Вы находитесь на странице: 1из 10

Fluid Phase Equilibria 322323 (2012) 5665

Contents lists available at SciVerse ScienceDirect

Fluid Phase Equilibria


journal homepage: www.elsevier.com/locate/fluid

Viscosity prediction for natural gas processing applications


H. Motahhari, M.A. Satyro, H.W. Yarranton
Department of Chemical and Petroleum Engineering, University of Calgary, 2500 University Dr. NW, Calgary, AB, Canada T2N 1N4

a r t i c l e

i n f o

Article history:
Received 2 February 2012
Received in revised form 1 March 2012
Accepted 1 March 2012
Available online 14 March 2012
Keywords:
Viscosity
Natural gas
Equation of state
Expanded Fluid viscosity correlation
Simulation

a b s t r a c t
The Expanded Fluid (EF) viscosity correlation was used to model the viscosity of mixtures commonly
encountered in natural gas processing. This correlation provides viscosity values as a function of uid density and characterizes each pure compound with three uid-specic parameters: c2 , so and c3 , when using
experimental densities (Version 1) and two parameters, c2 and so , when using a cubic equation of state
(Version 2). In particular, the original EF correlation for hydrocarbons was adapted for non-hydrocarbon
components, including: carbon dioxide, hydrogen sulde, nitrogen, helium, water, methanol, ethylene
glycol, diethylene glycol and triethylene glycol. A temperature dependency was introduced for parameter
c2 for components with signicant hydrogen bonding such as water and methanol. For all other components, a xed c2 was found to be adequate. Both versions of the correlation were t to experimental data
for the non-hydrocarbon components with overall average absolute relative deviation (AARD) below 6%.
The viscosities of several sweet and sour natural gas mixtures were predicted with overall AARDs of 6.3%
and 5.1% for Versions 1 and 2, respectively. The viscosities of aqueous solutions both of methanol and of
glycols were also modeled. A binary interaction parameter was required to t the data for all aqueous
mixtures in Version 2 but only for mixtures of water and methanol in Version 1. The overall AARDs of the
calculated viscosities of the aqueous solutions of methanol and glycols by Versions 1 and 2 were 6% and
8.7%, respectively.
2012 Elsevier B.V. All rights reserved.

1. Introduction
The accurate design and operation of natural gas processes
requires the prediction of the phase behavior and transport properties of natural gases of different phases and compositions. Natural
gases include non-hydrocarbon components such as water, hydrogen sulde, carbon dioxide, nitrogen, and helium [1]. In addition,
gas dehydration may involve solvents such as ethylene glycol (EG),
diethylene glycol (DEG) and triethylene glycol (TEG). This study
focuses on the prediction of the viscosity of natural gases and glycols for process and reservoir simulation.
In general, viscosity is required as a function of temperature, pressure, and composition. A suitable viscosity model for
implementation in simulators must: (1) trace the full range of single phase properties in the gas, liquid, critical, and supercritical
regions; (2) be fast; (3) predict both pure component and mixture
viscosities; (4) be continuous across the critical point and consistent with the phase equilibrium model.
Although numerous viscosity models are available in the literature, no single model is applicable consistently to all natural
gas processing applications. The viscosity of the glycolwater

Corresponding author. Tel.: +1 403 220 6529.


E-mail address: hyarrant@ucalgary.ca (H.W. Yarranton).
0378-3812/$ see front matter 2012 Elsevier B.V. All rights reserved.
doi:10.1016/j.uid.2012.03.006

solutions are commonly calculated using the correlations provided


by the glycol manufacturers in the form of graphical charts or equations [1]. Unfortunately little information about the accuracy of
these correlations is available. Natural gas viscosity is commonly
calculated using several empirical correlations or semi-theoretical
models which are reviewed below.
A commonly used empirical correlation is the LGE
(LeeGonzalezEakin) correlation [2], along with its variations
[3,4], which is based on experimental viscosity data of natural
gases. This correlation is valid to within 3% of the measured
values in the range of the temperature, density and composition
of the natural gas mixtures used in its development.
Semi-theoretical models have a theoretical basis but include
parameters that must be determined from experimental data. The
Lucas viscosity model [5] is based on the kinetic theory of gases at
dilute conditions and employs empirical modications for dense
gas conditions. However, the model is not suitable for liquids such
as gas condensates. The TRAPP program [6] is based on the twoparameter corresponding states theory originally developed by
Hanley et al. [79]. It is fully predictive for selected pure hydrocarbons and simple non-hydrocarbon components and their mixtures.
SUPERTRAPP is a modication of TRAPP which can be tuned for
better results for cyclic hydrocarbon compounds and branchedalkanes. Unlike the Lucas model, the TRAPP models can predict
viscosity in the liquid phase as well as gas phase; however, their

H. Motahhari et al. / Fluid Phase Equilibria 322323 (2012) 5665

applicability for components like water has not yet been investigated.
Another, recently developed, semi-theoretical model is Friction
Theory (F-theory) [10,11] which relates viscosity to the friction
forces between uid layers arising from internal pressure. A modication of the one-parameter version of this model coupled with a
cubic equation-of-state predicted the viscosity of natural gas mixtures to within 10% of the experimental values [12]. In addition, the
model coupled with a cubic equation-of-state was tested for modeling the viscosity of selected pure alcohols with 7 tuned parameters
[13] for each alcohol. However, the only adaption of the model
for water was based on the WagnerPrauss reference equation of
state [14] with 18 tuned adjustable parameters [15]. To the authors
knowledge, a single consistent application of F-theory has not yet
been demonstrated for all aspects of the natural gas processing.
Another approach is to use viscosity mixing rules to calculate the
viscosity of hydrocarbon mixtures as well as aqueous solutions of
glycols. However, as Poling et al. [16] emphasized, these models are
limited to mixtures in which the components are at reduced temperatures less than 0.7. An example is the GrunbergNissan mixing
rule [17] which is modication of Arrhenius mixing rules with
generalized binary interaction parameters [18]. For glycol/alcohol
mixtures with water, the McAllister mixing rule [19,20] can be used
but has multiple adjustable parameters for ne tuning of the calculations to the experimental values. Teja and Rice [21] developed a
mixing rule model based on corresponding states theory with two
reference uids, and used the pure components as the reference
for each binary mixture. With one adjustable binary interaction
parameter, they tted the viscosity of binary mixtures, including hydrocarbon mixtures and aqueous solutions of alcohols, with
maximum average deviations of 23% [21]. The model was applied
to aqueous solutions of glycols using two adjustable binary interaction parameters with deviations within 8.3% of the measured values
[22].
Recently, Yarranton and Satyro [23] introduced the Expanded
Fluid (EF) correlation for viscosity modeling of hydrocarbons. This
correlation relates viscosity to the uid density. The method was
rst developed with measured density data [23] but was later
extended [24] for use with calculated densities from the Advanced
PengRobinson (APR) equation of state (EoS) [25]. Both versions of
the correlation were successfully applied to model the viscosity of
the pure hydrocarbons including n-alkanes, branched alkanes, aromatics and cyclic over a wide range of pressure and temperature
in gas and liquid phases. The correlation also predicted the viscosity of over 100 mixtures of hydrocarbon components with overall
average deviations less than 8%, in both gas and liquid phases [26].
Although the EF correlation tted the experimental viscosity data of
pure hydrocarbons and predicted the viscosity of hydrocarbon mixtures, its applicability for natural gas processing applications with
non-hydrocarbon constituents has not been studied. Since, the correlation was built empirically based on hydrocarbon viscosity data;
its validity for non-hydrocarbons has yet to be determined.
The objective of this study is to investigate the applicability of
the EF correlation for the mixtures commonly dealt with in the
natural gas processing applications. First, the correlation is tested
for the pure non-hydrocarbon components including carbon dioxide, hydrogen sulde, nitrogen, helium, water, methanol, ethylene
glycol, diethyelene glycol and triethylene glycol. Then, the applicability of the correlation is studied for sweet and sour natural gas
mixtures and aqueous solutions of methanol and glycols. Both Version 1 (based on measured density) and Version 2 (based on EoS
density) are evaluated. Version 2 is intended for simulator applications. However, Version 2 is not well suited for testing model
adaptations for non-hydrocarbons because errors in the EoS densities may skew the interpretation. Therefore, any adaptations to the
model are made with Version 1 and then applied to Version 2.

57

2. Expanded Fluid viscosity correlation


The EF viscosity correlation is based on the empirical observation that, as the uid expands, its viscosity decreases due to the
greater distance between the molecules [27]. In this correlation,
the viscosity of the uid is formulated as a departure function from
the dilute gas viscosity as follows [23]:
 o = c1 (exp(c2 ) 1)

(1)

where  and o are the uid and its dilute gas viscosity, c1 and c2
are tting parameters. is the correlating parameter between the
viscosity and the uid expansion as follows:
=

(2)

exp{(s /) 1} 1

where  is the uid density, s is the density of the uid in the


compressed state and n is an empirical exponent which improves
the predictions near the critical region. By denition, the molecules
of the uid at compressed state are too packed to move in viscous
ow; therefore, the viscosity approaches innity.
There are two versions of the correlation: Version 1 uses measured density values [23]; Version 2 uses density values calculated
with the Advanced PengRobinson (APR) equation-of-state (EoS)
[24]. A detailed description of this cubic equation of state can be
found elsewhere [25] and a brief review is provided in Appendix A.
For both versions of the correlation, the parameters n and c1 have
xed values:

Version 1:
Version 2:

n = 0.65 [dimensionless]
n = 0.4872 [dimensionless]

c1 = 0.165 [mPa s]
c1 = 0.4214 [mPa s]

To improve viscosity predictions at elevated pressures, the compressed state density is formulated as a function of pressure:
Version 1 :

s =

so
exp(c3 P)

(3)

Version 2 :

s =

so
1 c4 (1 exp(c3 P))

(4)

where so is the compressed state density in vacuum, c3 (in kPa1 )


and c4 are pressure dependency parameters, and P is pressure in
kPa. In Version 1, c3 is a tting parameter. In Version 2, the parameters c3 and c4 are correlated to molecular weight [24]:
c3 = 1.435 106 MW 0.4267

[kPa1 ]

(5)

For MW 97 : c4 = 0.015 + 0.00042 |50 MW | [dimensionless]


(6)
For MW < 97 :

c4 = 0.035 [dimensionless]

(7)

After xing and correlating the parameters, the following uidspecic parameters remain for each component:
Version 1:
Version 2:

c2 , c3 and so
c2 and so

Note, for Version 1, correlations were developed for parameter


c2 for pure hydrocarbons [23] reducing the number of adjustable
parameters to two. The uid-specic parameters for a wide range
of hydrocarbons were determined in previous works [23,24,26].
The inputs to the correlation are the uid density (measured or
calculated by APR EoS), pressure, and the dilute gas viscosity. The
dilute gas viscosity of the uid is determined from the following
empirical correlation [28]:
o = A + BT + CT 2 + DT 3

(8)

where A, B, C and D are tting parameters specic for each pure


component. The numerical values of these parameters are taken
from Yawss handbook [28].

58

H. Motahhari et al. / Fluid Phase Equilibria 322323 (2012) 5665


Table 1
Summary of the pure non-hydrocarbon experimental data used in this study.

2.1. Mixing rules


The uid-specic parameters of the single-uid mixture are calculated using the following mixing rules [26]:

o
s,mix
=

nc
nc 

wi wj
i=1 j=1

 wi wj
c2,mix
=
o
2
s,mix
nc

nc

i=1 j=1

c3,mix =

1
1
o + o
s,i
s,j

c2,j
c2,i
o + o
s,i
s,j

1
(1 ij )


(1 ij )

nc

1
w
i

i=1

(9)

(10)

(11)

c3,i

where nc is the number of components in the mixture, wi is the


mass fraction of the component i in the mixture, and ij is a binary
interaction parameter. Note, Eq. (11) is only used for Version 1 of
the correlation.
The binary interaction parameter can be used to tune the correlation when experimental viscosity data for the mixture are
available. The default value of ij is zero. It was shown [26] that
Version 1 of the correlation predicted the viscosity of hydrocarbon
mixtures with an AARD of 5% with ij set to zero. However, non-zero
values of ij were required to tune Version 2 [26]. Based on viscosity data for 79 binary hydrocarbon mixtures, the following general
correlations were developed for binary interaction parameters for
use with Version 2 [26]:
ijo = 0.319fMW

(12)

0.3
0.3
)fMW + 0.04fWK
ijo = (0.244 0.45fWK

(13)

where

ijo

is the reference ij at atmospheric condition and:

fMW = 1

fWK = 1

MWi MWj

(14)

MWi + MWj

WKi WKj

(15)

WKi + WKj

where MWi and WKi are the molecular weight and Watson-K factor
values of component i of mixture, respectively. Either of Eqs. (12)
or (13) were shown to provide mixture viscosity predictions with
an AARD of 8% for Version 2. However, Eq. (12) performed better
for mixtures of components with signicantly different molecular weights, and Eq. (13) provided better results for mixtures of
n-alkanes with aromatic or naphthenic components [26].
The binary interaction parameter at elevated pressures is given
by [26]:
ij = ijo exp((P 101.325))

(16)

where P is pressure in kPa, and is the decay rate of ij from


atmospheric pressure to high pressure conditions. The optimized
value of for use with Eqs. (12) and (13) are 1.65 105 and
1.31 105 kPa1 , respectively.
The mixture dilute gas viscosity is calculated using Wilkes
method [29] as follows:
o,mix =

 xi o,i

i

(17)

x
j j ij

where
ij =

[1 + (o,i /o,j )

0.5

(MWi /MWj )

[8(1 + MWi /MWj )]

0.5

0.25 2

(18)

Component

NP

Temperature (K)

Max. pressure (kPa)

Source

Hydrogen sulde
Carbon dioxide
Helium
Nitrogen
Water
Methanol
Ethylene glycol
Diethylene glycol
Triethylene glycol

18
520
376
1490
1795
575
95
35
24

190483
2031100
14.2918.3
631989
2561139
176.1511.1
273428
273428
273428

4400
450,000
83,000
194,000
389,000
200,000
atm.
atm.
atm.

[3036]
[3739]
[37]
[37]
[37]
[37]
[37]
[37]
[37]

where xi and o,i are the mole fraction and dilute gas viscosity of
the component i of the mixture, respectively.
3. Results and discussion
The correlation is rst tested for pure non-hydrocarbon components and a modication for hydrogen bonding uids is introduced.
Then, the validity of the correlation is tested by comparing predicted viscosities of mixtures commonly found in natural gas
processes to experimental values.
3.1. Pure components
There are two categories of pure non-hydrocarbon components encountered in natural gas processing. The rst category is
simple non-hydrocarbons which are usually constituents of produced natural gas and include nitrogen, carbon dioxide, hydrogen
sulde and helium. These components are either non-polar and
interact via London dispersion forces or polar with polarpolar
interactions. The second category is associating non-hydrocarbon
components normally used to treat natural gas streams such as
methanol and glycols. Also included in this category is water which
is a constituent of natural gas but is also an associating species.
The molecules of these components are polar and form hydrogen
bonds with neighboring molecules due to the presence of an O H
group. The application of the EF correlation to each category of
non-hydrocarbons is presented below.
In both cases, the required input properties for the pure components are the dilute gas viscosity and the density at given pressure
and temperature. Table 1 summarizes the viscosity data compiled
for each component. Density data were also compiled from the
same sources but were only available for more limited ranges of
pressure and temperature. Version 1 of the correlation is tested
rst using the datasets which include both measured density and
viscosity. Any conceptual modications to the correlation are introduced at this point. Then, Version 2 of the correlation is tested using
densities calculated with the APR EoS.
3.1.1. Simple non-hydrocarbons
Version 1 of the correlation was tted to data by adjusting the
uid-specic parameters of c2 , c3 , and so to the values given in
Table 2. Table 3 provides a summary of the average absolute deviation (AAD), the maximum absolute deviation (MAD), the average
absolute relative deviation (AARD), the maximum absolute relative
deviation (MARD), and the bias of the tted correlation. The maximum deviation was observed for helium at very low temperatures.
The overall AARD of 4.2% and MARD of 34% are both comparable
to the deviations reported for the application of the correlation
to pure hydrocarbon components. Therefore, it is concluded that
the correlation framework, based on the non-polar hydrocarbon
components data, is sufcient for these simple non-hydrocarbons.
The uid specic parameters, c2 and so , were then determined
for Version 2, Table 2. Note that the general correlations of Eqs.

H. Motahhari et al. / Fluid Phase Equilibria 322323 (2012) 5665

59

Table 2
Fluid specic parameters to use with Versions 1 and 2 of the correlation for pure non-hydrocarbon components.
Component

Version 1

Version 2

c2
Hydrogen sulde
Carbon dioxide
Helium
Nitrogen
Water
Methanol
Ethylene glycol
Diethylene glycol
Triethylene glycol

b 103 (K1 )

0.188
0.236
0.0517
0.1147
0.1463
0.1463
0.23
0.2487
0.2977

c3 106 (kPa1 )

1194.59
1617.66
300
1012.39
1197
1045
1193.69
1195.04
1214.28

23.1
4.22

99.519
0.6301

so (kg/m3 )

c2
0.0437
0.0644
0.0149
0.0244
0.0674
0.0674
0.1274
0.1147
0.1417

0.187
0.1
0.1
0.3
0.449

0.6

Viscosity (mPa.s)

0.4

0.3

Viscosity (mPa.s)

Version 1
Version 2
260 K
280 K
308 K

0.2

0.1

b 103 (K1 )

35.090
2.672

50000

100000 150000 200000

1092.76
1571.97
286.16
938.41
1368.18
1156
1201.53
1175.86
1189.40

MD-sat. curve
exp-sat. liquid
exp-gas (1 bar)
Version 1
Version 2

0.06

a
0

18.5
10.2

so (kg/m3 )

b
0.006
100

200

300

400

500

600

Temperature (K)

Pressure (kPa)

Fig. 1. EF correlation tted to experimental data of: (a) compressed carbon dioxide, data from [3739]; (b) hydrogen sulde on co-existence curve and 1 bar pressure, data
from [3137].

(5)(7) were used for the pressure dependency parameters. A summary of the deviations for Version 2 of the correlation is provided
in Table 3. Version 2 of the correlation ts the data with the same
accuracy of the Version 1 with overall AARD and MARD of 4.2% and
31%, respectively. However, the accuracy of the Version 2 at elevated pressures is slightly less than Version 1 because the pressure
dependent parameters are correlated in Version 2 but adjustable
in Version 1. The deviation at higher pressures is shown for carbon
dioxide in Fig. 1a. The t of Version 2 can be improved if the pressure dependency parameters (c3 and c4 ) were adjusted as well, but
the generality of the correlation for mixtures would be lost.
Note that both versions of the correlation were tted to limited
experimental data for hydrogen sulde. Much of these data were
calculated at the liquid-vapor coexistence curve using molecular
dynamics (MD) simulation [36]. The calculated values are reportedly in good agreement with experimental data [40]. The only
exceptions were saturated gas data which are clearly inconsistent
with the experimental dilute gas data and were excluded from the
analysis. Fig. 1b shows that the t of both versions of the correlation
to the viscosity data of hydrogen sulde on the co-existence curve.

3.1.2. Associating non-hydrocarbons


Unlike the simple non-hydrocarbon components, the tting
of Version 1 of the correlation to viscosity data of associating non-hydrocarbon components was not straightforward.
Although Version 1 could t the limited atmospheric data
points of the glycols with AARD below 5%, it failed to adequately t water data. Less signicantly, the tted correlation for
methanol over-predicted the viscosity in the vicinity of the critical
region.
These results are not surprising because there is hydrogen bonding in these uids while the correlation was originally developed
for molecules interacting via dispersion forces only. The strength,
length, and number of hydrogen bonds between the molecules
depend on pressure and temperature [41]. The response of viscosity to uid expansion will be different than for a uid dominated by
dispersion forces. The effect of hydrogen bonding is most obvious
in water. For instance, the density behavior of saturated water is
different than other liquids with a maximum occurring at 4 C and
negative thermal expansions at temperatures below 4 C. Also, at
temperatures above 50 C the viscosity of water increases under

Table 3
Summary of the deviations for Versions 1 and 2 of the correlation for pure non-hydrocarbon components.
Component

Hydrogen sulde
Carbon dioxide
Helium
Nitrogen
Water
Methanol
Ethylene glycol
Diethylene glycol
Triethylene glycol

Version 1

Version 2

MARD (%)

AARD (%)

Bias (%)

MARD (%)

AARD (%)

Bias (%)

15
9.7
34
24
48
17
12
23
7.1

3.5
3.3
4.1
5.8
5.3
2.1
3.0
5.0
4.5

1.3
2.5
3.1
4.5
2.6
0.2
0.6
0.1
0.3

13
14
31
20
25
30
54
41
41

3.3
5.6
5.0
2.8
1.9
5.4
5.8
8.6
14

2.7
3.3
0.7
0.6
0.2
0.4
0.8
0.0
0.0

60

H. Motahhari et al. / Fluid Phase Equilibria 322323 (2012) 5665

10

10
256 K
263 K
273 K
288 K
303 K
323 K
343 K
373 K
473 K
673 K
Version 1
Version 2

exp data
Version 2

0.1

Viscosity (mPa.s)

Viscosity (mPa.s)

Version 1

0.1

a
0.01
200

0.01
300

400

500

600

700

200000

Temperature (T)

400000

600000

Pressure (kPa)

Fig. 2. EF correlation tted to experimental data [37] of water: (a) saturated liquid; (b) compressed uid.

c2 = c2 + Kc2 exp(c2 T )

(19)

where c2 and Kc2 and c2 are tting parameters. The default value
for Kc2 and c2 is zero for all components except those with signicant hydrogen bonding such as water and methanol. Therefore, by
denition for all hydrocarbons and simple non-hydrocarbons, the
value of c2 is equal to the value of the non-temperature dependent
c2 .
Version 1 of the correlation was then tted to the experimental data of water and methanol by adjusting the parameters so , Kc2
and c2 with c2 xed at 0.1463. Version 1 t the limited available data of glycols with a xed c2 (Kc2 and c2 set to zero). The
uid specic parameters for the associating non-hydrocarbons and
a summary of the deviations of the tted correlation are given in
Tables 2 and 3, respectively. The AARD are less than 5% while the
MARD of 48% occurs for water. The correlation ts the data of water
on the co-existence curve with a maximum deviation of 14%, Fig. 2a.
Fig. 2b shows that the maximum MARD occurs for compressed liquid water at temperatures below 50 C whereas the water viscosity
decreases under compression. The correlation is not able to model
this behavior.
Having established the new temperature dependent expression
for parameter c2 , Version 2 of the correlation was then tted to
the experimental data of pure associating non-hydrocarbon uids.
Tables 2 and 3 report the uid-specic parameters and model deviations, respectively, for Version 2. As with Version 1, the parameters
so , Kc2 and c2 were adjusted for methanol and water with c2 xed
at 0.0647. Only so and c2 were adjusted to t the glycol data (Kc2
and c2 were set to zero). The overall AARD and MARD for Version
2 are 7.1% and 54%, respectively.
While the highest deviations in Version 1 were observed for
water, they are observed for glycol in Version 2, Table 3. The AARDs
for the glycols are below 14% with Version 2 but below 5% with
Version 1. For instance, for diethylene glycol, the Version 2 model
viscosities deviate at higher temperatures, Fig. 3. The main contribution to the deviations for the glycols is inaccurate density
predictions by APR EoS. Using measured densities as an input to

Version 2 reduces the overall AARD for the glycols from 9.5 to 7.7%;
that is, to nearly the same error as observed for Version 1.
Note, Version 2 of the correlation t the high pressure data of
water at temperatures below 50 C more accurately than Version
1, Fig. 2b. The improvement is attributed to the combined effects
of: (1) the form of the pressure dependency expression in Version 2
which provides a better t for water; and (2) the underestimation of
compressed liquid water density by the APR EoS which accidentally
compensates for the effect of the unusual density trends in water.
3.2. Mixtures
3.2.1. Natural gas mixtures
Table 4 provides the composition of 11 natural gases mixtures
[2,4346] used to evaluation the proposed correlation. Not included
in Table 4 but used in this study are data for sour natural gas mixtures from Elsharkawy [47]. This dataset consists of compositions
and single viscosity data points for 17 different gas samples at a
given temperature and pressure. To broaden the evaluation for sour
mixtures, 8 data points from an MD simulation [46] for mixtures of
methane, carbon dioxide and hydrogen sulde were also used.
Viscosity predictions were made with both versions of the
correlation where possible but Version 1 is limited to mixtures
with experimental density data. The correlation parameters for
the hydrocarbon components of these mixtures were taken from
[23,24] for Versions 1 and 2 of the correlation, respectively. The

exp data

100

Version 1
Version 2

Viscosity (mPa.s)

compression as found with other liquids; however, at lower temperatures its viscosity increases under compression [42].
A modication to the model was required to the model the viscosity of uids with signicant hydrogen bonding, such as water
and methanol. The parameter c2 is the proportionality of the uid
viscosity to its expansion and indirectly represents the effect of the
intermolecular forces on viscosity. Hence, a temperature dependent c2 is appropriate to model the temperature dependency of
hydrogen bonding and its effect on the viscosity. The following formulation of the temperature dependent c2 was determined to t
the data:

10

1
200

300

400

500

Temperature (T)
Fig. 3. The t of the Versions 1 and 2 of the correlation to experimental viscosity
data [37] of dietheylene glycol using xed c2 .

H. Motahhari et al. / Fluid Phase Equilibria 322323 (2012) 5665

15

15
G1
G5

G2
G6

G3
G7

G4
G8

5
0
-5
-10

40000

G3
G7

G4
G8

5
0
-5

-15
200

-15
20000

G2
G6

-10

a
0

G1
G5

10

Deviation (%)

10

Deviation (%)

61

60000

300

400

500

Temperature (K)

Pressure (kPa)

Fig. 4. Deviations of the predicted viscosity (Version 2) for natural gas mixtures versus: (a) pressure; (b) temperature.

Table 4
Composition of the natural gas mixtures used in this study.

N2
CO2
H2 S
He
n-C1
n-C2
n-C3
C4
C5
C6
C7
C8
C9
C10+

G1 [43]

G2 [44]

5.6
0.66

1.83

84.84
8.4
0.5

94.67
3.5

G3 [2]

G4 [2]

G5 [2]

G6 [2]

G7 [45]

G8 [45]

3.2

1.4
1.4

4.8
0.9

0.55
1.7

0.66
2.19

1.39
1.00

86.33
6.8
2.4
0.91
0.22
0.1
0.04

0.03
71.71
14
8.3
2.67
0.39
0.09
0.01

0.03
80.74
8.7
2.9
1.7
0.13
0.06
0.03

Viscosity (mPa.s)

parameters for non-hydrocarbon components were taken from


Table 2. Note, the sour natural gas mixtures of Elsharkawy [47]
contain plus-fractions of heavier hydrocarbons which were modeled as the equivalent n-alkanes based on their molecular weight.
Although, both general correlations for ij (Eqs. (12) and (13)) in
Version 2 were used in this study, only results with Eq. (12) are
shown since the results with Eq. (13) were not signicantly different.
Table 5 provides a summary of the AARD, MARD, and bias of the
predictions. For sweet natural gas mixtures (G1G8), Fig. 4 shows
that the predictions are in good agreement with the measured values over a broad range of conditions with overall AARDs of 6.4 and
2.8% for Versions 1 and 2, respectively. Some comparisons with
other models were possible. For instance, the predictions for G1

0.018

0.016

0.016

263 K
294 K
323 K
354 K
241 K
Version 1
Version 2

0.012
0.01
0.008
0.006
0

a
5000

10000

Pressure (kPa)

15000

0.02
92.20
4.34
0.54
0.33
0.07
0.06
0.04
0.00
0.00
0.00

G9 [46]

G10 [46]

G11 [46]

0.6

0.4

0.63
0.27

0.4

0.6

0.1

are comparable to the values estimated by Assael et al. [43] using


the relatively complex mixing rules of the VesoicWakeham viscosity model [48,49] coupled with the special density correlation
of AGA8-DX92 [50]. Also, the performance of Version 2 for the G7
and G8 mixtures is superior to the performance of the most of the
models studied by Langelandsvik et al. [45]. Among these models,
only LGE-3 (AARD of 1.4%) and SUPERTRAPP (AARD of 2.2%) provided predictions comparable to Version 2 (AARD of 1.9%) for these
two gas samples.
The predictions with Version 1 for mixtures G1G8 are slightly
less accurate than Version 2. The less accurate predictions occurred
where the uid transitions from dilute to dense gas behavior, as
shown in Fig. 5a and b for natural gas mixture G1 and methane
at 323 K, respectively. Version 1 of the correlation originally

0.018

0.014

0.01
80.01
9.31
4.96
2.00
0.51
0.18
0.12
0.04
0.01
0.01

91.46
3.1
1.4
1.17
0.28
0.26
0.08

Viscosity (mPa.s)

Component

0.014
0.012
exp data

0.01

Version 1
Version 2

0.008

0.006
0

5000

10000

15000

20000

Pressure (kPa)

Fig. 5. Comparison of the predictions from Versions 1 and 2 of the EF correlation with measured viscosity data of: (a) natural gas mixture G1 [43]; (b) methane at 323 K [37].

62

H. Motahhari et al. / Fluid Phase Equilibria 322323 (2012) 5665

Table 5
Summary of the deviations of the predictions by Versions 1 and 2 of the correlation for natural gas mixtures.
Version 1
MARD (%)
G1
11
8.1
G2
11
G3
G4
15
G5
14
G6
15
8.6
G7
13
G8
6.3
G9
G10
5.5
G11
6.8
Sour natural gas samples [47]

Version 2
AARD (%)

Bias (%)

MARD (%)

6.5
3.5
8.3
3.2
8.8
7.5
6.3
9.7

6.5
3.5
8.3
0.3
8.8
7.6
6.3
9.7

3.5

3.5

3.9
3.1
5.6
9.1
7.5
11
4.3
4.6
5.1
4.0
14
51

developed and tested for hydrocarbons mostly in liquid phases with


limited density and viscosity data near the critical region. Fine tuning of the parameters n and c1 could improve the results for Version
1 near the critical region but such an exercise is beyond the scope
of this study. In any case, the Version 1 predictions are still within
15% of the measured values.
For the sour gas mixtures (G9G11), Versions 1 and 2 predicted
the viscosity of the binary and ternary mixtures containing H2 S
with overall AARDs of 5.1% and 6.2%, respectively. All data of these
mixtures are synthetic values from molecular dynamics (MD) simulations [47]. The deviations were calculated by comparing the
correlation predictions to the mean of the simulated viscosity values by different intermolecular potential elds in MD simulations.
The MD simulated values were scattered with an average and maximum deviations of 2.9% and 7.2% around the mean values, which
are comparable to the deviations of predictions by EF correlation.
The predictions for gas samples of Elsharkawy [47] were made
only with Version 2 because density data were not available. As
reported in Table 5, the overall AARD for these mixtures is 20%,
which is considerably higher than other natural gas mixtures. The
higher deviations are attributable to the simplistic characterization
used for the heavy plus-fractions in these mixtures. Fig. 6 shows
that the deviation of the predictions increases as the mass fraction of the plus-fraction increases. The predictions for mixtures
with no plus-fraction were within 7% of the experimental values,
comparable to the sweet gas mixtures.

3.2.2. Associating mixtures


The proposed correlation was tested on aqueous solutions
of ethylene glycol, diethylene glycol triethyelene glycol and
methanol. Data of viscosity and density of all mixtures were compiled from NIST database [37] and are in atmospheric pressure
conditions. The temperature range of the data is 273450 K.
Mixtures of these associating uids are highly non-ideal and
have signicant excess volumes. Predictions from Version 2 using
untuned EoS density values are signicantly skewed due to the
errors in the calculated density. Therefore, it is essential to tune
the EoS density calculation to achieve accurate viscosity predictions. The APR equation of state was tuned to the experimental
density data by adjusting parameters Aij and Bij of the volume shift

AARD (%)

Bias (%)

1.8
1.8
3.1
4.1
4.4
3.5
1.7
2.1

0.7
1.8
2.8
2.1
4.3
3.1
1.4
2.1

9.6
20

9.6
9.3

0.18
0.16

Viscosity (mPa.s)

Mixture

0.14
0.12
0.1
0.08
0.06
0.04

exp data

0.02

Version 2

0.2

0.4

0.6

0.8

Mass Fraction of HC Plus-fraction


Fig. 6. Deviation of the predicted viscosity (Version 2) for sour natural gas mixtures
[47] as a function of the heavy hydrocarbon plus-fraction content.

mixing rule (Eq. (A.14) in Appendix A). The tted parameters are
given in Table 6.
Table 7 provides a summary of the AARD, MARD, and the bias
of the predictions of the correlation for these mixtures. Note, ij
was set to zero unless otherwise stated. The AARDs for Version 1
of the correlation are all less than 10% except for mixtures of water
and methanol with an AARD of 21%. The deviations for Version 2
with ij = 0 are considerably higher with an overall AARD of 26%.
The reduced accuracy of Version 2 for mixtures is consistent with
results previously obtained for the hydrocarbon mixtures [26].
Version 2 was then tted to the data using a binary interaction
parameter, Table 7, reducing the overall AARD to 8.7%. Unfortunately, the general ij correlations developed for the hydrocarbons
(Eqs. (12) and (13)) are not applicable for these mixtures of associating species. Both equations give ij values that increase with
increasing molecular weight difference. However, the optimized
ij values for glycols mixtures with water have the opposite trend.
Fig. 7a and b demonstrates the performance of the correlation
for aqueous solutions of diethylene glycol and methanol, respectively. Both versions of the correlation are qualitatively correct.
However, although Version 1 correctly predicts a maximum in the

Table 6
Adjusted volume translation parameters of APR EoS for the aqueous solutions of glycols and methanol.
Parameter

Ethyelene glycol(1) + water(2)

Diethyelene glycol(1) + water(2)

Triethyelene glycol(1) + water(2)

Methanol(1) + water(2)

A12
A21
B12
B21

0.219
0.194
4.16
2.76

0.288
0.240
1.40
1.13

0.348
0.279
0.150
3.04

0.0466
0.051
1.05
1.10

H. Motahhari et al. / Fluid Phase Equilibria 322323 (2012) 5665

63

Table 7
Summary of deviations of the predicted viscosity by the EF correlation for aqueous solutions of glycols and methanol.
Mixture

Version 1
ij

EG + water
DEG + water
TEG + water
Methanol + water

29
17
10
42
0.067 12

AARD (%)

Bias (%)

MARD (%)

AARD (%)

Bias (%)

ij

MARD (%)

AARD (%)

Bias (%)

9.8
4.6
5.6
21
4.0

6.4
0.4
3.3
21
2.3

62
62
64
52

28
26
25
28

26
25
18
29

0.069
0.066
0.055
0.179

40
24
40
19

11
6.8
13
3.9

3.4
0.7
0.7
0.6

40

2
exp data

1.8

35

Version 1

30

Version 2, ij=-0.066

1.6

Viscosity (mPa.s)

Viscosity (mPa.s)

Version 2 with tuned ij

Version 2 (ij = 0)

MARD (%)

25
20
15
10

1.4
1.2
1
0.8
0.6

exp data
Version 1, ij=0
Version 1, ij=-0.067
Version 2, ij=-0.179

0.4
0.2

0
0

0.5

0.5

Methanol Mass Fraction

Diethylene Glycol Mass Fraction

Fig. 7. Comparison of the viscosty predictions by Versions 1 and 2 of the EF corrlelation with experimental data of: (a) diethylene glycol + water at 293 K; (b) methanol + water
at 293 K.

viscosity of the water + methanol mixture, the maximum value is


signicantly under-estimated. A good t can be obtained using
a binary interaction parameter of ij = 0.067 which reduced the
AARD to 4.1%. Version 2 tted the data using ij = 0.179 to give an
AARD of 3.9%. In general for Version 2, tuning of binary interaction
parameter is required to provide an accurate viscosity model for
mixtures. For Version 1, the binary interaction parameter was only
required for the most extreme case of hydrogen bonding, mixtures
of water and methanol.

4. Conclusion
The previously developed Expanded Fluid viscosity correlation
was extended to natural gas processing applications; in particular,
to include non-hydrocarbon components such as carbon dioxide, hydrogen sulde, nitrogen, helium, water, methanol, ethylene
glycol, diethylene glycol and triethylene glycol. Both Version 1
(measured density input) and Version 2 (EoS based density input)
were evaluated. In both cases, for uids with signicant hydrogen bonding, such as water and methanol, the c2 parameter of the
correlation was modied from a constant to the following:
c2 = c2 + Kc2 exp(c2 T )
where c2 and Kc2 and c2 are tting parameters. The default value
of Kc2 and c2 is zero for all components except those with signicant hydrogen bonding.
The modied correlation was tted to experimental data for
pure non-hydrocarbons with overall average absolute relative deviations (AARD) below 6%. The maximum absolute relative deviation
(MARD) for predictions with Versions 1 and 2 were 48% and 54%,
respectively. For all of the non-hydrocarbon components evaluated
in this study, a xed c2 value (with Kc2 and c2 equal to 0) was sufcient, except for methanol and water. Version 1 of the correlation
is not recommended for compressed water (P > 10 MPa) at temperatures below 50 C due to an unrealistic trend in the predicted
viscosity values.

Versions 1 and 2 of the correlation predicted the viscosity of


several sweet and sour natural gas mixtures with overall AARDs
of 6.3 and 5.1%, respectively. The MARDs of the predictions with
Versions 1 and 2 were 15 and 51%, respectively. The higher deviations occurred for gas samples with heavy plus-fractions due to an
over-simplied characterization of the plus fraction.
The viscosity of aqueous solutions of glycols and methanol were
also predicted using Version 1 of the correlation with MARD and
AARD of 29 and 6%, respectively. The viscosity interaction parameter (ij ) was set to zero for all binaries except for water/methanol
where a single value of 0.067 was used for all temperature and
compositions. The predictions with Version 2 with ij = 0 were less
accurate with MARD and AARD of 64 and 26%, respectively. Binary
interaction parameters were determined for all binaries reducing
the MARD and AARD of the calculated viscosity values to 40 and
8.7%, respectively.
List of symbols

attractive term in EoS, kPa m6 /kmol2


attractive term at critical point in EoS, kPa m6 /kmol2
cross term in volume translation for the mixtures, Eq.
(A.13),
tting parameter in dilute gas viscosity correlation, Eq.
A
(8), mPa s
Aij
adjustable constant in aij , Eq. (A.14),
A i , B i , C i tting parameters in (T) for polar components, Eq. (A.6),

b
covolume in EoS, m3 /kmol
B
tting parameter in dilute gas viscosity correlation, Eq.
(8), mPa s K1
Bij
adjustable constant in aij , Eq. (A.14), K1
c1
tting parameter in EF correlation, Eq. (1), mPa s
tting parameter in EF correlation, Eq. (1),
c2
c2
tting parameter in temperature dependent c2 , Eq. (19),

a
ac
aij

64

c3
c4
C
D
fc
fMW
fWK
kij
Kc2
MW
n
nc
P
R
s
T
Tr
V
VAPR
VT
wi
WK
xi

(T)

ij
ijo
c2


o

s
so

H. Motahhari et al. / Fluid Phase Equilibria 322323 (2012) 5665

pressure dependency parameter in EF correlation, Eqs. (3)


and (4), kPa1
pressure dependency parameter in EF correlation, Eq. (4),

tting parameter in dilute gas viscosity correlation, Eq.


(8), mPa s K2
tting parameter in dilute gas viscosity correlation, Eq.
(8), mPa s K3
correction factor in VT , Eq. (A.10), m3 /kmol
correlating function, Eq. (14),
correlating function, Eq. (15),
interaction parameter in EoS,
tting parameter in temperature dependent c2 , Eq. (19),

molecular weight, kg/kmol


empirical exponent in EF correlation, Eq. (2),
number of components in the mixture,
pressure, kPa
universal gas constant, kPa m3 /(kmol K)
volume translation, m3 /kmol
absolute temperature, K
reduced temperature,
molar volume, m3 /kmol
molar volume calculated by APR EoS, m3 /kmol
translated molar volume, m3 /kmol
mass fraction of the component i,
Watson-K factor,
mole fraction of the component i,
decay rate of ij from 101.325 kPa to high pressure, Eq.
(16), kPa1
empirical scaling function to adjust vapor pressures in
EoS,
correlating parameter between viscosity and uid expansion, Eq. (1),
binary interaction parameter in EF correlation, Eq. (10),
reference ij at atmospheric condition, Eq. (12) and (13),

tting parameter in temperature dependent c2 , Eq. (19),


K1
isothermal compressibility, Eq. (A.11),
viscosity, mPa s
dilute gas viscosity, mPa s
density, kg/m3
compressed state density, Eq. (2), kg/m3
compressed state density in vacuum, Eqs. (3) and (4),
kg/m3

Acknowledgments
The authors are grateful for nancial support from the sponsors
of the NSERC Industrial Research Chair in Heavy Oil Properties and
Processing, including NSERC, Schlumberger, Shell Canada Energy
Ltd., and Petrobras. We also thank Virtual Materials Group for providing VMGSim process simulation software and Mr. Carl Landra
from Virtual Material Group for his advice on tuning APR EoS for
density predictions of aqueous solutions of associating uids.

where V is the molar volume, R is the universal gas constant, T is


the absolute temperature, b is the covolume, and a is the attractive
term given by:
a = ac (T )

where ac is the attractive term at critical point and (T) is an empirical dimensionless scaling function of temperature. This empirical
function is correlated to acentric factor for non-polar or slightly
pure components such as hydrocarbons.
For hydrocarbons, (T) is given by:

i = 1 + fi (1 Tri )

fi = 0.37464 + 1.54226i 0.26992i2

For < 0.5 :

(A.4)

fi = 0.3796 + 1.4850i 0.1644i2

For => 0.5 :


+ 0.01666i3

(A.5)

For polar compounds the APR EoS uses the following empirical
function of reduced temperature [52] suggested by Mathias and
Copeman [53]:

(T ) = 1 + Ai (1

Tri ) + Bi (1

Tri ) + Ci (1

Tri )

(A.6)

The constants A i , B i and C i are determined by tting the vapor


pressure predicted by the equation of state against experimental
vapor pressure data.
The parameters a and b are calculated for the mixtures using the
following mixing rules:
a=
b=




(1 kij )

ai aj xi xj

xi bi

(A.7)
(A.8)

where xi is the mole fraction of component i and kij is the interaction


parameter between components i and j and determined based on
experimental vapor-liquid equilibrium data.
To correct the deciency of simple cubic EoSs in density calculations, the volume translation concept [54] is implemented in APR
EoS using the method suggested by Mathias et al. [55] as follows:
V T = V APR + s + fc

0.41 
0.41 +

(A.9)

where VT is the translated molar volume, VAPR is the molar volume


as calculated by the APR EoS, s is the volume translation, fc is a
correction factor given by:
fc = Vc (3.946b + s)

(A.10)

and is the isothermal compressibility given by:


=

V2
RT

P
V

(A.11)
T

The volume translation for the mixtures is given by:

si xi + sExcess (T, x )

(A.12)

The APR EoS is based on the Peng and Robinson equation of state
[51]:
a
RT
P=
2
V b
V + 2bV b2

(A.3)

where Tri is the reduced temperature of component i and fi is


given by:

smix =
Appendix A. Advanced PengRobinson equation of state
(APR EoS)

(A.2)

sExcess (T, x ) =

nc

i=1

xi ln

nc


xj aij

(A.13)

j=1

The cross term aij is given by:


(A.1)

aij = eAij +(Bij /T )

(A.14)

H. Motahhari et al. / Fluid Phase Equilibria 322323 (2012) 5665

where Aij and Bij are adjustable constants that can be used to match
liquid densities of mixtures. The default value of these constants are
zero unless otherwise set.
References
[1] A.J. Kidnay, W.R. Parrish, D.G. McCartney, Fundamentals of Natural Gas Processing, CRC Press, Boca Raton, FL, USA, 2011.
[2] A.L. Lee, M.H. Gonzalez, B.E. Eakin, The viscosity of natural gases, J. Petrol.
Technol. 18 (8) (1966) 9971000.
[3] C. Whitson, M. Brule, Phase behavior, monograph, Soc. Petrol. Eng. 20 (2000)
26.
[4] F.E. Londono, R.A. Archer, T.A. Blasingame, Simplied Correlations for Hydrocarbon Gas Viscosity and Gas DensityValidation and Correlation of Behavior
Using a Large-Scale Database, Paper No. 75721 at the SPE Gas Technol. Symp.,
Calgary, Alberta, Canada, 2002.
[5] K. Lucas, Phase Equilibria and Fluid Properties in the Chemical Industry,
Dechema, Frankfurt, 1980, p. 573.
[6] M.L. Huber, Transport properties o f uids, their correlation, prediction and estimation, in: J. Millat, J.H. Dymond, C.A. Nieto de Castro (Eds.), IUPAC, Cambridge
Univ. Press, Cambridge, 1996 (Chapter 12).
[7] H.J.M. Hanley, Prediction of the viscosity and thermal conductivity coefcients
of mixtures, Cryogenics 16 (11) (1976) 643651.
[8] J.F. Ely, H.J.M. Hanley, Transport properties. 1. Viscosity of uids and mixtures,
Ind. Eng. Chem. Fundam. 20 (4) (1981) 323332.
[9] J.F. Ely, H.J.M. Hanley, A Computer Program for the Prediction of Viscosity and
Thermal Conductivity in Hydrocarbon Mixtures (TRAPP), NBS Technical Note
1039, 1981.

[10] S.E. Quinones-Cisneros,


C.K. Zberg-Mikkelsen, E.H. Stenby, The friction theory
(f-theory) for viscosity modeling, Fluid Phase Equilib. 169 (2000) 249276.

C.K. Zberg-Mikkelsen, E.H. Stenby, One parameter fric[11] S.E. Quinones-Cisneros,


tion theory models for viscosity, Fluid Phase Equilib. 178 (2001) 116.

[12] C.K. Zberg-Mikkelsen, S.E. Quinones-Cisneros,


E.H. Stenby, Viscosity prediction of natural gas using the friction theory, Int. J. Thermophys. 23 (2) (2002)
437454.

[13] C.K. Zberg-Mikkelsen, S.E. Quinones-Cisneros,


E.H. Stenby, Viscosity modeling
of associating uids based on the friction theory: pure alcohols, Fluid Phase
Equilib. 194197 (2002) 11911203.
[14] W. Wagner, A. Pruss, The IAPWS Formulation 1995 for the thermodynamic
properties of ordinary water substance for general and scientic use, J. Phys.
Chem. Ref. Data 31 (2) (2002) 387535.

[15] S.E. Quinones-Cisneros,


U.K. Deiters, Generalization of the friction theory for
viscosity modeling, J. Phys. Chem. B 110 (2006) 1282012834.
[16] B.E. Poling, J.M. Prausnitz, J.P. OConnell, The Properties of Gases and Liquids,
5th ed., McGraw-Hill, New York, 2000.
[17] L. Grunberg, A.H. Nissan, Mixture Law for Viscosity, Nature 164 (1949) 799800.
[18] J.D. lsdale, J.C. MacGillivray, G. Cartwright, Prediction of Viscosity of Organic
Liquid Mixtures by a Group Contribution Method, Natl. Eng. Lab. Rept., East
Kilbride, Glasgow, Scotland, 1985.
[19] R.A. McAllister, The viscosity of liquid mixtures, AIChE J. 6 (1960) 427431.
[20] M. Dizechi, E. Marschall, Viscosity of some binary and ternary liquid mixtures,
J. Chem. Eng. Data 27 (3) (1982) 358363.
[21] A.S. Teja, P. Rice, Generalized corresponding states method for the viscosities
of liquid mixtures, Ind. Eng. Chem. Fundam. 20 (1981) 7781.
[22] T. Sun, A.S. Teja, Density, Viscosity, and thermal conductivity of aqueous ethylene, diethylene, and triethylene glycol mixtures between 290 K and 450 K, J.
Chem. Eng. Data 48 (2003) 198202.
[23] H.W. Yarranton, M.A. Satyro, Expanded uid-based viscosity correlation for
hydrocarbons, Ind. Eng. Chem. Res. 48 (2009) 36403648.
[24] M.A. Satyro, H.W. Yarranton, Expanded uid based viscosity correlation for
hydrocarbons using an equation of state, Fluid Phase Equilib. 298 (2010) 111.
[25] VMGSim, Version 5.0 Users Manual, Virtual Materials Group, Inc., Calgary,
Alberta, 2009.
[26] H. Motahhari, M.A. Satyro, H.W. Yarranton, Predicting the viscosity of asymmetric hydrocarbon mixtures with the expanded uid viscosity correlation,
Ind. Eng. Chem. Res. 50 (22) (2011) 1283112843.
[27] J.H. Hildebrand, Viscosity and Diffusivity A Predictive Treatment, John Wiley
and Sons, New York, 1977.

65

[28] C.L. Yaws, Transport Properties of Chemicals and Hydrocarbons, William


Andrew Inc., Norwich, NY, USA, 2008.
[29] C.R. Wilke, A viscosity equation for gas mixtures, J. Chem. Phys. 18 (1950)
517519.
[30] A.O. Rankine, C.J. Smith, The viscosities and molecular dimensions of methane,
sulphuretted hydrogen and cyanogen, Philos. Mag. 42 (1921) 615620.
[31] A.K. Pal, A.K. Barua, Viscosity and intermolecular potentials of hydrogen sulde,
sulfur dioxide and ammonia, Trans. Faraday Soc. 63 (1967) 341346.
[32] P.K. Bhattacharyya, A.K. Ghosh, A.K. Barua, Dipoledipole interaction and viscosity of polar gases, J. Phys. 3 (1970) 526535.
[33] A.K. Pal, P.K. Bhattacharyya, Viscosity of binary polargas mixtures, J. Chem.
Phys. 51 (1969) 828831.
[34] P.K. Bhattacharyya, Viscosity of binary polargas mixtures: CH3 ClH2 S and
CH3 ClSO2 , J. Chem. Phys. 53 (1970) 893895.
[35] I.V. Runovskaya, A.D. Zorin, G.G. Devyatykh, Viscosity of condensed volatile
inorganic hydrides of group IIIVI elements, Russ. J. Inorg. Chem. 15 (1970)
13381339.
[36] C. Nieto-Draghi, A.D. Mackie, A.J. Bonet, Transport coefcients and dynamic
properties of hydrogen sulde from molecular simulation, J. Chem. Phys. 123
(2005) 014505-1014505-8.
[37] NIST Standard Reference Database, NIST/TRC Source Database, WinSource, Version 2008.
[38] A. Padua, W.A. Wakeham, J. Wilhelm, The viscosity of liquid carbon dioxide,
Int. J. Thermophys. 15 (1994) 767777.
[39] P.S. van der Gulik, Viscosity of carbon dioxide in the liquid phase, Physica A 238
(1997) 81112.
[40] K.A.G. Schmidt, S.E. Quinones-Cisneros, J.J. Carroll, B. Kvamme, Hydrogen sulde viscosity modelling, Energy Fuels 22 (2008) 34243434.
[41] R.C. Dougherty, Temperature and pressure dependence of hydrogen bond
strength: a perturbation molecular orbital approach, J. Chem. Phys. 109 (17)
(1998) 73727378.
[42] M. Yves, The Hydrogen Bond and the Water Molecule: The Physics and Chemistry of Water, Aqueous and Bio Media, Elsevier, Amsterdam, Netherlands,
2007.
[43] M.J. Assael, N.K. Dalaouti, Vesovic, Viscosity of natural-gas mixtures: measurements and prediction, Int. J. Thermophys. 22 (1) (2001) 6171.
[44] H. Nabizadeh, F. Mayinger, Viscosity of binary mixtures of hydrogen and natural gas (hythane) in the gaseous phase, High Temp. High Press. 31 (1999)
601612.
[45] L.I. Langelandsvik, S. Solvang, M. Rousselet, I.N. Metaxa, M.J. Assael, Dynamic
viscosity measurements of three natural gas mixtures comparison against
prediction models, Int. J. Thermophys. 28 (2007) 11201130.
[46] G. Galliero, C. Nieto-Draghi, C. Boned, J.B. Avalos, A.D. Mackie, A. Baylaucq,
F. Montel, Molecular dynamics simulation of acid gas mixtures: a comparison between several approximations, Ind. Eng. Chem. Res. 46 (2007)
52385244.
[47] A.M. Elsharkawy, Predicting volumetric and transport properties of sour gases
and gas condensates using EOSs, corresponding state models, and empirical
correlations, J. Pet. Sci. Technol. 21 (2003) 17591787.
[48] V. Vesovic, W.A. Wakeham, The prediction of the viscosity of dense gas mixtures, Int. J. Thermophys. 10 (1) (1989) 125132.
[49] V. Vesovic, W.A. Wakeham, Prediction of the viscosity of uid mixtures over
wide ranges of temperature and pressure, Chem. Eng. Sci. 44 (10) (1989)
21812189.
[50] M. Jaeschke, P. Schley, Calculation of the compressibility factor of natural gases
according to the AGA8-DC92 equation of state: 1. Set up of the equation, gwfGas/Erdgas 137 (7) (1996) 339.
[51] D.Y. Peng, D.B. Robinson, A new two-constant equation of state, Ind. Eng. Chem.
Fundam. 15 (1976) 5964.
[52] M.A. Satyro, The role of thermodynamic modeling consistency in process simulation, in: 8th World Congress of Chemical Engineering, Palais des Congress,
Montreal, August 2327, 2009.
[53] P.M. Mathias, T.W. Copeman, Extension of the PengRobinson equation of state
to complex mixtures: evaluation of the various forms of the local composition
concept, Fluid Phase Equilib. 13 (1983) 91108.
[54] A. Peneloux, E. Rauzy, R. Freze, A consistent correction for
RedlichKwongSoave volumes, Fluid Phase Equilib. 8 (1982) 723.
[55] P.M. Mathias, T. Naheiri, E.M. Oh, A density correction for the PengRobinson
equation of state, Fluid Phase Equilib. 47 (1989) 7787.

Вам также может понравиться