You are on page 1of 22

Home Search Collections Journals About Contact us My IOPscience

A new non-iterative inversion method for electrical resistance tomography

This content has been downloaded from IOPscience. Please scroll down to see the full text.

2002 Inverse Problems 18 1809

(http://iopscience.iop.org/0266-5611/18/6/323)

View the table of contents for this issue, or go to the journal homepage for more

Download details:

IP Address: 128.6.218.72
This content was downloaded on 20/08/2015 at 12:08

Please note that terms and conditions apply.


INSTITUTE OF PHYSICS PUBLISHING INVERSE PROBLEMS
Inverse Problems 18 (2002) 1809–1829 PII: S0266-5611(02)37066-7

A new non-iterative inversion method for electrical


resistance tomography
A Tamburrino and G Rubinacci
DAEIMI, Laboratory of Computational Electromagnetics and Electromagnetic Non-Destructive
Evaluation, Università degli Studi di Cassino, via G di Biasio, 43-03043 Cassino, Italy

E-mail: tamburrino@unicas.it

Received 16 May 2002, in final form 22 October 2002


Published 8 November 2002
Online at stacks.iop.org/IP/18/1809

Abstract
In this paper, the inverse problem of resistivity retrieval is addressed in the
frame of electrical resistance tomography (ERT). The ERT data is a set of
measurements of the dc resistances between pairs of electrodes in contact with
the conductor under investigation. This paper is focused on a non-iterative
inversion method based on the monotonicity of the resistance matrix (and of
its numerical approximations). The main features of the proposed inversion
method are its low computational cost requiring the solution of O(n) direct
problems, where n is the number of parameters used to represent the unknown
resistivity, and its very simple numerical implementation.

1. Introduction

Electrical resistance tomography (ERT) deals with the problem of reconstructing the resistivity
of a conducting domain D from the knowledge of the potentials and currents on a discrete set
of electrodes placed at the boundary ∂ D of D.
The problem has been discussed by many authors from the mathematical, numerical and
experimental points of view. The knowledge of the Dirichlet-to-Neumann map, i.e. the relation
between scalar potential and current density at the boundary, uniquely identifies the resistivity
for error-free data [1–5].
Many approaches for solving this inverse problem are based on the well known first-order
approximation consisting in linearizing the operator mapping the resistivity into the boundary
data. We refer to [6] for a list of references, also covering other classes of methods such
as iterative minimization for various error functionals, which are mainly based on Newton’s
method [7].
In [8] we enlarged the set of retrievable resistivities (with respect to methods based on the
first-order approximation) in the frame of a quadratic approach. Moreover, we also gave an
estimate of the maximum number of linearly independent equations that can be derived from
0266-5611/02/061809+21$30.00 © 2002 IOP Publishing Ltd Printed in the UK 1809
1810 A Tamburrino and G Rubinacci

the model and we showed that no local minima occur when the ratio between the number of
linearly independent equations and the unknown parameters is sufficiently large [8, 9].
This paper makes two main contributions. The first of these regards the proof of the
monotonicity of the resistance matrix1 . Specifically, we show that2
ρ1 (r )  ρ2 (r ) in D ⇒ Rρ1  Rρ2 (1)
where Rρk is the resistance matrix related to the resistivity ρk . Similar results have been
previously obtained for the Neumann-to-Dirichlet map [10] under the assumption that the
resistivity and its reciprocal are elements of L ∞ (D). Here we prove the monotonicity of
the resistance matrix, which involves a mixed boundary condition problem (see section 2),
and for a resistivity ρk that is allowed to be either vanishing or infinite in some volume
contained in D. In addition, we introduce two complementary numerical formulations and
solve the forward problem by a finite element method based on nodal and edge-element shape
functions [8]. Moreover, we show that the numerically computed resistance matrix inherits
the same monotonicity property.
The second main contribution of the paper is that it presents a non-iterative method,
based on monotonicity, for solving the inverse problem. Specifically we address the problem
of retrieving the shape of the domain  ⊂ D of a homogeneous inclusion of resistivity ρc
embedded in a homogeneous conductor of resistivity ρ0 . This non-iterative inversion method
reconstruct two sets ˜ int and 
˜ ext that satisfy ˜ int ⊆  ⊆ 
˜ ext for error-free measurements.
This method, which to our knowledge is original, is an improvement of the method that
we proposed in [11] where only  ˜ ext was retrieved. The improvement is relevant because
numerical simulations have shown that  ˜ int approximates  better than  ˜ ext .
Non-iterative methods appear to be an interesting alternative to iterative methods based
on the minimization of an error functional related to the distance, in a suitable norm,
between numerically computed and measured boundary values. These iterative approaches,
representing the state of the art for non-linear inverse problems, require the solution of the
direct problem for several assigned tentative shapes of the inclusion and can be very expensive
in terms of computer time. Moreover, the convergence cannot be guaranteed.
The majority of studies devoted to non-iterative approaches are based on an idea of Colton
and Kirsch developed in the frame of inverse scattering problems [12, 13]. In those papers,
they characterize the support of an obstacle with the range of a suitable operator. Non-iterative
methods include the work of Ikeata to recover the convex hull of polygonal cavities [14]
and the work of Brühl [15, 16]. Brühl introduces a dipole singularity, associated with some
Neumann function, for the Laplace equation. Its boundary values are contained in the range
of the Dirichlet-to-Neumann map if and only if the dipole singularity lies in . All these
methods are based on elegant mathematical constructions which require very careful numerical
implementations. In fact, most of them are based on the knowledge of the Dirichlet-to-
Neumann map and need an infinite number of different measurements to determine this map.
Moreover, as stated in [16], a robust numerical convergence criterion of some infinite series,
the terms of which are subject to noise, is required.
In this paper, we construct a non-iterative inversion method starting from a completely
different standpoint. Specifically, from the monotonicity of the resistance matrix (see (1)),
we obtain that, if Rρ1 − Rρ2 is not positive semi-definite, then ρ1 (r )  ρ2 (r ) in D is false.
1 We recall that the resistance matrix Rρk relates the currents and voltages at the electrodes located on the boundary
of the conductor D by V = Rρk I , where I  [i1 , . . . , i M ]T ,V  [v1 , . . . , v M ]T , ik ∈ R is the current injected at
the kth electrode, vk ∈ R is the voltage of the kth electrode with respect to the (M + 1)th electrode and M + 1 is the
number of electrodes. Similarly, the conductance matrix relates V and I by I = Gρk V .
2 Note that, if A and B are square matrices, A  B means that A − B is a positive semi-definite matrix.
A new non-iterative inversion method for electrical resistance tomography 1811

Therefore, by checking if R̃− Rk is positive semi-definite, where R̃ is the measured resistance
matrix (the data) and Rk is the resistance matrix related to the resistivity

ρc in k
ρk (r ) =
ρ0 in D\k ,
it is possible to reconstruct the shape of the anomaly . In addition, this approach can be
naturally extended to non-homogeneous background and/or inclusion, which is the case when
ρ0 and/or ρc depend on the spatial coordinate r .
This paper is organized as follows. In section 2 the mathematical and numerical models for
the steady ohmic conduction are described. Starting from the two complementary formulations,
several different numerical models for computing the resistance matrix are described. In
section 3 the monotonicity is shown for the resistance matrix and its numerical approximations.
Firstly the case when the resistivity and its reciprocal are elements of L ∞ (D) is addressed.
Then, we examine the case of perfectly insulating or conducting anomalies. In section 4 the
inversion method is proposed and validated by numerical examples.
Finally, we notice that any inequality involving the resistance matrix has an equivalent
form in terms of the conductance matrix. Therefore the proposed approach can be applied to
the conductance matrix as well.

2. Formulation of the direct problem

In this section we briefly describe the mathematical modelling of the direct problem, in line
with [8] where additional details can be found. The first subsection refers to the modelling
of steady ohmic conduction when voltages or currents at the electrodes are fixed. The second
subsection refers to the modelling of the resistance and conductance matrices.
The reference problem is schematically described in figure 1 where a three-dimensional
bounded ohmic material is hosted in a non-conductive medium and is accessible by means
of a finite number of electrodes located at its boundary. We assume that each electrode can
be held to be equipotential (we neglect the contact resistance), in the frame of the so-called
‘shunt model’, although this is not restrictive for our analysis. The ‘complete model’, where the
electrodes are modelled by an impedance condition, turns out to be very accurate for predicting
experimental data [17]. The existence and uniqueness of the solution for the complete model
has been proved in [18].

2.1. Field equations, variational and weak form characterizations


The field equations describing the steady ohmic conduction are
∇ ×E =0 in D (2)
∇ ·J =0 in D (3)
ρ(r )J (r ) = E (r ) in D (4)
J ·n=0 on S⊥ (5)
E ×n=0 on S1 ∪ · · · ∪ S M+1 (6)

E · d l = vk , k = 1, . . . , M (7)
γk

J · n dS = i k , k = 1, . . . , M (8)
Sk
1812 A Tamburrino and G Rubinacci

Figure 1. Reference configuration: cross section of the conductive body together with the
electrodes.

where E and J are the electric field and the current density, respectively, ρ is the resistivity, D is
the conducting domain, Sk is the part of the boundary S = ∂ D in contact with the kth electrode,
S⊥ is the part of the boundary in contact with the host medium (S = S⊥ ∪ S1 ∪ · · · ∪ SM+1 ), n
is the inward normal on ∂ D, γk is a line contained in D connecting Sk to SM+1 , i k ∈ R is the
current at the kth electrode and vk ∈ R is the voltage of the kth electrode with respect to the
(M + 1)th.
To take into account the continuity of the tangential component of the electric field and of
the normal component of the current density at the interfaces where the conductivity has first
kind discontinuity, (2) and (3) must be considered in the weak form (see [19]).
When 0 < ρmin  ρ  ρmax a.e. in D, the electric field E belongs to the functional space
E defined as (see [19])
E  {E ∈ L2rot (D)|∇ × E = 0, n × E = 0 on S1 ∪ · · · ∪ S M+1 } (9)
whereas the current density J belongs to the functional space J defined as
J  {J ∈ L2div (D)|∇ · J = 0, n · J = 0 on S⊥ } (10)
where
L2rot (D)  {a ∈ L2 (D)|∇ × a ∈ L2 (D)}
L2div (D)  {a ∈ L2 (D)|∇ · a ∈ L 2 (D)}.

It is worth noting that the line integral γk E · dl is well defined for any E ∈ E and the surface

integral Sk J · n dS is well defined for any J ∈ J (see [19]).
The variational characterizations of problem (2)–(8) in terms of electric scalar potential
(E = −∇ϕ) and electric vector potential (J = ∇ × T ) are

find ϕ ∈ ΦV minimizing σ
∇ϕ
2 dr (11)
 D

find T ∈ A I minimizing ρ
∇ × T
2 dr (12)
D
A new non-iterative inversion method for electrical resistance tomography 1813

where σ = ρ −1 is the conductivity, the spaces ΦV and A I are defined as


ΦV  {ϕ ∈ Φ|ϕ | S M+1 = 0, ϕ | Sk = vk , k = 1, . . . , M} (13)
  


A  T ∈ A
I 
∇ × T · n dS = i k , k = 1, . . . , M (14)
Sk

Φ  {ϕ ∈ L 2grad (D)|n × ∇ϕ = 0 on S1 ∪ · · · ∪ S M+1 }


A  {T ∈ L2rot (D)|n · ∇ × T = 0 on S⊥ }
L 2grad (D)  {ϕ ∈ L 2 (D)|∇ϕ ∈ L2 (D)}

and V  [v1 , . . . , v M ]T , I  [i 1 , . . . , i M ]T are the vectors of the electrode voltages and


currents.
The Euler equations (weak form) of (11) and (12) are

find ϕ ∈ Φ such that
V
σ ∇ϕ · ∇ϕ dr = 0, ∀ϕ ∈ Φ0 (15)
D
find T ∈ A I such that ρ∇ × T · ∇ × T dr = 0, ∀ T ∈ A0 (16)
D

where Φ0 and A0 are the spaces obtained by assuming vk = 0 and i k = 0 ∀k = 1, . . . , M


in (13) and (14), respectively.
It is worth noting that Φ0 and A0 are linear spaces and
ΦV = ϕ V + Φ0 (17)
A =a +A
I I 0
(18)
where ϕ V and a I are arbitrary elements of ΦV and A I , respectively.
The numerical formulation referred to throughout the paper for computing the electric
scalar potential is obtained from (15) (or, equivalently, from (11)) by replacing Φ0 with a
finite-dimension linear subspace Φ0m ⊂ Φ0 :

find ϕm ∈ Φm such that
V
σ ∇ϕm · ∇ϕ dr = 0, ∀ϕ ∈ Φ0m (19)
D

find ϕm ∈ ΦmV minimizing σ
∇ϕ
2 dr (20)
D

where ΦmV  ϕ V + Φ0m . The scalar potential is approximated by means of standard


isoparametric shape functions defined onto a finite element mesh used to discretize D.
Analogously, the numerical formulation for computing the electric vector potential is
obtained from (16) (or, equivalently, from (12)) by replacing A0 with a finite-dimension linear
subspace A0m ⊂ A0 :

find Tm ∈ Am such that
I
ρ∇ × Tm · ∇ × T dr = 0, ∀T ∈ A0m (21)
D

find Tm ∈ AmI minimizing ρ
∇ × T
2 dr (22)
D

where AmI= a + I
A0m .
The electric vector potential Tm is approximated by means of an
edge-element-based shape function defined onto the finite element mesh used to discretize D.
The uniqueness of Tm is enforced by using a tree–cotree decomposition of the finite element
mesh [20].
1814 A Tamburrino and G Rubinacci

2.2. The resistance and conductance matrices


In linear problems, as in the present case, the electrode currents and voltages are related by a
linear transformation:
V = Rρ I (23)
where Rρ is the well known resistance matrix that is symmetrical when ρ is a symmetrical
tensor. Moreover, the resistance matrix is related to the ohmic power dissipated in the conductor
by

I Rρ I =
T
ρ
∇ × T
2 dr . (24)
D
This follows from (4), (23), J = ∇ × T and the identity

V I=
T
E · J dr (25)
D
which, as is well known, can be obtained by exploiting ∇ × E = 0 and ∇ · J = 0 in D.
Assuming as usual that ρ > 0, Rρ turns out to be a positive definite matrix, as follows
from (24). Therefore, Rρ is invertible and its inverse is the symmetrical and positive definite
conductance matrix
I = Gρ V . (26)
As for the resistance matrix, the conductance matrix is related to the ohmic losses in the
conductor: 
V Gρ V =
T
ρ −1
∇ϕ
2 dr . (27)
D
This follows from (4), (26), (25) and E = −∇ϕ.
The above summary of the basic definitions and properties of the resistance and
conductance matrices forms the guidelines for building consistent numerical methods for their
computation.
Specifically, we define the T -based numerical approximation of the resistance matrix RρA
as the symmetrical matrix giving the numerically computed ohmic power obtained by replacing
T on the rhs of (24) with its numerical approximation, i.e. RρA is defined through3

I A Rρ I A =
T A
ρ
∇ × Tm
2 dr (28)
D
where Tm is the solution of (21) or (22) when the currents vector is I A . In addition, we define
the conductance matrix GρA and the electrode voltages vector V A as
GρA  (RρA )−1 (29)
V A  RρA I A .
In this way, (23), (25) and (26) are automatically imposed. It is worth noting that in a T -based
numerical scheme the currents are given (see (22)) whereas the voltages are a numerical result.
Analogously, within the numerical scheme based on the electric vector potential
formulation, we define the -based numerical approximation of the conductance matrix
through4

T 
V Gρ V = ρ −1
∇ϕm
2 dr (30)
D
3 I A is the given vector of electrode currents. The subscript A is specified as a reminder that we are considering
quantities in the electric vector potential numerical formulation.
4 V is the given vector of electrode voltages. The subscript  is specified as a reminder that we are considering

quantities in the electric scalar potential numerical formulation.
A new non-iterative inversion method for electrical resistance tomography 1815

where ϕm is the solution of (19) or (20) when the voltage vector is V . In addition, we define
the resistance matrix R
ρ and the currents vector I as

R  −1
ρ  (G ρ ) (31)
I  G
ρ V .
Another numerical approximation of Rρ can be obtained by combining the scalar and
vector potential formulations, as proposed in [8] where we showed that
R
ρ  Rρ  Rρ .
A

Specifically, we define the resistance matrix Rρ R as


RρA + R
ρ
Rρ R  . (32)
2
The matrix Rρ R is the baricentric matrix in the set of matrices M satisfying R
ρ  M  Rρ .
A

We verified numerically that the numerical errors affecting Rρ and Rρ tend to compensate.
A

Thus Rρ R tends to be a better approximation of Rρ . Consistently with (32) we define the


conductance matrix Gρ R as
Gρ R  ( Rρ R )−1 .
Dually, it is possible to prove that
GρA  Gρ  G
ρ.
Thus we introduce Gρ G , another approximation of Gρ , defined as
G
ρ + Gρ
A
Gρ G 
2
and, consistently, Rρ G :
Rρ G  ( Gρ G )−1 .

3. Monotonicity for the resistance matrix

In this section we analyse the monotonicity of the resistance matrix. Specifically, we prove
that
ρ1 (r )  ρ2 (r ) in D ⇒ R̂ρ1  R̂ρ2 (33)
where R̂ρk is either the actual resistance matrix or the resistance matrix numerically computed
with one of the methods described in section 2.2. We allow ρk to be either vanishing or infinite
in some domains of D. Specifically, we assume that
0 < ρmin  ρk (r )  ρmax < +∞ a.e. in D\(0k ∪ ∞
k )
and that ρk is vanishing in the domain 0k and infinite in the domain ∞
k .
We first prove (33), assuming that 0k and ∞k are void. Then we prove the monotonicity
of the resistance matrix, assuming that ρ1 and ρ2 are infinite (or vanishing) in the domains
1 and 2 , respectively, and equal elsewhere. These two particular cases make it possible
to illustrate the mathematical techniques for dealing with the more general situation. It is
worth noting that the numerical models are obtained by replacing a functional space with a
finite-dimensional linear subspace. This allows the same techniques to be used to analyse the
monotonicity for both the actual resistance matrix and its numerical approximations.
In the following we will extensively use the following proposition:
Proposition 1. Let A and B be two real, symmetrical and invertible matrices. It turns out
that A  B ⇔ A−1  B −1 .
1816 A Tamburrino and G Rubinacci

3.1. Monotonicity for resistivities bounded a.e.


Here we assume that
0 < ρmin  ρk (r )  ρmax < +∞ a.e. in D, k = 1, 2. (34)
We initially derive some basic inequalities and, then, other similar inequalities by matrix
algebra. These results are based on well known principles and, to our knowledge, have been
first derived for the resistance matrix defined under the boundary conditions (5) and (6) in [11].
The basic inequalities following directly from the variational characterizations (12), (20)
and (22) and from (24), (28) and (31) are
ρ1 (r )  ρ2 (r ) in D ⇒ Rρ1  Rρ2 (35)
ρ1 (r )  ρ2 (r ) in D ⇒ RρA1  RρA2 (36)
ρ1 (r )  ρ2 (r ) in D ⇒ R 
ρ1  R ρ2 . (37)
To prove (35) let us fix the electrode current vector I . Let T1 and T2 be the solutions of
(12) when the resistivity is ρ1 and ρ2 , respectively. Then

I T R ρ1 I = ρ1
∇ × T1
2 dr
D
 ρ2
∇ × T1
2 dr
 D

 ρ2
∇ × T2
2 dr
D
= I T R ρ2 I
where the first and last lines follow from (24), the second line from the assumption ρ1  ρ2
in D and the third line from (12). The thesis follows from the arbitrariness of I . It is worth
noting that the problem with resistivity ρ1 and that with resistivity ρ2 share the same electrode
current vector I . This implies that the minimization involved in (12) is performed on the same
set A I in both cases.
The proof of (36) can be obtained by using the same technique whereas the proof of (37)
can be obtained firstly by showing that ρ1 (r )  ρ2 (r ) in D ⇒ G 
ρ1  Gρ2 and then taking
into account definition (31) and proposition 1.
Finally, matrices Rρ R and Rρ G satisfy the following inequalities:
ρ1 (r )  ρ2 (r ) in D ⇒ Rρ1 R  Rρ2 R (38)
ρ1 (r )  ρ2 (r ) in D ⇒ Rρ1 G  Rρ2 G . (39)
To prove (38) we notice that, when ρ1 (r )  ρ2 (r ) in D, then
RρA1 + R
ρ1
Rρ1 R =
2
RρA2 + R
ρ1

2
RρA2 + R
ρ2

2
= Rρ2 R ,
where the first inequality comes from (36) and the second from (37). The proof of (39) can be
obtained by using the same technique.
A new non-iterative inversion method for electrical resistance tomography 1817

3.2. Monotonicity for resistivities with perfectly insulating or conducting inclusions


In this section we tackle the problem of deriving inequalities involving the resistance and
conductance matrices related to an inclusion in c ⊂ D. We analyse the problems of a
perfectly insulating inclusion and a perfectly conducting inclusion.
We assume that ρ, the resistivity of the conductor in the absence of the inclusion, satisfies
0 < ρmin  ρ(r )  ρmax < +∞ a.e. in D. In addition, we assume that c , the closure of
c , satisfies condition c ∩ ∂ D = ∅, i.e. the inclusion is not in contact with any electrode
and does not completely surround any electrode. This last assumption avoids the presence of
either infinite or vanishing entries in the resistance and conductance matrices.
Specifically we show that for perfectly insulating inclusions it turns out that
1 ⊆ 2 ⇒ R̂1  R̂2 (40)
where R̂k is either the actual resistance matrix or the numerically computed resistance matrix,
corresponding to the resistivity

ρ(r ) for r ∈ D\k
ρk (r ) =
+∞ for r ∈ k .
Dually, for perfectly conducting inclusions it turns out that
1 ⊆ 2 ⇒ R̂1  R̂2 (41)
where R̂k is either the actual resistance matrix or the numerically computed resistance matrix,
corresponding to the resistivity

ρ(r ) for r ∈ D\k
ρk (r ) =
0 for r ∈ k .
The results of this section cannot be obtained from those of section 3.1 because in c we are
assuming either a vanishing or an infinite resistivity ((34) is not satisfied). The combination
of the techniques of this section and 3.1 gives the complete mathematical framework for
deriving (33) in its generality.
In the following we modify the variational formulation when a perfectly insulating or
conducting inclusion is present.

3.2.1. Variational formulations in the presence of an inclusion. Here we consider the


variational formulations for perfectly insulating and conducting inclusions, in view of the
numerical solution of the inverse problem. Specifically, we take into account that the numerical
solution can be efficiently obtained, using the finite element method, by discretizing the whole
domain D and adding some linear constraints specifying the anomaly in c . This may appear
cumbersome if compared to the natural approach consisting of discretizing D\c only and
imposing proper boundary conditions on ∂c . However, when many direct problems have
to be solved for different choices of c , as needed for solving the inverse problem, the latter
approach entails remeshing the domain D\c and recomputing the stiffness matrices. For
this reason the approach based on the discretization of the whole domain D is very efficient
in terms of computational cost, as highlighted in [22] for a zero-thickness perfectly insulating
crack and in [23] for perfectly insulating volumetric cracks.
The variational formulations modelling a perfectly insulating inclusion are

find ϕ ∈ ΦV minimizing σ
∇ϕ
2 dr (42)
D\c

find T ∈ AI c minimizing ρ
∇ × T
2 dr (43)
D\c
1818 A Tamburrino and G Rubinacci

where σ = ρ −1 and
A
I
c
 {T ∈ A I |∇ × T = 0 in c }. (44)
The ohmic power dissipated in c is vanishing. Thus the resistance and conductance
matrices can be obtained by means of

I T Rc I = ρ
∇ × T
2 dr (45)
D\c

V T Gc V = σ
∇ϕ
2 dr (46)
D\c
where T and ϕ are solutions of (42) and (43), respectively.
We stress again that knowledge of the solutions in c is not strictly required (it does
not affect the resistance and conductance matrices, see (45) and (46)). However, having
variational formulations where the solution spaces consist of functions defined on D instead
of on D\c makes it possible to overcome unnecessary technical difficulties in deriving the
proof of inequalities for the resistance and conductance matrices.
The numerical formulations for problems (42) and (43) can be obtained by replacing ΦV
with ΦmV and A I with AmI in (42) and (44), respectively, as in section 2.1.
However, a few additional considerations are required. First of all we recall that we
approximate the scalar potential by means of the linear combination of standard isoparametric
shape functions defined on the finite element mesh discretizing D. Each shape function
is associated with a node of the mesh and its support consists of those elements having in
common that node. The vector potential is approximated by means of the linear combination
of edge-element-based shape functions defined on the finite element mesh. Each shape function
is associated with an edge of the mesh and its support consists of those elements having in
common that edge. The uniqueness of the vector potential is enforced by using a tree–cotree
decomposition of the finite element mesh [20]. Then, we assume that c is the union of
elements of the finite element mesh. This assumption5 allows us to avoid the remeshing of
D\c . Finally, in order to guarantee c ∩ ∂ D = ∅, we assume that c is the union of elements
that are not on the boundary ∂ D.
With reference to the numerical formulation of problem (42), we notice that the solution
space ΦmV can be naturally reduced to a smaller space. In fact, the scalar unknowns
corresponding to nodes contained in the interior of c are not involved in the minimization.
With reference to problem (43), we notice that the constraint ∇ × T = 0 in c can be satisfied
in AmI (it is enough set to zero the coefficient of the shape functions associated with edges that
belong to c ). Therefore, the set obtained by replacing A I with AmI in the definition of AI c is
not empty. Therefore, the numerical formulations for (42) and (43) are6

find ϕm ∈ ΦmV minimizing σ
∇ϕ
2 dr (47)
D\c

find Tm ∈ Am,I
c
minimizing ρ
∇ × T
2 dr (48)
D\c
where
  


I 
Am,
I
c
 T ∈ Am 
∇ × T
2 dr = 0 .
c
5 Under this assumption, the mesh M obtained by removing from the original mesh M (the finite element mesh
of D) the elements corresponding to Vc , is a mesh for D\Vc . Consequently, a shape function defined onto M can be
represented in D\Vc as a linear combination of shape functions defined onto M.
6 Notice that the constraint ∇ × T = 0 in V has been replaced with 2
c Vc
∇ × T
dr = 0.
A new non-iterative inversion method for electrical resistance tomography 1819

We stress again that this allows the numerical problem to be solved indifferently on the
whole domain D (setting proper linear constraints) or on D\c .
Similarly, the variational formulations modelling a perfectly conducting inclusion are

find ϕ ∈ Φc minimizing
V
σ
∇ϕ
2 dr (49)
D\c

find T ∈ A I minimizing ρ
∇ × T
2 dr (50)
D\c
where
Φ
V
c
 {ϕ ∈ ΦV |∇ϕ = 0 in c }, (51)
and the related numerical formulations are 
find ϕm ∈ Φm,
V
c
minimizing σ
∇ϕ
2 dr (52)
D\c

find Tm ∈ AmI minimizing ρ
∇ × T
2 dr (53)
D\c
where   
V

Φm,
V
c
 ϕ ∈ Φm 
∇ϕ
2 dr = 0 .
c
Notice that (45) and (46) hold in the same form also for a perfectly conducting inclusion.
Finally, the numerically computed resistance and conductance matrices RA c and G c can
be obtained by replacing T and ϕ with their numerical approximations Tm and ϕm in (45) and
(46) whereas the numerical approximations Rc R , Rc G are obtained by combining the
two complementary formulations as in section 2.2.

3.2.2. Inequalities for perfectly insulating and perfectly conducting inclusions. With
reference to perfectly insulating inclusions, we notice that, as for volumetric identification
(see section 3.1), (40) can be achieved in generality by simply showing
1 ⊆ 2 ⇒ R1  R2 (54)
1 ⊆ 2 ⇒ RA 1  RA 2 (55)
1 ⊆ 2 ⇒ R 
1  R2 . (56)
To prove (54) let us fix the electrode current vector I . Let T1 and T2 be the solutions of
(43) for a crack in 1 and 2 , respectively, it turns out that

I T R2 I = ρ
∇ × T2
2 dr
D\2

= ρ
∇ × T2
2 dr
D\1

 ρ
∇ × T1
2 dr
D\1

= I T R1 I
where the first and last lines follow from the definition of the resistance matrix, the second
line from ∇ × T2 = 0 in 2 ⊃ 1 , the third line from T2 ∈ AI 2 ⊂ AI 1 and the variational
characterization (43) of T1 . Finally, (54) follows from the arbitrariness of I . The proof of (55)
and (56) can be obtained by using the same technique.
In a similar way it is possible to prove inequality (41) that holds for perfectly conducting
inclusions.
1820 A Tamburrino and G Rubinacci

Figure 2. (a) The finite element mesh (48 × 48 × 1 hexahedral elements) discretization of the
domain D. (b) The subdivision (8 × 8 × 1 hexahedral elements) of D into τk . A τk is made of
6 × 6 × 1 elements of the finite element mesh.

4. Inversion method

In this section we propose a non-iterative inversion method based on inequality (33) for solving
the problem of shape identification of volumetric cracks. This method reconstructs two sets
˜ ext and  ˜ int . For error-free measurements, these two sets satisfy ˜ int ⊆  ⊆  ˜ ext ⊆ D
where  is the region containing the cracks and D is the conducting domain.
This method is described by steps of increasing difficulty, firstly showing the basic
idea underlying the inversion method for a simple but practical configuration consisting of
retrieving homogeneous cracks in a homogeneous conductor from noiseless measurements of
the resistance matrix. Then we tackle the problem of the measurement noise and the error
introduced by representing the inclusion volume as the union of elements of the finite element
mesh.
In the following, we assume that ρ0 and ρc are the background resistivity and the inclusion
resistivity, respectively, that 0 < ρ0 < ρc  +∞ and that D is discretized by a finite element
mesh. In addition, we assume that D is subdivided into N ‘small’ non-overlapped parts
τ1 , . . . , τ N , and that τk is made of the union of contiguous elements of the finite element mesh
(see figure 2).
Let ρ S (r ) be the resistivity defined as

ρ0 for r ∈ D\S
ρ S (r ) 
ρc for r ∈ S,

where S ⊆ D, and let R̂ S be the resistance matrix (either the actual resistance matrix or the
resistance matrix numerically computed) related to ρ S .
Let R̃ be the data, i.e. the resistance matrix processed by the inversion algorithm.

4.1. Fundamentals of the inversion method


In this section we assume that the crack volume  is the union of a number of τk and that R̃
is noise free7 , i.e. R̃ = R̂ .
7 When the resistance matrices are numerically computed, the assumption R̃ = R̂ means that the data are free from
the model error, i.e. the data are elements of the range of the mapping S → R̂ S .
A new non-iterative inversion method for electrical resistance tomography 1821

Figure 3. The resistance matrix of the two systems reduces to the same scalar value [ρ0 (L − ) +
ρc ] A−1 , where L is the length of the resistor and A is its cross-sectional area, independent of the
position of the inclusion.

We notice that from (33) it turns out that


1 ⊆ 2 ⊆ D ⇒ R̂1 − R̂2  0 (57)
or, by exchanging the role of 1 and 2 ,
2 ⊆ 1 ⊆ D ⇒ R̂1 − R̂2  0. (58)
Therefore, we have

Proposition 2. If either 1 ⊆ 2 or 2 ⊆ 1 the sign of R̂1 − R̂2 is well defined8


or, by reversing proposition 2,

Proposition 3. If the sign of R̂1 − R̂2 is not well defined then 1  2 and 2  1 .

It is worth noting in (57) that R̂1 − R̂2  0 does not imply that 1 ⊆ 2 . For example,
consider the resistance of the system in figure 3. The resistance does not depend on the position
x of the crack. Thus, although R̂1 = R̂2 , the two sets 1 and 2 are unrelated.
The inversion algorithm is based on proposition 3. The main idea is to use proposition 3
for reducing the region containing the cracks. We introduce the set of test resistance matrices
R̂τ1 , . . . , R̂τ N and, thanks to proposition 3, if the sign of R̃ − R̂τk is not well defined then τk
is not part of . Therefore we define  ˜ ext , the first estimate of , as9

˜ ext 
 τk , Iext  {k|sk = 1, k = 1, . . . , N}
k∈Iext

where sk is the sign index of R̃ − R̂τk defined as10



j λk, j
sk  
j |λk, j |

˜ ext can be
and λk, j is the j th eigenvalue of R̃ − R̂τk . It is worth noting that, by construction, 
larger than , i.e.  ⊆  ˜ ext .
Then, starting from  ˜ ext we construct a set  ˜ int that is contained in :

˜ int  τk , Iint  {k ∈ Iext |tk < 1}
k∈Iint

˜ ext is not part


where tk is the sign index for R̂˜ ext \τk − R̃. It is worth noting that, if τk ⊆ 
˜
of , then R̂˜ ext \τk − R̃ is positive semi-definite. Therefore, int ⊆  by construction.
8 The sign of a matrix A is well defined when either A  0 or A  0.
9 Note that  is never contained in τk , thus R̃ − R̂τk is never a negative semi-definite matrix.
10 The sign index s assumes values in [−1, 1]. s is 1 for a positive semi-definite matrix and −1 for a negative
k k
semi-definite matrix. sk ∈ (−1, 1) for a not positive nor negative semi-definite matrix.
1822 A Tamburrino and G Rubinacci

4.2. Treatment of noise


Here we modify the proposed method to tackle two different problems. Namely, the problem
of measurement noise11 and the problem of approximating the shape of the crack as the union
of τk . We consider these two problems jointly because the replacement of the crack volume
with the union of τk is equivalent to the introduction of a noise source affecting the measured
resistance matrix.
The noise affecting the measured matrix R̃ may corrupt the eigenvalues of R̃ − R̂τk or
R̂˜ ext \τk − R̃ and, consequently, the sign indices sk and tk . The effect of noise on sk is harmful
because it can be responsible for excluding from Iext the index of a τk contained in the crack
volume.
To deal with the measurement noise we slightly change the definition of  ˜ int
˜ ext , Iext , 
and Iint into

˜ ext,ε  τk (59)
k∈Iext,ε

Iext,ε  {k|sk  1 − ε, k = 1, . . . , N} (60)



˜ int,ε 
 τk
k∈Iint,ε

Iint,ε  {k ∈ Iext |tk < 1 − ε}


where ε ∈ (0, 1) is a threshold.
To reconstruct  ˜ ext we solve the one-parameter minimization problem

εext = arg min


R̃ − R̂˜ ext,ε
2 (61)
ε∈(0,1)

where
·
is a suitable norm, for instance the Frobenius norm, and set
˜ ext = 
 ˜ ext,εext
Iext = Iext,εext . (62)
Once Iext has been found by (62), we solve a second one-parameter minimization problem:
εint = arg min
R̃ − R̂˜ int,ε
2
ε∈(0,1)

and set
˜ int = 
 ˜ int,εint .

Notice that the number of direct problems that have to be solved to obtain ˜ ext is at most
2N + 1. Specifically, the solution of N direct problems is required to obtain the matrices R̂ τk
and the solution of at most N + 1 direct problems is required if the minimization (61) is carried
out exhaustively (the mapping ε → R̂˜ ext,ε is a piecewise constant mapping assuming at most
N + 1 different values (see (59) and (60)). Similarly, the number of direct problems solved
˜ ext is at most 2N + 1. Therefore, the maximum number of direct problems solved
to obtain 
during the inversion procedure is at most 4N + 2. This number can be reduced to 3N + 2 if the
matrices R̂ τk are precomputed and stored.

11When the resistance matrices are numerically computed, the measurement noise includes also the error due to the
numerical method.
A new non-iterative inversion method for electrical resistance tomography 1823

4.3. Generalization to non-homogeneous background and/or inclusion


The approach can be immediately extended to the case of non-homogeneous background and/or
inclusion. Specifically, let the background resistivity ρ0 and the inclusion resistivity ρc be two
functions of the position satisfying
0 < ρ0 (r )  ρ0U < ρcL  ρc (r )  ρcU  +∞, ∀r ∈ D
where ρ0 ,ρc and ρc are constants. When an inclusion occurs in , the resistivity of the
U L U

conductor is 
ρ0 (r ), for r ∈ D\
ρ (r ) =
ρc (r ), for r ∈ .
In the non-homogeneous case, we simply replace the test matrix R̂τk with the resistance
matrix related to the resistivity ρτk defined as
 U
ρ0 for r ∈ D\
ρτk (r ) 
ρcL for r ∈ 
and the test matrix R̂˜ ext \τk with the resistance matrix related to the resistivity

˜ ρ0U for r ∈ (D\ ˜ ext ) ∪ τk
ρτk ext (r ) 
ρcU ˜
for r ∈ ext \τk .

4.4. Numerical example


The numerical example regards a square plate (10 cm × 10 cm × 1 cm) with a background
conductivity ρ0 = 1  m probed by Ne = 16 equispaced electrodes of rectangular shape (see
figure 1). For the sake of simplicity we assume that the resistivity does not depend upon the z
coordinate (the z axis is orthogonal to the plate) and that the electrodes (width 1.25 cm, height
1 cm) extend for the whole height of the plate.
Under these assumptions, the current density does not depend upon the z coordinate and
does not have a component along the z axis; it is thus possible to use a three-dimensional finite
element mesh consisting of a single layer along the z axis although the geometry is fully three
dimensional. We assume that the inclusion is made of a conducting material with resistivity
ρc = 2  m.
The choice of the number of degrees of freedom used to represent the unknown resistivity
in the domain D depends on the number of electrodes Ne and the noise level. These three
parameters have been related by the heuristic estimate in [10] which was developed for a disc.
Nevertheless, we use that estimate for the square plate to evaluate roughly the magnitude of the
number of electrodes and number of degrees of freedom to represent the unknown. Specifically,
the resolution δ (the linear size of the smallest inclusion) and the number of electrodes are given
by [10]

δ e 1/2
≈ (63)
L 2|µ|

2|µ| 1/2
Ne ≈ 2 +1 (64)
e
where L is the length of the edge of the plate (10 cm), e is the accuracy of the measurements
and µ  (ρ0 − ρc )/(ρ0 + ρc ). In [10] e is the maximum (normalized) error on the voltages,
i.e.

δ V
 e
I
, ∀I (65)
where δ V is the error affecting the voltage vector and
·
is the Euclidean norm.
1824 A Tamburrino and G Rubinacci

Here we model the measurement noise as multiplicative noise, i.e. we assume that
R̃i j = Ri j (1 + ξi j ) (66)
where Ri j is the noiseless resistance matrix and R̃i j is the measured resistance matrix affected
by multiplicative noise that is mathematically described by the random variable ξi j . In addition,
we assume that the random variables ξi j are statistically independent and uniformly distributed
in the interval [− , ].
Notice that the definition of e given in (65) is useful in the case of additive noise on the
voltage vector. In the present case (multiplicative noise) we replace e in (63) and (64) with an
effective ee f f obtained by means of a statistical average (see the appendix)

R
F
ee f f = √ √ (67)
3 Ne − 1
where
·
F is the Frobenius norm.
Since we have assigned the number of electrodes Ne , and assuming R is the resistivity
in the absence of inclusion12 , we obtain from (63) and (64) that δ/L ≈ 8 and from (64) and
(67) that ≈ 4 × 10−5 . Therefore, we represent the unknown on a 8 × 8 uniform grid (see
figure 2(b)). Notice that the value 4 × 10−5 is the maximum noise level compatible with the
prescribed resolution, i.e. any smaller than 4 × 10−5 is an acceptable noise level.
The resistance matrix has been numerically computed by using (32) on a finite element
mesh made up of 48 × 48 × 1 regular hexahedral elements (see figure 2(a)).
Complementary formulations are important to estimate and reduce the numerical errors
affecting the numerical solutions, as shown in [8] by means of numerical simulations. As an
example, figure 4 shows the plots of the relative errors

R ρ − R
F


RρA − R∗
F
Rρ R − R∗
F
e  , e A
 , e R  (68)

R∗
F
R∗
F
R∗
F
as the mesh is refined. In (68) R∗ is the resistance matrix computed by using (32) on the finer
mesh made of 48 × 48 × 1 elements and ρ is the resistivity described at the top of figure 5(a).
As appears from figure 4, the complementary formulation gives a solution that converges
more rapidly than the solution based on the scalar potential or the electric vector potential13.
The numerical results concerning the inversion method have been organized into three
parts. In the first part we present the effectiveness of the inversion method, in the second part
the known limits and in the third part the robustness against noise.
Figures 5 report reconstructions of the shapes of different classes for = 2 × 10−5 .
Specifically, figure 5(a) reports the reconstructions  ˜ ext and ˜ int for a simply connected
domain close to being a convex set, figure 5(b) reports the reconstructions for an inclusion
consisting of three disjoint simply connected domains, figure 5(c) reports the result for a
non-convex but simply connected inclusion and 5(d) for a multiply connected inclusion.
Firstly, we notice that, for an inclusion that does not consist of a simply connected domain
close to being a convex set, the first step of the inversion procedure gives a reconstruction ˜ ext
that covers the object and that it is connected. In these cases the second step is essential in
order to remove the parts that do not belong to the inclusion.
We highlight that the second step of the procedure is critically dependent on the first
step. Specifically, if the first step fails to reconstruct a set including the inclusion, i.e.  is
12 In the present case (|µ| = 1/3) the order of magnitude of
R
F does not depend on the inclusion shape. When
the inclusion fills D, it turns out that
R
F is only twice that obtained in the case of the absence of inclusion, and
moreover, it corresponds to ≈ 2 × 10−5 .
13 Notice that a discretization finer than 48 × 48 × 1 should be used to numerically approximate the actual resistance

matrix within = 4 × 10−5 . Therefore, the parameter represents the noise level with respect to the synthetic data
and not with respect to the real data.
A new non-iterative inversion method for electrical resistance tomography 1825

Figure 4. The relative errors e (¾), e A () and e R (♦) as functions of the discretization parameter
Nmesh . The finite element mesh for computing the relative errors is made of Nmesh × Nmesh × 1
elements.

Figure 5. Reconstruction of different shapes. Upper line: the set  ˜ ext . Lower line: the set 
˜ int .
White represents the background and grey (dark) represents the true shape. Black represents a
pixel that belongs to the reconstruction but not to the true shape. Note that for this level of noise
( = 2 × 10−5 ) the set  ˜ int matches the true shape exactly (error-free reconstruction).

not contained in  ˜ ext , then the sign index tk (see section 4.1) may be useless because the set
˜
ext \τk never includes the inclusion.
Figure 6 reports some reconstructions showing the limits of the inversion method.
Specifically, we notice that the method does not work properly when  contains an inclusion
that is not a simply connected domain.
Finally, figures 7–9 report the reconstructions when the level of noise is increased up to
three orders of magnitude with respect to the theoretical estimated value ( = 2 × 10−5 ) of
1826 A Tamburrino and G Rubinacci

Figure 6. Reconstruction of different shapes. Upper line: the set  ˜ ext . Lower line: the set 
˜ int .
White represents the background and grey (light or dark) represents the true shape. Black represents
a pixel that belongs to the reconstruction but not to the true shape and light grey represents a pixel
that belongs to the true shape but not to the reconstruction. The level of noise is = 2 × 10−5 .

Figure 7. Reconstruction of different shapes for = 2 × 10−4 . Upper line: the set  ˜ ext . Lower
˜ int . White represents the background, grey (light or dark) represents the true shape.
line: the set 
Black represents a pixel that belongs to the reconstruction but not to the true shape and light grey
represents a pixel that belongs to the true shape but not to the reconstruction.

the maximum level of acceptable noise. It is worth noting that shape (a), the single inclusion
close to a convex set, can be recovered (within one error pixel) up to = 2 × 10−2 by the first
step, giving ˜ ext . In addition, shapes (c) and (d), consisting of single ‘strongly’ non-convex
inclusions, can be recovered up to = 2 × 10−4 .
A new non-iterative inversion method for electrical resistance tomography 1827

Figure 8. Reconstruction of different shapes for = 2 × 10−3 . Upper line: the set  ˜ ext . Lower
˜ int . White represents the background and grey (light or dark) represents the true
line: the set 
shape. Black represents a pixel that belongs to the reconstruction but not to the true shape and light
grey represents a pixel that belongs to the true shape but not to the reconstruction.

(a) (b) (c) (d)

Figure 9. Reconstruction of different shapes for = 2 × 10−2 . Upper line: the set  ˜ ext . Lower
˜ int . White represents the background and grey (light or dark) represents the true
line: the set 
shape. Black represents a pixel that belongs to the reconstruction but not to the true shape and light
grey represents a pixel that belongs to the true shape but not to the reconstruction.

5. Conclusions

A new inversion method for the identification of the shape of inclusions has been developed,
based on the monotonicity of the resistance and conductance matrices. The monotonicity has
thus been thoroughly analysed and reference has also been made to the numerically computed
matrices. To this end, several numerical models have been presented which are consistent
in the sense that the numerically computed resistance and conductance matrices satisfy the
same monotonicity property as the actual matrices. Moreover, the issues of accuracy (the
average models give increased accuracy by combining the two complementary formulations)
1828 A Tamburrino and G Rubinacci

and efficiency (in shape identification of cracks, the solution of the direct problem does not
require remeshing as the crack shape is changed) have been tackled with great care.
The monotonicity property satisfied by the resistance and conductance matrices, which
represent the key to the proposed inversion method, makes it possible to infer information on
the unknown resistivity. This information can be obtained by performing a simple test which
consists of computing the eigenvalues of the difference between the measured resistance matrix
and a trial resistance matrix. Moreover, the computational cost for computing the eigenvalues
increases as O(n 3 ), where n + 1 is the number of electrodes, and it is generally small because
n is not normally greater than a few dozen.
Finally, the inversion algorithm, based on a fully non-linear unknown-data model, is a
straightforward application of the monotonicity of the resistance and conductance matrices.
Its strength lies in its local nature, i.e. by using a proper trial resistivity configuration, we can
infer information on a given position of the specimen. Moreover, the number of direct problems
solved during the inversion procedure increases linearly with the number of pixels (voxels)
used for representing the unknown map of resistivity distribution of crack configuration.

Acknowledgments

This work has been supported in part by the Italian Space Agency and MIUR.

Appendix

In this appendix is computed ee f f .


First of all, notice that each component of the voltage noise is a zero mean random variable.
In fact, from (66) it turns out that

N e −1

δ Vi = ξik Rik Ik
k=1

and thus

N e −1

δ Vi = ξik Rik Ik = 0
k=1

where · is the statistical average operator.


The measurement of the resistance matrix is usually performed by injecting a unitary
current into the kth electrodes and measuring the electrode voltages under the condition that
no current is injected in the remaining electrodes, apart from the sink electrode, of course. In
this configuration ek2 , the mean square noise affecting the vector of the electrode voltages, is

N e −1
N e −1
N e −1
2 N
e −1

ek2 = δ Vi 2 = (ξik Rik Ik )2 = R2ik Ik2 ξik2 = R2 .


i=1 i=1 i=1
3 i=1 ik
Then we define ee f f as the root mean square of the ek :

Ne −1 Ne −1
1
N e −1
1
R
F
ee f f =  ek = √ √
2  R2ik = √ √ .
Ne − 1 i=1 3 N e − 1 k=1 i=1 3 Ne − 1

Finally, we notice that the same result can be achieved assuming that the current vector I is
random uniformly distributed over the unitary sphere.
A new non-iterative inversion method for electrical resistance tomography 1829

References

[1] Kohn R and Vogelius M 1984 Determining conductivity by boundary measurements Commun. Pure Appl. Math.
37 289–98
[2] Sylvester J and Uhlmann G 1987 Global uniqueness theorem for an inverse boundary value problem Ann. Math.
125 153–69
[3] Nachman A I 1988 Reconstruction from boundary measurements Ann. Math. 128 531–76
[4] Nachman A I 1995 Global uniqueness for a two-dimensional inverse boundary value problem Ann. Math. 142
71–96
[5] Isakov V 1993 Uniqueness and stability in multi-dimensional inverse problems Inverse Problems 9 579–621
[6] Cheney M, Isaacson D and Newell J C 1999 Electrical impedance tomography SIAM Rev. 41 85–101
[7] Pinheiro P A T, Loh W W and Dickin F J 1998 Three-dimensional reconstruction algorithm for electrical
resistance tomography IEE Proc. Sci. Meas. Technol 145 85–93
[8] Tamburrino A, Ventre S and Rubinacci G 2000 Electrical resistance tomography: complementarity and quadratic
models Inverse Problems 16 1585–618
[9] Pierri R and Tamburrino A 1997 On the local minima problem in conductivity imaging via a quadratic approach
Inverse Problems 13 1547–68
[10] Gisser D G, Isaacson D and Newell J C 1990 Electric current computed tomography and eigenvalues SIAM J.
Appl. Math. 50 1623–34
[11] Rubinacci G, Tamburrino A, Ventre S and Villone F 2002 Shape identification in conductive material by electrical
resistance tomography Electromagnetic Non-destructive Evaluation vol 6 (Amsterdam: IOS Press) pp 13–20
[12] Colton D and Kirsch A 1996 A simple method for solving inverse scattering problems in the resonance region
Inverse Problems 12 383–93
[13] Kirsch A 1998 Characterization of the shape of the scattering obstacle using the spectral data of the far field
operator Inverse Problems 14 1489–512
[14] Ikeata M and Ohe T 2002 A numerical method for finding the convex hull of polygonal cavities using the
enclosure method Inverse Problems 18 111–24
[15] Brühl M 2001 Explicit characterization of inclusion in electrical impedance tomography SIAM J. Math. Anal.
32 1327–41
[16] Brühl M and Hanke M 2000 Numerical implementation of two noniterative methods for locating inclusions by
impedance tomography Inverse Problems 16 1029–42
[17] Cheng K-S, Isaacson D, Newell J C and Gisser D G 1989 Electrodes model for electrical current computed
tomography IEEE Trans. Biomed. Eng. 36 918–24
[18] Somersalo E, Cheney M and Isaacson D 1992 Existence and uniqueness for electrode models for electric current
computed tomography SIAM J. Appl. Math. 52 1023–40
[19] Bossavit A 1998 Computational Electromagnetism, Variational Formulations, Edge Elements, Complementarity
(Boston, MA: Academic)
[20] Albanese R and Rubinacci G 1998 Finite element methods for the solution of 3D eddy current problems Adv.
Imaging Electron Phys. 102 1–86
[21] Rikabi J, Bryant C F and Freeman E M 1988 An error based approach to complementary formulations of static
field solutions Int. J. Numer. Methods Eng. 26 1963–87
[22] Albanese R, Rubinacci G and Villone F 1999 An integral computational model for crack simulation and detection
via eddy-currents J. Comput. Phys. 152 736–55
[23] Albanese R, Rubinacci G, Tamburrino A and Villone F 2000 Phenomenological approaches based on an integral
formulation for forward and inverse problems in eddy current testing Int. J. Appl. Electromag. Mech. 12
115–37

You might also like