12
 ELECTRIC PROPERTIES OF TISSUES D  AMIJAN MIKLAVC ˇ IC ˇ N  ATAS ˇ  A  P  A VS ˇ ELJ University of Ljubljana Ljubljana, Slovenia FRANCIS X. H  ART University of the South Sewanee, Tennessee 1. INTRODUCTION The electrical properties of biological tissues and cell sus- pensions have been of interest for over a century for many reasons. The y dete rmi ne the path ways of current ow thr ough the body and, thu s, are very imp ortant in the analysis of a wide range of biomedical applications such as func tional electric al stimulation and the diagno sis and treatment of various physiological conditions with weak electric currents, radio-frequency hyperthermia, electro- cardi ograp hy , and body compo sition. On a more funda- mental level, knowledge of these electrical properties can lead to an understanding of the underlying basic biological processes. Indeed, biological impedance studies have long been important in electrophysiology and biophysics; one of the rst demonstrations of the existence of the cell mem- brane was based on dielectric studies on cell suspensions (1). To analyze the response of a tissue to electric stimula- tion, we need data on the specic conductivities and rel- ative permittivities of the tissues or organs. A microscopic description of the response is complicated by the variety of cell shapes and their distrib ution inside the tissue as well as the dif fer ent pro per ties of the ext rac ell ula r med ia. Therefore, a macroscopic approach is most often used to cha racter ize el d dis tri but ions in bio logica l sys tems. Moreover, even on a macroscopic level, the electrical prop- ertie s are complic ated. They can depend on the tissue ori- en tati on re lati ve to the appl ie d eld (dir ec ti on al anisotropy), the frequency of the applied eld (the tissue is neither a perfect dielectric nor a perfect conductor), or they can be time- and space-dependent (e.g., changes in tissue conductivity during electropermeabilization). 2. BIOLOGICAL MATER IALS IN AN ELECT RIC FIELD The electrical properties of any material, including bio- logical tissue, can be broadly separated into two catego- ri es: conduc ti ng and insulat ing. In a conduc tor , the electric charges move freely in response to the application of an electric eld, whereas in an insulator (dielectric), the charges are xed and not free to move. A more detailed discu ssion of the funda mental processes underly ing the electrical properties of tissue can be found in Foster and Schwan (2). If a conductor is placed in an electric eld, charges will move within the conductor until the interior eld is zero. In the case of an insulator, no free charges exist, so net migr ation of charg e does not occur. In polar materi als, however, the positive and negative charge centers in the molecules do not coincide. An electric dipole moment,  p, is said to exist. An applied eld,  E 0 , tends to orient the di- poles and produces a eld inside the dielectric,  E  p , which opposes the applied eld. This process is called polariza- tion. Most materials contain a combination of orientable dipoles and relatively free charges so that the electric eld is reduced in any material relati ve to its free-spac e value. The net eld inside the material,  E, is then  E ¼ E 0   E  p :  ð1Þ The net eld is lowered by a significant amount relative to the applied eld if the material is an insulator and is es- sentially zero for a good conductor. This reduction is char- acteri zed by a fac tor  e r , whic h is call ed the relative permittivity or dielectric constant, according to  E ¼  E 0 e r :  ð2Þ In pract ice, most materials, incl uding biologica l tissue, actual ly displ ay some character istics of both insulators and conductors bec ause they contain dipoles as wel l as charges that can move, but in a restricted manner. For materi als that are hetero geneo us in struct ure, charges may become trapped at interfaces. As positive and nega- tive ions move in opposite directi ons under the appli ed eld, internal charge separations can then result within the material, producing an effective internal polarization that acts like a very large dipole. On a mac ros copic lev el, we describe the mat eri al as having a permittivity,  e, and a conductivity,  s. The per- mitti vity characteri zes the material’s abilit y to trap or store charge or to rotate molecular dipoles, whereas the conductivity describes its ability to transport charge (3). The permittivity also helps to determine the speed of light in a materi al so that fr ee spa ce has a pe rmit ti vi ty e 0  ¼ 8:85 10 12 F=m. For other media, e ¼ e r e 0 :  ð3Þ The energy stored per unit volume in a material,  u, is u ¼  e  E 2 2  ;  ð4Þ and the power dissipated per unit volume,  p, is  p ¼  s  E 2 2  :  ð5Þ We can rep resent these tendencies by usi ng a cir cui t model to describe the tissue (1,4). Consider a sample of material that has a thickness,  d, and cross-sectional area,  A. If the material is an insulator, then we treat the sample as a capacitor with capacitance C ¼ e   A=d:  ð6Þ 1 Wiley Encyclopedia of Biomedical Engineering,  Copyright  & 2006 John Wiley & Sons, Inc.

Electric Properties

Embed Size (px)

Citation preview

  • ELECTRIC PROPERTIES OF TISSUES

    DAMIJAN MIKLAVCICNATASA PAVSELJUniversity of LjubljanaLjubljana, Slovenia

    FRANCIS X. HARTUniversity of the SouthSewanee, Tennessee

    1. INTRODUCTION

    The electrical properties of biological tissues and cell sus-pensions have been of interest for over a century for manyreasons. They determine the pathways of current flowthrough the body and, thus, are very important in theanalysis of a wide range of biomedical applications such asfunctional electrical stimulation and the diagnosis andtreatment of various physiological conditions with weakelectric currents, radio-frequency hyperthermia, electro-cardiography, and body composition. On a more funda-mental level, knowledge of these electrical properties canlead to an understanding of the underlying basic biologicalprocesses. Indeed, biological impedance studies have longbeen important in electrophysiology and biophysics; one ofthe first demonstrations of the existence of the cell mem-brane was based on dielectric studies on cell suspensions(1).

    To analyze the response of a tissue to electric stimula-tion, we need data on the specific conductivities and rel-ative permittivities of the tissues or organs. A microscopicdescription of the response is complicated by the variety ofcell shapes and their distribution inside the tissue as wellas the different properties of the extracellular media.Therefore, a macroscopic approach is most often used tocharacterize field distributions in biological systems.Moreover, even on a macroscopic level, the electrical prop-erties are complicated. They can depend on the tissue ori-entation relative to the applied field (directionalanisotropy), the frequency of the applied field (the tissueis neither a perfect dielectric nor a perfect conductor), orthey can be time- and space-dependent (e.g., changes intissue conductivity during electropermeabilization).

    2. BIOLOGICAL MATERIALS IN AN ELECTRIC FIELD

    The electrical properties of any material, including bio-logical tissue, can be broadly separated into two catego-ries: conducting and insulating. In a conductor, theelectric charges move freely in response to the applicationof an electric field, whereas in an insulator (dielectric), thecharges are fixed and not free to move. A more detaileddiscussion of the fundamental processes underlying theelectrical properties of tissue can be found in Foster andSchwan (2).

    If a conductor is placed in an electric field, charges willmove within the conductor until the interior field is zero.In the case of an insulator, no free charges exist, so netmigration of charge does not occur. In polar materials,

    however, the positive and negative charge centers in themolecules do not coincide. An electric dipole moment, p, issaid to exist. An applied field, E0, tends to orient the di-poles and produces a field inside the dielectric, Ep, whichopposes the applied field. This process is called polariza-tion. Most materials contain a combination of orientabledipoles and relatively free charges so that the electric fieldis reduced in any material relative to its free-space value.The net field inside the material, E, is then

    EE0 Ep: 1

    The net field is lowered by a significant amount relative tothe applied field if the material is an insulator and is es-sentially zero for a good conductor. This reduction is char-acterized by a factor er, which is called the relativepermittivity or dielectric constant, according to

    E E0er

    : 2

    In practice, most materials, including biological tissue,actually display some characteristics of both insulatorsand conductors because they contain dipoles as well ascharges that can move, but in a restricted manner. Formaterials that are heterogeneous in structure, chargesmay become trapped at interfaces. As positive and nega-tive ions move in opposite directions under the appliedfield, internal charge separations can then result withinthe material, producing an effective internal polarizationthat acts like a very large dipole.

    On a macroscopic level, we describe the material ashaving a permittivity, e, and a conductivity, s. The per-mittivity characterizes the materials ability to trap orstore charge or to rotate molecular dipoles, whereas theconductivity describes its ability to transport charge (3).The permittivity also helps to determine the speed of lightin a material so that free space has a permittivitye08:85 1012 F=m. For other media,

    e ere0: 3

    The energy stored per unit volume in a material, u, is

    u eE2

    2; 4

    and the power dissipated per unit volume, p, is

    p sE2

    2: 5

    We can represent these tendencies by using a circuitmodel to describe the tissue (1,4). Consider a sample ofmaterial that has a thickness, d, and cross-sectional area,A. If the material is an insulator, then we treat the sampleas a capacitor with capacitance

    C e A=d: 61

    Wiley Encyclopedia of Biomedical Engineering, Copyright & 2006 John Wiley & Sons, Inc.

  • If it is a conductor, then we treat it as a conductor withconductance

    G s A=d: 7

    A simple model for a real material, such as tissue, wouldbe a parallel combination of the capacitor and conductor.Such a model is referred to as Debye-type. Other, morecomplicated models are sometimes used, as will be de-scribed later. If a constant (DC) voltage V is applied acrossthis parallel combination, then a conduction currentICGV will flow and an amount of charge QCV willbe stored.

    Suppose, instead, that an alternating (AC) voltage wasapplied to the combination:

    VtV0 cosot: 8

    Here, V0 is the amplitude of the voltage and o 2pf ,where f is the frequency of the applied signal. The chargeon the capacitor plates now is changing with frequency f.This change is associated with a flow of charge or currentin the circuit. We characterize this flow as a displacementcurrent:

    IddQ=dt oCV0 sinot: 9

    The total current flowing through the material is the sumof the conduction and displacement currents, which are 90degrees apart in phase because of the difference in thetrigonometric functions. This phase difference can be ex-pressed conveniently by writing

    VtV0eiot; where i1

    p10

    and taking its real part for physical significance. The totalcurrent is I Ic Id, hence

    IGV C dV=dt s ioeA V=d: 11

    The actual material, then, can be characterized as havingan admittance, Y, given by

    Y G ioC A=ds ioe; 12

    where indicates a complex-valued quantity. In terms ofmaterial properties, we define a corresponding, complex-valued conductivity

    s s ioe: 13

    Describing a material in terms of its admittance empha-sizes its ability to transport current. Alternatively, wecould emphasize its ability to restrict the flow of currentby considering its impedance, Z 1=Y, or, for a pureconductance, its resistance, R 1=G.

    Factoring ioe0 in Equation 11 yields

    I er is=oe0ioe0A=dCdV

    dt: 14

    We can define a complex-valued, relative permittivity

    e er isoe0

    e0r ie00r ; 15

    with e0r er and e00r s=oe0. The complex conductivity andcomplex permittivity are related by

    s ioe ioe0er : 16

    In physical terms, we can regard the conductivity of amaterial as a measure of the ability of its charge to betransported throughout its volume by an applied electricfield. Similarly, its permittivity is a measure of the abilityof its dipoles to rotate or its charge to be stored by an ap-plied external field. Note that if the permittivity and con-ductivity of the material are constant, the displacementcurrent will increase with frequency whereas the conduc-tion current does not change. At low frequencies, the ma-terial will behave like a conductor, but capacitive effectswill become more important at higher frequencies. Formost materials, however, these material properties are notconstant, but vary with the frequency of the applied sig-nal. s and e are frequency-dependent. Such a variation iscalled dispersion. Biological tissues exhibit several differ-ent dispersions over a wide range of frequencies.

    Dispersions can be understood in terms of the orienta-tion of the dipoles and the motion of the charge carriers. Atrelatively low frequencies, it is relatively easy for the di-poles to orient in response to the change in the appliedfield, whereas the charge carriers travel larger distancesover which a greater opportunity exists for trapping at adefect or interface. The permittivity is relatively high andthe conductivity is relatively low. As the frequency in-creases, the dipoles are less able to follow the changes inthe applied field, and the corresponding polarization dis-appears. In contrast, the charge carriers sample shorterdistances during each half-cycle and are less likely to betrapped. As frequency increases, the permittivity de-creases and, because trapping becomes less important,the conductivity increases (4,5). The dispersion can becharacterized by an angular relaxation frequencyor 2pfr or, equivalently, by a relaxation time Tr 1=fr.

    In a heterogeneous material, such as biological tissue,several dispersions are observed as illustrated in Fig. 1(4), which shows the variation with frequency of the com-plex permittivity of Equation 15. For frequencies belowabout 10kHz, the a dispersion is caused by counterion po-larization along cell membranes. The extremely high val-ues of permittivity reflect the trapping of charges atinternal interfaces and are not related to dipole orienta-tion. Note that even at the lowest frequencies a residual orDC conductivity s0 exists. The dispersion in the MHz fre-quency range originates in interfacial polarization of cellmembranes, which act as a barrier for passive ion trans-port between the inner and the outer cell media. Basically,the membrane can be modeled as a parallel combination ofa capacitor and a resistor. This so-called beta dispersionoccurs in the frequency range where the reactance of themembrane capacitance short-circuits the membrane re-

    2 ELECTRIC PROPERTIES OF TISSUES

  • sistance, so that the external electric field begins to pen-etrate into the cell interior. Additional contributions tothis b dispersion can develop because of the polarization ofproteins and other organic macromolecules. In the GHzrange (109Hz), g dispersion is caused by the polarization ofwater molecules.

    The relative importance of the permittivity and con-ductivity in determining the electrical properties of thetissue can be compared by taking the ratio of the displace-ment and conduction currents; Id=Icoe=s. For frequen-cies below the MHz range, this ratio is very low, even withthe large increase in permittivity of a dispersion. Hence, atlow frequencies, biological tissue is essentially conductivein nature.

    For a Debye-type response, which corresponds to par-allel RC elements, dispersion can be represented as

    er e1eS e11 iot is0=oe0 17

    and

    s s1s0 s11 iot ; 18

    where the time constant, t 1=RC. e1 and eS refer, re-spectively, to the relative permittivities at frequencies wellabove and well below the dispersion. s1 and s0 refer, re-spectively, to the conductivities at frequencies well aboveand well below the dispersion.

    The complexity of the dispersions illustrated in Fig. 1,however, cannot be readily described by three successive,simple Debye relaxations. The widths of the dispersions,in particular, are greater than predicted for simple RCparallel elements. A similar situation occurs for most ma-terials. Therefore, in place of a simple RC element, a moregeneral, empirical relation, the ColeCole response, isused in which

    er e1eS e11 iota is0=oe0 19

    and

    s s1 s0 s11 iota ; 20

    where a is a parameter that depends on the nature of thematerial. a is equal to 1 for a Debye-type dispersion andbecomes smaller as the width of the dispersion increases.

    Two physical interpretations exist for a factor. Someresearchers regard a wide dispersion as an indication ofnumerous Debye-type dispersions with a distribution ofsimple relaxation times. Other researchers regard thespread as an indication that the fundamental charge-transport and dipole-reorientation processes are essen-tially cooperative in nature and that, as the degree of co-operation increases, a becomes smaller than 1.

    The representation of the dispersion in a circuit modelcan then be achieved in two ways. First, several Debye-type RC elements connected in series could be used torepresent the single, broad dispersion. This process is un-wieldy. Second, and more commonly, the circuit model isgeneralized by the introduction of a Constant Phase El-ement (CPE) with a complex-valued impedance given by

    ZCPEAion; 21

    where A is a parameter and n a. This CPE impedancereduces to a simple resistance for n 0 and to a capacitivereactance for n 1. For a diffusive process, n0:5. TheColeCole dispersion can be represented in circuit termsas the parallel combination of a resistor and a CPE. Otherdispersions, such as ColeDavidson and HavrilakNe-gami, differ somewhat from Equations 19 and 20 and areused for many materials. However, the ColeCole model isgenerally used for biological materials.

    3. COMPLICATIONS IN DIELECTRIC MEASUREMENTS OFTISSUES

    The measurement of tissue dielectric properties can becomplicated because of several factors, such as tissue in-homogeneity, anisotropy, the physiological state of the tis-

    Figure 1. Typical frequency dependence of the com-plex permittivity of a heterogeneous material suchas biological tissues (4).

    ELECTRIC PROPERTIES OF TISSUES 3

  • sue, and electrode polarization. Therefore, caution mustbe used in the design of the measurement procedure.

    3.1. Inhomogeneity of Tissues

    Tissue is a very inhomogeneous material. The cell itself iscomprised of an insulating membrane enclosing a conduc-tive cytosol. A suspension of cells can be regarded at lowfrequencies simply as insulating inclusions in a conduct-ing fluid. The insulation is provided by the cell membrane.At frequencies in the MHz range, capacitive couplingacross this membrane becomes more important. Begin-ning in this range, the dispersive properties of the mem-brane and ultimately the cytosol must also be considered.For a thorough discussion of how the dielectric propertiesof cell suspensions vary with frequency, see Kotnik andMiklavcic (6,7) and Pavlin et al. (8).

    In tissue, the cells are surrounded by an extracellularmatrix, which can be extensive, as in the case of bone, orminimal, as in the case of epithelial tissue. Tissue does notcontain cells of a single size and function. For example,bone contains osteoblasts, osteocytes, and osteoclasts em-bedded in a collagen/hydroxyapatite matrix as well asbone marrow with stroma cells (9). The tissue is perfusedwith blood and linked to the central nervous system byneurons. It is thus difficult to extrapolate from the dielec-tric properties of a cell suspension to those of an intacttissue.

    A large discrepancy exists between various data onelectrical properties of biological materials found in theliterature. Why is there such a wide range of values ob-tained by different researchers? Excised samples carryalong with them various amounts of body fluids, and thelack of standardization of measurement techniques pre-sents its own difficulties and probably widens the range ofresistivity values. Moreover, there are seasonal, age, anddisease-linked changes as well as those that accompanythe physiological function of various biologic materials(10).

    3.2. Anisotropy of Tissues

    Some biological materials, such as bone and skeletal mus-cle, are distinctly anisotropic. Therefore, when referring topublished conductivity and permittivity values, we need tocheck the orientation of the electrodes relative to the ma-jor axis of the tissue (e.g., longitudinal, transversal, or acombination of both).

    Electrical anisotropy is often related to the physiolog-ical demands made on the tissue. Major bones and musclesof limbs are designed to produce and support significantlongitudinal forces. For example, muscles are composed offibers that are very large individual cells and are alignedin the direction of muscle contraction. Electrical conduc-tion along the length of the fiber is thus significantly eas-ier than conduction between the fibers in the extracellularmatrix because the extracellular matrix is less conductivethan the cell. Therefore, muscle tissue manifests typicalanisotropic electric properties (4). The longitudinal con-ductivity is significantly higher than the transverse con-ductivity even when path differences in the chargetransport are taken into account, especially in the low-

    frequency range (11). A similar anisotropy exists in thelong bones of the body where charge transport is easieralong the longitudinal axis than transverse to it.

    Moreover, tissue anisotropy is frequency-dependent.Namely, if the frequency of the current is high enough,the anisotropic properties disappear (specifically for mus-cle tissue, that happens in the MHz frequency range). Athigher frequencies, charge movement takes place overshorter distances so large-scale structures become lessimportant and capacitive coupling across membranes be-comes more important.

    A practical problem occurs when measuring the elec-trical properties of anisotropic materials: how to accu-rately align the applied electric field and tissue fibers (12).Namely, it has been shown that perfect alignment is cru-cial for obtaining accurate longitudinal and transversevalues. A study on skeletal muscle tissue shows (13) that a5 degree misalignment from true perpendicular or parallelorientations would result in an 18% overestimate in theperpendicular direction and a 0.4% underestimate in theparallel direction when measuring specific conductivity.

    3.3. Physiological Factors and Changes of Tissue

    Any changes in tissue physiology should produce changesin the tissue electrical properties (14). This principle hasbeen used to identify or monitor the presence of variousillnesses or conditions such as body fluid shift, blood flow,cardiac output, and muscular dystrophy (15) by variousimpedance diagnostic techniques, such as impedanceplethysmography, rheoencephalography, and thoracic im-pedance cardiograph. For a detailed discussion of the ap-plications of bioimpedance methods in medicine andbiotechnology, see (16,17).

    Tumors generally have higher water content than nor-mal cells because of cellular necrosis but also irregularand fenestrated vascularization. In addition, differencesmay exist in the membrane structure. Although an in-creased conductivity may be used to identify the presenceof tumors (18), parameters associated with the fitting ofthe overall dielectric spectrum to a circuit model may bemore reliable (19). In clinical practice, the presence of skinmay complicate the interpretation of impedance changesin tissue, such as the breast (20). The higher conductivityof tumors in the MHz frequency range could lead to theirselective targeting by radio-frequency hyperthermiatreatment (21).

    Fat is a poorer conductor of electricity than water.Changes in the percentage of body fat or water are re-flected in tissue impedance changes. For example, Biggs etal. (22) estimated the whole-body fat percentage by mea-surement of the resistivity of the upper arm and leg at50kHz. Van Kreel et al. (23) determined the total bodywater content by measurement of body impedance at sev-eral frequencies. This method can even be applied to in-dividual organs. For example, Schaefer et al. (24)correlated the complex permittivity of heart tissue withthe level of ischemia.

    In an extreme case, one can imagine that tissue deathor excision would result in significant changes in electricalproperties. Tissue metabolism decreases after the tissue

    4 ELECTRIC PROPERTIES OF TISSUES

  • has been excised and, often, the temperature falls. If thetissue is supported by temperature maintenance and per-fusion systems, the tissue may be stabilized for a limitedperiod of time in a living state in vitro (ex vivo). If the tis-sue is not supported, however, irreversible changes willoccur, followed by cell and tissue death (3). If the bloodflow is interrupted, metabolism continues, but in an an-aerobic way. Osmosis will cause cell swelling and tissuedamage. As a consequence, the extracellular pathwaysnarrow, which typically leads to an increase in the low-frequency impedance (o10kHz). The time of occurrence ofthese phenomena is different for different tissues. De-creased blood flow also accounts for changes in tissue re-sistivity, because blood is a good conductor (12).

    Conductivity changes caused by cell and tissue deathhave been studied by different researchers (10,12,25). Theresults show that in the first hour after the tissue samplehas been excised, the specific conductivity is almost con-stant, although the change depends on the tissue type.Liver tissue shows changes after only 30minutes from ex-cision, brain tissue after one hour, and the muscle tissuetwo hours after excision. In all cases, the conductivity in-creases with time (10). Changes in the frequency rangeabove 100Hz are lower and take a longer time to occur.However, because tissue impedance at low frequencies isalmost entirely ohmic, permittivity errors do not play anymajor role. For these reasons, considerable caution mustbe taken in the interpretation of electrical measurementsthat were performed on excised tissues.

    The electrical properties of tissue also depend on itstemperature. The mobility of the ions that transport thecurrent increases with the temperature as the viscosity ofthe extracellular fluid decreases. A general increase ofabout 2%/1C occurs in the conductivity of tissue (2) in thefrequency range below 1GHz, up to a temperature ofabout 401C. Above that point, the cell membrane beginsto deteriorate and allows the cytosol to leak into the ex-tracellular space. The rapid increase of conductivity withtemperature was suggested to be used to monitor the pro-gress of hyperthermia treatment (26).

    3.4. Electrode Polarization

    The measurement of tissue electrical properties, in vivo, iscomplicated (12). Two main sources of systematic errorexist, electrode polarization and lead inductance, whichbecome apparent at the lower and higher ends of the fre-quency range, respectively (27). Electrode polarization is amanifestation of molecular charge organization that oc-curs at the sample-electrode interface in the presence ofwater molecules and hydrated ions. In its simplest form,the phenomenon can be modeled as a frequency-depen-dent capacitor in series with a resistor. The effect in-creases with increasing sample conductivity, and itsconsequences are more pronounced on the capacitancethan the conductance of ionic solutions as well as biolog-ical samples.

    In a cell suspension, a counterion layer can form ateach electrode. The potential drop in this layer reduces theelectric field available to drive charge transport in thebulk suspension, resulting in apparently low suspension

    conductivity. As the frequency increases, the counterionlayer is less able to follow the changes in the applied sig-nal, the potential drop at the suspension/electrode inter-face decreases, and the apparent conductivity of thesuspension increases. The nature of the ions in the layeris determined by both the suspension and the material ofthe electrode. Hence, changing either the electrode mate-rial or the nature of the suspension will modify the mag-nitude and frequency response of this electrodepolarization.

    The process is more complicated in tissue. Insertion ofelectrodes can first cause the release of electrolytes fromthe surrounding tissue and, later, the development of apoorly conductive wound region may occur. This regioncan shield part of the electrode from the ionic current andthus reduce the polarization effects compared with anionic solution equivalent in conductivity to the intracellu-lar fluid. As with a cell suspension, the material of theelectrode plays an important part in determining its po-larization impedance, the relative importance of whichdecreases with increasing frequency. It is good practice tomeasure tissue impedance in vivo after waiting a suffi-cient time for the electrode polarization processes to sta-bilize. The waiting time will depend, in general, on thenature of the electrodes and the type of tissue. If possible,one should perform a preliminary experiment in whichimpedance spectra are taken every few minutes until nofurther changes occur in order to determine the waitingtime. A typical time might be on the order of 30minutes.

    Two different electrode setups are used to measure theelectric properties of biological materials; the two-elec-trode method and the four-electrode method (28).

    3.4.1. Two-electrode Method. This method is suitablefor alternating current measurements. We can not use itas such for direct current measurements because of theelectrode polarization that consequently gives incorrectresults for the conductivity of the sample between theelectrodes. For alternating current measurements, thefrequency range over which electrode polarization is im-portant depends to some extent on the system being mea-sured and the electrode material. For cell suspensions, it isimportant up to nearly 100 kHz, whereas for tissue mea-sured in vivo, it is significant only up to about 1 kHz. Byvarying the separation of the electrodes (29), the contri-bution of the electrode polarization can be determined andeliminated.

    3.4.2. Four-electrode Method. This method can be usedfor direct and alternating current measurements. Twopairs of electrodes are used: the outer (current) electrodesand the inner (voltage) electrodes. The current from thesource passes through the sample. Voltage electrodes ofknown separation are placed in the sample between thecurrent electrodes. By measuring the current as the volt-age drop across a resistor in series with the sample andthe voltage drop across the inner electrodes, one can de-termine the specific conductance of the sample betweenthe inner electrodes. The advantage of this method is thatthe polarization on the current electrodes has no influenceon the voltage difference between the voltage electrodes.

    ELECTRIC PROPERTIES OF TISSUES 5

  • Polarization at the voltage electrodes is negligible for di-rect and alternating currents because of the high inputimpedance of the measurement system. Some authors rec-ommend the two-electrode system with a polarization er-ror correction (12). They believe the four-electrode methoddoes not completely eliminate the polarization on voltageelectrodes. Some authors report 20% higher results whenmeasuring muscle resistance with the two-electrode tech-nique without error correction (30).

    With direct current measurements, care must be takento use small currents to avoid electrochemical injection ofions at the electrodes and nerve stimulation if the mea-surements are conducted in vivo.

    In the MHz frequency range and higher, lead induc-tance becomes an important factor as the inductive cou-plings between the leads to each other, to the measuringinstruments, and to nearby metal surfaces produce extra-neous voltage drops that increase with frequency. It isgood practice to identify and then reduce the effects ofelectrode polarization at low frequencies and lead induc-tance at high frequencies by replacing the sample to bemeasured by a saline solution for which the electricalproperties should not vary with frequency until the GHzfrequency range.

    4. DIELECTRIC PROPERTIES OF SOME TISSUES

    Large differences exist in electric properties of biologicalmaterials. These differences are determined, to a largeextent, by the fluid content of the material. For example,blood and brain conduct electric current relatively well.Lungs, skin, fat, and bone are relatively poor conductors.Liver, spleen, and muscle are intermediate in their con-ductivities. An excellent, detailed review of the electricalproperties of various tissues over a wide range of frequen-cies can be found in the series of three articles by Gabrielet al. (27,31,32). Of particular interest is their use of ColeCole response function, given by Equations 19 and 20, toparameterize the dielectric properties of tissue. Such pa-rameterization allows one to calculate a reasonable esti-mate for the conductivity and permittivity at anyfrequency for different tissue types.

    In the literature, we usually find data on specific con-ductivity and relative permittivity only at frequenciesabove 100Hz. For most tissues, data below that frequencyare very scarce or do not exist at all. The reason is not lackof interest, but because electrode effects can produce sig-nificant experimental errors for that frequency range. Theresults that have been published indicate that the imped-ance at frequencies under 100Hz is almost entirely resis-tive and that the capacitive component accounts for onlyaround 10% in most tissues. Between 100Hz and 100kHz,most tissues, with the exception of the anisotropic tissues,show almost no frequency-dependence.

    We will now briefly present the electric properties ofskeletal muscle, tumor, and skin and give the permittivityand specific conductivity ranges of some other tissues. Thebasic principles described earlier can be used to under-stand the differences in the electrical properties of varioustissues.

    4.1. Skeletal Muscle

    Data on skeletal muscle are the most abundant in the lit-erature. As a result of the anisotropy of this tissue, thedata are usually presented separately for the transverseand longitudinal directions, although some results withrandom orientation have been reported. In Figs. 25 wehave compiled some data from the literature (13,27,28,3034). The tissue samples were taken from different speciesand the measurements were made at different times afterexcision and with two different measurement methods(two and four-electrode technique). Therefore, the scatterof the data is rather high, especially in the low-frequencyrange. The anisotropy is also more pronounced in the low-frequency range. However, these differences are greater inthe conductivity data than in the permittivity data, wherefewer measurements are available.

    4.2. Tumor

    A tumor is an abnormal mass of tissue surrounded by oneor more normal body tissues. It has no useful function andgrows at the expense of healthy tissues. As noted previ-ously, many tumors have a significantly different electricalconductivity and permittivity from normal, surroundingtissues. This fact was attempted to be used in diagnosingtumors. For this reason, electrical impedance measuringand imaging systems are being designed and tested toscreen for tumors. A study by Smith et al. (35) on tumorsin liver tissue showed a significant difference in electricalproperties between healthy liver tissue and tumors. Re-sults show that tumor conductivity is 67.5 fold higherthan liver conductivity; the difference in permittivity val-ues is 25 fold. However, the dielectric properties of tu-mors cannot be generalized as large differences existbetween different tumor types and even between tumorsof the same type. Electrical properties also depend a greatdeal on the size or the development stage of the tumor.Namely, the tumor core can already exert tissue necrosis(35,36).

    4.3. Skin

    Skin is a very interesting tissue because of its highly in-homogeneous structure, which thus leads to inhomoge-neous dielectric properties. Generally, skin has threedifferent layers: the epidermis, dermis, and subcutaneoustissue. The epidermis is the outer layer of skin. The thick-ness of the epidermis varies in different types of skin. It isthe thinnest on the eyelids at 0.05mm and the thickest onthe palms and soles at 1.5mm (3). The epidermis containsdifferent layers, but the one that defines its dielectricproperties the most is the outermost layer, the stratumcorneum. That layer is composed of dead, flat skin cellsthat shed about every two weeks. Although it is very thin(typically around 20 mm), it contributes a great deal to thedielectric properties of the skin. Its high resistivity makesskin one of the most resistive tissues in the human body.Its main function is protection of the body from the exter-nal environmental factors. The lower-lying layers, hencethe rest of the epidermis (important in the immune re-sponse), the dermis (which gives firmness and elasticity),

    6 ELECTRIC PROPERTIES OF TISSUES

  • and the subcutaneous tissue (fat, connective tissue, largerblood vessels, and nerves), all have much lower resistivi-ties.

    Especially in the low-frequency range (under 10kHz),the impedance of skin is dominated by the stratum corn-eum even though this layer is very thin. Studies showthat, for frequencies under 10kHz, the share of stratumcorneum in the total impedance of skin is around 50% (37),but at 100kHz drops to around 10% (38).

    In Figs. 6 and 7, we can see the specific conductivityand relative permittivity, respectively, of intact skin whosedielectric properties are dominated by stratum corneum(diamonds) and lower-lying layers of skin (boxes) (39,40).The measurements were made first on intact skin. Thestratum corneum was then removed by cellulose adhesivetape stripping, and the measurements were repeated.

    4.4. Dielectric Properties Data Ranges of Some Body Tissues

    See Table 1 for data ranges of some body tissues(10,12,13,27,28,30,3236,3944).

    5. USES OF BIOIMPEDANCE MEASUREMENTS

    Today, bioimpedance measurements provide an importantmethod for the noninvasive investigation of tissue struc-ture and properties or for monitoring physiological change(i.e., static or dynamic human organism properties).One of the main problems one encounters using bioim-pedance measurements is still the reliability of the re-

    sults. The scatter of the data for the electrical parametersof tissues in Figs. 27 illustrates the problem of measure-ment reproducibility. Some of the scatter may result fromproblems in measuring technique, such as electrode po-larization, but some may simply be caused by the individ-ual variability among samples. At present, the relativeimportance of these two factors in determining the overallscatter is not clear. This scatter makes it difficult to es-tablish criteria of normality or reference value for partic-ular measurement results. However, in spite of the largedifferences between reported data on dielectric propertiesof different tissues, we can still find some very useful ap-plications based on the measurements of the differences orchanges in the specific conductivity or relative permit-tivity.

    As noted previously, tumor diagnosis is one of the im-portant applications of the measurements of bulk electri-cal properties of tumors and the surrounding tissues (35).Comparing dielectric properties of tumor and healthy sur-rounding tissues, one notices that a tumor has a muchhigher specific conductivity. The tumor tissue is more con-ductive at low frequencies because of its smaller volumefraction of intact cells, and at high frequencies because ofits higher water content and its irregular and fenestratedvascularization. As stated above, the dielectric propertiesof the tumor and normal tissue appear to be distinctlydifferent, by factors of 67.5 in the conductivity and 25 inthe permittivity. Pronounced differences in the bulk elec-trical properties of a tumor and the surrounding normaltissue, if consistently present, could lead to a variety of

    0,01

    0,1

    1

    10

    100

    1,E+01 1,E+02 1,E+03 1,E+04 1,E+05 1,E+06 1,E+07 1,E+08 1,E+09 1,E+10 1,E+11frequency (Hz)

    con

    duct

    ivity

    (S/m

    )Burger,van Dongen,1960,cow/horse,4el,1h,33degCBurger,van Dongen,1960,human,4el,1h,37degCBurger,van Dongen,1960,rabbit,4el,1h,37degCRush et al,1963,dog,4el,anesthetized,body tempEpstein,Foster,1983,dog,2/4el,fresh,body tempGielen et al,1984,rabbit,4el,in vivo,body tempBodakian,Hart,1994,beaf,2el,fresh meat,20degCBodakian,Hart,1994,2el,fresh chicken,20degCGabriel et al,1996,sheep,4el,37degCGielen et al,1984,rat,in vivo,37degC

    Figure 2. Specific conductivity for skeletal muscle, longitudinal direction.

    ELECTRIC PROPERTIES OF TISSUES 7

  • clinical applications. Moreover, because we only need tomeasure the dielectric properties of tumors relative tonormal surrounding tissue, the exact values for both arenot critical.

    Another application is monitoring of the changes inconductivity of the tissue during the process of the elect-ropermeabilization (4548). To effectively use electropora-tion in clinical applications, we need to detect whether thetarget tissue area has been permeabilized. This feedbackcould then be used to adjust the electroporation parame-ters during the treatment to make it more efficient. As thespecific conductivity of tissue increases when permeabili-zed, its measurement can be used as an indicator of thelevel of the electropermeabilization in the tissue. Again,we are only dealing with relative changes in tissue dielec-tric properties before and after tissue elect-ropermeabilization and not the exact absolute values. Afeasibility study for electrical impedance tomography as ameans to monitor tissue electroporation has been made(49). In this preliminary demonstration, the electroporat-ed regions in liver were clearly distinguishable. However,more work needs to be done to bring this technique toclinical practice.

    Electrical impedance tomography may also be used forother in vivo applications such as cryosurgery (50). Cryo-surgery is a surgical procedure that destroys tissue byfreezing it with a cryogen-cooled surgical probe that is incontact with the targeted tissue. However, although theextent of freezing can be monitored with an array of imag-ing techniques, the effective application of cryosurgery is

    still hampered by the fact that the extent of freezing doesnot necessarily correspond to the extent of tissue destruc-tion. Substantial changes in tissue electrical propertiescaused by tissue destruction can be monitored by means ofelectrical impedance tomography in order to evaluate theeffectiveness of the procedure.

    A more commercially orientated application is mea-surement of the dielectric properties of meat products thatcould serve as a monitoring tool of their storage and prep-aration history (34). Dielectric properties of beef andchicken were measured on both commercially purchasedand fresh meat, as well as thawed and cooked meat. Theresults show that the anisotropy of skeletal muscle islower in the commercially purchased samples. In addi-tion, the conductivity values are much higher for the com-mercial samples, particularly at low frequencies. Furtherchanges are produced by freezing and cooking, which in-dicates that dielectric spectroscopy can be used to deter-mine the storage/preparation history of meat products.

    With bioimpedance measurements, it is also possible toestimate the ratio of muscle to fat mass because fat haslower conductivity than muscle tissue (3). The intention isoften to determine total body water, extracellular/intra-cellular fluid balance, muscle mass, and fat mass. Appli-cation areas are as diverse as sports medicine, nutritionalassessment, and fluid balance in renal dialysis and trans-plantation. Other applications of the theories of measur-ing electric properties of biomaterials, ranging fromdiagnostic to therapeutic applications or laboratory pro-cedures, can be found in the literature (3,16,17).

    0,01

    0,1

    1

    10

    100

    1,E+01 1,E+02 1,E+03 1,E+04 1,E+05 1,E+06 1,E+07 1,E+08 1,E+09 1,E+10 1,E+11

    frequency (Hz)

    con

    duct

    ivity

    (S/m

    )Burger,van Dongen,1960,cow/horse,4el,1h,33degCBurger,van Dongen,1960,human,4el,1h,37degCBurger,van Dongen,1960,rabbit,4el,1h,37degCRush et al,1963,dog,4el,anesthetized,body tempEpstein,Foster,1983,dog,2/4el,fresh,body tempGielen et al,1984,rabbit,4el,in vivo,body tempBodakian,Hart,1994,beaf,2el,fresh meat,20degCBodakian,Hart,1994,2el,fresh chicken,20degCGabriel et al,1996,sheep,4el,37degCGielen et al,1984,rat,in vivo,37degC

    Figure 3. Specific conductivity for skeletal muscle, transverse direction.

    8 ELECTRIC PROPERTIES OF TISSUES

  • 1,E+00

    1,E+01

    1,E+02

    1,E+03

    1,E+04

    1,E+05

    1,E+06

    1,E+07

    1,E+08

    1,E+09

    1,E+01 1,E+02 1,E+03 1,E+04 1,E+05 1,E+06 1,E+07 1,E+08 1,E+09 1,E+10 1,E+11

    frequency (Hz)

    perm

    ittiv

    ityEpstein,Foster,1983,dog,2/4el,fresh,body temp

    Bodakian,Hart,1994,beaf,2el,fresh meat,20degCGabriel et al,1996,sheep,4el,37degC

    Gielen et al,1984,rat,in vivo,37degC

    Figure 4. Relative permittivity for skeletal muscle, longitudinal direction.

    1,E+00

    1,E+01

    1,E+02

    1,E+03

    1,E+04

    1,E+05

    1,E+06

    1,E+07

    1,E+08

    1,E+09

    1,E+01 1,E+02 1,E+03 1,E+04 1,E+05 1,E+06 1,E+07 1,E+08 1,E+09 1,E+10 1,E+11

    frequency (Hz)

    perm

    ittiv

    ity

    Epstein,Foster,1983,dog,2/4el,fresh,bodytemp

    Bodakian,Hart,1994,beaf,2el,fresh meat,20degCGabriel et al,1996,sheep,4el,37degC

    Gielen et al,1984,rat,in vivo,37degC

    Figure 5. Relative permittivity for skeletal muscle, transverse direction.

    ELECTRIC PROPERTIES OF TISSUES 9

  • 1,0E-05

    1,0E-04

    1,0E-03

    1,0E-02

    1,0E-01

    1,0E+00

    1,0E+01 1,0E+02 1,0E+03 1,0E+04 1,0E+05 1,0E+06 1,0E+07 1,0E+08 1,0E+09 1,0E+10 1,0E+11

    frequency (Hz)

    spec

    ific

    cond

    uctiv

    ity (S

    /m) skin with stratum corneum

    lower lying layers of skin

    Figure 6. Specific conductivity of intact skin with dominating stratum corneum (diamonds) andlower-lying layers of skin alone (boxes).

    1,0E+00

    1,0E+01

    1,0E+02

    1,0E+03

    1,0E+04

    1,0E+05

    1,0E+06

    1,0E+07

    1,0E+01 1,0E+02 1,0E+03 1,0E+04 1,0E+05 1,0E+06 1,0E+07 1,0E+08 1,0E+09 1,0E+10 1,0E+11

    frequency (Hz)

    rela

    tive

    perm

    ittiv

    ity

    skin with stratum corneumlower lying layers of skin

    Figure 7. Relative permittivity of intact skin with dominating stratum corneum (diamonds) andlower-lying layers of skin alone (boxes).

    10 ELECTRIC PROPERTIES OF TISSUES

  • BIBLIOGRAPHY

    1. K. R. Foster and H. P. Schwan, Dielectric properties of tissues.In: C. Polk and E. Postow, eds.,Handbook of Biological Effectsof Electromagnetic Fields. New York: CRC Press, 1996.

    2. K. R. Foster and H. P. Schwan, Dielectric properties of tissuesand biological materials: a critical review. Crit. Rev. Biomed.Eng. 1989; 17:25104.

    3. S. Grimnes and O. G. Martinsen, Bioimpedance & Bioelec-tricity Basics. San Diego, CA: Academic Press, 2000.

    4. J. P. Reilly, Applied Bioelectricity, From Electrical Stimula-tion to Electropathology. New York: Springer-Verlag, 1998.

    5. R. Pethig, Dielectric and Electronic Properties of BiologicalMaterial. New York: Wiley, 1979.

    6. T. Kotnik and D. Miklavcic, Second-order model of membraneelectric field induced by alternating external electric fields.IEEE Trans. Biomed. Eng. 2000; 47:10741081.

    7. T. Kotnik and D. Miklavcic, Theoretical evaluation of the dis-tributed power dissipation in biological cells exposed to elec-tric fields. Bioelectromagnetics 2000; 21:385394.

    8. M. Pavlin, T. Slivnik, and D. Miklavcic, Effective conductivityof cell suspensions. IEEE Trans. Biomed. Eng. 2002; 49:7780.

    9. R. S. Chiu and M. A. Stuchly, Electric fields in bone marrowsub-structures at power-line frequencies. IEEE Trans. Bio-med. Eng., in press.

    10. L. A. Geddes and L. E. Baker, The specific resistance of bio-logical material - a compendium of data for the biomedicalengineer and physiologist.Med. Biolog. Eng. 1967; 5:271293.

    11. F. X. Hart, N. J. Berner, and R. L. McMillen, Modelling theanisotropic electrical properties of skeletal muscle. Phys.Med. Biol. 1999; 44:413421.

    12. H. P. Schwan and C. F. Kay, Specific resistance of body tis-sues. Circ. Res. 1956; 4:664670.

    13. B. R. Epstein and K. R. Foster, Anisotropy in the dielectricproperties of skeletal muscle. Med. Biol. Eng. Comput. 1983;21:5155.

    14. E. Pacelat, R. Magjarevic, and V. Isgum, Measurement ofelectrode-tissue interface characteristics during high currenttranscranial pulse electrical stimulation. Measurement 2000;27:133143.

    15. M. Noshiro, T. Morimoto, H. Nagao, and H. Matsuda, Elec-trical impedance in the lower limbs of patients with Duch-enne muscular dystrophy: a preliminary study. Med. Biol.Eng. Comput. 1993; 31:97102.

    16. Bioelectrical impedance techniques in medicine. Crit. Rev.Biomed. Eng. 1996; 24(46).

    17. P. J. Riu, J. Rosell, R. Bragos, and O. Casas, Electrical Bio-impedance Methods: applications to Medicine and Biotech-nology. New York: The New York Academy of Sciences, 1999.

    18. D. Haemmerich, S. T. Staelin, J. Z. Tsai, S. Tungjitkusolmun,D. M. Mahvi, and J. G. Webster, In vivo electrical conductivityof hepatic tumours. Physiol. Meas. 2003; 24:251260.

    19. B. Blad, P. Wendel, M. Jonsson, and K. Lindstrom, An elec-trical impedance index to distinguish between normal andcancerous tissues. J. Med. Eng. Tech. 1999; 22:15.

    20. Y. Ultchin, U. Nachaliel, and A. Ori, Indirect calculation ofbreast tissue impedance values. Physiol. Meas. 2002; 23:177182.

    21. W. T. Joines, Y. Zhang, C. Li, and R. L. Jirtle, The measuredelectrical properties of normal and malignant human tissuesfrom 50 to 900MHz. Med. Phys. 1994; 21:547550.

    22. J. Biggs, K. Cha, and K. Horch, Electrical resistivity of theupper arm and leg yields good estimates of whole body fat.Physiol. Meas. 2001; 22:365376.

    23. B. K. van Kreel, N. Cox-Reyven, and P. Soeters, Determina-tion of total body water by multifrequency bio-electric imped-ance: development of several models.Med. Biol. Eng. Comput.1998; 36:337345.

    24. M. Schaefer, W. Gross, J. Ackemann, and M. M. Gebhard, Thecomplex dielectric spectrum of heart tissue during ischemia.Bioelectrochemistry 2002; 58:171180.

    25. D. Haemmerich, O. R. Ozkan, J.-Z. Tsai, S. T. Staelin, S.Tungjitkusolmun, D. M. Mahvi, and J. G. Webster, Changes inelectrical resistivity of swine liver after occlusion and post-mortem. Med. Biol. Eng. Comput. 2002; 40:2933.

    Table 1. Data Ranges of Specific Conductivities and Relative Permittivities of Some Other Tissues in the Low-FrequencyRange (10,12,13,27,28,30,3236,3944)

    Spec. Conductivity (S/m) Rel. Permittivity

    Tumor 0.220.4 60000 (at 1 kHz)Fat 0.020.04 10 000000 (at 10Hz)Muscle

    Transversal 0.040.14 1 50000040000000 (at 10Hz)Longitudinal 0.30.8 1000000066000000 (at 10Hz)

    Skin (dry) 0.000020.0002 14006600 (at 10Hz)Stratum corneum 0.0000125 10000 (at 2Hz)Lower-lying layers 0.227 1 200000 (at 2Hz)

    Bone 0.010.06 400001000000 (d.c.)Blood 0.430.7 3000 (at 1 kHz)Heart 0.060.4 7 00000020000000 (d.c.)Kidney 0.6 30000000 (d.c.)Liver 0.0230.2 1500000050000000 (d.c.)Lung (inflated) 0.0240.09 10000000 (d.c.)Spleen 0.043 45000000 (d.c.)Gray matter 0.033 50000000 (d.c.)White matter 0.023 30000000 (d.c.)

    Specific conductivities are given for direct current measurements (0Hz); measuring frequencies for relative permittivities are stated in brackets.

    ELECTRIC PROPERTIES OF TISSUES 11

  • 26. D. A. McRae and M. A. Esrick, Changes in electrical imped-ance of skeletal muscle measured during hyperthermia. Int.J. Hypertherm. 1993; 9:247261.

    27. S. Gabriel, R. W. Lau, and C. Gabriel, The dielectric proper-ties of biological tissues: II. Measurements in the frequencyrange 10Hz to 20GHz. Phys. Med. Biol. 1996; 41:22512269.

    28. H. C. Burger and R. van Dongen, Specific electric resistanceof body tissues. Phys. Med. Biol. 1960; 5:431447.

    29. F. X. Hart and W. R. Dunfee, In vivo measurement of the low-frequency dielectric spectra of frog skeletal muscle. Phys.Med. Biol. 1993; 38:10991112.

    30. S. Rush, J. A. Abildskov, and R. McFee, Resistivity of bodytissues at low frequencies. Circ. Res. 1963; 12:4050.

    31. S. Gabriel, R. W. Lau, and C. Gabriel, The dielectric proper-ties of biological tissues: III. Parametric models for the di-electric spectrum of tissues. Phys. Med. Biol. 1996; 41:22712293.

    32. C. Gabriel, S. Gabriel, and E. Corthout, The dielectric prop-erties of biological tissues: I. Literature survey. Phys. Med.Biol. 1996; 41:22312249.

    33. F. L. H. Gielen, W. Wallinga-de Jonge, and K. L. Boon, Elec-trical conductivity of skeletal muscle tissue: Experimentalresults from different muscles in vivo. Med. Biol. Eng. Com-put. 1984; 22: 569577.

    34. B. Bodakian and F. X. Hart, The dielectric properties of meat.IEEE Trans. Dielect. Elect. Insul. 1994; 1:181187.

    35. S. R. Smith, K. R. Foster, and G. L. Wolf, Dielectric propertiesof VX-2 carcinoma versus normal liver tissue. IEEE Trans.Biomed. Eng. 1986; 33:522524.

    36. A. J. Surowiec, S. S. Stuchly, J. R. Barr, and A. Swarup, Di-electric properties of breast carcinoma and the surroundingtissues. IEEE Trans. Biomed. Eng. 1998; 35:257263.

    37. Y. Yamamoto, T. Yamamoto, and T. Ozawa, Characteristics ofskin admittance for dry electrodes and the measurement ofskin moisturisation.Med. Biol. Eng. Comput. 1986; 24:7177.

    38. O. G. Martinsen, S. Grimnes, and E. Haug, Measuring depthdepends on frequency in electrical skin impedance measure-ments. Skin Res. Technol. 1999; 5:179181.

    39. T. Yamamoto and Y. Yamamoto, Electrical properties of theepidermal stratum corneum.Med. Biol. Eng. 1976; 14(2):151158.

    40. T. Yamamoto and Y. Yamamoto, Dielectric constant and re-sistivity of epidermal stratum corneum.Med. Biol. Eng. 1976;14(5):494499.

    41. H. P. Schwan, Electric characteristics of tissues. Biophysik.1963; 1:198208.

    42. F. X. Hart, The impedance spectroscopy of skeletal muscle.Proc. Tenth Electrotechnical and Computer Science Confer-ence ERK 2001, Invited lecture, Slovenia, 2001.

    43. H. P. Schwan and C. F. Kay, The conductivity of living tissues.Ann. NY Acad. Sci. 1957; 65:10071013.

    44. T. J. C. Faes, H. A. van der Meij, J. C. De Munck, and R. M.Heethaar, The electric resistivity of human tissues (100Hz10MHz): a meta-analysis of review studies. Physiol. Meas.1999; 20:R1R10.

    45. M. Pavlin, M. Kanduser, M. Rebersek, G. Pucihar, F. X. Hart,R. Magjarevic, and D. Miklavcic, Effect of cell electroporationon the conductivity of a cell suspension. Biophys. J., 2005;88:43784390.

    46. U. Pliquett and M. R. Prausnitz, Electrical impedance spec-troscopy for rapid and non-invasive analysis of skin electrop-oration. In: M. J. Jaroszeski, R. Gilbert, and R. Heller,Electrically Mediated Delivery of Molecules to Cells, Electro-chemotherapy, Electrogenetherapy and Transdermal Deliveryby Electroporation. Totowa, NJ: Humana Press, 2000.

    47. U. Pliquett, R. Langer, and J. C. Weaver, Changes in the pas-sive electrical properties of human stratum corneum due toelectroporation. BBA. 1995; 1239:111121.

    48. U. Pliquett and J. C. Weaver, Electroporation of human skin:simultaneous measurement of changes in the transport of twofluorescent molecules and in the passive electrical properties.Bioelectrochem. Bioenerget. 1996; 39:112.

    49. R. V. Davalos, B. Rubinsky, and D. M. Otten, A feasibilitystudy for electrical impedance tomography as a means tomonitor tissue electroporation for molecular medicine. IEEETrans. Biomed. Eng. 2002; 49:400403.

    50. R. V. Davalos and B. Rubinsky, Electrical impedance tomo-graphy of cell viability in tissue with application to cryosur-gery. J. Biomechan. Eng. 2004; 126:305309.

    12 ELECTRIC PROPERTIES OF TISSUES