Sie sind auf Seite 1von 160

CALORIMETRIC MEASUREMENTS OF THE MOLAR

HEAT CAPACITY AND THE MOLAR EXCESS

ENTHALPY FOR VARIOUS ALKANOLAMINES IN

AQUEOUS SOLUTIONS

A Thesis

Submitted to the Faculty of Graduate Studies and Research

In Partial Fulfillment of the Requirements

for the Degree of

Master of Applied Science

in Industrial Systems Engineering

University of Regina

by

Mayur Mundhwa

Regina, Saskatchewan

February, 2007

Copyright © 2007: M. Mundhwa


ABSTRACT

The experimental measurements of the molar heat capacity and molar excess enthalpy

for 2-((2-Aminoethyl)amino)ethanol, 3-Amino-1-Propanol, 2-(Methylamino)ethanol, 1-

Amino-2-Propanol, and N-N-Dimethylethanolamine in water were carried in a C80 heat

flow calorimeter at different temperatures and for the entire range of the mole fractions.

The estimated uncertainty in the measured values of the molar heat capacity and molar

excess enthalpy was found to be ± 2 %.

Among the five studied alkanolamines, 2-((2-Aminoethyl)amino)ethanol exhibited

the highest molar heat capacity values and 2-(Methylamino)ethanol exhibited the lowest.

The same solvent also exhibited the highest molar excess enthalpy values on the negative

side (exothermic mixing process), followed by N-N-Dimethylethanolamine, 2-

(Methylamino)ethanol, 3-Amino-1-Propanol, and 1-Amino-2-Propanol.

The molar excess heat capacities were calculated from the experimental molar heat

capacities and correlated as a function of the mole fractions employing the Redlich-Kister

equation. The molar excess enthalpy values were also correlated as a function of the mole

fractions employing the Redlich-Kister equation. The molar enthalpies were also

calculated at infinite dilution.

The molar excess enthalpy values were modeled using the solution theory models:

NRTL (Non Random Two Liquid theory) and UNIQUAC (UNIversal QUAsi Chemical

theory). The molar excess enthalpy values were also modeled using the group

contribution model: modified UNIFAC (UNIversal quasi chemical Functional group

Activity Coefficients - Dortmund). The modified UNIFAC was found to be the most

i
accurate and reliable model for the representation and prediction of the molar excess

enthalpy values.

The present study confirms that the molar heat capacities of alkanolamines are

dominated by –CH2, –OH and amine group contributions which increase with increasing

temperature. The study also confirmed that the molar excess enthalpy values are

dominated by the interaction between –OH and amine groups. The study observed the

molar excess enthalpy values also depend upon the positions of the –OH and –NH2

groups in the primary alkanolamines. Finally, this study confirmed by studying the molar

excess enthalpies of MAE and DMEA that the –CH3 group corresponds to the higher

molar excess enthalpy values on the negative side.

ii
ACKNOWLEDGEMENTS

I would like to express my gratitude to my supervisor, Dr. Amr Henni, whose

guidance, thought provoking suggestions, encouragement and financial assistance

supported me throughout my research. I would also like to thank all the people, who in

different ways helped me with my work on this Master’s thesis.

iii
DEDICATION

I would like to dedicate my thesis to my mother Rajeshriben, who taught me even the

largest task can be accomplished if it is done one step at a time and to my father

Arjunbhai, who taught me the best kind of knowledge to have is that which is learned for

its own sake. This thesis is also dedicated to my brother Himanshu.

Finally, this thesis is dedicated to all those who believe in the richness of learning.

iv
TABLE OF CONTENTS

ABSTRACT i
ACKNOWLEDGEMENT iii
DEDICATION iv
TABLE OF CONTENTS v
LIST OF TABLES viii
LIST OF FIGURES xiii
LIST OF SYMBOLS xvi

1. INTRODUCTION 1

1.1 Objectives 2
1.2 Systems studied 3
1.2.1 2-((2-Aminoethyl)amino)ethanol 4
1.2.2 3-Amino-1-Propanol 4
1.2.3 2-(Methylamino)ethanol 5
1.2.4 1-Amino-2-Propanol 5
1.2.5 N-N-Dimethylethanolamine 6

2. THERMODYNAMICS 8

2.1 Calorimetric methods 10


2.1.1 Adiabatic calorimeter 10
2.1.2 Isothermal calorimeter 11
2.1.3 Isothermal dilution calorimeter 12
2.1.4 Flow calorimeter 13
2.2 Correlation and prediction approaches 15
2.2.1 Empirical expressions 15
2.2.1.1 Molar excess heat capacity 15

v
2.2.1.2 Molar excess enthalpy 15
2.2.2 Solution theory methods 16
2.2.3 Group contribution methods 18
2.3 NRTL (Non-Random Two Liquid) model 19
2.4 UNIQUAC (UNIversal QUAsi-Chemical Theory) model 20
2.5 Modified UNIFAC (Dortmund) model 23
2.6 Determination of the parameters 25

3. EXPERIMENTAL STUDIES 27

3.1 Experimental apparatus 27


3.1.1 Overview of the C80 calorimeter 27
3.2 Calibration of the C80 calorimeter 31
3.2.1 Sensitivity calibration 31
3.2.2 Temperature calibration 32
3.3 Molar heat capacity (Cp) measurements 36
3.3.1 Methods of molar heat capacity (Cp) measurement 36
3.3.1.1 Continuous two steps method 38
3.3.1.2 Continuous three steps method 38
3.4 Molar excess enthalpy (HE) measurements 39
3.4.1 Membrane mixing cell 41
3.5 Sources of error 44
3.6 Verification of the C80 heat flow calorimeter 44
3.6.1 Molar heat capacity measurements 44
3.6.2 Molar excess enthalpy measurements 45
3.7 Solution preparation 50

4. EXPERIMENTAL RESULTS AND DISCUSSION 52

4.1 Materials 52
4.2 Molar heat capacity measurements 52

vi
4.2.1 2-((2-Aminoethyl)amino)ethanol (AEEA) 53
4.2.2 3-Amino-1-Propanol (AP) 57
4.2.3 2-(Methylamino)ethanol (MAE) 60
4.2.4 1-Amino-2-Propanol (MIPA) 64
4.2.5 N-N-Dimethylethanolamine (DMEA) 67
4.3 Comparison of Cp values of alkanolamines 71
4.4 Molar excess enthalpy measurements 73
4.4.1 2-((2-Aminoethyl)amino)ethanol (AEEA) 73
4.4.2 3-Amino-1-Propanol (AP) 81
4.4.3 2-(Methylamino)ethanol (MAE) 88
4.4.4 1-Amino-2-Propanol or Monoisopropanolamine (MIPA) 95
4.4.5 N-N-Dimethylethanolamine (DMEA) 102
4.5 Comparison of HE values of alkanolamines 109

5. CONCLUSION AND RECOMMENDATIONS 113

5.1 Conclusion 113


5.2 Recommendations 114

6. REFERENCES 115

7. APPENDICES 123

7.1 APPENDIX-A: Experimental data 123


7.2 APPENDIX-B: Molecular area and volume parameters 138

vii
LIST OF TABLES

Table (3.1) Sensitivity constants…………………………………………………………….…………………...32


Table (3.2) Temperature correction coefficients……………………………………………..…………..36
Table (3.3) Molar heat capacity data of MEA and MDEA………………………………..………..46
Table (3.4) Molar excess enthalpies of (DEA + H2O) and (MDEA + H2O)
systems..........................................................................................................................................................48
Table (4.1) Redlich-Kister equation fitting coefficients for the molar excess heat
capacity ( C pE ) (equation (4.3)), for AEEA (1) + water (2) mixtures from

(303.15 to 353.15) K……………………………………………………………….………………..56


Table (4.2) Redlich-Kister equation fitting coefficients for the molar excess heat
capacity ( C pE ) (equation (4.3)), for AP (1) + water (2) mixtures from

(303.15 to 353.15) K………………………………………………………………….……………..59


Table (4.3) Comparison of the molar heat capacity (Cp) values for pure MAE……….....63
Table (4.4) Redlich-Kister equation fitting coefficients for the molar excess heat
capacity ( C pE ) (equation (4.3)), for MAE (1) + water (2) mixtures from

(303.15 to 353.15) K………………………………………………………….……………………..63


Table (4.5) Redlich-Kister equation fitting coefficients for the molar excess heat
capacity ( C pE ) (equation (4.3)), for MIPA (1) + water (2) mixtures from

(303.15 to 353.15) K…………………………………………………………………….…………..66


Table (4.6) Comparison of the molar heat capacity (Cp) values for pure DMEA………70
Table (4.7) Redlich-Kister equation fitting coefficients for the molar excess heat
capacity ( C pE ) (equation (4.3)), for DMEA (1) + water (2) mixtures from

(303.15 to 353.15) K……………………………………………………….………………………..70


Table (4.8a) Redlich-Kister equation fitting coefficients for the molar excess enthalpy
(HE) (equation 4.8), for AEEA (1) + water (2) mixtures for (298.15,
313.15, and 323.15) K………………………………………………………………………………77

viii
Table (4.8b) Molar enthalpies of AEEA at infinite dilution in water, ( ΔH 1∞ ), and of

water at infinite dilution in AEEA, ( ΔH 2∞ ), for (298.15, 313.15, and


323.15) K……………………………………………………………………………..…………………...77
Table (4.9a) NRTL and UNIQUAC equations parameters for the molar excess enthalpy
data of AEEA (1) + water (2) for (298.15, 313.15, and 323.15) K……..…...80
Table (4.9b) Modified UNIFAC (Dortmund) equation parameters for the molar excess
enthalpy data of AEEA (1) + water (2) for (298.15, 313.15, and 323.15)
K……………………………………………………………………………………………………………….80
Table (4.10a) Redlich-Kister equation fitting coefficients for the molar excess enthalpy
(HE) (equation (4.9)), for AP (1) + water (2) mixtures for (298.15, 313.15,
and 323.15) K…………………………………………………………………………..……………….83
Table (4.10b) Molar enthalpies of AP at infinite dilution in water, ( ΔH 1∞ ), and of water

at infinite dilution in AP, ( ΔH 2∞ ), for (298.15, 313.15, and 323.15) K…...83


Table (4.11a) NRTL and UNIQUAC equations’ parameters for the molar excess
enthalpy data of AP (1) + water (2) for (298.15, 313.15, and 323.15)
K.........................................................................................................................................................................87
Table (4.11b) Modified UNIFAC (Dortmund) equations’ interaction parameters for the
molar excess enthalpy data of AP (1) + water (2) for (298.15, 313.15, and
323.15) K………………………………………………………………………..………………………...87
Table (4.12a) Redlich-Kister equation fitting coefficients for the molar excess enthalpy
(HE) (equation (4.10)), for MAE (1) + water (2) mixtures for (298.15,
313.15, and 323.15) K………………………………………………………………………………91
Table (4.12b) Molar enthalpies of MAE at infinite dilution in water, ( ΔH 1∞ ), and of

water at infinite dilution in MAE, ( ΔH 2∞ ), for 298.15 K……………….………...91

Table (4.12c) Molar enthalpies of MAE at infinite dilution in water, ( ΔH 1∞ ), and of

water at infinite dilution in MAE, ( ΔH 2∞ ), for (298.15, 313.15, and


323.15) K……………………………………………………………………………..…………………...91
Table (4.13a) NRTL and UNIQUAC equations parameters for the molar excess enthalpy
data of MAE (1) + water (2) for (298.15, 313.15 and 323.15) K………...…..94

ix
Table (4.13b) Modified UNIFAC (Dortmund) equations’ interaction parameters for the
molar excess enthalpy data of MAE (1) + water (2) for (298.15, 313.15
and 323.15) K…………………………………………………………………………..……………….94
Table (4.14a) Redlich-Kister equation fitting coefficients for the molar excess enthalpy
(HE) (equation 4.11), for MIPA (1) + water (2) mixtures for (298.15,
313.15, and 323.15) K………………………………………………………………………………97
Table (4.14b) Molar enthalpies of MIPA at infinite dilution in water, ( ΔH 1∞ ), and of

water at infinite dilution in MIPA, ( ΔH 2∞ ), for (298.15, 313.15, and


323.15) K…………………………………………………………………………………..……………...97
Table (4.15a) NRTL and UNIQUAC equations parameters for the molar excess enthalpy
data of MIPA (1) + water (2) for (298.15, 313.15, and 323.15) K…….…..101
Table (4.15b) Modified UNIFAC (Dortmund) equations’ interaction parameters for the
molar excess enthalpy data of MIPA (1) + water (2) for (298.15, 313.15,
and 323.15) K………………………………………………………………………….……………...101
Table (4.16a) Redlich-Kister equation fitting coefficients for the excess enthalpy (HE)
(equation 4.12), for DMEA (1) + water (2) mixtures for (298.15, 313.15,
and 323.15) K…………………………………………………………………………….…………...104
Table (4.16b) Molar enthalpies of DMEA at infinite dilution in water, ( ΔH 1∞ ), and of

water at infinite dilution in DMEA, ( ΔH 2∞ ), for 298.15 K………...…………..104

Table (4.16c) Molar enthalpies of DMEA at infinite dilution in water, ( ΔH 1∞ ), and of

water at infinite dilution in DMEA, ( ΔH 2∞ ), for (298.15, 313.15, and


323.15) K……………………………………………………………………….……………………….104
Table (4.17a) NRTL and UNIQUAC equations parameters for the molar excess enthalpy
data of DMEA (1) + water (2) for (298.15, 313.15, and 323.15) K…….…108
Table (4.17b) Modified UNIFAC (Dortmund) equations’ interaction parameters for the
molar excess enthalpy data of DMEA (1) + water (2) for (298.15, 313.15,
and 323.15) K……………………………………………………………….………………………...108
Table (A1) Experimental molar heat capacity ( C p /J·mole-1·K-1) values, for AEEA (1)

+ water (2) mixtures from (303.15 to 353.15) K……………………...…………….123

x
Table (A2) Experimental molar heat capacity ( C p /J·mole-1·K-1) values, for AP (1) +

water (2) mixtures from (303.15 to 353.15) K……………………………………….124


Table (A3) Experimental molar heat capacity ( C p /J·mole-1·K-1) values, for MAE (1) +

water (2) mixtures from (303.15 to 353.15) K……………………………………….125


Table (A4) Experimental molar heat capacity ( C p /J·mole-1·K-1) values, for MIPA (1)

+ water (2) mixtures from (303.15 to 353.15) K…………………………...……….126


Table (A5) Experimental molar heat capacity ( C p /J·mole-1·K-1) values, for DMEA (1)

+ water (2) mixtures from (303.15 to 353.15) K…………………………...……….127


Table (A6) Molar excess heat capacity ( C pE /J·mole-1·K-1), for AEEA (1) + water (2)

mixtures from (303.15 to 353.15) K………………………………………………………129


Table (A7) Molar excess heat capacity ( C pE /J·mole-1·K-1), for AP (1) + water (2)

mixtures from (303.15 to 353.15) K………………………………………………………129


Table (A8) Molar excess heat capacity ( C pE /J·mole-1·K-1), for MAE (1) + water (2)

mixtures from (303.15 to 353.15) K………………………………………………………130


Table (A9) Molar excess heat capacity ( C pE /J·mole-1·K-1), for MIPA (1) + water (2)

mixtures from (303.15 to 353.15) K………………………………………………………131


Table (A10) Molar excess heat capacity ( C pE /J·mole-1·K-1), for DMEA (1) + water (2)

mixtures from (303.15 to 353.15) K………………………………………………………132


Table (A11) Molar excess enthalpy (HE/J·mole-1), for AEEA (1) + water (2) mixtures
for (298.15, 313.15, and 323.15) K…………………………………………….…………..133
Table (A12) Molar excess enthalpy (HE/J·mole-1), for AP (1) + water (2) mixtures for
(298.15, 313.15, and 323.15) K……………………………………………………………...134
Table (A13) Molar excess enthalpy (HE/J·mole-1), for MAE (1) + water (2) mixtures for
(298.15, 313.15, and 323.15) K……………………………………………………………...135
Table (A14) Molar excess enthalpy (HE/J·mole-1), for MIPA (1) + water (2) mixtures
for (298.15, 313.15, and 323.15) K…………………………………………….…………..136
Table (A15) Molar excess enthalpy (HE/J·mole-1), for DMEA (1) + water (2) mixtures
for (298.15, 313.15, and 323.15) K…………………………………………….…………..137

xi
Table (B1) Molecular surface area and volume parameters for the UNIQUAC
Model……………………………………………………………………………………………………...138
Table (B2) Main groups and subgroups and the corresponding van der Waals
quantities for the modified UNIFAC (Dortmund) model………………………139

xii
LIST OF FIGURES

Figure (1.1) Chemical structures of the selected alkanolamines……………………………...…….7


Figure (3.1) Block diagram of C80 heat flow calorimeter………………………………..………….28
Figure (3.2) Sectional view of the C-80 heat flow calorimeter…………………………...……….29
Figure (3.3) Sensitivity calibration curve……………………………………………………………………..33
Figure (3.4) Temperature calibration with Indium at 0.4 K/min of scanning rate……….35
Figure (3.5) Standard cell for the molar heat capacity (Cp) measurement…………………...37
Figure (3.6) Three steps method for determination of molar heat capacity signals…......40
Figure (3.7) Molar excess enthalpy graph……………………………………………………………...…….42
Figure (3.8) Membrane mixing cell for the molar excess enthalpy measurement……….43
Figure (3.9) Comparison of the molar heat capacity data of pure MEA……………………...47
Figure (3.10) Comparison of the molar heat capacity data of pure MDEA………………..….47
Figure (3.11) Molar excess enthalpies of DEA (1) + water (2) system at 298.15 K……...49
Figure (3.12) Molar excess enthalpies of MDEA (1) + water (2) system at 313.15 K…..49
Figure (4.1) Molar heat capacity of AEEA in aqueous solution at various
temperatures……………………………………………………………………………………………...54
Figure (4.2) Molar excess heat capacity of AEEA in aqueous solution………………….……54
Figure (4.3) Molar heat capacity of AP in aqueous solution at various temperatures...58
Figure (4.4) Molar excess heat capacity of AP in aqueous solution.............................................58
Figure (4.5) Molar heat capacity of MAE in aqueous solution at various
temperatures……………………………………………………………………………………………...61
Figure (4.6) Molar excess heat capacity of MAE in aqueous solution…………………….…..61
Figure (4.7) Molar heat capacity of MIPA in aqueous solution at various
temperatures……………………………………………………………………………………………...65
Figure (4.8) Molar excess heat capacity of MIPA in aqueous solution………………….…….65
Figure (4.9) Molar heat capacity of DMEA in aqueous solution at various
temperatures……………………………………………………………………………………………...68
Figure (4.10) Molar excess heat capacity of DMEA in aqueous solution……………………...68
Figure (4.11) Comparison of Cp values for the five selected alkanolamines………………....72

xiii
Figure (4.12) Molar excess enthalpies for AEEA (1) + water (2) mixtures……………...……74
Figure (4.13) Molar excess enthalpies of binary mixtures AEEA, diethylamine, MEA,
diethanolamine and ethanol, in water at 298.15………………………………….……74
Figure (4.14) Molar excess enthalpy of AEEA (1) + water (2) represented by the NRTL
Model…………………………………………………………………………………….………………….78
Figure (4.15) Molar excess enthalpy of AEEA (1) + water (2) represented by the
UNIQUAC Model.................................................................................................................................78
Figure (4.16) Molar excess enthalpy of AEEA (1) + water (2) represented by the
UNIFAC Model………………………………………………….…………………………………….79
Figure (4.17) Molar excess enthalpies for AP (1) + water (2) mixtures……………….………..82
Figure (4.18) Molar excess enthalpy of AP (1) + water (2) represented by the NRTL
Model……………………………………………………………………………….……………………….85
Figure (4.19) Molar excess enthalpy of AP (1) + water (2) represented by the
UNIQUAC Model.................................................................................................................................85
Figure (4.20) Molar excess enthalpy of AP (1) + water (2) represented by the UNIFAC
Model………………………………………………………………………….…………………………….86
Figure (4.21) Molar excess enthalpies for MAE (1) + water (2) mixtures.................................89
Figure (4.22) Molar excess enthalpies for MAE (1) + water (2) mixtures at 298.15
K.........................................................................................................................................................................89
Figure (4.23) Molar excess enthalpy of MAE (1) + water (2) represented by the NRTL
Model…………………………………………………………………………………….………………….92
Figure (4.24) Molar excess enthalpy of MAE (1) + water (2) represented by the
UNIQUAC Model.................................................................................................................................92
Figure (4.25) Molar excess enthalpy of MAE (1) + water (2) represented by the
UNIFAC Model…………………………………………………………………….………………….93
Figure (4.26) Molar excess enthalpies for MIPA (1) + water (2) mixtures………………...….96
Figure (4.27) Molar excess enthalpy of MIPA (1) + water (2) represented by the NRTL
Model…………………………………………………………………………….………………………….99
Figure (4.28) Molar excess enthalpy of MIPA (1) + water (2) represented by the
UNIQUAC Model.................................................................................................................................99

xiv
Figure (4.29) Molar excess enthalpy of MIPA (1) + water (2) represented by the
UNIFAC Model……………………………………………………………………………………...100
Figure (4.30) Molar excess enthalpies for DMEA (1) + water (2) mixtures………….…….103
Figure (4.31) Molar excess enthalpies for DMEA (1) + water (2) mixtures at 298.15
K......................................................................................................................................................................103
Figure (4.32) Molar excess enthalpy of DMEA (1) + water (2) represented by the NRTL
Model……………………………………………………………………………………………………...106
Figure (4.33) Molar excess enthalpy of DMEA (1) + water (2) represented by the
UNIQUAC Model………………………………………………………………..…………………106
Figure (4.34) Molar excess enthalpy of DMEA (1) + water (2) represented by the
UNIFAC Model……………………………………………………………………………………...107
Figure (4.35) Comparison of the molar excess enthalpies of the five studied
alkanolamines at 298.15 K..........................................................................................................111
Figure (4.36) Comparison of the molar excess enthalpies of the five studied
alkanolamines at 313.15 K..........................................................................................................111
Figure (4.37) Comparison of the molar excess enthalpies of the five studied
alkanolamines at 323.15 K..........................................................................................................112

xv
LIST OF SYMBOLS

a NRTL and UNIQUAC energy interaction parameter

ai Redlich-Kister equation parameters

anm UNIFAC group interaction parameter between groups n and m, K

b NRTL and UNIQUAC energy interaction parameter, K

bnm UNIFAC group interaction parameter between groups n and m

cnm UNIFAC group interaction parameter between groups n and m, K-1

Cp molar heat capacity, J/mole/K

C p (T ) molar heat capacity of the substance at the desired temperature T,

J/mole/K

C p, reference (T ) molar heat capacity of the reference substance (sapphire) at the desired

temperature T, J/mole/K

C pE molar excess heat capacity, J/mole/K

G parameter in NRTL model

gE excess Gibbs free energy, J/mole

H molar enthalpy, J/mole

ΔH E change in molar excess enthalpy, J/mole

ΔH 1∞ molar enthalpy of alkanolamine at infinite dilution

ΔH 2∞ molar enthalpy of water at infinite dilution

HE molar excess enthalpy, J/mole

xvi
F
H blank heat flow of the blank cells

F
H reference heat flow of the reference material (sapphire)

F
H sample heat flow of the sample

mreference mass of the Reference (Sapphire)

msample mass of the Sample

M Solvent mass of alkanolamine

M Water mass of water

MWSolvent molecular weight of alkanolamine

MWWater molecular weight of water

N number of moles

q, q’ surface area parameter

Q objective function to be minimized by data regression in equation (2.27)

Q heat transfer, J/mole

Qk relative van der Waals surface area of subgroup k

r volume parameter

R universal gas constant, 8.314472(15) J/K/mole

ΔT difference between the final (Tf) and initial (Ti) temperatures, K

Tf final temperature, K

Ti initial temperature, K

x mole fraction

x1 mole fraction of alkanolamine

x2 mole fraction of water

xvii
Xm group mole fraction of group m in the liquid phase

z coordination number

Z variables in equation (2.27)

Acronyms

AAD average absolute deviation

AEEA 2-((2-aminoethyl)amino)ethanol

AP 3-amino-1-propanol

DEA diethanolamine

DEA diethylamine

DMEA dimethylethanolamine

MAE 2-(methylamino)ethanol

MEA monoethanolamine

MDEA methyldiethanolamine

MIPA monoisopropanolamine or 1-Amino-2-Propanol

NRTL non random two liquid

UNIFAC universal quasi chemical functional group activity coefficients

UNIQUAC universal quasi chemical

Greek Letters

α randomness factor in NRTL model

Γk group activity coefficient of group k in the mixture

Γk(i ) group activity coefficient of group k in the pure substance

γ activity coefficient

θ area fraction in UNIQUAC model

xviii
θm surface fraction of group m in the liquid phase

ν k(i ) number of structural groups of type k in molecule i

σZ standard deviation of the indicated data in equation (2.27)

σ standard deviation

τ energy interaction parameters in NRTL and UNIQUAC models

Φ* segment fraction in UNIQUAC model

Ψnm UNIFAC group interaction parameter between groups n and m

Superscripts

c combinatorial

cal calculated value

E excess property

exp experimental value

r residual

Subscripts

1 alkanolamine

12 interactions between alkanolamine and water

2 water

21 interactions between water and alkanolamine

e estimated data in equation (2.27)

i number of variables in equation (2.27)

i, j any species

ij interactions between i and j component

j number of data points in equation (2.27)

xix
ji interactions between j and i component

k number of data sets in equation (2.27)

m measured data in equation (2.27)

nm groups n and m

xx
1. INTRODUCTION

Alkanolamine solutions have been extensively studied over the last 25 years due to

their industrial importance with regard to natural gas, synthetic ammonia plants, fossil-

fuel-fired power plants, chemical and petrochemical industries and the removal of acid

gas impurities such as carbon dioxide (CO2), hydrogen sulfide (H2S) and sulfur dioxide

(SO2) from gas streams (Astarita et al., 1983).

It is very important to capture CO2 and other acid gases to minimize the possible

threat of Greenhouse Effect but it presents technological challenges. The main challenge

for CO2 capture is in the reduction of the overall cost by reducing both the energy and

capital costs (Astarita et al., 1983). The United States Department of Energy noted

existing technology for CO2 capture is very expensive (on the order of $ 150 per ton of

carbon). They also noted analysis performed by SFA Pacific, Inc. that capture CO2 by

existing technologies could increase the cost of electricity by 2.5 cents to 4 cents/kWh.

Recent investigations (Kvamsdal et al., 2004; de Koeijer, 2004) indicate that acid-

gases capture technology will remain competitive in the future. To make this technology

more competitive, it is important to find new, faster and more energy-efficient solvents.

These commercial solvents must have a high net cyclic capacity, high absorption rate for

acid-gases and good chemical stability. In order to minimize the absorption of

components other than the acid-gases by the solvent, it is important to control the solvent

distribution, water content and operating conditions of the process. Because the

efficiency and the cost of any absorption process depend directly on the various

thermodynamic properties of the solvent, investigation of these properties is crucial (Kirk

1
and Othmer, 1992). Some of the most important thermodynamic properties are the molar

heat capacity (Cp) and molar excess enthalpy (HE).

The molar heat capacities (Cp) of alkanolamine solutions are required for the

calculation of the heat duty in condenser, heat exchanger and reboiler used in the gas-

treating processes. In addition, Cp is directly linked with temperature derivatives to basic

thermodynamic functions like enthalpy and entropy. Cp values are also necessary in

evaluating the effect of temperature on phase and reaction equilibria. They also serve as a

sensitive indicator of phase transitions and assist in the understanding of changes in the

structure of liquid solutions (Sandler, 1999; Perry et al., 1999; Prausnitz et al., 1999).

The molar excess enthalpy (HE) values of alkanolamine solutions are required to test

the feasibility of existing solution theories as well as in formulating new solution

theories. The HE values are also used for modeling the multistage, multi-component

equilibria in absorption and stripping columns (Sandler, 1999; Perry et al., 1999;

Prausnitz et al., 1999).

1.1 Objectives

This study focused on the experimental findings of the molar heat capacity (Cp) and

molar excess enthalpy (HE) of five alkanolamines.

The objectives of this study include:

• To determine experimentally the molar heat capacity and molar excess enthalpy for

the following five alkanolamines:

1) 2-((2-Aminoethyl)amino)ethanol (H2N(CH2)2NH(CH2)2OH, AEEA)

2) 3-Amino-1-Propanol (H2NCH2CH2CH2OH, AP)

3) 2-(Methylamino)ethanol (CH3NHCH2CH2OH, MAE)

2
4) 1-Amino-2-Propanol (CH3CH(OH)CH2NH2, MIPA)

5) N-N-Dimethylethanolamine ((CH3)2NCH2CH2OH, DMEA)

The measurements for the molar heat capacity of the above mentioned five

alkanolamines in water were carried out at eleven different temperatures ranging

from (303.15 to 353.15) K and for the entire range of mole fractions. The

experimental measurements of the molar excess enthalpy for the above mentioned

five alkanolamine + water systems were carried out at three different temperatures

(298.15, 313.15 and 323.15) K and for the entire range of mole fractions.

• To calculate the molar excess heat capacities from the experimental molar heat

capacities and to correlate them as a function of the mole fractions employing the

Redlich-Kister equation.

• To correlate the experimental molar excess enthalpies as a function of the mole

fractions employing the Redlich-Kister equation and the solution theory models:

NRTL (Non Random Two Liquid theory) and UNIQUAC (UNIversal QUAsi

Chemical theory) and the Modified UNIFAC (UNIversal quasi chemical Functional

groups Activity Coefficient, Dortmund).

1.2 Systems studied

The thermodynamic properties of five solvents were studied. Among them, two were

primary, one secondary, one tertiary, and one with primary and secondary amine. The

selection criteria and background information are given in the following sections. No

literature data of the molar heat capacity are available for the five selected alkanolamines,

except for the pure MAE and pure DMEA at selective temperatures. Also, no literature

3
data of the molar excess enthalpy are available for the five selected (alkanolamine +

water) systems, except the MAE + water and DMEA + water systems at 298.15 K.

1.2.1 2-((2-Aminoethyl)amino)ethanol

2-((2-Aminoethyl)amino)ethanol (H2N(CH2)2NH(CH2)2OH, AEEA) is commonly

used in the production of fuel additives, lube oil additives, chelating agents, surfactants

and fabric softeners among other applications. AEEA is an organic compound with

primary and secondary amine groups, which together with a hydroxyl group, combine the

features of an ethyleneamine and an ethanolamine.

In 2006, Ma’mun et al. found in a screening test AEEA could be a potential absorbent

for capturing CO2 from post-combustion exhaust gases as it displays a high absorption

rate combined with high net cyclic capacity. The net cyclic capacity of AEEA is

significantly higher than that of the well known monoethanolamine (MEA). It maintains

its absorption power at higher loadings. In terms of vapor pressure, AEEA has a much

lower vapor pressure (0.969 kPa) as compared to the industry standard MEA (15.9 kPa)

at the regeneration temperature of 393.15 K (Wilson et al., 2002). No published data was

available for the molar heat capacity and molar excess enthalpy of AEEA + water

systems.

1.2.2 3-Amino-1-Propanol

3-Amino-1-Propanol (H2NCH2CH2CH2OH, AP) is a primary amine, and is

commonly used as an intermediate in the production of pharmaceuticals, cosmetics,

corrosion inhibitors, synthetic resins and plasticizers. Bavbek et al. (1999) studied the

kinetics between AP and CO2 at 303.15 K and found higher reaction rates for AP with

CO2. The main reason for selecting this amine was to investigate the effects of the

4
primary amine group and hydroxyl group on the thermodynamic properties in

comparison with the other selected alkanolamines. No literature data for molar heat

capacity and molar excess enthalpy were found for the AP + water systems.

1.2.3 2-(Methylamino)ethanol

2-(Methylamino)ethanol (CH3NHCH2CH2OH, MAE) is a secondary amine in which

a methyl group substitutes a hydrogen atom of amino group of monoethanolamine (MEA,

a primary amine). The methyl group theoretically enhances the reaction kinetics because

it increases the basicity of the amine, while not appreciably increasing hindrance around

the nitrogen atom. In 2002, Ali et al. showed that the electron donating methyl group on

the nitrogen atom increases the reaction rate between CO2 and MAE. MAE demonstrated

high absorption rates and CO2 capacities based on performance tests on a pilot plant

scale. In addition, it has excellent regeneration characteristics. Compared with MEA,

20% less regeneration energy is required (Mimura et al., 1997). Maham et al. (1997)

measured the molar heat capacities of 99 % pure MAE at (299.1, 322.8, 348.5, 373.2, and

397.8) K. The molar heat capacities measured in the present study at (299.11, 322.80, and

348.50) K were compared with those measured by Maham et al. (1997). The molar

excess enthalpies measured here at 298.15 K were compared with those measured by

Touhara et al. (1982) under the same temperature condition.

1.2.4 1-Amino-2-Propanol

1-Amino-2-Propanol or monoisopropanolamine (CH3CH(OH)CH2NH2, MIPA) is a

primary amine used as an absorbent of acid gases in the refinery of natural gas and

purification of ammonia. It is also used as an emulsifying agent due to its solubility in

water and low alkalinity, and as a component of insecticides, surfactants, rubber

5
chemicals, corrosion inhibitors and pigment dispersants. Camacho et al. (1997) analyzed

the process of CO2 absorption at high pressures in aqueous MIPA, in relation to the

thermal effects. Bavbek et al. (1999) studied the kinetics between MIPA and CO2. They

found the reaction rate of MIPA is high compared to secondary MAE. No published data

for molar heat capacity and molar excess enthalpy were located for MIPA + water

systems.

1.2.5 N-N-Dimethylethanolamine

Dimethylethanolamine ((CH3)2NCH2CH2OH, DMEA) is a tertiary amine used in the

synthesis of dyestuffs, pharmaceuticals and emulsifiers. The key purpose for selecting

DMEA was to investigate the effects of the different groups (amine, methyl, and

hydroxyl) on the thermodynamic properties in comparison to the selected primary and

secondary alkanolamines. Maham et al. (1997) measured the molar heat capacities of 99

% pure DMEA at (299.1, 322.8, 348.5, 373.2, and 397.8) K. The molar heat capacities

measured in this study at (299.11, 322.80, and 348.50) K were thus compared to those

measured by Maham et al. (1997). The molar excess enthalpies measured in this study at

298.15 K were compared with those measured by Touhara et al. (1982) at the same

temperature.

6
2-((2-Aminoethyl)amino)ethanol 3-Amino-1-Propanol

2-(Methylamino)ethanol

1-Amino-2-Propanol N-N-Dimethylethanolamine

Figure (1.1) Chemical structures of the selected alkanolamines

7
2. THERMODYNAMICS

This chapter discusses calorimetry, molar heat capacity and molar excess enthalpy, as

well as calorimetric methods and an evaluation of a number of existing approaches for

correlating the experimental molar excess heat capacity and molar excess enthalpy

values. There are a number of ways the experimental molar excess heat capacity and

molar excess enthalpy values are correlated and represented, e.g., by empirical

representations, solution theory and group contribution methods. Three selected models

(NRTL, UNIQUAC and Modified UNIFAC-Dortmund) are also presented.

The word calorimetry comes from the Latin word for heat and the Greek word for

measure. A calorimeter is an instrument in which a chemical or physical process takes

place, and by which the heat flow to or from the process is measured to obtain the

required thermodynamic properties. The calorimeter is well-insulated so ideally no heat

interacts with the surroundings. Thus, any heat liberated or absorbed during the studied

process must be measured by the calorimeter. The process of measuring the quantity of

heat exchanged during a chemical or physical process is known as calorimetry.

Thermodynamically, the relationship between the changes in temperature (ΔT) and heat

(Q) transferred during the process is given by:

Q = CpΔT = C (Tf - Ti) (2.1)

Where Q is the heat transfer as a result of change in temperature, the proportionality

constant Cp is the molar heat capacity of a substance, which is defined as the amount of

heat required to change the temperature of the substance by one degree, and ΔT is the

difference between the final (Tf) and initial (Ti) temperatures of a process.

8
The molar excess enthalpy is also an important property in the field of chemical

engineering thermodynamics. It is used to calculate heat duties during mixing and

separation processes and is defined as the change in enthalpy per mole of solution formed

when the pure components are mixed at constant temperature and pressure. When defined

this way, the calorimetrically measured heat of mixing is identical to the molar excess

enthalpy (HE). The excess property of a solution is a measure of the difference between

the property value of a real solution and the value it would have as an ideal solution at the

same temperature, pressure, and composition. Mathematically the change in molar excess

enthalpy is

ΔH E = ΔH − ΔH id (2.2)

(
ΔH E = (H − ∑ xi H i ) − H id − ∑ xi H i ) (2.3)

Where, H is the enthalpy of the real solution, Hi is the enthalpy of the pure chemical

species and xi is the mole fraction of the pure components. For an ideal solution the

enthalpy, Hid, can be written as:

H id = ∑ xi H i (2.4)

The formation of an ideal solution results in no change in molecular energies i.e.,

ΔH id = 0 . Therefore, equation (2.3) is reduced to:

ΔH E = H − ∑ xi H i = ΔH (2.5)

Further, ΔH E can be reduced to HE, by the definition of an excess property:

H E = H − H id = ΔH E (2.6)

From equations (2.5) and (2.6), the enthalpy change of a mixing process, or heat of

mixing, is equal to the molar excess enthalpy of the mixing process (Smith et al., 2001;

Sandler, 1999; Perry et al., 1999).

9
The measurements of molar heat capacity and molar excess enthalpy for different

substances are accomplished by various types of calorimeters depending on the

application, availability and accuracy of the data required. These calorimeters are

classified based on the methods applied.

2.1 Calorimetric methods

2.1.1 Adiabatic calorimeter

Originally designed by McGlashan, the adiabatic calorimeter is one of the most

successful calorimeters for the measurement of heat effects of solid and/or liquid systems

over a wide range of temperatures. This heat effect includes the heats of formation, heats

of chemical reaction, latent heat of phase changes, molar heat capacity, and heats of

mixing (Larkin and McGlashan, 1961; Armitage and Morcom, 1969; Hill and Swinton,

1980; van Miltenburg et al., 1987). In an adiabatic calorimeter the two boundaries are

separated by a perfect insulator, so the majority of the heat produced is retained by the

calorimeter container causing its temperature to rise. The ΔT is measured by an

appropriate thermometer. Such calorimeters are also referred to as integrating apparatus

because they respond to the total heat produced during the reaction, stirring, evaporation

etc. The best types of adiabatic calorimeters are those in which the temperature of the

external boundary is continuously adjusted to match that of the internal boundary, so that

the net heat transfer between the two boundaries is zero and the temperature between the

two boundaries is zero. Other methods to reduce heat transfer between the two

boundaries are to minimize the heat transfer coefficient and to minimize the time for heat

exchange. The main disadvantages of adiabatic calorimeters are they have relatively large

thermal inertia and the temperature of the calorimeter changes during the course of an

10
experiment. Thus, it is not satisfactory for kinetic studies of chemical or biological

transformations.

2.1.2 Isothermal calorimeter

An alternative method to conduct calorimetric experiments adiabatically is in an

isothermal environment. The isothermal calorimeter has two advantages over the

adiabatic calorimeter: (1) No corrections are required for the change in heat content of the

mixing vessel, and (2) No compensation is needed for heat losses to the surroundings.

The isothermal calorimeter was originally developed (Mrazek and Van Ness, 1961;

Savini et al., 1966) for mixtures and has been further improved by Marsh and co-workers

(Stokes et al., 1969; Ewing et al., 1970). Malcolm and co-workers (Malcolm and

Rowlinson, 1957; Kersaw and Malcolm, 1968) described the isothermal calorimeter

based on the Bunsen calorimeter. The stability and accuracy of Bunsen type calorimeters

have been discussed by Davies and Pritchard (1972).

When using the isothermal calorimeter, the temperature of the vessel must be

constant throughout the experiment. To maintain constant temperature, the heat produced

in the vessel is transferred as rapidly as possible to the outer jacket through a high

thermal conductivity medium. This type of calorimeter is capable of producing highly

precise and accurate results for different types of heat effect and is widely used for molar

excess enthalpy measurements (Alonso et al., 1987; Lim et al., 1994). There are two

ways to achieve a constant temperature. One is based on solid-liquid or liquid vapor

phase change. For example, solid-liquid phase change is used in the ice calorimeter

(Ginnings et al., 1953). The volume change involved is used as a measure of the heat

added or removed. The main disadvantage of this calorimeter is the measurements can

11
only be carried out at the phase change temperature of the calorimeter fluid. Another way

to keep the temperature constant is to use electrical heating to balance the removal of

heat, or to use electrical cooling (Peltier effect) to balance the addition of heat (Mann,

1954; Rossini, 1956). The major advantages of this type of calorimeter include easier

control of heat leakage and independence of results from the molar heat capacity of the

calorimeter since its temperature does not change.

2.1.3 Isothermal dilution calorimeter

The isothermal dilution calorimeter provides an alternative method of measuring the

molar excess enthalpy isothermally. For endothermic systems, one component is

injected slowly into a stirred vessel in which a second component has been placed. To

maintain the isothermal condition inside the calorimeter, electrical energy is added at an

appropriate rate. The addition of the second fluid may be terminated at any desired

composition and the heat of mixing may be determined from measurements of the

amount of liquid A originally in the vessel, as well as the amount of B added and the

amount of electrical energy added. Further addition of B allows the determination of the

heat of mixing at other compositions.

Usually the full composition range may be covered in two runs. For exothermic

experiments, electrical energy may be added continuously to raise the temperature

slightly above the thermostat temperature. As the second liquid is injected, the electrical

energy addition is reduced in such a way as to maintain isothermal operation.

An alternative method, and one which is vastly superior, is to use a Peltier

thermoelectric cooler to withdraw heat at a constant rate. In this method isothermal

conditions are maintained by continuously supplying electrical energy.

12
Dilution calorimeters normally use approximately 50 ml of each component to

cover the entire composition range. The precision of the measurements is about 0.2 %

of the maximum value of HE. A more detailed discussion of the various types of

isothermal dilution calorimeters is presented by Marsh (1978).

2.1.4 Flow calorimeter

The dilution calorimeter has one fluid flowing into a reservoir of another. The flow

calorimeter has both liquids injected into a mixing chamber at a steady and known rate.

Isothermal operation is achieved the same way for dilution calorimeter. The major

problem with the flow calorimeter lies in producing a steady, non-pulsed, but readily

variable flow rate. McGlashan and Stoeckli (1969) used a vapor pressure driven flow

system and achieved reproducibility of 0.25 % in the flow rates. They measured HE with

an accuracy of about one percent. The major drawback was their requirement of about

200 cm3 of each component.

Sturevant and Lyons (1969) developed a calorimeter requiring much smaller

samples but had a reproducibility of only two percent. Eliott and Wormald (1976)

constructed a flow calorimeter with a stated reproducibility of 0.2 %. Various

researchers have used modifications of the commercial isothermal flow calorimeter

(Gustin and Renon, 1973; Goodwin and Newshar, 1971; Harsted and Thomson, 1974).

Christensen et al. (1976) developed an isothermal flow calorimeter capable of

measuring both exothermic and endothermic heats capable of mixing over a large range

of pressures. Their modifications included the addition of a Peltier cooler and using a

long flat coil as the mixing vessel. Their precision was stated to be 0.4 percent.

13
Flow calorimeters have the advantage of being easily adaptable to measurements over

a wide range of temperatures and pressures and of being able to cover the full

composition range in only a few hours. The flow calorimeter is one of the best choices

for determining the heat of mixing for most liquid systems. The flow rates can be varied

in a continuous, known manner so that a complete HE composition curve can be produced

quickly with a precision of around 1 % (Rowlinson and Swinton, 1982). With flow

techniques, it is possible to avoid several of the difficulties that may arise in batch

calorimeters of more conventional design. In particular, it is unnecessary to use mercury

to separate the component liquids before mixing and the mixing can readily take place in

the absence of any vapor space. In addition, any desired composition is obtained directly

by setting the relative flow-rates of the components. Also, transient effects during the

initial formation of the mixture are unimportant (Tanaka et al., 1975). Typically, one

mole of each of the pure components is required to produce a complete set of results.

This type of calorimeter is not suitable for viscous fluids or for liquids with large

differences in density.

In practice, the choice of calorimeter depends on the research requirements. As for

the nature of the calorimetric experiment, a batch type calorimeter is usually suitable for

processes such as chemical reaction, phase change, and bio-processes; while a flow

calorimeter is much more efficient in measuring the heat of mixing of liquid systems.

This advantage appears to be especially important when the experiments involve multi

component systems.

14
2.2 Correlation and prediction approaches

The primary aim of the correlation is to maximize the extent of usage of the

experimental data. In addition to this, due to the difficulties involved in the experimental

measurements and the increasing time consumption with each additional component of a

multi-component mixture, it is highly desirable to develop practical approaches capable

of predicting the physical and chemical properties of multicomponent systems.

2.2.1 Empirical expressions

2.2.1.1 Molar excess heat capacity

In this study, the experimental molar excess heat capacity data of the aqueous

alkanolamines were correlated as a function of the mole fractions using the Redlich-

Kister equation (Redlich and Kister, 1948). This equation has the following form for

binary systems:

n
C pE /J·mole -1 ·K -1 = x1 x 2 ∑ a i ( x1 − x 2 )
i
(2.7)
i =0

Where, C pE is the molar excess heat capacity, x1 and x2 are the mole fractions of

alkanolamines and water respectively, and ai are the coefficients that are obtained from

least squares analysis of the dependence of C pE on x1 or x2.

2.2.1.2 Molar excess enthalpy

Molar excess enthalpy or heat of mixing values of the alkanolamine + water systems

were also correlated as a function of the mole fractions using the Redlich-Kister equation:

n
H E / J·mole -1 = x1 x 2 ∑ ai ( x1 − x 2 )
i
(2.8)
i =0

15
Where, H E is the molar excess enthalpy, x1 and x2 are the mole fractions of

alkanolamines and water respectively, and ai are the coefficients that are obtained from

least squares analysis of the dependence of H E on x1 or x2.

The most popular approaches for the heat of mixing correlation and prediction are

based on solution theory and group contribution models.

The majority of the following discussions have been presented by Prausnitz et al.

(1999).

2.2.2 Solution theory methods

Well known solution theories are the regular theory, the quasi-lattice theory, the two-

liquid theory, and the associated solution theory. Solution theories represent the behavior

of a solution in terms of its intermolecular forces and structure. This approach involves

one or more characteristic parameters which are fitted to the experimental data.

In 1946 and 1953 Wohl developed an equation based on the regular solution theory,

which is defined as one in which the components mix with no excess entropy, SE,

provided that there is no volume change upon mixing, i.e., SE = 0 and VE = 0. The main

advantage of this equation is that some rough physical significance can be assigned to the

equation parameters. This type of equation provides a good representation of interactions

of the molecules. Because all its parameters contribute different groups of molecules, this

equation is viewed as an empirical or semi-empirical equation. In other words, the greater

the number of molecules in a group, the greater the number of parameters are required.

Also, Wohl’s expansion for the excess Gibbs energy can be extended systematically to

multicomponent solutions.

16
In 1964, Wilson proposed an equation based on the local composition concept. He

replaced the mole fractions in Flory-Huggins equation with the local compositions. The

equation provides better representations of the non-ideal behavior of real systems rather

than that proposed by Wohl type with the same number of parameters and it is also

capable of representing multicomponent behavior with binary parameters. The main

disadvantage of Wilson’s equation is that it can not handle liquid-liquid immiscibility.

Wang (1996) noted that Wilson equation fails to provide reasonable results for systems

with relatively high HE values.

In 1968-69, Renon and Prausnitz developed the Non-Random Two-Liquid (NRTL)

model based on the two-liquid theory. In this model, they assumed that in a binary system

the liquid has a structure made up of cells of molecules of 1 and 2 types, each surrounded

by assortments of the same molecules, with each of the surrounding molecules in turn

surrounded in a similar manner. The accuracies of this model, for the correlation and the

prediction are similar to the Wilson equation but the model can be used for liquid-liquid

immiscible systems. The main disadvantage of NRTL model over Wilson equation is that

it requires three parameters for each pair of constituents. In 1978, Cruz and Renon

extended this model to electrolyte solutions. It provided good results in the representation

of the molar excess enthalpies of alcohol solutions when the model was combined with

associated solution models (Nagata et al., 1984 and 1985).

In 1975, Abrams and Prausnitz developed the Universal Quasi-Chemical

(UNIQUAC) equation with the concept of two-liquid model and local compositions.

They assumed expression of activity coefficients has two parts, one is the combinatorial

part, which is due to differences in size and shape of the molecules and the second is the

17
residual part, due to energetic interactions. They also showed the Wohl type of equation

and the Wilson and NRTL equations are special cases of the UNIQUAC equation when

suitable values of the parameters are assigned. The major characteristic of the UNIQUAC

equation are:

• Applicable to multicomponent mixtures in terms of binary parameters only

• Applicable to liquid-liquid equilibria

• Temperature dependency is valid over a moderate range

• Superior representation of mixtures of different molecular sizes

The main disadvantages of this equation are in the algebraic complexity of the

equation and often the representation of data is poorer than some simpler equations.

The development of solution theories relies on basic understanding of the solution at

the molecular level. But this method demonstrated great potential and promising future

despite being ideal.

2.2.3 Group contribution methods

There is one other approach for the representation of experimental data which is

based on the concept of group contribution. This approach is based on the assumption

that a real solution is composed of the constituent groups of its components. In 1925,

Langmuir suggested this concept for the calculation of thermodynamic properties. A

systematic development known as the analytical solution of groups (ASOG) was

established by Wilson and Deal (1962), which has been modified by Vera and Vidal in

1984 by an analytical group solution method, known as Simplified Group Method

Analysis (SIGMA). Another method known as the Universal Functional Activity

Coefficient (UNIFAC) has been developed by Frendenslund et al. in 1975. This equation

18
or model is based on the UNIQUAC method, the difference being in the residual term,

which contributes to the interaction of groups in the UNIFAC method. By reducing the

interactions between molecules to the interactions between their functional groups,

UNIFAC allows one to represent a large variety of experimental data with a reduced set

of parameters and to predict with reasonable confidence the behavior of systems for

which experimental data are unavailable.

Another good group contribution model known as DISQUAC (Dispersive-Quasi-

Chemical) model is developed by using simple extension of the quasi-chemical theory.

This model uses the structure-dependent interaction parameters. Each contact, either

polar or non-polar, is characterized by dispersive interchange energy and the polar

contacts are characterized by two additional parameters, the quasi-chemical interchange

energy and the coordination number.

Any group-contribution model requires very basic information about the constituent

groups of each component, but on the other hand it requires more time for calculations.

For the representation of experimental HE data, both the solution theory-related models

and group contribution models are superior.

2.3 NRTL (Non-Random Two Liquid) model

As mentioned previously in this chapter, the Non-Random Two-Liquid model is

based on the local composition concept and two-fluid theory. Renon and Prausnitz

(1968a, 1968b, and 1969) extended the concept of local composition which was

developed by Wilson (1964) for the two fluid theory.

The original form of the NRTL model has the following form for an excess Gibbs

energy:

19
gE ⎛ τ G τ 12 G12 ⎞
= x1 x 2 ⎜⎜ 21 21 + ⎟⎟ (2.9)
RT ⎝ x1 + x 2 G21 x 2 + x1G12 ⎠

, τ 21 = a 21 + 21 , G12 = exp(− α 12τ 12 ) , and G21 = exp(− α 21τ 21 )


b12 b
Where, τ 12 = a12 +
T T

The value of the non-randomness parameter, α , is regressed between 0 and 1 to

obtain more reliable and precise values of the parameters. The subscripts 1 and 2 denote

liquid components alkanolamine and water, respectively. For a binary system there are

five parameters a12, b12, a21, b21 and α , where the b terms represent temperature

dependency. The molar excess enthalpy, which indicates the temperature dependence of

the Gibbs free energy, can be determined by using the Gibbs-Helmholtz relation.

⎛ g Ex ⎞
d ⎜⎜ ⎟
HE ⎝ RT ⎟⎠
= (2.10)
R ⎛1⎞
d⎜ ⎟
⎝T ⎠

From equation (2.9) and (2.10), we can obtain the NRTL model for the molar excess

enthalpy:

HE ⎛ G b ( x (1 − ατ 21 ) + x 2 G 21 ) G12 b12 ( x 2 (1 − ατ 12 ) + x1G12 ) ⎞


= x1 x 2 ⎜⎜ 21 21 1 + ⎟
⎟ (2.11)
R ⎝ ( x1 + x G
2 21 ) 2
( x 2 + x G
1 12 ) 2

Gow (1993) used the partial excess enthalpy data with NRTL model to predict VLE

data. Hanks et al. (1971) also predicted the VLE data with NRTL by using excess

enthalpy data. Hanks et al. (1978) also showed the prediction of VLE data at higher

temperature using lower temperature excess enthalpy data.

2.4 UNIQUAC (UNIversal QUAsi-Chemical theory) model

Due to the lack of binary experimental data, it is difficult to obtain precise and

meaningful values of three binary parameters in the NRTL equation. In 1975, Abrams

20
derived an equation that extends the quasichemical theory developed by Guggenheim

(1952) for non-random mixtures to the mixtures or solutions of different molecular size.

This extension was called the UNIversal QUAsi-Chemical theory or, in short,

UNIQUAC.

The UNIQUAC model consists of two parts, (1) Combinatorial and (2) Residual.

The combinatorial part accounts for a combinatorial effects due to differences in

molecular size and shape and contains mole fraction, the average area fraction, and the

average segment fraction. These fraction values can be calculated from the pure-

component molecular structure constants depending on molecular size and external

surface areas which are from the van der Waals volume and surface areas of the

molecule. No adjustable binary parameters appear in this part and therefore no data or

empirical correlations are required to calculate the combinatorial contribution.

The residual part contributes inter-molecular forces and is computed from mole

fractions, area fractions and the energy of interactions between components. There are

two parameters representing the energy of interactions which are not measured or

tabulated, but must be determined empirically from vapor-liquid or liquid-liquid

equilibrium data.

The UNIQUAC equation is,

gE ⎛ gE ⎞ ⎛ gE ⎞
= ⎜⎜ ⎟⎟ + ⎜⎜ ⎟ (2.12)
RT ⎝ RT ⎠ combinatorial ⎝ RT ⎟⎠ residual

For a binary mixture,

⎛ gE ⎞ Φ* Φ* z ⎛ θ θ ⎞
⎜⎜ ⎟⎟ = x1 ln 1 + x 2 ln 2 + ⎜⎜ x1 q1 ln 1* + x 2 q 2 ln 2* ⎟⎟ (2.13)
⎝ RT ⎠ combinatorial x1 x2 2 ⎝ Φ1 Φ2 ⎠

21
⎛ gE ⎞
⎜⎜ ⎟⎟ ( ) (
= − x1 q1' ln θ1' + θ 2' τ 21 − x 2 q 2' ln θ 2' + θ1'τ 12 ) (2.14)
⎝ RT ⎠ residual

Where the z is the coordination number, and segment fraction, Φ * , and area fractions, θ

and θ ' , are given by

x1 r1 x 2 r2
Φ1* = Φ *2 = (2.15)
x1 r1 + x 2 r2 x1 r1 + x 2 r2

x1 q1 x2 q2
θ1 = θ1 = (2.16)
x1 q1 + x 2 q 2 x1 q1 + x 2 q 2

x1 q1' x 2 q 2'
θ1' = θ 2' = (2.17)
x1 q1' + x 2 q 2' x1 q1' + x 2 q 2'

b12 b21
a12 + a21 +
τ 12 = e T
τ 21 = e T
(2.18)

Parameters r, q, and q’ are pure-component molecular-structure constants depending

on the molecular size and the external surface areas. In the original equation q = q’. In

1978, Anderson adjusted q’ values for water and alcohols to obtain the optimum fit for

the systems containing water and alcohol components. The surface area parameters of the

selected alkanolamines are listed in Table (B1) of Appendix B. τ 12 and τ 21 are two

adjustable parameters. To obtain the expression of UNIQUAC equation for the excess

enthalpy, equation (2.10) is used. Only the residual term (equation 2.14) is used to obtain

the UNIQUAC equation for excess enthalpy as it contains the temperature dependency

term. From equation (2.14) and equation (2.10), the following form of the UNIQUAC

equation is derived.

HE ⎛ θ b τ ⎞ ⎛ θb τ ⎞
− = q1 x1 ⎜⎜ 2 21 21 ⎟⎟ + q 2 x 2 ⎜⎜ 1 12 12 ⎟⎟ (2.19)
R ⎝ θ1 + θ 2τ 21 ⎠ ⎝ θ 2 + θ1τ 12 ⎠

22
2.5 Modified UNIFAC (Dortmund) model

UNIFAC (UNIversal Functional Activity Coefficient) is based on the quasichemical

theory of liquid solutions proposed by Guggenheim (1952). In 1975, Abrams and

Prausnitz generalized this theory and in 1975 Fredenslund et al. applied this method to

functional groups within the molecules. This method has already been used in many areas

e.g. (1) for calculating the vapor-liquid equilibria (Fredenslund et al., 1977), (2) for

calculating the liquid-liquid equilibria (Magnussen et al., 1981), (3) for calculating the

solid-liquid equilibria (Gmehling et al., 1978), (4) for determining the activities in

polymer solutions (Oishi and Prauxnitz, 1978; Gottlieb and Herskowitz, 1981), (5) for

determining the vapor pressures of pure components (Jensen et al., 1981), (6) for

determining the influence of solvent on reaction rate (Lo and Paulaitis, 1981), (7) for

determining the flash points of solvent mixtures (Gmehling and Rasmussen, 1982), (8)

for determining the solubilities of gases (Nocon et al., 1983; Sander et al., 1983). This

literature is cited from Weidlich et al. (1987). Weidlich et al. noted, by considering a

couple of publications, (Thomas and Eckert, 1984) the results obtained by the UNIFAC

method for the calculation of activity coefficients at infinite dilution are unsatisfactory,

particularly for the systems with different sizes of molecules. They also noted the original

UNIFAC model is not able to calculate the excess enthalpy and thus temperature

dependence of the Gibbs excess energy with sufficient accuracy. For these reasons,

Weidlich et al. (1987) made some modifications in the original UNIFAC equation so the

vapor-liquid equilibria, activity coefficients at infinite dilution, and enthalpies of mixing

or heats of mixing can be calculated with sufficient accuracy using only one set of

parameters. The description of the original UNIFAC equation is given elsewhere

23
(Fredenslund et al., 1977). Here, only the temperature dependent term of the modified

UNIFAC (Dortmund) equation is considered as heats of mixing or molar excess enthalpy

data depend on temperature i.e. residual part of the modified UNIFAC (Dortmund)

equation. One of the main differences between the original UNIFAC and modified

UNIFAC (Dortmund) is the introduction of temperature-dependent interaction parameters

as:

⎡ a ⎤
Original UNIFAC: Ψnm = exp ⎢− nm ⎥ (2.20)
⎣ T ⎦

⎡ a + bnmT + c nmT 2 ⎤
Modified UNIFAC (Dortmund): Ψnm = exp⎢− nm ⎥ (2.21)
⎣ T ⎦

To obtain the expression of the modified UNIFAC (Dortmund) equation for the heat

of mixing or excess enthalpy, the Gibbs-Helmholtz equation (2.10) is used. The

expression for the UNIFAC model is:

gE
RT
(
= ∑ xi ln γ iR + ln γ iC ) (2.22)
i

Since the combinatorial part is temperature independent, only the residual part is used for

the calculation of excess enthalpy.

ln γ iR = ∑ν k(i ) (ln Γk − ln Γk(i ) ) (2.23)


k

⎛ ⎞
⎜ ⎛ ⎞ θ Ψ ⎟
ln Γk = Qk ⎜1 − ln⎜ ∑ θ m Ψmk ⎟ − ∑ m km ⎟ (2.24)
⎜ ⎝ m ⎠ m ∑ θ n Ψnm ⎟
⎝ n ⎠

From equation (2.10), (2.22), (2.23) and (2.24), we can obtain the following relations

⎡ ⎛ ∂ ln Γ ⎞ ⎛ ∂ ln Γk(i ) ⎞ ⎤
H E = − RT ∑∑ xiν k(i ) ⎢T ⎜ k
⎟ − T ⎜⎜ ⎟⎟ ⎥ (2.25)
i k ⎢⎣ ⎝ ∂T ⎠ P , x ⎝ ∂T ⎠ P , x ⎥⎦

24
Γk is the group activity coefficient of group k in the mixture and Γk(i ) is the group activity

coefficient of group k in the pure substance.

⎡ ⎤
⎢ Ψkm ∑ θ n Ψnm ⎥
⎛ ∂ ln Γk ⎞ (b + ln Ψmk + 2cmk T )Ψmk ⎡ ⎛ Ψkm ⎞ ⎤
T⎜ ⎟ = Qk ∑ θ m ⎢ mk + n
⎢bkm − bnm + ln⎜⎜ ⎟⎟ + (c km − c nm )2T ⎥ ⎥
⎝ ∂T ⎠ P , x m

⎢ ∑n n nk
θ Ψ ⎛ ⎞
⎜ ∑ θ n Ψnm ⎟
2
⎣ ⎝ Ψnm ⎠ ⎦⎥

⎢⎣ ⎝ n ⎠ ⎥⎦

(2.26)

where,

⎡ a + bnmT + c nmT ⎤ 2
Qm X m
∑ν j
m xj
Ψnm = exp⎢− nm ⎥ , θm = , and X m =
j

⎣ T ⎦ ∑ Qn X n n
∑∑ν
j n
n
j
xj

⎛ ∂ ln Γk( i ) ⎞
The above statement equation (2.26) also supplies to the expression T ⎜⎜ ⎟⎟
⎝ ∂T ⎠ P , x

van der Waals group parameters for the Modified UNIFAC (Dortmund) Method are

listed in Table (B2) of Appendix B.

2.6 Determination of the parameters

The parameters of the selected three models have been determined by using the

ASPEN PLUS software package, which has built in algorithms for the fitting of the

experimental data called the Data Regression System. ASPEN basically performs a non-

linear optimization by varying the parameters to minimize the value of Q in the following

objective function.


⎜ ⎛ ⎛Z −Z ⎞
2
⎞ ⎞⎟
Q = ∑k ⎜ ∑ j ⎜ ∑i ⎜⎜ e m
⎟⎟ ⎟ (2.27)
⎜ ⎜ σ ⎠i ⎟ ⎟⎟
⎝ ⎝ ⎝ Z
⎠ j ⎠k

where,

Q: Objective function to be minimized by data regression

25
Z: Variables

i: Number of variable

j: Number of data points

k: Number of data sets

e: Estimated data

m: Measured data

σZ : Standard deviation of the indicated data

The standard deviations are entered for every variable in each data set. In general, the

standard deviations were set to 0.05 for temperature, 0.1 % for mole fractions and 2 % for

the molar excess enthalpy values. To obtain the optimized values of the parameters, Britt-

Luecke’s algorithm was selected in ASPEN PLUS software.

26
3. EXPERIMENTAL STUDIES

This Chapter is an overview of the experimental apparatus i.e. Setaram C80

calorimeter, calibration of the C80 calorimeter, experimental procedures for measuring

the molar heat capacity and molar excess enthalpy, and verification of the C80

calorimeter along with the possible sources of error during the measurements. The

conclusion describes a calculation method for the preparation of the aqueous

alkanolamine solutions in terms of mole fractions.

3.1 Experimental apparatus

Measurements of the molar heat capacity and molar excess enthalpy were carried out

with “C80 Calvet Calorimeter” manufactured by Setaram Instrumentation (France). The

C80 calorimeter works on Tian-Calvet heat flow principles which were described in

detail by Calvet and Part (1963).

C80 calorimeter measures the rate at which the heat flows into a sample as a function

of temperature. The rate at which heat flows into the sample is the power required to keep

the temperature of the sample rising at a set-heating rate or scanning rate. The actual

measurement is the difference in power supplied to the cells and the calorimetric block.

The measured signal is the power required by the sample heater to maintain the

isothermal condition in both the sample and the reference cells. This signal reflects either

the specific heat or the heat of mixing of the sample.

3.1.1 Overview of the C80 calorimeter

Figures (3.1) and (3.2) show the block diagram and sectional view of the C-80

calorimeter, respectively. As shown in Figure (3.2), the massive aluminium block

27
Power TCP Disk Plotter
Supply Drives

3456A Computer
DVM
PT1

Amplifier DVM Printer


PT2
C1

C2

C1, C2: Calorimetric cells


PT1, PT2: Platinum temperature measuring and controlling probes
DVM: Digital voltmeter
TCP: Temperature controller and programmer

Figure (3.1) Block diagram of C80 heat flow calorimeter

28
Figure (3.2) Sectional view of the C-80 heat flow calorimeter

29
is placed at the center of the cylindrical calorimeter shell, which acts as a calorimetric

thermostat. Two cells identical in design (C1 and C2) are placed separately into two

identical cavities located symmetrically from the centerline of the calorimeter. One cell

acts as a sample or measurement cell and the other as a reference cell. Standard cells and

membrane mixing cells were used for the measurement of the molar heat capacity and

molar excess enthalpy respectively. Detailed descriptions along with the figures of both

these cells are given later in this chapter.

Each cell is surrounded by a separate thermopile known as fluxmeter. Both

fluxmeters are identical in design and are connected in opposition to give a differential

thermopile output. This arrangement cancels interfering signal disturbances caused by the

presence of residual temperature instability or temperature programming. The fluxmeters

thermally connect the two cells to the calorimetric block. Hence, any thermal exchange

between the cells and the calorimetric block leads the fluxmeter to output a proportional

signal, which is channeled to a digital voltmeter.

To measure and to control the temperatures of the two cells, two separate platinum

resistance probes are placed within the calorimeter. They are labeled PT1 and PT2. The

measuring probe PT1 is used to obtain the temperature of the cells within the calorimeter.

It is placed between the measurement cell and the reference cell. This probe has about

100 Ω of resistance at 0 °C and is connected to the temperature safety unit (Setaram

TS1). This unit converts the measured voltage and the current across PT1 into a resistance

and then to a temperature. If the temperature exceeds the set safety temperature, the

power to the calorimeter would be cut off automatically. The controlling probe PT2 is

used to control the temperature of the calorimetric block. This probe has about 20 Ω of

30
resistance at 0 °C and is connected to the temperature controller unit. The temperature

controller unit is also connected with the heater of the aluminum block to maintain the set

experimental temperature of the calorimeter. For cooling and thermal insulation, an air

gap is provided following a layer of insulating material surrounding the block. The C80

calorimeter is equipped with state of the art software (Setsoft 2000) for data collection

and processing.

3.2 Calibration of the C80 calorimeter

For the precise measurement of the molar heat capacity and the molar excess enthalpy

of a sample, the sensitivity and the temperature scales of the C80 calorimeter must first

be calibrated. Calibration verifies the validity of the instrument with the help of well

defined and tested standard procedures. For the C80 calorimeter, the International

Confederation for Thermal Analysis and Calorimetry (ICTAC) has recommended

standard procedures for the sensitivity as well as for the temperature calibration.

3.2.1 Sensitivity calibration

Sensitivity calibration of C80 calorimeter was performed by Joule Effect. The

calibration was performed for the entire temperature range of the C80 calorimeter i.e. 30
o
C to 300 oC at the scanning rate of 0.1 K/minute. Specially designed two Joule effect

calibration cells and a calibration unit (EJ3) supplied by the Setaram, with all the

necessary accessories were used for the calibration. Both cells are identical in design and

composed of a metal cylinder adjusted to fit inside the heat flux transducer.

Before starting the calibration, a measurement cell was connected to the calibration

unit (EJ3) using a Joule Effect cable. The main aim of this Joule Effect calibration

procedure was to supply a constant power to the calibration resistance heater for a known

31
period of time. The first step in this calibration procedure was to allow the calorimeter,

by giving the sufficient period of time (approximately 3 hours), to reach the thermal

equilibrium at the specified temperature. After the sufficient period of time the joule

effect calibration was initiated by pressing the impulsion start button on the EJ3 unit. The

set power in the EJ3 unit was then automatically supplied to the resistance heater of the

joule effect calibration cell. The usual values that were used for calibration for power (P)

is 10 mW, for the time of the joule effect (TEJ) is 2100 sec and for the total time (Ttotal)

is 4500 sec. The power supply setting of 10 mW was examined by measuring the voltage

and current and was found to be 9.998 mW. After initiation of the experiment, the

differential fluxmeter signal increased and eventually leveled off with time. The

sensitivity constants given in Table (3.1) were obtained from the results of the sensitivity

calibration experiment and were entered in the Setsoft 2000 software. Figure (3.3) shows

the sensitivity curve and the calculated sensitivity constants.

Table (3.1) Sensitivity constants

Sensitivity Constants Values ( μ V/mW)


a(0) +3.152393E+001
a(1) -2.194865E-003
a(2) -1.631050E-004
a(3) +3.116455E-007
a(4) -1.950394E-010

3.2.2 Temperature calibration

The temperature scale of the C80 calorimeter must be calibrated to compensate the

lag of actual temperature of the sample with that of the displayed temperature. This lag

occurs due to the time necessary for the heat to transfer from the heater into the sample.

32
Figure (3.3) Sensitivity calibration curve

33
Temperature calibration is done by measuring the display temperature of a phase

change of substance, which has a well documented precise transition temperature.

Sample of the pure substance was encapsulated in aluminum foil and placed inside the

cell. The melting point was found by measuring the rate of heat flow into the sample as a

function of temperature at the melting time. The difference between the observed

transition temperature and the actual transition temperature determines the amount of

adjustment necessary to determine an unknown temperature accurately.

In this experimental study, pure Tin and Indium were used for the purpose of the

temperature calibration. Before placing the sample inside the cell, both the measurement

and reference cells were filled to one third of their height with calcined aluminum oxide

(Al2O3). Since the Al2O3 does not have any phase transitions in the temperature range of

the C80 Calorimeter, the effect of Al2O3 is only to add thermal mass to the cells and

thereby increasing the total molar heat capacity of the loaded sample. The total amount of

heat absorbed by the sample during melting was calculated by measuring the area under

the peak as shown in Figure (3.4). The melting point is the temperature at which the

extrapolated baseline under the peak and the line describing the initial rise in heat flow

during melting, cross each other. When this temperature is expressed in terms of the

displayed temperature on the calorimeter during melting, the temperature is referred to as

the apparent melting temperature. The value obtained for the apparent melting point

depends on the heating rate. This temperature is measured for several heating rates (e.g.

0.4 K/min, 0.8 K/min & 1 K/min) for both Indium & Tin. The software is used to

calculate the temperature difference between the apparent temperature and the actual

temperature to the heating rate, to give the value of b0, b1, b2, and b3, which are the

34
Figure (3.4) Temperature calibration with Indium at 0.4 K/min of scanning rate

35
temperature correction coefficients. Temperature calibration data along with the

temperature correction coefficients are listed in Table (3.2).

Table (3.2) Temperature correction coefficients

Scanning Temperature
Theoretical Experimental Calculated
Sample rate Correction
Temp (°C ) Temp ( °C ) Temp (°C )
(K/min) Coefficients

Indium 0.4 156.598 158.5219 156.418


Indium 0.8 156.598 160.8989 156.671 b0 = 1.19235E+00
Indium 1.0 156.598 161.9948 156.706 b1 = -7.79682E-03
Tin 0.4 231.928 233.6261 232.108 b2 = 5.37130E+00
Tin 0.8 231.928 235.5007 231.855
b3 = -1.13404E-02
Tin 1.0 231.928 236.5286 231.820

3.3 Molar heat capacity (Cp) measurements

The C-80 heat flow calorimeter is capable of measuring heat capacities in the

temperature range from ambient (298.15 K) to 573.15 K. The measurements of the molar

heat capacity data were conducted using the standard cells and in the temperature range

from (303.15 to 353.15) K.

The standard cells used for the molar heat capacity measurements, illustrated in

Figure (3.5), consist of concentric stainless steel cylinders. The diameter and the height of

both cells are 17 mm and 80 mm respectively. The top portion of the cells is closed by a

lid with tapered bearing surface, which is supported by an O-ring.

3.3.1 Methods of molar heat capacity (Cp) measurement

There are two generally accepted methods in C80 Calorimeter to determine the molar

heat capacity at specific temperature.

36
Figure (3.5) Standard cell for the molar heat capacity (Cp) measurement

37
3.3.1.1 Continuous two steps method

In this method, prior to the molar heat capacity measurement of the sample, it is

necessary to conduct a reference run with both cells empty. The main purpose of this run

is to correct the minor asymmetry between the reference cell and the measurement cell

and also between the fluxmeters. For the measurement of the molar heat capacity data, a

sample weighting about 5.5 to 6 grams was placed inside the measurement or sample cell.

The sample cell was placed into the measurement chamber (C1) of the calorimeter. The

reference cell placed in the reference chamber (C2) of the calorimeter remained empty.

Both the measurement cell and the reference cell were allowed to come in isothermal

condition over a period of 3-4 hours. The scanning rate was set at 0.1 K/min for all

sample runs. After allowing a sufficient period of time to reach the isothermal condition

at 298.15 K, the temperature of both cells is increased at the scanning rate of 0.1 K/min.

During this time the calorimetric signal increases as a function of time. On attainment of

the higher set temperature, the calorimeter automatically reverted to an isothermal mode,

causing the calorimetric signal to return to the baseline. The calorimetric signal of the

sample under study can be obtained from this data by subtracting the reference signal

from the measurement signal.

3.3.1.2 Continuous three steps method

This method is widely used for the measurement of the molar heat capacity of liquids.

Becker et al. (2000), and Becker and Gmehling (2001) have applied this method to the

measurement of the molar heat capacity for nine and twelve organic substances

respectively. The scanning rate for the temperature was set at 0.1 K/min. The main

difference in this method, compared to the two steps method, is the addition of one more

38
run for the reference substance (e.g. Sapphire) i.e. the measurements of the molar heat

capacity were conducted with a sample run, a blank run and a third step as a reference

run. The remaining procedure was the same as in the two steps method. In all three steps,

the experimental condition was identical. Figure (3.6) shows the heat flow signals from

the blank, reference and sample runs. Molar heat capacity data were calculated by using

the following equation:

⎛ H sample
F
− H blank
F
⎞⎛ mreference ⎞
C p (T ) = ⎜ ⎟⎜ ⎟C (T ) (3.1)
⎜HF ⎟⎜ m sample ⎟ p ,reference
⎝ reference − H blank
F
⎠⎝ ⎠

where,

C p (T ) : Molar heat capacity of the substance at temperature T

F
H sample : Heat flow for the sample

F
H reference : Heat flow for the reference material (sapphire)

F
H blank : Heat flow for the blank cells

msample : Mass of the sample

mreference : Mass of the reference material (sapphire)

C p, reference (T ) : Molar heat capacity of the reference material (sapphire) at temperature T

3.4 Molar excess enthalpy (HE) measurements

The C-80 heat flow calorimeter is also used for the measurement of the molar excess

enthalpy or heat of mixing, heats of reactions and heats of solution.

In these experimental findings, the heat of mixing data were measured at three

different temperatures (298.15, 313.15, and 323.15) K. A membrane mixing cell was

used for the molar excess enthalpy measurements. Similar to the standard cells, two

39
Figure (3.6) Three steps method for determination of molar heat capacity signals

40
identical mixing cells are required for the measurements of heats of mixing, one is for the

measurement chamber in the calorimeter and another is for the reference chamber. In the

sample or measurement cell, both the amine and water were placed into the separate,

sealed compartments on either side of the membrane (aluminum foil) and were not mixed

until the experiment began. In the reference cell, the amine and water were premixed to

their final desired composition, separated into equal volumes and placed in the two

compartments, separated by an aluminum foil membrane. The reference cell serves to

cancel the heat supplied to maintain isothermal conditions and the negligible heat of

stirring. Both cells were lowered into the calorimeter block to heat to the temperature of

the experiment. Once both vessels have reached the experimental temperature, the

membrane was broken and the solutions were stirred lightly for one minute. Fluxmeters

measured the difference in heat flow between the vessels and the calorimetric block. The

calorimetric block had sufficient thermal mass to supply or receive heat to the vessels

without changing temperature. A curve was generated of differential heat flow versus

time as shown in Figure (3.7). The area under this curve was then integrated by the

computer to calculate the total molar excess enthalpy of the sample solution.

3.4.1 Membrane mixing cell

A membrane mixing cell, illustrated in Figure (3.8), is made from stainless steel. It is

made up of a body containing a shoulder in the centre and fitted with two threaded ends.

The lower part of the cell is used to house a container into which one of the reactive

agents is placed. The container is closed by a captive circular membrane between two

Teflon sealing rings. These dovetail into each other clamping the membrane and closing

the container.

41
Figure (3.7) Molar excess enthalpy graph

42
Figure (3.8) Membrane mixing cell for the molar excess enthalpy measurement

43
When the latter is screwed into the lower part of the body, the ring is pressed onto the

shoulder which seals the lower compartment of the container. The membrane is generally

made up of a 0.015 mm thick aluminum disc. The upper part is used to house the second

sample and a perforated lid, threaded with a Teflon (O) ring, closes this part. A moveable

rod crosses the lid, which can be maneuvered from outside the calorimeter, screws on to

the end of the rod and enables it to be pushed into the cell through the membrane and

turned. The bottom of the rod is threaded; it is used to take a sharp-edged impeller, which

enables the membrane to be torn, making it possible to agitate the reactive agents.

3.5 Sources of error

There are two sources which may cause errors during the experimental measurements

of the molar excess enthalpy: (1) Solvent sticks to the walls of the cell and never mixes

with the solution, and (2) Absorption of CO2 takes place during the weighing process of

the solvents and generates the heat of reaction in addition to the heat generated during the

mixing of the two components.

3.6 Verification of the C80 heat flow calorimeter

3.6.1 Molar heat capacity measurements

The purpose of testing the calorimeter before using it is to determine its reliability. In

this study to verify the reliability of C80 Calorimeter in the molar heat capacity

measurement, two well known alkanolamines: MEA (99 % pure) and MDEA (99+ %

pure) were used. Both solvents were purchased from Sigma Aldrich and were used

without further purification. The heat capacities of both solvents were measured in the

temperature range of (303.15 to 353.15) K at a scanning rate of 0.1 K/min and compared

with the available literature data.

44
Chiu et al. (1999) and Chen et al. (2001) measured the heat capacities of MEA (99 %

pure) and MDEA (99+ % pure) respectively from 303.15 K to 353.15 K using a

differential scanning calorimeter (DSC). They reported their data for 0.1 K/min scanning

rate intervals. As shown in Table (3.3), Figure (3.9) and Figure (3.10) the heat capacities

of MEA and MDEA measured with a C-80 calorimeter are in agreement with those

measured with a differential scanning calorimeter (DSC) by Chiu et al. (1999) and Chen

et al. (2001). The percentage absolute average deviation for the molar heat capacity data

of MEA and MDEA measured in this study is 0.23 % when compared to the values

measured by Chiu et al. (1999) and Chen et al. (2001). The estimated uncertainty of the

measured molar heat capacity data is ± 2 % over the entire range of mole fractions.

3.6.2 Molar excess enthalpy measurements

To verify the reliability of C80 heat flow calorimeter for molar excess enthalpy

measurements, two systems have been chosen from the literature to be investigated.

These systems are methyldiethanolamin + water, and diethanolamine + water. HE

measurements for the systems of MDEA + water, and DEA + water were reported by

Maham et al. (1997) for mole fractions from almost 0.05 to 0.9 at T = 298.15 K for DEA

and 313.15 K for MDEA. As shown in Table (3.4), Figure (3.11), and Figure (3.12), the

molar excess enthalpy values of these systems are in agreement with those measured by

Maham et al. (1997). The estimated uncertainty of the measured molar excess enthalpy

data is ± 2 % over the entire range of mole fractions.

45
Table (3.3) Molar heat capacity data of MEA and MDEA

MEA MDEA
Cp/(J·mole-1·K-1) Cp/( J·mole-1·K-1)
Temperature Temperature
(K) Chiu et al., Present (K) Chen et al., Present
(1999) Study (2001) Study

303.15 167 168 303.15 272 273


308.15 169 169 308.15 275 276
313.15 170 169 313.15 278 279
318.15 172 171 318.15 281 282
323.15 173 173 323.15 285 285
328.15 175 174 328.15 288 288
333.15 176 175 333.15 291 290
338.15 178 178 338.15 295 294
343.15 179 179 343.15 298 298
348.15 180 180 348.15 301 301
353.15 182 182 353.15 304 304

46
184

182

180

178
Cp /(J.mole-1.K-1)

176

174

172

170

168 Chiu et al. (1999)


Present study
166
300 310 320 330 340 350 360
T /K

Figure (3.9) Comparison of the molar heat capacity data of pure MEA

310

305

300
Cp /J mole-1 K-1

295

290

285

280

275 Chen et al. (2001)


Present study
270
300 310 320 330 340 350 360
T /K

Figure (3.10) Comparison of the molar heat capacity data of pure MDEA

47
Table (3.4) Molar excess enthalpies of (DEA + water) and (MDEA + water) systems

DEA + water at 298.15 K MDEA + water at 313.15 K


Maham et al. (1997) Present study Maham et al. (1997) Present study
HE HE HE HE
xDEA xDEA xMDEA xMDEA
(J·mole-1) ( J·mole-1) ( J·mole-1) ( J·mole-1)
0.0529 -678 0.0553 -685 0.0964 -1379 0.1004 -1409
0.0609 -768 0.0732 -896 0.1413 -1762 0.1501 -1803
0.0947 -1099 0.1001 -1141 0.2238 -2178 0.1986 -2074
0.3009 -2041 0.2005 -1780 0.2703 -2327 0.2593 -2271
0.3388 -2086 0.2943 -2079 0.3468 -2326 0.3681 -2308
0.3703 -2070 0.3400 -2135 0.4274 -2252 0.4312 -2237
0.3971 -2077 0.4007 -2082 0.5151 -1942 0.4941 -2072
0.4962 -1920 0.4969 -1916 0.6360 -1609 0.5930 -1743
0.5869 -1629 0.5945 -1609 0.7412 -1143 0.7778 -1048
0.6615 -1372 0.6597 -1389 0.8177 -838 0.8203 -869
0.7425 -1097 0.7421 -1091
0.8362 -712 0.8312 -721
0.9180 -355 0.9283 -302

48
0

-500

-1000
HE /J.mole-1

-1500

-2000
Maham et al. (1997)
Present study
-2500
0.0 0.2 0.4 0.6 0.8 1.0
x1

Figure (3.11) Molar excess enthalpies of DEA (1) + water (2) system at 298.15 K

-500
HE /(J mole-1)

-1000

-1500

-2000
Maham et al. (1997)
Present study
-2500
0.0 0.2 0.4 0.6 0.8 1.0
x1

Figure (3.12) Molar excess enthalpies of MDEA (1) + water (2) system at 313.15 K

49
3.7 Solution preparation

The solutions were prepared on the basis of mole fractions. An analytical balance

Ohaus (Model AP250D, Florham Park, NJ) was used to prepare gravimetrically the

binary mixtures of alkanolamine and deionized water with a precision of ± 0.1 mg. The

overall possible uncertainty in the mole fractions is around ± 0.0001. Amount of water

required for the specific mole fractions of solvent was determined as:

M Solvent
MWSolvent
Mole Fraction x = (3.2)
M Solvent M Water
+
MWSolvent MWWater

Where,

x: mole fraction of alkanolamine

M Solvent : Mass of alkanolamine

M Water : Mass of water

MWSolvent : Molecular weight of alkanolamine

MWWater : Molecular weight of water

N Solvent
x= (3.3)
N Solvent + N Water

Where,

N: Number of moles

M Solvent M Water
N Solvent = , and N Water = (3.4)
MWSolvent MWWater

x( N Solvent + N Water ) = N Solvent (3.5)

N Solvent
N Water = − N Solvent (3.6)
x
50
⎛1 ⎞
N Water = N Solvent ⎜ − 1⎟ (3.7)
⎝x ⎠

M Water M Solvent ⎛ 1 ⎞
= ⎜ − 1⎟ (3.8)
MWWater MWSolvent ⎝ x ⎠

M Solvent ⎛ 1 ⎞
M Water = MWWater ⎜ − 1⎟ (3.9)
MWSolvent ⎝ x ⎠

Equation (3.9) was used to calculate the required amount of water for the specific amount

of solvent for particular mole fraction (x).

51
4. EXPERIMENTAL RESULTS AND DISCUSSION

The molar heat capacities of the aqueous 2-((2-Aminoethyl)amino)ethanol (AEEA),

3-Amino-1-Propanol (AP), 2-(Methylamino)ethanol (MAE), 1-Amino-2-Propanol

(MIPA), and N-N-Dimethylethanolamine (DMEA) solutions were measured at eleven

temperatures ranging from (303.15 to 353.15) K and over the entire range of mole

fractions. The molar excess enthalpies of 2-((2-Aminoethyl)amino)ethanol (AEEA), 3-

Amino-1-Propanol (AP), 2-(Methylamino)ethanol (MAE), 1-Amino-2-Propanol (MIPA),

and N-N-Dimethylethanolamine (DMEA) in water were measured at three temperatures

(298.15, 313.15 and 323.15) K and over the entire range of mole fractions. As mentioned

earlier, the uncertainty in the measured Cp and HE values was found to be ± 2 %.

4.1 Materials

The alkanolamines used in this study were purchased from Sigma-Aldrich with purity

of 99 % for AEEA, ≥ 98 % for AP, 98+ % for MAE, ≥ 98 % for MIPA and 99.5+ % for

DMEA, and were used without further purification. The purity of all alkanolamines used

was based on mass %. A standard reference material, synthetic sapphire (α-Al2O3, 99.99

% pure by mass) was purchased from the National Institute of Standards and Technology

(Gaithersburg, USA).

4.2 Molar Heat Capacity Measurements

Experimentally measured values of the molar heat capacities of the five aqueous

alkanolamine solutions at (303.15, 308.15, 313.15, 318.15, 323.15, 328.15, 333.15,

338.15, 343.15, 348.15, and 353.15) K and for the entire range of mole fractions are

listed in Table (A1), Table (A2), Table (A3), Table (A4) and Table (A5), respectively.

52
4.2.1 2-((2-Aminoethyl)amino)ethanol (AEEA)

As shown in Figure (4.1), the molar heat capacity increases with increasing

temperature. Also the molar heat capacity increases with increasing mole fractions of

AEEA. Molar heat capacities of pure AEEA are represented as a linear function of the

temperature in the range from (303.15 to 353.15) K by:

C p , AEEA / J·mole-1 ·K -1 = 0.31 × T (K ) + 187 (4.1)

In equation (4.1), C p , AEEA represents the molar heat capacity of pure AEEA. The

percentage average absolute deviation (% AAD) for equation (4.1) is found to be 0.18.

The maximum value of C p , AEEA is found at 353.15 K.

The molar excess heat capacities ( C pE ) are calculated from the experimental molar

heat capacity values by:

C pE / J·mole-1 ·K -1 = C p − (x1C p ,1 + x2 C p , 2 ) (4.2)

where C p is the molar heat capacity of the mixture, C p ,1 and C p , 2 are the molar heat

capacities of pure alkanolamine and water respectively, and x1 and x2 are the mole

fractions of alkanolamine and water respectively. The molar heat capacity ( C p , 2 ) values

for water are taken from Osborne et al. (1939). Calculated values of the molar excess heat

capacity of AEEA (1) + water (2) mixtures are listed in Table (A6). Figure (4.2) shows

the concentration dependency of the molar excess heat capacities at various temperatures.

At all temperatures the molar excess heat capacity curves are positive with a maximum

around x1 = 0.2. The C pE value increases with increasing temperature for the entire range

of the mole fractions of AEEA. Figure (4.2) shows sharp changes in the values of the

molar excess heat capacity in water-rich regions.

53
350

300

0.1
250 0.2
Cp/(J mole-1 K-1)

0.3
0.4
200 0.5
0.6
0.7
150 0.8
0.9
Pure
100 AEEA

50
300 310 320 330 340 350 360
T /K
Figure (4.1) Molar heat capacity of AEEA in aqueous solution at various temperatures

50

40
CpE/(J mole-1 K-1)

30

20

10

0
0.0 0.2 0.4 0.6 0.8 1.0
x1

Figure (4.2) Molar excess heat capacity of AEEA in aqueous solution

54
Redlich-Kister relation is employed to correlate the molar excess heat capacity

values:

n
C pE / J·mole -1 ·K -1 = x1 x 2 ∑ a i ( x1 − x 2 )
i
(4.3)
i =0

The coefficients and the standard deviations for AEEA (1) + water (2) system are

presented in Table (4.1). All the fitting polynomials are examined by the F-test

(Bevington, 1969; Shoemaker et al., 1989; Harris, 1991).

55
Table (4.1) Redlich-Kister equation fitting coefficients for the molar excess heat
capacity ( C pE ) (equation (4.3)), for AEEA (1) + water (2) mixtures from
(303.15 to 353.15) K

T/K a0 a1 a2 a3 a4 σ / J·mole-1·K-1

303.15 15.4 -5.61 25.7 -33.0 -15.6 0.15

308.15 23.5 -7.51 36.5 -50.3 -13.6 0.31

313.15 29.1 -13.0 42.6 -51.6 -14.1 0.31

318.15 34.7 -17.5 52.7 -57.3 -20.5 0.34

323.15 40.7 -24.0 61.0 -62.1 -22.2 0.38

328.15 48.7 -31.3 66.0 -66.8 -13.1 0.36

333.15 56.7 -37.8 73.9 -79.1 -6.5 0.38

338.15 66.7 -44.1 82.9 -98.9 3.1 0.45

343.15 78.0 -50.9 97.4 -120 12.1 0.47

348.15 91.0 -59.0 115 -151 24.1 0.52

353.15 106 -70.7 131 -185 54.2 0.50

56
4.2.2 3-Amino-1-Propanol (AP)

As shown in Figure (4.3), the molar heat capacity increases for the entire range of the

mole fractions of AP as well as for pure AP with increasing temperature. The molar heat

capacity values of pure AP are represented as a linear function of the temperature in the

range from (303.15 to 353.15) K by:

C p , AP /J·mole-1 ·K -1 = 0.23 × T (K ) + 134 (4.4)

In equation (4.4), C p , AP represents the molar heat capacity of pure AP. The percentage

average absolute deviation (% AAD) for equation (4.4) is found to be 0.26. The

maximum value of C p , AP is found at 353.15 K.

The molar excess heat capacity ( C pE ) values are calculated from the measured molar

heat capacity values using the equation (4.2) and correlated as function of the mole

fractions employing the Redlich-Kister equation (4.3). The molar excess heat capacity

values and coefficients and the standard deviations for AP (1) + water (2) system are

presented in Table (A7) and Table (4.2) respectively. Figure (4.4) shows the

concentration dependency of the molar excess heat capacities at various temperatures. At

all temperatures the molar excess heat capacity curves are positive with a maximum

around x1 = 0.2. These C pE values increase with increasing temperature for the entire

range of mole fractions of AP.

57
240

220

200 0.1
0.2
0.3
Cp/(J mole-1K-1)

180 0.4
0.5
160 0.6
0.7
0.8
140 0.9
Pure
120 AP

100

80
300 310 320 330 340 350 360
T /K

Figure (4.3) Molar heat capacity of AP in aqueous solution at various temperatures

35

30

25
CpE/(J mole-1 K-1)

20

15

10

0
0.0 0.2 0.4 0.6 0.8 1.0
x1

Figure (4.4) Molar excess heat capacity of AP in aqueous solution

58
Table (4.2) Redlich-Kister equation fitting coefficients for the molar excess heat
capacity ( C pE ) (equation (4.3)), for AP (1) + water (2) mixtures from
(303.15 to 353.15) K

T/K a0 a1 a2 a3 a4 a5 σ /J·mol-1·K-1

303.15 12.9 -22.2 35.7 -7.19 9.26 -- 0.19

308.15 20.4 -24.7 35.8 -18.4 25.2 -- 0.14

313.15 25.6 -27.3 34.5 -18.6 31.7 -- 0.16

318.15 31.6 -29.2 32.8 -22.7 40.7 -- 0.18

323.15 37.6 -31.0 31.0 -29.5 50.6 -- 0.21

328.15 46.5 -33.1 23.5 -32.9 75.2 -7.57 0.17

333.15 54.2 -35.5 20.7 -39.3 93.1 -17.8 0.18

338.15 63.8 -38.6 20.1 -49.7 114 -28.4 0.20

343.15 75.1 -40.1 15.8 -66.7 148 -37.0 0.26

348.15 87.7 -45.3 12.2 -84.2 185 -54.1 0.32

353.15 103 -50.4 8.2 -108 237 -77.4 0.44

59
4.2.3 2-(Methylamino)ethanol (MAE)

As shown in Figure (4.5), the molar heat capacities increase for all mole fractions of

MAE as well as for pure MAE with increasing temperature. The molar heat capacities of

pure MAE are represented as a linear function of temperature in the range from (303.15

to 353.15) K by:

C p , MAE / J·mole-1 ·K -1 = 0.31 × T (K ) + 102 (4.5)

In equation (4.5), C p , MAE represents the molar heat capacity of pure MAE. The

percentage average absolute deviation (% AAD) for equation (4.5) is found to be 0.30.

The maximum value of C p , MAE is found at 353.15 K.

Maham et al. (1997) have measured the molar heat capacities of pure MAE (99 %

pure) at (299.1, 322.8, 348.5, 373.2, and 397.8) K. The Cp values, measured in this

present study at (303.15, 323.15, and 348.15) K, are compared with the values measured

by Maham et al. (1997) in Table (4.3). Values measured in this study are in agreement

with those measured by Maham et al. (1997).

The molar excess heat capacity ( C pE ) values are calculated from the experimental

molar heat capacity values using the equation (4.2) and correlated as function of mole

fractions employing Redlich-Kister equation (4.3). The molar excess heat capacity values

and coefficients and the standard deviations for MAE (1) + water (2) are presented in

Table (A8) and Table (4.4) respectively. Figure (4.6) shows the concentration

dependency of the molar excess heat capacities at various temperatures. At all

temperatures, the molar excess heat capacity curves are positive except for 303.15 K at x1

= 0.9, where it became a little bit negative. The maximum values of molar excess heat

60
220

200

0.1
180 0.2
Cp /(J mole-1 K-1)

0.3
160 0.4
0.5
0.6
140 0.7
0.8
0.9
120
Pure
MAE
100

80
300 310 320 330 340 350 360
T /K

Figure (4.5) Molar heat capacity of MAE in aqueous solution at various temperatures

40

30
CpE/(J mole-1K-1)

20

10

0
0.0 0.2 0.4 0.6 0.8 1.0
x1

Figure (4.6) Molar excess heat capacity of MAE in aqueous solution

61
capacity occurs between x1 = 0.1 and x1 = 0.3. For the temperatures (303.15, 308.15,

313.15, and 318.15) K, the maximum values of C pE are found between x1 = 0.1 and x1 =

0.2, while for the remaining temperatures, the maximum values of C pE occurred between

x1 = 0.2 and x1 = 0.3. These C pE values increase with increasing temperature.

62
Table (4.3) Comparison of the molar heat capacity (Cp) values for pure MAE

Maham et al., (1997) Present study


Temperature /K Cp / J·mol-1·K-1 Temperature /K Cp / J·mol-1·K-1
299.1 191.6 299.11 193
322.8 198.1 322.80 200
348.5 209.4 348.50 210

Table (4.4) Redlich-Kister equation fitting coefficients for the molar excess heat
capacity ( C pE ) (equation (4.3)), for MAE (1) + water (2) mixtures from
(303.15 to 353.15) K

T/K a0 a1 a2 a3 a4 σ /J·mol-1·K-1

303.15 17.4 -22.5 29.4 -62.7 30.3 0.16

308.15 25.4 -25.0 33.5 -68.5 34.7 0.23

313.15 31.2 -27.1 34.7 -68.4 35.1 0.23

318.15 38.5 -27.0 36.7 -72.8 33.5 0.21

323.15 45.7 -28.5 40.0 -74.8 34.3 0.16

328.15 54.0 -31.6 40.9 -78.2 44.3 0.24

333.15 64.4 -34.1 37.2 -82.1 60.4 0.32

338.15 77.2 -33.5 41.2 -99.4 65.2 0.44

343.15 92.5 -33.9 40.0 -110 94.9 0.45

348.15 109 -33.5 43.2 -128 130 0.54

353.15 123 -37.0 68.9 -162 133 0.66

63
4.2.4 1-Amino-2-Propanol (MIPA)

As shown in Figure (4.7), the molar heat capacities increase for all mole fractions of

MIPA as well as for pure MIPA with increasing temperature. The molar heat capacity

values of pure MIPA are represented as a linear function of temperature in the range from

(303.15 to 353.15) K by:

C p , MIPA / J·mole-1 ·K -1 = 0.39 ×T (K ) + 90.4 (4.6)

In equation (4.6), C p , MIPA represents the molar heat capacity of pure MIPA. The

percentage average absolute deviation (% AAD) for equation (4.6) is found to be 0.27.

The maximum value of C p, MIPA is found at 353.15 K.

The molar excess heat capacity ( C pE ) values are calculated from the measured molar

heat capacity values using the equation (4.2) and correlated as function of the mole

fractions employing Redlich-Kister equation (4.3). The molar excess heat capacity values

and the coefficients and the standard deviations for MIPA (1) + water (2) are presented in

Table (A9) and Table (4.5) respectively. Figure (4.8) shows the concentration

dependency of the molar excess heat capacities at various temperatures. For the

temperatures (298.15, 308.15, 313.15, 318.15 and 323.15) K, the molar excess heat

capacity curves are positive with a maximum at around x1 = 0.2, and for the remaining

temperatures, the molar excess heat capacity values become maximum around x2 = 0.3.

These C pE values increase with increasing temperature. Figure (4.8) shows sharp changes

in the molar excess heat capacities in water-rich region.

64
240

220

200 0.1
0.2
Cp /(J mole-1 K-1)

180 0.3
0.4
0.5
160
0.6
0.7
140 0.8
0.9
120 Pure
MIPA
100

80
300 310 320 330 340 350 360
T /K

Figure (4.7) Molar heat capacity of MIPA in aqueous solution at various temperatures

35

30

25
CpE/(J mole-1 K-1)

20

15

10

0
0.0 0.2 0.4 0.6 0.8 1.0
x1

Figure (4.8) Molar excess heat capacity of MIPA in aqueous solution

65
Table (4.5) Redlich-Kister equation fitting coefficients for the molar excess heat
capacity ( C pE ) (equation (4.3)), for MIPA (1) + water (2) mixtures from
(303.15 to 353.15) K

T/K a0 a1 a2 a3 a4 a5 σ /J·mol-1·K-1

303.15 12.1 -25.8 37.4 -29.8 4.65 0.31

308.15 20.2 -30.1 31.6 -34.8 26.6 0.30

313.15 26.9 -31.7 21.6 -36.4 43.1 0.29

318.15 33.3 -35.3 16.3 -36.7 52.0 0.25

323.15 40.8 -35.8 10.4 -48.8 63.2 12.6 0.21

328.15 49.2 -42.6 6.69 -32.0 74.5 -8.97 0.11

333.15 57.5 -51.5 5.52 -7.07 80.5 -45.5 0.01

338.15 67.6 -60.9 8.84 -1.82 85.0 -57.7 0.06

343.15 79.6 -75.1 7.46 15.6 96.1 -85.5 0.07

348.15 94.2 -92.3 -1.09 52.4 126 -148 0.06

353.15 112 -101 -8.32 25.9 156 -150 0.17

66
4.2.5 N-N-Dimethylethanolamine (DMEA)

As shown in Figure (4.9) the molar heat capacity values increase for all mole

fractions of DMEA as well as for pure DMEA as the temperature increases. The molar

heat capacity values of pure DMEA are represented as a function of the temperature in

the range from (303.15 to 353.15) K by:

C p , DMEA / J·mole-1 ·K -1 = 627 − 0.75 × T (K ) − 17261814 ÷ T 2 (K ) (4.7)

In equation (4.7), C p, DMEA represents the molar heat capacity of pure DMEA. The

percentage average absolute deviation (% AAD) for equation (4.7) is found to be 0.25.

The maximum value of C p , DMEA is found at 353.15 K.

Maham et al. (1997) have measured the molar heat capacity of pure DMEA (99 %

pure) at (299.1, 322.8, 348.5, 373.2, and 397.8) K. The Cp values measured in this study

at (303.15, 323.15, and 348.15) K are compared with the values measured by Maham et

al. (1997) in Table (4.6). Values measured in this study are in agreement with those

measured by Maham et al. (1997).

The molar excess heat capacities are calculated from the measured molar heat

capacities using the equation (4.2) and correlated as function of the mole fractions

employing Redlich-Kister equation (4.3). The molar excess heat capacity values and

coefficients and the standard deviations for DMEA (1) + water (2) are presented in Table

(A10) and Table (4.7) respectively. Figure (4.10) shows the concentration dependency of

the molar excess heat capacities at various temperatures. The maximum value of molar

excess heat capacity varies between x1 = 0.2 and x1 = 0.4 as temperature increases. At

303.15 K, maximum occurs at x1 = 0.2, for the temperatures ranging from (308.15 to

67
240

220

200 0.1
0.2
0.3
Cp/(J mole-1K-1)

180
0.4
0.5
160 0.6
0.7
140 0.8
0.9
120 Pure
DMEA

100

80
300 310 320 330 340 350 360
T /K

Figure (4.9) Molar heat capacity of DMEA in aqueous solution at various temperatures

60

50

40
CpE/(J mole-1K-1)

30

20

10

0
0.0 0.2 0.4 0.6 0.8 1.0
x1

Figure (4.10) Molar excess heat capacity of DMEA in aqueous solution

68
333.15) K, maximum occurs at x1 = 0.3 and for the temperatures ranging from (333.15 to

353.15) K, maximum value of molar excess heat capacity occurs at x1 = 0.4. The C pE

value increases with increasing temperature.

69
Table (4.6) Comparison of the molar heat capacity (Cp) values for pure DMEA

Maham et al. (1997) Present study


Temperature /K Cp / J·mol-1·K-1 Temperature /K Cp / J·mol-1·K-1
299.1 205.1 299.11 204
322.8 218.0 322.80 220
348.5 228.8 348.50 225

Table (4.7) Redlich-Kister equation fitting coefficients for the molar excess heat
capacity ( C pE ) (equation (4.3)), for DMEA (1) + water (2) mixtures from
(303.15 to 353.15) K

T/K a0 a1 a2 a3 a4 a5 σ /J·mol-1·K-1

303.15 48.9 -42.7 5.39 -8.12 111 -57.2 0.16

308.15 54.8 -43.5 9.56 -6.56 108 -49.0 0.17

313.15 62.1 -45.0 12.8 -6.77 106 -45.2 0.18

318.15 69.7 -45.9 20.7 4.91 93.9 -61.4 0.17

323.15 78.8 -45.8 24.2 8.36 91.4 -68.8 0.14

328.15 89.2 -42.6 33.9 6.80 81.2 -74.9 0.13

333.15 102 -41.0 48.6 18.2 65.6 -96.5 0.08

338.15 118 -39.2 62.9 39.0 60.3 -133 0.06

343.15 138 -42.2 78.9 89.7 61.1 -206 0.07

348.15 162 -44.4 86.0 129 60.4 -294 0.17

353.15 188 -48.7 101 160 46.8 -402 0.23

70
4.3 Comparison of Cp values of alkanolamines

As shown in Figure (4.11), among the five studied alkanolamines, AEEA + water

system has exhibited the maximum values of the molar heat capacity. Maham et al.

(1997) have analyzed by measuring the molar heat capacities of fourteen pure

alkanolamines and found the values of the molar heat capacity of alkanolamines are

dominated by –CH2 and –OH group contributions and the contributions increase with

increasing temperature. They also concluded that –NH group has exhibited its

contribution on the molar heat capacity in terms of the largest temperature dependency,

while the contribution of –N group is almost zero. In our study we concluded AEEA has

exhibited higher molar heat capacities because of the greater number of –CH2 group

present in the AEEA molecule and also because of the contribution of –NH group. If we

compare the values of the molar heat capacity of MAE and DMEA, we find that

additional methyl group present in DMEA has contributed towards the higher value of

the molar heat capacity of DMEA. This is also confirmed by comparing the molar heat

capacities of two primary amines AP and MIPA, as MIPA has exhibited higher values of

the molar heat capacities compared to AP because one methyl (–CH3) group is present in

MIPA molecule. Figure (4.11) shows that at temperature 338 K the values of the molar

heat capacity of DMEA and MIPA become very near and for higher temperatures the

values of molar excess heat capacity of MIPA exceeds the values of DMEA.

71
320

300

280
Cp /J mole-1K-1

AEEA
260 AP
MAE
240 MIPA
DMEA

220

200

180
300 310 320 330 340 350 360
T /K

Figure (4.11) Comparison of Cp values for the five selected alkanolamines

72
4.4 Molar excess enthalpy measurements

As mentioned earlier, the molar excess enthalpies were measured experimentally for

the five mentioned alkanolamines at three different temperatures (298.15, 313.15 and

323.15) K and over the entire mole fractions range of the alkanolamines. The

experimentally measured values are correlated as a function of the mole fractions using

Redlich-Kister equation. The enthalpies of alkanolamines and water at infinite dilutions

are also determined using the method proposed by Maham et al. (1997). Experimentally

measured molar excess enthalpies are modeled using both the “NRTL” and “UNIQUAC”

models. The molar excess enthalpies are also modeled using the modified functional

groups activity coefficient model: “Modified UNIFAC (Dortmund)”.

4.4.1 2-((2-aminoethyl)amino)ethanol (AEEA)

Molar excess enthalpies for (AEEA (1) + Water (2)) were measured experimentally at

three different temperatures (298.15, 313.15, and 323.15) K. The experimental values are

listed in Table (A11). Figure (4.12) shows the mole fractions dependency of the

experimental molar excess enthalpies at various temperatures. At all temperatures the

molar excess enthalpy curves are negative with a minimum around x1 = 0.35. Negative

molar excess enthalpies indicate a very strong exothermic mixing process between AEEA

and water. The exothermic mixing process between AEEA and water is due to the strong

interaction of the –OH group of water molecules with amine groups (–NH and –NH2) of

AEEA molecules.

The molar excess enthalpies are less negative at 298.15 K and 323.15 K compared to

the values at 313.15 K. Also as shown in the plots of molar excess enthalpy vs. mole

fractions at 298.15 K and 323.15 K in Figure (4.12), the molar excess enthalpies at

73
0

HE /(J mole-1) -1000

-2000

-3000

298.15 K
-4000
313.15 K
323.15 K
RK
-5000
0.0 0.2 0.4 0.6 0.8 1.0
x1
Figure (4.12) Molar excess enthalpies for AEEA (1) + water (2) mixtures

-1000
HE /(J mole-1)

-2000

-3000

AEEA (Present study)


-4000 DEtA (Copp & Everett, 1953)
MEA (Touhara et al., 1982)
DEA (Maham et al., 1997)
Ethanol (Larkin, 1975)
-5000
0.0 0.2 0.4 0.6 0.8 1.0
x1
Figure (4.13) Molar excess enthalpies of binary mixtures AEEA, diethylamine, MEA,
diethanolamine and ethanol, in water at 298.15

74
298.15 K are less negative than the values at 323.15 K till x1 = 0.4, but after x1 = 0.4,

molar excess enthalpies at 323.15 K are becoming more negative than the values at

298.15 K.

Alkanolamines are multifunctional compounds with amine and hydroxyl groups.

Figure (4.13) shows the effect of such groups on the molar excess enthalpies by

comparing the molar excess enthalpies of AEEA + water with the molar excess

enthalpies of ethanol + water (Larkin, 1975), diethylamine + water (Copp and Everett,

1953), monoethanolamine + water (Touhara et al., 1982) and diethanolamine + water

(Maham et al., 1997). Maham et al. (1997) concluded by comparing the magnitude of the

molar excess enthalpies of water + ethanol and water + diethylamine the interaction

between water (–OH group) and diethylamine (–NH group) is more dominant than the

interactions between –OH groups from water and ethanol molecules. Thus, ethanol and

amine groups have the opposite effect on the molar excess enthalpies of aqueous binary

mixtures. This is also concluded by comparing the values between monoethanolamine

and diethanolamine, as diethanolamine has exhibited lower molar excess enthalpies on

the negative side compared to the values of monoethanolamine because the two

molecules of ethanol are present in diethanolamine. This is also proved in our study in the

case of AEEA + water as the molar excess enthalpies of this system are quite high in

negative side than the values of water + ethanol. The molar excess enthalpies of AEEA +

water system are even more negative than the values of water + diethylamine. This is

because the strong interaction of both primary and secondary amine groups are present in

the AEEA molecule with water.

A Redlich-Kister relation was used to correlate the molar excess enthalpy data:

75
n
H E / J·mol -1 = x1 x 2 ∑ ai ( x1 − x 2 )
i
(4.8)
i =0

The coefficients and the standard deviations are presented in Table (4.8a). The calculated

molar excess enthalpies, by using equation (4.8), are plotted in Figure (4.12) as solid lines

for all three temperatures. The percentage average absolute deviations (% AAD) are 0.9,

2.7, and 3.9 for the temperatures 298.15 K, 313.15 K and 323.15 K respectively. All the

fitting polynomials were examined by the F-test (Bevington, 1969; Shoemaker et al.,

1989; Harris, 1991).

Both the molar enthalpy of solution of AEEA at infinite dilution ( ΔH 1∞ ) in water and

the molar enthalpy of solution of water at infinite dilution ( ΔH 2∞ ) in AEEA were

obtained using the method proposed by Maham et al (1997). Enthalpies at infinite

dilution for AEEA and water are listed in Table (4.8b).

Figures (4.14), (4.15), and (4.16) show the experimental HE of AEEA + water system

for all three temperatures along with the fitted curves of NRTL, UNIQUAC and UNIFAC

models respectively. The % AAD from the experimental data for the three models are

4.9, 6.3, and 2.3 respectively. Parameters of the three models are listed in Table (4.9a)

and Table (4.9b). Parameters of the three models were regressed simultaneously for the

three temperatures using the built-in algorithm called the Data Regression System (DRS)

of Aspen Plus software. Among the three models, UNIFAC provided the best fitting to

the experimental HE values for this system. Compared to the other studied alkanolamine

+ water systems, AEEA + water system has exhibited higher negative values of HE which

might be considered one of the reasons why NRTL and UNIQUAC models are not well

regressed compared to UNIFAC model for AEEA + water system.

76
Table (4.8a) Redlich-Kister equation fitting coefficients for the molar excess enthalpy
(HE) (equation 4.8), for AEEA (1) + water (2) mixtures for (298.15,
313.15, and 323.15) K

T/K a0 a1 a2 a3 a4 % AAD

298.15 -13843.3 10931.3 -4799.3 -1746.4 -494.1 0.9

313.15 -14920.2 8804.8 -5118.1 2439.7 -- 2.7

323.15 -14368.2 11724.4 -2428.0 -5773.8 -- 3.9

Table (4.8b) Molar enthalpies of AEEA at infinite dilution in water, ( ΔH 1∞ ), and of


water at infinite dilution in AEEA, ( ΔH 2∞ ), for (298.15, 313.15, and
323.15) K

T/K ΔH 1∞ /kJ·mole-1 ΔH 2∞ /kJ·mole-1

298.15 -28 -10

313.15 -31 -9

323.15 -23 -11

77
0

HE /(J mole-1) -1000

-2000

-3000

298.15 K
-4000
313.15 K
323.15 K
NRTL
-5000
0.0 0.2 0.4 0.6 0.8 1.0
x1
Figure (4.14) Molar excess enthalpy of AEEA (1) + water (2) represented by the NRTL
Model

-1000
HE /(J mole-1)

-2000

-3000

298.15 K
-4000
313.15 K
323.15 K
UNIQUAC
-5000
0.0 0.2 0.4 0.6 0.8 1.0
x1
Figure (4.15) Molar excess enthalpy of AEEA (1) + water (2) represented by the
UNIQUAC Model

78
0

-1000
HE /(J mole-1)

-2000

-3000

298.15 K
-4000
313.15 K
323.15 K
UNIFAC
-5000
0.0 0.2 0.4 0.6 0.8 1.0
x1
Figure (4.16) Molar excess enthalpy of AEEA (1) + water (2) represented by the
UNIFAC Model

79
Table (4.9a) NRTL and UNIQUAC equations parameters for the molar excess enthalpy
data of AEEA (1) + water (2) for (298.15, 313.15, and 323.15) K

NRTL UNIQUAC
Parameters values % AAD Parameters values % AAD
a12 0.405 a12 -2.738
a21 -1.736 a21 -0.455
6.3
b12 (K) -1305.53 4.9 b12 (K) 808.866
b21 (K) 252.226 b21 (K) 223.052
α 0.150

Table (4.9b) Modified UNIFAC (Dortmund) equation parameters for the molar excess
enthalpy data of AEEA (1) + water (2) for (298.15, 313.15, and 323.15) K

Groups (m)
CH2NH2 CH2NH CH2 H2O OH % AAD
(n)
anm/K 0.0000 -13653 313.37 -817.38 -75.630
CH2NH2 bnm 0.0000 -19.497 -2.7845 7.5325 -0.1511
-1
cnm/K 0.00000 0.17394 0.00815 -0.00414 0.00000
anm/K 18.080 0.0000 -1311.0 -988.51 -660.20
CH2NH bnm 10.114 0.0000 -1.5334 0.7069 1.7430
-1
cnm/K -0.01681 0.00000 0.02552 0.00429 0.00000
anm/K 1476.4 979.86 0.0000 2334.5 2777.0
CH2 bnm 2.2804 0.7340 0.0000 -3.6512 -4.6740 2.3
-1
cnm/K -0.02427 0.00427 0.00000 -0.00647 0.00155
anm/K 1283.5 10041 153.49 0.0000 1460.0
H2O bnm -5.7249 -8.9292 2.2779 0.0000 -8.6730
-1
cnm/K 0.00535 -0.06117 -0.00812 0.00000 0.01641
anm/K -923.70 -355.10 1606.0 -801.90 0.0000
OH bnm 2.4680 0.5800 -4.7460 3.8240 0.0000
-1
cnm/K 0.00000 0.00000 0.00092 -0.00751 0.00000

80
4.4.2 3-Amino-1-Propanol (AP)

Molar excess enthalpies for (AP (1) + water (2)) were measured experimentally at

three different temperatures (298.15, 313.15, and 323.15) K. The experimental values are

listed in Table (A12). Figure (4.17) shows the mole fractions dependency of the

experimental molar excess enthalpies at various temperatures. At all temperatures the

molar excess enthalpy curves are negative with a minimum around x1 = 0.4. Negative

molar excess enthalpies indicate an exothermic mixing process between AP and water.

The exothermic mixing process between AP and water is due to the strong interaction of

the –OH group of water molecules with amine (–NH2) group of AP molecules.

Figure (4.17) shows the molar excess enthalpies become less negative at lower

temperatures until x1 = 0.6. After x1 = 0.6 the values become less negative at higher

temperatures. The discussions of effects of the groups present in AP on molar excess

enthalpy are given in the later section of this chapter.

A Redlich-Kister relation was used to correlate the molar excess enthalpy data:

n
H E / J ⋅ mol -1 = x1 x 2 ∑ ai ( x1 − x 2 )
i
(4.9)
i =0

The coefficients and the standard deviations are presented in Table (4.10a). The

calculated molar excess enthalpies by equation (4.9) are plotted in Figure (4.17) as solid

lines for the three temperatures. The percentage average absolute deviations are 1.7, 1.9,

and 2.3 for the temperatures 298.15 K, 313.15 K and 323.15 K respectively. Both the

molar enthalpy of solution of AP at infinite dilution ( ΔH 1∞ ) in water and the molar

enthalpy of solution of water at infinite dilution ( ΔH 2∞ ) in AP were obtained using the

method proposed by Maham et al (1997), and are listed in Table (4.10b). Figure (4.18),

81
0

-500

-1000
HE /(J mole-1)

-1500

-2000

298.15 K
-2500 313.15 K
323.15 K
RK
-3000
0.0 0.2 0.4 0.6 0.8 1.0
x1
Figure (4.17) Molar excess enthalpies for AP (1) + water (2) mixtures

82
Table (4.10a) Redlich-Kister equation fitting coefficients for the molar excess enthalpy
(HE) (equation (4.9)), for AP (1) + water (2) mixtures for (298.15, 313.15,
and 323.15) K

T/K a0 a1 a2 a3 a4 % AAD

298.15 -9900.9 5380.2 -1555.2 -2449.5 -- 1.7

313.15 -9743.6 4953.8 -2613.9 -239.9 -- 1.9

323.15 -10396.5 4248.9 4940.6 2618.9 -12317.9 2.3

Table (4.10b) Molar enthalpies of AP at infinite dilution in water, ( ΔH 1∞ ), and of water


at infinite dilution in AP, ( ΔH 2∞ ), for (298.15, 313.15, and 323.15) K

T/K ΔH 1∞ /kJ·mole-1 ΔH 2∞ /kJ·mole-1

298.15 -14 -9

313.15 -17 -8

323.15 -25 -11

83
Figure (4.19) and Figure (4.20) show the experimental values of molar excess enthalpy of

AP + water system for all three temperatures along with the fitted curves of NRTL,

UNIQUAC and UNIFAC models respectively. The percentage average absolute

deviations (% AAD) from the experimental data for all three models are 3.9, 4.9, and 1.8

respectively. Parameters of the three models are listed in Table (4.11a) and Table (4.11b).

As mentioned in the previous section, parameters of the three models were regressed

simultaneously for the three temperatures using the built-in algorithm called the Data

Regression System (DRS) of Aspen Plus software. Among the three models, modified

UNIFAC provided the best fitting to the experimental molar excess enthalpy values of

AP + water system.

84
0

-500

-1000
HE /(J mole-1)

-1500

-2000

298.15 K
-2500 313.15 K
323.15 K
NRTL
-3000
0.0 0.2 0.4 0.6 0.8 1.0
x1
Figure (4.18) Molar excess enthalpy of AP (1) + water (2) represented by the NRTL
Model

-500

-1000
HE /(J mole-1)

-1500

-2000

298.15 K
-2500 313.15 K
323.15 K
UNIQUAC
-3000
0.0 0.2 0.4 0.6 0.8 1.0
x1
Figure (4.19) Molar excess enthalpy of AP (1) + water (2) represented by the
UNIQUAC Model

85
0

-500

-1000
HE /(J mole-1)

-1500

-2000

298.15 K
-2500 313.15 K
323.15 K
UNIFAC
-3000
0.0 0.2 0.4 0.6 0.8 1.0
x1
Figure (4.20) Molar excess enthalpy of AP (1) + water (2) represented by the UNIFAC
Model

86
Table (4.11a) NRTL and UNIQUAC equations’ parameters for the molar excess
enthalpy data of AP (1) + water (2) for (298.15, 313.15, and 323.15) K

NRTL UNIQUAC
Parameters values % AAD Parameters values % AAD
a12 -38.653 a12 -0.383
a21 17.681 a21 -1.183
3.9 4.9
b12 (K) -234.450 b12 (K) 201.667
b21 (K) -1109.81 b21 (K) 466.078
α 0.021

Table (4.11b) Modified UNIFAC (Dortmund) equations’ interaction parameters for the
molar excess enthalpy data of AP (1) + water (2) for (298.15, 313.15, and
323.15) K

Groups (m)
CH2 CH2NH2 H2O OH % AAD
(n)
anm/K 0.0000 2468.4 1519.4 2777.0
CH2 bnm 0.0000 19.247 -3.4654 -4.6740
cnm/K-1 0.00000 -0.08383 0.00111 0.00155
anm/K -466.48 0.0000 -935.70 -75.630
CH2NH2
bnm -3.2350 0.0000 4.0101 -0.1511
cnm/K-1 0.01949 0.00000 -0.00655 0.00000
1.8
anm/K -67.870 1997.5 0.0000 1460.0
H2O bnm -1.2059 -3.7534 0.0000 -8.6730
cnm/K-1 0.00756 0.00608 0.00000 0.01641
anm/K 1606.0 -923.70 -801.90 0.0000
OH bnm -4.7460 2.4680 3.8240 0.0000
cnm/K-1 0.00092 0.00000 -0.00751 0.00000

87
4.4.3 2-(Methylamino)ethanol (MAE)

Molar excess enthalpies for (MAE (1) + water (2)) were measured experimentally at

three different temperatures (298.15, 313.15, and 323.15) K. The experimental values are

listed in Table (A13). Figure (4.21) shows the mole fractions dependency of the

experimental molar excess enthalpies at various temperatures. At all temperatures the

molar excess enthalpy curves are negative with a minimum around x1 = 0.35. Negative

molar excess enthalpies indicate the exothermic mixing process between MAE and water

due to the strong interaction of the –OH group of water molecules with amine (–NH)

group of MAE molecules. Figure (4.21) shows, in the water rich region, the molar excess

enthalpies become slightly more negative at the higher temperatures. In the MAE rich

region there is no considerable difference between the values at higher and lower

temperatures except for a couple of values at 313.15 K for the x1 = 0.5833 and x1 =

0.6860, where molar excess enthalpies become higher than the values at 298.15 K and

323.15 K. Figure (4.22) shows the comparison of molar excess enthalpies measured in

this study at 298.15 K with that measured by Touhara et al. (1981). The measured values

are in agreement.

A Redlich-Kister relation was employed to correlate the molar excess enthalpies.

n
H E / J ⋅ mol -1 = x1 x 2 ∑ ai ( x1 − x 2 )
i
(4.10)
i =0

The coefficients and the standard deviations are presented in Table (4.12a). The

calculated molar excess enthalpies, by using equation (4.10), are plotted in Figure (4.21)

as solid lines for all three temperatures. The percentage average absolute deviations (%

AAD) are 1.9, 1.7, and 0.5 for the temperatures 298.15 K, 313.15 K and 323.15 K

respectively.

88
0

-500

-1000
HE /(J mole-1)

-1500

-2000

298.15 K
-2500 313.15 K
323.15 K
RK
-3000
0.0 0.2 0.4 0.6 0.8 1.0
x1
Figure (4.21) Molar excess enthalpies for MAE (1) + water (2) mixtures

-500

-1000
HE/J mole-1

-1500

-2000

-2500
Present study
Touhara et al. (1982)
-3000
0.0 0.2 0.4 0.6 0.8 1.0
x1
Figure (4.22) Molar excess enthalpies for MAE (1) + water (2) mixtures at 298.15 K

89
As displayed in Figure (4.21), the minimum value is shifted towards higher concentration

of MAE for higher temperatures. At 323.15 K, the minimum value of molar excess

enthalpy occurs at x1 = 0.4.

Both the molar enthalpy of solution of MAE at infinite dilution ( ΔH 1∞ ) in water and

the molar enthalpy of solution of water at infinite dilution ( ΔH 2∞ ) in MAE were obtained

using the method proposed by Maham et al. (1997), and are listed in Table (4.12c). Also

Table (4.12b) and Table (4.12c) show the comparison of both the molar enthalpy of

solution of MAE at infinite dilution ( ΔH 1∞ ) in water and the molar enthalpy of solution of

water at infinite dilution ( ΔH 2∞ ) in MAE at 298.15 K with the values obtained by

Touhara et al. (1981), and Maham et al. (1997) respectively.

Figure (4.23), Figure (4.24), and Figure (4.25) show the experimental values of molar

excess enthalpy of MAE + water system for all three temperatures along with the fitted

curves of NRTL, UNIQUAC and UNIFAC models respectively. The percentage average

absolute deviations (% AAD) from the experimental data for all three models are 3.5, 4.2,

and 2.1 respectively. Parameters of the three models are listed in Table (4.13a) and Table

(4.13b). Parameters of the three models were regressed simultaneously for the three

temperatures using the built-in algorithm called the Data Regression System (DRS) in

Aspen Plus software. Among the three models, UNIFAC (Dortmund) provided the best

fitting to the experimental molar excess enthalpy values of MAE + water system.

90
Table (4.12a) Redlich-Kister equation fitting coefficients for the molar excess enthalpy
(HE) (equation (4.10)), for MAE (1) + water (2) mixtures for (298.15,
313.15, and 323.15) K

%
T/K a0 a1 a2 a3 a4
AAD
298.15 -9993.0 5291.6 -2972.1 1117.1 -- 1.9

313.15 -9939.4 4206.3 -5784.4 3188.3 -- 1.7

323.15 -10434.5 4010.5 6043.0 4465.1 -16348.3 2.4

Table (4.12b) Molar enthalpies of MAE at infinite dilution in water, ( ΔH 1∞ ), and of


water at infinite dilution in MAE, ( ΔH 2∞ ), for 298.15 K

This Study Touhara et al. (1981)


T/K
ΔH 1∞ /kJ·mole-1 ΔH 2∞ /kJ·mole-1 ΔH 1∞ /kJ·mole-1 ΔH 2∞ /kJ·mole-1

298.15 -19 -6 -18.5 -6.2

Table (4.12c) Molar enthalpies of MAE at infinite dilution in water, ( ΔH 1∞ ), and of


water at infinite dilution in MAE, ( ΔH 2∞ ), for (298.15, 313.15, and
323.15) K

This Study Maham et al. (1997)


T/K
ΔH 1∞ /kJ·mole-1 ΔH 2∞ /kJ·mole-1 ΔH 1∞ /kJ·mole-1 ΔH 2∞ /kJ·mole-1

298.15 -19 -6 -18.4 -5.9

313.15 -19 -5 -- --

323.15 -29 -12 -- --

91
0

-500

-1000
HE /(J mole-1)

-1500

-2000

298.15 K
-2500 313.15 K
323.15 K
NRTL
-3000
0.0 0.2 0.4 0.6 0.8 1.0
x1
Figure (4.23) Molar excess enthalpy of MAE (1) + water (2) represented by the NRTL
Model

-500

-1000
HE /(J mole-1)

-1500

-2000

298.15 K
-2500 313.15 K
323.15 K
UNIQUAC
-3000
0.0 0.2 0.4 0.6 0.8 1.0
x1
Figure (4.24) Molar excess enthalpy of MAE (1) + water (2) represented by the
UNIQUAC Model

92
0

-500

-1000
HE /(J mole-1)

-1500

-2000

298.15 K
-2500 313.15 K
323.15 K
UNIFAC
-3000
0.0 0.2 0.4 0.6 0.8 1.0
x1

Figure (4.25) Molar excess enthalpy of MAE (1) + water (2) represented by the
UNIFAC Model

93
Table (4.13a) NRTL and UNIQUAC equations parameters for the molar excess enthalpy
data of MAE (1) + water (2) for (298.15, 313.15 and 323.15) K

NRTL UNIQUAC
Parameters Values % AAD Parameters values % AAD
a12 -67.695 a12 -0.092
a21 20.123 a21 -1.040
4.2
b12 (K) -159.767 3.5 b12 (K) 246.220
b21 (K) -1265.64 b21 (K) 418.965
α 0.019

Table (4.13b) Modified UNIFAC (Dortmund) equations’ interaction parameters for the
molar excess enthalpy data of MAE (1) + water (2) for (298.15, 313.15
and 323.15) K

Groups (m)
CH3NH CH2 H2O OH % AAD
(n)
anm/K 0.0000 -189.01 1154.1 -660.20
CH3NH
bnm 0.0000 -2.6535 8.1353 1.7430
-1
cnm/K 0.00000 0.08071 -0.02946 0.00000
anm/K 3856.4 0.0000 585.75 2777.0
CH2
bnm 0.7340 0.0000 -6.8013 -4.6740
-1
cnm/K -0.04163 0.00000 0.01258 0.00155
2.1
anm/K 5497.7 -189.78 0.0000 1460.0
H2O bnm -17.132 0.8523 0.0000 -8.6730
-1
cnm/K 0.01081 0.00992 0.00000 0.01641
anm/K -355.10 1606.0 -801.90 0.0000
OH bnm 0.5800 -4.7460 3.8240 0.0000
-1
cnm/K 0.00000 0.00092 -0.00751 0.00000

94
4.4.4 1-Amino-2-Propanol or Monoisopropanolamine (MIPA)

Excess enthalpies for (MIPA (1) + water (2)) were measured experimentally at three

different temperatures (298.15, 313.15, and 323.15) K. The experimental values are listed

in Table (A14). Figure (4.26) shows the mole fractions dependency of the experimental

molar excess enthalpies at various temperatures. At all temperatures the molar excess

enthalpy curves are negative with a minimum around x1 = 0.35. Negative molar excess

enthalpy values indicate the exothermic mixing process between MIPA and water. The

exothermic mixing process between MIPA and water is due to the strong interaction of

the –OH group of water molecules with –NH2 group of the MIPA molecules. The MIPA

+ water system exhibited the lowest negative values of molar excess enthalpy among the

five studied alkanolamine + water systems. Figure (4.26) shows no considerable

difference in the values of excess enthalpies at higher and lower temperatures. As

displayed in Figure (4.26), the minimum value is shifted towards higher concentration of

MIPA for higher temperatures. At 323.15 K, the minimum value of molar excess

enthalpy occur at x1 = 0.4.

A Redlich-Kister relation was used to correlate the molar excess enthalpy data:

n
H E / J ⋅ mol -1 = x1 x 2 ∑ ai ( x1 − x 2 )
i
(4.11)
i =0

The coefficients and the standard deviations are presented in Table (4.14a). The

calculated values of molar excess enthalpy, by using equation (4.11), are plotted in Figure

(4.26) as solid lines for all three temperatures. The percentage average absolute

deviations (% AAD) are 1.7, 3.5, and 1.0 for the temperatures 298.15 K, 313.15 K and

323.15 K respectively.

95
0

-500

-1000
HE /(J mole-1)

-1500

-2000

298.15 K
-2500 313.15 K
323.15 K
RK
-3000
0.0 0.2 0.4 0.6 0.8 1.0
x1

Figure (4.26) Molar excess enthalpies for MIPA (1) + water (2) mixtures

96
Table (4.14a) Redlich-Kister equation fitting coefficients for the molar excess enthalpy
(HE) (equation 4.11), for MIPA (1) + water (2) mixtures for (298.15,
313.15, and 323.15) K

T/K a0 a1 a2 a3 a4 a5 % AAD

298.15 -9158.3 4306.8 -2723.1 1382.7 -- -- 1.7

313.15 -9176.0 3640.8 -2921.1 3411.9 -- -- 3.5

323.15 -9318.7 3907.3 -160.6 1409.3 -3175.2 -760.0 1.0

Table (4.14b) Molar enthalpies of MIPA at infinite dilution in water, ( ΔH 1∞ ), and of


water at infinite dilution in MIPA, ( ΔH 2∞ ), for (298.15, 313.15, and
323.15) K

T/K ΔH 1∞ /kJ·mole-1 ΔH 2∞ /kJ·mole-1

298.15 -18 -6

313.15 -19 -5

323.15 -18 -7

97
Both the molar enthalpy of solution of MIPA at infinite dilution ( ΔH 1∞ ) in water and the

molar enthalpy of solution of water at infinite dilution ( ΔH 2∞ ) in MIPA were obtained

using the method proposed by Maham et al (1997), and are listed in Table (4.14b).

Figure (4.27), Figure (4.28), and Figure (4.29) show the experimental values of molar

excess enthalpy of MIPA + water system for all three temperatures along with the fitted

curves of NRTL, UNIQUAC and UNIFAC models respectively. The percentage average

absolute deviations (% AAD) from the experimental data for the three models are 2.0,

2.3, and 1.4 respectively. Parameters of the three models are listed in Table (4.15a) and

Table (4.15b). Parameters of the three models were regressed simultaneously for the three

temperatures by using the built-in algorithm called the Data Regression System (DRS) of

Aspen Plus software. All three models have provided the best fit to the experimental

molar excess enthalpy values of MIPA + water system with the modified UNIFAC being

the best fit among them.

98
0

-500

-1000
HE /(J mole-1)

-1500

-2000

298.15 K
-2500 313.15 K
323.15 K
NRTL
-3000
0.0 0.2 0.4 0.6 0.8 1.0
x1
Figure (4.27) Molar excess enthalpy of MIPA (1) + water (2) represented by the NRTL
Model

-500

-1000
HE /(J mole-1)

-1500

-2000

298.15 K
-2500 313.15 K
323.15 K
UNIQUAC
-3000
0.0 0.2 0.4 0.6 0.8 1.0
x1
Figure (4.28) Molar excess enthalpy of MIPA (1) + water (2) represented by the
UNIQUAC Model

99
0

-500

-1000
HE /(J mole-1)

-1500

-2000

298.15 K
-2500 313.15 K
323.15 K
UNIFAC
-3000
0.0 0.2 0.4 0.6 0.8 1.0
x1
Figure (4.29) Molar excess enthalpy of MIPA (1) + water (2) represented by the
UNIFAC Model

100
Table (4.15a) NRTL and UNIQUAC equations parameters for the molar excess enthalpy
data of MIPA (1) + water (2) for (298.15, 313.15, and 323.15) K

NRTL UNIQUAC
Parameters values % AAD Parameters Values % AAD
a12 -2.4170 a12 -1.017
a21 -36.662 a21 -0.1920
2.3
b12 (K) -683.459 2.0 b12 (K) 401.318
b21 (K) -9.01700 b21 (K) 221.013
α 0.136

Table (4.15b) Modified UNIFAC (Dortmund) equations’ interaction parameters for the
molar excess enthalpy data of MIPA (1) + water (2) for (298.15, 313.15,
and 323.15) K

Groups (m)
CH3 CH2NH2 H2O CH OH % AAD
(n)
anm/K 0.0000 1476.4 2927.2 0.0000 2777.0
CH3 bnm 0.0000 29.469 -6.9502 0.0000 -4.6740
cnm/K-1 0.00000 -0.11277 -0.00844 0.00000 0.00155
anm/K 598.39 0.0000 -1875.6 326.04 -75.630
CH2NH2 bnm 0.1121 0.0000 1.8385 -2.6348 -0.1511
cnm/K-1 0.03542 0.00000 0.01085 0.00336 0.00000
anm/K 155.28 6550.7 0.0000 -17.250 1460.0
H2O bnm 2.9069 -18.729 0.0000 0.8389 -8.6730 1.4
cnm/K-1 0.00562 -0.00790 0.00000 0.00090 0.01641
anm/K 0.0000 -164.04 1391.3 0.0000 2777.0
CH bnm 0.0000 4.9683 -3.6156 0.0000 -4.6740
cnm/K-1 0.00000 -0.01025 0.00114 0.00000 0.00155
anm/K 1606.0 -923.70 -801.90 1606.0 0.0000
OH bnm -4.7460 2.4680 3.8240 -4.7460 0.0000
cnm/K-1 0.00092 0.00000 -0.00751 0.00092 0.00000

101
4.4.5 N-N-Dimethylethanolamine (DMEA)

Molar excess enthalpies for (DMEA (1) + water (2)) were measured experimentally at

three different temperatures (298.15, 313.15, and 323.15) K. The experimental values are

listed in Table (A15). Figure (4.30) shows the mole fractions dependency of the

experimental molar excess enthalpies at various temperatures. At all temperatures the

molar excess enthalpy curves are negative with a minimum around x1 = 0.35. Negative

molar excess enthalpies indicate the exothermic mixing process between DMEA and

water due to the strong interaction of the –OH group of water molecules with –N group

of DMEA molecules. The DMEA + water system exhibited the second highest negative

values of molar excess enthalpy after AEEA + water system, among the five studied

alkanolamine + water systems. Figure (4.30) shows the molar excess enthalpies become

less negative as temperature increases.

A Redlich-Kister relation was used to correlate the molar excess enthalpy data:

n
H E / J ⋅ mol -1 = x1 x 2 ∑ ai ( x1 − x 2 )
i
(4.12)
i =0

The coefficients and the standard deviations are presented in Table (4.16a). The

calculated values, by using equation (4.12), are plotted in Figure (4.30) as solid lines for

all three temperatures. The percentage average absolute deviations (% AAD) are 0.8, 0.6,

and 3.4 for the temperatures 298.15 K, 313.15 K and 323.15 K respectively. Both the

molar enthalpy of solution of DMEA at infinite dilution ( ΔH 1∞ ) in water and the molar

enthalpy of solution of water at infinite dilution ( ΔH 2∞ ) in DMEA were obtained using

the method proposed by Maham et al. (1997), and are listed in Table (4.16b).

102
0

-500

-1000
HE /(J mole-1)

-1500

-2000

-2500
298.15 K
313.15 K
-3000
323.15 K
RK
-3500
0.0 0.2 0.4 0.6 0.8 1.0
x1
Figure (4.30) Molar excess enthalpies for DMEA (1) + water (2) mixtures

-500

-1000
HE /J.mole-1

-1500

-2000

-2500

-3000
Present study
Touhara et al., (1982)
-3500
0.0 0.2 0.4 0.6 0.8 1.0
x1
Figure (4.31) Molar excess enthalpies for DMEA (1) + water (2) mixtures at 298.15 K

103
Table (4.16a) Redlich-Kister equation fitting coefficients for the excess enthalpy (HE)
(equation 4.12), for DMEA (1) + water (2) mixtures for (298.15, 313.15,
and 323.15) K

T/K a0 a1 a2 a3 a4 a5 % AAD

298.15 -10779.8 5311.3 -1352.6 5657.0 -5167.8 -1793.3 0.8

313.15 -10230.0 6368.7 -1088.4 -1466.6 -5294.4 4615.4 0.6

323.15 -9603.1 4573.2 -4515.4 4998.2 -- -- 3.4

Table (4.16b) Molar enthalpies of DMEA at infinite dilution in water, ( ΔH 1∞ ), and of


water at infinite dilution in DMEA, ( ΔH 2∞ ), for 298.15 K

This Study Touhara et al. (1981)


T/K
ΔH 1∞ /kJ·mole-1 ΔH 2∞ /kJ·mole-1 ΔH 1∞ /kJ·mole-1 ΔH 2∞ /kJ·mole-1

298.15 -26 -8 -25.4 -6.3

Table (4.16c) Molar enthalpies of DMEA at infinite dilution in water, ( ΔH 1∞ ), and of


water at infinite dilution in DMEA, ( ΔH 2∞ ), for (298.15, 313.15, and
323.15) K

This Study Maham et al. (1997)


T/K
ΔH 1∞ /kJ·mole-1 ΔH 2∞ /kJ·mole-1 ΔH 1∞ /kJ·mole-1 ΔH 2∞ /kJ·mole-1

298.15 -26 -8 -24.1 -6.3

313.15 -26 -7 -- --

323.15 -24 -4 -- --

104
Figure (4.31) shows the comparison of molar excess enthalpies measured in this

study at 298.15 K with those measured by Touhara et al. (1981), values are in agreement.

Table (4.16b) and Table (4.16c) show the comparison of both the molar enthalpy of

solution of DMEA at infinite dilution ( ΔH 1∞ ) in water and the molar enthalpy of solution

of water at infinite dilution ( ΔH 2∞ ) in DMEA at 298.15 K with the values obtained by

Touhara et al. (1982) and Maham et al. (1997) respectively.

Figure (4.32), Figure (4.33), and Figure (4.34) show the experimental molar excess

enthalpies of DMEA + water system for all three temperatures along with the fitted

curves of NRTL, UNIQUAC and UNIFAC models respectively. The percentage average

absolute deviations (% AAD) from the experimental data for the three models are 1.8,

2.0, and 1.4 respectively. The parameters of the three models are listed in Table (4.17a)

and Table (4.17b). The parameters of the three models were regressed simultaneously for

the three temperatures using the built-in algorithm called the Data Regression System

(DRS) in Aspen Plus software. All three models have provided good fitting to the

experimental molar excess enthalpy values of DMEA + water system and with modified

UNIFAC the fit was best.

105
0

-500

-1000
HE /(J mole-1)

-1500

-2000

-2500
298.15 K
313.15 K
-3000
323.15 K
NRTL
-3500
0.0 0.2 0.4 0.6 0.8 1.0
x1
Figure (4.32) Molar excess enthalpy of DMEA (1) + water (2) represented by the NRTL
Model

-500

-1000
HE /(J mole-1)

-1500

-2000

-2500
298.15 K
313.15 K
-3000
323.15 K
UNIQUAC
-3500
0.0 0.2 0.4 0.6 0.8 1.0
x1
Figure (4.33) Molar excess enthalpy of DMEA (1) + water (2) represented by the
UNIQUAC Model

106
0

-500

-1000
HE /(J mole-1)

-1500

-2000

-2500
298.15 K
313.15 K
-3000
323.15 K
UNIFAC
-3500
0.0 0.2 0.4 0.6 0.8 1.0
x1
Figure (4.34) Molar excess enthalpy of DMEA (1) + water (2) represented by the
UNIFAC Model

107
Table (4.17a) NRTL and UNIQUAC equations parameters for the molar excess enthalpy
data of DMEA (1) + water (2) for (298.15, 313.15, and 323.15) K

NRTL UNIQUAC
Parameters values % AAD Parameters values % AAD
a12 0.131 a12 0.401
a21 2.021 a21 -1.925
1.8 2.0
b12 (K) -393.889 b12 (K) 127.816
b21 (K) -567.127 b21 (K) 613.564
α 0.961

Table (4.17b) Modified UNIFAC (Dortmund) equations’ interaction parameters for the
molar excess enthalpy data of DMEA (1) + water (2) for (298.15, 313.15,
and 323.15) K

Groups (m)
CH3N CH2 H2O CH3 OH % AAD
(n)
anm/K 0.0000 168.75 81.040 205.65 1876.0
CH3N bnm 0.0000 1.0201 -4.7503 -1.4436 11.500
-1
cnm/K 0.00000 0.03917 0.01136 0.00000 0.09000
anm/K -223.35 0.0000 2221.3 0.0000 2777.0
CH2 bnm 0.8103 0.0000 -1.2299 0.0000 -4.6740
-1
cnm/K -0.00206 0.00000 0.00803 0.00000 0.00155
anm/K 745.87 -189.78 0.0000 -17.250 1460.0
H2O bnm -3.4661 2.9399 0.0000 0.8389 -8.6730 1.4
-1
cnm/K -0.00012 0.00055 0.00000 0.00090 0.01641
anm/K -175.70 0.0000 1391.3 0.0000 2777.0
CH3 bnm 1.8570 0.0000 -3.6156 0.0000 -4.6740
-1
cnm/K 0.00000 0.00000 0.00114 0.00000 0.00155
anm/K 104.60 1606.0 -801.90 1606.0 0.0000
OH bnm -5.0140 -4.7460 3.8240 -4.7460 0.0000
-1
cnm/K 0.00885 0.00092 -0.00751 0.00092 0.00000

108
4.5 Comparison of HE values of alkanolamines

Figures (4.35), (4.36) and (4.37) show the comparison of molar excess enthalpies of

the five studied alkanolamines at 298.15 K, 313.15 K and 323.15 K respectively. Among

the five alkanolamines, AEEA + water system has exhibited the highest values of molar

excess enthalpy at 298.15 K on the negative side followed by DMEA + water, MAE +

water, AP + water and MIPA + water. At 313.15 K, AEEA + water system has also

exhibited the highest negative molar excess enthalpy values among the five studied

alkanolamines. As shown in Figure (4.36), MAE has exhibited slightly higher values of

molar excess enthalpies compared to DMEA in the amine rich region. Also at 313.15 K,

MIPA has exhibited slightly higher negative values of the molar excess enthalpies

compared to AP in the water rich region. As in the case of 323.15 K, AEEA + water

(Figure (4.37)) has exhibited the highest negative values of the molar excess enthalpies

and MAE has exhibited slightly higher negative values of molar excess enthalpies

compared to DMEA in both amine rich and water rich regions. At 323.15 K, AP has also

exhibited greater molar excess enthalpy values compared to MAE and DMEA in amine

rich region. An interesting observation in this study was the values of molar excess

enthalpy also depend upon the position of the –OH and -NH2 groups in the molecule of

alkanolamine, as in the case of AP and MIPA. According to Maham et al. (1997), –CH3

group contributes to the higher molar excess enthalpies but in this study we have

observed that MIPA, which contains one –CH3 group, has exhibited lower values of

molar excess enthalpies compared to AP. The reason could be the position of the –OH

group on the second carbon atom of the MIPA molecule, which has contributed to the

109
lower molar excess enthalpies on the negative side, compared to the contributions of the

position of the –OH group in AP on first carbon atom.

110
0

-1000

-2000
HE /J.mole-1

-3000

AEEA
AP
-4000
MAE
MIPA
DMEA
-5000
0.0 0.2 0.4 0.6 0.8 1.0
x1
Figure (4.35) Comparison of the molar excess enthalpies of the five studied
alkanolamines at 298.15 K

-1000

-2000
HE /J.mole-1

-3000

AEEA
AP
-4000
MAE
MIPA
DMEA
-5000
0.0 0.2 0.4 0.6 0.8 1.0
x1
Figure (4.36) Comparison of the molar excess enthalpies of the five studied
alkanolamines at 313.15 K

111
0

-1000

-2000
HE /J.mole-1

-3000

AEEA
AP
-4000
MAE
MIPA
DMEA
-5000
0.0 0.2 0.4 0.6 0.8 1.0
x1
Figure (4.37) Comparison of the molar excess enthalpies of the five studied
alkanolamines at 323.15 K

112
5. CONCLUSION AND RECOMMENDATIONS

5.1 Conclusion:

The molar heat capacities of the aqueous AEEA, AP, MAE, MIPA and DMEA

solutions were measured over a range of temperatures from (303.15 to 353.15) K and

over the entire range of mole fractions. The molar excess heat capacity ( C pE ) values for

the aqueous AEEA, AP, MIPA and DMEA solutions were positive at all temperatures

and mole fractions. At 303.15 K for x1 = 0.9, aqueous MAE solution has exhibited

negative value of the molar excess heat capacity ( C pE ), while for all other temperatures

and mole fractions, the molar excess heat capacity ( C pE ) values of aqueous MAE

solutions were positive. The molar excess heat capacities were correlated as a function of

mole fractions employing Redlich-Kister equation. Among the five studied

alkanolamines, AEEA has exhibited the highest values of the molar heat capacity and

MAE has exhibited the lowest values of the molar heat capacity. This study confirms the

observation made by Maham et al. (1997) that the values of molar heat capacity of the

alkanolamines are dominated by –CH2 and –OH group contributions and these

contributions increase with increasing temperature.

The molar excess enthalpies of the AEEA, AP, MAE, MIPA, and DMEA in water

were measured at three different temperatures (298.15, 313.15 and 323.15) K and over

the entire range of mole fractions. Molar enthalpies at infinite dilution were calculated for

the selected five alkanolamine (1) + water (2) systems for all three temperatures (298.15,

313.15 and 323.15) K.

113
The molar excess enthalpies were correlated as a function of mole fractions

employing the Redlich-Kister equation. The experimental molar excess enthalpy values

were modeled successfully by using the solution theory models: NRTL, and UNIQUAC

models and by using the group contribution model: Modified UNIFAC (Dortmund)

model. Among the three models, Modified UNIFAC (Dortmund) has exhibited lowest

percentage average absolute deviations from the experimental molar excess enthalpies,

followed by NRTL and UNIQUAC.

Among five studied alkanolamines, AEEA + water system has exhibited the highest

values of molar excess enthalpy in the negative side followed by DMEA + water, MAE +

water, AP + water, and MIPA + water system. This study confirms the conclusion made

by Maham et al. (1997) that the interaction between water (–OH group) and diethylamine

(–NH group) is more dominant than the interactions between –OH groups from water and

alcohol molecules. This study also confirms that –CH3 group has contributed the higher

values of molar excess enthalpy. One important observation made in this study is that the

values of the molar excess enthalpy also depend upon the position of the –OH and -NH2

groups in the molecule of alkanolamines as observed in AP and MIPA.

5.2 Recommendations:

• Measurements of the molar heat capacity and the molar excess enthalpy for the

mixed (chemical and physical) solvents at various temperatures

• Measurements of the molar heat capacity and the molar excess enthalpy for the

CO2 loaded alkanolamines at various temperatures and pressures

114
6. REFERENCES

Abrams, D. S.; Prausnitz, J. M. Statistical Thermodynamics of Liquid Mixtures: A New


Expression for the Excess Gibbs Energy of Partly or Completely Miscible
Systems. AIChE J. 1975, 21, 116-128.

Ali, S. H.; Merhant, S. Q.; Fahim, M. A. Reaction Kinetics of some Secondary


Alkanolamines with Carbon Dioxide in Aqueous Solutions by Stopped Flow
Technique. Sep. & Purif. Tech. 2002, 27, 121-136.

Alonso, R.; Guerrero, R.; Corrales, J. A. Excess Molar Enthalpies of (Methylcyclohexane


+ an Alkanol) at 323.15 K. I. Results for n-butanol, n-pentanol, and n-hexanol. J.
Chem. Thermodyn. 1987, 19, 1271-1273.

Anderson, T. F.; Prausnitz, J. M. Application of the UNIQUAC Equation to Calculation


of Multicomponent Phase Equilibria. 1. Vapor-Liquid Equilibria. Ind. Eng. Chem.
Proc. Des. Dev. 1978, 17, 552-561.

Astarita, G.; Savage, D.W.; Bisio, A. Gas Treating with Chemical Solvents; John Wiley
& Sons: New York, 1983.

Armitage, D. A.; Morcom, K. W. Thermodynamic Behavior of Mixture of


Hexafluorobenzene with Amines. Part.1 - Enthalpies of Mixing. Trans. Faraday
Soc. 1969, 65, 688-695.

Battler, J. R. Liquid-Liquid Equilibrium Surfaces from Binary Heats of Mixing; Ph.D.


Thesis: Rice University. Houston, Texas, 1984.

Bavbek, O.; Alper, E. Reaction Mechanism and Kinetics of Aqueous Solutions of


Primary and Secondary Alkanolamines and Carbon Dioxide. Turk. J. Chem. 1999,
23, 293-300.

Becker, L.; Aufderhaar, O.; Gmehling, J. Measurement of Heat Capacities for Nine
Organic Substances by Tian-Calvet Calorimetry. J. Chem. Eng. Data 2000, 45,
661-664.

Becker, L.; Gmehling, J. Measurement of Heat Capacities for 12 Organic Substances by


Tian-Calvet Calorimetry. J. Chem. Eng. Data 2001, 46, 1638-1642.

Bevington, P. R. Data Reduction and Error Analysis for the Physical Sciences; McGraw-
Hill: New York, 1969.

Bottoms, R. R. Organic Basis for Gas Purification. Ind. Eng. Chem. 1931, 23, 501-504.

115
Calvet, E.; Part, H.; Skinner, H. A. Recent Progress in Microcalorimetry; Pergamon
Press: New York, 1963.

Camacho, F.; Sanchez, S.; Pacheco, R. Absorption of Carbon Dioxide at High Partial
Pressures in 1-Amino-2-propanol Aqueous Solution. Considerations of Thermal
Effects. Ind. Eng. Chem. Res. 1997, 36, 4358-4364.

Chen, Y-J.; Shih, T-W.; Li, M-H. Heat Capacity of Aqueous Mixtures of
Monoethanolamine with N-Methyldiethanolamine. J. Chem. Eng. Data 2001, 46,
51-55.

Chiu, L-F.; Liu, H-F.; Li, M-H. Heat Capacity of Alkanolamines by Differential
Scanning Calorimetry. J. Chem. Eng. Data 1999, 44, 631-636.

Christensen, J. J.; Hanks, R. W.; Izatt, R.M. Hand book of Heats of Mixing; John Wiley
& Sons: New York, 1982.

Copp, J. L.; Everett, D. H. Excess Molar Enthalpies of Water and Diethylamine Mixtures
at 298.15 K. Discuss. Faraday Soc. 1953, 174-188.

Cruz, J. C.; Renon, H. A New Thermodynamic Representation of Binary Electrolyte


Solution Nonideality in the Whole Range of Concentration. AIChE J. 1978, 24,
817-830.

Davies, J. V.; Pritchard, H. O. The Properties of Diphenyl-Ether Calorimeters. J. Chem.


Thermodyn. 1972, 4, 9-22.

de Koeijer, G.; Solbraa, E. High Pressure Gas Sweetening with Amines for Reducing
CO2 Emission. 7th International Conference on Greenhouse Gas Control
Technologies, Vancouver, Canada, 2004.

Elliot, K.; Wormald, C. J. A Precision Differential Flow Calorimeter. J. Chem.


Thermodyn. 1976, 8, 881-893.

Ewing, M. B.; Marsh, K. N.; Stocks, R. H.; Tuxford, C. W. The Isothermal Displacement
Calorimeter: Design Refinements. J. Chem. Thermodyn. 1970, 2, 751-756.

Fredenslund, A.; Jones, R. L.; Prausnitz, J. M. Group-Contribution Estimation of Activity


Coefficients in Nonideal Liquid Mixtures. AIChE J. 1975, 21, 1086-1099.

Fredenslund, A.; Gmehling, J.; Rasmussen, P. Vapor-Liquid Equilibria Using UNIFAC;


Elsevier; Amsterdam, 1977.

Ginnings, D. C.; Furukawa, G. T. J. Heat Capacity Standard for the Range 14 to 1200 K.
American Chem. Soc. 1953, 75, 522-527.

116
Gmehling, J. G.; Anderson, T. F.; Prausnitz, J. M. Solid-Liquid Equilibria Using
UNIFAC. Ind. Eng. Chem. Fundam. 1978, 17, 269-273.

Gmehling, J.; Li, J.; Schiller, M. A Modified UNIFAC Model. 2. Present Parameter
Matrix and Results for Different Thermodynamic Properties. Ind. Eng. Chem.
Res. 1993, 32, 178-193.

Gmehling, J. G.; Rasmussen, P. Flash Points of Flammable Liquid Mixtures Using


UNIFAC. Ind. Eng. Chem. Fundam. 1982, 21, 186-188.

Goodwin, S. R.; Newsham, D. M. T. A Flow Calorimeter for Determination of Enthalpies


of Mixing of Liquids. J. Chem. Thermodyn. 1971, 3, 325-334.

Gottlieb, M.; Herskowitz, M. Estimation of the X Parameter for Poly(dimethylsiloxane)


Solutions by the UNIFAC Group Contribution Method. Macromolecules 1981,
14, 1468-1471.

Gow, A. S. Calculation of Vapor-Liquid Equilibria from Infinite-Dilution Excess


Enthalpy Data Using the Wilson or NRTL Equation, Ind. Eng. Chem. Res. 1993,
32, 3150-3161.

Gustin, J. L.; Renon, H. Heats of Mixing of Binary Mixtures of N-ethylpyrrolidone,


Ethanolamine, n-Heptane, Cyclohexane, and Benzene by Differential Flow
Calorimetry. J. Chem. Eng. Data 1973, 18, 164-166.

Guggenheim, E. A. Mixtures. Oxford: Oxford University Press, 1952.

Hanks, R. W.; Gupta, A. C.; Christensen, J. J. Calculation of Isothermal Vapor-Liquid


Equilibrium Data for Binary Mixtures from Heats of Mixing, Ind. Eng. Chem.
Fundam. 1971, 10, 504-509.

Hanks, R. W.; Tan, R. L.; Christensen, J. J. Limits of the Simultaneous Correlation of gE


and hE Data by the NRTL, LEMF, and Wilson’s Equations, Thermochim. Acta
1978a, 23, 41-55.

Hanks, R. W.; Tan, R. L.; Christensen, J. J. The Prediction of High Temperature Vapor-
Liquid Equilibria from Lower Temperature Heat of Mixing Data, Thermochim.
Acta 1978b, 27, 9-18.

Harris, D. C. Quantitative Chemical Analysis; W. H. Freeman and Company: New York,


1991.

Harsted, B. S.; Thomsen, E. S. Excess Enthalpies from Flow Microcalorimetry. J. Chem.


Thermodyn. 1974, 6, 549-555.

117
Huang, J. F.; Lee, L. S. A Modification of NRTL Model for Simultaneous Estimation of
Excess Enthalpy, Excess Gibbs Energy and Vapor-Liquid- Equilibrium. Journal
of the Chinese Inst. of Chem. Eng. 1995, 26, 237-252.

Huang, J. F.; Lee, L. S. Simultaneous Estimation of Excess Enthalpy, Excess Gibbs


Energy and Vapor-Liquid Equilibrium Using the Modified Wilson Model. Fluid
Phase Equilibria 1996, 121, 27-43.

Hill, R. J.; Swinton, F. L. The Thermodynamic Properties of Binary Mixtures Containing


Carbon Disulphide. II Excess enthalpies. J. Chem. Thermodyn. 1980, 12, 383-385.

Hill, R. J.; Swinton, F. L. The Excess Enthalpies of Some Mixtures Containing Carbon
Disulphide. J. Chem. Thermodyn. 1980, 12, 489-492.

Jensen, T.; Fredenslund, A.; Rasmussen, P. Pure-Component Vapor Pressures Using


UNIFAC Group Contribution. Ind. Eng. Chem. Fundam. 1981, 20, 239-246.

Kersaw, R. W.; Malcolm, G. N. Thermodynamics of Solutions of Polypropylene Oxide in


Chloroform and in Carbon Tetrachloride. Trans. Faraday Soc. 1968, 64, 323-336.

Kirk-Othmer. Encyclopedia of Chemical Technology; Vol 2, Fourth edition; John Wiley


& Sons: New York, 1992.

Kohl, A.; Nielson, R. Gas Purification; Fifth edition; Gulf Publishing Company: Texas,
1997.

Kvamsdal, H. M.; Maurstad, O.; Jordal, K.; Bolland, O. Benchmark of gas-turbine cycles
with CO2 capture. 7th International Conference on Greenhouse Gas Control
Technologies, Vancouver, Canada, 2004.

Langmuir, I. The Distribution and Orientation of Molecules. Third Colloid Symposium


Monograph, The Chemical Catalog Company, Inc.: New York, 1925.

Larkin, J. A. Thermodynamic Properties of Aqueous Nonelectrolyte Mixture. 1. Excess


Enthalpy for water + ethanol at 298.15 K. J. Chem. Thermodyn. 1975, 7, 137-148.

Larkin, J. A.; McGlashan, M. L. A New Calorimeter for Heat of Mixing. The Heat of
Mixing of Benzene with Carbon Tetrachloride. J. Chem. Soc. 1961, 3425-3432.

Lim, K. H.; Whiting, W. B.; Smith, D. H. Excess Enthalpies and Liquid-Liquid


Equilibrium Phase Compositions of the Nonionic Amphiphile 2-Butoxyethanol
and Water. J. Chem. Eng. Data 1994, 39, 399-403.

Lo, H. S.; Paulaitis, M. E. Estimation of Solvent Effects on Chemical Reaction Rates


Using UNIFAC Group Contribution. AIChE J. 1981, 27, 842-844.

118
Magnussen, T.; Rasmussen, P.; Fredenslund, A. UNIFAC Parameter Table for Prediction
of Liquid-Liquid Equilibriums, Ind. Eng. Chem. Proc. Des. Dev. 1981, 20, 331-
339.

Maham, Y.; Hepler, L. G.; Mather, A. E.; Hakin, A. W.; Marriott, R. A. Molar Heat
Capacities of Alkanolamines from 299.1 to 397.8 K. J. Chem. Soc., Faraday
Trans. 1997, 93, 1747-1750.

Maham, Y.; Mather, A. E.; Hepler, L. G. Excess Molar Enthalpies of (Water +


Alkanolamine) Systems and Some Thermodynamic Calculations. J. Chem. Eng.
Data 1997, 42, 988-992.

Malcolm, G. N.; Rowlinson, J. S. The Thermodynamic Properties of Aqueous Solutions


of Polyethylene Glycol, Polypropylene Glycol and Dioxane. Trans. Faraday Soc.
1957, 53, 921-931.

Ma’mun, S.; Jakosbsen, J. P.; Svendsen, H. F.; Juliussen, O. Experimental and Modeling
Study of the Solubility of Carbon Dioxide in Aqueous 30 mass % 2-
2(aminoethyl)amino)ethanol solution. Ind. Eng. Chem. Res. 2006, 45, 2505-2512.

Mann, W. B. Use of Callendar’s Radio Balance for of the Energy Emission from
Radioactive Sources. J. Res. Natl. Bur. Stand. 1954, 52, 177-184.

Marsh, K. N. Chemical Thermodynamics. (Specialist Periodical Reports), Chemical


Society, London, 1978.

Matsumoto, Y.; Touhara, H.; Nakanishi, K.; Watanabe, N. Molar Excess Enthalpies for
Water + Ethanediol, + 1, 2-propanediol, and + 1, 3- propanediol at 298.15 K. J.
Chem. Thermodyn. 1977, 9, 801-805.

McGlashan, M. L.; Stoeckli, H. F. A Flow Calorimeter for Enthalpies of Mixing. J.


Chem. Thermodyn. 1969, 1, 589-594.

Mimura, T.; Simayoshi, H.; Suda, T.; Iijima, M.; Mituoka, S. Development of Energy
Saving Technology for Flue Gas Carbon Dioxide Recovery in Power Plant by
Chemical Absorption Method and Steam System. Energy Convers. Management
(Suppl.) 1997, 38, S57-S62.

Mrazek, R. V.; Van Ness, H. C. Heats of Mixing: Alcohol-Aromatic Binary Systems at


25 °C, 35 °C and 45 °C. AIChE J. 1961, 7, 190-195.

Nagata, I.; Tamura, K. Thermodynamics of Solutions of Ethanol in Non-Associating


Components. Thermochim. Acta. 1984, 77, 281-297.

Nagata, I.; Tamura, K. Thermodynamics of Solutions of Propanols in Non-Associating


Components. Thermochim. Acta. 1985, 87, 129-140.

119
Nagata, I.; Tamura, K. Excess Enthalpies and Complex Formation of Acetonitrile with
Acetone, Chloroform, and Benzene. Thermochim. Acta. 1981, 47, 315-331.

Nagata, I.; Tamura, K. Correlation and Prediction of Excess Thermodynamic Functions


of Strong Nonideal Liquid Mixtures. Ind. Eng. Chem. Proc. Des. Dev. 1974, 13,
47-53.

Nocon, G.; Weidlich, U.; Gmehling, J.; Menke, J.; Onken, U. Prediction of Gas
Solubilities by a Modified UNIFAC Equation. Fluid Phase Equilibria 1983, 13,
381-392.

Oishi, T.; Prausnitz, J. M. Estimation of Solvent Activities in Polymer Solutions Using a


Group-Contribution Method. Ind. Eng. Chem. Proc. Des. Dev. 1978, 17, 333-339.

Osborne, N. S.; Sitmson, H. F.; Ginnings, D. C. Measurements of Heat Capacity and


Heat of Vaporization of Water in the Range 0 oC to 100 oC. J. Res. Natl. Bur.
Stand. 1939, 23, 197-260.

Perry, R. H.; Green, D. W.; Maloney, J. O. Perry’s Chemical Engineers’ Handbook; 7th
ed., McGraw-Hill, USA, 1999.

Prausnitz, J. M., Molecular Thermodynamics of Fluid Phase Equilibria; 3rd ed.; Prentice
Hall PTR, 1999.

Prausnitz, J. M., Anderson, T. F., Grens, E. A., Eckert, C. A., Hsieh, R., O'Connell, J. P.,
Computer Calculations for Multicomponent Vapor-Liquid and Liquid-Liquid
Equilibria, Prentice-Hall, Englewood Cliffs, NJ, USA, 1980.

Redlich, O.; Kister, A. T. Algebraic Representation of Thermodynamic Properties and the


Classification of Solutions. Ind. Eng. Chem. 1948, 40, 345-348.

Renon, H.; Prausnitz, J. M. Local Composition Thermodynamic Excess Functions for


Liquid Mixtures. AIChE J. 1968a, 14, 135-144.

Renon, H.; Prausnitz, J. M. Liquid-Liquid and Vapor-Liquid Equilibria from Binary and
Ternary Systems with Dibutylketone, Dimethyl Sulfoxide, N-Hexane and 1-
Hexane. Ind. Eng. Chem. Proc. Res. Dev. 1968b, 7, 220-225.

Renon, H.; Prausnitz, J. M. Estimation of Parameters for the NRTL Equation for Excess
Gibbs Energies of Strongly Nonideal Liquid Mixtures. Ind. Eng. Chem. Proc.
Res. Dev. 1969, 8, 413-419.

Rossini, F. D. Experimental Thermochemistry. Vol. 1, Interscience, New York, 1956,


239-297.

120
Rowlinson, J. S.; Swinton, F. L. Liquid and Liquid Mixtures; 3rd edition; Butterworths:
1982.

Sander, B.; Skjold-Jørgensen, S.; Rasmussen, P. Gas Solubility Calculations. I. UNIFAC.


Fluid Phase Equilibria 1983, 11, 105-126.

Sandler S. I., Chemical and Engineering Thermodynamics; 3rd ed., John Wiley and Sons,
New York, NY, USA, 1999.

Savini, C. G.; Winterhalter, D. R.; Kovach, L. H.; Van Ness, H. C. Endothermic Heats of
Mixing by Isothermal Dilution Calorimetry. J. Chem. Eng. Data 1966, 11, 40-43.

Shoemaker, D. P.; Garland, C. W; Nibler, J. W. Experimental Physical Chemistry; 5th


edition: McGraw Hill, New York, 1989.

Smith, J. M., Van Ness, H. C., Abbott, M. M., Introduction to Chemical Engineering
Thermodynamics; 6th ed., McGraw-Hill, USA, 2001.

Stokes, R. H.; Marsh, K. N.; Tomlins, R. P. An Isothermal Displacement Calorimeter for


Endothermic Enthalpies of Mixing. J. Chem. Thermodyn. 1969, 1, 211-221.

Sturtevant, J. M.; Lyons, P. A. Application of Flow Calorimetry to the Determination of


Enthalpies of Mixing of Organic Liquids. J. Chem. Thermodyn. 1969, 1, 201-210.

Tanaka, R.; D’ Arcy, P. J.; Benson, G. C. Application of a Flow Microcalorimeter to


Determine the Excess Enthalpies of Binary Mixtures of Non-Electrolytes.
Thermochim. Acta. 1975, 11, 163-175.

Thomas, E. R.; Eckert, C. A. Prediction of Limiting Activity Coefficients by a Modified


Separation of Cohesive Energy Density Model and UNIFAC. Ind. Eng. Chem.
Proc. Des. Dev. 1984, 23, 194-209.

Touhara, H.; Okazaki, S.; Okino, F.; Tanaka, H.; Ikari, K.; Nakanishi, K.
Thermodynamic Properties of Aqueous Mixtures of Hydrophilic Compounds 2.
Aminoethanol and Its Methyl Derivates, J. Chem. Thermodyn. 1982, 14, 145-156.

van Miltenburg, J. C.; van den Berg, G. J. K.; van Bommel, M. J. Construction of an
Adiabatic Calorimeter. Measurements of Molar Heat Capacity of Synthetic
Sapphire and of n-Heptane. J. Chem. Thermodyn. 1987, 19, 1129-1137.

Vera, J. H.; Vidal, J. An Improved Group Method for Hydrocarbon-Hydrocarbon


Systems Arising from a Comparative Study of ASOG and UNIFAC. Chem. Eng.
Sci. 1984, 39, 651-661.

Wang, L. L. Excess Enthalpies of Binary and Ternary Mixtures. Ph.D. thesis 1996, The
University of Ottawa, Ottawa, Canada.

121
Weidlich, U.; Gmehling, J. A Modified UNIFAC Model. 1. Prediction of VLE, hE, and
γ ∞ . Ind. Eng. Chem. Res. 1987, 26, 1372-1381.

Wilson G. M. A New Expression for the Excess Free Energy of Mixing American Chem.
Soc. 1964, 86, 127-130.

Wilson G. M.; Deal, C. H. Activity Coefficients and Molecular Structure. Activity


Coefficients in Changing Environments-Solutions of Groups. Ind. Eng. Chem.
Fundam. 1962, 1, 20-23.

Wilson G. M.; VonNiederhausern, D. M. Critical Point and Vapor Pressure Measurement


for Nine Compounds by a Low Residence Time Flow Method. J. Chem. Eng.
Data 2002, 47, 761-764.

Wohl. K.; Thermodynamic Evaluation of Binary and Ternary Liquid Systems. Trans.
Amer. Inst. Chem. Eng. 1946, 42, 215-249.

Wohl. K.; Thermodynamic Evaluation of Binary and Ternary Liquid Systems. Chem.
Eng. Progr. 1953, 49, 218-219.

122
7. APPENDICES

7.1 APPENDIX A: Experimental data

This Appendix contains the experimental data of the molar heat capacity and the

molar excess enthalpy data for the five selected alkanolamine (1) + Water (2) systems.

This Appendix also contains the calculated values of the molar excess heat capacity for

the five selected alkanolamine (1) + water (2) systems.

Table (A1) Experimental molar heat capacity ( C p /J·mole-1·K-1) values, for AEEA (1) +
water (2) mixtures from (303.15 to 353.15) K

Mole fractions of AEEA (x1)

0.0999 0.1999 0.2995 0.4000 0.5002 0.5997 0.7011 0.8003 0.8979 1.0000

T/K Cp /J·mole-1·K-1

303.15 100 121 141 161 181 201 221 240 258 279

308.15 103 125 145 165 185 205 226 245 263 284

313.15 104 127 147 167 187 207 227 247 265 285

318.15 106 129 150 170 190 209 229 249 266 286

323.15 107 131 152 172 192 211 231 250 267 287

328.15 110 134 155 175 195 214 233 252 269 289

333.15 112 137 158 178 197 216 235 254 271 290

338.15 116 141 162 181 201 219 238 257 273 292

343.15 120 146 167 185 204 223 241 260 275 293

348.15 124 152 172 189 208 227 244 262 277 295

353.15 130 158 177 194 213 231 248 265 279 296

123
Table (A2) Experimental molar heat capacity ( C p /J·mole-1·K-1) values, for AP (1) +
water (2) mixtures from (303.15 to 353.15) K

Mole fractions of AP (x1)

0.0999 0.2001 0.3003 0.4015 0.4996 0.6006 0.7003 0.8003 0.8966 1.0000

T/K Cp /J·mole-1·K-1

303.15 94 108 120 131 143 155 168 180 193 204

308.15 96 111 123 135 147 159 171 184 196 207

313.15 97 112 124 136 148 160 172 185 197 208

318.15 98 113 126 138 150 162 173 186 198 208

323.15 99 115 128 140 152 163 175 187 199 209

328.15 101 117 130 142 154 166 177 189 200 210

333.15 103 119 132 145 157 169 179 191 202 211

338.15 106 122 135 148 160 172 182 193 204 213

343.15 110 126 139 151 163 175 185 196 206 214

348.15 114 130 143 155 167 179 188 198 208 216

353.15 119 135 147 160 172 183 192 201 210 217

124
Table (A3) Experimental molar heat capacity ( C p /J·mole-1·K-1) values, for MAE (1) +
water (2) mixtures from (303.15 to 353.15) K

Mole fractions of MAE (x1)

0.1000 0.1998 0.3003 0.3990 0.5002 0.6061 0.6987 0.8003 0.9004 1.0000

T/K Cp /J·mole-1·K-1

303.15 96 108 119 129 139 151 160 172 182 194

308.15 98 111 122 132 143 155 165 176 187 199

313.15 99 113 124 135 146 157 167 178 189 200

318.15 100 114 126 137 148 160 170 180 190 202

323.15 101 116 128 139 150 162 172 183 192 203

328.15 103 118 131 142 153 165 175 185 195 205

333.15 105 121 134 145 157 168 178 188 197 206

338.15 107 124 138 149 161 172 182 191 199 207

343.15 110 127 141 153 164 176 185 194 201 208

348.15 114 132 146 158 169 181 190 198 205 210

353.15 119 138 152 163 174 186 195 203 208 212

125
Table (A4) Experimental molar heat capacity ( C p /J·mole-1·K-1) values, for MIPA (1) +
water (2) mixtures from (303.15 to 353.15) K

Mole fractions of MIPA (x1)

0.1000 0.2007 0.2996 0.4001 0.4983 0.6013 0.6999 0.7996 0.8959 1.0000

T/K Cp /J·mole-1·K-1

303.15 95 110 121 133 145 158 170 183 196 209

308.15 97 113 124 137 149 162 174 187 200 213

313.15 98 114 126 139 151 164 176 188 201 214

318.15 99 116 128 141 154 166 177 190 202 215

323.15 100 117 130 144 156 168 180 192 204 217

328.15 102 119 133 147 159 171 182 194 206 218

333.15 104 121 136 150 162 174 185 197 209 221

338.15 106 125 140 154 166 178 188 200 211 223

343.15 109 128 144 158 170 181 192 203 214 225

348.15 112 133 149 163 175 185 195 206 216 228

353.15 117 138 155 169 180 190 199 209 219 230

126
Table (A5) Experimental molar heat capacity ( C p /J·mole-1·K-1) values, for DMEA (1) +
water (2) mixtures from (303.15 to 353.15) K

Mole fractions of DMEA (x1)

0.0997 0.1962 0.3003 0.3984 0.4983 0.6008 0.6980 0.7981 0.8938 1.0000

T/K Cp /J·mole-1·K-1

303.15 103 118 132 144 156 168 179 191 202 213

308.15 104 119 134 146 158 170 181 193 205 215

313.15 105 121 136 149 161 173 185 196 208 217

318.15 105 123 139 152 164 177 188 199 210 219

323.15 107 125 141 155 168 180 191 203 213 221

328.15 108 127 144 158 171 184 195 206 215 222

333.15 109 129 147 161 175 187 199 209 217 223

338.15 112 132 150 165 179 192 203 213 220 223

343.15 114 136 155 170 184 196 208 218 223 223

348.15 118 140 160 176 190 203 214 223 225 224

353.15 123 146 166 184 198 210 221 229 228 226

127
Table (A6) Molar excess heat capacity ( C pE /J·mole-1·K-1), for AEEA (1) + water (2)
mixtures from (303.15 to 353.15) K

Mole fractions of AEEA (x1)

0.0999 0.1999 0.2995 0.4000 0.5002 0.5997 0.7011 0.8003 0.8979

T/K C pE /J·mole-1·K-1

303.15 4.24 5.17 4.97 4.43 3.90 3.42 2.96 2.12 0.35

308.15 6.63 7.72 7.57 6.77 5.92 5.19 4.51 3.50 0.82

313.15 7.92 9.53 9.41 8.42 7.33 6.33 5.38 4.15 1.20

318.15 9.37 11.5 11.4 10.1 8.84 7.50 6.38 4.93 1.48

323.15 11.0 13.6 13.6 11.8 10.4 8.62 7.24 5.57 1.74

328.15 13.1 16.3 16.1 14.1 12.5 10.3 8.44 6.46 2.36

333.15 15.6 19.2 18.8 16.4 14.4 12.0 9.60 7.33 2.74

338.15 18.7 22.8 22.1 19.2 17.0 14.2 11.1 8.45 3.14

343.15 22.3 27.0 25.9 22.4 19.8 16.8 13.0 9.94 3.86

348.15 26.9 32.2 30.4 26.0 23.0 19.8 15.2 11.5 4.48

353.15 32.7 38.6 35.8 30.4 26.8 23.3 17.7 13.1 5.56

128
Table (A7) Molar excess heat capacity ( C pE /J·mole-1·K-1), for AP (1) + water (2)
mixtures from (303.15 to 353.15) K

Mole fractions of AP (x1)

0.0999 0.2001 0.3003 0.4015 0.4996 0.6006 0.7003 0.8003 0.8966

T/K C pE /J·mole-1·K-1

303.15 5.45 6.88 5.78 4.24 3.49 2.48 1.89 1.93 1.67

308.15 7.42 9.01 7.81 6.28 5.26 4.13 3.20 2.81 2.24

313.15 8.25 10.2 9.06 7.68 6.62 5.20 4.01 3.50 2.68

318.15 9.35 11.5 10.5 9.22 8.12 6.53 4.97 4.28 3.13

323.15 10.6 12.9 12.1 10.8 9.60 7.97 5.83 5.04 3.48

328.15 12.4 14.8 14.2 12.9 11.8 9.84 7.46 6.06 4.25

333.15 14.4 16.8 16.1 14.8 13.6 11.6 8.74 6.91 4.68

338.15 17.0 19.5 18.8 17.3 15.9 13.9 10.5 8.03 5.31

343.15 20.2 22.5 21.6 20.2 18.5 16.6 12.5 9.46 6.21

348.15 24.2 26.4 25.2 23.3 21.7 19.3 14.5 10.7 6.83

353.15 29.3 31.2 29.4 27.2 25.3 22.8 16.9 12.4 7.76

129
Table (A8) Molar excess heat capacity ( C pE /J·mole-1·K-1), for MAE (1) + water (2)
mixtures from (303.15 to 353.15) K

Mole fractions of MAE (x1)

0.1000 0.1998 0.3003 0.3990 0.5002 0.6061 0.6987 0.8003 0.9004

T/K C pE /J·mole-1·K-1

303.15 8.91 9.30 7.64 5.70 4.45 3.13 1.84 0.99 -0.17

308.15 10.5 11.2 9.84 7.95 6.41 4.88 3.36 2.26 0.48

313.15 11.2 12.4 11.3 9.42 7.87 6.10 4.50 3.04 0.93

318.15 12.2 13.8 13.0 11.2 9.59 7.99 6.03 4.11 1.46

323.15 13.2 15.4 14.7 13.0 11.4 9.65 7.67 5.16 2.15

328.15 14.8 17.3 17.0 15.1 13.3 11.5 9.31 6.26 2.93

333.15 16.5 19.4 19.5 17.9 15.9 13.5 11.2 7.75 3.86

338.15 18.8 22.2 22.6 21.2 19.1 16.4 13.9 9.66 4.65

343.15 21.8 25.5 26.2 24.8 22.7 20.2 17.2 12.1 6.54

348.15 25.5 29.6 30.3 28.7 26.5 24.3 20.9 14.9 8.73

353.15 30.2 34.7 34.9 32.9 30.2 27.2 23.8 17.5 9.72

130
Table (A9) Molar excess heat capacity ( C pE /J·mole-1·K-1), for MIPA (1) + water (2)
mixtures from (303.15 to 353.15) K

Mole fractions of MIPA (x1)

0.1000 0.2007 0.2996 0.4001 0.4983 0.6013 0.6999 0.7996 0.8959

T/K C pE /J·mole-1·K-1

303.15 6.57 8.12 5.90 4.44 3.17 2.21 1.26 0.51 0.25

308.15 8.31 10.1 7.94 6.55 5.27 3.83 2.34 1.43 0.90

313.15 9.15 11.1 9.27 8.21 7.02 5.17 3.36 2.12 1.35

318.15 10.0 12.3 10.9 9.89 8.66 6.42 4.18 2.76 1.66

323.15 11.0 13.5 12.6 11.8 10.5 7.96 5.48 3.59 2.18

328.15 12.4 15.2 14.8 14.1 12.4 9.68 6.86 4.54 2.81

333.15 13.8 16.9 17.2 16.4 14.4 11.4 8.17 5.46 3.19

338.15 15.9 19.7 20.2 19.3 16.9 13.5 9.62 6.51 3.68

343.15 18.3 22.9 23.8 22.8 19.9 15.7 11.1 7.45 4.08

348.15 21.6 26.5 27.8 26.9 23.7 18.2 13.0 8.76 4.64

353.15 25.8 31.2 32.8 31.6 27.9 22.0 15.6 9.93 5.04

131
Table (A10) Molar excess heat capacity ( C pE /J·mole-1·K-1), for DMEA (1) + water (2)
mixtures from (303.15 to 353.15) K

Mole fractions of DMEA (x1)

0.0997 0.1962 0.3003 0.3984 0.4983 0.6008 0.6980 0.7981 0.8938

T/K C pE /J·mole-1·K-1

303.15 14.0 15.4 15.1 13.8 12.2 9.71 7.47 5.29 3.83

308.15 14.3 16.4 16.5 15.3 13.6 11.2 8.78 6.48 4.75

313.15 15.1 17.8 18.3 17.1 15.4 13.0 10.2 7.70 5.54

318.15 15.8 19.1 20.0 19.0 17.4 14.8 12.0 9.31 6.30

323.15 16.8 20.7 22.0 21.2 19.7 17.1 14.0 11.0 7.23

328.15 17.9 22.5 24.2 23.7 22.3 19.8 16.7 13.2 8.41

333.15 19.3 24.8 26.9 26.8 25.6 23.0 20.1 16.0 9.93

338.15 21.4 27.7 30.4 30.7 29.5 27.2 24.3 19.8 12.2

343.15 24.2 31.2 34.8 35.8 34.5 32.1 29.2 24.6 15.1

348.15 27.6 35.5 39.7 41.8 40.7 38.0 34.5 29.0 16.9

353.15 32.4 40.9 45.5 48.3 47.1 44.0 40.2 33.0 17.7

132
Table (A11) Molar excess enthalpy (HE/J·mole-1), for AEEA (1) + water (2) mixtures
for (298.15, 313.15, and 323.15) K

298.15 K 313.15 K 323.15 K

x1 HE/J·mole-1 x1 HE/J·mole-1 x1 HE/J·mole-1

0.0298 -771 0.0299 -936 0.0301 -617

0.0496 -1273 0.0498 -1317 0.0498 -1029

0.0699 -1713 0.0701 -1720 0.0697 -1594

0.0992 -2241 0.1000 -2338 0.0992 -2037

0.1503 -3000 0.1495 -3051 0.1491 -2915

0.2015 -3447 0.2002 -3620 0.1988 -3414

0.2962 -3909 0.2956 -4119 0.3014 -3815

0.2982 -3918 0.3547 -4234 0.3527 -3938

0.3484 -4009 0.4001 -4150 0.3958 -3916

0.3994 -3955 0.4461 -3976 0.4971 -3600

0.4454 -3771 0.4893 -3784 0.5908 -3027

0.4962 -3489 0.6009 -3148 0.7056 -2241

0.5050 -3433 0.6723 -2669 0.7484 -1939

0.5922 -2865 0.7961 -1714 0.8999 -820

0.6969 -2192 0.9041 -888 -- --

0.7700 -1711 -- -- -- --

0.9077 -779 -- -- -- --

133
Table (A12) Molar excess enthalpy (HE/J·mole-1), for AP (1) + water (2) mixtures for
(298.15, 313.15, and 323.15) K

298.15 K 313.15 K 323.15 K

x1 HE/J·mole-1 x1 HE/J·mole-1 x1 HE/J·mole-1

0.0299 -404 0.0297 -483 0.0298 -643

0.0705 -927 0.0702 -1026 0.0718 -1213

0.1203 -1525 0.1200 -1569 0.1192 -1660

0.2000 -2130 0.2007 -2108 0.1996 -2144

0.2982 -2501 0.3015 -2541 0.3018 -2542

0.3452 -2595 0.3458 -2632 0.3493 -2645

0.4067 -2613 0.3982 -2647 0.3947 -2667

0.4989 -2462 0.5034 -2499 0.5038 -2498

0.5930 -2171 0.6045 -2125 0.6050 -2163

0.7515 -1531 0.7564 -1375 0.7594 -1423

0.8969 -711 0.9333 -481 0.8801 -753

134
Table (A13) Molar excess enthalpy (HE/J·mole-1), for MAE (1) + water (2) mixtures for
(298.15, 313.15, and 323.15) K

298.15 K 313.15 K 323.15 K

x1 HE/J·mole-1 x1 HE/J·mole-1 x1 HE/J·mole-1

0.0302 -530 0.0310 -535 0.0299 -742

0.0514 -902 0.0500 -902 0.0501 -1114

0.0695 -1163 0.0704 -1182 0.0700 -1359

0.1002 -1480 0.1011 -1591 0.1008 -1655

0.1485 -1919 0.1488 -1987 0.1504 -1975

0.1992 -2243 0.2003 -2302 0.2004 -2220

0.3021 -2671 0.2983 -2639 0.2985 -2568

0.3471 -2750 0.3491 -2678 0.3492 -2642

0.3996 -2732 0.3986 -2668 0.4023 -2650

0.4471 -2667 0.5035 -2513 0.5108 -2478

0.5017 -2508 0.5833 -2289 0.6095 -2142

0.5992 -2148 0.6860 -1919 0.6956 -1732

0.6960 -1707 0.8053 -1277 0.8039 -1169

0.7988 -1219 0.8956 -679 0.8815 -749

0.8939 -688 -- -- -- --

135
Table (A14) Molar excess enthalpy (HE/J·mole-1), for MIPA (1) + water (2) mixtures
for (298.15, 313.15, and 323.15) K

298.15 K 313.15 K 323.15 K

x1 HE/J·mole-1 x1 HE/J·mole-1 x1 HE/J·mole-1

0.0300 -497 0.0304 -560 0.0301 -478

0.0501 -777 0.0701 -1057 0.0500 -714

0.0697 -1010 0.1001 -1364 0.0705 -1022

0.1002 -1357 0.1500 -1761 0.0999 -1289

0.1505 -1755 0.1999 -2054 0.1507 -1679

0.2012 -2040 0.3021 -2443 0.1993 -1979

0.2994 -2393 0.3520 -2516 0.2999 -2304

0.3496 -2482 0.4015 -2502 0.3533 -2415

0.3993 -2478 0.5022 -2327 0.4045 -2454

0.4465 -2423 0.5910 -2039 0.5070 -2318

0.4940 -2320 0.6920 -1640 0.5849 -2081

0.6033 -2012 0.7864 -1184 0.6741 -1756

0.7155 -1527 0.8808 -730 0.8125 -1092

0.7954 -1130 -- -- 0.9073 -601

0.8847 -725 -- -- -- --

136
Table (A15) Molar excess enthalpy (HE/J·mole-1), for DMEA (1) + water (2) mixtures
for (298.15, 313.15, and 323.15) K

298.15 K 313.15 K 323.15 K

x1 HE/J·mole-1 x1 HE/J·mole-1 x1 HE/J·mole-1

0.0299 -701 0.0307 -701 0.0301 -661

0.0705 -1452 0.0695 -1314 0.0705 -1317

0.1008 -1854 0.1196 -1923 0.1217 -1823

0.1501 -2282 0.2038 -2444 0.2027 -2328

0.2052 -2587 0.2979 -2727 0.3001 -2631

0.3010 -2847 0.3490 -2782 0.3486 -2692

0.3490 -2895 0.4063 -2772 0.4001 -2682

0.4106 -2888 0.4925 -2584 0.4995 -2497

0.4544 -2816 0.5885 -2202 0.6093 -2046

0.5365 -2575 0.7046 -1686 0.7618 -1291

0.6599 -2008 0.7891 -1235 0.9197 -464

0.7996 -1247 0.9076 -607 -- --

0.9156 -557 -- -- -- --

137
7.2 APPENDIX B: Molecular area and volume parameters

Appendix B contains pure-component molecular-structure constants depending on the

molecular size and the external surface areas. Molecular surface area and volume

parameters for the UNIQUAC Model are listed in Table (B1). Table (B2) lists the

functional group area (Q) and volume (R) parameters for the Modified UNIFAC

(Dortmund) Model. The parameters in Table (B1) and Table (B2) were retrieved from the

ASPEN PLUS Data Regression Tool.

Table (B1) Molecular surface area and volume parameters for the UNIQUAC model

Surface area Surface area Volume


Alkanolamines
parameter (q) parameter (q’) parameter (r)
2-((2-Aminoethyl)amino)ethanol 3.836 3.836 4.45485

3-Amino-1-Propanol 2.900 2.900 3.24786

2-(Methylamino)ethanol 2.908 2.908 3.31246

1-Amino-2-Propanol 2.896 2.896 3.24984

N-N-Dimethylethanolamine 3.452 3.452 3.96638

Water 1.400 1.000 0.92000

138
Table (B2) Main groups and subgroups and the corresponding van der Waals quantities
for the modified UNIFAC (Dortmund) model

Main Group Subgroup R Q


CH3 0.6325 1.0608
CH2 0.6325 0.7081
CH2
CH 0.6325 0.3554
C 0.6325 0.0000
OH (prim) 1.2302 0.8927
OH OH (sec) 1.0630 0.8663
OH (tert) 0.6895 0.8345
Water H2O 1.7334 2.4561
CH3NH2 1.6607 1.6904
CH2NH2 1.6607 1.3377
CNH2
CHNH2 1.6607 0.9850
CNH2 1.6607 0.9850
CH3NH 1.3680 1.4332
CNH CH2NH 1.3680 1.0805
CHNH 1.3680 0.7278
CH3N 1.0746 1.1760
(C)3N
CH2N 1.0746 0.8240

Gmehling et al. (1993) have also published the van der Waals quantities for the

modified UNIFAC (Dortmund) model.

139

Das könnte Ihnen auch gefallen