Anda di halaman 1dari 15

Transport in Porous Media 51: 327341, 2003.

327
2003 Kluwer Academic Publishers. Printed in the Netherlands.

The Use of Ficks Law for Modeling Trace Gas


Diffusion in Porous Media

STEPHEN W. WEBB1 and KARSTEN PRUESS2


1 Environmental Technology Department, 6131/MS-0719, Sandia National Laboratories,
Albuquerque, NM 87185-0719, U.S.A.
2 Earth Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, CA 94720, U.S.A.

(Received: 7 June 2001; in final form: 23 October 2002)


Abstract. Two models for combined gas-phase diffusion and advection in porous media, the
advective-diffusive model (ADM) and the dusty-gas model (DGM), are commonly used. The ADM
is based on a simple linear addition of advection calculated by Darcys law and ordinary diffusion
using Ficks law with a porositytortuositygas saturation multiplier to account for the porous me-
dium. The DGM applies the kinetic theory of gases to the gaseous components and the porous
media (or dust) to develop an approach for combined transport due to diffusion and advection
that includes porous medium effect. The ADM and Ficks law are considered to be generally in-
ferior for gas diffusion in porous media, and the more mechanistic DGM is preferred. Under trace
gas diffusion conditions, Ficks law overpredicts the gas diffusion flux compared to the DGM.
The difference between the two models increases as the permeability decreases. In addition, the
difference decreases as the pressure increases. At atmospheric pressure, the differences are minor
(<10%) for permeabilities down to about 1013 m2 . However, for lower permeabilities, the dif-
ferences are significant and can approach two orders of magnitude at a permeability of 1018 m2 .
In contrast, at a pressure of 100 atm, the maximum difference for a permeability of 1018 m2 is
only about a factor of 2. A moleculewall tortuosity coefficient based on the DGM is proposed
for trace gas diffusion using Ficks law. Comparison of the Knudsen diffusion fluxes has also been
conducted. For trace gases heavier than the bulk gas, the ADM mass flux is higher than the DGM.
Conversely, for trace gases lighter than the bulk gas, the ADM mass flux is lower than the DGM.
Similar to the ordinary diffusion variation, the differences increase as the permeability decreases,
and get smaller as the pressure increases. At atmospheric pressure, the differences are small for
higher permeabilities (>1013 m2 ) but may increase to about 2.7 for He at lower permeabilities
of about 1018 m2 . A modified Klinkenberg factor is suggested to account for differences in the
models.

Key words: dusty-gas model, advective-diffusive model, Ficks law, Knudsen diffusion,
Klinkenberg, trace gas diffusion, gas diffusion, tortuosity.

Nomenclature
b Klinkenberg factor (Pa1 ).
DiK Knudsen diffusion coefficient for gas i (m2 /s).
Drat /D .
D12 1K
D12 binary diffusion coefficient in free space at the pressure and temperature conditions of
interest (m2 /s).
328 STEPHEN W. WEBB AND KARSTEN PRUESS

0
D12 binary diffusion coefficient in free space at 1 bar and 0 C (m2 /s).

D12 effective binary diffusion coefficient (Eq. (3)) (m2 /s).
F mass flux (kg/m2 s).
k effective permeability including slip (m2 ).
k0 intrinsic permeability (m2 ).
m molecular weight (kg/kg mol).
mrat m2 /m1 .
ND diffusive molar flux (kg mol/m2 s).
NT total (diffusive plus advective) molar flux (kg mol/m2 s).
P pressure (Pa).
P0 pressure of 1 bar (Pa).
R universal gas constant (J/kg mole K).
Sg gas saturation ().
T temperature (K).
x mole fraction ().
Greek Symbols
porous media factor ().
exponent on diffusion coefficient ().
viscosity (kg/m s).
density (kg/m3 ).
tortuosity ().
porosity ().
mass fraction ().
Subscripts
air air.
DGM dusty-gas model.
g gas phase.
i, j species or component.
r relative.
1 trace gas.
2 bulk gas.

1. Introduction
Gas-phase diffusion in porous media is an important factor in flow and transport in
the subsurface. Applications include contaminant migration and removal from low-
permeability layers in the subsurface where gas-phase diffusion may be the limiting
transport mechanism (Ho and Udell, 1992; Webb and Phelan, 1997), landfills and
waste repositories (Thibodeaux et al., 1986), evaporation and drying (Prat, 1993),
and chemical transport from buried landmines and unexploded ordnance (Webb
et al., 1999). Other porous media application include porous catalysts (Jackson,
1977) and gas nuclear reactors (Evans et al., 1961, 1962a, b, 1963).
A number of different models have been used to quantify gas diffusion pro-
cesses in porous media. Ficks law is the most popular approach to gas diffu-
sion due to its simplicity. However, application of Ficks law to porous media
MODELING TRACE GAS DIFFUSION 329

gas diffusion has recently been questioned by a number of investigators including


Thorstenson and Pollock (1989), Abirola et al. (1992) and Webb (1998). Another
approach often employed is the StefanMaxwell equations. This equation set is
simply an extension of Ficks law for a multicomponent mixture as discussed by
Bird et al. (1960). Baehr and Bruell (1990) compared the StefanMaxwell model
to Ficks law for transport of an organic vapor in a sand column and concluded
that Ficks law has serious limitations. Jaynes and Rowgowski (1983) compared
the equations for Ficks law and the StefanMaxwell equations and noted special
conditions where Ficks law is appropriate.
Ficks law and the StefanMaxwell equations apply to gas diffusion in open
spaces, not in porous media. While attempts have been made to define effective
diffusion parameters to account for the presence of the porous medium, the basic
transport equations are not altered. More recently, a model has been developed
that addresses many of the shortcomings of Ficks law or the StefanMaxwell
equations for gas transport in porous media. The dusty-gas model (DGM) takes the
gas transport equations a step further by including the effect of the porous media as
a dusty gas component of the gas mixture. The dusty gas is assumed to consist
of large molecules fixed in space the are treated as a component of the gas mixture.
The kinetic theory of gases is applied to this dusty-gas mixture. Knudsen diffu-
sion (moleculewall interaction) is inherently included in the model, and a porous
media factor similar to that described for the advective-diffusive model (ADM)
is used. The DGM, including numerous data-model comparisons, is discussed in
detail by Mason and Malinauskas (1983) and Cunningham and Williams (1980).
An important difference between the model is the coupling between processes.
In the ADM, advection and diffusion fluxes are linearly added together, which
while conceptually appealing is simply ad hoc. Knudsen diffusion is included by
modifying the permeability through the Klinkenberg parameter. This approach is
particularly popular as exemplified by Abriola and Pinder (1985) and Pruess et al.
(1999). In contrast, the DGM theoretically evaluates the coupling effects between
ordinary diffusion, Knudsen diffusion, and advection.
A number of investigators have compared these two approaches. Webb (1998)
compared the ADM and DGM models with Grahams laws (see Mason and
Malinauskas, 1983 or Cunningham and Williams, 1980), which are fundamental
relationships for diffusion, and evaluated both models against experimental data.
He found that the DGM is in agreement with Grahams laws, and comparisons
with experimental data are excellent. In contrast, the ADM does not agree with
Grahams laws, and comparisons with experimental data are much poorer for the
ADM than for the DGM, even for pure diffusion conditions. Massmann and Farrier
(1992) compared the two models including the effects of atmospheric pressure
variations and found similar results. Abriola et al. (1992) evaluated the ADM and
DGM for binary gas mixtures, concluding that Ficks law is not valid under many
conditions. Further assessment of the coefficients in the DGM for natural porous
media has been conducted by Abu-El-Shar and Abriola (1997), while Sleep (1998)
330 STEPHEN W. WEBB AND KARSTEN PRUESS

compared the DGM to Ficks law for the transport of organic vapors in numerical
simulations. In general, ADM and Ficks law are considered to be generally inferior
for gas diffusion in porous media, and the more mechanistic DGM is preferred.
An important facet of gas-phase diffusion is trace gas diffusion. For example,
Jury et al. (1983, 1984a, b, c) developed a screening model for pesticides, which
essentially treats the pesticide vapor as a trace gas. Webb et al. (1999) have con-
sidered the flux of chemical signatures from buried landmines and unexploded
ordnance, where the chemical signatures are trace gases. Trace gas diffusion is
also important in the transport of VOCs in the subsurface as considered by Jury
et al. (1990) and McCarthy and Johnson (1995). All of the above approaches, with
the exception of Webb et al. (1999), used Ficks law; Webb et al. (1999) used the
DGM. While the ADM (Ficks law) and the DGM have been compared for general
gas diffusion including advection, no evaluation has been performed for trace gas
diffusion conditions.
The present investigation concentrates on the modeling of trace gas diffusion
in porous media and the relationship between the Ficks law and the DGM. In
particular, based on data-model comparisons discussed above, Ficks law does not
adequately include the porous media effects for low permeability media. Modifi-
cation of Ficks law is proposed to include the porous media effects in the DGM.
Binary gas diffusion for all-gas conditions (Sg = 1.0) is considered for simplicity
where the trace gas is one of the gases.

2. Diffusion Models
2.1. ADVECTIVE - DIFFUSIVE MODEL
As discussed earlier, the ADM is based on an ad hoc linear addition of advection
calculated by Darcys law and ordinary diffusion using Ficks law. Slip effects,
or Knudsen diffusion, are included through a Klinkenberg parameter to define an
effective permeability for the advective flux. Porous medium effects for ordinary
diffusion are included through a porositytortuositygas saturation factor applied
to the diffusive flux in free space. This simple additive approach, while intuitively
appealing, ignores coupling between advective and diffusive mechanisms.
The ADM used in the present study is based on the implementation of Pruess
(1987, 1991) and Pruess et al. (1999). For all-gas conditions, the mass flux for
gas-phase component i, Fi , in a binary mixture, neglecting gravity, is
k
Fi = g i Pg D12 g i (1)
g

where D12 = D21 is assumed.


Slip, or Knudsen diffusion, is included through the Klinkenberg factor, b, or
 
b
k = k0 1 + (2)
Pg
MODELING TRACE GAS DIFFUSION 331

The effective diffusion coefficient is given by


 
0 P0 T + 273.15
D12 = Sg D12 = Sg D12 = D12 (3)
P 273.15
where the exponent is typically 1.8 for an airwater vapor mixture (Pruess, 1987).
The term Sg is commonly referred to as the porous media factor, or .
For isobaric conditions (Pg = 0), there is no advection, and for a binary mix-
ture the ADM reduces to Ficks law, or
F1 = D12 g 1 (4a)

F2 = D12 g 2 (4b)

For a binary mixture, 1 = 2 , so


F1 = F2 (5)
and the mass fluxes of the two components are equal and opposite.

2.2. DUSTY- GAS MODEL


The DGM was developed to describe gas transport through porous media including
the coupling between the various transport mechanisms. The term dusty-gas is
used because the porous medium is assumed to consist of large molecules fixed
in space. Similar to the ADM, the DGM considers advection, Knudsen diffusion,
and ordinary diffusion. One of the key aspects of the DGM is the combination
of diffusion (ordinary and Knudsen) and advection. Ordinary and Knudsen diffu-
sion are combined through addition of momentum transfer based on kinetic theory
arguments, and diffusive fluxes (ordinary plus Knudsen) are added to advective
fluxes based on ChapmanEnskog kinetic theory. The DGM is discussed in detail
by Mason and Malinauskas (1983) and Cunningham and Williams (1980).
Ignoring thermal diffusion, which is typically small, the DGM can be written
either in terms of the diffusive molar flux, N D , or the total molar flux (diffusive plus
advective), N T (Thorstenson and Pollack, 1989). The two expression for all-gas
conditions are:

n
xi NjD xj NiD NiD Pg xi xi Pg
= + (6a)
j=1
Dij DiK RT RT
j=i

  
n
xi NjT xj NiT NiT Pg xi k0 Pg xi Pg
= + 1+ (6b)
j=1
Dij DiK RT DiK g RT
j=i

The first term on the left-hand side of both equations considers moleculemolecule
interactions and is based on the StefanMaxwell equations. The second term
332 STEPHEN W. WEBB AND KARSTEN PRUESS

considers moleculeparticle (porous media) interactions, while the right-hand side


is the driving force for diffusion and advection, which includes mole fraction and

pressure gradients. The effective binary diffusion coefficient, D12 , is the same as
that for the ADM (Equation (3)) in that it is simply the binary diffusion coefficient
modified by effective gas porosity and tortuosity factors (Mason and Malinauskas,
1983).
There are many alternative forms of the DGM that can be derived by combin-
ing various forms of the equations. In particular, as detailed by Thorstenson and
Pollock (1989), the total mass flux of component 1 in an isothermal binary system
can be written as

F1 = m1 N1T

[D1K D12 (Pg /RT )x1 + D1K (D12 + D2K )x1 (Pg /RT )]
= m1
(D12 + x1 D2K + x2 D1K )
k0 Pg Pg
x1 m1 (7)
g RT

The flux of component 1 consists of a diffusive flux (first term) and an advec-
tive flux (second term). The diffusive flux has ordinary diffusion (mole fraction
gradient) and Knudsen diffusion (pressure gradient) components, similar to the

ADM equation. The effective ordinary diffusion coefficient, D12 , is calculated in
the same manner as described in the ADM section. Calculation of the Knudsen
diffusion coefficient, DiK , is discussed later in this paper.
Note that in the special case of isobaric conditions (Pg = 0), the advective and
Knudsen diffusion fluxes are zero. However, this does not mean that the Knudsen
diffusion coefficients are not important. The ordinary diffusion flux is dependent
on both Knudsen ordinary diffusion coefficients; in the ADM, only the ordinary
diffusion coefficient is used. The Knudsen diffusion coefficients characterize the
impact of the porous media (moleculewall interactions) on ordinary diffusion,
which is not included in the ADM.

2.3. COMPARISON OF MODELS


The equations for the two models look quite different. The ADM uses a mass
fraction gradient as its driving force for diffusion, while the DGM employs a mole
fraction gradient. One can easily show that the advective flux calculated by both
models is equal if the flux from Knudsen diffusion (Klinkenberg factor) is included
as a separate term (Webb, 1998). The DGM is rewritten in terms of a mass fraction
gradient for ease of comparison to Ficks law. In addition, the molar density is re-
written in terms of the mass density. These operations yield the following equations
for the diffusive fluxes (ordinary and Knudsen) without the advection term
MODELING TRACE GAS DIFFUSION 333

ADM:

k0 b
F1 = D12 g 1 g 1 Pg (8)
g Pg

DGM:

D1K D12 g (x1 m1 + x2 m2 )
F1 =
1
(D12 + x1 D2K + x2 D1K ) m2

D1K (D12 + D2K )x1
m1 Pg (9)
RT (D12 + x1 D2K + x2 D1K )

Both expressions have contributions from mass fraction and pressure gradients,
although the coefficients in front of the two driving forces are considerably differ-
ent. For example, in the ADM, the coefficient for the ordinary diffusion expression
(mass fraction gradient term) is a constant, while the DGM expression is a function
of the mole fraction.
Now consider a trace gas, labeled gas 1, such that

x1 , x2 1.0

The flux given by the above equations will be developed for the trace gas. Using
this limiting behavior, the two flux equations become
ADM (no changes):

k0 b
F1 = D12 g 1 g 1 Pg (10)
g Pg

DGM:

D1K D1K (D12 + D2K )
F1 = D g 1 g 1 Pg (11)
(D12 + D1K ) 12
Pg (D12 + D1K )

Looking at the mass fraction gradient expression first, the two expressions can be
made to be identical for trace gas diffusion if another coefficient is added to Ficks
law (first term of ADM). The ordinary diffusion flux is proportional to the mass
gradient as indicated by Ficks law; the factor out front just needs to be modified.
This added coefficient will be denoted as DGM to designate it as a tortuosity factor
due to wallmolecule interactions in the DGM, or

D1K
DGM = = (1 + Drat )1 (12)
(D12 + D1K )

where

D12
Drat = (13)
D1K
334 STEPHEN W. WEBB AND KARSTEN PRUESS

The pressure gradient, or Knudsen diffusion, terms are much different. As dis-
cussed by Thorstenson and Pollock (1989), the Knudsen diffusion coefficient and
the Klinkenberg parameter are related by
k0 bi
DiK = (14)
i
Knudsen diffusion coefficients are needed for the bulk gas (gas 2) and the trace gas
(gas 1). The Knudsen diffusion coefficient is a function of the molecular weight
of the gas due to the variation of the mean free path. The coefficients for different
gases are related by
 1/2
mi
Dj K = DiK (15)
mj
The resulting variation in the Klinkenberg parameter, b, for different gases from
the above expressions agrees with the data shown in Klinkenberg (1941).
The DGM expression can be further rewritten as

k0 bg (D12 (m2 /m1 )1/2 + D1K )
F1 = DGM D12 g 1 g 1 Pg (16)
g Pg (D12 + D1K )
A second correction factor can be defined as the Klinkenberg factor for the DGM,
bDGM , where
1/2
(D12 (m2 /m1 )1/2 + D1K ) 1 + mrat Drat
bDGM = = (17)
(D12 + D1K ) 1 + Drat
where
m2
mrat = (18)
m1
and the final DGM expression for trace gas diffusion is

k0 b
F1 = DGM D12 g 1 bDGM g 1 Pg (19)
g Pg
With this first correction factor, DGM , Ficks law can be used to predict trace gas
diffusion in a binary mixture. With the second correction factor, bDGM , the ADM
can be used to calculate trace gas Knudsen diffusion in a binary mixture.

2.3.1. Evaluation of Parameter Values


The Klinkenberg factor for air can be estimated from Heid et al. (1950) correlation
for air at 25 C as presented by Thorstenson and Pollock (1989), or

bair = 0.11k00.39 (20)


MODELING TRACE GAS DIFFUSION 335

where bair is in Pa and k0 is the intrinsic permeability in m2 . The data used in this
correlation are from oil-field cores with permeability values between about 1012
and 1017 m2 . Subsequently, Jones and Owens (1980) measured permeabilities on
low-permeability gas sands between 1014 and 1019 m2 ; their correlation can be
expressed as
bair = k00.33 (21)
Between 1014 and 1017 m2 , which is where the permeabilities for the data over-
lap, the Klinkenberg factors from both correlations are quite similar. For simplicity,
the Heid et al. (1950) correlation is used in the present evaluation.
The Klinkenberg and Knudsen diffusion coefficients are proportional to the
square root of the absolute temperature. The temperature of Heid et al. (1950)
correlation is reported to be at 25 C by Thorstenson and Pollock (1989).

3. Results and Discussion


Typical values of DGM and bDGM will be evaluated in this section. Assumptions
made in this evaluation include air as the bulk gas at 25 C, and an air viscosity of
1.837105 Pa s (Bird et al., 1960). The pressure is varied between 1 and 100 atm.
Assumed properties for the porous media include a porosity of 0.4, a gas saturation
of 1.0, and a tortuosity of 0.74 as calculated by Millington and Quirk (1961) model.

3.1. DGM

The value of DGM depends on D12 and DiK . D12 , the effective ordinary diffusion
coefficient, depends on the tortuosity, porosity, gas saturation, pressure, temper-
ature, and the diffusing gases as shown in Equation (3). DiK , the Knudsen dif-
fusion coefficient, depends on the diffusing gases and the intrinsic permeability.
Because the value of the Klinkenberg coefficient is based on data from porous
media experiments, the numbers are effective values and do not need additional
correction.
In the results that follow, trace gas diffusion in air is presented for intrinsic
permeabilities ranging from 1010 to 1018 m2 . Three different trace gases in air in
air are considered with widely different molecular weights. They are helium (m =
4.003), water vapor (m = 18.016), and trichloroethylene (TCE) (m = 131.389).

The porous media ordinary diffusion coefficients, D12 are listed in Table I along
with the Knudsen diffusion coefficients for air at various intrinsic permeabilities at
1 atm. As noted earlier, the ordinary diffusion coefficients are inversely propor-
tional to pressure as indicated by Equation (3). In this evaluation, the reference
pressure, P0 , is equal to 105 Pa. Note that the porous media ordinary diffusion coef-
ficient is the free space value times the tortuosityporositygas saturation product,
or 0.296. For airwater vapor, the diffusivity as presented by Massman (1998) has
been used. For airHe and airTCE, the Fuller et al. model as presented by Reid
336 STEPHEN W. WEBB AND KARSTEN PRUESS

Table I. Diffusion coefficients



Porous medium ordinary diffusion coefficients 1 atm and 25 C (m2 /s), D12
Water vaporair 7.58 106
Heliumair 2.02 105
TCE-air 2.44 106

Knudsen diffusion coefficients for air (m2 /s), DK
k0 = 1010 m 2 4.76 103
k0 = 1014 m2 1.73 105
k0 = 1018 m2 6.27 108

et al. (1987) has been used. The Knudsen diffusion coefficient has been calculated
for air and then modified for the trace gas of interest by Equation (15).
Figure 1 shows the variation of the Knudsen diffusion coefficient for air.
The Knudsen diffusion coefficient varies widely from 4.76103 m2 /s at 1010 m2
down to 6.27 108 m2 /s at 1018 m2 ; from Equation (15), the corresponding val-
ues for water vapor are 6.03 103 and 7.9 108 m2 /s. For comparison, the air
water vapor porous medium ordinary diffusion coefficient, D12 , is 7.58106 m2 /s.
Drat varies from 1.26 103 to 95.0 for water vapor, or about five orders of mag-
nitude over the range of permeability.

Figure 1. Knudsen diffusion coefficient for air.


MODELING TRACE GAS DIFFUSION 337

Figure 2 shows the value of DGM for trace gas diffusion of helium, water va-
por, and TCE in air as a function of permeability and pressure. At 1 atm, DGM
approaches unity for permeabilities greater than 1013 m2 . However, as the per-
meability decreases, DGM decreases dramatically, reaching a value of about 0.01

Figure 2. DGM and bDGM factors as a function of pressure, intrinsic permeability and trace
gas.
338 STEPHEN W. WEBB AND KARSTEN PRUESS

at a permeability of 1018 m2 . At 100 atm, DGM is near unity for permeabilities


greater than 1016 m2 . The minimum value of DGM at this pressure is 0.5.
Thus, for media with permeabilities of 1012 m2 or larger, Ficks law and the
DGM will predict essentially the same ordinary diffusion mass flux for trace gas
diffusion. However, for lower permeability media at 1 atm, the differences can be
up to two orders of magnitude, with the DGM mass flux being significantly smaller
than the Ficks law mass flux.
The variation in DGM for the various trace gases is minor compared to the over-
all variation. For TCE, the ordinary diffusion coefficient is significantly smaller
than for water vapor. However, the Knudsen diffusion coefficient is also much
smaller as is evident from Equation (15). Therefore, the value Drat shows only a
minor change between water vapor and TCE. Similarly, both diffusion coefficients
are larger in the case of He.
The large variation with permeability is driven by the significant variation in
the Knudsen diffusion coefficient. The differences in the predicted diffusion mass
fluxes are not due to the failure of the form of Ficks law for trace gas diffusion;
ordinary diffusion is still directly proportional to the mass or mole flux gradient.
Rather, it is the failure of the definition of the porous media factor in Equation (3).
The porous media factor in the ADM does not consider the effects of wallmolecule
interactions, which are included in the DGM separately.
Support for the use of a much lower total tortuosity coefficient, which includes
DGM , comes from the low-permeability graphite (2.131018 m2 ) experiment con-
ducted by Evans et al. (1962a, b, 1963) in the development of the DGM. While
their experimental data are not directly applicable because the data are not for
trace gas diffusion, the magnitude can be used to qualitatively evaluate the present
model. Based on gas diffusion measurements, the porositytortuosity factor of the
graphite in their experiments was 1.42104 . The porosity of the sample was 0.11,
so the tortuosity factor is 1.29 103 . Based on Millington and Quirk (1961),
the porous media tortuosity coefficient is about 0.05, and the measured value is
about a factor of 40 lower. This reduction is primarily due to the wallmolecule
interactions, which are included in DGM . The value of DGM in the case presented
here is 0.015, so the total tortuosity factor is 7.5104 , or within a factor of 2 of the
data. Therefore, the data of Evan et al. (1962a, b, 1963) support the present model.

3.2. bDGM
Similar to DGM , the value of bDGM has been evaluated as a function of permeability
and trace gas as a function of pressure. The variation of bDGM is also shown in
Figure 2. The variation is small compared to the variation in the ordinary dif-
fusion ratio, although the molecular weight makes a significant difference of the
magnitude of the factor. At 1 atm, the values range from 2.65 to 1.0 for He (low
permeability to high permeability variation), 1.27 to 1.0 for water vapor, and 0.47
to 1.0 for TCE. The values become closer to 1.0 as the pressure increases.
MODELING TRACE GAS DIFFUSION 339

4. Conclusions
Ficks law and the DGM have been compared for trace gas conditions. The porous
media permeability and trace gas has been varied to quantify the differences. The
conclusions from this study are:
1. Ordinary diffusion mass fluxes calculated by Ficks law should be corrected by
the addition of DGM , which addresses moleculewall interactions. At 1 atm,
the value of DGM is about unity at higher permeabilities (>1013 m2 ). For
lower permeabilities, Ficks law dramatically overpredicts the gas diffusion
flux. The value of DGM decreases to about 0.01 at a permeability of 1018 m2 .
The difference is much smaller as the pressure increases. At 100 atm, the value
of DGM at a permeability of 1018 m2 is 0.5. The variation of the ratio for
different trace gases is small.
2. A correction factor for Knudsen diffusion is also proposed to account for wall
molecule effects. The variation is much smaller than for ordinary diffusion. At
1 atm, the ratio of the mass fluxes of the DGM to the ADM is about 1.0 at
higher permeabilities (>1013 m2 ). For lower permeabilities, the ratio depends
on the molecular weight of the trace gas. At a permeability of 1018 m2 and
1 atm of air, the DGM to ADM mass flux ratio is about 2.65 for He, 1.27 for
water vapor, and about 0.47 for TCE. The effect of pressure is small.
With these added terms, Ficks law and the ADM can be used to predict trace gas
diffusion in binary systems including wallmolecule interactions.

Acknowledgements
The authors wish to thank Mike Itamura and Eric Lindgren of Sandia National
Laboratories, as well as the anonymous reviewers, for their comments, which
greatly improved the paper. This work was supported by the Strategic Environ-
mental Research and Development Program (SERDP) for application to unex-
ploded ordnance chemical sensing and the Defense Advanced Research Projects
Agency (DARPA) for application to buried landmines. Sandia is a multiprogram
laboratory operated by Sandia Corporation, a Lockheed Martin Company, for the
United States Department of Energy under Contract DE-AC04-94AL85000.

References
Abriola, L. M. and Pinder, G. F.: 1985, A multiphase approach to the modeling of porous media
contamination by organic compounds. 1. Equation development, Water Resour. Res. 21, 1118.
Abriola, L. M., Fen, C.-S. and Reeves, H. W.: 1992, Numerical simulation of unsteady organic vapor
transport in porous media using the dusty gas model, in: Wyre (ed.), Subsurface Contamination
by Immiscible Fluids, Balkema, Rotterdam, pp. 195202.
Abu-El-Shar, W. and Abriola, L. M.: 1997, Experimental assessment of gas transport mechanisms
in natural porous media: Parameter evaluation, Water Resour. Res. 33, 505516.
340 STEPHEN W. WEBB AND KARSTEN PRUESS

Baehr, A. L. and Bruell, A. J.: 1990, Application of the StefanMaxwell equations of determine
limitations of Ficks law when modeling organic vapor transport in sand columns, Water Resour.
Res. 26, 11551163.
Bird, R. B., Stewart, W. E. and Lightfoot, E. N.: 1960, Transport Phenomena, Wiley, New York.
Cunningham, R. E. and Williams, R. J. J.: 1980, Diffusion in Gases and Porous Media, Plenum Press,
New York.
Evans III, R. B., Watson, G. M. and Mason, E. A.: 1961, Gas diffusion in porous media at uniform
pressure, J. Chem. Phys. 35, 20762083.
Evans III, R. B., Watson, G. M. and Mason, E. A.: 1962a, Gas diffusion in porous media. II. Effect
of pressure gradients, J. Chem. Phys. 36, 18941902.
Evans III, R. B., Watson, G. M. and Truitt, J.: 1962b, Interdiffusion of gases in a low permeability
graphite at uniform pressure, J. Appl. Phys. 33, 2682.
Evans III, R. B., Watson, G. M. and Truitt, J.: 1963, Interdiffusion of gases in a low permeability
graphite. II. Influence of pressure gradients, J. Appl. Phys. 34, 2020.
Heid, J. G., McMahon, J. J., Nielson, R. F. and Yuster, S. T.: 1950, Study of the permeability of rocks
to homogeneous fluids, API Drilling Prod. Pract., 230244.
Ho, C. K. and Udell, K. S.: 1992, An experimental investigation of air venting of volatile liquid
hydrocarbon mixtures from homogeneous and heterogeneous porous media, J. Contaminant
Hydrol. 11, 291316.
Jackson, R.: 1977, Transport in Porous Catalysts, Chem. Eng. Monograph 4, Elsevier, New York.
Jaynes, D. B. and Rogowski, A. S.: 1983, Applicability of Ficks law to gas diffusion, Soil Sci. Soc.
Am. J. 47, 425430.
Jones, F. O. and Owens, W. W.: 1980, A laboratory study of low-permeability gas sands, J. Petrol.
Technol., 16311640.
Jury, W. A., Spencer, W. F. and Farmer, W. J.: 1983, Behavior assessment model for trace organics
in soil. I. Model description, J. Environ. Qual. 12, 558564.
Jury, W. A., Farmer, W. J. and Spencer, W. F.: 1984a, Behavior assessment model for trace organics
in soil. II. Chemical classification and parameter sensitivity, J. Environ. Qual. 13, 567572.
Jury, W. A., Spencer, W. F. and Farmer, W. J.: 1984b, Behavior assessment model for trace organics
in soil. III. Application of screening model, J. Environ. Qual. 13, 573579.
Jury, W. A., Spencer, W. F. and Farmer, W. J.: 1984c, Behavior assessment model for trace organics
in soil. IV. Review of experimental evidence, J. Environ. Qual. 13, 580586.
Jury, W. A., Russo, D., Streile, G. and Abd, H.: 1990, Evaluation of volatilization by organic
chemicals residing below the surface, Water Resour. Res. 26, 1320.
Klinkenberg, L. J.: 1941, The permeability of porous media to liquids and gases, API Drilling Prod.
Pract., 200213.
Mason, E. A. and Malinauskas, A. P.: 1983, Gas Transport in Porous Media: The Dusty-Gas Model,
Chem. Eng. Monograph 17, Elsevier, New York.
Massman, W. J.: 1998, A review of the molecular diffusivities of H2 O, CO2 , CH4 , CO, O3 , SO2 ,
NH3 , N2 O, NO, and NO2 in Air, O2 and N2 Near STP, Atmos. Environ. 32, 11111127.
Massmann, J. and Farrier, D. F.: 1992, Effects of atmospheric pressures on gas transport in the Vadose
zone, Water Resour. Res. 28, 777791.
McCarthy, K. A. and Johnson, R. L.: 1995, Measurement of trichloroethylene diffusion as a function
of moisture content in sections of gravity-drained soil columns, J. Environ. Qual. 24, 4955.
Millington, R. J. and Quirk, J. M.: 1961, Permeability of porous solids, Trans. Faraday Soc. 57,
12001207.
Prat, M.: 1993, Percolation model of drying under isothermal conditions in porous media, Int. J.
Multiphase Flow 19, 691704.
Pruess, K.: 1987, TOUGH Users Guide, NUREG/CR-4645, SAND86-7104, LBL-20700, US
Nuclear Regulatory Commission.
MODELING TRACE GAS DIFFUSION 341

Pruess, K.: 1991, TOUGH2 A General-Purpose Numerical Simulator for Multiphase Fluid and
Heat Flow, LBL-29400, Lawrence Berkeley Laboratory.
Pruess, K., Oldenburg, C. and Moridis, G.: 1999, TOUGH2 Users Guide, Version 2.0, LBNL-43134,
Lawrence Berkeley National Laboratory.
Reid, R. C., Prausnitz, J. M. and Poling, B. E.: 1987, The Properties of Gases & Liquids, 4th edn,
McGraw-Hill, New York.
Sleep, B. E.: 1998, Modeling transient organic vapor transport in porous media with the dusty gas
model, Adv. Water Resour. 22, 247256.
Thibodeaux, L. J., Springer, C. and Hildebrand, G.: 1986, Transport of chemical vapors through soil:
a landfill cover simulation experiment, in: AIChE Meeting, August 2427, Boston, MA.
Thorstenson, D. C. and Pollock, D. W.: 1989, Gas transport in unsaturated zones: multi-component
systems and the adequacy of Ficks laws, Water Resour. Res. 25, 477507.
Webb, S. W.: 1998, Gas diffusion in porous media evaluation of an advective-dispersive formulation
and the dusty-gas model for binary mixtures, J. Porous Media 1, 187199.
Webb, S. W. and Phelan, J. M.: 1997, Effect of soil layering on NAPL removal behavior in soil-heated
vapor extraction, J. Contaminant Hydrol. 27, 285308.
Webb, S. W., Pruess, K., Phelan, J. M. and Finsterle, S. A.: 1999, Development of a mechanistic
model for the movement of chemical signatures from buried landmines/UXO, in: International
Symposium on Aerospace/Defense Sensing, Simulation, and Controls, SPIE, Orlando, FL, April
1317.

Anda mungkin juga menyukai