Anda di halaman 1dari 7

High-Pressure Acid-Gas Viscosity

Correlation
G. Galliro, C. Boned, and A. Baylaucq, University of Pau, and F. Montel, SPE, Total

Summary
Acid gases containing hydrogen sulfide (H2S) are often encountered in the petroleum industry. However, reliable experiments on
their thermophysical properties in reservoir conditions, on viscosity in particular, are scarce. From a modeling point of view, H2S
and carbon dioxide (CO2) are polar compounds and as such are
often considered rather difficult to model accurately.
In this work, we propose a correlation with a strong physical
background based on a corresponding-states (CS) approach to predict the viscosity from the temperature and the density of a large
variety of systems for all stable thermodynamic states (gas, liquid,
and supercritical). In particular, this correlation is applicable to predict the viscosity of sour/acid-gas mixtures, whatever the thermodynamic conditions. This approach is based on the Lennard-Jones (LJ)
fluid model, which has been studied extensively thanks to moleculardynamics (MD) simulations over a wide range of thermodynamic
conditions. This fluid model can be extended to deal with polar
molecules such as CO2 or H2S without a loss of accuracy.
First, we demonstrate that the proposed physically based correlation is able to provide an excellent estimation of the viscosity
[with average absolute deviations (AADs) below 5%] of pure
compounds, including normal-alkanes, CO2, or even H2S, whatever
the thermodynamic conditions (gas, liquid, or supercritical).
Then, using a one-fluid approximation and a set of combining
rules, the correlation is applied to various fluid mixtures in a fully
predictive way (i.e., without any additional fitted parameters).
Using this scheme, the deviations between predictions and measurements are as low as those on pure fluids using temperature and
density as inputs. The viscosity of natural- and acid-gas mixtures
at reservoir conditions is shown to be very well predicted by the
proposed scheme. In addition, it is shown that this correlation can
also be applied to predict reasonably the viscosity of asymmetric
high-pressure mixtures, even in the liquid phase.
This physically based approach is easy to include in any simulation software as long as, apart from temperature and density,
the only inputsthe molecular parameters of each speciescan
be estimated from the critical temperature and the critical volume
when not known.
Introduction
Natural gases containing CO2 or H2S, the so-called acid/sour gases,
are often encountered in the petroleum industry (Ungerer et al.
2005). Nevertheless, reliable experiments on their thermophysical
properties are scarce. In the case of H2S, practically no data exist
because of its high toxicity. This lack of information is even more
pronounced concerning transport properties, such as viscosity
(Galliro and Boned 2008; Liley et al. 1988; Schmidt et al. 2008),
at typical petroleum-reservoir conditions (i.e., high pressures and
high temperatures). A predictive, physically based model is highly
required as long as high-pressure viscosity is one of the key physical properties for the design of acid-gases disposal.
Molecular simulation, which can be considered a numerical
experiment on a model fluid, is one of the valuable alternatives to
experiments for gathering physical information on such systems of

Copyright 2010 Society of Petroleum Engineers


This paper (SPE 121484) was accepted for presentation at the EUROPEC/EAGE
Conference and Exhibition, Amsterdam, 811 June 2009, and revised for publication.
Original manuscript received for review 27 February 2009. Revised manuscript received for
review 18 September 2009. Paper peer approved 29 October 2009.

682

interest. Using this simulation technique allows one to obtain exact


results on a molecular model representing the fluid studied without
any limitations on the state studied. It has been shown that such an
approach is efficient in describing the thermophysical properties
of fluids, such as acid-gas mixtures (Galliro et al. 2007; Ungerer
et al. 2005). Nevertheless, molecular simulations require a rather
long time to obtain results and so cannot yet be easily used routinely from an engineering point of view. In order to circumvent
this problem, it is possible to use the molecular-simulation results
on a well-defined fluid model to construct a correlation/theory on
the basis of this model. If this fluid model is representative of some
real fluids, then this correlation/theory can be applied to predict
properties of real fluids by use of a CS approach.
The LJ fluid model is a simple two-parameters molecular model
representing simple nonassociative real fluids fairly well, at least
concerning their viscosity (Galliro et al. 2006a). This fluid model
exactly respects, by definition, a CS behavior. Thus, using a large
database of MD simulations of the viscosity of the LJ fluid, we
have constructed a correlation that is able to represent the MD
results accurately (Galliro et al. 2005a). MD simulations have
been performed on gas, liquid, and supercritical states.
In addition, by using temperature-dependent molecular parameters, the LJ fluid model can also be used to represent molecules
exhibiting a dipole moment, such as H2S. In previous works (Galliro et al. 2007; Galliro and Boned 2008), we have shown that
such a fluid model is able to provide results on thermophysical
properties of H2S consistent with the scarce database and with
more-complex molecular models.
In this paper, we will show that this correlation developed to
describe the viscosity of the LJ fluid is able to provide a very good
estimation of the viscosity of a large variety of species, including
H2S in particular, whatever the thermodynamic state (gas, liquid, and
supercritical), using temperature and density as inputs. In addition, it
will be shown that this fully predictive scheme is able to provide a
very good estimation (i.e., an AAD below 5%) of the viscosity of various mixtures: acid gas, natural gas, or even more-complex mixtures,
such as asymmetric ones without additional fitted parameters.
Models and Correlation
The Fluid Model. Nonpolar Pure Compounds. To describe the
molecule, the LJ uid model has been employed. In this model,
molecules are represented by simple spheres, without internal
degrees of freedom, interacting through an LJ potential (Allen and
Tildesley 1987). In this nonpolar-uid model, each species is completely characterized by its molecular weight M and two molecular
parameters, the potential depth  and the molecular diameter .
These molecular parameters,  and , are closely related to the
critical temperature Tc and the critical molar volume vc.
In a first approximation, the molecular parameters of each species can be obtained with

k BTc
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (1)
1.2593

and

v
 = 0.305 c
NA

1/ 3

, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (2)

where NA is the Avogadro number and kB is the Boltzmann constant.


September 2010 SPE Journal

In this work, Eq. 1 is used systematically to deduce the  of each


studied compound. For , Eq. 2 should be employed directly only
when there is no previously optimized (on viscosity) value available for a given compound. In this work, some optimized values of
 are provided for some typical compounds of petroleum fluids.

and

Dipolar Compounds. The LJ uid model does not include polar


or association effects. It can be used directly to describe such
complex compounds but will be able to provide reasonable results
only for a rather limited temperature range (Galliro et al. 2007;
Yoshimura et al. 2009; Zberg-Mikkelsen et al. 2006). Nevertheless, for a not-too-polar uid such as H2S, it is possible to use
the LJ model combined with an isotropic multipolar potential to
take into account dipolar interactions (Galliro et al. 2007). In
that approximation, it is possible to represent a dipolar molecule
by a simple LJ model with temperature-dependent molecular
parameters. Between Molecule i and Molecule j, these temperature-dependent molecular parameters are

So, by using the vdW1, a mixture is represented by a single


pseudocompound having the molecular parameters Mx, x, and x,
defined by Eqs. 6 through 8. It should be noted that, instead of Eqs.
9 and 10, the usual Lorentz-Berthelot combining rules (Ungerer
et al. 2005) can be used. Nevertheless, we noticed during this
work that they generally induce less-accurate predictions of the
viscosity than those yielded by Eqs. 9 and 10 if they are used with
the scheme proposed in this work.
Reduced Variables. In order to apply a CS scheme, it is necessary to define the reduced (dimensionless) variables in which all
fluids are equivalent in terms of both thermodynamic and transport
properties. In the following, the reduced variables will be noted
with a superscript asterisk. For the LJ fluids, this scaling procedure is performed using the molecular parameters Mx, x, and x,
representing the studied fluid, and is defined using Eqs. 6 through
10. The reduced thermodynamic variables used as inputs in this
work are defined in this frame as

ijpol = ij F 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (3)
and

 ijpol =

 ij
F1 6

, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (4)

with

3 + 3
ij = ii ii 3 jj jj . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (10)
2 ij

T* =

k BT *  x3
, =
, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (11)
x
Mx

where  is the mass density. The reduced viscosity is

i2 2j
F = 1+
, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (5)
12 k BT ij ij6
where i is the dipole moment of Molecule i that is provided for
some compounds in Poling et al. (2000). When a dipole moment
 is known in Debyes, it should be multiplied by 3.1622761025
to be used in Eq. 5.
Mixtures. The scheme proposed in this work is based on the
CS law (i.e., it is assumed that, with adequate scaling that we will
discuss in the next subsection, all fluids follow the same trend in
the scaled/reduced space). The LJ fluid model applied on a pure
fluid respects two-parameter CS behavior exactly. In LJ mixtures,
a single set of molecular parameters representing the mixture
should be defined to use a CS approach. To do so, a simple van
der Waals one-fluid approximation (vdW1) has been used in this
work. The vdW1 is known to provide generally reasonable results
(Vidal 2003; Galliro et al. 2005a) even if, in some asymmetric
mixtures, it may exhibit deficiencies (Galliro et al. 2005b, 2006b).
In an N-components LJ mixture, the vdW1 is
N

M x = xi Mi , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (6)
i =1

 x3 = xi x j ij3 , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (7)
i =1 j =1

* =

where  is the dynamic viscosity.


The knowledge of the dependence of * on T * and * is sufficient to define the viscosity of the LJ fluid model exactly.
Viscosity Correlation. To use the LJ uid as a model to predict
the viscosity of real uid, it is necessary to determine the relation between * and the thermodynamic state (using T * and *
as inputs). To do so, we assume that the reduced viscosity can be
expressed as a sum of two independent contributions:

 * ( T * ,  * ) = 0* ( T * ) +  * ( T * ,  * ) , . . . . . . . . . . . . . . . . . . (13)
where 0* represent the low-density viscosity (function of T * only)
and * is the residual viscosity (function of T * and *). At the
critical point, the viscosity diverges (Vesovic et al. 1990). Nevertheless, in this work, we have not taken into account this contribution because its effect is limited to a very small area close to the
critical point (Vesovic et al. 1990).
Low-Density Contribution. The first-order approximation of
the Chapman-Enskog approach (Poling et al. 2000) applied to
the LJ fluid provides an excellent estimation of the low-density
reduced viscosity. This relation is

and

0* =
N

x x3 = xi x j ij ij3 , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (8)


i =1 j =1

where xi is the molar fraction of Component i. Cross-molecular


parameters between different species have to be defined by combining rules. In this work, it is proposed to define the cross-molecular parameters between Compound i and Compound j as
 3 +  3jj
 ij = ii
2

 x2
, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (12)
Mx x

5 Ac
16 v

T*
, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (14)


where Ac is a numerical parameter taken equal to 0.95 (Galliro


et al. 2005a) and v is the collision integral that can be estimated
with a good precision for 0.3 T * 100 by the Neufeld et al.
(1972) correlation:

( )

v = a1 T *

a2

+ a3e a4T + a5e a6T . . . . . . . . . . . . . . . . . . . . . . (15)

1/ 3

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (9)

September 2010 SPE Journal

The numerical parameters ai appearing in Eq. 15 are provided in


Table 1.
683

TABLE 1NUMERICAL PARAMETER VALUES IN EQS. 15 A ND 16


i

ai

1.16145

0.14874

0.52487

bi

0.062692

4.095577

8.743269 10

Residual Contribution. MD simulation allows computation


of the viscosity of the LJ fluid in dense states. In a previous
work (Galliro et al. 2005a), we performed a large number of
MD simulations of the LJ fluid in various thermodynamic states
covering gas, liquid, and supercritical conditions (i.e., for states
covering from 0.6 T * 6 and * 1.275 with a reduced critical
point located at Tc* 1.3 and c* 0.3 ). Using the MD-simulations
results, we have developed a correlation that is able to provide an
excellent estimation of the residual viscosity by employing

) (

) (Tb ) (e

 * = b1 e b2 1 + b3 e b4  1 +
*

5
* 2

b6  *

1 , . . . . .(16)

in which the numerical parameters bi have been adjusted on MD


results and are provided in Table 1.
Thus, using Eqs. 13 through 16, one can write the complete LJ
viscosity correlation as

* =

4.75
*

16 a1T * a2 + a3e a4T + a5e a6T

) (

T*


+ b1 e b2 1 + b4 e b5 1 +
*

b5 b6*
e 1 , . . . . . . . . (17)
T *2

in which all numerical parameters, ai and bi, are provided in Table 1.


Application of the Proposed Scheme. In order to apply the
proposed scheme to estimate the viscosity of a real fluid, the following steps are needed:
Estimation of the molecular parameters of the compounds
using Tables 2 and 3, or Eqs. 1 and 2 (plus Eqs. 3 and 5 if dipolar
compounds)

0.7732
6

11.12492

2.16178
2.542477 10

2.43787
6

14.863984

Application of the vdW1 (Eqs. 6 through 10) to deduce Mx,


x, and x
Calculation of the reduced thermodynamic conditions T * and
* using Eq. 11, for which the viscosity is estimated
Computation of * using the viscosity correlation, Eq. 17,
and the parameters in Table 1
Calculation of the viscosity using Eq. 12 in reverse (i.e., using
* M xx
=
)
 x2
As an example, we have applied this scheme to a natural gas
(Assael et al. 2001) composed of xC1 = 0.8484, xC2 = 0.084, xC3 =
0.005, xN2 = 0.056, and xCO2 = 0.0066. For T = 353.388 K and
p = 13.96 MPa, they have measured  = 95.5 kg/m3 and  = 16.37
106 Pas.
First, the molecular parameters of each compound of this
mixture have been taken from Tables 2 and 3. Then, using the
vdW1 relations (Eqs. 6 through 10), we have obtained that, x =
1347.7 J/mol, x = 3.692 1010 m, and Mx = 0.01822 kg/mol. Next,
the reduced thermodynamic conditions have been deduced using
Eq. 11T * = 2.18 and * = 0.159. From those reduced thermodynamic conditions, using Eq. 17, we have calculated that * =
* M x x
0.2739. Finally, using  =
, we have deduced the vis x2
cosity predicted by the correlation. The calculation has given  =
16.54 106 Pas, which is only 1% over the experimental value.
Results on Real Fluids
Pure Fluids. As a preliminary step, we have tried to obtain optimized molecular diameters  of some compounds,  being deduced
from Eq. 1 and the critical temperature given in Poling et al. (2000).
For each compound studied, the  value has been adjusted so that
the correlation proposed provides viscosities as close as possible to

TABLE 2OPTIMIZED MOLECULAR PARAMETERS OF SOME NORMAL-ALKANES


n-alkane

(10

10

m)

(Jmol )

TminTmax (K)

pminpmax (MPa)

AAD (%)

max

(%)

Database

C1

3.6325

1258.1

200500

0.1100

1.3

3.6

Lemmon et al. (2007)

C2

4.2093

2015.8

200500

0.160

3.4

8.6

Lemmon et al. (2007)

C3

4.6717

2442

200450

0.130

6.5

10.7

Lemmon et al. (2007)

C4

5.0741

2806.7

250450

0.130

5.2

8.7

Lemmon et al. (2007)

C5

5.4009

3101

303383

0.1100

5.5

10.4

Audonnet and Pdua


(2001)

C6

5.7051

3351.2

303348

0.1250

4.9

9.9
9.4

Oliveira and Wakeham


(1992)
Baylaucq et al. (1999)

C7

5.9916

3566.4

293343

0.1100

4.5

C10

6.7285

4078.1

293373

0.1140

3.3

9.6

Canet et al. (2002)

C16

7.9044

4773.3

298348

0.1150

5.4

9.9

Tanaka et al. (1991)

TABLE 3OPTIMIZED MOLECULAR PARAMETERS OF SOME COMPOUNDS OF PETROLEUM INTEREST


Compounds

(10

10

m)

(Jmol )

TminTmax (K)

pminpmax (MPa)

AAD (%)

max

(%)

Database

H2S +dipole

3.688

2320

200500

0.16

2.9

11.9

H2 S

3.667

2355

200500

0.16

4.5

18.7

Liley et al. (1988)

CO2

3.6641

2007.9

250500

0.1100

4.3

8.7

Lemmon et al. (2007)


Lemmon et al. (2007)

684

Liley et al. (1988)

N2

3.5734

833.1

200500

0.1100

1.3

2.5

O2

3.3269

1020.6

200500

0.180

2.1

4.0

Lemmon et al. (2007)

C6 H6

5.1202

3710.7

283393

0.1200

4.7

11.5

Dymond et al. (1981)

C7 H8

5.3897

3906.8

293373

0.1140

2.6

6.5

Baylaucq et al. (2003)

September 2010 SPE Journal

15

Methane
Carbon dioxide
Hydrogen sulfide

100*(1-corr/exp)

10
5
0
5
10
15
0

200

400

600

800

1000 1200

Density (kgm )
Fig. 1Deviations provided by the proposed correlation for the
major compounds appearing in acid gas.

the data provided in the most reliable recent literature. Temperature


and density have been used as inputs. The adjustment has been
made by minimizing the maximum absolute deviation max between
correlation results and those coming from the literature.
Normal-Alkane. By using Eq. 2 to determine the molecular
diameter of normal-alkane, the viscosity of long chains may not
be predicted correctly. As an example, if Eq. 2 is used for estimating  of the C16, the scheme proposed yields an AAD equal
to 137% compared with the Tanaka et al. (1991) values (18 data
points from T = 298 to 348 K and p = 0.1 to 150 MPa). If the
molecular diameter is optimized on the data, the AAD decreases
to 5.4%; see Table 2.
In fact, in very dense states, the viscosity values are highly
sensitive to the estimation of the molecular diameter of long chains.
For the C16, the relative difference between  deduced from Eq. 2
and the value given in Table 2 is less than 2%.
So, using up-to-date viscosity databases provided in the literature, the molecular parameters of various normal-alkanes have
been adjusted for an as-large-as-possible range of thermodynamic
conditions. All optimized parameters are provided in Table 2. As
shown in Table 2, the LJ fluid model is able to provide a very good
estimation of the viscosity of normal-alkane over a wide range of
thermodynamic conditions, even for rather long chains with at most
only one adjustable molecular parameter per species.
Other Important Compounds. In acid-gas mixtures, apart
from the n-alkane, there is also a need to model H2S and CO2. In
addition, because the proposed scheme is dedicated as well to be
applicable for a wide range of mixtures of petroleum interest, other
usual compounds have been included. Optimized molecular parameters are provided in Table 3. For H2S, the molecular parameters
have been taken from Galliro et al. (2007), the dipole moment 
being equal to 0.9 Debye.
As expected, the less complex (i.e., the more spherical and the
less polar) the molecule, the better the viscosity results provided
by the correlation; see Tables 2 and 3. Nevertheless, even for quite
complex molecules (e.g., toluene), the results provided by the proposed scheme remain very good compared with the simplicity of
the molecular model. It should be noted that optimized molecular
parameters of some other compounds are also provided in ZbergMikkelsen et al. (2006) and Yoshimura et al. (2009).
Concerning C1, H2S, and CO2, which are the most important
compounds in acid-gas mixtures, the correlation is able to provide very good results for the whole range of thermodynamic
conditions; see Fig. 1. Concerning CO2, because no quadrupolar
moment was included in the proposed model (Galliro et al. 2007),
the deviations are larger than for methane. Concerning H2S, care
should be taken because the database provided by Liley et al.
(1988) may suffer from deviations from the real values as high as
20% (Liley et al. 1988).
September 2010 SPE Journal

Simple Binary Mixtures. The application to mixtures of the


scheme proposed in this work is straightforward as long as there
are no adjustable extra parameters, which is usually the case in most
schemes (Poling et al. 2000). Before testing it on a more-complex
mixture, the correlation has been applied on three simple dense
mixtures for which densities and viscosities have been measured.
The first simple mixture studied is composed of C1 and CO2
for three molar fractions of methane, 0.243, 0.464, and 0.755.
Densities and viscosities (159 data points) have been measured by
Dewitt and Thodos (1966) at temperatures ranging from 323 to 473
K (supercritical) and pressures ranging from 3.3 to 68 MPa. For
this mixture, the proposed scheme yields an AAD of 2.3% with
a max of 8.6%. The deviations obtained are slightly larger for a
low content of methane than for high content, which is consistent
with the fact that the correlation performs better for pure methane
than for pure CO2; see Tables 2 and 3. As a point of comparison,
if the classical Lee, Gonzalez, Eakin (1966) (LGE) approach is
employed for estimating the viscosity of this mixture, it yields an
AAD of 76% with a max of 591%. This is not surprising as long
as the LGE correlation, dedicated to natural gas, is known to fail
when nonhydrocarbon compounds are present in a non-negligible
amount (Elsharkawy 2004) and in dense states. As will be shown,
our scheme does not suffer from these limitations.
The second mixture analyzed is composed of C1 and C2 for three
molar fractions of methane, 0.345, 0.5, and 0.685. Viscosities and
densities (135 data points) have been measured by Diller (1984)
at temperatures ranging from 200 to 300 K and pressures ranging
from 1.5 to 30 MPa. The correlation yields an AAD of 2.7% with
a max of 8.9% when compared with experimental values. As for the
previous mixture, deviations generally decrease when the content
of methane increases.
The third mixture is composed of C1 and N2 and has been
measured by Diller (1982) for three molar fractions, 0.317, 0.5,
and 0.714. Measurements have been performed at temperatures
ranging from 200 to 300 K and pressures from 1.6 to 30 MPa. On
that system, the correlation yields an AAD of 2.3% with a max of
8.8% when compared with experimental values.
It is interesting to note that, knowing the densities, the proposed
correlation is able to provide a very good estimation of the viscosities of these simple mixtures for a large range of thermodynamic
states without any extra parameters adjusted on mixture data. In
addition, results obtained on these systems are as good as those on
pure fluids, indicating that the one-fluid approximation employed
here (vdW1) is efficient at least for such simple mixtures.
Natural- and Acid-Gas Mixtures. A more interesting case is the
one involving natural and acid gases. Nevertheless, to our knowledge only a few sets of acid-gas viscosity for mixtures containing
H2S were published (Elsharkawy 2003, 2004). Concerning natural
gas, there is a more comprehensive literature, even if there are not
many data published for high pressures.
In this study, we have selected three sets of recent measurements
on natural gases. The first mixture is the one analyzed by Nabizadeh
and Mayinger (1999). Measurements have been performed at T =
298 to 400 K and p = 0.1 to 6.5 MPa (52 data points). The second
mixture is the one studied in Assael et al. (2001), which covers a
temperature range of 240 to 450 K and a pressure range of 0.1 to
14 MPa (40 data points). The third system is composed of three
different samples of a natural gas (from the North Sea) studied by
Langelandsvik et al. (2007). The temperatures range from 263 to 303
K, and the pressures range from 5 to 25 MPa (45, 34, and 45 data
points). All these natural gases have a content of methane greater
than 0.8 in molar fraction.
As shown by the results in Fig. 2, the correlation is able to
provide a very good estimation of all natural gases studied here
with the largest deviations for the mixtures studied by Langelandsvik et al. (2007). It should be noted that the mixtures studied by
Langelandsvik et al. (2007) contain more compounds than the two
others and are less well defined. In fact, it seems that our correlation tends to slightly and systematically underestimate the viscosity
of natural gas for the highest pressures (Fig. 2). More precisely, on
the mixture of Nabizadeh and Mayinger (1999), we obtain an AAD
685

100*(1-corr/exp)

100*(1-corr/exp)

Nabizadeh and
Mayinger (1999)

Langelandsvik et al. (2007), samp. 1


Langelandsvik et al. (2007), samp. 2
Langelandsvik et al. (2007), samp. 3

Assael et al. (2001)

0
0
(a)

20

40

60

80 100 120 140 160

50

100

Density (kgm )

150

200

250

300

350

Density (kgm3 )

(b)

Fig. 2Deviations provided by the proposed scheme for different natural gases.

Asymmetric Liquid Mixtures. Generally, viscosity prediction


is very difcult to achieve when the species involved are very
asymmetric (i.e., differ largely in size). This is mainly because
the behavior, from the viscosity point of view, of a small molecule
surrounded by large molecules is completely different from the
symmetric situation (i.e., a large molecule surrounded by small
ones). Such asymmetry in the behavior is not easy to tackle through
a simple one-uid approximation.
As a first example, we have applied our correlation on a weakly
asymmetric quaternary mixture composed of normal-alkane in the
liquid state studied by Wu et al. (1998). This mixture is composed of
normal alkanes, C7, C8, C11, and C13 in various proportions at ambient
pressure and at temperatures varying from 293.15 to 313.15 K (28
data points). As for the previous gaseous systems, the results yielded
by the correlation are in good agreement with the measurement, with
an AAD of 6.3% and a max of 9%. Interestingly, this deviation is of
the order of magnitude of those on the pure compounds, even if in
the liquid phase the viscosity varies more abruptly than in the gas
phase (case of the previous mixtures studied).
In our group, the densities and viscosities of mixtures of methane and toluene have been measured (Baylaucq et al. 2003) for a
large range of thermodynamic states, T = 293 to 373 K and p = 0.1
to 140 MPa. This asymmetric mixture has been studied for xC1 =
0.25, 0.37, 0.5, 0.64, and 0.95, and it has been shown in Baylaucq
et al. (2005) that the viscosity of such mixture was difficult to
predict accurately by conventional approaches. Results provided
by our fully predictive scheme are provided in Fig. 3.
Results shown in Fig. 3a clearly demonstrate that the scheme
proposed is able to provide a reasonable estimation of viscosity, even in a liquid asymmetric mixture without any adjustable

10

20

10

0
5
10
15
20
25

xC1=0.25
xC1=0.37
xC1=0.5
xC1=0.64
xC1=0.95

200 300 400 500 600 700 800 900 1000


(a)

Density (kgm )

100*(1-corr/exp)

100*(1-corr/exp)

of 0.97% with a max of 1.69 % and an AAD of 0.83% with a max


of 3.39 % for the mixture of Assael et al. (2001). Concerning the
mixtures of Langelandsvik et al. (2007), we obtain an overall AAD
of 4.04% with a max of 6.9%. As a point of comparison with usual
correlations, for those three mixtures, the LGE approach yields
AADs of 2.17, 2.21, and 1.35%, respectively. This means that,
even for natural gas, our correlation is able to provide viscosity
estimation as good as that of the LGE scheme, which is dedicated
specifically to natural gas.
Elsharkawy (2003) has provided densities and viscosities of
three different acid-gas mixtures containing H2S. The first mixture
(Gas 11) is mainly composed of xH2S = 0.23 and xC1 = 0.76 at T =
352.55 K and p = 34.47 MPa. The second one (Gas 13) is principally composed of xH2S = 0.49 and xC1 = 0.45 at T = 322.05 K
and p = 17.24 MPa. The third one (Gas 14) is mainly composed
of xH2S = 0.7 and xC1= 0.2 at T = 352.55 K and p = 9.4 MPa. The
other minor compounds are CO2, N2, and alkanes. They have been
taken into account for the calculation.
Applying the correlation, we obtained viscosities equal to
0.0317, 0.0362, and 0.0198 mPas, respectively; the LGE correlation yields 0.031, 0.036, and 0.018 mPas, respectively; whereas
the experimental values are 0.03, 0.03, and 0.022 mPas, respectively. Thus, even if there is a non-negligible deviation on the
second mixture, the values predicted by our scheme are consistent
compared with the experimental values. This is not surprising
because both methane and H2S are well represented by the fluid
model employed in this work; see Tables 1 and 2. Furthermore,
as shown by the previous results on the simple mixtures studied,
the scheme proposed is very efficient in dealing with mixtures of
similar sizes in a purely predictive way.

0
10
xC0 =0.15

20

xC0 =0.3
2

xC0 =0.5

30

xC0 =0.85
2

40
550
(b)

600

650

700

750

800

850

900

Density (kgm3 )

Fig. 3Deviations provided by the proposed scheme for (a) methane/toluene mixture and (b) CO2/decane mixture.
686

September 2010 SPE Journal

parameters. For the system studied here, we obtain an AAD of 5.9%


and a max of 26.3%, which is not worse than for more-complex
approaches (Baylaucq et al. 2005). When looking more carefully at
Fig. 3a, it appears that the results for xC1 = 0.25, 0.5, and 0.64 are
excellent; those for xC1 = 0.37 are reasonable; and those for xC1 =
0.95 yield the largest deviations of all for the highest densities. This
is somewhat surprising because we expect that results should have
been very good for high methane content, because pure methane is
well modeled. Such differences may come from intrinsic limitations
of the vdW1 (Galliro et al. 2005b) when there is a small amount
of the lightest compound in an asymmetric mixture.
The last asymmetric mixture studied is composed of CO2 and
C10. The viscosities of this mixture have been measured by Barrufet et al. (1996) for temperatures ranging from 310 to 403 K and
pressures up to 12 MPa. As shown in Fig. 3b, the correlation is
able to provide reasonable results except for xCO2 = 0.85, for which
the deviations reach 31.6%. This trend is consistent with those on
the methane/toluene mixture. For that precise mixture (Fig. 3b),
the correlation overestimates viscosity for the lowest densities,
whereas it is the contrary for the highest densities.
From these results, it can be concluded that, even for asymmetric
mixtures, the proposed purely predictive scheme is able to provide a
good estimation of viscosity whatever the thermodynamic state, at
least from an engineering point of view. A very interesting feature
of the proposed correlation is that it provides results on mixtures
that are generally as good as those on pure fluids without any additional fitted parameters. Nevertheless, it should be noted that such
an approach should be taken with great care when dealing with
associative compounds, including water. A simple LJ model cannot
tackle the very subtle physics of hydrogen bonding (ZbergMikkelsen et al. 2006).
Conclusions
In this paper, a fully predictive method for natural- and acid-gases
viscosity calculation is presented. The model is based on MD simulations in gas, liquid, and supercritical states. Molecular simulations
were performed on the LJ fluid to obtain a large, fully consistent database. The correlation, built on this fluid model in reduced variables,
is fully predictive and uses temperature and density as inputs.
A CS scheme was developed to allow application of the correlation on real pure fluids according to two molecular parameters,
the molecular diameter and the potential depth, for many compounds. Application to pure components usually requires only the
knowledge of the critical temperature and the critical volume. For
complex molecules, the molecular diameter should be adjusted
on available experimental viscosity data (knowing the density),
and a database of optimized molecular parameters is provided in
this work. In addition, by using temperature-dependent molecular
parameters, the method was extended to deal with polar compounds, such as H2S. This correlation is able to provide the viscosity of a large variety of compounds, including normal-alkane, H2S,
or CO2, with an excellent accuracy (i.e., AADs below 5%) on a
very wide range of thermodynamic states.
The method was extended to mixtures using the vdW1 without
any extra adjustable parameters. This fully predictive method was
tested extensively against available data on multicomponent mixtures in gas, supercritical, and liquid states. It is shown that this
approach is able to predict the viscosity of natural- and acid-gas
mixtures with accuracy as good as that on pure fluids, whatever the
composition. This approach is as efficient as the usual LGE correlation on natural gas at low density but can be applied with success also to dense systems and to nonmethane-dominant mixtures
(e.g., CO2- or H2S-rich mixtures), contrary to the LGE correlation.
Finally, it is shown that this correlation can also be applied to predict reasonably the viscosity of asymmetric high-pressure mixtures
even in the liquid phase without any fitted parameter.
Nomenclature
ai = numerical parameters for the low-density viscosity
bi = numerical parameters for the residual viscosity
kB = Boltzmann constant, Jmol1K1
September 2010 SPE Journal

Mi
N
NA
p
T
v
xi
i




v

= molecular weight of Component i, kgmol1


= number components in a mixture
= Avogadro number
= pressure, Pa
= temperature, K
= molar volume, m3mol1
= molar fraction of Component i
= dipole moment of Molecule i, Debye
= potential depth, Jmol1
= dynamic viscosity, Pas
= density, kgm3
= molecular diameter, m
= collision integral

Acknowledgments
We gratefully acknowledge Total S.A. for allowing us to publish
this work.
References
Allen, M.P. and Tildesley, D.J. 1987. Computer Simulation of Liquids.
Oxford, UK: Oxford Science Publications, Oxford University Press.
Assael, M.J., Dalaouti, N.K., and Vesovic V. 2001. Viscosity of Natural-Gas
Mixtures: Measurements and Prediction. Int. J. Thermophys. 22 (1):
6171. doi: 10.1023/A:1006784814390.
Audonnet, F. and Pdua, A.A.H. 2001. Simultaneous measurement of density
and viscosity of n-pentane from 298 K to 383 K and up to 100 MPa using
a vibrating-wire instrument. Fluid Phase Equilibria 181 (12): 147161.
doi: 10.1016/S0378-3812(01)00487-3.
Barrufet, M.A., Salem, S.K.E.-S., Tantawy, M., and Iglesias-Silva, G.A. 1996.
Liquid Viscosities of Carbon Dioxide + Hydrocarbons from 310 K to 403
K. J. Chem. Eng.Data 41 (3): 436439. doi: 10.1021/je950256y.
Baylaucq, A., Boned, C., Dauge, P., and Xans, P. 1999. Viscosity Versus Pressure, Temperature and Composition of the Ternary System Heptane +
Methylcyclohexane + 1-Methylnaphtalene. Comparative Analysis of Some
Models. Physics and Chemistry of Liquids 37 (5): 597626. doi: 10.1080/
00319109908035939.
Baylaucq, A., Boned, C., Canet, X., and Zberg-Mikkelsen, K. 2003. High-Pressure (up to 140 MPa) Dynamic Viscosity of the Methane and Toluene System: Measurements and Comparative Study of Some Representative Models. Int. J. Thermophys. 24 (3): 621638. doi: 10.1023/A:1024023913165.
Baylaucq, A., Boned, C., Canet, X., Zberg-Mikkelsen, C., QuionesCisneros, S., and Zhou, H. 2005. Dynamic Viscosity Modeling of
Methane+n-Dencane and Methane + Toluene Mixtures: Comparative
Study of Some Representative Models. Petroleum Science and Technology 23 (2): 143157. doi: 10.1081/LFT-200028122.
Canet, X., Baylaucq, A., and Boned, C. 2002. High-Pressure (up to 140
MPa) Dynamic Viscosity of the Methane + Decane System. Int. J.
Thermophys. 23 (6): 14691486. doi: 10.1023/A:1020781715494.
Dewitt, K.J. and Thodos, G. 1966. Viscosities of binary mixtures in the
dense gaseous state: The methane-carbon dioxide system. Can. J.
Chem. Eng. 44 (3): 148155. doi: 10.1002/cjce.5450440305.
Diller, D.E. 1982. Measurements of the viscosity of compressed gaseous
and liquid nitrogen + methane mixtures. Int. J. Thermophys. 3 (3):
237243. doi: 10.1007/BF00503319.
Diller, D.E. 1984. Measurements of the viscosity of compressed gaseous
and liquid methane + ethane mixtures. J. Chem. Eng. Data 29 (2):
215221. doi: 10.1021/je00036a035.
Dymond, J.H., Robertson, J., and Isdale, J.D. 1981. Transport properties of
nonelectrolyte liquid mixturesIV. Viscosity coefficients for benzene,
perdeuterobenzene, hexafluorobenzene, and an equimolar mixture of
benzene + hexafluorobenzene from 25 to 100C at pressures up to the
freezing pressure. Int. J. Thermophys. 2 (3): 223236. doi: 10.1007/
BF00504186.
Elsharkawy, A.M. 2003. Predicting Volumetric and Transport Properties
of Sour Gases and Gas Condensates Using EOSs, Corresponding State
Models, and Empirical Correlations. Petroleum Science and Technology
21 (1112): 17591787. doi: 10.1081/LFT-120024560.
Elsharkawy, A.M. 2004. Efficient methods for calculations of compressibility, density and viscosity of natural gases. Fluid Phase Equilibria
218 (1): 113. doi: 10.1016/j.fluid.2003.02.003.
687

Galliro, G. and Boned, C. 2008. Dynamic viscosity estimation of hydrogen


sulfide using a predictive scheme based on molecular dynamics. Fluid
Phase Equilibria 269 (12): 1924. doi: 10.1016/j.fluid.2008.04.017.
Galliro, G., Boned, C., and Baylaucq, A. 2005a. Molecular Dynamics
Study of the Lennard-Jones Fluid Viscosity: Application to Real Fluids.
Ind. Eng. Chem. Res. 44 (17): 69636972. doi: 10.1021/ie050154t.
Galliro, G., Boned, C., Baylaucq, A., and Montel, F. 2005b. Influence of
the mass ratio on viscosity in Lennar-Jones mixtures: The one-fluid
model revisited using nonequilibrium molecular dynamics. Fluid Phase
Equilibria 234 (12): 5663. doi: 10.1016/j.fluid.2005.05.016.
Galliro, G., Boned, C., Baylaucq, A., and Montel, F. 2006a. Molecular
dynamics comparative study of Lennard-Jones
-6 and exponential
-6
potentials: Application to real simple fluids (viscosity and pressure).
Phys. Rev. E 73 (6): 061201. doi: 10.1103/PhysRevE.73.061201.
Galliro, G., Boned, C., Baylaucq, A., and Montel, F. 2006b. The van der
Waals one-fluid model for viscosity in Lennard-Jones fluids: Influence
of size and energy parameters. Fluid Phase Equilibria 245 (1): 2025.
doi: 10.1016/j.fluid.2006.03.006.
Galliro, G., Nieto-Draghi, C., Boned, C., Avalos, J.B., Mackie, A.D.,
Baylaucq, A., and Montel, F. 2007. Molecular dynamics simulation of
acid gas mixtures: A comparison between several approximations. Ind.
Eng. Chem. Res. 46 (15): 52385244. doi: 10.1021/ie061616l.
Langelandsvik, L.I., Rousselet, M., Metaxa, I.N., and Assael, M.J. 2007.
Dynamic Viscosity Measurements of Three Natural Gas MixturesComparison against Prediction Models. Int. J. Thermophys. 28 (4):
11201130. doi: 10.1007/s10765-007-0270-3.
Lee, A.L., Gonzalez, M.H., and Eakin, B.E. 1966. The Viscosity of Natural
Gases. J Pet Technol 18 (8): 9971000; Trans., AIME, 237. SPE-1340PA. doi: 10.2118/1340-PA.
Lemmon, E.W., Huber, M.L., and McLinden, M.O. 2007. NIST Standard Reference Database 23: Reference Fluid Thermodynamic and
Transport Properties-REFPROP, Version 8.0. Gaithersburg, Maryland:
Standard Reference Data Program, National Institute of Standards and
Technology.
Liley, P.E., Makita, T., and Tawaka, Y. 1988. Properties of Inorganic and
Organic Fluids, Vol. V-1,175182. New York: Cindas Data Series on
Material Properties, Taylor & Francis.
Nabizadeh, H. and Mayinger, F. 1999. Viscosity of hydrogen and natural
gas (hythane) in the gaseous phase. High Temperatures High Pressures 31 (6): 601612. doi: 10.1068/htwu363.
Neufeld, P.D., Janzen, A.R., and Aziz, R.A. 1972. Empirical Equations
to Calculate 16 of the Transport Collision Integrals (l,s)* for the Lennard-Jones (126) Potential. The Journal of Chemical Physics 57 (3):
11001102. doi: 10.1063/1.1678363.
Oliveira, C.M.B.P. and Wakeham, W.A. 1992. The viscosity of five liquid
hydrocarbons at pressures up to 250 MPa. Int. J. Thermophys. 13 (5):
773790. doi: 10.1007/BF00503906.
Poling, B.E., Prausnitz, J.M., and OConnell, J.P. 2000. The Properties of
Gases and Liquids, fifth edition (hardback). New York: McGraw-Hill
Professional.
Schmidt, K.A.G., Quiones-Cisneros, S.E., Caroll, J.J., and Kvamme, B.
2008. Hydrogen Sulfide Viscosity Modeling. Energy & Fuels 22 (5):
34243434. doi: 10.1021/ef700701h.
Tanaka, Y., Hosokawa, H., Kubota, H., and Makita, T. 1991. Viscosity and
density of binary mixtures of cyclohexane with n-octane, n-dodecane
and n-hexadecane under high pressures. Int. J. Thermophys. 12 (2):
245264. doi: 10.1007/BF00500750.
Ungerer, P., Tavitian, B., and Boutin, A. 2005. Applications of Molecular
Simulation in the Oil and Gas IndustryMonte Carlo Methods. Paris:
Editions Technip.
Vesovic, V., Wakeham, W.A., Olchowy, G.A., Sengers, J.V., Watson, J.T.R.,
and Millat, J. 1990. The Transport Properties of Carbon Dioxide. J.
Phys. Chem. Ref. Data 19 (3): 763. doi: 10.1063/1.555875.
Vidal, J. 2003. Thermodynamics: Applications in Chemical Engineering
and the Petroleum Industry. Paris: Editions Technip.

688

Wu, J.W., Shan, Z., and Asfour, A.A. 1998. Viscometric Properties of
Multicomponent Liquid n-Alkane Systems. Fluid Phase Equilibria 143
(12): 263274. doi: 10.1016/S0378-3812(97)00269-0.
Yoshimura, M., Boned, C., Baylaucq, A., Galliro, G., and Ushiki, H. 2009.
Influence of the chain length on the dynamic viscosity of some amines: Measurements and comparative study of some models. The Journal of Chemical
Thermodynamics 41 (3): 291300. doi: 10.1016/j.jct.2008.08.006.
Zberg-Mikkelsen, C.K., Watson, G., Baylaucq, A., Galliro, G., and
Boned, C. 2006. Comparative experimental and modeling studies of
the viscosity behavior of ethanol +C7 hydrocarbon mixtures versus
pressure and temperature. Fluid Phase Equilibria 245 (1): 619. doi:
10.1016/j.fluid.2006.01.030.

Guillaume Galliro holds a PhD degree from the University of


Bordeaux, France. email: guillaume.galliero@univ-pau.fr. After
two years in a postdoctoral position at the Technical University
of Denmark and the University of Pau, France, he became an
assistant professor at the University of Marne la Valle, France,
in 2005. In 2007, Galliero took a full researcher position at the
Centre National de la Recherche Scientifique (CNRS), affiliated with the University of Pau. Since November 2009, he has
been a full professor in the Laboratory of Complex Fluids at the
University of Pau, France. Gallieros research interests encompass simulation and modeling of thermophysical properties
of fluids as well as multiscale/multiphysics study of heat- and
mass-transfer simulations of multicomponent mixtures. He is
an expert in molecular-simulations techniques. Galliero has
published numerous articles in scientific journals and contributed to many international meetings. Christian Boned holds a
PhD degree from the University of Pau and Pays de lAdour.
email: christian.boned@univ-pau.fr. From 1967 to 1988, he was
an assistant and then an associate professor at the university.
Since 1988, he has been a full professor in the Laboratory of the
Complex Fluids, associated with CNRS and Total. Boned is the
head of the Viscosity, Transport Properties team. His research
team focuses primarily on experimental measurements, especially those involving large temperature and pressure intervals.
His expertise is in density and viscosity at high pressure. Boneds
research interest encompasses both experimental and theoretical aspects, such as free-volume theory of dynamic viscosity and molecular-dynamics representation of viscosity. He is the
coauthor of more than 130 articles published in international
journals. Boned is a member of the International Association on
Transport Properties and has been co-organizer of several meetings. Antoine Baylaucq holds a PhD degree from the University
of Pau and Pays de lAdour, France, and worked at the university as an assistant professor until August 2009. email: antoine.
baylaucq@total.com. In September 2009, he joined Total as
head of the Fluids and Organic Geochemistry Laboratory.
Baylaucqs areas of research interest are sampling, pressure/
volume/temperature (PVT), fluid properties, fluid behavior, fluid
composition, gas injection, heavy oil, oil shale, gas shale, wax,
asphaltene, and transport properties. Franois Montel holds
a PhD degree from the Ecole Suprieure de Physique et de
Chimie Industrielles in Paris. email: francois.montel@total.com.
He was also an assistant professor there for three years. Montel
then joined Elf in 1978 and became head of the PVT laboratory and an acknowledged technical expert. Montel now
works as an expert in fluid thermodynamics for Total Exploration
and Production. He has been published numerous times in scientific journals. Montel is an associate professor at the Ecole
Nationale Suprieure du Ptrole et des Moteurs in Rueil and
a course instructor in applied thermodynamics at Imperial
College. His research interests include catagenesis, migration,
trapping, alteration, sampling, PVT, fluid properties, fluid distribution, reservoir connectivity, fluid behaviour, equation of state
modeling, phase equilibria, gas injection, CO2 sequestration,
heavy oil, oil shale, gas shale, wax, asphaltene, hydrates, and
sulphur.

September 2010 SPE Journal

Anda mungkin juga menyukai