630
Views
7
CrossRef citations to date
0
Altmetric
Articles

A three-dimensional image reconstruction algorithm for electrical impedance tomography using planar electrode arrays

, &
Pages 471-491 | Received 24 Sep 2015, Accepted 15 Mar 2016, Published online: 06 Apr 2016

Abstract

We present a three-dimensional non-iterative reconstruction algorithm developed for conductivity imaging with real data collected on a planar rectangular array of electrodes. Such an electrode configuration as well as the proposed imaging technique is intended to be used for breast cancer detection. The algorithm is based on linearizing the conductivity about a constant value and allows real-time reconstructions. The performance of the algorithm was tested on numerically simulated data and we successfully detected small inclusions with conductivities three or four times the background lying beneath the data collection surface. The results were fairly stable with respect to the noise level in the data and displayed very good spatial resolution in the plane of electrodes.

AMS Subject Classifications:

1. Introduction

Electrical Impedance Tomography (EIT) is a technology used to image the distribution of electrical properties such as conductivity and/or permittivity within an object using measurements of electric currents and voltages on its surface. Since different materials display different electrical properties, EIT can be used as a method of industrial, geophysical and medical imaging (see, for example, [Citation1] and the references therein). The application of EIT considered in this paper is breast cancer imaging.

Breast cancer is routinely investigated by palpation, X-ray mammography or ultrasound imaging with sensitivity rates of up to 90%. The diagnoses, however, yield rather unspecific results. Only one in five biopsies of suspicious lesions leads to a malignant histological diagnosis,[Citation2] causing unnecessary distress for the patient and significant delays in establishing a diagnosis. Research is therefore aimed at developing alternative imaging techniques to diagnose malignant breast tumours more accurately and possibly earlier. Since in vivo studies have discovered a difference of three times or more in the specific electrical conductivity and permittivity between healthy and cancerous tissue,[Citation3Citation5] imaging the electrical properties of breast tissues could improve the specificity of mammograms. The advantages (portability, low cost, little or zero patient discomfort, no known patient risk and no known side effects) of an impedance imaging system over traditional X-ray mammography could make this technology a welcome addition to the tools available in the fight against breast cancer. Although much research has been devoted to both the theoretical and the practical developments of Electrical Impedance Mammography,[Citation6Citation19] to date, the technique has not reached clinical acceptance because of its sensitivity to measurement errors, high computational demands and practical issues: errors in electrode positions or boundary shape, high and uncontrollable contact impedance of the skin (variations of 20% or more).

Several EIT mammographic sensors have been developed recently at the University of Mainz in collaboration with Oxford Brookes University. In contrast to most previous EIT instruments designed for breast cancer detection,[Citation9] but similar to devices studied by [Citation6,Citation7,Citation16,Citation18,Citation20,Citation21], these mammographic sensors are planar. Detailed descriptions of earlier prototypes can be found in [Citation22,Citation23]. The latest design consists of a planar sensing head with 36 disk electrodes of equal size arranged in a rectangular array of 20 outer (active) electrodes where the external currents are injected, and 16 inner (passive) electrodes where the induced voltages are measured, see Figure . To avoid any problems due to the unknown contact impedance, the voltages are not measured at the active electrodes, but at the passive electrodes, very high impedance voltage measurements are taken, and the problem of the unknown contact impedance does not arise. These voltage measurements are relative to the systems ground. Moreover, the device has a fixed geometry and the positions of the electrodes are exactly known.

Almost all previous EIT devices [Citation6,Citation7,Citation9,Citation16,Citation18,Citation20,Citation21] use the same electrodes for current injection and voltage measurement. The excitation current is injected (extracted) at one pair of electrodes at a time and the resulting voltage is measured at all or some of the remaining electrodes. The novelty of our EIT devices, and hence of the image reconstruction methods proposed consists precisely in the distinct use of active and passive electrodes. The active electrodes are used only for current injection while the passive electrodes only for voltage measurements.

Figure 1. Layout of the rectangular electrode array of the latest prototype developed at the University of Mainz in collaboration with Oxford Brookes University:

are the active electrodes used for current injection, and
are the passive electrodes used for potential measurements.

Figure 1. Layout of the rectangular electrode array of the latest prototype developed at the University of Mainz in collaboration with Oxford Brookes University: Display full size are the active electrodes used for current injection, and Display full size are the passive electrodes used for potential measurements.

In two-dimensions, two different non-iterative algorithms for imaging the conductivity at the surface using the tomographs designed at the University of Mainz were described in [Citation22,Citation23]. In both cases numerical reconstructions had very good spatial resolution, and the algorithms were robust with respect to errors in the data. A three-dimensional iterative reconstruction method that enforced sparsity and used an adapted complete electrode model was also applied to these devices in [Citation24]. Although it produced good images, this iterative procedure proved to be quite demanding computationally and sensitive to the choice of the regularization parameter.

The purpose of the current paper is to respond to these issues i.e. the two-dimensional nature of the first two algorithms and the numerical sensitivity of the sparsity algorithm, by developing a simple, direct and rapid three-dimensional reconstruction algorithm to image the region beneath the rectangular electrode array. The proposed inversion technique is intended to be used in the future for conductivity imaging using real data collected by the specific EIT device described above. The reconstruction method is similar to an approach described in [Citation20] and is based on linearizing the conductivity distribution about a constant approximation in order to reduce the computational demands.

The structure of the paper is as follows. In Section 2, we present the mathematical formulation of the inverse problem and of the method proposed to image the conductivity. Section 3 is dedicated to the numerical implementation of the reconstruction algorithm. The performance of the algorithm is illustrated by showing a number of reconstructions from simulated data in Section 4.

2. Mathematical formulation

Since the EIT sensor is much smaller than the human body to which it is applied, and given the rapid decay of the induced potentials as we move away from the top surface (see Equation (Equation3)), the mathematical analysis of breast cancer detection can be considered to be the inverse conductivity problem of EIT on an unbounded domain, specifically the lower half space. Let Ω={(x,y,z):z<0}R3 be a conductive object with boundary Γ={(x,y,z):z=0} and QL(Ω) be a set of uniformly strictly positive parameters. Suppose that σQ is an isotropic scalar electrical conductivity distribution and that there are no current sources inside Ω. A set of L electrodes is placed on Γ in a rectangular array. Let {el}l=1M be the set of passive electrodes (voltage measurement) and {el}l=M+1L be the active electrodes (current injection). If low-frequency currents are applied to the active electrodes, the electric potential u satisfies the generalized Laplace equation(1) ·σ(x,y,z)u(x,y,z)=0inΩ,(1)

subject to the boundary condition(2) σ(x,y,0)uz(x,y,0)=j(x,y)onΓ,(2)

where j is the induced current density distribution.

In this paper, we restrict our analysis to currents jL2(Γ) with bounded support and to weak solutions uH of Equation (Equation1) in the lower half-space Ω for which the following point-wise estimates hold uniformly with respect to the direction of x=(x,y,z):(3) ux=O1/xasx,(3)

and(4) uζx=O1/xasxforζ=x,y,z.(4)

Here, L2(Γ) denotes the space of L2-functions with vanishing integral mean on Γ and H=H1,α(Ω) is a completion of suitable C-functions with respect to the norm ·1,α parametrized by the weight function ραx,y,z=1+xα with α>1. Further details on the definition of the weighted Sobolev spaces H1,α(Ω) and on the derivation of these asymptotic estimates can be found in [Citation25].

The weak formulation for the boundary value problem (Equation1)–(Equation2) and its properties are well known. However, we will use the same formulation as in [Citation25], i.e. a function uH is called a weak solution of this Neumann boundary problem if(5) Tσ,u,v=Lj(v),for anyvH,(5)

where the trilinear form T:L(Ω)×H×HR and the linear form Lj:HR are defined by(6) Tσ,u,v=Ωσu·vdV,(6) (7) Lj(v)=ΓjvdS.(7)

Hence, the solution of the Neumann boundary value problem (Equation1)–(Equation2) is given by u=Tσ-1Lj, where Tσ-1L(H,H), the inverse of T(σ,·,·), exists and is bounded and continuous.[Citation26,Citation27]

Since in our application the voltages are not measured at the active electrodes, the ave-gap electrode model [Citation12] can be used. We, therefore, assume that the current density is uniformly distributed over each active electrode and that it is zero outside of the support of all active electrodes. Hence, we approximate j in Equation (Equation2) by(8) j(x,y)=Il/Al,(x,y)onel,l=M+1,,L0,otherwise,(8)

where Il is the current sent to the lth active electrode el and Al is the area of el.

In our experimental set-up, a basis of current patterns {I1,,IL-M-1} is applied to the set of L-M active electrodes {el}l=M+1L. For each current pattern Ik=IM+1k,,ILk, k=1,,L-M-1, the resulting potentials Uk=U1k,,UMk are measured at the set of M passive electrodes {el}l=1M. The ave-gap model predicts the voltage measured on each electrode as the average of the solution uk to (Equation1), (Equation2) and (Equation8) with j=jk over the surface of the electrode:(9) Ulk=1Aleluk(x,y,0)dS,l=1,,M.(9)

The inverse problem is to estimate the conductivity σ(x,y,z) from all L-M-1 linearly independent sets of surface measurements.

The reconstruction algorithm presented in this paper is based on the assumption that the spatially varying conductivity is a small perturbation from a constant and known background conductivity σ0, i.e.(10) σ(x,y,z)=σ0+δσ(x,y,z),(10)

where σ=σ0 near Γ. Let us also express the corresponding potential u in terms of u0, the solution of (Equation1), (Equation2) and (Equation8) for σ=σ0, as(11) u(x,y,z)=u0(x,y,z)+δu(x,y,z).(11)

It is shown in [Citation26,Citation27] that both the parameter-to-solution map(12) Aσ:QH,σTσ-1Lj.(12)

and the complete parameter-to-solution map (i.e. forward operator for a single measurement with current density j)(13) Aj:QL2,α(Γ),σγTσ-1Lj,(13)

are analytic on each uniformly strictly positive open set Q. It therefore follows that δu=Oδσ in Ω¯. Here, γ:HL2,α(Γ) is the trace operator and L2,α is the weighted L2-space with respect to ρα.

Our method now follows the approach of Calderón [Citation28]. Inserting Equations (Equation10) and (Equation11) into (Equation1) and simplifying leads to the following linearized equation(14) ·σ0δu+·δσu0=Oδσ.(14)

We choose a test function v0H that is the solution of the following boundary value problem:(15) ·σ0v0(x,y,z)=0,inΩ,(15) (16) σ0v0z(x,y,0)=j~(x,y)=I~l/Al,(x,y)onel,l=1,,M0,otherwise,,(16)

where I~l is a simulated current applied at the lth passive electrode. More explicitly, u0 is the (weak) potential created in Ω if it consisted of a material of constant conductivity σ0, while v0 is the induced (weak) potential when currents are applied to this object of uniform conductivity distribution but reversing the roles of the active and passive electrodes. An illustration of the thought experimental set-up corresponding to the homogeneous forward problem (Equation15)–(Equation16) is presented in Figure . There are M-1 linearly independent simulated current patterns {I~1,,I~M-1}, where I~i=I1i,,IMi, that can be applied to the M passive electrodes.

We now multiply the expression in Equation (Equation14) by this test function v0 and integrate over Ω. Applying Gauss’s divergence theorem [Citation29] and taking into account the asymptotic behaviour of weak solutions of (Equation1) and (Equation15) given by Equations (Equation3)–(Equation4) yields(17) Γn·v0σ0δu+δσu0dS=Ωσ0v0·δu+δσv0·u0dV+O(δσ),(17)

where n is the outward normal unit vector.

Furthermore, multiplying Equation (Equation15) by δuH and repeating the same steps as above, we obtain(18) Γn·σ0δu0v0dS=Ωσ0δu·v0dV.(18)

Inserting this result in Equation (Equation17) gives(19) Γn·v0σ0δu+δσu0dS=Γn·σ0δuv0dS+Ωδσv0·u0dV+O(δσ).(19)

Since both u and u0 satisfy the same Neumann boundary condition (Equation2), it follows that n·σ0δu+δσu0Γ+O(δσ)=0 and Equation (Equation19) becomes(20) Γn·σ0δuv0dS=-Ωδσv0·u0dV+O(δσ).(20)

Within this setting, for fixed Neumann data n·σ0u0Γ=j, Equation (Equation20) expresses in weak form the change in the Neumann-to-Dirichlet map which occurs due to a change in σ0 for some test function v0H that is the solution of Equation (Equation15) with Neumann trace n·σ0v0Γ=j~L2(Γ).

Equation (Equation20) is true for any uk and u0k, (weak) solutions of Equation (Equation1) for a given current pattern jk applied to the active electrodes, and for any v0i, (weak) solution of Equation (Equation15) subject to a current density j~i corresponding to the ith simulated current pattern I~i applied at the passive electrodes and given by Equation (Equation16). Thus, the linearized problem (Equation20) can be approximated by the following system of equations(21) l=1Melu0k-ukj~idSΩδσv0i·u0kdV,k=1,,L-M-1andi=1,,M-1.(21)

Using the ave-gap model for the electrodes, the inverse problem becomes that of finding the perturbation δσ satisfying the following system of equations(22) l=1MU0,lk-UlkI~li=Ωδσv0i·u0kdV,k=1,,L-M-1andi=1,,M-1,(22)

where(23) U0,lk=1Alelu0k(x,y,0)dS.(23)

The left-hand side of (Equation22) contains only measured and simulated quantities, Ulk and U0,lk, respectively, and will be denoted by B(ik), i.e.(24) B(i,k)=l=1MU0,lk-UlkI~li.(24)

A further simplification that can be introduced to aid the image reconstruction process comes from the fact that our objective is to identify a small object(tumour) of uniform conductivity lying within a region (breast tissue) of uniform, but different and known, conductivity. We may therefore assume that the conductivity is piecewise constant. In this case the half-space can be approximated by voxels, {Vn}n=1, and thusδσ(x,y,z)=n=1δσnχn(x,y,z),

where χn is the characteristic function over the nth voxel, i.e.χn(x,y,z)=1,(x,y,z)Vn,0,otherwise,

However, in practice, we only have a finite number of measurements and, hence, we can reconstruct the conductivities of a finite number N of voxels, where N(L-M-1)×(M-1). In this case, Equation (Equation22) reduces to an over-determined linear system of equations(25) B(i,k)=n=1NA(i,k,n)δσn,k=1,,L-M-1andi=1,,M,(25)

where(26) A(i,k,n)=Vnv0i·u0kdV.(26)

Note that the matrix A is independent of the measured voltage data and it can be computed in advance and stored for use with other reconstructions in the same geometry.

Since, the linearized system of Equations (Equation25) inherits ill-posedness from the original non-linear inverse conductivity problem, the matrix A is ill-conditioned and regularization is required. The system of Equations (Equation25) is therefore solved by means of generalized inverse and truncated singular value decomposition was used as the regularization scheme.[Citation30] In this way, very small singular values (i.e. smaller than a certain threshold ϵ) will be neglected and will not enter in the reconstruction.

3. Numerical implementation

3.1. Electrode and voxel configuration

As described earlier, our planar array of electrodes is rectangular and consists of L=36 circular electrodes of radius rl=3.5mm, l=1,,L, which are equally spaced. The distance between the centres of two adjacent electrodes is d=12mm. There are M=16 passive (inner) electrodes and L-M=20 active (outer) electrodes. The terms outer and inner are relative to the positions of the electrodes in the array. As seen in Figure , all the electrodes located on the boundary of the rectangular array are active, while the remaining ones are passive. At this point, it is important to note that this electrode configuration not only provides a simpler geometry than the hexagonal pattern of the earlier prototypes in [Citation22,Citation23], but for some electrical impedance imaging problems in geophysics, archaeology, medical diagnosis and industrial plant control, an appropriate electrode geometry may be that of a rectangular array of electrodes placed on a surface plane. For example, rectangular electrode configurations have been employed before for medical applications such as: breast cancer detection,[Citation6,Citation7,Citation20,Citation31] respiratory monitoring, functional imaging of the digestive system and peripheral venography,[Citation32] or for engineering and environmental studies (e.g. [Citation33] and the references therein).

For the numerical implementation of our algorithm, we use a layered voxel configuration similar to the one introduced in [Citation20] and it is depicted in Figure . There are 52 inner voxels in each layer:

  • 36 small inner voxels aligned immediately under the electrodes of dimensions 12mm×12mm×2mm, and

  • 16 outer large voxels arranged around the boundary of the electrode array of dimensions 24mm×24mm×2mm (i.e. the same height, 2mm, but four times the volume of a small voxel).

Although our interest is to image the conductivity in the region under the electrode array, the voxel configuration has to model, nevertheless, an unbounded domain. The reason for introducing the outer large voxels is to achieve this in a practical manner. As mentioned in Section 2, the total number of voxels should satisfy N(L-M-1)×(M-1)=285. Hence, since there are 52 voxels per layer, we cannot consider more than 5 layers of voxels (i.e. N=5×52=260285). Note that the number of voxels per layer is determined by the geometry of the electrode array, while the total number of layers is constrained by the number of active and passive electrodes, i.e. the total number (=285) of possible combinations of measured and simulated currents. The height of each voxel layer, however, is not fixed and it can be adjusted to allow conductivity reconstructions up to the required depths. Most of breast tumours are located near the skin surface.[Citation34] The results of a clinical study of single-breast lesions located at least 0.5 cm from the skin surface and pectoralis margin and at least 2.0cm from the nipple showed that the mean distance from the skin surface to the lesion was 0.9 cm (range, 0.5-1.7 cm).[Citation35] Therefore, in this approach we chose a voxel configuration which allows reconstructions up to a depth of 1 cm (=5 layers×2 mm).

Figure 2. Voxel configuration: are the active electrodes, are the passive electrodes, and P1-P5 denote the positions in the xy-plane where inclusions were placed in our numerical simulations.

Figure 2. Voxel configuration: are the active electrodes, are the passive electrodes, and P1-P5 denote the positions in the xy-plane where inclusions were placed in our numerical simulations.

3.2. Computation of matrix A

Before we discuss the construction of matrix A in (Equation26), we need to address the problem of computing the potentials u0 and v0 for a homogeneous medium.

Let P=(x,y,z) be an interior point in Ω, Q0=(x,y,0) be a point on the boundary Γ and Ql=(xl,yl,0) be the centre of lth electrode. For a constant conductivity distribution σ0, the solutions of the two theoretical forward problems, defined by (Equation1), (Equation2) and (Equation8) with j=ji and (Equation15)–(Equation16) with j~=j~k, are straightforward [Citation29]:(27) u0k(x,y,z)=12πσ0Γ1rPQ0jk(x,y)dxdy=12πσ0l=M+1LIlkAlel1rPQ0dxdy,k=1,,L-M-1,(27)

and, respectively,(28) v0i(x,y,z)=12πσ0Γ1rPQ0j~i(x,y)dxdy=12πσ0l=1MI~liAlel1rPQ0dxdy,i=1,,M-1,(28)

where rPQ0=|P-Q0|=(x-x)2+(y-y)2+z2.

In order to compute the entries of the matrix A, we have to discretize the volume integral in (Equation26). To this end, we follow the same approach as in [Citation20] and we divide each of the voxels into m subvoxels Vnj (j=m=1 for the inner voxels and j=1,,m=4 for the outer ones). Thus, there will be 100 subvoxels of equal volume per layer. The volume of the jth subvoxel of the nth voxel Vnj is, therefore, Volnj=12×12×2=288mm3. Using this discretization, and the expressions for u0k and v0i given by Equations (Equation27) and (Equation28), we obtain the following expression for the entries of the matrix A by evaluating the integral (Equation26) at the points Pnj, the centres of subvoxels Vnj:(29) A(i,k,n)=j=1mVolnj(2πσ0)2l=1Ml=M+1LI~liIlkAlAl×Pel1rPQ0dxdyP=Pnj·Pel1rPQ0dxdyP=Pnj.(29)

In order to evaluate these entries, we have to calculate(30) Pel1rPQ0dxdy=Pel1(P-Ql)-(Q0-Ql)dxdy,forl=1,,L.(30)

These quantities were computed analytically by converting the integrals to cylindrical coordinates as follows. Let (r,θ,z) and (s,ϑ,0) be the cylindrical coordinates of P~=P-Ql and Q~0=Q0-Ql, respectively. Specifically, Q~=(x~,y~,0), where x~=x-xl=scos(ϑ) and y~=y-yl=ssin(ϑ). Then, rPQ0 can be expressed in cylindrical coordinates as(31) rPQ0=|P~-Q~|=r2+s2-2rscos(ϑ-θ)+z2.(31)

It is clear from symmetry that the following integral is independent of the polar angle θ, i.e.(32) el1rPQ0dxdy=0rl02π1r2+s2-2rscos(ϑ-θ)+z2sdsdϑ=0rl02π1r2+s2-2rscosϑ+z2sdsdϑ.(32)

Hence,(33) Pel1rPQ0dxdy=err+ezz0rl02π1rPQ0sdsdϑ,(33)

with(34) r0rl02π1rPQ0sdsdϑ=2rlrκ2Eκ-(2-κ)Kκ,(34)

and(35) z0rl02π1rPQ0sdsdϑ=12κrrl4zKκ-2i×(r-rl-iz)Π2rlr+rl-izκ+(-r+rl-iz)Π2rlr+rl+izκ,(35)

where κ=4rrl(r+rl)2+z2, and K(κ), E(κ) and Π(n|κ) are the complete elliptic integrals of first, second and third kind, respectively. Further details about these special functions can be found in [Citation36]. Both the r-component and the z-component of the gradient given by Equations (Equation34) and (Equation35), respectively, are real valued functions. However, due to numerical inaccuracies in the computer algorithms used to evaluate these functions, there are always some residual imaginary parts in the order of the machine precision. To overcome this issue, only their real parts are considered in further computations.

Note that although in our implementation we preferred to use the above closed form expressions, all these integrals could also be evaluated to great accuracy by using a numerical quadrature method such as boundary elements (see, for example, [Citation37]). This would be more consistent with the numerical experiments where the forward model is a bounded domain but at the cost of increased computation time.

3.3. Construction of matrix B

For the construction of matrix B given by Equation (Equation24), we need the measured and simulated voltages, Ulk and U0,lk, respectively.

In practice, the voltages Ulk are known experimental data. However, as no real data is yet available, in this paper we simulate the measured voltages numerically as explained in Section 4.

In order to find the values of the simulated voltages U0,lk given by Equation (Equation23), we need to compute first u0kx,y=u0kx,y,z=0 by evaluating the expression in Equation (Equation27) at z=0. To this end, we use the same change of variables as in section 3.2 and Equation (Equation32). Note that due to the symmetry considerations of the integral (Equation32), the solution u0k is also independent of the polar angle θ. Hence, u0kr,θ,z=0=u0kr, and we obtain(36) u0kr=12πσ0l=M+1LIlkAl0rl02π1r2+s2-2rscosϑsdsdϑ,k=1,,L-M-1.(36)

The integral in Equation (Equation36) can be computed analytically, i.e.(37) 0rl02π1r2+s2-2rscosϑsdsdϑ=2r+rlE4rrlr+rl2-2r-rlK4rrlr+rl2.(37)

Once the currents applied at the active electrodes Ilk are known from measurements, u0k can be estimated in a straightforward way using Equations (Equation36)–(Equation37) at different points on the surfaces of passive electrodes el, l=1,,M, and then obtain U0,lk by evaluating numerically the integral in Equation (Equation23).

The simulated currents I~i=I~1i,,I~16i, i=1,,M-1=15, are assumed to be standard linearly independent trigonometric current patterns, i.e.(38) I~li=cosi(l-1)2πM,i=1,,M2,sini-M2(l-1)2πM,i=M2,,M-1,(38)

and l=1,,M.

4. Numerical examples

In this section, we present some reconstructions obtained by applying the above reconstruction algorithm to simulated data. The set-up of our numerical tests mimicked closely the laboratory experiments presented in [Citation23,Citation24]. We considered a rectangular tank of length L=15cm, width W=15cm and height H=7.5cm which contained an isotropic medium of conductivity approximately equal to that of healthy breast tissue, σ0=200 mS/m. The rectangular array of electrodes was placed at the centre of the top of the tank (i.e. z=0 cm). The dimensions of the tank were much larger than those of the electrode array (7.2 cm×7.2 cm) and of the voxel configuration (12 cm×12 cm×1 cm) used for conductivity reconstructions, thus approximating an infinite half space.

Figure 3. Conductivity reconstructions for a cylindrical object of radius R25 of conductivity 800 mS/m placed at position P1 and z=-2 mm. The thick continuous line in (e) marks the position and the height of the inclusion along the z-axis.

Figure 3. Conductivity reconstructions for a cylindrical object of radius R25 of conductivity 800 mS/m placed at position P1 and z=-2 mm. The thick continuous line in (e) marks the position and the height of the inclusion along the z-axis.

Figure 4. Conductivity reconstructions for a cylindrical object of radius R25 of conductivity 800 mS/m placed at position P1 and z=-3 mm. The thick continuous line in (e) marks the position and the height of the inclusion along the z-axis.

Figure 4. Conductivity reconstructions for a cylindrical object of radius R25 of conductivity 800 mS/m placed at position P1 and z=-3 mm. The thick continuous line in (e) marks the position and the height of the inclusion along the z-axis.

Figure 5. Conductivity reconstructions for a cylindrical object of radius R25 of conductivity 800 mS/m placed at position P1 and z=-6 mm. The thick continuous line in (e) marks the position and the height of the inclusion along the z-axis.

Figure 5. Conductivity reconstructions for a cylindrical object of radius R25 of conductivity 800 mS/m placed at position P1 and z=-6 mm. The thick continuous line in (e) marks the position and the height of the inclusion along the z-axis.

Figure 6. Conductivity reconstructions for a cylindrical object of radius R35 of conductivity 800 mS/m placed at position P2 and z=-3 mm. The thick continuous line in (e) marks the position and the height of the inclusion along the z-axis.

Figure 6. Conductivity reconstructions for a cylindrical object of radius R35 of conductivity 800 mS/m placed at position P2 and z=-3 mm. The thick continuous line in (e) marks the position and the height of the inclusion along the z-axis.

Figure 7. Conductivity reconstructions for a cylindrical object of radius R25 of conductivity 800 mS/m placed at position P2 and z=-3 mm. The thick continuous line in (e) marks the position and the height of the inclusion along the z-axis.

Figure 7. Conductivity reconstructions for a cylindrical object of radius R25 of conductivity 800 mS/m placed at position P2 and z=-3 mm. The thick continuous line in (e) marks the position and the height of the inclusion along the z-axis.

Figure 8. Conductivity reconstructions for a cylindrical object of radius R15 of conductivity 800mS/m placed at position P2 and z=-3 mm. The thick continuous line in (e) marks the position and the height of the inclusion along the z-axis.

Figure 8. Conductivity reconstructions for a cylindrical object of radius R15 of conductivity 800mS/m placed at position P2 and z=-3 mm. The thick continuous line in (e) marks the position and the height of the inclusion along the z-axis.

Figure 9. Conductivity reconstructions for a cylindrical object of radius R25 of conductivity 800 mS/m placed at position P3 and z=-3 mm.

Figure 9. Conductivity reconstructions for a cylindrical object of radius R25 of conductivity 800 mS/m placed at position P3 and z=-3 mm.

Figure 10. Conductivity reconstructions for a cylindrical object of radius R25 of conductivity 800 mS/m placed at position P4 and z=-3 mm.

Figure 10. Conductivity reconstructions for a cylindrical object of radius R25 of conductivity 800 mS/m placed at position P4 and z=-3 mm.

Figure 11. Conductivity reconstructions for a cylindrical object of radius R35 of conductivity 800 mS/m placed at position P1 and z=-4 mm in the presence of a resistive medium (50 mS/m) in Voxel Layer 1. The thick continuous line in (e) marks the position and the height of the inclusion along the z-axis.

Figure 11. Conductivity reconstructions for a cylindrical object of radius R35 of conductivity 800 mS/m placed at position P1 and z=-4 mm in the presence of a resistive medium (50 mS/m) in Voxel Layer 1. The thick continuous line in (e) marks the position and the height of the inclusion along the z-axis.

Figure 12. Conductivity reconstructions for two cylindrical objects of radii R25 and R15 of conductivities 800 mS/m placed at positions P5 and P6, respectively, and z=-3 mm.

Figure 12. Conductivity reconstructions for two cylindrical objects of radii R25 and R15 of conductivities 800 mS/m placed at positions P5 and P6, respectively, and z=-3 mm.

Figure 13. Conductivity reconstructions for two cylindrical objects, one of radius R25 and of conductivity 600 mS/m, and the other of radius R15 and of conductivity 800 mS/m, placed at positions P5 and P6, respectively, and z=-3 mm.

Figure 13. Conductivity reconstructions for two cylindrical objects, one of radius R25 and of conductivity 600 mS/m, and the other of radius R15 and of conductivity 800 mS/m, placed at positions P5 and P6, respectively, and z=-3 mm.

Cylindrical inclusions of radii R15=1.5 mm, R25=2.5 mm and/or R35=3.5 mm, heights h=5 mm and conductivities σ=800 mS/m or 600mS/m were then placed on below positions P1-P5 in the xy-plane (see Figure ) and at different depths z. Note that the largest cylindrical object R35 has the same radius as the electrodes, while the other two, R25 and R15, have much smaller radii. P1, P2 and P5 in Figure are positions directly below a passive electrode, while P4 and P5 are positions between the passive electrodes.

To simulate the measured values of the potential on the boundary, we first had to solve the direct problem (Equation1)–(Equation8). In order to avoid inverse crimes and to test the robustness of our inversion techniques, we used EIDORS [Citation38] as a forward solver. EIDORS is a finite element software package which has no connection with the reconstruction method under consideration. Current patterns similar to those defined in Equation (Equation38) were applied at the active electrodes, Ik=I17k,,I36k, k=1,,L-M-1=19. The direct problem was solved for each of the 19 different current patterns and we obtain the values of the corresponding voltages at the passive electrodes Uk=U1k,,U16k, k=1,,19, by interpolating the numerical solution at points on the surface of the electrodes and evaluating numerically the integral in Equation (Equation9). This was the data used by our reconstruction algorithm. The conductivities of all N=260 voxels were obtained by inverting the over-determined linear system of Equations (Equation25). In the numerical examples considered, singular values smaller than ϵ=10-5 were cut off (i.e. 35 singular values were used in the reconstructions).

In EIT a noise level of 1% is reasonable in many circumstances, but in some medical applications greater accuracy can be achieved.[Citation39] However, since our reconstruction method is quite stable with respect to the noise level in the data, in all the numerical examples discussed below, we present reconstruction results for data with 2% additive Gaussian random errors. Note that in all figures presented below the colour ranges are the same for each subplot. Moreover, the homogeneity properties of the proposed algorithm are good as, although not displayed here, we were able to reconstruct a uniform distribution of conductivity of 200 mS/m using simulated data for the tank with no inclusions.

In Figures , we present the conductivity reconstructions for the medium size cylindrical object of radius R25 and conductivity 800mS/m (i.e. four times higher than the background conductivity) placed below position P1 (i.e. below an electrode and next to an active electrode) and at depths z=-2, -3 and -6mm, respectively, from the array of electrodes. The position of the inclusion in the xy-plane was successfully recovered in all three cases but the reconstructed conductivity values were smaller than the actual conductivity values. This difficulty in recovering the amplitude of high contrast conductivities is a common feature of EIT linearization methods, see for example [Citation20,Citation22,Citation40]. We also found that estimating the depth of the inhomogeneity was more ill-posed than reconstructing its position in the xy-plane. For an inclusion placed at this position, the values of the reconstructed conductivities at position P1 were larger in the upper voxel layers with a maximum value attained in the second layer irrespective of the depths of the inclusions. However, the deeper the object, the smaller the values of the reconstructed conductivity. This suggests that some information about the depth location of the inclusions is present in the data, but as seen from Figures and there is limited information about objects’ heights as it seems as if the inclusion extends over all five voxel layers. This conclusion is in agreement with our findings in [Citation24].

To test the spatial resolution of our algorithm, we placed cylindrical inclusions of conductivity 800 mS/m and of different radii R35, R25 and R15 at position P2 (i.e. below an electrode, but further away from the active electrodes) and depth z=-3 mm. The numerical results can be found in Figures , respectively. Similar to the previous numerical experiments, the position of the inclusion in the plane of electrodes was well characterized. The values of the reconstructed conductivities at position P2 were also smaller than the true ones, but overall larger in the upper voxel layers with a maximum value attained in the first layer in this case. Moreover, as expected, the smaller the size of the inhomogeneity, the smaller were the values of the reconstructed conductivity. Note that, by using the electrode array and the experimental set-up under consideration as well as the proposed three-dimensional reconstruction algorithm, we managed to detect inhomogeneities which are much smaller in size (i.e. a cylinder of radius R15 and height 5 mm whose conductivity is only four times higher than the background) and up to larger depths than in [Citation20,Citation23,Citation24].

When a cylindrical inhomogeneity of radius R25 and of conductivity 800 mS/m was placed on a position between two electrodes, P3, or in the middle of four neighbouring electrodes, P4, then the conductivity of all two or four adjacent voxels was much larger than that of the background, see Figures and .

Next we demonstrate that the detection of a small conductive inhomogeneity is not affected by the presence of a more resistive tissue layer (i.e. the skin) at the surface. To this end, we included a 2 mm thick resistive layer of conductivity 50 mS/m layer directly under the surface. A cylindrical object of radius R35 was positioned at P1 and z=-4 mm. As seen from the numerical reconstructions presented in Figure , the presence of the object is visible and its position in the xy-plane is well characterized.

The final simulations included in the paper consist of reconstructions of two conductive objects of radii R25 and R15 placed at positions P5 and P6, respectively, and at the same depth (z=-3 mm). Firstly, in Figure we present the results obtained when the two inclusions have the same conductivity (800 mS/m). In this case, the presence of the smaller cylindrical object (R15) is slightly shielded by the larger object (R25). Then, in Figure we show the reconstructions of objects of different conductivities (600 mS/m and 800 mS/m, respectively) when the presence of the smaller object is more pronounced.

5. Conclusion

In this article, we have presented a three-dimensional non-iterative reconstruction method developed for conductivity imaging in breast cancer detection using real data from a planar EIT device developed at the University of Mainz in collaboration with Oxford Brookes University. The head of the sensor contains both active electrodes, where standard trigonometric current patterns are applied, and passive electrodes, where the induced voltages are measured, arranged in a rectangular array. A finite region beneath the surface was discretized into voxels of different sizes depending on their position relative to the electrode array and their conductivities were determined from the data measured on the electrode array. The reconstruction algorithm is based on linearizing the conductivity about a constant value. It is simple, direct and fast, and it allows reconstructions in real-time.

The performance of the algorithm was tested on numerically simulated data. Small inclusions of various conductivities placed at several depths were detected and their positions in the plane of the electrode array were successfully recovered. Although the depth resolution is rather poor, the reconstructions have good spatial resolution in the xy-plane and are quite stable with respect to the noise level in the data. The most relevant feature is the fact that it can detect smaller objects up to larger depths than the other two-dimensional non-iterative approaches developed for similar planar EIT devices in [Citation23,Citation24]. In future, we plan to obtain reconstructions from real data and try to improve the depth resolution of the algorithm by possibly imposing a priori and/or a posteriori sparsity constraints on the reconstructed conductivity values in each voxel.

Acknowledgements

The authors would like to express their gratitude to Professor Khaled Hayatleh for a number of useful conversations concerning this work.

Notes

No potential conflict of interest was reported by the authors.

References

  • Borcea L. Electrical impedance tomography. Inverse Probl. 2002;18:R99–R136.
  • Lee CH, Dershaw DD, Kopans D, et al. Breast cancer screening with imaging: recommendations from the society of breast imaging and the ACR on the use of mammography, breast MRI, breast ultrasound, and other technologies for the detection of clinically occult breast cancer. J. Am. Coll. Radiol. 2010;7:18–27.
  • Rigaud B, Morucci JP, Chauveau N. Bioelectrical impedance techniques in medicine. Part I: bioimpedance measurement. Second section: impedance spectrometry. Crit. Rev. Biomed. Eng. 1996;24:257–351.
  • Jossinet J. Variability of impeditivity in normal and pathological breast tissue. Med. Biol. Eng. Comput. 1996;34:346–350.
  • Jossinet J. The impeditivity of freshly excised human breast tissue. Physiol. Meas. 1998;19:61–75.
  • Cherepenin V, Karpov A, Korjenevsky A, et al. A 3D electrical impedance tomography (EIT) system for breast cancer detection. Physiol Meas. 2001;22:9–18.
  • Cherepenin VA, Karpov AY, Korjenevsky AV, et al. Three-dimensional EIT imaging of breast tissues: system design and clinical testing. IEEE Trans. Med. Imaging. 2002;21:662–667.
  • Kerner TE, Paulsen KD, Hartov A, et al. Electrical impedance spectroscopy of the breast: clinical imaging results in 26 subjects. IEEE Trans Med Imaging. 2002;21:638–645.
  • Zou Y, Guo Z. A review of electrical impedance techniques for breast cancer detection. Med End Phys. 2003;25:79–90.
  • Kim BS, Isaacson D, Xia H, et al. A method for analyzing electrical impedance spectroscopy data from breast cancer patients. Physiol Meas. 2007;28:S237–S246.
  • Poplack SP, Tosteson TD, Wells WA, et al. Electromagnetic breast imaging: results of a pilot study in women with abnormal mammograms. Radiology. 2007;243:350–359.
  • Choi MH, Kao T-J, Isaacson D, Saulnier GJ, et al. A reconstruction algorithm for breast cancer imaging with electrical impedance tomography in mammography geometry. IEEE Trans. Biomed. Eng. 2007;54:700–710.
  • Trokhanova OV, Okhapkin MB, Korjenevsky AV. Dual-frequency electrical impedance mammography for the diagnosis of non-malignant breast disease. Physiol. Meas. 2008;29:S331–S344.
  • Murphy EK, Isaacson D, Saulnier GJ, et al. Analysis of forward solvers for electrical impedance tomography in a mammography geometry. J. Phys. Conf. Ser. 2010;224:012033.
  • Ardrey DB, Murphy EK, Isaacson D, et al. Electrical impedance tomography using the finite element method in the mammography geometry. In: Proceedings of IEEE 37th Annual Northeast Bioengineering Conference; Troy, NY; 2011. p. 1–2.
  • Sze G. Detection of breast cancer with electrical impedance mammography [PhD thesis]. Sussex: University of Sussex; 2012.
  • Pak DD, Rozhkova NI, Kireeva MN, et al. Diagnosis of breast cancer using electrical impedance tomography. Biomed. Eng. 2012;46:154–157.
  • Bilal R. Investigation of undesired errors relating to the planar array system of electrical impedance mammography for breast cancer detection [PhD thesis]. Sussex: University of Sussex; 2012.
  • Al Amin A, Parvin S, Kadir MA, et al. Classification of breast tumour using electrical impedance and machine learning techniques. Physiol. Meas. 2014;35:965–974.
  • Mueller JL, Isaacson D, Newell JC. A reconstruction algorithm for electrical impedance tomography data collected on rectangular electrode arrays. IEEE Trans. Biomed. Eng. 1999;46:1379–1386.
  • Azzouz M, Hanke M, Oesterlein C, et al. The factorization method for electrical impedance tomography data from a new planar device. Int. J. Biomed. Imaging. 2007;2007. 7 p. Article ID 83016.
  • Hähnlein C, Schilcher K, Sebu C, et al. Conductivity imaging with interior potential measurements. Inverse Probl. Sci. Eng. 2011;19:729–750.
  • Georgi K-H, Hähnlein C, Schilcher K, et al. Conductivity reconstructions using real data from a new planar electrical impedance device. Inverse Probl. Sci. Eng. 2013;21:801–822.
  • Gehre M, Kluth T, Sebu C, et al. Sparse 3D reconstructions in electrical impedance tomography using real data. Inverse Probl. Sci. Eng. 2014;22:31–44.
  • Lukaschewitsch M, Maass P, Pidcock MK, et al. The asymptotic behaviour of weak solutions to the forward problem of electrical impedance tomography on unbounded three dimensional domains. Math. Meth. Appl. Sci. 2009;32:206–222.
  • Lukaschewitch M. Inversion of geoelectric boundary data, a non-linear ill-posed problem [PhD thesis]. Potsdam: University of Potsdam; 1999.
  • Lukaschewitch M, Maass P, Pidcock M. Tikhonov regularization for electrical impedance tomography on unbounded domains. Inverse Probl. 2003;19:585–610.
  • Calderón AP. On an inverse boundary value problem. In: Seminar on numerical analysis and its applications to continuum physics. Rio de Janeiro: Sociedade Brasileira de Matematica; 1980. p. 67–73.
  • Kervokian J. Partial differential equations: analytical solutions techniques. 2nd ed. New York (NY): Springer-Verlag New York; 2000.
  • Hansen PC. Regularization tools: a Matlab package for analysis and solution of discrete ill-posed problems. Numer. Algorithms. 1994;6:1–35.
  • Assenheimer M, Laver-Moskovitz O, Malonek D, et al. The T-SCANTM technology: electrical impedance as a diagnostic tool for breast cancer detection. Physiol. Meas. 2001;22:1–8.
  • Kotre CJ. Subsurface electrical impedance imaging: measurement strategy, image reconstruction and in vivo results. Physiol. Meas. 1996;17:A197–204.
  • Kemna A, Binley A, Slater L. Crosshole IP imaging for engineering and environmental applications. Geophysics. 2004;69:97–107.
  • Tutt B. New partial-breast radiation therapy regimen uses protons. Oncolog. 2013;58:4–5.
  • March D, Coughlin B, Barham R, et al. Breast masses: removal of all US evidence during biopsy by using a handheld vacuum-assisted device-initial experience. Radiology. 2003;227:549–555.
  • Gradshteyn IS, Ryzhik M. Table of integrals, series and products. 4th ed. New York (NY): Academic Press Inc.; 1965.
  • Brebbia CA, Telles JCF, Wrobel LS. Boundary element techniques. Berlin: Springer-Verlag; 1984.
  • EIDORS: Electrical impedance tomography and diffuse optical tomography reconstruction software [computer software]. Available from: http://eidors3d.sourceforge.net/.
  • Terzopoulos N, Hayatleh K, Hart B, et al. A novel bipolar-drive circuit for medical applications. Physiol. Meas. 2005;26:N21–N27.
  • Knudsen K, Lassas M, Mueller J, et al. Reconstructions of piecewise constant conductivities by the D-bar method for electrical impedance tomography. J. Phys.: Conf. Ser. 2008;124:012029.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.