Skip to main content
Log in

Steady flows driven by sources of random strength in heterogeneous aquifers with application to partially penetrating wells

  • Original Paper
  • Published:
Stochastic Environmental Research and Risk Assessment Aims and scope Submit manuscript

Abstract

Average steady source flow in heterogeneous porous formations is modelled by regarding the hydraulic conductivity K(x) as a stationary random space function (RSF). As a consequence, the flow variables become RSFs as well, and we are interested into calculating their moments. This problem has been intensively studied in the case of a Neumann type boundary condition at the source. However, there are many applications (such as well-type flows) for which the required boundary condition is that of Dirichlet. In order to fulfill such a requirement the strength of the source must be proportional to K(x), and therefore the source itself results a RSF. To solve flows driven by sources whose strength is spatially variable, we have used a perturbation procedure similar to that developed by Indelman and Abramovich (Water Resour Res 30:3385–3393, 1994) to analyze flows generated by sources of deterministic strength. Due to the linearity of the mathematical problem, we have focused on the explicit derivation of the mean head distribution G d (x) generated by a unit pulse. Such a distribution represents the fundamental solution to the average flow equations, and it is termed as mean Green function. The function G d (x) is derived here at the second order of approximation in the variance σ2 of the fluctuation \(\varepsilon \left({{\mathbf{x}}}\right) = 1- \frac{K{\left({{\mathbf{x}}}\right)}}{K_{A}}\) (where K A is the mean value of K(x)), for arbitrary correlation function ρ(x), and any dimensionality d of the flow domain. We represent G d (x) as product between the homogeneous Green function G (0) d (x) valid in a domain with constant K A , and a distortion term Ψ d (x) = 1 + σ2ψ d (x) which modifies G (0) d (x) to account for the medium heterogeneity. In the case of isotropic formations ψ d (x) is expressed via one quadrature. This quadrature can be analytically calculated after adopting specific (e.g.. exponential and Gaussian) shape for ρ(x). These general results are subsequently used to investigate flow toward a partially-penetrating well in a semi-infinite domain. Indeed, we construct a σ2-order approximation to the mean as well as variance of the head by replacing the well with a singular segment. It is shown how the well-length combined with the medium heterogeneity affects the head distribution. We have introduced the concept of equivalent conductivity K eq(r,z). The main result is the relationship \(\frac{K^{\rm eq}\left(r,z\right)} {K_{A}} = 1-\sigma^{2}\psi^{\left(w\right)}\left(r,z\right)\) where the characteristic function ψ(w)(r,z) adjusts the homogeneous conductivity K A to account for the impact of the heterogeneity. In this way, a procedure can be developed to identify the aquifer hydraulic properties by means of field-scale head measurements. Finally, in the case of a fully penetrating well we have expressed the equivalent conductivity in analytical form, and we have shown that \(K^{({\rm efu})}\leq K^{\rm eq}\left(r\right) \leq K_{A}\) (being \(K^{({\rm efu})}\) the effective conductivity for mean uniform flow), in agreement with the numerical simulations of Firmani et al. (Water Resour Res 42:W03422, 2006).

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7

Similar content being viewed by others

Abbreviations

C h (x,y):

head-covariance

C Kh (x,y):

cross-covariance between the conductivity and head

d :

domain dimensionality

Δ:

well length

Δ2 :

Laplacian operator

δ(x):

Dirac delta function

E :

constant of Euler–Mascheroni

Ei(x):

exponential integral

erfc(x):

complementary error function

ɛ(x):

residual of K(x)/K A

ϕ(x):

source function

\({\widetilde{f}}\left({{\mathbf{k}}}\right)\) :

Fourier transform of f(x)

Φ(k):

spectrum (i.e., Fourier transform of ρ(x))

G (0) d (x):

homogeneous Green function pertaining to Ω(d)

G d (x):

mean Green function in Ω(d)

G (2) d (x):

second order correction to the Green function in Ω(d)

\({\overline{G}}^{\left(0\right)}\left({{\mathbf{x}}};{{\mathbf{x}}}^{\prime}\right)\) :

homogeneous Green function pertaining to \(\overline{\Upomega}\)

\({\overline{G}}^{\left(2\right)}\left({{\mathbf{x}}};{{\mathbf{x}}}^{\prime}\right)\) :

second order correction of the Green function pertaining to \({\overline{\Upomega}}\)

H(x):

Heaviside step-function

h(x):

pressure head

h n (x):

n-order correction to the pressure head

h 1(x):

head fluctuation

h w :

head boundary condition at the well

I :

horizontal integral scale of heterogeneity

I v :

vertical integral scale of heterogeneity

Lν(x):

ν-order Struve function

λ:

anisotropy ratio

K(x):

hydraulic conductivity

K A :

arithmetic mean of K(x)

K G :

geometric mean of K(x)

K eq(x):

equivalent conductivity

\(K^{({\rm efu})}\) :

effective conductivity in mean uniform flows

K ν(x):

modified ν order Bessel function

κ (x):

normalized equivalent conductivity

Ω(d) :

unbounded flow domain of d dimensionality

\({\overline{\Upomega}}\) :

three-dimensional semi-infinite flow domain with impervious boundary

Ψ d (x):

characteristic heterogeneity function

ψ d (x):

normalized second order correction to the Green Function

ψ q d (x):

normalized second order correction due to a flux-type boundary condition

ψ h d (x):

normalized second order correction due to a head-type boundary condition

\(\psi^{\left({\rm w}\right)}\left({\mathbf{x}}\right)\) :

normalized second order correction to the head in partially-penetrating well

ψ *3 (λ ):

asymptotic value of ψ3(x)

ρ(x):

autocorrelation function of ɛ(x)

ρ Y (x):

autocorrelation function of Y(x)

Q w :

well discharge

\({\overline{Q}}_{w}\) :

discharge per unit well length

r :

radial distance

r w :

well radius

σ2 :

variance of ɛ(x)

σ 2 Y :

variance of Y(x)

σ 2 h :

head-variance in well-flow

W a,b (x):

Whittaker function

x :

vectorial distance

ξ K :

coefficient of variation of K(x)

z :

depth

Y(x):

log-conductivity

〈〉:

ensemble average operator

∇:

gradient

References

  • Ababou R, Wood EF (1990) Comment on Effective groundwater model parameter values: Influence of spatial variability of hydraulic conductivity, leakance, and recharge by J.J. Gomez-Hernandez and S.M. Gorelick. Water Resour Res 26:1843–1846

    Google Scholar 

  • Beran MJ (1968) Statistical continuum theories. Interscience, New York

    Google Scholar 

  • Bruggeman GA (1999) Analytical solutions of geohydrological problems. Elsevier, Amsterdam

    Book  Google Scholar 

  • Butler JJ (1991) A stochastic analysis of pumping test in laterally nonuniform media. Water Resour Res 27:2401–2414

    Article  Google Scholar 

  • Cassiani G, Kabala ZJ (1998) Hydraulics of a partially-penetrating well: solution to a mixed-type boundary value problem via dual integral equation. J Hydrol 211:100–111

    Article  Google Scholar 

  • Chen CS, Chang CC (2002) Use of cumulative volume of constant-head injection test to estimate aquifer parameters with skin effects: field experiment and data analysis. Water Resour Res 38. doi:10.1029/2001WR000300

  • Choy TC (1999) Effective medium theory. Clarendon Press, Oxford

    Google Scholar 

  • Dagan G (1978) A note on packer, slug, and recovery tests in unconfined aquifers. Water Resour Res 14:929–934

    Article  Google Scholar 

  • Dagan G (1979) The generalization of Darcy’s law for nonuniform flow. Water Resour Res 15:1–7

    Article  Google Scholar 

  • Dagan G (1982) Analysis of flow through heterogeneous random aquifers 2. Unsteady flow in confined formations. Water Resour Res 18:1571–1585

    Article  Google Scholar 

  • Dagan G (1989) Flow and Transport in Porous Formations. Springer, Hedelberg

    Google Scholar 

  • Desbarats AJ (1992) Spatial averaging of transmissivity in heterogeneous fields with flow toward a well. Water Resour Res 28:757–767

    Article  Google Scholar 

  • Desbarats AJ (1994) Spatial averaging of hydraulic conductivities under radial flow conditions. Math Geol 26:1–21

    Article  Google Scholar 

  • Fiori A, Indelman P, Dagan G (1998) Correlation structure of flow variables for steady flow toward a well with application to highly anisotropic heterogeneous formations. Water Resour Res 34:699–708

    Article  Google Scholar 

  • Firmani G, Fiori A, Bellin A (2006) Three-dimensional numerical analysis of steady state pumping tests in heterogeneous confined aquifers. Water Resour Res 42:W03422. doi:10.1029/2005WR004382

    Article  Google Scholar 

  • Gelhar LJ, Axness CL (1983) Three-dimensional stochastic analysis of macrodispersion in aquifers. Water Resour Res 19:161–180

    Article  Google Scholar 

  • Gomez-Hernandez JJ, Gorelick SM (1989) Effective groundwater model parameter values: influence of spatial variability of hydraulic conductivity, leakance, and recharge. Water Resour Res 25:405–419

    Article  Google Scholar 

  • Gradshteyn IS, Ryzhik IM (1965) Tables of integrals, series, and products. Academic, New York

    Google Scholar 

  • Guadagnini A, Neuman SP (1999) Nonlocal and localized analyses of conditional mean steady state flow in bounded randomly nonuniform domains, 1. Theory and computational approach. Water Resour Res 35:2999–3018

    Article  Google Scholar 

  • Guadagnini A, Riva M, Neuman SP (2003) Three-dimensional steady state flow to a well in a randomly heterogeneous bounded aquifer. Water Resour Res 39:1048. doi:10.1029/2002WR001443

    Article  Google Scholar 

  • Hantush MS (1964) Hydraulics of wells. In: Ven Te Chow (ed) Advances in hydroscience. Academic, New York

  • Hiller CK, Levy BS (1994) Estimation of aquifer diffusivity from analysis of constant head pumping test data. Ground Water 32:47–52

    Article  Google Scholar 

  • Indelman P (2001) Steady-state flow in heterogeneous porous media. Transport Porous Media 45:105–127

    Article  Google Scholar 

  • Indelman P (2002) On mathematical models of average flow in heterogeneous formations. Transport Porous Media 48:209–224

    Article  Google Scholar 

  • Indelman P (2003) Transient well-type flows in heterogenous formations. Water Resour Res 40. doi:10.1029/2002WR001407

  • Indelman P, Abramovich B (1994) Nonlocal properties of nonuniform averaged flows in heterogeneous media. Water Resour Res 30:3385–3393

    Article  Google Scholar 

  • Indelman P, Dagan G (2004) A note on well boundary condition for flow through heterogeneous formations. Water Resour Res 40:W03601. doi:10.1029/2003WR002602

    Article  Google Scholar 

  • Indelman P, Fiori A, Dagan G (1996) Steady flow toward wells in heterogeneous formations: mean head and equivalent conductivity. Water Resour Res 32:1975–1983

    Article  Google Scholar 

  • Jones L, Lemar T, Tsai C-T (1992) Results of two pumping tests in Wisconsin age weathered till in Iowa. Ground Water 30:529–538

    Article  Google Scholar 

  • Leblanc DR, Garabedian S P., Hess KM, Gelhar LW, Quadri RD, Stollenwerk KG, Wood WW (1991) Large scale natural gradient tracer test in sand and gravel, Cape Code, 1. Experimental design and observed tracer movement. Water Resour Res 27:895–910

    Article  Google Scholar 

  • Lessoff S, Indelman P, Dagan G (2000) A note on the influence of a constant velocity boundary condition on flow and transport in heterogeneous formations. Water Resour Res 36:3095–3101

    Article  CAS  Google Scholar 

  • Matheron G (1967) Elements pour une Theorie des Milieux Poreux. Masson, Paris

    Google Scholar 

  • Murdoch LC, Franco J (1994) The analysis of constant drawdown wells using instantaneous source functions. Water Resour Res 30:117–124

    Article  Google Scholar 

  • Muskat M (1937) The flow of homogeneous fluids through porous media. McGraw-Hill, New York

    Google Scholar 

  • Naff RL (1991) Radial flow in heterogeneous porous media: An analysis of specific discharge. Water Resour Res 27:307–316

    Article  Google Scholar 

  • 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

    Article  Google Scholar 

  • Neuman SP, Tartakovsky D, 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 by Shlomo P. Neuman and Shlomo Orr, Water Resour Res 32:1479–1480

    Article  Google Scholar 

  • Oliver DS (1993) The influence of nonuniform transmissivity and storativity on drawdown. Water Resour Res 29:169–178

    Article  Google Scholar 

  • Riva M, Guadagnini A, Neuman SP, Franzetti S (2001) Radial flow in a bounded randomly heterogeneous aquifer. Transp Porous Media 45:139–193

    Article  Google Scholar 

  • Rubin Y (2003) Applied Stochastic Hydrogeology. Oxford University Press, Oxford

    Google Scholar 

  • Sahimi M (2003) Heterogeneous materials I. Springer, Heidelberg

    Google Scholar 

  • Sanchez-Vila X (1997) Radially convergent flow in heterogeneous porous media. Water Resour Res 33:1633–1641

    Article  Google Scholar 

  • Sanchez-Vila X, Axness CL, Carrera J (1999) Upscaling trasmissivity under radially convergent flow in heterogeneous media. Water Resour Res 35:613–621

    Article  Google Scholar 

  • Schad H, Teutsch G (1994) Effects of the investigation scale on pumping test results in heterogeneous porous aquifers. J Hydrol 159:61–77

    Article  Google Scholar 

  • Shvidler MI (1966) The source-type solution of the problem of unsteady flow in random porous media. Fluid Dyn, pp 95–98

  • Streltsova TD (1988) Well testing in heterogeneous formations. Wiley, New York

    Google Scholar 

  • Sudicky EA (1986) A natural gradient experiment on solute transport in a sand aquifer: spatial variability of hydraulic conductivity and its role in the dispersion process. Water Resour Res 22:2069–2082

    Article  CAS  Google Scholar 

  • Tartakovsky DM, Neuman SP (1998) Transient flow in bounded randomly heterogeneous domains, 1, Exact conditional moment equations and recursive approximations. Water Resour Res 34:1–12

    Article  Google Scholar 

  • Zhang D (2002) Stochastic methods for flow in porous media. Academic, New York

    Google Scholar 

Download references

Acknowledgments

This study was supported by grant TIDe (MIUR # 9260). Comments from two anonymous reviewers were greatly appreciated. We are also indebted to Gabriella Pisanti for the useful suggestions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Gerardo Severino.

Appendices

Appendix A: Derivation of the mean Green function

The mean Green function results from the first of Eq. (11a) as G d (x) = G (0) d (x) + σ2 G (2) d (x), being G (0) d (x) = FT−1[k −2]. The second order correction G (2) d (x) is expressed as difference (see the first of (12)) between G (2) d,q (x), which accounts for a flux-type condition, and G (2) d,h (x) that adjusts the previous one to satisfy the condition of given head. The terms G (2) d,q (x) and G (2) d,h (x) are calculated as follows:

$$ G_{d,q}^{\left(2\right)}\left({{\mathbf{x}}}\right) =\hbox{FT}^{-1} \left[{\widetilde{g}}\left({{\mathbf{k}}}\right) {\widetilde{G}}_{d}^{\left(0\right)}\left(k\right) \right] \qquad G_{d,h}^{\left(2\right)}\left({{\mathbf{x}}}\right) = \hbox{FT}^{-1}\left[{\widetilde{s}}\left({{\mathbf{k}}}\right) {\widetilde{G}}_{d}^{\left(0\right)}\left(k\right) \right]. $$
(68)

The problem is therefore reduced to compute the inverse (68) with \({\widetilde{g}}\left({{\mathbf{k}}}\right)\) and \({\widetilde{s}}\left({{\mathbf{k}}}\right)\) given by (11b). The correction G (2) d,q (x) was derived by Indelman (2001), and we recall it for the sake of completeness

$$ G_{d,q}^{\left(2\right)}\left({{\mathbf{x}}}\right) = \frac{G_{d}^{\left( 0\right)}\left(x\right)}{d}+ \frac{d} {2} \int\limits_{\Upomega_{\left(d\right)}}d {{\mathbf{x}}}^{\prime}\rho \left({{\mathbf{x}}}^{\prime}\right) \left\vert {{\mathbf{x}}}- {{\mathbf{x}}}^{\prime}\right\vert^{2}\left[\left(\frac{x^{\prime 2}-\cdot {{\mathbf{x}}}^{\prime}}{x^{\prime}\left\vert {{\mathbf{x}}}- {{\mathbf{x}}}^{\prime}\right\vert}\right)^{2}-\frac{1} {d}\right] {\overline{\overline{G}}}_{d}^{\left(0\right)}\left(x^{\prime}\right) {\overline{\overline{G}}}_{d}^{\left(0\right)}\left(\left\vert {{\mathbf{x}}}-{{\mathbf{x}}}^{\prime}\right\vert \right) $$
(69)

where \({\overline{\overline{G}}}_{d}^{\left(0\right)}\left(x^{\prime}\right)\) is defined by (19). Instead, the explicit expression of the term G (2) d,h (x) is presented here for the first time.

In order to calculate G (2) d,h (x) we make use of the second of (13), i.e.,

$$ G_{d,h}^{\left(2\right)}\left({{\mathbf{x}}}\right) =\int\limits_{\Upomega_{\left(d\right)}}d{{\mathbf{x}}}^{\prime}s\left({{\mathbf{x}}} ^{\prime}\right) G_{d}^{\left(0\right)}\left(\left\vert {{\mathbf{x}}}-{{\mathbf{x}}}^{\prime}\right\vert \right). $$
(70)

Then, substituting (15) into (70) leads (after integrating by parts) to

$$ G_{d,h}^{\left(2\right)}\left({{\mathbf{x}}}\right) =-\int\limits_{\Upomega_{\left(d\right)}}d{{\mathbf{x}}}^{\prime}\rho \left({{\mathbf{x}}}^{\prime}\right) \frac{\partial}{\partial x_{m}^{\prime}}G_{d}^{\left(0\right)}\left(x^{\prime}\right) \left[\frac{\partial}{\partial y_{m}}G_{d}^{\left(0\right)}\left(y\right) \right]_{y=\left\vert {{\mathbf{x}}}-{{\mathbf{x}}}^{\prime}\right\vert}. $$
(71)

By noting that \(\frac{\partial}{\partial x_{m}} = \frac{x_{m}} {x} \frac{d}{dx},\) one has

$$ G_{d,h}^{\left(2\right)}\left({{\mathbf{x}}}\right) =\int\limits_{\Upomega_{\left(d\right)}}d{{\mathbf{x}}}^{\prime}\rho \left({{\mathbf{x}}}^{\prime}\right) \left(x^{\prime 2}-{{\mathbf{x}}}\cdot {{\mathbf{x}}}^{\prime}\right) \frac{\overline{\overline{G}}_{d}^{\left(0\right)}\left(x^{\prime}\right)} {x^{\prime}} \frac{{\overline{\overline{G}}}_{d}^{\left(0\right)}\left(\left\vert {{\mathbf{x}}}- {{\mathbf{x}}}^{\prime}\right\vert \right)}{\left\vert {{\mathbf{x}}}-{{\mathbf{x}}}^{\prime}\right\vert}. $$
(72)

Finally, dividing (69) and (72) by G (0) d (x) leads to (18a), (18b).

Appendix B: Derivation of the second-order correction in the case of a fully penetrating well

By letting Δ → ∞ in (48) one has

$$ \left\langle h_{2}\left(r\right) \right\rangle = \frac{-b} {6\pi} \left[\alpha +\ln \frac{r}{2} + \frac{1} {2} \int\limits_{0}^{\infty}d\zeta \Upupsilon \left(\zeta^{2}+r^{2}\right) \right], $$
(73)

with \({\Upupsilon} \left(x\right)\) given by (43). It is convenient to represent the integral appearing into (73) as \(\int\limits_{0}^{\infty}d\zeta {\Upupsilon} \left(\zeta^{2}+r^{2}\right) ={{\mathcal{I}}}_{1}\left(r\right) -{{\mathcal{I}}}_{2}\left(r\right),\) being

$$ {{\mathcal{I}}}_{1}\left(r\right) =\left\{\begin{array}{ll}\int\limits_{0}^{\infty}d\zeta \left(\frac{2} {\sqrt{\zeta^{2}+r^{2}}} + \sqrt{\zeta^{2}+r^{2}}-1\right) \exp \left(-\sqrt{\zeta^{2}+r^{2}}\right) & \hbox{exponential} \\ \\ \int\limits_{0}^{\infty}d\zeta \left(\frac{2}{\sqrt{\zeta^{2}+r^{2}}}-\pi \sqrt{\zeta^{2}+r^{2}}\right) \exp \left[-\frac{\pi} {4}\left(\zeta^{2}+r^{2}\right) \right] & \hbox{Gaussian} \end{array}\right. $$
(74)

and

$$ {{\mathcal{I}}}_{2}\left(r\right) =\left\{\begin{array}{ll} \int\limits_{0}^{\infty}d\zeta \left(\zeta^{2}+r^{2}\right) \hbox{Ei} \left(-\sqrt{\zeta^{2}+r^{2}}\right) & \hbox{exponential}\\\\ \frac{\pi^{2}}{2} \int\limits_{0}^{\infty}d\zeta \left(\zeta^{2}+r^{2}\right) \hbox{erfc}\left[\frac{1}{2}\sqrt{\pi \left(\zeta^{2}+r^{2}\right)} \right] & \hbox {Gaussian}. \end{array}\right. $$
(75)

In order to evaluate \({{\mathcal{I}}}_{1}\left(r\right)\) we introduce the new variable \(u=\sinh^{-1}\left(\frac{\zeta}{r}\right)\) so that it yields \({{\mathcal{I}}}_{1}\left(r\right) =\exp \left(-\frac{\pi} {8} r^{2}\right) \left[\left(1-\frac{\pi}{4}r^{2}\right) {\hbox{K}}_{0}\left(\frac{\pi}{8}r^{2}\right) - \frac{\pi}{4}r^{2} {\hbox{K}}_{1}\left(\frac{\pi}{8}r^{2}\right) \right]\) (e.g., Gradshteyn and Ryzhik 1965) for Gaussian, and \({{\mathcal{I}}}_{1}\left(r\right) =\left(2+r^{2}\right) {\hbox{K}}_{0}\left(r\right)\) (e.g., Gradshteyn and Ryzhik 1965) for exponential autocorrelation.

To calculate \({{\mathcal{I}}}_{2}\left(r\right)\) we first integrate (75) by parts

$$ {{\mathcal{I}}}_{2}\left(r\right) =\left\{\begin{array}{ll}-\frac{1} {3} \int\limits_{0}^{\infty}d\zeta \frac{3 r^{2}+\zeta^{2}} {\zeta^{2}+r^{2}} \zeta^{2}\exp \left(-\sqrt{\zeta^{2}+r^{2}}\right) & \hbox{exponential} \\ \\ \frac{\pi^{2}} {6} \int\limits_{0}^{\infty}d\zeta \frac{3 r^{2}+\zeta^{2}} {\sqrt{\zeta^{2}+r^{2}}} \zeta^{2}\exp \left[-\frac{\pi} {4} \left(\zeta^{2}+r^{2}\right) \right] & \hbox{Gaussian}. \end{array}\right. $$
(76)

Hence, by carrying out the quadratures in (76), and after some algebraic derivations (which we omit for brevity), one obtains (56)–(57).

Appendix C: Derivation of the head variance for a fully penetrating well

Starting from the general expression (58) of the head fluctuation, and by applying the method of images, the head variance for a fully-penetrating can be written as

$$ \begin{aligned} \sigma_{h}^{2}\left(r\right) &= \pi^{-1}\int \int \int\limits_{-\infty}^{+\infty} \int\limits_{-\infty}^{+\infty}d{{\mathbf{r}}}^{\prime}d{{\mathbf{r}}}^{\prime \prime} dz^{\prime}dz^{\prime \prime}{{\mathcal{G}}}\left({{\mathbf{x}}}^{\prime}; {{\mathbf{r}}}\right) {{\mathcal{G}}}\left({{\mathbf{x}}}^{\prime \prime};{{\mathbf{r}}} \right) \frac{\partial}{\partial x_{m}^{\prime}}h^{\left(0\right)}\left(r^{\prime}\right) \frac{\partial}{\partial x_{n}^{\prime \prime}} h^{\left(0\right)}\left(r^{\prime \prime}\right)\\ &\quad \times \frac{\partial^{2}}{\partial x_{m}^{\prime}\partial x_{n}^{\prime \prime}} \int\limits_{-\infty}^{+\infty}\int\limits_{-\infty}^{+\infty}dk^{\prime}dk^{\prime \prime}\exp \left[-j\left(k^{\prime}z^{\prime}+k^{\prime \prime} z^{\prime \prime}\right) \right] \left\langle {\widetilde{\varepsilon}} \left({{\mathbf{r}}}^{\prime}, k^{\prime}\right) {\widetilde{\varepsilon}} \left({{\mathbf{r}}}^{\prime \prime},k^{\prime \prime}\right) \right\rangle, \quad \left(m,n=1,2\right)\end{aligned} $$
(77)

being

$$ {{\mathcal{G}}}\left({{\mathbf{x}}};{\varvec{\alpha}}\right) = \frac{1} {4\pi} \left(\frac{1}{\sqrt{\left\vert {{\mathbf{x}}}_{h}- {\varvec{\alpha}}\right\vert^{2}+z^{2}}}- \frac{1} {\sqrt{x_{h}^{2}+z^{2}}}\right) \qquad {{\mathbf{x}}}\equiv \left({{\mathbf{x}}}_{h},z\right). $$
(78)

In Eq. (77) we have represented the fluctuation ɛ by means of its FT along the stationarity axis, i.e., \(\varepsilon \left({{\mathbf{r}}}, z\right) =\int\limits_{-\infty}^{+\infty}\frac{dk}{\sqrt{2\pi }}\exp \left(-jkz\right) {\widetilde{\varepsilon}}\left({{\mathbf{r}}}, k\right).\) By using the stationarity property \(\left\langle {\widetilde{\varepsilon}}\left({{\mathbf{r}}}^{\prime}, k^{\prime}\right) {\widetilde{\varepsilon}}\left({{\mathbf{r}}}^{\prime \prime},k^{\prime \prime}\right) \right\rangle =\sqrt{2\pi} \sigma^{2} \delta \left(k^{\prime}+k^{\prime \prime}\right) \widetilde{\rho}\left(\left\vert {{\mathbf{r}}}^{\prime}-{{\mathbf{r}}}^{\prime \prime}\right\vert, k^{\prime \prime}\right),\) and carrying out the quadratures over z′ and z″, one obtains

$$ \frac{\sigma_{h}^{2}\left(r\right)}{\sigma^{2}} =\int \int \int\limits_{0}^{\infty}\frac{d{{\mathbf{r}}}^{\prime}d{{\mathbf{r}}}^{\prime \prime}dk}{\left(2\pi \right)^{5/2}}{{\mathcal{K}}}_{k}^{r}\left ({{\mathbf{r}}}^{\prime}\right) {{\mathcal{K}}}_{k}^{r}\left({{\mathbf{r}}}^{\prime \prime}\right) \frac{\partial}{\partial x_{m}^{\prime}}h^{\left(0\right)}\left(r^{\prime}\right) \frac{\partial}{\partial x_{n}^{\prime \prime}}h^{\left(0\right)}\left(r^{\prime \prime}\right) \frac{\partial^{2}\widetilde{\rho}\left(\left\vert {\bf r}^{\prime}-{\bf r}^{\prime \prime}\right\vert ,k\right)}{\partial x_{m}^{\prime}\partial x_{n}^{\prime \prime}} $$
(79)

where \({{\mathcal{K}}}_{k}^{r}\left({\varvec{\alpha}}\right) = \hbox{K}_{0}\left(k\left\vert {\varvec{\alpha}}-{{\mathbf{r}}}\right\vert \right) -\hbox{K}_{0}\left(k\alpha \right).\) Noting that

$$ \frac{\partial^{2}{\widetilde{\rho}}\left(\left\vert {{\mathbf{r}}}^{\prime}- {\bf r}^{\prime \prime}\right\vert ,k\right)}{\partial x_{m}^{\prime}\partial x_{n}^{\prime \prime}} = \left. -\frac{\alpha_{m}\alpha_{n}}{\alpha^{2}} \frac{\partial^{2}} {\partial \alpha^{2}} {\widetilde{\rho}} \left(\alpha ,k\right) + \alpha^{-1}\left(\frac{\alpha_{m}\alpha_{n}} {\alpha^{2}} -\delta_{m,n}\right) \frac{\partial}{\partial \alpha} {\widetilde{\rho}} \left(\alpha ,k\right) \right\vert_{{\varvec{\alpha}} = {{\mathbf{r}}}^{\prime} - {{\mathbf{r}}}^{\prime \prime}}, $$
(80)

and accounting for \(\frac{\partial}{\partial x_{m}}h^{\left(0\right)}\left(r\right) = \frac{x_{m}}{r}\frac{d} {dr} h^{\left(0\right)}\left(r\right),\) leads to

$$ \sigma_{h}^{2}\left(r\right) = \frac{\sigma^{2}}{\left(2\pi \right)^{5/2}} \int \int \int\limits_{0}^{\infty} \frac{d{{\mathbf{r}}}^{\prime} d{{\mathbf{r}}}^{\prime \prime}dk}{r^{\prime}r^{\prime \prime}}\frac{d} {dr^{\prime}} h^{\left(0\right)}\left(r^{\prime}\right) \frac{d} {dr^{\prime \prime}} h^{\left(0\right)}\left(r^{\prime \prime}\right) {{\mathcal{K}}}_{k}^{r}\left({{\mathbf{r}}}^{\prime}\right) {{\mathcal{K}}}_{k}^{r}\left({{\mathbf{r}}}^{\prime \prime}\right) {{\vartheta}}_{k}\left({{\mathbf{r}}}^{\prime}, {{\mathbf{r}}}^{\prime \prime}\right), $$
(81)

being the function \({{\vartheta}}_{k}\left({{\mathbf{r}}}^{\prime}, {{\mathbf{r}}}^{\prime \prime}\right)\) dependent on the shape of ρ, i.e.,:

$$ {{{\vartheta}}}_{k}\left({{\mathbf{r}}}^{\prime}, {{\mathbf{r}}}^{\prime \prime}\right) = \left. \frac{\overline{\chi}}{\alpha^{2}} \frac{\partial ^{2}}{\partial \alpha^{2}} {\widetilde{\rho}}\left(\alpha ,k\right) -\alpha ^{-1}\left({{\mathbf{r}}}^{\prime}\cdot {{\mathbf{r}}}^{\prime \prime}+ \frac{\overline{\chi}}{\alpha^{2}}\right) \frac{\partial}{\partial \alpha} {\widetilde{\rho}}\left(\alpha, k\right) \right\vert_{\alpha =\left\vert {{\mathbf{r}}}^{\prime}-{{\mathbf{r}}}^{\prime \prime}\right\vert} $$
(82)
$$ {\overline{\chi}} = \left(r^{\prime 2}-{{\mathbf{r}}}^{\prime}\cdot {{\mathbf{r}}} ^{\prime \prime}\right) \left(r^{\prime \prime 2}-{{\mathbf{r}}}^{\prime}\cdot {{\mathbf{r}}}^{\prime \prime}\right). $$
(83)

Finally, switching to polar coordinates \({{\mathbf{r}}}^{\prime}\equiv r^{\prime}\left(\cos \theta^{\prime},\sin \theta^{\prime}\right),\) and \({{\mathbf{r}}}^{\prime \prime}\equiv r^{\prime \prime}\left(\cos \theta^{\prime \prime},\sin \theta^{\prime \prime}\right)\) into (81) leads to (62).

Rights and permissions

Reprints and permissions

About this article

Cite this article

Severino, G., Santini, A. & Sommella, A. Steady flows driven by sources of random strength in heterogeneous aquifers with application to partially penetrating wells. Stoch Environ Res Risk Assess 22, 567–582 (2008). https://doi.org/10.1007/s00477-007-0175-5

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00477-007-0175-5

Keywords

Navigation