A Helmholtz Free Energy Formulation of The Thermodynamic Properties of The Mixture (Water + Ammonia)

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 35

A Helmholtz Free Energy Formulation of the

Thermodynamic Properties of the Mixture


{Water + Ammonia}
Cite as: Journal of Physical and Chemical Reference Data 27, 63 (1998); https://doi.org/10.1063/1.556015
Submitted: 04 March 1997 . Published Online: 15 October 2009

Reiner Tillner-Roth, and Daniel G. Friend

ARTICLES YOU MAY BE INTERESTED IN

Survey and Assessment of Available Measurements on Thermodynamic Properties of the


Mixture
Journal of Physical and Chemical Reference Data 27, 45 (1998); https://
doi.org/10.1063/1.556014

The IAPWS Formulation 1995 for the Thermodynamic Properties of Ordinary Water Substance
for General and Scientific Use
Journal of Physical and Chemical Reference Data 31, 387 (2002); https://
doi.org/10.1063/1.1461829

Thermodynamic properties of ammonia


Journal of Physical and Chemical Reference Data 7, 635 (1978); https://
doi.org/10.1063/1.555579

Journal of Physical and Chemical Reference Data 27, 63 (1998); https://doi.org/10.1063/1.556015 27, 63

© 1998 American Institute of Physics and American Chemical Society.


A Helmholtz Free Energy Formulation of the Thermodynamic Properties of
the Mixture ˆWater 1 Ammonia‰

Reiner Tillner-Rotha… and Daniel G. Friend


National Institute of Standards and Technology, Physical and Chemical Properties Division, 325 Broadway, Boulder, Colorado 80303

Received March 4, 1997; revised manuscript received October 28, 1997

A thermodynamic model incorporating a fundamental equation of state for the Helm-


holtz free energy of the mixture $water1ammonia% is presented which covers the ther-
modynamic space between the solid–liquid–vapor boundary and the critical locus. It is
also valid in the vapor and liquid phases for pressures up to 40 MPa. It represents
vapor–liquid equilibrium properties with an uncertainty of 60.01 in liquid and vapor
mole fractions. Typical uncertainties in the single-phase regions are 60.3% for the den-
sity and 6200 J mol21 for enthalpies. Details of the data selection and the optimization
process are given. The behavior of the fundamental equation of state is discussed in all
parts of the thermodynamic space. © 1998 American Institute of Physics and American
Chemical Society. @S0047-2689~98!00401-2#
Key words: ammonia-water, equation of state, Helmholtz free energy, thermodynamic properties.

Contents 2. Coefficients of the reducing functions Eqs. ~9!


1. Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64 and ~10!, and the departure function, Eq. ~13!. . . . 66
2. The Fundamental Equation of State . . . . . . . . . . . . 65 3. Relations between the reduced Helmholtz free
2.1 Ideal-Gas Properties. . . . . . . . . . . . . . . . . . . . . 65 energy and thermodynamic properties. . . . . . . . . . . 67
2.2 The Residual Part of the Helmholtz Free 4. Saturation properties of $water1ammonia% . . . . . . 78
Energy. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66 5. Properties of $water1ammonia% in the one-phase
3. The Database. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66 region. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
4. The Optimization Process. . . . . . . . . . . . . . . . . . . . 68
5. Discussion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 List of Figures
5.1. Vapor and Liquid Compositions. . . . . . . . . . . 69 1. Distribution of selected ( p,T,x) and (p,T,x,y)
5.2. Saturated Liquid Density. . . . . . . . . . . . . . . . . 69 data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
5.3. Caloric Properties of the Saturated Liquid. . . 70 2. Distribution of selected experimental data in the
5.4. Pressure and Density at the Solid–Liquid– single-phase regions. . . . . . . . . . . . . . . . . . . . . . . . . 68
Vapor Boundary. . . . . . . . . . . . . . . . . . . . . . . . 72 3. Distribution of selected saturated liquid densities,
5.5. The Two-Phase Envelope at High enthalpies and isobaric heat capacities. . . . . . . . . . 68
Temperatures. . . . . . . . . . . . . . . . . . . . . . . . . . . 72 4. Deviations between liquid mole fractions x and
5.6. Properties of the Compressed Liquid. . . . . . . . 73 values from Eq. ~1!. . . . . . . . . . . . . . . . . . . . . . . . . 69
5.7. Properties of Superheated Vapor. . . . . . . . . . . 73 5. Deviations between vapor mole fractions y and
5.8. The Supercritical Region. . . . . . . . . . . . . . . . . 74 values from Eq. ~1!. . . . . . . . . . . . . . . . . . . . . . . . . 70
6. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74 6. Distribution coefficient K of ammonia in dilute
7. Acknowledgments. . . . . . . . . . . . . . . . . . . . . . . . . . 75 solutions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
8. Appendices. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75 7. Deviations between saturated liquid densities
8.1. Appendix A: Linearization of VLE Data. . . . . 75 and values from Eq. ~1!. . . . . . . . . . . . . . . . . . . . . . 71
8.2. Appendix B: Tables of Thermodynamic 8. Deviations between saturated liquid enthalpies
Properties. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76 and values from Eq. ~1!. . . . . . . . . . . . . . . . . . . . . . 71
9. References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76 9. Deviations between isobaric heat capacities and
values from Eq. ~1!. . . . . . . . . . . . . . . . . . . . . . . . . 71
List of Tables 10. Equilibrium pressures and saturated liquid
1. Coefficients of the ideal-gas part of the Helmholtz densities at the solid–liquid–vapor boundary from
free energy, Eq. ~5!. . . . . . . . . . . . . . . . . . . . . . . . . 65 Eq. ~1!. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
11. Dew-bubble curves and coexisting densities at
©1998 by the U.S. Secretary of Commerce on behalf of the United States. high temperatures. . . . . . . . . . . . . . . . . . . . . . . . . . . 72
All rights reserved. This copyright is assigned to the American Institute of 12. Enthalpy and entropy on the two-phase boundary
Physics and the American Chemical Society.
Reprints available from ACS; see Reprints List at back of issue.
at high temperatures. . . . . . . . . . . . . . . . . . . . . . . . . 73
a! 13. Deviations between liquid densities and values
Author to whom correspondence should be sent; Permanent address: Uni-
versität Hannover, Institut für Thermodynamik, Callinstrasse 36, 30167 from Eq. ~6!. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
Hannover, Germany. 14. Molar excess enthalpy H̄ E in the liquid. . . . . . . . . 74

0047-2689/98/27„1…/1/34/$14.00 63 J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998


64 R. TILLNER-ROTH AND D. G. FRIEND

15. Deviations between vapor densities and values 1. Introduction


from Eq. ~1!. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
16. Compressibility factor Z and reduced enthalpy The mixture $water1ammonia% is an important working
H̄/(R m T n ) in the supercritical region for three fluid in absorption–refrigeration cycles. Recently, this mix-
reduced temperatures. ~——! pure components; ture has also been considered as a working fluid in future
~– – – –! x51/3; ~-•-•-! x52/3. . . . . . . . . . . . . . . . . 74 power generation plants based on the Kalina cycle1. For de-
sign and simulation of such machinery, an accurate descrip-
Nomenclature tion of the thermodynamic properties of the mixture $water
1ammonia% for a wide range of pressure, temperature, and
a, Q coefficient
composition is needed.
Ā molar Helmholtz free energy Several thermodynamic models have been published in the
C̄ p molar isobaric heat capacity past. Old sources correlate experimental data by graphical
C̄ v molar isochoric heat capacity methods ~Kracek,2 Scatchard et al.3! and report tables for
Ḡ molar Gibbs free energy saturation properties. A recent source of tables which is
H̄ molar enthalpy widely used today is the work of Macriss et al.4 from the
K distribution coefficient Institute of Gas Technology ~IGT!. These tables cover the
kT , kV factors range up to 5 MPa for vapor–liquid equilibrium ~VLE! and
M molar mass single-phase properties.
p pressure An early wide-ranging equation of state was developed by
Rm universal gas constant Schulz5 in 1971. His model consists of two separate equa-
s uncertainty tions for the Gibbs free energy G5G( p,T,x) of the vapor
S sum of squares and of the liquid phases for pressures up to 2.5 MPa. Ziegler
and Trepp6 extended this model to pressures up to 5 MPa. A
S̄ molar entropy
very recent modification reported by Ibrahim and Klein7 is
T temperature
valid for temperatures above 230 K and pressures up to 11
ū molar internal energy MPa.
V̄ molar volume In addition to the Gibbs free energy models, numerous
w speed of sound other approaches are found in the literature. A Benedict–
x liquid or overall mole fraction of ammonia Webb–Rubin ~BWR!-type equation of state similar to that
y vapor mole fraction of ammonia used by Lee and Kesler8 is presented by Park and Sonntag.9
Y thermodynamic property They chose this model because of its capability to predict
Z compressibility factor properties of $water1ammonia% up to 20 MPa and 650 K,
a , b , g, ti , where experimental data are scarce for the VLE and virtually
di , ei exponents nonexistent for other properties. El-Sayed and Tribus,10 for
wi fugacity coefficient of component i example, took the approach of establishing separate equa-
F reduced Helmholtz free energy tions for different properties which were then combined in
d inverse reduced volume or reduced density cycle calculations. Ammonia–water mixtures can also be
q reduced temperature treated as weak electrolyte solutions based on the work of
j mass fraction Edwards et al.11 Such an approach is given by Kurz12 for
% molar density example. It yields good results for the VLE at low and mod-
t inverse reduced temperature erate temperatures and pressures.
There are several other models reported in the
Subscripts literature,13–17 but most of them are applicable only in a re-
c critical point parameter stricted range or allow calculation of only a limited number
n reducing quantity of thermodynamic properties of $water1ammonia%. None of
01 pure component 1 ~water! these models, however, has been fitted to new measurements
02 pure component 2 ~ammonia! in the single- and two-phase regions. Most of the models
0i pure component i available so far are based on a limited set of experimental
tr triple point parameter data, and it is sometimes unclear whether the selection of
data resulted from a comprehensive analysis. Furthermore,
correlations for $water1ammonia% often use tabulated val-
Superscripts ues, such as those published by Macriss et al.4 or by Scat-
+ ideal gas chard et al.,3 rather than original experimental data. No cor-
E excess relation based on the results of a thorough data analysis
r residual could be found.
8 saturated liquid In this work, a fundamental equation of state for the Helm-
9 saturated vapor holtz free energy of $water1ammonia% is developed accord-

J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998


HELMHOLTZ FREE ENERGY FORMULATION OF THE MIXTURE ˆWATER1AMMONIA‰ 65

ing to the approach of Tillner-Roth.18 The equation incorpo- TABLE 1. Coefficients of the ideal-gas part of the Helmholtz free energy, Eq.
~5!
rates accurate wide-ranging equations of state for the pure
components and describes the entire thermodynamic space of i a +i ui i a +i ti
the mixture by a single mathematical expression. All thermo-
dynamic properties can be calculated from this model and, 1 27.720 435 – 9 216.444 285 –
2 8.649 358 – 10 4.036 946 –
therefore, all types of measurements of thermodynamic prop- 3 3.006 32 – 11 21.0 –
erties can be used as input data during the optimization pro- 4 0.012 436 1.666 12 10.699 55 1/3
cess. 5 0.973 15 4.578 13 21.775 436 23/2
Prior to the development of this correlation, data available 6 1.279 50 10.018 14 0.823 740 34 27/4
in the literature were compiled. Simultaneously with the de- 7 0.969 56 11.964
8 0.248 73 35.600
velopment of the new equation of state, all experimental data
were evaluated and the most consistent data were identified.
The results of the data compilation and analysis are de-
scribed separately by Tillner-Roth and Friend.19
pure components F°0i are different due to different reducing
parameters T n,0i and V̄ n,0i used in the pure fluid equations:
2. The Fundamental Equation of State

A fundamental equation of state for the Helmholtz free F° ~ T, V̄ ,x ! 5 ~ 12x ! F°01 ~ t 01 , d 01! 1xF°02 ~ t 02 , d 02!
energy A5U2TS for a binary mixture has been developed 1 ~ 12x ! ln~ 12x ! 1x ln x. ~3!
by Tillner-Roth.18 It is written in terms of the reduced Helm-
holtz free energy as
For convenience, uniform variables,

5F5F° ~ t °, d °,x ! 1F r~ t , d ,x ! . ~1!
R mT t °5T°n /T and d °5 V̄ °n / V̄ , ~4!
F is split into an ideal part F°, depending on the dimension-
less variables t °5T°n /T, d °5 V̄ °n / V̄ , and mole fraction x of are introduced. Consequently, the coefficients of the pure
ammonia, and a residual part F r, depending on t 5T n /T, fluid ideal-gas functions are transformed, and the ideal-gas
d 5 V̄ n / V̄ , and x. In Eq. ~1!, Ā is the molar Helmholtz free function is written as
energy and R m 58.314471 J mol21 K21 is the universal gas
constant given by Moldover et al.20 F
F°~t°,d°,x!5lnd°1~12x! a°11a°2 t°1a°3 lnt°1ln~12x!

GF
Pure fluid equations of state in terms of the dimensionless
8
Helmholtz free energy form the basis of this model. For wa-
ter, the fundamental equation of state developed by Prub and 1 ( a°i ln@12exp~2uit°!# 1x a°91a°10 t°
i54
Wagner21 is used, which was adopted as the IAPWS Formu-
lation in 1995 for the thermodynamic properties of ordinary
water substance for general and scientific use by the Interna-
tional Association for the Properties of Water and Steam. For
1a°11 ln t°1ln x1 (
14

i512
G
a°i ~ t ° ! t i . ~5!

ammonia, the fundamental equation of state developed by T°n 5500 K and 1/V̄ °n 515 000 mol m23 were chosen arbi-
Tillner-Roth et al.22 is used. Both formulations are given in trarily. The coefficients for Eq. ~5! are given in Table 1.
terms of F( t , d ) and are based on the current temperature They differ from the original coefficients21,22 because of the
scale ~ITS-90!. variable transformation.
The reference values for enthalpy ~or internal energy! and
2.1. Ideal-Gas Properties
entropy of the mixture equation of state ~EOS! are fixed be-
Ideal-gas properties of $water1ammonia% are represented cause of preset reference values of the pure fluid equation of
s
by the ideal-gas part F° which is derived from the ideal-gas state. They can be changed by changing the coefficients a 1
parts F +01 and F +02 of the water and ammonia equations of
s s s
and a 2 , which correspond to the water EOS, and a 9 and a 10
state. They are combined at constant T and constant V̄ ac- related to the ammonia equation of state. However, this
cording to would change the enthalpy and entropy values of the pure
fluid equations of state as well. In the present work it has
F° ~ T, V̄ ,x ! 5 ~ 12x ! F 01~ T, V̄ ! 1xF 02~ T, V̄ !
s s
been chosen to set

1 ~ 12x ! ln~ 12x ! 1x ln x. ~2!


u 8 50 and s 8 50
The last two terms result from the entropy of mixing of the
ideal-gas mixture. Because Eq. ~ 2! is evaluated at fixed T for saturated liquid at the triple-point temperatures of both
and V̄ , the dimensionless variables t 0i 5T n,0i /T and pure components, 273.16 K for water and 195.495 K for
d 0i 5 V̄ n,0i / V̄ used to evaluate the ideal-gas function for the ammonia.

J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998


66 R. TILLNER-ROTH AND D. G. FRIEND

TABLE 2. Coefficients of the reducing functions, Eqs. ~9! and ~10!, and the departure function, Eq. ~13!

Equation of state constants


water ammonia
T c,015647.096 K T c,025405.40 K
% c,015322 kg m23 % c,025225 kg m23
p c,01522.064 MPa p c,02511.36 MPa
M 1 50.018 015 268 kg mol21 M 2 50.017 030 26 kg mol21
Reducing functions
k V 51.239 5117, k T 50.964 8407, a 51.125 455, b 50.897 8069
Departure function
g 50.524 8379
i ai ti di ei i ai ti di ei

1 21.855 822E-02 3/2 4 – 8 21.368 072E-08 4 15 1


2 5.258 010E-02 1/2 5 1 9 1.226 146E-02 7/2 4 1
3 3.552 874E-10 13/2 15 1 10 27.181 443E-02 0 5 1
4 5.451 379E-06 7/4 12 1 11 9.970 849E-02 21 6 2
5 25.998 546E-13 15 12 1 12 1.058 4086E-03 8 10 2
6 23.687 808E-06 6 15 2 13 20.196 3687 15/2 6 2
7 0.258 6192 21 4 1 14 20.777 7897 4 2 2

2.2. The Residual Part of the Helmholtz Free Energy The reducing functions contain a total of four adjustable pa-
rameters a , b , k T and k V which have to be fitted to experi-
The residual part F r has the form mental data. The values are listed in Table 2. For the reduc-
F r~ t , d ,x ! 5 ~ 12x ! F r01~ t , d ! 1xF r02~ t , d ! 1DF r~ t , d ,x ! . ing constants of the pure fluids their critical parameters are
~6! used as listed also in Table 2.
For the departure function DF r, the following expression
F r01 and F r02 are the residual contributions of the water
has been developed:
and the ammonia equations of state given by Prub and
Wagner21 and Tillner-Roth et al.22 They are combined at DF r~ t , d ,x !
6

constant reduced variables t and d . An empirical departure


x ~ 12x g !
5a 1 t d 1
t1 d1
( a i exp~ 2 d e ! t t d d
i52
i i i

function DF r is needed to describe the thermodynamic prop-


erties of $water1ammonia% accurately. It is evaluated at the 13

same reduced variables t and d as the residual parts of the 1x ( a i exp~ 2 d e ! t t d d


i57
i i i

pure fluids.
To ensure that the mixture model is valid also for the pure 1a 14x 2 exp~ 2 d e 14! t t 14d d 14. ~13!
components, t and d are functions of composition,
The optimization process and the formulation of the equation
T n~ x ! V̄ n ~ x ! of state are described in Section 4. Coefficients and expo-
t~ x !5 and d ~ x ! 5 , ~7!
T V̄ nents are listed in Table 2. Relations between thermody-
namic properties and the fundamental equation of state are
and they must fulfill the conditions summarized in Table 3.
x→0: T n ~ x ! →T c,01 and V̄ n ~ x ! → V̄ c,01 ,

x→1: T n ~ x ! →T c,02 and V̄ n ~ x ! → V̄ c,02 . ~8! 3. The Database


For the functions T n (x) and V̄ n (x), the expressions
During the last 150 years, numerous experimental studies
T n ~ x ! 5 ~ 12x ! 2 T c,011x 2 T c,0212x ~ 12x a ! T c,12 ~9! on the thermodynamic properties of $water1ammonia% were
and carried out. Available experimental data are summarized by
Tillner-Roth and Friend.19 The majority of measurements is
V̄ n ~ x ! 5 ~ 12x ! 2 V̄ c,011x 2 V̄ c,0212x ~ 12x b ! V̄ c,12 concerned with VLE properties such as ( p,T,x), (p,T,x,y),
~10! or ( p,T,y) data. Few sources report single-phase measure-
are used where ments. Since the accuracy of VLE data varies strongly
among different sources, an assessment of all these measure-
T c,125k T 21 ~ T c,011T c,02! ~11! ments is essential in order to determine the most reliable
and data. Furthermore, the selected VLE data have to be thermo-
dynamically consistent with available single-phase data to
V̄ c,125k V 21 ~ V̄ c,011 V̄ c,02! . ~12! allow the formulation of an accurate equation of state.

J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998


HELMHOLTZ FREE ENERGY FORMULATION OF THE MIXTURE ˆWATER1AMMONIA‰ 67

TABLE 3. Relations between the reduced Helmholtz free energy and thermodynamic properties

Compressibility factor Z p( t , d ,x) V̄


511 d F rd
R mT
Molar internal energy Ū( t , d , t °, d °,x)
5 t °F t+ ° 1 t F tr
R mT
Molar enthalpy H̄( t , d , t °, d °,x)
511 d F rd 1 t °F t+ ° 1 t F tr
R mT
Molar entropy S̄ ( t , d , t °, d °,x)
5 t °F t+ ° 1 t F tr 2F°2F r
Rm
Molar isochoric heat capacity C̄ v ( t , d , t °, d °,x)
52 t ° 2 F t+ ° t ° 2 t 2 F tt
r
Rm
Molar isobaric heat capacity C̄ p ( t , d , t °, d °,x) C̄ v (11 d F rd 2 d t F dr t ) 2
5 1
Rm R m 112 d F rd 1 d 2 F rdd
Speed of sound w 2 ( t , d , t °, d °,x)M (11 d F rd 2 d t F dr t ) 2
5112 d F rd 1 d 2 F rdd 1
R mT (C̄ v /R m )
Fugacity coefficients ln@Zw1(t,d,x)#5Fr1 d F rd 2xF w

ln@Zw2(t,d,x)#5Fr1 d F rd 1(12x)F w

Abbreviations:

Ft+ °5 S D
]F°
]t°
S D S D S D
d°,x
,Ft+ °t°5
]2F°
]t°2 d°,x
,Ftr 5
]Fr
]t
d ,x
,F rd 5
]Fr
]d
t ,x

F tt
r
5 S D
] F
2

]t2
r

S D S D S D
d ,x
,F dr t 5
] F
2

]t]d
r

x
, F rdd 5
] F
2

]d2
r

t ,x
F rx 5
]F
]x
r

d,t
,

F G
F w 5F rx 1
d dV̄ n
V̄ n dx
F rd 1
t dT n r
F
T n dx t

The data analysis was carried out simultaneously with the ~vii! 146 enthalpies of the saturated liquid of Zinner,34
development of the equation of state presented here. Starting ~viii! 98 isobaric heat capacities reported by Hildenbrand
with the entire set of experimental data, interim results for and Giauque,35 Chan and Giauque,36 and by Wrewsky
the equation of state were used to compare measurements and Kaigorodoff.37
and to eliminate data which showed large scatter or signifi-
The isobaric heat capacities from References 35 and 36 were
cant systematic deviations. This process was repeated several
times with more stringent criteria imposed at each step. found to be inconsistent with the enthalpies measured by
Since the Helmholtz free energy model established here con- Zinner,34 as shown by Tillner-Roth and Friend.19 Since these
sists of a single expression describing the whole thermody- were the only caloric data at low temperatures, they were
namic space, systematic deviations of data from the model nevertheless included, although with a low weight, to give
are also an indicator for thermodynamic inconsistencies be- the equation of state at least some support in the low tem-
tween measurements of different properties, especially be- perature liquid region. In addition to the selected data, the
tween single-phase and two-phase properties. (p,T,x,y) data of Gillespie et al.38 were generally included
The results of the data analysis are given by Tillner-Roth for comparisons during the optimization process. However,
and Friend19 for all available sets of measurements. From since they overlap the selected data, they were not used in
these results, the following experimental data were selected fitting the coefficients.
to establish the fundamental equation of state for $water In contrast to other correlations, no tabulated data like
1ammonia%: those of Macriss et al.4 were used in this work, because the
accuracy of tabulated values may be affected by the choice
~i! 402 (p,T,x) data of Sassen et al.,23 Postma,24 and weight of experimental input data used to generate the
Mollier,25 and Perman,26 numbers.
~ii! 265 (p,T,x,y) data of Smolen et al.,27 Polak and At high temperatures and pressures, only experimental
Lu,28 and Iseli,29 data for the (p,T,x) and ( p,T,x,y) behavior are available.
~iii! 57 saturated liquid densities of Jennings,30 Coexisting densities in the critical region were obtained from
~iv! 1208 liquid (p, V̄ ,T,x) data of Harms-Watzenberg,31 a modified Leung–Griffiths model which was established si-
~v! 599 vapor ( p, V̄ ,T,x) data of Harms-Watzenberg31 multaneously with the present equation of state. Details of
and of Ellerwald,32 the Leung–Griffiths formulation for $water1ammonia% will
~vi! 92 excess enthalpies in the liquid of Staudt,33 be published in a separate paper.39 The Leung–Griffiths for-

J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998


68 R. TILLNER-ROTH AND D. G. FRIEND

FIG. 2. Distribution of selected experimental data in the single-phase re-


gions.

~iv! checking the behavior of the interim equation in the


entire thermodynamic space.
Two optimization methods were employed. A modified
FIG. 1. Distribution of selected (p,T,x) and (p,T,x,y) data. Marquard–Fletcher algorithm developed by Dennis et al.40
was used for all nonlinear optimizations. The structure of the
departure function was determined using the implementation
malism was chosen because it is based on scaling law theory of Wagner’s regression analysis,41 which was developed at
and is able to accurately describe VLE behavior of mixtures the Institute of Thermodynamics, University of Hannover,
in the critical region. Equilibrium data ( p,T,x,y,% 8 ,% 9 ) Germany.42
were calculated from an interim Helmholtz free energy equa- A weighted sum of squares

S D
tion up to about three-fourths of the critical pressure. These
i 2Y i
Y exp calc 2
data served as input during a subsequent optimization of the
Leung–Griffiths model. Additionally, experimental ( p,T,x)
S5 (i si
~14!
and (p,T,x,y) data were also considered in this stage. Den-
sities from the improved Leung–Griffiths model in the criti-
cal region were included during a subsequent improvement
of the Helmholtz free energy model.
VLE data ~Fig. 1! cover the whole two-phase region be-
tween the solid–liquid–vapor boundary and the critical line.
Single-phase properties ~Fig. 2! are restricted to temperatures
between 200 K and 413 K in the liquid reaching up to 40
MPa. In the vapor, available data cover temperatures be-
tween 323 K and 523 K at pressures up to 10 MPa and
densities up to 1.5 mol dm23 . Leung–Griffiths data
(p,T,x,y,% 8 ,% 9 ) were used for four temperatures between
450 K and 600 K. They are shown in Fig. 3 together with
selected saturated liquid densities and enthalpies.

4. The Optimization Process

Several tasks had to be simultaneously accomplished dur-


ing the optimization process:
~i! fitting all adjustable parameters to the selected experi-
mental data,
~ii! finding a structure for the reducing functions T n (x)
and V̄ n (x) and the departure function DF r, FIG. 3. Distribution of selected saturated liquid densities, enthalpies and
~iii! assessing and modifying the database, isobaric heat capacities.

J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998


HELMHOLTZ FREE ENERGY FORMULATION OF THE MIXTURE ˆWATER1AMMONIA‰ 69

was used for the objective function. For the nonlinear opti-
mization, it was formulated according to the direct method
described by Ahrendts and Baehr.43 In this case, Y i can be
any thermodynamic property and s i is the corresponding to-
tal uncertainty estimated from experimental uncertainties ac-
cording to the Gaussian error propagation law.
The sum of squares for the regression analysis is formu-
lated differently. Since this optimization method is used to
find the structure of the departure function, only the contri-
bution of an experimental value to DF r can be used for the
quantity Y i . For a given set of reducing functions, this con-
tribution can be extracted from any experimental property
which depends linearly on the coefficients of the departure
function. The quantity Y for a (p, V̄ ,T,x) value, for example,
is calculated from

S D FS D S D
FIG. 4. Deviations between liquid mole fractions x and values from Eq. ~1!.
] DF r 1 p V̄ ] F r01
Y5 5 21 2 ~ 12x !
] d t d R mT ]d t

S DG
After a few steps of optimization, it became clear that
] F r02
2x . ~15! these terms did not sufficiently describe the composition de-
]d t pendence of DF r. Therefore, terms of the form
x 2 ~ 12x g ! d m t n exp~ 2 d e ! and x 3 ~ 12x g ! d m t n exp~ 2 d e !
In this case, Y is linearly related to the first derivative of DF r
with respect to d . Similar linear relations can be derived for were also included, bringing the total number of candidate
other properties. terms to about 850. Subsequently, terms with e.2 were no
Nonlinear properties have to be linearized before they can longer selected. Therefore, the number of terms was re-
be considered during the regression analysis. Regarding the stricted to about 650.
data situation for $water1ammonia%, this is important for The search for the functional form of the departure func-
VLE data which form the major part of the database. It is tion was initially based only on ( p, V̄ ,T,x) data and on lin-
essential to use this experimental information during the lin- earized VLE properties which were obtained from the in-
ear regression analysis in order to formulate an accurate de- terim equation. As the equation became more accurate,
parture function. The linearization of VLE data is described linearized enthalpy and heat capacity data were also in-
in Appendix A. cluded.
In a first step, only the coefficients k T and k V of the re-
ducing functions T n (x) and V̄ n (x) were optimized with the
departure function DF r in Eq. ~6! set to zero. With this first 5. Discussion
equation, the available experimental data were assessed in
order to identify unreliable measurements which were subse- In the following sections, the behavior of the equation of
quently discarded from the database. It was also used for the state in the different regions of the thermodynamic space is
linearization of VLE data to be included in the subsequent examined. In addition to comparisons with selected experi-
search for the departure function. The assessment of experi- mental data, the general behavior of density, enthalpy, en-
mental data was repeated each time an improved equation of tropy, and heat capacity is discussed in regions where experi-
state was found. mental data are scarce or unavailable. The equation of state
The structure of the departure function was determined is also extrapolated into the supercritical region. Compari-
using Wagner’s regression analysis.41 Wagner’s method al- sons for data which are not discussed here are given by
lows the correlator to select from a bank of terms those terms Tillner-Roth and Friend19 where the same equation of state
which represent a given data set with the lowest possible has been used.
value of the objective function.
The initial departure function was formulated from a bank 5.1. Vapor and Liquid Compositions
of about 400 terms of the forms Experimental VLE data for the ( p,T,x,y) behavior form
the major part of available measurements. In Figs. 4 and 5,
x ~ 12x g ! d m t n and x ~ 12x g ! d m t n exp~ 2 d e ! the liquid and vapor mole fractions x and y are compared
with values which were obtained from flash calculations for
where 1,m,5, 20.5,n,20, and 1,e,4. The exponent given p and T. Absolute deviations of liquid mole fractions
g remains constant during the regression analysis, and is are plotted in Fig. 4 versus temperature, pressure and liquid
optimized only in the nonlinear process. composition for the data sets selected for the optimization.

J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998


70 R. TILLNER-ROTH AND D. G. FRIEND

FIG. 6. Distribution coefficient K of ammonia in dilute solutions.

deviations increase up to 60.03 when approaching the criti-


cal locus.
The good representation is partially due to the fact that the
vapor mole fraction is close to 1 over a large range of tem-
perature and pressure. In this region, the vapor phase behav-
ior is dominated by the ammonia equation of state, which is
very accurate in the gaseous phase. In the range of interme-
diate and low vapor mole fractions, the dew curve is very flat
FIG. 5. Deviations between vapor mole fractions y and values from Eq. ~1!. at low temperatures. Small uncertainties in pressure result in
considerable deviations for the vapor composition. Accord-
ingly, a scatter of 60.02 or less in y is a remarkable result.
In the range of very dilute ammonia solutions, only a few
At temperatures below 250 K, a slight systematic devia- data are available. The distribution coefficient K5y/x of
tion is observed reaching 20.02 in x for the lowest tempera- ammonia is plotted over liquid mole fraction in Fig. 6 and is
tures around 200 K. Above 250 K up to around 390 K liquid compared with experimental results at two different pres-
mole fractions are represented within 60.01, which is within sures and for one temperature. The K values obtained from
the scatter of experimental data. Systematic deviations are (p,T,x,y) measurements of Polak and Lu28 agree very well
observed for some of the data measured by Sassen et al.23 with the values obtained from Eq. ~1!, especially at 0.448
Sassen’s measurements at x50.189 and x50.346 show an MPa where they are represented almost within the experi-
offset of 20.01 in x compared to the data of Iseli29 and mental scatter. At 0.1013 MPa they are slightly higher than
Smolen et al.27 in their overlapping temperature range. The K values from Eq. ~1!. Data of Dvorak and Boublik44 are
mole fractions of these two series of Sassen et al., therefore, compared for T5363.15 K. At liquid mole fractions around
were shifted to x50.199 and x50.355 during the optimiza- 0.01 agreement is good, while at lower mole fractions the
tion, so that they smoothly connect to the other selected data. experimental distribution coefficients are higher than the cal-
However, the data shown in Fig. 4 are Sassen’s original val- culated results. Furthermore, the uncertainty seems to in-
ues. At high temperatures close to the critical temperature crease for the more dilute systems.
line, deviations in x increase up to 0.04 in mole fraction. For
$water1ammonia% the dew-bubble curves are extremely flat 5.2. Saturated Liquid Density
in the vicinity of the critical line ~see Fig. 11! and large
deviations in x or y correspond to only small deviations of Measured saturated liquid densities cover temperatures be-
pressure. No other systematic pattern is observed for the de- tween 243 K and 500 K. Densities at high temperatures are
viations when they are plotted over pressure or liquid mole available only for water-rich mixtures and, thus, do not ex-
fraction x, except for the deviations at low and high tempera- tend into the critical region. Measurements of King et al.,45
tures discussed before. Representation of experimental data Jennings,30 and Harms-Watzenberg31 are compared with re-
is mostly within 60.01. sults from the equation of state in Fig. 7. Densities are gen-
Comparisons for vapor compositions y are given in Fig. 5 erally represented within 61% over the entire range of tem-
for selected data sets. The new EOS represents the vapor perature and composition. None of the three data sets shows
mole fractions generally within 60.01 or better, with the a systematic pattern in the deviations, except the densities of
exception of data at high temperatures and pressures where Jennings30 at x50.8 which are probably too high.

J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998


HELMHOLTZ FREE ENERGY FORMULATION OF THE MIXTURE ˆWATER1AMMONIA‰ 71

FIG. 8. Deviations between saturated liquid enthalpies and values from Eq.
FIG. 7. Deviations between saturated liquid densities and values from Eq. ~1!.
~1!.

mole fractions between 0.3 and 0.4 and positive around


5.3. Caloric Properties of the Saturated Liquid x50.7. This behavior is due to the inconsistency between
For the saturated liquid, Zinner34 published a smoothed enthalpies and heat capacities, which were measured be-
table of enthalpy data instead of the original measurements. tween x50.333 and x50.67.
These tabulated values were used during the correlation pro- Heat capacities are discussed in Fig. 9. The negative de-
cess to establish Eq. ~1! because no other enthalpy data were viations of up to 10% observed for the data of Giauque and
available in this range. In addition to these enthalpies, liquid his co-workers35,36 reflect the inconsistency between these
isobaric heat capacities were reported by two groups35,36 at heat capacities and the saturated liquid enthalpies of Zinner.
very low temperatures. These heat capacities were probably Deviations for the heat capacities of Wrewsky and
measured under atmospheric pressure and, therefore, do not
exactly overlap the enthalpies. However, they are located
very close to the saturation boundary and, therefore, signifi-
cantly influence the representation of the caloric properties of
the saturated liquid.
Unfortunately, the enthalpies and heat capacities are in-
consistent, as described by Tillner-Roth and Friend.19 Since
no other enthalpy or heat capacity measurements are avail-
able in this range, it is not clear which data set is more
reliable. This inconsistency can be resolved only by new
reliable measurements of heat capacity or enthalpy. Since
enthalpies and heat capacities were included during the op-
timization to prevent the development of nonphysical behav-
ior at low temperatures, systematic deviations are observed
for both properties where experimental data overlap.
In Fig. 8, Zinner’s enthalpies are compared with results
from Eq. ~1!. They are represented within 6300 J/mol over
the entire temperature range of his experiments; this is about
twice our estimate of the experimental uncertainty. The larg-
est deviations occur at low temperatures, where the tabulated
data of Zinner may be an extrapolation of his underlying
experimental results. At temperatures above 370 K, devia-
tions are within 6150 J mol21 and thus correspond to the
experimental uncertainty. Deviations are negative at liquid FIG. 9. Deviations between isobaric heat capacities and values from Eq. ~1!.

J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998


72 R. TILLNER-ROTH AND D. G. FRIEND

FIG. 10. Equilibrium pressures and saturated liquid densities at the solid–
liquid–vapor boundary from Eq. ~1!.

Kaigorodoff,37 measured at higher temperatures, are only


within 64%. These data seem to be more consistent with
Zinner’s results than the heat capacities from Refs. 35 and
36.
FIG. 11. Dew-bubble curves and coexisting densities at high temperatures.
5.4. Pressure and Density at the
Solid–Liquid–Vapor Boundary

The solid–liquid–vapor temperature as a function of mix- three-phase boundary, equilibrium pressures were calculated
ture composition is described by the correlation given by along lines of constant composition. These lines show a
Tillner-Roth and Friend.19 When the reduced temperature physically reasonable shape in the entire range of composi-
T tr(x)/T n (x) is calculated at this boundary, values down to tion.
0.28 are obtained. The lower temperature limits of the pure An interesting result is obtained for the saturated liquid
fluid equations of state are given by the respective triple density along the solid–liquid–vapor boundary. Starting
point temperatures corresponding to T tr,01 /T c,0150.43 for from pure water, the density decreases for increasing ammo-
water and T tr,02 /T c,0250.48 for ammonia. Since the mixture nia content until a minimum is reached around 200 K and
model uses pure fluid equations combined at constant re- x'0.28. Towards the eutectic point at x50.33, Eq. ~1! pre-
duced variables, the pure fluid equations must be extrapo- dicts an increasing density. Since no experimental data are
lated considerably beyond their lower temperature limits for available in this region, no conclusion can be drawn whether
calculation of properties close to the three-phase locus. Any this prediction is correct or not. For liquid mole fractions
spurious behavior due to extrapolation has to be compen- increasing above x50.33, liquid density decreases again to-
sated for by the departure function, in order to achieve an wards the triple-point liquid density of ammonia at x51.
adequate representation of thermodynamic properties of the
$water1ammonia% mixture close to its three-phase locus.
At low temperatures, Eq. ~1! has been fitted to experimen- 5.5. The Two-Phase Envelope at High
tal (p,T,x) data of Postma24 down to 200 K and to liquid Temperatures
densities from Harms-Watzenberg31 reaching down to
243.15 K. The behavior at very low temperatures is illus- From the set of selected data, only the (p,T,x) data of
trated by Fig. 10. Equilibrium pressures and liquid densities Sassen et al.23 and some (p,T,x,y) data of Iseli29 at high
shown in this graph were obtained from VLE calculations for ammonia concentrations extend to the critical locus. Devia-
a given x and a temperature obtained from the triple-point tions of vapor and liquid composition were already discussed
temperature equation established by Tillner-Roth and in Figs. 4 and 5. In this section, the general form of the
Friend.19 The results are shown in Fig. 10. For comparison, two-phase envelope is the topic of interest.
triple-point pressures measured by Postma24 are also in- Properties on the two-phase boundary are shown in Figs.
cluded. No density measurements are available below 243 K. 11 and 12 in the near-critical range for five temperatures
Postma’s equilibrium pressures are represented fairly well between 400 K and 600 K. Coexisting densities and
over the full range of the solid–liquid–vapor boundary. (p,T,x,y) data from the modified Leung–Griffiths model es-
Triple-point pressures are extremely low (,50 Pa! near the tablished by Rainwater and Tillner-Roth39 for the critical re-
eutectic point at x50.33. In addition to the pressures on the gion of $water1ammonia% are included in Fig. 11.

J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998


HELMHOLTZ FREE ENERGY FORMULATION OF THE MIXTURE ˆWATER1AMMONIA‰ 73

FIG. 12. Enthalpy and entropy on the two-phase boundary at high tempera-
tures.

FIG. 13. Deviations between liquid densities and values from Eq. ~6!.
Slight differences between results from Eq. ~1! and the
Leung–Griffiths model are observed for 500 K and 550 K
close to the critical locus. Vapor mole fractions from the
Leung–Griffiths model are higher by about 0.03 to 0.05,
liquid mole fractions are higher by up to 0.02 than values on average they are represented well. Other data, which all
from Eq. ~1! for these isotherms. Agreement between the two overlap the results of Harms-Watzenberg, are discussed by
models for liquid and vapor compositions is good at 400 K, Tillner-Roth and Friend.19
450 K, and 600 K. In addition to density, the enthalpy of mixing has also
Differences between coexisting densities correspond to the been measured in the liquid phase. Results of Staudt33 are
deviations observed for the (p,T,x,y) behavior. Agreement compared with results from Eq. ~1! in Fig. 14. The enthalpy
is excellent for the isotherms at 400 K, 450 K, and 600 K, of mixing is almost independent of pressure and temperature
while the coexistence curves at 500 K and 550 K from Eq. in the range of available experimental data between 298 K
~1! are shifted to slightly lower ammonia mole fractions and 373 K and at pressures between 2 MPa and 12 MPa.
compared to the Leung–Griffiths data. Reasonable behavior Higher deviations occur only at intermediate mole fractions
is also observed for entropy and enthalpy on the two-phase around x50.5, where differences of about 2250 J mol21 are
envelope, Fig. 12. The curves for all isotherms go smoothly observed at high temperatures. The excess enthalpy is par-
from the saturated vapor to the saturated liquid. It has to be ticularly well represented for x.0.6.
emphasized again that no experimental densities or caloric
properties are available to verify the accuracy of these pre-
dictions.

5.7. Properties of Superheated Vapor

Two sets of ( p, V̄ ,T,x) properties are available in the va-


5.6. Properties of the Compressed Liquid
por phase.31,32 In Fig. 15 density deviations are plotted for
Equation ~1! has been fitted to the extensive set of liquid both sets over p, T, and x. Most data are represented within
densities measured by Harms-Watzenberg.31 His results are 60.5% in the entire range of experiments. Deviations for the
compared with Eq. ~1! in Fig. 13. The data are represented data of Harms-Watzenberg31 generally increase with increas-
within 60.8% with a few exceptions where his results ap- ing pressure, reaching 11.3% at high temperatures. Eller-
proach the critical region at high ammonia mole fractions. At wald’s data32 do not show this behavior. Deviations for his
temperatures below 300 K, the data show larger scatter, but densities remain within 60.5% at pressures up to 10 MPa.

J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998


74 R. TILLNER-ROTH AND D. G. FRIEND

FIG. 16. Compressibility factor Z and reduced enthalpy H̄/(R m T n ) in the


E supercritical region for three reduced temperatures. ~——-! pure compo-
FIG. 14. Molar excess enthalpy H̄ in the liquid.
nents; ~– – – –! x51/3; ~-•-•! x52/3.

5.8. The Supercritical Region


are plotted over the reduced density d in Fig. 16. The reduc-
No experimental data are available in the supercritical re-
ing functions, Eqs. ~9! and ~10!, were used to calculate
gion. Therefore, only the general behavior of the equation of
state can be analyzed. For this purpose, the compressibility d (x)5 V̄ n (x)/ V̄ and q 51/t 5T/T n (x). Results are shown
for the pure components and for the mixtures at x51/3 and
factor Z5p V̄ /(R m T) and the reduced enthalpy H̄/(R m T n )
x52/3 for three reduced temperatures q 5T/T n 51/t at 1.1,
1.5, and 2.0.
Reasonable behavior is obtained for all three temperatures.
The compressibility factor of the mixtures shows about the
same behavior as for the pure fluids, although it is slightly
higher for the mixtures around d 51.2 than for the pure com-
ponents. This might be a property of the mixture, but it is
also possible that there is a small structural error in the equa-
tion of state, due to the lack of experimental data in this
region.
Enthalpies for the mixtures are located between the enthal-
pies for the pure components for q up to 1.5. At q 52, the
enthalpy of the mixture is closer to the enthalpy of pure
ammonia, especially around d 51.5. This behavior corre-
sponds to the observations for the compressibility factor Z.
Experimental data in the supercritical region are required to
determine whether or not this result is correct.
Overall, the behavior of the equation of state is reasonable
even when extrapolated to 23T n which corresponds to
810 K for ammonia and to 1295 K for pure water. However,
since no experimental data are available in this region, no
estimates of accuracy can be given for supercritical states.

6. Conclusions

A fundamental equation of state for the Helmholtz free


FIG. 15. Deviations between vapor densities and values from Eq. ~1!. energy of $water1ammonia% has been established from

J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998


HELMHOLTZ FREE ENERGY FORMULATION OF THE MIXTURE ˆWATER1AMMONIA‰ 75

which all thermodynamic properties can be derived by ther- equation of state is used to determine the properties remain-
modynamic relations. The entire thermodynamic space from ing to complete the full information ( p,T,x,y, V̄ 8 , V̄ 9 ) for a
vapor to liquid, including VLE, is described by a single vapor–liquid equilibrium state. For a ( p,T,x) value, for ex-
mathematical expression. ample, y, V̄ 8 , and V̄ 9 have to be calculated.
The Helmholtz free energy model covers the entire two- From the first two equations two independent conditions
phase region between the solid–liquid–vapor boundary and can be determined for the saturated vapor and saturated
the critical locus. The uncertainty of liquid and vapor mole liquid which are identical to those for single-phase
fractions is about 60.01 except in the vicinity of the critical
(p, V̄ ,T,x) data as described in Section 4. Using the formula
locus, where it can increase up to 0.04. Enthalpies and den-
from Table 3 for calculation of the fugacity coefficient, Eqs.
sities show a reasonable behavior on the whole two-phase
~A3! and ~A4! transform into
envelope, although experimental data for these properties are
available only in a limited range of temperature and compo-
12x Z9
sition. ln 52ln 1F r92F r8 1 d 9 F rd 9 2 d 8 F rd8
Experimental data in the single-phase regions are 12y Z8
restricted to subcritical temperatures, for the liquid below
420 K and 40 MPa and for the vapor for pressures below 2y ~ F rx9 1 t 8x F rt9 1 d 9x F rd9 !
10 MPa. Typical uncertainties are 60.3% for the density and
1x ~ F rx8 1 t x F rt8 1 d x8 F rd8 ! , ~A5!
6200 J/mol for enthalpy in this range. No experimental data 8
are available for supercritical temperatures. Extrapolation
x Z9
into the supercritical area gives reasonable results, but the ln 52ln 1F r9 2F r8 1 d 9 F rd9 2 d 8 F rd8
accuracy in these regions is unknown. y Z8
In conclusion, the new equation of state represents the
currently available measurements mostly within the limits of 1 ~ 12y !~ F rx9 1 t x9 F rt9 1 d x9 F rd9 !
experimental error, but more experimental data are needed in
order to verify the reliability of calculated thermodynamic 2 ~ 12x !~ F rx8 1 t 8x F rt8 1 d 8x F rd8 ! . ~A6!
properties especially in the single-phase regions.
Partial differentials are abbreviated as in Table 3. Reduced
variables for the saturated vapor are given by
7. Acknowledgments
V̄ n ~ y ! T n~ y !
This work has been carried out at NIST from April 1995 d 95 and t 9 5 ;
V̄ 9 T
to March 1996. R. T.-R. thanks W. M. Haynes and M. O.
McLinden for the opportunity to work at NIST as a guest those of the saturated liquid are
researcher. Partial support from the U.S. Department of En-
ergy, Geothermal Division, is gratefully acknowledged. V̄ n ~ x ! T n~ x !
d 85 and t 8 5 .
V̄ 8 T
8. Appendices
The derivatives of the reduced variables with respect to x are
8.1. Appendix A: Linearization of VLE Data abbreviated as

S D S D
The vapor–liquid equilibrium of a binary mixture is cal-
]t ]d
culated by solving simultaneously the four equations t x5 and d x 5 .
]x T
]x %
p5p ~ V̄ 8 ,T,x ! , ~A1!
Upon introducing the general structure of the residual
p5p ~ V̄ 9 ,T,y ! , ~A2! Helmholtz free energy of the binary mixture,
~ 12x ! w 1 ~ V̄ 8 ,T,x ! 5 ~ 12y ! w 1 ~ V̄ 9 ,T,y ! , ~A3! F r5 ~ 12x ! F r011xF r021DF r, ~A7!
x w 2 ~ V̄ 8 ,T,x ! 5y w 2 ~ V̄ 9 ,T,y ! . ~A4! the contributions of the departure function can be separated
w i is the fugacity coefficient of component i calculated as from the contributions of the pure fluid equations in Eqs.
given in Table 3; x and y are the liquid and vapor mole ~A5! and ~A6!:
fractions of component 2.
12x Z9
When an interim equation is known from which the VLE Y 1 5ln 1ln 2A5DF r9 2DF r8 2yDF rx9 1xDF rx8
can be calculated, these four equations can be linearized and 12y Z8
the information of a VLE measurement can be included in
the linear regression analysis to determine the functional 1 ~ d 9 2y d 9x ! DF rd9 2 ~ d 8 2x d 8x ! DF rd8 2y t 9x DF rt9
form of DF r. Regardless of the type of experimental VLE
data, (p,T,x), (p,T,x,y), or ( p,T,y) values, the interim 1x t x8 DF rt8 , ~A8!

J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998


76 R. TILLNER-ROTH AND D. G. FRIEND

x Z9 chosen. Compositions are given in terms of mass fractions j


Y 2 5ln 1ln 2B5DF r9 2DF r8 1 ~ 12y ! DF rx9 of ammonia. At temperatures below 0 °C, the solid region is
y Z8 encountered and no VLE exist for most tabulated vapor com-
positions. At high temperatures, the VLE region is restricted
2 ~ 12x ! DF rx8 1 @ d 9 1 ~ 12y ! d 9x # DF rd9
by the critical line. In the single-phase region, properties
2 @ d 8 1 ~ 12x ! d x8 # DF rd8 1 ~ 12y ! t x9 DF rt9 were calculated on isobars as functions of temperature and
mass fraction of ammonia.
2 ~ 12x ! t 8x DF rt8 . ~A9!
A and B are determined only from the pure fluid equations
and are given by
9. References
A5F r019 2F r018 1 ~ d 9 2y d 9x !@~ 12y ! F r01,d 9 1yF r02,d 9 # 1
A. I. Kalina and M. Tribus, Thermodynamics of the Kalina Cycle and the
Need for Improved Properties Data, Proceedings of the 12th International
2 ~ d 8 2x d 8x !@~ 12x ! F r01,d 8 1xF r02,d 8 # Conference Properties of Water and Steam ~IAPWS!, edited by H. J.
White et al. ~Begell House, 1995!, p. 841.
2y t x 8 @~ 12y ! F r01,t 9 1yF r02,t 9 #
2
F. C. Kracek, J. Phys. Chem. 34, 499 ~1930!.
3
G. Scatchard, L. F. Epstein, J. Warburton, Jr., and P. J. Cody, J. ASRE 53,
413 ~1947!.
1x t 8x @~ 12x ! F r01,t 8 1xF 02,
r
t8# , ~A10! 4
R. A. Macriss, B. E. Eakin, R. T. Ellington, and J. Huebler, Physical and
and Thermodynamic Properties of Ammonia–Water Mixtures, ~Inst. Gas
Tech., Chicago, Illinois, 1964!, Res. Bull. No. 34.
B5F r029 2F r028 1 @ d 9 1 ~ 12y ! d 9x #@~ 12y ! F r01,d 9 1yF r02,d 9 #
5
S. Schulz, Equations of state for the system ammonia-water for the use
with computers, Proceedings of the XIIth International Congress. Refrig-
2 ~ d 8 1 ~ 12x ! d 8x !@~ 12x ! F r01,d 8 1xF r02,d 8 # eration, Washington, 1971, Vol. 2, p. 431.
6
B. Ziegler and C. Trepp, Int. J. Refrig. 7, 101 ~1984!.
7
O. M. Ibrahim and S. A. Klein, ASHRAE Trans. 99, 1495 ~1993!.
1 ~ 12y ! t 9x @~ 12y ! F r01,t 9 1yF r02,t 9 # 8
B. I. Lee and M. G. Kesler, AIChE J. 21, 510 ~1975!.
9
Y. M. Park and R. E. Sonntag, ASHRAE Trans. 97, 150 ~1991!.
2 ~ 12x ! t x @~ 12x ! F r01,t 8 1xF r02,t 8 # . ~A11!
8 10
Y. M. El-Sayed and M. Tribus, Proc. ASME Ann. Winter Meet. 1, 89
~1985!.
The two quantities Y 1 and Y 2 are linear with respect to the 11
T. J. Edwards, G. Maurer, J. Newman, and J. M. Prausnitz, AICHE J. 24,
coefficients of the departure function as long as a linear com- 966 ~1978!.
bination of terms is used for DF r and as long as all variables 12
F. Kurz, Dissertation, Universität Kaiserslautern, 1994.
t , d and the composition of the saturated liquid and saturated V. Abovsky, Fluid Phase Equil. 116, 170 ~1996!.
13
14
D. G. Friend and A. L. Olson, Standard thermophysical properties of the
vapor are given. ammonia-water binary fluid, Proceedings of the 12th International Con-
For (p,T,x) properties it proved to be effective to calcu- ference Properties of Water Steam, 1995, edited by H. J. White et al.
late V̄ 8 for the experimental values of p, T, and x from an ~Begell House, 1995!, p 854.
15
K. E. Herold, K. Han, and M. J. Moran, ASME Proc. AES 4, 65 ~1988!.
interim Helmholtz equation of state. The vapor mole fraction 16
Z. Duan, N. Mo” ller, and J. H. Weare, J. Solid State Chem. 25, 43 ~1996!.
y is obtained from a flash calculation for given p and T; V̄ 9 17
J. Patek and J. Klomfar, Int. J. Refrig. 18, 228 ~1995!.
18
is calculated with the same equation using the experimental R. Tillner-Roth, Die thermodynamischen Eigenschaften von R 152a, R
134a und ihren Gemischen - Messungen und Fundamentalgleichungen,
p and T, and the calculated mole fraction y. For (p,T,x,y)
Forsch.-Ber. DKV No. 41 ~DKV, Stuttgart, 1993!.
properties, only V̄ 8 and V̄ 9 have to be calculated using the 19
R. Tillner-Roth and D. G. Friend, J. Phys. Chem. Ref. Data 27, 45 ~1998!.
20
experimental (p,T,x) or ( p,T,y). Although the calculated M. R. Moldover, J. P. M. Trusler, T. J. Edwards, J. B. Mehl, and R. S.
Davis, J. Res. National Bur. Stand. 93, 85 ~1988!.
properties may not be very reliable during the first step, their 21
A. Prub and W. Wagner, Eine neue Fundamentalgleichung für das fluide
accuracy increases with the accuracy of the interim equation Zustandsgebiet von Wasser für Temperaturen von der Schmelzlinie bis zu
of state. The quantities Y 1 and Y 2 have to be recalculated 1273 K bei Drücken bis zu 1000 MPa, Fortschr.-Ber. VDI 6, No. 320
whenever an improved equation has been found, and a new ~VDI, Düsseldorf, 1995!.
22
R. Tillner-Roth, F. Harms-Watzenberg, and H. D. Baehr, Eine neue Fun-
search for the departure function is initiated. Recalculation of damentalgleichung für Ammoniak, Proceedings of the 20th DKV-Tagung
the quantities A and B is also necessary whenever the coef- Heidelberg, Germany, Vol. II, 1993, p. 167.
23
ficients of the reducing functions T n (x) and V̄ n (x) have been C. L. Sassen, R. A. C. van Kwartel, H. J. van der Kooi, and J. de Swaan
Aarons, J. Chem. Eng. Data 35, 140 ~1990!.
readjusted. 24
S. Postma, Rec. Trav. Chim. 39, 515 ~1920!.
25
H. Mollier, Z. VDI 52, 1315 ~1908!.
8.2. Appendix B: Tables of Thermodynamic
26
E. P. Perman, J. Chem. Soc. 79, 718 ~1901!.
27
T. M. Smolen, D. B. Manley, and B. E. Poling, J. Chem. Eng. Data 36,
Properties 202 ~1991!.
28
J. Polak and B. C. Y. Lu, J. Chem. Eng. Data 20, 182 ~1975!.
Tables of thermodynamic properties are provided for the 29
M. Iseli, Dissertation, ETH Zürich, 1985.
two-phase envelope, Table 4, and for the single-phase re- 30
B. H. Jennings, ASHRAE Trans. 71, 21 ~1965!.
gion, Table 5. For the two-phase region, pressure, density,
31
F. Harms-Watzenberg, Fortschr.-Ber. VDI 3, No. 380 ~VDI, Düsseldorf,
1995!.
enthalpy, and entropy are given as functions of temperature 32
M. Ellerwald, Dissertation, Universität Dortmund, 1981.
and composition. Values are listed for the bubble-point curve 33
H. J. Staudt, Dissertation, Universität Kaiserslautern, 1984.
and for the dew-point curve. Mass based units have been 34
K. Zinner, Z. Kälteind. 41, 21 ~1934!.

J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998


HELMHOLTZ FREE ENERGY FORMULATION OF THE MIXTURE ˆWATER1AMMONIA‰ 77

35 41
D. L. Hildenbrand and W. F. Giauque, J. Am. Chem. Soc. 75, 2811 W. Wagner, Eine mathematisch statistische Methode zum Aufstellen ther-
~1953!. modynamischer Gleichungen - gezeigt am Beispiel der Dampfdruckkurve
36
J. P. Chan and W. F. Giauque, J. Phys. Chem. 68, 3053 ~1964!. reiner fluider Stoffe, VDI Fortschr.-Ber. 3, No. 39 ~VDI, Düsseldorf,
37
M. Wrewsky and A. Kaigorodoff, Z. Phys. Chem. 112, 83 ~1924!. 1974!.
38 42
P. C. Gillespie, W. V. Wilding, and G. M. Wilson, AIChE Symp. Ser. 83, M. Bischoff, Diplom thesis, Universität Hannover, 1988.
97 ~1987!. 43
J. Ahrendts and H. D. Baehr, Forsch. Ing.-Wesen 45, 1 ~1979!.
39
J. C. Rainwater and R. Tillner-Roth ~unpublished!. 44
K. Dvorak and T. Boublik, Coll. Czech. Chem. Commun. 28, 1249
40
J. E. Dennis, M. Gay, and R. E. Welsch, An Adaptive Nonlinear Least- ~1963!.
45
Squares algorithm, Tech. Summary Report 2010, Math. Res. Center, H. H. King, J. L. Hall and G. C. Ware, J. Am. Chem. Soc. 52, 5128
Madison, 1979. ~1930!.

J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998


78 R. TILLNER-ROTH AND D. G. FRIEND
TABLE 4. Saturation properties of $water1ammonia%
J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998
HELMHOLTZ FREE ENERGY FORMULATION OF THE MIXTURE ˆWATER1AMMONIA‰ 79
TABLE 4. Saturation properties of $water1ammonia%—Continued
J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998
80 R. TILLNER-ROTH AND D. G. FRIEND
TABLE 4. Saturation properties of $water1ammonia%—Continued
J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998
HELMHOLTZ FREE ENERGY FORMULATION OF THE MIXTURE ˆWATER1AMMONIA‰ 81
TABLE 4. Saturation properties of $water1ammonia%—Continued
J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998
82 R. TILLNER-ROTH AND D. G. FRIEND
TABLE 4. Saturation properties of $water1ammonia%—Continued
J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998
HELMHOLTZ FREE ENERGY FORMULATION OF THE MIXTURE ˆWATER1AMMONIA‰ 83
TABLE 4. Saturation properties of $water1ammonia%—Continued
J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998
84 R. TILLNER-ROTH AND D. G. FRIEND
TABLE 4. Saturation properties of $water1ammonia%—Continued
J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998
HELMHOLTZ FREE ENERGY FORMULATION OF THE MIXTURE ˆWATER1AMMONIA‰ 85
TABLE 4. Saturation properties of $water1ammonia%—Continued
J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998
86 R. TILLNER-ROTH AND D. G. FRIEND
TABLE 5. Properties of $water1ammonia% in the one-phase region
J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998
HELMHOLTZ FREE ENERGY FORMULATION OF THE MIXTURE ˆWATER1AMMONIA‰ 87
TABLE 5. Properties of $water1ammonia% in the one-phase region—Continued
J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998
88 R. TILLNER-ROTH AND D. G. FRIEND
TABLE 5. Properties of $water1ammonia% in the one-phase region—Continued
J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998
HELMHOLTZ FREE ENERGY FORMULATION OF THE MIXTURE ˆWATER1AMMONIA‰ 89
TABLE 5. Properties of $water1ammonia% in the one-phase region—Continued
J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998
90 R. TILLNER-ROTH AND D. G. FRIEND
TABLE 5. Properties of $water1ammonia% in the one-phase region—Continued
J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998
HELMHOLTZ FREE ENERGY FORMULATION OF THE MIXTURE ˆWATER1AMMONIA‰ 91
TABLE 5. Properties of $water1ammonia% in the one-phase region—Continued
J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998
92 R. TILLNER-ROTH AND D. G. FRIEND
TABLE 5. Properties of $water1ammonia% in the one-phase region—Continued
J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998
HELMHOLTZ FREE ENERGY FORMULATION OF THE MIXTURE ˆWATER1AMMONIA‰ 93
TABLE 5. Properties of $water1ammonia% in the one-phase region—Continued
J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998
94 R. TILLNER-ROTH AND D. G. FRIEND
TABLE 5. Properties of $water1ammonia% in the one-phase region—Continued
J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998
HELMHOLTZ FREE ENERGY FORMULATION OF THE MIXTURE ˆWATER1AMMONIA‰ 95
TABLE 5. Properties of $water1ammonia% in the one-phase region—Continued
J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998
96 R. TILLNER-ROTH AND D. G. FRIEND
TABLE 5. Properties of $water1ammonia% in the one-phase region—Continued
J. Phys. Chem. Ref. Data, Vol. 27, No. 1, 1998

You might also like