkhoeck" in Heenvliet


1976-1982 Grammar school "Blaise Pascal" in Spijkenisse (diploma 1982)
1982-1988 Delft University of Technology, study Chemical Engineering (master degree:
1988)

Masters project: PHASEGRAPH - Classification of solid-liquid-


equilibria, Inorganic and Physical Chemistry Group,
Section Applied Thermodynamics and Phase
Equilibria

1988-1992 Delft University of Technology, PhD research Chemical Engineering (PhD


degree: 1992)

PhD subject: Critical phenomena in binary fluid mixtures. Classification


of phase equilibria with the Simplified-Perturbed-Hard-
Chain Theory.

3. Work experience

- Industrial experience at Koninklijk Shell Laboratorium Amsterdam: Assessment of excess


models for the purpose of writing software to describe the phase equilibria of metal alloys
(1986)
- Laboratory assistant at the International Environmental Engineering Institute (1985, 1986,
1987)
Curriculum Vitae 241

- PhD research at Delft University of Technology (including laboratory assistance 2nd-year


students and a stay of 5 months at the Ruhr Universtät Bochum, Germany, 1990-1991)
- Scientific assistant Delft University of Technology: Investigation of liquid-liquid
immiscibility loops and high-pressure immiscibility.

4. Publications

- A. van Pelt. Phasegraph, Klassifikatie van vaste stof-vloeistof-evenwichten. Master thesis,


Delft University of Technology, (1988).
- A. van Pelt. Toepassing van MAPLE in de thermodynamika. Centre of Expertise
Computer Algebra Nederland, Nieuwsbrief 5, (1990)
- A. van Pelt, C.J. Peters and J. de Swaan Arons. Liquid-liquid immiscibility loops predicted
with the simplified-perturbed-hard-chain Theory. J. Chem. Phys. 95(10), (1991), 7569-
7575.
- A. van Pelt, C.J. Peters and J. de Swaan Arons. Application of the Simplified- Perturbed-
Hard-Chain Theory for pure components near the critical point. Fluid Phase Equilibria 74,
(1992), 67-83.
- A. van Pelt and Th.W. de Loos. Connectivity of critical lines around the van Laar point in
T,x-projections. J. Chem. Phys. 97(2), (1992), 1271-1281.
- A. van Pelt. Critical phenomena in binary fluid mixtures. Recl. Trav. Chim. Pays-Bas,
112(4), (1993), 272.
- A. van Pelt, C.J. Peters and J. de Swaan Arons. Calculation of critical lines in binary
mixtures with the Simplified-Perturbed-Hard-Chain Theory. Fluid Phase Equilibria 84,
(1993), 23-47.
- A. van Pelt, U.K. Deiters, C.J. Peters and J. de Swaan Arons. The limiting behaviour at
high temperatures of the Simplified-Perturbed-Hard-Chain Theory. Fluid Phase Equilibria
90, (1993), 45-56.
- A. van Pelt, P.H.E. Meijer, C.J. Peters and J. de Swaan Arons. Mathematical double points
according to the Simplified-Perturbed-Hard-Chain Theory. J. Chem. Phys. 99(12), (1993),
9920-9929.
- A. van Pelt, G.X. Jin and J.V. Sengers. Critical scaling laws and a classical equation of
state. Int. J. Thermophys. 15(4), (1994), 687-697.
- A. van Pelt, C.J. Peters, J. de Swaan Arons and U.K. Deiters. The global phase behaviour
of the Simplified-Perturbed-Hard-Chain Theory. J. Chem. Phys. 102(8), (1995), 3361-
3375.
- A. van Pelt and J.V. Sengers. Thermodynamic properties of 1,1-Difluoroethane (R152a) in
the critical region. J. of Supercritical Fluids 8, (1995), 81-99.

5. Future employment

- Postdoc position at the Institute of Physical Science and Technology at the University of
Maryland in College Park (Maryland, USA): Crossover approach to global critical
phenomena in fluids.
242 Curriculum Vitae
Curriculum Vitae 243

You might also like