Vous êtes sur la page 1sur 8

ISSN 0040-5795, Theoretical Foundations of Chemical Engineering, 2009, Vol. 43, No. 3, pp. 337344. Pleiades Publishing, Ltd.

., 2009.
Original Russian Text I.V. Derevich, E.G. Grechko, V.A. Pershukov, 2009, published in Teoreticheskie Osnovy Khimicheskoi Tekhnologii, 2009, Vol. 43, No. 3, pp. 352359.

Calculating the Dynamic Viscosity of Individual Substances


Using the PengRobinson Equation
I. V. Derevicha, E. G. Grechkoa, and V. A. Pershukovb
a

Moscow State University of Engineering Ecology, Staraya Basmannaya ul. 21/4, Moscow, 105066 Russia
b OOO Alltek, Moscow, Russia
e-mail: nchmt@iht.mpei.ac.ru
Received June 27, 2007

AbstractA method for calculating the dynamic viscosity of individual compounds in the liquid and gaseous
states has been developed using Eyring and Frankls excited zone formation hypothesis. In this method, the
activation energy is estimated in terms of the deviation of the enthalpy of the substance from the ideal-gas
approximation. This deviation is calculated using the PengRobinson equation f state. The expressions obtained
for viscosity contain a constant, whose value is the same for a wide variety of compounds. The results of these
calculations are compared with experimental data for the viscosity of liquids and vapors along the saturation
line and isobars, including supercritical pressures.
DOI: 10.1134/S0040579509030154

INTRODUCTION
The accuracy of calculations for hydrodynamic processes in various applications depends on the reliability
of the determination of the dynamic viscosities of the
working media in the liquid and gaseous states. Sufficiently complete experimental data are presently available
only for a small number of substances. For a wide variety
of substances, tables of experimental data have been compiled only for narrow temperature and pressure ranges.
However, developers of new technologies have to perform
calculations for various working substances over wide
ranges of state parameters. In this situation, the available
experimental data are obviously insufficient and numerical simulation methods are coming to the foreground in
viscosity determination. The methods of statistical physics
are insufficiently developed to predict the collective
effects responsible for momentum transfer in liquid matter. It is, therefore, pertinent to search for empirical and
semiempirical methods for predicting the thermophysical
properties of substances.
There are several semiempirical approaches to the calculation of the dynamic viscosities of liquids and vapors
(see, e.g., [1]). For rarefied gases, whose state can be
viewed as being ideal, dynamic viscosity is found using
Chapman and Enskogs kinetic theory of gases [2]. In
practical calculations, the collision integral is approximated with empirical formulas. This allows the collisions
of polar gas molecules to be taken into consideration. The
data calculated using this theory are in satisfactory agreement with experimental data both for individual rarefied
gases and for their mixtures. The effect of compression on
the viscosity of gases is taken into account in terms of the

residual viscosity hypothesis. The dimensionless difference between the viscosity of the compressed gas and the
viscosity of the rarefied gas is a universal function
assumed to depend on the reduced density of the substance. The residual viscosity hypothesis is also used to
approximate the viscosities of liquids and supercritical fluids [1, 3]. Note that different authors report very different
formulas and empirical coefficients for residual viscosity.
The residual viscosity hypothesis is employed in the simulation of the dynamic viscosities of organic mixtures,
such as petroleum [3]. In this case, the amount of empirical information necessary for determining the unknown
coefficients is much larger.
The viscosity of liquids is also simulated using several
approaches. Examples of purely empirical approaches are
the OrickErbar and van VelzenCardosoLangenkamp
methods (see, e.g., [1]). Either method uses two empirical
constants, whose values have been determined for many
pure organic liquids. Within these empirical methods, a
technique based on the group constituents of the chemical
formula has been developed for estimating the constants.
However, the data calculated by these methods deviate
substantially from experimental data. Reliable simulation
of viscosity is possible only for supercooled liquids with a
reduced temperature below 0.75. For higher temperatures,
it is necessary to use other empirical formulas, such as the
Lecu and Steels formula [1]. Purely empirical formulas
can be generalized to the viscosity of solutions. The
practical value of various empirical methods for calculating the viscosity of organic compounds is assessed,
e.g., in a review by Mehrotra et al. [4]. Empirical methods are also developed using the corresponding states
principle, which is based on van der Waals hypothesis

337

338

DEREVICH et al.

that the properties of substances in terms of reduced thermodynamic parameters show similar behaviors [5].
Rigorous lattice gas models, which use methods of statistical physics, are applied to the self-consistent calculation of the shear viscosity and compressibility of a number
of dense gases [6]. Present-day ultrahigh-performance
computers allow transfer coefficients to be simulated by
molecular dynamics methods (see, e.g., [7]).
Semiempirical models based on ideas of statistical
physics and on equations of state have been developed in
recent years. Qun-Fang et al. [8] utilize Eyring and
Frankls hypothesis that momentum transfer in liquids
takes place through the formation of an excited zone in
which the concentration of molecules is lower than that in
the main volume of the liquid [9, 10]. The probability of
the formation of this exited zone is estimated using activation theory. The activation energy is assumed to be proportional to the heat of evaporation of the pure liquid. This
approached proved to be very fruitful. The technique of
Qun-Fang et al. [8] involves two empirical constants in the
calculation of the dynamic viscosity of the liquid. As a
result, the model can satisfactorily predict the viscosity of
fluids in a wide temperature range up to the critical point
and in a wide pressure range, including the supercritical
state of matter. The excited zone model is also used to predict the viscosity of solutions, including petroleum [11,
12]. Quinones-Cisneros et al. [13] simulate the viscosity
of liquids as the sum of the viscosity of the rarefied gas and
the viscosity due to repulsive and attractive forces. The
repulsive and attractive forces are significant at elevated
pressures, and their contribution is taken into account
using an equation of state, e.g., the SoaveRedlich
Kwong equation. Some empirical constants of the model
suggested by Quinones-Cisneros et al. [13] are derived
from experimental data for the critical isotherm. The
resulting relationships can safely be applied to both liquids
and supercritical fluids. This technique was extended to
the viscosity of solutions of organic liquids [14]. An
approach using the analogy between the pressuretemperaturevolume diagram and the pressuretemperature
dynamic viscosity diagram [15] serves as the basis for
simultaneous calculation of the dynamic viscosities of the
liquid and vapor in wide ranges of temperature and pressure. A cubic equation similar to the PengRobinson
equation has been formulated, which involves dynamic
viscosity as the analogue of volume in classical equations
of state. The empirical functions describing the effects of
attraction and repulsion have been determined for this
equation. The data calculated using this equation are in
satisfactory agreement with experimental data for light
paraffinic hydrocarbons. The approach suggested in [15]
is extended to the viscosity of mixtures by applying the
one-liquid approximation with pseudocritical thermodynamic parameters. Use of equations of state in the evaluation of the energy parameters governing momentum transfer in solutions is promising for the estimation of the viscosity of nonideal mixtures (see, e.g., [16]). In this case,
the viscosity of the mixture is correlated with the Gibbs

mixing energy calculated using the equation of state with


preset mixing rules.
Here, we assume that momentum transfer in the liquid
is due to the formation of an excited zone in which the
molecule concentration is lower than that in the main volume of the liquid. Earlier, this approach was used in the
simulation of the effect of dissolved light gases on the viscosity of petroleum [17]. The formation of an excited zone
requires some energy to be spent. The value of this energy
is estimated from the deviation of the enthalpy of the real
substance from the ideal-gas approximation. This deviation is calculated using the PengRobinson equation. In
the prediction of the viscosity of a compressed gas, it is
assumed that the residual viscosity is also correlated with
the deviation of the real enthalpy from the ideal-gas
approximation. In the supercritical region of thermodynamic parameters, the viscosity of the fluid is combined
from the viscosities of the liquid and the compressed gas.
This approach allows the dynamic viscosity of both the
liquid and vapor to be determined over wide ranges of
temperature and pressure using a single procedure. The
data thus calculated will be compared with experimental
data for paraffinic and isoparaffinic compounds and for a
number of simple gases for which adequate tables of
experimental data are available. In the calculation, we will
use an approximation of viscosity at the critical point and
an empirical constant whose value is invariable for, e.g.,
the paraffins. When the dynamic viscosity data for the critical point are sufficiently accurate, this empirical constant
is universal for the substances considered in this study.
The calculated data will be compared with experimental
data obtained for various substances on the saturation line
and on isobars in a wide pressure range.
MODEL OF THE VISCOSITY
OF THE LIQUID AND VAPOR
The considerations presented below are not a rigorous theory of the viscosity of liquids or compressed
gases. The formulas set up in this work are semiempirical, and the descriptions of theoretical constructions
mainly serve to illustrate the functional dependences
suggested.
The viscosity model is based on the hypothesis that
momentum transfer in the liquid is due to the formation
of a macroscopic excited zone with a lower molecule
concentration [9, 10]. In the absence of thermal events,
the minimum work necessary for the formation of an
excited zone in the equilibrium approximation is as follows [18]:
A min = U + P o V ,

(1)

where Po is the pressure in the medium and U and V are


the local changes in the internal energy and in the volume of the medium in the place of excited zone formation.
We will assume that the thermodynamic state of the
zone with a lower molecule concentration can be

THEORETICAL FOUNDATIONS OF CHEMICAL ENGINEERING

Vol. 43

No. 3 2009

CALCULATING THE DYNAMIC VISCOSITY OF INDIVIDUAL SUBSTANCES

approximated by the ideal gas state. The minimum


value of the work necessary for the formation of a
vacancy (Eq. (1)) will then be equal to the deviation of
the enthalpy of the substance from the enthalpy of the
ideal gas:
A min = H H = H .
id

id

(2)

The value of Hid in (2) can be calculated using an


equation of state describing both the vaporous and liquid states of the substance.
According to the theory of thermodynamic fluctuations, the vacancy forms via the activation mechanism.
We assume that the effective energy barrier that must be
surmounted for vacancy formation is proportional to
the minimum work (Eq. (2)). Accordingly, the probability of the formation of the excited zone is proportional to
id

we

H
-----------RT

Here, is an empirical coefficient. It is constant for


each particular substance, and its value is adjusted by
analyzing experimental information.
The randomly appearing vacancies play a dual role.
On the one hand, the decrease in the probability of
vacancy formation causes an increase in the friction
factor. On the other hand, the vacancies themselves
serve as momentum dissipation centers and the number
density of vacancies is proportional to the density of the
substance. Based on these considerations, it is possible
to write the dynamic viscosity of the liquid in the following functional form:
id

l e

H
-----------RT

(3)

At the critical point (T = Tc, P = Pc), the viscosity


takes its critical value:
id

c c e

Hc
-----------RT

(4)

A comparison between (3) and (4) leads to the following simple formula for the viscosity of the liquid:
H H c
- .
l = c r exp ----------------------------

RT
id

339

When the gas can be taken to be ideal, Hid/RT 0, the


contribution from the second term on the right-hand side
of (6) is small, and the viscosity of the substance is equal
to the viscosity of the rarefied gas. As the temperature
approaches the critical point, l and v
c, as is clear
from Eqs. (5) and (6).
The calculation of dynamic viscosity throughout the
possible range of thermodynamic parameters is carried
out as follows.
At T < Tc, the liquid and vapor states can exist in the
system. In this case, if the pressure is below the saturation
vapor pressure (P < Ps(T)), then = v (Eq. (6)). At P
Ps(T), the viscosity of the substance is equal to the viscosity of the liquid: = l (Eq. (5)).
For an overheated vapor under subcritical conditions
(T Tc, P Pc), the viscosity of the substance is calculated
using the formula for the vapor: = v (Eq. (6)).
For a supercritical fluid (P > Pc), an approximating formula describing the intermediate state of this fluid is used
throughout the temperature range:
n 1/n

= ( l + v ) ,
n

where n is the exponent (n  1).


The dynamic viscosity at the critical point can be
calculated using the following empirical approximation
(see, e.g., [15]):
c = 7.7T c P c .
1/6

1/2

2/3

(7)

Here, the critical temperature is in Kelvin degrees, the


molar mass is in g/mol, and the critical pressure is in bars (1
bar = 105 Pa). Accordingly, the dynamic viscosity at the critical point is in micropoises (1 P = 107 Pa s).
Quinones-Cisneros et al. [14] collected a set of experimental data for the critical viscosity of hydrocarbons and
simple gases.
In order to make the above procedure comprehensible to
a wide variety of users, we will provide a detailed description of the relevant formulas and the methods of determination of the necessary thermodynamic parameters.

id

The formula for the viscosity of the compressed gas


(presented below) is a purely intuitional construction
based on well-known trends of the viscosity of gases at
elevated temperatures. For gases, momentum dissipation is due to the chaotic motion of molecules, whose
intensity is proportional to the gas temperature. Using
the residual viscosity hypothesis [1, 3], the viscosity of
the gas can be approximated as
H H c
id
id
- .
v = v + ( c v )T r exp ----------------------------

RT
id

VISCOSITY OF THE RAREFIED GAS

(5)

id

(6)

The calculation of viscosity in the ideal-gas approximation is based on the kinetic theory of gases developed by Chapman and Enskog [2]:
T
id
-.
v = 26.69 -----------2
v

(8)

In formula (8), viscosity is in P, is the molar mass


(g/mol), T is absolute temperature (K), is the radius
of the molecule in the hard-sphere approximation (),
and v is the dimensionless collision integral depending on temperature and on the polarity of the molecules.

THEORETICAL FOUNDATIONS OF CHEMICAL ENGINEERING

Vol. 43

No. 3

2009

340

DEREVICH et al.

For nonpolar molecules, the collision integral is


determined using the following empirical formula [1]:
E'
C'
A'

v = ------B + --------------------------- + -------------------------- ,


T* exp ( D ' T* ) exp ( F ' T * )

(9)

(10)

where Pc is the critical pressure (atm).


The collision integral for polar molecules is estimated using the following formula, taking into account
expression (9):
2

(11)

Here, is a dimensionless parameter related to


polarity by the formula
2

dp
-3 ,
= ----------2

(12)

where dp is the dipole moment of the molecule (D).


Formulas (8)(12) are convenient for computerassisted calculations and allow dynamic viscosity to be
calculated in the ideal-gas approximation.
PENGROBINSON EQUATION OF STATE
In this section, we will report a computational procedure and formulas for determination of the necessary
model parameters using the PengRobinson equation
of state. This equation is written as follows [19]:
a ( , T r )
RT
-.
P = ------------ -----------------------------------------------V b V (V + b) + b(V b)

(13)

Here, the coefficients a and b are related with the


critical thermodynamic parameters of the substance in
the following way:
a =

2
RT c
0.45724 ---------,
2
Pc

RT c
b = 0.07780 ---------.
Pc

(14)

where Tr = T/Tc is the reduced temperature.


The equation of state (13) is a cubic polynomial of
the c compressibility factor Z = PV/(RT):
Z ( 1 B )Z + ( A 3B 2B )Z
3

where is Pitzers acentric factor and Tc is the critical


temperature (K).
The characteristic radius of the molecule () is estimated using the empirical formula

0.2

v = v + ------------- ,
T*

m ( ) = 0.37464 + 1.54226 0.26992 ,

( AB B B ) = 0.

----------- = ( 2.3551 0.0874 ),


kBT c

= ( 2.3551 0.0874 ) ( T c /P c ) ,

( , T r ) = [ 1 + m ( ) ( 1 T r ) ] ,
2

where A' = 1.16145, B' = 0.14874, C' = 0.52487, D' =


0.77320, E' = 2.16178, and F' = 2.43787 are empirical
constants; T = kBT/ is dimensionless absolute temperature; kB = 1.3806581023 J/K is the Boltzmann constant;
and is the characteristic value of the attractive potential (erg).
The /(kBTc) ratio is related to Pitzers acentric factor as

1/3

The function (, Tr) in (13) has the following unified form:

(15)

Here,
2

A = aP/R T ,

B = bP/RT .

Near the saturation line, the cubic equation (15) has


three real roots. The largest real root corresponds to the
vaporous state (Zv); the smallest root, to the liquid state
(Zl). At the critical point, all of the three roots are equal.
Far from the phase equilibrium line, Eq. (15) has a single real root. The molar volume of the substance in the
liquid or vaporous state is derived from the compressibility factor:
RT
V = Z -------.
P
The density of the substance is determined using the
molar mass :

= ---.
V
The thermodynamic parameters along the saturation
line and the saturation vapor pressure are also derived
from the equation of state (Eqs. (13), (15)). The liquid
vapor thermodynamic equilibrium condition reduces to
the equality of fugacity coefficients:
( Z v , T , P ) = ( Z l, T , P ),

(16)

The fugacity coefficient appearing in (16) is determined


using an equation of state:
V

1
RT
ln ( ) = Z 1 ln ( Z ) ------- P ------- dV ,

RT
V

(17)

where V is a sufficiently large gas volume corresponding to the ideal-gas approximation.


For the PengRobinson equation of state (13), formula (17) takes the following form:

THEORETICAL FOUNDATIONS OF CHEMICAL ENGINEERING

Vol. 43

No. 3 2009

CALCULATING THE DYNAMIC VISCOSITY OF INDIVIDUAL SUBSTANCES

106, Pa s

341

107, P s

104
103

1000

102

101

10

100

100

150

200

250

300

350
T, K

Fig. 1. Dynamic viscosity of propane along the saturation


line. The points represent experimental data [21].

ln ( ) = Z 1 ln ( Z B )
Z + ( 1 + 2 )B
A
-------------- ln ----------------------------------- .
Z + ( 1 2 )B
2 2B

(18)

The saturation vapor pressure Ps(T) at a given temperature is calculated as the root of Eq. (16) in the interval in which Eq. (15) has three real roots. The initial
approximation here is the LeeKesler approximation
for the saturation vapor pressure [20]. Various values of
the fugacity coefficient (18) calculated for the liquid
and vapor compressibilities are substituted into the
right- and left-hand sides of Eq. (16).
The deviation of the enthalpy from the ideal-gas
approximation is also determined using an equation of
state:
H

id

id

= H H = RT PV
V

P
P T ------ dV .
T V

(19)

For the PengRobinson equation (13), the deviation


of the enthalpy from the ideal-gas approximation (19)
is
H
------------ = 1 Z
RT
A
1 d
Z + ( 1 + 2 )B
+ -------------- 1 + --- -------- ln ----------------------------------- .

dT
r
2 2B
Z + ( 1 2 )B
id

(20)

Here, the derivative of the function (, Tr) is written as


d ( , T r )

------------------------ = m ( ) ----- .
Tr
dT r

10

50

100

150

200

250
T, C

Fig. 2. Dynamic viscosity of n-heptane along the saturation


line. The points represent experimental data [21].

Note that the deviation of a liquid or compressed gas


from the ideal-gas state is always positive: Hid/RT > 0.
Formulas (13)(20) allow all the thermodynamic
parameters of the substance in the liquid and vapor
phases to be calculated using the equation of state.
RESULTS OF THE CALCULATIONS
We will present a few plots to compare the data calculated within the above model to experimental data
available from the literature. The accuracy of the calculations is governed by the dynamic viscosity at the critical
point and by the value of the parameter , which appears in
formulas (5) and (6). In the estimation of the critical viscosity using the approximating formula (7), varies over the
interval 0.17 0.2. When experimental critical viscosity
data are used in the calculation, satisfactory results are
obtained at a fixed value of = 0.2. The parameter n in the
approximating formula for the supercritical region was set
to be 11 based on comparisons with experimental data for
the substances examined. Note that the saturation vapor
pressure, the phase equilibrium line, and all thermodynamic
properties were calculated using only the equation of state.
Thus, the viscosity calculation procedure is fully consistent
with thermodynamic calculations within the equation-ofstate method.
Figures 1 and 2 compare the calculated data with experimental viscosity data for propane and n-heptane along the
saturation line. Clearly, the theoretical and experimental
data are in satisfactory agreement. Figures 3 and 4 illustrate
the behaviors of the dynamic viscosities of carbon dioxide
and nitrogen along the saturation line. Note that our computational procedure allows the liquidvapor equilibrium
curve and the dynamic viscosities of both phases to be
determined using a unified approach.
The temperature dependence of the dynamic viscosity
of nitrogen on isobars is shown in Fig. 5. It is clear that our

THEORETICAL FOUNDATIONS OF CHEMICAL ENGINEERING

Vol. 43

No. 3

2009

342

DEREVICH et al.

106, P s

106, P s

102
102

101
101
200

225

250

275

300
T, K

Fig. 3. Dynamic viscosity of carbon dioxide along the saturation line. The points represent experimental data [22].

107, P s

70

80

90

100

110

120

130
T, K

Fig. 4. Dynamic viscosity of nitrogen along the saturation


line. The points represent experimental data [23].

107, P s
1
2
3
4

103

104

103
102
102
75

100

125

150

175

200

225
T, K

Fig. 5. Dynamic viscosity of nitrogen on isobars. The points


represent experimental data [21]. P = (1) 0.1, (2) 3, (3) 5,
and (4) 7 MPa.

calculations satisfactorily predict the viscosity jump due to


the change of the physical state of the substance. The deviation of the calculated data from the experimental data at
low temperatures (7080 K) might arise from the experimental error. It follows from Fig. 5 that our approach allows
the dynamic viscosity to be calculated in the supercritical
region, where the changes in viscosity are not accompanied
by a phase transition.
Figures 6 and 7 illustrate the comparison between
theoretical and experimental dynamic viscosity data for
n-octane. n-Octane is a well-studied model compound
used for example, as the reference in the LeeKesler
equation. The calculated and experimental data are in

100

100

150

200

250

300

350

300
T, C

Fig. 6. Viscosity of n-octane along the saturation line. The


points represent experimental data [18].

good agreement. Note that our computational procedure


is accurate at supercritical pressures as well.
For further illustration of our approach, we will compare
theoretical and experimental data for isobutane on the saturation line (Fig. 8). The satisfactory agreement between the
theoretical and experimental data suggests that the dynamic
viscosity is determined by the standard set of critical thermodynamic parameters and by Pitzers acentric factor.
For all of the compounds examined, we calculated
the standard deviation between theoretical and experimental data. On the saturation line, the standard error
does not exceed 6% for the vapor and 8% for the liquid.
For the supercritical state, the maximum standard error
is below 9%.

THEORETICAL FOUNDATIONS OF CHEMICAL ENGINEERING

Vol. 43

No. 3 2009

CALCULATING THE DYNAMIC VISCOSITY OF INDIVIDUAL SUBSTANCES

107, P s
1
2
3
4

103

343

model varies within a limited range. When experimental


critical viscosity data are used, this empirical parameter
becomes universal. The calculated data are compared with
representative experimental data for paraffins and simple
gases on saturation lines and isobars, including supercritical pressures.
The computational procedure can be extended to the
viscosity of organic mixtures.
ACKNOWLEDGMENTS
This work was supported by the Russian Foundation
for Basic Research, project no. 07-08-12292-OFI.

102
50

100

150

200

250

300

350
T, C

Fig. 7. Dynamic viscosity of n-octane on isobars. The points


represent experimental data [21]. P = (1) 0.1, (2) 1, (3) 3,
and (4) 6 MPa.

108, P s

104

103

150

200

250

300

350
T, C

Fig. 8. Dynamic viscosity of isobutane along the saturation


line. The points represent experimental data [21].

CONCLUSIONS
Based on the hypothesis that momentum transfer in the
liquid takes place via the formation of an excited zone
with a lower molecule concentration, we suggested a simple procedure for calculating the dynamic viscosities of
the liquid and gas over wide ranges of pressure and temperature. The energy barrier to the formation of the excited
zone was assumed to be proportional to the deviation of
the enthalpy of the substance from the enthalpy of the
ideal gas. The viscosity calculations for the liquid and gas
were extended to supercritical pressures. In the estimation
of the dynamic viscosity at the critical point using the
approximating formula, the empirical parameter of the

NOTATION
Aminminimum work necessary for the formation
of a vacancy, J;
dpdipole moment of the molecule, D;
kBBoltzmann constantsnt, J/K;
Ppressure, Pa;
Runiversal das constant, J/(mol K);
Tabsolute temperature, K;
Vmolar volume, m3/mol;
wprobability of the formation of a vacancy;
Zcompressibility factor of the substance;
dimensionless parameter of the model;
characteristic value of the attractive potential,
erg;
dynamic viscosity, Pa s;
dimensionless parameter taking into account the
polarity of the molecules;
molar weight, g/mol;
density, kg/m3;
radius of the molecule, ;
fugacity coefficient;
v collision integral;
Pitzers acentric factor.
SUBSCRIPTS AND SUPERSCRIPTS
cvalue at the thermodynamic critical point;
lliquid state;
ssaturation line;
0parameter of the medium;
rreduced thermodynamic parameter;
vvaporous state;
idideal gas.
REFERENCES
1. Reid, R., Prausnitz, J., and Sherwood, T., The Properties
of Gases and Liquids, New York: McGraw-Hill, 1977.

THEORETICAL FOUNDATIONS OF CHEMICAL ENGINEERING

Vol. 43

No. 3

2009

344

DEREVICH et al.

2. Chapman, S. and Cowling, T.G., Mathematical Theory


of Non-Uniform Gases, Cambridge: Cambridge Univ.
Press, 1952.
3. Al-Syabi, Z., Danesh, A., Tohidi, B., et al., A Residual
Viscosity Correlation for Predicting the Viscosity of Petroleum Reservoir Fluids over Wide Ranges of Pressure and
Temperature, Chem. Eng. Sci., 2001, vol. 56, p. 6997.
4. Mehrotra, A.K., Monnery, W.D., and Svrcek, W.Y., A
Review of Practical Calculation Methods for the Viscosity of Liquid Hydrocarbons and Their Mixtures, Fluid
Phase Equilib., 1996, vol. 117, p. 344.
5. McCarty, R.D., Corresponding States for Transport
Properties, Fluid Phase Equlib., 2007, vol. 256, p. 42.
6. Komarov, V.N. and Tovbin, Yu.K., Self-Consistent Calculation of the Compressibility and Viscosity of Dense
Gases in the Lattice-Gas Model, Teplofiz. Vys. Temp.,
2003, vol. 41, no. 2, p. 217 [High Temp. (Engl. Transl.),
vol. 41, no. 2, p. 181].
7. Kelkar, M.S., Rafferty, J.L., Maginn, E.J., and Siepmann, J.I., Prediction of Viscosities and VaporLiquid
Equilibria for Five Polyhydric Alcohols by Molecular
Simulation, Fluid Phase Equilib., 2007, vol. 260, p. 218.
8. Qun-Fang, L., Yu-Chun, H., and Rui-Sen, L., Correlation
of Viscosities of Pure Liquids in a Wide Temperature
Range, Fluid Phase Equilib., 1997, vol. 140, p. 221.
9. Eyring, H. and Jhon, M.S., Significant Liquid Structures,
London: Wiley-Interscience, 1969.
10. Frankl, Ya.I., Kineticheskaya teoriya zhidkostei (Kinetic
Theory of Liquids), Leningrad: Nauka, 1975.
11. Fan, W. and Lei, Q., Generalized Correlation for Predicting the Kinematic Viscosity of Liquid Petroleum Fractions, Fluid Phase Equilib., 1999, vol. 166, p. 125.
12. Qunfang, L. and Yu-Chung, H., Correlation of Viscosity
of Binary Liquid Mixtures, Fluid Phase Equilib., 1999,
vol. 154, p. 153.
13. Quinones-Cisneros, S.E., Zeberg-Mikkelsen, C.K., and
Stenby, E.H., The Friction Theory (f-Theory) for Viscos-

14.
15.

16.

17.

18.
19.
20.
21.

22.
23.

ity Modeling, Fluid Phase Equilib., 2000, vol. 169,


p. 249.
Quinones-Cisneros, S.E., Zeberg-Mikkelsen, C.K., and
Stenby, E.H., One Parameter Friction Theory Models for
Viscosity, Fluid Phase Equilib., 2001, vol. 178, p. 1.
Fan, T.-B. and Wang, L.-S., A Viscosity Model Based on
PengRobinson Equation of State for Light Hydrocarbon Liquids and Gases, Fluid Phase Equilib., 2006, vol.
247, p. 59.
Macias-Salinas, R., Garcia-Sanchez, F., and EliosaJimenez, G., An Equation-of-State-Based Viscosity
Model for Non-Ideal Liquid Mixtures, Fluid Phase
Equilib., 2003, vol. 210, p. 319.
Derevich, I.V. and Gromadskaya, R.S., Effect of Dissolved Gases on the Viscosity of Petroleum, Teor. Osn.
Khim. Tekhnol., 2002, vol. 36, no. 6, p. 583 [Theor. Found.
Chem. Eng. (Engl. Transl.), vol. 36, no. 6, p. 583].
Landau, L.D. and Lifshits, E.M., Teoreticheskaya fizika
(Theoretical Physics), vol. 5: Statisticheskaya fizika
(Statistical Physics), Moscow: Nauka, 1976.
Peng, D.-Y. and Robinson, D.D., A New Two-Constant
Equation of State, Ind. Eng. Chem. Fundam., 1976,
vol. 15, p. 58.
Lee, B.I. and Kesler, M.G., A Generalized Thermodynamic Correlation Based on Three-Parameter Corresponding States, AIChE J., 1975, vol. 21, no. 3, p. 510.
Vargaftik, N.B., Spravochnik po teplofizicheskim svoistvam gazov i zhidkostei, (A Handbook of Thermophysical
Properties of Gases and Liquids), Moscow: Nauka,
1972.
Fenghour, A., Wakeham, W.A., and Vesovich, V., The
Viscosity of Carbon Dioxide, J. Phys. Chem. Ref. Data,
1998, vol. 27, no. 1, p. 31.
Stephan, K., Krauss, R., and Laesecke, A., Viscosity and
Thermal Conductivity of Nitrogen in a Wide Range of
Fluid States, J. Phys. Chem. Ref. Data, 1987, vol. 16,
no. 4, p. 993.

THEORETICAL FOUNDATIONS OF CHEMICAL ENGINEERING

Vol. 43

No. 3 2009

Vous aimerez peut-être aussi