Top Banner
Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory F. Cornaton * , P. Perrochet CHYN, Institute of Geology, University of Neucha ˆtel, Emile-Argand 11, CH-2007, Neucha ˆtel, Switzerland Received 22 October 2004; received in revised form 12 October 2005; accepted 13 October 2005 Available online 1 December 2005 Abstract We present a methodology for determining reservoir groundwater age and transit time probability distributions in a deterministic manner, considering advective–dispersive transport in steady velocity fields. In a first step, we propose to model the statistical dis- tribution of groundwater age at aquifer scale by means of the classical advection–dispersion equation for a conservative and non- reactive tracer, associated to proper boundary conditions. The evaluated function corresponds to the density of probability of the random variable age, age being defined as the time elapsed since the water particles entered the aquifer. An adjoint backward model is introduced to characterize the life expectancy distribution, life expectancy being the time remaining before leaving the aquifer. By convolution of these two distributions, groundwater transit time distributions, from inlet to outlet, are fully defined for the entire aquifer domain. In a second step, an accurate and efficient method is introduced to simulate the transit time distribution at discharge zones. By applying the reservoir theory to advective–dispersive aquifer systems, we demonstrate that the discharge zone transit time distribution can be evaluated if the internal age probability distribution is known. The reservoir theory also applies to internal life expectancy probabilities yielding the recharge zone life expectancy distribution. Internal groundwater volumes are finally identified with respect to age and transit time. One- and two-dimensional theoretical examples are presented to illustrate the proposed math- ematical models, and make inferences on the effect of aquifer structure and macro-dispersion on the distributions of age, life expec- tancy and transit time. Ó 2005 Elsevier Ltd. All rights reserved. Keywords: Age; Life expectancy; Transit time; Reservoir theory; Dispersion; Laplace transforms; Finite elements 1. Introduction The knowledge of groundwater age distributions is of prime interest in many environmental issues since they depend on the intrinsic characteristics of the overall transport properties of an aquifer and its sub-systems. For instance, an important fraction of young water within a water sample can often be taken as the signa- ture of a reservoir with good turnover property. On the opposite, a considerable component of old water may reflect a poorly recharged aquifer, and/or signifi- cant internal mixing processes. The impact of a contam- ination hazard on groundwater quality can be investigated using groundwater age, since the age distri- bution provides direct information on the time required for a water particle, or a conservative tracer, to reach any critical zone that is to be protected. The fate of a solute being transported in groundwater partially depends on the time spent by the water molecules during their migration throughout the aquifer system. Reactive transport of a specific substance is also linked to groundwater age; the time spent flowing through any kind of mineral heterogeneity being a conditional factor for the development of any potential reactions. With the 0309-1708/$ - see front matter Ó 2005 Elsevier Ltd. All rights reserved. doi:10.1016/j.advwatres.2005.10.009 * Corresponding author. Fax +41 32 7182603. E-mail addresses: [email protected] (F. Cornaton), pier- [email protected] (P. Perrochet). Advances in Water Resources 29 (2006) 1267–1291 www.elsevier.com/locate/advwatres
25

Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

Apr 20, 2023

Download

Documents

Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

Advances in Water Resources 29 (2006) 1267–1291

www.elsevier.com/locate/advwatres

Groundwater age, life expectancy and transit time distributionsin advective–dispersive systems: 1. Generalized reservoir theory

F. Cornaton *, P. Perrochet

CHYN, Institute of Geology, University of Neuchatel, Emile-Argand 11, CH-2007, Neuchatel, Switzerland

Received 22 October 2004; received in revised form 12 October 2005; accepted 13 October 2005Available online 1 December 2005

Abstract

We present a methodology for determining reservoir groundwater age and transit time probability distributions in a deterministicmanner, considering advective–dispersive transport in steady velocity fields. In a first step, we propose to model the statistical dis-tribution of groundwater age at aquifer scale by means of the classical advection–dispersion equation for a conservative and non-reactive tracer, associated to proper boundary conditions. The evaluated function corresponds to the density of probability of therandom variable age, age being defined as the time elapsed since the water particles entered the aquifer. An adjoint backward modelis introduced to characterize the life expectancy distribution, life expectancy being the time remaining before leaving the aquifer. Byconvolution of these two distributions, groundwater transit time distributions, from inlet to outlet, are fully defined for the entireaquifer domain. In a second step, an accurate and efficient method is introduced to simulate the transit time distribution at dischargezones. By applying the reservoir theory to advective–dispersive aquifer systems, we demonstrate that the discharge zone transit timedistribution can be evaluated if the internal age probability distribution is known. The reservoir theory also applies to internal lifeexpectancy probabilities yielding the recharge zone life expectancy distribution. Internal groundwater volumes are finally identifiedwith respect to age and transit time. One- and two-dimensional theoretical examples are presented to illustrate the proposed math-ematical models, and make inferences on the effect of aquifer structure and macro-dispersion on the distributions of age, life expec-tancy and transit time.� 2005 Elsevier Ltd. All rights reserved.

Keywords: Age; Life expectancy; Transit time; Reservoir theory; Dispersion; Laplace transforms; Finite elements

1. Introduction

The knowledge of groundwater age distributions is ofprime interest in many environmental issues since theydepend on the intrinsic characteristics of the overalltransport properties of an aquifer and its sub-systems.For instance, an important fraction of young waterwithin a water sample can often be taken as the signa-ture of a reservoir with good turnover property. Onthe opposite, a considerable component of old water

0309-1708/$ - see front matter � 2005 Elsevier Ltd. All rights reserved.doi:10.1016/j.advwatres.2005.10.009

* Corresponding author. Fax +41 32 7182603.E-mail addresses: [email protected] (F. Cornaton), pier-

[email protected] (P. Perrochet).

may reflect a poorly recharged aquifer, and/or signifi-cant internal mixing processes. The impact of a contam-ination hazard on groundwater quality can beinvestigated using groundwater age, since the age distri-bution provides direct information on the time requiredfor a water particle, or a conservative tracer, to reachany critical zone that is to be protected. The fate of asolute being transported in groundwater partiallydepends on the time spent by the water molecules duringtheir migration throughout the aquifer system. Reactivetransport of a specific substance is also linked togroundwater age; the time spent flowing through anykind of mineral heterogeneity being a conditional factorfor the development of any potential reactions. With the

Page 2: Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

1268 F. Cornaton, P. Perrochet / Advances in Water Resources 29 (2006) 1267–1291

age information, inferences can be made on the aquiferphysical characteristics, as well as on the chemical trans-formations that contaminants may undergo. The agecan also be of importance for estimating historicalaspects related e.g. to the agricultural practices, or theland use of a particular region, which are expected tohave lead to groundwater contamination. As it is thecase throughout this work, groundwater age can be con-sidered as a fully conservative tracer. Consequently, theworst scenario with regard to a contamination case isthus chosen by solving the age transport problem.

Generally, a water sample shows a mixture of ages,which may range between several orders of magnitude,as a consequence of the reservoir geometrical andhydro-dispersive characteristics (spatial repartition ofthe hydraulic and transport parameters). Groundwaterage must, therefore, be regarded as a statistical, or prob-abilistic distribution, rather than considering it as a sin-gle absolute or average value. The dating methodscommonly provide an average value over the water sam-ple for the age of groundwater after recharge, which intheory does not represent hard data. In fact, the meanof an unknown distribution, here the distribution ofages, is not necessarily a reliable value for the mostlikely of this distribution. By far the most frequentlyused dating methods are based on the measurementsof natural tracers, such as the isotopes of radon, carbonor oxygen, and on the measurements of man-inducedatmospheric concentrations of elements such as tritium(3H), helium (3He), chloride (36Cl), krypton (85Kr), orchlorofluorocarbons (CFC), which have increased stea-dily between the 1940s and the early 1990s. The mostefficient methods for dating recent waters are the onesbased on 3H/3He, 85Kr, and CFC-measurements[63,55,25], which are known to provide age dates overa period of 40 years with an accuracy of 20% or less[11]. CFC-based ages provide a good resolution forgroundwater with relatively young components [8,6],but they do not account for the time spent by water mol-ecules within the unsaturated zone, such that in the caseof deep water table aquifers the travel time durationwithin the unsaturated zone must, therefore, be esti-mated. Absolute dating techniques involve decay ofradionuclides in groundwater, for which 14C or 36Clare classical elements that are used for dating oldgroundwater, e.g. in deep and large sedimentary basins.Groundwater age is very often used to make inferenceson aquifer parameters, groundwater recharge, flowpaths and flow velocities. However, with many of thesedating methods, the interpretation of ages is achieved bymeans of simple conceptual models and fitting the data.This involves significant simplifications of the flow andtransport processes, which may lead to erroneous inter-pretations about e.g. the past release of contaminants inaquifers. As will be discussed in this paper, mixing anddispersion are major factors, which can lead to unrepre-

sentative mean age measurements. Dating methodsremain however very useful for calibrating numericalmodels [56,58], which attempt to simulate the flow pat-terns, the flow rates, and the distribution of ages ingroundwater.

To quantify the distribution of ages in aquifers, sev-eral types of mathematical models have been developedduring the past decades of research in this field. Never-theless, in many cases age distributions are more or lessarbitrarily chosen and not deterministically calculated.Analytical lumped-parameter type models have beenhowever extensively used in the simulation of environ-mental tracer data [45,47,74,9,57,1], such as isotopicdata, which are commonly interpreted with advectivetransit time models, although isotope transport doesnot necessarily undergo advective processes only. Speci-fied transit time distributions describing piston-flow,exponential mixing, combined piston-exponential mix-ing, or dispersive mixing, are chosen to solve the inverseproblem by fitting the model on measured tracer outputdata. This procedure calls for significant simplifications,which can often not be justified, such as for instanceneglecting the reservoir structure, as well as the spatialvariability of infiltration rates [10] and aquifer flowand transport parameters. Amin and Campana [1] pro-posed to model the groundwater age mixing process bymeans of a three-parameter gamma function whichaccounts for various states of mixing ranging betweenno mixing (piston flow model) and perfect mixing (expo-nential model). Robust verifications of the applicabilityof lumped parameter models can hardly be found[36,44,28]. Transit time distributions are often obtainedwith numerical solutions by making the assumption ofpure advective motion of the groundwater particles,the particle-tracking technique being the most popularone. Purely kinematic ages ignore the effect of dispersionand mixing on age transport [22,12], and often reveal tobe ill posed in complex heterogeneous systems [71], forwhich the 3-D implementation is subject to severe tech-nical problems. The importance of including age dilu-tion processes such as dispersion and matrix diffusion,when comparison is made between modeled and mea-sured ages, has been pointed out by many authors[66,45,46]. Moreover, particle-tracking does not allowcalculation of transit time distributions since groundwa-ter volumes are not associated to simulated ages.

More elaborated quantitative approaches considerage as a mass that is transported by groundwaterthrough volume-averaged temporal moment equations[64,37,35,71], in which the product of groundwater agewith its mass (age mass qA) is the conserved quantity.Harvey and Gorelick [37], and Varni and Carrera [71]derived a set of recursive temporal moment equations,which are sequentially solved in order to simulate thetransit time distributions from the n calculatedmoments. According to Harvey and Gorelick [37], the

Page 3: Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

F. Cornaton, P. Perrochet / Advances in Water Resources 29 (2006) 1267–1291 1269

first five moments that characterize the accumulatedmass, the mean, the variance, the skewness and the kur-tosis of a breakthrough curve, respectively, may providesufficient information to summarize the entire distribu-tion. Since many natural systems reveal a multi-modal-ity of the age distribution within the reservoir and atthe discharge zones, and since the shape of this distribu-tion is a priori unknown, an infinity of moments wouldtherefore theoretically be required to construct the entiredistribution.

In the literature one can find many terms, whichrelate to a specific time spent by water molecules withinthe aquifer. Use will be made throughout this work ofthe notion of transit time as the total residence time ofwater molecules within the aquifer, i.e. the age of thesemolecules when they exit the aquifer. The notion of tra-vel time is rather used to characterize the time spent totravel between two arbitrary locations in the aquifer.Travel time probabilities have been a subject of highinterest in many studies characterizing solute transportin sub-surface hydrology [16,18,19,38,39]. The traveltime probability is commonly defined as the responsefunction to an instantaneous unit flux impulse [21,39].In their transfer function approach of contaminanttransport through unsaturated soil units, Jury and Roth[39] model tracer breakthrough curves with one-dimen-sional travel time probability functions. Shapiro andCvetkovic [60], Dagan [19], and Dagan and Nguyen[20] derived the forward travel time probability for amass of solute by using the Lagrangian concept of par-ticle displacement in porous media. The derivation offorward and backward models for location and traveltime probability has become a classical mathematicalapproach for contaminant transport characterizationand prediction [69,43,50,51]. The spreading of a con-taminant mass is analyzed by following the randommotion of solute particles, and to do so, the advec-tion–dispersion equation (ADE) is assimilated to theFokker–Planck (or forward Kolmogorov) equation.The expected resident concentration of a conservativetracer is taken as the probability density function forthe location of a particle, at any time after havingentered the system.

The aim of the present work is to provide a generaltheoretical framework to model complete groundwaterage distributions at aquifer scale in a deterministicway. The concept of age variability is associated to theconcept of random variables and their distributions byusing classical elements of probability, allowing theintroduction of mathematical definitions for age, lifeexpectancy and transit time statistical distributions. For-ward and backward ADEs for conservative tracers areused to simulate the above-mentioned distributions ataquifer scale. By manipulating the ADEs, the reservoirtheory [27] is expressed in order to characterize rechargeand discharge zones transit time distributions with

refined accuracy. The proposed models are illustratedand discussed by means of analytical and numericalanalysis of one- and two-dimensional theoretical flowconfigurations.

2. The ‘ages’ of groundwater as space-dependent

random variables

In this section we present the models allowing the cal-culation of the statistical distribution of groundwaterage, life expectancy and transit time in arbitraryaquifers.

2.1. Definitions

The characterization of groundwater molecules withrespect to a travel time within an aquifer system is fullydependent on the spatial reference from which this timeis ‘‘measured’’. Usually the groundwater age is definedas a relative quantity with respect to a starting locationwhere age is assumed to be zero. Use will be madethroughout this work of three variations of terminology.For a given spatial position in the reservoir, the age (A)relates to the time elapsed since the water moleculesentered the system at the recharge limits, where age iszero. For the same spatial position, the life expectancy(E) is defined as the time required for the water mole-cules to reach an outlet limit of the system. Life expec-tancy is therefore zero at an outlet. The transit time(T) finally refers to the total time required by the samewater molecules to migrate from an inlet zone (T = E)to an outlet zone (T = A). In a REV, the three variablesA, E and T are random variables, characterized by prob-ability density functions (pdf) that can be regarded asthe statistical occurrence of water molecules with respectto time, which could be observed in a groundwater sam-ple if any analytical procedure would allow suchmeasurements.

2.2. Age probability

The typical heterogeneity of aquifer systems involvesstrongly varying flow velocity fields, with multi-scalecoherence lengths. The spatial variability of velocityand transport parameters induces a spreading of thecontaminant mass. The tensor of macro-dispersion inthe classical ADE accounts for the uncertainty in thetransport prediction induced by mixing. Various studiesrelate to how the ADE is limited by the impact of phys-ical and chemical heterogeneities on solute transport,such that up-scaling is not always satisfying [17,67,34].If such heterogeneities are present at aquifer scale thetransport parameters should be time-dependent, but thistime dependency may be relaxed when the correlationscales of the transport parameter random fields are finite

Page 4: Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

1270 F. Cornaton, P. Perrochet / Advances in Water Resources 29 (2006) 1267–1291

[34,49]. The ADE with time-independent parametersholds only when the solute particles have enough timeto be distributed by dispersion between the flow lines.Since we are interested in solving the age transport prob-lem at aquifer scale, we make the assumption that theADE with time-independent transport parameters (theparameters have reached their asymptotic values) canmodel the evolutional transport of the groundwaterage and life expectancy distributions under steady-flowconditions. The modelled process applies to conserva-tive and non-reactive tracers.

Let us consider an aquifer domain X in the three-dimensional space, with hydraulic recharge boundaryC�, discharge boundary C+, and impermeable boundaryC0, as illustrated in Fig. 1. The boundary C+ corre-sponds to the open boundary of the system, throughwhich a free exit of age mass is expected. With respectto the above-mentioned considerations on contaminantspreading, it is convenient to describe the groundwatersample age distribution as a random variable associatedto a probabilistic model. The age probability distribu-tion at a position x in X can be evaluated by solvingthe ADE when a unit pulse of conservative tracer is uni-formly applied on the recharge area C�. The resultingbreakthrough curve is the probabilistic age distribution[21,39]. Making use of this property, we propose tomodel groundwater age and life expectancy pdfs by for-ward and backward transient-state transport equations,under steady-state hydraulic conditions. The pre-solu-tion of the velocity field is performed by the followingsteady-state groundwater flow equation:

r � q ¼ qI � qO; q ¼ �KrH in X; ð1Þwhere q is the Darcy flux vector [LT�1], which is validfor ideal tracers, H is the hydraulic head [L], qI and qO

are fluid source and sink terms [T�1], respectively, andK is the tensor of hydraulic conductivity [LT�1]. Theage pdf is then obtained by solving the following for-ward boundary value problem:

Γ+

Γ0

Γ0

Γ0

Γ0

Γ−

Γ−

Γ−

T = A + E

A

E

(a)

T = A

T = E

T = E

T = E

Fig. 1. Schematic illustration of a groundwater reservoir X, with indicated nonormal flow field; (b) life expectancy problem with reversed flow field. Theposition and direction of water molecules at a given time after their release. Tvariable transit time (T) as the sum of the two random variables age (A) an

o/gA

ot¼ �r � qgA þr �DrgA þ qIdðtÞ � qOgA in X;

ð2aÞ

gAðx; 0Þ ¼ gAðx;1Þ ¼ 0 in X; ð2bÞ

JAðx; tÞ � n ¼ ðq � nÞdðtÞ on C�; ð2cÞ

JAðx; tÞ � n ¼ 0 on C0; ð2dÞ

where gA(x, t) is the transported age pdf [T�1], JA(x, t)is the total age mass flux vector [LT�2], D is the tensorof macro-dispersion [L2T�1], x = (x, y, z) is the vectorof Cartesian coordinates [L], t is time [T], / = /(x) isporosity or mobile water content [�], n is a normal out-ward unit vector, and d(t) is the time-Dirac delta func-tion [T�1], which ensures a pure flux impulse on C�.The source term qId(t) is meant for simulating a poten-tial internal production of water (3-D) or 2-D hori-zontal aquifer configurations with an areal rechargeintensity qI. The sink term qOgA may result fromany internal extraction of groundwater. The tensor ofmacro-dispersion D = /D* = D(x) in Eq. (2a) is definedby Bear [3]:

D ¼ ðaL � aTÞq� q

kqk þ aTkqkIþ /DmI; ð3Þ

where aL and aT are the longitudinal and transversalcoefficients of dispersivity [L], respectively, Dm is thecoefficient of molecular diffusion [L2T�1], and I is theidentity matrix. The total age mass flux vector JA(x, t)is classically defined by the sum of the convective anddispersive fluxes:

JAðx; tÞ ¼ qgAðx; tÞ �DrgAðx; tÞ. ð4Þ

2.3. Life expectancy probability

The life expectancy probability distribution satisfiesthe adjoint backward model of Eq. (2a):

Γ+

Γ0

Γ0

Γ0

Γ0

Γ−

Γ−

Γ−

T = A + E

A

E

(b)

-flow (C0), inlet (C�) and outlet (C+) boundaries: (a) age problem witharrow heads on the symbolized flowlines (dashed lines) stand for thehe black dot stands for a small water sample, to illustrate the random

d life expectancy (E).

Page 5: Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

F. Cornaton, P. Perrochet / Advances in Water Resources 29 (2006) 1267–1291 1271

o/gE

ot¼ r � qgE þr �DrgE � qIgE in X; ð5aÞ

gEðx; 0Þ ¼ gEðx;1Þ ¼ 0 in X; ð5bÞJEðx; tÞ � n ¼ �ðq � nÞdðtÞ on Cþ; ð5cÞ�DrgEðx; tÞ � n ¼ 0 on C0; ð5dÞ

where gE(x, t) is the transported life expectancy pdf, andwhere the total life expectancy mass flux vector JE(x, t)is

JEðx; tÞ ¼ �qgEðx; tÞ �DrgEðx; tÞ. ð6ÞEq. (5a) is the formal adjoint of Eq. (2a) [32,2], the so-called �backward-in-time� equation [69,73,72], also back-ward Kolmogorov equation [40]. Given the forwardequation, the backward equation is technically obtainedby reversing the sign of the flow field, and by adaptingthe boundary conditions [50,51,72]. On the impermeableboundary C0, a third-kind condition (Cauchy) in the for-ward equation becomes a second-kind condition (Neu-mann) in the backward equation, and vice versa. Asecond-kind condition in the forward model will also be-come a third-kind condition in the backward model [33,p. 146]. The advection term is known to be not self-ad-joint (it should be written in the form q Æ $gE in Eq. (5a))unless flow is divergence free [72]. However, the back-ward equation can still handle non-divergence free flowfields by means of the important sink term �qIgE

appearing in Eq. (5a). This sink term has been derivedin Cornaton [13] from the vertical averaging process ofthe general 3-D backward ADE, and is consistent withthe analysis of Neupauer and Wilson [51,52]. Rechargeby internal sources (3-D or 2-D vertical) or by arealfluxes (fluid source for 2-D horizontal) is introducedby the first-order decay type term �qIgE, which is a con-sequence of the reversed flow field. Internal sources pro-duce a sink of life expectancy probability, while internalsinks (term qOgA in Eq. (2a)) do not appear in the back-ward model since a fluid sink may not influence the lifeexpectancy pdf. The boundary C� corresponds to theopen boundary of the system (since flow is reversed)through which a free exit of life expectancy mass isexpected.

The simulation of life expectancy with (5) is valid inthe case of steady-state velocity fields only. Transient-state velocity fields would require another appropriateformulation of (5). For a steady-state hydraulic situa-tion, if q approaches zero in some regions of the reser-voir, e.g. like in aquitards in which transport isdiffusion-dominant, then (5) still holds because of theirreversible nature of dispersion. The amount of agemass diffused between an aquifer and an aquitard is pro-portional to concentration differences between the twomedia, and is the same in both the forward and back-ward problems. In the boundary value problems (2)and (5) the classical homogeneous Neumann boundarycondition (�D$g Æ n = 0) at the outlet limit for the age

problem (at the inlet limit for the life expectancy prob-lem) is not used in order to allow a natural age/lifeexpectancy gradient through the open boundaries.Instead, the normal projection of the dispersive flux isevaluated implicitly at the boundaries. The evaluationprocedure in the framework of the finite element methodis described in Cornaton et al. [14]. This kind of bound-ary condition, which is referred to as Implicit Neumann

condition, enables continuity of the total mass flux atoutlet. The Implicit Neumann condition is a generalizedversion of the Free Exit condition for parabolic equa-tions proposed by Frind [31]. As discussed by Naumanand Buffham [48], Parker and van Genuchten [53], Kreftand Zuber [42], and Bear and Verruitj [4], total mass fluxcontinuity at outlet permits upgradient solute movementby dispersion.

Eqs. (2) and (5) simulate the forward and backwardtransport resulting from a unit pulse input. The space–time evolution of the water molecules is described bythe distributions gA(x, t) and gE(x, t). Both differentialequations deal with conditional probabilities that char-acterize the statistical occurrence of water molecules asa function of age and life expectancy. Location proba-bility is related to resident concentration [41,18,39],and describes the position x of water molecules at agiven time after their release in the system. On the otherhand, travel time probability is related to flux concentra-tion [41,18,39], and characterizes for a position x theamount of time spent within X since the water moleculesentered the system (in the right-hand side of Eq. (5c),age is at the flux concentration gf

A ¼ dðtÞ). Resident con-centration relates to the mass of solute per unit porousvolume while flux concentration is defined as the solutemass flux per unit water flux. A flux concentration is thephysical representation of the mean of the microscopicfluid concentrations weighted by the associated micro-scopic fluid velocities [53]. The multi-dimensional rela-tion between flux and resident concentrations can befound in [65], and is formally the projection of the totalmass flux on the flow velocity direction. Accordingly,the flux concentration form of the random variableage is

gfA ¼

JA � qkqk2

¼ gA þJd

A � qkqk2

¼ gA �DrgA � qkqk2

ð7Þ

with Jd = �D$g denoting the dispersive part of J. Byanalogy, the flux concentration form of the variable E

can be defined as

gfE ¼ �

JE � qkqk2

¼ gE �Jd

E � qkqk2

¼ gE þDrgE � qkqk2

. ð8Þ

The age pdf at a position x in X characterizes the prob-ability per unit time for the time spent since recharge,and is a flux concentration, evaluated enforcing Eq.(7). The resident concentration of age characterizes a

Page 6: Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

1272 F. Cornaton, P. Perrochet / Advances in Water Resources 29 (2006) 1267–1291

volume-averaged age density function, which describesthe density of probability of finding water molecules atthe position x in X, at time t.

2.4. Transit time probability

In Eqs. (2a) and (5a), the dependent variables areprobability density functions of continuous time ran-dom variables. The behavior of these random variablescan also be described by cumulative distribution func-tions (cdf). Let U be one of these two variables (U = A

or U = E), with u the associated values the variable U

may take at a given position x of space. The cdf GU(x, u)and the pdf gU(x, u) of the variable U are commonlydefined as

GU ðx; uÞ ¼ P �1 6 U 6 u½ � ¼Z u

�1gU ðx; sÞds ð9Þ

with

gU ðx; uÞ ¼oGUðx; uÞ

ou; GU ðx;�1Þ ¼ 0;

GU ðx;1Þ ¼ 1;ð10Þ

where P denotes the probability event on U, or numberof occurrences with U 6 u ratioed to the total number ofoccurrences, and s is a dummy variable for integration.The probability functions property (10) together withthe boundary conditions in Eqs. (2) and (5) ensure thatage and life expectancy gA and gE are directly related toprobabilities, and since concentration can be modelledby the ADE, then probability too can be modelled bythe ADE. For a given position x in X, the ages ofgroundwater molecules are described by the pdf gA,which measures the density of probability of having anage t. The same molecules are also characterized bythe pdf gE, which measures the density of probabilityof having a life expectancy t. Introducing now the ran-dom variable transit time T, with density of probabilitygT, the water molecules can be described by their inten-sity of probability of flowing throughout the system at atime t. The variable T is a random variable correspond-ing to the sum of the two random variables A and E (seeFig. 1). Hence the statistical distribution of T is the pdfof the sum of A and E, gT = gA+E. This problem can besolved if the joint pdf gA,E of A and E, which character-izes the joint behavior of A and E, is known [5]:

gT ðx; tÞ ¼ gAþEðx; tÞ ¼Z þ1

�1gA;Eðx; s; t � sÞds. ð11Þ

The joint quantity gA,E(x, a, e)dade relates to the prob-ability that A lies in the small interval a to a + da, andthat E lies in the small interval e to e + de. Assumingthat A and E are stochastically independent variables(for the same spatial position, A depends on the initialpoint while E depends on the end point, and under asteady-state flow regime E may not be influenced by

the memory of the past evolution), and since gA andgE are zero for negative values of their arguments, thejoint pdf gA,E simplifies in gAgE [5]. The probability den-sity function gT in Eq. (11) can then be obtained usingthe convolution integral:

gT ðx; tÞ ¼Z t

0

gAðx; sÞgEðx; t � sÞds ð12Þ

from which the cdf of T can be calculated enforcing Eq.(9). The fact that gA and gE are zero for negative valuesof their arguments allows applying the convolution inte-gral from 0 to t. Since the pdfs gA and gE give the ageand life expectancy probability of occurrence at each po-sition x in X, both the maximum age as well as the max-imum life expectancy correspond to the maximumtransit time. Consequently, the time variable t can equiv-alently refer to all specific values of age, life expectancy,and transit time. The convolution integral (12) statesthat the probability that the variable T lies in a smallinterval around t is proportional to the product of theprobability that the variable A lies in the interval t tot + dt and a factor proportional to the probability thatthe variable E lies in a small interval around t � s, thevalue of E ensuring that A and E sum to T. This productis then summed over all possible values of time t (fromthe minimum age to the maximum age) to yield the tran-sit time pdf at a position x in space. The derived distri-bution of T = A + E in Eq. (12) can also be viewed as atransfer function convolution process, the input distri-bution being the age pdf gA, and the signal transferringfunction being the life expectancy gE.

To our point of view, Eq. (12) is an important resultof the present work. The field of gT characterizes theevolution of groundwater molecules throughout theaquifer domain by specifying the amount of time fromrecharge to discharge. At a given position in the reser-voir, the temporal evolution of the groundwater mole-cules can be characterized by the three pdfs gA, gE andgT. Each function contains specific information on atime of residence, the nature of which is a function ofthe spatial references that are chosen for evaluation.For instance, gA is conditioned by the inlet limit C�,where the variable A is nil, while gE is conditioned bythe outlet limit C+, where the variable E is nil. For thevariable T, the pdf gT is conditioned by the fact thatT = A at outlet, and that T = E at inlet.

If gA and gE are resident concentrations, so is gT. IfgA and gE are flux concentrations, so is gT. Applying aLaplace transform to Eqs. (7) and (8) and convoluting,the transit time resident and flux concentrations arefound to be linked by the following relation:

gfT ¼ gf

AgfE

¼ gT �1

qk k2gAJ

d

E � gEJd

A þðJd

A � Jd

EÞqqk k2

" #� q; ð13Þ

Page 7: Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

F. Cornaton, P. Perrochet / Advances in Water Resources 29 (2006) 1267–1291 1273

where gU ðx; sÞ denotes the s-transform state of the func-tion gU(x, u), U = A, E or T. Eq. (13) shows that transittime flux concentration is dependent on the transit timeresident concentration, but also on the tensor productbetween the age and life expectancy dispersive fluxes,and on the convolution product of the age and lifeexpectancy flux and resident concentrations. For a zerodispersion case, gf = g for the random variables A, E

and T. Consider the semi-infinite 1-D domain of charac-teristic length L (the outlet is supposed at the positionx = L) and uniform velocity v along the x-axis, as illus-trated in Fig. 2. The age flux pdf at the position x is ob-tained as the solution of Eq. (2a) using the boundarycondition gA ¼ gf

A ¼ dðtÞ at x = 0, and the age residentpdf is obtained as the solution of Eq. (2a) using theboundary condition (2c) at x = 0. These solutions are gi-ven in dimensionless form in Appendix A. They showthe trivial fact that for a one-dimensional flow configu-ration, the transit time flux pdf is unique and indepen-dent on the spatial coordinate, as illustrated in Fig. 2a.For example, the age flux pdf at X = 1/4 in Fig. 2a isequal to the life expectancy flux pdf at X = 3/4, andthe convolution of both distributions gives the transittime flux pdf, which equals both the age flux pdf at out-let (X = 1) and the life expectancy flux pdf at inlet(X = 0). The average age x/v and the average life expec-tancy (L � x)/v sum to the average transit time L/v,which is independent on x in a rectilinear flow line. Ina similar way, the age resident pdf at the position X

00

0.5

1

1.5

2

2.5

ALT

0 0.5 10

0.5

1

1.5

2

2.5

0 0.5 10

1

2

3

4

ALT

0 0.5 1 1.5 2 2.5 30

1

2

3

4

(b)

(a)

Dimensionless time Τ Dimensio

Dimensionless space X Dimensio

Dim

ensi

onle

ss f

lux

pdf

Dim

ensi

onle

ss r

esid

ent p

df

Τ = 1/2 Τ

X = 1/4 X =

Fig. 2. Age, life expectancy and transit time dimensionless pdfs in a 1-D domnormalized by the average turnover time s0 = L/v, x by L, and Pe = Lv/D. Thtime.

equals the life expectancy resident pdf at the position1 � X, as shown in Fig. 2b. For a fixed value of time,the transit time resident pdf is constant. This shows that,in 1-D, the intensity of probability of finding water mole-cules at any position in the domain that transit at time t

or less, given that t is fixed, is always identical. The tran-sit time resident pdf gives the intensity of probability forthe spatial position of water molecules, for a given valueof transit time. In 1-D, this intensity of probability isuniform since the trajectory is unique. The residentage and life expectancy curves show the typical apparentdiscontinuity in concentration at inlet (resident pdf of A)and at outlet (resident pdf of E). These discontinuitieshave a drawback on the transit time resident pdf, whichis not necessarily equal to the age resident pdf at outlet,and equivalently not necessarily equal to the life expec-tancy resident pdf at inlet. This can be attributed to dis-persion effects at the boundaries. Since the Cauchycondition is homogeneous for T = 0+ at inlet for theage pdf problem, and at outlet for the life expectancypdf problem, backward movement of water moleculesby dispersion (i.e. in the reserved direction of velocity)is put down by non-zero age and life expectancy residentconcentrations at the boundaries, the magnitude ofwhich is higher the higher the dispersivity. If we con-sider e.g. the inlet boundary, the age and life expec-tancy concentrations are both not nil. The convolutionof both pdfs is, therefore, not equal to any of them(age pdf at outlet, life expectancy pdf at inlet) since both

0 0.5 10

0.5

1

1.5

2

2.5

0.5 1

geife expectancyotal transit time

0 0.5 1 1.5 2 2.5 30

1

2

3

4

1.5 2 2.5 3

geife expectancyotal transit time

nless time Τ Dimensionless time Τ

nless space X Dimensionless space X

Τ = 3/2= 1

X = 3/4 1/2

ain for a Peclet number Pe = 20: (a) flux pdfs; (b) resident pdfs. Time ise average age and the average life expectancy sum to the average transit

Page 8: Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

1274 F. Cornaton, P. Perrochet / Advances in Water Resources 29 (2006) 1267–1291

concentrations can have a significant value at inlet andoutlet at the same time.

The spatial distribution of the transit time pdf is alsoruled by a differential equation. Combining Eqs. (2a)and (5a) after a Laplace transform, the following equa-tion can also be obtained:

r � qgT ¼L�1 Sd

� �; ð14aÞ

Sd ¼ gEr � Jd

A � gAr � Jd

E; ð14bÞ

where L�1 denotes the inverse Laplace transform. Eq.(14a) is the transit time pdf differential equation. It isof steady-state kind, with a source term that accountsfor the divergence of the age and life expectancy disper-sive fluxes. For the pure advective case in divergence freeflow fields, Eq. (14a) simplifies in q Æ $gT(x, t) = 0. Thislocal condition states that the flux vector and the transittime gradient have to be always perpendicular, as arequirement for keeping transit times constant alongthe flow paths. The resolution of Eq. (14a), which is ofhyperbolic kind, is linked to technical difficulties, e.g.for the evaluation of the source term (14b) over the do-main, and is not beneficial since the pdfs gA(x, t) andgE(x, t) must be evaluated in a preliminary step.

2.5. Mean age, mean life expectancy and mean

transit time

The average values of the probability density func-tions gA, gE and gT are defined by their first order tem-poral moment, the nth moment being

ln½gU � ¼Z þ1

�1ungU ðx; uÞdu ¼ ð�1Þn o

ngU ðx; s ¼ 0Þosn

ð15Þwith U = A, E, or T. Applying the convolution theoremin the Laplace domain results in the transformed equa-tion (12), gT ðx; sÞ ¼ gAðx; sÞgEðx; sÞ. Accounting for thepdf property gU ðx; s ¼ 0Þ ¼ 1 and enforcing Eq. (15)for n = 1 and 2 yields the average transit time and itsvariance:

hT i ¼ hAi þ hEi; ð16Þ

l2½gT � ¼ l2½gA� þ l2½gE� þ 2hAihEi; ð17Þ

r2½gT � ¼ r2½gA� þ r2½gE� ð18Þ

with hAi = hAi(x) = l1[gA], hEi = hEi(x) = l1[gE] andhTi = hTi(x) = l1[gT]. The average age and average lifeexpectancy of a water sample sum up to the averagetransit time. Since the variables A and E are supposedto be stochastically independent, the variances r2 ofthe age and life expectancy pdfs also sum up to the var-iance of the transit time pdf. Using the operator in Eqs.(15), (2a) and (5a) can be transformed into their nth nor-malized moment ln form. The recursive application ofthe operator (15) to Eqs. (2a) and (5a) yields

�r � qln½gA� þ r �Drln½gA� � qOln½gA�þ /nln�1½gA� ¼ 0 ð19Þ

for the forward nth moment ADE, and

r � qln½gE� þ r �Drln½gE� � qIln½gE� þ /nln�1½gE� ¼ 0

ð20Þfor the backward nth moment ADE. The nth momentforms of the ADEs (2a) and (5a) are only dependenton the (n � 1)th moment ln�1. For instance, sincel0[gA] = 1 and gA(x, 0) = 0, the first moment form ofEq. (2a) (for n = 1) corresponds to the mean age equa-tion as defined by Goode [35], in which the mean ageis the average over a sample of water molecules of thetime elapsed since recharge:

�r � qhAi þ r �DrhAi � qOhAi þ / ¼ 0 in X. ð21Þ

The first moment form of Eq. (5a) gives the backwardadjoint mean life expectancy equation:

r � qhEi þ r �DrhEi � qIhEi þ / ¼ 0 in X. ð22ÞFinally, the mean transit time equation is deduced bysubtracting Eq. (21) and (22):

q � rhT i ¼ hSdi; ð23aÞhSdi ¼ r �DrhAi � r �DrhEi; ð23bÞ

where the divergence of the advection term has beendeveloped in order to annihilate the fluid sourceand sink terms. The boundary value problems (2) and(5) involve that the finite moments of the age and lifeexpectancy pdfs exist for homogeneous boundary condi-tions. By definition, the mean groundwater age in a stea-dy-flow reservoir, or mean residence time, can bedetermined from the average of the normalized flux con-centration response to a narrow flux input uniformly ap-plied on the recharge limits, since this breakthroughcurve corresponds to the water molecules residence timedistribution [21,41,39]. The mean groundwater age canthen be calculated by prescribing at all inflow bound-aries a solute mass that is proportional to the water flux[37]. Consequently, Eqs. (21) and (22) can be solved byassigning hAi = 0 on the inlet limits C�, and hEi = 0 onthe outlet limits C+, respectively. Mean age and meanlife expectancy are continuously generated duringgroundwater flow, since porosity acts as a source termin Eqs. (21) and (22). This source term indicates thatgroundwater is aging one unit per unit time, in average.The mean age and mean life expectancy equations canbe easily handled by numerical codes that solve ADEs,by distributing a source term equal to porosity, and byreversing the velocity field for the case of Eq. (22). Eq.(23a) would require the boundary conditions hTi = hEion C�, and hTi = hAi on C+. However, mean transittime can rather be obtained by solving Eqs. (21) and(22), and by post-processing Eq. (16). If dispersion is

Page 9: Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

F. Cornaton, P. Perrochet / Advances in Water Resources 29 (2006) 1267–1291 1275

nil Eq. (23a) is simply q Æ $hTi = 0, the solution of whichis comparable to the stream lines in a flow model, andassociates to each line the advective transit time from in-let to outlet.

According to Parker and van Genuchten [53,54],Kreft and Zuber [41,42], and Sposito and Barry [65],temporal moments have a real physical meaning if theyare related to flux concentration functions, while spatialmoments must characterize resident concentration func-tions. Flux and resident concentration depend them-selves on the measurement technique [65,59]. An agedate as deduced from isotopic measurements corre-sponds to an age average over the number of moleculesin the water sample, and may often be close to a residentconcentration. Mean age computations using Eq. (21)are, therefore, well suited for fitting isotopic age dates.

2.6. Theoretical illustration example

The numerical solutions presented in this work havebeen performed using the Laplace Transform Galerkin(LTG) finite element technique [68], which allows elimi-nating the time-derivative in Eqs. (2a) and (5a). The

2

1

0

2

4

6

8

(d)

h1

h2

h3h4

h4 h3

h2

h 1,h 2,

h 3,h 4

h1

0 100 200012345

(c)

(b)

(a)

pdfs

[1/

year

]

Time [years]

AgeLife expectancyTotal transit time

0 2 4 6 8 10

h1 h2

Inlet: q = I

Fig. 3. Age, life expectancy and transit time computations in a 2-D verticalrepresentation by forward particle-tracking with marker-isochrones each 1 ye(c) spatial distribution of mean age (solid lines), mean life expectancy (daincrement; (d) age, life expectancy and transit time pdfs at some observationsDm = 0.

Cauchy type boundary conditions (2c) and (5c) implythat Eqs. (2a) and (5a) must be formulated accordingto their divergence form [23], such that a total mass fluxcontinuity at the boundaries must be properly handled.The assembled linear system of equations is solved bythe accurate ILUT-based sparse iterative solver [62] withcomplex arguments and GMRES [61] or BiCGSTAB[70] acceleration. The numerical inversion of the Laplacetransformed functions is performed by the algorithm ofCrump [15]. The quotient-difference algorithm is used toaccelerate the convergence of the Fourier series [24].This algorithm proved to be very efficient to treat inver-sion at the neighborhood of discontinuities or sharpfronts, and the required computational effort, which islinearly proportional to the 2n + 1 number of discreteLaplace variables, is highly diminished with respect toother acceleration methods.

We use here the simple example of a theoretical verti-cal aquifer section (see Fig. 3) to illustrate age, life expec-tancy and transit time computations. The configurationof the flow problem corresponds to the typical caseknown as �well-mixed reservoirs�, that generate an expo-nential-like transit time distribution at outlet. This aspect

12 12

0

2

4

6

8v1

v2

v3 v4

v 1,v 2,

v 3,v 4

v2v3

v4

v1

300 400 500

pdfs

[1/

year

]

Time [years]

AgeLife expectancyTotal transit time

0 2 4 6 8 10

h3 h4

v1

v2

v3

v4

0.5 m/y

Out

let:

H =

5 m

theoretical aquifer: (a) geometry, boundary conditions and flow fieldar; (b) backward particle-tracking with marker-isochrones each 1 year;shed lines) and mean transit time (dotted lines) with 1 year of time-points. Parameters: K = 10�4 m/s, / = 0.35, aL = 0.1 m, aT = 0.001 m,

Page 10: Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

1276 F. Cornaton, P. Perrochet / Advances in Water Resources 29 (2006) 1267–1291

will be discussed in more details in Section 4.2. The aqui-fer is homogeneous, and is uniformly recharged on top bya constant infiltration rate. The outlet, which could rep-resent a trench, is simulated by means of a prescribedconstant hydraulic head from the top to the bottom ofthe section. We solve the boundary value problems (2)and (5), as well as Eqs. (21) and (22). The model is discret-ized into 100,000 bilinear quadrangles of size Dx = 0.5 mand Dz = 0.25 m. Fig. 3a shows the flow field by means ofa forward particle-tracking representation. Fig. 3b givesthe backward particle-tracking solution, which repre-sents the purely advective life expectancy spatial distribu-tion. In Fig. 3c, we have plotted the solutions of Eqs. (21),(22) and (16). Mean age and mean life expectancy are invery good agreement with particle-tracking solutions.Isolines of mean age are horizontally distributed in rela-tion to the flow configuration: velocity increasesupstream to downstream (increase of hydraulic gradienttowards outlet), but also becomes more and more hori-zontally distributed. Travel paths to reach a same depthare longer downstream than upstream, but since velocityincreases downstream, it creates a compensating effectand the corresponding travel times finally become simi-lar, in average. Mean life expectancy gradients are mainlyhorizontal, and gradually increasing upstream. Meantransit time provides a good representation of the flowfield; its distribution is very close to the flow lines shownin Fig. 3a. However, the mean transit time solution ownsthe additional information on the total time required bywater molecules to travel from inlet to outlet. Finally,in Fig. 3d are given some observed pdfs (see Fig. 3a forthe location of the observation points). Successive agepdfs along a horizontal profile confirm the vertical gradi-ent of mean age shown in Fig. 3c, each age pdf at a samedepth being very comparable to one other. Along a verti-cal profile, successive age pdfs are more and more spreadin relation to the effect of longitudinal dispersion, whichis bigger the longer the travel path. The behavior of thelife expectancy pdfs is similar to that of age, but in thereversed direction of velocity. The transit time pdfs alonga horizontal profile are very similar to each other; theyare simply shifted along the axis of time. This reveals asame amount of dispersion affecting transport of watermolecules along each travel path from inlet to outlet. Ver-tically, the functions gT(x, t) show more and more disper-sion with increasing depth, in relation to the verticalmixing of water molecules.

3. Generalized reservoir theory

Recent studies [28,29] proposed a direct approach tocalculate outlet transit time distributions by applying thereservoir theory (RT) [27,7] to the average groundwaterage field, resulting from a pure advective transport solu-tion. In the following, we show how the RT can be gen-

eralized to advective–dispersive systems by manipu-lating the forward and backward transport equations.We first introduce some characteristic reservoir timeprobability distributions. Two equivalent mathematicalformulations are then proposed, in order to evaluatethe discharge zone transit time pdf as a function of thereservoir internal physical properties.

3.1. Characteristic reservoir distributions

When a specific age distribution is assigned to eachelementary water volume in the reservoir, the volumeof mobile water can be classified in a cumulative mannerwith respect to the age occurrence in the reservoir. LetMA(t) be the cumulated amount of water molecules inthe reservoir with an age inferior or equal to a particularvalue t, and mA(t) be the corresponding probabilityfunction, or reservoir internal age cdf. The functionmA(t) is the porous volume normalized function MA(t),which is evaluated by integrating over X the probabilityof finding water molecules with an age t or less, assum-ing that each molecule has entered the system on C�:

mAðtÞ ¼MAðtÞ

M0

¼ 1

M0

ZX

/GAðx; tÞdX

¼ 1

M0

ZX

/Z t

0

gAðx; sÞds

� �dX ð24Þ

with M0 being the total amount of mobile water (aquiferporous volume). The function mA(t) cumulates the prob-ability of finding water molecules which have travelleduntil a position x before time t. The function MA(t) iszero at origin and tends towards the total porous vol-ume M0 at infinity. The internal age frequency distribu-tion function wA(t) is the pdf associated to mA(t), andfrom Eq. (10) it follows

wAðtÞ ¼omAðtÞ

ot¼ 1

M0

ZX

/gAðx; tÞdX. ð25Þ

Note that Eq. (25) corresponds to the zeroth spatial mo-ment of the age pdf, which is equivalent to the age mass[59]. The function wA(t) gives the probability density offinding elements in X that have reached the age t, andwA(t)dt is the probability that A lies in the interval[t, t + dt]. Thus, the quantity M0wA(t) represents thenumber of elements per unit time (or flow rate fraction)that are in the interval [t, t + dt], and is equivalent to thezeroth spatial moment of the age pdf gA(x, t). We similarlydefine the internal life expectancy cdf mE(t) and pdf wE(t):

mEðtÞ ¼MEðtÞ

M0

¼ 1

M0

ZX

/GEðx; tÞdX

¼ 1

M0

ZX

/Z t

0

gEðx; sÞds

� �dX; ð26Þ

wEðtÞ ¼omEðtÞ

ot¼ 1

M0

ZX

/gEðx; tÞdX. ð27Þ

Page 11: Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

F. Cornaton, P. Perrochet / Advances in Water Resources 29 (2006) 1267–1291 1277

The function mE(t) cumulates the probability that awater molecule in the reservoir will reach the outlet be-fore time t.

We finally consider the internal distribution of thegroundwater molecules transit time pdf gT as deducedfrom the convolution integral in Eq. (12), to introducethe functions mT(t) and W(t) as the internal transit timecdf and pdf, respectively:

mT ðtÞ ¼MT ðtÞ

M0

¼ 1

M0

ZX

/GT ðx; tÞdX

¼ 1

M0

ZX

/Z t

0

gT ðx; sÞds

� �dX; ð28Þ

WðtÞ ¼ omT ðtÞot

¼ 1

M0

ZX

/gT ðx; tÞdX. ð29Þ

The function MT(t) corresponds to the mass of mobilewater in the reservoir having a transit time inferior orequal to t. The function W(t) characterizes the probabil-ity density of finding water molecules within the reservoirthat have a transit time inferior or equal to t, and thequantity M0W(t)dt gives the amount of water moleculesthat travel through X within the time interval [t,t + dt].

3.2. The transit time pdf of discharge and

recharge zones

A reservoir discharge zone is a particular portion offinite size, which intercepts the groundwater moleculesthat contribute to the outflow rate. These water mole-cules have contrasted arrival times that convergingwater fluxes mix together before flowing out. At a givenposition on a discharge boundary, the transit time pdf isthe age pdf. To characterize the contribution of each ageflux event on an outflow boundary in terms of transittime probability, it is convenient to average the ageprobability fluxes on C+. Under steady-flow regime,the representative transit time distribution uA(t) of thereservoir outlet zone can be defined as a flux averagedconcentration [59], i.e. uA(t) is evaluated as the flowrate-normalized sum on C+ of the total age mass fluxresponse function JA resulting from a unit flux impulseon C�:

uAðtÞ ¼1

F 0

ZCþ

JA � n dC ¼ 1

F 0

ZCþ

½qgA �DrgA� � ndC;

ð30Þ

where n is a normal unit vector pointing outward theboundary, and where F0 represents the total dischargeflow rate throughout the bounded domain, which atsteady-state is evaluated by

F 0 ¼Z

qðxÞ � ndC ¼Z

C�

qðxÞ � ndC. ð31Þ

Note that for the sake of simplicity, internal sources andsinks are neglected, qI = qO = 0. While flux concentra-

tions are defined with respect to a control plane ortho-gonal to the velocity vector direction, the outlet transittime probability function in Eq. (30) is defined by theprojection of the total age flux on the arbitrary-shapedboundary C+. Inserting Eq. (7) into Eq. (30) producesthe equivalent relation:

uAðtÞ ¼1

F 0

ZCþ

qgfA � I� q� q

qk k2

!DrgA

" #� ndC. ð32Þ

With Eq. (32), one can see that, if the velocity vector qpoints in the direction of the outward normal unit vectorn, then the dispersive term inside the brackets vanishes,and the pdf uA(t) is equal to the total steady-flow rate-normalized sum of the flux-weighted age flux concentra-tion pdfs on the outlet limit. This is always the case inone dimension. The cross product term in Eq. (32) re-veals also that since the outflow limit is of arbitraryshape, then when velocity does not point in the directionof n a dispersive correction term is required. This is re-lated to the fact that flux concentrations are defined withrespect to a control plane which is orthogonal to thevelocity direction.

The discharge flow rate can also be described as acumulative function of the transit time values. We intro-duce the function fA(t) as the transit time cdf of the res-ervoir outlet, i.e. fA(t) is the probability that themolecules flow out with a transit time t or less, such thatit corresponds to the normalized cumulated outflowfunction FA(t):

fAðtÞ ¼F AðtÞ

F 0

¼Z t

0

uAðsÞds. ð33Þ

Note that the function fA(t) can also be viewed as theintegral on C+ of the total mass flux deduced from theresident concentration solutions of the ADE (2a) for aunit step-input of mass flux on C�. By analogy, the lifeexpectancy pdf and cdf of the reservoir inlet C� can sim-ilarly be defined by

uEðtÞ ¼1

F 0

ZC�

JE � ndC

¼ � 1

F 0

ZC�

½qgE þDrgE� � ndC

¼ � 1

F 0

ZC�

qgfE þ I� q� q

qk k2

!DrgE

" #� ndC;

ð34Þ

fEðtÞ ¼F EðtÞ

F 0

¼Z t

0

uEðsÞds. ð35Þ

Eq. (13) indicates that the transit time flux pdf equalsboth the life expectancy flux pdf at inlet, and the age fluxpdf at outlet. As a matter of fact, assuming neither addi-tion nor subtraction of mass during transport, steady-flow conditions and stationary state, then, theoretically,every probability flux JA on C+ has an identical counter-

Page 12: Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

1278 F. Cornaton, P. Perrochet / Advances in Water Resources 29 (2006) 1267–1291

part JE on C� and vice versa. Therefore, Eqs. (30), (32)and (34) relate to the same and unique function,u(t) = uA(t) = uE(t), and thus f(t) = fA(t) = fE(t). In fact,each element added in the reservoir at the inlet must exitat some position xo of the outlet sooner or later. Eachelement added at the position xo at the outlet must travelthe same distance upstream, and thus spend the sametime-span within the reservoir, backward in time, beforereaching a position somewhere at the inlet limit.

At the reservoir outlet, the arrival times gA are dis-tributed along the discharge boundary, implying mixingand superposition of the information carried by eachbreakthrough curve. Moreover, within the reservoirmixing processes can also be important, and the trueminimum and maximum ages can be diluted. To char-acterize an outlet representative transit time distribu-tion, we must ensure that the minimum and themaximum ages are captured. Technical problems canoften occur when solving Eq. (30) or Eq. (34), becausemass flux line/surface integration is required. If the out-let is of small size, then the capture of calculated break-through curves, or the identification of particle arrivals,reveals to be ill-posed, mainly because of the loss ofinformation due to the mixing of converging fluxes.Hence, numerical methods will generally require a highlevel of refinement in the neighborhood of these inte-gration limits, which rapidly becomes a computationallimiting factor. Because the transit time pdf u(t) onthe inlet limit is identical on the outlet limit, discretiza-tion methods imply that the number of observationnodes should be the same at the inlet and at the outlet,in order to be able to recover the same breakthroughcurves. In other words, the temporal resolution of thecurve, when a counting of the individual arrival timesis performed, is a direct function of the spatial refine-ment in the vicinity of exit zones. The same restrictionaffects other simulation methods, such as the random-walk procedure.

In the following, we propose an alternative approachthat is relaxed from the above-mentioned practicalproblems. Eqs. (2) and (5) are considered to simulatethe age and life expectancy probability distributions inthe reservoir X. Integrating Eq. (2a) over X, and makinguse of the divergence theorem (�X$ Æ FdX = �C F Æ ndC)results inZ

½qgA �DrgA� � ndCþ o

ot

ZX

/gA dX

¼ �Z

C�

½qgA �DrgA� � ndC. ð36Þ

Normalizing Eq. (36) by the steady-flow rate F0 andaccounting for Eq. (25) and (30), Eq. (36) can be turnedinto the following form:

uAðtÞ þ s0

owAðtÞot

¼ � 1

F 0

ZC�

½qgA �DrgA� � ndC ð37Þ

with the quantity s0 being the turnover time commonlydefined at steady state as the ratio of porous volume toflow rate:

s0 ¼M0

F 0

. ð38Þ

Similarly, the integration of the ADE (5a) has the form

�Z

½qgE þDrgE� � ndCþ o

ot

ZX

/gE dX

¼Z

C�

½qgE �DrgE� � ndC. ð39Þ

Normalizing Eq. (39) by F0 and accounting for Eq. (27)and (34) yields

uEðtÞ þ s0

owEðtÞot

¼ 1

F 0

ZC�

½qgE �DrgE� � ndC. ð40Þ

The boundary integrals in the right-hand sides of Eqs.(37) and (40) can be simplified by accounting for theboundary conditions (2c) and (5c). For instance, insert-ing the boundary condition (2c) into Eq. (37) reduces theboundary integral to �F0d(t), and inserting the bound-ary condition (5c) into Eq. (40) reduces the boundaryintegral to F0d(t), and Eqs. (37) and (40) become

uAðtÞ þ s0

owAðtÞot

¼ dðtÞ; ð41aÞ

uEðtÞ þ s0

owEðtÞot

¼ dðtÞ. ð41bÞ

With Eq. (41a) we have recovered the RT formulation ofEriksson [27], in which the outlet zone transit time pdfuA(t) is proportional to the first derivative of the inter-nal age pdf wA(t), thus characterizing the probabilityfor the water molecules of being removed from X perunit time. Eq. (41b) is an equivalent formulation whichrelates the recharge boundary life expectancy distribu-tion to the internal life expectancy distribution. Sincethe functions uA(t) and uE(t) are equal, it follows fromEqs. (41a) and (41b) that wA(t) = wEE(t), which allowswriting the following general RT formulation:

uðtÞ þ s0

owðtÞot¼ dðtÞ. ð42Þ

Eq. (42) generalizes the RT to advective–dispersive sol-ute transport processes, and is valid for both age and lifeexpectancy. If dispersion is set to zero, the function w(t)can be evaluated by integration of the field gA(x, t) =d(t � hAi(x)) as proposed by Etcheverry and Perrochet[28,29]. Similarly, the function w(t) can be evaluatedby integration of the field gE(x, t) = d(t � hEi(x)). Sincew(t) = wA(t) = wE(t), it follows from Eq. (9) thatM(t) = MA(t) = ME(t). This points out the importancein allowing age and life expectancy dispersive fluxescrossing naturally the outlet and inlet boundary por-tions, because the use of the homogeneous Neumanncondition on C+ and may lead to different results for

Page 13: Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

F. Cornaton, P. Perrochet / Advances in Water Resources 29 (2006) 1267–1291 1279

u(t) if contrasted boundary configurations and flow con-ditions exist. The third-kind boundary conditions (2c)and (5c) are the most meaningful conditions for solvingthe age and life expectancy problems. They ensure apure total flux pulse input entering the system at inletfor the age problem, and at outlet for the life expectancyproblem. These conditions become homogeneous fort > 0 (zero flux), and do not allow backward mass lossesby dispersion. This would not be the case when using aDirichlet type condition, which may lead to incorrectsolute mass balances. For the one-dimensional case,the use of the Dirichlet condition permits the simulationof the age and life expectancy pdfs, but directly for fluxconcentration pdfs (see Appendix A).

The fundamental relation between the outlet transittime cdf f(t) and the internal age pdf w(t) given by theRT is obtained after integration of Eq. (42):

F 0 � F ðtÞ ¼ M0wðtÞ ¼oMðtÞ

otð43Þ

or

f ðtÞ þ s0wðtÞ ¼ 1. ð44ÞEq. (43) indicates that the outflow of water moleculesthat leave the system through C+ with an age older thant is balanced by the number of elements per unit timewithin X that reach the age t, i.e. in the interval[t, t + dt]. The corresponding volume of groundwaterreaching the age t in X is M0w(t)dt. In other words,the flow rate fraction of age t or less at the outlet is afunction of the spatial occurrence in the reservoir ofwater molecules of age t or less. Since the function f(t)is zero at the origin, it follows that the value of w(t) atorigin is the turnover rate r0 (w(0) = r0 = 1/s0), inde-pendently of the level of dispersion. Since f(t) is a cumu-lative function, then w(t) must be monotonicallydecreasing and M(t) must be an increasing function withmonotonically decreasing increments [27]. The pdf w(t)is constant from zero to the minimum age tmin at outlet,with w(0, . . . , tmin) = r0, which throws light on the factthat the probability per unit time of finding elements

0

1

2

3

4

5

6

7

0 1 2 30

1

2

3

0 1

(a)

Dimensionless time Dimensionl

Pe = 50 Pe = ∞Pe =

Pe = 500

Pe = ∞

Pe = 5 Pe = 5

Pe = 50

Dim

ensi

onle

ss

Dim

ensi

onle

ss

Fig. 4. Reservoir theory pdfs for a 1-D semi-infinite flow domain, as a funcoutlet transit time pdf; (b) internal age (or life expectancy) pdf; (c) internal trax by a characteristic length L, and Pe = Lv/D.

in X that have reached the age t 6 tmin is certain. FromEq. (43) it follows that M(t) has a constant derivativeequal to M0w(0) = F0 until the minimum age tmin isreached at outlet. Note that the same considerationscan be made for f(t) = fE(t), w(t) = wE(t) and M(t) =ME(t).

Fig. 4 illustrates in 1-D the outlet (or inlet) transittime pdf, the internal age (or life expectancy) pdf, andthe internal transit time pdf resulting from the analyticalresolution of Eqs. (25), (29) and (42) (see Appendix A).Under pure convective transport conditions (dashedlines in Fig. 4), the function u(t) equals the piston-flowtransit time pdf d(t � s0), and the function w(t) is theHeaviside function H(s0 � t)/s0, such that w(t) =w(0) = 1/s0 until tmin = s0, and w(t) = 0 after tmin. Thefirst temporal moment of the pdf u(t) (average transittime at outlet st) is dispersion-independent, and equalsthe turnover time s0. The average internal transit timesit and the average internal age si (first temporalmoments of W(t) and w(t)) are dispersion dependent.Increasing longitudinal dispersion (low Peclet numbers)generates short arrival times and tailing effects (see thevariances of the pdfs in Appendix A), and thus old arri-val times at the outlet as well as old ages within thedomain, which are visible on the three functions u(t),w(t) and W(t) for a range of Peclet numbers. The func-tion w(t) is constant from 0 until the minimum age atoutlet. Since the Cauchy type condition prevents back-ward losses by dispersion at x = 0, the value of w(t) atthe origin is always 1/s0 for any Peclet number(Fig. 4b). From the spatial organization of age or lifeexpectancy occurrence it is possible to predict the transittime distribution of a reservoir outlet (or equivalentlythe life expectancy distribution of a reservoir inlet),without the need of �counting� the arrivals of the watermolecules at a boundary of finite size. Thus, the pdfu(t) defined in Eqs. (30) or (34) as a pure boundaryproperty becomes a property of the reservoir internalstructure and hydro-dispersive characteristics. Theinformation that can be lost when u(t) is directly evalu-ated at the reservoir exit zone (or inlet zone) is recovered

2 3

(b)

0

1

2

3

4

5

6

7

0 1 2 3

(c)

ess time Dimensionless time

Dim

ensi

onle

ss

Pe = 50

500

Pe = 5

Pe = 500

Pe = ∞

tion of the Peclet number for Pe = 5, 10, 25, 50, 100, 250 and 500: (a)nsit time pdf. Time is normalized by the average turnover time s0 = L/v,

Page 14: Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

1280 F. Cornaton, P. Perrochet / Advances in Water Resources 29 (2006) 1267–1291

with Eq. (42). A far more accurate evaluation of u(t) isthus achieved, for which the main operation (domainintegrals (25) and (27)) is not time-consuming and caneasily be implemented for one-, two- and three-dimen-sional systems. A convenient way to compute Eq. (42)is to work in the Laplace space, since it allows handlingall time-dependent quantities in a quasi-analytical way.

3.3. Temporal moments of the reservoir theory

probability density functions

A direct consequence of the RT is that the expectedvalue of the mean residence time, i.e. the average transittime st at outlet or inlet, equals the reservoir mean turn-over time s0. This property can be found by calculatingthe first temporal moment of the transit time pdf u(t),and by making use of Eq. (44):

st ¼Z þ1

0

tuðtÞdt ¼Z þ1

0

½1� f ðtÞ�dt

¼ s0

Z þ1

0

wðtÞdt ¼ s0 ¼M0

F 0

; ð45Þ

where use has been made of the pdf property in Eq. (10).Since the reservoir is considered under steady-flow con-ditions, internal average time characteristics can be de-fined. The temporal moments of the global functionsw(t) and W(t) have physical significance since residentconcentration has been integrated in space. The meaninternal age sia and the mean internal life expectancysie are deduced from the first temporal moment of thefunction w(t). Integrating by parts and making use ofthe relation (42) results in

si ¼Z þ1

0

twðtÞdt

¼ t2wðtÞ2

����þ1

0

� 1

2

Z þ1

0

t2 owðtÞot

dt

¼ 1

2s0

Z þ1

0

t2uðtÞdt ¼ l2½u�2s0

¼ s0

21þ r2½u�

s20

� �. ð46Þ

Since wA(t) = wE(t) = w(t), we have set sia = sie = si.Using Eq. (16), the average internal transit time sit is de-duced from the first temporal moment of the functionW(t), and is found to be equal to twice the value of si:

sit ¼Z þ1

0

tWðtÞdt ¼ 1

M0

ZX

/Z þ1

0

tgT ðx; tÞdt dX

¼ 1

M0

ZX

/hT idX

¼ 1

M0

ZX

/hAidXþ 1

M0

ZX

/hEidX ¼ 2si. ð47Þ

Using Eq. (42), integrating by parts and inserting Eq.(46), the second temporal moment and the variance ofthe distribution u(t) reveal themselves as being functionsof s0 and si (or sit) only:

l2½u� ¼Z þ1

0

t2uðtÞdt ¼ 2s0si ¼ s0sit; ð48aÞ

r2½u� ¼ s0ð2si � s0Þ ¼ s0ðsit � s0Þ. ð48bÞ

Finally note that by using the same technique than inEq. (46), one can show than the second moment ofw(t) is a function of the third moment of u(t), l3[u] =3s0l2[w].

3.4. Accuracy of the RT approach

The accuracy of the RT method compared to a clas-sical evaluation of the transit time pdf u(t) at the outletlimit is illustrated in Fig. 5, using a four-layered verticalaquifer. The outlet zone is of very small size (Fig. 5a),which forces the information on age to be highly mixed.As attested by Fig. 5b, the individual age pdfs at outletcan be of very different shape. The differences betweenthe two evaluation methods are very important. Theflux-weighted sum of the age mass fluxes monitored atthe outlet nodes suffers from a loss of informationinduced by the mixing of converging fluxes at the outletsurroundings. With the RT, this information is recov-ered, since we ensure that each contribution to the out-flow rate is accounted for. According to Eq. (45), theaverage residence time st must equal the turnover times0; it is clear with this example that this property isnot satisfied by straight application of Eq. (30), whilethe RT provides a very accurate solution (Fig. 5c).

4. Direct evaluation of aquifer water volumes versus

age, life expectancy and transit time

The specific groundwater volumes related to a givenrange of ages or residence times are important quantitiesto consider when addressing aquifer management strat-egies. Assessing the long-term evolution of groundwaterchemistry, or defining corrective measures aiming atrestoring groundwater quality after a contaminationevent, indeed requires appropriate age and volume-related information. In this section, we complete theframework of the RT by analyzing the transit time cdff(t) to show how complex aquifer porous volumes candirectly be quantified as a function of age, life expec-tancy, and transit time.

4.1. Characteristic groundwater volume functions

Provided the residence time distribution u(t) and thetotal discharge F0 are known, the characterization ofinternal groundwater volumes can be done in a rela-tively straightforward manner by a simple analysis ofthe function F(t) as defined in Eq. (33), F ðtÞ ¼F 0

R t0uðuÞdu. In Fig. 6, the theoretical shape of the

cumulated outflow function F(t) (or cumulated inflow

Page 15: Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

15915

9.5

158.

5

158

157.

5

157

156.

5 156

155.5

155

15 4.5

154

153.5

153

152.5

152

0 150 300 450 600 750 900 1050 1200 1350 15000

50

100

150

log10(K) [m/s]: -4 -3.85 -3.75 -3.65 -3.55 -3.45 -3.35 -3.25 -3.15 -3

H = 160 m H = 150 m(a)

Time [days]

Out

letn

odal

pdfs

[1/d

ay]

0 2500 5000 7500 10000 125000

0.001

0.002

0.003

0.004(b)

Time [days]

pdfs

[1/d

ay]

0 2500 5000 7500 10000 125000

0.0002

0.0004

0.0006

0.0008

0.001

Transit time pdf (RT eval.)Transit time pdf (flux eval.)

τ0 = 3989.972 daysτt from RT = 3989.811 daysτt at boundary = 5235.103 days

(c)

Fig. 5. Theoretical four-layered aquifer illustrating the accuracy of the RT compared to the classical direct evaluation at the outlet of the functionu(t): (a) model geometry, boundary conditions and head solution in meters; (b) age pdfs monitored at outlet, from which a flux-weighted averageevaluation of the transit time pdf is performed; (c) outlet transit time pdf evaluated at outlet and using the RT. Transport parameters: aL = 2.5 m,aT = 0.01aL, Dm = 0.

F. Cornaton, P. Perrochet / Advances in Water Resources 29 (2006) 1267–1291 1281

function) is represented. The areas tA, tT, and t0 inFig. 6a characterize different groundwater volumes thatcan be defined as functions of time t (Fig. 6b). The areatA(t) represents the total groundwater volume V in X ofage inferior or equal to t but that will experience a tran-sit time superior to t:

tAðtÞ ¼ V X A 6 t and T > tf g ¼ t½F 0 � F ðtÞ� ¼ M0twðtÞ.ð49Þ

In Eq. (49), the pdf w(t) is expressed given the relation-ship (43) between w(t) and F(t). When the minimumtransit time tmin is not nil, the function tA(t) containsthe volume of age inferior or equal to time tmin, whichis the area tA1(t) = tmin(F0 � F(t)), and the volume ofage inferior or equal to t and superior to time tmin, whichis the area tA2(t) = (t � tmin)(F0 � F(t)). The areatT = tT(t) is the volume of groundwater that flowsthrough X with a transit time t or less:

Fig. 6. Theoretical cumulated outflow function, internal age pdf (scaled bycumulated outflow function F(t), with the indicated areas tA, tT and t0 represetime; (b) aquifer porous volumes M(t), tA(t) and tT(t) as a function of age a

tT ðtÞ ¼ V XfT 6 tg ¼ tF ðtÞ � t0ðtÞ ð50Þwith the area t0(t) being the amount of exfiltrated waterhaving travelled from inlet to outlet during an observa-tion time period t,

t0ðtÞ ¼ V Cþ A 6 tf g ¼Z t

0

F ðuÞdu ¼ tF 0 �MðtÞ; ð51Þ

and where use has been made of Eqs. (25) and (43) toexpress the function M(t). The quantity tF(t) in Eq.(50) is the total amount of groundwater water in X thatreaches the age t or less on C+, plus the volume flowingout with an age t or less. The amount of groundwaterwater tT(t) is nil until the minimum transit time tmin,and reaches the total porous volume at the maxi-mum transit time tmax. The function tT(t) containsthe volume of age inferior or equal to time tmin, whichis the area tT1(t) = tminF(t) in Fig. 6a, and the vol-ume of age superior to time tmin, which is the area

the porous volume) and internal groundwater volume functions: (a)nting characteristic internal groundwater volumes relative to a value ofnd transit time.

Page 16: Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

1282 F. Cornaton, P. Perrochet / Advances in Water Resources 29 (2006) 1267–1291

tT2(t) = (t � tmin)F(t) � t0(t) in Fig. 6a. Note also thatthe total amount of groundwater of age tmin or less is gi-ven by the sum tA1 + tT1 = tminF0.

Since the amount of groundwater tT(t) is the internalvolume that will leave the reservoir up to time t, it equalsthe internal transit time cumulative distribution func-tion, tT(t) = MT(t), and it represents a part of the func-tion M(t). The complementary part is the amount ofgroundwater water of age t or less and of transit timesuperior to t, namely the function tA(t) defined in Eq.(49):

MðtÞ ¼ V X A 6 tf g ¼ tAðtÞ þ tT ðtÞ. ð52ÞWith Eq. (52), one can express tA(t) as the difference be-tween two increasing functions that both tend to theporous volume M0 at infinity. The function tA(t) is thuszero at the origin and at infinity. Since tT(t) is zero untilthe minimum transit time tmin, tA(t) must equal M(t) be-tween 0 and tmin. During this time-span, which can betaken as the signature of badly recharged and/or advec-tion-dominated systems for significant values of tmin,these two functions have a constant derivative (seeFig. 6b) equal to the steady-flow rate F0 = M0w(0).The behavior of this function (number of peaks, com-pared durations of increasing and decreasing parts) isinstructive as to the volumetric proportions of ground-water remaining a long time in the system, or flowingquickly to the outlet. The time after which the functiontA(t) starts to decrease gives information on the impor-tance of the water volumes in the aquifer with long orshort transit times. If this time is relatively young, thenthe aquifer may present a good turn-over property, andvice versa.

Differentiating Eq. (52) with respect to time, andaccounting for Eqs. (25), (29) and (42), yields

uðtÞ ¼ s0

tWðtÞ. ð53Þ

This fundamental relation includes all the features of theRT in the most compact form. Compared to the stan-dard rule (42), Eq. (53) is a great improvement. It is sim-pler and may provide the transit time distribution u(t)with much higher resolution and accuracy, in relationwith the fact that no differentiation between u(t) andW(t) is required.

Note that if F(t) characterizes the inlet cumulatedinflow rate, i.e. F(t) = F0fE(t), then each aquifer porousvolume defined above as a function of age becomes afunction of life expectancy. By analyzing the system out-let transit time cdf, the quantification of complexgroundwater volumes with respect to age, life expec-tancy, and transit time, is straightforward, and providesimportant practical system insights. An obvious applica-tion is the groundwater resources protection and vulner-ability assessments. The underground storage of toxicwastes also requires the knowledge of groundwater vol-

umes that are potentially subject to contamination, andassumptions on the time spent by these volumes withinthe reservoir until they reach an outlet area.

4.2. Characteristic times versus aquifer particularities

Following Bolin and Rhode [7], some aquifer config-uration particular cases can be drawn from the firstmoments of u(t) and w(t).

The first considerations correspond to the case forwhich the average transit time and the average internalage are identical, st = si. Using Eqs. (45) and (46), thecondition for having st = s0 = si is

R10

t½uðtÞ�wðtÞ�dt ¼ 0, for which a sufficient condition is uðtÞ ¼wðtÞ ¼ 1

s0expð� t

s0Þ. The exponential form of the transit

time pdf is a direct consequence of the first-order lineardifferential Eq. (42): if one of the two pdfs u(t) and w(t)has an exponential form, then the other must be identi-cal. This is the exponential model, often termed the well-mixed model, which is mathematically equivalent to theunit response function of a well-mixed reservoir. Inchemical engineering, this model is used for reactorsinside which the age distribution of the elements is uni-form, i.e. there is a perfect mixing of the elements.Although the exponential model is widely used byhydrogeologists to simulate isotopic data, its applicationin aquifer systems must be handled carefully, since itinvolves a large number of poorly realistic assumptionson the aquifer structure and recharge conditions. Eriks-son [26] interpreted the exponential distribution of agesin groundwater as a consequence of an exponentialdecrease of porosity and permeability with depth.Luther and Haitjema [44] argued that the conditionson the validity of the exponential residence time distri-bution in porous media (horizontal, un-stratified andhomogeneous aquifer with respect to porosity /,recharge rate I, and saturated thickness H), can berelaxed if the parameters /, I, and H vary in a piecewiseconstant way, such that the ratio /H/I remains constantthroughout the domain. This ratio characterizes the sys-tem turnover time, which means that each water sampletaken from the reservoir must lead to a mean age thatequals the aquifer mean turnover time. In nature how-ever, such system configurations and conditions onmean age may hardly be found. Etcheverry [30] showedthat a simple linear variation of the thickness H signifi-cantly influences the shape of the theoretical exponentialresidence time distribution.

We now consider the case where st > si. This case cor-responds to the situation for which only few water mol-ecules leave the aquifer rapidly after having entered.Confined aquifer conditions and/or very distantrecharge and discharge zones are typical settings leadingto these features. For such a configuration, the outlettransit time pdf u(t) is generally small or nil for youngages, attesting the existence of a minimum duration

Page 17: Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

F. Cornaton, P. Perrochet / Advances in Water Resources 29 (2006) 1267–1291 1283

for travelling from inlet to outlet. The functions tA(t)and M(t) tend to remain identical until the minimumtransit time is reached. After this date, the functiontA(t) should decrease rapidly, reflecting the fact thatafter the shortest travel distance from source to sinkhas been covered, the outflow of older water moleculesis concentrated over a short time-span, in relation tothe relative uniformity of the travel distances withinthe system. A narrow triangle-shaped function tA(t) istypical of aquifers with significant minimum transittimes and bad turnover property. Note also that fromEq. (46) one can see that for systems with large turnovertime, or relatively low hydro-dispersive properties (i.e.when r[u] is small), the average age in the reservoirmay also be smaller than the turnover time, with thelowest possible value si ¼ s0

2for the piston-flow configu-

ration only (r[u] = 0).Finally, the case st < si corresponds to the situation

for which important amounts of water molecules enterthe aquifer and flow out relatively rapidly, while suffi-cient amounts of water stay long enough to increasethe value of si. For such a configuration, u(0) must bebigger than w(0), and the two curves must both decreaseand cross each other at a certain date, after which w(t) ishigher than u(t). This situation may be encounteredwhen the source and sink zones are close to each other,or when superficial recharge is uniformly distributed,such that the fraction of young water is important atoutlet, but when the heterogeneity of the velocity distri-bution is such that long travel paths might lead to oldages within the domain. Karstic systems are typicalmedia where the case st < si is encountered, in relationto the effect of the high velocities in the karstic network,which can carry water molecules more or less indepen-dently of the surrounding low permeability matrix. Con-sidering Eq. (46), the case st < si can occur in systemswith small turnover time, or relatively high hydro-dis-persive properties (large r[u]).

5. Analysis of dispersion and aquifer geometry

effects on age distributions

In the following, 2-D theoretical examples are pre-sented to illustrate the proposed methods. Analyticaland numerical solutions are provided in order to testthe sensitivity of the probability and density functionsto the advective and dispersive parameters, as well asto the aquifer structure.

5.1. Two-dimensional single flow system aquifer

A 2-D half-circular crown-shaped reservoir is used tosimulate a single-flow system aquifer, homogeneouswith respect to porosity / and hydraulic conductivityK. A positive head difference is applied between the

recharge and discharge areas (Fig. 7a). This geometryis well suited for the derivation of analytical solutions[29]. Due to homogeneity and symmetry, the flow linesremain parallel to each other, flow being one-dimen-sional along the flow line coordinate. The spatial distri-butions of mean age and mean life expectancy can thusbe solved analytically (see Appendix B), and reveal to besymmetric (Fig. 7a), yielding a co-centrical distribu-tion of mean transit time. The porous volumes tA

(=tA1 + tA2) and tT (=tT1 + tT2), quantified using thecumulative outflow function F(t), can easily be identifiedin Fig. 7 by using the mean age and the mean transittime isochrones. Note that the special case of this theo-retical aquifer allows a good representation of thegroundwater internal volumes with the mean values ofthe global distributions. However, the mean of theage, life expectancy, and transit time distributions wouldgenerally give a rough representation only, when e.g.significant dispersive processes take place. Etcheverryand Perrochet [29] analyzed the pure advective caseand found that u(t) is proportional to 1/t, and thatw(t) is proportional to ln(1/t) between the minimumand the maximum transit time (see Fig. 8a). The func-tion W(t) is a constant from the minimum to the maxi-mum transit time, which indicates that the aquifervolumes ranging between a unit increase of transit timeremain constant (the aquifer volumes between two iso-contours in Fig. 7b are all the same). The case aT = 0in Fig. 8a was analyzed by means of analytical solutions(Appendix B). For this system, the pdfs u(t), w(t) andW(t) show generally moderate fluctuations with respectto longitudinal dispersion. The main effect of the aL

coefficient can be seen in the increase of the spreadingof the probability distributions. The outlet transit timepdf shows younger arrival times when aL increases,but also a longer tail. The functions w(t) and W(t) areaffected the same way by longitudinal dispersion. Theaverage internal age si and the average internal transittime sit are dispersion dependent; they increase linearlywith aL, indicating that dispersion leads to a higher aver-age age of the system. Note that the second temporalmoments of the internal age and transit time distribu-tions increase proportionally with the square of aL,while the variance of u(t) increases linearly with aL.Contrasted behavior appears when the coupled effectof longitudinal and lateral dispersivity is taken intoaccount (Fig. 8b). The case aT 5 0 was analyzed usingnumerical solutions, by increasing the ratio a = aL/aT.The tailing effect decreases from a starting situation(a =1 in Fig. 8b) with increasing values of lateral dis-persivity (decrease of a ratio in Fig. 8b). Lateral disper-sivity induces the mixing of ages, old water moleculesmoving laterally between the flow lines and replacingyounger molecules, and vice versa. The internal ground-water volume functions also attest of the above-men-tioned dispersion effects. For example, the function

Page 18: Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

(a)

(b)

Fig. 8. RT solutions for the 2-D vertical half-circular aquifer: (a) analytical solutions (see Appendix B) for the pdfs u(t), w(t) and W(t), and for theinternal groundwater volume functions (in % of M0) M(t), tA(t) and tT(t), as a function of the aL coefficient (aT = Dm = 0); (b) numerical solutions asa function of the ratio a = aL/aT, with aL = 50 m.

Fig. 7. Analytical solutions for mean age, mean life expectancy and mean transit time for the 2-D vertical half-circular aquifer (Appendix B): (a)mean age (solid lines) and mean life expectancy (dashed lines) isochrones with 50 days increment; (b) mean transit time isochrones with 50 daysincrement. Parameters: DH = H2 � H1 = 100 m, K = 10�4 m/s, / = 0.2, inner radius r0 = 250 m, outer radius R = 1000 m, minimum transit timetmin = 143 days, maximum transit time tmax = 2285 days, turnover time s0 = st = 772.5 days.

1284 F. Cornaton, P. Perrochet / Advances in Water Resources 29 (2006) 1267–1291

tA(t) shows long tails for large values of aL, reinforcingthe observed behavior at outlet of the transit time pdffor old ages. The internal increase of water volumes thatrequire long transit times to exit the aquifer due to anincrease of aL has of course its consequence at outletwith older arrival times. As with the pdfs u(t), w(t)and W(t), this tailing effect decreases when lateral disper-sion is added.

5.2. Vertical multi-layer aquifer

In this second theoretical example, we consider afour-layered vertical aquifer system, as illustrated inFig. 9a. From the top to the bottom of the model, the

layers have decreasing thicknesses and increasing porevelocities. The domain is discretized into 30,000 homo-geneous bilinear quadrangles. The total porous volumeM0 is 65312.5 m3. A constant input flow rate entersthe system along the left limit, using imposed fluxes ofvarying intensity proportional to the different layershydraulic conductivities. The system turnover time iss0 = 162.267 days. The outlet is simulated at the top ofthe right side by a constant hydraulic head along a rela-tive small zone of 15 m. The permeability–porosity cou-ples have been set in a way that the pore velocitycontrasts involve specific ages within each layer (veloci-ties vi in Fig. 9a), creating specific ages within each layer,as illustrated in Fig. 9b. The fact that the layers thick-

Page 19: Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

161 160 159 158 157 156 155154 153

H = 150 m

1250 m

150 m

10 50 100 150 200 250

v1

v2

v3v4

(a)

(b)

F0

= 4

02.5

m3 /d

Fig. 9. 2-D vertical multi-layered aquifer: (a) boundary conditions and Laplacian flow field in meters; (b) mean age isochrones in days, illustratingthe pore velocity contrast effect on age transport (aL = Dx = Dz = 2.5 m, aT = 0, Dm = 0). Average pore velocities (in m/day): v1 = 4.0, v2 = 6.5,v3 = 10.5, v4 = 23.5.

F. Cornaton, P. Perrochet / Advances in Water Resources 29 (2006) 1267–1291 1285

ness diminishes with depth, while the influx intensityincreases, is meant to create arrival time peaks at outletof comparable orders of magnitude. The temporalmoments si and sit are good indicators of the dynamicsof the global system. Because they are volume-averagedquantities, their magnitude will depend on the waterquantities of contrasted age and transit time, whichare directly related to the flow and transport dynamics.For this example, si and sit show only small variationswith respect to dispersion (Table 1), because even ifvelocities between layers are contrasted, flow in the sys-tem is generally rapid. Longitudinal dispersion has theeffect of making the system older, by creating long tailsthat can be observed at the reservoir outlet, but also onthe internal age and internal transit time distributions(Fig. 10a and b). Spreading in the direction of velocityis of course the main cause for this. Lateral dispersionpresents the property of homogenizing the ages by mix-ing water molecules of different ages. For a given level oflongitudinal dispersion, increasing the ratio a = aL/aT

(Fig. 11) makes the system younger (tailing is lowered),with diminishing values of the mean internal age si andthe mean internal transit time sit (Table 1). Longitudinaldispersion induces spreading of solute particles, andthus variability around the plume center of gravity.When lateral dispersion is added, the mixing of watermolecules between flow lines homogenizes the ages.

Table 1Dispersive parameters related to Figs. 10 and 11, and calculated pdfstatistics (units in meters and days)

Parameters Fig. 10 Fig. 11

a = aL/aT 10 10 10 1 50 10aL 5.0 10.0 20.0 5.0 5.0 5.0aT 0.5 1.0 2.0 0.0 0.1 0.5

Statisticsr[u] 75.710 75.955 79.871 81.766 79.786 75.940si 98.902 98.915 101.010 102.156 100.905 98.916

When longitudinal spreading is important, the exchangesurface along the plume body is extended, and mixingcan then be expected to be of high magnitude. The inter-nal transit time pdf W(t) is of particular interest for thecharacterization of the internal organization of the flowdynamics. Its shape and particularly the number ofmodes it shows has a direct consequence on the shapeof the outlet transit time pdf. If we take the exampleof the case a =1 (aT = 0) in Fig. 11, the functionW(t) exhibits as many peaks as the outlet transit timepdf u(t). However, the magnitude of these peaks are dif-ferent for the two functions, and discern the transit timefrequencies at outlet from those of the system pore vol-ume. The peak of maximum intensity is the first one foru(t), and the third one for W(t). If one looks at the timevalue corresponding to the first peak of u(t), say t1, theninside the domain the density of probability that thewater molecules have a transit time equal to t1 is smallerthan the same density of probability at the reservoir out-let. This is due to the fact that W(t) deals with volume-average probabilities while u(t) deals with flux-averageprobabilities. The four consecutive peaks of the outlettransit time pdf correspond to the four families of watermolecules, which transit within the four layers of thesystem, even though some parts of the curve attest tosome mixing effects (the density of probability is notnecessarily zero between two peaks). This simple theo-retical example shows the complexity that is to beexpected for the nature of the age and transit time distri-butions. It underlines how the significance of averageage can lead to erroneous interpretations, regardingthe system dynamics (e.g. inferences on average velocity)or hydrogeological problems related to risk and vulner-ability assessment. Whenever the outlet transit time pdfis multi-modal, then the average transit time can defi-nitely not be used as a reliable quantity representativeof the distribution since, as illustrated here, mean agesat outlet are mostly related to the lowest densities ofprobability (see Figs. 10 and 11, Table 1). Once more

Page 20: Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

(a)

(c)

(b)

(d)

Fig. 10. Calculated pdfs, cdfs and groundwater volumes for the 2-D vertical multi-layered aquifer as a function of dispersion, for the ratio a = aL/aT = 10: (a) outlet transit time pdf; (b) internal age pdf and internal transit time pdf; (c) outlet cumulated outflow function; (d) groundwater volumesversus time (in % of M0).

(a)

(c)

(b)

(d)

Fig. 11. Calculated pdfs, cdfs and groundwater volumes for the 2-D vertical multi-layered aquifer as a function of the ratio a = aL/aT: (a) outlettransit time pdf; (b) internal age pdf and internal transit time pdf; (c) outlet cumulated outflow function; (d) groundwater volumes versus time (in %of M0).

1286 F. Cornaton, P. Perrochet / Advances in Water Resources 29 (2006) 1267–1291

Page 21: Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

F. Cornaton, P. Perrochet / Advances in Water Resources 29 (2006) 1267–1291 1287

this consideration points out to the interpretation ofaverage age measurements, but also mean age simula-tions, which should be done very carefully. Generally,attention is given to hydrodynamic dispersion, heteroge-neity, and long distance travelling induced mixing[35,71,72], which are factors that represent a source ofuncertainty for the age dating methods and environmen-tal tracer data interpretations. Here one can see that thegeological, the structural and the hydraulic boundaryconfigurations can by themselves be responsible ofunrepresentative mean ages, even in advection-domi-nated transport regimes. Even if the entire groundwaterage distribution cannot be measured in the field, meanage dates can be of great help for model calibration pur-poses. However, mean ages should a priori not representan absolute simulation answer, and the knowledge ofthe entire age distribution should be of prime interestin most cases.

6. Summary and conclusions

(1) The probability distributional evolution ofgroundwater age and life expectancy has been simulatedby forward and backward transient advection–disper-sion type equations, according to proper boundary con-ditions. For a given position in space, the age and thelife expectancy pdfs are complementary distributions.Straight application of the principle of superpositionto these distributions results in an integrated distribu-tion of transit times, which characterizes the probabilityof having a given transit time from the recharge zone tothe discharge zone, together with the quantities associ-ated to that transit time. The transit time pdf is simplythe convolution integral of the age pdf and the lifeexpectancy pdf, and tells about the entire history ofwater molecules since recharge until discharge. Age, lifeexpectancy, and transit time pdfs provide different kindof information, and each of them can reveal to be moreadvantageous than the other one depending on thehydrogeological insights to be provided. For instance,life expectancy and transit time distributions are wellsuited for vulnerability assessment problems (e.g. well-head protection, or underground storage of toxic waste),and they allow the mapping of different regions within arecharge zone, in terms of residence time in the aquiferand associated properties.

(2) The results of Eriksson [27] have been recoveredby manipulating the advection–dispersion equation,extending thus the RT to systems with significant disper-sive components. In the classical RT neither advectionnor dispersion was considered. We have demonstratedhere that the RT is still valid in systems with spatiallyvarying velocity fields and non-negligible dispersive/dif-fusive effects. The RT is a simple one-dimensional for-mulation, time being the only dependent variable. The

outlet transit time distribution is derived from the inter-nal age distribution, and therefore has a much morerefined resolution. In so doing, mixing of convergingflow patterns near the outlet is ignored, and the maxi-mum transit time is never smaller than the maximumage in the reservoir. The RT formulation also appliesto the internal distribution of life expectancy, and canbe used to evaluate recharge zones life expectancy pdfs.It has also been shown that the RT can be expressed in amore compact form, which relates the outlet transit timedistribution to its internal counterpart, the internal dis-tribution of transit times from inlet to outlet.

(3) From the reservoir characteristic distributions,fundamental additional transient information on watervolumes and water fluxes can be gained. From the outlettransit time cdf, specific groundwater porous volumescan directly be identified and quantified with respect toage, life expectancy and transit time. These functionscan be very useful for aquifer characterization andintrinsic vulnerability assessments. They can be used toeasily evaluate the magnitudes of young and old ground-water volumes in the aquifer.

(4) Using analytical and numerical solutions for the-oretical aquifer configurations, some effects of macro-dispersion on simulated age and transit time distribu-tions could be underlined. For instance, longitudinaldispersion can have a significant aging effect, while lat-eral dispersion rejuvenates the system through trans-verse mixing. Like in previous studies, it has beenshown that the average age resulting from dating-meth-ods, or direct simulations, can lead to erroneous inter-pretations. It is a well-known fact that dispersion caninduce a high variability of the age distribution aroundthe average. Here we have shown that not only the mix-ing processes could make the average age insignificant,but also the geological structure and the geometry ofthe flow patterns.

(5) The proposed methodology can equivalently beimplemented in one, two and three dimensions, andhas considerable technical and numerical advantages,which may be pivotal in handling very large systems.In fact, when the outlet transit time pdf is defined by inte-grating all hydro-dispersive properties over the entireflow field, the level of refinement required by a stabletransport model is generally sufficient. The RT ensuresthat the minimum and the maximum age in the reservoirare captured at the outlet, which is hardly the case withtraditional methods, mainly because of the mixing ofconverging fluxes near the outlet. In the present work,the RT has been combined with ADEs, and the equa-tions were solved using the LTG technique. However,no restriction appears for the use of the methodology ifage and life expectancy pdfs are calculated by othertransport models, such as random-walk simulators.

(6) The presented models have been developed for theglobal aquifer system, regardless the number of individ-

Page 22: Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

1288 F. Cornaton, P. Perrochet / Advances in Water Resources 29 (2006) 1267–1291

ual inlet and outlet zones. At this stage, the RT has beenrendered applicable to the whole system. In a subse-quent article (this issue), we generalize the RT to anyobservation zone, and to systems with several inletsand outlets.

Acknowledgements

The authors would like to acknowledge the Swiss Re-search National Fund for financially supporting this re-search under Grant no. 2100-064927. The authors alsothank Professor H.-J. Diersch and one anonymous re-viewer for constructive criticism of this manuscript.

Appendix A. Reservoir theory for a 1-D semi-infinite

domain

The age resident pdf is found by solving the one-dimensional form of the age pdf ADE (2a) with the Cau-chy type condition vgA(0, t) � DogA(0, t)/ox = vd(t), andwith a zero concentration gradient at a point at infinity.The corresponding age flux pdf is the flux concentrationgf

A, which is deduced using the one-dimensional form ofEq. (7), gf

A ¼ gA � Dv

ogAox . These solutions are e.g. given in

[39]. Using the dimensionless variables

X ¼ xL; T ¼ v

Lt; Pe ¼ vL

D;

where L is a characteristic length defining the supposedoutlet position, and where Pe is the Peclet number, theage resident and flux pdfs read

gAðX ; T Þ ¼ffiffiffiffiffiPepffiffiffiffiffiffi

pTp exp � PeðX � T Þ2

4T

!

� Pe2

expðXPeÞerfc

ffiffiffiffiffiPepðX þ T Þ2ffiffiffiffiTp

� �; ðA:1Þ

gfAðX ; T Þ ¼

XffiffiffiffiffiPep

2ffiffiffipp

T32

exp � PeðX � T Þ2

4T

!. ðA:2Þ

The life expectancy resident and flux pdfs can be de-duced from (A.1) and (A.2) by substitution of X by1 � X and X by 1, respectively:

gEðX ; T Þ ¼ffiffiffiffiffiPepffiffiffiffiffiffi

pTp exp � Peð1� X � T Þ2

4T

!

� Pe2

exp½ð1� X ÞPe�erfc

ffiffiffiffiffiPepð1� X þ T Þ

2ffiffiffiffiTp

� �;

ðA:3Þ

gfEðX ; T Þ ¼

ð1� X ÞffiffiffiffiffiPep

2ffiffiffipp

T32

exp � Peð1� X � T Þ2

4T

!.

ðA:4Þ

The transit time flux pdf is calculated by straight appli-cation in the time domain of the convolution integral inEq. (12):

gfT ðX ; T Þ ¼

Z T

0

gfAðX ; T 0Þgf

EðX ; T � T0ÞdT 0

¼ffiffiffiffiffiPep

2ffiffiffipp

T32

exp � Peð1� T Þ2

4T

!

¼ gfAð1; T Þ ¼ gf

Eð0; T Þ. ðA:5ÞThe transit time flux pdf in Eq. (A.5) could have beendeduced by substitution of X by 1 in Eq. (A.2), or bysubstitution of X by 0 in Eq. (A.4). The transit time res-ident pdf is found by convoluting (A.1) and (A.3) in theLaplace domain:

gT ðX ; sÞ ¼ gAðX ; sÞgEðX ; sÞ

¼ 4

1þ cð Þ2exp

Pe2

1� c½ �� �

; ðA:6Þ

where s is the Laplace variable, g is s-transformed stateof the function g, and with c ¼

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffið1þ 4s=PeÞ

p. The inver-

sion of Eq. (A.6) yields

gT ðX ; T Þ ¼ gT ðT Þ

¼ Pe 1þ Peð1þ T Þ2

� �expðPeÞerfc

Peð1þ T Þ2ffiffiffiffiffiffiffiffiPeTp

� �

� Pe2TffiffiffiffiffiffiffiffiffiffiffipPeTp exp � Peð1� T Þ2

4T

!; ðA:7Þ

where use has been made of the shifting theorem

L�1 f ðasþ bÞf g ¼ e�bta

af

ta

� ;

and of the following Laplace transform:

L�1 expð�cffiffisp Þ

ðbþffiffispÞ2

( )

¼ ð1þ bcþ 2b2tÞ expðb2t þ bcÞerfcc

2ffiffitp þ b

ffiffitp� �

� 2btffiffiffiffiffiptp exp � c2

4t

� �.

The internal age, internal life expectancy and internaltransit time pdfs are obtained from Eqs. (25), (27) and(29):

wðT Þ ¼Z 1

0

gAðX ; T ÞdX ¼Z 1

0

gEðX ; T ÞdX

¼ 1

2erfc

ffiffiffiffiffiPepðT � 1Þ

2ffiffiffiffiTp

� ��

� expðPeÞerfc

ffiffiffiffiffiPepðT þ 1Þ

2ffiffiffiffiTp

� ��; ðA:8Þ

WðT Þ ¼Z 1

0

gT ðX ; T ÞdX ¼ gT ðT Þ. ðA:9Þ

Page 23: Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

F. Cornaton, P. Perrochet / Advances in Water Resources 29 (2006) 1267–1291 1289

In 1-D, the reservoir theory is equivalent to the massconservation relation between flux and resident concen-trations given by Jury and Roth [39]:

uðT Þ ¼�owðT ÞoT

¼� o

oT

Z 1

0

gðX ;T ÞdX

¼ffiffiffiffiffiPep

2ffiffiffipp

T32

exp �Peð1� T Þ2

4T

!¼ gf

Að1;T Þ ¼ gfEð0;T Þ.

ðA:10Þ

The dimensionless outlet transit time pdf is simply theage flux concentration at X = 1, or equivalently the lifeexpectancy flux concentration at X = 0. The mean ofthe pdfs u(T), w(T), and W(T) read

st ¼ l1½u� ¼ 1; si ¼ l1½w� ¼1

2þ 1

Pe;

sit ¼ l1½W� ¼ 1þ 2

Pe¼ 2si;

ðA:11Þ

and the spreading of these pdfs is measured by theirvariance:

r2½u� ¼ 2

Pe; r2½w� ¼ ðPeþ 6Þ2

12Pe2;

r2½W� ¼ 2Peþ 6

Pe2.

ðA:12Þ

Appendix B. Reservoir theory in a single flow system

Consider the crown-shaped aquifer geometry inFig. 7. Since the flow lines remain parallel to each otherin the entire domain, we assume a 1-D transport processat each point x = hr on the flow line coordinate. We fol-low the same procedure than in Appendix A, given thatthe porous volume is M0 ¼ /pðR2 � r2

0Þ=2, the steady-flow rate F0 = KDH ln(R/r0)/p, the dispersion coefficientD(r) = aLv(r), and the pore velocity v(r) = q(r)//= KDH//pr. The age resident and flux pdfs are deducedfrom Eqs. (A.1) and (A.2) by substituting x by hr, andthe life expectancy resident and flux pdfs are deducedby replacing x by (p � h)r. The transit time flux pdf isfinally obtained by replacing x by pr in Eq. (A.2). Theaverage age, life expectancy and transit time are givenby the first moment of the corresponding flux pdfs:

hAiðh; rÞ ¼ h/pr2

KDH; ðB:1Þ

hEiðh; rÞ ¼ ðp� hÞ/pr2

KDH; ðB:2Þ

hT iðrÞ ¼ hAiðh; rÞ þ hEiðh; rÞ ¼ /p2r2

KDH. ðB:3Þ

Since on the outlet line (h = p) velocity points in thedirection of the outward unit vector, the discharge zonetransit time pdf u(t) can be evaluated by averaging the

age flux pdfs on the outlet line, or equivalently by aver-aging the life expectancy flux pdfs on the inlet line(h = 0). The internal age (or life expectancy) pdf w(t),and the internal transit time pdf W(t) are calculated fol-lowing the domain integration in Eqs. (25), (27) and(29), by integrating between the inner radius r0 and theouter radius R, and by integrating between 0 and p(see Fig. 7). The average transit time st equals the turn-over time s0 (see Eq. (45)):

st ¼ s0 ¼M0

F 0

¼ /p2ðR2 � r20Þ

2KDH lnðR=r0Þ. ðB:4Þ

The average internal age (or life expectancy) si and inter-nal transit time sit = 2si vary linearly with aL:

si ¼ l1½w� ¼/p

12KDHðR2� r20Þ½8ðR3� r3

0ÞaLþ 3pðR4� r40Þ�.

ðB:5ÞThe variance of u(t) can then be deduced from Eqs.(B.4) and (B.5) by enforcing Eq. (46).

References

[1] Amin IE, Campana ME. A general lumped parameter model forthe interpretation of tracer data and transit time calculation inhydrologic systems. J Hydrol 1996;179:1–21.

[2] Arnold L. Stochastic differential equations: theory and applica-tions. New York: John Wiley; 1974.

[3] Bear J. Dynamics of fluids in porous media. New York: Elsevier;1972.

[4] Bear J, Verruitj A. Modeling groundwater flow and pollu-tion. Boston: D. Reidel; 1987.

[5] Benjamin JR, Cornell CA. Probability, statistics, and decision forcivil engineers. New York: McGraw-Hill Book Co.; 1970.

[6] Boehlke JK, Denver JM. Combined use of groundwater dating,chemical, and isotopic analysis to resolve the history and fate ofnitrate contamination in two agricultural watersheds, AtlanticCoastal Plain, MD. Water Resour Res 1996;31(9):2319–39.

[7] Bolin B, Rodhe H. A note on the concept of age distribution andtransit time in natural reservoirs. Tellus XXV 1973;1:58–62.

[8] Busenberg E, Plummer LN. Use of chlorofluoromethanes (CCl3Fand CCl2F2) as hydrologic tracers and age-dating tools; Thealluvium and terrace system of central Oklahoma. Water ResourRes 1992;28(9):2257–83.

[9] Campana ME, Simpson E. Groundwater residence times andrecharge rates using a discrete-state compartment model and 14Cdata. J Hydrol 1984;72:171–85.

[10] Campana ME. Generation of ground-water age distributions.Ground Water 1987;25(1):51–8.

[11] Cook P, Solomon G. Recent advances in dating young ground-water: chlorofluorocarbons, 3H/3He and 85Kr. J Hydrol 1997;191:245–65.

[12] Cordes C, Kinzelbach W. Continuous groundwater velocity fieldsand path lines in linear, bilinear, and trilinear finite elements.Water Resour Res 1992;28:2903–11.

[13] Cornaton F. Deterministic models of groundwater age, lifeexpectancy and transit time distributions in advective–dispersivesystems. Ph.D. thesis, University of Neuchatel, Switzerland, 2003.

[14] Cornaton F, Perrochet P, Diersch H-JG. A finite elementformulation of the outlet gradient boundary condition for

Page 24: Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

1290 F. Cornaton, P. Perrochet / Advances in Water Resources 29 (2006) 1267–1291

convective–diffusive transport problems. Int J Numer MethodsEng 2004;61:2716–32.

[15] Crump KS. Numerical inversion of Laplace transforms usingFourier series approximation. J Ass Comp Mech1976;23(1):89–96.

[16] Dagan G. Stochastic modelling of groundwater flow by uncon-ditional and conditional probabilities, 2, The solute transport.Water Resour Res 1982;18(4):835–48.

[17] Dagan G. Solute transport in heterogeneous porous formations. JFluid Mech 1984;145:151–77.

[18] Dagan G. Theory of solute transport by groundwater. Ann RevFluid Mech 1987;19:183–215.

[19] Dagan G. Flow and transport in porous formations. Berlin,Heidelberg, Germany: Springer; 1989.

[20] Dagan G, Nguyen V. A comparison of travel time and concen-tration approaches to modelling transport by groundwater. JContam Hydrol 1989;4:79–91.

[21] Danckwerts PV. Continuous flow systems: distribution of resi-dence times. Chem Eng Sci 1953;2(1):93–102.

[22] Davis SN, Bentley HW. Dating groundwater, a short review. In:Nuclear and chemical dating techniques: Lloyd Curie. Am ChemSoc Symp Ser 1982;176:187–222.

[23] Diersch H.-J.G. About the difference between the convective andthe divergence form of the transport equation. FEFLOW Soft-ware White Papers Vol. I, 1998, WASY GmbH Berlin.

[24] De Hoog FR, Knight JH, Stokes AN. An improved method fornumerical inversion of Laplace transforms. SIAM J Sci StatistComput 1982;3:357–66.

[25] Dunkle SA, Plummer LN, Busenberg E, Phillips PJ, Denver JM,Hamilton PA, et al. Chlorofluorocarbons (CCl3F and ClCl2F2) asdating tools and hydrologic tracers in shallow groundwater of theDelmarva Peninsula, Atlantic Coastal Plain, United States. WaterResour Res 1993;29:3837–60.

[26] Eriksson E. The possible use of Tritium for estimating ground-water storage. Tellus 1958;10:472–8.

[27] Eriksson E. Compartment models and reservoir theory. Ann RevEcol Syst 1971;2:67–84.

[28] Etcheverry D, Perrochet P. Reservoir theory, groundwater transittime distributions and lumped parameter models. IAEA-SM-361/44, Vienna, 1999.

[29] Etcheverry D, Perrochet P. Direct simulation of groundwatertransit time distributions using the reservoir theory. Hydrogeol J2000;8:200–8.

[30] Etcheverry D. Une approche deterministe des distributions destemps de transit de l�eau souterraine par la theorie des reservoirs.Ph.D. thesis, University of Neuchatel, Switzerland, 2001.

[31] Frind E. Solution of the advection–dispersion equation with freeexit boundary. Numer Methods Partial Differential Equations1988;4:301–13.

[32] Garabedian PR. Partial differential equations. New York: JohnWiley; 1964.

[33] Gardiner CW. Handbook of stochastic methods for physics,chemistry and natural sciences. Berlin: Springer; 1983.

[34] Gelhar LW, Axness CL. Three-dimensional stochastic analysis ofmacro dispersion in aquifers. Water Resour Res1983;19(1):161–80.

[35] Goode DJ. Direct simulation of groundwater age. Water ResourRes 1996;32:289–96.

[36] Haitjema HM. On the residence time distributions in idealizedgroundwatersheds. J Hydrol 1995;172:127–46.

[37] Harvey CF, Gorelick SM. Temporal moment generating equa-tions: modeling transport and mass transfer in heterogeneousaquifers. Water Resour Res 1995;31:1895–912.

[38] Jury WA. Simulation of Solute transport Using a TransferFunction Model. Water Resour Res 1982;18(2):363–8.

[39] Jury WA, Roth K. Transfer functions and solute transport throughsoil: theory and applications. Basel: Birkhauser Publ.; 1990.

[40] Kolmogorov AN. Uber die analytischen Methoden in derWahrscheilichkeitsrechnung. Math Anal 1931;104:415–58.

[41] Kreft A, Zuber A. On the physical meaning of the dispersionequation and its solutions for different initial and boundaryconditions. Chem Eng Sci 1978;33:1471–80.

[42] Kreft A, Zuber A. Comments on ‘‘Flux-averaged and volume-averaged concentrations in continuum approaches to solutetransport’’ by J.C. Parker and M.Th. van Genuchten. WaterResour Res 1986;22(7):1157–8.

[43] LaBolle EM, Quastel J, Fogg GE. Diffusion theory for transportin porous media: transition-probability densities of diffusionprocesses corresponding to advection–dispersion equations.Water Resour Res 1998;34(7):1685–93.

[44] Luther KH, Haitjema HM. Numerical experiments on theresidence time distributions of heterogeneous groundwatersheds.J Hydrol 1998;207:1–17.

[45] Maloszewski P, Zuber A. Determining the turnover time ofgroundwater systems with the aid of environmental tracers, 1.Models and their applicability. J Hydrol 1982;57:207–31.

[46] Maloszewski P, Zuber A. Influence of matrix diffusion andexchange reactions on radiocarbon ages in fissured carbonateaquifers. Water Resour Res 1991;27:1937–45.

[47] Maloszewski P, Zuber A. Principles and practice of calibrationand validation of mathematical models for the interpretation ofenvironmental tracer data in aquifers. Adv Water Res1993;16:173–90.

[48] Nauman EB, Buffham BA. Mixing in continuous flow sys-tems. New York: John Wiley; 1983.

[49] Neumann SP, Winter CL, Newman CM. Stochastic theory offield-scale Fickian dispersion in anisotropic porous media. WaterResour Res 1987;23:453–66.

[50] NeupauerR,WilsonJL.Adjointmethodforobtainingbackward-in-time location and travel time probabilities of a conservativegroundwater contaminant. Water Resour Res 1999;35(11):3389–98.

[51] Neupauer R, Wilson JL. Adjoint-derived location and travel timeprobabilities for a multidimensional groundwater system. WaterResour Res 2001;37(6):1657–68.

[52] Neupauer R, Wilson JL. Backward probabilistic model ofgroundwater contamination in non-uniform and transient flow.Adv Water Res 2002;25:733–46.

[53] Parker JC, van Genuchten MTh. Flux-averaged and volume-averaged concentrations in continuum approaches to solutetransport. Water Resour Res 1984;20(7):866–72.

[54] Parker JC, van Genuchten MTh. Reply Water Resour Res1986;22(7):1159–60.

[55] Plummer LN, Michel RL, Thurman EM, Glynn PD. Environ-mental tracers for age dating young ground water. Regionalground-water quality. New York: Alley W.M., Van NostrandReinhold; 1993. p. 255–94.

[56] Reilly TE, Plummer LN, Phillips PJ, Busenberg E. The use ofsimulation and multiple environmental tracers to quantifygroundwater flow in a shallow aquifer. Water Resour Res1994;30:421–33.

[57] Richter J, Szymezak P, Abraham T, Jordan H. Use of combina-tion of lumped parameter models to interpret groundwaterisotopic data. J Contam Hydrol 1993;14(1):1–13.

[58] Robertson WD, Cherry JA. Tritium as an indicator of rechargeand dispersion in a groundwater system in Central Ontario. WaterResour Res 1989;25:1097–109.

[59] Rubin Y. Applied stochastic hydrology. New York: OxfordUniversity Press; 2003.

[60] Shapiro AM, Cvetkovic VD. Stochastic analysis of solute arrivaltime in heterogeneous porous media. Water Resour Res1988;24(10):1711–8.

[61] Saad Y, Schultz M. GMRES: A generalized minimal residualalgorithm for solving nonsymmetric linear systems. SIAM J SciStatist Comput 1986;7:856–69.

Page 25: Groundwater age, life expectancy and transit time distributions in advective–dispersive systems: 1. Generalized reservoir theory

F. Cornaton, P. Perrochet / Advances in Water Resources 29 (2006) 1267–1291 1291

[62] Saad Y. ILUT: A dual threshold incomplete ILU factorization.Linear Algebra Appl 1994;1:387–402.

[63] Solomon DK, Poreda RJ, Schiff SL, Cherry JA. Tritium andHelium 3 as groundwater-age tracers in the Borden aquifer. WaterResour Res 1992;28:741–55.

[64] Spalding DB. A note on mean residence-times in steady flows ofarbitrary complexity. Chem Eng Sci 1958;9:74–7.

[65] Sposito G, Barry DA. On the Dagan model of solute transport ingroundwater: foundational aspects. Water Resour Res1987;23(10):1867–75.

[66] Sudicky EA, Frind E. Carbon 14 dating of groundwater inconfined aquifers: implication of aquitard diffusion. Water ResourRes 1981;17:1060–4.

[67] Sudicky EA. A natural-gradient experiment on solute transport ina sand aquifer: spatial variability of hydraulic conductivity and itsrole in the dispersion process. Water Resour Res 1986;22:2069–82.

[68] Sudicky EA. The Laplace transform Galerkin technique: a time-continuous finite element theory and application to mass trans-port in groundwater. Water Resour Res 1989;25(8):1833–46.

[69] Uffink GJM. Application of the Kolmogorov�s backward equa-tion in random walk simulation of groundwater contaminanttransport. Contaminant Transport in Groundwater. Balkema,Rotterdam: H.E. Kobus and Kinzelbach; 1989.

[70] van der Vorst HA. Bi-CGSTAB: a fast and smoothly convergingvariant of Bi-CG for the solution of non-symmetric linear systems.SIAM J Sci Statist Comput 1992;12:631–44.

[71] Varni M, Carrera J. Simulation of groundwater age distributions.Water Resour Res 1998;34(12):3271–81.

[72] Weissmann GS, Zhang Y, LaBolle E, Fogg GE. Dispersion ofgroundwater age in an alluvial aquifer system. Water Resour Res2002;38(10). doi:10.1029/2001WR000907.

[73] Wilson JL, Liu J. Field validation of the backward-in-timeadvection dispersion theory. In: Proceedings of the 1996 HSRC/WERC joint conference on the environment, great plains–rockymountain hazard. Substance Cent., Manhattan, KS, 1997.

[74] Zuber A. Mathematical models for the interpretation of environ-mental radioisotopes in groundwater systems. In: Fritz P, FontesJ-Ch, editors. Handbook of environmental isotope geochemistry,vol. 2, 1986.