Numerical method of moments for solute transport in a ... - Springer Link

2 downloads 0 Views 539KB Size Report
Sep 30, 2006 - numerical finite difference method is implemented to obtain the solutions ... method of moment for flow to conditionally simulate the distributions ...
Stoch Environ Res Risk Assess (2007) 21:665–682 DOI 10.1007/s00477-006-0078-x

ORIGINAL PAPER

Numerical method of moments for solute transport in a nonstationary flow field conditioned on hydraulic conductivity and head measurements Jichun Wu Æ Bill X. Hu

Published online: 30 September 2006  Springer-Verlag 2006

Abstract The unconditional stochastic studies on groundwater flow and solute transport in a nonstationary conductivity field show that the standard deviations of the hydraulic head and solute flux are very large in comparison with their mean values (Zhang et al. in Water Resour Res 36:2107–2120, 2000; Wu et al. in J Hydrol 275:208–228, 2003; Hu et al. in Adv Water Resour 26:513–531, 2003). In this study, we develop a numerical method of moments conditioning on measurements of hydraulic conductivity and head to reduce the variances of the head and the solute flux. A Lagrangian perturbation method is applied to develop the framework for solute transport in a nonstationary flow field. Since analytically derived moments equations are too complicated to solve analytically, a numerical finite difference method is implemented to obtain the solutions. Instead of using an unconditional conductivity field as an input to calculate groundwater velocity, we combine a geostatistical method and a method of moment for flow to conditionally simulate the distributions of head and velocity based on the measurements of hydraulic conductivity and head at some points. The developed theory is applied in several case studies to investigate the influences of the measurements of hydraulic conductivity and/or the hydraulic head on the variances of the predictive head J. Wu  B. X. Hu (&) Department of Earth Sciences, Nanjing University, Nanjing, China e-mail: [email protected] J. Wu  B. X. Hu Department of Geological Sciences, Florida State University, Tallahassee, FL 32306, USA

and the solute flux in nonstationary flow fields. The study results show that the conditional calculation will significantly reduce the head variance. Since the hydraulic head measurement points are treated as the interior boundary (Dirichlet boundary) conditions, conditioning on both the hydraulic conductivity and the head measurements is much better than conditioning only on conductivity measurements for reduction of head variance. However, for solute flux, variance reduction by the conditional study is not so significant. Keywords Stochastic analysis  Groundwater transport  Solute flux  Conditional simulation  Data of hydraulic conductivity and hydraulic head  Moment equation

1 Introduction It has been well recognized that natural media are generally heterogeneous and groundwater flow and solute transport through the media will be significantly influenced by the heterogeneity. However, owing to the current measurement technology and financial limitations, measurements of a medium’s parameters can be conducted only at limited points or locations for almost all environmental projects. This situation leads to uncertainties of the parameter spatial distributions, which precludes using traditional deterministic approaches to accurately predict flow and transport processes. Therefore, it has become quite common to approach the subsurface flow and transport problems stochastically. In the stochastic formalization,

123

666

uncertainty is represented by probability or by related quantities like statistical moments. Medium properties are treated as random space functions and their spatial variations can be described by spatial correlation functions. Stochastic perturbation approach is the most commonly used method in the subsurface stochastic hydrology literature to study flow and solute transport in heterogeneous media and many theories have been developed (Dagan 1989; Gelhar 1993; Cushman 1997; Zhang 2002). These theories explore the mechanisms of solute dispersion, and provide relationships between the dispersivity and the heterogeneity of hydraulic conductivity. However, most of the theories were developed based on restrictive assumptions, such as spatially stationary parameter distributions and the absence or negligibility of boundary influences. These assumptions greatly simplify the mathematical calculation, but limit applications of these theories to groundwater flow and solute transport in complicated subsurface environments. Spatial stationary distribution of hydraulic conductivity is probably the most important assumption made in many stochastic studies. This assumption can be violated if either mean value of the hydraulic conductivity varies spatially or its autocovariance depends on the actual locations. For example, distinct geological layers, zones, and facies may cause the hydraulic conductivity field of a given geological formation to be spatially nonstationary. A stationary stochastic process will become nonstationary after conditioning on measurement. Many previous studies on nonstationary flow and transport focus on the nonstationarity caused by conditioning. Conditional simulation has important implications for stochastic solute transport calculations because the calculation of solute concentration is based on conditional conductivity and head fields, obtained from site-specific measurements, rather than generic effective parameters, such as macrodispersivity. At any given site, the conditional mean of the solute concentration is the unbiased estimate of the actual concentration with minimum variance (Jazwinski 1970). The conditional concentration variance is a convenient measurement of the accuracy of the mean concentration estimate. Graham and McLaughlin (1989) applied a covariance moment method to derive conditional moments of solute concentration and velocity on measurements of the hydraulic conductivity, head, and concentration. The covariance moment method was later applied to groundwater flow and solute transport in a stationary medium with spatially random recharge Li and Graham (1998). Neuman (1993) applied a unified Eulerian–Lagrangian method to develop a theory for

123

Stoch Environ Res Risk Assess (2007) 21:665–682

conservative transport in a random velocity field with internal and external boundaries. The theory provides a framework for estimation of mean concentration and solute flux conditioned on local hydraulic measurements. This theory was later applied to study groundwater flow and solute transport conditioning on log-conductivity and/or hydraulic head (Zhang and Neuman 1995a–c). It is also shown from many other studies that it is theoretically possible (Neuman and Orr 1993; Neuman et al. 1996) and computationally feasible (Guadagnini and Neuman 1999a, b; Hernandez et al. 2002) to render optimum unbiased predictions of flow (head and velocity) in randomly heterogeneous media and to assess predictive uncertainty conditional on measurements of medium parameters. Stochastic study on solute transport in a medium with a nonstationary conductivity field is a relatively new research area. Rubin and Dagan (1992) applied a Lagrangian approach to derive the general framework for solute transport conditioning on conductivity measurements. A Monte Carlo simulation method was used to calculate the conditional trajectory covariances. Under the assumption of Gaussian distribution for solute distribution, the conditional trajectory covariances were related to one parcel’s accumulative distribution function through a control plane (CP) downstream of the solute source. Ezzedine and Rubin (1996) applied a geostatistical method to conditionally estimate the spatially distributed solute concentration, where the unconditional results are calculated through a stationary transport theory. Indelman and Rubin (1996) presented a Lagrangian theory for solute transport in nonstationary velocity fields and applied it to study solute transport in a medium with stationary conductivity covariance but a linear trend of the mean log permeability. Butera and Tanda (1999) extended the analytical transport theory to radiant flow system. Similarly, Destouni et al. (2001) applied an analytical approach to study the solute transport in a stationary conductivity field with unidirectional, nonuniform mean groundwater flow. All these studies used either Monte Carlo simulation method or modified analytical method to solve solute transport in conditional and/or nonstationary flow fields. Recently, a general framework for solute flux through a nonstationary medium has been developed (Zhang et al. 2000). The solute flux is defined as mass of solute per unit time and unit area. The solute flux is related to the flux-averaged concentration by dividing the former by the groundwater flux (Dagan et al.1992; Andricevic and Cvetkovic 1998). The flux-averaged concentration is consistent with common procedures for measuring concentrations in laboratory columns,

Stoch Environ Res Risk Assess (2007) 21:665–682

soils, and aquifers. The solute flux method is to predict the solute travel time and transverse displacement through a control plane located at a fixed distance downstream of the solute source. The solute flux statistics (mean and variance) are derived through a Lagrangian approach and are expressed in terms of the probability density functions (PDFs) of particle travel time and transverse displacement. These PDFs are evaluated through the first two moments of travel time and transverse displacement and the assumed distribution functions. In Zhang et al.’s (2000) study, the general framework for solute flux can be used for two or three-dimensional flow, but the method for the moments of solute travel time and transverse displacement are only suitable for a two-dimensional flow case. Hu et al. (2003) extended the numerical method of moments to study solute transport in three-dimensional nonstationary flow fields. Both Zhang et al.’s (2000) and Hu et al.’s (2002) study results show that for groundwater flow and solute transport through a nonstationary conductivity field, the standard deviations of the hydraulic head and solute flux are very large in comparison to their mean values. To reduce the uncertainties, conditioning the flow and transport calculations on field measurements is the only approach. In this study, we developed a conditional method of moments for groundwater flow and solute transport in a heterogeneous conductivity field. The measurements of hydraulic conductivity and head are used to reduce the variances of the head and the solute flux. Zhang et al.’s (2000) general framework for solute flux in a nonstationary flow field is adopted, but a new method is developed for the moments of solute transverse locations. Geostatistical and conditional methods are applied to study the hydraulic conductivity, hydraulic head, and groundwater velocity conditioning on the measurements of conductivity and head. The conditional study results of velocity covariances are then used as input data for solute transport calculation. A finite difference method is implemented to numerically solve the moment equations for flow and transport. The developed method is used to study the influences of the measured hydraulic conductivity data and/or the measured head data on the predictive uncertainty of the head and the solute flux prediction in nonstationary flow fields in several synthetic case studies.

667

is a Cartesian coordinate vector. Groundwater seepage velocity, V(x), satisfies the continuity equation,  (nV) = 0, and Darcy’s law, V = –(K/n) h, where n is the effective porosity, and h is the hydraulic head. The V(x) is considered to be nonstationary, caused by the medium nonstationarity and/or bounded domain. A solute of total mass M is released into the flow field at time t = 0, over the injection area A0 located at x = 0, either instantaneously or with a known release rate quantified by a rate function, /(t) [T–1]. We denote with q0(a) [M/L2 ] an areal density of injected solute mass at the location a 2A0. With Da denoting an elementary area at a, the mass particle, q0Da, is advected by the random spatially nonstationary groundwater velocityR field V. The total advected solute mass isM ¼ A0 q0 da: If the solute mass is uniformly distributed over A0, then q0 = M/A0. For t > 0, a solute plume is formed and advected downstream by the flow field toward a y-plane (perpendicular to x-coordinate), or control plane (CP), located at some distance from the source, through which the solute mass flux is to be predicted or measured. For nonreactive solute, integrating the solute flux for a single particle, Dq ” q0(a)da/(t – s)d(y – g), over the injection area A0, averaging over the sampling area A(y) centered at y, yields the solute mass flux component orthogonal to the CP at location x as (Dagan et al. 1992; Cvetkovic et al. 1992; Cvetkovic and Dagan 1994; Andricevic and Cvetkovic 1998; Zhang et al. 2000), Z Z 1 qðt; y; x; AÞ ¼ q0 ðaÞ/ðt  sÞdðy0  gÞdy0 da ð1Þ A A0

where s ” s(x; a) is the travel time of the advective particle from a to the CP at x, and g ” g(x; a) is the transverse locations of a particle passing through the CP. The quantities s and g in Eq. 1 are random variables and are functions of the underlying random velocity field. The expected value of q in Eq. 1 in the case of point sampling (i.e., A fi 0) can be expressed as (Dagan et al. 1992; Cvetkovic et al. 1992; Cvetkovic and Dagan 1994; Zhang et al. 2000), Z Z1 hqðt; xÞi ¼ q0 ðaÞ/ðt  sÞf1 ½sðx; aÞ A0

2 Stochastic formulation of solute mass flux We consider incompressible groundwater flow in a two-dimensional, heterogeneous aquifer with spatially variable hydraulic conductivity K(x), where x ” (x, y)

A

0

¼ t; gðx; aÞ ¼ ydsda

ð2Þ

where f1[s(x, a), g(x, a)] denotes the joint PDF of travel time s for a particle from a to reach x and the corresponding transverse displacement g. The variance of the solute flux for the point sampling is evaluated as

123

668

Stoch Environ Res Risk Assess (2007) 21:665–682

r2q ðt; xÞ ¼ hq2 ðt; xÞi  hqðt; xÞi2

ð3Þ

with Z Z Z1Z1 hq ðt; xÞi ¼ q0 ðaÞq0 ðbÞ/ðt  s1 Þ/ðt  s2 Þ 2

A0 A0

0

0

 f2 ½s1 ðx; aÞ ¼ t; g1 ðx; aÞ ¼ y; s2 ðx; bÞ ¼ t; g2 ðx; bÞ ¼ yds1 ds2 dadb

ð4Þ

where f2[s1(x, a), g1(x, a); s2(x, b), g2(x, b)] is the twoparticle joint PDF of travel time and transverse displacement. For the special case of stationary flow the PDFs do not depend on the absolute locations of the starting point a(ax, ay, az) and CP location, but depend on their relative distance. In stationary flows, s (x; a) and g(x; a) are uncorrelated, at least up to the first order in the variance of log-conductivity (Dagan et al. 1992), and thus f1[s, g] = f1[s ]f1[g]. This assumption was later extended to the two-particle joint PDF (Andricevic and Cvetkovic 1998), that is, f2[s1, g1; s2, g2] = f2[s1, g1]f2[s2, g2]. However, this assumption is not suitable to solute transport in a nonstationary flow (Zhang et al. 2000). The solute discharge is another quantity of interest defined as the total solute mass flux over the entire CP at x, Z Z Qðt; xÞ ¼ q0 ðaÞ/ðt  sÞdðy  gÞdady ð5Þ A0 CP

where y is a point in the CP. Its mean and variance are given as Z Z1 hQðt; xÞi ¼ q0 ðaÞ/ðt  sÞf1 ½sðx; aÞdsda A0

ð6Þ

0

r2Q ðt; xÞ ¼ hQ2 ðt; xÞi  hQðt; xÞi2

ð7Þ

Z Z Z1Z1 hQ ðt; xÞi ¼ q0 ðaÞq0 ðbÞ/ðt  s1 Þ/ðt  s2 Þ

and f2, or an infinite number of statistical moments. The approach used in this study is to evaluate a finite number of statistical moments and assume certain functions for f1 and f2. For solute transport in a stationary, log-normal conductivity field without boundary influence, the study results indicate that the travel time, s, satisfies a log-normal distribution and transverse displacement, g, as a normal distribution (Bellin et al. 1994; Cvetkovic et al. 1996; Andricevic and Cvetkovic 1998). These assumptions were recently applied to a typical nonstationary conductivity field (Zhang et al. 2000; Wu et al. 2003), a zonally stationary field, where a medium includes several distinct geological units, each of which has a stationary log-normal conductivity distribution, but its statistical moments are different from its neighboring units. Later, this assumption was further applied to a dual-scale heterogeneous medium (Hu et al. 2003), where the medium is composed of many small pieces of two or more materials, and the conductivity distribution within each piece is assumed to be lognormal. Based on the assumptions, the solute flux, mean, and variance, calculated from the method of moment, are consistent with the results by Monte Carlo simulation. In this study, we assumed the unconditional conductivity distribution is log-normal; a solute parcel’s distributions conditioning on the measurements of conductivity and head satisfy lognormal in the longitudinal direction and normal in the transverse direction. It should be pointed out that the assumptions are made only for a single parcel; the plume distributions in the longitudinal and transverse directions may not satisfy the assumptions since the convection speeds of the parcels may be different. Based on the assumptions, the PDFs can be evaluated from the first two moments of s(x; a) and g(x; a), including their joint moments. In the Lagrangian frame, s(x; a) and g(x; a) can be related to the velocity field through (Andricevic and Cvetkovic 1998; Zhang et al. 2000)

2

A0 A0 0

0

 f2 ½s1 ðx; aÞ; s2 ðx; bÞds1 ds2 dadb

ð8Þ

where f1[s(x, a)] is the marginal PDF of f1[s(x, a)], and f2 [s1(x, a); s2(x, b)] is the marginal PDF of f2[s1(x, a), g1(x, a); s2(x, b), g2(x, b)].

ds 1 ¼ ; dx V1 ðx; gÞ

123

ð9Þ

where Vi(x, g) (i = 1, 2) is the Lagrangian velocity. Since the Eulerian velocity V(x,y) is a random variable, so are the Lagrangian velocity V(x, g), travel time s(x; a), and transverse displacement g(x; a). In the integral form, the travel time can be expressed as

3 Joint probability density functions To evaluate the statistical moments of solute flux, one needs to know the one- and two-particle PDFs f1

dg V2 ðx; gÞ ¼ dx V1 ðx; gÞ

sðx ¼ L; aÞ ¼

ZL ax

dx V1 ½x; gðx; aÞ

ð10Þ

Stoch Environ Res Risk Assess (2007) 21:665–682

669

We decompose the Eulerian velocity as V(x,y) = U(x,y) + u(x,y), where U is the ensemble mean velocity, and u is a zero-mean velocity fluctuation. For the Lagrangian velocity V[x, g(x; a)], both the particle transverse position g and the velocity V at this location are random variables. Zhang et al. (2000) expanded the Lagrangian velocity around its mean path [x, Æg(x; a) æ] in a Taylor series, Vi ðx; gÞ ¼ Ui ðx; hgiÞ þ ui ðx; hgiÞ  @Ui ðx; gÞ þ    ði ¼ 1; 2Þ þ g0 @g 

ð11Þ

g¼hgi

where g¢ = g – Ægæ with Æg¢æ ” 0. It should be noted that the mean velocity in the nonstationary field is generally not constant. Substituting Eq. 11 into 10, one has, to the first order, ZL

hsðL; aÞi ¼

ax

s0 ðL;aÞ ¼ 

ax

U2 ðx;hgiÞ 1 whereAðx; hgiÞ ¼  U 2 ðx;hgiÞ ; Bðx; hgiÞ ¼ U ðx;hgiÞ and 1 1

 @U2 ðx;gÞ  Cðx;hgiÞ ¼ g¼hgi @g  U2 ðx;hgiÞ @U1 ðx;gÞ   g¼hgi =U1 ðx;hgiÞ: U1 ðx;hgiÞ @g

ð17Þ

Multiplying (16) by g¢(x2; b) and taking ensemble, we obtain dhg0 ðx1 ; aÞg0 ðx2 ; bÞi ¼ Aðx1 ; hg1 iÞhu1 ðx1 ; hg1 iÞg0 ðx2 ; bÞi dx1 þ Bðx1 ; hg1 iÞhu2 ðx1 ; hg1 iÞg0 ðx2 ; bÞi þ Cðx1 ; hg1 iÞhg0 ðx1 ; aÞg0 ðx2 ; bÞi

dx U1 ðx; hgiÞ

ð17aÞ

ð12Þ

with the initial condition (here we assume the initial positions of parcels are deterministic)

and ZL

dg0 ðx; aÞ ¼ Aðx; hgiÞu1 ðx; hgiÞ þ Bðx; hgiÞu2 ðx; hgiÞ dx ð16Þ þ Cðx; hgiÞg0 ðx; aÞ

hg0 ðx1 ; aÞg0 ðx2 ; bÞij lx ¼0 or x 1

1 ½u1 ðx;hgiÞþb1 ðx;hgiÞg0 dx 2 U1 ðx;hgiÞ

ð13Þ

where b1 ðx;hgiÞ ¼ f½@U1 ðx;yÞ=@ygjy¼hgi : If the flow field is stationary, b1 will be zero, and (13) returns to the Dagan’s results (1984, 1992b). From (13), we can obtain the covariance of Æs(L; a) æ as

rs1 s2 ðL; a; L; bÞ ¼

ZL ZL ax bx

dx1 dx2 2 U1 ðx1 ; hg1 iÞU12 ðx2 ; hg2 iÞ

 ½hu1 ðx1 ; hg1 iÞu1 ðx2 ; hg2 iÞi

þb1 ðx1 ; hg1 iÞb1 ðx2 ; hg2 iÞhg01 g02 i



ð17bÞ

Zx1

 dv hu1 ðv; hgiÞg0 ðx2 ; bÞi U12 ðv; hgiÞ ax  þ b1 ðv; hgi; hniÞhg0 ðv; aÞg0 ðx2 ; bÞi ð18Þ

ð14Þ

Zhang et al. (2000) used the integral form to obtain the mean and covariance of g¢. Here, we use a differential equation form, which is more general, to obtain the expressions of Æg(x; a)æ and various correlation functions related to g¢. The statistical moments of g, to the first order, are obtained as dhgðx; aÞi U2 ðx; hgiÞ ¼ dx U1 ðx; hgiÞ

0

hs ðx1 ; aÞg ðx2 ; bÞi ¼ 

þb1 ðx2 ; hg2 iÞhu1 ðx1 ; hg1 iÞg02 i

¼0

Similarly, we can also obtain the expressions of Æu1 g¢æ and Æu2 g¢æ. The velocity correlations, Æui ujæ (i = 1, 2; j = 1, 2), can be obtained from the flow equations and are treated as known quantities or input data here. Therefore, there are three differential equations for the three unknowns, Æg¢g¢æ, Æu1 g¢æ, and Æu2 g¢æ, so the equations are solvable. Owing to the complexity of these equations, we solve them numerically. From (13), the joint moment Æs¢(x1; a)g¢(x2; b)æ is obtained as 0

þb1 ðx1 ; hg1 iÞhu1 ðx2 ; hg2 iÞg01 i

2 ¼0

ð15Þ

We can obtain r2s (L, a) and r2g(L, a) by letting b = a in Eqs. 14 and 17. As pointed out by Zhang et al. (2000), these expressions are derived under the condition that the coefficient of variation of velocity is smaller than 1. This condition may be satisfied for many practical subsurface flows where the variance of log hydraulic conductivity is moderately large. If s and g obey log-normal and normal distributions, respectively, their joint PDF for one particle can be expressed as

123

670

Stoch Environ Res Risk Assess (2007) 21:665–682

f1 ½sðx; aÞ; gðx; aÞ ¼

which are then used as the input data for transport calculations. Since the moment equations used here are similar to Zhang and Winter’s (1999) work, the expressions will not be given in this article.

1

pffiffiffiffiffiffiffiffiffiffiffiffiffi exp 2psrg ðx; aÞrln s ðx; aÞ 1  r2 ( " 1 ðg  hgðx; aÞiÞ2   2ð1  r2 Þ r2g

ðg  hgðx; aÞiÞðln s  hln sðx; aÞiÞ rg ðx; aÞrln s ðx; aÞ #) ðln s  hln sðx; aÞiÞ2 þ ð19Þ r2ln s ðx; aÞ

2r

where r2ln s ðx; aÞ ¼ ln½r2s ðx; aÞ þ hsðx; aÞi2   2 lnhsðx; aÞi; h ln sðx; aÞi ¼ 2 ln hs ðx; aÞi  12 ln½r2s ðx; aÞ þ hsðx; aÞi2 ; and   r ¼ hs0 ðx; aÞg0 ðx; aÞi= rg ðx; aÞrln s ðx; aÞhsðx; aÞi : Similarly, the joint two-particle PDF of (s 1, g1) and (s 2, g2) can be expressed as (Zhang et al. 2000) f2 ½s1 ðx; aÞ; g1 ðx; aÞ; s2 ðx; bÞ; g2 ðx; bÞ   1 1 T 1 ¼ exp  X R X 2 ð2pÞ2 s1 s2 jRj1=2

ð20Þ

where X = [(ln s1 – Æln s1æ), (g1 – Æg1æ), (ln s2 – Æln s2æ), (g2 – Æg2æ)]T and S is the covariance matrix given as 0

r2ln s1 B r B ln s1 g1 R¼B @ rln s1 ln s2 rln s1 g2

rln s1 g1 r2g1 rln s2 g1 rg1 g2

rln s1 ln s2 rln s2 g1 r2ln s2 rln s2 g2

where hln si i ¼ 2 lnhsi i  12 ln½r2si þ hsi i2   2 lnhsi i;rln si gj ¼ hs0i g0j i=hsi i; ln½hsi ihsj i þ hs0i s0j i  ln½hsi ihsj i:

1 rln s1 g2 rg1 g2 C C C rln s2 g2 A r2g2 2

; r2ln si

hsi i and

ð21Þ

Spatial nonstationarity of groundwater flow may be generated from the presence of nonstationary medium features (e.g., distinct geologic layers, zones, or faces), the presence of finite boundaries, and/or the fluid pumping and injecting. Here, we consider flow nonstationarity stemming from the combined effects of the nonstationarity in the hydraulic conductivity and the finite boundary. The nonstationarity in the conductivity field may manifest in three ways: (1) the mean ÆY æ may vary spatially, (2) the two-point covariance CY(x, v) may depend on the actual locations of x and v rather than only their separation distance, and (3) the nonstationarity may arise from a conditioning study even the conductivity field is assumed to be stationary before the conditioning. The conditional mean and covariance of a nonstationary hydraulic conductivity field, Y(x), for given n local measurements, Y(xi), at x1, x2, ..., xn are (e.g., Zhang 2002) hYðxÞic ¼ hYðxÞi þ

n X

ai ðxÞ½Yðxi Þ  hYðxÞi

ð22Þ

i¼1

ln½r2si þ

¼ rln si ln sj ¼

As shown above, the calculations of the various moments of travel time and transverse location require the mean velocity and velocity covariance as input data. Osnes (1998) used semi-analytical methods to obtain velocity covariances in a two-dimensional, bounded domain. James and Graham (1999) developed a numerical method for accurately approximating head and flux covariances and cross-covariances in a finite two- and three-dimensional domain using the mixed finite element method. Zhang (1998) and Zhang and Winter (1999) also developed a numerical moment for groundwater flow in a stationary conductivity field in a bounded domain. In all these studies, the perturbation method is applied, and the results are of firstorder accuracy in terms of the variance of log hydraulic conductivity. In this study, Zhang and Winter’s (1999) method is applied to obtain the velocity moments,

123

4 Conditional statistics of a nonstationary conductivity field

CYc ðx; vÞ ¼ CY ðx; vÞ 

n X

ai ðxÞCY ðxi ; vÞ

ð23Þ

i¼1

where the moments with the superscript ‘‘c’’ are conditional ones, and those without this superscript are the unconditional counterparts. The coefficients ai are solutions of the following equations: n X

ai ðxÞCY ðxi ; xj Þ ¼ CY ðx; xj Þ for

16j6n

ð24Þ

i¼1

One can see that the conditional moments are generally nonstationary even if their unconditional counterparts are. In this study, the method is used to generate conditional conductivity covariance functions for stationary and nonstationary conductivity fields. The calculation results are then used as the input data for the flow and solute transport calculation. In general, one expects the reduction of the conditional permeability variance renders the deduction of the variances of hydraulic head and solute flux.

Stoch Environ Res Risk Assess (2007) 21:665–682

671

H1=10

Using method of moment to calculate unconditional mean and variance of hydraulic head in a random conductivity field has been widely studied (Dagan 1989; Gelhar 1993; Zhang 2002). For the unconditional stochastic method, actual measurements are not directly utilized in the obtaining the moments of head, although they may be used to infer the statistical moments of conductivity. To provide site-specific estimates and reduce estimation uncertainty of hydraulic head, conditional study on hydraulic head and groundwater flow has been extensively conducted in the last two decades (Dagan 1984, 1989; Dagan et al. 1992; Graham and McLaughlin 1989; Rubin and Dagan 1992; Neuman and Orr 1993; Neuman 1993; Zhang and Neuman 1995a–c; Ezzedine and Rubin 1996; Guadagnini and Neuman 1999a, b). The measurements may enter into the stochastic conditional solution at different stages. Here, we use the conditional conductivity statistics calculated from Eqs. 22–24 as input data to solve the

H2=0 q1=0

CP

10 9 8 7 6

Y

Central Point

5 4

Source location

3 2 1 0

0

2

q2=0

4

6

8

10

x

Fig. 1 Sketch of the study domain

Fig. 2 The distribution of the variances of a hydraulic head and b velocity, along the central line x2 = 5.0 m for cases 1–3

0.4

a

Case 1 Case 2 Case 3

0.3

2

σ h 0.2

0.1

0.0 0

2

4

x

6

8

10

0.4

b

Case 1 Case 2 Case 3

0.3

2 σ V1 0.2

0.1

0.0 0

2

4

x

6

8

10

123

672

Stoch Environ Res Risk Assess (2007) 21:665–682

flow equation for hydraulic head. Since the conditional conductivity field is no longer stationary, in this study Zhang and Winter’s (1999) nonstationary method is applied to calculate the conditional mean and variance of hydraulic head. The head measurements are treated as interior boundary conditions (Dirichlet boundary) at the measured points, which means that the head values at the measurement points are deterministic prescribed, i.e., the measured head values are assumed to be the mean values at the measured points and the head variances at the measured points are zero. Therefore, when conditioning on both hydraulic conductivity and hydraulic head at same measurement point, the variance of hydraulic conductivity (r2Y), the valance of hydraulic head (r2h), and the cross covariance between hydraulic conductivity and hydraulic head (CYh) are all equal to zero.

Fig. 3 The breakthrough curves of a expected value ÆQ æ and b standard deviation rQ of total solute flux through the CP for cases 1–3

5 Illustrative examples In the previous sections, the theoretical framework has been developed for the estimation of solute flux in a nonstationary flow field. The main required data for the calculation are the statistical moments of the conductivity field. In this section, we use several case studies to show the influences of the measurements of hydraulic conductivity and head on the means and variances of head and solute flux predictions. 5.1 Transport in a nonstationary flow field with uniform hydraulic gradient In this case study, the study domain is 10 · 10 m2 (Fig. 1). The unconditional conductivity field of the medium is statistically homogeneous. The mean and variance of the log-conductivity field are ÆYæ = 0 and

0.5

a

Case 1 Case 2

0.4

Case 3

0.3 Q 0.2

0.1

0.0 0

5

10

15

20

25

t 0.5

b

Case 1 Case 2

0.4

Case 3

0.3 σQ

0.2

0.1

0.0 0

123

5

10

15 t

20

25

30

Stoch Environ Res Risk Assess (2007) 21:665–682

673

r2Y= 0.4, respectively. Although the covariance of Y may take any other form, here we consider, for simplicity, only the exponential function,  h i1=2  CY ðx  vÞ ¼ r2Y exp  ðxi  vi Þ2 =k2i

H = 2 m centered at the point (0.5, 5.0 m), and the CP is located at x1 = 9.4 m. To focus on the influence of conditional study on the variances of head, velocity and solute flux, we consider a special case for the measured values of hydraulic conductivity and head in this subsection: the measured values at points are the same as their expectations (or unconditional mean values). Based on these measured values, the mean conductivity field and mean hydraulic head distribution will not change by conditioning, so the mean hydraulic gradient is uniform in the study domain. In this subsection, we investigate the influences on variances of head, velocity, and solute flux by conditioning on measurements of conductivity and head individually and jointly and at different points. Several cases are proposed for the purpose of comparison. Case 1 is the unconditional case: no measurements in

ð25Þ

where ki (i = 1,2) is the correlation length along the xi direction. For the sake of simplicity, here we assume k1 = k2 = k = 1.0 m. We impose constant heads on the vertical sides of the domain, H = 10.0 m on the left side and H = 0.0 m on the right side, and assume the other two sides are impermeable boundaries. A finite difference method is implemented for the various calculations for flow and transport. The whole domain is uniformly discretized into 40 · 40 square elements with a cell size of 0.25 · 0.25 m2. The line source with unit mass is of size

Fig. 4 The distribution of the variances of a hydraulic head and b velocity, along the central line x2 = 5.0 m for cases 3–6

0.2

a

Case 3 Case 4 Case 5 Case 6

σ h2 0.1

0.0 0

3

4

6

8

10

x 0.4

b

Case 3 Case 4 Case 5

0.3

Case 6

2

σ V1 0.2

0.1

0.0 0

2

4

6

8

10

x

123

674

Stoch Environ Res Risk Assess (2007) 21:665–682

Fig. 5 The breakthrough curves of a expected value ÆQæ and b standard deviation rQ of total solute flux through the CP for cases 3–6

0.7

a

Case 3

0.6

Case 4 Case 5 Case 6

0.5 0.4

Q 0.3 0.2 0.1 0.0 0

5

10

15

20

25

t 0.5

b

Case 3 Case 4

0.4

Case 5 Case 6

0.3 σQ

0.2

0.1

0.0 0

5

10

15

20

25

t

the simulation domain. The flow nonstationarity is resulted only from the finite boundaries. The results of this case are used as references for the other cases. In cases 2 and 3, we consider the measurements only at one point, the domain’s central point (x1 = 5.0 m, x2 = 5.0 m). We only have conductivity measurement in case 2 and have both the measurements of conductivity and head in case 3. In cases 4–6, there are measurements of both conductivity and head at 4, 9 and 19 points, respectively. These measurement points in each case are uniformly distributed along the central line, x2 = 5.0 m. Since the measured values of hydraulic head in any points are chosen to be the same as the expected values in the unconditional case, the distribution of the mean head and the mean velocity are the same for all six cases, and will not be presented here. Figure 2a and b presents the variances of hydraulic head and velocity in the longitudinal direction, r 2h andr2v1 ; along the central line x2 = 5.0 m for cases 1–3, respectively. It is shown

123

from Fig. 2a that r 2h does not decrease significantly by conditioning on conductivity measurement, but dramatically decreases by conditioning on both conductivity and head measurements. When conditioning on conductivity measurements, only r 2Y becomes zero at the measurement point; the decrease of r 2h is only caused by the decrease of r 2Y in the whole domain. On the other hand, when conditioning on both conductivity and head measurements, both r 2Y and r 2h become Table 1 The assumed measured data at the central point of the study domain Case number

Assumed measured data

7 8 9 10 11 12

Yc Yc Yc Yc Yc Yc

= = = = = =

0.5 – 0.5 0.5, Hc = 5.3 m 0.5, Hc = 4.7 m – 0.5, Hc = 5.3 m – 0.5, Hc = 4.7 m

Stoch Environ Res Risk Assess (2007) 21:665–682

675

Figure 3a and b shows the breakthrough curves of ÆQæ and r2Q through the CP for cases 1–3, respectively. It is shown from Fig. 3a and b that the conditional calculation decreases the solute dispersion with a higher peak value of ÆQ æ and also slightly decreases r2Q. When conditioning on the measured data of hydraulic conductivity, the r2Y decreased in the whole domain, therefore the solute dispersion decreased. Further, the solute results in cases 2 and 3 are the same, which means that conditional study on head measurement will not influence transport calculation. For these two cases, the conditional conductivity field and the mean hydraulic gradient are the same, so the movement and dispersion of the solute are very similar. In the above case studies, there is only one measurement point. Next, we investigate the influence of the number of measurement points on flow and transport predictions. We assume there are 4, 9 and 19 measurement points for cases 4–6, respectively, and the measurement points are uniformly

zero at the measurement point, which renders a dramatical decrease of r 2h in the whole domain. Figure 2b shows thatr2v1 decreases significantly near the measurement point by conditioning on conductivity only, because ther2v1 is greatly effected by the r 2Y. However,r2v1 does not decrease to zero even at the measurement point, because CYh and r2h are not zero at the measured point. Interestingly, by conditioning on both conductivity and head in case 3,r2v1 decreases in all the study area much less than that in case 2 except at the measurement point, wherer2v1 reaches zero since r2Y, CYh, and r2h all become zero. Further, in comparison with the unconditional case,r2v1 slightly increases one correlation length away from the measurement point. Figure 2b also shows thatr2v1 does not become zero at the left and right described head boundaries, because both the hydraulic conductivity and hydraulic gradient in the mean flow direction are not deterministic. Further, the finite difference method adopted in the calculation will also affect the result. Fig. 6 The distribution of the mean values of a hydraulic head and b velocity, along the central line x2 = 5.0 m for cases 1, 2, 7 and 8

10.0

a

Case 1 Case 2

8.0

Case 7 Case 8

6.0

h 4.0

2.0

0.0 0

2

4

6

8

10

x 2.0

b

Case 1 Case 2 Case 7 Case 8

1.5

V1 1.0

0.5

0.0 0

2

4

x

6

8

10

123

676

Stoch Environ Res Risk Assess (2007) 21:665–682

Fig. 7 The distribution of the variances of a hydraulic head and b velocity, along the central line x2 = 5.0 m for cases 1, 2, 7 and 8

0.4

a

Case1 Case 2 Case 7

0.3

Case 8

σ h2 0.2

0.1

0.0 0

2

4

6

8

10

x 0.4

b

Case1 Case 2 Case 7

0.3

Case 8

σ V2 0.2 1

0.1

0.0 0

2

4

6

8

10

x

distributed along the central line, x2 = 5.0 m. At these points, both conductivity and head are measured, and the measured conductivity and head values are all assumed to be the same as the unconditional, expected values. Figure 4a and b presents the distributions of r 2h andr2v1 along the central line x2 = 5.0 m for cases 3–6, respectively. It is shown from Fig. 4 that r 2h andr2v1 decrease to zero at all the measurement points, and r2h andr2v1 at other points decrease dramatically with the increase of measurement points since the increase of the measurement points will decrease the uncertainties of conductivity and head distributions, which will result in the decrease of the velocity uncertainty. Figure 5a and b shows the breakthrough curves of ÆQ æ and r2Q for cases 3–6, respectively. From the figures, one can see that with the increase of the number of measurement points, the much more reduced r2Y, r2h, and r2v1 in the whole domain result in the increase of

123

ÆQæ’s peak value, which means the decease of mean solute dispersion, and decrease of r2Q. Therefore, the increase of measurements directly reduces the uncertainties of spatial conductivity and head distributions, which in turn changes the mean breakthrough curve and decreases the prediction uncertainty. It should be pointed that the conditional study will generate to the non-distribution of the conductivity field, which will lead to the shifts of mean concentration and concentration variance peaks. 5.2 Transport in a nonstationary flow field with a nonuniform hydraulic gradient In the last subsection, we assume the measured values of conductivity and head are the same as their unconditional expectation values at the measurement points. However, the measurement values of head and conductivity at a specific point are generally not the same

Stoch Environ Res Risk Assess (2007) 21:665–682

677

Fig. 8 The breakthrough curves of a expected value ÆQæ and b standard deviation rQ of total solute flux through the CP for cases 1, 2, 7 and 8

0.5

a Case 1 Case 2

0.4

Case 7 Case 8

0.3

Q 0.2

0.1

0.0 0

5

10

15

20

25

t 0.5

b Case 1 Case 2

0.4

Case 7 Case 8

0.3

σQ 0.2

0.1

0.0 0

5

10

15

20

25

30

t

as their expectation values. In this subsection, we consider a more general case of nonstationarity, nonstationary covariance function of log-conductivity with non-uniform distributions of mean log-conductivity, and mean gradient hydraulic head due to the conditional study. The study domain, hydraulic boundary conditions, sources size, and location, and CP location are all the same as the case in the last subsection. For the sake of illustration, in the following case studies, we consider only one measurement point located at the central point of the study domain (e.g., x1 = 5.0, x2 = 5.0). To investigate the influences of the measured head and conductivity values at the point as well as their combinations on flow and transport, six additional cases, case 7–12, are proposed in this subsection. The measured parameter values of the six cases are shown in Table 1. Figure 6a and b presents the distributions of Æh æ and ÆV1æ along the central line x2 = 5.0 m for cases 1, 2, 7

and 8, respectively. It is shown from Fig. 6a and b that by conditioning on different measured conductivity values, the distribution of Æh æ changes slightly along the central line, but the mean velocity changes dramatically along the line, especially at the measured point, because ÆY æ will have a peak (or sink) value at the measured point after the conditional study, which will significantly change the mean velocity at the point. Figure 7a and b presents the distribution of the variances of hydraulic head and velocity along the central line x2 = 5.0 m for cases 1, 2, 7 and 8, respectively. Figure 7 shows that the distribution of head variance is slightly influenced by conditioning on different measured conductivity values at the central point, whereas the velocity variance is significantly affected in the whole line. Conditioning on the bigger value of hydraulic conductivity results in the bigger velocity variances. The distributions of the variances of hydraulic head and velocity are all symmetrical to the measured point.

123

678

Stoch Environ Res Risk Assess (2007) 21:665–682

Fig. 9 The distribution of the mean values of a hydraulic head and b velocity, along the central line x2 = 5.0 m for cases 3 and 9–12

10.0

a

Case 3 Case 9 Case 10

8.0

Case 11 Case 12

6.0

h 4.0

2.0

0.0 0

2

4

6

8

10

x

2.0

b

Case 3 Case 9 Case 10

1.5

Case 11 Case 12

V1 1.0

0.5

0.0 0

2

4

6

8

10

x

Figure 8a and b shows the breakthrough curves of expected values and variances of Q through the CP for cases 1, 2, 7 and 8, respectively. Figure 8 shows that conditioning on different values of measured conductivity at the same point will significantly affect the mean and variance of solute flux. Conditioning on a higher value of hydraulic conductivity renders the earlier arrival time, the higher peak value and the smaller solute dispersion, and vice versa. This is because conditioning on a higher value of hydraulic conductivity renders a higher mean velocity, which results in faster movement and smaller dispersion of the solute. Comparing the values of ÆQ æ and rQ, one can also see that the relative uncertainty, rQ/ÆQ æ, gets its smallest value in case 8 and gets its largest value in case 1. Cases 3, 9, 10, 11 and 12 are grouped together to study the influence of the various combinations of the

123

measured conductivity and head values on Æhæ, ÆV1æ, r2h,r2v1 ; ÆQæ and rQ. Figure 9a and b presents the distributions of Æh æ and ÆV1æ along the central line x2 = 5.0 m for the five cases, respectively. It is shown from the figures that the joint effect of different values of measured hydraulic conductivity and head at the central point slightly affects the distribution of Æh æ, but dramatically changes the distribution of ÆV1æ, especially in the vicinity of the measurement point. Further, the mean velocity distribution is unsymmetrical for cases 9–12. Figure 10a and b shows the distribution of r2h and r2v1 along the central line x2 = 5.0 m for the five cases. From the figures, one can see that both r2h andr2v1 vary significantly with the variations of measured hydraulic conductivity and head at the central point. The distributions of r2h andr2v1 are not symmetrical in cases 9–12. Figure 11a and b presents the breakthrough curves of

Stoch Environ Res Risk Assess (2007) 21:665–682

679

ÆQæ and rQ for the five cases. It is shown from the figures that jointly conditioning on different values of measured hydraulic conductivity and head at the central point significantly affects the solute transport process. Comparing the values of ÆQ æ and rQ in Fig. 10a and b, one can find that rQ/ÆQ æ in case 9 (or 11) is smaller than that in case 10 (or 12), which means the higher the measured head value at the central point, the lower the relative uncertainty. Similarly, comparing the results of case 9 (or 10) with case 11 (or 12), one can also find the higher the measured conductivity value at the central point, the lower the relative uncertainty.

solute flux through a control plane at a specific distance downstream of a contaminant source. The measurements of hydraulic conductivity and head were used for the conditional flow calculation, and the calculation results were then served as the input data for the solute transport prediction. The purpose of the conditional study is to reduce the variance or uncertainty around the solute prediction, which is based on mean solute flux calculation. Conditional calculation and/or limited boundary hydraulic condition will make a flow field spatially nonstationary even if the conductivity distribution in the medium is stationary. To study solute transport in a nonstationary field, Zhang et al.’s 2000 general framework for solute flux in a nonstationary flow field was adopted, but a new method was developed for the calculation of solute transverse moments. Based on the unconditional distribution of the log-conductivity and the measurements of conductivity, a geostatistical method was applied to conditionally study the

6 Summary and discussion In this study, a conditional transport theory was developed for solute flux in a nonstationary flow field. The theory is to predict the mean and variance of

Fig. 10 The distribution of the variances of a hydraulic head and b velocity, along the central line x2 = 5.0 m for cases 3 and 9–12

0.3

a

Case 3 Case 9 Case 10 Case 11 Case 12

0.2

σ h2 0.1

0.0 0

2

4

6

8

10

x 0.4

b

Case 3 Case 9 Case 10 Case 11

0.3

Case 12

σ V2 0.2 1

0.1

0.0 0

2

4

x

6

8

10

123

680

Stoch Environ Res Risk Assess (2007) 21:665–682

Fig. 11 The breakthrough curves of a expected value ÆQæ and b standard deviation rQ of total solute flux through the CP for cases 3 and 9–12

0.5

a

Case 3 Case 9 Case 10

0.4

Case 11 Case 12

0.3

Q 0.2

0.1

0.0 0

5

10

15

20

25

t 0.5

b

Case 3 Case 9 Case 10

0.4

Case 11 Case 12

0.3

σQ 0.2

0.1

0.0 0

5

10

15

20

25

30

t

distributions of conductivity, which is then used for the flow and transport calculations. For hydraulic head measurement, the measurement points are treated as specified internal boundaries. The various moments for flow and solute transport were developed analytically through the perturbation method. Owing to the complexity of the calculation, a finite difference method is implemented to numerically solve these moment equations. Using the developed method, several case studies have been conducted to investigate the influences of the measurements of conductivity and head on the flow and transport predictions, especially the decrease of the variances of head, velocity and solute flux. The results of the case studies demonstrate that conditioning the flow and solute calculations on the measured data will significantly affect the calculation results of the hydraulic head, flow, and solute flux. Conditioning on both the hydraulic conductivity and the head measurements will more significantly decrease head variance than conditioning on conductivity

123

measurements only. However, this different conditioning will not significantly influence the mean and variance of solute flux. Generally speaking, an increase in the number of measurements will result in smaller head variance, especially in the areas around the measurement points, less the mean solute dispersion, and smaller relative solute flux variance. Further, the different values of measurements at the same points will make the means and variances of hydraulic head and solute flux different. In summary, the number of the measured points, the sort of measured data (conductivity or/and head), and the measured values of parameters are all important factors for prediction of hydraulic head and solute flux. In the study, two important assumptions have been taken to develop in the development of theoretical framework in this study, first-order approximations and assumed distribution functions for solute flux in longitudinal and transverse direction. In Sect. 3, we discussed the assumptions for the flux parcel distributions

Stoch Environ Res Risk Assess (2007) 21:665–682

in longitudinal and transverse directions and concluded the assumptions are reasonable if the unconditional conductivity field satisfies log-normal distribution and the variance is not large, r 2y < 1.0. The first-order (small variance) approximations are invoked in deriving the general expressions of the travel time and transverse displacement moments as well as in obtaining the flow moments. Zhang et al. 2000 discussed the influence of first-order (small variance) approximations on flow and transport. They concluded that the first-order stochastic theories for flow and transport in a stationary conductivity field are quite robust and applicable to a moderate variance. Further, conditioning on conductivity and head measurements accounting reduces the underlying variance. Therefore, the small-variance assumption may be even more justified for the approaches that take medium into account conditioning. Acknowledgements This work was partially funded by the Department of Energy’s Yucca Mountain Project under contract between DOE and the University and Community College System of Nevada, partially funded by NSFC (project number 40272106), and partially funded by the Teaching and Research Award Program for Outstanding Young Teacher (TRAPOYT) of MOE, P.R.C.

References Andricevic R, Cvetkovic V (1998) Relative dispersion for solute flux in aquifers. J Fluid Mech 361:145–174 Bellin A, Rubin Y, Rinaldo A (1994) Eulerian–Lagrangian approach for modeling of flow and transport in heterogeneous geological formations. Water Resour Res 30:2913– 2925 Butera I, Tanda MG (1999) Solute transport analysis through heterogeneous media in non-uniform in the average flow by a stochastic approach. Transp Porous Med 36:255–291 Cushman JH (1997) The physics of fluids in hierarchical porous media: Angstroms to Miles. Kluwer, Dorchet Cvetkovic V, Dagan G (1994) Transport of kinetically sorbing solute by steady random velocity in heterogeneous porous formations. J Fluid Mech 265:189–215 Cvetkovic V, Shapiro AM, Dagan G (1992) A solute flux approach to transport in heterogeneous formation. 2. Uncertainty analysis. Water Resour Res 28:1377–1388 Cvetkovic V, Cheng H, Wen XH (1996) Analysis of nonlinear effects on tracer migration in heterogeneous aquifers using Lagrangian travel time statistics. Water Resour Res 32:1671–1681 Dagan G (1984) Solute transport in heterogeneous porous formations. J Fluid Mech 145:151–177 Dagan G (1989) Flow and transport in porous formations. Springer, Berlin Heidelberg New York Dagan G, Cvetkovic V, Shapiro AM (1992) A solute flux approach to transport in heterogeneous formations. 1. The general framework. Water Resour Res 28:1369–1376 Destouni G, Simic E, Graham W (2001) On the applicability of analytical methods for estimating solute travel time statistics

681 in nonuniform groundwater flow. Water Resour Res 37: 2303–2308 Ezzedine S, Rubin Y (1996) A geostatistical approach to the conditional estimation of spatially distributed solute concentration and notes on the use of tracer data in the inverse problem. Water Resour Res 32(4):853–861 Gelhar LW (1993) Stochastic subsurface hydrology. PrenticeHall, Englewood Cliffs Graham W, McLaughlin D (1989) Stochastic analysis of nonstationary subsurface solute transport. 1. Conditional moments. Water Resour Res 25:2331–2355 Guadagnini A, Neuman SP (1999a) Nonlocal and localized analyses of conditional mean steady state flow in bounded, randomly nonuniform domain. 1. Theory and computational approach. Water Resour Res 35:2999–3018 Guadagnini A, Neuman SP (1999b) Nonlocal and localized analyses of conditional mean steady state flow in bounded, randomly nonuniform domain. 2. Computational example. Water Resour Res 35:3019–3039 Hernandez AF, Neuman SP, Guadagnini A, and Carrera J (2002) Conditioning steady state mean stochastic flow equations on head and hydraulic conductivity measurements. In: Kovar K, Hrkal Z (eds) Acta Universitatis Carolinea Geologica, Prepublished proceedings of the ModelCARE’2002 conference 46:158–162 Hu BX, Wu JC, Zhang D, Shirley C (2002) A numerical method of moments for solute flux in nonstationary flow fields. In: Proceedings of the international congress on calibration and reliability in groundwater modeling. ModelCARE 2002. Acta Universitatis Carolinae-Geologica 46:109–112 Hu BX, Wu J, Panorska AK, Zhang D, Shirley C (2003) Groundwater flow and solute transport in a porous medium with multi-scale heterogeneity. Adv Water Resour 26:513– 531 Indelman P, Rubin Y (1996) Solute transport in nonstationary velocity fields. Water Resour Res 32:1259–1267 James AI, Graham WD (1999) Numerical approximation of head and flux covariances in three dimensions using mixed finite elements. Adv Water Resour 22:729–740 Jazwinski AH (1970) Stochastic processes and filtering theory. Academic, San Diego Li L, Graham W (1998) Stochastic analysis of solute transport in heterogeneous aquifers subject to spatially random recharge. J Hydrol 206:16–38 Neuman SP (1993) Eulerian–Lagrangian theory of transport in space-time nonstationary fields: exact nonlocal formalism by conditional moments and weak approximations. Water Resour Res 29:633–645 Neuman SP, Orr S (1993) Prediction of steady state flow in nonuniform geologic media by conditional moments: exact nonlocal formalism, effective conductivities, and weak approximation. Water Resour Res 29:341–364 Neuman SP, Tartakovsky DM, Wallstrom TC, Winter CL (1996) Correction to ‘‘Prediction of steady state flow in nonuniform geologic media by conditional moments: exact nonlocal formalism, effective conductivities, and weak approximation’’. Water Resour Res 32:1479–1480 Osnes H (1998) Stochastic analysis of velocity spatial variability in bounded rectangular heterogeneous aquifers. Adv Water Resour 21:203–215 Rubin Y, Dagan G (1992) Conditional estimation of solute travel time in heterogeneous formations—impact of transmissivity measurements. Water Resour Res 28:1033–1040 Wu J, Hu BX, Zhang D (2003) Application of nonstationary stochastic theories to solute transport in multi-scale geological media. J Hydrol 275:208–228

123

682 Zhang D (1998) Numerical solutions to statistical moment equations of groundwater flow in nonstationary, bounded heterogeneous media. Water Resour Res 34:529–538 Zhang D (2002) Stochastic methods for flow in porous media: coping with uncertainties. Academic, San Diego Zhang D, Neuman SP (1995a) Eulerian–Lagrangian analysis of transport conditioned on hydraulic data. 1. Analytical– numerical approach. Water Resour Res 31:39–51 Zhang D, Neuman SP (1995b) Eulerian–Lagrangian analysis of transport conditioned on hydraulic data. 2. Effects of log transmissivity and hydraulic head measurements. Water Resour Res 31:53–63

123

Stoch Environ Res Risk Assess (2007) 21:665–682 Zhang D, Neuman SP (1995c) Eulerian–Lagrangian analysis of transport conditioned on hydraulic data. 3. Spatial moments, travel time distribution, mass flow rate, and cumulative release across a compliance surface. Water Resour Res 31:65–75 Zhang D, Winter CL (1999) Moment equation approach to single phase fluid flow in heterogeneous reservoirs. Soc Petrol Eng J 4:118–127 Zhang D, Andricevic R, Sun AY, Hu BX, He G (2000) Solute flux approach to transport through spatially nonstationary flow in porous media. Water Resour Res 36:2107–2120