You are on page 1of 19

WATER RESOURCES RESEARCH, VOL. 39, NO. 3, 1051, doi:10.

1029/2001WR001146, 2003

When good statistical models of aquifer heterogeneity go bad:


A comparison of flow, dispersion, and mass transfer in connected
and multivariate Gaussian hydraulic conductivity fields
Brendan Zinn and Charles F. Harvey
Ralph M. Parsons Laboratory, Department of Civil and Environmental Engineering, Massachusetts Institute of Technology,
Cambridge, Massachusetts, USA

Received 20 December 2001; revised 6 June 2002; accepted 6 June 2002; published 12 March 2003.

[1] We describe the upscaled groundwater flow and solute transport characteristics of
two-dimensional hydraulic conductivity fields with three fundamentally different spatial
textures and consider the conditions under which physical mobile–immobile domain mass
transfer occurs in these fields. All three fields have near-identical lognormal univariate
conductivity distributions, as well as near-identical isotropic spatial covariance functions.
They differ in the pattern by which high- or low-conductivity regions are connected: the
first field has connected high-conductivity structures; the second is multivariate log-
Gaussian and, hence, has connected structures of intermediate value; and the third has
connected regions of low conductivity. We find substantially different flow and transport
behaviors in the three different fields. Flow and transport in the multivariate log-Gaussian
field are consistent with stochastic theory. The field with connected high-conductivity
paths has an effective conductivity greater than the geometric mean and large variations in
fluid velocity. It produces significant mass transfer behavior (i.e., tailing) when the
conductivity variance is large and, depending on the system parameters, this mass transfer
is driven by either diffusion or advection. In the field with connected low-conductivity
regions, the effective conductivity is below the geometric mean and transport is well
characterized by the advection–dispersion model with a dispersivity smaller than that in
the multivariate log-Gaussian field. Thus, physical mobile–immobile domain mass
transfer may occur in smooth hydraulic conductivity fields with univariate log-Gaussian
density functions if the variability in conductivity is sufficient and the high values are
more connected than modeled by the multivariate log-Gaussian distribution. INDEX
TERMS: 1829 Hydrology: Groundwater hydrology; 1831 Hydrology: Groundwater quality; 1832 Hydrology:
Groundwater transport

Citation: Zinn, B., and C. F. Harvey, When good statistical models of aquifer heterogeneity go bad: A comparison of flow,
dispersion, and mass transfer in connected and multivariate Gaussian hydraulic conductivity fields, Water Resour. Res., 39(3), 1051,
doi:10.1029/2001WR001146, 2003.

1. Introduction [Desbarats, 1990; Guswa and Freyberg, 2000; Dagan and


[2] Many researchers have considered how spatial hetero- Lessoff, 2001; LaBolle and Fogg, 2001].
geneity in hydraulic conductivity affects fluid flow and solute 2. Conductivity fields that are described by covariance
transport through porous media. This heterogeneity is often functions with long integral scales relative to the domain
characterized by a distribution, or histogram, of conductivity length. Such fields exhibit large-scale structures, flow
values, combined with a covariance, or variogram, function barriers and/or channels, that may control flow and
of separation distance. The histogram can be summarized by transport, driving behaviors that differ from flow and
its mean and variance, and the spatial covariance by its transport in conductivity fields with short integral scales.
integral scale. When these statistics differ between two [e.g., Di Federico et al., 1999].
conductivity fields, the flow and transport behaviors through 3. Conductivity fields that are composed of aligned
the two fields may be very different. Three well-known layers. These fields have effective conductivity values and
examples that are relevant to this paper are: solute spreading behavior that differs from statistically
1. Conductivity fields with a bi-modal distribution of isotropic fields [Matheron and de Marsily, 1980].
values. Solute transport through these fields may be [3] In this paper, we demonstrate that flow and transport
subjected to dual-porosity, or mass transfer, behavior, behaviors similar to those described in the three examples
creating solute plumes with longer, more dilute tails than above may all occur in isotropic log conductivity fields
plumes transported through unimodal conductivity fields. sharing the same Gaussian histogram and the same cova-
riance functions with integral scales that are a small fraction
Copyright 2003 by the American Geophysical Union. of the total domain length. Thus, conductivity fields with
0043-1397/03/2001WR001146 the same conventional spatial statistics may produce very
SBH 4-1
SBH 4-2 ZINN AND HARVEY: FLOW, DISPERSION, AND MASS TRANSFER

different groundwater fluxes, solute travel times, and solute This study also differs from previous studies in that it
spreading because of spatial patterns that are not charac- considers diffusion, allowing the work to be applied to both
terized by these conventional statistics. In other words, the small and large scales.
full univariate distribution of conductivity values and the [7] We focus our analysis on those characteristics of
spatial covariance function for these values may not provide conductivity fields that cause mass transfer behavior. Rate-
sufficient information to estimate effective flow and trans- limited mass transfer between mobile and immobile regions
port parameters. This can be particularly problematic for produces tailing, the slow release of dilute solute after the
field characterization, where data is often quite sparse and main body of a plume has advected past a particular location.
first and second-order spatial statistics are the only charac- We show that mass transfer can occur in conductivity fields
teristics easily ascertained. that have the same histogram and covariance as a multi-
[4] Conductivity fields with connected paths in extreme gaussian field, even though the multigaussian field does not
values may occur in a variety of geological formations. demonstrate significant mass transfer behavior. We also
Several papers address the occurrence of flowpaths at both show that this mass transfer behavior occurs by both
the small and large scales, as well as methods for character- diffusion and advection in low velocity regions. Very similar
izing such heterogeneity [e.g., Anderson, 1989; Koltermann solute tailing may be produced by either advection or
and Gorelick, 1996; Western et al., 2001]. Fogg [1986], diffusion through low-conductivity regions. Distinguishing
found that large scale structures in the Wilcox aquifer in between the two processes is important because tailing
Texas consisted of interconnected bodies of sands and clays, caused by slow advection will be affected by changes in
with flow controlled by the ‘‘continuity and interconnected- hydraulic gradients, such as increased gradients induced by
ness’’ of the sands. Labolle and Fogg [2001] determined higher pumping rates from remediation wells.
that high-conductivity channels dominated the flow behav- [8] While mass transfer is often not considered for con-
ior at the Livermore site. Connectedness also has a clear servative solutes in smooth conductivity fields, we demon-
presence in fracture flow, where, for example, Silliman and strate that it is possible simply by virtue of the structure of a
Wright [1988] concluded that connected higher conductivity field, provided that the conductivity variance is sufficiently
channels existed in fractured granite in Arizona. On smaller large. Here we simulate transport through fields with
scales, Journel and Alabert [1989] found connected chan- natural-log conductivity variances as large as 9. Large
nels in Berea Sandstone, as did Tidwell and Wilson [1999]. variances in ln(K) are often found in nature. An aquifer
Connectedness may be an important feature in many con- composed of equal volumes of silt regions with K = 106
ductivity fields, and therefore an important, if not dominant, m/s and sand regions with K = 103 m/s, would have a
control on solute transport. ln(K) variance of 12. The MADE site in Mississippi, which
[5] Both Wen and Gomez-Hernandez [1997] and Western has both small silt and sand regions, may have a centimeter-
et al. [2001] compared transport through two-dimensional scale ln(K) variance as large as 20 [Harvey and Gorelick,
conductivity fields that had similar spatial statistics but where 2000], although variance at the meter-scale (which averages
some fields contained large-scale connected features. Wen out small-scale variability) is considerably smaller, 4 as
and Gomez-Hernandez [1997] modified an isotropic multi- measured by borehole flowmeters [e.g., Feehley et al.,
gaussian (i.e., a multivariate normal log conductivity) field to 2000]. The Wilcox aquifer in Texas [Fogg, 1986] has high
create anisotropic non-multigaussian fields with layered variance (ln(K) variance  10), as does the Livermore site
structures of high or low conductivity. They showed that (ln(K) variance > 20), but at a large scale [LaBolle and
advective transport in these anisotropic fields differed from Fogg, 2001]. Well tests on the Culebra Dolemite, a key
advective transport in the isotropic multigaussian conductiv- facies at the WIPP site, showed a variation in hydraulic
ity field, despite identical histograms and nondirectional conductivity of six orders of magnitude [Meigs and Beau-
covariance functions (which averaged out the anisotropy in heim, 2001]. Our results agree in general terms with results
the field with connected structures). Western et al. [2001] from these sites, suggesting that our general conclusions
constructed two continuous fields with similar nondirectional about connectedness and high variance in conductivity
covariance functions; one with disordered statistically iso- fields are applicable to these field sites, and probably others.
tropic high-conductivity regions, and one crossed by two It is also important to note that many aquifers, such as the
large flow channels that created quick solute breakthrough. Cape Cod and Borden sites, show much lower hydraulic
They characterized the difference between the two fields with conductivity variability, and our results tend to be consistent
connectivity statistics, a function that charts the probability with transport results from these systems as well.
that a certain subset of conductivity values are all connected [9] Our results are determined through numerical gener-
to one another by values within that same subset (similar to ation of discrete, two-dimensional conductivity fields and
the concept of percolation thresholds). Both of these studies, simulation of flow and transport in these fields. We generate
which consider solute advection but not diffusion, best gridded hydraulic conductivity fields that differ from a
describe transport through large-scale heterogeneity where standard multigaussian field only in higher order cross-
the effects of small scale mixing can be ignored. moments, so that high or low conductivity values are more
[6] To our knowledge, this is the first study to compare connected. We then simulate flow and transport (both
upscaled flow and transport behavior in fields with the same advection and diffusion) in these fields, estimate effective
Gaussian log conductivity histogram and second order flow and transport parameters, and compare the parameters
statistics including the same directional covariance func- between the different fields.
tions with no artificially imposed large structures. We [10] Our investigation of the effects of connectedness is
compare behavior among statistically isotropic fields, rather intended to be simple and pragmatic. We undoubtedly leave
than between an isotropic field and an anisotropic field. open opportunities to further apply stochastic methods
ZINN AND HARVEY: FLOW, DISPERSION, AND MASS TRANSFER SBH 4-3

developed for multigaussian fields to connected fields. We dispersion coefficient describes solute spreading due to
do not intend to cover all of these considerations in a single variation in the velocity caused by spatial heterogeneity of
paper, rather we concentrate on the basic question: What conductivity at a smaller scale than is explicitly modeled.
model best explains breakthrough curves in fields with The first-order approximation of velocity variance in both
different characteristics of connectedness? the x and y directions ( parallel and perpendicular to flow,
respectively), for a two-dimensional, isotropic multigaus-
sian field from Gelhar [1993] is:
2. Background
2.1. Multigaussian Fields and Connectedness 3
s2vx  s2Y Keff i ð2aÞ
[11] The multigaussian distribution is the standard stat- 8
istical model used to describe spatially heterogeneous
hydraulic conductivity. Our use of the term multigaussian 1
s2vy  s2Y Keff i ð2bÞ
is shorthand for a multivariate Gaussian distribution, which 8
is detailed at length in numerous sources [e.g., Anderson,
1958]. The multigaussian model is relatively tractable, and where Y = ln(K) and sY2 is the variance of Y. Keff is the
is the basis of analytical solutions for effective conductivity effective conductivity (geometric mean) of K, and i is the
and macrodispersion fluxes [e.g., Dagan, 1989; Cushman, hydraulic gradient. The late-time macrodispersion coeffi-
1990; Gelhar, 1993]. A two-dimensional isotropic multi- cient for relatively small variance is derived from the
gaussian field is shown in Figure 1a. velocity statistics to be:
[12] One characteristic of multigaussian fields is that D ¼ Av  ls2Y v ð3Þ
extreme values in the field are the least connected [e.g.,
Journel and Deutch 1993], which is to say that extreme where A is the macrodispersivity, v indicates mean velocity,
values tend to cluster in isolated blobs, rather than form and the integral scale, l, is assumed to be finite. Many
channels or structures that span the entire length of the field. studies consider extensions and improvements of these
These characteristics have been quantified by Journel equations. Here, we found that the simple first-order results
[1983], Journel and Alabert [1989], and Gomez-Hernandez were sufficiently accurate to describe our numerical results
and Wen [1998] using indicator variograms. Xiao [1985] in multigaussian conductivity fields.
showed that the integral scale for the median indicator is [15] We also found it useful to compare some of our
maximal, and symmetrically decreases at higher or lower results to analytic results for macrodispersion in stratified
values. This pattern is visually evident in Figure 1a. In aquifers. For flow perfectly parallel to layered conductivity
multigaussian fields, values close to the mean tend to form variations, Gelhar et al. [1979] found that if the hydraulic
field-spanning structures. conductivity is described by a restricted class of covariance
2.2. Effective Conductivity functions that integrate to zero (i.e., hole effect covariance
functions), then the macrodispersion coefficient asymptoti-
[13] The effective hydraulic conductivity of a multigaus- cally approaches a value that increases quadratically with the
sian field appears to be well described by Matheron’s [1965] mean velocity and decreases with the diffusion coefficient:
conjecture [e.g., Dagan, 1989; Gelhar, 1993; Renard and de
Marsily, 1997], which, for an isotropic two-dimensional log v2
conductivity field, reduces to the geometric mean. This holds D¼dþB ð4Þ
d
true even for high variance multigaussian fields, but does not
hold true for other isotropic fields such as blobs of one where d is the apparent diffusion coefficient and B is a
conductivity embedded in a matrix of a different uniform coefficient that depends on the statistics of the conductivity
conductivity. Upper and lower bounds can be set on the layers. These results are analogous to classical Taylor – Aris
effective conductivity of any heterogeneous collection of dispersion in laminar flow through straight tubes or
conductivity values. The upper bound of the effective con- fractures [Taylor, 1953; Aris, 1956]. Matheron and
ductivity is the arithmetic mean, and the lower bound is the deMarsily [1980] showed that if the hydraulic gradient is
harmonic mean, which correspond to flow through a per- not exactly parallel to stratification, as often happens under
fectly layered system, parallel and perpendicular to the natural conditions, the macrodispersion coefficient ap-
layering, respectively. proaches an asymptote that also increases quadratically
with the mean velocity and decreases with the diffusion
2.3. Macrodispersion Coefficients coefficient, but in this case, the covariance function need
[14] Solute transport in porous media is often described not integrate to zero, so the result holds true for commonly
with the advective – dispersive model: used covariance functions.

@C   2.4. Mass Transfer Coefficients


¼ vrC þ r DrC ð1Þ
@t [16] Researchers have also employed rate-limited mobile –
immobile domain mass transfer models to describe solute
Where C is the solute concentration, v is the fluid velocity transport through heterogeneous media. In addition to
vector, and D is the dispersion coefficient tensor. A large describing intragranular diffusion and sorption, these mod-
body of literature relates the dispersion coefficient, or els are often used to describe transfer of solute in and out of
macrodispersion coefficient, to the variance and integral low-conductivity regions [e.g., Barenblatt, 1960]. Most
scale of the log conductivity field, as reviewed by Dagan studies consider physical mass transfer to be driven by
[1989], Gelhar [1993] and Cushman [1990]. The macro- diffusion, but some consider slow advection in and out of
SBH 4-4 ZINN AND HARVEY: FLOW, DISPERSION, AND MASS TRANSFER

Figure 1. Generation of a connected hydraulic conductivity field from a multigaussian field. (a)
Multinormal field generated by sequential Gaussian simulation. (b) The zero-mean field produced by
taking an absolute value transform followed by a normal-scores transform (equation (6)). The low
conductivity values are now the most connected. (c) The connected field generated by reversing the sign
of the values in field (b) and stretching the coordinate axis so that the field has the same correlation length
as field (a). See color version of this figure at back of this issue.

low permeability regions. Recently, Harvey and Gorelick results. In this aquifer, the variability in hydraulic conduc-
[2000] have used simple mass transfer models to explain tivity is large and the macrodispersion model failed to
nonreactive transport through the sedimentary aquifer at the describe the observed solute behavior as well as it has in
Macrodispersion Experiment (MADE) field site. Feehley et more homogeneous aquifers [e.g., Freyberg, 1986; Hess et
al. [2000] and Julian et al. [2001] have corroborated these al., 1992].
ZINN AND HARVEY: FLOW, DISPERSION, AND MASS TRANSFER SBH 4-5

[17 ] The first-order mass transfer equation may be


coupled with the advective-dispersive equation as:

@C @S  
þb ¼ vmobile rC þ r DrC ð5aÞ
@t @t

@S
¼ aðC  SÞ ð5bÞ
@t

where S is the concentration of the solute in the immobile


domain. This model assumes that there is a mobile domain
(i.e., a region through which solute advects at the ground-
water velocity in mobile region, vmobile), and an immobile
domain for which mass transfer with the mobile domain is
rate-limited. The ratio of rate-limited or nonequilibrium Figure 2. Comparison of spatial correlation functions for
immobile domain pore space to mobile domain pore space is the multigaussian field (Figure 1a) and the connected field
given by the capacity coefficient, b. The rate at which solute (Figure 1c) showing that they are nearly identical. Also
moves between these two domains is controlled by a, which shown are the theoretical correlation function of the
depends on the particular size and geometry of these regions. absolute value transform of the multigaussian field
More sophisticated diffusive mass transfer models account (application of equation (A10)) and the calculated covar-
for a range of mass transfer rate coefficients [Haggerty and iance of field (Figure 1b), indicating that these are similar.
Gorelick, 1995], and have been shown to better represent
field data [McKenna et al., 2001]. However, multirate mass
transfer entails much more complicated estimation, and a field with connected patterns of high conductivity, and; (3)
hence we chose to use a simpler model in this paper. a field with disconnected high conductivity, but connected
[18] Some researchers [e.g., Griffioen et al., 1998] have patterns of low conductivity. All three of these fields share
also considered advective mass transfer. Guswa and Frey- nearly identical univariate distributions and spatial cova-
berg [2000] recently demonstrated that advection through riance functions. Thus, all three conductivity fields share the
low permeability inclusions can lead to breakthrough curves same basic statistics, the first and second statistical
with tailing very similar to that caused by diffusion in and moments, but their structures differ in how the high and
out of low permeability regions. Thus, from a single experi- low values are connected.
ment, it could be very difficult to tell the two processes apart. [21] The isotropic multigaussian field (Figure 1a) consists
of a regular grid of 800 by 500 point values representing grid
2.5. Evolution of Transport Coefficients Over blocks generated by the sequential Gaussian simulation algo-
Travel Path rithm [Deutch and Journel, 1997] using a Gaussian cova-
[19] The macrodispersion literature, both from an Euler- riance function. The correlation of the field as a function of
ian [e.g., Gelhar, 1993] and Lagrangian [e.g., Dagan et al., separation distance is shown in Figure 2. The integral scale is
1992] point of view, describes solute transport as preasymp- approximately 1.1 percent of the length of the field, 9 blocks.
totic if the time-rate-of-growth of the second spatial moment [22] The connected (Figure 1c) and disconnected fields
of a plume has not stabilized. The travel distance for the were generated through a transformation of the multigaus-
effective dispersion coefficient to approach asymptotic sian field in four steps:
conditions has been calculated to be on the order of 10 1. The absolute value of the multigaussian field (zero-
correlation lengths [Dagan, 1984]. This notion of prea- mean, unit-variance) in Figure 1a was calculated. This
symptotic behavior does not apply directly to mass transfer transform shifts extreme values to become high values, and
models, as the spreading and tailing caused by mass transfer values originally close to the mean become low values.
occur only over timescales on the order of a1. In contrast 2. The histogram of the values in the field was converted
to the macrodispersive model, the mass transfer model back to a univariate Gaussian distribution by mapping the
predicts a nonconstant time-rate-of-growth of the second CDF value at each point to a standard normal CDF. This
spatial moments of a plume over timesscales on the order of transformation can be written explicitly as:
a1, even though all of the coefficients are constant. At    
pffiffiffi 1
timescales much larger than a1, transport may be modeled Y0 ¼ 2erf
Y
2erf pffiffiffi  1 ð6Þ
with a simple retardation factor, and the effects of rate- 2
limited mass transfer are no longer evident. Thus the tailing
behavior that the mass transfer model produces is a pre- where Y0 are the transformed values of ln(K) and Y are the
asymptotic effect. However, it is this effect that we wish to original values. This creates a field in which the extreme
understand and to model because of its importance for low values are connected and the high values form isolated
practical problems such as contaminant cleanup. blobs—the field shown in Figure 1b.
3. The block size of the field was increased so that the
integral scale matched that of the original multigaussian
3. Method
field. This provided the final disconnected field.
3.1. Generation of Connected Fields 4. The connected field (Figure 1c) was then generated
[20] We modeled transport behavior in three types of from the disconnected field by reflecting the values of the
conductivity fields: (1) a standard multigaussian field; (2) disconnected field around the mean, so connected patterns
SBH 4-6 ZINN AND HARVEY: FLOW, DISPERSION, AND MASS TRANSFER

of low conductivity become connected flow paths of high method when only advection is simulated. However, it can
conductivity. not be directly implemented when diffusion is considered.
[23] Appendix A gives an analytical solution for the [29] We combined the strengths of both approaches (at
covariance of the transformed field (after step 1, but before some cost in computational efficiency) by modeling diffu-
step 2), and shows how the integral scale is reduced by sion as random displacements during discrete time steps
the absolute value transform, which necessitates step 3. The while integrating the interpolated velocity overtime as in the
correlation functions for the multigaussian field and the boundary method. Instead of integrating the velocity over
connected fields, along with the analytic results, are shown the particle’s full path across the block domain, we inte-
in Figure 2. The results from Appendix A are in close grated through time, for an amount equal to the time step.
agreement with the calculated correlations from the realiza- This method more accurately tracks advective trajectories,
tion, indicating that the normal-scores transform (step 2) has while approximating the simultaneity of advection and
only a minor effect on the covariance function, despite the diffusion. Positions of particles within a block after a certain
fact that the transform is nonlinear. The integral scale of the amount of time has elapsed were solved analytically with
multigaussian field and the connected field differ by a factor linear interpolation using equation (12) from Pollock
of 1.86. Therefore, the separation distance between points [1988]. Diffusive movement (with a spatially homogeneous
was increased by the factor of 1.86 (step 3, shown above), diffusion coefficient) was incorporated by adding a zero-
and the number of rows and columns was reduced by a mean normally distributed random displacement pffiffiffiffiffiffiffiffiffiffiffi in both the
factor of 1.86 in order to maintain the same domain size. x and y direction, rx and ry , with s = 2dt. To insure
[24] An alternative approach is to use the covariance against excessively large advective displacements in blocks
function of the connected field (after steps 1 and 2) to with large velocities, we set t such that the maximum
generate a new multigaussian field. This approach provides velocity in the entire field multiplied by t did not exceed a
the possibility of producing identical, rather than nearly distance of one half of a block width.
identical, covariance functions. However, for a single ran- [30] The numerical accuracy of the scheme was verified
dom field the variability between the experimental correla- against two known solutions: the solution for Taylor dis-
tion function and the function used to generate the field is persion in a parabolic velocity field [Taylor, 1953], and the
typically as large as the difference between our connected solution for advection and diffusion through a high-con-
and multigaussian fields. ductivity matrix with embedded circular impermeable
[25] Figure 1c shows the final connected field after all four regions [Goltz and Roberts, 1986], which includes sharp
of the steps above. The disconnected field is simply the interfaces between zero and finite velocities. The simula-
inverse, or reflected, version of the connected field, so the tions of the Taylor – Aris flow showed excellent agreement,
high-conductivity flow channels become low conductivity with both the first and second spatial moments of a particle
barriers. The connected and disconnected fields now have plume agreeing with the theoretical moments to within 0.1
histograms and experimental covariance functions (Figure 2) percent. The comparison between the theoretical moments
that are nearly identical to those of the multigaussian field. of the embedded cylinders scenario with our simulated
These ln(K) fields are used in all of the following simula- particle plume were also good, with about 1 percent error
tions of flow and transport. on the second moment, which we believe is caused in part
by the use of a finite number of particles, and in part by the
3.2. Simulation of Flow and Transport fact that the circular regions must be approximated with a
[26] A head gradient was applied across the field, with circular assemblage of squares.
constant head boundaries on the left and right borders, and [31] In each of the numerical experiments described below,
no-flow boundaries on the top and bottom, driving flow 1000 particles were initially distributed on a vertical line
from left to right. The steady state, two-dimensional flow located 19 correlation lengths from the left boundary. We
equation was solved with a finite difference approximation. considered two distributions of particles along this vertical
The velocity field was calculated along block edges by line. In the first method, the density of particles is proportional
taking the difference in the heads of the two adjoining to the velocity at a given point. In the second method, the
blocks, multiplying by the mean conductivity, and dividing particles are evenly spaced. The first, flux-proportional,
by a uniform porosity. method has the advantage that it better simulates how par-
[27] Although we will work with dimensionless parame- ticles would be introduced into an aquifer from a well screen,
ter groups, it is, perhaps, useful to think of the following or into an experimental column or box along a constant head
possible parameter values: a length of 8 m (800 1cm square boundary. It is analogous to a Dirac pulse in the flux concen-
blocks), effective conductivity of 1.8*106 m/s (16 cm/ tration [Kreft and Zuber, 1978]. The second method, in which
day), gradient of 0.02, and porosity of 0.3. This leads to a particles begin evenly spaced, has the advantage that compar-
mean flow velocity for the multigaussian field (shown in ison to theoretical results from the advective – dispersive
Figure 1a) of approximately 1.2*107 m/s (1.1 cm/day). model is easier, with an initial condition that is simply a Dirac
Note that flow velocities in the other fields will be different, pulse in the resident concentration. Vanderborght and Ver-
as discussed in section 4. eecken [1998] determined that for a two dimensional system
[28] Conservative solute transport (advection and diffu- with an exponential variogram and ln(K) variance of approx-
sion) was simulated with particle tracking by combining two imately 2.5, the choice of initial conditions could affect
commonly used methods—the constant time step method transport parameters to a distance of about three correlation
[e.g., Goode, 1990] and the ‘‘boundary to boundary’’ lengths. Theoretical predictions with the mass transfer model
method [Pollock, 1988]. The boundary method is more can easily be made with either initial distribution of particles.
accurate and computationally efficient than the time step The velocity-proportional case corresponds to an initial con-
ZINN AND HARVEY: FLOW, DISPERSION, AND MASS TRANSFER SBH 4-7

dition with zero immobile solute, and the evenly spaced case where the nondimensional parameters are:
corresponds to an initial condition where the mobile and x
immobile concentrations are at equilibrium. x0 ¼ ð10aÞ
L
[32] Breakthrough was recorded at another vertical line
that was 67 correlation lengths from the injection line (e.g.,
600 blocks in the multigaussian field, 323 blocks in the v
t0 ¼ t ð10bÞ
connected and disconnected fields). This allows comparison L
of consistent tracking distances in the three fields, and is
also a distance large enough to assure that the solute plume A
passed through many correlation lengths of heterogeneity. Pe ¼ ð10cÞ
l
3.3. Estimation of Effective Parameters From
Breakthrough Curves L
Da ¼ a ð10dÞ
[33] The solute transport model used to fit the simulated v
breakthrough curves incorporates both dispersion and
mobile – immobile domain mass transfer. The immobile Note that the capacity coefficients (b’s) are already
domain is divided into one subdomain that contributes to dimensionless. Pe is the common notation for the Peclet
rate-limited mass transfer over the timescale of observa- number, while Da, the Damköhler number, is similar to the
tions, and a second noncontributing subdomain in which parameter by Bahr and Rubin [1987] except that their
mass transfer is too slow to significantly affect solute Damköhler number is formulated for kinetic chemical
concentrations over the timescale of observations. Hence sorption instead of diffusive mass transfer. Also note that
our model is a two-site model, where one portion of the equation (9a) includes a nondimensional factor that is the
immobile domain has a finite mass transfer rate coefficient, ratio of the integral scale to the length of the domain (l/L).
and another portion has a rate coefficient of zero. The We address the importance of this parameter in a later
mobile domain velocity is then: section of the paper.
  [36] An analytical solution to the first-order mass transfer
vmobile ¼ vð1 þ btot Þ ¼ v 1 þ bnoncont þ bnoneq ð7Þ equation is fit to the simulated breakthrough curves to
where v is the average velocity in the aquifer as a whole, estimate the four effective parameters: Pe, bnoncont, bnoneq,
and btot, the ratio of all immobile domain pore space to the and Da. The values of Da and bnoneq, in particular, will
mobile domain pore space, is the sum of the capacity quantitatively indicate the importance of nonequilibrium
coefficients for both the noncontributing subdomain, mass transfer, while the sum of bnoneq and bnoncont will
bnoncont, and the nonequilibrium subdomain, bnoneq. indicate the magnitude of the immobile domain.
[34] In this paper, we will employ the one-dimensional [37] The distribution of arrival times of particles is con-
form of equation 5, which (incorporating equations (3) and verted to a discrete series of concentration values by
(7)) is: summing all of the particles that come through the break-
through point between two times into a single total value.
@C @S   @C This total mass is then divided by the flux and the total time
þ bnoneq ¼ v 1 þ bnoncont þ bnoneq elapsed over which the particles were summed in order to
@t @t @x
  @2C give a concentration at the midpoint of each time interval.
þ Av 1 þ bnoncont þ bnoneq ð8aÞ Once we have the results in terms of concentration, we
@x2
estimate the four parameters, with the Eulerian analytical
@S solution of Goltz and Roberts [1987] modified for our
¼ aðC  SÞ ð8bÞ
@t parameter groups:
8 9
This form is similar to that used by Harvey and Gorelick > >
Me x0 =2LlPe < 0
x  ðs Þ 0 =
[2000] to model transport at the MADE site, with the Cðx0 ; s0 Þ ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi exp  qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð11Þ
exceptions that they incorporated a retardation factor to 2ðs0 Þ Ll Peð1 þ btot Þ >
: l >
;
L Peð1 þ btot Þ
describe pore-scale immobile regions, and they did not
consider a noncontributing domain. Because the mean
groundwater velocity was not known at the MADE site, the Where C(x0, s0) is solute concentration in the Laplace
effects of such a noncontributing domain, if it existed, domain, and (s) is defined by:
would have been accounted for by a larger estimate of the sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
mean velocity. 0 1 þ btot Dabnoneq s0
[35] Equation 8 can then be nondimensionalized by nor-  ðs Þ ¼ þ 0 þ s0 ð12Þ
4 Ll Pe s þ Da
malizing the distance by the length of the domain, L, and time
by L/v, the mean time for fluid to flow across the domain:
s0 is the normalized Laplace parameter, s0 = sLv, and M is the
@C @S   @C initial solute mass. The concentration in time is then
þ bnoneq 0 ¼  1 þ bnoncont þ bnoneq
@t 0 @t @x0 calculated by inverse Laplace transform of C(x, s0) using the
l   @2C algorithm from Hollenbeck [1998, available at ftp://
þ Pe 1 þ bnoncont þ bnoneq ð9aÞ ftp.mathworks.com/pub/contrib/v4/math/invlap.m] by way
L @x02
of de Hoog et al. [1982]. The theoretical breakthrough
@S curve is fit to the simulation breakthrough curve by simple
¼ DaðC  SÞ ð9bÞ
@t least squares minimization.
SBH 4-8 ZINN AND HARVEY: FLOW, DISPERSION, AND MASS TRANSFER

primary direction of flow) over any transect perpendicular to


flow and dividing by the hydraulic gradient. The resulting
effective conductivities, normalized by the geometric mean
conductivity, are plotted in Figure 3 as a function of the ln(K)
variances of the fields. This Figure compares the effective
conductivities for each of the three fields (connected, dis-
connected, and multigaussian) to the geometric, arithmetic,
and harmonic mean conductivities, which are the same for
all three fields because the fields share the same histogram.
[40] The effective conductivity for the multigaussian field
agrees well with theory—for all variance values it remains
close to the geometric mean, the theoretical prediction. The
effective conductivities of connected and disconnected
fields deviate significantly from the geometric mean. For
the connected field, the effective conductivity exceeds the
geometric mean, and for the disconnected field is below the
Figure 3. The effective conductivity for the three different geometric mean. These differences increase as the variance
conductivity fields: multigaussian, connected, and discon- increases.
nected. These are shown as compared to the arithmetic, 4.1.2. Velocity Variability
geometric, and harmonic means of the conductivity values. [41] The connected field shows by far the largest velocity
The values are all normalized by the geometric mean of the variation (Figure 4), and one that steadily increases with
conductivity values. increasing conductivity variance. The multigaussian field
matches the theoretical prediction quite well in the x
direction and at low variance in the y direction, but does
[38] We found that considering the entire breakthrough not match as well at high variance in the y direction. Higher-
curve, rather than just temporal or spatial moments of the order approximations for velocity variance have shown
solute plume, provided more accurate parameter estimates. some potential for deviation from the first order model in
Estimating mass transfer parameters from moments requires directions perpendicular to flow [e.g., Deng and Cushman,
the calculation of higher order moments, which are subject 1998]. These results suggest that solute spreading, and
to increasing errors as more weight is applied to extreme hence effective dispersivities, should increase in all three
values [Harvey and Gorelick, 1995]. field types as the variance increases, and that the connected
field may have higher dispersivities than the multigaussian
4. Results field, which may be higher than the disconnected field.
4.1. Part I—Characteristics of the Flow Fields 4.2. Part II—Solute Transport Characteristics
4.1.1. Effective Conductivity 4.2.1. Mass Transfer Behavior
[39] The effective conductivity of each field is calculated [42] The simulated breakthrough curve and model fits in
by averaging the specific discharge in the x direction (the Figure 5 demonstrate that using the mass transfer model can

Figure 4. Coefficient of variation in velocity for the three different spatial patterns of conductivity as a
function of the spatial variance of ln(K). Also shown is the theoretical prediction of the coefficients of
variation. The left figure shows the variation in the direction of flow (x direction) and the right-hand
figure shows the variation perpendicular to the direction of flow ( y direction).
ZINN AND HARVEY: FLOW, DISPERSION, AND MASS TRANSFER SBH 4-9

Figure 6. A simulated breakthrough curve, the best fit


using the complete model equation (equation 9), and two
fits made without considering a noncontributing immobile
Figure 5. Comparison of breakthrough curve fits using the domain—one fits the peak of the breakthrough curve, one
advective – dispersive model and the single-rate mass matches the tail. Neither do both, whereas the complete
transfer model. The breakthrough curve was simulated in model better fits the curve at all times.
the connected conductivity field with a variance of 9 and no
diffusion. The breakthrough curves are shown in the form of are plotted against normalized time. The figures show
cumulative mass fraction breakthrough. The mass transfer breakthrough curves for various values of normalized dif-
model provides a better fit, especially in the tail. fusion coefficient:

dL
d0 ¼ ð13Þ
significantly improve the model fit to simulated solute vl2
transport. In the case of the connected conductivity field
with a ln(K) variance of 9 and no diffusion, both models where d is the diffusion coefficient of the solute particles.
adequately fit the peak of the breakthrough curve. However, This normalized diffusion coefficient represents the ratio of
the macrodispersive model alone poorly fits the tail of the the rate of diffusion in and out of structures with
breakthrough curve (Figure 5). This tail is frequently a characteristic length, l, to the rate of advection through
problem in real-world situations, and thus its character- the field. If the value is greater than 1, then the mean time to
ization is of primary importance. The tail is a significant diffuse in and out of statistical structures in the field is
percentage of the total mass, and breakthrough of this mass smaller than the mean time to advect across the field.
continues well after the macrodispersive model predicts [45] The breakthrough curve for the connected field
insignificant concentrations, as is clear in the cumulative (Figure 8) shows tailing in the high variance cases, with
breakthrough curve. the most significant tailing occurring with low and inter-
[43] Attempting to fit the breakthrough curve with mediate diffusion coefficients. The multigaussian field has
bnoncont set to zero (Figure 6) demonstrates the importance very small tails in its breakthrough curves in the high
of modeling a subdomain that does not affect solute over the
timescale of the numerical experiment. Figure 6 shows that
either the peak of the curve (the sharp ascent at the front of
the curve) or the tail of the curve can be matched, but
without the use of bnoncont, the complete curve cannot be
reproduced. Thus, there appears to be a significant fraction
of the domain through which very little solute passes over
the time-scale of this particular simulation. Combining a
nonequilibrium and noncontributing immobile domain, we
can match the breakthrough curve quite well, as can be seen
from Figures 5 and 6.
4.2.2. Transport Through Connected, Multigaussian,
and Disconnected Conductivity Fields
[44] The breakthrough curves (Figure 7) are substantially
different for the three different fields, although these fields
share the same basic statistics. Much of the differences
among these breakthrough curves is explained simply by
the difference in mean fluid velocity (effective conductiv- Figure 7. Breakthrough curves for the three different
ity), discussed in the previous section. However, there are patterns of hydraulic conductivity with variance of 9 and no
also significant differences in spreading and tailing, as diffusion present. Demonstrates different mean arrival
demonstrated in Figure 8, where the breakthrough curves times.
SBH 4 - 10 ZINN AND HARVEY: FLOW, DISPERSION, AND MASS TRANSFER

Figure 8. Cumulative breakthrough curves for the three different patterns of hydraulic conductivity
after normalizing time by mean fluid travel time. Breakthrough curves are plotted for combinations of
differing normalized diffusion coefficient (columns) and sY2 (rows) for each of the three patterns of
hydraulic conductivity.

variance case, which disappears for high diffusion coeffi- [47] Because the parameters are in nondimensional groups,
cients. The disconnected field does not appear to have these results can be applied over a large range of length and
significant tailing. The difference between the curves for timescales. However, it is useful to check that the values
different conductivity fields diminishes for high values of plotted in Figure 9 may represent practical situations. For
the diffusion coefficient. example, if we consider transport over a length of 6m, for a
4.2.3. Breakthrough Curve Analysis—Parameter groundwater velocity of 3.5*107 m/s (3 cm/day), an appa-
Estimates rent diffusion coefficient of 0.35 cm2/day (4*1010 m2/s,
[46] The qualitative observations described above are approximately the diffusion coefficient in porous media for
quantified by estimating transport and mass transfer para- many dyes), and a representative length of heterogeneities
meters from the simulated breakthrough curves. The four of 9 cm, then the normalized diffusion coefficient, d0
parameters estimated were: the Peclet number, Pe (the non- (equation (13)), is approximately 1, which falls near the
dimensionalized form of the macrodispersivity), the two center of each plot.
capacity coefficients, bnoncont and bnoneq, and the Damköhler 4.2.3.1. Dispersivity (Pe)
number, Da (the nondimensionalized form of the nonequili- [48] The normalized dispersivity results in Figure 9 are
brium mass transfer rate coefficient). A complete summary consistent with the trends in velocity variances. For all three
of the estimated parameters is shown in Figure 9, which is a fields, the estimated dispersivity increases with variance in
series of contour plots of the four parameters for the three ln(K) and hence with velocity variance. Also, the estimated
different patterns for hydraulic conductivity (multigaussian, dispersivity in the connected field is greater than the multi-
connected, and disconnected). The parameters shown are gaussian field, which in turn is greater than the disconnected
those estimated with the flux-proportional starting condi- field, again consistent with velocity variance.
tions. The x axis of each plot indicates the normalized [49] The first-order approximation (equation (3)) for the
diffusion coefficient (d0). The y axis of each plot indicates dispersivity in a multigaussian field is simply the variance
the ln(K) variance of the conductivity field. The contour in ln(K) when normalized by the correlation length. Because
plots show the estimated parameter for the particular field we have already normalized our dispersivity estimates by
type as a function of these two variables. the correlation length, the contours of our estimates should
ZINN AND HARVEY: FLOW, DISPERSION, AND MASS TRANSFER SBH 4 - 11

Figure 9. Estimated transport parameter values for each of the three different patterns of hydraulic
conductivity. Each plot shows contours of an estimated parameter as a function of the ln(K) variance, sY2
( y axis) and the normalized diffusion coefficient d0(x axis).

intersect the ln(K) axis at the value of the contour lines, if coefficient and increase with increasing field conductivity
the first-order multigaussian result applies. For the con- variance.
nected field, this is clearly not the case. The dispersivity 4.2.3.2. Noncontributing Capacity Coefficient (Bnoncont)
values are significantly larger than the multigaussian theory [51] The noncontributing capacity coefficient (second
predicts. For the disconnected field, the estimated disper- column of Figure 9) indicates the proportion of the flow
sivities are much lower at high ln(K) variance than pre- field from which solute is excluded. In conjunction with the
dicted. For the multigaussian field, the results are much nonequilibrium capacity coefficient, it also indicates how
closer to the theory, but do not perfectly agree. This much faster the average advective flow of solute in the field
discrepancy may be caused by the inclusion of mass transfer is relative to the average fluid velocity in the field.
processes in our model, which will lead to lower estimates [52] The values of bnoncont are relatively large in the
of dispersivity because some of the spreading is accounted connected field at high variance. At a variance of 9 with
for by the mass transfer parameters, or by the flux-propor- no diffusion, the value of bnoncont reaches 1.7, indicating that
tional initial conditions. solute is not interacting with a large portion of the con-
[50] The estimated dispersivity for the connected field is ductivity field. As the diffusion coefficient increases, the
generally decreasing with an increasing diffusion coeffi- noncontributing capacity coefficient steadily decreases
cient, d0. This is similar to the results for dispersion in because diffusion mixes solute into a larger proportion of
stratified media [Matheron and deMarsily, 1980], which the domain. The multigaussian and disconnected fields
also has continuous paths of relatively high velocity. An show substantially smaller values of bnoncont.
exception exists for small diffusion coefficients and large 4.2.3.3. Nonequilibrium Capacity Coefficient (Bnoneq)
ln(K) variances, where the dispersivity, in fact, increases. [53] The nonequilibrium capacity coefficient (third col-
The disconnected field shows increasing dispersivity with umn of Figure 8) describes the ratio of the pore volume
increasing ln(K) variance, but exhibits little dependence on subject to nonequilibrium, or rate-limited, mass transfer, to
diffusion coefficient. The multigaussain field shows evi- the pore volume of the mobile domain. In concert with a
dence of dispersivities that increase with the diffusion value of Da on the order of 1, a high value of bnoneq
SBH 4 - 12 ZINN AND HARVEY: FLOW, DISPERSION, AND MASS TRANSFER

indicates that significant nonequilibrium mass transfer through low-conductivity regions rather than diffusion.
occurs. First, the nonequilibrium capacity coefficients were esti-
[54] By far the largest amount of nonequilibrium mass mated to be significantly large in simulations in which the
transfer occurs in the connected field, with b reaching a diffusion coefficient was zero. Without diffusion, the only
maximum value of 0.82 for a ln(K) variance of 9, and process driving mass transfer in and out of immobile
normalized diffusion coefficient of approximately 0.3. The regions in these simulations is advection. Second, the rate
disconnected field shows bnoneq values that are no greater coefficient did not increase in direct proportion to the
than 0.1, indicating only minimal nonequilibrium mass diffusion coefficient, as would be expected if mass transfer
transfer. The multigaussian field also shows relatively small were driven by diffusion.
bnoneq values. The highest value of bnoneq in the multi- [59] To examine the effects of advection on mass trans-
gaussian field is 0.2, which occurs in the multigaussian field fer for different diffusion coefficients, solute transport was
with variance of 9 and with normalized diffusion coeffi- simulated through the conductivity fields with the head
cients near 0.5. While much less significant than the values gradient increased by a factor of two, doubling all veloc-
for the connected field, this value is high enough to cause ities, while other conditions were held constant. The
the slight tailing noted earlier in the qualitative descriptions. importance of advection is then ascertained by examining
In all three fields the value of bnoneq increases with increas- the estimated rate coefficients, a, from the subsequent
ing ln(K) variance, consistent with the idea that as the breakthrough curves. If nonequilibrium behavior is entirely
variance of the fields increase, the proportion of conductiv- advection-dependant, the estimate of a will also increase
ity values that are sufficiently low to be treated as immobile by a factor of two, as the rate of transfer in and out of
also increases. immobile regions driven entirely by advection will scale
[55] Although the largest values of bnoneq in the con- identically with velocity. The assumptions for both fields,
nected field are found when d0 is greater than 0, but less than in both simulations, were that the mean conductivity was
1, significant values of bnoneq are also observed when the 8 cm/day, the total travel distance was 6 m, and the
normalized diffusion coefficient is 0, indicating advective- hydraulic gradient was either 0.01 or 0.02.
driven mass transfer. The relative importance of advective [60] The results for the two connected fields with ln(K)
and diffusive driven mass transfer is discussed later. variances of 6.25 and 9 (Figure 10) demonstrate that there is
4.2.3.4. Rate Coefficient (Da) an advective component to the nonequilibrium behavior in
[56] The mass transfer rate coefficient (nondimensional- both of these fields. In the absence of diffusion, estimated
ized as Da in the last column in Figure 9) indicates how rate coefficient values, a, double when the velocity is
quickly solute is moving in and out of effectively immobile doubled, indicating that mass transfer is driven only by
zones if nonequilibrium mass transfer is occurring. Values on advection. As the diffusion coefficient is increased, the
the order of 1 indicate that the timescale of mass transfer is estimates of a converge to the same value for both veloc-
comparable to the time to advect across the field. If Da is ities, indicating that diffusion is the dominant process,
much higher than 1, then mass transfer is occurring so quickly overwhelming advection. Parameters are shown in dimen-
that equilibrium mass transfer is present instead of nonequili- sional form because increasing the velocity by a factor of
brium mass transfer. If Da is much lower than 1, then the two causes the normalized coefficients of the two cases to
solute does not have sufficient time to move into the immo- vary. To prevent confusion, the estimates are shown as their
bile regions, and thus, during the timescale of the simulations, actual values.
these regions remain inaccessible. Such processes would be [61] We can gain some insight into conditions under
accounted for by the estimated bnoncont. The value of Da is which mass transfer is dominated by diffusion or advection
only meaningful when it is associated with a significant by considering the ratio of the estimated a’s for the two
immobile volume as indicated by a value of bnoneq signifi- different velocity conditions. A ratio that is less than 1.5
cantly larger than 0. Thus, the mass transfer rate coefficients suggests diffusion is more important. By this criterion, in
for the connected case are meaningful. For the multigaussian the connected field with variance of 6.25, advective mass
and disconnected fields, the value of Da is less important. transfer dominates up to a diffusion coefficient of 0.24 cm2/
[57] The estimated Da’s for the connected field increase day, which corresponds to a normalized diffusion coeffi-
with increasing variance in ln(K). This indicates that the cient in the base case of about 2. Though diffusion appears
effective rate that solute is transferred between mobile and to be the more dominant mass transfer process above this
immobile regions also increases as the contrast between threshold, the capacity coefficient is small (bnoneq < 0.2), so
high- and low-conductivity regions increases (and the there is actually very little rate-limited mass transfer at these
volume of the immobile region subject to nonequilibrium high diffusion coefficients. Thus, overall, advective mass
mass transfer increases). We also find that for high ln(K) transfer may be the dominant nonequilibrium process in the
variances, the rate coefficient is highest when there is no connected field with ln(K) variance of 6.25. For the con-
diffusion. These results are at odds with the notion that mass nected field with a variance of 9, advective mass transfer
transfer is caused exclusively by diffusion. If mass transfer dominates up until a diffusion coefficient of 0.21 cm2/day,
was caused by diffusion in and out of immobile regions of a which corresponds to a normalized diffusion coefficient of
size characterized by the correlation length, l, then the approximately 0.8. Since the immobile domain is still
estimated values of Da would increase linearly with increas- significant (bnoneq > 0.7) at this diffusion coefficient, dif-
ing diffusion coefficients (d0). fusive mass transfer plays an important role in this con-
4.2.4. Advective Mass Transfer ductivity field.
[58] Several results from the previous section indicate [62] We also found that increasing the velocity signifi-
that mass transfer is controlled in some cases by advection cantly increased the estimated dispersivity in the connected
ZINN AND HARVEY: FLOW, DISPERSION, AND MASS TRANSFER SBH 4 - 13

Figure 10. Estimated rate coefficients (a) as a function of the diffusion coefficient for the base
conditions compared with estimated rate coefficients where the velocity (hydraulic gradient) is doubled.
For zero diffusion the estimated rate coefficient doubled with a doubling of the velocity indicating
advective mass transfer.

field, but did not significantly affect the estimates of bnoneq equally spaced starting position. This may be due to the fact
and bnoncont (maximum change of 12%). For large diffusion that in the connected field, the flow channels are contiguous
coefficients (d0 = 3), the effective dispersivity increased by a throughout the entire domain. In the case where starting
factor of 3.2 for a doubling of the velocity. This increase positions are proportional to velocity, most particles start out
approaches the increase of a factor of 4 predicted for in fast moving streampaths, and in the absence of diffusion,
transport in layered systems by Gelhar et al. [1979] and stay in those paths. In the case with starting positions
Matheron and de Marsily [1980]. At lower diffusion coef- equidistant, the particles start out in and remain in a wider
ficients, we did not find as large an increase in the estimated variety of velocities, so the effective dispersivity is larger.
dispersion coefficient with velocity. We believe this is This explanation is supported by the fact that when a
because of the initial conditions that place particles only relatively small amount of diffusion is introduced (normal-
in high velocity paths, which is analyzed below. ized diffusion coefficient, d0, equal to 0.3), the difference in
4.2.5. Initial Conditions dispersivities drops to a factor of two, and decreases even
[63] In the simulations described so far, the spacing of further as more diffusion is added. Diffusion allows faster
particles introduced along the injection line was propor- mixing across flow paths, thereby negating the effects of the
tional to fluid flux, the approach of Desbarats [1990]. This starting positions. This explanation is also consistent with
starting condition was modeled with the mass transfer the result that the estimated fraction of the domain in which
model by setting the initial immobile concentration to zero. significant solute does not pass, bnoncont is significantly
We now consider an alternative initial positioning, equi- smaller with the initial conditions of evenly spaced particles.
distant placement along the vertical line. For this placement,
we change the initial conditions for the mass transfer model
to equal concentrations of immobile and mobile solute, i.e.,
equilibrium. It should be noted that Vanderborght and
Vereecken [1998], suggested that the flux proportional
method is generally preferable since it does not bias late-
time parameter estimates with early time anomalies. They
also suggested that this initial condition bias is removed
from the transport relatively quickly (a few correlation
lengths). However, they dealt with only multigaussian fields
at relatively low ln(K) variance (approximately 2.5), so
results in our fields may differ.
[64] The greatest effects on the breakthrough curve are seen
for a ln(K) variance of 9 for the connected field (Figure 11).
Little difference is seen at low variance between the different
starting conditions, or at high variance for the multigaussian
and disconnected fields, and the estimated parameters
change by less than 15%. However, for the connected field
with ln(K) variance of 9, the estimated dispersivity and Figure 11. Normalized breakthrough curves for the
bnoncont are significantly different between the two initial connected field with sY2 = 9 and a normalized diffusion
particle placements. At a diffusion coefficient of zero, the coefficient (d0) equal to 1, with initial particle positions
estimated dispersivity increased by a factor of three for the evenly spaced and proportional to velocity.
SBH 4 - 14 ZINN AND HARVEY: FLOW, DISPERSION, AND MASS TRANSFER

The initial equal spacing places particles in all parts of the initial bias in the initial velocities with this starting con-
domain, so at least initially, no region of the domain is dition. Also of note is that the multigaussian field does
outside of the path of particles. display some increase in bnoncont, but significantly less than
[65] It is important to note that the different parameter the connected field. This is consistent with our earlier
estimates for both bnoncont and the dispersivity do not finding that there is a small, but real, bnoncont in the high
indicate that different initial conditions drive fundamentally variance multigaussian field.
different behavior. Conductivity fields that are dominated
by mass transfer in one case are also dominated by it in the
other, while fields which are well-described by the advec- 5. Discussion
tive – dispersive model in the flux-proportional initial con- [70] Each of the three types of fields considered here has
dition will still be well-described by it when subjected to fundamentally different flow and transport behaviors.
equilibrium initial conditions. However, the differences in [71] In the connected field, the effective conductivity is
parameters indicate that the initial conditions can have some greater than the geometric mean, velocity variations are
influence on the evolution of the solute plume. much higher than the multigaussian equivalent, and solute
4.2.6. Evolution of Transport Parameters transport is subject to significant mass transfer. The high-
[66] Here we consider how estimated transport parameters conductivity zone forms a continuous network of paths
may change for breakthrough curves at different downstream through which fluid and solute move. Isolated low con-
distances. All of the simulations discussed thus far have dealt ductivity blobs are embedded in this network and form low
with the same travel distance (67 correlation lengths). Now velocity regions where solute is slowed or immobilized.
we consider estimated parameters 44 correlation lengths Flow around these blobs creates a large variance in the
downstream of the injection distance with the particles velocity perpendicular to the mean direction of flow, and the
started at the same location with the same initial starting sharp contrast between flow in the channels and the blobs
positions (both proportional to velocity and with equidistant creates a large variance in flow velocity in the mean
spacing). We ran the simulated flow and transport for all direction of flow. For relatively low contrast (sY2 below
three field types at the highest variance (ln(K) variance of 9), approximately 4), mass transfer effects are not evident and
where the largest differences, if any, should be found. transport can be accurately modeled by the advective-
[67] For the disconnected field, we found no significant dispersion equation, although mean plume velocity is still
change in any of the estimated parameters. For the multi- greater than in the multigaussian case. If the contrast
gaussian and connected fields, we found that the estimated between the high-conductivity paths and the low conduc-
dispersivities, rate coefficients, and nonequilibrium partition tivity blobs is larger (sY2 above approximately 4 but below
coefficients were all within close agreement (differences of approximately 8), then advection through low permeability
less than 10 percent) with the same parameters estimated at regions dominates tailing. This advective process creates
the further downstream breakthrough point. This held true large-scale behavior similar to diffusive mass transfer, but
for all values of the diffusion coefficient and for both initial the rates depend on head gradients rather than diffusion
conditions. coefficients. If the contrast between the high and low
[68] However, a larger noncontributing partition coeffi- conductivities becomes even larger, the size of the effec-
cient, bnoncont, was estimated from breakthrough curves tively immobile regions increases and the lowest-conduc-
recorded at 44 correlation lengths, as opposed to 67 corre- tivity regions become areas with no significant fluid flow. In
lation lengths, for both the connected and multigaussian this high variance case (sY2 above approximately 8), mass
fields when initial particle positions were proportional to transfer is driven by both diffusion and advection.
velocity. The largest increase was estimated for no diffu- [72] At higher variances (sY2 above approximately 6), the
sion—in the multigaussian field, bnoncont increased by 31 effective dispersion coefficient for the connected field also
percent, and in the connected field it increased by 77 shows characteristics similar to the effective dispersion
percent. The difference in estimated bnoncont decreased as coefficient in layered media [Matheron and de Marsily,
diffusion coefficient (d0) increased, until at the largest value 1980]. The effective dispersion coefficient decreases with
of diffusion coefficient (d0 = 3), the estimates for both cases an increasing diffusion coefficient, and the dispersion coef-
were nearly identical. When the starting positions of the ficient increases with velocity at a rate greater than linear,
particles were equally spaced, we found no substantial although not at the quadratic rate found for perfectly layered
difference in the estimated values of bnoncont. media. Thus, Matheron and de Marsily’s theory, which
[69] We attribute this increased estimate of bnoncont to the considered flow that was not perfectly aligned with strat-
same phenomena discussed in the previous section: the ification, appears to also have some application to flow
velocity-proportional initial conditions place the bulk of channels that are not straight.
the particles in the fastest parts of the mobile domain, and [73] Flow and transport in the connected field has many
the farther the particles travel, the more this initial bias is of the characteristics that are attributed to conductivity
averaged out. Thus the average velocity of the particles in fields with layering, bimodal histograms, or large integral
the mobile domain will be higher the closer the break- scales, described in the introduction. Yet, the connected
through point is to the starting point. This is consistent with field is isotropic, univariate lognormal, and has an integral
the result that higher diffusion coefficients decrease the scale much smaller than the domain length. The connected
difference, since diffusion mixes particles out of fast paths field has behaviors similar to a layered field because the
more quickly. It is also consistent with the lack of signifi- high-conductivity regions, although isotropic, form contig-
cant differences in the estimated values of bnoncont when the uous preferential channels for flow. It also may reproduce
particles are started equally spaced, as there is no significant some of the behavior of nonstationary fields (i.e., field with
ZINN AND HARVEY: FLOW, DISPERSION, AND MASS TRANSFER SBH 4 - 15

integral scales larger than the domain size) because the all have nearly identical lognormal univariate conductivity
high-conductivity structures span the entire domain. Finally, distributions and nearly identical isotropic spatial cova-
the connected field can reproduce behaviors, such as non- riance functions. Under many existing stochastic models,
equilibrium mass transfer, that are often attributed to fields these fields would be identical in their predicted upscaled
with bimodal distributions, such as low conductivity blobs flow and solute transport behavior. A secondary conclusion
embedded in a matrix of uniformly higher conductivity. is that significant rate-limited mass transfer may occur in
This is because the connected field also creates regions of smooth conductivity fields if the high values are well
low velocity embedded in channels of high velocity, even connected. This provides some theoretical basis for the
though the univariate distribution of ln(K) is Gaussian, and application of mass transfer models to transport of conser-
hence unimodal. vative solutes in sedimentary aquifers with relatively high
[74] In the disconnected field, the effective conductivity conductivity variance and connected high conductivities,
is less than the geometric mean, the velocity variance is such as the MADE site in Mississippi. Below, we list
small, and mass transfer behavior does not occur. The specific conclusions.
contiguous structure of low conductivity areas in this field 1. Spatial patterns of hydraulic conductivity can be
forces flow through low-conductivity regions, so there are constructed with connected structures of either high or low
few, if any, isolated low velocity regions where mass conductivity, that have the same univariate probability
transfer occurs. Flow lines are relatively straight, and the density function and isotropic covariance function as a
variance in fluid velocity is relatively low. Although the multigaussian field that does not have connected structures
effective conductivity and dispersivity values are lower in of extreme values. Although the integral scale of these
this disconnected field than in the equivalent multigaussian connected fields is much smaller than the domain size, the
field, the upscaled solute behavior in this type of field is connected structures of either low or high values span the
similar to classical macrodispersion in a multigaussian field. entire domain.
[75] Flow and transport characteristics in the multigaus- 2. Matheron’s conjecture that the effective conductivity
sian field agree well with existing theory and are well in a two-dimensional conductivity field is the geometric
modeled with the advective – dispersive equation. Thus, this mean does not apply to the connected conductivity fields
field serves primarily as a control for numerical experiments considered here, although they are isotropic with lognormal
in the connected and disconnected fields. Although the univariate conductivity distributions. The effective conduc-
multigaussian field shows modest amounts of mass transfer tivity is larger than the geometric mean for the field with
at high variance, this behavior is not significant compared to connected high values of conductivity, and smaller for the
the connected field. As with the disconnected case, there is field with connected low values.
no continuous high-velocity zone. Values close to the mean 3. Velocity variability, and hence the effective dispersion
form a connected structure, spanning the domain, which coefficient, is substantially higher in the conductivity field
isolates high-conductivity blobs. A small volume of low- with connected high values than in multigaussian field. The
conductivity blobs also exists and is of sufficient size and field with connected low values has substantially smaller
conductivity contrast to cause mass transfer, but this is a effective dispersion coefficients than the multigaussian
relatively minor part of the entire field. field.
[76] The finding that mass transfer behaviors occur in 4. Rate-limited mass transfer may be a significant
fields with an effective conductivity greater than the geo- process in conductivity fields with connected structures of
metric mean is consistent with the results of Guswa and high values. At variances in ln(K) above approximately 6,
Freyberg [2002]. They studied flow and transport through a the majority of the field considered here is best modeled as
homogeneous medium with elliptical inclusions (of varying immobile. A practical implication is that when the effective
shape, size, conductivity contrast, etc.) and also found that conductivity is significantly larger than often assumed, mass
mass transfer occurs when the mean effective conductivity transfer behavior may occur.
is greater than the geometric mean, but that transport is well 5. Rate-limited mass transfer in hydraulic conductivity
described by the advective – dispersive mode when the fields with connected high values can be driven by a
effective conductivity is at or below the geometric mean. complex interaction of advection and diffusion. For the
[77] There are a variety of issues we did not consider. cases considered here, mass transfer is primarily driven by
First, we have not studied the effective transverse dispersion advection below variances in ln(K) of approximately 8. For
of the solute movement through the various fields. Second, all values of conductivity variance, the largest total
we neglected pore-scale, or mechanical, dispersion, and immobile domains are found when no diffusion is
consider only small-scale mixing by diffusion. Third, we simulated. However, at higher variances, diffusion also
did not consider multiple rates of mass transfer for our plays a significant, and in some cases dominant, role. For
effective model. A multirate mass transfer model could higher values of the diffusion coefficient, the effective mass
improve the breakthrough curve fits. However, our simple transfer rate coefficients were reduced (contrary to the
model provides qualitatively good fits and the parameter notion that the rate of mass transfer is driven by the rate of
estimates give a clear quantitative characterization of the diffusion), and the total proportion of the domain found to
dominant transport processes in each scenario. be effectively immobile was also reduced.
6. Effective dispersion coefficients in the conductivity
field with connected high values had characteristics similar
6. Conclusions to those of Matheron and de Marsily’s [1980] dispersion
[78] Our general conclusion is that very different flow coefficients for transport through layers, which also have
and transport behaviors can occur in conductivity fields that continuous paths of high relative fluid velocity. The
SBH 4 - 16 ZINN AND HARVEY: FLOW, DISPERSION, AND MASS TRANSFER

effective dispersion coefficient generally decreased with an vmobile [L/T] effective fluid velocity in that part of the
increasing diffusion coefficient and increased with velocity conductivity field which is considered to be
at a rate greater than linear. mobile: vmobile = v(1 + btot)
[79] Most hydrogeologists will agree that large continu- x [L] distance along primary direction of flow
ous flow channels have large effects on groundwater flow x0 [ – ] distance normalized by the length of the
and solute transport. Here we demonstrate that significant domain: x0 = Lx
effects of flow channeling may occur in statistically iso- a [T1] mass transfer rate coefficient, i.e., the
tropic hydraulic conductivity fields that share the same effective rate at which solute moves in and
basic statistics as stationary multigaussian fields with out of the immobile domain
integral scales much less than the domain length. In bnoncont [ – ] portion of the immobile domain which
particular, we find that rate-limited mass transfer may be does not significantly interact with the solute:
a significant process during solute transport through a btot = bnoncont + bnoneq
hydraulic conductivity field with a texture of connected bnoneq [ – ] portion of the immobile domain which
high values, and that this mass transfer is driven by a causes nonequilibrium mass transfer beha-
complex interaction of both advection and diffusion. These vior: btot = bnoncont + bnoneq
results suggests that information on the connectedness of btot[ – ] total capacity coefficient:
geologic media may be necessary not only to choose btot = total volume of immobile domain
total volume of mobile domain
parameters for flow and transport models, but also to l [L] integral scale of the conductivity field
2
choose the form of the transport model. In conductivity sY [L2/T2] variance of the natural logarithm of hydraulic
fields with nearly identical conventional statistics, solute conductivity
spreading may be well modeled by Fickian dispersion, or a
mass transfer model may be required to adequately predict Appendix A: Derivation of the Covariance
tailing, and the mass transfer rate coefficient may depend Function for the Absolute Value Transform of a
on the hydraulic gradient, or on the apparent diffusion
Multigaussian Field
coefficient. Characterizing patterns of connectedness in the
field is an extremely difficult problem. The results here [80] Here we derive the spatial correlation function for
show that representing spatial heterogeneity as multigaus- the absolute value transform of a stationary multinormal
sian may not be a conservative assumption. field. We show that the length scale of heterogeneities in the
transformed field is less than that of the original field by
deriving an analytic relationship for the covariance of the
7. Notation transformed field as a function of the covariance of the
A [L] (macro)dispersivity original multigaussian field.
C [M/L3] solute concentration in the mobile domain
A1. Mean and Variance of the Transformed Field
d [L2/T] diffusion coefficient
d0 [ –] the ratio of the rate of diffusion in and out of [81] Let x be a normally distributed (m = 0, s2 = 1)
structures of length scale l to the rate of random variable, and y be the absolute value of x, y = jxj.
advection through the domain, or the time to Then the expected value (i.e., the mean) of y is:
advect through the domain to the time to Z 1  2
0 1 x
diffuse in and out of structures with a length m ¼ E½y ¼ jxj pffiffiffiffiffiffi exp dx
0 dL dt0 2p 2s2
scale of l: d ¼ vl2 ¼ l2 . This is an assigned 1

characteristic of each simulation, designed Z 1  2 rffiffiffi


1 x 2
to denote the magnitude of diffusion. ¼ 2 x pffiffiffiffiffi
ffi exp dx ¼ ðA1Þ
2p 2s2 p
D [L2/T] dispersion coefficient: D = A vmobile 0

Da [ –] the ratio of the estimated rate of mass


Similarly, the variance of y is:
transfer to the rate of advection through the
L 0
domain: Da = a v ¼ at This is an estimated Z 1  2
  1 x
parameter. s02 ¼ E y2  m02 ¼ jxj2 pffiffiffiffiffiffi exp dx  m02
1 2p 2s2
K [L/T] hydraulic conductivity
Keff[L/T] effective upscaled hydraulic conductivity of 2
¼ s2  m02 ¼ 1  ðA2Þ
a field of conductivity values p
L [L] the length of the domain (conductivity field)
from injection to breakthrough point since s2 = 1.
Pe [ –] ratio of the rate of dispersion to the rate of
advection over the correlation length scale A2. Covariance of the Transformed Field
Pe = Al [82] We would like to find the correlation function r0(h)
s [T1] the Laplace parameter for two transformed variables y1 and y2 at locations sepa-
s0 [ –] the normalized Laplace parameter: s0 = sLv rated by a distance h as a function of the correlation function
S [M/L3] solute concentration in the immobile domain r(h) for the untransformed variables x and x . By definition
1 2
t [T] time the correlation coefficient r0(h) is:
t0 [ –] time normalized by the mean time of fluid qffiffi
to advect across length scale of the domain: 02 E ½y y  2
0 E ½ y y
1 2  m 1 2 p
t = tLv r0 ðhÞ ¼ ¼ 2
ðA3Þ
v [L/T] mean upscaled fluid velocity s02 1p
ZINN AND HARVEY: FLOW, DISPERSION, AND MASS TRANSFER SBH 4 - 17

where the only difference is that the correlation coefficient


is negative (due to the absolute value transform affecting
only one of the two variables) by the relation:

Covða; bÞ ¼ Covða; bÞ ðA8Þ

The solution of the first integral may be written:


pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
rðhÞ 1  r2 ðhÞ rðhÞ arcsinðrðhÞÞ
I1 ¼ I4 ¼ ¼ þ ðA9Þ
4 2p 2p

and the second integral is:


pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
rðhÞ 1  r2 ðhÞ rðhÞ arcsinðrðhÞÞ
I2 ¼ I 3 ¼ ¼ þ ðA10Þ
4 2p 2p

Incorporating (A9) and (A10) into (A5), and substituting


into (A3) gives:
 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 
Figure A1. Spatial correlation after absolute value 2 1 þ 1  r2 ðhÞ þ rðhÞ arcsinðrðhÞÞ
0
transformation as a function of the original spatial r ðhÞ ¼ ðA11Þ
p2
correlation. The dashed line indicates the nontransformed
case (r = r0) for comparison. [83] Thus, the correlation coefficient for the transformed
variable y is a function of the correlation coefficient of the
We will find E[y1 y2] as a function of the cdf of the original variable at the same separation distance h, and is
multigaussian distribution of the untransformed variables x1 independent of the correlation at other separation distances.
and x2: Figure A1 shows that the correlation coefficient for the
transformed variable is always less than that of the untrans-
 2 
1 x  2x1 x2 rðhÞ þ x22 formed variable. Thus the connected and disconnected
f ðx1 ; x2 ; rðhÞÞ  pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi exp  1 fields have reduced spatial correlation, and hence reduced
2p 1  r2 ðhÞ 2ð1  r2 ðhÞÞ
integral scales.
ðA4Þ

To account for the absolute value, E[y1 y2] is written as the [84] Acknowledgments. Funding for this work was provided by the
sum of four integrals, which account for the four possible National Science Foundation (EAR-9875995), and by the Department of
Energy, Basic Energy Sciences division (DE-FG02-ooER15029). We would
sign combinations of x1 and x2 (both x1 and x2 positive, like to thank Roy Haggerty and Patricia Culligan and two anonymous
both negative, x1 positive and x2 negative, and x1 negative reviewers for their very helpful comments.
and x2 positive):
Z 0 Z 0
References
E½yy0 ¼ jx1 jjx2 j f ðx1 ; x2 ; rðhÞÞdx1 dx2 Anderson, M. P., Hydrogeologic facies models to delineate large-scale
1 1 spatial trends in glacial and glaciofluvial sediments, Geol. Soc. Am. Bull.,
101(4), 501 – 511, 1989.
Z 1 Z 0 Anderson, T., An Introduction to Multivariate Statistics, John Wiley, New
þ jx1 jjx2 j f ðx1 ; x2 ; rðhÞÞdx1 dx2 York, 1958.
0 1 Aris, R., On the dispersion of a solute in a fluid flowing through a tube,
Proc. R. Soc. London, Ser. A, 235, 67 – 77, 1956.
Z 0 Z 1 Bahr, J. M., and J. Rubin, Direct comparison of kinetic and local equili-
þ jx1 jjx2 j f ðx1 ; x2 ; rðhÞÞdx1 dx2 brium formulations for solute transport affected by surface reactions,
1 0 Water Resour. Res., 23(3), 438 – 452, 1987.
Barenblatt, G. E., I. P. Sheltov, and I. N. Kochina, Basic concept of the
Z 1 Z 1 theory of seepage of homogeneous liquids in fissured rocks, J. Appl.
þ jx1 jjx2 j f ðx1 ; x2 ; rðhÞÞdx1 dx2 ðA5Þ Math. Mech., 24, 1286 – 1303, 1960.
0 0 Cushman, J. H., (Ed.), Dynamics of Fluids in Hierarchical Porous Media,
Academic, San Diego, Calif., 1990.
The first and fourth integrals are the same and may be Dagan, G., Solute transport in heterogeneous porous formations, J. Fluid
written: Mech., 145, 151 – 177, 1984.
Dagan, G., Theory of Flow and Transport in Porous Formations, Springer-
Z 1 Z 1 Verlag, New York, 1989.
I1 ¼ I4 ¼ x1 x2 f ðx1 ; x2 ; rðhÞÞdx1 dx2 ðA6Þ Dagan, G., and S. C. Lessoff, Solute transport in heterogeneous formations
0 0 of bimodal conductivity distribution, 1, Theory; 2, Applications, Water
Resour. Res., 37(3), 465 – 480, 2001.
Since x1 and x2 are interchangeable, the second and third Dagan, G., V. D. Cvetkovic, and A. Shapiro, A solute flux approach in
integrals are also equal and may be written: transport in heterogenous formations, 1, The general framework, Water
Resour. Res., 28(5), 1369 – 1376, 1992.
Z 1 Z 1 De Hoog, F. R., J. H. Knight, and A. N. Stokes, An improved method for
I2 ¼ I3 ¼ x1 x2 f ðx1 ; x2 ; rðhÞÞdx1 dx2 ðA7Þ numerical inversion of Laplace transforms, J. Sci. Stat. Comput., 3(3),
0 0 357 – 366, 1982.
SBH 4 - 18 ZINN AND HARVEY: FLOW, DISPERSION, AND MASS TRANSFER

Deng, F.-W., and J. H. Cushman, Higher-order corrections to the flow Journel, A. G., Nonparametric-estimation of spatial distributions, Math.
velocity covariance tensor. Revisited, Water Resour. Res., 34(1), 103 – Geol., 15(3), 445 – 468, 1983.
106, 1998. Journel, A. G., and F. Alabert, Non-gaussian data expansion in the earth
Desbarats, A. J., Macrodispersion in sand – shale sequences, Water Resour. sciences, Terra Nova, (1), 123 – 134, 1989.
Res., 26(1), 153 – 163, 1990. Journel, A. G., and C. V. Deutch, Entropy and spatial disorder, Math. Geol.,
Deutch, C. V., and A. G. Journel, Geostatistical Software Library and 25(3), 329 – 355, 1993.
User’s Guide (Applied Geostatistics Series), Oxford Univ. Press, New Julian, H. E., J. M. Boggs, C. M. Zheng, and C. E. Feehley, Numerical
York, 1997. simulation of a natural gradient tracer experiment for the natural attenua-
Di Federico, V., S. P. Neuman, and D. M. Tartakovsky, Anisotropy, lacu- tion study: Flow and physical transport, Groundwater, 39(4), 534 – 545,
narity, and upscaled conductivity and its autocovariance in multiscale 2001.
random fields with truncated power variograms, Water Resour. Res., Koltermann, C. E., and S. M. Gorelick, Heterogeneity in sedimentary de-
35(10), 2891 – 2908, 1999. posits: A review of structure-imitating, process-imitating, and descriptive
Feehley, C. E., C. M. Zheng, and F. J. Molz, A dual-domain mass transfer approaches, Water Resour. Res., 32(9), 2617 – 2658, 1996.
approach for modeling solute transport in heterogeneous aquifers: Ap- Kreft, A., and A. Zuber, On the physical meaning of the dispersion equation
plication to the Macrodispersion Experiment (MADE) site, Water Resour. and its solutions for different initial and boundary conditions, Chem.
Res., 36(9), 2501 – 2515, 2000. Eng. Sci., 33, 1471 – 1480, 1978.
Freyberg, D. L., A natural gradient experiment on solute transport in a sand LaBolle, E. M., and G. E. Fogg, Role of molecular diffusion in contaminant
aquifer, 2, Spatial moments and the advection and dispersion of nonreac- migration and recovery in an alluvial aquifer system, Trans. Porous
tive tracers, Water Resour. Res., 22(13), 1986. Med., 42, 155 – 179, 2001.
Fogg, G., Groundwater flow and sand body interconnectedness in a thick, Matheron, G., Elements pour une theorie des milleuz poreux, Masson,
multiple-aquifer system, Water Resour. Res., 22(5), 679 – 694, 1986. Paris, 1965.
Gelhar, L. W., Stochastic Subsurface Hydrology, Prentice-Hall, Old Tappan, Matheron, G., and G. de Marsily, Is transport in porous media always
N. J., 1993. diffusive? A counterexample, Water Resour. Res., 16(5), 901 – 917,
Gelhar, L. W., and C. Axness, 3-Dimensional stochastic-analysis of macro- 1980.
dispersion in aquifers, Water Resour. Res., 19(1), 161 – 180, 1983. McKenna, S. A., L. C. Meigs, and R. Haggerty, Tracer tests in a fractured
Gelhar, L. W., A. L. Gutjahr, and R. L. Naff, Stochastic anslysis of mar- dolomite, 3, Double-porosity, multiple-rate mass transfer processes in
crodispersion in a stratified aquifer, Water Resour. Res., 15(6), 1387 – convergent flow tracer tests, Water Resour. Res., 37(5), 1143 – 1154,
1397, 1979. 2001.
Goltz, M. N., and P. Y. Roberts, 3-Dimensional solutions for solute trans- Meigs, L. C., and R. L. Beauheim, Tracer tests in a fractured dolomite, 1,
port in an infinite medium with mobile and immobile zones, Water Experimental design and observed tracer recoveries, Water Resour. Res.,
Resour. Res., 22(7), 1139 – 1148, 1986. 37(5), 1113 – 1128, 2001.
Goltz, M. N., and P. Y. Roberts, Using the method of moments to Parker, J. C., and M. T. V. Genuchten, Flux-averaged and volume averaged
analyze 3-dimensional diffusion-limited solute transport from temporal concentrations in continuum approaches to solute transport, Water Re-
and spatial perspectives, Water Resour. Res., 23(8), 1575 – 1585, 1987. sour. Res., 20(7), 866 – 872, 1984.
Gomez-Hernandez, J. J., and X.-H. Wen, To be or not to be multi-Gaussain? Pollock, D. W., Semianalytical computation of path lines for finite-differ-
A reflection on stochastic hydrogeology, Adv. Water Resour., 21(1), 47 – ence models, Groundwater, 26(6), 743 – 750, 1988.
61, 1998. Renard, P., and G. de Marsily, Calculating equivalent permeability: A re-
Goode, D. J., Particle velocity in block-centered finite difference ground- view, Adv. Water Resour., 20(516), 1997.
water flow models, Water Resour. Res., 26(5), 925 – 940, 1990. Silliman, S. E., and A. L. Wright, Stochastic analysis of paths of high
Griffioen, J. W., D. A. Barry, and J. Y. Parlange, Interpretation of two- hydraulic conductivity in porous media, Water Resour. Res., 24(11),
region model parameters, Water Resour. Res., 34(3), 373 – 384, 1998. 1901 – 1910, 1988.
Guswa, A. J., and D. L. Freyberg, Slow advection and diffusion through Taylor, G. I., Dispersion of soluble matter in solvent flowing slowly
low permeability inclusions, J. Contam. Hydrol., 46, 205 – 232, 2000. through a tube, Proc. R. Soc. London, Ser. A, 219, 186 – 203, 1953.
Guswa, A. J., and D. L. Freyberg, On using the equivalent conductivity to Tidwell, V. C., and J. L. Wilson, Permeability upscaling measured on a
characterize solute spreading in environments with low-permeability block of Berea Sandstone: Results and interpretation, Math. Geol., 31(7),
lenses, Water Resour. Res., 38(8), 1132, doi:10.1029/2001WR000528, 749 – 769, 1999.
2002. Vanderborght, J., and H. Vereecken, Solute transport in a heterogeneous soil
Haggerty, R., and S. M. Gorelick, Multiple-rate mass transfer for modeling for boundary and initial conditions: Evaluation of first-order approxima-
diffusion and surface reactions in media with pore-scale heterogeneity, tions, Water Resour. Res., 34(12), 3255 – 3270, 1998.
Water Resour. Res., 31(10), 2383 – 2400, 1995. Wen, X.-H., and J. J. Gomez-Hernandez, Numerical modeling of macro-
Harvey, C. F., and S. M. Gorelick, Temporal moment-generating equa- dispersion in heterogeneous media: A comparison of multi-Gaussian and
tions—Modeling transport and mass-transfer in heterogeneous aquifers, non-multi-Gaussian models, J. Contam. Hydrol., 30, 129 – 156, 1997.
Water Resour. Res., 31(8), 1895 – 1911, 1995. Western, A. W., G. Bloschl, and R. B. Grayson, Toward capturin g hydro-
Harvey, C. F., and S. M. Gorelick, Rate-limited mass transfer or macro- logically significant connectivity in spatial patterns, Water Resour. Res.,
dispersion: Which dominates plume evolution at the Macrodispersion 37(1), 83 – 97, 2001.
Experiment (MADE) site?, Water Resour. Res., 36(3), 637 – 650, 2000. Xiao, H., A description of the behavior of indicator variograms for a bi-
Hess, K. M., S. H. Wolf, and M. A. Celia, Large-scale natural gradient variate normal distribution, Master’s thesis, Stanford Univ., 1985.
tracer test in sand and gravel, Cape Cod, Massachusetts, 3, Hydraulic
conductivity variability and calculated macrodispersivities, Water Resour.
Res., 28(8), 2011 – 2027, 1992. 

Hollenbeck, K. J., C. F. Harvey, R. Haggerty, and C. J. Werth, A method for C. F. Harvey and B. Zinn, Ralph M. Parsons Laboratory, Department of
estimating distributions of mass transfer rate coefficients with application Civil and Environmental Engineering, Massachusetts Institute of Technol-
to purging and batch experiments, J. Contam. Hydrol., 37, 367 – 388, 1999. ogy, Cambridge, MA 02139, USA. (charvey@mit.edu)
ZINN AND HARVEY: FLOW, DISPERSION, AND MASS TRANSFER

Figure 1. Generation of a connected hydraulic conductivity field from a multigaussian field. (a)
Multinormal field generated by sequential Gaussian simulation. (b) The zero-mean field produced by
taking an absolute value transform followed by a normal-scores transform (equation (6)). The low
conductivity values are now the most connected. (c) The connected field generated by reversing the sign
of the values in field (b) and stretching the coordinate axis so that the field has the same correlation length
as field (a).

SBH 4-4

You might also like