pii: s0309-1708(00)00023-3

13
In situ evaluation of solute retardation using single-well push–pull tests M.H. Schroth a, * , J.D. Istok a , R. Haggerty b a Department of Civil, Construction, and Environmental Engineering, 202 Apperson Hall, Oregon State University, Corvallis, OR 97331, USA b Department of Geosciences, 104 Wilkinson Hall, Oregon State University, Corvallis, OR 97331, USA Abstract More ecient methods are needed for the in situ evaluation of solute sorption to aquifer sediments. The objective of this study was to develop a simplified method for estimating retardation factors for injected solutes from ‘‘push–pull’’ test extraction phase breakthrough curves (BTCs). Sensitivity analyses based on numerical simulations were used to evaluate the method performance for a variety of test conditions. Simulations were conducted for varying retardation factors, aquifer parameters and injection phase durations, for tests performed under nonideal transport conditions such as nonlinear equilibrium and linear nonequilibrium sorption, and for a test performed in a physically heterogeneous aquifer. Predicted retardation factors showed errors 614% in tests performed under ideal transport conditions (physically homogeneous aquifer with spatially uniform dispersivity that does not vary from solute to solute, spatially uniform linear equilibrium sorption). The method performed more poorly for solutes with large retardation factors R > 20 and for tests conducted under nonideal transport conditions, and is expected to perform poorly in aquifers with highly heterogeneous sorption. In an example application, we used the method to estimate the distribution coecient for 85 Sr using data from a field test performed by Pickens JF, Jackson RE, Inch KJ, Merritt WF. (Water Resour Res 1981;17:529– 44). Reasonable agreement was found between distribution coecients obtained using the simplified method of estimation and those obtained by Pickens et al. (1981). Ó 2000 Elsevier Science Ltd. All rights reserved. Keywords: Solute transport; Sorption; Retardation; Nonuniform flow; Single-well test; Numerical simulation 1. Introduction Quantitative information on sorption is needed to predict solute transport in the subsurface for applica- tions to site characterization, risk assessment, and re- medial design. The term sorption is used here to describe any reversible reaction between a solute (hereafter re- ferred to as sorbing solute) and aquifer solids, e.g., due to ion exchange or adsorption/desorption reactions, that causes a reduced migration velocity and a delayed ar- rival time (retardation) for the sorbing solute relative to a conservative, i.e., a nonsorbing solute, hereafter re- ferred to as tracer. Sorption processes have been described by a variety of equilibrium and nonequilibrium models (e.g., [11,35]) containing one or more system specific parameters. Conventionally, these parameters have been obtained by small-scale laboratory batch and column studies per- formed with aquifer material collected from boreholes. However, it is not clear that sorption parameters ob- tained by laboratory studies are representative of in situ conditions (e.g., [14]). On the other hand, large-scale field transport experiments (e.g., [20,27,29]), which provide more representative values for sorption parameters than laboratory studies, are too time- consuming and expensive for routine use. For these reasons, fast and low-cost field methods for determining sorption parameters are needed. Single-well injection–withdrawal tests, which we call ‘‘push–pull’’ tests, have been used for the quantitative determination of a wide range of aquifer physical, bio- logical, and chemical characteristics. In a push–pull test, a prepared test solution containing one or more solutes is injected (‘‘pushed’’) into the aquifer using an existing well; the test solution/groundwater mixture is then ex- tracted (‘‘pulled’’) from the same location. Aquifer characteristics are determined by an analysis of solute breakthrough curves (BTCs) obtained by measuring www.elsevier.com/locate/advwatres Advances in Water Resources 24 (2001) 105–117 * Corresponding author. Present address: Swiss Federal Institute of Technology (ETH), Zurich, Institute of Terrestrial Ecology, Grab- enstr. 3, CH-8952 Schlieren, Switzerland. Tel.: +41-1-633-6039; fax: +41-1-633-1122. E-mail address: [email protected] (M.H. Schroth). 0309-1708/01/$ - see front matter Ó 2000 Elsevier Science Ltd. All rights reserved. PII: S 0 3 0 9 - 1 7 0 8 ( 0 0 ) 0 0 0 2 3 - 3

Upload: others

Post on 17-Nov-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: PII: S0309-1708(00)00023-3

In situ evaluation of solute retardation using single-well push±pulltests

M.H. Schroth a,*, J.D. Istok a, R. Haggerty b

a Department of Civil, Construction, and Environmental Engineering, 202 Apperson Hall, Oregon State University, Corvallis, OR 97331, USAb Department of Geosciences, 104 Wilkinson Hall, Oregon State University, Corvallis, OR 97331, USA

Abstract

More e�cient methods are needed for the in situ evaluation of solute sorption to aquifer sediments. The objective of this study

was to develop a simpli®ed method for estimating retardation factors for injected solutes from ``push±pull'' test extraction phase

breakthrough curves (BTCs). Sensitivity analyses based on numerical simulations were used to evaluate the method performance for

a variety of test conditions. Simulations were conducted for varying retardation factors, aquifer parameters and injection phase

durations, for tests performed under nonideal transport conditions such as nonlinear equilibrium and linear nonequilibrium

sorption, and for a test performed in a physically heterogeneous aquifer. Predicted retardation factors showed errors 614% in tests

performed under ideal transport conditions (physically homogeneous aquifer with spatially uniform dispersivity that does not vary

from solute to solute, spatially uniform linear equilibrium sorption). The method performed more poorly for solutes with large

retardation factors �R > 20� and for tests conducted under nonideal transport conditions, and is expected to perform poorly in

aquifers with highly heterogeneous sorption. In an example application, we used the method to estimate the distribution coe�cient

for 85Sr using data from a ®eld test performed by Pickens JF, Jackson RE, Inch KJ, Merritt WF. (Water Resour Res 1981;17:529±

44). Reasonable agreement was found between distribution coe�cients obtained using the simpli®ed method of estimation and those

obtained by Pickens et al. (1981). Ó 2000 Elsevier Science Ltd. All rights reserved.

Keywords: Solute transport; Sorption; Retardation; Nonuniform ¯ow; Single-well test; Numerical simulation

1. Introduction

Quantitative information on sorption is needed topredict solute transport in the subsurface for applica-tions to site characterization, risk assessment, and re-medial design. The term sorption is used here to describeany reversible reaction between a solute (hereafter re-ferred to as sorbing solute) and aquifer solids, e.g., dueto ion exchange or adsorption/desorption reactions, thatcauses a reduced migration velocity and a delayed ar-rival time (retardation) for the sorbing solute relative toa conservative, i.e., a nonsorbing solute, hereafter re-ferred to as tracer.

Sorption processes have been described by a varietyof equilibrium and nonequilibrium models (e.g., [11,35])containing one or more system speci®c parameters.

Conventionally, these parameters have been obtained bysmall-scale laboratory batch and column studies per-formed with aquifer material collected from boreholes.However, it is not clear that sorption parameters ob-tained by laboratory studies are representative of in situconditions (e.g., [14]). On the other hand, large-scale®eld transport experiments (e.g., [20,27,29]), whichprovide more representative values for sorptionparameters than laboratory studies, are too time-consuming and expensive for routine use. For thesereasons, fast and low-cost ®eld methods for determiningsorption parameters are needed.

Single-well injection±withdrawal tests, which we call``push±pull'' tests, have been used for the quantitativedetermination of a wide range of aquifer physical, bio-logical, and chemical characteristics. In a push±pull test,a prepared test solution containing one or more solutesis injected (``pushed'') into the aquifer using an existingwell; the test solution/groundwater mixture is then ex-tracted (``pulled'') from the same location. Aquifercharacteristics are determined by an analysis of solutebreakthrough curves (BTCs) obtained by measuring

www.elsevier.com/locate/advwatres

Advances in Water Resources 24 (2001) 105±117

* Corresponding author. Present address: Swiss Federal Institute of

Technology (ETH), Zurich, Institute of Terrestrial Ecology, Grab-

enstr. 3, CH-8952 Schlieren, Switzerland. Tel.: +41-1-633-6039; fax:

+41-1-633-1122.

E-mail address: [email protected] (M.H. Schroth).

0309-1708/01/$ - see front matter Ó 2000 Elsevier Science Ltd. All rights reserved.

PII: S 0 3 0 9 - 1 7 0 8 ( 0 0 ) 0 0 0 2 3 - 3

Page 2: PII: S0309-1708(00)00023-3

solute concentrations at the well during the extractionphase.

Push±pull tracer tests have been used to study pro-cesses of advection and dispersion during nonuniform¯ow. A general form of the advection±dispersion equa-tion for nonuniform radial ¯ow was developed byHoopes and Harleman [15] to study the displacementand mixing of groundwater by waste water injected intoa single well. Type curves for the analysis of radial-¯ow®elds were presented by Sauty [30]. Methods to estimatelongitudinal dispersivity from push±pull tests were de-veloped by Mercado [22] and Gelhar and Collins [6].Pickens and Grisak [25] used push±pull tests to measure®eld-scale dispersion in a strati®ed aquifer. Based on theearlier work conducted by Leap and Kaplan [19], Hallet al. [13] derived equations to determine the e�ectiveporosity and regional groundwater velocity from push±pull test results. In addition, push±pull tests have provenuseful as a diagnostic tool to quantify matrix di�usion infractured rock (e.g., [10,21]).

Push±pull tests have also been used to quantify mi-crobial metabolic activity including denitri®cation [33],petroleum hydrocarbon degradation [28], and aerobicrespiration, denitri®cation, sulfate reduction, andmethanogenesis [17,31]. Simpli®ed methods for deter-mining zero- and ®rst-order reaction-rate coe�cientsfrom push±pull tests were recently developed by Hagg-erty et al. [12] and Snodgrass and Kitanidis [32].

Only a few studies have addressed the use of push±pull tests to characterize sorption of injected solutes toaquifer sediments. Drever and McKee [3] performedlarge-scale push±pull tests in an attempt to determineLangmuir isotherm sorption parameters for cations in-troduced during in situ coal gasi®cation and uraniumrecovery. However, in their analysis, Drever and McKee[3] did not speci®cally take into account the radial-¯ow®eld in the vicinity of the injection/extraction well duringthe test. Instead sorption parameters were determinedby matching the tails of extraction phase BTCs to typecurves prepared by numerical simulation. Pickens et al.[26] performed push±pull tests to determine sediment±water distribution coe�cients for 85Sr in a strati®edaquifer. However, Pickens et al. [26] used only injectionphase data from observation wells and did not attemptto estimate distribution coe�cients from extractionphase BTCs.

The objective of this study was to develop a simpli®edmethod for estimating solute retardation factors for in-jected solutes from push±pull test BTCs. Sensitivityanalyses were used to evaluate the method performancefor a variety of simulated test conditions. Simulationswere conducted for varying solute retardation factors,aquifer parameters and injection phase durations, fortests performed under nonideal transport conditionssuch as nonlinear equilibrium sorption and linear non-equilibrium sorption, and for a test performed in a

physically heterogeneous con®ned aquifer. In an exam-ple application, we used the method to estimate the re-tardation factor and distribution coe�cient for 85Srusing data from a ®eld test performed by Pickens et al.[26].

2. Solute transport theory and simpli®ed method develop-

ment

The governing equation for one-dimensional solutetransport subject to advection, dispersion and sorptionduring nonuniform, radial ¯ow in a homogeneous,con®ned aquifer is (adopted from [1,34])

oCot� qb

noSot� aLjvj o

2Cor2ÿ v

oCor; �1�

where C and S are solute aqueous phase and solid(sorbed) phase concentrations, respectively, qb is bulkdensity, n is e�ective porosity, aL is longitudinal dis-persivity, t is time, r is radial stance, and the averagepore water velocity v is

v � Q=2pbnr; �2�where Q is the pumping rate (positive during the injec-tion phase and negative during the extraction phase),and b is aquifer thickness. Note that Eq. (2) assumesthat the regional groundwater velocity is negligiblecompared to the imposed velocity due to pumping.Eq. (1) assumes that molecular di�usion is negligible andthat mechanical dispersion is a linear function of v (e.g.,[1]).

Prior to solving Eq. (1), a sorption model must bespeci®ed. While a large number of such models areavailable (e.g., [35]), we restrict our analysis at this pointto the linear equilibrium sorption isotherm

S � KdC; �3�where Kd is the distribution coe�cient, which we assumeto be spatially uniform. Additional information on thee�ects of the assumption of equilibrium sorption ontransport in a radial-¯ow ®eld is given in [34]. Substi-tuting Eq. (3) into Eq. (1), the solute retardation factorR can be de®ned as

R � 1� qb

nKd: �4�

Note that for a tracer Kd � 0 and therefore R � 1. UsingEq. (2), an ``e�ective'' velocity v* for a sorbing solutecan be de®ned as

v� � vR� Q

2pbnrR: �5�

During the injection phase of a push±pull test, a soluteÕsfrontal position r̂inj (de®ned here as the location whereC=Co � 0:5, where Co is the solute concentration in the

106 M.H. Schroth et al. / Advances in Water Resources 24 (2001) 105±117

Page 3: PII: S0309-1708(00)00023-3

injected test solution) is obtained by rearranging andintegrating Eq. (5)

r̂inj ������������������������Qinjtinj

pbnR� r2

w

r; �6�

where Qinj is the pumping rate during the injectionphase, tinj is the time since injection began, and rw is theinjection/extraction well radius. At the end of the in-jection phase, the frontal position attains a maximumvalue r̂max with

r̂max ����������������������Vinj

pbnR� r2

w

r�7�

with the total volume injected Vinj � QinjTinj, where Tinj isthe duration of the injection phase.

During the extraction phase, ¯ow is reversed and thefrontal position r̂ext is given by

r̂ext ����������������������������r̂2

max �Qexttext

pbnR

r; �8�

where Qext is the pumping rate during the extractionphase (not necessarily equal in magnitude to Qinj) andtext is the time since extraction began. Note that Eq. (8) isonly meaningful for jQextjtext6 Vinj.

Eqs. (6)±(8) reveal the inverse quadratic variation infrontal position with time due to the radial-velocity ®eld(Eqs. (2) and (5)). During the injection phase, a sorbingsolute �R > 1� will travel at a smaller e�ective velocity(Eq. (5)) and penetrate a smaller radial distance into theaquifer (Eqs. (6) and (7)) than a tracer (Fig. 1(a)).Conversely, the di�erence in frontal positions for asorbing solute and tracer will continuously decreaseduring the extraction phase (Eq. (8), Fig. 1(b)). Finally,when jQextjtext � Vinj, the frontal positions for the sorbingsolute and tracer will coincide at r � rw. Thus, unlike ina well-to-well test in which the di�erence in arrival timesfor a sorbing solute and a tracer can be used to estimate

a retardation factor, in a push±pull test, the arrival timesof a sorbing solute and tracer during the extractionphase are identical and hence provide no information onthe retardation factor of the injected sorbing solute.However, in the next section, we show that retardationfactors can be estimated from extraction phase BTCs byconsidering dispersion.

2.1. Simpli®ed method for estimating retardation factorsfrom push±pull tests

In this section, we develop a simpli®ed method forestimating solute retardation factors from push±pulltests based on a previously conducted analysis of lon-gitudinal dispersion during radial ¯ow [6].

Considering longitudinal dispersion only, Gelhar andCollins [6] derived approximate solutions to Eq. (1) withS� 0 by rewriting Eq. (1) in the form of a di�usionequation. Neglecting molecular di�usion, the approxi-mate solution for tracer concentration during the in-jection phase of a push±pull test is (adopted from [6])

CCo

� 1

2erfc r2

�(ÿ r̂2

inj

� 16

3aL r̂3

inj

��,ÿ r3

w

��1=2); �9�

where r̂inj is given by Eq. (6).During the subsequent extraction phase, the ap-

proximate solution for tracer concentration is given by

CCo

� 1

2erfc r2

�(ÿ r̂2

ext

� 16

3aL 2r̂3

max

��,ÿ r̂3

ext ÿ r3w

��1=2);

�10�where r̂max and r̂ext are given by Eqs. (7) and (8), re-spectively. Furthermore, neglecting the well radius inEq. (10), the approximate solution for tracer concen-tration at the injection/extraction well during the ex-traction phase is [6]

Fig. 1. Frontal positions �C=Co � 0:5� for a tracer (subscript ``tr'') and a co-injected sorbing solute (subscript ``sol'') during: (a) the injection phase

and (b) prior to (gray lines and labels) and during the extraction phase of a push±pull test conducted under ideal transport conditions in a

homogeneous con®ned aquifer.

M.H. Schroth et al. / Advances in Water Resources 24 (2001) 105±117 107

Page 4: PII: S0309-1708(00)00023-3

CCo

� 1

2erfc

Vext

Vinj

�8<: ÿ 1

�16

3

aL

r̂max

�2

,ÿ 1

���� ÿ Vext

Vinj

����1=2

� 1

�ÿ Vext

Vinj

�!#1=29=;; �11�

where Vext is the cumulative extracted volume�Vext � jQextjtext� and r̂max is given by Eq. (7) with rw � 0.Eqs. (7) and (11) may be used to determine aL from apush±pull test tracer BTC obtained at the injection/ex-traction well during the extraction phase of the test.First, the tracer frontal position at the end of the in-jection phase, r̂max;tr, is computed from Eq. (7) based onknown values of Vinj; b; n and R � 1. Then, Eq. (11) withr̂max � r̂max;tr is ®t to the tracer BTC to obtain an esti-mate for aL.

Under ideal transport conditions (physically homo-geneous aquifer with spatially uniform dispersivity thatdoes not vary from solute to solute, spatially uniformlinear equilibrium sorption) the dispersivity is a propertyof the aquifer and should therefore be identical for asorbing solute and a co-injected tracer. Di�erences in theshape of tracer and sorbing solute BTCs (Eq. (11)) maythen be attributed to retardation of the sorbing soluterelative to the tracer. Thus, we propose the followingmethod for estimating retardation factors for a sorbingsolute from push±pull test extraction phase BTCs: (1)determine r̂max;tr from Eq. (7) and aL from the tracerextraction phase BTC using Eq. (11); (2) keeping aL

®xed in Eq. (11), estimate the sorbing solute frontalposition at the end of the injection phase r̂max;sol by ®t-ting Eq. (11) to the sorbing solute BTC; and (3) em-ploying Eq. (7), an estimate of the retardation factor forthe sorbing solute (R�) can be obtained from

R� � �r̂max;tr=r̂max;sol�2: �12�

Note that this method does not require the use of a ¯owand transport model, and it can be easily implementedusing a spreadsheet program. Nonetheless, other meth-ods may be employed to estimate R from push±pull testBTCs. For example, numerical solutions to Eq. (1) maybe used to estimate R from sorbing solute BTCs usinginverse modeling.

3. Numerical methods

3.1. Overview

One- and two-dimensional simulations were con-ducted (1) to investigate the e�ects of varying retarda-tion factors on BTCs obtained at the injection/extraction well during the extraction phase of a push±pull test, and (2) to evaluate the performance of the

simpli®ed method for estimating retardation factors fora variety of simulated test conditions. Simulations wereperformed using the Subsurface Transport Over Multi-ple Phases (STOMP) code [36]. STOMP is a fully im-plicit volume-integrated ®nite di�erence simulator formodeling one-, two-, and three-dimensional ¯ow andtransport, which has been extensively tested and vali-dated against published analytical solutions as well asother numerical codes (e.g., [24]). For all simulations,the total variation diminishing scheme was employed[37], which consists of a third-order ¯ux limiter per-mitting oscillation-free numerical solutions while mini-mizing numerical dispersion for a wide range of gridPeclet numbers [8]. An additional simulation for con-ditions of linear nonequilibrium sorption was performedusing the STAMMT-R code [9], which solves transportand multirate mass-transfer equations for push±pull andwell-to-well tracer tests. The solutions assume homo-geneous media, but deal with a wide range of mass-transfer scenarios.

For all simulations, extraction phase BTCs were ob-tained by tracking tracer and sorbing solute ¯uxes at theinjection/extraction well. BTCs were then analyzed toobtain estimates for retardation factors for the sorbingsolutes and these were compared with the values of re-tardation factors used in the numerical simulations.

3.2. Simulations of tests conducted under ideal transportconditions

One-dimensional simulations of tests conducted un-der ideal transport conditions (con®ned aquifer withhomogeneous physical and chemical properties; linearequilibrium sorption) were performed for a tracer andsix sorbing solutes, for which values of Kd were selectedto give computed retardation factors ranging between 2and 100 (Eq. (4)).

The computational domain consisted of a line of 250nodes with a uniform node spacing of Dr � 1:0 cm. Thetime-step size during the simulations varied between 2and 10 s. Initial conditions were a constant hydraulichead for the aqueous phase and C � 0 and S � 0 for allsolutes. Time-varying third-type ¯ux boundary condi-tions at rw � 252:5 cm were used to represent pumpingat the injection/extraction well; constant head and zerosolute ¯ux boundary conditions at r � 252:5 cm wereused to represent aquifer conditions beyond the radiusof in¯uence of the well.

Simulations were conducted for a ``base case'' (BC)and a set of six additional cases, which had values of asingle parameter �aL; n or Tinj� larger or smaller than inthe BC. BC parameters �aL � 1:0 cm; n � 0:35; Tinj �12 h, b � 20 cm; vn � �0:83 cm=min at r � rw� werechosen to represent a typical test conducted in an un-consolidated sand. The duration of the injection phasewas selected to give r̂max � 93 cm for the tracer. Values

108 M.H. Schroth et al. / Advances in Water Resources 24 (2001) 105±117

Page 5: PII: S0309-1708(00)00023-3

for aL similar to the BC value of 1.0 cm were observedpreviously in laboratory experiments conducted oversimilar length scales with a sandy loam [16]. For theadditional cases, aL was varied from 0.1 to 10 cm, n wasvaried from 0.25 to 0.45, and Tinj was varied from 6 to24 h, while vn and b were identical for all simulationsduring both injection and extraction phases.

3.3. Simulations of tests conducted under nonideal trans-port conditions

One-dimensional simulations were performed fortests conducted under conditions of nonlinear equilib-rium sorption and linear nonequilibrium sorption. Inaddition, a two-dimensional simulation was performedfor a test conducted in a physically heterogeneous con-®ned aquifer.

For the simulation of a test conducted under condi-tions of nonlinear equilibrium sorption, we substitutedthe sorption term in Eq. (1) with the Langmuir sorptionisotherm (e.g., see [4])

S � SmaxKLC1� KLC

; �13�

where Smax is the maximum sorption capacity, and KL isthe Langmuir coe�cient. Note that for C � 1=KL;S � SmaxKLC, which represents the linear sorption iso-therm (Eq. (3)) with Kd � SmaxKL. The simulation wasperformed for a tracer, a solute subject to linear sorp-tion �R � 10�, and ®ve additional solutes subject toLangmuir-type sorption, for which SmaxKL was identicalto that of the linearly sorbing solute. Variation in thesorption behavior of solutes subject to Langmuir-typesorption was introduced by varying the solutesÕ Co, thusaltering C in Eq. (13). When expressed in terms of adimensionless number, �C0, with �C0 � CoKL, values em-ployed were in the range of 0:56 �C06 500. The com-putational domain, as well as initial and boundaryconditions for this simulation were identical to thosedescribed in Section 3.2.

For the simulation of a test conducted under condi-tions of linear nonequilibrium sorption, we substitutedthe sorption term in Eq. (1) with the ®rst-order chemicalnonequilibrium model (e.g., see [11,34])

oSot� k KdC� ÿ S�; �14�

where k is the ®rst-order rate coe�cient, which may beexpressed in dimensionless form as �k � k2pbna2

L=Q [34].The simulation was performed for a tracer, a solutesubject to linear equilibrium sorption �R � 10�, and ®veadditional solutes subject to nonequilibrium sorption.Values of Kd were identical for all sorbing solutes.Variation in the sorption behavior of solutes subject tononequilibrium sorption was introduced by varying �k inthe range of 10ÿ16 �k6 10ÿ6. This simulation was per-

formed using the STAMMT-R code with the compu-tational domain, initial and boundary conditionsidentical to those described in Section 3.2.

For the simulation of a test conducted across theentire saturated thickness of a heterogeneous con®nedaquifer, the two-dimensional computational domainconsisted of an array of 1200 nodes distributed in aradial (r)±vertical (z) coordinate system (where z is thevertical distance from the base of the aquifer). We useda variable node spacing of Dr � 5 cm for 0 cm <�r ÿ rw� < 50 cm;Dr � 10 cm for 50 cm < �r ÿ rw� <200 cm;Dr � 20 cm for 200 cm < �r ÿ rw� < 300 cm,and a uniform node spacing of Dz � 5 cm for0 cm6 z6 200 cm. The domain was assumed to containlayers and lenses of three di�erent isotropic porousmedia with aL � 1:0 cm and aT � 0:05 cm for all media.Layers of medium 1 (n � 0:35, hydraulic conductivityK � 4:33 cm=min) and medium 2 (n � 0:40; K �0:35 cm=min) were placed into the domain at random,followed by randomly replacing areas within layers ofmedium 1 and 2 with 15 lenses of medium 3 (n � 0:45,K � 0:0035 cm/min) (see Fig. 5(a) in Section 4). Thesimulation time step size varied between 2 and 10 s.Initial conditions were a constant hydraulic head for theaqueous phase and C � 0 and S � 0 for all solutes.Time-varying ®rst-type (Dirichlet) boundary conditions(resulting in vn � �0:83 cm=min at r � rw, averagedacross the entire thickness), together with time-varyingthird-type solute ¯ux boundary conditions were used atrw � 2:5 cm to represent pumping at the injection/ex-traction well. Zero ¯ux boundary conditions were im-posed along the aquifer base �z � 0 cm� and top�z � 200 cm�; constant head and zero solute ¯uxboundary conditions at r � 302:5 cm were used to rep-resent aquifer conditions beyond the radius of in¯uenceof the well. The injection/extraction of a tracer and sixsorbing solutes �26R6 100� was simulated along0 cm < z < 200 cm for Tinj � 12 h. Test conditions forthis simulation were selected such that r̂max;tr6 150 cm inany medium.

4. Results and discussion

4.1. Simulations of tests conducted under ideal transportconditions

Extraction phase BTCs for the BC simulation of atest in a physically and chemically homogeneous con-®ned aquifer with sorbing solutes subject to linearequilibrium sorption showed an almost simultaneousarrival of the frontal positions �C=Co � 0:5� for allsolutes near Vext=Vinj � 1:0 (Fig. 2), as predicted byEq. (8). However, no perfect coincidence of the BTCswas obtained at C=Co � 0:5 and Vext=Vinj � 1:0, as e.g.,predicted by Eq. (11), due to the variable dispersion in

M.H. Schroth et al. / Advances in Water Resources 24 (2001) 105±117 109

Page 6: PII: S0309-1708(00)00023-3

the nonlinear ¯ow ®eld. Similar results were obtainedfor all simulations conducted in a homogeneous con-®ned aquifer (not shown), verifying that di�erences inarrival times do not provide su�cient information toestimate retardation factors.

However, BTCs for solutes with increasing retarda-tion factors showed an increasing width of their dis-persed fronts (Fig. 2). The width of the dispersed frontcan be de®ned as the increment D�Vext=Vinj� between theintercepts of the tangent line at C=Co � 0:5 with hori-zontal lines at C=Co � 1:0 and 0.0 [22]. This observationis in agreement with our analysis of extraction phaseBTCs for sorbing solutes during radial ¯ow based on thetheory developed by Gelhar and Collins [6], wherechanges in the shape of the BTCs (Eq. (11)) are expectedas a function of R1=2 (Eq. (7)).

A set of BTCs identical to those presented in Fig. 2was obtained from multiple simulations of tracer push±pull tests using the same parameter set as for the BCsimulation except for Tinj, which was varied in fractionsof the duration of the injection phase of the BC simu-lation Tinj;bc (Fig. 3). Note that the fractions of Tinj;bc usedto generate Fig. 3 (2±100) are identical to retardationfactors chosen for the BC simulation (Fig. 2). Therefore,Fig. 3 displays the extraction phase BTCs for tracersthat were pushed to the same value of r̂max as the cor-responding sorbing solutes in Fig. 2 (Eq. (7)). Thus, theamount of variable dispersion during transport for atracer is identical to that of a sorbing solute with iden-tical r̂max. This may be explained as follows. If Eqs. (2)and (3) are inserted into Eq. (1), and we divide the right-hand side of Eq. (1) by R, then the hydrodynamic dis-persion coe�cient becomes aLjvj=R, thus indicating a

smaller dispersion coe�cient for a sorbing solute thanfor a tracer. However, dispersion will act upon a sorbingsolute R times longer than on the corresponding tracer,since the arrival time of a sorbing solute t̂sol at anylocation along the ¯ow path is related to that of a tracert̂tr at that location by t̂sol=t̂tr � R. Thus, prolonged dis-persion will exactly o�set the reduction in the dispersioncoe�cient for a sorbing solute. Note that this statementalso applies to ¯ow in a uniform ¯ow ®eld and that it isonly valid under ideal transport conditions, as de®nedearlier.

Using BC parameters in Eq. (7), we computedr̂max;tr � 92:9 cm, and we obtained aL � 0:98 cm by ®t-ting Eq. (11) to the tracer BTC (Fig. 2). The ®tted valueof aL is in good agreement with the simulated value ofaL � 1:0 cm. Similar results were obtained for simula-tions in which n and Tinj were varied, with estimates foraL ranging between 0.97 and 0.99. Using the simpli®edmethod of estimation for radial ¯ow, we obtained esti-mates of r̂max and predicted retardation factors R� for allsolutes during simulations of push±pull tests under idealtransport conditions (Table 1). Di�erences 614% be-tween values of R and R� were observed for all solutesfor the BC simulation as well as for simulations in whichn and Tinj were varied.

In general, di�erences between R and R� increasedwith increasing R (Table 1). This may be explained byconsidering the mean-squared error (MSE) betweensimulated and ®tted (Eq. (11)) BTCs, which also in-creased with increasing R (not shown). For example, forthe BC simulation, the MSE between ®tted and simu-lated tracer BTC was 1:66� 10ÿ4, and the MSE in-creased monotonically to 5:64� 10ÿ3 for the ®t of the

Fig. 3. Simulated tracer breakthrough curves obtained at the injection/

extraction well during the extraction phase of multiple push±pull tests

conducted under ideal transport conditions for varying total injection

times Tinj, shown as fractions of the total injection time for the BC

simulation Tinj;bc.

Fig. 2. Simulated breakthrough curves for a tracer and co-injected

sorbing solutes �26R6 100� obtained at the injection/extraction well

during the extraction phase of a push±pull test conducted under ideal

transport conditions using BC parameters.

110 M.H. Schroth et al. / Advances in Water Resources 24 (2001) 105±117

Page 7: PII: S0309-1708(00)00023-3

Tab

le1

Bes

tvalu

eso

ffr

on

tal

po

siti

on

sat

the

end

of

the

inje

ctio

np

hase�r̂ m

ax�a

nd

esti

mate

dso

lute

reta

rdati

on

fact

ors�R� �

for

solu

tes

insi

mu

lati

on

so

fp

ush

±p

ull

test

sco

nd

uct

edu

nd

erid

eal

tran

spo

rt

con

dit

ion

sa

Sim

ula

tio

nB

CL

ow

n(0

.25

)H

igh

n(0

.45)

Lo

wT

inj

(6h

)H

igh

Tin

j

(24

h)

Lo

wa L

(0.1

cm)

Hig

ha L

(10

cm)

BC

,gu

ess

n�

0.2

5

BC

,gu

ess

n�

0.4

5

Sim

ula

ted

Rr̂ m

ax

(cm

)

R�

r̂ max

(cm

)

R�

r̂ max

(cm

)

R�

r̂ max

(cm

)

R�

r̂ max

(cm

)

R�

r̂ max

(cm

)

R�

r̂ max

(cm

)

R�

r̂ max

(cm

)

R�

r̂ max

(cm

)

R�

Tra

cerb

92

.91

.00

11

01

.00

81.9

1.0

065.7

1.0

0131

1.0

092.9

1.0

092.9

1.0

0110

1.0

081.9

1.0

0

26

5.9

1.9

97

7.9

1.9

958.2

1.9

846.7

1.9

893.1

1.9

964.0

2.1

167.2

1.9

178.0

1.9

958.2

1.9

8

54

2.2

4.8

54

9.7

4.8

937.3

4.8

230.1

4.7

659.3

4.9

139.5

5.5

342.3

4.8

249.9

4.8

537.2

4.8

5

10

30

.39

.40

35

.69

.53

26.8

9.3

421.6

9.2

542.4

9.6

027.5

11.4

28.4

10.7

35.8

9.4

226.7

9.4

1

20

21

.81

8.2

25

.61

8.4

19.3

18.0

15.5

18.0

30.4

18.7

19.2

23.4

17.9

26.9

25.8

18.1

19.2

18.2

50

14

.04

4.0

16

.54

4.4

12.4

43.6

9.7

145.8

19.7

44.5

11.8

61.9

8.5

7117

16.6

43.8

12.4

43.6

10

09

.79

90

.01

1.7

88

.28.5

192.6

6.4

3104

14.1

86.4

8.0

2135

4.2

1487

11.6

89.8

8.6

390.1

aV

alu

eso

fR�

wer

eo

bta

ined

usi

ng

the

sim

pli

®ed

met

ho

do

fes

tim

ati

on

.B

Csi

mu

lati

on

para

met

ers

wer

ea L�

1:0

cm;

n�

0:3

5,

an

dT i

nj�

12

h.

bV

alu

eso

fr̂ m

ax;t

rw

ere

com

pu

ted

usi

ng

Eq

.(7

).

M.H. Schroth et al. / Advances in Water Resources 24 (2001) 105±117 111

Page 8: PII: S0309-1708(00)00023-3

R � 100 BTC. This phenomenon is due to the fact thatEqs. (9)±(11) are approximate solutions. As speci®ed byGelhar and Collins [6], these approximate solutions areonly accurate if �aL=Lo�1=2 � 1, where Lo is the totaldistance traveled by the solute front. In other words, ahigher accuracy is achieved when the solute dispersedzone width is small compared to Lo. However, Lo de-creased as a function of R (note that for a push±pull testLo � 2r̂max, Eq. (7)). Thus, less accurate ®ts of Eq. (11) toBTCs are expected for solutes with larger values of R.

The foregoing analysis is particularly relevant tosimulations in which aL was varied (Table 1). For thecase of aL � 10 cm, reasonably accurate predictions ofretardation factors were obtained only for solutes withR6 20. Di�erences between R and R� dramatically in-creased for sorbing solutes with R > 20 because the as-sumption of �aL=Lo�1=2 � 1 was increasingly violated.For example, we computed �aL=Lo�1=2 � 1:09 for thesolute with R � 100, and the poor agreement betweenthe simulated and the ®tted BTC was re¯ected in a MSEvalue of 7:51� 10ÿ2.

To corroborate this line of reasoning, we conductedan additional simulation in which Tinj was increased by afactor of 10. For this simulation, the numerical domainwas extended to a total length of 7.5 m. In case of thesolute with R � 100, we found R� � 85:9 with a MSE of2:49� 10ÿ2. Thus, for solutes with larger values of R,larger values of Lo and therefore r̂max (i.e., a longer in-jection phase duration or higher injection rate) are re-quired to more closely satisfy �aL=Lo�1=2 � 1, and henceto improve the agreement between R and R�.

Larger discrepancies between R and R� compared tothe BC were also observed in simulations withaL � 0:1 cm for solutes with larger values of R (Table 1).Causes for these discrepancies remain uncertain, assmall values of MSE ranging between 4:75� 10ÿ6 and1:66� 10ÿ4 indicated good agreement between simulat-ed and ®tted BTCs. A possible explanation is that nu-merical dispersion may have been more important inthese simulations due to the small value of aL.

So far, we assumed complete knowledge of parametervalues required in the computation of r̂max (Eq. (7)).However, the e�ective porosity n in particular is usuallynot known precisely. Similarly, the aquifer thickness bmay be uncertain. We investigated the e�ect of uncer-tainty in n or b on the method performance by reana-lyzing the BC simulation using values of n that wereboth lower (0.25) and higher (0.45) than the actualsimulated value (0.35, Table 1). (Note that changing n orb has an equivalent e�ect on r̂max (Eq. (7)).) Interest-ingly, results almost identical to the original BC analysiswere obtained for both low and high values of n. Dif-ferences 613% between R and R� were found for allvalues of R, and MSE values were similar to those in theBC analysis (not shown). The relative insensitivity of themethod to initial guesses of n may be explained as fol-

lows. Predicted values of r̂max for tracer and sorbingsolutes are equally a�ected by uncertainties in n (andwould be equally a�ected by uncertainties in b). There-fore, using the ratio of r̂max;tr and r̂max;sol to estimate re-tardation factors (Eq. (12)) e�ectively cancels out e�ectsof uncertainties in n or b.

4.2. Simulations of tests conducted under nonideal trans-port conditions

Extraction phase BTCs for the simulation of a testconducted under conditions of nonlinear equilibriumsorption (Fig. 4) were di�erent from those obtainedunder conditions of linear equilibrium sorption (Fig. 2).Increasing �C0 in Fig. 4 enhanced di�erences in BTCsbetween solutes subject to Langmuir-type sorption andthe linearly sorbing solute �R � 10�. As a result of in-creasing �C0, C also increased during the test and thesorption behavior gradually shifted from the linearsorption branch of the Langmuir isotherm�S � SmaxKLC for C � 1=KL� to the branch wheresorption is limited by the number of sorption sites�S � Smax for C � 1=KL�. The most extreme case forthis behavior is the BTC of the sorbing solute with�C0 � 500, which closely matched the tracer BTC in Fig. 4,indicating almost conservative transport of the sorbingsolute. Perhaps the most obvious visual di�erence inFig. 4 compared to Fig. 2 is that sorbing solute BTCs inFig. 4 do not all coincide near Vext=Vinj � 1. Rather, thelocation where sorbing solute BTCs cross the tracerBTC shifted gradually toward signi®cantly larger values

Fig. 4. Simulated breakthrough curves for a tracer, a solute subject to

linear equilibrium sorption �R � 10�, and ®ve solutes subject to non-

linear (Langmuir-type) equilibrium sorption obtained at the injection/

extraction well during the extraction phase of a push±pull test. Vari-

ations in the sorption behavior of solutes subject to nonlinear sorption

was introduced by varying the solutesÕ injection concentration, dis-

played here in dimensionless form as �C0.

112 M.H. Schroth et al. / Advances in Water Resources 24 (2001) 105±117

Page 9: PII: S0309-1708(00)00023-3

of Vext=Vinj. For example, for the sorbing solute with�C0 � 500, that location was near Vext=Vinj � 1:5. Thisphenomenon may serve as a qualitative indicator ofnonideal transport conditions during a push±pull test.

As a consequence of the di�erences in the shape ofBTCs obtained in Fig. 4, values of R* initially increasedfor solutes subject to Langmuir-type sorption �0:56�C06 5:0� (Table 2). This is in disagreement with pre-dicted Langmuir retardation factors, RL, given as (e.g.,[4])

RL � 1� qb

nSmaxKL

1� KLC� �2 !

; �15�

which should monotonically decrease with increasing �C0

and, hence, increasing C. In our simulation, this is trueonly for sorbing solutes with 506 �C06 500. In addition,agreement between simulated and ®tted (Eq. (11)) BTCswas increasingly poor for sorbing solutes with0:56 �C06 50, as witnessed by increasing MSR values(Table 2). Thus, our simpli®ed method for estimating Rappears not be generally applicable to cases of nonlinearequilibrium solute sorption. For cases in which Lang-muir-type sorption is suspected during a push±pull test,we therefore refer the reader to Drever and McKee [3],who have developed a method for estimating RL frompush±pull tests by matching tails of extraction phaseBTCs to type curves. Moreover, performing multiplepush±pull tests with varying �C0 may help assess thepresence of nonlinear sorption conditions.

Extraction phase BTCs for the simulation of a testconducted under conditions of linear nonequilibriumsorption (Fig. 5) also di�ered from those obtained underconditions of linear equilibrium sorption (Fig. 2). In Fig.5, increasing nonequilibrium conditions were simulatedusing decreasing values of �k. Note that �k represents aDamkohler number, expressing the rate of mass transferrelative to the rate of advection. Decreasing �k generallyenhanced di�erences in BTCs between solutes subject tononequilibrium sorption and the solute subject toequilibrium sorption �R � 10�. For the smallest value of

�k �10ÿ6�, the BTC of the respective sorbing solute ap-proached that of the tracer, indicating that sorption wasalmost completely blocked, i.e., oS=ot approached zeroin Eq. (14). Similar to the case of nonlinear equilibriumsorption (Fig. 4), sorbing solute BTCs in Fig. 5 did notall coincide near Vext=Vinj � 1, but the location wheresorbing solute BTCs crossed the tracer BTC shiftedgradually toward signi®cantly larger values of Vext=Vinj.Again, this phenomenon may serve as a qualitative in-dicator of nonideal transport conditions (here non-equilibrium conditions) during a push±pull test.

As a consequence of the shape of BTCs obtained inFig. 5, values of R� generally increased with decreasing �k(Table 2). A reasonable value of R� (1.92) was obtainedonly for the solute with �k � 10ÿ6. Increasing disagree-ment between simulated and ®tted (Eq. (11)) BTCs for

Table 2

Estimated solute retardation factors �R�� and MSR between simulated and ®tted breakthrough curves in simulations for push±pull tests conducted

under nonideal transport conditions of nonlinear equilibrium sorption, linear nonequilibrium sorption, and a test conducted in a physically

heterogeneous aquifera

Nonlinear sorption Nonequilibrium sorption Heterogeneous aquifer

Solute R* MSR Solute R* MSR Solute R* MSR

Tracer 1.00 1:65� 10ÿ4 Tracer 1.00 2:59� 10ÿ5 Tracer 1.00 6:34� 10ÿ3

R � 10 9.40 9:42� 10ÿ4 R � 10 9.90 1:55� 10ÿ4 R � 2 0.79 4:13� 10ÿ3

�C0b � 0.5 19.4 2:44� 10ÿ3 �kc� 10ÿ1 10.7 1:61� 10ÿ4 R � 5 0.91 3:07� 10ÿ3

�C0 � 1:0 31.2 4:41� 10ÿ3 �k � 10ÿ2 20.1 2:29� 10ÿ4 R � 10 1.16 2:58� 10ÿ3

�C0 � 5:0 54.7 1:74� 10ÿ2 �k � 10ÿ3 184 3:46� 10ÿ4 R � 20 2.06 2:88� 10ÿ3

�C0 � 50 2.74 1:13� 10ÿ2 �k � 10ÿ4 985 4:60� 10ÿ2 R � 50 5.77 5:20� 10ÿ3

�C0 � 500 0.86 6:90� 10ÿ4 �k � 10ÿ6 1.92 2:81� 10ÿ4 R � 100 14.5 9:18� 10ÿ3

a Values of R* were obtained using the simpli®ed method of estimation.b Dimensionless injection concentration.c Dimensionless ®rst-order rate coe�cient.

Fig. 5. Simulated breakthrough curves for a tracer, a solute subject to

linear equilibrium sorption �R � 10�, and ®ve solutes subject to linear

nonequilibrium sorption obtained at the injection/extraction well

during the extraction phase of a push±pull test. Variations in the

sorption behavior of solutes subject to nonequilibrium sorption was

introduced by varying the solutesÕ ®rst-order rate coe�cient, displayed

in dimensionless form as �k.

M.H. Schroth et al. / Advances in Water Resources 24 (2001) 105±117 113

Page 10: PII: S0309-1708(00)00023-3

sorbing solutes with decreasing �k was again witnessed byincreasing MSR values relative to that of the tracer(Table 2). Thus, similar to the case of nonlinear equi-librium sorption, our simpli®ed method for estimating Rappears not be generally applicable to cases of linearnonequilibrium sorption. However, nonequilibriumsorption e�ects during push±pull tests may be discernedby examining late-time concentrations, and they areparticularly obvious in a double-logarithmic plot of theextraction phase BTCs [10,21]. Furthermore, perform-ing multiple push±pull tests with varying �k may helpassess the presence of nonequilibrium sorption condi-tions.

Preferential ¯ow of injected solutes was observed inaquifer layers containing the medium with the highesthydraulic conductivity in simulations conducted for thecase when the injection/extraction was performed acrossthe entire saturated thickness of a physically heteroge-neous con®ned aquifer (Fig. 6(a)). Extreme di�erences inr̂max existed between solutes of di�erent R in di�erentlayers. For example, a value of r̂max;tr � 145 cm wasobserved in medium 1, whereas r̂max;sol was � 0 forR � 100 in the medium 3 lens located in the upper-leftcorner of the computational domain.

Consequently, extraction phase BTCs for simulationsconducted in a heterogeneous con®ned aquifer (Fig. 6(b))were substantially di�erent from those conducted in ahomogeneous aquifer (Fig. 2). For the simulation con-ducted in a heterogeneous aquifer, the width of thedispersed zone did not increase monotonically for allsolutes with increasing R, although a general trend oflarger dispersed widths for solutes with larger values ofR was observed (Fig. 6(b)). In particular, the tracer BTCshowed multiple crossovers with BTCs of sorbing sol-utes (with R6 20) both prior to and after Vext=Vinj � 1:0.The estimated value of aL (assuming an average n � 0:40in Eq. (7)) was 2.96 cm, which was signi®cantly di�erentfrom the simulated value of aL � 1:0 cm, re¯ecting theincreased amount of dispersion due to velocity di�er-ences between the layers (e.g., [5]). The distorted shapeof the tracer BTC led to poor agreement between sim-ulated and ®tted (Eq. (11)) BTCs (Table 2). In particu-lar, the MSE for the tracer BTC �6:34� 10ÿ3� was largerthan for all other solutes except for R � 100. This is aresult of the increased heterogeneity encountered by thetracer, which subsequently led to unreasonably smallpredicted retardation factors R� (Table 2).

In this context, it is useful to consider the e�ects ofmacrodispersion during a push±pull test in somewhatgreater detail. In their review paper, Gelhar et al. [7]classi®ed the push±pull test as a tool with low reliabilitywith respect to determining aL in the ®eld. The authorsargued that macrodispersion in a layered heterogeneousaquifer should be at least partially reversible during apush±pull test due to the reversal of the ¯ow directionbetween the injection and extraction phase. This could

lead to a possible underprediction of aL using Eq. (11).To investigate this phenomenon in our simulation, weobtained the tracer injection phase BTC by trackingsolute and water ¯uxes across the vertical axis at�r ÿ rw� � 80 cm (Fig. 6(a)). Using Eq. (9), we obtainedan estimate of aL (9.83 cm) that was much larger thanthe estimate of aL obtained from the tracer extractionphase BTC (2.96 cm), thus corroborating the argumentof Gelhar et al. [7].

The performance of the simpli®ed method of esti-mation is clearly discouraging for the simulation of apush±pull test conducted in a heterogeneous aquifer(Fig. 6). However, we note that this simulation was (forillustration purposes) conducted in an aquifer thatwas perfectly layered in the main ¯ow direction, thuscausing a signi®cant amount of macrodispersion over a

Fig. 6. Simulated push±pull test for a tracer and co-injected sorbing

solutes �26R6 100� conducted in a heterogeneous con®ned aquifer:

(a) frontal positions at the end of the injection phase (stipples, gray and

black shading represent zones of di�erent material properties) and (b)

breakthrough curves obtained at the injection/extraction well during

the extraction phase.

114 M.H. Schroth et al. / Advances in Water Resources 24 (2001) 105±117

Page 11: PII: S0309-1708(00)00023-3

relatively small solute travel distance. Certainly, if testswere conducted in an aquifer with a more random dis-tribution of hydraulic conductivity (e.g., [12]), wherelocal di�erences in r̂max for any one solute were less ex-treme than in Fig. 6(a), then we would expect a bettermethod performance.

It is also important to note that correlated sorptionand hydraulic conductivity ®elds are responsible for the®eld-scale e�ect of macrokinetics (e.g., [2,23]). Withmacrokinetics, higher ``e�ective'' dispersivities are gen-erally observed for sorbing solutes than for conservativetracers. Additionally, there is some indication that dis-persivity is a function of sorption, at least at the porescale (e.g., [18]). If either of these two conditions exist, itmay limit the applicability of our method to estimate R.In our method, di�erences in the shape of tracer andsorbing solute BTCs are attributed to di�erences in theretardation factor rather than to e�ective dispersivitiescaused by macrokinetics or sorption-enhanced disper-sion. No analysis is presented here to determine the ef-fects of macrokinetics or sorption-enhanced dispersionon the accuracy of our method to estimate R.

4.3. Example ®eld application

Pickens et al. [26] conducted a push±pull test toquantify the sorption of 85Sr in a sandy aquifer using 131Ias a tracer. The aquifer was composed of several hori-zontal layers and the test was conducted in an �8 mthick layer (con®ned below by a silty clay bed of �1 mthickness and above by �17 cm of silt and clay) con-sisting of well sorted very ®ne-to ®ne-grained sand. Sandporosities varied between 0.33 and 0.44 with an averagevalue of 0.38, and the bulk density was calculated to be1.7 g/cm3. The injection/extraction well was constructedof 10.4 cm diameter PVC pipe and screened across al-most the entire thickness of the selected aquifer layer. Atotal volume of Vinj � 244 m3 of test solution containing131I and 85Sr was injected over a period of Tinj � 94:32 hat a constant ¯ow rate Qinj � 2:587 m3=h, resulting in anaverage r̂max;tr of �5:0 m. Extraction pumping beganimmediately following the injection phase atQext � 2:282 m3=h for 405.6 h (equivalent to Vext=Vinj �3:6). During the injection phase, water samples werecollected from several depths using three multilevelsamplers installed at radial distances of 0.36, 0.66 and2.06 m from the injection/extraction well; during theextraction phase, water samples were collected from theinjection/extraction well discharge line.

Distribution coe�cients �Kd� for 85Sr, obtained frominjection phase BTCs at three separate depth intervals,ranged from 2.6 to 4.5 ml/g [26]. These values were ingood general agreement with values determined fromsediment core analyses �4:3 ml=g6Kd6 7:9 ml=g� andbatch experiments �2:8 ml=g6Kd6 7:8 ml=g�. Whileextraction phase BTCs for 131I and 85Sr were presented

(Fig. 7), Pickens et al. [26] did not report a quantitativeanalysis and considered the increase in width of thedispersed front for the 85Sr BTC compared to the 131IBTC to be the evidence for the presence of nonequilib-rium sorption e�ects.

Using the simpli®ed method of estimation for radial¯ow, we ®t Eq. (11) to the 131I BTC using r̂max � 5:0 mand estimated aL � 6:4 cm �MSE � 1:67� 10ÿ3�. Then,keeping aL ®xed, we ®t Eq. (11) to the 85Sr BTC anddetermined r̂max;Sr � 1:48 m �MSE � 1:98� 10ÿ3�. Thisis in good agreement with experimental results wherecomplete breakthrough of 85Sr was observed atr � 0:66 m, but no breakthrough of 85Sr was detected atr � 2:06 m [26]. Using Eq. (12), we estimated R��85

Sr� �11:4, and rearranging Eq. (4) we determined Kd �2:33 ml=g. Thus, reasonable agreement was found be-tween values of Kd determined from injection phaseBTCs by Pickens et al. [26] and the Kd value estimatedfrom extraction phase BTCs using the simpli®ed methodof estimation presented in this study. In particular, thevalue of Kd we determined (2.33 ml/g) matched closelythe smallest value of Kd (2.6 ml/g) reported by Pickenset al. [26], which was measured in a sublayer of theaquifer in which ¯ow and transport were fastest. Thisagrees with the observation of preferential ¯ow andtransport that occurred during the simulation of a push±pull test conducted in a heterogeneous (layered) aquifer(Fig. 6(a)).

Moreover, a reasonable estimate of Kd for 85Sr wasobtained by our method despite the fact that Pickenset al. [26] observed e�ects of macrokinetics during their®eld test (e�ective dispersivity values for 85Sr were

Fig. 7. Breakthrough curves for a tracer (131I) and a sorbing solute

(85Sr) obtained by Pickens et al. [26] at the injection/extraction well

during the extraction phase of a ®eld push±pull test. Solid lines show

®tted curves (Eq. (11)) employed to estimate R using the simpli®ed

method of estimation.

M.H. Schroth et al. / Advances in Water Resources 24 (2001) 105±117 115

Page 12: PII: S0309-1708(00)00023-3

typically a factor 2±5 larger than those obtained for 131Iduring the ®eld experiment). Apparently, nonideal ¯owconditions that cause macrodispersion and macro-kinetics were mild enough to allow the successful ap-plication of our simpli®ed method of estimation.Witness to this conclusion are both the fact that ob-tained values of MSR between measured and ®ttedBTCs were small and in the same range for tracer andsorbing solute, and that the point of coincidence of the131I and 85Sr BTCs was near Vext=Vinj � 1 (compareFigs. 2 and 7). Nonetheless, di�erences between the ex-perimental data and the ®tted BTCs (Fig. 7), particu-larly at late times, may be indicative of slightnonequilibrium sorption conditions during the experi-ment [10,21].

5. Summary and conclusions

The ability of the push±pull test to provide quanti-tative information on in situ solute retardation wasevaluated. A simpli®ed method for estimating retarda-tion factors from extraction phase BTCs was developedfor the case of radial ¯ow based on a previously pub-lished analysis of variable dispersion during nonuniform¯ow [6]. Sensitivity analyses were conducted to evaluatethe methodÕs performance for a variety of simulated testconditions, including tests performed under both idealand nonideal transport conditions. Finally, the methodwas applied to ®eld data presented by Pickens et al. [26]to estimate the distribution coe�cient of 85Sr in a sandyaquifer.

Sensitivity analyses based on numerical simulationsfor tests conducted under ideal transport conditions in aphysically and chemically homogeneous con®ned aqui-fer revealed errors 614% between simulated and pre-dicted values of R for a wide range of aquifer and testconditions. The method was most sensitive to changes inaquifer dispersivity, whereas an uncertainty in aquifere�ective porosity or thickness did not impede its suc-cessful use. In general, di�erences between simulatedand predicted values of R increased with increasing R.This was mainly attributed to the increased violation ofan assumption (regarding the total travel distance of thesolute front during a push±pull test) made in the de-velopment of the underlying approximate solutions tothe transport equation.

Conversely, poor agreement between simulated andpredicted values of R was often found for simulations ofpush±pull tests conducted under nonideal transportconditions such as nonlinear equilibrium and linearnonequilibrium sorption, and in a test performed in aphysically heterogeneous aquifer. Thus, the reader iscautioned against using this method to estimate retar-dation factors for either (1) very strongly or very slowlysorbing solutes, or (2) cases of severe aquifer hetero-

geneity (especially layering). Furthermore, no e�ort wasmade here to evaluate the potentially deleterious e�ectsof macrokinetics (resulting from negatively correlatedKd and K) on the method performance.

Nonetheless, when applied to a ®eld data set pre-sented by Pickens et al. [26], the method showedreasonable agreement with those authorsÕ results inestimating the distribution coe�cient of 85Sr duringtransport in a sandy aquifer. This is particularly en-couraging considering that nonideal transport condi-tions existed during the ®eld experiment, i.e., the testwas conducted in a heterogeneous aquifer that exhibitede�ects of macrokinetics. This indicates that push±pulltests and the simpli®ed method of estimation may beused successfully for the in situ evaluation of solutesorption under certain test and aquifer conditions.However, further analyses, in particular additional ap-plication of this method to ®eld data sets, will be re-quired to better understand both the methodÕscapabilities and limitations.

Acknowledgements

This research was funded by the US Department ofEnergy, Environmental Management Science Program,Grant No. DE-FG07-96ER14721. Special thanks toMart Oostrom and Mark White, Paci®c NorthwestNational Laboratory, for providing the STOMP simu-lator. We would also like to thank the three anonymousreviewers and the journal editor, S.M. Hassanizadeh, fortheir helpful review comments.

References

[1] Bear J. Hydraulics of groundwater. New York: McGraw-Hill;

1979.

[2] Chrysikopoulos CV, Kitanidis PK, Roberts PV. Generalized

Taylor±Aris moment analysis of the transport of sorbing solutes

through porous media with spatially peridiodic retardation factor.

Transport in Porous Media 1992;7:163±85.

[3] Drever JI, McKee CR. The push±pull test. A method for

evaluating formation adsorption parameters for predicting the

environmental e�ects on in situ coal gasi®cation and uranium

recovery. In Situ 1980;4:181±206.

[4] Fetter CW. Contaminant hydrogeology. Upper Saddle River, NJ:

Prentice-Hall; 1993.

[5] Gelhar LW. Stochastic subsurface hydrology. Englewood Cli�s,

NJ: Prentice-Hall; 1993.

[6] Gelhar LW, Collins MA. General analysis of longitudinal

dispersion in nonuniform ¯ow. Water Resour Res 1971;7:1511±

21.

[7] Gelhar LW, Welty C, Rehfeldt KR. A critical review of data on

®eld-scale dispersion in aquifers. Water Resour Res 1992;28:1955±

74.

[8] Gupta AD, Lake LW, Pope GA, Sepehrnoori K, King MJ. High-

resolution monotonic schemes for reservoir ¯uid ¯ow simulation.

In Situ 1991;15:289±317.

[9] Haggerty R, Fleming SW, McKenna SA, STAMMT-R, Solute

transport and multirate mass transfer in radial coordinates: a

116 M.H. Schroth et al. / Advances in Water Resources 24 (2001) 105±117

Page 13: PII: S0309-1708(00)00023-3

FORTRAN for modeling and analyzing radial single-well and

two-well tracer tests in formations exhibiting multiple rates of

di�usive mass transfer, Version 1.01. SAND99-0164, 2000, Sandia

National Laboratories, Albuquerque, New Mexico.

[10] Haggerty R, Fleming SW, Meigs LC, McKenna SA. Tracer

tests in fractured dolomite: 2. Analysis of mass transfer in single-

well injection±withdrawal tests. Water Resour Res 2000, in

review.

[11] Haggerty R, Gorelick SM. Multiple-rate mass transfer for

modeling di�usion and surface reactions in media with pore-scale

heterogeneity. Water Resour Res 1995;31:2383±400.

[12] Haggerty R, Schroth MH, Istok JD. Simpli®ed method of push±

pull test data analysis for determining in situ reaction rate

coe�cients. Ground Water 1998;36:314±24.

[13] Hall SH, Luttrell SP, Cronin WE. A method for estimating

e�ective porosity and ground-water velocity. Ground Water

1991;29:171±4.

[14] Harmon TC, Semprini L, Roberts PV. Simulating solute transport

using laboratory-based sorption parameters. J Environ Eng

1992;118:666±89.

[15] Hoopes JA, Harleman DRF. Dispersion in radial ¯ow from a

recharge well. J Geophys Res 1967;72:3595±607.

[16] Istok JD, Field JA, Schroth MH, Sawyer TE, Humphrey MD.

Laboratory and ®eld investigation of surfactant sorption using

single-well, push±pull tests. Ground Water 1999;37:589±98.

[17] Istok JD, Humphrey MD, Schroth MH, Hyman MR, OÕReilly

KT. Single-well, push±pull test for in situ determination of

microbial activities. Ground Water 1997;35:619±31.

[18] Lafolie F, Hayot C, Schweich D. Experiments on solute transport

in aggregated porous media: are di�usions within aggregates and

hydrodynamic dispersion independent. Transport in Porous

Media 1997;29:281±307.

[19] Leap DI, Kaplan PG. A single-well tracing method for esti-

mating regional advective velocity in a con®ned aquifer:

theory and preliminary laboratory veri®cation. Water Res 1988;

23:993±8.

[20] Mackay DM, Freyberg DL, Roberts PV, Cherry JA. A natural

gradient experiment on solute transport in a sand aquifer 1.

Approach and overview of plume movement. Water Resour Res

1986;22:2017±29.

[21] Meigs LC, Beauheim RL, McCord JT, Tsang YW, Haggerty R,

Design, modelling, and current interpretations of the H-19 and H-

11 tracer tests at the WIPP site, NEA/EC GEOTRAP Workshop.

In: Field tracer experiments: role in the prediction of radionuclide

migration. Nuclear Energy Agency, Paris, Cologne, Germany,

1997. p. 157±69.

[22] Mercado A, Underground water storage study: Recharge

and mixing tests ant Yavne 20 well ®eld. Technical

Report No. 12, TAHAL ± Water Planning for Israel Ltd., Tel

Aviv, 1966.

[23] Miralles-Wilhelm F, Gelhar LW. Stochastic analysis of sorption

macrokinetics in heterogeneous aquifers. Water Resour Res

1996;32:1541±9.

[24] Nichols WE, Aimo NJ, Oostrom M, White MD, STOMP:

subsurface transport over multiple phases. Application guide.

PNNL-11216, 1997, Paci®c Northwest National Laboratory,

Richland, WA.

[25] Pickens JF, Grisak GE. Scale-dependent dispersion in a strati®ed

granular aquifer. Water Resour Res 1981;17:1191±211.

[26] Pickens JF, Jackson RE, Inch KJ, Merritt WF. Measurement of

distribution coe�cients using a radial injection dual-tracer test.

Water Resour Res 1981;17:529±44.

[27] Ptak T, Schmid G. Dual-tracer transport experiments in a

physically and chemically heterogeneous porous aquifer: e�ective

transport parameters and spatial variability. J Hydrol

1996;183:117±38.

[28] Reinhard M, Shang S, Kitanidis PK, Orwin E, Hopkins GD,

Lebron CA. In situ BTEX biotransformation under enhanced

nitrate- and sulfate-reducing conditions. Environ Sci Technol

1997;31:28±36.

[29] Roberts PV, Goltz MN, Mackay DM. A natural gradient

experiment on solute transport in a sand aquifer 3. Retardation

estimates and mass balances for organic solutes. Water Resour

Res 1986;22:2047±58.

[30] Sauty JP. Identi®cation des param�etres du transport hydrodis-

persif dans les aquif�eres par interpr�etation de tracßages en

�ecoulement cylindrique convergent ou divergent. J Hydrol

1978;39:69±103.

[31] Schroth MH, Istok JD, Conner GT, Hyman MR, Haggerty R,

OÕReilly KT. Spatial variability in in situ aerobic respiration and

denitri®cation rates in a petroleum-contaminated aquifer. Ground

Water 1998;36:924±37.

[32] Snodgrass MF, Kitanidis PK. A method to infer in situ reaction

rates from push±pull experiments. Ground Water 1998;36:645±50.

[33] Trudell MR, Gillham RW, Cherry JA. An in situ study of the

occurrence and rate of denitri®cation in a shallow uncon®ned

sand aquifer. J Hydrol 1986;83:251±68.

[34] Valocchi AJ. E�ect of radial ¯ow on deviations from local

equilibrium during sorbing solute transport through homoge-

neous soils. Water Resour Res 1986;22:1693±701.

[35] van Genuchten MT, Cleary RW, Movement of solutes in soil:

computer-simulated and laboratory results. In: G.H. Bolt editor.

Soil Chemistry, Physico-Chemical Models. New York: Elsevier;

1982, pp. 349±86.

[36] White MD, Oostrom M, STOMP: subsurface transport over

multiple phases. Theory Guide. PNNL-11217, 1996, Paci®c

Northwest National Laboratory, Richland, WA.

[37] White MD, Oostrom M, STOMP: subsurface transport over

multiple phases. User's Guide. PNNL-11218, 1996, Paci®c

Northwest National Laboratory, Richland, WA.

M.H. Schroth et al. / Advances in Water Resources 24 (2001) 105±117 117