Controls on induced polarization in sandy unconsolidated sediments and application to aquifer characterization

Similar documents
Hydrological geophysical relationships

THE INDUCED POLARIZATION METHOD

IP interpretation in environmental investigations

Impedance Spectroscopy on Carbonates

Resistivity & IP methods

Pore radius distribution and fractal dimension derived from spectral induced polarization

Site Characterization & Hydrogeophysics

Scholars Research Library. Induced polarization interpretation for subsurface characterisation: A case study of Obadore, Lagos State

11/22/2010. Groundwater in Unconsolidated Deposits. Alluvial (fluvial) deposits. - consist of gravel, sand, silt and clay

ELECTRICAL RESISTIVITY TOMOGRAPHY

1. Resistivity of rocks

Simultaneous Measurement of Capillary Pressure and Dielectric Constant in Porous Media

STATISTICAL COMPARISON OF ELECTRICAL EFFICIENCY THEORY TO ARCHIE S EQUATIONS AND EFFECTIVE MEDIUM THEORY by Charles R.

water L v i Chapter 4 Saturation

Geophysical Characterization and Monitoring of Groundwater/Surface-Water Interaction in the Hyporheic Corridor at the Hanford 300 Area

Rock/water interaction in dielectric properties: Experiments with hydrophobic sandstones

RELATIONSHIP BETWEEN CAPILLARY PRESSURE AND RESISTIVITY INDEX

SASKATCHEWAN STRATIGRAPHY GLACIAL EXAMPLE BOULDERS IN GLACIAL DEPOSITS

Geophysics Course Introduction to DC Resistivity

Essentials of Geology, 11e

Permeability and fluid transport

10. GEOTECHNICAL EXPLORATION PROGRAM

Assessing the Tier 2 Trigger for Fractured Sedimentary Bedrock Sites

Improved Exploration, Appraisal and Production Monitoring with Multi-Transient EM Solutions

Geophysics of Exploration for Water. Geoelectric methods III. edited by Peter Vass

NEW SATURATION FUNCTION FOR TIGHT CARBONATES USING ROCK ELECTRICAL PROPERTIES AT RESERVOIR CONDITIONS

Characterization of Sand Formation from a Crude Oilfield

PETROPHYSICAL EVALUATION CORE COPYRIGHT. Saturation Models in Shaly Sands. By the end of this lesson, you will be able to:

Geology 228/378 Applied & Environmental Geophysics Lecture 8. Induced Polarization (IP) and Nuclear Magnetic Resonance (NMR)

Groundwater Hydrology

Variation of Moisture Content as a Parameter of Study by Induced Polarization Technique in Soil Sample of Coastal Andhra Pradesh

Module for: Resistivity Theory (adapted/modified from lectures in PETE 321 (Jensen/Ayers))

A new approach to fitting induced-polarization spectra

Tu 23P1 06 Mapping Possible Flowpaths of Contaminants through Surface and Cross-borehole Spectral Timedomain Induced Polarization

Estimating the hydrocarbon volume from elastic and resistivity data: A concept

A combination of effective medium theory and mixture theory to model electrical conductivity response of shaly sand reservoirs.

CHARACTERIZATION OF SOIL PROFILE OF DHAKA CITY USING ELECTRICAL RESISTIVITY TOMOGRAPHY (ERT)

Prof. Stephen A. Nelson EENS 111. Groundwater

Electrical Properties of Moist Limestone Samples 7. In The Frequency Range 1Hz-10 Hz From Abu Rawash Area

Soils, Hydrogeology, and Aquifer Properties. Philip B. Bedient 2006 Rice University

Reservoir Rock Properties COPYRIGHT. Sources and Seals Porosity and Permeability. This section will cover the following learning objectives:

Electrical prospecting involves detection of surface effects produced by electrical current flow in the ground.

RADIONUCLIDE DIFFUSION IN GEOLOGICAL MEDIA

An empirical method for estimation of anisotropic parameters in clastic rocks

Correlation Between Resistivity Index, Capillary Pressure and Relative Permeability

C) D) 3. Which graph best represents the relationship between soil particle size and the rate at which water infiltrates permeable soil?

USE OF GEOPHYSICAL SURVEYS FOR FILL CHARACTERIZATION AND QUANTITY ESTIMATION AT BROWNFIELD SITES A CASE HISTORY. Abstract

ROCK PHYSICS DIAGNOSTICS OF NORTH SEA SANDS: LINK BETWEEN MICROSTRUCTURE AND SEISMIC PROPERTIES ABSTRACT

A Risk-based Groundwater Modelling Study for Predicting Thermal Plume Migration from SAGD Well-pads

Saturation Modelling: Using The Waxman- Smits Model/Equation In Saturation Determination In Dispersed Shaly Sands

1. Water in Soils: Infiltration and Redistribution

Geophysics for Environmental and Geotechnical Applications

Finding Large Capacity Groundwater Supplies for Irrigation

APPLICATION OF ELECTRICAL RESISTIVITY TOMOGRAPHY FOR SAND UNDERWATER EXTRACTION

MOUNT POLLEY MINING CORPORATION TECHNICAL REPORT ON MULTI-ELECTRODE RESISTIVITY AND SEISMIC REFRACTION SURVEYS MOUNT POLLEY TAILINGS DAM PROJECT

Downloaded 02/05/15 to Redistribution subject to SEG license or copyright; see Terms of Use at

SPE These in turn can be used to estimate mechanical properties.

5. Which surface soil type has the slowest permeability rate and is most likely to produce flooding? A) pebbles B) sand C) silt D) clay A) B) C) D)

RESISTIVITY IMAGING AND BOREHOLE INVESTIGATION OF THE BANTING AREA AQUIFER, SELANGOR, MALAYSIA. A.N. Ibrahim Z.Z.T. Harith M.N.M.

Key words: Spectral induced polarization (SIP); IP imaging; Near-surface geology. INTRODUCTION

SAGEEP. DoWN IN THE DUMPS. Workshop Notes

Monitoring of In-Situ Chemical Oxidation (ISCO) with Time Series Geophysical Surveys

HYDROGEOPHYSICS. Susan Hubbard. Lawrence Berkeley National Laboratory. 1 Cyclotron Road, MS Berkeley, CA USA.

INVESTIGATION ON THE EFFECT OF STRESS ON CEMENTATION FACTOR OF IRANIAN CARBONATE OIL RESERVOIR ROCKS

Influence of pore fluid chemistry on the complex conductivity and induced polarization responses of Berea sandstone

LECTURE 10. Module 3 : Field Tests in Rock 3.6 GEOPHYSICAL INVESTIGATION

Chapter 1 Introduction

Uncertainty analysis for the integration of seismic and CSEM data Myoung Jae Kwon & Roel Snieder, Center for Wave Phenomena, Colorado School of Mines

POTASH DRAGON CHILE GEOPHYSICAL SURVEY TRANSIENT ELECTROMAGNETIC (TEM) METHOD. LLAMARA and SOLIDA PROJECTS SALAR DE LLAMARA, IQUIQUE, REGION I, CHILE

Estimation of Chargeability and Permeability using Spectral Induced Polarization Data

Part II: Self Potential Method and Induced Polarization (IP)

Case Study: University of Connecticut (UConn) Landfill

Applied Geophysics for Environmental Site Characterization and Remediation

Resistivity-IP mapping for landfill applications

Shear Wave Velocity Estimation Utilizing Wireline Logs for a Carbonate Reservoir, South-West Iran

Hysteresis in the electrical resistivity of partially saturated sandstones

Geoelectricity. ieso 2010

GY 402: Sedimentary Petrology

Chapter 4 Influences of Compositional, Structural and Environmental Factors on. Soil EM Properties

2D Resistivity Imaging Investigation of Solid Waste Landfill Sites in Ikhueniro Municipality, Ikpoba Okha Local Government Area,Edo State, Nigeria.

Fractal dimension of pore space in carbonate samples from Tushka Area (Egypt)

MEASURING THE PERMEABILITY OF OPEN-FRAMEWORK GRAVEL. A thesis submitted in partial fulfillment of the requirements for the degree of Master of Science

CONTENTS 1. INTRODUCTION. 2. THE D.C. RESISTIVITY METHOD 2.1 Equipment 2.2 Survey Procedure 2.3 Data Reduction

Groundwater Resource Evaluation in Support of Dewatering a South Carolina Limestone Quarry

The Role of Magnetotellurics in Geothermal Exploration

Azimuthal Resistivity to Characterize Fractures in a Glacial Till. Mark Boris, University of Saskatchewan Jim Merriam, University of Saskatchewan

Student, Department of Earth Sciences, IIT Roorkee, Uttarakhand, India, b

VIBRATION-INDUCED CONDUCTIVITY FLUCTUATION (VICOF) TESTING OF SOILS *

PREDICTION OF ACID MINE DRAINAGE POTENTIAL FROM COAL MINES

GeoCanada 2010 Working with the Earth

Variety of Cementation Factor between Dolomite and Quartzite Reservoir

Determination of the Laminar, Structural and Disperse Shale Volumes Using a Joint Inversion of Conventional Logs*

This paper was prepared for presentation at the International Symposium of the Society of Core Analysts held in Toronto, Canada, August 2005

Estimating Permeability from Acoustic Velocity and Formation Resistivity Factor

Study on Estimation of Hydraulic Conductivity of Porous Media Using Drag Force Model Jashandeep Kaur, M. A. Alam

Unsaturated Flow (brief lecture)

Capillary pressure as a unique function of electric permittivity and water saturation

Geophysical Journal International

Hamed Aber 1 : Islamic Azad University, Science and Research branch, Tehran, Iran. Mir Sattar Meshin chi asl 2 :

GEOLOGICAL LOG INTERPRETATION TUTORIAL

Transcription:

GEOPHYSICS, VOL. 68, NO. 5 (SEPTEMBER-OCTOBER 2003); P. 1547 1558, 8 FIGS., 1 TABLE. 10.1190/1.1620628 Controls on induced polarization in sandy unconsolidated sediments and application to aquifer characterization L. D. Slater and D. R. Glaser ABSTRACT Resistivity and induced polarization (IP) measurements (0.1 1000 Hz) were made on clay-free unconsolidated sediments from a sandy, alluvial aquifer in the Kansas River floodplain. The sensitivity of imaginary conductivity σ, a fundamental IP measurement, to lithological parameters, fluid conductivity, and degree of saturation was assessed. The previously reported power law dependence of IP on ace area and grain size is clearly observed despite the narrow lithologic range encountered in this unconsolidated sedimentary sequence. The grain-size σ relationship is effectively frequency independent between 0.1 and 100 Hz but depends on the representative grain diameter used. For the sediments examined here, d 90, the grain diameter of the coarsest sediments in a sample, is well correlated with σ. The distribution of the internal ace in the well-sorted, sandy sediments investigated here is such that most of the sample weight is likely required to account for the majority of the internal ace. We find the predictive capability of the Börner model for hydraulic conductivity (K ) estimation from IP measurements is limited when applied to this narrow lithologic range. The relatively weak dependence of σ on fluid conductivity (σ w ) observed for these sediments when saturated with an NaCl solution (0.06 10 S/m) is consistent with competing effects of ace charge density and ace ionic mobility on σ as previously inferred for sandstone. Importantly, IP parameters are a function of saturation and exhibit hysteretic behavior over a drainage and imbibition cycle. However, σ is less dependent than the real conductivity σ on saturation. In the case of evaporative drying, the σ saturation exponent is approximately half of the σ exponent. Crosshole IP imaging illustrates the potential for lithologic discrimination of unconsolidated sediments. A fining-upward sequence correlates with an upward increase in normalized chargeability M n, a field IP parameter proportional to σ. The hydraulic conductivity distribution obtained from the Börner model discriminates a hydraulically conductive sand gravel from overlying medium sand. INTRODUCTION Improved interpretation of induced polarization (IP) surveys requires better understanding of the physicochemical controls on IP measurements and the significance of measured IP parameters. Induced polarization is attributed to the presence of interfaces at which local charge concentration gradients result in a delayed voltage response in earth materials upon current stimulation (for review, see Ward, 1990; Schön, 1996). Two distinct interfaces are commonly recognized and attributed to IP signals. Electrode (or metallic) polarization occurs at interfaces between zones of electronic and ionic conduction, such as the interface between a metallic mineral grain and electrolyte (for review, see Ward and Fraser, 1967). The large IP anomalies associated with electrode polarization have been utilized in mineral prospecting for disseminated sulfides. Membrane (or boundary layer) polarization occurs at interfaces between zones of unequal ionic transport properties associated with clays distributed through a rock matrix (Marshall and Madden, 1959). Induced polarization is hence a method for mapping clay content (Vacquier et al., 1957; Ogilvy and Kuzmina, 1972). The presence of organic contaminants, particularly in clays, may also generate significant IP anomalies (Olhoeft, 1985; Vanhala et al., 1992; Börner et al., 1993). Manuscript received by the Editor July 26, 2002; revised manuscript received April 12, 2003. Formerly University of Missouri Kansas City, Kansas City, Missouri 64110; presently Rutgers University, Department of Earth and Environmental Sciences, 195 University Avenue, Room 407, Newark, New Jersey 07102. E-mail: Islater@andromeda.rutgers.edu. University of Missouri Kansas City, 420 Robert H. Flarshiem Hall, 5100 Rockhill Road, Kansas City, Missouri 64110. E-mail: dannyglaser@ hotmail.com. c 2003 Society of Exploration Geophysicists. All rights reserved. 1547

1548 Slater and Glaser Small yet measurable IP signatures are also observed in the absence of metals, clay particles, or organic contaminants (Vanhala and Soininen, 1995; Schön, 1996; Vanhala, 1997). In this case, the interface causing the IP response is that between a silica mineral grain and electrolyte. The ace of silicate minerals carries a net negative charge, which is balanced by counter ions attracted to the ace. These counter ions form two layers: a fixed (Stern) layer chemically absorbed to the grain ace and a diffuse layer extending into the pore space. Upon application of an electric field, charges migrate tangentially (conduction) and are displaced normally/tangentially (polarization). This interface response depends on the frequency of the electric field, ionic mobility, temperature, and grain mobility (Revil and Glover, 1998). The significance of this polarization in relation to physical properties of clay-free, near-ace sediments is studied in this paper. Understanding induced polarization requires a model for the low-frequency electrical properties of earth materials. Theoretically based three-phase (dry mineral grain, pore fluid, and interface phase) effective medium theories are available (e.g., Bussian, 1983; Samstag and Morgan, 1991). A simpler approach that is easily adaptable to IP interpretation is based on modeling parallel ion transport paths through pore-filled fluids and at the grain fluid interface (Marshall and Madden, 1959; Waxman and Smits, 1968; Vinegar and Waxman, 1984; Börner et al., 1996; Schön, 1996; Lesmes and Frye, 2001). Lesmes and Frye (2001) review how standard measures of IP are related to such a parallel pathway model. In this paper we report low-frequency electrical measurements on clay-free unconsolidated sediments obtained from a sequence of alluvial deposits. We show that previously identified relationships between IP measurements and lithological parameters are discernible even when lithologic variability is restricted to a narrow range as encountered in a sequence of sandy deposits. However, results from this study suggest that the correlation between a representative grain diameter and IP parameters will depend on the particle-size characteristics of the material. We also show that IP parameters are a function of saturation and exhibit hysteretic behavior over a drainage and imbibition cycle. The results of a crosshole survey are included to illustrate how lithologic variability and hydraulic conductivity might be inferred from high-resolution IP surveys. LOW-FREQUENCY ELECTRICAL PROPERTIES Electrical model Measured complex conductivity σ is expressed in terms of conductivity phase shift φ and magnitude σ, or by the real σ and imaginary σ components, σ = σ e iφ = σ + iσ. (1) The real component is a representation of current flow in phase with the applied electric field, while the imaginary component is the displacement of current 90 out of phase with the applied field. There are two pathways for electric charge transfer within a medium: electrolytic (volumetric) conduction occurring through fluids in interconnected pore spaces and ace (interface) conduction occurring within an ionic double layer at the grain fluid interface. A simple model for the low-frequency electrical properties of earth materials is obtained by adding electrolytic and ace conduction pathways in parallel (e.g., Marshall and Madden, 1959; Vinegar and Waxman, 1984; Börner et al., 1996; Lesmes and Frye, 2001). At low frequencies (<1000 Hz), the complex conductivity is then σ = σ el + σ (ω) = [σ el + σ (ω)] + iσ (ω), (2) where σ el is the conductivity of the electrolytic path through the pore space and where σ and σ are frequency (ω) dependent real and imaginary components of the complex ace conductivity σ, respectively. Thus, σ represents interface current flow in phase with the applied electric field and σ represents diffusive interface polarization. This model assumes that measured imaginary conductivity is only a function of the ace (interface) conductivity, whereas the real conductivity is a function of both electrolytic and ace (interface) conductivity mechanisms. In clay-free sediments σ el σ such that σ σ el. The electrolytic conductivity σ el is considered independent of frequency when measured below 1000 Hz and given by σ el = ( 1 F ) σ w S n = σ w φ m eff Sn, (3) where σ w is the fluid conductivity, F is the formation factor related to the tortuosity of the electrolytic conduction path in a saturated medium, φ eff is the effective porosity for electrolytic current flow, m is a function of effective grain diameter (Sen et al., 1981), S is the degree of saturation, and n is a material specific saturation exponent quantifying the dependence of σ el on saturation (Archie, 1942; McNeill, 1990). Surface conduction The specific ace conductance s in a quartz-electrolyte system is a function of the ace area S 0, ace ionic charge density 0, and ace ionic mobility μ s : = eμ s 0 S 0, (4) s where e is the electronic charge. (Revil and Glover, 1998; Revil et al., 1998; Lesmes and Frye, 2001). Ionic charge density and ionic mobility generally depend on fluid chemistry. At low frequencies σ approaches the dc ace conductivity σ (dc), which is related to s via a geometric factor with units of length and is a characteristic of the geometry of the medium (Johnson et al., 1986; Pride, 1994; Revil and Glover, 1998). Johnson et al. (1986) define a geometric parameter, which they envisage as a weighted volume ace area ratio and a measure of the dynamically interconnected pore size such that lim σ = σ (dc) = 1 ( ) 2 s. (5) ω 0 F Revil and Cathles (1999) provide an expression for in natural earth materials as a function of a representative grain diameter R r : R r = m(f 1). (6) Consequently, in saturated earth materials σ is primarily determined by s, which is the product of o, μ s, and S 0 and is

Controls on induced polarization 1549 inversely related to a representative or average grain diameter R r through. Induced polarization parameters measured using field instruments include chargeability M, phase angle φ, and percent frequency effect (PFE). These equivalent parameters measure the ratio of interface polarization to the combined strength of electrolytic and ace conduction. The chargeability and PFE field parameters can be scaled by the conductivity magnitude to obtain normalized field parameters: normalized chargeability M n and metal factor M F, respectively (Keller, 1959; Marshall and Madden, 1959). These normalized field parameters are equivalent to σ and hence are sensitive to ace (interface) properties as in equations (4) (6). Lesmes and Frye (2001) review field IP and normalized IP parameters and their relationship to the model for low-frequency electrical properties described here. Previous experimental work Relationships between measures of the internal ace and imaginary conductivity terms have been shown in sandstone (e.g., Knight and Nur, 1987; Börner and Schön, 1991) and unconsolidated sediments (Börner et al., 1996; Slater and Lesmes, 2002). Börner and Schön (1991) identify a strong relationship between internal ace area normalized at total volume (S tot ) and ace conductivity. When porosity varies widely, the internal ace normalized to the pore volume (S por ) may correlate more closely with the ace conductivity (Börner et al., 1996). Schön (1996) reports an inverse relationship between median grain diameter (d 50 ) and σ in unconsolidated sediments. Slater and Lesmes (2002) illustrate an inverse relationship between grain size for which 10% of the sample is finer (d 10 ) and σ when applied to artificial sand clay mixtures and natural unconsolidated sediments. The dependence of σ on structural parameters encouraged the prediction of hydraulic conductivity K using IP measurements (Börner et al., 1996; De Lima and Niwas, 2000; Slater and Lesmes, 2002). These studies suggest that orderof-magnitude K estimates are obtainable for unconsolidated sediments and sandstone when K varies over three or more orders of magnitude. The Börner model, based on a modified (electrical) Kozeny-Carmen equation, is an attractive electrical approach to K estimation. It utilizes a complex conductivity measurement in conjunction with a fluid conductivity measurement to predict K. Estimates of formation factor F and S por are required in the electrical version of the Kozeny-Carmen model (Börner et al., 1996). When σ el σ, σ σ el and F of a saturated medium can be estimated if σ w is known [Equation (3)]. The measured imaginary conductivity σ provides an electrical estimate of S por.thebörner model is given as K = a FS c por[el] = a F ( ) 10 5 σ c, (7) [1Hz] where S por[el] is the electrically estimated S por, a = 0.00001, and c ranges between 2.8 and 4.6, depending on material type. A general version of the model permitting different parameters for the relationship between specific ace and imaginary conductivity is K = A F ( ) σ c, (8) [1Hz] where A is a fitting parameter. Since ace charge density and ace ionic mobility are both modified by fluid chemistry, σ is observed to vary with σ w (Börner et al., 1993, 1996; Revil and Glover, 1998; Lesmes and Frye, 2001). Relative to σ el, ace conductivity exhibits weak σ w dependence (Vinegar and Waxman, 1984; Revil et al., 1998; Lesmes and Frye, 2001; Slater and Lesmes, 2002). Lesmes and Frye (2001) attribute the σ behavior of Berea Sandstone saturated with NaCl to the competing effects of ace charge density and ace ionic mobility with increasing pore solution concentration. Very few studies of the dependence of σ on degree of saturation are reported. Vinegar and Waxman (1984) investigate changes in σ with water saturation for an oil water system. They find that σ exhibits power law dependence on the degree of water saturation. However, the saturation exponent for σ is approximately half the exponent for σ. STUDY SITE The City of Olathe well field is located on the Kansas River floodplain near Desoto, Kansas. Installation of a horizontal collector well is planned to enhance yield and provide adequate water supply for the city. Limestone bedrock is overlain with a 14-m-thick sand aquifer (fining-upward sequence) capped by a 3 4-m-thick silt layer. The range of saturated aquifer thickness throughout the river valley is 0 30 m, averaging 13 m at the field site. Depth to the water table is approximately 10 m. Four wells were drilled in 2001 for sample retrieval and piezometer installation as part of an aquifer pump test to constrain horizontal well design. A Rotosonic vibratory drilling method was used to advance the core barrel and retrieve an unconsolidated sediment sample. This method was developed to obtain undisturbed core samples from unconsolidated aquifers. However, the coarse, granular matrix of this aquifer was of insufficient integrity to support a sample on removal. Consequently, only disturbed samples of aquifer material were obtainable for laboratory tests. Piezometers constructed from PVC tubing were screened over the full vertical distance of the aquifer and installed in the wells. Electrodes for crosshole electrical imaging were installed on piezometers in two wells spaced 10 m apart. Ideally, we would have used nonpolarizing electrodes, such as Pb-PbCl junctions or platinum electrodes (e.g., Weller et al., 1996). However, excessive cost and awkward implementation in boreholes ruled against this approach. Dahlin et al. (2002) show that, by using signal processing to correct for the electrode polarization at potential electrodes, stainless steel electrodes can give IP measurements of similar quality to that obtained with Pb-PbCl electrodes. This processing procedure is standard to the IRIS Syscal R1 instrument used in this study. Stainless steel mesh electrodes were wrapped around the outside of the piezometer at 0.64-m intervals between 4 and 18 m depth. Electrical contact in the unsaturated zone was achieved by grouting over the full length above the water table. Contact resistance between electrode and formation ranged from 0.1 to 10 k-ohm, with higher values in the unsaturated zone.

1550 Slater and Glaser Laboratory measurements EXPERIMENT The inherent difficulty in retrieving sandy unconsolidated sediments often limits laboratory electrical studies to repacked samples. The degree to which laboratory relationships apply at the field scale is a consideration. One can reasonably assume that grain-size distribution and ace area of the repacked sample will be representative of the in-situ material, given careful sample handling. These are the primary physical properties influencing the IP measurements discussed here. The total porosity, effective porosity, grain orientation, and possible anisotropy in grain orientation of repacked samples are unlikely to represent the in-situ condition because they will depend on the packing procedure. Laser particle-size analysis was used to determine the grainsize distribution for all samples. Twelve samples from the aquifer, six per well, were selected as representative of the range of grain-size characteristics encountered in this aquifer. Figure 1 shows the grain-size curves for four of the twelve FIG. 1. Grain-size distribution obtained from laser particle-size analysis for 4 of the 12 samples used in this study. The selected samples span the range of grain size encountered in these materials. samples used in this study. Selected samples shown in Figure 1 span the range of grain sizes in this study and illustrate that the aquifer is composed of well-sorted sands. Specific ace area was measured using a nitrogen BET method. Effective porosity φ eff is difficult to measure and was not determined in this study. However, total porosity φ tot, likely to be close to φ eff in these loosely packed sediments, was determined from wet and oven-dried weight measurements. Mineralogy was investigated using X-ray diffraction, and major/minor constituents were identified with the help of a computer program. Quartz was the only major mineral present in the twelve samples. Albite, microcline, and albide were detected as trace constituents in some samples. The results of laser particle-size analysis indicate that clay-size particles constitute negligible sample mass for all samples (Figure 1). Thus, the samples are considered effectively devoid of clay minerals and clay particles. Hydraulic conductivity K was measured using standard flow tests for sandy unconsolidated sediments (American, 1997). Measurements were made at four discharge velocities, and mean K and standard deviation σ dev were estimated. Electrical measurements were made for six NaCl solutions with σ w varying between 0.01 and 10 S/m. Formation factor F was estimated from the gradient of σ σ el versus σ w [Equation (3)]. This estimate assumes that σ is insignificant, deemed appropriate for these coarse, clay-free materials and verified by linearity in the plot of σ versus σ w. Table 1 summarizes sample physical properties and associated error estimates. A modified permeameter was used for simultaneous measurement of electrical and hydraulic properties on saturated, repacked samples (Figure 2). Four electrode electrical measurements were obtained with a National Instruments NI 4551 dynamic signal analyzer. Phase shift between current stimulusvoltage signal φ, i.e., between channel 1 and channel 2 in Figure 2, and conductivity magnitude σ were measured at 40 frequencies, spaced at equal logarithmic intervals, between 0.1 and 1000 Hz. Current was injected at silver coil electrodes located at sample ends. Sample voltage was measured using silver silver chloride point electrodes located just outside the current flow path. An AD620 preamplifer boosted the input impedance on the voltage channel and prevented current leakage into the circuitry (Figure 2a). The phase response of the Table 1. Measured physical properties of unconsolidated sediments from the Olathe study site, Kansas River floodplain: d 10 = grain diameter for which 10% of sample is finer by weight; d 90 = grain diameter for which 90% of sample is finer by weight; φtot = total porosity; Ss = specific ace area; F = formation factor (ND indicates not determined). d 10 d 90 S s Sample (m) (m) φ tot (m 2 /g) F N1 22.5 25 4.74E 05 2.89E 04 0.46 2.83 2.93 N1 26.5 35 6.77E 05 7.81E 04 0.41 0.54 5.25 N1 37 40 2.98E 04 1.41E 03 0.37 1.29 5.59 N1 40 43.5 3.29E 04 1.24E 03 0.35 0.91 ND N1 44 51 1.27E 04 1.15E 03 0.40 0.59 3.43 N1 51 53 3.11E 04 1.15E 03 0.41 0.99 5.35 N2 19 25 1.89E 04 5.41E 04 0.35 0.74 4.81 N2 25 26.5 2.16E 05 4.90E 04 0.38 1.06 4.49 N2 27 35 7.75E 05 1.03E 03 0.39 1.76 4.00 N2 45 47 5.40E 04 1.58E 03 0.32 0.66 7.32 N2 47 55 1.55E 05 1.25E 03 0.38 1.12 5.30 N2 55 60 7.50E 04 1.61E 03 0.33 0.44 8.66 error ±3.4% ±3.4% ±0.01 ±0.01 ±0.01

Controls on induced polarization 1551 preamplifier, which depends on frequency and sample resistance, was removed by obtaining a calibration on a precision resistor matching the sample resistance (Figure 2b). Real and imaginary conductivity were calculated from equation (1). Accuracy in the conductivity estimates was primarily determined by error in the phase measurement. Repeatibility tests indicate that relative errors are generally less than 5% and 1% for σ and σ, respectively (Slater and Lesmes, 2002). Prior to σ measurements, samples were flushed until outflow σ w matched inflow σ w, at which point pore σ w was considered known. Changes in pore σ w from ion exchange between the mineral ace and free electrolyte were not quantifiable but were assumed small because of the low cation exchange capacity (CEC) of coarse, clay-free sediments. Figure 3 shows the real and imaginary conductivity plotted as a function of frequency for all 12 samples. As σ σ el in these samples, σ is essentially frequency independent (Figure 3a). The frequency dependence of the imaginary conductivity (σ = σ ) is illustrated in Figure 3b. For most of the samples, σ is only weakly frequency dependent between 0.1 and 100 Hz. Some samples show power law dependence across the entire frequency range (e.g., N1 37 40 and N2 25 26.5) as previously reported for unconsolidated sediments and weakly consolidated sandstone (Börner et al., 1996). Other samples exhibit increased frequency dependence, with σ (ω) increasing more rapidly above 100 Hz (e.g., N2 45 47 and N2 47 55). These differences in σ (ω) may reflect differences in the grain/pore-size distribu- tion between samples. Lesmes and Morgan (2002) suggest that the relative proportion of grains with a certain radius determine the magnitude of σ (ω) at a particular frequency. Smaller repacked samples (length = 63 mm; radius = 25 mm) were constructed to investigate the dependence of σ on saturation in these materials. Degree of saturation was varied in two ways. Evaporative drying by exposure of the sample to air was one approach. Degree of saturation was determined from the weight loss associated with evaporative drying. These small samples were also interfaced with a high-precision syringe pump permitting accurate withdrawal and infusion of water. This allowed investigation of the dependence of σ on saturation over a drainage and imbibition cycle. Field measurements Crosshole electrical measurements were made between the two boreholes installed with electrodes at the Olathe well field. All measurements were made using an electrode configuration whereby voltage and current pairs were split between the two boreholes (bipole bipole configuration). Bing and Greenhalgh (2000) numerically illustrate that this electrode geometry provides good image resolution relative to other frequently used crosshole electrode configurations. The measurement sequence consisted of 1380 measurements and 1380 corresponding reciprocals (whereby the current and voltage pairs are exchanged), addressing 48 electrodes (24 per borehole). Relative reciprocal errors were generally less than 3% for the conductivity magnitude, with outliers discarded from the inversion. Chargeability error is better quantified by the absolute reciprocal error, as many low chargeability (<5 mv/v) measurements are obtained. Chargeability reciprocal errors were generally less than 1 mv/v, with outliers again discarded from the inversion. The data set was inverted by A. Binley using code developed at Lancaster University (see Kemna and Binley, 1996; Kemna et al., 1999). This 2D algorithm, based on an Occam s procedure (degroot-hedlin and Constable, 1990), implicitly assumes that the electrical structure in the third dimension is constant. The real world is rarely two dimensional. However, this approximation is relatively appropriate for such a sedimentary sequence, at least over the horizontal sensitivity of the electrical measurements (<10 m). Lithologic logs from four wells drilled over 20 m at the site, as well as ace resistivity measurements, support this assumption. The inversion requires measurements of σ and φ as input data. However, because of the equivalence between field and normalized IP parameters, we substituted measures of M for φ. The output from the algorithm is then the distribution of σ and a scaled phase angle, i.e., M. Similarly, the distribution of scaled measures of σ, simply defined as σ field, and σ, being the normalized chargeability M n (Lesmes and Frye, 2001), were obtained. RESULTS Laboratory results FIG. 2. (a) Schematic of instrumentation used to obtain low-frequency (0.1 1000 Hz) electrical measurements for sandy, unconsolidated sediments, (b) Example of measured phase φ for a sample calibration resistor and the corrected sample. In conductivity space, a positive phase is a capacitive effect. Figure 4 illustrates the dependence of measured electrical parameters on sample structural properties. Structural controls on electrolytic conductivity are assessed using F, which is the ratio of σ w to σ el [Equation (3)] and hence the ratio of σ w to σ when σ el σ (i.e. at high σ w ). Controls on magnitude of ace

1552 Slater and Glaser conductivity are assessed using σ (1Hz) to σ (1Hz) [equation (2)]. which is considered equal Formation factor versus φ tot (Figure 4a) follows an Archie relationship [equation (3)] with m = 1.6 ± 0.1, typical for unconsolidated sediments (e.g., Schön, 1996). The weak correlation coefficient reflects the limited φ tot range (0.32 0.46) for these materials. The imaginary conductivity (i.e., σ ) approximates a linear dependence on S tot (Figure 4b) and also correlates with S por. For our samples σ is slightly better correlated with S tot than with S por. The correlation between grain size and σ depends on the percentage of the sample weight used to define the representative grain size. Figure 4c shows only a weak relationship between σ and d 10 for samples measured in this study. However, the σ dependence on grain diameter is clear when d 50 (the grain diameter for which 50% of the sample is finer by weight) or, better still, d 90 (the grain diameter for which 90% of the sample is finer by weight) is used as the representative grain-size parameter (Figures 4d and 4e). The correlation between representative grain-size parameter and σ is effectively frequency independent between 1 and 100 Hz, and R 2 (defining the strength of the correlation) increases smoothly from d 10 to d 90. Figure 5 illustrates the predictive capability of the Börner model for K estimation [equation (8)] when applied to these sediments. Least-squares regression of our data set yields c = 1.4 and A = 2.3 10 5 with K in m/day and σ in S/m. Hydraulic conductivity predicted using the Börner model is plotted versus that measured with flow tests. The hydraulic conductivity of five of the eleven samples is well predicted. For the remaining six samples the predictive capability of the model is a little less than an order of magnitude. Note that the measured hydraulic conductivity spans only two orders of magnitude in these sediments. Figure 6 shows the dependence of measured σ and σ on σ w for four NaCl-saturated samples. The linear σ increase with increasing σ w reflects the dominance of σ el over σ in these samples and follows from equation (3). All four samples exhibit an increase in imaginary conductivity (σ = σ ) with increasing σ w at low σ w. An apex is observed, after which σ decreases with further increase in σ w. The position of the apex varies between samples but lies within 0.2 1.0 S/m. Figure 6 illustrates that σ is much less σ w dependent than σ el. Within the σ w range of 0.1 1.0 S/m, often typical of natural groundwater, σ is approximately constant for these samples. FIG. 3. Plots of (a) σ (ω) and (b) σ (ω) over the frequency range 0.1 1000 Hz for the 12 sandy unconsolidated sediments examined in this study. For these clay-free samples, σ σ el and is thus frequency independent to the first order.

Controls on induced polarization 1553 Figure 7 illustrates the dependence of σ and σ on saturation for sample N2 55 60. Figure 7a shows the change in the conductivity terms as the sample is dried by evaporation from S = 1toS = 0.15. As a first approximation we fit single σ and σ saturation exponents to the entire saturation range. The σ σ el saturation exponent n [equation (3)] is 1.3 ± 0.1, and the saturation exponent for σ is 0.6 ± 0.1. Similar saturation dependence is also observed for two additional measured samples, with the σ saturation exponent consistently about half the σ saturation exponent. Figure 7b illustrates the dependence of σ and σ on saturation as water content in sample N2 27 35 is varied through a pressure drainage and imbibition cycle over the range S = 1toS = 0.28. Hysteresis in both the σ and σ responses to saturation is observed, with higher values occurring during imbibition from S = 0.3 0.7. The smaller sampling interval in saturation space achieved with the pressure drainage and imbibition experiments illustrates that saturation exponents vary over the saturation range. FIG. 4. Observed relationships between electrical measurements (1 Hz) and physical properties of unconsolidated sandy sediments. (a) Comparison of measured data (circles) and relationship given in equation (3), assuming φ tot φ eff (exponent m from least-squares minimization = 1.6 ± 0.1, R 2 = 0.58). (b) Squares: approximate linearity between σ and S tot, showing the best-fit relationship σ = 1 10 5 (S tot ) 1.1±0.2 (R 2 = 0.66); triangles: relationship between σ and S por, showing best-fit relationship σ = 3 10 6 (S por ) 1.2±0.3 (R 2 = 0.59). (c) Observed relationship between σ and d 10. (d) Observed correlation between σ and d 50, showing the best-fit relationship σ = 10 8 (d 50 ) 1±0.3 (R 2 = 0.54). (e) Observed correlation between σ and d 90, showing the best-fit relationship σ = 10 8 (d 90 ) 1.1±0.2 (R 2 = 0.65).

1554 Slater and Glaser FIELD RESULTS Figure 8 shows the results of electrical imaging correlated with summary lithologic logs. Sample locations and d 90 values obtained in the laboratory are also shown for comparison. The lithology consists of a general fining-upward sequence of sandy sediments overlain by silty clay. The groundwater ace falls within the fine medium sand unit at 10 m depth, approximately 2 m below the interface with the overlying fine sand. FIG. 5. Comparison of hydraulic conductivity predicted using the Börner model (K pred ) and that measured from flow tests on laboratory samples (K meas ). The straight line shows the 1:1 relation. The M n image shows a clear correlation with lithology in the sandy sediments and is consistent with the laboratory findings. There is an increase in M n throughout the fining-upward sequence and, given the difference in scale between laboratory and field measurements, a reasonable inverse correlation between M n and d 90 obtained from samples. Boundaries between major lithologic units at approximately 8 and 14 m correlate with shifts in the M n image. Note that M n is not strongly affected by the groundwater ace. The σ field image generally shows weaker correlation with lithology, although the interface between the fine sand and fine medium sand is well resolved. The overlying silty clay exhibits high σ field, presumably as a result of high σ. (Laboratory measurements of this material were not made.) High σ field in the fine sand may reflect high σ w of capillary-held pore fluid, supplied by ion-rich water percolating through the overlying silt/clay. The groundwater ace at 10 m affects σ el and hence σ field, causing an image structure within the fine medium sand that is unrelated to lithology. The interface between saturated and unsaturated fine medium sand causes a decrease in σ field above 10 m. Figure 8c is an image of the predicted hydraulic conductivity distribution based on the Börner model [equation (8)] using model parameters estimated from the laboratory measurements. The image is truncated above the groundwater ace as the model assumes saturation. The formation factor for each pixel in the image was estimated from the ratio of σ w measured in a well (0.09 ms/m) at the site to the pixel real conductivity (σ field ). This calculation assumes that measured σ w is representative of the pore fluid throughout the image zone and that σ field σ el. The validity of the first assumption is somewhat FIG. 6. Measured σ and σ (1 Hz) as a function of σ w for an NaCl solution: (a) N1 26.5 35, (b) N1 51 53, (c) N2 19 25, (d) N2 27 35 ( =σ, =σ ).

Controls on induced polarization 1555 FIG. 7. (a) Measured σ and σ (1 Hz) as a function of saturation during evaporative drying of sample N2 55 60. The σ σ el saturation exponent n [equation (7)] is 1.3 ± 0.1 (R 2 = 0.98), and the saturation exponent for σ is 0.6 ± 0.1 (R 2 = 0.96). (b) Measured σ and σ (1 Hz) as a function of saturation during pressure drainage (triangles) and subsequent imbibition (circles) for sample N2 27 35. FIG. 8. Inversion result for 2D crosshole electrical survey at the Olathe field site correlated with site lithology, (a) M n with d 90 values obtained from laser particle-size analysis superimposed. (b) σ field. (c) Predicted hydraulic conductivity obtained with the Börner model. The M n color scale covers three orders of magnitude, whereas the σ field color scale covers only one order of magnitude. The groundwater ace falls within the fine medium sand unit at 10 m depth, approximately 2 m below the interface with the overlying fine sand.

1556 Slater and Glaser uncertain. The latter assumption is reasonable given the absence of clays and the relatively high fluid conductivity. Value M n was directly substituted for σ in the model. The basal sand and gravel layer is imaged as a hydraulically conductive zone, with K pred reaching 100 m/day. The overlying fine medium sand layer is imaged as a region with K pred 1 m/day. DISCUSSION Significance of laboratory measurements Our results illustrate the controls on induced polarization in sandy, unconsolidated sediments. The internal ace area is a fundamental lithological parameter controlling σ in these sediments. We find that σ is better correlated with the ace area normalized to the total volume S tot than with the ace area normalized to the pore volume S por.börner et al. (1996) find that using S por rather than S tot improves the correlation with σ when porosity varies considerably. Our result probably reflects the limited porosity range typical of sandy unconsolidated sediments (A. Weller, 2002, personal communication). Observed relationships between σ and representative grain-size parameters will vary, depending on how the ace area is distributed across the grain-size distribution. Consider first a material in which less than 10% of its weight accounts for most of its internal ace. In this case a strong correlation between σ and d 10 is expected. Sands with significant clay content fit this category because most of the ace area is associated with the finer clay particles (S o 1/R r ). The artificial sand clay mixtures and glacial till samples described in Slater and Lesmes (2002) match this description and illustrate a strong correlation between σ and d 10. In the case of coarse, well-sorted sandy samples described in our study, most of the sample weight is required to account for the majority of the internal ace. The imaginary conductivity then correlates with a representative grain diameter for which the majority of the sample is finer (d 90 ), as we observe. Relationships observed in Figures 4c 4e thus reflect the distribution of the specific ace across the grain size in these samples. Presumably, the strong correlation between σ and d 50 shown in Schön (1996) reflects the wide range of material type for which the median grain size accounts for most of the internal ace. Relationships between σ and effective grain diameter determined for natural materials thus depend on the grain-size distribution and how this controls the distribution of the internal ace. Consequently, any relationship between σ and effective grain diameter is site dependent. The predictive capability of the Börner model for K estimation from IP measurements is somewhat restricted by the limited K range typical of unconsolidated, sandy sediments. Börner et al. (1996) conclude that order-of-magnitude K estimates were achievable with this model when applied to a large range of material types. In our study, K varies by less than two orders of magnitude, and the predictive capability of the model when applied to these sediments is slightly better than an order of magnitude. The characteristic shape of the σ dependence on σ w observed for unconsolidated, sandy sediments is very similar to that observed by Lesmes and Frye (2001) for Berea Sandstone. They interpret this shape in terms of competing effects of ace charge density and ace ionic mobility on ace conductivity. Adopting this interpretation, we find that below about 0.5 S/m the increase in ace charge density with increasing σ w dominates the decrease in ace ionic mobility with increasing σ w. Consequently, σ increases with increasing σ w. Above about 1 S/m the effect of decreasing ace ionic mobility with σ w increase dominates, and the relationship between σ and σ w reverses. Fluid conductivity generally exerts a weak control on σ, relative to its control on σ σ el, in these clay-free unconsolidated sediments. Within a typical σ w range for natural groundwater, σ is approximately invariant in sandy, unconsolidated sediments. However, little is known about how the σ dependence on σ w varies with clay mineralogy, clay content, and clay distribution (addressed by Wildenschild et al., 2001) throughout the matrix. Clay is characterized by high CEC, which may enhance changes in ace chemistry with varying σ w. We thus recognize a need to expand studies of the controls of fluid chemistry on ace conductivity to a wider range of earth materials, particularly clay containing unconsolidated sediments. We show that ace conductivity is saturation dependent in sandy, unconsolidated sediments. During evaporative drying we observe that the saturation exponent for σ is approximately half that for σ, i.e., σ is much less dependent on saturation. Vinegar and Waxman (1984) report a similar relationship between real and imaginary saturation exponents for sandstone samples when oil is the nonconducting, porefilling medium. Hysteresis in the σ versus saturation curve (as observed in our data) is well documented and generally attributed to differing distributions of pore fluid resulting from drainage and imbibition processes (e.g., Longeron et al., 1989; Knight, 1991). To the best of our knowledge, we show the first results that illustrate hysteresis in the σ versus saturation response during drainage and imbibition. The cause of the σ dependence on saturation is uncertain and complicated by the fact that ace conduction may exist at the air fluid interface, as well as the solid fluid interface, in unsaturated sediments (Knight, 1991). A detailed discussion of mechanisms causing the dependence of σ our paper. Relevance to other applications on saturation is beyond the scope of Our laboratory studies indicate that ace conductivity in sandy, unconsolidated sediments is strongly dependent on lithological properties and is less dependent on fluid conductivity and saturation of the pore space. In contrast, σ measured in resistivity and electromagnetic surveys is highly dependent on fluid conductivity and saturation, which may limit successful interpretation of these geophysical measurements in terms of lithologic variability. Our discussion thus favors the use of IP for investigating lithologic variability in near-ace, unconsolidated sediments. Relationships between σ and lithologic parameters exist even when lithologic variability is relatively small, such as in the sandy aquifer studied here. Our field study provides an example of the potential significance of the current study to geophysical investigations. Fieldscale crosshole electrical imaging in this sandy aquifer demonstrates that subtle lithologic variability is resolvable from an IP image. Any dependence of IP on saturation, which would manifest itself as a decrease in M n above the groundwater ace, is minor relative to the influence of lithologic variability. In

Controls on induced polarization 1557 contrast, the conductivity image (i.e., σ field ) is less well correlated with the site lithology and is significantly influenced by the position of the groundwater ace. This field example also demonstrates how the Börner model might be used to depict order-of-magnitude or greater hydraulic conductivity variability from crosshole IP surveys in sandy sediments. The prediction of hydraulic conductivity from geophysical measurements continues to receive considerable attention. The Börner model is an attractive approach, as it has a semitheoretical basis and only requires an estimate of the groundwater conductivity in addition to the geophysical measurement. However, the predictive capability of this model in unconsolidated sandy aquifers is limited by the relatively narrow range of K in such materials. CONCLUSIONS This study illustrates that previously identified relationships between IP measurements and lithological parameters are discernible even when lithologic variability is restricted to a narrow range, as is typically encountered in a sequence of sandy deposits. However, the correlation between a representative grain diameter and IP measurements depends on the particlesize characteristics of the material. In the case of well-sorted, sandy sediments, IP measurements correlate most closely with measures of the coarsest fraction of the material. We find the predictive capability of the Börner model for K estimation from IP measurements is limited when applied to this narrow lithologic range of material type. Importantly, IP measurements are a function of saturation and exhibit hysteretic behavior over a drainage and imbibition cycle. Our results suggest that competing effects of ace charge density and ace ionic mobility on ace conductivity may explain the dependence of IP measurements on fluid conductivity in these materials. Results of crosshole electrical imaging of an alluvial sequence of sandy sediments illustrate how high-resolution IP measurements might be used to generate images of lithologic variability. The hydraulic conductivity distribution obtained from the Börner model successfully discriminates a hydraulically conductive sand/gravel from overlying medium sand. Further studies are required to better define the nature of ace conductivity in clay containing sediments as well as in unsaturated materials. ACKNOWLEDGMENTS This work was supported by National Science Foundation award EAR-0073680 and American Chemical Society award PRF-36265-G2. Andrew Binley (Lancaster University, UK) performed the complex conductivity inversion using code developed at Lancaster. We thank D. Stous and C. Marone (Burns and MacDonnell Engineers) for permitting access to the Olathe well field and providing samples for laboratory analysis. Rutgers University graduate student Craig Ulrich collected the data obtained for unsaturated samples. University of Missouri Kansas City undergraduate student Isaiah Utne provided valuable lab and field support. L. S. is grateful to David Lesmes (Boston College) for many insightful discussions on the nature of IP mechanisms. Andreas Weller (Technische Universität Clausthal) provided constructive review comments that were critical in improving the interpretation of this data set. This paper also benefited from reviews provided by David Nobes and an associate editor. REFERENCES American Society for Testing and Materials (ASTM), 1997, Annual book of ASTM standards: Soil and rock I: D 420 D 4914. Archie, G. E., 1942, The electrical resistivity log as an aid in determining some reservoir characteristics: Trans. Am. Inst. Min. Metall. Petr. Eng., 146, 54 62. Bing, Z., and Greenhalgh, S. A., 2000, Configurations for cross-hole resistivity tomography: Geophys. Prosp., 48, 887 912. Börner, F. D., and Schön, J. H., 1991, A relation between the quadrature component of electrical conductivity and the specific ace area of sedimentary rocks: The Log Analyst, 32, 612 613. Börner, F. D., Gruhne, M., and Schön, J., 1993, Contamination indications derived from electrical properties in the low frequency range: Geophys. Prosp., 41, 83 98. Börner, F. D., Schopper, J. R., and Weller, A., 1996, Evaluation of transport and storage properties in the soil and groundwater zone from induced polarization measurements: Geophys. Prosp., 44, 583 602. Bussian, A. E., 1983, Electrical conductance in a porous medium: Geophysics, 48, 1258 1268. Dahlin, T., Leroux, V., and Nissen, J., 2002, Measuring techniques in induced polarisation imaging: J. Appl. Geophys., 50, 279 298. degroot-hedlin, C., and Constable, S. C., 1990, Occam s inversion to generate smooth, two-dimensional models from magnetotelluric data: Geophysics, 55, 1613 1624. De Lima, O. A. L., and Niwas, S., 2000, Estimation of hydraulic parameters of shaly sandstone aquifers from geoelectrical measurements: J. Hydrol., 235, 12 26. Johnson, D. L., Koplik, J., and Schwartz, L. M., 1986, New pore-size parameter characterizing transport in porous media: Phys. Rev. Lett., 57, 2564 2567. Keller, G. V., 1959, Analysis of some electrical transient measurements on igneous, sedimentary and metamorphic rocks, in Wait, J. R., Ed., Overvoltage research and geophysical applications: Pergamon Press, Inc. Kemna, A., and Binley, A., 1996, Complex electrical resistivity tomography for contaminant plume delineation: 2nd Ann Mtg., Environ. Eng. Geophys. Soc., Eur. Section, Proceedings, 196 199. Kemna, A., Binley, A., Ramirez, A., and Daily, W., 1999, Complex resistivity tomography for environmental applications: Chem. Eng. J., 77, 11 18. Knight, R., 1991, Hysteresis in the electrical resistivity of partially saturated sandstones: Geophysics, 56, 2139 2147. Knight, R. J., and Nur, A., 1987, The dielectric constant of sandstones, 60 Hz to 4 MHz: Geophysics, 52, 644 654. Lesmes, D. P., and Frye, K. M., 2001, The influence of pore fluid chemistry on the complex conductivity and induced-polarization responses of Berea Sandstone: J. Geophys. Res., 106, No. B3, 4079 4090. Lesmes, D. P., and Morgan, F. D., 2001, Dielectric spectroscopy of sedimentary rocks: J. Geophys. Res., 106, No. B7, 13329 13346. Longeron, D. G., Argaud, M. J., and Feraud, J. P., 1989, Effect of overburden pressure and the nature and microscopic distribution of fluids on electrical properties of rock samples: SPE Form. Eval., 4, June, 194 201. Marshall, D. J., and Madden, T. R., 1959, Induced polarization: A study of its causes: Geophysics, 24, 780 816. McNeill, J. D., 1990, Use of electromagnetic methods for groundwater studies, in Ward, S. H., Ed., Geotechnical and environmental geophysics 1: Soc. Expl. Geophys., 191 218. Ogilvy, A. A., and Kuzmina, E. N., 1972, Hydrologic and engineering geologic opportunities for employing the method of induced potentials: Geophysics, 37, 839-861. Olhoeft, G. R., 1985, Low-frequency electrical properties: Geophysics, 50, 2492 2503. Pride, S. R., 1994, Governing equations for the coupled electromagnetics and acoustics of porous media: Phys. Rev. B., 50, 15678 15696. Revil, A., and Cathles, L. M., III, 1999, Permeability of shaly sands: Water Resource Res., 35, 651 662. Revil, A., and Glover, P. W. J., 1998, Nature of ace electrical conductivity in natural sands, sandstones, and clays: Geophys. Res. Lett., 25, 691 694. Revil, A., Cathles, L. M., III, Losh, S., and Nunn, J. A., 1998, Electrical conductivity in shaly sands with geophysical applications: J. Geophys. Res., 103, B10, 23925 23936. Samstag, F. J., and Morgan, F. D., 1991, Induced polarization of shaly sands: Salinity domain modeling by double embedding of the effective medium theory: Geophysics, 56, 1749 1756.

1558 Slater and Glaser Schön, J. H., 1996, Physical properties of rocks Fundamentals and principles of petrophysics: Peramon Press. Sen, P. N., Scala, C., and Cohen, M. H., 1981, A self-similar model for sedimentary rocks with application to the dielectric constant of fused glass beads: Geophysics, 46, 781 795. Slater, L., and Lesmes, D. P., 2002, Electical-hydraulic relationships observed for unconsolidated sediments: Water Resource, Res., 38, No. 10, 31-1 31-10. Vacquier, V., Holmes, R., Kintzinger, P. R., and Lavergne, M., 1957, Prospecting for ground water by induced electrical polarization: Geophysics, 22, 660 687. Vanhala, H., 1997, Mapping oil contaminated sand and till with the spectral induced polarization method: Geophys. Prosp., 45, 303 326. Vanhala, H., and Soininen, H., 1995, Laboratory technique for measurement of spectral induced polarization response of soil samples: Geophys. Prosp., 43, 655 676. Vanhala, H., Soininen, H., and Kukkonen, I., 1992, Detecting organic chemical contaminants by spectral induced-polarization method in a glacial till environment: Geophysics, 57, 1014 1017. Vinegar, H. J., and Waxman, M. H., 1984, Induced polarization of shaly sands: Geophysics, 49, 1267 1287. Ward, S. H., 1990, Resistivity and induced polarization methods, in Ward, S. H., Ed., Geotechnical and environmental geophysics: Investigations in geophysics, 5, Soc. Expl. Geophys., 169 189. Ward, S. H., and Fraser, D. C., 1967, Conduction of electricity in rocks, in Mining geophysics 2: Soc. Expl. Geophys., Waxman, M. H., and Smits, L. J. M., 1968, Electrical conductivities in oil-bearing shaly sands: Soc. Petr. Eng. J., 243, 107 122. Weller, A., Gruhne, M., Seichter, M., and Börner, F. D., 1996, Monitoring hydraulic experiments by complex conductivity tomography: Euro. J. Environ. Eng. Geophys., 1, 209 222. Wildenschild, D., Roberts, J. J., and Carlberg, E., 2001, On the relationship between microstructure and electrical and hydraulic properties of sand-clay mixtures: Geophys. Res. Lett., 27, 3085 3088.