Mechanisms of Anomalous Dispersion in Flow Through ... › 2016 › 10 ›...

12
Mechanisms of Anomalous Dispersion in Flow Through Heterogeneous Porous Media Alina Tyukhova, 1, 2 Marco Dentz, 3, * Wolfgang Kinzelbach, 1 and Matthias Willmann 1 1 Institute of Environmental Engineering, ETH Zurich, Zurich, Switzerland. 2 Department of Civil and Environmental Engineering, Massachusetts Institute of Technology, Cambridge, MA, USA 3 Spanish National Research Council (IDAEA-CSIC), Barcelona, Spain (Dated: October 19, 2016) We study the origins of anomalous dispersion in heterogeneous porous media in terms of the medium and flow properties. To identify and quantify the heterogeneity controls, we focus on porous media which are organized in assemblies of equally sized conductive inclusions embedded in a constant conductivity matrix. We study the behavior of particle arrival times for different conductivity distributions and link the statistical medium characteristics to large scale transport using a continuous time random walk (CTRW) approach. The CTRW models particle motion as a sequence of transitions in space and time. We derive an explicit map of the conductivity onto the transition time distribution. The derived CTRW model predicts solute transport based on the conductivity distribution and the characteristic heterogeneity length. In this way, heavy tails in solute arrival times and anomalous particle dispersion as measured by the centered mean square displacement are directly related to the medium properties. These findings shed light on the mechanisms of anomalous dispersion in heterogeneous porous media, and provide a basis for the predictive modeling of large scale transport. I. INTRODUCTION Anomalous dispersion has been widely observed in transport through heterogeneous porous media [1–5]. It manifests itself in heavy tails in solute arrival time dis- tributions, or breakthrough and the non-linear evolution of the second centered moments of solute distributions. Anomalous dispersion can be caused by different physical processes, chemical heterogeneity [6, 7], the interplay of physical heterogeneity and diffusion [8, 9], and physical heterogeneity alone. Here, we concentrate on the impact of physical heterogeneity in the distribution of hydraulic conductivity [4, 5, 10, 11]. In highly heterogeneous fields fast flow concentrates along highly permeable preferen- tial paths consisting of connected structures of large hy- draulic conductivity[12–14]. The spatial complement to the flow channels form disconnected zones of slow advec- tive velocities where solutes are delayed. The interplay of fast channels and slow advection in disconnected zones leads to anomalous dispersion. The impact of these mechanisms on large scale trans- port can be described in terms of continuous times ran- dom walks (CTRW). The CTRW [15, 16] has found ap- plications for the modeling of anomalous diffusion in a wide range of physical systems [2, 17–20]. Berkowitz and Scher [2, 21], have realized that the CTRW provides the dynamics needed to characterize non-Fickian hydrody- namic transport in heterogeneous porous and fractured media. The CTRW describes particle movements as a random walk in space and time as [9, 22] x n+1 = x n + n , t n+1 = t n + n v n , (1) * [email protected] with n the transition length and v n the particle veloc- ity. The spatial jumps and waiting times may be inde- pendent or correlated random variables [23]. The tran- sition times in the CTRW are given in terms of the par- ticle velocities v n , whose statistics have typically been estimated by using particle tracking simulations in the detailed heterogeneous flow [23–26]. The multirate mass transfer (MRMT) framework models the interplay of fast channels and slow advection by a mobile-immobile ap- proach. Fast channels define the mobile zone, regions of slow advection are represented as immobile. The mobile and immobile regions are connected through linear mass transfer [27–29] characterized by the memory function. The latter encodes the distribution of residence times in the immobile zones, which in principle is related to the statistics of slow advection. CTRW and MRMT have a similar phenomenological basis and both model his- tory dependent transport dynamics. In fact, it has been shown [30–32] that both models are under certain condi- tions mathematically equivalent. For both modeling ap- proaches, the relation between the (statistical) medium and flow properties and large scale transport is of central importance [2, 26, 33–35]. Oftentimes, the distribution of transition times (CTRW) and the memory function (MRMT) are estimated on the basis of coupled flow and transport simulations or from experimental data, for ex- ample breakthrough curves. For some systems such as diffusion in quenched random trap models [36], or hydro- dynamic transport under linear retention due to physi- cal and chemical medium heterogeneity, can the medium properties and geometry be directly linked to the aver- age non-Fickian transport behavior [37–39]. The objec- tive here is to investigate the quantitative link between the hydraulic conductivity distribution and large scale transport in the framework of CTRW and thus elucidate the heterogeneity controls on anomalous dispersion.

Transcript of Mechanisms of Anomalous Dispersion in Flow Through ... › 2016 › 10 ›...

Page 1: Mechanisms of Anomalous Dispersion in Flow Through ... › 2016 › 10 › inclusionsctrw_preprint.pdfpredictive modeling of large scale transport. I. INTRODUCTION Anomalous dispersion

Mechanisms of Anomalous Dispersion in Flow Through Heterogeneous Porous Media

Alina Tyukhova,1, 2 Marco Dentz,3, ∗ Wolfgang Kinzelbach,1 and Matthias Willmann1

1Institute of Environmental Engineering, ETH Zurich, Zurich, Switzerland.2Department of Civil and Environmental Engineering,

Massachusetts Institute of Technology, Cambridge, MA, USA3Spanish National Research Council (IDAEA-CSIC), Barcelona, Spain

(Dated: October 19, 2016)

We study the origins of anomalous dispersion in heterogeneous porous media in terms of themedium and flow properties. To identify and quantify the heterogeneity controls, we focus onporous media which are organized in assemblies of equally sized conductive inclusions embeddedin a constant conductivity matrix. We study the behavior of particle arrival times for differentconductivity distributions and link the statistical medium characteristics to large scale transportusing a continuous time random walk (CTRW) approach. The CTRW models particle motionas a sequence of transitions in space and time. We derive an explicit map of the conductivityonto the transition time distribution. The derived CTRW model predicts solute transport basedon the conductivity distribution and the characteristic heterogeneity length. In this way, heavytails in solute arrival times and anomalous particle dispersion as measured by the centered meansquare displacement are directly related to the medium properties. These findings shed light on themechanisms of anomalous dispersion in heterogeneous porous media, and provide a basis for thepredictive modeling of large scale transport.

I. INTRODUCTION

Anomalous dispersion has been widely observed intransport through heterogeneous porous media [1–5]. Itmanifests itself in heavy tails in solute arrival time dis-tributions, or breakthrough and the non-linear evolutionof the second centered moments of solute distributions.Anomalous dispersion can be caused by different physicalprocesses, chemical heterogeneity [6, 7], the interplay ofphysical heterogeneity and diffusion [8, 9], and physicalheterogeneity alone. Here, we concentrate on the impactof physical heterogeneity in the distribution of hydraulicconductivity [4, 5, 10, 11]. In highly heterogeneous fieldsfast flow concentrates along highly permeable preferen-tial paths consisting of connected structures of large hy-draulic conductivity[12–14]. The spatial complement tothe flow channels form disconnected zones of slow advec-tive velocities where solutes are delayed. The interplayof fast channels and slow advection in disconnected zonesleads to anomalous dispersion.

The impact of these mechanisms on large scale trans-port can be described in terms of continuous times ran-dom walks (CTRW). The CTRW [15, 16] has found ap-plications for the modeling of anomalous diffusion in awide range of physical systems [2, 17–20]. Berkowitz andScher [2, 21], have realized that the CTRW provides thedynamics needed to characterize non-Fickian hydrody-namic transport in heterogeneous porous and fracturedmedia. The CTRW describes particle movements as arandom walk in space and time as [9, 22]

xn+1 = xn + `n, tn+1 = tn +`nvn, (1)

[email protected]

with `n the transition length and vn the particle veloc-ity. The spatial jumps and waiting times may be inde-pendent or correlated random variables [23]. The tran-sition times in the CTRW are given in terms of the par-ticle velocities vn, whose statistics have typically beenestimated by using particle tracking simulations in thedetailed heterogeneous flow [23–26]. The multirate masstransfer (MRMT) framework models the interplay of fastchannels and slow advection by a mobile-immobile ap-proach. Fast channels define the mobile zone, regions ofslow advection are represented as immobile. The mobileand immobile regions are connected through linear masstransfer [27–29] characterized by the memory function.The latter encodes the distribution of residence times inthe immobile zones, which in principle is related to thestatistics of slow advection. CTRW and MRMT havea similar phenomenological basis and both model his-tory dependent transport dynamics. In fact, it has beenshown [30–32] that both models are under certain condi-tions mathematically equivalent. For both modeling ap-proaches, the relation between the (statistical) mediumand flow properties and large scale transport is of centralimportance [2, 26, 33–35]. Oftentimes, the distributionof transition times (CTRW) and the memory function(MRMT) are estimated on the basis of coupled flow andtransport simulations or from experimental data, for ex-ample breakthrough curves. For some systems such asdiffusion in quenched random trap models [36], or hydro-dynamic transport under linear retention due to physi-cal and chemical medium heterogeneity, can the mediumproperties and geometry be directly linked to the aver-age non-Fickian transport behavior [37–39]. The objec-tive here is to investigate the quantitative link betweenthe hydraulic conductivity distribution and large scaletransport in the framework of CTRW and thus elucidatethe heterogeneity controls on anomalous dispersion.

Page 2: Mechanisms of Anomalous Dispersion in Flow Through ... › 2016 › 10 › inclusionsctrw_preprint.pdfpredictive modeling of large scale transport. I. INTRODUCTION Anomalous dispersion

2

To this end, we consider transport in the flow throughheterogeneous porous media which are organized in as-semblies of equally sized conductive inclusions embeddedin a constant conductivity matrix. The conductivities in-side the inclusions are constant and distributed betweenthe inclusions. This type of media serves as models forheterogeneous porous media characterized by finite cor-relation length and arbitrary conductivity point distri-butions. Eames and Bush [40] studied solute dispersionin such media and derived expressions for the dispersioncoefficients. Fiori et al. [41, 42, 43] studied anomaloustransport in media consisting of inclusions with lognor-mal distributions of hydraulic conductivity and derivedsemi-analytical expressions for solute travel times. Thesesolutions have been implemented into a time-domain ran-dom walk approach, which is similar to the CTRW [44].

In this paper we investigate anomalous transport andits heterogeneity controls through detailed numericalsimulations of flow and particle transport in realizationsof the model porous medium. The next section intro-duces the flow and transport model, defines the heteroge-neous model medium, and discusses the flow properties.Section III derives the CTRW model to quantify anoma-lous dispersion, and the relations between the transitiontime distribution, particle and flow velocities, and the hy-draulic conductivity distribution. Section IV applies thedeveloped model to predict first passage time distribu-tions obtained from direct numerical simulations of flowand particle transport in realization of the heterogeneousmodel media for power-law and lognormal distributionsof hydraulic conductivity. Section V uses the developedCTRW approach to study the dispersion properties inhighly heterogeneous porous media.

II. FLOW AND TRANSPORT MODEL

In this section we outline the transport and flow mod-els and define the heterogeneous porous medium modelunder consideration.

A. Transport

We consider particle transport in the absence of mi-croscale dispersion and focus solely on the impact of het-erogeneous advection on the dispersion of an advectedscalar c(x, t). Its evolution in a divergence-free flow v(x)with ∇ · v(x) = 0 is governed by the Liouville equation

∂c(x, t)

∂t= −v(x) · ∇c(x, t). (2)

The coordinate vector is x = (x, y)>, where the super-script > denotes the transpose. We take a Lagrangianview point in order to derive the effective equations ofmotion of the dispersed scalar and start from the equiv-

FIG. 1. Illustration of the model medium consisting of a con-ductive matrix with k0 = 1 and less conductive disc-shapedinclusions of radius r0 = 1 and conductivity k, which is dis-tributed according to pk(k). The inclusions are embeddedin a rectangular unit cell of size `0. Different colors denotedifferent conductivity values.

alent advection equation

dx(t)

dt= v[x(t)], (3)

which describes the evolution of scalar particles, whosedensity is denoted by c(x, t). We employ a stochasticframework to quantify the average transport behavior inthe heterogeneous flow field v(x). This means v(x) isconsidered a realization of an ensemble of random flowfields whose statistical properties are discussed in thenext section.

In this framework, we study the ensemble transportbehavior in terms of the first passage times at a planelocated at x = xc,

τ(xc) = inf {t|x(t) ≥ xc} (4)

The probability density function (PDF) of first passagetimes is given by

f(t, xc) = 〈δ[t− τ(xc)]〉, (5)

where the angular brackets denote the combined averageover all particles launched in a given realization and theaverage over the ensemble of random media. Note thatf(t, xc) is equivalent to the solute breakthrough curvemeasured at the position xc.

Page 3: Mechanisms of Anomalous Dispersion in Flow Through ... › 2016 › 10 › inclusionsctrw_preprint.pdfpredictive modeling of large scale transport. I. INTRODUCTION Anomalous dispersion

3

B. Flow

We consider here flow through heterogeneous porousmedia that are composed of a homogeneous matrix ofhydraulic conductivity k0 and equally sized discs of ra-dius r0. Each sphere is embedded in a unit cell of length`0. The conductivities k of the disc-shaped inclusions areassigned randomly from the PDF pk(k). Without loss ofgenerality, we set k0 = 1 and r0 = 1 in the following. Thevolume fractions of the disc-shaped inclusions is given byχ = π/`20. A realization of the random media under con-sideration is illustrated in Figure 1.

The flow velocity v(x) through this medium is givenby the Darcy equation [45],

v(x) = −k(x)∇h(x), (6)

where k(x) denotes hydraulic conductivity and h(x) hy-draulic head. As outlined in the following, hydraulic con-ductivity is modeled as a spatial random field, this meansv(x) is a realization of an ensemble of random flow fieldscharacterized by certain statistical properties. Flow isdriven here by a uniform mean hydraulic gradient ∇h(x)that is aligned with the x–axis of the coordinate system,see also Appendix A 1.

1. Single Inclusion

In order to characterize the flow in the randommedium, we first consider flow through an isolated unitcell embedded in an infinite porous matrix. The steady-state flow potential function is given by [40, 46]

Φ(r) = um

(1 +

1− k1 + k

1

r2

)r1 (7a)

for r ≡ |r| > 1 and

Φ(r) =2umkr11 + k

(7b)

for r ≤ 1. Note that r = (r1, r2)> refers to the coordi-nate system with the origin in the center of the circularinclusion; um denotes the velocity in the matrix at infin-ity. The velocity field is given by u(r) = ∇Φ(r). In thematrix outside the inclusion, we have

uo(r) = um

(1 +

1− k1 + k

r22 − r21r4

)e1

− um1− k1 + k

2r1r2r4

e2 (8a)

for r > 1. Inside the inclusion, the flow velocity is con-stant and given by

ui(r) =2umk

1 + ke1 (8b)

for r ≤ 1 with ei the unit vector aligned with the i–direction of the coordinate system. Darcy velocities in-side the inclusion are uniform and aligned with the di-rection of the mean pressure gradient.

2. Distribution of Inclusions

We consider now the properties of the flow in themedium illustrated in Figure 1. The conductivitieswithin the inclusions are drawn independently from thedistribution pk(k). In order to transfer the informa-tion on the flow for the single inclusion to the randommedium, some remarks are in order. As the flow poten-tial (7) decreases as r−2, we assume that interaction be-tween the discs can be neglected. Thus, in the following,we use expressions (8a) and (8b) as an approximationfor the velocity field in the unit cell. The characteristicmatrix velocity um and the effective background conduc-tivity ke are still to be determined. They are imposedby the boundary conditions and the medium geometry.In order to determine ke, we consider the average flowvelocity q, which is given by the effective Darcy equa-tion [47, 48]

q = −ke|∇h|, (9)

where ke is the effective conductivity of the medium and∇h is the hydraulic gradient, which here is aligned withthe one-direction of the coordinate system. The Maxwellformula gives for ke [49, 50]

ke =1 + χΛ

1− χΛ, Λ =

∞∫0

dkpk(k)k − 1

k + 1, (10)

where χ = πr2

`2 is the volume fraction of the inclusions.For strong conductivity contrasts between the inclusionsand the matrix, that is 〈k〉 � 1 we may approximateΛ ≈ −1. In this case the effective conductivity is

ke ≈ 1− χ1 + χ

(11)

In general, we evaluate the integral in (10) using the fullconductivity distribution pk(k).

As outlined in Ref. [49], Maxwell’s approximation givesgood estimates for ke also for non-dilute distributions ofdiscs. The average flow velocity is now given by q =−ke∇h. Note that the average velocity q is referred tothe bulk of the medium, while um refers to the matrixdomain, this means to the area outside the inclusions.The total flux is partitioned between the inclusion andthe matrix as

q = (1− χ)um + 2umχ〈k/(1 + k)〉, (12)

Thus, we obtain for the characteristic matrix velocity umthe expression

um =−ke∇h

1− χ+ 2χ〈k/(1 + k)〉.(13)

The flow velocities in a unit cell of the heterogeneousmedium illustrated in Figure 1 then are given by (8) with

Page 4: Mechanisms of Anomalous Dispersion in Flow Through ... › 2016 › 10 › inclusionsctrw_preprint.pdfpredictive modeling of large scale transport. I. INTRODUCTION Anomalous dispersion

4

um given by (13). Specifically, the velocity ui inside aninclusion is given by (8b) as

ui =2umk

1 + k. (14)

The PDF pi(v) of the ui can be directly related to thePDF of conductivity values pk(k) as

pi(v) =2um

(2um − v)2pk

(v

2um − v

). (15)

III. STOCHASTIC PARTICLE-BASEDTRANSPORT MODEL

We focus first on particle transport in streamline co-ordinates, this means we consider particle movements asa function of distance s along a streamline,

dt(s)

ds=

1

v(s), v(s) = |v [x(s)] |, (16)

where we set x(s) ≡ x[t(s)]. Particle motion in terms ofthe distance s along the streamline reads as

dx(s)

ds= ev(s), ev(s) =

v[x(s)]

v(s). (17)

We focus now on the particle movement along the x–axis of the coordinate system and choose as the coarse-graining scale the size `0 of a unit cell. Thus, the particlemotion can be described by

xn+1 = xn + `0, tn+1 = tn + τn, (18)

where the transition time for a unit cell is given by

τn =

sn+`0∫sn

ds′

v(s′). (19)

Due to the random nature of the permeability distribu-tion illustrated in Figure 1, subsequent trapping times τncan be considered random and independent. Thus, theensemble particle motion, this means particle trajecto-ries sampled between different realizations of the randommedium in Figure 1 forms a CTRW which is fully char-acterized by the PDF ψ(t) of transit times τn. Beforedetermining the particle velocities and transit time dis-tribution in terms of the permeability distributions, webriefly recall the CTRW description of particle transport.

A. Continuous Time Random Walk

The PDF of horizontal particle positions averaged overparticles and medium realizations is given by c(x, t) =〈δ(x− xnt

)〉, where the number of steps nt to reach time

t by the process (18) is given by nt = max(n|tn ≤ t).Thus, c(x, t) can be expanded as

c(x, t) =

t∫0

dt′R(x, t′)

∞∫t−t′

dt′′ψ(t′′), (20)

R(x, t) = δ(x)δ(t) +

t∫0

dt′ψ(t− t′)R(x− `, t′). (21)

The first passage times (4) read now in terms of theCTRW (18) as [51]

τ(xc) = tnxc, nxc = min(n|xn ≥ xc). (22)

Since the spatial increment is constant equal to `0, thenumber of steps to reach xc is given by nxc

= dxc/`0e,where the upper braces denote the ceiling function. Thus,the PDF of first passage times, f(t, r) = 〈δ[t− τ(r)]〉 canbe expanded as

f(t, xc) =

t∫0

dt′f(t′, xc − `0)ψ(t− t′). (23)

Note that (18)–(23) describe a CTRW for the averageparticle dynamics in the flow through the heterogeneousmedium that is fully parameterized in terms of the distri-bution of hydraulic conductivity. Based on this CTRW,we also analyze the dispersion behavior in the heteroge-neous porous medium. To this end, we determine thesecond centered moment in the average flow direction,which is defined by

σ2(t) = 〈(xnt− 〈xnt

〉)2〉, (24)

where nt = sup(n|tn ≤ t) is the number of steps neededto reach time t in the process (18). In the following,we discuss the particle velocities in the unit cell and thecorresponding transit times.

B. Particle Velocities and Transit Times

We discuss here the distribution of the particle veloc-ities entering the CTRW model outlined in the previoussection and its relation to the medium properties. Fur-thermore, we determine the transition time distributionthat corresponds to the distribution of particle velocities.a. Velocity Distributions We simplify flow velocities

in that we do not account for variability in the velocitythrough the matrix, and set it equal to its average um.According to (19), the CTRW approach outlined in theprevious section requires the particle velocities v(s) sam-pled spatially along streamlines as an input. Their PDFps(v) is obtained in terms of the relative particle fluxesthat pass through matrix and inclusions. The fluxes Qiand Qm through inclusions and matrix are determinedat the vertical centerline of a unit cell such that

Qi(ui) = 2ui, Qm(ui) = (`0 − 2)(2um − ui) (25)

Page 5: Mechanisms of Anomalous Dispersion in Flow Through ... › 2016 › 10 › inclusionsctrw_preprint.pdfpredictive modeling of large scale transport. I. INTRODUCTION Anomalous dispersion

5

where we expressed the matrix velocity at the center line,which is given by 2um/(1 +k) in terms of ui. Recall thatthe disc radius here is r0 = 1. Thus, the conditional fluxdensity Q(v|ui) is

Q(v|ui) = Qm(ui)δ(v − um) +Qi(v)δ(v − ui). (26)

The global flux density Q(v) = 〈Q(v|ui)〉 is obtained byaveraging over the ensemble of inclusion velocities ui suchthat

Q(v) = 〈Qm(ui)〉δ(v − um) + 〈Qi(ui)〉vpi(v)

〈ui〉. (27)

The PDF ps(v) of particle velocities is obtained by nor-malizing the flux density Q(v) such that

ps(v) = (1− α)δ(v − um) + αvpi(v)

〈ui〉. (28)

where we defined

α =〈Qi(ui)〉

〈Qm(ui)〉+ 〈Qi(ui)〉. (29)

b. Transition Time Distribution The PDF of tran-sition times consists of the distribution of transit timesτm through the matrix, denoted by ψm(t) and transittimes τi through the inclusions, denoted by ψi(t). Thecharacteristic transit time through the matrix is givenby τ0 = `0/um, while the minimum transition time isrelated to the maximum flow velocity 2um. Thus it isτ0/2. In order to account for the flow variability in thematrix and its effect on particle dispersion, the distri-bution ψm(t) of transition times through the matrix ismodeled by a truncated exponential distribution as

ψm(t) = τ−10 exp[−(t− τ0/2)/τ0]H(t− τ0/2), (30)

where H(t) denotes the Heaviside step function. Thetransit times through the inclusions are estimated herein terms of an effective transition length `i of the inclu-sion. This effective transition length depends in generalon the conductivity contrast. For a high conductivitycontrast, i.e., k � 1, the velocity changes abruptly atthe interface between inclusion and matrix. At the hor-izontal centerline, the velocity contrast at a distance ∆from the interface can be approximated by uo(r1, 0)/ui ≈1 + ∆(1 − k)/k. Thus, for small 〈k〉 � 1, the effectivelength `i is approximated by the average transition lengthacross the disc-shaped inclusion as

`i =2

π

π∫0

dϕ cos(ϕ) =4

π. (31)

Recall that the dimensionless inclusion radius is r0 = 1and that the transition length depends on the distancefrom the centerline. For lower conductivity contrastsbetween matrix and inclusion, the velocity outside theinclusion is similar to the inclusion velocity and we set

li = `0. The transit time through the inclusion then isgiven by τi = `i/ui. As derived above the PDF of particlevelocities in the inclusions is given by the flux weightedvpi(v)/〈ui〉. Thus, we obtain for ψi(t)

ψi(t) =`2i

t3〈ui〉pi(`i/t). (32)

The transition time distribution over a unit cell is thengiven by

ψ(t) = (1− α)ψm(t) + αψi(t− t′). (33)

The distribution of long transition time is dominatedby the low end of the conductivity spectrum, at whichwe can set

ui ≈ 2umk, (34)

see (8b). Thus, for k � 1, the inclusion velocity is lin-early related to the inclusion conductivity so that thevelocity PDF at small velocities can be obtained fromthe PDF of conductivities as

pi(v) ≈ 1

2umpk[v/(2um)]. (35)

This allows to map the PDF of conductivity through (32)directly to the PDF of transition times,

ψi(t) ≈`2i

2umt3pk[`i/(2umt)], (36)

or in other words allows to express a transport attributein terms of a medium property.

IV. FIRST-PASSAGE TIME DISTRIBUTIONS

In the following, we investigate the impact of broadconductivity distributions on the long-time behavior off(t, xc) and thus on the character of anomalous trans-port. To this end, we consider power-law distributionsthat behave as pk(k) ∝ k−γ for small conductivity valuesas well as broad lognormal distributions.

Within the CTRW approach derived in the previoussection, first passage time distributions can be obtainedstraightforwardly from the Laplace transforms of (23)as f∗(λ, xc) = ψ∗(λ)nxc . Using the Laplace transformof (33) and (30)

f∗(λ, xc) =

[(1− α)

exp(−λτ0/2)

1 + λτ0+ αψ∗i (λ)

]nxc

. (37)

In the limit of times much larger than τ0, and equivalentlyλτ0 � 1, we approximate the latter by

f∗(λ, xc) ≈ exp {nxcln [1− α+ αψ∗i (λ)]} . (38)

Page 6: Mechanisms of Anomalous Dispersion in Flow Through ... › 2016 › 10 › inclusionsctrw_preprint.pdfpredictive modeling of large scale transport. I. INTRODUCTION Anomalous dispersion

6

A. Power-Law Conductivity Distribution

We consider the doubly truncated power-law conduc-tivity distribution

pk(k) =1− γ

1− k1−γc

k−1c

(k

kc

)−γ, (39)

for kc ≤ k ≤ 1. For illustration, we consider the valuesγ = −3/2 and γ = −1/2. The corresponding velocitydistribution is given by (15). As outlined above, we fo-cus here on the asymptotic behavior in order to studythe anomalous character of particle transport. The long-time behavior is dominated by the small particle veloci-ties and thus through (34) to small hydraulic conductivi-ties. Thus (39) implies here that the velocity distributionbehaves as pi(v) ∝ (v/um)−γ , see (35), and from (36)that the transition time PDF scales as

ψi(t) ∝ (t/τa)γ−3 (40)

for t � τc, where we defined the cut-off time scale τc =`i/(2umkc), which corresponds to the lower conductivitycut-off kc. Furthermore, we define the time scale τa =`i/(2um), which corresponds to the upper conductivitycut-off of 1. We consider here 0 < γ < 2. For 1 < γ < 2,the Laplace transform of the transition time distributioncan be expanded as [22]

ψ∗i (λ) = 1− aγ(λτa)2−γ (41)

for τ−1c � λ� τ−1a , while for 0 < γ < 1, we obtain [22]

ψ∗i (λ) = 1− λτav + bγ(λτa)2−γ , (42)

where we defined the mean transition time τav = `i/〈ui〉.Inserting the expansion (41) into (38), we obtain

f∗(λ, xc) ≈ exp[−nxc

αaγ(λτa)2−γ], (43)

which is a skewed Levy-stable density. Its inverseLaplace transform has the scaling form f(t, xc) =(αaγxc)

−1/(2−γ)f0[t/(αaγxc)1/(2−γ)]. The scaling func-

tion f0(t) has the Laplace transform f∗0 (λ) =exp[−(λτa)2−γ ]. The long time behavior of f(t, xc) isgiven by f(t, xc) ∝ tγ−3. Inserting the expansion (42)into (38), we obtain

f∗(λ, xc) ≈ exp{−nxcα[λτav − bγ(λτa)2−γ ]

}, (44)

Thus, the long-time behavior for the f(t, xc) is also givenby f(t, xc) ∝ tγ−3.

Figure 2 shows the first-passage time distributionsobtained from direct numerical simulations of particletransport in the flow through heterogeneous conductiv-ity fields characterized by the point distribution (39) forγ = 3/2 and γ = 1/2. The simulation data are comparedto the predictions of the corresponding CTRW model de-scribed in Section III. The late time power-law scalingis indicated by the dashed lines. The CTRW model pro-vides an accurate prediction of the late time scaling of the

10-10

10-9

10-8

10-7

10-6

10-5

10-4

10-3

104

105

106

107

fptd

t

10-10

10-9

10-8

10-7

10-6

10-5

10-4

10-3

104

105

106

fptd

t

FIG. 2. First passage time distributions obtained from (trian-gles) direct numerical simulation of particle transport in theheterogeneous porous medium and (solid lines) the predictionof the CTRW model, for the power-law k-distribution (39)with kc = 10−5 and (top panel) γ = 3/2 or respectively(bottom panel) γ = 1/2. The dashed lines indicate thepower-law behavior ∝ tγ−3 expected in the intermediate timeregime. The CTRW simulation parameters are given in Ap-pendix A 2. The direct numerical simulations are describedin Appendix A 1.

first passage time distributions. The late time scaling isdirectly related to the behavior of pk(k) for k � 1. Thepeak behavior is captured satisfactorily by the truncatedexponential distribution of particle transit times in thematrix.

B. Lognormal Conductivity Distribution

We consider now the truncated lognormal distributionof conductivities

pk(k) =

√2

πσ2

exp[− (ln k−µ)2

2σ2

]kerfc(µ/

√2σ2)

(45)

for 0 < k < 1. The transition time distribution ψi(t) isgiven by (32) in terms of the velocity distribution throughthe inclusions. As pointed out above, for k � 1 we may

Page 7: Mechanisms of Anomalous Dispersion in Flow Through ... › 2016 › 10 › inclusionsctrw_preprint.pdfpredictive modeling of large scale transport. I. INTRODUCTION Anomalous dispersion

7

set ui ≈ 2umk, see (8b). This allows to relate the velocityPDF pi(v) to the conductivity PDF according to (35) andthe transition time PDF ψi(t) to pk(k) through (36). Itfollows that the transition time PDF ψi(τ) has itself theform of a truncated lognormal distribution. Thus, un-like in the previous section, where the CTRW predicts apower-law behavior of the first passage time distributionas a consequence of the generalized central limit theoremhere this is not the case.

Figure 3 shows first-passage time distributions ob-tained from direct numerical simulations in conductivityfields characterized by the lognormal conductivity distri-bution (45) for σ2 = 11.4 and µ = −9.23 and µ = 2.3.The CTRW model developed in Section III provides agood prediction of the tailing behavior and captures thepeak behavior satisfactorily.

Edery et al. [26] proposed to fit a power-law to thelow-k end of pk(k) corresponding to the time regime, forwhich a prediction is desired. The resulting power-lawapproximation may then be used to make an approxi-mation on the tailing behavior of the first-passage timedistribution. Note that (45) may be expanded into apower-law around any k0 < 1 as

pk(k) ∝ k−γ , γ =σ2 − µ+ ln k0

σ2. (46)

This can be readily seen by expanding ln pk(k) aroundln k0 up to linear order. For small k, the first-passagetime τi through the inclusions are related to k as τi =`i/(2umk). This means first passage times of the orderof a t0 correspond to

k0 ∼ `i/(2umt0). (47)

Thus, (46) together with (47) may be used as a roughapproximation for the tail scaling around a first-passagetime t0. The dashed lines in Figure 3 indicate the scal-ing approximation around t0 = 105. For σ2 = 11.4 andµ = −9.23, we obtain from (46) that γ = 1.1. Forσ2 = 11.4 and µ = −2.3 we obtain γ = 0.4. These esti-mates of γ provide a good approximation to the tailingbehavior around t0 = 105. Note, however, that unlike inthe previous section, here, the first passage time distri-butions do not exhibit power-law tails because the firstpassage times, here are obtained through the summationof lognormally distributed random variables.

V. DISPERSION

In the previous section, we have demonstrated the pre-dictive capabilities of the CTRW model derived in Sec-tion III for global particle transport in a heterogeneousmedium characterized by a broad distribution of hy-draulic conductivities. In this section, we use this modelin order to analyze particle dispersion and its controls interms of the distribution of hydraulic conductivities.

10-11

10-10

10-9

10-8

10-7

10-6

10-5

10-4

10-3

104

105

106

107

fptd

t

10-13

10-12

10-11

10-10

10-9

10-8

10-7

10-6

10-5

10-4

10-3

104

105

106

107

fptd

t

FIG. 3. First passage time distributions obtained from (trian-gles) direct numerical simulation of particle transport in theheterogeneous porous medium, see Appendix A 1 and (solidlines) the prediction of the CTRW model, for the lognormalk-distribution (45) with σ2 = 11.4 and (top panel) µ = −9.23or respectively (bottom panel) µ = −2.3. The CTRW sim-ulation parameters are given in Appendix A 2. The dashedlines indicate the approximate power-law behavior obtainedfrom (46) and (47).

Dispersion is measured in terms of the centered meansquare displacement

κ(t) = m2(t)−m1(t)2. (48)

The first and second displacement moments are definedin the CTRW framework as

m1(t) = 〈xnt〉, m2(t) = 〈x2nt〉, (49)

where nt = sup(n|tn ≤ t). We obtain explicit Laplacespace expressions for m∗1(λ) and m∗2(λ) in terms of thetransition time distribution ψ∗(λ) [22, 52]

m∗1(λ) =`0λ2

λψ∗(λ)

1− ψ∗(λ)(50)

m∗2(λ) =`20λ2

λψ∗(λ)

1− ψ∗(λ)+ 2

`20λ3

λ2ψ∗(λ)

[1− ψ∗(λ)]2. (51)

The Laplace transform of the transition time distribution

Page 8: Mechanisms of Anomalous Dispersion in Flow Through ... › 2016 › 10 › inclusionsctrw_preprint.pdfpredictive modeling of large scale transport. I. INTRODUCTION Anomalous dispersion

8

here is given by

ψ∗(λ) = (1− α)ψ∗m(λ) + αψ∗i (λ). (52)

For transition time distributions characterized by finitemean and mean square transition times, κ(t) increaseslinearly with time, κ(t) = 2Det. The effective disper-sion coefficient is given in terms of the first and secondmoments of ψ(t) as [22]

De =`202

〈τ2〉 − 〈τ〉2

〈τ〉3. (53)

For the composite transition time distribution (33) thei–th moment 〈τ i〉 of the transition time is given by

〈τ i〉 = (1− α)〈τ im〉+ α〈τ ii 〉. (54)

For the first and second moments of the mobile transitiontime, we obtain from (30)

〈τm〉 = τ0 + τ0/2, 〈τ2m〉 = τ0/22 + 2τ0(τ0 + τ0/2).

(55)

For the truncated power-law distribution (39), thetransition time distribution ψi(t) can be approximatedby the truncated power-law

ψi(t) =2− γτa

(t/τa)γ−3

1− (τc/τa)γ−2(56)

for τa < t < τc. The cut-off time τc is related to thesmallest conductivity value as τc = `i/(2umkc).

A. Intermediate Time Regime

We first consider dispersion in the intermediate timeregime τa � t � τc. It behaves as ψ(t) ∝ tγ−3. In thistime regime, dispersion is anomalous and characterizedby the following scalings. For 1 < γ < 2, one obtains [22]

κ(t) ∝ t4−2γ . (57)

For 0 < γ < 1, one finds

κ(t) ∝ t1+γ . (58)

The second centered moment of the particle distributionincreases superdiffusively.

B. Long Time Regime

Now we investigate the dependence of the effective dis-persion coefficient (53) on the cut-off time scale τc andequivalently on the minimum conductivity scale kc. Thefirst and second moments of the transition time distribu-tion ψi(t) are obtained from (56) as

〈τi〉 = τa2− γ1− γ

1− (τc/τa)γ−1

1− (τc/τa)γ−2(59a)

〈τ2i 〉 = τ2a2− γγ

(τc/τa)γ − 1

1− (τc/τa)γ−2. (59b)

We furthermore note that α ∝ 〈ui〉 ≈ 2um〈k〉, wherethe mean conductivity over the inclusions is obtainedfrom (39) as

〈k〉 =1− γ2− γ

1− k2−γc

1− k1−γc

(60)

This means, for 0 < γ < 1, the mean conductivity con-verges to the finite value 〈k〉 = (1 − γ)/(2 − γ) in thelimit kc → 0, while for 1 < γ < 2 it goes toward 0 as〈k〉 ∝ kγ−1c . In the following, we quantify the behaviorof the dispersion coefficient (53) in the limit of τc � τa.a. 0 < γ < 1 We first consider the case 0 < γ < 1.

In this case, we obtain for the transition time momentsin leading order

〈τi〉 = τa2− γ1− γ

. (61)

〈τ2i 〉 = τ2a2− γγ

(τc/τa)γ . (62)

For the mean conductivity, we obtain

〈k〉 =1− γ2− γ

. (63)

This means, both α and 〈τ〉 are constant in the limit oflarger τc. Thus, dispersion coefficient (53) is dominatedby 〈τ2i 〉, which increases as τγc . Thus, in leading order,we can write

De =`202

2− γγ

ατ2a〈τ〉2

(τc/τa)γ ∝ k−γc (64)

This means, the effective dispersion coefficient increasesmonotonically with increasing conductivity contrast be-tween the inclusions.b. 1 < γ < 2 Unlike in the case γ < 1, here kc → 0 is

a singular limit for the conductivity distribution becausethe normalizability of pk(k) depends on the finiteness ofkc. We obtain for the mean conductivity 〈k〉 in the limitkc � 1 the leading behavior

〈k〉 =γ − 1

2− γkγ−1c . (65)

This means, the mean inclusion conductivity, and there-fore the mean velocity through the inclusions go to 0 inthe limit kc → 0. In this limit, the inclusions are on av-erage impermeable. Based on (65), we may now write αas

α = α(τc/τa)1−γ , α = α(τc/τa)γ−1, (66)

where α → constant in the limit kc → 0. For the meanand mean squared transition times through the inclu-sions, we obtain from (59) in leading order for τc � τa

〈τi〉 = τa2− γγ − 1

(τc/τa)γ−1 (67)

〈τ2i 〉 = τ2a2− γγ

(τc/τa)γ . (68)

Page 9: Mechanisms of Anomalous Dispersion in Flow Through ... › 2016 › 10 › inclusionsctrw_preprint.pdfpredictive modeling of large scale transport. I. INTRODUCTION Anomalous dispersion

9

Thus, we obtain in leading order for the effective disper-sion coefficient

De =`202

2− γγ

τ2a〈τ〉3

ατc/τa ∝ k−1c . (69)

This means, it increases linearly with the cut-off timescale τc and is inversely proportional to the minimumconductivity kc.

Note that the scaling behaviors (64) and (69) areuniversal for any conductivity distribution that showsa (truncated) power-law behavior for kc < k � 1 aspk(k) ∝ k−γ . Note also that for any conductivity distri-bution that has the scaling form

pk(k) =1

kcfk(k/kc), (70)

with fk(k) a scaling function that is normalized to 1, onefinds that α ∝ kc, 〈τi〉 ∝ k−1c and 〈τ2i 〉 ∝ k−1c . Thus inthe limit of kc → 0 one finds here that De ∝ k−1c , seealso Ref. [40].

VI. SUMMARY AND CONCLUSIONS

We investigated the mechanisms of anomalous disper-sion in the flow through heterogeneous porous media. Tothis end, we considered a medium composed of equallysized inclusions embedded in a porous matrix. The hy-draulic conductivities of the inclusions are broadly dis-tributed with tails toward small values. Such media canbe seen as idealizations of highly heterogeneous porousmedia that have a characteristic correlation scale. Theconductivities in the inclusions is mapped onto the flowvelocities through an analytical expression. The back-ground velocity is related to the effective conductivityof the medium, which is obtained from the Maxwell for-mula.

Based on these results we formulate the purely advec-tive movement of solute particles in terms of travel dis-tance along streamlines, which renders the equations ofmotion a CTRW, in which the transition length is fixedthrough the characteristic heterogeneity length scale andthe transition time is related to the particle velocities.We define unit cells as a domain that contains a singleinclusion. When crossing a unit cell particles can by-pass the inclusion or pass through. The probabilitiesfor the respective path are given by the fluxes thoughmatrix and inclusions. Thus the PDF of particle veloci-ties is obtained from the PDF of flow velocities throughflux weighting. This provides a direct link between themedium properties through the explicit relation betweenflow velocities and hydraulic conductivity.

The derived CTRW model is then used to predict firstpassage time distributions obtained from particle track-ing simulations in the flow through heterogeneous mediacharacterized by different heavy-tailed distributions ofhydraulic conductivities. Specifically, we consider trun-cated power-law behaviors at small conductivities and

broad truncated lognormal distributions. The power-lawin the conductivity is directly mapped onto a power-lawof transition times, which predicts the power-law behav-ior observed in the direct numerical simulations as a con-sequence of the generalized central limit theorem. Alsofor the lognormal conductivity distributions, we observebroad distributions of first-passage times. Here however,they are not power-laws because they result from an ad-dition of lognormally distributed random variables. Nev-ertheless, for very broad conductivity distributions, thelognormal PDF may be approximated by a power-lawwith an exponent determined by the mean and varianceof the log-hydraulic conductivity. This may be used todescribe the tailing behavior in certain time ranges basedon the expressions derived for power-law distributions.The heavy tails are eventually tempered at time scalesthat correspond to the smallest hydraulic conductivityvalues.

Based on the derived CTRW model, we investigatethe dispersion behavior in terms of the second centeredmoments of the particle distribution, or centered meansquare displacement. The CTRW predicts anomalousdispersion characterized by non-linear evolutions of thecentered mean square displacement. Specifically, for thetruncated power-law distribution of hydraulic conductiv-ity CTRW predicts a power-law evolution. At asymp-totically long times, much larger than the cut-off timescale, dispersion becomes normal as a consequence of thecentral limit theorem. The corresponding dispersion co-efficients are quantified in terms of the conductivity dis-tributions.

The medium under consideration is d = 2 dimensional.The developed CTRW model, however, can be straight-forwardly generalized to d = 3 dimensional media basedon similar analytical expressions for the flow velocity in-side the inclusions [41]. Furthermore, the present studyconsiders purely advective transport. Diffusion into low-conductivity inclusions would introduce an additionalcut-off scale for the transition time distribution if thecharacteristic diffusion time over the inclusion is smallerthan the largest advection time scale. These behaviorsare subject of ongoing research.

In conclusion, the derived CTRW model provides apredictive description of transport through highly hetero-geneous media based on the distribution of hydraulic con-ductivity and characteristic heterogeneity length scales.The concrete heterogeneity model is a caricature ofhighly heterogeneous porous media characterized by fi-nite correlation scales. Thus, the presented results shedlight on the fundamental mechanisms of anomalous dis-persion and its relations with the medium and flow prop-erties.

Acknowledgments

The authors gratefully acknowledge financial supportby the Swiss National Science Foundation (SNF) through

Page 10: Mechanisms of Anomalous Dispersion in Flow Through ... › 2016 › 10 › inclusionsctrw_preprint.pdfpredictive modeling of large scale transport. I. INTRODUCTION Anomalous dispersion

10

grant 200021 132304 and by Nagra, Wettingen. MD ac-knowledges the support of the European Research Coun-cil (ERC) through the project MHetScale (617511).

Appendix A: Numerical Simulations

In the following, we give some details on the numericalsimulations. First, we describe the setup of the direct nu-merical simulations of flow and particle transport in theconductivity fields characterized by low-conductivity in-clusions embedded in a higher conductive matrix. Then,we give the details for the particle tracking simulationthat implement the CTRW model developed in Sec-tion III.

1. Flow and Transport Simulations

We consider a regular field size of 512 by 512 cells, con-sisting of a highly conductive matrix with k0 = 1 and aset of circular low conductivity inclusions with radii r0 =15 and centers in (xi, yj) : yj = 32j + si, xi = 32i + 16,where si is a random shift (Figure 1). Hydraulic con-ductivities k inside inclusions are independent identicallydistributed random variables with the PDF pk(k).

Flow is driven by a constant hydraulic head gradientbetween the inlet boundary at x = 0 and the outletboundary at x = 512, |∇h| = ∆h/512 = [h(512, y) −h(0, y)]/512 = −0.1. The flow field is solved numericallyusing a finite difference scheme with a unit discretization∆x = ∆y = 1 [53]. This means 30 cells per inclusion di-ameter. Such a fine discretization is required due to thehigh conductivity contrast between inclusions and ma-trix.

Transport is solved by particle tracking based on theadvection equation (3) using the scheme of Pollock [54].The Pollock algorithm interpolates the flow velocitywithin a finite difference cell bi-linearly to guarantee thatthe divergence of the flow velocity is 0 [55]. This interpo-lation is necessary because the finite difference methodgives only the flow velocities perpendicular to the cell fa-cies. The trajectory within the cell is then determinedanalytically by integration of the advection equation (3),which gives the Pollock integral. The particle trackingsimulations use 105 particles, which are injected propor-tional to the flux at the surface at x = 30. The observa-tion plain is the right boundary of the domain at x = 512.The first passage times are determined according to (4)through an average over the 105 particles in single re-alizations and between 103 realization for each randomfield under consideration.

2. CTRW Simulations

The CTRW model is based on the stochastic recursionrelation (18). The first passage times are determined

according to (23). The transition time distribution ψ(t) isgiven by (33). It requires the parameters α given by (29),the matrix velocity um given by (13), which depends onthe effective conductivity ke given by (10), the averageinclusion velocity 〈ui〉and the effective length `i. Thevolume fraction of the inclusion is χ = 0.69. The fluxweighted velocity distributions vipi(v)/〈vi〉 are generatedby the rejection method. To this end, we consider thecorresponding flux-weighted conductivity distribution

pk(k) =k

1 + kpk(k)/〈k/(1 + k)〉 (A1)

and compare it to cpk(k), where c is chosen such thatthat cpk(k) ≥ pk(k). The CTRW simulations reportedin Figures 2 and 3 use 107 particles.

a. Power-law Conductivity Distribution

For the power-law conductivity distribution (39) weobtain the following parameters. For γ = 3/2, the ef-fective conductivity is ke = 0.186, the matrix velocity isum = 0.059, the mean inclusion velocity is 〈ui〉 = 0.00036the trapping frequency is α = 0.036. The effective im-mobile length is here set to li = 4r/π = 19 because themean conductivity in the inclusion is 〈k〉 = 0.0032� 1.

For γ = 1/2, the effective conductivity is ke = 0.436,the matrix velocity is um = 0.072, the mean inclu-sion velocity is 〈ui〉 = 0.031 the trapping frequency isα = 0.8. The effective immobile length is here set toli = 32 because the mean conductivity in the inclusionis 〈k〉 = 0.334. Regarding the reasoning for the choice ofthe effective transition length li, see also the discussionin Section III B.

b. Log-normal Conductivity Distribution

For the lognormal conductivity distribution (45) weobtain the following parameters. For µ = −9.23 and σ2 =11.4, the effective conductivity is ke = 0.19, the matrixvelocity is um = 0.059, the mean inclusion velocity is〈ui〉 = 0.00073 the trapping frequency is α = 0.085. Theeffective immobile length is here set to li = 4r0/π = 19because the mean conductivity in the inclusion is 〈k〉 =0.0077� 1.

For µ = −2.3 and σ2 = 11.4, the effective conductivityis ke = 0.29, the matrix velocity is um = 0.064, the meaninclusion velocity is 〈ui〉 = 0.012 the trapping frequencyis α = 0.62. The effective immobile length is here setto li = 23 because the mean conductivity in the inclu-sion is 〈k〉 = 0.14, this means the contrast between thematrix conductivity and average inclusion conductivityis relatively low, see also the discussion in Section III B.

Page 11: Mechanisms of Anomalous Dispersion in Flow Through ... › 2016 › 10 › inclusionsctrw_preprint.pdfpredictive modeling of large scale transport. I. INTRODUCTION Anomalous dispersion

11

[1] Y. Hatano and N. Hatano, “Dispersive transport of ionsin column experiments: An explanation of long-tailedprofiles,” Water Resour Res 34 (5), 1027–1033 (1998).

[2] B. Berkowitz, A. Cortis, M. Dentz, and H. Scher, “Mod-eling non-fickian transport in geological formations as acontinuous time random walk,” Rev. Geophys. 44 (2006),10.1029/2005RG000178.

[3] S. P. Neuman and D. M. Tartakovsky, “Perspective ontheories of anomalous transport in heterogeneous media,”Adv. Water Resour. 32, 670–680 (2008).

[4] P. K. Kang, T. Le Borgne, M. Dentz, O. Bour, andR. Juanes, “Impact of velocity correlation and distri-bution on transport in fractured media: field evidenceand theoretical model,” Water Resour. Res. 51, 940–959(2015).

[5] Aldo Fiori and Matthew W. Becker, “Power law break-through curve tailing in a fracture: The role of advec-tion,” J. Hydrol. 525, 706–710 (2015).

[6] A Bellin, A Rinaldo, WJP Bosma, SEATM Vanderzee,and Y Rubin, “Linear equilibrium adsorbing solute trans-port in pphysically and chemically heterogeneous porousformations. 1. analytical solutions,” Water Resour. Res.29, 4019–4030 (1993).

[7] D. Bolster and M. Dentz, “Anomalous dispersion inchemically heterogeneous media induced by long-rangedisorder correlation,” J. Fluid Mech. 695, 366–389(2012).

[8] I. Neretnieks and A. Rasmuson, “An approach to mod-eling radionuclide migration in a medium with stronglyvarying velocity and block sizes along the flow path,”Water Resour. Res. 20, 1823–1836 (1984).

[9] A. Comolli, J. Hidalgo, C. Moussey, and M. Dentz,“Non-fickian transport under heterogeneous advectionand mobile- immobile mass transfer,” Transp. PorousMedia 113 (2016), 10.1007/s11242-016-0727-6.

[10] M. W. Becker and A. Shapiro, “Tracer transport in frac-tured crystalline rock: Evidence of nondiffusive break-through tailing,” Water Resour. Res. 36, 1677–1686(2000).

[11] G. Di Donato, E. Obi, and M. J. Blunt, “Anoma-lous transport in heterogeneous media demonstrated bystreamline-based simulation,” Geophys. Res. Lett. 30,1608 (2003).

[12] MJ Ronayne and SM Gorelick, “Effective permeability ofporous media containing branching channel networks,”Phys Rev. E 73 (2006), 10.1103/PhysRevE.73.026305.

[13] A. R. Tyukhova, W. Kinzelbach, and M. Willmann, “De-lineation of connectivity structures in 2-d heterogeneoushydraulic conductivity fields,” Water Resour. Res. 51,5846–5854 (2015).

[14] A.R. Tyukhova and M. Willmann, “Connectivity metricsbased on the path of smallest resistance,” Advances inWater Resources 88, 14–20 (2016).

[15] E.W. Montroll and G.H. Weiss, “Random walks on lat-tices. ii,” J. Math. Phys. 6, 167 (1965).

[16] H. Scher and M. Lax, “Stochastic transport in a dis-ordered solid. I. Theory,” Phys. Rev. B 7, 4491–4502(1973).

[17] M. F. Shlesinger, B. J. West, and J. Klafter, “Levydynamics of enhanced diffusion: Application to turbu-lence,” Phys. Rev. Lett. 58, 1100–1103 (1987).

[18] Ralf Metzler and Joseph Klafter, “The random walk’sguide to anomalous diffusion: a fractional dynamics ap-proach,” Phys. Rep. 399, 1–77 (2000).

[19] J. Klafter and I. Sokolov, “Anomalous diffusion spreadsits wings,” Phys. World 18, 29–32 (2005).

[20] Simon Thalabard, Giorgio Krstulovic, and Jeremie Bec,“Turbulent pair dispersion as a continuous-time randomwalk,” J. Fluid Mech. 755, R4 (2014).

[21] B. Berkowitz and H. Scher, “Anomalous transport in ran-dom fracture networks,” Phys. Rev. Lett. 79, 4038–4041(1997).

[22] M. Dentz, A. Cortis, H. Scher, and B. Berkowitz, “Timebehavior of solute transport in heterogeneous media:transition from anomalous to normal transport,” Adv.Water Resour. 27, 155–173 (2004).

[23] Tanguy Le Borgne, Marco Dentz, and Jesus Carrera,“Lagrangian statistical model for transport in highly het-erogeneous velocity fields,” Phys. Rev. Lett. 101 (2008),10.1103/PhysRevLett.101.090601.

[24] B Berkowitz and H Scher, “Theory of anomalous chem-ical transport in random fracture networks,” Phys. Rev.E 57, 5858–5869 (1998).

[25] P. K. Kang, M. Dentz, T. Le Borgne, and R. Juanes,“Spatial markov model of anomalous transport throughrandom lattice networks,” Phys. Rev. Lett. 107, 180602(2011).

[26] Y. Edery, A. Guadagnini, H. Scher, and B. Berkowitz,“Origins of anomalous transport in heterogeneous me-dia: Structural and dynamic controls,” Water ResourcesResearch 50, 1490–1505 (2014), cited By 17.

[27] R. Haggerty and S.M. Gorelick, “Multiple-rate mass-transfer for modeling diffusion and surface-reactions inmedia with pore-scale heterogeneity,” Water Resour. Res.31, 2383–2400 (1995).

[28] J. Carrera, X. Sanchez-Vila, I. Benet, A. Medina,G. Galarza, and J. Guimera, “On matrix diffusion: for-mulations, solution methods, and qualitative effects,”Hydrogeology Journal 6, 178–190 (1998).

[29] R. Haggerty, S. A. McKenna, and L. C. Meigs, “On thelate time behavior of tracer test breakthrough curves,”Water Resour. Res. 36, 3467–3479 (2000).

[30] M. Dentz and B. Berkowitz, “Transport behavior of apassive solute in continuous time random walks andmultirate mass transfer,” Water Resour. Res. 39, 1111(2003).

[31] R. Schumer, D. A. Benson, M. M. Meerschaert, andB. Baeumer, “Fractal mobile/immobile solute trans-port,” Water Resour Res 39(10), 1296 (2003).

[32] D. A. Benson and M. M. Meerschaert, “A simpleand efficient random walk solution of multi-rate mo-bile/immobile mass transport equations,” Adv. WaterResour. 32, 532–539 (2009).

[33] M. Willmann, J. Carrera, and X. Sanchez-Vila, “Trans-port upscaling in heterogeneous aquifers: What physicalparameters control memory functions?” Water Resour.Res. 44 (2008), 10.1029/2007WR006531.

[34] Y. Zhang, C. T. Green, and G. E. Fogg, “The impactof medium architecture of alluvial settings on non-fickiantransport,” Adv. Water Resour. 54, 78–99 (2013).

[35] Alina R. Tyukhova and Matthias Willmann, “Conserva-tive transport upscaling based on information of connec-

Page 12: Mechanisms of Anomalous Dispersion in Flow Through ... › 2016 › 10 › inclusionsctrw_preprint.pdfpredictive modeling of large scale transport. I. INTRODUCTION Anomalous dispersion

12

tivity,” Water Resources Research 52, n/a–n/a (2016).[36] J. P. Bouchaud and A. Georges, “Anomalous diffusion

in disordered media: Statistical mechanisms, models andphysical applications,” Phys. Rep. 195, 127–293 (1990).

[37] M. Dentz and A. Castro, “Effective transport dynamicsin porous media with heterogeneous retardation proper-ties,” Geophys. Res. Lett. 36, L03403 (2009).

[38] M. Dentz and D. Bolster, “Distribution- versuscorrelation-induced anomalous transport in quenchedrandom velocity fields,” Phys. Rev. Lett. 105, 244301(2010).

[39] P. K. Kang, Dentz. M., and R. Juanes, “Predictability ofanomalous transport on lattice networks with quencheddisorder,” Phys. Rev. E 83, 030101(R) (2011).

[40] I. Eames and J. W. M. Bush, “Longitudinal dispersionby bodies fixed in potential flow.” Proc. R. Soc. Lond. A455, 3665–3686 (1999).

[41] A. Fiori, I. Jankovic, and G. Dagan, “Modeling flowand transport in highly heterogeneous three-dimensionalaquifers: Ergodicity, gaussianity, and anomalous behav-ior - 2. approximate semianalytical solution,” Water. Re-sour. Res. 42, W06D13 (2006).

[42] A. Fiori, I. Jankovic, Dagan G., and V. Cvetkovic, “Er-godic transport through aquifers of non-gaussian log con-ductivity distribution and occurrence of anomalous be-havior,” Water Resour. Res. 43, W09407 (2007).

[43] A. Fiori, G. Dagan, I. Jankovic, and A. Zarlenga,“The plume spreading in the made transport experiment:Could it be predicted by stochastic models?” Water Re-sour. Res. 49, 2497–2507 (2013).

[44] V. Cvetkovic, A. Fiori, and G. Dagan, “Solute transportin aquifers of arbitrary variability: A time-domain ran-

dom walk formulation,” Water Resources Research 50,5759–5773 (2014), cited By 6.

[45] J. Bear, Dynamics of fluids in porous media (AmericanElsevier, New York, 1972).

[46] S. W. Wheatcraft and F. Winterberg., “Steady state flowpassing through a cylinder a permeability different fromthe surrounding medium,” Water Resour. Res. 21, 1923–1929 (1985).

[47] P. Renard and G. de Marsily, “Calculating equivalentpermeability: A review,” Adv. Water Resour. 20, 253–278 (1997).

[48] X. Sanchez-Vila, A. Guadagnini, and J. Carrera, “Rep-resentative hydraulic conductivities in saturated ground-water flows,” Rev. Geophys. 44, RG3002 (2006).

[49] S. Torquato, Random Heterogeneous Materials (Springer,2002).

[50] A. Fiori, I. Jankovic, and G. Dagan, “Effective conduc-tivity of heterogeneous multiphase media with circularinclusions,” Phys. Rev. Lett. (2005).

[51] M. Dentz, P. K. Kang, and T. Le Borgne, “Continuoustime random walks for non-local radial solute transport,”Adv. Wat. Res. 82, 16–26 (2015).

[52] M. F. Shlesinger, “Asymptotic solutions of continuous-time random walks,” J. Stat. Phys. 10, 421–434 (1974).

[53] M.G. McDonald and A.W. Harbaugh, A modular three-dimensional finite-difference groundwater flow model.(Washington : Scientific Publications, 1984).

[54] David W. Pollock, “Semianalytical computation of pathlines for finite difference models,” Ground Water 26, 743–750 (1988).

[55] C. Cordes and W. Kinzelbach, “Continuous groundwa-ter velocity fields and path lines in linear, bilinear, andtrilinear finite elements,” Water Resour. Res. (1992).