You are on page 1of 15

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/245407141

Upscaling transportof adsorbing solutes in porous media

Article  in  Journal of Porous Media · January 2010


DOI: 10.1615/JPorMedia.v13.i5.10

CITATIONS READS
21 168

2 authors:

Amir Raoof S. M. Hassanizadeh


Utrecht University Utrecht University
68 PUBLICATIONS   563 CITATIONS    281 PUBLICATIONS   8,613 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Simulation of fluid flow through thin porous layers and their interfaces View project

Sustainability Strategic Theme View project

All content following this page was uploaded by Amir Raoof on 07 January 2014.

The user has requested enhancement of the downloaded file.


Journal of Porous Media 13(5), 395–408 (2010)

UPSCALING TRANSPORT OF ADSORBING SOLUTES


IN POROUS MEDIA

Amir Raoof ∗ & S. Majid Hassanizadeh

Universiteit Utrecht, Budapestlaan 4, 3584 CD, Utrecht, Netherlands



Address all corespondence to A. Raoof E-mail: raoof@geo.uu.nl

Original Manuscript Submitted: 12/11/2008; Final Draft Received: 4/13/2009

Adsorption of solutes in porous media is commonly modeled as an equilibrium process. Indeed, one may safely assume
that within the pore space, the concentration of adsorbed solute at a point on the grain surface is algebraically related
to the concentration in the fluid next to the grain. The same, however, cannot be said about average concentrations. In
fact, during solute transport, concentration gradients develop within the pore space, and these could potentially give
rise to a scale-dependent adsorption process. The main objective of this research is to develop relationship between pore-
scale adsorption coefficient and corresponding upscaled adsorption parameters. Two approaches are used: Theoretical
averaging and numerical upscaling. In the averaging approach, equilibrium adsorption is assumed at the pore-scale
and solute transport equations are averaged over REV. This leads to explicit expressions for macro-scale adsorption rate
constants as a function of micro-scale parameters. In the numerical approach, first we simulate solute transport within
a single tube undergoing equilibrium adsorption at the pore wall, and then flux averaged concentration breakthrough
curves are obtained. These are used to determine the upscaled adsorption rate constants as functions of pore-scale
hydraulic and adsorption parameters. Results of the two approaches agree very well.

KEY WORDS: pore scale, kinetic adsorption, equilibrium adsorption, upscaling, solute transport

1. INTRODUCTION 1988; Pagitsas et al., 1986; Ryan et al., 1980; Sahimi,


1988; Shapiro and Brenner, 1986, 1987, 1988). Often,
The first continuum scale for the description of flow and discrepancies between observation and theory arise be-
transport processes is the pore scale. Commonly, physic- cause the large-scale description of transport does not ac-
ochemical processes are reasonably well understood and count for some important aspects of small-scale behavior
described by simple relationships at the pore scale. For (Bryant and Thompson, 2001; Raje and Kapoor, 2000;
example, mixing due to diffusion is described by Fick’s Gramling et al., 2002). Commonly, small-scale effects are
law, equilibrium dissolution is explained by Raoult’s law, lumped into empirical terms or coefficients that depend on
and adsorption at the grain surface is assumed to fol- porous media properties and structure (Guo and Thomp-
low an equilibrium relationship. In practice, however, son, 2001). Hence, inclusion of subscale mass transfer
we make measurements, and model transport phenom- effects in the description of porous systems is essential
ena, on scales much larger than the pore scale. Much to develop theoretically sound equations for describing
effort is being made to understand and quantify links be- modeling mass transfer at the larger scales.
tween these scales; in particular, the dependence of mass Because of limitations in measurements at small scale
transfer on hydrodynamic properties of the system (e.g., and in observation frequency, experiments alone cannot
Quintard and Whitaker, 1995; Wood et al., 2000, 2004, provide sufficient qualitative and quantitative understand-
2007; Edwards and Davis, 1995; Edwards et al., 1993; ing of transport; theoretical work is also needed. In
Kechagia et al., 2002; Mauri, 1991; Mojaradi and Sahimi, porous media, upscaling may start either at the molecular

1091-028X/10/$35.00 Copyright °
c 2010 Begell House, Inc. 395
396 Raoof & Hassanizadeh

level (e.g., Murdoch and Hassanizadeh, 2002) or at the coupling can be the reason for much faster laboratory-
pore scale (e.g., Whitaker, 1969, 1986; Hassanizadeh and measured reaction rates of many minerals than those ob-
Gray, 1979). The most common approach is upscaling served in the field (White and Brantley, 2003; Maher et
from the pore scale, where the principal idea is to aver- al., 2004). Then the question arises as to what extent pore-
age the pore-scale transport and reaction processes over a scale reaction models and parameters are applicable at the
representative support volume to produce a macroscopic macro scale.
model of the reactive transport. Recently, Meile and Tuncay (2006) addressed this
Reaction processes such as adsorption, mineral disso- question for the case of mineral dissolution and homoge-
lution, or homogeneous reactions could greatly influence nous reaction with the aid of a pore-scale numerical
transport of dissolved matter in the soil and groundwa- model. They found that macro-scale descriptions of these
ter (Serrano, 2003; Bolt, 1979; Weber et al., 1991; van processes are different from pore-scale descriptions be-
der Zee, 1990; Acharya et al., 2005). These processes cause of the effect of small-scale gradients in concentra-
typically occur at solid-fluid boundaries or fluid-fluid in- tion fields. To investigate these effects, they numerically
terfaces. Commonly, adsorption is modeled as a (linear) generated virtual porous media using random placement
equilibrium process (Weber et al., 1991). The equilib- of identical spherical particles and solved diffusion and
rium assumption means that the chemical potential of the reaction in the resulting pore spaces. They showed that
solute in the fluid next to the solid grain is equal to the upscaled values of reaction and dissolution rates depend
chemical potential of the solute attached to the grain. This on type of reaction, pore geometry, and macroscopic con-
assumption is probably acceptable in most interactions at centration gradient. They found that differences between
the pore scale, that is, at the pore boundaries. The same, two scales become more significant for surface reactions
however, cannot be said at the larger scales, where we as compared to homogeneous reactions. A limitation in
work with average solute concentrations. Owing to vari- the work of Meile and Tuncay (2006) is that they con-
ation of concentration within the pore and mass transfer sidered only diffusion transport and neglected advection.
at the grain surface, the solute concentration close to the Other modeling studies have shown that the role of advec-
wall will be different from the average concentration in tion on distribution of chemicals at the pore level is very
the bulk fluid (Binning and Celia, 2008; Meile and Tun- important (e.g., Bryant and Thompson, 2001; Knutson et
cay, 2006; Li et al., 2007; Kechagia et al., 2002). Consid- al., 2001; Robinson and Viswanathan, 2003; Szecsody et
ering this issue, the adsorbed solute concentration, which al., 1998).
could be in equilibrium with the solution concentration Li et al. (2007) studied the effect of pore-scale concen-
close to the adsorbed site, may not be in equilibrium with tration gradients on a mineral dissolution rate influenced
the average concentration of solution. Instead, one often by advection. They introduced two kinds of models for
has to use a so-called kinetic relationship between average minerals that could dissolve at different rates. First, they
concentration of dissolved and adsorbed solutes. This is developed a Poiseuille flow model that coupled the re-
particularly the case if the porous medium has a micro- action rate to both advection and diffusion within a pore
scopically heterogeneous structure such as in aggregated space. Next they developed a well-mixed reactor model
porous media or fractured media. that assumed complete mixing within the pore. They have
Studies of reactive solute transport commonly involve shown that concentration gradients could cause scale de-
a series of batch experiments, through which reaction pa- pendence of reaction rates. Significant concentration gra-
rameters are obtained. Given the fact that in batch experi- dients would develop when diffusion is slower than the
ments, there are no spatial variations in the concentration advection process, provided that rates of advection and re-
field (as it is a well-mixed system), the reaction models action are comparable. This shows the effect of pore-scale
and reaction parameters actually pertain to the pore scale. gradients and residence times on transport of reactive so-
In an open system, where diffusion and/or hydrodynamic lutes. The effect of residence times on reactive transport
dispersion occur, gradients in concentration exist, as ob- was also addressed by Robinson and Viswanathan (2003),
served both in experiments (e.g., Rashidi et al., 1996; who showed the importance of pore-scale gradients, espe-
Kapoor et al., 1998; Taylor 1953; Aris, 2005) and through cially for nonlinear reactions; solute pulses of short dura-
numerical simulations (e.g., Li et al., 2006; Cao and Ki- tion; and systems with broad residence time distribution
tanidis, 1998; Shapiro and Adler, 1997). As such, the curves. Characteristic timescales of reaction processes
reaction rates are an outcome of coupling between the re- pose constraints for transport models (Mo and Friedly,
action and hydrodynamic processes (Li et al., 2007); this 2000; Cao and Kitanidis, 1998).

Journal of Porous Media


Upscaling Transport of Adsorbing Solutes 397

Experimental studies (e.g., Guo and Thompson, 2001) wish to find relations for upscaled transport coefficients
as well as pore-scale numerical models (Knutson et al., for a solute that, at the pore scale, undergoes diffusion
2001) have shown the dependence of mass transfer coef- and advection as well as first-order equilibrium reaction at
ficients (e.g., in dissolution process) on the hydrodynam- the solid surface. First, the problem of solute transport is
ics of porous media. For example, Knutson et al. (2001) formulated at the pore scale, considering equilibrium ad-
found that the dimensionless mass transfer coefficient of sorption. Then the equations are averaged, and upscaled
dissolution increased with Peclet number, Pe. equations are obtained. As a result, explicit expressions
Another method for the upscale adsorption process in are derived for the mass exchange between fluid and solid
porous media is the homogenization technique. Auriault phases in terms of average concentrations. Next, steady
and Lewandowska (1996) have used the homogenization state flow and transient adsorption in a single tube are
technique (double asymptotic developments) to derive a simulated numerically. The resulting breakthrough curves
macroscopic equation for the average concentration field from this single-tube model are compared to the solution
of a solute that, at micro scale, is undergoing adsorp- of the 1-D continuum scale transport equation to estimate
tion. They found effective parameters (e.g., dispersion- upscaled adsorption parameters.
adsorption tensor) that characterize the medium at the
macro scale. Through various characteristic dimension- 2. THEORETICAL UPSCALING OF
less parameters, they have shown the practical importance ADSORPTION IN POROUS MEDIA
of processes in different flow/transport regimes. The
macroscopic parameters were found to be dependent only 2.1 Formulation of the Pore-Scale Transport
on the microscopic transport parameters and microscopic Problem
geometrical properties. However, they found conditions
Consider the transport of a solute in the pore space of a
under which the macroscopic
¡ ¢ model does not exist, for ex-
granular soil. Processes affecting transport are considered
ample, P e ¿ O ε−1 where ε, the homogenization pa-
to be advection, diffusion, and chemical reaction within
rameter, is the ratio of the micro-scale heterogeneities size
the water phase plus adsorption to the solid phase at the
to the macro-scale domain size. Because under some con-
pore boundaries. The general form of the equation of
ditions the problem could not be homogenized, they de-
mass balance for the solute reads:
scribed nonhomogenizable and homogenizable domains
in terms of characteristic dimensionless parameters. The ∂ci ¡ ¢
existence of nonhomogenizable domains demands cau- + ∇. ci v + ∇.ji = rb i (1)
∂t
tion in applying their method to real situations.
Van Duijn et al. (2008) applied the homogeniza- where ci (ML−3 ) is mass concentration of solute i
tion technique to study adsorptive solute transport in in a pore; v (LT−1 ) is interstitial water velocity; rb i
a capillary slit. They found upscaled equations using (ML−3 T−1 ) is the rate of chemical reactions with other
the asymptotic expansion technique in terms of the ra- solutes, and ji (ML−2 T−1 ) is the diffusive flux of the so-
tio of characteristic transversal and longitudinal lengths lute. The diffusive flux ji is given by Fick’s first law:
under dominant Peclet and Domkohler numbers. They
have distinguished different characteristic timescales such ji = −D0i ∇ci (2)
as longitudinal, transversal, desorption, and adsorption
where D0i (L2 T−1 ) is the molecular diffusion coefficient
timescales. They have derived effective models for vari-
of solute i in water.
ous conditions such as linear adsorption-desorption, non-
In principle, one should solve this equation within the
linear reactions, and equilibrium adsorption. They have
pore space of the soil subject to boundary conditions. At
verified their method through comparison of solutions
a point on the pore boundary, adsorption causes a flux of
with numerical solutions of original problems with and
solute from the fluid to the solid phase; this gives rise to
without adsorption. They did simulations with high Peclet
an increase of the mass density of adsorbed solutes. The
numbers (larger than 104 ). They found excellent agree-
rate of adsorption is equal
¡ to the¢solute mass flux normal
ment between results of their models and original prob-
to the pore boundary, ci v + ji · n. So the following
lems.
condition at the pore boundary holds:
This research is aimed at developing upscaled rela-
tionships for adsorption in terms of pore-scale properties ∂si
through both theoretical and numerical averaging. We = (ci v + ji ) · n|s (3)
∂t

Volume 13, Number 5, 2010


398 Raoof & Hassanizadeh

where si (ML−2 ) is the mass of adsorbed solute per unit Note that this volume-averaged definition of the flow ve-
area of the solid grains, n is a unit vector normal to the locity is admissible only if the fluid mass density vari-
pore wall, and |s denotes evaluation of the preceding ations are small (see Hassanizadeh and Gray, 1979, for
quantity within the pore but at the solid surface. Because discussion). We continue:
si is unknown, an additional equation is needed to have Average Diffusion Flux:
a determinate system. That extra equation comes from Z
the continuity of chemical potential at the grain surface. i 1
j = f ji dV (7)
This condition leads to an equilibrium relationship, and a V
linear approximation yields: Vf

¯ Average Chemical Reaction Rate:


si = kDi
ci ¯s (4)
Z
¯ 1
where ci ¯s is the solute concentration of fluid at the pore ri = f ri dV (8)
i
V
wall and kD (L) is an equilibrium, pore-scale distribution Vf
coefficient.
The set of Eqs. (1)–(4), together with conditions at the Average Adsorbed Mass Fraction:
outside boundaries of the porous medium and an appro- Z
i 1
priate set of initial conditions, completely specify the so- s = s s si dA (9)
lute transport problem at the pore scale. ρ V
fs A

2.2 Averaging of Pore-Scale Equations where ρs (ML−3 ) denotes the solid mass density.
Note that the average mass density of the sorbed solute
Now we would like to upscale these equations to the is now defined in the form of mass fraction (mass of so-
macro scale. To do so, we need to define an averaging vol- lute per unit mass of grains) si (MM−1 ), as is common in
ume, commonly denoted as the representative elementary solute transport.
volume (REV) (Bachmat and Bear, 1987). This is a well- Averaging of Eq. (1) over V f yields the following
known concept and has been extensively discussed in the macro-scale equation (see, e.g., Whitaker, 1969, 1986;
porous medium literature (see, e.g., Bear, 1988; Bachmat Hassanizadeh, 1986; Gray and Hassanizadeh, 1998):
and Bear, 1986; Hassanizadeh and Gray, 1979). Let us
denote the space occupied by REV with V . Volume V ∂nci ¡ ¢ ¡ ¢
bi
+ ∇ · nci v + ∇ · nJi = nri − U (10)
is in turn composed of two subvolumes: V f occupied by ∂t
the fluid phase and V s occupied by the solid grains. The
boundary of solid grains is denoted with Af s . The su- where v is the average flow velocity and ri is the reaction
perscripts f and s are used to denote the fluid and solid rate; the macro-scale adsorption term Ub i (ML−3 T−1 ) is
phases, respectively. Note that the averaging volume V is defined by:
taken to be invariant in time and space, whereas the sub- Z
volumes V f and V s may vary both in time and space. 1 ¡ i ¢
b i
U = c (v − w) + ji · ndA (11)
We shall integrate the fluid Eq. (1) over V f , and the V
interface condition (3) will be integrated over Af s . First, Af s

we need to define average properties:


where w is the velocity of the solid-fluid interface. Note
Average Mass Concentration: that Ji denotes the macro-scale hydrodynamic dispersion
Z vector and accounts for diffusion as well as the mixing of
i 1
c = f ci dV (5) solutes as a result of pore-scale velocity variations:
V
Vf
i
Ji = ncie
v + nj (12)
Average Flow Velocity:
Z where ev denotes the pore-scale velocity deviations, de-
f 1 fined as:
v = f vdV (6)
V v = v − vf
e (13)
V f

Journal of Porous Media


Upscaling Transport of Adsorbing Solutes 399

Next, the microscopic boundary condition (3) is av- of definitions (4) and (9) leads to the following macro-
eraged overAf s . The time average theorem and defini- scale relationship:
tion (9) are needed. The result is the following macro- i i
scale differential equation for the averaged adsorbed so- b i = − D0 (1 − n) ρs s̄i + SD0 f (c̄i )
U (19)
i
kD d d
lute concentration:
∂ (1 − n) ρs s̄i bi where S (L−1 ) denotes the solid grain specific area, S =
=U (14)
∂t Af s /V . Substitution of this result into the adsorbed mass
Equations (10) and (14) provide two equations to be balance (14) yields the standard linear kinetic adsorption
solved for c̄i and s̄i , provided that an appropriate rela- equation:
tionship is found for the adsorption rate U b i . This term
∂ (1−n) ρs s̄i b i
represents the exchange of mass between solid grains and = U = nkatt f (c̄i )−(1−n) ρs kdet si (20)
∂t
the fluid phase. Such a mass exchange could be due to
thawing, freezing, dissolution, precipitation, or adsorp- where the kinetic rate coefficients, katt and kdet (T−1 ),
tion. Here we consider the case of adsorption only and are defined by:
assume that adsorbed mass has no effect on the fluid or SD0i
katt = (21a)
solid mass density. Phase change, dissolution, and pre- nd
cipitation are neglected. Di
Because there is no phase change, the normal water kdet = i 0 (21b)
kD d
flux at the grain boundary, (v − w) · n in Eq. (11), will be
identically zero. Then, substitution of Fick’s law (Eq. (2)) The mass balance equation for the solutes now becomes:
in (2) Eq. (11) yields: ∂nc̄i ¡ ¢ ¡ ¢
Z i
+ ∇ · nc̄i v̄ i + ∇. nJ i = nr̄i
bi = − 1 ∂c ∂t (22)
U D0i dA (15)
V ∂n − nkatt f (c̄i ) + (1 − n) ρs kdet s̄i
Af s
which, together with Eq. (20), forms a set of two equa-
where ∂(.)/∂n denotes the derivative in the direction nor- tions to be solved for c̄i and s̄i .
mal to the grain surface. We now assume that ∂ci /∂n can
be approximated as follows:
¯ 2.3 Kinetic versus Equilibrium Effects
∂ci ¯¯ ³ ¯ ¯ ´
i¯ i¯
= c s − c pore /d (16) It is shown here that even if the adsorption process can be
∂n ¯Af s described by a linear isotherm at the pore scale, in gen-
¯

where c s denotes the¯ solute concentration in the fluid eral, it has a kinetic nature at the macro scale. The ques-
tion is when we can assume an equilibrium isotherm at
at the pore wall and ci ¯pore denotes the solute concentra-
the macro scale. That would be the case, of course, if the
tion within the pore at ¯some distance d. We now employ
kinetic process is very fast. This can be studied best if
Eq. (4) to eliminate ci ¯s from (16) and substitute the re-
we write the kinetic Eq. (20) in an alternative form that is
sult back into (15):
commonly employed:
Z i Z
D i
D ¯ ¡ i ¢
bi = −
U 0
si dA+ 0 ci ¯pore dA (17) Ub i = κ (1 − n) ρs KD f ( c̄i ) − s̄i (23)
i
V kD d Vd
Af s Af s i
where κ and KD are the macro-scale kinetic rate coef-
¯There, is of course, no information on the value
¯ of ficient and distribution coefficient, respectively, and are
ci ¯pore . However, it is plausible to assume that ci ¯pore is defined as:
concentration, c̄i . This means Di
a function of average fluid ¯ κ= i0 (24a)
that we may replace ci ¯pore by f (c̄i ) in Eq. (17) and set: kD d
¯ i
i SkD
ci ¯pore = f (c̄i ) (18) KD = (24b)
(1 − n) ρs
Use of this approximation (Eq. (18)), recognizing the fact From (24a), it is evident that the kinetic rate coefficient is
that ci is a constant within the averaging volume, and use the ratio of the micro-scale diffusion mass flux D0i ∆ci /d

Volume 13, Number 5, 2010


400 Raoof & Hassanizadeh

to the amount of mass that is to be adsorbed, that is, and (2) an equivalent upscaled 1-D model for the cross-
i
kD ∆ci . Obviously, the faster the diffusion process, the sectionally averaged concentration. These models allow
larger the kinetic coefficient will be (and thus the smaller us to investigate some of the assumptions made in our
the kinetic effect). On the other hand, the larger the upscaling approach and also to verify results of that ap-
amount of mass to be adsorbed, the more important the proach.
kinetic effect will be.
It is interesting to note that Eqs. (24) allow us to write
3.1 Flow and Transport at Pore Scale (Single-
the macro-scale kinetic rate coefficient as a function of the
Tube Model)
macro-scale distribution coefficient; this can be achieved
i
by eliminating kD between the two equations in (24a). Consider a long single tube with a constant circular cross
We obtain: section with radius R0 . We assume fully developed steady
SD0i state laminar flow in the tube (Poiseuille flow) so that the
κ= s i
(25)
ρ (1 − n)d KD velocity distribution is given by (Daugherty and Franzini,
The inverse of the kinetic coefficient, κ , is the char- 1965):
−1
" µ ¶2 #
acteristic timescale of the kinetic process. Now, if this r
v(r) = 2v̄ 1 − (27)
characteristic time is of the same order of magnitude as, R0
or bigger than, the residence time of solutes in pores, d/v
(i.e., for κ−1 ≥ d/v, then the kinetic process will be im- where r is the radial coordinate, v(r) is the local fluid ve-
portant. This condition can be represented by the dimen- locity, v̄ is the average flow velocity, and R0 is the radius
sionless number σD : of the cylinder. In the single-tube model, adsorption oc-
curs only at the wall of the tube (Fig. 1). In this study,
ρs (1 − n)vKD i
changes in the radius of the tube due to the adsorption
σD = i
(26)
SD0 process are neglected.
The mass transport in the tube is given by the following
Thus, if σD ≥ 1, then kinetic effects are important. equation:
Note that if the flow velocity is very small, then the kinetic
µ ³ r ´2 ¶ ∂c · 2
effects become negligible. In the limiting case of no flow, ∂c ∂ c
as is the case in batch experiments, the equilibrium rela- + 2v̄ 1 − = D0
∂t R ∂z ∂z 2
tionship (24) applies with no approximation. Thus batch µ ¶¸ (28)
1 ∂ ∂c
experiments can be used to obtain the macro-scale distri- + r
r ∂r ∂r
bution coefficient.
In the following section, we will perform numeri-
where D0 (L2 T−1 ) is the molecular diffusion coefficient
cal experiments to explore the assumptions leading to
and z is the axial direction along the tube.
Eqs. (24a) and also find an approximate value for d in
In the tube, the wall acts as an adsorbent, and therefore
Eq. (16).
the adsorption relation appears in the boundary condition
of the differential Eq. (28). Adsorption to the wall may
be described by the following equation, which prescribes
3. NUMERICAL UPSCALING OF ADSORBING that the diffusive mass flux to the wall is the only source
SOLUTE TRANSPORT of accumulation at the wall:
¯
Perhaps the simplest step in upscaling is to replace the ∂s ∂c ¯¯
= −D0 (29)
3-D flow and concentration fields within the pore (or a ∂t ∂r ¯s
tube) by 1-D fields, whereby velocity and concentration
are averaged over the pore cross section. As mentioned in where s (ML−2 ) is adsorbed mass per unit area. This is
Section 1, this upscaling has been considered for homo- the same as Eq. (3), where the first Fick’s law is assumed.
geneous reactions as well as dissolution. Here we treat Similar to Section 2, we assume that linear equilibrium
the upscaling of adsorptive solute transport. adsorption holds at the wall, characterized by the distri-
To analyze the scale dependence of the adsorption bution coefficient value kD (L):
process, we have developed two models: (1) a single-
tube model to simulate details of transport within a pore s = kD c|s (30)

Journal of Porous Media


Upscaling Transport of Adsorbing Solutes 401

FIG. 1: Conceptual representation for the single-tube model. The velocity profile is assumed to be parabolic

Next we introduce the following dimensionless variables The standard package Flex-PDE (Flex-PDE, 2005) has
and parameters: been used to numerically solve this set of equations. We
have simulated solute transport for a range of values of
∗ c ∗ r ∗ z ∗ v̄t
c = , r = , z = , t = parameters P ep and κ. The solution of Eqs. (32)–(34)
c0 R0 R0 R0 ∗ ∗ ∗ ∗
(31) results in concentration filed c (z , r , t ) for different
v̄R0 s k D values of Peclet number, P ep , and dimensionless dis-
P ep = , s∗ = , κ=
D0 c0 R0 R0 tribution coefficient, κ. This concentration field and its
where c0 is the concentration at the inlet boundary. cross-sectional average, c̄∗ (z ∗ , t∗ ), may be considered to
So we will end up with pore-scale dimensionless mass be equivalent to “observation data.” We then use these
transport equation: observed data to obtain the relationship for upscaled ad-
sorption parameters.
∗ ∗
· 2 ∗
∂c ∂c 1 ∂ c
+ 2(1 − r∗2 ) ∗ =
∂t∗ ∂z P ep ∂z ∗2
µ ∗
¶¸ (32) 3.2 Flow and Transport at 1-D Tube Scale
1 ∂ ∗ ∂c
+ ∗ ∗ r One can obtain the first level of upscaling by averaging
r ∂r ∂r∗
the pore-scale concentration over the cross section of the
where P ep is the pore-scale Peclet number, which ex- tube. This results in the 1-D average concentration field
presses the ratio between the magnitude of the advective c̄∗ (z ∗ , t∗ ), to which we refer as the 1-D tube scale. Now,
and diffusive transport terms. The dimensionless bound- the question arises what the governing equation should be
ary conditions for our model are as follows: for this average concentration. To answer this question,
we examine a typical breakthrough curve of average con-
at z ∗ = 0, c∗ = 1.0 (33a) centration for an adsorbing solute at a position far farm
∂c∗ the tube inlet (e.g., at z ∗ = 100), as shown in Fig. 2.
at z ∗ = ∞, = 0 (33b) It is evident that the breakthrough curve shows a non-
∂z ∗
ideal behavior with a nonsymmetric derivative. The tail-
∂c∗
at r∗ = 0, = 0 (33c) ing observed here suggests that at this scale, adsorption
∂r∗
should be described as a kinetic process. The governing
∗ ∂c∗ ∂s∗ equations for 1-D transport of kinetically adsorbing solute
at r = 1, = −P ep ∗ (33d)
∂r∗ ∂t are as follows:
Substitution of Eq. (30) in the boundary condition (33d) ∂c∗ ∂c∗ 1 ∂ 2 c∗ ∂s∗
results in: ∗
+ ∗
= ∗2
− ∗ (35)
∂t ∂z P e ∂z ∂t
∂c∗ ∂c∗
= −P e p κ at r∗ = 1 (34)
∂r∗ ∂t∗ ∂s∗ ∗ ∗

= katt c∗ − kdet s∗ (36)
It is evident that P e and κ are the two parameters in ∂t
p
this set of equations that control the transport and reaction where P e = v̄R0 /DL , which involves the average flow
processes within the tube. velocity v̄ and the longitudinal dispersion coefficient DL .

Volume 13, Number 5, 2010


402 Raoof & Hassanizadeh

FIG. 2: Breakthrough curve of average concentration (solid line) resulting in non-Gaussian derivative (dashed line);
based on simulations from single-tube model with continuous input

Note that P e is the tube-scale Peclet number; it is differ- studied upscaling of mineral dissolution within a single
ent from the pore-scale P ep , which is based on the diffu- pore. Reactive flow experiments were performed in a
sion coefficient (see Eq. (31)). cylindrical pore, 500 µm in diameter and 4000 µm long,
∗ ∗
Furthermore, katt and kdet are dimensionless adsorp- drilled in a single crystal of calcite. They employed the
tion and desorption rate coefficients, respectively. These single-tube model to simulate the concentration of Ca2+
are related to dimensional coefficients katt and kdet resulting from dissolution of calcite. The kinetic calcite
through the following relationships: dissolution formula (from Chou et al., 1989) was assumed
to hold at the wall of tube. Then, the flux-average concen-
katt R0 kdet R0

katt = ∗
, kdet = (37) tration of Ca2+ , calculated from the single-tube model,
v̄ v̄ was compared to the measured Ca2+ concentration from

These equations contain three parameters: P e, katt , the experiment for different pH values and flow condi-

and kdet . We have evaluated these parameters by fit- tions. They found an excellent agreement between mod-
ting the solution of this set of equations to the average eling results and results of experiment. This, we believe,
breakthrough concentration, c̄∗ (z ∗ , t∗ ), obtained from the is an indication of the applicability of our procedure in
single-tube model for various values of pore-scale Peclet upscaling from pore to tube scale.
number (P ep ) and κ. This procedure results in relation-

ships between the three upscaled parameters (P e, katt , 3.2.1 Upscaled Peclet Number (Pe)

and kdet ) and their corresponding pore-scale parameters
(P ep and κ). We have used the cross-sectional averaged Here we assumed that the dispersion is not affected by
concentrations from the single-tube model to find corre- the adsorption process. Therefore we evaluate the up-
sponding upscaled adsorption parameters. A similar ap- scaled Peclet number (Pe) for the case of a nonadsorb-
proach was used by Li et al. (2008) for upscaling of disso- ing solute (i.e., κ = 0). This is done by fitting the break-
lution processes under steady state flow conditions. They through curve from the single-tube model to the solution

Journal of Porous Media


Upscaling Transport of Adsorbing Solutes 403

of the 1-D transport equation (Eq. (35)) with no adsorp- have been checking the covariance matrix to make sure
∗ ∗
tion (i.e., katt = kdet = 0). Figure 3 shows the result- that parameters are not correlated.
ing graph, where P e is plotted as a function of pore-scale By repeating the procedure for a range of pore-scale
Peclet number (P ep ). In the same figure, the Taylor dis- parameters (1 < P ep < 300, and 0.1 < κ < 20) and
∗ ∗
persion formula (Taylor, 1953; Aris, 2005) for upscaled founding the corresponding upscaled katt and kdet we
Peclet number in a tube is also plotted (Eq. (38)). The ex- could find relations between these sets of parameters. Fig-
∗ ∗
cellent agreement indicates the accuracy of our numerical ures 5 and 6 show the plots of katt and kdet as a function
code in capturing the transport within the tube and also of P ep and κ.
shows that the Taylor assumption is valid for this prob- From relations shown in Figs. 5 and 6, we found the
∗ ∗
lem. We use Eq. (38) to obtain upscaled Peclet number best-fit formulas for katt and kdet :
(P e) in simulation of adsorbing solute in the next section:
∗ 4.0(1 − e−3κ )
katt = (40a)
1 1 P ep P e0.95
p
= + (38)
Pe P ep 48 9.0

kdet = (40b)
∗ ∗
(0.5 + 4.5κ)P e0.95
p
3.2.2 Upscaled Adsorption Parameters (katt and kdet )
Next, we define the upscaled distribution coefficient (KD )
Since we employed the Taylor formula to estimate the up- by:
scaled Peclet number, there are only two upscaled param- ∗ ¡ ¢
∗ ∗ katt
eters, katt and kdet , left to be determined as a function of KD = ∗
= 0.4 1 − e−3κ (0.5 + 4.5κ) (41)
pore-scale parameters: kdet


According to Eq. (41), upscaled distribution coefficient
katt = f (P ep , κ) KD is independent of Peclet number. It is practically a
(39)

kdet = f (P ep , κ) linear function of pore-scale distribution coefficient κ, as
shown in Fig. 7.
The CXTFIT curve-fitting program (Toride et al., 1995) The linearity of upscaled distribution coefficient KD
was used for the purpose of solving Eqs. (35) and (36) as a function of pore-scale distribution coefficient κ also
and fitting the breakthrough curves. Figure 4 shows an shows the validity of our upscaling process. Both KD and
example of fitting the breakthrough curves and the cor- κ are a measure of the capacity of the porous medium to
responding parameters. Through the fitting process, we absorb the mass, and thus they should be linearly related.

FIG. 3: Graph of P e versus P ep . The points are the result of fitting of the breakthrough curve of average concentra-
tion (with κ = 0) to Eq. (35) to find P e. The solid line shows the function by the Taylor formula of Eq. (38)

Volume 13, Number 5, 2010


404 Raoof & Hassanizadeh

FIG. 4: The resulting breakthrough curves for pore-scale (circles) and 1-D upscale (solid line) models. In this
illustration, the pore-scale parameters are P ep = 35 and κ = 5.0, and the corresponding upscaled parameters by
∗ ∗
fitting are found to be katt = 0.14 and kdet = 0.05. Using Eq. (38), P e will be 1.3

(a) (b)

FIG. 5: The relation between macro-scale kdet as a function of pore-scale κ and P ep

(a) (b)

FIG. 6: The relation between macro-scale katt as a function of pore-scale κ and P ep

Journal of Porous Media


Upscaling Transport of Adsorbing Solutes 405

∗ ∗
FIG. 7: The relation between upscaled distribution coefficient (KD = katt /kdet ) as a function of pore-scale dimen-
sionless distribution coefficient κ

4. DISCUSSION OF RESULTS D0i


kdet = i
(45b)
kD wR0
Equations (40) can be converted to dimensional forms us-
ing Eqs. (31) and (37); this gives: These equations agree quite well with Eqs. (42a), which
³ were obtained through numerical averaging. Indeed, if
kD ´
4.0 1 − e−3 R0 v̄ D00.95
0.05
we acknowledge the fact that for a tube, porosity is unity
katt = (42a) and specific surface S = 2/R0 , Eqs. (45) reduce to the
R01.95 following forms:
avg 2D0
9.0 v̄
0.05
D00.95 katt = (46a)
kdet = ³ ´ (42b) wR02
0.5 + 4.5 kRD0 R01.95 avg D0
kdet = i wR
(46b)
We observe that katt and kdet are only weak functions kD 0

of velocity and that they strongly depend on the geome- where superscript avg stands for “averaging method.” Re-
try of the pore space and the diffusion coefficient as well call that wR0 denotes the radial position
¯ of a point in the
as the pore-scale distribution coefficient. We could com- pore where point concentration ci ¯pore is equal to average
pare Eqs. (42a) with corresponding equations derived us- concentration c̄i (see Eq. (43)).
ing the averaging method in Section 2.2. For¯the case of a Now, in Eqs. (42a), if we neglect the dependence
single tube, it is plausible to assume that ci ¯pore at a po- on velocity and the exponential term and let (0.5 +
sition somewhere between the pore center and pore wall 4.5kD /R0 ) equate to 4.5kD /R0 , we obtain the following
(denoted by wR0 , where 1.0 > w > 0) will be equal approximations:
to the average fluid concentration, c̄i . This means that in 4Di
Eq. (19), we may set: katt = 20 (47a)
R0
f (c̄) = c̄ and d = wR0 (43) 2D0i
kdet = (47b)
kD R0
As a result, Eq. (19) leads to the following macro-scale We notice that the two sets are the same if w = ¯ 0.5,
kinetic adsorption relationship: which means that we should expect concentration ci ¯pore
b i = nkatt c̄i − (1 − n) ρs kdet s̄i
U (44) appearing in Eq. (16) to be equal to the average concen-
tration c̄i at r = 0.5R0 . This possibility has been verified
where the kinetic rate coefficients katt and kdet (T−1 ) are by plotting concentration breakthrough curves at r = 0,
defined by: r = 0.5R0 , and r = R0 as well as average concentration
¯
SD0i c̄i in Fig. 8; the agreement between c̄i and ci ¯r=0.5R is ex-
katt = (45a) cellent. This result suggests that Eqs. (46) provide a very
nwR0

Volume 13, Number 5, 2010


406 Raoof & Hassanizadeh

FIG. 8: Comparison between average concentration breakthrough curve and breakthroughs of concentrations at three
different positions in the tube

good approximation for upscaled kinetic adsorption coef- Aris, R., On the dispersion of a solute in a fluid flowing
ficients. The parameters w must be seen as an empirical through a tube, Proc. R. Soc. London, Ser. A, vol. 235,
pp. 67–77, 2005.
factor that will be different for different pore structures.
Auriault, J. L. and Lewandowska, J., Diffusion/adsorption/ ad-
vection macrotransport in soils, Eur. J. Mech. A, vol. 15,
5. CONCLUSION pp. 681–704, 1996.
Bachmat, Y. and Bear, J., Macroscopic modelling of transport
In this work, we have shown that even if there is equi- phenomena in porous media. 1: The continuum approach,
librium adsorption at the pore wall (or at the grain sur- Transp. Porous Media, vol. 1, pp. 213–240, 1986.
face), one may need to employ a kinetic description at the Bachmat, Y. and Bear, J., On the concept and size of a repre-
sentative elementary volume (REV), in Advances in Trans-
larger scale. This result was obtained through theoreti- port Phenomena in Porous Media, pp. 3–19, Martinus Ni-
cal averaging from pore to REV scale as well as through jhof, Dordrecht, Netherlands, 1987.
numerical averaging from pore to the tube scale. Both Bear, J., Dynamics of Fluids in Porous Media, Courier Dover,
approaches result in formulas for macro-scale kinetic ad- New York, 1988.
sorption parameters as a function of pore-scale parame- Binning, P. J. and Celia, M. A., Pseudokinetics arising from
the upscaling of geochemical equilibrium, Water Resour.
ters such as Peclet number and distribution coefficient. Res., vol. 44(7), pp. 145–155, 2008.
Formulas from two approaches agree very well. The up- Bolt, G., Movement of solutes in soil: Principles of ad-
scaled adsorption parameters are found to be only weak sorption/exchange chromatography, Soil Chem. B, vol. 5,
functions of velocity; they strongly depend on geometry pp. 285–348, 1979.
of the pore and diffusion coefficient in the solution as well Bryant, S. L. and Thompson, K. E., Theory, modeling and ex-
periment in reactive transport in porous media, Curr. Opin.
as the pore-scale distribution coefficient. Colloid Interface Sci., vol. 6, pp. 217–222, 2001.
The relations for upscaled transport coefficients are ap- Cao, J. and Kitanidis, P., Pore-scale dilution of conserva-
propriate for using in pore-network models where tube- tive solutes: An example, Water Resour. Res., vol. 34,
scale relationships are needed to model adsorbing so- pp. 1941–1949, 1998.
lute transport in individual pore throats. Using the pore- Chou, L., Garrels, R., and Wollast, R., Comparative study of
the kinetics and mechanisms of dissolution of carbonate
network model, we could scale up from the pore scale to minerals, Chem. Geol., vol. 78, pp. 269–282, 1989.
the scale of REV or even the core scale. Daugherty, H. and Franzini, J., Steady flow of incompressible
fluids in pipes, in Fluid Mechanics with Engineering Ap-
plications, 564, McGraw-Hill, New York, 1965.
REFERENCES
Edward, D. A. and Davis, A., Diffusion and convective disper-
Acharya, R., Zee, S., and Leijnse, A., Transport modeling of sion through arrays of spheres with surface adsorption, dif-
nonlinearly adsorbing solutes in physically heterogeneous fusion, and unequal solute partitioning, Chem. Eng. Sci.,
pore networks, Water Resour. Res., vol. 41, W02020, 2005. vol. 50, pp. 1441–1454, 1995.

Journal of Porous Media


Upscaling Transport of Adsorbing Solutes 407

Edwards, D. A., Shapiro, M., and Brenner, H., Dispersion and pp. 1091–1123, 2002.
reaction in two-dimensional model porous media, Phys. Pagitsas, M., Nadim, A., and Brenner, H., Macrotransport pro-
Fluids A, vol. 5, pp. 837–848, 1993. cesses in the presence of bulk and surface chemical reac-
Flex-PDE, A flexible solution system for partial differential tions, J. Chem. Phys., vol. 85, pp. 4038–4044, 1986.
equations, 1996–2005, http://www.flexpde.com, PDE So- Quintard, M. and Whitaker, S., Aerosol filtration: An analy-
lutions Inc., Sunol, CA, 2005. sis using the method of volume averaging, J. Aerosol Sci.,
Gramling, C. M., Harvey, C. F., and Meigs, L. C., Reactive vol. 26, pp. 1227–1255, 1995.
transport in porous media: A comparison of model predic- Raje, D. S. and Kapoor, V., Experimental study of bimolecu-
tion with laboratory visualization, Environ. Sci. Technol., lar reaction kinetics in porous media, Environ. Sci.;echnol.,
vol. 36, pp. 2508–2514, 2002. vol. 34, pp. 1234–1239, 2000.
Gray, W. G. and Hassanizadeh, S. M., Macroscale continuum Rashidi, M., Peurrung, L., Tompson, A., and Kulp, T., Ex-
mechanics for multiphase porous-media flow including perimental analysis of pore-scale flow and transport in
phases, interfaces, common lines and common points, Adv. porous media, Adv. Water Resour., vol. 19, pp. 163–180,
Water Resour., vol. 21, pp. 261–281, 1998. 1996.
Guo, G. and Thompson, K. E., Experimental analysis of lo- Robinson, B. A. and Viswanathan, H. S., Application of the
cal mass transfer in packed beds, Chem. Eng. Sci., vol. 56, theory of micromixing to groundwater reactive transport
pp. 121–132, 2001. models, Water Resour. Res., vol. 39(11), p. 1313, 2003.
Hassanizadeh, S., Derivation of basic equations of mass trans- Ryan, D., Carbonell, R., and Whitaker, S., Effective diffusiv-
port in porous media: 1. Macroscopic balance laws, Adv. ities for catalyst pellets under reactive conditions, Chem.
Water Resour., vol. 9, pp. 196–206, 1986. Eng. Sci., vol. 35, pp. 10–16, 1980.
Hassanizadeh, S. and Gray, W., General conservation equa- Sahimi, M., Diffusion-controlled reactions in disordered
tions for multi-phase systems: 1. Averaging procedure, porous media–I. Uniform distribution of reactants, Chem.
Adv. Water Resour., vol. 2, pp. 131–144, 1979. Eng. Sci., Elsevier, vol. 43(11), pp. 2981–2993, 1988.
Kapoor, V., Jafvert, C. T., and Lyn, D. A., Experimental study Sahimi, M. and Jue, V. L., Diffusion of large molecules in
of a bimolecular reaction in Poiseuille flow, Water Resour. porous media, Phys. Rev. Lett., vol. 62, pp. 629–632,
Res., vol. 34, pp. 1997–2004, 1998. 1989.
Kechagia, P. E., Tsimpanogiannis, I. N., Yortsos, Y. C., and Serrano, S. E., Propagation of nonlinear reactive contaminants
Lichtner, P. C., On the upscaling of reaction-transport pro- in porous media, Water Resour. Res., vol. 39(8), p. 1228,
cesses in porous media with fast or finite kinetics, Chem. 2003.
Eng. Sci., vol. 57, pp. 2565–2577, 2002. Shapiro, M. and Adler, P. M., Coupled transport of multi-
Knutson, C. E., Werth, C. J., and Valocchi, A. J., Pore-scale component solutes in porous media, J. Eng. Math., vol. 31,
modeling of dissolution from variably distributed non- pp. 357–378, 1997.
aqueous phase liquid blobs, Water Resour. Res., vol. 37, Shapiro, M. and Brenner, H., Taylor dispersion of chemically
pp. 2951–2963, 2001. reactive species: Irreversible first-order reactions in bulk
Li, L., Peters, C. A., and Celia, M. A., Upscaling geochemi- and on boundaries, Chem. Eng. Sci., vol. 41, pp. 1417–
cal reaction rates using pore-scale network modeling, Adv. 1433, 1986.
Water Res., vol. 29, pp. 1351–1370, 2006. Shapiro, M. and Brenner, H., Chemically reactive generalized
Li, L., Steefel, C. I., and Yang, L., Scale dependence of min- Taylor dispersion phenomena, AIChE J., vol. 33, pp. 1155–
eral dissolution rates within single pores and fractures, 1167, 1987.
Geochimica et Cosmochimimica Acta, Elsevier, vol. 72(2), Shapiro, M. and Brenner, H., Dispersion of a chemically re-
pp. 360–377, 2008. active solute in a spatially periodic model of a porous
Maher, K., DePaolo, D. J., and Lin, J. C. F., Rates of silicate medium, Chem. Eng. Sci., vol. 43, pp. 551–571, 1988.
dissolution in deep-sea sediment: In situ measurement us- Szecsody, J. E., Zachara, J. M., Chilakapati, A., Jardine, P. M.,
ing 234U/238U of pore fluids, Geochim. Cosmochim. Acta, and Ferrency, A. S., Importance of flow and particle-scale
vol. 68, pp. 4629–4648, 2004. heterogeneity on CoII/IIIEDTA reactive transport, J. Hy-
Mauri, R., Dispersion, convection, and reaction in porous me- drol., vol. 209, pp. 112–136, 1998.
dia, Phys. Fluids A, vol. 3, pp. 743–756, 1991. Taylor, G., Dispersion of soluble matter in solvent flowing
Meile, C. and Tuncay, K., Scale dependence of reaction rates slowly through a tube, Proc. R. Soc. London, Ser. A,
in porous media, Adv. Water Resour., vol. 29, pp. 62–71, vol. 219, pp. 186–203, 1953.
2006. Toride, N., Leij, F., and van Genuchten, M. T., The CXTFIT
Mo, Z. and Friedly, J., Local reaction and diffusion in porous code for estimating transport parameters from laboratory
media transport models, Water Resour. Res., vol. 36, or field tracer experiments, version 2.0, Res. Rep 137, 121,
pp. 431–438, 2000. Riverside, California, 1995.
Mojaradi, R. and Sahimi, M., Diffusion-controlled reactions in van der Zee, S. E., Analytical traveling wave solutions for
disordered porous media: 2. Nonuniform distribution of re- transport with nonlinear and nonequilibrium adsorption,
actants, Chem. Eng. Sci., vol. 43, pp. 2995–3004, 1988. Water Resour. Res., vol. 26, pp. 2563–2578, 1990.
Murdoch, A. and Hassanizadeh, S., Macroscale balance rela- van Duijn, C. J., Mikelic, A., Pop, I. S., and Rosier, C., Effec-
tions for bulk, interfacial and common line systems in tive dispersion equations for reactive flows with dominant
multiphase flows through porous media on the basis of Peclet and Damkohler numbers, Adv. Chem. Eng., vol. 34,
molecular considerations, Int. J. Multiphase Flow, vol. 28, pp. 1–45, 2008.

Volume 13, Number 5, 2010


408 Raoof & Hassanizadeh
View publication stats

Weber, Jr., W., McGinley, P., and Katz, L., Sorption phenom- pp. 479–506, 2003.
ena in subsurface systems: Concepts, models and effects Wood, B. D., Quintard, M., and Whitaker, S., Jump conditions
on contaminant fate and transport, Water Res., vol. 25, at non-uniform boundaries: The catalytic surface, Chem.
pp. 499–528, 1991. Eng. Sci., vol. 55, pp. 5231–5245, 2000.
Whitaker, S., Advances in theory of fluid motion in porous me- Wood, B. D., Quintard, M., and Whitaker, S., Estimation of
dia, Ind. Eng. Chem., vol. 61, pp. 14–28, 1969. adsorption rate coefficients based on the Smoluchowski
Whitaker, S., Flow in porous media: I. A theoretical derivation equation, Chem. Eng. Sci., Oxford; New York: Pergamon
of Darcy’s law, Transp. Porous Media, vol. 1, pp. 3–25, Press, vol. 59(10), pp. 1905–1921, 2004.
1986. Wood, B. D., Radakovich, K., and Golfier, F., Effective reac-
White, A. F. and Brantley, S. L., The effect of time on the tion at a fluid-solid interface: Applications to biotrans-
weathering of silicate minerals: Why do weathering rates formation in porous media, Adv. Water Resour., vol. 30,
differ in the laboratory and field? Chem. Geol., vol. 202, pp. 1630–1647, 2007.

Journal of Porous Media

You might also like