24
This article was downloaded by: [58.7.214.162] On: 30 January 2014, At: 22:33 Publisher: Taylor & Francis Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK Communications in Soil Science and Plant Analysis Publication details, including instructions for authors and subscription information: http://www.tandfonline.com/loi/lcss20 Calibration Functions for Estimating Soil Moisture from GPR Dielectric Constant Measurements Remke L. Van Dam a b a Michigan State University, Department of Geological Sciences , East Lansing , Michigan , USA b Queensland University of Technology, Science and Engineering Faculty, Institute for Future Environments , Brisbane , Queensland , Australia Accepted author version posted online: 04 Nov 2013.Published online: 30 Jan 2014. To cite this article: Remke L. Van Dam (2014) Calibration Functions for Estimating Soil Moisture from GPR Dielectric Constant Measurements, Communications in Soil Science and Plant Analysis, 45:3, 392-413, DOI: 10.1080/00103624.2013.854805 To link to this article: http://dx.doi.org/10.1080/00103624.2013.854805 PLEASE SCROLL DOWN FOR ARTICLE Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and should be independently verified with primary sources of information. Taylor and Francis shall not be liable for any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of the Content. This article may be used for research, teaching, and private study purposes. Any substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &

bPlant Analysis Communications in Soil Science andrvd/pub/csspa_2014.pdf · tant part of the interlinked global freshwater system. Because the vadose zone is ... number of θ–ε

Embed Size (px)

Citation preview

This article was downloaded by: [58.7.214.162]On: 30 January 2014, At: 22:33Publisher: Taylor & FrancisInforma Ltd Registered in England and Wales Registered Number: 1072954 Registeredoffice: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Communications in Soil Science andPlant AnalysisPublication details, including instructions for authors andsubscription information:http://www.tandfonline.com/loi/lcss20

Calibration Functions for Estimating SoilMoisture from GPR Dielectric ConstantMeasurementsRemke L. Van Dam a ba Michigan State University, Department of Geological Sciences ,East Lansing , Michigan , USAb Queensland University of Technology, Science and EngineeringFaculty, Institute for Future Environments , Brisbane , Queensland ,AustraliaAccepted author version posted online: 04 Nov 2013.Publishedonline: 30 Jan 2014.

To cite this article: Remke L. Van Dam (2014) Calibration Functions for Estimating Soil Moisture fromGPR Dielectric Constant Measurements, Communications in Soil Science and Plant Analysis, 45:3,392-413, DOI: 10.1080/00103624.2013.854805

To link to this article: http://dx.doi.org/10.1080/00103624.2013.854805

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the“Content”) contained in the publications on our platform. However, Taylor & Francis,our agents, and our licensors make no representations or warranties whatsoever as tothe accuracy, completeness, or suitability for any purpose of the Content. Any opinionsand views expressed in this publication are the opinions and views of the authors,and are not the views of or endorsed by Taylor & Francis. The accuracy of the Contentshould not be relied upon and should be independently verified with primary sourcesof information. Taylor and Francis shall not be liable for any losses, actions, claims,proceedings, demands, costs, expenses, damages, and other liabilities whatsoever orhowsoever caused arising directly or indirectly in connection with, in relation to or arisingout of the use of the Content.

This article may be used for research, teaching, and private study purposes. Anysubstantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &

Conditions of access and use can be found at http://www.tandfonline.com/page/terms-and-conditions

Dow

nloa

ded

by [

58.7

.214

.162

] at

22:

33 3

0 Ja

nuar

y 20

14

Communications in Soil Science and Plant Analysis, 45:392–413, 2014Copyright © Taylor & Francis Group, LLCISSN: 0010-3624 print / 1532-2416 onlineDOI: 10.1080/00103624.2013.854805

Calibration Functions for Estimating Soil Moisturefrom GPR Dielectric Constant Measurements

REMKE L. VAN DAM

Michigan State University, Department of Geological Sciences, East Lansing,Michigan, USA, and Queensland University of Technology, Science andEngineering Faculty, Institute for Future Environments, Brisbane, Queensland,Australia

Ground-penetrating radar (GPR) is widely used for assessment of soil moisturevariability in field soils. Because GPR does not measure soil water content directly,it is common practice to use calibration functions that describe its relationship with thesoil dielectric properties and textural parameters. However, the large variety of mod-els complicates the selection of the appropriate function. In this article an overview ispresented of the different functions available, including volumetric models, empiricalfunctions, effective medium theories, and frequency-specific functions. Using detailedinformation presented in summary tables, the choice for which calibration function touse can be guided by the soil variables available to the user, the frequency of the GPRequipment, and the desired level of detail of the output. This article can thus serve as aguide for GPR practitioners to obtain soil moisture values and to estimate soil dielectricproperties.

Keywords Ground-penetrating radar, hydrogeology, soil water

Introduction

Fresh water is an essential commodity and a vital resource for many ecosystems on earth.However, with the demands for water increasing from all parts of society, the availability ofsufficient clean water is under threat from numerous environmental problems (Entekhabiet al. 1999; Vörösmarty et al. 2000). Although a large part of the available fresh wateris stored in subsurface aquifers, unsaturated zone processes such as recharge and evapo-transpiration make it actively interact with the atmosphere. This unsaturated, or vadose,zone forms a natural buffer between the groundwater and the atmosphere and is an impor-tant part of the interlinked global freshwater system. Because the vadose zone is directlyaffected by many human activities, it is of great importance to understand and monitor thespatial variability and temporal changes in the vadose zone on both global and local scales.

In recent decades, ground-penetrating radar (GPR) has become a popular method forthe study of soil water conditions and dynamics at a range of scales, including those relatedto contamination, water (re)distribution, permafrost, and climate change (Huisman et al.2003; Annan 2005; Rubin and Hubbard 2005; Mahmoudzadeh 2013). GPR offers differ-ent techniques for soil moisture estimation, such as normal move-out or common midpoint

Received 12 August 2012; accepted 26 July 2013.Address correspondence to Remke L. Van Dam, Michigan State University, Department of

Geological Sciences, 206 Natural Science, East Lansing, MI 48824, USA. E-mail: [email protected]

392

Dow

nloa

ded

by [

58.7

.214

.162

] at

22:

33 3

0 Ja

nuar

y 20

14

Estimating Soil Moisture from GPR 393

measurements (Greaves et al. 1996; Turesson 2006; Mangel et al. 2012), vertical radar pro-filing and cross-borehole imaging (Hubbard, Rubin, and Majer 1997; Cassiani, Strobbia,and Gallotti 2004; Kuroda, Jang, and Kim 2009), common offset depth profiling (Lunt,Hubbard, and Rubin 2005; Moysey 2010; Schmelzbach, Tronicke, and Dietrich 2012),measurements for estimation of the groundwave velocity (Hubbard, Grote, and Rubin2002; Galagedara et al. 2005; Hamann et al. 2013), and surface reflectivity measurements(Redman et al. 2002; Weihermuller et al. 2007). Recent new approaches include the use ofmultichannel GPR (Bradford 2008; Gerhards et al. 2008) and analysis of early-time signalattributes (Pettinelli et al. 2007; Di Matteo, Pettinelli, and Slob 2013).

The behavior of the electromagnetic energy transmitted by GPR antennas is criti-cally dependent on the dielectric properties of the soil. The dielectric medium propertiescontrol factors such as the attenuation losses and propagation of electromagnetic energy(Miller, Hendrickx, and Borchers 2004; Annan 2005), as well as the reflective contrastbetween an object of study and the surrounding (soil) medium, and the surface reflectivity(Mätzler 1998). The dielectric properties of a material are a function of, among others,texture, bulk density, mineralogy, temperature, organic-matter content, soil water content,and fluid salinity (Hasted 1973; Topp, Davis, and Annan 1980; Wensink 1993; Van Damand Schlager 2000). A further complicating factor is that all these variables change spa-tially and sometimes also temporally in the soil system (Wilson et al. 2003; Van Dam et al.2005).

Because GPR does not measure soil water content directly, it is common practiceto estimate the volumetric water content (θ ) using calibration functions that describe itsrelationship with the soil dielectric properties (ε) and textural parameters. Over the pastseveral decades a large number of θ–ε models has been proposed, many of which areeither purely empirical in nature or are based on volumetric mixing laws. It is still a grandchallenge to select the best calibration function for each occasion, likely in part becausecurrent literature offers few overviews of available options. As a consequence, GPR usersoften resort to using one of only a few typical calibration functions, without consideringwhether other ones may be more appropriate.

This article presents a review of a number of calibration functions for the predictionof dielectric properties of field soils, based on information on water content and texturalvariables. Many of these functions can be easily inverted so that water content can beobtained from dielectric properties. This article is not intended as a theoretical overviewof porous media dielectrics; more on this topic can be found in recent review papers (e.g.,Chelidze and Gueguen 1999; Cosenza et al. 2009). Because of the large number of optionsin literature, this study presents an overview of some of the key calibration functions; it isnot intended as an all-encompassing list. Characteristics, applications, and advantages anddisadvantages are discussed. For easy comparison, key aspects of the different functions(input, output, frequency range) are presented in summary tables.

Background

The interaction of electromagnetic energy with matter is affected by the characteristicsof the material and by the frequency of the electromagnetic energy. Frequency-dependentdielectric properties can be characterized in terms of losses of energy due to relaxationmechanisms that operate at different frequencies. The relaxations are caused by differentforms of atomic- or molecular-scale resonance (Santamarina and Fam 1997). In a soil mix-ture the relaxation mechanisms may be attributed to the solid material and the pore water,as well as to interfacial phenomena. Hasted (1973) summarized some of the different typesof relaxation mechanisms that play a role in wet soils. Because most GPR systems operate

Dow

nloa

ded

by [

58.7

.214

.162

] at

22:

33 3

0 Ja

nuar

y 20

14

394 R. L. Van Dam

in a frequency range between about 50 MHz and 1.2 GHz, bound water relaxation is themajor resonance mechanism of interest. Other relevant resonance mechanisms below andabove the typical GPR frequencies, respectively, are Maxwell-Wagner relaxation around1 MHz and free-water relaxation around 10 GHz (Hilhorst and Dirksen 1994). It must benoted, however, that although bound water has proven to be a very convenient fitting param-eter, it is a somewhat ill-defined concept. Indeed, many functions that use bound water toexplain θ–ε data could also be explained through structural and geometrical effects (e.g.,Sen 1984).

The dielectric permittivity is a complex function with real and imaginary compo-nents, defined as ε∗ = ε′ − jε′′, where j is the square root of –1. The real part of thepermittivity (ε′) is a measure for the total polarizability of the material, which is the ratioof the electric-field storage capacity to that of free space. The real part of the dielectricpermittivity is often expressed as the relative permittivity, dielectric constant, or apparentdielectric permittivity, which are denoted (interchangeably) by the symbols ε′, εr, or Ka.The relative permittivity is a frequency-dependent variable and decreases with increasingfrequency (Powers 1997). The imaginary part (ε′′) of the dielectric permittivity repre-sents the energy absorption by polarization (or relaxation) losses and the ionic conductionlosses. The dielectric losses include dispersive losses, as well as free-water relaxation andbound-water relaxation losses (Hasted 1973).

At frequencies between 0.1 and 1.5 GHz, ε∗ for free water is only weakly frequencydependent and dielectric losses are generally low (Topp, Davis, and Annan 1980; Davis andAnnan 1989). The situation is more complex in fine-textured soils containing clay minerals(Kelleners et al. 2005; Logsdon 2005), and in the presence of ferrimagnetic minerals (VanDam et al. 2013). Also, at these frequencies ε′ and ε′′ are very sensitive to changes insoil water conductivity above about 10 mS/m (Wensink 1993; Knoll and Knight 1994).At frequencies below around 50 MHz, ε∗ depends strongly on soil type (Campbell 1990;Roth et al. 1990). For these low frequencies, the relaxation losses are dominated by theconductivity (Hilhorst and Dirksen 1994). At frequencies above about 1 to 1.5 GHz thedielectric losses increase with increasing free water content, even for low conductivityvalues (Nguyen et al. 1997).

Accurate determination of the relationship between dielectric and hydrogeologicalproperties of soil material is essential for calibration of measurements and model data.Time-domain reflectometry (TDR) is one of the most common electromagnetic tech-niques for measurements of soil water content and soil electrical conductivity. Originally alaboratory-based technique, TDR has become a well-established and widely used methodfor water content determination of field soils at centimeters to decimeter scales (Topp andDavis 1985; Jones, Wraith, and Or 2002; Robinson et al. 2003; Cerny 2009). Several ofthe calibration functions discussed in this paper used TDR to derive or test their relation-ships. The TDR data collection is easily automated, which makes it an ideal technique formonitoring soil moisture variability at high temporal resolutions. However, because of thesmall measurement volume of TDR probes, a large number of probes are needed for goodspatial coverage (Huisman et al. 2002).

Calibration Functions

Volumetric Mixing Models

Volumetric mixing models describe the dielectric properties of a soil based on the relativeamounts of the different soil constituents and their individual dielectric characteristics.The basic input parameters to all models include solid matter, pore space, and volumetric

Dow

nloa

ded

by [

58.7

.214

.162

] at

22:

33 3

0 Ja

nuar

y 20

14

Estimating Soil Moisture from GPR 395

water content. Depending on the model, additional input variables such as organic matterand bound water may provide improved detail and fitting options for specific conditions(Table 1). Usually, frequency-dependent behavior is not taken into account. Most modelshave been calibrated by TDR, which uses a wide frequency range. Over the years differentvolumetric mixing models have been proposed; the basic Liechtenecker-Rother mixingmodel for a material with n components can be written as

εαm =

n∑i=1

viεαi , (1)

where vi is the volume fraction of the ith soil constituent, and α is a fitting parame-ter. Figure 1 demonstrates the important effect of the scaling factor α on the correlationbetween water content and dielectric properties for a water-saturated two-phase medium.The α parameter can theoretically vary from –1 for an electrical field perpendicular to lay-ering to +1 for an electrical field parallel to layering (see insets in Figure 1). For α = 0.5,Eq. (1) becomes the widely used complex refractive index (CRI) or exponential model,which is assumed to be representative of a true isotropic soil medium (Birchak et al. 1974;Roth et al. 1990). Other empirical values and ranges for α reported in the literature include0.33 (Landau and Lifshitz 1960; Zegelin and White 1994), 0.4 to 0.8 for multiphase mix-tures (Jacobsen and Schjonning 1995), 0.46 for three-phase systems (Roth et al. 1990),0.65 for four-phase systems including bound water (Dobson et al. 1985), and 0.67 to 1.0 forpure clays (illite, benthonite) and a vertisol with 84% clay (Dirksen and Dasberg 1993).Based on measurements on quartz–kaolinite mixtures and silt loam soils, Ponizovsky,Chudinova, and Pachepsky (1999) present a mixing model with multiple adjustable param-eters. They found that a model with three adjustable parameters (α, vbw, εb) fitted data forall tested samples (Table 1).

Although the scaling factor α is largely empirical, several attempts have been madeto give it a more physical basis (e.g., Zakri, Laurent, and Vauclin 1998). The value of α ispartly related to the dominant orientation of the sediment grains and pore geometry (e.g.,Brovelli and Cassiani 2008). As shown by Jones and Friedman (2000), the presence ofaligned ellipsoidal particles, for example, in bedding planes of sedimentary deposits, hasan effect on the effective permittivity. It has been shown that the value of α also (inversely)correlates with the measurement frequency (Hilhorst 1998).

Volumetric mixing models can consist of as many components as desired, but usuallycontain three (solids, liquid, air) or four in case of (1) the presence of significant amounts ofclay or organic matter or (2) the presence of a mixture of two liquids such as oil and water.The addition of a fourth component can account for water that is physically and chemi-cally bound to the particles, which has a significantly lower relative permittivity than freewater. In the case of materials with a large specific surface area, such as clay and organicmatter, the bound water component is important to correctly predict the relation betweenthe dielectric properties and total water content of a soil (Figure 2). In four-phase dielectricmixing models, the relative permittivity of the bound water component is represented by asingle value (usually around 3.2 to 5.5). The bound water volume can be calculated usingθbw = lδρbSSA, where l is the number of molecular water layers, each with thickness δ, ρb

is the dry bulk density of the soil, and SSA is the specific surface area of the soil (Dirksenand Dasberg 1993).

In reality, the relative permittivity of the bound water layer is a function of distancefrom the solid surface (Conway, Bockris, and Ammar 1951) and ranges from a minimum(εb,min = 3.2–5.5) near the solid-phase surface toward a maximum where the permittivity

Dow

nloa

ded

by [

58.7

.214

.162

] at

22:

33 3

0 Ja

nuar

y 20

14

Tabl

e1

Ove

rvie

wof

volu

met

ric

mix

ing

mod

els

Inpu

taC

alib

rate

dfo

r

Nam

ean

dre

fere

nce

fra

nge

(GH

z)Te

xtur

eO

ther

Out

putb

fra

nge

(GH

z)N

o.of

soils

No.

ofsa

mpl

esSo

ilty

pes

Bir

chac

ket

al.(

1974

)–

v a,v

,α,ε

iK

a4–

60

2C

rush

edlim

esto

ne,b

ento

nite

clay

Boh

land

Rot

h(1

994)

–v a

,vs

v fw

,εi,

αK

a∫ T

DR

rang

e17

34O

Sc ,USD

Ad:S

aL,S

a,C

lL,L

,Si

L,S

aClL

,SiC

lL,S

iCl,

LSa

Dob

son

etal

.(19

85)e

1.4–

18v a

,vs

v fw

,vbw

,εi,

αε

′ ,ε

′′1.

4–18

55

USD

A:S

aL,L

,SiL

,SiL

,SiC

lH

illho

rste

tal.

(200

0)∞

v iv i

,Si,

εi

ε∗

––

––

Max

wel

l–D

eL

oor

In:D

eL

oor

(196

8)–

v a,v

sv f

w,v

bw,ε

i,ε

b,A

Ka

––

––

In:B

ohla

ndR

oth

(199

4)–

v a,v

sv f

w,v

bw,ε

i,α

Ka

∫ TD

Rra

nge

1734

OSc ,U

SDA

d:S

aL,S

a,C

lL,L

,Si

L,S

aClL

,SiC

lL,S

iCl,

Lsa

Poni

zovk

syet

al.(

1999

)–

v a,v

sv f

w,v

bw(m

ax),

εi,

α∫ T

DR

rang

e7

7U

SDA

d:S

a,L

SaSa

Cl,

Cl,

SiL

Rot

het

al.(

1990

)–

v a,v

sT

w,θ

Ka

∫ TD

Rra

nge

1013

Peat

,USD

Ad:S

aL,C

lL,L

,SiL

,Si

Cl,

LSa

Wan

gan

dSc

hmug

ge(1

980)

–v a

,vs

v fw

,vbw

,εi

ε′ ,

ε′′

1.4–

5–

––

aSy

mbo

lsm

ean

v a,

volu

me

air

frac

tion

inso

il;v s

,vo

lum

eso

lidfr

actio

nin

soil;

v c,

volu

me

clay

frac

tion

inso

il;v i

,vo

lum

eof

ithso

ilco

mpo

nent

,vo

lum

etri

cw

ater

cont

ent;

Tw

,te

mpe

ratu

reof

soil

wat

er;

v bw,

volu

me

boun

dw

ater

frac

tion

inso

il;v b

w(m

ax),

max

imum

volu

me

boun

dw

ater

frac

tion

inso

il;v f

w,

volu

me

free

wat

erfr

actio

nin

soil;

α,e

mpi

rica

lvar

iabl

e;ε

i,re

lativ

epe

rmitt

ivity

ofith

soil

com

pone

nt;ε

b,e

ffec

tive

rela

tive

perm

ittiv

ityne

arbo

unda

ries

;A,d

epol

ariz

atio

nel

lipso

idfa

ctor

;S i

,dep

olar

izat

ion

fact

orof

ithso

ilco

mpo

nent

.bSy

mbo

lsm

ean

Ka,r

elat

ive

perm

ittiv

ity;ε

’,re

alpa

rtof

diel

ectr

icpe

rmitt

ivity

;ε”,

imag

inar

ypa

rtof

diel

ectr

icpe

rmitt

ivity

;ε∗ ,

com

plex

diel

ectr

icpe

rmitt

ivity

.c O

S:or

gani

cso

il.dU

SDA

text

ure

clas

sific

atio

n.Sa

,san

d;L

Sa,l

oam

ysa

nd;

SaL

,san

dylo

am;

SaC

lL,s

andy

clay

loam

;L

,loa

m;

ClL

,cla

ylo

am;

SaC

l,sa

ndy

clay

;C

l,cl

ay;

SiC

l,si

ltycl

ay;S

iClL

,silt

ycl

aylo

am;S

iL,s

iltlo

am;S

i,si

lt.e A

lso

inB

ohla

ndR

oth

(199

4).

396

Dow

nloa

ded

by [

58.7

.214

.162

] at

22:

33 3

0 Ja

nuar

y 20

14

Estimating Soil Moisture from GPR 397

0

20

40

60

80

0 0.2 0.4 0.6 0.8 1

App

aren

t die

lect

ric p

erm

ittiv

ity

Volumetric water content (cm3/cm3)

α = 0.3

α = –1

α = 0.8

α = 1

Figure 1. Diagram illustrating the effects of the value of empirical parameter α, ranging from −1 to1, on the results of a two-phase dielectric mixing model. The relationship was calculated for amedium with 40% porosity. The curves for α values 0.3 and 0.8 (common extremes reported inthe literature for soil material) have been indicated. The insets show the electric equivalents of thetwo extremes of the complex refractive mixing (α = −1 and α = 1), illustrating the effect of grainorientation.

equals that of free water (εb,max = 80). The effective permittivity (εeff) of the mixture ofbound and free water can be estimated using (Friedman 1998)

εeff = dwεb,max

dw + 1

λln

(εb,max − (

εb,max − εb,min)

e−λdw

εb,min

) , and (2)

0.1

0

0

5

10

15

20

25

0 0.1 0.2 0.3 0.4

App

aren

t die

lect

ric p

erm

ittiv

ity

Volumetric water content (cm3/cm3)

Figure 2. Diagram showing the effect of the amount of bound water on a four-phase dielectricmixing model as in Eq. (1), with α = 0.5. The modeled medium consists of 60% grains (relativepermittivity = 4) and 40% porosity, filled with air (1), free water (80), and bound water (5.5). Themaximum volumetric bound water content (cm3/cm3) ranges in four incremental steps from 0 to 0.1.

Dow

nloa

ded

by [

58.7

.214

.162

] at

22:

33 3

0 Ja

nuar

y 20

14

398 R. L. Van Dam

0

20

40

60

80

0 0.1 0.2 0.3 0.4

App

aren

t die

lect

ric p

erm

ittiv

ity

Volumetric water content (cm3/cm3)

Figure 3. Effects of specific surface area and bulk density on the effective permittivity of the waterphase, calculated using Eqs. (2) and (3). Gray lines represent coarse-grained soils with a bulk densityof 1.6 g/cm3 and specific surface areas of 1, 10, 25, and 50 m2/kg, respectively. Black lines representfine-grained soils with a bulk density of 1.2 g/cm3 and specific surface areas of 50, 100, 250, and500 m2/kg, respectively (after Friedman 1998).

dw = θ/ρbSSA. (3)

Here, ρb is the dry bulk density (g/cm3), SSA is the specific surface area (cm2/g) of thematerial, λ is an empirical decay factor taken to be 108 cm−1, and dw is the thickness of thewater film.

Figure 3 shows how the effective permittivity of the total water phase changes asa function of volumetric water content and textural characteristics. With an increase inspecific surface area, the larger contribution of bound water leads to a decrease in εeff.Also, it is apparent that for soils with equal SSA (and equal particle density), an increasein bulk density decreases the effective permittivity (Figure 3). Figure 4 compares the θ–ε

relationship for loamy sand (ρb = 1.47 g/cm3 and SSA = 5.2·105 cm2/g) and a silty claysoil (ρb = 1.19 g/cm3 and SSA = 2.52·106 cm2/g) with that of a three-phase mixturewithout bound water. It shows that not taking the bound water into account leads to anoverestimation of the relative permittivity.

Empirical Functions

Empirical calibration functions are mathematical descriptions of the relationship betweendielectric properties and other characteristics of a medium, especially volumetric watercontent and texture information. There is no physical basis for such functions and theymay therefore be valid only for the soil types that were used to develop the relationship.Many empirical functions have originated in the fields of hydrology and crop and soilsciences and are typically based on network-analyzer measurements using time-domainreflectometry (TDR), waveguides, or coaxial transmission lines.

Dow

nloa

ded

by [

58.7

.214

.162

] at

22:

33 3

0 Ja

nuar

y 20

14

Estimating Soil Moisture from GPR 399

0

5

10

15

20

25

0 0.1 0.2 0.3 0.4

App

aren

t die

lect

ric p

erm

ittiv

ity

Volumetric water content (cm3/cm3)

Figure 4. Apparent dielectric permittivity as a function of volumetric water content for a sandyloam (solid black line) and silty clay (dashed black line) (from Dobson et al. 1985). The effectivepermittivity of water was calculated using Eq. (2) and a three-phase dielectric mixing model as inEq. (1), with α = 0.5. The modeled medium consists of 60% grains (εr = 4) and 40% porosity,filled with air (εr = 1), free water (εr = 80), and bound water (εb,min = 5.5). The result of a simplethree-phase mixing model (solids, air, free water), without the effect of bound water, is shown forcomparison (solid gray line).

The calibration function that has seen the most widespread use is the classic Toppequation, which uses a third-order polynomial to describe the relation between soilvolumetric water content (θ ) and bulk or apparent relative permittivity (Ka) for measure-ments taken below the relaxation frequency of free water (Topp, Davis, and Annan 1980).In its classic form θ is estimated from the apparent relative permittivity; here it is presentedinverted for easier comparison with other calibration functions:

Ka = 3.03 + 9.3θ + 146θ2 − 76.7θ3 (4)

The regression is an average of TDR measurements integrated over a frequency range ofaround 1 MHz to 1 GHz for four mineral soils ranging from clay to sandy loam (Table 2).The regression has been proven to be a reliable estimator of water content or dielectricproperties for mineral soils. Figure 5 shows the good correlation between the Topp equationand an empirical function derived from TDR measurements of nine mineral soils with claycontents ranging from 2 to 46% (Roth, Malicki, and Plagge 1992).

The bulk density and a large clay or organic-matter content can all have a profoundeffect on the relation between θ and Ka (Roth et al. 1990; Dirksen and Dasberg 1993;Malicki, Plagge, and Roth 1996; Curtis and Narayanan 1998). The increase in specificsurface area and bound water associated with most clays leads to a reduction in the bulkdielectric properties of clay-rich samples. Several authors have therefore presented empir-ical functions similar to Eq. (4) for situations when the Topp equation does not performwell. Lines 3 and 4 in Figure 5 represent alternatives to the Topp equation for a marinesilty-clay soil with 45% illite–kaolinite clay and a Vertisol with 86% smectite clay, respec-tively (Dirksen and Dasberg 1993). Likewise, soils high in organic matter usually havea large specific surface area, so that the Topp equation does not predict the volumetricwater content correctly (Malicki, Plagge, and Roth 1996). Empirical functions derived

Dow

nloa

ded

by [

58.7

.214

.162

] at

22:

33 3

0 Ja

nuar

y 20

14

Tabl

e2

Ove

rvie

wof

empi

rica

lcal

ibra

tion

func

tions

Inpu

taC

alib

rate

dfo

r

Nam

ean

dre

fere

nce

fra

nge

(GH

z)Te

xtur

eO

ther

Out

putb

fra

nge

(GH

z)N

o.of

soils

No.

ofsa

mpl

esSo

ilty

pes

Her

kelr

ath

etal

.(19

91)

Om

––

θK

a∫ T

DR

rang

e1

5O

Sc

Led

ieu

etal

.(19

86)

linea

r–

–θ

,ρb

Ka

∫ TD

Rra

nge

??

Loa

mM

alic

kiet

al.(

1996

)B

D–

–θ

,ρb

Ka

∫ TD

Rra

nge

1834

OSc ,U

SDA

d:S

a,Sa

L,S

aClL

,C

l,C

lL,L

,SiL

,SiC

lL,S

iCl

Pers

son

etal

.(20

02)

AN

N–

Cl,

Si,S

,ρb,O

mK

a∫ T

DR

rang

e5

10U

SDA

§:Sa

,Lsa

,SaL

,SaC

lLR

oth

etal

.(19

92)

Om

––

θK

a∫ T

DR

rang

e9

18O

Sc ,USD

Ad

:Sa,

SaL

,LSa

,Sa

ClL

,Cl,

ClL

,L,S

iL,S

iCl

Topp

etal

.(19

80)

––

θK

a∫ 0.

001–

12

4U

SDA

d:S

aL,C

lL,C

laSy

mbo

lsm

ean

Cl,

clay

%;S

a,sa

nd%

;Si,

silt

%;O

m,o

rgan

icm

atte

r%

;ρb,d

rybu

lkde

nsity

;θ,v

olum

etri

cw

ater

cont

ent.

bSy

mbo

lsm

ean

Ka,a

ppar

entr

elat

ive

perm

ittiv

ity.

c OS,

orga

nic

soil.

dU

SDA

text

ure

clas

sific

atio

n,as

inTa

ble

1.

400

Dow

nloa

ded

by [

58.7

.214

.162

] at

22:

33 3

0 Ja

nuar

y 20

14

Estimating Soil Moisture from GPR 401

0

10

20

30

40

0 0.2 0.4 0.6

App

aren

t die

lect

ric p

erm

ittiv

ity

Volumetric water content (cm3/cm3)

1

2 3

4

5

6

Figure 5. Effects of soil type on empirical models describing the relationship between apparentbulk dielectric permittivity and volumetric water content. Line 1 (black dashed line) represents theTopp-model (Topp et al. 1980). Line 2 represents measurements of mineral soils (Roth et al. 1992).Lines 3 and 4 represent empirical models for a Dutch marine silty-clay soil with 45% illite-kaoliniteclay and a Kenyan vertisol with 86% smectite clay, respectively (Dirksen and Dasberg 1993). Lines5 and 6 are empirical models derived for organic-rich soils by Roth et al. (1992) and Herkelrath et al.(1991), respectively.

from dielectric measurements of samples high in organic-matter content (curves 5 and6 in Figure 5) show that Eq. (4) may overpredict the apparent dielectric permittivity by upto about 50% (Herkelrath, Hamburg, and Murphy 1991; Roth, Malicki, and Plagge 1992).

An alternative means for the empirical determination of the relationship between watercontent and bulk relative permittivity of soil is via artificial neural networks (ANNs),although ANNs do not produce a universal predictive function and may need to be recal-ibrated for each new sample set. Using 10 samples (sand, loamy sand, sandy loam, sandyclay loam) from five different soils in Denmark, Persson et al. (2002) demonstrated thatANNs could improve the accuracy of predicting the relationship between the soil bulkrelative permittivity and soil water content.

Effective Medium Approach

Effective medium theory assumes that natural sample heterogeneity can be representedby a basic geomorphological unit. The unit is an elementary cell filled with water (matrix)and contains a subspherical heterogeneity that corresponds to solids and/or air (Sen, Scala,and Cohen 1981; Friedman 1998; Miyamoto, Annaka, and Chikushi 2005; Cosenza et al.2009). As a consequence of this conceptualization, effective medium theories specificallyaddress inclusion (heterogeneity) shape rather than size. By combining the effects of elec-tric fields in each of the cell components, the effective property of the cell is calculated. Thesimplest effective medium theories (or composite spheres models) are those where the het-erogeneity is contained in a cell filled with the matrix fluid, neglecting any electromagneticinteractions between the two (Maxwell–Garnett rule). Another approach is the Bruggemanrule where the heterogeneity and matrix fluid are symmetrical inclusions in cells filled withthe unknown effective medium (Cosenza et al. 2009). A disadvantage of the composite

Dow

nloa

ded

by [

58.7

.214

.162

] at

22:

33 3

0 Ja

nuar

y 20

14

402 R. L. Van Dam

spheres model is that it is only accurate for known geometries and is difficult to implementfor heterogeneous and multiple-phase materials (Nguyen et al. 1997). Here, calibrationfunctions based on the effective medium theory have been grouped with volumetric mixingmodels (Table 1), as some of their characteristics are similar.

The Maxwell–De Loor model (De Loor 1968) assumes disc-shape inclusions with ran-dom distribution and orientation. This model has been used to describe dielectric propertiesof four-phase mixtures using (Bohl and Roth 1994; Dobson et al. 1985):

εm = εh +3∑

i=1

vi

3(εi − εh)

3∑j=1

(1

1 + Aj εi/εb − 1

). (5)

Here, εh, εi, and εb are the relative permittivity of the host medium (solids), the permittivityof the inclusions, and the effective permittivity near boundaries, respectively; vi representsthe volume fraction of the inclusions; and Aj refers to so-called depolarization ellipsoidfactors.

A different equation is based purely on the depolarization factors of different soil con-stituents (Hilhorst et al. 2000). In this approach the measured relative permittivity is relatedto the volume-weighted sum of the permittivities of the individual material constituents.The depolarization factor (S) is introduced to account for electric-field refractions at thematerial interfaces:

ε =n∑

i=1

εiSivi (6)

Here, vi is the volume fraction of the ith soil constituent. The depolarization factor S isrelated to the electric field refraction in soil, which is in turn a function of the shape andsurface roughness of the grains. Theoretically, the depolarization factor can be calculatedfor all materials. However, in practice this remains difficult for all but homogeneous mate-rials with regularly shaped grains. For randomly oriented spheroids and a frequency abovethat of bound water relaxation effects, the depolarization factors for air and solid matter Sa

and Ss can be approximated by 1 (Hilhorst 1998), while Sw can be estimated by

Sw (θ) = 1

3 (2φ − θ)(7)

where φ is the porosity. A frequency domain sensor was used to measure the dielectricproperties of glass beads and fine sands (Hilhorst and Dirksen 1994; Hilhorst et al. 2000).Both measured values and the calculated θ–ε relationship [using Eq. (6)] show reasonableto good overlap with data presented by Topp, Davis, and Annan (1980).

Frequency-Specific Empirical Functions

Many studies have documented frequency-dependent (complex) dielectric properties ofsoil samples (Table 3), highlighting the difficulty of using one single model to describethe θ–ε relationship for GPR data. As these studies indicate, using different GPR antennafrequencies may require that the frequency-dependent nature of dielectric properties andloss mechanisms be taken into account. Although the Topp equation [Eq. (4)] often per-forms very well for mineral soils over the GPR frequency spectrum, frequency-specific

Dow

nloa

ded

by [

58.7

.214

.162

] at

22:

33 3

0 Ja

nuar

y 20

14

Estimating Soil Moisture from GPR 403

Table 3Studies of frequency-dependent dielectric soil properties

ReferenceFrequency range

(GHz)

No. ofsamples[soils] Soil types

Curtis et al. (1995) 0.45–26.5 30[12]a USDAb: Sa, SaL, Si, SiCl,SiClL, Cl

Dirksen and Dasberg (1993)∫

0.01–1 11 USDAb: SiL, L, Cl, SiClHeimovaara et al. (1994) 0.001–0.15 3 USDAb: SiL, LSa, SiClLKnoll and Knight (1994) 0.0001, 0.001, 0.01 11 Artificial mixtures of sand

and clayNguyen et al. (1997) 1–0.75 1 SandWensink (1993) 0.001–3 11 Clay, silt, peat

aJ.O. Curtis, personal communication.bUSDA texture classification, as in Table 1.

measurements by Curtis (2001) using a network analyzer and a large number of samplesin a coaxial transmission line demonstrate that only at 100 MHz can the Topp equationbe replicated (line 4 in Figure 6). At both lower and higher frequencies the Topp equa-tion loses accuracy. At 50 MHz, measurements of the θ–ε relationship for six soils showa reasonable overlap (line 3 in Figure 6) with the Topp equation (Campbell 1990). Belowaround 50 MHz both ε′ and ε′′ depend strongly on soil type, soil water temperature, andmeasurement frequency. Based on measurements of 6 soils at 1 and 5 MHz, Campbell(1990) shows that Eq. (4) overestimates the volumetric water content at those frequencies

0

10

20

30

40

0 0.2 0.4 0.6

App

aren

t die

lect

ric p

erm

ittiv

ity

Volumetric water content (cm3/cm3)

1

5

1

4

2

5

3

Figure 6. Effects of measurement frequency on the relationship between apparent bulk dielectricpermittivity and volumetric water content. Line 1 is the Topp model (black dashed line). Lines2 and 3 represent polynomial fits to measurements on a broad range of soil types, conducted at1 and 50 MHz, respectively (Campbell 1990). Line 4 represents a polynomial fit to measurementsconducted at 100 MHz (Curtis 2001). Line 5 was derived from Curtis (2001) for measurementsconducted at 1 GHz.

Dow

nloa

ded

by [

58.7

.214

.162

] at

22:

33 3

0 Ja

nuar

y 20

14

404 R. L. Van Dam

(results for measurements at 1 MHz given by line 2 in Figure 6). At frequencies of 1 GHz(line 5 in Figure 6), Eq. (4) underpredicts the volumetric water content (overpredicts therelative permittivity) by up to around 15% (Curtis 2001).

Relaxation-Time Functions. The Cole–Cole model relates characteristic relaxation times tofrequency-dependent behavior of a material and describes the induced polarization effectsof a mixture as a function of frequency, f (Cole and Cole 1941). The complex dielectricpermittivity can be described as

ε∗ (f ) =[ε∞ + εs − ε∞

1 + (jf/

frel)1−β

]− jσdc

2π f ε0(8)

where εs and ε∞ are the static value of the dielectric permittivity and the high-frequencylimit of the real dielectric permittivity, respectively; ε0 is the dielectric permittivity of freespace (8.854·10−12 F/m); frel is the dielectric relaxation frequency of the material; σ dc isthe electrical conductivity; and β is an empirical parameter (0–1) to describe the spreadin relaxation frequencies. For distilled water or other pure liquids with a single relaxationfrequency, β is zero, resulting in the original Debye model (Debye 1929).

An adaptation to the Debye model by Hilhorst et al. (2000) combines Eqs. (6) and (7)with a relationship between soil matric pressure and dielectric relaxation frequency. Wateris bound to the soil matrix by a number of forces, some of which reduce its energy status(Koorevaar, Menelik, and Dirksen 1983). The degree of bonding decreases with distancefrom the particle surface. The soil matric pressure (pm) can be used as a measure for thetightness of bond between the soil particles and the water molecules and can be used todetermine the relaxation frequency of soil water according to (Hilhorst 1998)

frel = frel,0epmVRT (9)

where V is the partial molar volume of water, R is the universal gas constant, and T isabsolute temperature. This equation confirms that the relaxation frequency of bound wateris several orders of magnitude smaller than that of free water under atmospheric pressure(frel,0 ≈ 10 GHz). By substitution in the Debye equation, the permittivity of a water layerwith volume dθ and with matric pressure pm can be written as

ε∗(f ) =[

Sw (θ)εs − ε∞

1 + jf /frel (pm)+ Mf→∞

]dθ (10)

where Mf→∞ accounts for high-frequency electric field discontinuities. As in a compositespheres model, the total permittivity can be found by integrating over all water layers withinthe soil, with soil matric pressures ranging from dry [pm(θ = 0)] to saturated [pm(θ )].The model can in principle be used to calculate complex dielectric soil properties for awide range of frequencies. However, the complex and difficult-to-obtain input variablessuch as the matric pressure and relaxation frequency as well as the empirical nature ofthe depolarization factor S make this model less suitable for routine estimations of soildielectric properties.

Several authors have proposed simpler frequency-specific empirical functions to pre-dict dielectric properties of soils (Table 4). These approaches often use a volumetric mixingmodel as their base and include information of physical background of dielectric behavior.As previously for the empirical calibration functions, these models have been developedand calibrated for a specific set of soils and thus may only be valid for the data that wereused to develop the relationship.

Dow

nloa

ded

by [

58.7

.214

.162

] at

22:

33 3

0 Ja

nuar

y 20

14

Tabl

e4

Ove

rvie

wof

freq

uenc

y-sp

ecifi

cem

piri

calc

alib

ratio

nfu

nctio

ns

Inpu

taC

alib

rate

dfo

r

Nam

ean

dre

fere

nce

fra

nge

(GH

z)Te

xtur

eO

ther

Out

putb

fra

nge

(GH

z)N

o.of

soils

No.

ofsa

mpl

esSo

ilty

pes

Bri

sco

etal

.(19

92)

0.45

,1.2

5,5.

3,9.

3–

θK

a0.

45,1

.25,

5.3,

9.3

33

USD

Ac :S

aL,C

lL,

SiC

lC

ampb

ell(

1990

)0.

001,

0.05

–θ

Ka

0.00

1–0.

056

6Sa

nd,s

ilt,c

lay

Col

ean

dC

ole

(194

1)∞

–ε

s,ε

∞,ε

0,f

rel,

βε

′ ,ε

′′ ,σ

eff

∞–

––

Cur

tis(2

001)

0.1,

0.5,

1–

θK

a0.

1–1

50–7

520

0–25

0d–

Deb

ye(1

929)

∞–

εs,

ε∞

,ε0,f

rel

ε′ ,

ε′′ ,

σef

f∞

––

–D

obso

net

al.(

1985

)1.

4–18

Cl,

Saθ

,ρε

′ ,ε

′′ ,σ

eff

1.4–

185

5U

SDA

c :SaL

,L,S

iL,

SiC

lH

allik

aine

net

al.(

1985

)1.

4–18

(9in

cr.)

Cl,

Saθ

ε′ ,

ε′′

1.4–

185

5U

SDA

c :SaL

,L,S

iL,

SiL

,SiC

lH

ilhor

st(1

998)

0.00

1–10

0v a

,vs

εs,

ε∞

,vfw

,fre

l,p m

,Si

ε′ ,

ε′′

0.02

,FD

sens

or,

and∫ 0.

01–1

711

USD

Ac :S

iL,S

aL,C

l,Si

ClL

,ClL

,SiC

l+pu

resa

nd(1

)an

dcl

ay(3

)Pe

plin

skie

tal.

(199

5)0.

3–1.

3C

l,Sa

θ,ρ

b,ρ

′ ,ε

′′ ,σ

eff

0.3–

1.3

44

Art

ifici

alm

ixtu

res

ofsa

nd,s

ilt,c

lay

aSy

mbo

lsm

ean

Cl,

clay

%;

Sa,s

and

%;ρ

b,d

rybu

lkde

nsity

s,de

nsity

ofso

lids;

θ,v

olum

etri

cw

ater

cont

ent;

v a,v

olum

eai

rfr

actio

nin

soil;

v fw

,vol

ume

free

wat

erfr

actio

nin

soil;

v s,v

olum

eso

lidfr

actio

nin

soil;

εs,

stat

icdi

elec

tric

perm

ittiv

ity;ε

0,d

iele

ctri

cpe

rmitt

ivity

ofva

cuum

;ε∞

,die

lect

ric

perm

ittiv

ityat

infin

itefr

eque

ncy;

p m,

mat

ric

pres

sure

;Si,

depo

lari

zatio

nfa

ctor

ofi-

thso

ilco

mpo

nent

;fre

l,di

elec

tric

rela

xatio

nfr

eque

ncy;

β,s

prea

din

rela

xatio

nfr

eque

ncie

s.bSy

mbo

lsm

ean

ε′ ,

real

part

ofdi

elec

tric

perm

ittiv

ity;ε

”,im

agin

ary

part

ofdi

elec

tric

perm

ittiv

ity;σ

eff,

effe

ctiv

edi

elec

tric

cond

uctiv

ity;K

a,a

ppar

entr

elat

ive

perm

ittiv

ity.

c USD

Ate

xtur

ecl

assi

ficat

ion,

asin

Tabl

e1.

dJ.

O.C

urtis

,per

sona

lcom

mun

icat

ion.

405

Dow

nloa

ded

by [

58.7

.214

.162

] at

22:

33 3

0 Ja

nuar

y 20

14

406 R. L. Van Dam

Above 1.4 GHz. Based on network analyzer measurements of soils over a frequency rangeof 1.4 to 18 GHz, Hallikainen et al. (1985) published a general empirical function thatcalculates the real (ε′) and imaginary (ε′′) parts of the permittivity as a function of clay(C), sand (S), and volumetric water content (θ ):

ε′orε′′ = (a + bS + cC) + (d + eS + fC)θ + (g + hS + iC)θ2 (11)

In this equation the coefficients a to i vary with frequency. For all frequencies, an increasein clay content leads to an increase in the ratio θ over ε′. For three soil types with sand/clayratios of 3.8, 2.3, and 0.1, respectively, a frequency increase from 4 to 18 GHz correlatespositively with the θ–ε′ ratio and inversely with the θ–ε′′ ratio. For dry soils, however,there is virtually no effect of frequency on either ε′ or ε′′ (Hallikainen et al. 1985; Mätzler1998).

Another high-frequency model, by Dobson et al. (1985), was developed from experi-mental observations between 1.4 to 18 GHz on four different soils (Table 4) with a varietyof water contents. This model calculates the real and imaginary parts of the permittivityaccording to

ε′ =[

1 + ρb

ρs

(εα

s − 1)+ θβ′ε′α

fw − θ

]1/α, and (12)

ε′′ =[θβ ′′

ε′′fw

α]1/α

, (13)

where the real and imaginary parts of the free water permittivity are defined as

ε′fw = εw∞ + εw0 − εw∞

1 + (2π f τw)2 , and (14)

ε′′fw = 2π f τw(εw0 − εw∞)

1 + (2π f τw)2 + σeff

2πε0f

ρs − ρb

ρsθ. (15)

In these equations, α = 0.65, which was based on an empirical fit to experimental data,and Tw is the relaxation time of water. Variables β ′ and β ′′ and the dielectric permittivityof solid material were empirically determined as β ′ = 127.48 − 0.519S − 0.152C, β ′′ =1.33797 − 0.603S − 0.166C, and εs = (1.01 + 0.44ρs)2 − 0.062. The effective dielectricconductivity (σ eff) is given as

σeff = −1.645 + 1.939ρb − 2.013S + 1.594C (16)

Using this pedotransfer function, the real (ε′) and imaginary (ε′′) parts of the dielectriccoefficient (ε) of a soil are found using the known fractions of sand (S) and clay (C), thebulk density (ρb), and the particle density (ρs) for different soil water contents.

Below 1.3 GHz. The model of Peplinski, Ulaby, and Dobson (1995) is complementary tothe model described previously and has been developed from experimental observationsbetween 0.3 to 1.3 GHz on a variety of artificial mixtures of sand, silt, clay, and watercontents. It is identical to the higher frequency model (Dobson et al. 1985) except that thereal part of the dielectric permittivity is given by

ε′ = 1.15

[1 + ρb

ρs(εα

s − 1) + θβ′ε′αfw − θ

]1/α

− 0.68 (17)

Dow

nloa

ded

by [

58.7

.214

.162

] at

22:

33 3

0 Ja

nuar

y 20

14

Estimating Soil Moisture from GPR 407

and the conductivity by

σeff = 0.0467 + 0.2204ρb − 0.4111S − 0.6614C (18)

The Dobson–Peplinski models have been successfully tested for different soil types(Miller, Hendrickx, and Borchers 2004) and use inputs that are often readily available insoil databases. The poor overlap between the two models, however, illustrates a significantlevel of uncertainty in the 1.3- to 1.4-GHz range.

Discussion

With the abundance of calibration functions available to correlate soil dielectric propertieswith volumetric water content and other physical material properties, it is hard to choosethe best function for a given situation. Indeed, because of the large number of differentparameters used in these calibration functions, comparing them in a quantitative way is adifficult task. Only the most basic variables θ and ε are part of all functions. The choicefor which calibration function to use depends on (1) the type of input information availableto the user, (2) the frequency of the GPR equipment, and (3) the desired level of detail ofthe output. In the following, these three items are discussed in more detail for the differentfunctions that were presented in the summary tables.

Input

To find the dielectric properties of a soil, some functions will not require more inputthan the volumetric water content [e.g., Eq. (4)], whereas others require input that can berelatively easily acquired using field or laboratory measurements or from soil databases.Common input parameters are the relative volumes and dielectric properties of air, solids,and clay fractions in a soil sample. Other functions require input that is empirical innature, which cannot easily be determined experimentally. In such cases, it is critical that acalibration function is selected that has been tested on relevant sample material and at fre-quencies comparable to the user’s GPR equipment All empirical functions (Table 2) andmost volumetric mixing models (Table 1) require relatively few input parameters. Valuesfor the spread in relaxation frequencies, necessary to describe the behavior of heteroge-neous mixtures in some models (e.g., Debye 1929; Hilhorst 1998), need to be measured foreach specific sample. As a result, the use of such models to describe frequency-dependentdielectric properties of field soils may require recalibration for each sample.

Frequency Range

Because dielectric properties and, thus, the θ–ε relationship, are frequency dependent(Table 3), it is of great importance to consider for what frequency range each calibrationfunction was developed. Frequency is an input variable in only very few models, becausemost were developed for a wide frequency range using TDR instrumentation. Those thatdo include frequency as an input parameter thus enable prediction of frequency-dependentbehavior. Although most of the functions in Tables 1, 2, and 4 give only apparent dielectricpermittivity as output, some do predict both the real and imaginary parts of the permittiv-ity. The relaxation-based models are well suited for describing all frequency-dependentrelaxation phenomena, but their complexity and difficult-to-obtain input variables aredrawbacks.

Dow

nloa

ded

by [

58.7

.214

.162

] at

22:

33 3

0 Ja

nuar

y 20

14

408 R. L. Van Dam

Empirical functions have been mostly developed from controlled laboratory TDR andcoaxial transmission line experiments, the results of which are most accurate for the low100s of MHz. For higher and lower frequencies alternative empirical functions have beenproposed by Hallikainen et al. (1985) and Campbell (1990), respectively. Volumetric mix-ing and empirical models usually do not incorporate frequency-dependent behavior of soildielectric properties. The empirical function by Hallikainen et al. (1985), the “De Loor”volumetric mixing model in Dobson et al. (1985), and frequency-specific calibration func-tions by Dobson et al. (1985) and Peplinski, Ulaby, and Dobson (1995), among a fewothers, are exceptions. The frequency-specific empirical model by Hilhorst (1998) canbe used to calculate complex dielectric soil properties for a wide range of frequencies.However, the complex and difficult-to-obtain input variables make this model less suitablefor routine estimations of soil dielectric properties.

Output

Most empirical functions have been developed for a specific frequency range and oftena limited number of soils (Table 2). Nevertheless, despite these limitations, it has beenshown that if care is taken regarding the type of material that the model has been calibratedfor, these functions can be used with satisfying accuracy (Van Dam et al. 2002; Lunt,Hubbard, and Rubin 2005). The output of volumetric mixing models is limited by theuse of the poorly defined empirical factor α. Both Dobson et al. (1985) and Bohl andRoth (1994) compare a number of mixing models for predicting the relationships betweensoil water content and dielectric soil properties. They concluded that simple three- andfour-phase CRI mixing models were adequate to describe mineral soils. For organic soilsonly four-phase mixing models and the Maxwell–De Loor model provided good results.Frequency-specific calibration functions describe the θ–ε relationship of soils in significantdetail. However, the empirical nature of these models makes them potentially inaccuratefor soil types not used to develop the relationships. Also, these models should be used withcaution when operating at the extremes of their frequency ranges.

Conclusions

This article presented a review of methods for the prediction of dielectric properties of fieldsoils with the goal of providing users of ground-penetrating radar for soil moisture mea-surements with an overview of calibration functions. The methods have been grouped intovolumetric, empirical, effective medium, and frequency-specific approaches. The majorcharacteristics of each model, such as the input and output variables, the range of fre-quencies the model can be used at, and the soil types the models have been calibrated for,have been presented in tables and should provide the reader with information to choosebetween models. This review has shown that among the many different calibration func-tions available, none are suitable for all situations. Many calibration functions are verylimited in their number of input variables, whereas others are too complicated for routineuse. Frequency-specific empirical functions for the prediction of soil dielectric propertiescombine the advantages of empirical and physical models to deal with behavior for whichno good theoretical models are available.

Funding

The author acknowledges U.S. Army Research Office grant (DAAD19-02-1-027).

Dow

nloa

ded

by [

58.7

.214

.162

] at

22:

33 3

0 Ja

nuar

y 20

14

Estimating Soil Moisture from GPR 409

References

Annan, A. P. 2005. GPR methods for hydrogeological studies. In Hydrogeophysics, ed. Y. Rubin andS. S. Hubbard, 185–213. Amsterdam, Netherlands: Springer.

Birchak, J. R., C. G. Gardner, J. E. Hipp, and J. M. Victor. 1974. High dielectric constant microwaveprobes for sensing soil moisture. In Proceedings of the Institute of Electrical and ElectronicsEngineers 62:93–98.

Bohl, H., and K. Roth. 1994. Evaluation of dielectric mixing models to describe the θ (ε) relation. InProceedings of the Symposium and Workshop on Time Domain Reflectometry in Environmental,Infrastructure and Mining Applications (Special Publication SP 19-94), 309–319. Washington,DC: U.S. Bureau of Mines.

Bradford, J. H. 2008. Measuring water content heterogeneity using multifold GPR with reflectiontomography. Vadose Zone Journal 7:184–193.

Brisco, B., T. J. Pultz, R. J. Brown, G. C. Topp, M. A. Hares, and W. D. Zebchuck. 1992. Soilmoisture measurement using portable dielectric probes and time domain reflectometry. WaterResources Research 28:1339–1346.

Brovelli, A., and G. Cassiani. 2008. Effective permittivity of porous media: A critical analysis of thecomplex refractive index model. Geophysical Prospecting 56:715–727.

Campbell, J. E. 1990. Dielectric properties and influence of conductivity in soils at 1 to 50 megahertz.Soil Science Society of America Journal 54:332–341.

Cassiani, G., C. Strobbia, and L. Gallotti. 2004. Vertical radar profiles for characterization of a deepvadose zone. Vadose Zone Journal 3:1093–1105.

Cerny, R. 2009. Time-domain reflectometry method and its application for measuring moisturecontent in porous materials: A review. Measurement 42:329–336.

Chelidze, T. L., and Y. Gueguen. 1999. Electrical spectroscopy of porous rocks: A review, I:Theoretical models. Geophysical Journal International 137:1–15.

Cole, K. S., and R. H. Cole. 1941. Dispersion and adsorption in dielectrics I: Alternating currentcharacteristics. Journal of Chemical Physics 9:341–351.

Conway, B. E., J. O. M. Bockris, and I. A. Ammar. 1951. The dielectric constant of the solution inthe diffuse and Helmholtz double layers at a charged interface in aqueous solution. Transactionsof the Faraday Society 47:756–766.

Cosenza, P., A. Ghorbani, C. Camerlynck, F. Rejiba, R. Guérin, and A. Tabbagh. 2009. Effectivemedium theories for modelling the relationships between electromagnetic properties andhydrological variables in geomaterials: A review. Near Surface Geophysics 7:563–578.

Curtis, J. O. 2001. Moisture effects on the dielectric properties of soils. IEEE Transactions onGeoscience and Remote Sensing 39:125–128.

Curtis, J., and R. Narayanan. 1998. Effects of laboratory procedures on soil electrical propertymeasurements. IEEE Transactions on Instrumentation and Measurement 47:1474–1480.

Curtis, J. O., C. A. Weiss Jr., and J. B. Everett. 1995. Effect of soil composition on complex dielectricproperties. Vicksburg, Miss.: U.S. Army Corps of Engineers.

Davis, J. L., and A. P. Annan. 1989. Ground-penetrating radar for high resolution mapping of soiland rock stratigraphy. Geophysical Prospecting 37:531–551.

De Loor, G. P. 1968. Dielectric properties of heterogeneous mixtures containing water. Journal ofMicrowave Power 3:67–73.

Debye, P. 1929. Polar molecules. New York: Chemical Catalog Company.Di Matteo, A., E. Pettinelli, and E. Slob. 2013. Early-time GPR signal attributes to estimate soil

dielectric permittivity: A theoretical study. IEEE Transactions on Geoscience and RemoteSensing 51:1643–1654.

Dirksen, C., and S. Dasberg. 1993. Improved calibration of time domain reflectometry for soil watercontent measurements. Soil Science Society of America Journal 57:660–667.

Dobson, M. C., F. T. Ulaby, M. T. Hallikainen, and M. A. El-Rayes. 1985. Microwave dielectricbehavior of wet soil, part II: Dielectric mixing models. IEEE Transactions on Geoscience andRemote Sensing GE-23:35–46.

Dow

nloa

ded

by [

58.7

.214

.162

] at

22:

33 3

0 Ja

nuar

y 20

14

410 R. L. Van Dam

Entekhabi, D., G. R. Asrar, A. K. Betts, K. J. Beven, R. L. Bras, C. J. Duffy, T. Dunne, R. D.Koster, D. P. Lettenmaier, D. B. McLaughlin, W. J. Shuttleworth, M. T. van Genuchten, M. Y.Wei, and E. F. Wood. 1999. An agenda for land surface hydrology research and a call forthe second international hydrological decade. Bulletin of the American Meteorological Society80:2043–2058.

Friedman, S. P. 1998. A saturation degree-dependent composite spheres model for describingthe effective dielectric constant of unsaturated porous media. Water Resources Research34:2949–2961.

Galagedara, L. W., G. W. Parkin, J. D. Redman, P. von Bertoldi, and A. L. Endres. 2005. Fieldstudies of the GPR ground wave method for estimating soil water content during irrigation anddrainage. Journal of Hydrology 301:182–197.

Gerhards, H., U. Wollschlager, Q. H. Yu, P. Schiwek, X. C. Pan, and K. Roth. 2008. Continuous andsimultaneous measurement of reflector depth and average soil-water content with multichannelground-penetrating radar. Geophysics 73:J15–J23.

Greaves, R. J., D. P. Lesmes, J. Mo Lee, and M. N. Toksoz. 1996. Velocity variations and watercontent estimated from multi-offset, ground-penetrating radar. Geophysics 61:683–695.

Hallikainen, M. T., F. T. Ulaby, M. C. Dobson, M. A. El-Rayes, and L. Wu. 1985. Microwavedielectric behavior of wet soil, part I: Empirical models and experimental observations. IEEETransactions on Geoscience and Remote Sensing GE-23:25–34.

Hamann, G., J. Tronicke, C. M. Steelman, and A. L. Endres. 2013. Spectral velocity analysis for thedetermination of ground-wave velocities and their uncertainties in multi-offset GPR data. NearSurface Geophysics 11:167–176.

Hasted, J. B. 1973. Aqueous dielectrics. London: Chapman and Hall.Heimovaara, T. J., W. Bouten, and J. M. Verstraten. 1994. Frequency domain analysis of time

domain reflectometry waveforms, 2: A four-component complex dielectric mixing model forsoils. Water Resources Research 30:201–209.

Herkelrath, W. N., S. P. Hamburg, and F. Murphy. 1991. Automatic, real-time monitoring of soilmoisture in a remote field area with time domain reflectometry. Water Resources Research27:857–864.

Hilhorst, M. A. 1998. Dielectric characterisation of soil (PhD thesis, Wageningen AgriculturalUniversity, Wageningen, Netherlands).

Hilhorst, M. A., and C. Dirksen. 1994. Dielectric water content sensors: Time domain ver-sus frequency domain. In Proceedings of the Symposium and Workshop on Time DomainReflectometry in Environmental, Infrastructure and Mining Applications. (Special PublicationSP 19-94), 23–33. Washington, DC: U.S. Bureau of Mines.

Hilhorst, M. A., C. Dirksen, F. W. H. Kampers, and R. A. Feddes. 2000. New dielectric mixtureequation for porous materials based on depolarization factors. Soil Science Society of AmericaJournal 64:1581–1587.

Hubbard, S. S., Y. Rubin, and E. Majer. 1997. Ground-penetrating-radar-assisted saturation andpermeability estimation in bimodal systems. Water Resources Research 33:971–990.

Hubbard, S. S., K. Grote, and Y. Rubin. 2002. Mapping the volumetric soil water content of aCalifornia vineyard using high-frequency GPR ground wave data. Leading Edge 21:552–559.

Huisman, J. A., J. J. J. C. Snepvangers, W. Bouten, and G. B. M. Heuvelink. 2002. Mapping spa-tial variation in surface soil water content: Comparison of ground-penetrating radar and timedomain reflectometry. Journal of Hydrology 269:194–207.

Huisman, J. A., S. S. Hubbard, J. D. Redman, and A. P. Annan. 2003. Measuring soil water contentwith ground penetrating radar: A review. Vadose Zone Journal 2:476–491.

Jacobsen, O. H., and P. Schjonning. 1995. Comparison of TDR calibration functions for soil waterdetermination. In Proceedings of the Symposium Time-Domain Reflectometry Applications inSoil Science, Tjele, Denmark, SP report 11, 25–33.

Jones, S. B., and S. P. Friedman. 2000. Particle shape effects on the effective permittivity ofanisotropic or isotropic media consisting of aligned or randomly oriented ellipsoidal particles.Water Resources Research 36:2821–2833.

Dow

nloa

ded

by [

58.7

.214

.162

] at

22:

33 3

0 Ja

nuar

y 20

14

Estimating Soil Moisture from GPR 411

Jones, S. B., J. M. Wraith, and D. Or. 2002. Time domain reflectometry measurement principles andapplications. Hydrological Processes 16:141–153.

Kelleners, T. J., D. A. Robinson, P. J. Shouse, J. E. Ayers, and T. H. Skaggs. 2005. Frequency depen-dence of the complex permittivity and its impact on dielectric sensor calibration in soils. SoilScience Society of America Journal 69:67–76.

Knoll, M. D., and R. Knight. 1994. Relationships between dielectric and hydrogeologic propertiesof sand–clay mixtures. In Fifth International Conference on Ground Penetrating Radar, 45–61.Waterloo, Ontario: Waterloo Centre for Groundwater Research.

Koorevaar, P., G. Menelik, and C. Dirksen. 1983. Elements of soil physics. Amsterdam, Netherlands:Elsevier.

Kuroda, S., H. Jang, and H. J. Kim. 2009. Time-lapse borehole radar monitoring of an infiltrationexperiment in the vadose zone. Journal of Applied Geophysics 67:361–366.

Landau, L. D., and E. M. Lifshitz. 1960. Electrodynamics of continuous media. New York: Pergamon.Ledieu, J., P. De Ridder, P. De Clerck, and S. Dautrebande. 1986. A method of measuring soil

moisture by time-domain reflectometry. Journal of Hydrology 88:319–328.Logsdon, S. D. 2005. Soil dielectric spectra from vector network analyzer data. Soil Science Society

of America Journal 69:983–989.Lunt, I. A., S. S. Hubbard, and Y. Rubin. 2005. Soil moisture content estimation using ground-

penetrating radar reflection data. Journal of Hydrology 307:254–269.Mahmoudzadeh, M. R. 2013. Off- and on-ground GPR techniques for field-scale soil moisture

mapping. Geoderma 200–201:55–66.Malicki, M. A., R. Plagge, and C. H. Roth. 1996. Improving the calibration of dielectric TDR soil

moisture determination taking into account the solid soil. European Journal of Soil Science47:357–366.

Mangel, A. R., S. M. J. Moysey, J. C. Ryan, and J. A. Tarbutton. 2012. Multi-offset ground-penetrating radar imaging of a lab-scale infiltration test. Hydrology and Earth System Sciences16:4009–4022.

Mätzler, C. 1998. Microwave permittivity of dry sand. IEEE Transactions on Geoscience and RemoteSensing 36:317–319.

Miller, T. W., J. M. H. Hendrickx, and B. Borchers. 2004. Radar detection of buried landmines infield soils. Vadose Zone Journal 3:1116–1127.

Miyamoto, T., T. Annaka, and J. Chikushi. 2005. Extended dual composite sphere model fordetermining dielectric permittivity of andisols. Soil Science Society of America Journal 69:23–29.

Moysey, S. M. J. 2010. Hydrologic trajectories in transient ground-penetrating-radar reflection data.Geophysics 75:WA211–WA219.

Nguyen, B., A. M. Geels, J. Bruining, and E. C. Slob. 1997. Calibration measurements of dielectricproperties of porous media. SPE Journal 4:353–359.

Peplinski, N. R., F. T. Ulaby, and M. C. Dobson. 1995. Dielectric properties of soils inthe 0.3–1.3 GHz range. IEEE Transactions on Geoscience and Remote Sensing 33:803–807.

Persson, M., B. Sivakumar, R. Berndtsson, O. H. Jacobsen, and P. Schjønning. 2002. Predicting thedielectric constant–water relationship using artificial neural networks. Soil Science Society ofAmerica Journal 66:1424–1429.

Pettinelli, E., G. Vannaroni, B. Di Pasquo, E. Mattei, A. Di Matteo, A. De Santis, and P. A. Annan.2007. Correlation between near-surface electromagnetic soil parameters and early-time GPRsignals: An experimental study. Geophysics 72:A25–A28.

Ponizovsky, A. A., S. M. Chudinova, and Y. A. Pachepsky. 1999. Performance of TDR calibrationmodels as affected by soil texture. Journal of Hydrology 218:35–43.

Powers, M. H. 1997. Modeling frequency-dependent GPR. The Leading Edge 16:1657–1662.Redman, J. D., J. L. Davis, L. W. Galagedara, and G. W. Parkin. 2002. Field studies of GPR

air launched surface reflectivity measurements of soil water content. In Ninth InternationalConference on Ground Penetrating Radar, 156–161. Santa Barbara, CA: SPIE.

Dow

nloa

ded

by [

58.7

.214

.162

] at

22:

33 3

0 Ja

nuar

y 20

14

412 R. L. Van Dam

Robinson, D. A., S. B. Jones, J. M. Wraith, D. Or, and S. P. Friedman. 2003. A review of advancesin dielectric and electrical conductivity measurement in soils using time domain reflectometry.Vadose Zone Journal 2:444–475.

Roth, C. H., M. A. Malicki, and R. Plagge. 1992. Empirical evaluation of the relationship betweensoil dielectric constant and volumetric water content as the basis for calibrating soil moisturemeasurements by TDR. Journal of Soil Science 43:1–13.

Roth, K., R. Schulin, H. Flühler, and W. Attinger. 1990. Calibration of time domain reflectometry forwater content measurement using a composite dielectric approach. Water Resources Research26:2267–2273.

Rubin, Y., and S. S. Hubbard. 2005. Hydrogeophysics. Amsterdam, Netherlands: Springer.Santamarina, J. C., and M. Fam. 1997. Dielectric permittivity of soils mixed with organic and inor-

ganic fluids (0.02 GHz to 1.30 GHz). Journal of Environmental and Engineering Geophysics2:37–52.

Schmelzbach, C., J. Tronicke, and P. Dietrich. 2012. High-resolution water content estimationfrom surface-based ground-penetrating radar reflection data by impedance inversion. WaterResources Research 48:W08505.

Sen, P. N. 1984. Grain shape effects on dielectric and electrical properties of rocks. Geophysics49:586–587.

Sen, P. N., C. Scala, and M. H. Cohen. 1981. A self-similar model for sedimentaryrocks with application to the dielectric constant of fused glass beads. Geophysics 46:781–795.

Topp, G. C., and J. L. Davis. 1985. Measurement of soil water content using time-domainreflectometry (TDR): A field evaluation. Soil Science Society of America Journal 49:19–24.

Topp, G. C., J. L. Davis, and A. P. Annan. 1980. Electromagnetic determination of soilwater content: Measurements in coaxial transmission lines. Water Resources Research 16:574–582.

Turesson, A. 2006. Water content and porosity estimated from ground-penetrating radar andresistivity. Journal of Applied Geophysics 58:99–111.

Van Dam, R. L., and W. Schlager. 2000. Identifying causes of ground-penetrating radar reflec-tions using time-domain reflectometry and sedimentological analyses. Sedimentology 47:435–449.

Van Dam, R. L., E. H. Van Den Berg, S. Van Heteren, C. Kasse, J. A. M. Kenter, and K. Groen. 2002.Influence of organic matter in soils on radar-wave reflection: Sedimentological implications.Journal of Sedimentary Research 72:341–352.

Van Dam, R. L., B. Borchers, and J. M. H. Hendrickx. 2005. Strength of landmine signaturesunder different soil conditions: Implications for sensor fusion. International Journal of SystemsScience 36:573–588.

Van Dam, R. L., J. M. H. Hendrickx, N. J. Cassidy, R. E. North, M. Dogan, and B. Borchers. 2013.Effects of magnetite on high-frequency ground-penetrating radar. Geophysics 78 (5):H1–H11.

Vörösmarty, C. J., P. Green, J. Salisbury, and R. B. Lammers. 2000. Global water resources:Vulnerability from climate change and population growth. Science 289:284–288.

Wang, J. R., and T. J. Schmugge. 1980. An empirical model for the complex dielectric permittivityof soils as a function of water content. IEEE Transactions on Geoscience and Remote SensingGE-18:288–295.

Weihermueller, L., J. A. Huisman, S. Lambot, M. Herbst, and H. Vereecken. 2007. Mapping thespatial variation of soil water content at the field scale with different ground penetrating radartechniques. Journal of Hydrology 340:205–216.

Wensink, W. A. 1993. Dielectric properties of wet soils in the frequency range 1–3000 MHz.Geophysical Prospecting 41:671–696.

Wilson, D. J., A. W. Western, R. B. Grayson, A. A. Berg, M. S. Lear, M. Rodell, J. S.Famiglietti, R. A. Woods, and T. A. McMahon. 2003. Spatial distribution of soil moisture over6 and 30 cm depth, Mahurangi River Catchment, New Zealand. Journal of Hydrology 276:254–274.

Dow

nloa

ded

by [

58.7

.214

.162

] at

22:

33 3

0 Ja

nuar

y 20

14

Estimating Soil Moisture from GPR 413

Zakri, T., J. Laurent, and M. Vauclin. 1998. Theoretical evidence for “Lichtenecker’s mixtureformulae” based on the effective medium theory. Journal of Physics D: Applied Physics31:1589–1594.

Zegelin, S. J., and I. White. 1994. Calibration of TDR for applications in mining, grains andfruit storage and handling. In Proceedings of the Symposium and Workshop on Time DomainReflectometry in Environmental, Infrastructure and Mining Applications (Special PublicationSP 19-94), 115–129. Washington, D.C.: U.S. Bureau of Mines.

Dow

nloa

ded

by [

58.7

.214

.162

] at

22:

33 3

0 Ja

nuar

y 20

14