Sunteți pe pagina 1din 171

PRO/II Component

and Thermophysical
Properties
Reference Manual

The software described in this manual is furnished under a license


agreement and may be used only in accordance with the terms of that
agreement.
Information in this document is subject to change without notice.
Simulation Sciences Inc. assumes no liability for any damage to any
hardware or software component or any loss of data that may occur as
a result of the use of the information contained in this manual.

Copyright Notice

Copyright 1994 Simulation Sciences Inc. All Rights Reserved. No


part of this publication may be copied and/or distributed without the
express written permission of Simulation Sciences Inc., 601 S. Valencia
Avenue, Brea, CA 92621, USA.

Trademarks

PRO/II is a registered mark of Simulation Sciences Inc.


SIMSCI is a service mark of Simulation Sciences Inc.
Printed in the United States of America.

Credits

Contributors:
Althea Champagnie, Ph.D.
John Cunningham
Allan Harvey, Ph.D.
John Tanger, Ph.D.
C.H. Twu, Ph.D.
Layout:
Kris Oca
On-line Document Conversion:
Mark Norton
Peter Stepman
Althea Champagnie, Ph.D.

Table of Contents

1.1

List of Tables
List of Figures

TOC-5
TOC-6

Introduction

Int-1
Int-1

What is in This Manual?

Int-1

Who Should Use This Manual?


Finding What you Need

Int-1
Int-1

Component Data

I-3

1.1.1

Defined Components
Component Libraries

I-4
I-4

Using DATAPREPTM
Fixed Properties

I-6
I-6

Temperature-dependent Properties

I-6

Properties From Structure


Petroleum Components
General Information
Property Generation--SIMSCI Method

I-8
I-9
I-9
I-9

Property Generation--CAVETT Method

I-13

Property Generation--Lee-Kesler Method


Assay Processing
General Information

I-16
I-18
I-18

Cutpoint Sets (Blends)


Interconversion of Distillation Curves

I-19
I-21

Cutting TBP Curves

I-25

Generating Pseudocomponent Properties


Vapor Pressure Calculations

I-30
I-30

1.1.2

1.1.3

1.2

General Information

Thermodynamic Methods

I-37

1.2.1

Basic Principles
General Information

I-38
I-38

Phase Equilibria
Enthalpy

I-38
I-41

Entropy

I-43

Density
Application Guidelines
General Information
Thermodynamic Expert System (TES)

I-44
I-45
I-45
I-45

Refinery and Gas Processes

I-46

Natural Gas Processing


Petrochemical Applications

I-49
I-52

Chemical Applications

I-54

1.2.2

PRO/II Component and Thermophysical Properties


Reference Manual

Table of Contents

TOC-1

1.2.3

1.2.4

Generalized Correlation Methods


General Information
Ideal (IDEAL)

I-58
I-58
I-58

Chao-Seader (CS)
Grayson-Streed (GS)

I-60
I-61

Erbar Modification to Chao-Seader (CSE) and


Grayson-Streed (GSE)
Improved Grayson-Streed (IGS)

I-61
I-62

Curl-Pitzer (CP)

I-62

Braun K10 (BK10)


Johnson-Grayson (JG)

I-63
I-64

Lee-Kesler (LK)

I-64

API
Rackett

I-65
I-65

COSTALD

I-66

Equations of State
General Information

I-69
I-69

General Cubic Equation of State

I-69

Alpha Formulations
Mixing Rules

I-71
I-73

Soave-Redlich Kwong (SRK)


Peng-Robinson (PR)

I-74
I-74

Soave-Redlich-Kwong Kabadi-Danner (SRKKD)

I-75

Soave-Redlich-KwongPanagiotopoulos-Reid
(SRKP) and Peng-Robinson PanagiotopoulosReid (PRP)

I-76

Soave-Redlich-Kwong Modified PanagiotopoulosReid (SRKM) and Peng-Robinson Modified


Panagiotopoulos-Reid (PRM)
I-77

1.2.5

1.2.6

TOC-2

Table of Contents

Soave-Redlich-Kwong SimSci (SRKS)


Soave-Redlich-Kwong Huron-Vidal (SRKH) and
Peng-Robinson Huron-Vidal (PRH)

I-77

HEXAMER
UNIWAALS

I-80
I-83

Benedict-Webb-Rubin-Starling

I-84

Lee-Kesler-Plcker (LKP)
Free Water Decant
General Information
Calculation Methods

I-85
I-88
I-88
I-88

Liquid Activity Coefficient Methods


General Information
Margules Equation

I-90
I-90
I-93

van Laar Equation

I-94

Regular Solution Theory


Flory-Huggins Theory

I-95
I-96

Wilson Equation

I-97

NRTL Equation
UNIQUAC Equation

I-98
I-99

UNIFAC

I-101

Modifications to UNIFAC
Fill Methods

I-104
I-107

I-79

May 1994

1.2.7

1.2.8

1.2.9

1.2.10

1.2.11

1.2.12

Henrys Law

I-110

Heat of Mixing Calculations


Vapor Phase Fugacities
General Information
Equations of State

I-111
I-113
I-113
I-114

Truncated Virial Equation of State

I-114

Hayden-OConnell
Special Packages
General Information

I-116
I-117
I-117

Alcohol Package (ALCOHOL)


Glycol Package (GLYCOL)

I-117
I-120

Sour Package (SOUR)

I-122

GPA Sour Water Package (GPSWATER)


Amine Package (AMINE)

I-125
I-127

Electrolyte Mathematical Model


Discussion of Equations

I-131
I-131

Modeling Example

I-133

Electrolyte Thermodynamic Equations


Thermodynamic Framework

I-135
I-135

Equilibrium Constants

I-135

Aqueous Phase Activities


Vapor Phase Fugacities

I-136
I-139

Organic Phase Activities


Enthalpy

I-143
I-143

Aqueous Liquid Phase

I-144

Molar Volume and Density


Solid-Liquid Equilibria
General Information

I-144
I-147
I-147

vant Hoff Equation


Solubility Data

I-147
I-148

Fill Options for Solubility Data

I-148

Transport Properties
General Information

I-150
I-150

PURE Methods
PETRO Methods

I-150
I-151

TRAPP Correlation

I-155

Special Methods for Liquid Viscosity


Liquid Diffusivity

I-157
I-159

Index

PRO/II Component and Thermophysical Properties


Reference Manual

Idx-1

Table of Contents

TOC-3

List of Tables

TOC-4

1.1.1-1

PRO/II Library Component Properties . . . . . . . . . . . . . I-5

1.1.1-2

PRO/II Temperature-dependent Property Equations and


Extrapolation Conventions . . . . . . . . . . . . . . . . . . . I-7

1.1.1-3

PRO/II Vapor Pressure Equations . . . . . . . . . . . . . . . . I-7

1.1.2-1

Values of Constants for Equations (14)-(17) . . . . . . . . . . I-11

1.1.3-1

Primary TBP Cutpoint Set . . . . . . . . . . . . . . . . . . . . I-19

1.1.3-2

Blending Example . . . . . . . . . . . . . . . . . . . . . . . . I-20

1.1.3-3

Values of Constants a, b, c . . . . . . . . . . . . . . . . . . . I-23

1.1.3-4

Values of Constants a, b . . . . . . . . . . . . . . . . . . . . . I-25

1.2.2-1

Methods Recommended for Low Pressure Crude Systems . . I-47

1.2.2-2

Methods Recommended for High Pressure Crude Systems . . I-47

1.2.2-3

Methods Recommended for Reformers and Hydrofiners . . . . I-48

1.2.2-4

Methods Recommended for Lube Oil and Solvent


De-asphalting Units . . . . . . . . . . . . . . . . . . . . . . . I-48

1.2.2-5

Methods Recommended for Natural Gas Systems . . . . . . . I-50

1.2.2-6

Methods Recommended for Sour Water Systems . . . . . . . I-51

1.2.2-7

Recommended Ranges for Amine Systems . . . . . . . . . . . I-52

1.2.2-8

Methods Recommended for Light Hydrocarbons . . . . . . . . I-52

1.2.2-9

Methods Recommended for Aromatics . . . . . . . . . . . . . I-53

1.2.2-10

Methods Recommended for Aromatic/Non-aromatic


Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . I-54

1.2.2-11

Methods Recommended for Alcohol Systems . . . . . . . . . I-54

1.2.2-12

Methods Recommended for Non-ionic Chemical Systems . . . I-56

1.2.2-13

Methods Recommended for Ionic Chemical Systems . . . . . I-56

1.2.2-14

Methods Recommended for Environmental Applications . . . I-57

1.2.2-15

Methods Recommended for Solid Applications . . . . . . . . I-57

1.2.4-1

Some Cubic Equations of State . . . . . . . . . . . . . . . . . I-69

1.2.4-2

Constants for Two-parameter Cubic Equations of State . . . . I-70

1.2.4-3

Alpha Formulations . . . . . . . . . . . . . . . . . . . . . . . I-72

1.2.5-1

Components Available in the SIMSCI Water Solubility Method . I-89

1.2.6-1

Margules Equation . . . . . . . . . . . . . . . . . . . . . . . I-93

1.2.6-2

van Laar Equation . . . . . . . . . . . . . . . . . . . . . . . . I-94

1.2.6-3

Regular Solution Theory . . . . . . . . . . . . . . . . . . . . I-95

1.2.6-4

Flory-Huggins Theory . . . . . . . . . . . . . . . . . . . . . . I-96

1.2.6-5

Wilson Equation . . . . . . . . . . . . . . . . . . . . . . . . . I-97

Table of Contents

May 1994

1.2.65

Wilson Equation . . . . . . . . . . . . . . . . . . . . . . . . . I-97

1.2.6-6

NRTL Equation . . . . . . . . . . . . . . . . . . . . . . . . . . I-98

1.2.6-7

UNIQUAC Equation . . . . . . . . . . . . . . . . . . . . . . . I-99

1.2.6-8

UNIFAC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . I-101

1.2.8-1

Components Available for ALCOHOL Package . . . . . . . . . I-118

1.2.8-2

Components Available for GLYCOL Package . . . . . . . . . . I-120

1.2.8-3

Application Guidelines for Amine Systems . . . . . . . . . . . I-129

1.2.11-1

vant Hoff . . . . . . . . . . . . . . . . . . . . . . . . . . . . I-148

1.2.12-1

Transport Properties . . . . . . . . . . . . . . . . . . . . . . . I-150

1.2.12-2

Stream Average Properties . . . . . . . . . . . . . . . . . . . I-151

1.2.12-3

TRAPP Components (3.3 versions) . . . . . . . . . . . . . . . I-156

List of Figures
1.1.3-1

Cutting TBP Curves . . . . . . . . . . . . . . . . . . . . . . . I-27

1.1.3-2

Matching Lightends to TBP Curve . . . . . . . . . . . . . . . . I-30

1.2.6-1

Flowchart for FILL Methods . . . . . . . . . . . . . . . . . . . I-109

1.2.8-1

Binary Interaction Data in the Alcohol Databank . . . . . . . . I-119

1.2.8-2

Binary Interaction Data in the Glycol Databank . . . . . . . . . I-121

PRO/II Component and Thermophysical Properties


Reference Manual

Table of Contents

TOC-5

This page left intentionally blank.

TOC-6

Table of Contents

May 1994

Introduction
General
Information

The PRO/II Component and Thermophysical Properties Reference Manual


provides details on the basic equations and calculation techniques used in the
PRO/II simulation program. It is intended as a complement to the PRO/II
Keyword Input Manual, providing a reference source for the background behind
the various PRO/II calculation methods.

What is in
This Manual?

This manual contains the correlations and methods used to calculate thermodynamic and physical properties, such as the Soave-Redlich-Kwong (SRK)
cubic equation of state for phase equilibria. This manual also contains information on the definition of pure components and petroleum fractions.
For each method described, the basic equations are presented, and appropriate references provided for details on their derivation. General application
guidelines are provided, and, for many of the methods, hints to aid solution
are supplied.

Who Should Use


This Manual?

For novice, average, and expert users of PRO/II, this manual provides a good
overview of the property calculation methods used to simulate a single unit
operation or a complete chemical process or plant. Expert users can find
additional details on the theory presented in the numerous references cited
for each method. For the novice to average user, general references are also
provided on the topics discussed, e.g., to standard textbooks.
Specific details concerning the coding of the keywords required for the
PRO/II input file can be found in the PRO/II Keyword Input Manual.
Detailed sample problems are provided in the PRO/II Application Briefs
Manual and in the PRO/II Casebooks.

Finding What
you Need

A Table of Contents and an Index are provided for this manual. Crossreferences are provided to the appropriate section(s) of the PRO/II Keyword
Input Manual for help in writing the input files.

PRO/II Component and Thermophysical Properties


Reference Manual

Introduction

Int-1

Symbols Used in This Manual


Symbol

Meaning

Indicates a PRO/II input coding note. The number beside the


symbol indicates the section in the PRO/II Keyword Input
Manual to refer to for more information on coding the
input file.
Indicates an important note.

Indicates a list of references.

Int-2

Introduction

May 1994

Section 1.1

1.1

Component Data

Component Data
PRO/II allows the user to specify pure-component physical property data for a
given simulation. Pure component data are usually associated with either a predefined component in a data library, a user-defined (non-library) component, or a
petroleum pseudocomponent.
Properties for defined components can be accessed in a variety of ways. They can be
retrieved from an on-line databank or library, estimated from structural or other
data, or input by the user as non-library components. User input can be used to
override properties retrieved from the libraries.
Properties for pseudo or petroleum components are derived from generalized
correlations based on minimal data, usually the normal boiling point, molecular
weight, and standard density. Hydrocarbon streams defined in terms of assay data
(including distillation data) can be converted to discrete pseudocomponents by a
number of assay processing methods.

PRO/II Component and Thermophysical Properties


Reference Manual

I-3

Component Data

1.1.1

Section 1.1

Defined Components
Component
Libraries

Table 1.1.1-1 lists the property data available in the built-in component libraries for
predefined components. These libraries include the PROCESS library (the physical
property library used as the default in PROCESS, PIPEPHASE, HEXTRAN, and
early versions of PRO/II), the SIMSCI library (a fully documented physical property bank), the DIPPR (Design Institute for Physical Property Research) library
from the American Institute of Chemical Engineers, and the OLILIB library of
electrolyte species, which contains a subset of the library component properties
listed in the following sections.
Most of the fixed properties used in a simulation can be found in the input
reprint of the simulation. The coefficients of the correlations used for the
temperature-dependent properties stored in the libraries are not shown because they are usually covered by contractual agreements which disallow
their display in a simulation.
References

I-4

Defined Components

1.

PPDS, Physical Property Data Service, jointly sponsored by the National


Physical Laboratory, National Engineering Laboratory, and the Institution
of Chemical Engineers in the UK.

2.

DIPPR, Design Institute for Physical Property Data, sponsored by the


American Institute of Chemical Engineers.

May 1994

Section 1.1

Component Data
Table 1.1.1-1: PRO/II Library Component Properties
Fixed Properties and Constants

Temperature-dependent Properties

Acentric Factor

Enthalpy of Vaporization

Carbon Number

Ideal Vapor Enthalpy

Chemical Abstract Number

Liquid Density

Chemical Formula

Liquid Thermal Conductivity

Critical Compressibility Factor

Liquid Viscosity

Critical Pressure

Saturated Liquid Enthalpy

Critical Temperature

Solid Density

Critical Volume

Solid Heat Capacity

Dipole Moment

Solid Vapor Pressure

Enthalpy of Combustion

Surface Tension

Enthalpy of Fusion

Vapor Pressure

Flash Point

Vapor Thermal Conductivity

Free Energy of Formation

Vapor Viscosity

Freezing Point (normal melting point)


Gross Heating Value
Heat of Formation
Hydrogen Deficiency Number
Liquid Molar Volume
Lower Heating Value
Molecular Weight
Normal Boiling Point
Rackett Parameter
Radius of Gyration
Solubility Parameter
Specific Gravity
Triple Point Temperature
Triple Point Pressure
UNIFAC Structure
van der Waals Area and Volume

PRO/II Component and Thermophysical Properties


Reference Manual

Defined Components

I-5

Component Data

Using DATAPREPTM

Section 1.1

SimSci provides an interactive program, DATAPREP, for review and manipulation of the pure component fixed and temperature-dependent properties available for defined components in libraries. DATAPREP is PC based.
Detailed descriptions for both the fixed and temperature-dependent library
properties mentioned above are contained in Appendix A of the DATAPREP
User Guide.
A comprehensive summary of the available data for each component, for any
release of the libraries, can also be generated by using DATAPREP.
DATAPREPs functionality also includes the ability to generate keyword
input file inserts containing component properties for non-library components. Private pure component libraries can also be made using DATAPREP.
In addition to the USER.LB1 and USER.LB2 files that can be used directly
by PRO/II on a PC, ASCII input files are generated for input to the library
manager program for use on other platforms. Please refer to the DATAPREP
User Guide for further information about its capabilities.
Reference
DATAPREP User Guide, 1991, Simulation Sciences Inc.

Fixed Properties

As explained in the above section, these properties are described in Appendix A of the DATAPREP User Guide. Some things to be aware of are that
the specific gravities of permanent gases are often given relative to air, without any annotations in the output, and liquid molar volumes can be
extrapolated from a condition very different from 77 F (25 C), if the
component doesnt naturally exist as a liquid at 77 F.

Temperaturedependent
Properties

The temperature-dependent correlations available for use in PRO/II are listed


in Section 17, Component Properties, of the PRO/II Keyword Input Manual.
The equations that are typically used to represent a property are listed in Table 1.1.1-2. While temperature-dependent library properties are fitted and are
usually very accurate at saturated, subcritical conditions, caution must be
used in the superheated or supercritical regions.
Because of the form of some of the allowable temperature-dependent equations,
extrapolation beyond the minimum and maximum temperatures is not done using
the actual correlation. PRO/II has adopted the rules shown in Table 1.1.1-2, based
on the property, for extrapolation of the temperature-dependent correlations.

I-6

Defined Components

May 1994

Section 1.1

Component Data
Table 1.1.1-2: PRO/II Temperature-dependent Property Equations and
Extrapolation Conventions
Temperature-dependent
Property

Recommended
Equations

Extrapolation
Method

Vapor Pressure

14, 20, 21, 22

ln(Prop.) vs. 1/T

Liquid Density

1, 4, 16, 32

Prop. vs. T

Ideal Vapor Enthalpy

1, 17, 41

Prop. vs. T

Enthalpy of Vaporization

4, 15, 36, 43

Prop. vs. T

Saturated Liquid Enthalpy

1, 42, 35

Prop. vs. T

Liquid Viscosity

13, 20, 21

ln(Prop.) vs. 1/T

Vapor Viscosity

1, 19, 26, 27

Prop. vs. T

Liquid Thermal Conductivity

1, 4, 34

Prop. vs. T

Vapor Thermal Conductivity

1, 19, 33

Prop. vs. T

Surface Tension

1, 15, 30

Prop. vs. T

Solid Thermal Conductivity

Prop. vs. T

Solid Density

Prop. vs. T

Solid Cp or Enthalpy

Prop. vs. T

Solid Vapor Pressure

20

ln(Prop.) vs. 1/T

Another note of caution concerns the use of equations 20 and 21 in modeling


component vapor pressures. These equations are actually combinations of
two or more traditionally used vapor pressure equations (e.g., Antoine). It is
intended that the user apply only subsets of the available coefficients with
these equations corresponding to the more traditional equations. Table
1.1.1-3 gives some examples of this mapping.

Table 1.1.1-3: PRO/II Vapor Pressure Equations


Equation 20 / 21 Coefficients
Common VP Equations (#)

C1

C2

Clapeyron (20 or 21)

Antoine (21)

Riedel (20)

Frost-Kalkwarf (21)

Reidel-Plank-Miller (20)

PRO/II Component and Thermophysical Properties


Reference Manual

C3

C4

C5

C6

C7

x
x
x

Defined Components

I-7

Component Data

Properties From
Structure

I-8

Defined Components

Section 1.1

Properties for defined components, either library or non-library, may be


estimated if the user supplies a component structure and invokes the FILL
option in the component data category of input. This procedure primarily
uses the methods of Joback and is good for components with molecular
weights below 400 and components with less than 20 unique structural
groups. More accurate results are obtained for components containing just
one type of functional group. For example, amine properties would be more
accurate than those predicted for an ethanol amine, which would contain
functional groups for both an alcohol and an alcohol amine. This feature is
available in DATAPREP and in PRO/II for all versions subsequent to v3.3.
Complete documentation of the estimation techniques used is contained in
Section 3.7, Prediction, of the DATAPREP User Guide.

May 1994

Section 1.1

1.1.2

Component Data

Petroleum Components
General
Information

Petroleum components (often called pseudocomponents) are either defined


on a one-by-one basis on PETROLEUM statements or generated from one or
more streams given in terms of assay data. The processing of assays is described in Section 1.1.3, Assay Processing. Each individual pseudocomponent is typically a narrow-boiling cut or fraction. Component properties are
generated based on two of the following three properties:
Molecular weight.
Normal boiling point (NBP).
Standard liquid density.
If only two are supplied, the third is computed with the SIMSCI method (or
with another method if requested with the MW keyword). These methods are
described in the sections below.
From those three basic properties, the program estimates all other properties
needed for the calculation of thermophysical properties. Three different sets of
characterization methods are provided. These are known as the CAVETT, SIMSCI, and Lee-Kesler methods. The Cavett methods developed in 1962 have been
the default in all versions of PRO/II up to and including the 3.5 series. The SIMSCI methods use a combination of published (Black and Twu, 1983; Twu, 1984)
and proprietary methods developed by SimSci. These are the default for all
PRO/II versions subsequent to the 3.5 series. The LK option accesses methods
developed by Lee and Kesler in 1975 and 1976.

Property
Generation-SIMSCI Method

Critical Properties and Acentric Factor


The SIMSCI characterization method was developed by Twu in 1984. It
expresses the critical properties (and molecular weight) of hydrocarbon components as a function of NBP and specific gravity. The correlation is expressed as a perturbation about a reference system of normal alkanes. The
critical temperature (in degrees Rankine) is given by:

Tc = Tc 1 + 2fT 1 2fT

(1)

12
12
fT = SGT0.362456 Tb + 0.03982850.948125 Tb SGT

(2)

PRO/II Component and Thermophysical Properties


Reference Manual

Petroleum Components

I-9

Component Data

Section 1.1

SGT = exp5SG SG 1

(3)
3

7 2

(4)

Tc = Tb (0.533272 + 0.19101710 Tb + 0.77968110 Tb

0.2843761010T3b + 0.9594681028 Tb 13)1

SG = 0.843593 0.128624 3.36159 13749.5

(5)

12

where:
SG =

specific gravity

Tb =

normal boiling point, degrees Rankine

1 - T b / T c

SG = specific gravity correction


f=

correction factor

SG =

specific gravity

subscript T refers to the temperature


subscript c refers to the critical conditions
superscript refers to the reference system
The critical volume (in cubic feet per pound mole) and the critical pressure
(in psia) are given by similar expressions:

Vc = Vc 1 + 2fV 1 2fV

12

fV = SGV 0.466590 Tb

(6)

1 2
+ 0.182421 + 3.01721 Tb SGV

(7)

SGV = exp 4SG SG 1


(8)

(9)

3
14
Vc = 1 0.419869 0.505839 1.56436 9481.70

Pc = Pc Tc Tc Vc Vc 1 + 2fP 1 2fP

(10)

12

fP = SGP [ 2.53262 46.1955 Tb

1 2

+ 11.4277 + 252.140 Tb

0.00127885Tb

(11)

+ 0.00230535Tb SGP]

SGP = exp0.5SG SG 1

Pc = 3.83354 + 1.19629

where:

12

2
4
+ 34.8888 + 36.1952 + 104.193

V=

molar volume, ft3/lbmole

P=

pressure, psia

(12)
2

(13)

subscripts V and P refer to the volume and pressure

I-10

Petroleum Components

May 1994

Section 1.1

Component Data

The acentric factor for the SIMSCI method is estimated with the use of a generalized Frost-Kalkwarf vapor equation developed at SimSci. The equation is
given by:
lnPR = A1 + A4 + A2 + A5 TR + A3 + A6 lnTR + A7PRTR

(14)

where:
A1 to A7 =

constants given in Table 1.1.2-1

PR =

reduced pressure (P/Pc)

TR =

reduced temperature (T/Tc ),

a parameter evaluated at the NBP and given by:

lnP A P T 2 f T
R,b
7 R,b R,b
R,b
=
f TR,b

(15)

where:
subscripts R,b indicate reduced properties evaluated at the
normal boiling point
Functions f and f are given by:
f (TR) = A1 + A2 TR + A3 lnTR

(16)

f (TR) = A4 + A5 TR + A6 lnTR

(17)

The values of the seven constants in these equations are shown in Table 1.1.2-1.
Table 1.1.2-1: Values of
Constants for Equations (14)-(17)
A1

10.2005

A2

-10.6317

A3

-5.58058

A4

2.09167

A5

-2.09167

A6

1.70214

A7

0.4312

To compute the acentric factor, the parameter is determined using equation


(15) and the known (or already estimated) values for the critical temperature
and pressure and the normal boiling point (NBP). This is then used in equation
(14) to compute the reduced vapor pressure at a reduced temperature of 0.7,
which is then used in the definition of the acentric factor:
= log10 PR,T

PRO/II Component and Thermophysical Properties


Reference Manual

=0.7

(18)

Petroleum Components

I-11

Component Data

Section 1.1

Other Fixed Properties


The heat of formation is computed from a proprietary correlation developed by
SimSci. The solubility parameter is estimated from the following equation:
HV RT 1 2
=

VL

(19)

The molar latent heat of vaporization, HV, is computed from the KistiakowskyWatson method described later on in this section, while VL is the liquid molar
volume at 25 C.
Temperature-dependent Properties
The ideal-gas enthalpy (needed for equation-of-state calculations) is calculated from the method of Black and Twu developed in 1983. The method was
an extension of work done by Lee and Kesler and involved fitting a wide variety of ideal-gas heat capacity data for hydrocarbons from the API 44 project
and other sources. The equation (which produces enthalpies in Btu/lb and
uses temperatures in degrees Rankine) is as follows:
(20)

A3 2 A4 3
T +
T
2
3

H = A1 + A2 T +
A2 = C1 + C7C4

(21)

A3 = C2 + C7C5

(22)

A4 = C3 + C7C6

(23)

C1 = 0.33886 + 0.02827 K

(24)
2

C2 = (0.9291 1.1543K + 0.0368K )10


7

(25)

C3 = 1.665810

(26)

C4 = (0.26105 0.59332K)

(27)

C5 = 4.9210

(28)
7

(29)

C6 = (0.536 0.6828K)10

C7 = [(12.8 K) (10 K) 10K]

(30)

The Watson characterization factor, K, is defined as:


K=

1 3

NBP
SG

(31)

where:
NBP = normal boiling point in degrees Rankine
SG =

I-12

Petroleum Components

specific gravity

May 1994

Section 1.1

Component Data

The constant A1 in equation (20) is determined so as to give an enthalpy of zero


at the arbitrarily chosen zero for enthalpy, which is the saturated liquid at 0 C.
The latent heat of vaporization as described below (to get from saturated liquid to saturated vapor) and the SRK equation of state (to get from saturated
vapor to ideal gas) are used to compute the enthalpy departure between this
reference point and the ideal-gas state.
The vapor pressure is calculated from the reduced vapor-pressure equation
(14) used above in the calculation of the acentric factor. The latent heat of
vaporization is also calculated from equation (14). The saturated liquid enthalpy is calculated by computing the departure from the ideal-gas enthalpy,
as a sum of the latent heat and the enthalpy departure (computed with the
SRK equation of state) for the saturated vapor. The saturated liquid density is
computed by applying the Rackett equation (see Section 1.2.3, Generalized
Correlation Methods) to saturated temperature and pressure conditions as
predicted from the vapor-pressure equation (14).

Property
Generation-CAVETT Method

Critical Properties and Acentric Factor


Optionally, the user may choose to compute critical properties from the methods developed in 1962 by Cavett. This option is called the CAVETT method.
The equations are:
2

Tc = 768.07121 + 1.7133693Tb 0.0010834003Tb 0.0089212579(API)Tb


6 3

(32)

2 2

+ 0.3889058410 Tb + 0.5309492(API)Tb + 0.32711610 (API) Tb


3

5 2

(33)

log10Pc = 2.8290406 + 0.9412010910 Tb 0.3047474910 Tb


4

8 3

0.208761110 (API)Tb + 0.1518410310 Tb + 0.1104789910 (API)Tb


7

2 2

0.4827159910 (API) Tb + 0.1394961910 (API) Tb

where:
Tc =

critical temperature in degrees Rankine

Pc =

critical pressure in psia

Tb =

normal boiling point in degrees Fahrenheit

API =

API gravity

When the CAVETT characterization options are chosen, the acentric factor is
computed by a method due to Edmister (1958):
=

3 log10 Pc
1
7 (Tc Tb) 1

PRO/II Component and Thermophysical Properties


Reference Manual

(34)

Petroleum Components

I-13

Component Data

Section 1.1

In equation (34), Pc is in atmospheres. Finally, the critical volume is estimated from the following equation:
Zc =

15

Pc Vc

RTc

(35)

= 0.291 0.08

PRO/II Note: For more information on using the CAVETT characterization


method, see Section 15, Petroleum Component Properties, of the PRO/II
Keyword Input Manual.
Other Fixed Properties
When the CAVETT characterization option is chosen, the heat of formation
and solubility parameter are calculated exactly as in the SIMSCI method
above.
Temperature-dependent Properties
Ideal-gas enthalpies (in Btu/lb-mole) are computed with the following equations:

H = A + BT + CT + DT

(36)

A = a5MW a4 + a0

(37)

B = 459.67 (1379.01a1 2a2) + a3

(38)

C = a2 1379.01a1

(39)

11
8
D = a1 = 9.17510 API 5.63310 MW

(40)

a2 = 3.10710

5.83210

a3 = 1.702510

API MW

1.48910

5
API API 2.85710 API

+ 4.29310 3.08410
3

(41)
(42)

K K + 0.088) K 0.819 MW

a4 = 459.67 459.67 459.67a1 a2 + a3

(43)

a5 = (0.351K 8.953) K + 43.402 K + 188.25

(44)

+ 3.53510

3
API 0.053API 5.95610 K + 3.544 API

where:
T=

temperature in degrees Rankine

MW = molecular weight
API =

API gravity

K=

Watson K-factor defined by equation (31)

The constant a0 in equation (37) is determined so as to be consistent with the


arbitrary zero of enthalpy, which is the saturated liquid at 0 C.

I-14

Petroleum Components

May 1994

Section 1.1

Component Data

Vapor pressures (in psia) are computed from a generalized Antoine equation:
ln P = A +

A = ln(14.696)

B=

(45)

B
T 80.0

(46)

B
Tb 80.0

ln Pc ln(14.696)
1
1

Tc 80.0 Tb 80.0

(47)

Temperatures (including the critical temperature Tc and normal boiling point


Tb) are in degrees Rankine.
The saturated liquid density (in lb/ft3) is computed as follows:
= A + BT + CT

(48)

A = 1.328 L,60

(49)

B=

C=

0.3076 L,60
Tc

(50)

0.3989 L,60

(51)

2
Tc

where:
L,60 = liquid density at 60 F, calculated from the specific gravity
and the density of water
Temperatures are in degrees Rankine
The latent heat of vaporization (in Btu/lb-mole) is calculated from a combination of the Watson equation (Watson, 1943, Thek and Stiel, 1966), for the temperature variation of the heat of vaporization, and the expression of
Kistiakowsky (1923), for the heat of vaporization at the normal boiling point:
Tc T
Hvap = H0

Tc Tb

(52)

H0 = Tb 7.58 + 4.571 lnTb

(53)

0.38

The critical temperature Tc, normal boiling point Tb, and temperature T are
all in degrees Rankine.
The saturated liquid enthalpy is estimated with the correlation of Johnson
and Grayson. This method is discussed in Section 1.2.3, Generalized Correlation Methods. A constant is added so that the saturated liquid enthalpy is
zero at 0 C.

PRO/II Component and Thermophysical Properties


Reference Manual

Petroleum Components

I-15

Component Data

Property
Generation--LeeKesler Method

Section 1.1

Critical Properties and Acentric Factor


Kesler and Lee used the following equations in 1976 to correlate critical temperatures and critical pressures of hydrocarbons:
(54)
5

Tc = 341.7 + 811 SG + (0.4244 + 0.1174 SG)Tb + (0.4669 3.2623 SG)10 Tb


(55)
2
3
lnPc = 8.3634 0.0566 SG 0.24244 + 2.2898 SG + 0.11857 SG 10 Tb +

1.4685 + 3.648 SG + 0.47227 SG2107 T2b 0.42019 + 1.6977 SG21010 T3b

where:
Tc, Tb = critical and normal boiling temperatures (both in degrees
Rankine)
Pc =

critical pressure in psia

SG =

specific gravity

The acentric factor is estimated from an equation in an earlier work by Lee


and Kesler (1975):
6

lnPR,b 5.92714 + 6.09648 TR,b + 1.28862 lnTR,b 0.169347 TR,b

(56)

15.2518 15.6875 TR,b 13.4721 lnTR,b + 0.43577 TR,b


6

where:
subscripts R,b indicate reduced properties evaluated at the normal
boiling point
The critical volume is then estimated from the following equation:
Zc =

Pc Vc
= 0.2905 0.085
RTc

(57)

Other Fixed Properties


When the Lee-Kesler characterization option is chosen, the heat of formation
and the solubility parameter are calculated exactly as in the SIMSCI method
described previously.
Temperature-dependent Properties
Ideal-gas enthalpies (in Btu/lb-mole) are computed by integrating the following equation for the ideal-gas heat capacity:
(58)

2
4
7 2
Cp = 0.33886 + 0.02827K 0.9291 1.1543K + 0.0368K 10 T 1.665810 T
4
7 2
(CF)0.26105 0.59332 (4.56 9.48)10 T (0.536 0.6828)10 T

I-16

Petroleum Components

May 1994

Section 1.1

Component Data

The factor CF is given by:


CF = [(12.8 K)(10 K) (10)]

(59)

where:
K=

Watson K-factor defined by equation (31)

T=

temperature in degrees Rankine

acentric factor as calculated by equation (56)

The constant of integration is determined so as to give an enthalpy of zero at the


arbitrarily chosen basis for enthalpy, which is the saturated liquid at 0 C.
When the Lee-Kesler characterization option is chosen, the vapor pressure,
saturated liquid density, saturated liquid enthalpy, and latent heat of vaporization are all calculated by the methods used for CAVETT characterization, as
described in the previous section.
References
1.

Black, C., and Twu, C.H., 1983, Correlation and Prediction of


Thermodynamic Properties for Heavy Petroleum, Shale Oils, Tar Sands
and Coal Liquids, paper presented at AIChE Spring Meeting, Houston,
March 1983.

2.

Cavett, R.H., 1962, Physical Data for Distillation Calculations Vapor-Liquid Equilibria, 27th Mid-year Meeting of the API Division of
Refining, 42[III], 351-357.

3.

Edmister, W.C., 1958, Applied Hydrocarbon Thermodynamics, Part 4:


Compressibility Factors and Equations of State, Petroleum Refiner,
37(4), 173.

4.

Kesler, M.G., and Lee, B.I., 1976, Improve prediction of enthalpy of


fractions, Hydrocarbon Proc., 53(3), 153-158.

5.

Kistiakowsky, W., 1923, Z. Phys. Chem., 107, 65.

6.

Lee, B.I., and Kesler, M.G., 1975, A Generalized Thermodynamic


Correlation Based on Three-Parameter Corresponding States, AIChE J.,
21, 510-527.

7.

Thek, R.E., and Stiel, L.I., 1966, A New Reduced Vapor Pressure Equation,
AIChE J., 12, 599-602.

8.

Twu, C.H., 1984, An Internally Consistent Correlation for Predicting the


Critical Properties and Molecular Weights of Petroleum and Coal-tar
Liquids, Fluid Phase Equil., 16, 137-150.

9.

Watson, K.M., 1943, Ind. Eng. Chem., 35, 398.

PRO/II Component and Thermophysical Properties


Reference Manual

Petroleum Components

I-17

Component Data

1.1.3

Section 1.1

Assay Processing
General
Information

Hydrocarbon streams may be defined in terms of laboratory assay data. Typically, such an assay would consist of distillation data (TBP, ASTM D86,
ASTM D1160, or ASTM D2887), gravity data (an average gravity and possibly a gravity curve), and perhaps data for molecular weight, lightends components, and special refining properties such as pour point and sulfur content.
This information is used by PRO/II to produce one or more sets of discrete
pseudocomponents which are then used to represent the composition of each
assay stream.
The process by which assay data are converted to pseudocomponents can be analyzed in terms of several distinct steps. Before each of these is examined in detail, it will be useful to list briefly each step of the process in order:
The user defines one or more sets of TBP cutpoints (or accepts the default set of cutpoints that PRO/II provides). These cutpoints define the
(atmospheric) boiling ranges that will ultimately correspond to each
pseudocomponent. Multiple cutpoint sets (also known as blends) may
also be defined to better model different sections of a process.
Each set of user-supplied distillation data is converted to a TBP (True
Boiling Point) basis at one atmosphere (760 mm Hg) pressure.
The resulting TBP data are fitted to a continuous curve and then the program cuts each curve to determine what percentage of each assay
goes into each pseudocomponent as defined by the appropriate cutpoint
set. Gravity and molecular weight data are similarly processed so that
each cut has a normal boiling point, specific gravity, and molecular
weight. During this step, the lowest-boiling cuts may be eliminated or
modified to account for any lightends components input by the user.
Within each cutpoint set, all assay streams using that set (unless they are
explicitly excluded from the blending - this is described later) are combined to get an average normal boiling point, gravity, and molecular
weight for each of the pseudocomponents generated from that cutpoint
set. These properties are then used to generate all other properties (critical properties, enthalpy data, etc.) for that pseudocomponent.
Note: Special refinery properties such as cloud point and sulfur content may
also be defined within assays. The distribution of these properties into pseudocomponents and their subsequent processing by the simulator is outside the
scope of this chapter but will be covered in a later document.

I-18

Assay Processing

May 1994

Section 1.1

Cutpoint Sets
(Blends)

Component Data

Defining Cutpoints
In any simulation, there is always a primary cutpoint set, which defaults
as shown in Table 1.1.3-1.
Table 1.1.3-1: Primary TBP Cutpoint Set
TBP Range, oF

Number of
Components

Width, oF

100-800

28

25

800-1200

50

1200-1600

100

The primary cutpoints shown in Table 1.1.3-1 may be overridden by supplying a new set for which no name is assigned. In addition, secondary sets
of cutpoints may be supplied by supplying a set and giving it a name. The
blend with no name (primary cutpoint set) always exists (even if only named
blends are specifically given); there is no limit to the number of named
blends (secondary cutpoint sets) that may be defined. The user may designate
one cutpoint set as the default; if no default is explicitly specified, the primary cutpoint set will be the default. Each cutpoint set (if it is actually used
by one or more streams) will produce its own set of pseudocomponents for
use in the flowsheet.

15

PRO/II Note: For more information on specifying TBP cutpoint sets, see Section
15, Petroleum Component Properties, of the PRO/II Keyword Input Manual.
Association of Streams With Blends
Each assay stream is associated with a particular blend. By default, an assay
stream is assigned to the default cutpoint set. A stream may be associated
with a specific secondary cutpoint set by explicitly specifying the name of
that cutpoint set (blend) in association with the stream. If the assay stream is
associated with a blend name not given for any cutpoint set previously defined, a new blend with that name is created using the same cutpoints as the
primary cutpoint set. The user may also specify that a stream use a certain set
of cutpoints but not contribute to the blended properties of the pseudocomponents generated from that set (this might be appropriate if an estimate were
being supplied for a recycle stream, for example). This is done by selecting
the XBLEND option, which excludes the stream in question from the blending. The default is for the stream to be included in the blending for the purposes of pseudocomponent property generation; this is called the BLEND
option. It is not allowed for the XBLEND option to be used on all streams associated with a blend, since at least one stream must be blended in to define
the pseudocomponent properties. The blending logic is best illustrated by an
example:

PRO/II Component and Thermophysical Properties


Reference Manual

Assay Processing

I-19

Component Data

Section 1.1

Suppose that two secondary cutpoint sets A1 and A2 were defined, and that
A1 was designated as the default. This means that three sets actually exist,
since the primary cutpoint set supplied by PRO/II still exists (though it is no
longer the set with which streams will be associated by default). Now, suppose the following streams (where extraneous information like the initial conditions is not shown) are given:
Table 1.1.3-2: Blending Example
Stream

Blend Option

Blend Name

S1

none given
(defaulted to
BLEND)

none given
(defaulted to A1)

S2

XBLEND

none given
(defaulted to A1)

S3

XBLEND

A1

S4

BLEND

A2

S5

BLEND

B1

S6

XBLEND

B1

S7

BLEND

B2

Streams S1 and S2 will use the pseudocomponents defined by secondary cutpoint set A1, since it is the default. S3 will also use A1s pseudocomponents
since it is specified directly. The pseudocomponents in blend A1 will have
properties determined only by the cuts from stream S1, since the XBLEND
option was used for S2 and S3. Stream S4 will use the pseudocomponents defined by cutpoint set A2. Streams S5 and S6 will go into a new blend B1
which will use the cutpoints of the primary cutpoint set. Since XBLEND is
used for stream S6, only stream S5s cuts will be used to determine the properties of the pseudocomponents in blend B1. Finally, stream S7 will use another new blend, B2, also with the cutpoints from the primary cutpoint set.
Since it is a different blend, however, the pseudocomponents from blend B2
will be completely distinct (even though they will use the same cutpoint
ranges) from those of blend B1.

15

I-20

Assay Processing

PRO/II Note: For more information on blending options for assay streams, see
Section 15, Petroleum Component Properties, of the PRO/II Keyword Input
Manual.

May 1994

Section 1.1

Component Data

Application Considerations
The selection of cutpoints is an important consideration in the simulation of
hydrocarbon processing systems. Too few cuts can result in poor representation of yields and stream properties when distillation operations are
simulated; moreover, desired separations may not be possible because of
component distributions. On the other hand, the indiscriminate use of cuts
not needed for a simulation serves only to increase the CPU time unnecessarily. It is wise to examine the cut definition for each problem in light of simulation goals and requirements. The default primary cutpoint set in PRO/II
represents, in our experience, a good selection for a wide range of refinery
applications.
In some circumstances, it may be desirable to use more than one cutpoint set
in a given problem. This multiple blends functionality is useful when different portions of a flowsheet are best represented by different TBP cuts; for
example, one part of the process may have streams that are much heavier
than another and for which more cutpoints at higher temperatures would be
desirable. It is also useful when hydrocarbon feeds to a flowsheet differ in
character; for example, different blends might be used to represent an aromatic stream (producing pseudocomponents with properties characteristic of
aromatics) and a paraffinic stream feeding into the same flowsheet. The extra
detail and accuracy possible with this feature must be balanced against the increase in CPU time caused by the increased number of pseudocomponents.

Interconversion of
Distillation Curves

Types of Distillation Curves


Assays of hydrocarbon streams are represented by distillation curves. A distillation curve represents the amount of a fluid sample that is vaporized as
the temperature of the sample is raised. The temperature where the first vaporization takes place is referred to as the initial point (IP), and the temperature at which the last liquid vaporizes is called the end point (EP). Each data
point represents a cumulative portion (usually represented as volume percent) of the sample vaporized when a certain temperature is reached.
Estimation of thermophysical properties for the pseudocomponents requires
(among other things) a distillation curve that represents the true boiling point
(TBP) of each cut in the distillation. However, rigorous TBP distillations are
difficult and not well standardized so it is common to perform some other
well-defined distillation procedure; standard methods are defined by the
American Society for Testing and Materials (ASTM). The ASTM procedures
most commonly used for hydrocarbons are D86, D1160, and D2887.

PRO/II Component and Thermophysical Properties


Reference Manual

Assay Processing

I-21

Component Data

Section 1.1

ASTM D86 distillation is typically used for light and medium petroleum
products and is carried out at atmospheric pressure. D1160 distillation is
used for heavier petroleum products and is often carried out under vacuum,
sometimes at absolute pressures as low as 1 mm Hg. The D2887 method uses
gas chromatography to produce a simulated distillation curve; it is applicable
to a wide range of petroleum systems. D2887 results are always reported by
weight percent; other distillations are almost always reported on a volume
percent basis. More details on these distillation procedures may be found in
the API Technical Data Book; complete specifications are given in volume 5
(Petroleum Products and Lubricants) of the Annual Book of ASTM Standards.
Conversion of D1160 Curves
PRO/II converts D1160 curves to TBP curves at 760 mm Hg using the threestep procedure recommended in the API Technical Data Book:
Convert to D1160 at 10 mm Hg using API procedure 3A4.1 (which in
turn references procedure 5A1.13). This procedure is expressed as a way
to estimate a vapor pressure at any temperature given the normal boiling
point, but the same equations may be solved to yield a normal boiling
temperature given the boiling temperature at another pressure. The equations used are as follows:
log10P =

3000.538X 6.76156
for X > 0.0022 (P < 2 mm Hg)
43X 0.987672

log 10P =

2663.129 X 5.994296
0.0013 X 0.0022
95.76X 0.972546

(1)

(2)

(2 P 760 mm Hg)
log10P =

2770.085X 6.412631
for X < 0.0013 (P > 760 mm Hg)
36X 0.989679

(3)

where:
P* =

vapor pressure in mm Hg at temperature T (in degrees


Rankine)

The parameter X is defined by:


Tb
X=

0.0002867Tb

(4)

T
748.1 0.2145Tb

where:
Tb =

I-22

Assay Processing

boiling point (in degrees Rankine) at a pressure of 760 mm


Hg

May 1994

Section 1.1

Component Data

For conversions where neither pressure is 760 mm Hg, the conversion may
be made by applying the above equations twice in succession, using 760 mm
Hg as an intermediate point:
Convert to TBP at 10 mm Hg using API Figure 3A2.1 (which has been
converted to equation form by SimSci).
Convert to TBP at 760 mm Hg using API procedure 3A4.1.
Conversion of D2887 Curves
PRO/II converts D2887 simulated distillation data to TBP curves at 760 mm Hg
using the two-step procedure recommended in the API Technical Data Book:
Convert to D86 at 760 mm Hg using API procedure 3A3.1. This procedure
converts D2887 Simulated Distillation (SD) points (in weight percent) to
D86 points (in volume percent) using the following equation:
b c

(5)

D86 = a(SD) F

where:
D86 and SD = the ASTM D86 and ASTM D2887 temperatures in degrees Rankine at each volume percent (for D86) and the
corresponding weight percent (for SD), and a, b, and c
are constants varying with percent distilled according
to Table 1.1.3-3.
Table 1.1.3-3: Values of Constants a, b, c
Percent
Distilled

6.0154

0.7445

0.2879

10

4.2262

0.7944

0.2671

30

4.8882

0.7719

0.3450

50

24.1357

0.5425

0.7132

70

1.0835

0.9867

0.0486

90

1.0956

0.9834

0.0354

100

1.9073

0.9007

0.0625

The parameter F in equation (5) is calculated by the following equation:


0.05434

F = 0.009524(SD 10%)

(6)

0.6147

(SD 50%)

where:
SD10% and SD50% = D2887 temperatures in degrees Rankine at the 10% and
50% points, respectively

PRO/II Component and Thermophysical Properties


Reference Manual

Assay Processing

I-23

Component Data

Section 1.1

Convert to TBP at 760 mm Hg using API procedure 3A1.1, which is


described in the section Conversion of D86 Curves with New (1987) API
Method below.
Conversion of D86 Curves
PRO/II has three options for the conversion of D86 curves to TBP curves at 760
mm Hg. These are the currently recommended (1987) API method, the older
(1963) API method, and the Edmister-Okamoto correlation. In addition, a correction for cracking may be applied to D86 data; this correction was recommended
by the API for use with their older conversion procedure, but is not recommended for use with the current (1987) method. The conversion of D86
curves takes place in the following steps:
If a cracking correction is desired, correct the temperatures above 475 F as
follows:
Tcorr = Tobs + D ; log10D = 1.587 + 0.00473Tobs

(7)

where:
Tcorr, Tobs = the corrected and observed temperatures, respectively,
in degrees Fahrenheit.
If necessary, convert the D86 curve at pressure P to D86 at 760 mm Hg
with the standard ASTM correction factor:
T760 = TP + 0.00012(760 P)(460 + TP)

(8)

where:
TP =

D86 temperature in Fahrenheit at pressure P

T760 = D86 temperature in Fahrenheit at 760 mm Hg


Convert from D86 at 760 mm Hg to TBP at 760 mm Hg using one of the
three procedures below.
a)

Conversion of D86 Curves with New (1987) API Method


By default, PRO/II converts ASTM D86 distillation curves to TBP curves
at 760 mm Hg using procedure 3A1.1 (developed by Riazi and Daubert in
1986) recommended in the 5th edition of the API Technical Data Book.
The equation for this procedure is as follows:
TBP = a(D86)

(9)

where a and b are constants varying with percent of liquid sample distilled
as given in Table 1.1.3-4:

I-24

Assay Processing

May 1994

Section 1.1

Component Data
Table 1.1.3-4: Values of Constants a, b
Percent Distilled

0.9167

1.0019

10

0.5277

1.0900

30

0.7429

1.0425

50

0.8920

1.0176

70

0.8705

1.0226

90

0.9490

1.0110

100

0.8008

1.0355

b) Conversion of D86 Curves with Old (1963) API Method


This method, while no longer the default, is still available for users whose
flowsheets may be tuned to the results using the old method. This method
was recommended (and shown in graphical form) in older editions of the
API Technical Data Book. The graphical correlation has been converted to
equation form by SimSci.
c)

Cutting TBP Curves

Conversion of D86 Curves with Edmister-Okamoto Method


Edmister and Okamoto (1959) developed a method which is still widely
used for converting ASTM D86 curves to TBP curves. If the EdmisterOkamoto method is specified as the conversion method, their procedure
(converted from the original graphical form to equations by SimSci) is
used for conversion of D86 to TBP curves.

Fitting of Distillation Curves


Before a curve is cut into pseudocomponents, the distillation data must be fitted to a continuous curve. This is necessary because the supplied data points
will not in general correspond to the desired cutpoints. PRO/II offers three
methods for fitting distillation curves.
The default is the cubic spline method (known as the SPLINE option). A cubic spline function is used to fit all given volume percents between the first
and last points. Beyond those bounds, points 1 and 2 and points N and N-1
are used to define a normal distribution function to extrapolate to the 0.01%
and 99.99% points, respectively. If only two points are supplied, the entire
curve is defined by the distribution function fit. This extrapolation feature is
particularly valuable when extrapolating heavy ends distillations which often
terminate well below 50 volume percent. This method in general results in an
excellent curve fit. The only exception is when the distillation data contain a
significant step function (such a step is often the unphysical result of an error
in obtaining or reporting the data); in that case, the step creates an instability
that tends to propagate throughout the entire length of the curve. Should this
happen, the input data should be checked for validity.

PRO/II Component and Thermophysical Properties


Reference Manual

Assay Processing

I-25

Component Data

Section 1.1

The quadratic fit method (known as the QUADRATIC option) provides a successive quadratic approximation to the shape of the input assay curve. This
method is recommended in the rare case (see above) where a cubic spline fit
is unstable.
The Probability Density Function (PDF) method (known as the PDF option)
is different in that it does not necessarily pass through all the points input by
the user. Instead, it fits a probability density function to all points supplied.
The resulting curve will maintain the probability-curve shape characteristic
of petroleum distillations, while minimizing the sum of the squares of the differences between the curve and the input data. If desired, the curve may be
constrained to pass through either or both of the initial point and end point.
The PDF method is recommended whenever it is suspected that the distillation data are noisy, containing significant random errors.
It is worth noting that the choice of curve-fitting procedure will also have a
slight impact on the distillation interconversions described in the previous section. That is because most of the conversion procedures work by doing the
conversion at a fixed set of volume percents, which must be obtained by interpolation and sometimes extrapolation, using some curve-fitting procedure.
Division into Pseudocomponents
Once a smooth distillation curve is obtained, the volume percent distilled at each
cutpoint is determined. The differences between values at adjacent cutpoints define the percent of the streams volume that is assigned to the pseudocomponent
defined by the interval between two adjacent cutpoints. For example, using the
default set of cutpoints shown in Table 1.1.3-1, the first pseudocomponent would
contain all material boiling between 100 F and 125 F, the second would contain the material boiling between 125 F and 150 F, and so forth. Material boiling above the last cutpoint (1600 F) would be combined with the last
(1500-1600) cut, while (with the exception of lightends as discussed below)
material boiling below 100 F would be combined with the first cut. If the distillation data do not extend into all of the cut ranges (in this example, if the initial
point were higher than 125 F or if the end point were lower than 1500 F), the
unused cuts are omitted from the simulation.
The normal boiling point (NBP) of each cut is determined as a volume-fraction average (or, in rare cases where TBP, D86, or D1160 distillations are
entered on a weight basis, as a weight-fraction average) by integrating across
the cut range. For small cut ranges, this will closely approach other types of
average boiling points. These average boiling points are used (possibly after
blending with cuts from other assay streams in the flowsheet) as correlating
parameters when calculating other thermophysical properties for each
pseudocomponent.

I-26

Assay Processing

May 1994

Section 1.1

Component Data

These procedures are demonstrated in Figure 1.1.3-1 for a fictitious assay


with an IP of 90 F being cut according to the default cutpoint set (Table
1.1.3-1); for simplicity only the first ten percent of the curve is shown. In addition to its range, the first cut picks up the portion boiling below 100 F, and
its average boiling point (about 110 F in this case) is determined by integrating the curve from the IP to the 125 F point. The second cut is assigned the
material boiling from 125 F to 150 F, which is integrated to get a NBP of
approximately 138 F. The third and subsequent cuts are generated in a similar manner.
Figure 1.1.3-1:
Cutting TBP Curves

Gravity Data
PRO/II requires the user to enter an average gravity (either as a Specific
Gravity, API Gravity, or Watson K-factor) for each assay. If a Watson K is
given, it is converted to a gravity using the TBP data for the curve. Entry of a
gravity curve is recommended but not required.
If a user-supplied gravity curve does not extend to the 95% point, quadratic
extrapolation is used to generate an estimate for the gravity at the 100%
point. A gravity for each cut is determined at its mid-point, and an average
gravity for the stream is computed. If this average does not agree with the
specified average, the program will either normalize the gravity curve (if
data are given up to 95%) or adjust the estimated 100% point gravity value to
force agreement. Since the latter could in some cases result in unreasonable
gravity values for the last few cuts, the user should consider providing an estimate of the 100% point gravity value and letting the program normalize the
curve, particularly when gravity data are available to 80% or beyond.

PRO/II Component and Thermophysical Properties


Reference Manual

Assay Processing

I-27

Component Data

Section 1.1

If no gravity curve is given, the program will generate one from the specified
average gravity. The default method for doing this is referred to as the WATSONK method. For a pure component, the Watson K-factor is defined by the
following equation:
K=

13

NBP
SG

(10)

where:
NBP = normal boiling point in degrees Rankine
SG =

specific gravity at 60 F relative to H2O at 60 F

For a mixture (such as a petroleum cut), the NBP is traditionally replaced by a


more complicated quantity called the mean average boiling point (MeABP). For
this purpose, however, it is sufficient to simply use the volume-averaged boiling point computed from the distillation curve. The gravity curve is generated by assuming a constant value of the Watson K, applying equation (10)
to each cut to get a gravity, averaging these values, and then adjusting the assumed value of the Watson K until the resulting average gravity agrees with
the average gravity input by the user.
Another method (known as the PRE301 option) is available primarily for compatibility with older versions. It is similar to the preferred method described
above, except that the average Watson K is estimated from the 10, 30, 50, 70,
and 90 percent points on a D86 curve (which can be obtained from the TBP
curve by reversing one of the procedures in the previous section) and then
applied to the NBP of each TBP cut to generate a gravity curve. This curve is
then normalized to produce the specified average gravity.
The preferred method (constant Watson K applied to TBP curve) is justified
by the observation that, for many petroleum crude streams, the Watson K of
various petroleum cuts above light naphtha tends to remain fairly constant.
For other types of petroleum streams, however, this assumption is often incorrect. Hence, for truly accurate simulation work, the user is advised to supply
gravity curves whenever possible.
Molecular Weight Data
In addition to the NBP and specific gravity, simulation with assays requires
the molecular weight of each cut. These may be omitted completely by the
user, in which case they are estimated by the program.
The user may supply a molecular weight curve, which is quadratically interpolated and extrapolated to cover the entire range of pseudocomponents. Optionally, the user may also supply an average molecular weight. In that case,
the molecular weight value for the last cut is adjusted so that the curve
matches the given average, or if the 100% value is provided, the entire molecular weight curve is normalized to match the given average.

I-28

Assay Processing

May 1994

Section 1.1

Component Data

If no molecular-weight data are supplied, the molecular weights are estimated; the default method is a proprietary modification (known as the SIMSCI method) of the method developed by Twu (1984). This method is a
perturbation expansion with the normal alkanes as a reference fluid. Twus
method was originally developed to be an improvement over Figure 2B2.1 in
older editions of the API Technical Data Book. That figure relates molecular
weight to NBP and API gravity for NBPs greater than 300 F. The SIMSCI
method matches that data between normal boiling points of 300 F and 800
F, and better extrapolates outside that temperature range.
The unaltered old API method is (API63) is also available.
A newer API method, called the extended API method (known as the EXTAPI
option), is also available. This is API procedure 2B2.1, and it is an extension of
the earlier API method which better matches known pure-component data below
300 F. The equation is as follows:
(11)
4
3
1.26007
4.98308
MW = 20.486 exp (1.165 10 Tb 7.78712 SG + 1.1582 10 Tb SG) Tb
SG

where:
SG =

specific gravity of the pseudocomponent

Tb =

normal boiling point in degrees Rankine

Lightends Data
Hydrocarbon streams often contain significant amounts of light hydrocarbons
(while there is no universal definition of light, C6 is a common upper limit).
Simulation of such systems is more accurate if these components are considered
explicitly rather than being lumped into pseudocomponents. If the distillation
curve is reported on a lightends-free basis, the light components can be fed to
the flowsheet in a separate stream and handled in a straightforward manner. Typically, however, the lightends make up the initial part of the reported distillation
curve, and adjustment of the cut-up curves is required to avoid double-counting
the lightends components.
By default, the program matches user-supplied lightends data to the TBP curve.
The user-specified rates for all lightends components are adjusted up or down, all
in the same proportion, until the NBP of the highest-boiling lightends component
exactly intersects the TBP curve. All of the cuts from the TBP curve falling into the
region covered by the lightends are then discarded and the lightends components
are used in subsequent calculations. This procedure is illustrated in Figure 1.1.3-2,
where lightend component flows are adjusted until the highest-boiling lightend
(nC5 in this example) has a mid-volume percent (point a) that exactly coincides
with the point on the TBP curve where the temperature is equal to the NBP of nC5.
The cumulative volume percent of lightends is represented by point b, and the
cuts below point b (and the low-boiling portion of the cut encompassing that point)
are discarded.

PRO/II Component and Thermophysical Properties


Reference Manual

Assay Processing

I-29

Component Data

Section 1.1

Figure 1.1.3-2:
Matching Lightends
to TBP Curve

Alternatively, the lightends may be specified as a fraction or percent (on a


weight or liquid-volume basis) of the total assay or as a fixed lightends flowrate.
In these cases, the input numbers for the lightends components can be normalized to determine the individual component flowrates. A final alternative is to
specify the flowrate of each lightends component individually.

Generating
Pseudocomponent
Properties

Once each curve is cut, the program processes each blend to produce average
properties for the pseudocomponents from each cutpoint interval in that
blend. All the streams in a given blend (except for those for which the
XBLEND option was used) are totaled to get the weights, volumes, and
moles for each cutpoint interval. Using the above totals, the average molecular weight and gravity are calculated for each cut range. Finally, the normal
boiling point for each pseudocomponent is calculated by weight averaging
the individual values from the contributing streams.
Once the normal boiling point, gravity, and molecular weight are known for
each pseudocomponent, all other properties (critical properties, enthalpies,
etc.) are determined according to the characterization method selected by the
user (or defaulted by the program). These methods are described in Section
1.1.2, Petroleum Components.

Vapor Pressure
Calculations

I-30

Assay Processing

While not a part of the programs actual processing of assay streams, many
problems involving hydrocarbon systems will involve a specification on
some vapor pressure measurement. The two most common of these are the
True Vapor Pressure (TVP) and the Reid Vapor Pressure (RVP). PRO/II allows specification of these quantities from several unit operations, and they
may be reported in output in the Heating/Cooling Curve (HCURVE) utility
or as part of a user-defined stream report.

May 1994

Section 1.1

Component Data

True Vapor Pressure (TVP) Calculations


The TVP of a stream is defined as the bubble-point pressure at a given reference temperature. By default, that reference temperature is 100 F, but this
may be overridden by the user. The user may specify a specific thermodynamic system to be used in performing all TVP calculations in the flowsheet;
by default, the calculation for a stream is performed using the thermodynamic system used to generate that stream.
Reid Vapor Pressure (RVP) Calculations
The RVP laboratory procedure provides an inexpensive and reproducible measurement correlating to the vapor pressure of a fluid. The measured RVP is usually within 1 psi of the TVP of a stream. It is always reported as psi, although
the ASTM test procedures (except for D5191 which, as mentioned below, uses
an evacuated sample bomb) actually read gauge pressure. Since the air in the
bomb accounts for approximately 1 atm, the measured gauge pressure is a rough
measure of the true vapor pressure. Six different calculation methods are available. Within each calculation method, the answer will depend somewhat on the
thermodynamic system used. As with the TVP, the thermodynamic system for
RVP calculations may be specified explicitly or, by default, the thermodynamic
system used to generate the stream will be used.
The APINAPHTHA method calculates the RVP from Figure 5B1.1 in the API
Technical Data Book, which represents the RVP as a function of the TVP and
the slope of the D86 curve at the 10% point. The graphical data have been converted to equation form by Simsci. This method is the default for PRO/IIs RVP
calculations. It is useful for many gasolines and other finished petroleum products, but it should not be used for oxygenated gasoline blends.
The APICRUDE method calculates the RVP from Figure 5B1.2 in the API
Technical Data Book, which represents the RVP as a function of the TVP and
the slope of the D86 curve at the 10% point. The graphical data have been
converted to equation form by SimSci. It is primarily intended for crude oils.
The ASTM D323-82 method (known as the D323 method) simulates a standard ASTM procedure for RVP measurement. The liquid hydrocarbon portion
of the sample is saturated with air at 33 F and 1 atm pressure. This liquid is
then mixed at 100 F with air in a 4:1 volume ratio. Since the test chamber is
not dried in this procedure, a small amount of water is also added to simulate
this mixture. The mixture is flashed at 100 F at a constant volume (corresponding to the experiment in a sealed bomb), and the gauge pressure of the
resulting vapor-liquid mixture is reported as the RVP. Both air and water
should be in the component list for proper use of this method.
The obsolete ASTM D323-73 method (known as the P323 method) is available for compatibility with earlier versions of the program.

PRO/II Component and Thermophysical Properties


Reference Manual

Assay Processing

I-31

Component Data

Section 1.1

The ASTM D4953-91 method (known as the D4953 method) was developed
by the ASTM primarily for oxygenated gasolines. The experimental method
is identical to the D323 method, except that the system is kept completely
free of water. The algorithm for simulating this method is identical to that for
D323, except that no water is added to the mixture. Air should be in the component list for proper use of this method.
The ASTM D5191-91 method (known as the D5191 method) was developed as an
alternative to the D4953 method for gasolines and gasoline-oxygenate blends. In
this method, the air-saturated sample is placed in an evacuated bomb with five
times the volume of the sample, and then the total pressure of the sample is measured. In the simulator, this is accomplished by flashing, at constant volume, a mixture of 1 part sample (at 33 F and 1 atm) and 4 parts air (at the near-vacuum
conditions of 0.01 psia and 100 F). The resulting total pressure is then converted to a dry vapor pressure equivalent (DVPE) using the following equation:
DVPE = 0.965X A

(12)

where:
X=

the measured total pressure

A=

0.548 psi (3.78 kPa)

This number is then reported as the RVP. Air should be in the component list
for proper use of this method.
Comments on RVP and TVP Methods
Because of the sensitivity of the RVP (and the TVP) to the light components
of the mixture, these components should be modeled as exactly as possible if
precise values of RVP or TVP are important. This might mean treating more
light hydrocarbons as defined components rather than as pseudocomponents;
oxygenated compounds blended into gasolines should also be represented as
defined components rather than as part of an assay. It is also important to apply a thermodynamic method that is appropriate for the stream in question
(see Section 1.2.2, Application Guidelines). The thermodynamics becomes
particularly important for oxygenated systems, which are not well-modeled
by traditional hydrocarbon methods such as Grayson-Streed. These systems
are probably best modeled by an equation of state such as SRK with the SimSci alpha formulation and one of the advanced mixing rules (see Section
1.2.4, Equations of State). It is important to have binary interaction parameters between the oxygenates and the hydrocarbon components of the system.
PRO/IIs databanks contain many such parameters, but others may have to be
regressed to experimental data or estimated.
One should not be too surprised if calculated values for RVP differ from an
experimental measurement by as much as one psi. Part of this is due to the
uncertainty in the experimental procedure, and part is due to the fact that the
lightends composition inside the simulation may not be identical to that of
the experimental sample.

I-32

Assay Processing

May 1994

Section 1.1

Component Data

One of the less appreciated effects in experimental measurements is the presence of water, not only in the sample vessel, but also in the air in the form of
humidity. The difference between the D323 (a wet method) RVP and the
D4953 (a dry method) RVP will be approximately the vapor pressure of
water at 100 F (about 0.9 psi), with the D323 RVP being higher. Both of
these calculations assume that dry air is used in the procedure. The presence
of humidity in the air mixed with the sample can alter the D323 results, lowering the measured RVP because of the decreased driving force for vaporization of the liquid water. In the extreme case of 100% humidity, the D323
results will be nearly identical with the D4953 results. Therefore, a wet
test performed with air that was not dry would be expected to give results intermediate between PRO/IIs D323 and D4953 calculations. The results from
the D5191 method (both in terms of the experimental and calculated numbers) should in general be very close to D4953 results.
The primary application guideline for which RVP calculational model to use
is, of course, to choose the one that corresponds to the experimental procedure applied to that stream. Secondary considerations include limitations of
the individual methods. The APINAPHTHA and APICRUDE methods are
good only for hydrocarbon naphtha and crude streams, respectively. The
D323 method (and its obsolete predecessor, P323) is intended for hydrocarbon streams; the presence of water makes it less well-suited for use with
streams containing oxygenated compounds. The D4953 and D5191 methods
are both better suited for oxygenated systems, and calculations with these
methods should give similar results.
References
1.

American Petroleum Institute, 1988, Technical Data Book - Petroleum


Refining, 5th edition (also previous editions), American Petroleum
Institute, Washington, DC.

2.

American Society for Testing of Materials, Annual Book of ASTM


Standards, section 5 (Petroleum Products, Lubricants, and Fossil Fuels),
ASTM, Philadelphia, PA (issued annually).

3.

Edmister, W.C., and Okamoto, K.K., 1959, Applied Hydrocarbon


Thermodynamics, Part 12: Equilibrium Flash Vaporization Calculations
for Petroleum Fractions, Petroleum Refiner, 38(8), 117.

4.

Twu, C.H., 1984, An Internally Consistent Correlation for Predicting the


Critical Properties and Molecular Weights of Petroleum and Coal-tar
Liquids, Fluid Phase Equil., 16, 137-150.

PRO/II Component and Thermophysical Properties


Reference Manual

Assay Processing

I-33

This page left intentionally blank.

I-34

Assay Processing

May 1994

Section 1.2

1.2

Thermodynamic Methods

Thermodynamic Methods
PRO/II offers numerous methods for calculating thermodynamic properties such
as K-values, enthalpies, entropies, densities, gas and solid solubilities in liquids,
and vapor fugacities. These methods include:
Generalized correlations, such as the Chao-Seader K-value method and the API
liquid density method.
Equations of state, such as the Soave-Redlich-Kwong method for calculating
K-values, enthalpies, entropies, and densities.
Liquid activity coefficient methods, such as the Non-Random Two-Liquid
(NRTL) method for calculating K-values.
Vapor fugacity methods, such as the Hayden-OConnell method for dimerizing
species.
Special methods for calculating the properties of specific systems of components such as alcohols, glycols, and sour water systems.
Solid-liquid equilibria methods such as the vant Hoff method for calculating
the solubility of a solid in a liquid.
In addition, the electrolyte version of PRO/II contains a number of thermodynamic methods to handle systems containing aqueous ionic species.

PRO/II Component and Thermophysical Properties


Reference Manual

I-37

Thermodynamic Methods

1.2.1

Section 1.2

Basic Principles
General
Information

When modeling a single chemical process or an entire chemical plant, the use of
appropriate thermodynamic methods and precise data is essential in obtaining a
good design. PRO/II contains numerous proven thermodynamic methods for the
calculation of the following thermophysical properties:
Distribution of components between phases in equilibrium (K-values).
Liquid-phase and vapor-phase enthalpies.
Liquid-phase and vapor-phase entropies.
Liquid-phase and vapor-phase densities.

Phase
Equilibria

When two or more phases are brought into contact, material is transferred
from one to another until the phases reach equilibrium, and the compositions
in each phase become constant. At equilibrium for a multicomponent system,
the temperature, pressure, and chemical potential of component i is the same
in every phase, i.e.:

(1)

(2)

(3)

T =T ==T

P =P ==P

i = i = = i

where:
T=

system temperature

P=

system pressure

the chemical potential

, , ..., represent the phases


The fugacity of a substance is then defined as:
0
0
i i = RT lnfi / f i

(4)

where:
fi =
0

I-38

Basic Principles

fugacity of component i

fi =

standard-state fugacity of component i at T, P

0i =

standard-state chemical potential of component i at T, P

May 1994

Section 1.2

Thermodynamic Methods

It follows from (3) and (4) that the fugacities in each phase must also be equal:

(5)

fi = fi = = fi , i = 1,2, n

The fugacity of a substance can be visualized as a corrected partial pressure such that the fugacity of a component in an ideal-gas mixture is equal
to the component partial pressure.
For vapor-liquid equilibrium calculations, the ratio of the mole fraction of a component in the vapor phase to that in the liquid phase is defined as the K-value:
Ki yi / xi

(6)

where:
Ki =

K-value, or equilibrium ratio

yi =

mole fraction in the vapor phase

xi =

mole fraction in the liquid phase

For liquid-liquid equilibria, a corresponding equilibrium ratio or distribution


coefficient is defined:
I

(7)

II

KDi x i / x i

where:
KDi =

liquid-liquid distribution coefficient

I, II represent the two liquid phases


The vapor-phase fugacity coefficient of a component, Vi , is defined as the
ratio of its fugacity to its partial pressure, i.e.:
V

(8)

i fi / yiP

where:
Vi =

vapor-phase fugacity coefficient of component i

If a liquid activity coefficient method is used in the liquid phase calculation,


then the activity coefficient of the liquid phase can be related to the liquid
fugacity by the following relationship:
L

OL

f i = i xi f i

(9)

where:
Li =
0L

fi =

PRO/II Component and Thermophysical Properties


Reference Manual

liquid-phase activity coefficient


standard-state fugacity of pure liquid i

Basic Principles I-39

Thermodynamic Methods

Section 1.2

With this definition of liquid fugacity, iL 1 as xi 1. The standard-state


fugacity is as follows:
P L

0L
sat sat
f i = Pi i exp sat vi / RT dP
Pi

(10)

where:
Psat
i = saturated vapor pressure of component i at T
R=

gas constant

vLi

liquid molar volume of component i at T and P

sat
i

= fugacity coefficient of pure component i at T and Pisat

Equation (10) provides two correction factors for the pure liquid fugacity. The
fugacity coefficient, isat, corrects for deviations of the saturated vapor from
ideal-gas behavior. The exponential correction factor, known as the Poynting correction factor, corrects for the effect of pressure on the liquid fugacity. The
Poynting correction factor is usually negligible for low and moderate pressures.
Combining equations (6), (8), and (9) yields:
L OL

Ki = i f i / P i

(11)

Combining equations (7) and (9) yields:


L

KDi = i II / i I

(12)

If an equation of state is applied to both vapor and liquid phases, the vaporliquid K-values can be written as:
V

Ki = i / i

(13)

The liquid-liquid equilibria can be written as:


L

KDi =

i II

(14)

iI

Equations (11), (12), (13), and (14) are used to calculate the distribution of
components between phases.
For vapor-liquid equilibria, equation-of-state methods may be used to calculate
the fugacity coefficients for both liquid and vapor phases using equation (13).
One important limitation of equation-of-state methods is that they have to be applicable over a wide range of densities, from near-zero density for gases to high
liquid densities, using constants obtained from pure-component data. Equations
of state are not very accurate for nonideal systems unless combined with component mixing rules and alpha formulations (see Section 1.2.4, Equations of State)
appropriate for those components.

I-40

Basic Principles

May 1994

Section 1.2

Thermodynamic Methods

Equation (11) may be solved by using equation-of-state methods to calculate vapor fugacities combined with liquid activity methods to compute liquid activity
coefficients (see Section 1.2.6, Liquid Activity Methods). Liquid activity methods are most often used to describe the behavior of strongly nonideal mixtures.
References

Enthalpy

1.

Prausnitz, J. M., Lichtenthaler, R. N., and Gomes de Azevedo, E., 1986, Molecular
Thermodynamics of Fluid-Phase Equilibria, 2nd Ed., Prentice-Hall, N.Y.

2.

Sandler, S.I., 1989, Chemical and Engineering Thermodynamics, 2nd ed.,


John Wiley & Sons, New York.

3.

Smith, J.M. and Van Ness, H.C., 1987, Introduction to Chemical Engineering
Thermodynamics, 4th ed., McGraw-Hill, New York.

4.

Van Ness, H.C. and Abbott, M.M., 1982, Classical Thermodynamics of


Nonelectrolyte Solutions: With Applications to Phase Equilibria, McGrawHill, New York.

The enthalpy of a system, H, is defined in terms of the internal energy of the


system, U as follows:
(15)

H = U + PV

where:
H=

enthalpy of the system of nT moles

U=

internal energy of the system of nT moles

V=

total volume of the system

At constant temperature and pressure, the internal energy of the system is related to the volume by:
dU = T(P / T)V,n P dV
T

(16)

The enthalpy of the system is then given by:


H=

(17)

P T(P / T)V,n dV + RT(z 1) + H


T

where:
H* =

mixture ideal gas enthalpy = ni h0i

hi0 =

molar enthalpy of ideal gas i at temperature T

z=

compressibility factor PV nTRT

PRO/II Component and Thermophysical Properties


Reference Manual

Basic Principles I-41

Thermodynamic Methods

Section 1.2

PRO/II provides two distinct approaches to the calculation of enthalpy. For the
majority of thermodynamic systems of methods, enthalpy is calculated as a departure from the ideal-gas enthalpy of the mixture. Enthalpy departure functions
for both vapor and liquid phases are calculated by an equation of state or corresponding states model.
For liquid activity coefficient thermodynamic systems, however, PRO/II by
default invokes the LIBRARY thermodynamic method for vapor and liquid
enthalpy calculations.
The LIBRARY method consists of two correlations. The first correlates saturated-liquid enthalpy as a function of temperature, and the second correlates
latent heat of vaporization, also as a function of temperature. At temperatures
below the critical, vapor enthalpy is calculated by adding the latent heat to the
saturated liquid enthalpy at the system temperature. In other words, the vapor
enthalpy is the saturated vapor enthalpy at the system temperature. For both
phases, the pressure is implicitly the saturated vapor pressure at the system
temperature. No other pressure correction term is applied.
For almost all library components, the correlations in use for liquid enthalpy
can be used safely up to a reduced temperature, T/Tc , of approximately 0.9.
Tc is the temperature at the critical point, beyond which vapor and liquid become indistinguishable.

Note: The normal boiling point of a library component typically occurs when
the reduced temperature, Tr, is approximately equal to 0.7.
In general, the use of liquid activity coefficient models is not recommended
for system pressures above 1000 kPA. Below these conditions, the use of LIBRARY enthalpy methods will not introduce significant errors, provided that
the system temperature is below the critical temperatures of all components
present in significant quantities.
Quite often, however, we would like to use a liquid activity coefficient model
when permanent gases are present in the mixture. As the system temperature is
usually above the critical temperature of these gases, there is no standard-state
liquid fugacity at system conditions, so we replace that term by the Henrys Law
constant. However, the problem of adding the supercritical components contribution to the liquid enthalpies remains.
For the liquid-phase contribution, PRO/II extrapolates the components saturated liquid enthalpy curve linearly from the critical temperature. Above the
critical this extrapolation uses the slope of the library enthalpy tangent to the
liquid saturation curve at the normal boiling point. At temperatures above the
critical, there is no distinction between vapor and liquid phases and the vapor
enthalpy is set equal to the extrapolated liquid enthalpy. The point at which
the slope for linear extrapolation is obtained is chosen quite arbitrarily; as
mentioned, we use the normal boiling point temperature.

I-42

Basic Principles

May 1994

Section 1.2

Thermodynamic Methods

Note: At temperatures near Tc, the enthalpy of the saturated vapor for a pure component exhibits a decrease with temperature. This can lead to the computation of a
negative value of the constant-pressure heat capacity cp when using the LIBRARY
method for vapor enthalpies. This is entirely an artifact of the fact that the saturation curve is not a constant-pressure path. The printout of a negative heat capacity
is therefore a sign that the temperature is too high to be using LIBRARY vapor
enthalpies, and the user should switch to another method.
For low pressure and temperatures well below the lowest critical, LIBRARY enthalpies are often satisfactory. For high pressures or temperatures above the
critical of a component, it will usually be better to use an equation of state
for vapor and, possibly, liquid enthalpies. Beware, however, if a liquid activity coefficient method was selected for K-value; in such systems the traditional
cubic equation of state may not be capable of describing the liquid phase nonideality, and it is therefore unlikely that the equation of state will predict the correct
liquid-phase enthalpy. In this situation, one of the more advanced cubic equations using an alpha formulation, which correctly predicts pure-component vapor pressures is a better choice. As the contribution to the liquid enthalpy of
dissolved supercritical components is usually small, the LIBRARY method can
usually safely be used for liquid enthalpies. Ideal-gas based enthalpies and saturation enthalpies can be used in combination for vapor and liquid, respectively,
for defined components, because the ideal-gas enthalpy datum has been fixed
relative to the saturated-liquid enthalpy datum (HL = 0 at T = 273.15 K). For
components that are sub-critical at 273.15 K, the SRK vapor enthalpy departure
function, which applied to the ideal-gas enthalpy, gives the equivalent results as
adding the latent heat to zero-liquid enthalpy. For components that are supercritical at 273.15 K, using an alpha formulation will give consistent results between
departure-based and library enthalpies.

Entropy

The entropy of a system, S, is defined in terms of the enthalpy, as follows:

S = S R ln R ln P / Pref H H / T

(18)

and,

S = nisi R ni ln xi
0

(19)

where:
=

fugacity coefficient of mixture

Pref =

reference pressure of 1 atmosphere

mixture ideal gas entropy

si =

molar entropy of ideal gas i

ni =

moles of component i

xi =

mole fraction of component i

S =

PRO/II Component and Thermophysical Properties


Reference Manual

Basic Principles I-43

Thermodynamic Methods

Section 1.2

The ideal molar entropy is related to the ideal molar enthalpy by:
si =
0

T
Tref

(20)

0
c dT T =
/ hi / TP dT / T
pi
T
ref

where:
Tref =

reference temperature, 1 degree Rankine in PRO/II

cpi* =

ideal gas heat capacity of component i

Ideal gas entropy at the reference temperature is set equal to zero.


As for enthalpy computations, liquid and vapor entropies are calculated in
PRO/II using either an equation-of-state method such as SRKM or a generalized correlation method such as Curl-Pitzer.

Density

Cubic equation-of-state methods are generally not very accurate in predicting liquid densities. More accurate predictive methods have been developed
especially for liquid mixtures. Such methods include the API and Rackett
correlation methods. These methods are described in detail in Section 1.2.3,
Generalized Correlation Methods.
Vapor densities are computed in PRO/II using the following formulae:
v = zRT / P

(21)

v = MW / v

(22)

where:
v =

vapor density

MW = molecular weight
v=

molar vapor volume

z=

compressibility factor

Vapor densities can be predicted quite accurately using equation-of-state


methods, in addition to generalized correlation methods. The IDEAL
vapor-density method corresponds to z=1.

I-44

Basic Principles

May 1994

Section 1.2

1.2.2

Thermodynamic Methods

Application Guidelines
General
Information

Choosing an appropriate thermodynamic method for a specific application is


an important step in obtaining an accurate process simulation. Normally,
there may be any number of thermodynamic methods suitable for a given application. The user is left to use his or her best judgement, experience, and
knowledge of the available thermodynamic methods to choose the best
method.
It is important to note that, for many thermodynamic methods, the PRO/II
databanks contain adjustable binary parameters obtained from fitting published experimental and/or plant data. The thermodynamic method chosen
should ideally be used only in the temperature and pressure ranges at which
the parameters were regressed. Ideally, for each simulation, actual experimental or plant data should be regressed in order to obtain the best interaction parameters for the application.
There are several places where the user can find information and guidelines
on using the thermodynamic methods available in PRO/II. These are:
PRO/II Keyword Input Manual.
PRO/II Application Briefs Manual.
PRO/II Thermodynamic Expert System.
The first source outlines the PRO/II keywords needed to code the thermodynamic category in a PRO/II input file and provides some information on the
limitations and strengths of each thermodynamic method. The second source
shows how PRO/II is used to simulate many refinery, chemical, petrochemical, and solid processing applications using the thermodynamic methods appropriate to each system. The third source, available only in the PC (DOS)
version of PRO/II, is described below.

Thermodynamic
Expert System (TES)

The Thermodynamic Expert System (TES) is available only to PRO/II PC


(DOS) users who access PRO/II through the Graphics User Interface (GUI).
This system utilizes nearly 300 proven heuristics to determine the best thermodynamic method for a particular simulation. The following information is
used by the TES:

PRO/II Component and Thermophysical Properties


Reference Manual

Application Guidelines I-45

Thermodynamic Methods

Section 1.2

The list of components in the simulation.


The presence of any assay streams.
The expected temperature and pressure ranges.
The number of expected liquid phases.
The presence or absence of a vapor phase.
The water handling option chosen (i.e., whether water should be decanted or handled rigorously).
This option should be used if available.

Refinery and Gas


Processes

These processes may be subdivided into the following:


Low pressure crude systems (vacuum towers and atmospheric stills).
High pressure crude systems (including FCCU main fractionators and
coker fractionators).
Reformers and hydrofiners.
Lube oil and solvent de-asphalting units.
Low Pressure Crude Units
Low pressure crude units generally contain less than 3 volume percent light
ends. Moreover, the petroleum fractions present in the feed exhibit nearly
ideal behavior. For these units, the characterization of the petroleum fractions is far more important than the thermodynamic method used. The user
should try different assay and characterization methods first if the simulation
results do not match the plant data.
Since these units contain a small amount of light ends, the Braun K10
(BK10) method should be used quickly as a first attempt and will likely give
acceptable answers. The BK10 method does, however, provide only gross estimates for the K-values for H2 and is not recommended for streams containing a significant amount of H2. For such systems, and for other systems
where the BK10 results are not satisfactory, the Grayson-Streed (GS),
Grayson-Streed Erbar (GSE), or Improved Grayson-Streed (IGS) methods
should be chosen. These methods contain special coefficients for hydrogen
and methane and, as such, provide better predictions for streams containing
small amounts of H2 at low pressures. It is important to note that the predefined thermodynamic systems GS, GSE, and IGS use the Curl-Pitzer (CP)
method for calculating enthalpies. For systems containing heavy ends such
as vacuum towers, however, the saturated vapor is often at reduced temperatures of less than 0.6. This is the lower limit of the Curl-Pitzer enthalpy
method. For these units, therefore, substituting the Lee-Kesler (LK) method
for Curl-Pitzer enthalpies may improve the results.

I-46

Application Guidelines

May 1994

Section 1.2

20.4

Thermodynamic Methods

PRO/II Note: For more information on using predefined systems of thermodynamic methods, see Section 20.4, Predefined Systems, of the PRO/II Keyword
Input Manual.
In addition, the top of many of these low pressure units often contain significant amounts of light components such as methane. Under these conditions,
an equation-of-state method such as Soave-Redlich-Kwong (SRK) or PengRobinson (PR) will provide better answers than the BK10 or Grayson-Streed
methods.
Table 1.2.2-1:
Methods Recommended for Low Pressure Crude Systems
BK10

Gives fast and acceptable answers.

GS/GSE/IGS

Generally more accurate than BK10, especially for streams


containing H2. Use LK enthalpies instead of CP enthalpies
for vacuum towers.

SRK/PR

Provides better results when light ends dominate.

High Pressure Crude Units


High pressure crude units generally contain greater amounts of light ends
than low pressure units. Still, for these units, as for the low pressure crude
units, the characterization of the petroleum fractions remains far more important than the thermodynamic method used. The user should again try different assay and characterization methods first if the simulation results do not
match the plant data. Since these units contain larger amounts of light ends,
the GS, GSE, IGS, SRK, or PR methods should be used and will likely give
acceptable answers.
For FCCU main fractionators, the petroleum fractions are much more hydrogen deficient than are crude fractions. Since most characterization correlations are derived from crude petroleum data, it is expected that the results
will be less accurate than for crude fractions.
See Section 1.1.3 of this manual, Assay Processing, for more information on
characterizing petroleum fractions.
Table 1.2.2-2:
Methods Recommended for High Pressure Crude Systems
GS/GSE/IGS

Quicker but generally less accurate than SRK or PR,


especially for streams containing light ends. Use LK
enthalpies instead of CP enthalpies for better results.

SRK/PR

Provides better results when light ends dominate.

PRO/II Component and Thermophysical Properties


Reference Manual

Application Guidelines I-47

Thermodynamic Methods

Section 1.2

Reformers and Hydrofiners


These units contain streams with a high hydrogen content. The GraysonStreed method, which contains special liquid activity curves for methane and
hydrogen, may be used to provide adequate answers. For the SRK and PR
methods, the PRO/II databanks contain extensive binary interaction parameter data for component pairs involving hydrogen. These methods provide results comparable or better than the GS methods. Moreover, these methods
are more accurate than GS methods in predicting the hydrogen solubility in
the liquid phase. If the user wishes to obtain the most accurate prediction of
hydrogen solubility in the hydrocarbon liquid phase, he/she should use the
SimSci modified SRK or PR methods, SRKM or PRM.

Table 1.2.2-3:
Methods Recommended for Reformers and Hydrofiners
GS/GSE/IGS

Quicker but generally less accurate than SRK or PR,


especially for predicting the hydrogen content of the liquid
phase.

SRK/PR

Provides better results than GS methods.

SRKM/PRM

Provides better results than SRK/PR when predicting the


hydrogen content of the liquid phase.

Lube Oil and Solvent De-asphalting Units


These units contain streams with nonideal components such as H2S and mercaptans. The SimSci modified SRK or PR methods, SRKM or PRM, are recommended, but only if user-supplied binary interaction data are available. If
no binary interaction data specifically regressed for the system are available,
then the data in the PRO/II databanks can be used, and the SRK or PR methods are recommended.

Table 1.2.2-4:
Methods Recommended for Lube Oil and Solvent De-asphalting Units

I-48

Application Guidelines

SRKM/PRM

Recommended when user-supplied binary interaction data


are available.

SRK/PR

Recommended when no user-supplied binary interaction


data are available.

May 1994

Section 1.2

Natural Gas
Processing

Thermodynamic Methods

Natural gas systems often contain inerts such as N2, acid or sour gases such
as CO2, H2S, or mercaptans, and water, along with the usual light hydrocarbon components. Natural gas streams may be treated by a number of methods, e.g., sweetening using amines or dehydration using glycol.
For natural gas systems containing less than 5% N2, CO2, or H2S, and no
other polar components, SRK, PR, or Benedict-Webb-Rubin-Starling
(BWRS) methods provide excellent answers. The SRK and PR binary interaction parameters between these lower molecular weight molecules and
other components are estimated by correlations based on the molecular
weight of the hydrocarbon molecule. For small amounts of these components, this is satisfactory. The BWRS equation of state also contains many
binary interaction parameters for component pairs involving lower weight
components supplied in Dechema. Unlike cubic equations of state such as
SRK or PR, the BWRS equation of state does not satisfy the critical constraints, and so does not extrapolate well into the critical region.
For natural gas systems containing more than 5% N2, CO2, or H2S, but no other
polar components, equation-of-state methods such as SRK or PR are still recommended, although the binary parameters estimated by molecular weight correlations may not produce the best results. The user should provide binary
interaction parameters for component pairs involving these lower molecular
weight components if possible.
For natural gas systems containing water at low pressures, equation-of-state
methods such as SRK or PR may be used, along with the default water decant option, to predict the behavior of these systems.
For these systems at high pressures, where the solubility of hydrocarbon in
water is significant, the default water decant option, which predicts a pure
water phase, is unacceptable. In this case, equation-of-state methods containing advanced mixing rules such as SRKM, PRM, or SRKS, or the KabadiDanner modification to SRK (SRKKD) should be used to predict the
vapor-liquid-liquid behavior of these systems. These methods provide the
best answers if all the relevant binary interaction parameters are available.
For the SRKKD method in particular, PRO/II contains binary interaction
parameters for component pairs involving N2, H2, CO2, CO, and H2S. For
SRKM, PRM, or SRKS methods, the user should make sure that all relevant
binary interaction data are entered.

PRO/II Component and Thermophysical Properties


Reference Manual

Application Guidelines I-49

Thermodynamic Methods

Section 1.2

For natural gas systems containing polar components such as methanol, the
SRKM, PRM, or SRKS methods are recommended to predict the vapor-liquid-liquid behavior of these systems. Interaction parameters for the important binaries are required if accurate answers are to be obtained.

Table 1.2.2-5:
Methods Recommended for Natural Gas Systems
SRK/PR/
BWRS

Recommended for most natural gas and low pressure natural gas +
water systems.

SRKKD

Recommended for high pressure natural gas + water systems.

SRKM/PRM/
SRKS

Recommended for natural gas + polar components.

The common processes used to treat natural gas streams may be sub-divided
into the following:
Glycol dehydration systems.
Sour water systems.
Amine systems.
Glycol Dehydration Systems
The predefined thermodynamic system GLYCOL has been specially created
for these systems. This system uses the predefined system SRKM but invokes the GLYCOL databank. This databank contains binary interaction parameters for component pairs involving TEG and, to a lesser extent, DEG
and EG. These data have been regressed in the temperature and pressure
range normally seen in glycol dehydrators:
Temperature: 80-400 F.
Pressure: up to 2000 psia.
See Section 1.2.8, of this manual, Special Packages, for more information on
this method.
Sour Water Systems
The standard version of PRO/II contains two methods, SOUR and GPSWATER,
for predicting the VLE behavior of sour water systems. These methods are described in more detail in Section 1.2.8, of this manual, Special Packages. The
electrolyte version of PRO/II also contains a rigorous sour water method which
is described in Section 1.2.10 of this manual, Electrolyte Thermodynamic Equations. The recommended temperature, pressure and composition ranges for each
method are given in Table 1.2.2-6 below.

I-50

Application Guidelines

May 1994

Section 1.2

Thermodynamic Methods
Table 1.2.2-6:
Methods Recommended for Sour Water Systems
SOUR

Recommended Ranges:
68 < T (F) < 300
P(psia) < 1500
wNH3 + wCO2 + wH2S < 0.30

GPSWATER

Recommended Ranges:
68 < T(F) < 600
P(psia) < 2000
wNH3 < 0.40
PCO2 + PH2S < 1200 psia

Electrolyte
Version of
PRO/II

Recommended when strong electrolytes such as caustic


are used, or when pH control or accurate prediction of HCN
or phenol phase distribution is important. Recommended
Ranges:
32 < T(F) < 400
P(psia) < 3000
xdissolved gases < 0.30

Amine Systems
Amine systems used to sweeten natural gas streams may be modeled in
PRO/II using the AMINE special package (see Section 1.2.8, Special Packages). Data is provided for amines MEA, DEA, DGA, DIPA, and MDEA. Results obtained for MEA and DEA are accurate enough for use in final design
work. However, results for DIPA systems are not suitable for final design
work. For MDEA or DGA systems, the results may be made to more closely
fit plant data by the use of a dimensionless residence time correction.

25.5

PRO/II Note: For more information on using the dimensionless residence time
correction factor for MDEA or DGA systems, see Section 25.5, Amine, of the
PRO/II Keyword Input Manual.
The recommended temperature, pressure, and loading ranges (gmoles sour
gases per gmole amine) for each amine system available in PRO/II are given
in Table 1.2.2-7.

PRO/II Component and Thermophysical Properties


Reference Manual

Application Guidelines I-51

Thermodynamic Methods

Section 1.2
Table 1.2.2-7:
Recommended Ranges for Amine Systems

Petrochemical
Applications

MEA

Recommended Ranges:
25 < P(psig) < 500
T(F) < 275
wamine ~ 0.15 - 0.25
0.5-0.6 gmole gas/gmole amine

DEA

Recommended Ranges:
100 < P(psig) < 1000
T(F) < 275
wamine ~ 0.25 - 0.35
0.45 gmole gas/gmole amine

DGA

Recommended Ranges:
100 < P(psig) < 1000
T(F) < 275
wamine ~ 0.55 - 0.65
0.50 gmole gas/gmole amine

MDEA

Recommended Ranges:
100 < P(psig) < 1000
T(F) < 275
wamine ~ 0.50
0.40 gmole gas/gmole amine

DIPA

Recommended Ranges:
100 < P(psig) < 1000
T(F) < 275
wamine ~ 0.30
0.40 gmole gas/gmole amine

Common examples of these processes are the following:


Light hydrocarbon applications.
Aromatic systems.
Aromatic/non-aromatic systems.
Alcohol dehydration systems.
Light Hydrocarbon Applications
Most light hydrocarbon mixtures at low pressures may be modeled well by
the SRK or PR equations of state. The BWRS equation of state, which was
developed for light hydrocarbon mixtures, is also recommended, but not near
the critical region. At high pressures, the SRKKD equation of state should be
used to best predict the water solubility in the hydrocarbon phase. The
COSTALD liquid density method was developed expressly for light
hydrocarbon mixtures. This method is over 99.8% accurate in predicting the
liquid densities of these mixtures and should be requested by the user.
Table 1.2.2-8:
Methods Recommended for Light Hydrocarbons

I-52

Application Guidelines

SRK/PR/
BWRS

Recommended for systems of similar light hydrocarbons at


low pressures.

SRKKD

Recommended at higher pressures.

COSTALD

Recommended for liquid density.

May 1994

Section 1.2

Thermodynamic Methods

Aromatic Systems
Mixtures of aromatic components, such as aniline and nitrobenzene at low
pressures less than 2 atmospheres, exhibit close to ideal behavior. Ideal methods can therefore be used to predict phase behavior and compute enthalpies,
entropies, and densities. At pressures above 2 atmospheres, the GraysonStreed or SRK or PR methods provide good results in the prediction of phase
equilibria. The SRK or PR equations of state should provide better results,
but with a small CPU penalty.
Table 1.2.2-9:
Methods Recommended for Aromatics
IDEAL

Recommended for systems at low pressures below 2 atm.

GS/SRK/PR

Recommended at pressures higher than 2 atm.

IDEAL/API/
COSTALD

Recommended for liquid density. The COSTALD method is


best at high temperatures and if light components such as
CH4 are present.

Aromatic/Non-aromatic Systems
Systems of mixtures of aromatic and non-aromatic components are highly
nonideal. Liquid activity methods such as NRTL or UNIQUAC, or equation-ofstate methods with advanced mixing rules such as SRKM or SRKS can be used
to model these systems. Both types of methods can be used to successfully
model aromatic/non-aromatic mixtures, provided that all the binary interaction
data for the components in the system are provided. The PRO/II databanks contain an extensive variety of interaction data for the NRTL, UNIQUAC, SRKM,
and SRKS methods. One advantage to using the liquid activity methods NRTL
or UNIQUAC, however, is that the FILL option may be used to fill in any missing interaction parameters using UNIFAC. All library components in the PRO/II
databanks have UNIFAC structures already defined. PRO/II also will estimate
UNIFAC structures for petro components based on their Watson K and molecular weight values, and the user may supply UNIFAC structures for components
not in the PROII databanks.

24.10

PRO/II Note: For more information on using the FILL option to fill in missing
interaction parameters for liquid activity methods, see Section 24.10, Filling in
Missing Parameters, of the PRO/II Keyword Input Manual.
When gases such as H2, N2, or O2 are present in small quantities (up to about 5
mole %), the Henrys Law option may be used to calculate the gas solubilities.
Once the Henrys Law option is selected by the user, PRO/II arbitrarily defines
all components with critical temperatures less than 400 Kelvin as solute components, though the user may override these selections.

24.11

PRO/II Note: For information on using the SOLUTE option to define Henrys
components, see Section 24.11, Henrys Law for Non-condensible Components, of the PRO/II Keyword Input Manual.

PRO/II Component and Thermophysical Properties


Reference Manual

Application Guidelines I-53

Thermodynamic Methods

Section 1.2

For large amounts of supercritical gases, an equation-of-state method with an


advanced mixing rule should be used to predict the phase behavior.
Table 1.2.2-10:
Methods Recommended for Aromatic/Non-aromatic Systems
SRKM/PRM

Recommended at high pressures or when > 5 mole %


supercritical gases are present.

NRTL/
UNIQUAC

Recommended with the FILL option when binary


interaction parameters are not available or with the HENRY
option when < 5 mole % supercritical gases are present.

Alcohol Dehydration Systems


The PRO/II special package ALCOHOL is recommended for systems containing alcohols with water. This package uses a special databank of NRTL
parameters containing interaction parameters expressly regressed under temperature and pressure conditions commonly found in dehydration systems.
The NRTL method is suggested if user-supplied interaction data are to be
used.

25.1

PRO/II Note: For information on using the ALCOHOL method, see Section
25.1, Alcohol, of the PRO/II Keyword Input Manual.
Table 1.2.2-11:
Methods Recommended for Alcohol Systems

Chemical
Applications

ALCOHOL

Recommended for all alcohol dehydration systems.

NRTL/
UNIQUAC

Recommended when user-supplied data are provided.

Non-ionic Systems
These systems, which typically contain oxygen, nitrogen, or halogen derivatives of hydrocarbons, such as amides, esters, or ethers, are also similar to
non-hydrocarbon systems found in petrochemical applications. For low pressure systems, a liquid activity coefficient method is recommended. For single
liquid phase systems, the WILSON, NRTL, or UNIQUAC methods are
equally good, provided all interaction parameters are provided. PRO/II databanks contain extensive parameters for NRTL and UNIQUAC, but the user
must supply interaction data for the WILSON method. The WILSON method
is the simplest and requires the least CPU time.

I-54

Application Guidelines

May 1994

Section 1.2

Thermodynamic Methods

For systems with two liquid phases, the NRTL or UNIQUAC methods
should be used, provided that at least some interaction data are available. The
FILL option can be used to fill in any missing interaction data using the UNIFAC method. If no interaction data are available, the UNIFAC method
should be used since the PRO/II databanks contain a large amount of group
interaction data for both VLE and LLE applications. For moderate pressure
systems up to 10 atmospheres, a liquid activity method can still be used,
provided that the interaction parameters used are still valid in that pressure
range. For example, if the system pressure were much higher than the pressure at which the interaction parameters were regressed, the vapor-phase
fugacity may be taken into account in modeling the phase behavior. If the
PHI option is selected, the liquid-phase Poynting correction factor is automatically selected also.
It is also important to note that all the interaction parameters in the PRO/II
databanks, except for dimerizing components such as carboxylic acids, were
regressed without including any vapor-phase non-ideality. This means that
the PHI option should be used for carboxylic acid systems at all pressures,
but should only be used for most components at high pressures. For systems containing components such as carboxylic acids that dimerize in the vapor phase,
the Hayden-OConnell fugacity method may be used to calculate all vapor-phase
properties such as fugacity, enthalpy, and density. For components such as hydrogen fluoride which forms hexamers in the vapor phase, PRO/II contains an equation of state specially created for such systems, HEXAMER. This method is
recommended for processes such as HF alkylation or the manufacture of refrigerants such as HFC-134a. For all other components, an equation-of-state method
such as SRK or PR may be used to calculate vapor-phase fugacities.

24

PRO/II Note: For information on using the PHI option to calculate the
vapor-phase fugacity when using a liquid activity method, see Section 24,
Liquid Activity Methods, of the PRO/II Keyword Input Manual.
When supercritical gases are present in small quantities (generally less than
5 mole %), the Henrys Law option should be used to compute gas solubilities. For high pressure systems, greater than 10 atmospheres, or for systems
with large quantities of supercritical gas, an equation-of-state method using
an advanced mixing rule, such as SRKM or PRM, should be used. The UNIWAALS equation-of-state method uses UNIFAC structure information to
predict phase behavior. This method is useful when interaction data are not
available and, unlike a liquid activity method such as UNIFAC, is able to
handle supercritical gases.

PRO/II Component and Thermophysical Properties


Reference Manual

Application Guidelines I-55

Thermodynamic Methods

Section 1.2
Table 1.2.2-12:
Methods Recommended for Non-ionic Chemical Systems
WILSON

Recommended for single liquid phase slightly nonideal


mixtures. If all interaction data are not available, use the
FILL=UNIFAC option.

NRTL/
UNIQUAC

Recommended for all nonideal mixtures. Use with the FILL


option when binary interaction parameters are not available
or with the HENRY option when < 5 mole % supercritical
gases are present. For moderate pressures, use the PHI
option for vapor phase non-idealities.

SRKS/
SRKM/PRM/
UNIWAALS

Recommended for high pressure systems or when > 5


mole % supercritical gases are present.

HOCV

Recommended for vapor fugacity and enthalpy and density


calculations in systems containing dimerizing components
such as carboxylic acids. Use with a liquid activity method .

HEXAMER

Recommended for systems containing hexamerizing


components such as HF.

Ionic Systems
A special version of PRO/II expressly made for aqueous electrolytes is recommended when modeling these systems. This version combines the PRO/II
flowsheet simulator with rigorous electrolyte thermodynamic algorithms developed by OLI Systems, Inc. Chemical systems which may be modeled by this
special version include amine, acid, mixed salts, sour water, caustic, and Benfield systems. See Sections 1.2.9, Electrolyte Mathematical Model, and 1.2.10,
Electrolyte Thermodynamic Equations for further details.
Table 1.2.2-13:
Methods Recommended for Ionic Chemical Systems
PRO/II Electrolyte Version

Environmental Applications
These systems typically involve stripping dilute pollutants out of water. By themselves, liquid activity methods such as NRTL do not model these dilute systems
with much accuracy. A better approach is to use a liquid activity method in combination with Henrys constants at the process temperature to model these dilute
aqueous systems. PRO/II contains Henrys Law constants for many components
such as HCl, SO2, and ethanediol in water. Some additional Henrys constants
for chlorofluorocarbons (CFCs) and hydrofluorocarbons (HFCs) in water are
also available in the PRO/II databanks. Other sources for Henrys Law data
include the U.S. Environmental Protection Agency.

I-56

Application Guidelines

May 1994

Section 1.2

Thermodynamic Methods
Table 1.2.2-14:
Methods Recommended for Environmental Applications
Liquid Activity Method + Henrys Law Option

Solid Applications
Solid-liquid equilibria for most systems can be represented in PRO/II by the
vant Hoff (ideal) solubility method or by using user-supplied solubility data.
In general, for those systems where the solute and solvent components are
chemically similar and form a near-ideal solution, the vant Hoff method is
appropriate. For nonideal systems, solubility data should be supplied. For
many organic crystallization systems, which are very near ideal in behavior,
the vant Hoff SLE method provides good results. The VLE behavior can usually be adequately represented by IDEAL or any liquid activity methods.
Precipitation of solid salts and minerals from aqueous solutions can be calculated more rigorously by using the electrolyte version of PRO/II.

26.1

PRO/II Note: For information on using the ideal solid-liquid equilibria method
VANTHOFF, see Section 26.1, Vant Hoff Solubility, of the PRO/II Keyword
Input Manual.
Table 1.2.2-15:
Methods Recommended for Solid Applications
Ideal or Liquid
Activity
Method(VLE)
+ VANT HOFF
Method (SLE)

Recommended for most solid systems involving organics.

PRO/II
Electrolyte
Version

Recommended for solid salt and mineral precipitation


from aqueous solutions.

PRO/II Component and Thermophysical Properties


Reference Manual

Application Guidelines I-57

Thermodynamic Methods

1.2.3

Section 1.2

Generalized Correlation Methods


General
Information

Ideal
(IDEAL)

Vapor-liquid equilibria can be predicted for hydrocarbon mixtures using


various general correlation methods. Examples of these are those developed
by Chao and Seader or Grayson and Streed. Vapor-liquid equilibria can also
be predicted by convergence pressure correlations such as the K10 charts developed by Cajander et al. Densities, enthalpies, and entropies can also be
calculated using a number of correlation methods such as Lee-Kesler and
COSTALD.
Ideal K-values are generally applicable to systems which exhibit behavior
close to ideality in the liquid phase. Mixtures of similar fluids often exhibit
nearly ideal behavior. In an ideal solution at constant temperature and pressure, the fugacity of every component is proportional to its mole fraction.
For every component i, the following fundamental thermodynamic equilibrium relationship holds:
L

(1)

fi = fi

where:
superscript L refers to the liquid phase
superscript V refers to the vapor phase
fi =

fugacity of component i

In the vapor phase, the fugacity is assumed to be equal to the partial pressure:
(2)

fi = yi P

where:
yi =

vapor mole fraction

P=

system pressure

In the liquid phase for an ideal liquid (ignoring correction factors that are
usually small):
L

sat

(3)

fi = xi fpure i = xi Pi

where:
xi =
f

liquid mole fraction

pure i =pure-component i liquid fugacity


Pisat = vapor pressure of component i at the

I-58

Generalized Correlation Methods

system temperature

May 1994

Section 1.2

Thermodynamic Methods

Raoults law thus holds:


(4)

sat

yi P = xi Pi

The ideal K-value is therefore given by:


sat

(5)

Ki = yi / xi = Pi / P

Note that there is no compositional dependency of the K-values. They are only a
function of temperature (due to the dependence of Pisat on T) and pressure.
Ideal vapor densities are obtained from the ideal gas law:
= P / RT

(6)

where:
= vapor molar density of mixture
Ideal liquid densities are obtained from pure-component saturated-liquid density
correlations.
Ideal liquid enthalpies are obtained from pure-component liquid enthalpy correlations, and the corresponding vapor enthalpies are obtained by adding in
the effect of the known latent heat of vaporization of the component.
Ideal entropies are calculated from the ideal enthalpy data using the following equation:
Si = cpi dT / T = Hi / T dT / T
P

(7)

where:

22.1

Si =

ideal entropy

cpi =

ideal heat capacity of component i

Hi =

ideal enthalpy

Tref =

reference temperature (1 degree Rankine)

T=

temperature of mixture

PRO/II Note: For more information on using the IDEAL and LIBRARY methods,
see Section 22.1, IDEAL and LIBRARY, of the PRO/II Keyword Input Manual.

PRO/II Component and Thermophysical Properties


Reference Manual

Generalized Correlation Methods I-59

Thermodynamic Methods

Chao-Seader
(CS)

Section 1.2

Chao and Seader calculated liquid K-values for the components of nonideal
mixtures using the relationship:
Ki = yi / xi =

0L
i fi

(8)


i P

where:
fi0L =

the standard-state fugacity of component i in the pure


liquid phase

i =

the activity coefficient of component i in the equilibrium


liquid mixture

i =

the fugacity coefficient of component i in the equilibrium


vapor mixture

It was shown that i could be calculated from molar liquid volumes and solubility
parameters, using the Scatchard-Hildebrand equation, with regular liquid solution
assumed. The Redlich-Kwong equation of state (see Section 1.2.4, Equations of
State) was used to evaluate . Chao and Seader presented a generalized correlation
for fi0L/P, the fugacity coefficient of pure liquid i in real and hypothetical states.
In the development of their correlation for vapor-liquid K-values, Chao and
Seader used the framework of Pitzers modified form of the principle of corresponding states for the pure-liquid fugacity coefficients, giving values of
fi0L/P as a function of reduced temperature, reduced pressure, and acentric
factor for both real and hypothetical liquids:
OL
OL
OL
ln fi / P = ln fi / P + ln fi / P

0
1

(9)

where:
= acentric factor
The first term on the right hand side of equation (9) represents the fugacity
coefficient of simple fluids. The second term is a correction accounting for the
departure of the properties of real fluids from those of simple fluids.

22.3

PRO/II Note: For more information on using the Chao-Seader method, see Section 22.3, Chao-Seader, of the PRO/II Keyword Input Manual.
Limitations of the Chao-Seader method are given below:
For all hydrocarbons;
Pressure: up to 2000 psia, but not exceeding 0.8 of the critical pressure
of the system.
Temperature: -100 F to 500 F and pseudoreduced temperature, Tr,
of the equilibrium liquid mixture less than 0.93. The pseudoreduced
temperature is based on the molar average of the critical temperatures of the components.
Concentration: up to 20 mole % of other dissolved gases in the
liquid.

I-60

Generalized Correlation Methods

May 1994

Section 1.2

Thermodynamic Methods

The method is not suitable for other non-hydrocarbon components such


as N2, H2S, CO2, etc.
Reference
Chao, K. C., and Seader, J. D., 1961, A Generalized Correlation of
Vapor-Liquid Equilibria in Hydrocarbon Mixtures, AIChE J., 7(4),
598-605.

Grayson-Streed
(GS)

22.2

Grayson and Streed modified the Chao-Seader correlation in 1963 by fitting


data over a wider range of conditions and, hence, deriving different constants for
the equations giving the fugacity coefficients of the pure liquids. Special coefficients for hydrogen and methane are supplied because typical application temperatures are far above the critical points of these two components. Grayson and
Streeds modifications have extended the application range for hydrocarbon systems up to 800 F and 3000 psia. The lower limits imposed by Chao and
Seader still apply.
PRO/II Note: For more information on using the Grayson-Streed method, see
Section 22.2, Grayson-Streed, of the PRO/II Keyword Input Manual.

Reference
Grayson, H. G., and Streed, C. W., 1963, Vapor-Liquid Equilibria for
High Temperature, High Pressure Hydrocarbon-Hydrocarbon Systems, 6th
World Congress, Frankfurt am Main, June 19-26.

Erbar Modification
to Chao-Seader
(CSE) and GraysonStreed (GSE)

22.4

In 1963, Erbar and Edmister developed a new set of constants for the ChaoSeader liquid fugacity coefficient, specifically for N2, H2S, and CO2, in
order to improve the prediction of the K-values of these gases. At the same
time, new solubility parameter and molar volume values were found for
these components.
PRO/II Note: For more information on using the Erbar-modified Chao-Seader
and Grayson-Streed methods, see Section 22.4, Modifications to GraysonStreed and Chao-Seader, of the PRO/II Keyword Input Manual.
A limitation of this modified method, however, is that the H2S correlation
cannot be used in any cases where an azeotrope may exist (e.g., H2S/C3H8
mixtures), as the azeotrope will not be predicted.
Reference
Erbar, J. H., and Edmister, W. C., 1963, Vapor-Liquid Equilibria for
High Temperature, High Pressure Hydrocarbon-Hydrocarbon Systems, 6th
World Congress, Frankfurt am Main, June 19-26.

PRO/II Component and Thermophysical Properties


Reference Manual

Generalized Correlation Methods I-61

Thermodynamic Methods

Section 1.2

Improved GraysonStreed (IGS)

For hydrocarbon-water mixtures, the Grayson-Streed and Erbar-modified


Grayson-Streed methods accurately predict the phase behavior of the hydrocarbon-rich phase, but are incapable of predicting the composition of the
water-rich phase. A new method has been developed in which a separate set
of solubility parameters was used in the water-rich phase, and a new set of
liquid fugacity coefficients developed for N2, H2O, H2S, CO, and O2. This
new method is known as Improved Grayson-Streed. It was found that the
Grayson-Streed liquid fugacity coefficient for the simple fluid decreases
rapidly as Tr increases above 2.5, and can in fact become negative. The liquid fugacity coefficient for the simple fluid was therefore replaced by that
for hydrogen at reduced temperatures of 2.5 and above.

22.4

PRO/II Note: For more information on using the Improved Grayson-Streed


method, see Section 22.4, Modifications to Grayson-Streed and Chao-Seader,
of the PRO/II Keyword Input Manual.

Curl-Pitzer (CP)

This correlation may be used to predict both liquid and vapor enthalpies and
entropies. It computes the enthalpy deviation using the principle of
corresponding states, i.e., in terms of the reduced temperature, reduced
pressure, and acentric factor. The critical temperature and pressure for the
mixture is computed using the mixture rules of Stewart, Burkhart, and Voo.
The mixture acentric factor used is the molar average value.
The Curl-Pitzer method is limited to nonpolar mixtures and may be used for
Pr up to 10, and Tr from 0.35 to 4.0 for liquids, and Tr from 0.6 to 4.0 for
vapors. For systems containing heavy ends, the saturated vapor is sometimes
at a reduced temperature of less than 0.6. In this case, the CP correlation extrapolates reasonably, producing satisfactory results.

22.5

PRO/II Note: For more information on using the Curl-Pitzer method, see Section 22.5, Curl-Pitzer, of the PRO/II Keyword Input Manual.
The Curl-Pitzer method is generally useful for refinery hydrocarbons and in
oil absorption gas plants.
References

I-62

1.

Stewart, Burkhart, and Voo, 1959, Prediction of Pseudo-Critical Constants


for Mixtures, Paper presented at AIChE Meeting, Kansas City.

2.

American Petroleum Institute, 1970, Technical Data Book - Petroleum


Refining, 2nd Ed., Procedure 7B3.1.

3.

American Petroleum Institute, 1970, Technical Data Book - Petroleum


Refining, 2nd Ed., Procedure 7H2.1, 7-201 - 7-202.

Generalized Correlation Methods

May 1994

Section 1.2

Braun K10 (BK10)

Thermodynamic Methods

The K-value of each component is a function of the system temperature,


pressure, and the composition of the vapor and liquid phases. For natural gas
systems, the convergence pressure can be used as the parameter that represents the composition of the vapor and liquid phases in equilibrium. The
convergence pressure is, in general, the critical pressure of a system at a
given pressure. At a given temperature, and as the system pressure increases,
the K-values of all components converge to unity when the system pressure
reaches the convergence pressure.
The Braun K10 charts developed by Cajander et al. in 1960 show the low pressure equilibrium ratio, arbitrarily taken at 10 psia system pressure and 5000 psi
convergence pressure. For many hydrocarbon systems, no experimental data are
available. For these cases, the equilibrium K-values may be predicted from
vapor pressure:
sat

(10)

K10 = P

/ 10

Psat =

saturated vapor pressure in psia.

where:
The relationship given in equation (10) only holds for K-values less than 2.5.
For nonhydrocarbons like H2, the K-value is assumed to be 10 times as large
as the methane value. For N2, O2, and CO, the K-values are assumed to be
identical to that of methane. The K-values for CO2 and H2S are assumed to
be identical to that of propylene.
For petroleum fractions in which the form of the vapor pressure curve is unknown, a rough K10 chart is developed from the normal boiling point of the
fraction. The following method is used:
On the appropriate K10 chart, the point K10 = 14.7/10 = 1.47 is plotted at
the atmospheric boiling point.
The whole K10 curve can then be sketched in by similitude to the known
K10 curves for homologous hydrocarbons.
The K10 charts apply to mixtures that behave ideally at low pressures, e.g.,
for mixtures of one molecule type such as mixtures of paraffins and olefins.
For mixtures of naphthalenes mixed with olefins and paraffins, the accuracy
of BK10 is slightly poorer. Large errors can be expected for mixtures of aromatics with paraffins, olefins, or naphthalenes, which cause nonidealities and
form azeotropes.

22.6

PRO/II Note: For more information on using the Braun K10 method, see Section 22.6, Braun K10, of the PRO/II Keyword Input Manual.

Reference
Cajander, B. C., Hipkin, H. G., and Lenior, J. M., 1960, Prediction of
Equilibrium Ratios from Nomographs of Improved Accuracy, J. Chem.
Eng. Data, 5, 251-259.

PRO/II Component and Thermophysical Properties


Reference Manual

Generalized Correlation Methods I-63

Thermodynamic Methods

Johnson-Grayson
(JG)

Section 1.2

This correlation may be used to predict both liquid and vapor enthalpies. It is
essentially an ideal-enthalpy correlation, using saturated liquid at 0 C as
the datum for the correlation (-200 F in versions 3.5 and earlier).
Vapor-phase corrections are calculated using the Curl-Pitzer correlation.
Pressure effects are not considered for the liquid phase.
Johnson-Grayson is useful for systems containing heavy ends between 0 F
and 1200 F. However, it can be extrapolated to higher temperatures. The correlation should not be used if the mixture is C4-C5 or lighter.

22.7

PRO/II Note: For more information on using the Johnson-Grayson method, see
Section 22.7, Johnson-Grayson, of the PRO/II Keyword Input Manual.

Reference
Johnson, and Grayson, 1961, Enthalpy of Petroleum Fractions,
Petroleum Refiner, 40(2), 123-29.

Lee-Kesler (LK)

22.8

This correlation may be used to predict both liquid and vapor enthalpies, entropies, and densities. This correlation uses the three-parameter corresponding-states theory, which essentially states that all fluids having the same
acentric factor must have the same properties at the same reduced temperature and pressure. Special mixing rules have been used to calculate the mixture reduced properties. For most fluids, the Lee-Kesler method is 98%
accurate in predicting the gas-phase compressibility factors. The method also
gives reasonable results for slightly polar mixtures. This method is not recommended for highly polar mixtures, or those which form strongly associative hydrogen bonds. However, the Lee-Kesler method provides accurate
results for polar fluids at low temperatures near the saturated-vapor region.
The Lee-Kesler method is not recommended for calculating liquid densities
of hydrocarbons heavier than C8.
PRO/II Note: For more information on using the Lee-Kesler method, see Section 22.8, Lee-Kesler, of the PRO/II Keyword Input Manual.

References

I-64

1.

American Petroleum Institute, 1975, Technical Data Book-Petroleum


Refining, 3rd Ed., 2-1 - 7-4.

2.

Lee, B. I., and Kesler, M. G., 1975, A Generalized Thermodynamic


Correlation Based on Three-Parameter Corresponding States, AIChE J., 21,
510-27.

3.

Kesler, M. G., and Lee, B. I., 1976, Improved Prediction of Enthalpy of


Fractions, Hydrocarbon Proc., 53, 153-158.

Generalized Correlation Methods

May 1994

Section 1.2

Thermodynamic Methods

API

This correlation may be used to predict liquid densities. An initial density is


calculated at 60 F using the weight average of the components. The reduced
temperature and pressure of the stream at 60 F and 14.7 psia are computed
using Kays rule, i.e., the reduced temperature and pressure are assumed to
be a linear function of the liquid mole fraction. A density factor C is then
read from Figure 6A2.21 in the API Technical Data Book. A second correction factor is then determined, corresponding to the reduced temperature and
pressure, at the actual fluid conditions. Finally, the actual liquid density is
calculated according to:
L

(11)

act = 60 Cact / C60

where:

Lact = actual liquid density


L60 =

liquid density at 60 F

Cact = actual correction factor


C60 =

22.9

correction factor at 60 F

PRO/II Note: For more information on using the API method, see Section
22.9, API Liquid Density, of the PRO/II Keyword Input Manual.
The API method works well for most hydrocarbon systems, provided that the
reduced temperature is less than 1.0.
Reference
American Petroleum Institute, 1978, Technical Data Book - Petroleum
Refining, 5th Ed., 6-45 - 6-46.

Rackett

This correlation may be used to predict liquid densities. The saturated-liquid


density is obtained from:

(12)

i
Vsi = RTci / Pci Zrai

2 7
i = 1 + (1Tri) / for Tri 0.75
3

i = 1.6 +

6.9302610
for Tri > ~ 0.75
Tri 0.655

Vsi =

saturated liquid volume

Zrai =

Rackett parameter for component i

where:

Tci, Pci =
Tri =

PRO/II Component and Thermophysical Properties


Reference Manual

critical temperature and pressure for component i

reduced temperature for component i

Generalized Correlation Methods I-65

Thermodynamic Methods

22.10

Section 1.2

PRO/II Note: For more information on using the Rackett liquid density method,
see Section 22.10, Rackett Liquid Density, of the PRO/II Keyword Input
Manual.
The PRO/II databanks contain Rackett parameters for many components.
However, if Rackett parameters are not available, PRO/II will use the critical
compressibility factor, zc.
For mixtures, there are two ways to use the Rackett equation. The most
straightforward, known as the RACKETT method in PRO/II, is to use equation (12) for the molar volume of each pure component and then mix the
volumes together linearly. A second approach is the One-Fluid" Rackett
method (known as the RCK2 method), in which mixing rules are used to determine effective critical parameters for the mixture and then equation (12) is
used to determine the mixture density. For most mixtures, the difference between these two methods will not be significant.
References

COSTALD

1.

Rackett, H. G., 1970, Equation of State for Saturated Liquids, J. Chem. Eng.
Data, 15, 514.

2.

Spencer, C. F., and Danner, R. P., 1972, Improved Equation for Prediction
of Saturated Liquid Density, J. Chem. Eng. Data, 17, 236-241.

3.

Spencer, C. F., and Adler, S. B., 1978, A Critical Review of Equations for
Predicting Saturated Liquid Density, J. Chem. Eng. Data, 23, 82-89.

The corresponding-states liquid density model predicts the liquid densities of


LNG-like fluids. This accurate and reliable method is over 99.8% accurate in
predicting the densities of light hydrocarbon mixtures. This model uses two characteristic parameters for each pure component in the mixture - a characteristic
volume, V*, and a tuned acentric factor, SRK. The acentric factor is chosen
such that the SRK equation of state best matches the vapor-pressure data. Typically, this tuned acentric factor varies little in value from the standard acentric
factor. The saturated volume is given by:
(0)
()

Vs / V = Vr 1 SRK Vr

(0)
Vr

k 3
= 1 + Ak (1 Tr) / , 0.25 < Tr < 0.95

(13)
(14)

I-66

Generalized Correlation Methods

May 1994

Section 1.2

Thermodynamic Methods

Vr = Bk Tr / Tr 1.00001 , 0.25 < Tr < 1.0


()

(15)

where:
Vs =

saturated molar volume

V* =

characteristic volume

Vr =

reduced volume

Ak, Bk = COSTALD parameters


SRK = SRK tuned acentric factor
For mixtures, the following mixing rules are used:

Tcm = xixjVij Tcij / Vm

i j

Vm = 1 / 4

(16)

2 3

1 3
x V + 3xiVi / xiVi /
i i
i
i
i

(17)

(18)

1/2

Vij Tcij = Vi Tci Vj Tcj

SRK = xiSRKi

(19)

where:
subscript m refers to mixture properties.
For compressed pure liquids and liquid mixtures, the original work was extended by Thomson et al. in 1982, adding a pressure correction of the form:
sat
V = Vs 1 Cln (B + P) / (B + P )

(20)

where:
B, C are constants, dependent on composition

22.11

Psat =

saturated vapor pressure, obtained from a generalized vapor


pressure relationship.

V=

molar volume

PRO/II Note: For more information on using the COSTALD liquid density
method, see Section 22.11, Costald Liquid Density, of the PRO/II Keyword
Input Manual.
The COSTALD method is valid for aromatics and light hydrocarbons up to
reduced temperatures of 0.95. PRO/II databanks contain COSTALD characteristic volume, V*, for many components. However, if the characteristic volume is not available, PRO/II will use the critical volume of the pure
component, Vc. For petroleum and assay components, however, PRO/II will
back calculate a characteristic volume, if missing, in order to provide a correct specific gravity for the pseudocomponent.

PRO/II Component and Thermophysical Properties


Reference Manual

Generalized Correlation Methods I-67

Thermodynamic Methods

Section 1.2

References

I-68

1.

Hankinson, R. W., and Thomson, G. H., 1979, A New Correlation for Saturated
Densities of Liquids and Their Mixtures, AIChE J., 25o, 653-663.

2.

Thomson, G. H., Brobst, K. R., and Hankinson, R. W., 1982, An Improved


Correlation for Densities of Compressed Liquids and Liquid Mixtures,
AIChE J., 28, 671-676.

Generalized Correlation Methods

May 1994

Section 1.2

1.2.4

Thermodynamic Methods

Equations of State
General
Information

General Cubic
Equation of State

Equations of state for phase-equilibrium calculations are applicable to wide


ranges of temperature and pressure conditions. They can also be used to calculate all the related thermodynamic properties such as enthalpy and entropy. The
reference state for both the vapor and liquid phase is the ideal gas, and deviations from the ideal-gas state are determined by calculating fugacity coefficients
for both phases. For cubic equations of state in particular, critical and supercritical conditions can be predicted quite accurately. By using an appropriate temperature-dependent function to describe the attractive forces between molecules,
volume function, and mixing rule, cubic equations of state have been shown to
be quite successful in predicting vapor-liquid equilibria for highly nonideal
systems.
A general two-parameter cubic equation of state can be expressed by the
equation:
2
2
P = RT / (vb) a(T) / v + ubv + wb

(1)

where:
P=

pressure

T=

absolute temperature

v=

molar volume

u,w =

constants, typically integers

The values of u and w determine the type of cubic equation of state. Table
1.2.4-1 shows three of the best known of these. The van der Waals equation
developed in 1873 is obtained by setting u=w=0. By setting u=1 and w=0,
the Redlich-Kwong equation (1949) is obtained. Peng and Robinson developed their equation of state in 1976 by setting u=2 and w=-1.
Table 1.2.4-1: Some Cubic Equations of State

Equation of state

van der Waals (vdW)

Redlich-Kwong (RK)

-1

Peng-Robinson (PR)

The parameters a and b at the critical temperature (ac and bc) are found by
setting the first and second derivatives of pressure with respect to volume
equal to zero at the critical point. Application of these constraints at the critical point to equation (1) yields:

PRO/II Component and Thermophysical Properties


Reference Manual

Equations of State

I-69

Thermodynamic Methods

Section 1.2

u + (u1) / 3 + (u1) / 27 Bc + u + w 2 / 3(u1) 1 / 9(u1) Bc +


1 / 3 + 1 / 9(u1) Bc 1 / 27 = 0

(2)

Ac = 3Zc + uBc + (uw) Bc

(3)

Zc = 1 / 3 (u1) Bc / 3

(4)

where:
Ac = Pcac/R2Tc2
Bc = Pcbc/RTc
subscript c refers to the critical point
The critical constraints result in three expressions for three unknowns, Ac, Bc,
and Zc. These unknowns depend on the values of u and w. Actually, Ac and Bc
are the only true unknowns appearing in these equations, because Pc, Tc, and Vc
(and hence Zc) are properties of a substance, having numerical values independent of any equation of state. In solving these three equations, Vc is in fact
treated as a third unknown. Table 1.2.4-2 lists these constants for the van der
Waals, Redlich-Kwong, and Peng-Robinson equations of state.
Table 1.2.4-2: Constants for Two-parameter Cubic Equations of State

Ac

Bc

Zc

Equation of state

0.42188

0.1250

0.3750

van der Waals (vdW)

0.42747

0.0866403

0.3333

Redlich-Kwong (RK)

0.45724

0.0778

0.3074

Peng-Robinson (PR)

References

I-70

Equations of State

1.

Abbott, M. M., 1973, Cubic Equations of State, AIChE J., 19, 596-601.

2.

van der Waals, J. D., 1873, Over de Constinuiteit van den gas-en Vloeistoftoestand, Doctoral Dissertation, Leiden, Holland.

3.

Redlich, O., and Kwong, N. S., 1949, On the Thermodynamics of Solutions.


v: An Equation of State. Fugacities of Gaseous Solutions, Chem. Rev., 44,
233.

4.

Peng, D. Y., and Robinson, D. B., 1976, A New Two-constant Equation of


State for Fluids and Fluid Mixtures, Ind. Eng. Chem. Fundam., 15, 58-64.

May 1994

Section 1.2

Alpha
Formulations

Thermodynamic Methods

The temperature-dependent parameter a(T) can be rewritten as:


a(T) = (T) a(Tc)

(5)

In equation (5), (T) is a temperature-dependent function which takes into


account the attractive forces between molecules. The accuracy of the equation of
state for pure-component vapor pressures (and therefore to a large extent for mixture phase equilibria) depends on the form of the alpha formulation, (T), from
equation (5). The real-gas behavior approaches that of the ideal gas at high temperatures, and this requires that goes to a finite number as the temperature
becomes infinite. Three basic requirements for the temperature-dependent alpha
function must therefore all be satisfied:
1.

The function must be finite and positive for all temperatures.

2.

The function must equal unity at the critical point.

3.

The function must approach a finite value as the temperature approaches


infinity.

For the Redlich-Kwong equation of state, which works well for the vapor
phase at high temperatures, (T) is given by:
(T) = Tr

(6)

1 / 2

PRO/II allows the user to utilize a choice of 11 different alpha formulations


for cubic equations of state (SRK, PR, modified SRK or PR, or UNIWAALS). Table 1.2.4-3 shows the 11 available formulations for (T).

PRO/II Component and Thermophysical Properties


Reference Manual

Equations of State

I-71

Thermodynamic Methods

Section 1.2

Table 1.2.4-3: Alpha Formulations


Form

Equation

Reference
Soave (1972)

01

= 1 + C11 T 0.5
r

02

3
= C1 + C21 T C
r

C2

= 1 + (1 Tr) C1 +
Tr

Soave (1979)

03

04

2
= exp C1 1 T C
r

Boston-Mathias (1980)

05

2 1)
2
= T 2(C
exp C1 1 T 2C
r
r

Twu (1988)

06

3(C2 1)
2C3
=TC
exp C1 1 T C
r
r

Twu-Bluck-CunninghamCoon (1991)
(Recommended by SimSci)

2C1
= exp
1 Tr( C1 + 1) 2

1
+
C
1

Alternative for form (04)

07

08

C2
3
=TC
r exp C11 T r

Alternative for form (06)

09

0.5
0.5
= 1 + C11 T 0.5
r + C21 T r + C31 T r

10

= 1 + C11 Tr0.5 + C21 Tr 0.7 Tr

11

= exp C1 1 Tr + C21 T 0.5


r

Peng-Robinson (1980)

3 2

Mathias-Copeman (1983)

Mathias (1983)

Melhem-Saini-Goodwin
(1989)

where:
C1, C2, C3 = constants
Tr = T/Tc = reduced temperature

23.5

I-72

Equations of State

PRO/II Note: For more information on using alpha functions, see Section 23.5,
Cubic Equation of State Alpha Formulations, of the PRO/II Keyword Input
Manual.

May 1994

Section 1.2

Thermodynamic Methods

References

Mixing Rules

1.

Soave, G., 1972, Equilibrium Constants from a Modified Redlich-Kwong


Equation of State, Chem. Eng. Sci., 35, 1197.

2.

Soave, G, 1979, Application of a Cubic Equation of State to VaporLiquid Equilibria of Systems Containing Polar Components, Inst. Chem.
Eng. Symp. Ser., No. 56, 1.

3.

Boston, J. F., and Mathias, P. M., 1980, Phase Equilibria in a Third


Generation Process Simulator, Proc. of the 2nd Inter. Conf. on Phase
Equil. & Fluid Properties in the Chemical Process Industries, Berlin
(West), March 17-21.

4.

Twu, C. H., 1988, A Modified Redlich-Kwong Equation of State for Highly


Polar, Supercritical Systems, Inter. Symp. on Thermodynamics in Chemical Engineering and Industry, May 30-June 2.

5.

Twu, C.H., Bluck, D., Cunningham, J.R., and Coon, J.E., 1991, A Cubic
Equation of State with a New Alpha Function and New Mixing Rule,
Fluid Phase Equil., 69, 33-50.

6.

Mathias, P. M., and Copeman, T. W., 1983, Extension of the PengRobinson Equation of State to Complex Mixtures, Fluid Phase Equil., 13,
91-108.

7.

Mathias, P. M., 1983, A Versatile Phase Equilibrium Equation of State, Ind.


Eng. Chem. Proc. Des. Dev., 22, 358-391.

8.

Melhem, G. A., Saini, R., and Goodwin, B. M., 1989, A Modified Peng-Robinson Equation of State, Fluid Phase Equil., 47, 189-237.

The accuracy of correlating vapor-liquid equilibrium data using a cubic equation


of state can be improved further by choosing an appropriate mixing rule for calculating a and b in equation (1) for mixtures. The original mixing rule was derived from the van der Waals one-fluid approximation:
a = xi xj aij
i

(7)

b = xi bi

(8)

where:
xi =

mole fraction of component i.

The binary interaction parameter, kij, is introduced into the mixing rule to correct the geometric mean rule of parameter a in the general cubic equation of
state (1):

PRO/II Component and Thermophysical Properties


Reference Manual

Equations of State

I-73

Thermodynamic Methods

Section 1.2
(9)

1 2
aij = (aiaj) / (1kij)

where:
kij = kji =

binary interaction parameter.

The original mixing rule is capable of representing vapor-liquid equilibria


for nonpolar and/or slightly polar systems using only one (possibly temperature-dependent) binary interaction parameter.

Soave-Redlich
Kwong (SRK)

In 1972, to improve the prediction of the vapor pressure of pure components,


and thus multicomponent vapor-liquid equilibria, Soave proposed the following form of (T):
(10)

1 2

(T) = 1 + M1Tr /

M = 0.480 + 1.574 0.176

(11)

where:
Tr =

reduced temperature, T/Tc

acentric factor

The constants in (11) were obtained from the reduction of vapor-pressure data
for a limited number of common hydrocarbons. This limits the use of the SRK
equation of state to nonpolar components. This equation of state does not
accurately predict the behavior of polar components or light gases such as hydrogen. However, the simplicity of equations (10) and (11), and its accuracy
for calculating vapor pressures at temperatures higher than the normal boiling point for hydrocarbons allowed it to gain widespread popularity in industry. PRO/II contains correlations for the kijs of hydrocarbons with N2, O2, H2,
H2S, CO2, mercaptans and other sulfur compounds.

Peng-Robinson
(PR)

23.2

I-74

Equations of State

The form of (T) proposed by Peng and Robinson in 1976 is the same as
that proposed in 1972 by Soave. The numerical values for the constants in
equation (11) are different because the volume function is different and because a somewhat different set of data was used.
PRO/II Note: For more information on using SRK and PR equations of state in
PRO/II, see Section 23.1, Soave-Redlich-Kwong, and Section 23.2, PengRobinson, of the PRO/II Keyword Input Manual.

May 1994

Section 1.2

Soave-RedlichKwong KabadiDanner (SRKKD)

Thermodynamic Methods

While the K-values between the hydrocarbon-rich liquid phase and vapor
phase can be accurately predicted by most cubic equations of state, the
K-values involving the water-rich liquid phase are not. In order to apply
cubic equations of state to water-hydrocarbon systems, Kabadi and Danner
in 1985 proposed a two-parameter mixing rule for the SRK equation of state.
This proposed mixing rule is composition-dependent and is designed expressly for water and well-defined hydrocarbon systems:
(12)

1 2
2
a = xixj(aiaj) / (1kij) + awi xw xi
i

iw

0.8
awi = Gi1Trw

(13)

Gi = gj

(14)

where:
Trw = T/Tcw
awi =

interaction parameter between hydrocarbons and water in


the hydrocarbon-rich phase

gj =

hydrocarbon group contribution from group j

Gi =

sum of group contributions from the different structural


groups forming a hydrocarbon molecule i

To provide estimates for water/hydrocarbon equilibria when no data are available, Kabadi and Danner developed a procedure for estimating the binary interaction parameters kij and Gi. Within a homologous series of hydrocarbons,
kij was found to be approximately constant, and recommended values were
given for seven hydrocarbon classes. A group contribution method was proposed for estimating Gi.
One limitation of this method, however, is that the solubility of hydrocarbon in
the aqueous phase is predicted only within an order of magnitude.

20.6

PRO/II Note: See Section 21, Application Guidelines, and Section 20.6, FreeWater Decant Considerations, of the PRO/II Keyword Input Manual for information on how to handle hydrocarbon-water systems in PRO/II.

Reference
Kabadi, V. N., and Danner, R. P., 1985, A Modified Soave-Redlich-Kwong
Equation of State for Water-Hydrocarbon Phase Equilibria, Ind. Eng.
Chem. Proc. Des. Dev., 24, 537-541.

PRO/II Component and Thermophysical Properties


Reference Manual

Equations of State

I-75

Thermodynamic Methods

Soave-RedlichKwong
PanagiotopoulosReid (SRKP) and
Peng-Robinson
PanagiotopoulosReid (PRP)

Section 1.2

In 1986 Panagiotopoulos and Reid proposed an asymmetric mixing rule containing two parameters for the SRK and PR equations of state (denoted as
SRKP and PRP). The interaction parameter they proposed to be used in
equation (7) is given by:
aij = (ai aj) 1 / 2( 1kij) + (kijkji) xi

(15)

The two adjustable interaction parameters are kij and kji. This asymmetric
definition of the binary interaction parameters significantly improves the
accuracy in correlating binary data for polar and nonpolar systems. This
mixing rule has been used to test several systems, including low pressure
nonideal systems, high pressure systems, three-phase systems, and systems
with supercritical fluids. The results in all cases reported are in good agreement with experimental data.

Reference
Panagiotopoulos, A. Z., and Reid, R. C., 1986, A New Mixing Rule for Cubic
Equations of State for Highly Polar Asymmetric Systems, ACS Symp. Ser.
300, American Chemical Society, Washington, DC, 71-82.

The Panagiotopoulos-Reid mixing rule, however, is fundamentally inconsistent for multicomponent systems. This inconsistency is exhibited in two
(related) flaws:

I-76

Equations of State

1.

The dilution of the mixture with additional components (reducing all the mole
fractions, xi) nullifies the effect of the second binary parameter kji. In the
limit of an infinite number of components so that all the xi approach zero,
the mixing rule reduces to the original van der Waals mixing rule,
equation (9).

2.

The mixing rule is not invariant to dividing a component into a number of


identical pseudocomponents. For example, if methane in a mixture is divided arbitrarily into alpha and beta methane, the calculated properties of the mixture will be slightly different.

May 1994

Section 1.2

Soave-RedlichKwong Modified
PanagiotopoulosReid (SRKM) and
Peng-Robinson
Modified
PanagiotopoulosReid (PRM)

Thermodynamic Methods

SimSci has modified equation (15) in a way that eliminates the first of the
two flaws noted above. This improvement provides better predictions of
properties for multicomponent systems:

aij = (aiaj)1 / 2 (1kij) + (kij kji) (xi / xi + xj)cij

(16)

Equation (16) is identical to equation (15) for binary systems if c12 = 1. The
expression for aji, which is similar to equation (16), can be obtained by interchanging subscripts i and j. The four adjustable interaction parameters are kij
and kji, and cij and cji. For binary nonpolar systems, where deviations from
ideality are not large, or are only weakly asymmetric, only two parameters,
k12 and k21, are sufficient to fit the data (i.e., c12 = c21 = 1). In this case, equation (16) becomes identical to the mixing rule proposed (also for the purpose
of overcoming the first flaw mentioned above) by Harvey and Prausnitz in
1989. For binary polar or polar-nonpolar systems, where the nonideality is
large or strongly asymmetric, it may be necessary to include the additional
parameters c12 and c21. In particular, for binary polar-nonpolar systems,
which have the greatest deviation from ideality, c12 is not set equal to c21.
For binary polar systems however, c12 can generally be set equal to c21.
Reference
Harvey, A.H., and Prausnitz, J.M., 1989, Thermodynamics of High-Pressure
Aqueous Systems Containing Gases and Salts, AIChE J., 35, 635-644.

Soave-RedlichKwong SimSci
(SRKS)

In 1991, Twu et al. proposed another modified mixing rule that eliminated
both of the inconsistencies of the Panagiotopoulos-Reid mixing rule noted
above. For a binary system, the mixing rule can be expressed in the following form for a12:

2
1 2
a12 = (a1a2) / (1k12 + (H12G12x2 / x1+G12x2)

(17)

H12 = (k21 k12)

(18)

G12 = exp(12H12)

(19)

PRO/II Component and Thermophysical Properties


Reference Manual

Equations of State

I-77

Thermodynamic Methods

Section 1.2

The four adjustable parameters are: k12, k21, 12, and 21. Again, as for the
SRKM equation of state, for binary nonpolar systems, where deviations from
ideality are not large or are only weakly asymmetric, only two parameters, k12
and k21, are sufficient to fit the data (i.e., 12 = 21 = 1). For binary polar or polarnonpolar systems, where the nonideality is large or strongly asymmetric, it may be
necessary to include the additional parameters 12 and 21. In particular, for binary polar-nonpolar systems, which have the greatest deviation from ideality,
12 is not set equal to 21. For binary polar systems however, 12 can generally be set equal to 21. Twu et al. have derived the activity coefficients from
the SRKS equation of state and have found that, for a binary system, k12 or
k21 are directly related to the infinite-dilution activity coefficients 1 or 2,
respectively. The values of k12 and k21 are therefore determined when both
values of the infinite-dilution activity coefficients are known for a binary
system. The physical meaning of the binary parameters k12 and k21 is that
they are used to locate the infinite-dilution activity coefficients in a binary
system containing components 1 and 2. After both end points of the liquid
activity coefficients are found, the parameters 12 and 21 are then required
to describe the shapes of the liquid activity coefficient curves for
components 1 and 2 in the finite range of concentration. In general, for real
systems, kij is not equal to kji, and ij and ji are not equal to zero. The conventional mixing rule, obtained by setting k12 = k21 and 12 = 21 = 0 for a
binary system, either results in a compromise of the phase-equilibrium
representation, or fails to correlate highly asymmetric systems.
For a multicomponent system, equation (17) can be generalized as:
(20)
1 2
1 3 1 3
1 6 3

am = xixj (aiaj) / (1kij) + xi Hij / Gij / (aiaj) / xj / Gijxj

i j
i
j
j

Hij = (kji kij)

(21)

Gij = exp (ijHij)

(22)

where:
subscript m refers to the multicomponent mixture.
Reference
Twu, C.H., Bluck, D., Cunningham, J.R., and Coon, J.E., 1991, A Cubic
Equation of State with a New Alpha Function and New Mixing Rule,
Fluid Phase Equil., 69, 33-50.

I-78

Equations of State

May 1994

Section 1.2

Soave-RedlichKwong HuronVidal (SRKH) and


Peng-Robinson
Huron-Vidal
(PRH)

Thermodynamic Methods

The previous SRK and PR mixing rule modifications include compositiondependence for applying these equations of state to complex mixtures. A
more complicated way to represent the phase behavior of strongly nonideal
systems is to develop the relationship between the mixing rule and excess
Gibbs free energy, such that the infinite-pressure Gibbs free energy could be
expressed by a NRTL-like method (see Section 1.2.6, Liquid Activity Coefficient Methods). This approach was proposed by Huron and Vidal in 1979.
The general equation relating excess Gibbs free energy to fugacity coefficients is given by:
E
g = RT ln xi ln i

(23)

where:
gE =

excess Gibbs free energy per mole

fugacity coefficient of the mixture

i =

fugacity coefficient of pure component i

At infinite pressure, the excess Gibbs free energy is calculated using the
Redlich-Kwong equation of state and linear mixing rules for the parameter b
from the general cubic equation of state. At infinite pressure, equation (23)
then becomes:
g = a / b (aixi / bi) ln(2)

(24)

where:
gE =

the excess Gibbs free energy at infinite pressure

Equation (24) can be rewritten to produce a new mixing rule for the cubic
equation of state parameter a:
a = b (aixi / bi) g / ln (2) ln(2)

(25)

The excess Gibbs free energy can be calculated by any liquid activity
method. Huron and Vidal chose to use the NRTL liquid activity method to
calculate gE:
E
g = xi xj Gji cji / xkGki

i
k
j

(26)

cij = gji gii

(27)

Gji = bjexpajicji / RT

(28)

PRO/II Component and Thermophysical Properties


Reference Manual

Equations of State

I-79

Thermodynamic Methods

Section 1.2

The only difference between the classical NRTL equation and equations (2628) given above are the definition of the local composition as corrected volume fractions, which leads to the introduction of the volume parameter bj in
the calculation of Gji. Substituting for gE in equation (25) yields:
a = bxi (ai / bi) 1 / ln(2) (xj Gji cji) / xkGki

i
j
k

(29)

By regressing experimental data to obtain the parameters in the modified


NRTL expression, excellent representation of vapor-liquid equilibria can be
made for several systems. The Huron-Vidal mixing rules are highly empirical
in nature. However, the prediction of equilibria at low densities is reasonable, and the equation of state can be expected to yield better results at
higher pressures, because the mixing rules have been derived at the infinitepressure limit of the excess Gibbs free energy. One limitation of this model
is that it cannot directly utilize parameters for the NRTL method correlated
from low temperature data. This is because an excess Gibbs energy model
from an equation of state at infinite pressure cannot be equated with an activity coefficient excess Gibbs energy model at low pressure.

23.3

PRO/II Note: See Section 23.3, Modified Soave-Redlich-Kwong and Peng-Robinson, of the PRO/II Keyword Input Manual for more information on using
the modified SRK and PR methods in PRO/II.

Reference
Huron, M. J., and Vidal, J., 1979, New Mixing Rules in Simple Equations of
State for Representing Vapor-Liquid Equilibria of Strongly Non-ideal Mixtures, Fluid Phase Equil., 3, 255-271.

HEXAMER

I-80

Equations of State

Hydrogen fluoride is an important chemical used in many vital processes, including HF alkylation, and in the manufacture of refrigerants and other halogenated compounds. Unlike hydrocarbons, however, hydrogen fluoride is
polar and hydrogen bonded, and therefore self-associates not only in the liquid phase, but also in the vapor phase. Experimental evidence strongly suggests that the HF vapor exists primarily as a mixture of monomer and
hexamer. In addition, evidence points to the hexamer existing in the form of
a cyclic benzene-like species. This behavior results in significant departures
from ideality, especially in calculating fugacity coefficients, vapor compressibility factors, heat of vaporization, and enthalpies.

May 1994

Section 1.2

Thermodynamic Methods

Twu et al. (1993), developed a cubic equation of state with a built-in chemical equilibrium model to account for HF association. The cubic equation of
state incorporating association is given by:
P = nr

nr =
v=

(30)

a(T)
RT

(v b) v2 + ubv + wb2

(31)

nT

n0
(32)

V
n0

where:
a(T) = (T)a(Tc ) =

Redlich-Kwong equation-of-state
parameter which refers to the monomer

b=

Redlich-Kwong equation-of-state parameter which refers to


the monomer

v=

molar volume

V=

total volume

nr =

extent of association

nT =

total number of moles of monomer and hexamers

n0 =

the number of moles that would exist in the absence of


association

Note: Only 1 hexamerizing component (HF) may be present when using the
HEXAMER method.
The values of a(Tc) and b can be obtained from the critical constants for the
Redlich-Kwong equation of state (see Table 1.2.4-2) and the critical temperature
and pressure for HF. The alpha function, (T), is obtained by matching the equation of state to HF vapor pressure data. Comparing equation (30) above to the
general two-parameter equation of state given by equation (1), it can be seen that
the only difference is the term nr, which accounts for the contribution of association. The value of nr is 1.0 when there is no association, and approaches 1/6
when there is complete hexamerization. As the temperature increases, the extent
of hexamerization should decrease, i.e., the value of nr should increase.
The total number of moles of monomer and hexamer, nT, and the total number of moles that would exist in the absence of association, n0, are related by:
(33)

n0
=
iz , i = 1,6
nT i
i

where:
zi =

the true mole fraction of species i

ni =

number of moles of species i

PRO/II Component and Thermophysical Properties


Reference Manual

Equations of State

I-81

Thermodynamic Methods

Section 1.2

The hexamerization equilibrium reaction is written as:


6(HF) (HF)6

(34)

The corresponding chemical equilibrium constant for this reaction, which is


a function of temperature only, is defined as:
K(T) =

6 z6 1
6 6 5
1 z1 P

(35)

where:
K=

equilibrium constant

1 =

fugacity coefficient of the true monomer species

6 =

fugacity coefficient of the true hexamer species

z1 =

true mole fraction of the monomer species

z6 =

true mole fraction of the hexamer species

P=

total pressure

The fugacity coefficients in equation (35) are found from the cubic equation of
state using classical thermodynamics. Then, by substituting equation (33) into
equation (35) and using the overall material balance, this reduces to:
6

(36)

K z1 = (6 5 z1) (1 z1)

where:
5

RT
K = K
= the reduced equilibrium constant
v
b
Once the equilibrium constant K is known, equation (36) can be solved to obtain a value for z1 and a corresponding value for z6.
The equilibrium constant for HF hexamerization can be calculated from the
following relationship:
log10 K = 43.65 +

(37)

8910

where:
K=

equilibrium constant, 1/mmHg5

T=

temperature, K

Twu et al. have shown that, at the critical point, the values of z1, the true
mole fraction of monomer, and nr are given by:
z1(Tc) = 0.7006

(38-39)

nr (Tc) = 0.4005

I-82

Equations of State

May 1994

Section 1.2

Thermodynamic Methods

So, even at the critical point, there is still a considerable amount of the hexamer species present.
Mixture properties are computed by using the SRKS mixing rule, equation
(20), discussed previously.

23.7

PRO/II Note: See Section 23.7, Associating Equation of State, of the PRO/II
Keyword Input Manual for more information on using the HEXAMER method
in PRO/II.

Reference
Twu, C. H., Coon, J. E., and Cunningham, J. R., 1993, An Equation of State
for Hydrogen Fluoride, Fluid Phase Equil., 86, 47-62.

UNIWAALS

In the UNIWAALS model proposed by Gupte et al. in 1986, a cubic equation of


state is combined with an excess Gibbs free energy model. By using this approach, the same parameters of the excess Gibbs free energy model based on
low pressure VLE data can be extended to apply to high pressures by using the
equation of state. This is a valuable method because group interaction parameters from group contribution methods such as UNIFAC (see Section 1.2.6, UNIFAC) are readily available for numerous groups. The equations for the
UNIWAALS method are developed by equating the gE derived from the van der
Waals equation of state to the gE derived from UNIFAC at the system temperature and pressure. This equality produces the following mixing rule:
a / RTb = fPv / RT lnP(vb) / RT xiln(P(vibi) / RT +

(40)

f xiai / (fiRTBi) fg / RT
E

f=b/v

(41)

fi = bi / vi

(42)

where:
vE =

excess volume

The mixing rule for the a/b parameter contains the mixture (v) and pure (vi)
fluid volumes. The volumes of the pure components are obtained for the liquid phase at the given temperature and pressure conditions. The parameter b
for the mixture is calculated using the original mixing rule developed for the
RK equation of state given in equation (8), and UNIFAC is used to calculate
gE/RT. Subsequently, the van der Waals equation of state and equation (40)
are solved simultaneously to obtain the mixture volume, v, and a/RTb.

PRO/II Component and Thermophysical Properties


Reference Manual

Equations of State

I-83

Thermodynamic Methods

23.4

Section 1.2

PRO/II Note: See Section 23.4, UNIWAALS, of the PRO/II Keyword Input
Manual, for more information on using the UNIWAALS method in PRO/II.
Several limitations to this method should be noted:
1.

For the calculation of the parameter a, the mixture and pure-component


liquid volumes (v and vi) are required, even if the liquid phase does not
actually exist at the given temperature and pressure.

2.

The mixture parameter v is volume dependent, and thus pressure and volume become related through a differential equation, rather than through a
conventional algebraic equation.

3.

The critical constraints of the UNIWAALS equation of state are no longer satisfied by the values of the parameters a and b at the critical temperature. The
resulting equation of state is no longer a cubic equation of state, and analytical
solution of the equation of state is impossible.

4.

The fugacity coefficients are cumbersome to evaluate.

5.

The accuracy of the UNIWAALS model is not better than that of the UNIFAC model at low temperatures, and the accuracy deteriorates with increasing temperatures.

Reference
Gupte, P. A., Rasmussen, P., and Fredenslund, A., 1986, A New
Group-Contribution Equation of State for Vapor-Liquid Equilibria, Ind.
Eng. Chem. Fundam., 25, 636-645.

Benedict-WebbRubin-Starling

The Benedict-Webb-Rubin equation of state was first proposed in 1940 to predict liquid and vapor properties at high temperatures, and to correlate vapor-liquid equilibria for light hydrocarbon mixtures. This original (BWR) equation of
state however provided poor results at low temperatures and around the critical
point. To improve the accuracy of this equation in predicting thermodynamic
properties for light hydrocarbons in the cryogenic liquid, gas, and dense fluid regions, and at high temperatures, the BWR equation was modified by Starling in
1973 to give the following form:
2
3
4
2
P = RT + B0RT A0C0 / T + D0 / T E0 / T +

(43)

(bRT a d / T) + (a+d / T) + c / T (1+ )exp( )


3

The eleven parameters for pure components (B0, A0, etc.) are generalized as
functions of component acentric factor, critical temperature, and critical density. The mixing rules for the eleven mixture parameters are analogous to the
mixing rules used for the BWR equation. The single binary interaction parameter for the BWRS equation of state is built into the mixing rules. The
BWRS equation of state can correlate pure-component properties for light
hydrocarbons very accurately when experimental data covering entire ranges
are available.

I-84

Equations of State

May 1994

Section 1.2

23.6

Thermodynamic Methods

PRO/II Note: See Section 23.6, Benedict-Webb-Rubin-Starling, of the PRO/II


Keyword Input Manual for information on using the BWRS method in PRO/II.
Limitations to the BWRS equation of state are given below:
1.

Because the equation is generalized in terms of critical temperatures, critical density, and acentric factor, it has difficulty predicting properties for
heavy hydrocarbons and polar systems.

2.

The BWRS equation does not satisfy the critical constraints, and therefore
the equation is inferior to cubic equations of state when applied to the critical and supercritical regions.

3.

The BWRS equation is less predictive than cubic equations of state for
mixture calculations.

4.

Unlike cubic equations of state, BWRS cannot be solved analytically and


normally requires more CPU time.

References

Lee-Kesler-Plcker
(LKP)

1.

Benedict, M., Webb, G. R., and Rubin, L. C., 1940, An Empirical Equation
for Thermodynamic Properties of Light Hydrocarbons and Their Mixtures.
I. Methane, Ethane, Propane, and Butane, J. Chem. Phys., 8, 334-345.

2.

Starling, K. E., 1973, Fluid Thermodynamic Properties for Light Petroleum


Systems, Gulf Publishing Company, Houston, TX.

The LKP equation (available in versions 3.5 and later) is based on the
Benedict-Webb-Rubin equation of state and on Pitzers extended theory of
corresponding states. Thermodynamic data are correlated as a function of
critical temperature and pressure and the acentric factor as follows:

Z = Zo + Zr Zo

(44)

where:
Z=

compressibility factor

acentric factor

subscripts o, r denote Simple and Reference fluids, respectively.


The work of Plcker et al., introduces new mixing rules which are purported by
the authors to better handle mixtures of asymmetric molecules. This is accomplished by the introduction of an exponent, , into the mixing rules.

PRO/II Component and Thermophysical Properties


Reference Manual

Equations of State

I-85

Thermodynamic Methods

Section 1.2

The mixing rules proposed here are:


Tc, mix =

(45)

zj zk Vcjk Tcjk

Vc, mix j

Vc, mix = zj zk Vcjk


j

mix

(46)

= zj j

(47)

where:
Vc =

the molar critical volume

Tc =

the critical temperature

z=

mole fraction in liquid or vapor phase

the acentric factor

The cross coefficients are given by:


12

Tcjk = Tcj Tck

Vcjk =

(48)

Kjk

(49)

3
1 1 3
1 3
Vcj + Vck

where:
Kjk is an adjustable binary parameter, characteristic of the j-k binary,
independent of temperature, density, and composition.
The pseudo-critical pressure is found by:
Tc, mix
Pc, mix = 0.2905 0.085 mix R

vc, mix

(50)

When is zero, the mixing rules are similar to those of Prausnitz and Gunn;
when is 1.0, the mixing rules become the van der Waals mixing rules, as
used by Leland et al. For symmetric mixtures, is zero; for strongly asymmetric mixtures, is a positive value less than unity. Based on an analysis of
experimental data, the authors suggest using a value of 0.25 when a specific
determination is not available. PRO/II uses a default value of 0.25.
Adjustable binary parameters, Kijs are also used in the mixing rules. Values
reported by Plcker et al. have been incorporated into PRO/II. The LKP
method is claimed by the authors to be superior to Starlings BWRS equation
for highly asymmetric systems. The method is not accurate around the critical point because the mixture critical constants are empirical, and do not represent the true critical point. Therefore, the authors recommend that the
method not be used above a reduced temperature of 0.96.

I-86

Equations of State

May 1994

Section 1.2

Thermodynamic Methods

References
1.

Lee, B.I., and Kesler, M.G., 1975, A Generalized Thermodynamic


Correlation Based on Three-Parameter Corresponding States, AIChE J., 21,
510-527.

2.

Leland, T.W., and Mueller, W.H., 1959, Applying the Theory of


Corresponding States to Multicomponent Mixtures, Ind. Eng. Chem., 51,
597-600.

3.

Pitzer, K.S., and Hultgren G.O., 1958, The Volumetric and Thermodynamic
Properties of Fluids, V. Two Component Solutions, J. Am. Chem. Soc., 80,
4793-96.

4.

Plcker, U., Knapp, H., and Prausnitz, J.M., 1978, Calculation of High-Pressure Vapor-Liquid Equilibria from a Corresponding States Correlation
with Emphasis on Asymmetric Mixtures, Ind. Eng. Chem. Proc. Des. Dev.,
17, 324-332.

5.

Prausnitz, J.M., and Gunn, R.D., 1958, Volumetric Properties of Nonpolar


Gaseous Mixtures, AIChE J., 4, 430-35.

6.

Prausnitz, J.M., and Gunn, R.D., 1958, Pseudocritical Constants from


Volumetric Data for Gas Mixtures, AIChE J., 4, 494.

PRO/II Component and Thermophysical Properties


Reference Manual

Equations of State

I-87

Thermodynamic Methods

1.2.5

Section 1.2

Free Water Decant


General
Information

In many hydrocarbon-water mixtures, including those found in refinery and


gas processing plants, the water phase formed is nearly immiscible with the
liquid hydrocarbon phase. For such systems, the water phase can be assumed
to decant as a pure aqueous phase. This reduces the number of computations
involved with rigorous VLLE methods. The water-decant method as implemented in PRO/II follows these steps:
Water vapor is assumed to form an ideal mixture with the hydrocarbon
vapor phase.
The water partial pressure is calculated using one of two methods.
The pressure of the system, P, is calculated on a water-free basis by
subtracting the water partial pressure.
A pure water liquid phase is formed when the partial pressure of water
reaches its saturation pressure at that temperature.
The amount of water dissolved in the hydrocarbon-rich liquid phase is
computed using one of a number of water solubility correlations.

Note: The free water decant option may only be used with the Soave-RedlichKwong, Peng-Robinson, Grayson-Streed, Chao-Seader, Improved GraysonStreed, Erbar modifications to Grayson-Streed and Chao-Seader, Braun K10,
or Benedict-Webb-Rubin-Starling methods. Water decant is automatically activated when one of these methods is selected.

Calculation
Methods

The amount of water dissolved in the hydrocarbon-rich liquid phase can be


computed once the water K-values, Kw, are known. These are calculated using the following relationship:
Kw = P w / P xw

(1)

where:
Pw =

water partial pressure at temperature T

xw =

solubility of water in the hydrocarbon-rich liquid phase

P=

system pressure

The water partial pressure is calculated using either the ASME steam tables
or Chart 15-14 in the GPSA Data Book. The GPSA Data Book option is recommended for natural gas mixtures above 2000 psia. Two methods are available for calculating water properties:

I-88

Free Water Decant

May 1994

Section 1.2

Thermodynamic Methods

Water properties can be calculated assuming saturated vapor and liquid


conditions.
Water properties can be calculated using steam tables based on the
Keenan and Keyes equation of state.
The water solubility, xw, can be computed by one of three methods in PRO/II:
The default method developed by SimSci. In this method, water solubility is calculated for individual hydrocarbons and light gases given in
Table 1.2.5-1. The SimSci method uses a correlation based on the
number of carbon and hydrogen atoms present in the component. For
pseudocomponents, the water solubility is calculated as a function of the
Watson (UOP) K-factor.
The second method uses Figure 9A1.2 in the API Technical Data Book
to compute water solubility in kerosene. PRO/II will automatically
invoke this option if the SIMSCI solubility option is chosen, and a
component not included in Table 1.2.5-1 is present in the system.
The third method employs the equation-of-state method that is being
used for calculating the K-values of the other components present in the
system to compute the water K-value. Missing binary interaction parameters for the water-hydrocarbon component pairs are estimated using
the Soave-Redlich-Kwong Kabadi-Danner equation of state. This
method is only valid for SRK or PR equations of state.

Table 1.2.5-1:
Components Available in the SIMSCI Water Solubility Method

20.6

Paraffins

Naphthenes

Unsaturated Hydrocarbons

Aromatics

Methyl Mercaptan

CS2

NH3

Argon

CO2

Helium

HCl

H2S

N2

NO

O2

SO2

PRO/II Note: For more information on using the free-water decant option, see
Section 20.6, Free Water Decant Considerations, of the PRO/II Keyword Input
Manual.

PRO/II Component and Thermophysical Properties


Reference Manual

Free Water Decant

I-89

Thermodynamic Methods

1.2.6

Section 1.2

Liquid Activity Coefficient Methods


General
Information

Liquid activity coefficient methods for phase-equilibrium calculations differ


at a fundamental level from equation-of-state (EOS) methods. In EOS methods, fugacity coefficients (referring to an ideal-gas state) are computed for
both vapor and liquid phases. In activity coefficient methods, the reference
state for each component in the liquid phase is the pure liquid at the temperature and pressure of the mixture. It is often more convenient and accurate to
use this approach when the liquid phase is a mixture of components which
do not differ greatly in volatility; it is also often easier to describe strongly
nonideal systems with a liquid activity coefficient model than with an equation of state.
The thermodynamic methods of liquid mixtures within an activity coefficient
framework are covered in standard textbooks, a few of which are referenced
at the end of this section. The activity coefficient is introduced in the way
the fugacity of component i in the liquid phase is written:
L

(1)

oL

f i = i xi f i

where:
fLi =

fugacity of component i in liquid phase

foL
i

standard-state liquid fugacity of component i

xi =

mole fraction of component i in liquid

i =

liquid-phase activity coefficient of component i

The standard-state fugacity f oL


i is defined as that of the pure liquid i at the
temperature and pressure of the mixture. With this definition, i approaches
one in the limit xi 1. The standard-state fugacity may be related to the vapor pressure of component i as follows:

f i = Pi i exp
0L

sat

sat

P
sat

Pi

L
vi / RT dP

(2)

where:
P=

system pressure

Psat
i =

vapor pressure of component i at the system temperature

R=

gas constant

T=

system temperature

vLi =

liquid molar volume of component i at T and P

sat
i = fugacity coefficient of pure component i at temperature T
and pressure Psat
i

I-90

Liquid Activity Coefficient Methods

May 1994

Section 1.2

Thermodynamic Methods

The exponential term in equation (2) is the Poynting factor which accounts for
the effect of pressure on the liquid fugacity. If the pressure does not exceed a few
atmospheres, this correction can generally be neglected. Since liquid volumes
do not depend greatly on pressure, equation (2) can be simplified to:
f i = Pi i exp (P Pi ) vi RT

sat

sat

sat

(3)

When liquid activity coefficients are used, any method may be used to compute the vapor-phase fugacity. An ideal gas is often assumed, but in general
vapor fugacities may be written as:
V

(4)

fi = i yiP

where:
fVi =

fugacity of component i in vapor phase

yi =

mole fraction of component i in vapor

fugacity coefficient of component i in vapor

V
i

For an ideal gas, the fugacity coefficient Vi equals one, but it may also be
computed from an equation of state or other correlation.
Equations (1) and (4) are fundamentally different in the way they describe
liquid and vapor fugacities, respectively. The two equations do not in general
match at the vapor-liquid critical point, where vapor and liquid phases become indistinguishable. Phase-equilibrium calculations near a vapor-liquid
critical point must be carried out with some other method such as an equation of state.
The familiar vapor-liquid K-value is defined as the ratio of yi to xi and can be
obtained by combining equations (1) and (4):
(5)

oL

yi i f i
Ki = = V
xi P
i

At low and moderate pressures, the Poynting correction is often ignored and
equation (5) becomes:
sat

Ki =

(5a)

sat

i i Pi
V

iP

Unless there is vapor-phase association (as is the case with carboxylic acids,
for example), the fugacity coefficients may also be ignored at low and moderate pressures. Equation (5) then simplifies to:
(5b)

sat

i Pi
Ki =
P

PRO/II Component and Thermophysical Properties


Reference Manual

Liquid Activity Coefficient Methods

I-91

Thermodynamic Methods

Section 1.2

For most low-pressure systems, the regression of experimental vapor-liquid


equilibrium data will produce essentially the same parameters if equation
(5a) or (5b) is used in place of the full equation (5). This is not necessarily
the case at higher pressures and for systems where vapor-phase nonideality is
important. Significant errors can be introduced when the regression and calculations using the regressed parameters employ differing sets of simplying
assumptions. In general, calculations should be performed using the same assumptions about vapor fugacities and the Poynting factor as those employed
in fitting the parameters. An important exception to this rule is the case
where parameters were fitted at low pressure, but the calculations are at a substantially higher pressure. In such a case, it is best to employ nonideal vaporphase fugacities and the Poynting correction in the calculation even if they were
not used in the original fit.
Liquid activity coefficients are derived from expressions for the excess Gibbs
energy of a liquid mixture. The defining equation is:
1 G
RT ni

T,P,n

(6)

ln i =

ji

where:
GE =

excess Gibbs energy of liquid mixture

ni =

moles of component i in liquid

The following sections describe the expressions available for describing liquid-phase activity coefficients.
References

I-92

1.

Prausnitz, J.M., Lichtenthaler, R.N. and Gomes de Azevedo, E., 1986,


Molecular Thermodynamics of Fluid-Phase Equilibria, 2nd ed., PrenticeHall, Englewood Cliffs, NJ.

2.

Sandler, S.I., 1989, Chemical and Engineering Thermodynamics, 2nd ed.,


John Wiley & Sons, New York.

3.

Smith, J.M. and Van Ness, H.C., 1987, Introduction to Chemical


Engineering Thermodynamics, 4th ed., McGraw-Hill, New York.

4.

Van Ness, H.C. and Abbott, M.M., 1982, Classical Thermodynamics of


Nonelectrolyte Solutions: With Applications to Phase Equilibria,
McGraw-Hill, New York.

5.

Walas, S.M., 1985, Phase Equilibria in Chemical Engineering, Butterworth, Boston.

Liquid Activity Coefficient Methods

May 1994

Section 1.2

Thermodynamic Methods

Margules
Equation
Table 1.2.6-1: Margules Equation
Required Pure-component Properties
Vapor pressure

Application Guidelines
Temperature

Use at or near
temperatures where
parameters were
fitted.

The oldest empirical correlations for liquid activity coefficients, such as the
Margules equation (1895), are derived from simple polynomial expansions.
The most popular form of the Margules equation was proposed by Redlich
and Kister (1948). When that expansion is truncated after the quadratic term,
the resulting three-parameter correlation is known as the four-suffix Margules equation. The resulting expression for the activity coefficient is:
2

ln i = ( 1 xi ) Ai + 2Bi Ai Di xi + 3Di xi

where:

(7)

Ai = xj aij
j=1
N

Bi = xj aji
j=1
N

Di = xj dij
j=1

dij = dji

Thus, for each ij binary pair in a multicomponent system, the parameters are
aij, aji, and dij. No temperature dependence is included in this implementation; one should therefore be cautious about using this equation at temperatures differing substantially from the range in which the parameters were
fitted.
References
1.

Margules, 1895, Sitzber., Akad. Wiss. Wien, Math. Naturw., (2A),


104, 1234.

2.

Redlich, O. and Kister, A. T., 1948, Algebraic Representation of


Thermodynamic Properties and the Classification of Solutions, Ind. Eng.
Chem. 40, 345-348.

PRO/II Component and Thermophysical Properties


Reference Manual

Liquid Activity Coefficient Methods

I-93

Thermodynamic Methods

Section 1.2

van Laar Equation


Table 1.2.6-2: van Laar Equation
Required Pure-component Properties

Application Guidelines

Vapor pressure

Components

Use for chemically similar


components.

Table 1.2.6-2 van Laar Equation

Another old correlation which is still frequently used is the van Laar equation. It
may be obtained by discarding ternary and higher-order terms in an alternative
expansion of the excess Gibbs energy (known as Wohls equation), though that is
not how van Laar derived it originally. The resulting expression for the activity coefficient is:
N

ln i =

1
ail Zl aij Zi Zj
2

l=1

j=1


j=1 k=1
j,kl

(8)
ajk

aij
aji

Zj Zk

where:
Zl =

xl

xj ail
j

li

Two parameters, aij and aji, are required for each binary. As with the Margules equation, no method is included for making the parameters temperature dependent. It should also be noted that the van Laar equation, because of
its functional form, is incapable of representing maxima or minima in the relationship between activity coefficient and mole fraction. In practice, however, such maxima and minima are relatively rare.
References

I-94

1.

van Laar, J. J., 1910, The Vapor Pressure of Binary Mixtures, Z. Phys.
Chem., 72, 723-751.

2.

Wohl, K., 1946, Thermodynamic Evaluation of Binary and Ternary Liquid


Systems, Trans. AIChE, 42, 215-249.

Liquid Activity Coefficient Methods

May 1994

Section 1.2

Thermodynamic Methods

Regular Solution
Theory
Table 1.2.6-3: Regular Solution Theory
Required Pure-component Properties
Vapor pressure

Application Guidelines
Components

Liquid molar volume,


Solubility parameter

Not recommended for polar


components and solutions
containing fluorocarbons.

Hildebrand defined a regular solution as one in which the excess entropy


vanishes when the solution is mixed at constant temperature and constant volume. This is nearly the case for most solutions of nonpolar compounds, provided the molecules do not differ greatly in size. The excess Gibbs energy is
then primarily determined by the attractive intermolecular forces. Scatchard
and Hildebrand made a simple assumption relating mixture interactions to
those in pure fluids; the result is a simple theory in which the activity coefficients are a function of pure-component properties only. The important property is the solubility parameter, which is related to the energy required to
vaporize a liquid component to an ideal-gas state. The activity coefficient expression is:
L
2
RT ln i = v i i j i

(9)

j=

xj vj

xk vk

where:
vLi =

liquid molar volume of component i

i =

solubility parameter of component i

There are no adjustable parameters in regular solution theory. It is useful for


mixtures of nonpolar components, but it should not be used for highly
nonideal mixtures, especially if they contain polar components. Solubility parameters have been tabulated for numerous compounds, and these parameters
are included for most components in PRO/IIs library.
Reference
Hildebrand, J.H., Prausnitz, J. M., and Scott, R. L., 1970, Regular and
Related Solutions, Van Nostrand Reinhold Co., New York.

PRO/II Component and Thermophysical Properties


Reference Manual

Liquid Activity Coefficient Methods

I-95

Thermodynamic Methods

Section 1.2

Flory-Huggins
Theory
Table 1.2.6-4: Flory-Huggins Theory
Required Pure-component Properties
Vapor pressure

Application Guidelines
Components

Liquid molar volume,


Solubility parameter

Best for components which


are chemically similar and
which differ only in size
(e.g., polymer solutions).

Table 1.2.64 Flory-Huggins Theory

The Flory-Huggins model may be considered a correction to the Regular Solution Theory for the entropic effects of mixing molecules which differ greatly in
size. It is therefore suitable for polymer/solvent systems, especially if the molecules involved are nonpolar. In this simplest implementation of the theory, there
are no binary parameters. The activity coefficient expression is:
reg

ln i = lni

L
vL
vi
i
+ ln L + 1 L
v
v

(10)

where:
reg
i =
L

vi =
L

v =

activity coefficient from regular solution theory


liquid molar volume of component i
liquid molar volume of solution

References

I-96

1.

Flory, P. J., 1942, Thermodynamics of Higher Polymer Solutions,


J.Chem.Phys., 10, 51.

2.

Huggins, M. L., 1942, Thermodynamic Properties of Solutions of LongChain Compounds, Ann. N.Y. Acad. Sci., 43, 9.

3.

Misovich, M. J., Grulka, E. A., and Banks, R. F., 1985, Generalized


Correlation for Solvent Activities in Polymer Solutions, Ind. Eng. Chem.
Proc. Des. Dev., 24, 1036.

Liquid Activity Coefficient Methods

May 1994

Section 1.2

Thermodynamic Methods

Wilson Equation
Table 1.2.6-5: Wilson Equation
Required Pure-component Properties

Application Guidelines

Vapor pressure

Components

Liquid molar volume

Useful for polar or


associating components in
nonpolar solvents. Cannot
be used if liquid-liquid
immiscibility exists.

Table 1.2.65 Wilson Equation

The Wilson equation was the first to incorporate the concept of local composition. The basic idea is that, because of differences in intermolecular forces, the
composition in the neighborhood of a specific molecule in solution will differ from
that of the bulk liquid. The two parameters per binary are, at least in principle, associated with the degree to which each molecule can produce a change in the composition of its local environment. The expression for the activity coefficient is:
N

j=1

k=1

ln i = 1 ln xj Aij

(11)

xk Aki
N

xj Akj
j=1

where:
L

Aij =

vi

L
vj
L

Aij =

vi

L
vj

exp

aij

exp

aij

Aij = aij

RT

(when unit of aij is K)

(when unit of aij is KCAL or KJ)


(when unit of aij is NODIME)

vLi = the liquid molar volume of component i


aij represents a characteristic energy of interaction between species i and j.
While there is no explicit temperature dependence in the Wilson equations
parameters, the derivation is such that the equation may be used with some confidence over a wider range of temperatures than either the Margules or van Laar
equations. It is also much more successful in correlating mixtures containing polar components. The Wilson equation cannot describe local maxima or minima
in the activity coefficient. Its single significant shortcoming, however, is that it is
mathematically unable to predict the splitting of a liquid into two partially miscible phases. It is therefore completely unsuitable for problems involving liquidliquid equilibria.

PRO/II Component and Thermophysical Properties


Reference Manual

Liquid Activity Coefficient Methods

I-97

Thermodynamic Methods

Section 1.2

References
1.

Holmes, M. H. and van Winkle, M., 1970, Wilson Equation Used to


Predict Vapor Compositions, Ind. Eng. Chem., 62(1), 22-31.

2.

Orye, R. V. and Prausnitz, J. M., 1965, Multicomponent Equilibria with


the Wilson Equation, Ind. Eng.Chem., 57(5), 18-26.

3.

Wilson, G. M., 1964, Vapor-Liquid Equilibrium XI. A New Expression for


the Excess Free Energy of Mixing, J. Amer. Chem. Soc., 86, 127.

NRTL Equation
Table 1.2.6-6: NRTL Equation
Required Pure-component Properties

Application Guidelines

Vapor pressure

Components

Useful for strongly nonideal


mixtures and for partially
miscible systems.

The NRTL (non-random two-liquid) equation was developed by Renon and


Prausnitz to make use of the local composition concept, while avoiding the
Wilson equations inability to predict liquid-liquid phase separation. The resulting equation has been quite successful in correlating a wide variety of
systems. The expression for the activity coefficient is:

ji Gji xj
ln i =

Gki xk

xj Gij

Gkj xk


ij

xk kj Gkj
k

Gkj xk

(12)

where:
ij = aij +

bij

ij = aij +

bij

RT

+
+

cij
T

cij
2 2

(when unit is K)

(when unit is KCAL or KJ)

R T

Gji = exp(ji ji) , ji = ji + ji T

I-98

Liquid Activity Coefficient Methods

May 1994

Section 1.2

Thermodynamic Methods

Three parameters, ij, ji, and ij = ji, are required for each binary. These
parameters may be made temperature-dependent as described above. If ij is
to be represented with only one constant, it has been found empirically that
better results over a range of temperatures are obtained if only bij is used and
aij = cij = 0. The parameter does not vary greatly from binary to binary,
and it is often satisfactory to fix it at 0.3 for vapor-liquid systems and 0.2 for
liquid-liquid systems.
References
1.

Renon, H. and Prausnitz, J. M., 1968, Local Composition in Thermodynamic Excess Functions for Liquid Mixtures, AIChE J., 14, 135-144.

2.

Harris, R. E., 1972, Chem. Eng. Prog., 68(10), 57.

UNIQUAC
Equation
Table 1.2.6-7: UNIQUAC Equation
Required Pure-component Properties
Vapor pressure

Application Guidelines
Components

van der Waals


area and volume

Useful for nonelectrolyte


mixtures containing polar
or nonpolar components
and for partially miscible
systems.

The UNIQUAC (universal quasi-chemical) equation was developed by Abrams


and Prausnitz based on statistical-mechanical considerations and the latticebased quasi-chemical model of Guggenheim. As in the Wilson and NRTL equations, local compositions are used. However, local surface-area fractions are
used as the primary composition variable instead of volume fractions. Each
molecule i is characterized by a volume parameter ri and a surface-area parameter qi.
The excess Gibbs energy (and therefore the logarithm of the activity coefficient) is divided into a combinatorial and a residual part. The combinatorial
part depends only on the sizes and shapes of the individual molecules; it contains no binary parameters. The residual part, which accounts for the energetic interactions, has two adjustable binary parameters. The UNIQUAC
equation has, like the NRTL equation, been quite successful in correlating a
wide variety of systems. The expression for the activity coefficient is:
C

(13)

ln i = ln i + ln i

_
M
i z
i
i
C
x l
ln i = ln + qi ln + li
xi j j
xi 2
i
j=1

PRO/II Component and Thermophysical Properties


Reference Manual

Liquid Activity Coefficient Methods

(14)

I-99

Thermodynamic Methods

Section 1.2

M
M
R

ln i = qi 1 ln j ji
j=1
j=1

j ij
M

(15)

k kj
k=1

where:
ij = exp

Uij

(when unit is K)

ij = exp

Uij

RT

(when unit is KCAL or KJ)

Uij = aij + bijT


x i qi

i =

xi qi
i=1

lj =

i =

_
z
(r qj) (rj 1)
2 j
xi ri
M

xj rj
j=1

qi =
ri =

Awi
9

2.5 x 10
Vwi
15.17

_
z = 10

Awi =

van der Waals area of molecule i

Vwi =

van der Waals volume of molecule i

Two parameters, Uij and Uji, are required for each binary; they may be made
temperature-dependent as described above. If no temperature-dependence is
used for Uij, better results over a range of temperatures are normally obtained by using aij and setting bij = 0.

I-100

Liquid Activity Coefficient Methods

May 1994

Section 1.2

Thermodynamic Methods

References
1.

Abrams, D. S. and Prausnitz, J. M., 1975, Statistical Thermodynamics


of Mixtures: A New Expression for the Excess Gibbs Free Energy of
Partly or Completely Miscible Systems, AIChE J., 21, 116-128.

2.

Anderson, T. F. and Prausnitz, J. M., 1978, Application of the UNIQUAC


Equation to Calculation of Multicomponent Phase Equilibria. 1. Vapor-Liquid Equilibria, Ind. Eng. Chem. Proc. Des. Dev., 17, 552-561.

3.

Anderson, T. F. and Prausnitz, J. M., 1978, Application of the UNIQUAC


Equation to Calculation of Multicomponent Phase Equilibria. 2. Liquid-Liquid Equilibria, Ind. Eng. Chem. Proc. Des. Dev., 17, 561-567.

4.

Maurer, G. and Prausnitz, J. M., 1978, On the Derivation and Extension of


the UNIQUAC Equation, Fluid Phase Equilibria, 2, 91-99.

UNIFAC
Table 1.2.6-8: UNIFAC
Required Pure-component Properties

Application Guidelines

Vapor pressure

Pressure

Up to 10 atmospheres

van der Waals


area and volume

Temperature

32 - 300 F

Components

All components well below


their critical points

The UNIFAC (universal functional activity coefficient) method was developed in 1975 by Fredenslund, Jones, and Prausnitz. This method estimates
activity coefficients based on the group contribution concept following the
Analytical Solution of Groups (ASOG) model proposed by Deer and Deal in
1969. Interactions between two molecules are assumed to be a function of
group-group interactions. Whereas there are thousands of chemical compounds of interest in chemical processing, the number of functional groups is
much smaller. Group-group interaction data are obtained from reduction of
experimental data for binary component pairs.
The UNIFAC method is based on the UNIQUAC model, which represents
the excess Gibbs energy (and logarithm of the activity coefficient) as a combination of two effects. Equation (13) is therefore used:
C

ln i = ln i + ln i

The combinational term, ln C


i , is computed directly from the UNIQUAC
equation (14) using the van der Waals area and volume parameter calculated
from the individual structural groups:

PRO/II Component and Thermophysical Properties


Reference Manual

Liquid Activity Coefficient Methods

I-101

Thermodynamic Methods

Section 1.2
_
i
i z
i
c
ln i = ln + qi ln + li
xi
xi 2
i

NOC

xj lj

j=1

where:
xi ri
i =NOC

xj rj
j=1

x i qi
i =NOC

xj qj
j=1

_
z
li = (ri qi) (ri 1)
2
NOG

ri = k Rk
i

k=1
NOG

qi = k Qk
i

k=1

where:
NOC = number of components
NOG = number of different groups in the mixture
_
z=
lattice coordination number = 10
i

vk =

number of functional groups of type k in molecule i

Rk =

volume parameter of functional group k

Qk =

area parameter of functional group k

xi =

mole fraction of component i in the liquid phase

The group volume and area parameters are obtained from the atomic and molecular structure.
Rk =
Qk =

(16)

Vwk
15.17

(17)

Awk
2.5x10

where:

I-102

Vwk =

van der Waals volume of group k

Awk =

van der Waals area of group k

Liquid Activity Coefficient Methods

May 1994

Section 1.2

Thermodynamic Methods

The residual term, ln Ri , is given by:


(18)

NOG

R
i
ln i = ln k ln k

k=1

where:
k =
i

k =

residual activity coefficient of group k in the mixture


residual activity coefficient of group k in a reference solution
containing only molecules of group type i. This quantity is required so that Ri 1 as xi 1

The residual activity coefficient is given by:

ln k = Qk 1 ln m mk

m
m

m km

n nm
n

(19)

where:
m, n = 1, 2, ... NOG
The parameter mk is given by:
amk
mk = exp

RT

(20)

where:
amk =

binary interaction parameter for groups m and k

The binary energy interaction parameter amk is assumed to be a constant and


not a function of temperature. A large number of interaction parameters between structural groups, as well as group size and shape parameters have
been incorporated into PRO/II.
References
1.

Deer, E.L., and Deal, C.H., 1969, Inst. Chem. Eng. Symp. Ser., 32(3), 40.

2.

Fredenslund, Aa., Jones, R.L., and Prausnitz, J.M., 1975, Group


Contribution Estimation of Activity Coefficients in Nonideal Liquid
Mixtures, AIChE J., 27, 1086-1099.

3.

Skjold-Jrgensen, S., Kolbe, B., Groehling, J., and Rasmussen, P., 1979,
Vapor-Liquid Equilibria by UNIFAC Group Contribution. Revision and
Extension, Ind. Eng. Chem. Proc. Des. Dev., 18, 714-722.

4.

Gmehling, J., Rasmussen, P., and Fredenslund, Aa., 1983, Vapor-Liquid


Equilibria by UNIFAC Group Contribution. Revision and Extension,
Ind. Eng. Chem. Proc. Des. Dev., 22, 676-678.

5.

Hansen, H.K., Rasmussen, P., Fredenslund, Aa., Schiller, M., and


Gmehling, J., 1991, Vapor-Liquid Equilibria by UNIFAC Group
Contribution. 5. Revision and Extension, Ind. Eng. Chem. Res., 30,
2352-2355.

PRO/II Component and Thermophysical Properties


Reference Manual

Liquid Activity Coefficient Methods

I-103

Thermodynamic Methods

Modifications
to UNIFAC

Section 1.2

The UNIFAC method provides good order-of-magnitude estimates. The


accuracy of the method can be improved by incorporating a temperaturedependent form for the binary group energy interaction parameter.
UFT1 - Lyngby modified UNIFAC
Researchers at Lyngby developed a three-parameter temperature-dependent
form for the binary interaction parameter. The parameter mk is given by:
Amk
mk = exp

RT

(21)

Amk = amk + bmk (T To) + Cmk Tln

+ T To
T

(22)

To

where:
amk, bmk, cmk =
To =

binary interaction parameters

298.15 K

The combinatorial part of the logarithm of the activity coefficient is given by:
c

ln i = ln

(23)

wi
wi
+1
xi
xi

(24)

ri 3 xi

wi = NOC

2
3

ri

xi

i=1

Reference
Larsen, B.L., Rasmussen, P., and Fredenslund, Aa., 1987, A Modified UNIFAC Group Contribution Model for Prediction of Phase Equilibria and
Heats of Mixing, Ind. Eng. Chem. Res., 26, 2274-2286.

UFT2 - Dortmund modified UNIFAC


For this modified method, the temperature-dependent form of Amk is given
by:
2

Amk = amk + bmk T + cmk T

I-104

Liquid Activity Coefficient Methods

(25)

May 1994

Section 1.2

Thermodynamic Methods

The combinational part of the logarithm of the activity coefficient is given


by:
ln

c
i

_
wi zqi ln i
i
wi
= ln + 1
+1
xi
xi 2 i
Qi

(27)

ri 4 xi

wi = NOC

3
4

rj

(26)

xj

j=1

where:
z = lattice coordination number = 10
xi ri
i = NOC

xj rj
j=1

xi qi
i = NOC

xj qj
j=1

Reference
Weidlich, V., and Gmehling, J., 1987, A Modified UNIFAC Model:
1. Prediction of VLE, hE, and , Ind. Eng. Chem. Res., 26, 1372-1381.

UFT3
For this modified UNIFAC method, the temperature-dependent form of Amk
is given by:
(28)

Amk = amk + bmk T + cmk T

The combinatorial and residual parts of the activity coefficient are identical
to those described previously for the UNIFAC method.
Reference
Torres-Marchal, C., and Cantalino, A.L., 1986, Industrial Applications
of UNIFAC, Fluid Phase Equil., 29, 69-76.

PRO/II Component and Thermophysical Properties


Reference Manual

Liquid Activity Coefficient Methods

I-105

Thermodynamic Methods

Section 1.2

UNFV - Free volume modification to UNIFAC


This method was developed for modeling polymer systems. The liquid activity coefficient is given by a combination of the same combinatorial and residual terms as UNIFAC, plus a free volume effect term:
c

(29)

FV

ln i = ln i + ln i + ln i

13

V i

FV
i
ln i = 3 Ci ln
Ci 1
1
V 3 1
V

m
m

1 1


V i 3

(30)

where:

Vi=

Vwi
15.17 bri
NOG

1
ri =
Mi

k Rk
i

k=1

Vi wi + Vj wj

Vm=

15.17b (ri wi + rj wj)

b = 1.28

where:
Vi =

volume per gram of solvent i

Mi =

molecular weight of solvent i

wi =

weight fraction of component i

Ci =

number of effective degrees of freedom per molecule of


solvent i = 3.3

Reference
Oishi, T., and Prausnitz, J.M., 1978, Estimation of Solvent Activities
in Polymer Solutions Using a Group-Contribution Method, Ind. Eng.
Chem. Proc. Des. Dev., 17, 333-339.

I-106

Liquid Activity Coefficient Methods

May 1994

Section 1.2

Fill Methods

24.10

Thermodynamic Methods

The ability of a liquid activity method to accurately predict vapor-liquid


equilibria and/or vapor-liquid-liquid equilibria depends to a great degree on
whether or not binary interaction parameters are available for that method.
PRO/II contains a proven mechanism for filling in missing binary interaction
parameters for liquid activity methods. When a liquid activity method such
as NRTL is selected for phase-equilibrium calculations, and the FILL option
is selected, PRO/II uses the following mechanism in order to obtain the binary interaction data the model needs:
1.

Any user-supplied binary interaction parameters or mutual solubility, infinite dilution, or azeotropic data are used in preference to any other data.

2.

The VLE and LLE databanks which contain binary interaction parameters
are then searched for data.

3.

The SimSci azeotropic databank is searched for appropriate azeotropic


data, which are then regressed to provide binary interaction data.

4.

For VLE calculations, if steps 1 through 3 do not supply the required parameters, then the group contribution methods, UNIFAC or its modification UFT1 or the regular solution method or the Flory-Huggins method,
may be used to provide estimates for the interaction parameters. For LLE
calculations, UNIFAC or the modified UNIFAC method UFT1 is used to
supply the parameter estimates.

PRO/II Note: For more information on using the FILL option, see Section 24.10,
Filling In Missing Parameters, of the PRO/II Keyword Input Manual.
5.

Finally, if binary interaction parameters are still missing after steps 1-4 are
followed, then all missing parameters are set equal to zero.

Figure 1.2.6-1 shows the mechanism used by PRO/II to backfill missing


binary parameters for VLE or VLLE calculations.
For VLLE calculations, in order to avoid conflicts between VLE and LLE binary interaction data, PRO/II follows a number of strict rules when filling in
these binary interaction data.
If no VLE or LLE interaction data are supplied by the user, PRO/II uses
the following order in searching for interaction data for both VLE and
LLE calculations:
1. The LLE databank.
2. The VLE databank.
If the user supplies VLE interaction data only, then PRO/II uses the following order in searching for binary parameters for both VLE and LLE
calculations:
1.
2.
3.

The user-supplied values given on the KVALUE(VLE) statement.


The LLE databank.
The VLE databank.

PRO/II Component and Thermophysical Properties


Reference Manual

Liquid Activity Coefficient Methods

I-107

Thermodynamic Methods

Section 1.2

If the user supplies LLE interaction data only, or both VLE and LLE interaction data, for LLE calculations, the databanks are searched in the order:
1. The user-supplied data given on the KVALUE(LLE) statement.
2. The LLE databank.
3. The VLE databank.
If the user supplies LLE interaction data only, or both VLE and LLE interaction data, for VLE calculations, the databanks are searched in the order:
1.
2.

20.7

I-108

The user-supplied data given on the KVALUE(VLE) statement.


The VLE databank.

PRO/II Note: For more information on specifying VLLE methods, see Section
20.7, Vapor-Liquid-Liquid Equilibrium Considerations, of the PRO/II Keyword Input Manual.

Liquid Activity Coefficient Methods

May 1994

Section 1.2

Thermodynamic Methods

Figure 1.2.6-1:
Flowchart for FILL
Methods

PRO/II Component and Thermophysical Properties


Reference Manual

Liquid Activity Coefficient Methods

I-109

Thermodynamic Methods

Henrys Law

24.11

Section 1.2

When liquid activity methods are used, the standard-state fugacity for a component is the fugacity of the pure liquid. This standard state is not convenient, however, for dissolved gases, especially if the temperature is above the
critical temperature of the solute in question. For supercritical gases and also
for trace solutes such as organic pollutants in water, it is more convenient to
use a standard state defined at infinite dilution. This standard-state fugacity
is the Henrys constant.
PRO/II Note: For more information on using Henrys Law, see Section 24.11,
Henrys Law for Non-condensible Components, of the PRO/II Keyword Input
Manual.
Thermodynamically, the Henrys constant of a solute i in a solvent j is defined as
the infinite-dilution limit of the ratio of the fugacity to the mole fraction:
(31)

fi
x
1 i

Hi,j = lim
xj

xi 0

Unless the pressure is high or there is vapor-phase association, the fugacity fi can
be replaced by the partial pressure yiP. The K-value can then be expressed as:
Ki =

(32)

Hi,j
P

This relationship is strictly true only in the infinite-dilution limit, but K-values from Henrys law generally remain accurate at solute mole fractions up
to approximately five percent.
PRO/II correlates Henrys constants to the following functional form:
ln Hi,j = C1 +

C2
T

+ C3ln T + C4P

(33)

The temperature dependence in equation (33) is that expected from a thermodynamic analysis, provided the solvents critical point is not approached too
closely. Thermodynamics also predicts that the effect of pressure on the effective
Henrys constant at conditions beyond infinite dilution is linear in pressure (with
C4 proportional to the solutes partial molar volume). The pressure correction is
negligible at low and moderate pressures; if the pressure is sufficiently high for
that term to become important, it is likely that better results could be obtained by
an equation of state with an advanced mixing rule.
When the HENRY option is specified, components with critical temperatures below 400 K are automatically designated as solute components by PRO/II. The user
may, however, override these designations as desired. PRO/II has an extensive databank of Henrys constants for supercritical gases in various solvents and also for
many organic compounds in water. Henrys constants may also be input by the
user. If no Henrys constant is given for a solute, PRO/II substitutes the solutes vapor pressure (extrapolated if necessary). This substitution is good only
for nearly ideal solutions. In particular, if no Henrys constant is available for an
organic solute in water, it is better to remove the organic from the list of solutes
and allow the liquid activity method (with interaction parameters filled in via
UNIFAC, if necessary) to compute the K-value.

I-110

Liquid Activity Coefficient Methods

May 1994

Section 1.2

Thermodynamic Methods

Note: The temperature dependence of Henrys constants is very important. Especially for organic solutes in water, often only a single value at 25 C is reported. Calculations using this input value at significantly different temperatures (for example,
steam stripping near 100 C) are likely to produce unrealistic answers (for example,
drastically overestimating the amount of steam required). In such cases, the user can
obtain a better answer by assuming that the temperature dependence of the solutes
Henrys constant is the same as for its vapor pressure. The slope of ln Psat versus 1/T
becomes C2 in equation (33), and (with C3=C4=0) the 25 C point can then be used
to solve for C1.
The Henrys constant of a solute in a mixture of solvents is estimated from
the following mixing rule:
(34)

ln Hi,mix = xjln Hi,j


j=1

where the sum is taken over all solute species j, and the mole fractions xj
used in the sum are computed on a solute-free basis.
Reference
Prausnitz, J.M., Lichtenthaler, R.N., and Gomes de Azevedo, E., 1986,
Molecular Thermodynamics of Fluid-Phase Equilibria, 2nd edition, Prentice-Hall, Englewood Cliffs, NJ, Chapter 8.

Heat of Mixing
Calculations

For many liquid mixtures, the enthalpy may be accurately approximated as a


mole fraction sum of pure-component enthalpies (see Section 1.2.3, Ideal). For
some systems, however, the excess enthalpy, or heat of mixing, is not negligible
and should be accounted for if accurate prediction of the liquid enthalpy is important. It should be noted that SimScis equations of state and generalized correlations produce a heat of mixing as a natural part of their calculations. Therefore,
explicit calculation of the heat of mixing is only used in conjunction with the
IDEAL method for liquid enthalpy, which is normally used with liquid activity
coefficient K-value methods.
Gamma Method
Thermodynamics allows the excess enthalpy to be computed directly from
the activity coefficients in a mixture and their temperature dependence. This
is known as the GAMMA option, and the equation is:
(35)

ex
2 lni
H = xi RT

where:
Hex = excess heat of mixing

PRO/II Component and Thermophysical Properties


Reference Manual

Liquid Activity Coefficient Methods

I-111

Thermodynamic Methods

Section 1.2

Despite the attractiveness of this direct thermodynamic computation, experience has shown that the activity-coefficient parameters which correlate phase
equilibria do not in general produce very accurate values for excess enthalpies. GAMMA is a viable option when no other method is available, but the
resulting heats of mixing may only be accurate to within a factor of two.
Redlich-Kister Expansion
Experimental data for heats of mixing for binary systems are most often represented by an expansion about an equimolar mixture:
ex

(36)

H = a12 + b12X + c12X + d12X + e12X + f12X + g12X + h12X

where:
X = x1 x2
In equation (36), known as the Redlich-Kister expansion, a12 represents the
excess enthalpy of a 50-50 binary mixture. Higher-order terms correlate
asymmetry in the curve of excess enthalpy versus composition.
SimScis databanks contain regressed values of the coefficients in equation
(36) for approximately 2200 binary mixtures. In addition, these parameters
may be regressed to heat-of-mixing data with SimScis REGRESS program
and then entered through input by the user.

24.14

PRO/II Note: For more information on using the heat-of-mixing corrections to


liquid-enthalpy calculations, see Section 24.14, Redlich-Kister, Gamma Heat
of Mixing, of the PRO/II Keyword Input Manual.
The empirical nature of the Redlich-Kister expansion means that there is some
degree of arbitrariness in the way it is extended to mixtures. SimSci offers two
options, known as RK1 and RK2. Both have the same basic form:
(37)

H = xixj (aij + bijX + cijX + dijX + eijX + fijX + gijX + hijX )


ex

where:
X = xi xj
xi xj
X =
xi + xj

(form RK1)
(form RK2)

Note: Which mixture rule is better for a multicomponent system (they are
equivalent for binaries) depends upon the system, and there are no general
guidelines. RK2 is, however, somewhat preferable from the standpoint of
theoretical consistency.

I-112

Liquid Activity Coefficient Methods

May 1994

Section 1.2

1.2.7

Thermodynamic Methods

Vapor Phase Fugacities


General
Information

In vapor-liquid equilibrium calculations, it becomes necessary to calculate


separately the fugacity of each component in the vapor and liquid phases. Each
of the two phases may require a different technique. For example, liquid-phase
nonidealities may be described by activity coefficients, while deviations from
ideal-gas behavior in the vapor phase are described by fugacity coefficients. The
vapor-phase fugacity coefficients may be obtained through the use of an equation of state. The fugacity coefficients are obtained from classical thermodynamics as follows:
P
ln i = 1 / RT
V ni

T,V,nj

RT / V dV lnz

i = fi / yiP

(1)

(2)

where:
i =

fugacity coefficient of component i

fi =

fugacity of component i

R=

gas constant

T=

system temperature

P=

system pressure

ni =

number of moles of i

yi =

mole fraction of i in the vapor phase

V=

volume of system

z=

compressibility factor of the mixture

In equation (1), the partial derivatives of P with respect to ni must be evaluated


using an appropriate equation of state. Therefore the problem of calculating
fugacities of components in a gaseous mixture is equivalent to the problem of establishing a reliable equation of state for such a mixture. Once such an equation
of state is found, the fugacities can be derived from equations (1) and (2).

PRO/II Reference Manual


Volume I: Component and Thermophysical Properties

Vapor Phase Fugacities I-113

Thermodynamic Methods

Equations of State

Section 1.2

Equations of state are powerful methods for calculating vapor-phase fugacities at


low and high densities. The analytical expression of the fugacity coefficient can be
derived from a cubic equation of state. The derivation of the fugacity coefficient
from a cubic equation of state is straightforward because the cubic equation of state
in pressure is volume-explicit. Cubic equations of state are usually applied to systems comprising mixtures of nonpolar or weakly polar components. The usefulness of a cubic equation of state can be greatly enhanced by using an advanced
alpha function and an advanced mixing rule. These modified cubic equations of
state can be suitable for systems containing polar components.

Note: See Section 1.2.4 of this manual, Equations of State, for more information on using equations of state.

In addition, a cubic equation of state, when incorporated with a chemical


theory of association, is suitable for systems containing polar and hydrogenbonding molecules. These include carboxylic acids which form monomerdimer pairs and hydrogen fluoride. Such methods include the Associating
Hexamer Equation of State and the Hayden OConnell method, discussed in
a later section.

23.7

PRO/II Note: For more information on using the SimSci-developed Associating


Hexamer Equation of State method for vapor fugacities, see Section 23.7, Associating Hexamer Equation of State, of the PRO/II Keyword Input Manual.
The equation-of-state methods are generally more reliable for calculating vapor-phase fugacity coefficients than any other method, except for dimerizing
components, where the Hayden-OConnell method should be used.

Truncated Virial
Equation of State

Many equations of state have been proposed for calculating vapor fugacities, as
mentioned in the previous section, but almost all of them are empirical in nature.
The virial equation of state for gases has a sound theoretical foundation and is
free of arbitrary assumptions. The virial equation gives the compressibility factor as a power series in the reciprocal molar volume:
2

(3)

z = Pv / RT = 1 + B / v + C / v + D / v +

where:
v=

molar volume

B, C, D, .. =

second, third, fourth, etc. virial coefficients

The virial coefficients are a function of temperature and composition only.


For low or moderate vapor densities, the virial equation can be truncated
after the second virial coefficient and converted to a pressure-explicit form:
z = 1 + BP / RT

I-114

Vapor Phase Fugacities

(4)

May 1994

Section 1.2

Thermodynamic Methods

The compositional dependence of B for a mixture containing N components


is given by:
N

B=
i

(5)

yiyjBij
j

where:
Bii =

second virial coefficient for pure component i

Bij =

second virial cross coefficient

The cross coefficients characterize the interaction between one molecule of


component i and one of component j. They may be obtained from mixture data,
though often they are estimated from the pure component coefficients.
OConnell and Prausnitz developed a correlation for the reduced second
virial coefficient (both pure component and cross coefficients) which consists of three generalized functions:
1.

One for nonpolar contributions to the second virial coefficient.

2.

One for polar interactions based on the dipole moment.

3.

An association function for substances which exhibit specific forces such


as hydrogen bonds.

Use of this correlation requires the critical temperature, critical pressure,


critical volume, acentric factor, dipole moment, and association constant for
each component present. Missing dipole moments and association constants
are assumed to be zero. One limitation of this method is that, as the virial
equation of state is an expansion about the compressibility factor of an ideal
gas, higher-order terms cannot be neglected in high density regions. The
virial equation of state can provide reliable estimates of vapor-phase fugacity
coefficients at low pressures or high temperatures only.

24.13

PRO/II Note: For more information on using the Truncated Virial fugacity
method, see Section 24.13, Truncated Virial Vapor Fugacity, of the PRO/II
Keyword Input Manual.

Reference
OConnell, J. P., and Prausnitz, J. M., 1967, Empirical Correlation of Second
Virial Coefficients for Vapor-Liquid Equilibrium Calculations, Ind. Eng.
Chem. Proc. Des. Dev., 6, 245-250.

PRO/II Reference Manual


Volume I: Component and Thermophysical Properties

Vapor Phase Fugacities I-115

Thermodynamic Methods

HaydenOConnell

Section 1.2

The truncated virial equation of state described above is useful for predicting
deviations from ideality in those systems where moderate attractive forces yield
fugacity coefficients not far removed from unity. However, in systems containing
carboxylic acids, two acid molecules tend to form a dimer, resulting in large negative deviations from vapor ideality even at very low pressures.
To account for dimerization, Hayden and OConnell, in 1975, developed an
expression for the fugacity coefficient based on the chemical theory of vapor
imperfection. The chemical theory assumes that there are dimerization
equilibria for a binary mixture of components A and B:
KA2
A1 + A1 A2
KB2
B1 + B1 B2
KA B
A1 + B1 AB

(6)

(7)

(8)

where:
A1, B1 = monomers
A2, B2 = dimers
AB =

cross dimer

Hayden and OConnell related second virial coefficients to the dimerization


equilibrium constants, KA2, KB2, and KAB, and developed generalized second
virial coefficients for simple and complex systems. Properties required to use
this correlation are: the critical temperature, critical pressure, mean radius of
gyration, dipole moment, association parameter, and solvation parameter. Association and solution parameters for common associating components are
available in PRO/IIs databanks.

24.12

PRO/II Note: For more information on using the Hayden-OConnell fugacity


method, see Section 24.12, Hayden-OConnell Vapor Fugacity, of the PRO/II
Keyword Input Manual.
This method is a reliable generalized method for calculating vapor fugacities
up to moderate pressures, especially for systems where no data are available.
Reference
Hayden, J. G., and OConnell, J. P., 1975, A Generalized Method for
Predicting Second Virial Coefficients, Ind. Eng. Chem. Proc. Des. Dev.,
14, 209-216.

I-116

Vapor Phase Fugacities

May 1994

Section 1.2

1.2.8

Thermodynamic Methods

Special Packages
General
Information

PRO/II contains a number of thermodynamic methods specifically developed


for special industrial applications. Data packages are available for the following applications:
Alcohol dehydration systems.
Glycol systems.
Sour water systems.
Amine systems.
For many applications, databanks containing binary interactions specifically
regressed for components commonly found in the application have been developed
and incorporated into PRO/II. For example, for alcohol dehydration systems, a
special alcohol databank, in combination with the NRTL K-value method, is used
to calculate the K-values. For other applications, such as the SOUR or GPSWAT
method for sour systems containing NH3, H2S, CO2, and H2O, a K-value method
has been specifically developed for phase equilibrium calculations.

Alcohol
Package
(ALCOHOL)

The alcohol data package uses the NRTL liquid activity method to calculate
phase equilibria (see Section 1.2.6 of this manual, Liquid Activity Methods).
This system uses a special set of NRTL binary interaction parameters for
systems containing alcohols, water, and other polar components. The binary
parameters have been obtained by the regression of experimental data for alcohol systems. The recommended temperature and pressure ranges for the
ALCOHOL data package are as follows:
Temperature:
122-230 F for H2O-alcohol systems.
150-230 F for all other systems.
Pressure: up to 1500 psia.
The vapor enthalpy and density and the vapor and liquid entropies are
calculated using the SRKM equation of state (see Section 1.2.4 of this manual, Equations of State), while the liquid enthalpy and density are calculated
using ideal methods (see Section 1.2.3, Generalized Correlation Methods).
Table 1.2.8-1 shows the components present in the ALCOHOL databank for
which there are binary interaction parameters available.

PRO/II Reference Manual


Volume I: Component and Thermophysical Properties

Special Packages I-117

Thermodynamic Methods

Section 1.2
Table 1.2.8-1: Components Available for ALCOHOL Package
Component

Formula

LIBID

Miscellaneous
Acetaldehyde
Sulfolane

C2H4O
C4H8O2S

ACH
SULFLN

H2
N2
O2
CO2

H2
N2
O2
CO2

C5H12
C5H12
C5H10
C6H14
C6H12
C6H14
C6H12
C6H6
C6H12
C7H16
C7H16

IC5
NC5
CP
2MP
1HEXENE
NC6
MCP
C6H6
CH
24DMP
3MHX

C7H14
C7H16
C7H14
C7H8
C8H18

1T2MCP
NC7
MCH
TOLU
24DMHX

C8H16

1T2C4MCP

Light Gases
Hydrogen
Nitrogen
Oxygen
Carbon Dioxide
Hydrocarbons
Isopentane
n-Pentane
Cyclopentane
2-Methylpentane
1-Hexene
n-Hexane
Methylcyclopentane
Benzene
Cyclohexane
2,4-Dimethylpentane
3-Methylhexane
1-Trans-2-Dimethyl
cyclopentane
n-Heptane
Methylcyclohexane
Toluene
2,4-Dimethylhexane
1-Trans-2-Cis-4-Trimethylcyclopentane

Figure 1.2.8-1 shows the binary interaction parameters present in the ALCOHOL databank. It should be noted that for all pairs not denoted by x, the
missing binary interaction parameter can be filled in, based on the hierachy
discussed in Section 1.2.7, Liquid Activity Methods.

25.1

I-118

Special Packages

PRO/II Note: For more information on using the ALCOHOL method, see Section 25.1, Alcohol, of the PRO/II Keyword Input Manual.

May 1994

Section 1.2

Thermodynamic Methods

Figure 1.2.8-1:
Binary Interaction Data
in the Alcohol Databank

X
X X
X X X
X X X X
X X X X X
X X X X X X
X X X X X X X
X X X X X X X X
X X X X X X X X X
X X X X X X X X X X
X X X X X X X X X X X
X X X X X X X X X X X X
X X X X X X X X X X X X X
X X X X X X X X X X X X X X
X X X X X X X X X X X X X X X
X X X X X X X X X X X X X X X X
X X X X X X X X X X X X X X X X X
X X X X X X X X X X X X X X X X X X
X X X X X X X X X X X X X X X X X X X
X X X X X X X X X X X X X X X X X X X X
X X X X
X X X X
X X X X
X X X X

X
X

X X X X
X X X X

X
X

X
X

X X X X
X X X X

X
X

X
X

X
X

X
X X

X X
X X

X X X X
X X X X

X X X X X X X X X X X X X X X X X X X X X
X X X X
X X X X
X X X X

X X X X
X X X X

X X X X
X X X X

X
X X
X X X

X X X X X

X X X X
X X X X X

X X
X X

X X X X X X X X X X X X X X X X X X X X X X X
X
X X X X X X X X X X X X
X
X
X X X X X X

X X X

X X X X X
X X X X X

X X X X X X X X X X X X X X X X X X X X X X X X X X X X X

X X X X X X
X X X X X X X
X X X X X X X X
X X X X X X X X X

X X X X X X X X X X X X

Hydrogen
Nitrogen
Oxygen
Carbon Dioxide
Isopentane
n-Pentane
2-Methylpentane
n-Hexane
2,4-Dimethylpentane
3-Methylhexane
n-Heptane
2,4-Dimethylhexane
Cyclopentane
Methylcyclopentane
Cyclohexane
1-T-2DM-Cyclopentane
Methylcyclohexane
1-T,2-C-4 TMCP
1-Hexene
Benzene
Toluene
DIPE
DEE
DME
Acetic Acid
Formic Acid
MEK
Acetone
Ethyl Acetate
Methyl Formate
Acetaldehyde
Sulfolane
n-Pentanol
3-M-1-Butanol
n-Butanol
Isobutanol
Sec. Butanol
Tert-Butanol
n-Propanol
Isopropanol
Ethanol
Methanol

Nitrogen
Oxygen
Carbon Dioxide
Isopentane
n-Pentane
2-Methylpentane
n-Hexane
2, 4-Dimethylpentane
3-Methylhexane
n-Heptane
2, 4-Dimethylhexane
Cyclopentane
Methylcyclopentane
Cyclohexane
1-T-2DM-Cyclopentane
Methylcyclohexane
1-T,2-C-4 TMCP
1-Hexene
Benzene
Toluene
DIPE
DEE
DME
Acetic Acid
Formic Acid
MEK
Acetone
Ethyl Acetate
Methyl Formate
Acetaldehyde
Sulfolane
n-Pentanol
3-M-1-Butanol
n-Butanol
Isobutanol
Sec-Butanol
Tert-Butanol
n-Propanol
Isopropanol
Ethanol
Methanol
Water

PRO/II Reference Manual


Volume I: Component and Thermophysical Properties

Special Packages I-119

Thermodynamic Methods

Glycol
Package
(GLYCOL)

Section 1.2

The glycol data package uses the SRKM equation of state to calculate phase
equilibria for glycol dehydration applications (see Section 1.2.4 of this manual, Equations of State). This system uses a special set of SRKM binary
interaction data and alpha parameters for systems containing glycols (particularly TEG), water, and other components. The binary parameters and alpha
parameters have been obtained by the regression of experimental data for
glycol systems. The recommended temperature and pressure ranges for the
GLYCOL package are:
Temperature: 80-400 F.
Pressure: up to 2000 psia.
Other thermodynamic properties such as the vapor and liquid enthalpy, entropy, and vapor density are calculated using the SRKM equation of state,
while the liquid density is calculated using the API method (see Section
1.2.3, Generalized Correlation Methods).
Table 1.2.8-2 shows the components present in the GLYCOL databank for
which there are binary interaction parameters available.

Table 1.2.8-2: Components Available for GLYCOL Package


Components
Hydrogen
Nitrogen
Oxygen
Carbon Dioxide
Hydrogen Sulfide
Methane
Ethane
Propane
Isobutane
n-Butane
Isopentane
Pentane
Hexane
Heptane
Cyclohexane
Methylcyclohexane
Ethylcyclohexane
Benzene
Toluene
o-Xylene
m-Xylene
p-Xylene
Ethylbenzene
Ethylene Glycol
Diethylene Glycol
Triethylene Glycol
Water

I-120

Special Packages

Formula
H2
N2
O2
CO2
H2S
CH4
C2H6
C3H8
C4H10
C4H10
C5H12
C5H12
C6H14
C7H16
C6H12
C7H14
C8H16
C6H6
C7H8
C8H10
C8H10
C8H10
C8H10
C2H6O2
C4H10O3
C6H14O4
H2O

LIBID
H2
N2
O2
CO2
H2S
C1
C2
C3
IC4
NC4
IC5
NC5
NC6
NC7
CH
MCH
ECH
BNZN
TOLU
OXYL
MXYL
PXYL
EBZN
EG
DEG
TEG
H2O

May 1994

Section 1.2

Thermodynamic Methods

Figure 1.2.8-2 shows the binary interaction parameters, denoted by x,


present in the glycol databank. Interaction parameters denoted by o are
supplied from the SRK databank. It should be noted that for all pairs not denoted by x or o, the missing binary interaction parameters are estimated
using a molecular weight correlation or are set equal to 0.0.

Nitrogen
Oxygen
Carbon Dioxide
Hydrogen Sulfide
Methane
Ethane
Propane
Isobutane
N-butane
Isopentane
Pentane
Hexane
Heptane
Cyclohexane
Methylcyclohexane
Ethylcyclohexane
Benzene
Toluene
o-xylene
m-xylene
p-xylene
Ethylbenzene
Ethylene Glycol
Diethylene Glycol
Triethylene Glycol
Water

X
X O
X O
X O
X O

O
O O

X O
X O

O O X
O O X X

X O
X O

O O O O O
O O O O O O

X O
O O O
O
X O
O O X X O
X O X O O X X O
X O
O O

X O O O O
O
O O O

X O
X

O
O

X X X
X O O
X
X

O O
O
O

O O
O O

O O O

X
X

X
X
X X X
X X

X
X X X X X X X X X X X X X X X X X X X X
X
X X X X X X X X X X X X X X X X X X X X X X X X X X

Hydrogen
Nitrogen
Oxygen
Carbon Dioxide
Hydrogen Sulfide
Methane
Ethane
Propane
Isobutane
N-butane
Isopentane
Pentane
Hexane
Heptane
Cyclohexane
Methylcyclohexane
Ethylcyclohexane
Benzene
Toluene
o-xylene
m-xylene
p-xylene
Ethylbenzene
Ethylene Glycol
Diethylene Glycol
Triethylene Glycol

Figure 1.2.8-2:
Binary Interaction Data
in the Glycol Databank

25.2

PRO/II Note: For more information on using the GLYCOL method, see Section 25.2, Glycol, of the PRO/II Keyword Input Manual.

PRO/II Reference Manual


Volume I: Component and Thermophysical Properties

Special Packages I-121

Thermodynamic Methods

Sour Package
(SOUR)

Section 1.2

This sour water package uses the SWEQ (Sour Water EQuilibrium) method
developed by Wilson for a joint API/EPA project. Phase equilibria for sour
water components NH3, H2S, CO2, and H2O are modeled using a modified
van Krevelen approach. The van Krevelen model assumes that H2S and CO2
exist in solution only as ionized species. This is true only for solutions containing an excess of NH3 or other basic gas. This limitation has been removed in the SWEQ method by considering the chemical equilibrium
between ionic species of H2S or CO2 and their undissociated molecules in
the liquid phase.
In the SWEQ model, the partial pressure in the vapor phase for H2S or CO2 is
given by:
pH2S = HH2S CH2S

(1)

pCO2 = HCO2 CCO2

(2)

where:
pi =

partial pressure of component i

Hi =

Henrys Law constant for component i in water

Ci =

concentration of component i in the liquid phase, gmoles/


kg solution

The SWEQ model uses Henrys Law constants for each component in solution
as a function of temperature and composition of the undissociated molecular species in the liquid phase. The Henrys constants for H2S and CO2 were obtained
from data published by Kent and Eisenberg, who developed a model for predicting H2S-CO2-MEA-H2O and H2S-CO2-DEA-H2O systems. The Henrys Law
constants used in the SWEQ model for equations (1) and (2) are:
4

(3)

ln(HCO2) = 18.33 2.48951 x 10 / T + 2.23996 x 10 / T


9

12

9.0918 x 10 / T + 1.2601 x 10 / T

(4)
5

11

ln( HH2S) = 100.684 2.46254 x 10 / T + 2.39029 x 10 / T 1.01898 x 10 / T +


13

1.5973 x 10 / T 5.0 x 10 CNH3 + (0.965 486 / T)CCO2

where:
T=

I-122

Special Packages

system temperature, degrees Rankine

May 1994

Section 1.2

Thermodynamic Methods

The Henrys Law constant for water was obtained by correlating H2O vapor
pressure data from the A.S.M.E. steam tables over the range 25 C to 150 C:
(5)

lnHH2O = 14.466 6.9966 x 10 / T 77.67

The Henrys Law constant for NH3 was taken from data published by
Edwards et al.:
4

ln( HNH3) = 178.339 1.55179 x 10 / T 25.6767ln(T) + 1.966 x 10 T +

(6)

( 131.4 / T 0.1682)CNH3 + 6.0 x 10 2CCO2 + CH2S

The chemical equilibria of all the main reactions in the liquid phase due to the dissociation of the sour gas molecules are considered in the model. The reaction equilibrium constants, Ki, are correlated as functions of temperature, composition of
undissociated sour gas molecules in the liquid phase, and ionic strength.
0

(7)

0.4

ln Ki = ln Ki + aCH2S + bCCO2 + cI

where:
Ki =

equilibrium constant of reaction i

Ki0 =

equilibrium constant at infinite dilution for all species

a,b,c =

constants

I=

ionic strength = 1 / 2 CjZ2j

Zj =

ionic charge of species j

The reaction equilibrium constants at infinite dilution, Ki0, are given in the
form first proposed by Kent and Eisenberg:
0

ln Ki = A + B / T + C / T + D / T + E / T

(8)

where:
A,B,C,D,E = constants
The constants used in the SWEQ model for equations (7) and (8), obtained by
the regression of experimental data, are given in the original EPA report. The
original SWEQ method was developed for pressures less than 50 psia where
nonidealities in the vapor phase are not important. Corrections for vapor-phase
nonidealities using SRKM have been incorporated into PRO/II, thus extending
the applicable pressure range to 1500 psia.

PRO/II Reference Manual


Volume I: Component and Thermophysical Properties

Special Packages I-123

Thermodynamic Methods

Section 1.2

The phase behavior of all other components present in the system is modeled
using the SRKM equation of state (see Section 1.2.4 of this manual, Equations of State). The following limits apply to the SOUR method as
implemented in PRO/II:
Temperature: 67-300 F.
Pressure: up to 1500 psia.
Composition: wNH3 + wCO2 + wH2S <0.30.
where:
wi =

weight fraction of component i

Note: NH3 and water must be present when using the SOUR method.
Other thermodynamic properties such as the vapor enthalpy, vapor and liquid
entropy, and vapor density are calculated using the SRKM equation of state,
while the liquid enthalpy and density are calculated using ideal methods (see
Section 1.2.3, Generalized Correlation Methods).

25.3

PRO/II Note: For more information on using the SOUR method, see Section
25.3, Sour, of the PRO/II Keyword Input Manual.

References

I-124

Special Packages

1.

Wilson, G. M., 1980, A New Correlation for NH3, CO2, H2S Volatility Data
from Aqueous Sour Water Systems, EPA Report EPA-600/2-80-067.

2.

van Krevelen, D. W., Hoftijzer, P. J., and Huntjens, F. J., 1949, Rec. Trav.
Chim., 68, 191-216.

3.

Black, C., 1958, Vapor Phase Imperfections in Vapor-Liquid Equilibria, Ind.


Eng. Chem., 50, 391-402.

4.

Kent, R. L., and Eisenberg, B., 1976, Better Data for Amine Treating,
Hydrocarbon Processing, Feb., 87-90.

5.

Handbook of Chemistry and Physics, 1971, 51st Edition, The Chemical


Rubber Co.

6.

Edwards, T. J., Newman, J., and Prausnitz, J. M., Thermodynamics of Aqueous


Solutions Containing Volatile Weak Electrolytes, 1975, AIChE J., 21, 248-259.

May 1994

Section 1.2

GPA Sour Water


Package
(GPSWATER)

Thermodynamic Methods

This sour water package uses the method developed for the Gas Processors
Association in 1990 for sour water systems containing components NH3, H2S,
CO2, CO, CS2, MeSH, EtSH, and H2O. This model uses the SWEQ model (see
above) as a precursor, extending the temperature range of applicability to 600
F. The total pressure limit is increased to 2000 psia by allowing for vapor
phase non-idealities and accounting for pressure effects in the liquid phase,
using a Poynting correction factor.
As in the SWEQ model, the chemical equilibria for all the reactions involving the NH3, H2S, and CO2 in water are considered. The components CO,
methyl mercaptan (MeSH), and ethyl mercaptan (EtSH) are treated as
Henrys Law components (see Section 1.2.6, Liquid Activity Coefficient
Methods). Reactions considered include:
Water:
1

(9)

H2O H + OH

Ammonia:
2

NH3 + H NH4

(10)

Hydrogen Sulfide:
3

(11)

H2S + OH H2O + HS

Bisulfide:
4

(12)

HS + OH H2O + S

Carbon Dioxide:
5

(13)

CO2 + OH HCO3

Bicarbonate:

(14)
2

HCO3 + OH H2O + CO3

PRO/II Reference Manual


Volume I: Component and Thermophysical Properties

Special Packages I-125

Thermodynamic Methods

Section 1.2

Carbon Dioxide and Ammonia:


7

(15)

NH3 + CO2 H2NCOOH


8

(16)

H2NCOOH + OH H2NCOO + H2O


9

(17)

H2NCOOH + NH3 H2NCONH2 + H2O

The chemical equilibrium constants, Ki, are correlated as functions of temperature and composition. In addition, the effect of inert gases such as N2 and H2 on
phase equilibria is also considered. In the liquid phase, pressure effects are accounted for by the use of a Poynting correction factor, and electrostatic effects
are incorporated into the calculated liquid activity coefficients.
Vapor-phase nonidealities are computed using a truncated virial equation of
state. The virial equation used is truncated after the third virial coefficient as
follows:
2

(18)

z=1+B/v+C/v

where:
B, C are the second and third virial coefficients
v=

molar volume

z=

compressibility factor

Phase equilibria for all other components present in the system are modeled
using the SRKM equation of state (see Section 1.2.4 of this manual, Equations of State).
The following limits apply to the GPSWATER method:
Temperature: 68-600 F.
Pressure: up to 2000 psia.
Composition:
wNH3 < 0.40.
pCO2 + pH2S < 1200 psia
where:
wi =

weight fraction of component i

pi =

partial pressure of component i in the vapor phase

Note: NH3, CO2, H2S, and water must be included in the component list
when using the GPSWATER method.

I-126

Special Packages

May 1994

Section 1.2

Thermodynamic Methods

Other thermodynamic properties such as the vapor enthalpy, vapor and liquid entropy, and vapor density are calculated using the SRKM equation of state, while
the liquid enthalpy and density are calculated using ideal methods (see Section
1.2.3, Generalized Correlation Methods).

25.4

PRO/II Note: For more information on using the GPSWATER method, see
Section 25.4, GPA Sour Water, of the PRO/II Keyword Input Manual.

Reference
Wilson, G. M., and Eng, W. W. Y., 1990, GPSWAT: GPA Sour Water
Equilibria Correlation and Computer Program, GPA Research Report RR118, Gas Processors Association.

Amine Package
(AMINE)

The PRO/II simulation program contains a method to model the removal of H2S
and CO2 from natural gas feeds using alkanolamines. Alkanolamines are formed
by ammonia reacting with an alcohol. Amines are considered to be either primary, secondary, or tertiary, depending on whether 1 or 2 or 3 of the hydrogen
atoms have been replaced on the ammonia molecule. PRO/II provides data for
the primary amine monoethanolamine (MEA), secondary amines diethanolamine (DEA), diglycolamine (DGA), and diisopropanolamine (DIPA), and the
tertiary amine methyldiethanolamine (MDEA). MEA and DEA are frequently
used amines in industry.
In aqueous solutions, H2S and CO2 react in an acid-base buffer mechanism with
alkanolamines. The acid-base equilibrium reactions are written as chemical dissociations following the approach taken by Kent and Eisenberg:
Water:
1

(19)

H2O H + OH

Hydrogen Sulfide:
2

(20)

H2S H +HS

Bisulfide:

HS H + S

PRO/II Reference Manual


Volume I: Component and Thermophysical Properties

(21)

Special Packages I-127

Thermodynamic Methods

Section 1.2

Carbon Dioxide:
4

(22)

H2O + CO2 H + HCO3

Bicarbonate:

(23)

HCO3 H +CO3

Alkanolamine:
6

(24)

RR NH2 + H + RR NH

where:
Ki =

equilibrium constant for reaction i

RRNR =

alkanolamine

R represents an alkyl group, alkanol, or hydrogen


In addition to the acid-base reactions above, CO2 also reacts directly with
primary and secondary alkanolamines to form a stable carbamate, which can
revert to form bicarbonate ions.
Carbamate Reversion to Bicarbonate:

(25)

RR NCOO + H2O RR NH + HCO3

Tertiary amines such as MDEA are not known to form stable carbamates. In
an aqueous solution with MDEA, CO2 forms bicarbonate ions by reaction
(22) only.

Note: CO2, H2O, and H2S must be included in the component list when using
the AMINE method.
The chemical equilibrium constants, Ki, are represented by the following
equation:
ln Ki = A + B T + C T + D T
2

I-128

Special Packages

(26)

May 1994

Section 1.2

Thermodynamic Methods

The equlibrium constant for the protonated amine dissociation reaction given
in reaction (24) is corrected to the pure amine reference state. This is done
by relating the constant to the infinite-dilution activity coefficient of the
amine in water estimated from experimental data for the system amine-water.
The liquid enthalpy is calculated using ideal methods and adding a correction for the heat of reaction as follows, using either a modified ClausiusClapeyron equation or fits of data from the Gas Processors Association:
Hr = lnKT1 / KT2 R / 1 / T1 1 / T2

(27)

where:
Hr = heat of reaction
R=

gas constant

KT1, KT2 =

K-values at temperatures T1 and T2

The vapor-phase enthalpy and density, and liquid- and vapor-phase entropy
are calculated using the SRKM equation of state (see Section 1.2.4 of this
manual, Equations of State). Ideal methods are used to calculate the liquidphase density (see Section 1.2.3, Generalized Correlation Methods).
For MEA and DEA systems, data have been regressed from a large number of
sources, resulting in good prediction of phase equilibria for these systems. For
systems containing DIPA, a limited amount of experimental data were available,
and so the DIPA results are not recommended for final design purposes. For
MDEA and DGA systems, the user is allowed to input a residence time correction to allow the simulation results to more closely match plant data. The application ranges suggested for amine systems are shown in Table 1.2.8-3.

Table 1.2.8-3: Application Guidelines for Amine Systems

Amine

25.5

MEA

DEA

DGA

MDEA

DIPA

Pressure, psig

25-500

1001000

1001000

1001000

1001000

Temperature, F

<275

<275

<275

<275

<275

Concentration, wt %
amine

~15-25

~25-35

~55-65

~50

~30

Acid gas loading,


gmole gas/gmole
amine

0.5-0.6

0.45

0.50

0.4

0.4

PRO/II Note: For more information on using the AMINE method, see Section
25.5, Amine, of the PRO/II Keyword Input Manual.

PRO/II Reference Manual


Volume I: Component and Thermophysical Properties

Special Packages I-129

Thermodynamic Methods

Section 1.2

References

I-130

Special Packages

1.

Kent, R. L., and Eisenberg, B., 1976, Better Data for Amine Treating,
Hydrocarbon Processing, Feb., 87-90.

2.

Maddox, R. N., Vaz, R. N., and Mains, G. J., 1981, Ethanolamine Process
Simulated by Rigorous Calculation, Hydrocarbon Processing, 60, 139-142.

May 1994

Section 1.2

1.2.9

Thermodynamic Methods

Electrolyte Mathematical Model

Discussion of
Equations

The mathematical model employed in PRO/II Electrolytes is a deterministic


set of nonlinear algebraic equations. The equation set is composed of:
Equilibrium Expressions
For each vapor-liquid, liquid-liquid, solid-liquid, and liquid intraphase equilibrium, there is an equation of the form:

K=

iP

iR

( )niP (m )niP
iP
iP

(1)

( )niR (m )niR
iR
iR

where:
K=

thermodynamic equilibrium constant: a function of temperature and pressure

iP, iR =

activity coefficient or, for vapors, fugacity coefficient


of the ith product and reactant, respectively; a function
of temperature, pressure, and composition

niP, niR =

stoichiometric coefficient of the ith product and reactant, respectively

miP, miR =

molality or, for vapors, partial pressure of the ith product and reactant, respectively

Note: When H2O (the solvent) appears in the equilibrium expression, its
contribution is expressed as aH2O, the activity of water. All pure solid
component activities are assumed to be one. The fugacity coefficient, i is defined as fi/yiP or fi/Pi, where f denotes fugacity, P represents total pressure, and
Pi stands for partial pressure. The adopted convention is that i approaches one
as total pressure approaches zero. The activity ai is given by imi, where i approaches one as the molality of all solutes approaches zero.

PRO/II Reference Manual


Volume I: Component and Thermophysical Properties

Electrolyte Mathematical Model

I-131

Thermodynamic Methods

Section 1.2

Note: Equilibrium constants often are written entirely in terms of fugacities, in


which case thermodynamics requires that K is a function of temperature only.
When activities are used intead of fugacities, as in the present treatment, then
K is affected by the choice of standard state. Since the standard state for solutes
is infinite dilution in H2O, a pressure dependence is introduced from the pressure dependence of the water properties. This effect is negligible except at conditions approaching waters critical point. Consequently, over the temperature
and pressure validity range for the electrolyte thermodynamic methods (0-200
C; 0-200 atm), K is treated only as a function of temperature.
An Electroneutrality Equation
NC

(2)

NA

| Zi | mi = | Zi | mi
i=1

i=1

where:
Zi =

species charge

NC, NA =

number of cations and anions, respectively

Equations For Solutions Involving a Second Liquid Phase,


(liquid-liquid equilibrium)
A

(3)

ai = ai

i = 1,NM

ai =

activity of ith species

A, O =

represent aqueous and organic phases, respectively

NM =

number of molecular species distributing between phases

where:

Note: The required number of material balances, i.e. NB, completes the model
and assures that the number of equations and the number of unknowns are
equal. Normally these balances include an overall, a vapor phase, an organic
phase, and several component balances.
Thus, assuming NK equilibrium equations, the model has NK + NB + NM + 1
equations. The customary unknowns are:
The moles of H2O in the aqueous liquid phase plus all ionic and molecular species molalities.
The vapor phase: species mole fractions plus overall vapor fraction.
The organic liquid phase: species mole fractions plus overall organic
phase fraction (if second liquid phase is present).
The solid phase composition: moles precipitated for all solid species.

I-132

Electrolyte Mathematical Model

May 1994

Section 1.2

Thermodynamic Methods

As noted above, the number of NB equations required is that number which


assures that the number of equations equals the overall number of unknowns.
This is a natural consequence of the phase rule.

Modeling
Example

To better understand this modeling concept, consider the aqueous system represented by H2O-CO2-NACL. The reactions considered will be:
H2Ovap H2O

(4)

CO2vap CO2aq

(5)

+1

(6)

H2Oaq H + OH

+1

CO2aq + H2Oaq H + HCO3


+1

HCO3 H + CO3
+1

NACLppt NA

(7)

(8)

(9)

+ CL

Based upon the general model described earlier, this leads to:
Equilibrium Expressions
KH2Ovap =

KCO2vap =

KH2Oaq =

KCO2aq =

(10)

aH2O

H2Ovap yH2Ovap P
CO2aq mCO2aq

(11)

CO2vap yCO2vap P

H+1 mH+1 OH1 mOH1

(12)

aH2O
H+1 mH+1 HCO31 mHCO31

KHCO3ion =

KNACLppt=

(13)

CO2aq mCO2aq aH2O


H+1 mH+1 CO32 mCO32

(14)

HCO31 mHCO31
NA+1 mNA+1 CL1 mCL2

(15)

aNACLppt

Electroneutrality Equation
mH+1 + mNA+1 = mHCO31 + 2mCO32 + mOH1 + mCL1

(16)

Liquid-liquid Equilibrium Equations


There are none in this example.

PRO/II Reference Manual


Volume I: Component and Thermophysical Properties

Electrolyte Mathematical Model

I-133

Thermodynamic Methods

Section 1.2

Material Balance Equations


Overall material balance:
3H2Oin + 3CO2in + 2NACLin = 2NACLppt + 3H2O + (H2O / 55.51)
(3mCO2aq + 5mHCO31 + 4mCO32 + mH+1 + 2mOH1 +
mCL1 + mNA+1) + V(3yH2Ovap + 3yCO2vap)

(17)

where:
in = inflow or feed component in units of moles
V = overall vapor fraction on a mole basis
In equation (17), the products H2O and NACLppt are in units of moles.
Vapor balance
yH2Ovap + yCO2vap = 1

(18)

where:
y = mole fraction for vapor species
Sodium balance
NACLin = (H2O / 55.51)(mNA+1) + NACLppt

(19)

Chlorine balance
NACLin = (H2O / 55.51)(mCL1) + NACLppt

(20)

Carbon balance
CO2in = (H2O / 55.51)(mCO32 + mHCO31 + mCO2aq) + V (yCO2vap)

(21)

Equations (10) -- (21) are the required 12 equations. Assuming that the temperature and pressure are known, and further assuming that deterministic formulations are available for the equilibrium constants, activity coefficients,
fugacity coefficients, and the activity of water, then the corresponding 12 unknowns (calculated variables) are:
V, yH2Ovap, yCO2vap, mH+1 mOH1, mCO32,
mHCO31, mNA+1, mCL1, mCO2aq, moles NACLppt,
and moles water.

164

I-134

PRO/II Note: A summary of the pregenerated electrolyte models available in


PRO/II is given in Section 164, Electrolyte Models, of the PRO/II Keyword Input Manual. Additional models may be generated by the user with the PRO/II
Electrolyte Utility Package (EUP).

Electrolyte Mathematical Model

May 1994

Section 1.2

1.2.10

Thermodynamic Methods

Electrolyte Thermodynamic Equations

Thermodynamic
Framework

The mathematical model described in the previous section utilizes several


thermodynamic quantities. Specifically these are:
Equilibrium constants -- normally strong functions of temperature and
weaker functions of pressure.
Aqueous-phase activity coefficients -- normally strong functions of temperature and composition and weaker functions of pressure.
Vapor-phase fugacity coefficients -- normally significant functions of temperature, pressure and composition, particularly at elevated pressures.
Organic liquid-phase activities -- normally strong functions of temperature and composition and weaker functions of pressure.
The formulations used by PRO/II Electrolytes for each of these quantities are
described below.
In addition, the thermodynamic framework includes formulations for the calculation of enthalpies and densities for aqueous liquid, organic liquid, vapor,
and solid phases. These latter formulations are also presented below.

163

Equilibrium
Constants

PRO/II Note: For information on using the electrolyte thermodynamic methods


in PRO/II, see Section 163, Electrolyte Thermodynamic Data, of the PRO/II
Keyword Input Manual.
By considering basic thermodynamic relationships and assuming a constant heat
capacity of reaction, the following general equation can be derived:
ln K (T) =

G H 1 1 Cp Tr Tr

ln + 1
R T Tr
RTr
R T T

(1)

where:
T=

temperature in Kelvins

Tr =

reference temperature, 298.15 K

G = free energy of the reaction in the standard state at the reference temperature (and pressure, 1 bar). This is derived
from the standard-state Gibbs free energies of formation,
Gf , by first, summing the product of the reaction coefficient times Gf over all reactants and then, over all products. Next, the sum for the reactants is subtracted from the
sum for the products to obtain G .

PRO/II Reference Manual


Volume I: Component and Thermophysical Properties

Electrolyte Thermodynamic Equations

I-135

Thermodynamic Methods

Section 1.2

H = corresponding standard state heat of reaction at the reference


conditions. Obtained from the standard-state enthalpy of formation, Hf , using the same general procedure as used for
G .
Cp= corresponding standard state heat capacity of reaction at the
reference conditions. Obtained from the standard state heat capacity, Cp, using the same general procedure as used in G.
R =

gas constant

The derivation of the relationship given in equation (1) can be found in the
Handbook of Aqueous Electrolyte Thermodynamics1. Values of Gf , Hf ,
and Cp for reaction species are usually available in the critically evaluated
data compilations of the National Bureau of Standards2 or the Russian Academy of Sciences3.
The chosen standard states for the thermodynamic framework are as follows:
Aqueous solutes - hypothetical, infinitely dilute solution at unit molality;
Solvent - pure fluid; Gaseous species - hypothetical 1 bar ideal gas; Solid
species - pure solid. These are the same standard states as used by the NBS2.
For some reactions, where sufficient measurements are available for K as a
function of temperature, fitted functions of temperature are used instead of
equation (1).

Aqueous Phase
Activities

The key to successful simulation of aqueous systems is to accurately predict the


reaction equilibria described in the previous section. Greater precision is added
by a good description of the following correction factors:
Activity coefficients of ions in solution.
Activity coefficients of molecules in solution.
Activity of water.
In PRO/II Electrolytes, these quantities can be represented in terms of a number of alternative as well as complementary formulations. The common element of all of these formulations is that they involve the interaction of pairs
of species in solution. Two general assumptions are made:
Interactions between like-charged ions are not significant.
Higher-level interactions (involving more than two species) are not
significant.

I-136

Electrolyte Thermodynamic Equations

May1994

Section 1.2

Thermodynamic Methods

Ions
For ions the formulation used is:
ln i = DHi + BZi + Pi

(2)

DHi = Debye-Hckel term for long-range, ion-ion interactions,


defined as:
(3)

2 1 2
A Zi I /

DHi =

1 2
1+I /

where:
A=

Debye-Hckel constant, a known function of temperature


and solvent density.1
NI

I=

ionic strength = 1 / 2

Z2i mi; where m = molality

i=1

Zi =

charge on ion i

The Debye-Hckel term predicts the long-range or electrostatic effects. For


dilute solutions of ionic strength less than 0.1, this is the only term needed.
BZi =

Bromley-Zemaitis4,5 term for short-range, ion-ion


interactions, defined as:
(4)

NO

Zi + Zj
BZi =
ij mj
2

j=1

where:
ij =

(0.06 + 0.6 Bij) Zi Zj

NO =

(1 + 1.5 / Zi Zj I)

+ Bij + Cij I + Dij I

(5)

number of ions with charge opposite to that of the ion being represented.

Bij,Cij,Dij =

Pi =

three interaction coefficients for each cation-anion interaction. These are each made 3-parameter functions of temperature. Thus, for each cation-anion interaction, there
are 9 coefficients that must be established.

Pitzer6,7 term for short-range, ion-molecular interactions,


defined as:

NM
2

Z1
Pi = BPij + 2 BPSjmj
4I

j=1

(6)

(0)

(1)
1 2
1 2
BPij = ij + ij 1 + 2I / 1 exp 2I / / 2I

(7)

where:

and

PRO/II Reference Manual


Volume I: Component and Thermophysical Properties

Electrolyte Thermodynamic Equations

I-137

Thermodynamic Methods

Section 1.2
(8)

NS

BPSj = 0.86859 BPPjk mk


k=1

and
(1)
1 2

1 2
BPPjk = jk 1 1 + 2I / + 2I exp 2I /

(9)

where:
NM =

number of molecular species in solution

NS =

number of species in solution

(1)
(0)
ij , ij

two interaction coefficients for each ion-molecule or


molecule-molecule interaction.

Note: Each of the ij interaction coefficients is a 3-parameter function of


temperature. Thus, for each interaction, there are 6 coefficients that must be
established.
Molecules Other Than Water
For molecules other than water, the Setschenow equation is used. Where
ij(0) and ij(1) parameters are available, the preferred formulation is from
Pitzer6,7. The Setschenow and Pitzer relations are:
NI

ln i = bi I + bj mj

(10)

j=1

where:
bi =

Setschenow coefficient for the neutral species

bj =

Salting-out coefficient particular to each ion

NI =

the number of ionic species in solution

and
NS

ln i = 2 BPij mi

(11)

j=1

BPij is defined by equation (7) above.


Water Activity
The water activity for multicomponent systems is obtained from an integrated form of the Gibbs-Duhem equation, together with a mixing rule suggested by Meissner and Kusik8. The formulation can be represented as:

I-138

Electrolyte Thermodynamic Equations

May1994

Section 1.2

Thermodynamic Methods
(12)

NC NA

ln a

H2O

1
AZ

Zi Zj ij mi mj

i=1 j=1
NM

0.01801
i=1

NS

m + 2

BPW
m
m

i
ij
i
j

j=1

where:
NC

2
AZ = Zi mi
i=1

NA
Z2 m

i
i

i=1

(13)

(0)
(1)
1 2

BPWij = ij + ij exp 2I /

(14)

NC =

the number of cation species in solution

NA =

the number of anion species in solution

The above formulations are, in cases where the necessary interaction coefficients
have been fit to cover the conditions being simulated, quite adequate for predicting systems in which water is the principal solvent.

Vapor Phase
Fugacities

Four alternative methods are provided:


Ideal, all fugacity coefficients are assumed to be 1.0.
Nothnagel method, generally valid up to 20 atmospheres.
Nakamura method, generally valid up to 200 atmospheres.
Soave-Redlich-Kwong (SRK) method, valid over a wide range of conditions
and generally recommended when vapor-phase non-ideality is important.
Nothnagel Method
Nothnagel et al.9 developed a method for calculating fugacity coefficients in
mixtures at moderate pressures. The main feature of the method is the inclusion of dimerization effects on the second virial coefficient.
The equation of state is written as:
P=

(15)

nT R T
V nT bm

where:
bm =

size parameter for mixture, cm3 / mole

nT =

number of moles of true species

P=

pressure, atmospheres

R=

gas constant, 82.056 cm3 atm / mole K

T=

temperature, Kelvins

V=

total volume, cm3

PRO/II Reference Manual


Volume I: Component and Thermophysical Properties

Electrolyte Thermodynamic Equations

I-139

Thermodynamic Methods

Section 1.2

nT is obtained from the solution of the dimerization equilibria described below. bm is given by a sum over all true species:
n bm = ni bi

(16)

where:
bi =

size parameter for true species i, cm3 / mole

ni =

number of moles of true species i

If i is a dimer formed by monomers A and B, bi is given by:


(17)

1 1 3
1 3
bi = b / + b /
B
8 A

Values of b for monomers are tabulated in the original reference.


For component j, the fugacity coefficient is given by:
zj
j = exp bj P / RT
y

(18)

where:
bj =

size parameter for monomer j, cm3 / mole

yj =

apparent mole fraction of component j

zj =

true mole fraction of component j monomer

The true mole fractions zj are computed from the dimerization equilibria.
Each dimerization is described by an equilibrium constant Kij:
Kij =

zij ij
P
zi zj i j

(19)

1 3 1 3

ij = exp bi / + bj / P / 8RT

The dimerization equilibrium constants are related to the enthalpy and entropy of dimerization:
lnRTKij = Hij / RT Sij / R

(20)

(21)

where:
Hij = enthalpy of dimerization, cal/mole
Sij = entropy of dimerization, cal/mole
R=

gas constant, 82.056 cm3 atm/mole K (left side of equation


(21)), 1.987 cal/mole K (right side of equation (21))

The correlations for Hij and Sij, along with the necessary parameters for
178 components, may be found in the original paper9.

I-140

Electrolyte Thermodynamic Equations

May1994

Section 1.2

Thermodynamic Methods

Nakamura Method
Nakamura et al.10 proposed the following perturbed-hard-sphere equation of
state for gas mixtures:
P=

2
3
RT 1 + +
a

3
v
v
(v
+ c)
(1

(22)

where:
P=

pressure, atmospheres

R=

gas constant, .082056 liter atm/mole K

T=

temperature, Kelvins

v=

molar volume, liter / mole

reduced density, b/4v

b=

parameter signifying the hard-core size of the molecule, liter / mole

a=

parameter signifying the attractive force strength, atm/mole

c=

constant, liter / mole

In addition to equations for the reduced enthalpy difference and entropy difference, the following equation was presented for calculating the fugacity coefficient of a species k in the gas mixture:
n
5 ( 1) m c
4 32 b k 4 2 2
M
2
ln k =
+
yj akj

2 b
3
RT
v
(
m
+
1)

M (1 )
j=1
m=1
(1 )

4
a ck (1)m (m + 1) c
M
M
RTv
v
m=1 (m + 2)

1
+ ln z
2

+ 1 +

(23)

With the P, R, T, v, and terms defined earlier, the following definitions and
calculations apply:
n=

number of species

bk = exp [2.30259 ( T)]

(24)

where:
=

pure-component parameter

pure-component parameter
(25)

b = yi bi
M

i=1

where:
yi =

vapor mole fraction of i

ck =

pure-component parameter

PRO/II Reference Manual


Volume I: Component and Thermophysical Properties

Electrolyte Thermodynamic Equations

I-141

Thermodynamic Methods

Section 1.2
(26)

c = yi ci
M

i=1

z=

compressibility factor =
n

Pv
RT

a = yi yj aij

(27)
(28)

i=1 j=1

aij = ij +

ij
T

pure-component parameter

ij =

interaction parameter

(29)

where:

ij = ij + ij

(30)

ij = ij + ij

(31)

ij =

i j

(32)

ij =

i j

(33)

ij =
0

1 0
0
+ j
2 i

(34)

i = i = 0.0 for non polar gas

where:
i =

pure component parameter

ij =

interaction parameter

Values for the pure-component and interaction parameters used in equations


(21) through (34) are given in the original paper by Nakamura et al. for the
following components: Ar, CH4, C2H4, C2H6, C3H6, C3H8, CO, CO2, H2,
H2O, H2S, N2, NH3, and SO2.
SRK Method
Calculation of fugacity coefficients from the Soave-Redlich-Kwong equation of
state is described in Section 1.2.4, Equations of State, of this manual.

Note: For both the Nothnagel and Nakamura options, if an electrolyte model
contains a volatile species not covered in the original paper by Nakamura et
al. for the method (MEA and HClO are examples of such species) the fugacity
coefficients all default to 1.0 (ideal gas). The SRK option does not suffer from
this limitation.

I-142

Electrolyte Thermodynamic Equations

May1994

Section 1.2

Thermodynamic Methods

Organic Phase
Activities

Activities of components in an organic liquid phase (if one exists) are obtained
from the Kabadi-Danner modification to the SRK equation of state. This method
is described in Section 1.2.4, Equations of State, of this manual.

Enthalpy

The pressure dependence of the enthalpy for vapor, liquid, and solid phases
is neglected in this thermodynamic framework. However, this does not introduce significant uncertainties in enthalpy calculations over the stated model
validity ranges for temperature (0 to 200 C) and pressure (up to 200 bars).
Because enthalpy is treated as pressure-independent, PRO/II Electrolytes users will receive a warning about potential failure of the flash when pressure
is varied to meet a duty specification.
Vapor and Solid Phases
The enthalpy of the vapor or solid phase at the temperature and solution composition of interest is evaluated using:
NC

(35)

Hvap or Hsol = yi Hi
i=1

where:
Hi, yi = the standard state molar enthalpy and the mole fraction, respectively, of the ith vapor or solid component
NC =

the total number of components present in the vapor or


solid phase.

At the temperature of interest, Hi is evaluated using:

Hi = Hfi,T +
r

T
Tr

(36)

Cp dT
i

where:
Cpi, H fi,Tr = the standard state, isobaric, molar heat capacity, and the
standard state molar enthalpy of formation for the ith
vapor or solid component
Tr =

the reference temperature of 298K.

Values of Cpi for vapor and solid species are obtained from empirical functions of temperature, which are given by:

Cp = Ai + BiT + CiT

(37)

(for vapor)

(38)

Cp = Ai + BiT + Ci / T (for solid)


i

where:
Ai , Bi , C i =

PRO/II Reference Manual


Volume I: Component and Thermophysical Properties

temperature-independent, but phase-dependent


constants characteristic of the ith vapor or solid
components.

Electrolyte Thermodynamic Equations

I-143

Thermodynamic Methods

Aqueous Liquid
Phase

Section 1.2

The enthalpy of the aqueous solution, Haq, at the temperature and solution
composition of interest is evaluated using:
(39)

NI

NM

Haq = Xw Hw + xiHi + xiHi


i=1

i=1

where:
Hw =

the molar enthalpy of pure water at the temperature of interest and at the vapor/liquid saturation pressure of H2O

xi =

the mole fraction of the ith solute species

NM, NI =

the total number of molecular and ionic solute species,


respectively

The enthalpy of pure water is obtained from the equation of state for H2O
given by Haar, Gallagher, and Kell11. At the temperature and solution composition of interest, Hi is evaluated using:

2 lni
Hi = Hi RT

(40)

where:
i =

the activity coefficient of the ith solute species

R =

the gas constant

Values of i and (lni / T) are obtained from equations (2) through (9) for
ionic species and from equations (10) and (11) for molecular solutes. Values
of Hi are obtained from equation (36), using values of Cpi and H fi,Tr for
aqueous species and using the following relations to represent Cpi for ionic
and molecular solutes:

Cp = Cp ,T (for ionic solutes)


i

(41)

Cp = Ai + BiT + CiT (for molecular solutes)

(42)

where:
Ai, Bi, and Ci =

Molar Volume and


Density

temperature-independent, but phase-independent constants


characteristic of the ith molecular solute

Vapor Phase
The density of the vapor phase is evaluated using the equation of state which
corresponds to the chosen vapor fugacity method.

I-144

Electrolyte Thermodynamic Equations

May1994

Section 1.2

Thermodynamic Methods

Aqueous Liquid Phase


The molar volume of the aqueous solution, vaq, at the temperature, pressure,
and solution composition of interest is evaluated using:
NM

NI

vaq = xw vw + xi vi,T + xi vi,T

i=1

(43)

i=1

where:
vw =

the molar volume of pure water at the temperature and pressure of interest, as given by the HGK11 equation of state

vi,Tr =

the standard-state molar volume of the ith molecular or


ionic aqueous solute species, at the reference temperature

Organic Liquid Phase


The density of an organic liquid phase is calculated using the Rackett
method for liquid density. This method is described in Section 1.2.3, Generalized Correlations, of this manual.
Solid Phase
The molar volume of the solid phase, vsol, at the temperature of interest is
evaluated using:
NC

(44)

vsol = xi vi,T

i=1

where:
vi,Tr = the standard state molar volume of the ith pure solid component at the reference temperature

PRO/II Reference Manual


Volume I: Component and Thermophysical Properties

Electrolyte Thermodynamic Equations

I-145

Thermodynamic Methods

Section 1.2

References
1. Zemaitis, J.F. Jr., Clark, D.M., Rafal, M., and Scrivner, N.C., 1986, Handbook of Aqueous Electrolyte Thermodynamics, AIChE.
2a. Wagman, D.D., et al., 1968-1973, Selected Values of Chemical
Thermodynamic Properties, NBS Tech Note, 270-3 to 8.
2b. Chase, M.W. Jr., Davies, C.A., Downey, J.R. Jr., Frurip, D.J., McDonald,
R.A., and Syverud, A.N., 1985, JANAF Thermochemical Tables, 3rd edn.,
J. Phys. Chem. Ref. Data, 14, Supplement no. 1, 1856 pp.
2c. Wagman. D. D., et al., 1982, The NBS Tables of Chemical Thermodynamic
Properties, J. Phys. Chem. Ref. Data, 11, Supplement no. 2, 392 pp.
3. Glushko, V.P., editor, 1965-1981, Thermal Constants of Compounds, Russian Academy of Sciences, Vols. I-X.
4. Zemaitis, J.F., Jr., 1980, Predicting Vapor-Liquid-Solid Equilibria in Multicomponent Aqueous Solutions of Electrolytes, Thermodynamics of Aqueous Systems with Industrial Applications, S.A. Newman, ed., ACS
Symposium Series, 133, 227-246.
5. Bromley, L.A., 1973, Thermodynamic Properties of Strong Electrolytes in
Aqueous Solutions, AIChE J., 19, 313-320.
6. Pitzer, K.S., 1979, Theory: Ion Interaction Approach, Activity Coefficients
in Electrolyte Solutions, 1, 157-208, R.M. Pytkowicz, ed., CRC Press,
Boca Raton, FL.
7. Pitzer, K.S., 1980, Thermodynamics of Aqueous Electrolytes at Various
Temperatures, Pressures and Compositions, Thermodynamics of Aqueous
Systems with Industrial Applications, S.A. Newman, ed., ACS Symposium
Series, 133, 451-466.
8. Meissner, H.P., and Kusik, C.L., 1973, Aqueous Solutions of Two or More
Strong Electrolytes - Vapor Pressures and Solubilities, Ind. Eng. Chem.
Proc. Des. Dev., 12, 205-208.
9. Nothnagel, K.H., Abrams, D.S., and Prausnitz, J.M., 1973, Generalized Correlation of Fugacity Coefficients in Mixtures at Moderate Pressures, Ind.
Eng. Chem. Proc. Des. Dev., 12, 25-35.
10. Nakamura, R., Breedveld, G.J.F., and Prausnitz, J.M., 1976, Thermodynamic Properties of Gas Mixtures Containing Polar and Nonpolar Components, Ind. Eng. Chem. Proc. Des. Dev., 15, 557-564.
11. Haar, L., Gallagher, J.S., and Kell, G.S., 1984, NBS/NRC Steam Tables,
Hemisphere Press, Washington D.C., 320 pp.

I-146

Electrolyte Thermodynamic Equations

May1994

Section 1.2

1.2.11

Thermodynamic Methods

Solid-Liquid Equilibria

General
Information

The solubility of solids in liquids can be described by the vant Hoff (idealsolubility) equation. This is sufficient for many systems where non-idealities
are small. Alternatively, solubility data, correlated as a function of temperature, may be entered directly. Precipitation of solid salts and minerals from
aqueous solutions may be calculated rigorously using PRO/II Electrolytes.
This capability is described separately in Sections 1.2.9, Electrolyte Mathematical Model, and 1.2.10, Electrolyte Thermodynamic Equations, of this manual.

vant Hoff
Equation

The simplest description of the solubility of a solid in a liquid phase is obtained


by assuming that the activity coefficient of the solute in the liquid phase is one.
The solubility is then entirely determined by the ratio of the pure solids fugacity
to its standard-state fugacity in the liquid phase, which is that of a pure subcooled liquid. This ratio is one at the solutes triple point where the solubility
also becomes one. At lower temperatures, it can be calculated with fair accuracy
using the heat of melting; a more accurate estimate results if the heat capacity
change of melting is known. A full derivation of the ideal solubility (or vant
Hoff, after the Dutch chemist who first proposed it) equation is given by Prausnitz et al. The result is:
ln xi =

Hm T
Cp Tt
Cp Tt
1 +
ln
1

RT Tt
R T
R
T

(1)

where:
Hm = enthalpy change of melting
Cp = heat capacity change of melting
Tt =

triple-point temperature

In practice, the more easily accessible melting temperature is usually used instead of the triple-point temperature. The difference is almost always negligible. The ideal-solubility equation predicts the same solubility for a given
solute regardless of solvent composition. It is therefore primarily useful for
systems where the solute and solvent are of a similar chemical nature and form
a nearly ideal solution. For example, the solubility of aromatic hydrocarbons
in benzene is well described by equation (1).

PRO/II Reference Manual


Volume I: Component and Thermophysical Properties

Solid-Liquid Equilibria

I-147

Thermodynamic Methods

Section 1.2
Table 1.2.11-1 vant Hoff
Required pure component properties
Triple-point
temperature (Tt)

(or melting
temperature if Tt
is not available)

Application Guidelines
Components

Solute and solvent should


be of a similar chemical
nature (i.e. form a nearideal solution).

Enthalpy of melting

Solubility Data

For systems where sufficient data exist, solid solubilities may be entered by
the user in the form of a correlation of solubility versus temperature. This
correlation has the same functional form as the vant Hoff equation:
ln xij = Aij + Bij T + Cij ln(T)

(2)

where:
xij =

the equilibrium solubility of solute i in solvent j at


temperature

For solubility of a solid solute i in a mixed solvent, theory dictates that the
mixing rule should have the following form:
ln xi = Zj ln xij

(3)

In equation (3) the sum is over all solvent species, and Zj is the mole fraction
of solvent component j normalized to a solute-free basis. If the normal liquid mole fractions are denoted by z, this is written as:
Zj =

Fill Options for


Solubility Data

zj

(4)

1 zi

In a multicomponent mixture, data may be missing for one or more of the i,j
pairs appearing in the sum in equation (3). Three options are provided for
filling in missing values of xij. The default option is to fill in missing values
with xij as calculated by equation (1), the vant Hoff equation.

Note: When the vant Hoff equation is used as a FILL option, the Cp terms
are ignored.

The FILL = ONE option uses values of one (complete miscibility) for the solutes solubility in the missing solvents. The FILL = FREE option causes the
missing solvent or solvents to be ignored in the solubility calculation. In other
words, if a solvent k is missing solubility data for the solute, the sum in equation
(3) is only taken over those solvents for which data exist, and the mole fractions
in that sum are renormalized to a k-free (as well as solute-free) basis:

I-148

Solid-Liquid Equilibria

May 1994

Section 1.2

Thermodynamic Methods
Zj =

(5)

zj
1 zi zk
k

In equation (5) the sum is over all solvents k for which there are no solubility
data. Note that equation (5) is meaningless if no solvent in the mixture has
solubility data. If FILL = FREE is specified in such a case, the calculations
are defaulted to the vant Hoff equation.
Reference
Prausnitz, J.M., Lichtenthaler, R.N., and Gomes de Azevedo, E., 1986,
Molecular Thermodynamics of Fluid-Phase Equilibria, 2nd edition, PrenticeHall, Englewood Cliffs, NJ, Chapter 9.

PRO/II Reference Manual


Volume I: Component and Thermophysical Properties

Solid-Liquid Equilibria

I-149

Thermodynamic Methods

1.2.12

Section 1.2

Transport Properties

General
Information

The following transport properties are calculated and/or used by PRO/II:

Table 1.2.12-1: Transport Properties


Liquids

Vapors

Viscosity

Viscosity

Thermal Conductivity

Thermal Conductivity

Diffusivity

In addition, PRO/II will calculate the vapor-liquid surface tension for a


stream. Most library components include saturated vapor and liquid values
for viscosity, thermal conductivity, and surface tension as part of the stored
physical property data. Several correlations have also been included in
PRO/II which predict the above properties for hydrocarbon mixtures. Liquid
diffusivity is used by some of the unit operations but is not stored in the
component library. The transport methods are described in the sections that
follow.
With the exception of the PURE and TRAPP methods, there are no special
provisions to characterize the hydrocarbon type (paraffinic, olefinic, etc.).
While these non-characterizing methods may be used for all hydrocarbon
types, or petroleum fractions, the best accuracy is to be expected for paraffins, with a degradation in accuracy for olefins or aromatics. Non-hydrocarbon transport properties are best represented by the properties from structure
methods discussed in Section 1.1, Component Data.

27
PURE Methods

I-150

Transport Properties

PRO/II Note: For more information on specifying PURE methods, see Section
27, Transport and Special Properties, of the PRO/II Keyword Input Manual.
The user may choose to compute transport properties as a weighted average
of pure-component values. These PURE methods (also known as LIBRARY
methods) require that the property in question be available for each component in the mixture with the exception of petroleum pseudocomponents. For
a pseudocomponent, the property (if not supplied by the user) is calculated
using the PETRO method below. The pure-component properties at the temperature of interest are combined to calculate stream average properties according to the following mixing rules:

May 1994

Section 1.2

Thermodynamic Methods
Table 1.2.12-2: Stream Average Properties
Stream Property

Additive Basis

Liquid Thermal Conductivity


2
2
m = wi 1

wi = Weight fraction

(1)

yi = Mole fraction

(2)

xi = Mole fraction

(3)

yi = Mole fraction

(4)

xi = Mole fraction

(5)

Vapor Thermal Conductivity

yi i (MWi)

13

m =

yi (MWi)1 3
i

Liquid Viscosity
m = xi 1i 3
i

Vapor Viscosity

yii (MWi)

12

m =

yi (MWi)1 2
i

Surface Tension
m = xi i
i

The user may also provide individual component values as a function of temperature either in tabular or equation forms.

PETRO Methods

27

PRO/II Note: For more information on specifying the PETRO methods, see Section 27 of the PRO/II Keyword Input Manual, Transport and Special Properties.
Liquid Viscosity
The method selected for liquid viscosity is dependent on the reduced temperature, which is in turn calculated by Kays rule. When the system is near
the critical point (0.98<Tr<1.0), the method developed by Letsou and Steil is
used. For the range 0.76<Tr<0.98, equations relating viscosity to reduced
temperature and acentric factors developed by Letsou and Steil (as cited in
section 9-12 in Reid, Prausnitz and Poling) are used. These methods were developed from data on simple liquid hydrocarbons.
For low temperatures in which the temperature is below the normal boiling
point, a method based on the Arrhenius relation and the Thomas equation
was developed by SimSci. This method relates liquid density at the normal
boiling point, the critical temperature, and the system temperature. The density at
the normal boiling point is estimated using the equation of Gunn and Yamada.

PRO/II Reference Manual


Volume I: Component and Thermophysical Properties

Transport Properties

I-151

Thermodynamic Methods

Section 1.2

For intermediate ranges (Tb<T<0.76Tc), the Andrade equation is used to determine the viscosity, based on reference values at the normal boiling point and
a reduced temperature of 0.76. The reference values are computed by the
methods described above.
After the mixture viscosity has been estimated using one of the above methods,
it is corrected for the effects of pressure using the method of Kouzel (also given
in the API Technical Data Book, pp. 11-47, 2nd edition). The vapor pressure,
needed for this method, is estimated with the Reidel equation.
If water decant is active for the thermodynamic method, water viscosities are
taken from the component data library. A wet viscosity is then calculated
for water/hydrocarbon streams by combining the hydrocarbon and water viscosities with the following mixing rule:
(6)

1 3
13
wet = xaq aq + xhc hc

where:
=

viscosity

x=

mole fraction

subscripts aq and hc refer to the aqueous and hydrocarbon portions


of the stream
References

I-152

Transport Properties

1.

Andrade, E. N., 1930, The Viscosity of Liquids, Nature, 125, 309-310.

2.

Gunn, R. D., and Yamada, T., 1971, A Corresponding States Correlation of


Saturated Liquid Volume, AIChE J., 17, 1341-1345.

3.

Kendall, J., and Monroe, K. P., 1917, The Viscosity of Liquids, II, The
Viscosity - Composition Curve for Ideal Liquid Mixtures, J. Amer. Chem.
Soc., 39, 1787-1802.

4.

Kouzel, 1965, Hydrocarbon Proc., 44, 120.

5.

Letsou, A. and Stiel, L. I., 1973, Viscosity of Saturated Nonpolar Liquids at


Elevated Pressures, AIChE J., 19, 409-411.

6.

Partington, J. R., 1949, An Advanced Treatise in Physical Chemistry, 2,


Longmans, London.

7.

Reid, R. C., Prausnitz, J. M., and Poling, B. E., 1987, The Properties of
Gases and Liquids, 4th edition, McGraw-Hill, New York.

8.

Reidel, L., 1954, Eine Neue Universelle Dampfdrukformel, Chem. Ing. Tech.,
26, 83.

9.

Thomas, L. H., 1946, The Dependence of Viscosities of Liquids on Reduced


Temperature and a Relation Between Viscosity, Density, and Chemical
Constitution, J. Chem. Soc., Part II, 573-579.

May 1994

Section 1.2

Thermodynamic Methods

Liquid Thermal Conductivity


The method of Sato and Reidel is used to calculate liquid thermal conductivities. Figure 12A4.1 (Page 12-11) in the API Technical Data Book is used to
correct thermal conductivities for pressure effects.
If water decant is active for the thermodynamic method, water thermal conductivities are taken from the component data library. A wet thermal conductivity is then calculated for water/hydrocarbon streams by combining the
hydrocarbon and water viscosities with the following mixing rule:
wet =

(7)

1
w 2 + w 2
hc hc
aq aq

12

where:
=

thermal conductivity

w=

weight fraction in the liquid phase

subscripts aq and hc refer to the aqueous (water) and hydrocarbon


portions of the stream
References
1.

American Petroleum Institute, 1978, Data Book, 5th edition.

2.

Reidel, L., 1950, The Determination of the Thermal Conductivity and the
Specific Heat of Various Mineral Oils, Chem. Ing. Tech., 21, 349.

Surface Tension
The surface tension for hydrocarbon liquids is estimated with procedure 10A3.1
(Page 10-17) in the API Technical Data Book (3rd edition). Surface tension for
water is extracted from the component data library. Wet surface tension values for decant hydrocarbon systems are computed with the formula:
wet = xaq aq + xhc hc

(8)

where:
=

surface tension

x=

mole fraction in the liquid phase

subscripts aq and hc refer to the aqueous and hydrocarbon portions


of the stream

PRO/II Reference Manual


Volume I: Component and Thermophysical Properties

Transport Properties

I-153

Thermodynamic Methods

Section 1.2

Vapor Viscosity
Vapor viscosities at low pressures are computed by the method of Thodos et
al. They are then corrected for pressure effects using the equation of Dean
and Stiel. Water vapor viscosities are taken from the component library.
Wet vapor viscosities for decant systems are calculated by the following
combinatorial formula:
12

wet =

1 2

(9)

yaq aq MWaq + yhc hc MWhc


1 2

1 2

yaq MWaq + yhc MWhc

where:
=

mole fraction in the vapor phase

MW = molecular weight
subscripts aq and hc refer to the aqueous and hydrocarbon portions
of the stream
References
1.

Thodos, G., and Yoon, 1970, Viscosity of Nonpolar Gaseous Mixtures


at Normal Pressures, AIChE J., 16, 300-304.

2.

Dean, D.G., and Stiel, L.S., 1965, The Viscosity of Nonpolar Gas Mixtures at
Moderate and High Pressures, AIChE J., 11, 526-532.

3.

Herning, F., and Zipperer, L., 1936, Calculation of the Viscosity of Technical
Gas Mixtures from the Viscosities of the Individual Gases, Gas-U. Wasserfach., 79, 69.

Vapor Thermal Conductivity


The Roy-Thodos method is used to determine vapor thermal conductivities.
The function of temperature used is the one that is presented by the authors
for saturated hydrocarbons. This method is corrected for pressure effects using the equations of Stiel and Thodos.
Data for water are retrieved from the component data library. Wet conductivities are predicted for water/hydrocarbon systems as follows:
13

wet =

1 3

yaq aq MWaq + yhc hc MWhc


1 3

(10)

1 3

yaq MWaq + yhc MWhc

where:
y=

mole fraction in the vapor phase

MW = molecular weight
subscripts aq and hc refer to the aqueous and hydrocarbon
portions of the stream.

I-154

Transport Properties

May 1994

Section 1.2

Thermodynamic Methods

References

TRAPP
Correlation

1.

Perry, R.H., and Green, D., 1984, Perrys Chemical Engineers Handbook,
6th edition, McGraw-Hill, New York.

2.

Roy, D., and Thodos, G., 1979, Thermal Conductivity of Gases, Ind.
Eng.Chem.Fundam., 9, 71-79.

3.

Stiel, L.I., and Thodos, G., 1964, The Thermal Conductivity of Nonpolar Substances in the Dense Gaseous and Liquid Regions., AIChE J., 10, 26.

The TRAPP method predicts viscosities and thermal conductivities for pure
hydrocarbon components and mixtures of hydrocarbons and light inorganics
for both vapor and liquid phases.

Note: Since this method does not predict surface tension, this property is computed from the PETRO correlation described earlier in this section.
A brief description of the method is presented here. For additional details,
see the Ely and Hanley reference paper.
The TRAPP method uses a one-fluid conformal model coupled with the
extended corresponding-states approach developed by Leland and his co-workers. Mathematically, the viscosity of a fluid mixture or component is given by:
x (, T ) = o (o, T o)F

(11)

and the thermal conductivity is represented by:


(, T) = o (o, To)F (i, T)

(12)

where:
=

viscosity

thermal conductivity

F=

a dimensional factor

the fluid density

subscripts x and o denote unknown and reference fluids


superscript refers to the potential or translational contribution
superscript refers to the contribution due to internal degrees
of freedom
This method is applicable to the full range of densities and temperatures, from
the dilute gas to the dense liquid. The required constants for each component
are the critical constants (Tc, Pc and Vc) and the acentric factor. The method is
currently restricted to the list of components given in Table 1.2.12-3.

Note: The authors of the TRAPP method recommend it not be used when water
is present. It should also be used with care at reduced temperatures above 0.925.

PRO/II Reference Manual


Volume I: Component and Thermophysical Properties

Transport Properties

I-155

Thermodynamic Methods

Section 1.2
Table 1.2.12-3: TRAPP Components (3.3 versions)
Component Name

ID Name

Component Name

ID Name

METHANE

METHANE

PROPYLENE

PRLN

ETHANE

ETHANE

1-BUTENE

BUT1

PROPANE

PROPANE

cis-2-BUTENE

BTC2

ISOBUTANE

IC4

trans-2-BUTENE

BTT2

n-BUTANE

BUTANE

ISOBUTENE

IBTE

ISOPENTANE

2MB

1,3-BUTADIENE

13BD

n-PENTANE

PENTANE

1-PENTENE

PNT1

NEOPENTANE

22PR

cis-2-PENTENE

PTC2

n-HEXANE

HEXANE

trans-2-PENTENE

PTT2

2-METHYLPENTANE

2MP

2-METHYL-1-BUTENE

2BT1

3-METHYLPENTANE

3MP

3-METHYL-1-BUTENE

3BT1

2,2-DIMETHYLBUTANE

22MB

2-METHYL-2-BUTENE

2BT2

2,3-DIMETHYLBUTANE

23MB

1-HEXENE

HXE1

n-HEPTANE

HEPTANE

1-HEPTENE

HPT1

n-OCTANE

OCTANE

PROPADIENE

ALEN

n-NONANE

NONANE

1,2-BUTADIENE

12BD

n-DECANE

DECA

BENZENE

BNZN

n-UNDECANE

UNDC

TOLUENE

TOLU

n-DODECANE

DDEC

m-XYLENE

MXYL

n-TRIDECANE

TRDC

o-XYLENE

OXYL

n-TETRADECANE

TDCN

p-XYLENE

PXYL

n-PENTADECANE

PNDC

ETHYLBENZENE

EBZN

n-HEXADECANE

HXDC

CARBON DIOXIDE

CO2

n-HEPTADECANE

HPDC

CARBON MONOXIDE

CO

CYCLOPENTANE

CP

HYDROGEN

H2

METHYLCYCLOPENTANE

MCP

HYDROGEN SULFIDE

H2S

CYCLOHEXANE

CH

OXYGEN

O2

METHYLCYCLOHEXANE

MCH

NITROGEN

N2

ETHYLCYCLOPENTANE

ECP

SULFUR DIOXIDE

SO2

ETHYLCYCLOHEXANE

ECH

WATER

H2O

ETHYLENE

ETLN

References

I-156

Transport Properties

1.

Leland, T. W., Robinson, J. S., and Suther G.A., 1968, Statistical


Thermodynamics of Mixtures of Molecules of Different Sizes, Trans.
Farad. Soc., 64, 1447-1460.

2.

Ely, J. F., and Hanley, H. J. M., 1981, Prediction of Viscosity and Thermal
Conductivity in Hydrocarbon Mixtures-Computer Program TRAPP, Proceedings of 60th Annual Convention, Gas Processors Association.

May 1994

Section 1.2

Special Methods
for Liquid
Viscosity

Thermodynamic Methods

SIMSCI (Twu) Correlation


For petroleum fractions, viscosity prediction methods are usually based on
boiling point and specific gravity. The API Technical Data Book (1978) expresses the Watson correlation in a nomograph form. However, the graphical form is unsuitable for simulation purposes and cannot be extrapolated.
Abbott et. al., (1970, 1971) give a correlation which agrees with the API
data quite well but breaks down when extrapolated.
The SimSci or Twu correlation is based on a perturbation expansion method.
Here, the properties of a real system (petroleum fractions, in this case) are expanded about the values of a reference system (chosen to be n-alkanes). The
kinematic viscosity of a petroleum fraction at 210 F is expressed in the following manner:
450
o 450
= ln 2 +
ln2 +
Tb
Tb

f2 = xSG 21.1141

x = 1.99873

(SG)

(14)

12
Tb

56.7394
12

(13)

1 + 2f2

1 2f2

(15)

Tb
SG = SG SG

(16)

Similar equations for kinematic viscosity at 100 F are given below:


450
o 450 1 + 2f1
= ln 1 +
ln 1 +
Tb
Tb 1 2f1

f1 = 1.33932 xSG 21.1141

(17)

(18)

(SG)
12

Tb

You can then determine the relationship between viscosities at any two temperatures, given the kinematic viscosity at the 2 temperatures from equations
(13-18) above and using the generalized relationship developed by Wright
(and modified by Twu), given in equations (19-22) below.
lnln(Z) = A + Bln (T)
Z = + 0.7 + e
B=

(19)
2

(20)

(1.47 1.84 0.51 )

(lnlnZ1 lnlnZ2)

(21)

(lnT1 lnT2)

lnlnZ = lnlnZ 1 + B(lnT ln T 1)

(22)

where:
=

kinematic viscosity

PRO/II Reference Manual


Volume I: Component and Thermophysical Properties

Transport Properties

I-157

Thermodynamic Methods

Section 1.2

References
1.

Abbott, M.M., Kaufman, T.G., and Domash, L., 1971, A Correlation for
Predicting Liquid Viscosities of Petroleum Fractions, Can. J. Chem. Eng.,
49, 379.

2.

Twu, C.H., 1985, Internally Consistent Correlation for Predicting Liquid Viscosities of Petroleum Fractions, Ind. Eng. Chem. Proc. Des. Dev., 24, 1287.

3.

Wright, W.A., 1969, An Improved Viscosity-temperature Chart for


Hydrocarbons, J. of Materials, 4, 19.

API Method
The API method used in PRO/II is very similar to the SIMSCI correlation discussed above. Instead of using one reference fluid, two reference fluids are
used. Watson plots show that the logarithmic function of viscosity at the
same boiling point temperature is a linear function of API gravity. Hence,
this relationship is extended as follows:
r1

r1

ln() = ln( ) +

API API

ln(r1) ln(r2)
r2
r1

API API

(23)

where:
=

kinematic viscosity of the petroleum fraction

API =

API gravity

superscripts r1,r2 refer to reference fluids 1 and 2


Since all the calculations are made at the same boiling point, equation (23)
can be simplified based on the definition of Watson characterization factor
and of API gravity as:
r2
= r1

(1 2)(K10)

(24)

r1

where:
K=

the Watson Characterization factor.

The viscosity of the reference fluid (either fluid 1 or fluid 2) can be expressed as:
ln( ) = C1 +
r

C2
3
6
+ C3ln(Tb) + C4Tb + C5T6 + C6Tb
T6

(25)

where:
C1, C2, C3, C4, C5, C6 = empirical constants.

I-158

Transport Properties

May 1994

Section 1.2

27

Thermodynamic Methods

PRO/II Note: For more information on specifying the SIMSCI (TWU) or API
liquid viscosity methods, see Section 27, Transport and Special Properties, of
the PRO/II Keyword Input Manual.

Reference
Twu, C. H., 1986, A Generalized Method for Predicting Viscosities of
Petroleum Fractions, AIChE J., 32, 2091-2094.

Liquid Diffusivity

PRO/IIs Dissolver unit requires the diffusivity of the solute in the solvent
liquid if the user does not supply mass transfer data. PRO/II can calculate
these diffusivities based on user-input data or can estimate them from the
Wilke-Chang correlation.
User-supplied Diffusivity Data
When the user supplies diffusivity data, the user must specify which component (or components) is to be considered the solute. Diffusivity data can then
be supplied for each solute in each solvent in the following form:
lnDij = c1 + c2 T + c3lnT

(26)

where:
Dij =

diffusivity (in m2/sec) of solute i in solvent j

T=

temperature in Kelvins

c1, c2, c3 =

constants

For the diffusivity of a solute in a mixture of solvents, the following mixing


rule is used:
lnDi,m = Xj ln Dij

(27)

where:
Di,m = diffusivity (in m2/sec) of solute i in solvent mixture
Xj

= mole fraction of solvent component j on a solute-free basis

In equation (27) the sum is over all solvent components j.

PRO/II Reference Manual


Volume I: Component and Thermophysical Properties

Transport Properties

I-159

Wilke-Chang Correlation
Wilke and Chang (1955) developed a method for estimating the diffusivity of
a solute at infinite dilution in a binary mixture. SimSci has adapted this
method slightly and put it in multicomponent form. The Wilke-Chang correlation is of limited accuracy, but it will generally provide a diffusivity that is
good to within 50%. The correlation should not be used in cases where the
solute is water or where the solute is electrolytic. The equation is as follows:
12

Di,m =

7.4x10

12

j Xj (jMj)

0.6
m Vi

(28)

where:
Di,m = diffusivity (in m2/sec) of solute i in solvent mixture
Xj

Mj =

mole fraction of solvent component j on a solute-free basis


molecular weight of solvent component j

temperature in Kelvins

viscosity of solvent mixture in centipoise

Vi =

molar volume of solute i at its normal boiling point in


cm3/mole

The sum in equation (28) is taken over all solvents j. The factor accounts
for association of the solvent. It is taken as 2.6 for water, 1.9 for methanol,
1.5 for ethanol, 1.20 for propyl alcohols and n-butanol, and 1.0 for all other
solvents.
Reference
Wilke, C.R., and Chang, P., 1955, Correlation of Diffusion Coefficients
in Dilute Solutions, AIChE J., 1, 264-270.

I-160

Transport Properties

May 1994

Index
A
Acentric factor
I-5, I-11, I-16
Activity coefficient
I-39
Alcohol data package
I-117
application guidelines
I-54
binary interaction databank table
I-119
See also NRTL
recommended ranges
I-117
Alpha formulation
I-71
equation forms
I-72
function requirements
I-71
Amine data package
I-127
application guidelines
I-51
equilibrium reactions
I-127
heat of reaction correction
I-129
recommended ranges
I-129
required components
I-128
API Data Book
I-62, I-64 - I-65
API gravity
I-27
API liquid density
I-41, I-87
recommended ranges
I-65
Application guidelines
I-45
high pressure crude units
I-47
hydrofiners
I-48
low pressure crude units
I-46
lube oil units
I-48
reformers
I-48
solids
I-57
solvent de-asphalting units
I-48
Assay streams
I-19, I-136
Associating equation of state
I-80
equilibrium constant for HF association
I-82
extent of association
I-81
hexamerization equilibrium reaction
I-82
mixture properties
I-83
Redlich-Kwong cubic equations of state
I-81
ASTM distillation data
See Distillation data, ASTM D1160
See Distillation data, ASTM D2887
See Distillation data, ASTM D86

B
Benedict-Webb-Rubin-Starling
application guidelines
recommended ranges
Binary interaction parameters

I-84
I-50, I-52
I-85
I-73, I-86

PRO/II Component and Thermophysical Properties


Reference Manual

Blends
See Cutpoints, multiple sets
Braun K10
Bromley-Zemaitis term

I-47, I-63
I-137

C
Carbon Number
I-5
Chao-Seader
I-60
recommended ranges
I-60
Characterization
CAVETT
I-13
Lee-Kesler
I-16
SIMSCI
I-9
Chemical potential
I-38
Component data
library components
I-4
non-library components
I-3
petroleum components
I-9
Component properties
extrapolation conventions
I-7
fixed properties
I-6
properties from structure
I-8
temperature correlation equations
I-7
temperature-dependent properties
I-6
COSTALD liquid density
I-66
application guidelines
I-52 - I-53
characteristic volume
I-66
Critical compressibility factor
I-5
Critical pressure
I-5, I-16
Critical temperature
I-5, I-9 - I-10, I-16
Critical volume
I-5, I-10, I-16
Cubic equations of state
See Equation-of-state methods
Curl-Pitzer
I-62
application guidelines
I-46
Cutpoints
application guidelines
I-21
multiple sets
I-19, I-21
secondary set
I-19

D
DATAPREP
Debye-Hckel constant
Decanting free water
See Free water decanting
Density
definition
ideal

I-6
I-137

I-5
I-44
I-44
Index

Idx-1

Diffusivity
See Transport properties
Dimerization
Dimers
See Dimerization
Dipole moment
DIPPR library
Distillation data
API 1963 method
API 1987 method
ASTM D1160
ASTM D2887
ASTM D86
Edmister-Okamoto method
Dortmund modified UNIFAC, UFT2

I-56, I-116

I-5
I-4
I-25
I-24
I-22
I-23
I-24
I-25
I-104

E
Edmister-Okamoto
See Distillation data
Electrolyte mathematical model
Electrolyte thermodynamics
aqueous phase activities
density
enthalpy
equilibrium constants
molar volume
Nakamura method
Nothnagel method
organic phase activities
SRK method
vapor-phase fugacities
Electrolytes
application guidelines
mathematical model
thermodynamics
Enthalpy
basis
definition
heat of mixing
ideal
ideal-gas
liquid vapor
solid
Entropy
definition
ideal
Equation-of-state methods
alpha formulations
Benedict-Webb-Rubin-Starling
general cubic equation of state
Lee-Kesler-Plcker
mixing rules
Peng-Robinson
Peng-Robinson Huron-Vidal

Idx-2

Index

I-131
I-135
I-136
I-144
I-143
I-135
I-144
I-141
I-139
I-143
I-142
I-139
I-56 - I-57
I-131
I-135
I-5
I-43
I-41
I-111
I-41
I-16, I-143
I-144
I-143
I-41
I-43
I-44
I-69
I-71
I-84
I-69
I-85
I-73, I-78
I-74
I-79

Peng-Robinson PanagiotopoulosReid I-76 - I-77


Redlich-Kwong
I-69
Soave-Redlich-Kwong
I-74
Soave-Redlich-Kwong Huron-Vidal
I-79
Soave-Redlich-Kwong Kabadi-Danner
I-75
Soave-Redlich-Kwong modified
I-77
Soave-Redlich-Kwong PanagiotoupolosReid
I-76
Soave-Redlich-Kwong SimSci
I-77
UNIWAALS
I-83
van der Waals
I-69
Equilibrium
liquid-liquid
I-39 - I-40
solid-liquid
I-147
vapor-liquid
I-40
vapor-liquid electrolyte
I-39
Erbar-modified Chao-Seader
I-61
Erbar-modified Grayson-Streed
I-61
Estimating binary interaction parameters
I-107
azeotropic data
I-107
ideal data
I-107
infinite dilution data
I-107
mutual solubility data
I-107

F
FILL methods
See Filling in missing interaction data
See also Filling in missing solubility data
Filling in missing interaction data
I-107, I-109
databank search order
I-107
Filling in missing solubility data
I-148
Flash point
I-5
Flory-Huggins
I-96
See also Liquid activity methods
Free volume modification to UNIFAC,
UNFV
I-106 - I-107
Free water decanting
I-88
calculation methods
I-88
equation of state water solubility method I-89
kerosene water solubility method
I-89
SimSci water solubility method
I-89
Fugacity
I-38
coefficient
I-39

G
Gamma method
See Enthalpy, heat of mixing
Generalized correlation methods
API liquid density
Braun K10
Chao-Seader
COSTALD liquid density
Curl-Pitzer
Erbar modification to Chao-Seader

I-111
I-58
I-87
I-63
I-60
I-66
I-62
I-61
May 1994

Erbar modification to Grayson-Streed


I-61
Grayson-Streed
I-61
IDEAL
I-58
Improved Grayson-Streed
I-62
Johnson-Grayson
I-64
Lee-Kesler
I-64
Rackett liquid density
I-65
Glycol data package
I-120
binary interaction databank
I-121
GPA sour water data package
See GPSWATER data package
GPSWATER data package
I-125
equilibrium reactions
I-125
recommended ranges
I-126
required components
I-126
Gravity data
I-27
Grayson-Streed
application guidelines
I-47 - I-48, I-53, I-61
Gross heating value
I-5
Group interaction data
I-101
Dortmund modified UNIFAC, UFT2
I-104
free volume modified UNIFAC,
UNFV
I-105 - I-107
Lyngby modified UNIFAC, UFT1
I-104

H
Hayden-OConnell fugacity
application guidelines
dimerization equilibrium reactions
Heat capacity
See Specific heat capacity
Heat of combustion
Heat of formation
Heat of fusion
Heat of mixing
See Enthalpy, heat of mixing
See also Gamma method
See Redlich-Kister expansion
Heat of vaporization
See Latent heat of vaporization
Henrys law gas solubility
application guidelines
application ranges
Henrys constants
solute components
HEXAMER
See Associating equation of state
Hexamer equation of state
Hexamerization
See also Hexamer equation of state
Huron-Vidal SRK and PR
Hydrogen deficiency number

I-116
I-56
I-116

I-5
I-16
I-5
I-111

I-110
I-57
I-110
I-110
I-110

I-80
I-56, I-80
I-79
I-5

PRO/II Component and Thermophysical Properties


Reference Manual

I
IDEAL and LIBRARY
application guidelines
densities
enthalpies
entropies
K-values
Improved Grayson-Streed
application guidelines
Ionic solutions
Ionic systems
See Ionic solutions

I-58
I-53
I-59
I-59
I-59
I-59
I-62
I-47 - I-48
I-56, I-137

J
Johnson-Grayson

I-64

K
K-values
Kabadi-Danner SRK
application guidelines
mixing rule

I-39
I-75
I-50, I-52
I-75

L
Latent heat of vaporization
I-5, I-15
Lee-Kesler
application guidelines
I-47, I-64
Lightends data
I-29
matching
I-29, I-86
Liquid activity methods
I-90
Dortmund modified UNIFAC, UFT2
I-104
Flory-Huggins
I-96
free volume modified UNIFAC,
UNFV
I-106 - I-107
Lyngby modified UNIFAC, UFT1
I-104
Margules
I-93
modified UNIFAC, UFT3
I-105
NRTL
I-98
regular solution
I-95
UNIFAC
I-101
UNIQUAC
I-99
van Laar
I-94
Wilson
I-97
Liquid molar volume
I-5, I-145
Liquid-liquid equilibria
See Equilibrium, liquid-liquid
LLE
See Equilibrium, liquid-liquid
Local composition
I-97
Lower heating value
I-5
Lyngby modified UNIFAC, UFT1
I-104

Index

Idx-3

RK1

M
Margules
application guidelines
See also Liquid activity methods
Melting temperature
Molar volume of liquid
See Liquid molar volume
Molecular weight
Molecular weight data
API method
extended API (EXTAPI)
SimSci method

I-93
I-93
I-148

I-5
I-28
I-29
I-29
I-29

N
Normal boiling point
Normal melting point
NRTL I-98
application guidelines

I-5
I-5
I-54, I-56, I-98

P
Panagiotopoulos-Reid SRK and PR
mixing rule
Peng-Robinson
application guidelines
Petroleum components
acentric factor
CAVETT method
critical properties
Lee-Kesler method
SIMSCI method
Pitzer term
Pollutants
Poynting correction
PROCESS library
Pseudocomponents

I-76
I-76
I-74
I-47 - I-48, I-50
I-9, I-13, I-16
I-13
I-9, I-13, I-16
I-16
I-9
I-137
I-56
I-40, I-91
I-4
I-26

R
Rackett liquid density
one-fluid version
Rackett parameter
Radius of gyration
Raoults law
Redlich-Kister expansion
See Enthalpy, heat of mixing
Redlich-Kwong equation
Regular solution
application guidelines
Reid vapor pressure (RVP)
APICRIDE method
APINAPHTHA method
ASTM D323-73 method
ASTM D323-82 method
ASTM D4953-91 method
ASTM D5191-91 method
Idx-4

Index

I-65
I-66
I-5
I-5
I-59
I-112
I-71
I-95
I-95
I-31
I-31
I-31 - I-32
I-31
I-31
I-32
I-32

See Enthalpy, heat of mixing


RK2
See Enthalpy, heat of mixing

S
SIMSCI library
I-4
SimSci modified SRK and PR
I-77
application guidelines
I-48, I-50, I-54, I-56
mixing rule
I-76
Soave-Redlich-Kwong
I-74
alpha formulation
I-72, I-74
Huron-Vidal modification
I-79
Kabadi-Danner modifications
I-75
Panagiotopoulos-Reid modification
I-76
SimSci modification
I-77
Soave-Redlich-Kwong SimSci
I-77
Solid molar volume
I-145
Solid-liquid equilibrium (SLE)
I-147 - I-148
FILL options
I-148
vant Hoff solubility
I-147
Solubility parameter
(Hildebrand)
I-5, I-12, I-16, I-95 - I-96
Sour water data package
I-122
application guidelines
I-50
Henrys Law constants
I-122
recommended ranges
I-124
Specific gravity
I-5, I-27
Specific heat capacity
I-5, I-16
Surface tension
See Transport properties
SWEQ
See Sour water data package

T
Thermal conductivity
See Transport properties
Thermodynamic Expert System (TES)
I-45
Transport properties
I-150
API method
I-158
diffusivity
I-159
PETRO methods
I-151
PURE methods
I-150
surface tension
I-5
thermal conductivity
I-5, I-153 - I-155
TRAPP methods
I-155
TWU method
I-157
viscosity
I-5, I-151, I-154, I-157
Wilke-Chang method
I-160
Triple-point pressure
I-5
Triple-point temperature
I-5, I-148
True boiling point (TBP) data
cubic spline
I-25
cutting into pseudocomponents
I-26
probability density function (PDF)
I-26
May 1994

quadratic fit
True vapor pressure (TVP)
application guidelines
Truncated virial fugacity
virial coefficients

I-26
I-31
I-32
I-114
I-114

U
UNIFAC
I-101
application guidelines
I-55
data estimation, FILL
I-107
Dortmund modified UNIFAC, UFT2
I-104
free volume modified UNIFAC,
UNFV
I-106 - I-107
Lyngby modified UNIFAC, UFT1
I-104
modified UNIFAC, UFT3
I-105
UNIQUAC
I-99
application guidelines
I-54, I-56
UNIWAALS
application guidelines
I-55 - I-56, I-83
User-supplied solid solubility
I-148

W
Water decant
See Free water decanting
Water handling methods
See Free water decanting
Watson K-factor
Wilke-Chang method
Wilson
application guidelines

I-27
I-160
I-56, I-97

V
van der Waals area and volume
van der Waals equation
van Laar
application guidelines
vant Hoff solid solubility
application guidelines
Vapor phase fugacity
equations of state
Hayden-OConnell
Nakamura
Nothnagel
truncated virial
Vapor pressure
Antoine equation
Vapor-liquid equilibria
See Equilibrium, vapor-liquid
Viscosity
See Transport properties
VLE
See Equilibrium, vapor-liquid

I-5
I-69
I-94
I-94
I-147
I-57
I-113
I-114, I-142
I-116
I-141
I-139
I-114
I-5
I-15

PRO/II Component and Thermophysical Properties


Reference Manual

Index

Idx-5

This page left intentionally blank.

Idx-6

Index

May 1994

S-ar putea să vă placă și