Paper The following article is Open access

Anisotropic conductivity tensor imaging in MREIT using directional diffusion rate of water molecules

, , , and

Published 19 May 2014 © 2014 Institute of Physics and Engineering in Medicine
, , Citation Oh In Kwon et al 2014 Phys. Med. Biol. 59 2955 DOI 10.1088/0031-9155/59/12/2955

0031-9155/59/12/2955

Abstract

Magnetic resonance electrical impedance tomography (MREIT) is an emerging method to visualize electrical conductivity and/or current density images at low frequencies (below 1 KHz). Injecting currents into an imaging object, one component of the induced magnetic flux density is acquired using an MRI scanner for isotropic conductivity image reconstructions. Diffusion tensor MRI (DT-MRI) measures the intrinsic three-dimensional diffusion property of water molecules within a tissue. It characterizes the anisotropic water transport by the effective diffusion tensor. Combining the DT-MRI and MREIT techniques, we propose a novel direct method for absolute conductivity tensor image reconstructions based on a linear relationship between the water diffusion tensor and the electrical conductivity tensor. We first recover the projected current density, which is the best approximation of the internal current density one can obtain from the measured single component of the induced magnetic flux density. This enables us to estimate a scale factor between the diffusion tensor and the conductivity tensor. Combining these values at all pixels with the acquired diffusion tensor map, we can quantitatively recover the anisotropic conductivity tensor map. From numerical simulations and experimental verifications using a biological tissue phantom, we found that the new method overcomes the limitations of each method and successfully reconstructs both the direction and magnitude of the conductivity tensor for both the anisotropic and isotropic regions.

Export citation and abstract BibTeX RIS

Content from this work may be used under the terms of the Creative Commons Attribution 3.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

The diffusion tensor MRI (DT-MRI) has been used to measure the diffusivity of water in tissues such as brain white matter and skeletal muscle by applying the diffusion-sensitizing magnetic field gradient (Cleveland et al 1976, Chenevert et al 1990, Moseley et al 1990, LeBihan 1991). The DT-MRI can determine the nerve fiber tract orientation within the brain white matter using the effective diffusion tensor. Adopting the effective macroscopic anisotropic tensor modeled by a two-phase anisotropic medium (Sen and Torquato 1989), the electrical conductivity tensor and the water diffusion tensor have been analyzed in terms of the intra- and extra-cellular transport coefficients with the same eigenvectors. Based on the model, Tuch et al (2001) derived a linear relationship between the eigenvalues of the conductivity tensor and those of the water diffusion tensor.

Using the linear relationship, a conductivity tensor map of the human brain has been investigated in several studies: independent estimation of the extra-cellular conductivity and diffusivity (Tuch et al 2001, Sekino et al 2003), use of the volume-constraint model (Wolters et al 2006), estimation of the signal attenuation in the cortex and the corpus callosum using the stimulated echo acquisition mode sequence (Sekino et al 2009) and estimation of the volume fraction in each compartment through a multi-compartment model (Wang et al 2008). There was also an attempt to experimentally find the cross-property relation among electrical conductivity, diffusion and T2 (Oh et al 2006). However, these methods lack a reliable way of experimentally estimating the extra-cellular conductivity values needed to completely recover the conductivity tensor in terms of its direction and magnitude.

There have been several previousstudies to produce conductivity images by using an MRI scanner as a tool to acquire the internal measurements of the current-induced magnetic field (Zhang 1992, Woo et al 1994, Ider and Birgul 1998, Kwon et al 2002, Seo et al 2003, Oh et al 2003, Muftuler et al 2004, Ozdemir et al 2004, Oh et al 2005). The externally injected low-frequency current produces the internal current density J = (Jx, Jy, Jz) and magnetic flux density B = (Bx, By, Bz) distributions, where the conductivity information is embedded. When the main field of the MRI scanner is in the z-direction, the induced z-component, Bz, causes changes in the MR phase image. Without rotating the object, therefore, one can extract an image of only Bz from the MR phase image (Joy et al 1989). Magnetic resonance electrical impedance tomography (MREIT) can now produce images of the electrical conductivity distribution with injection currents of about 5 mA or less (Birgul et al 2006, Hamamura et al 2006, Woo and Seo 2008, Hamamura and Muftuler 2008, Hasanov et al 2008, Kim et al 2009, Seo and Woo 2011).

Since Bz reflects J as its volume integral and the normal component of J is continuous at the interface of a conductivity change, most MREIT image reconstruction algorithms using only Bz need at least two independent injection currents to produce isotropic conductivity images. However, biological tissues such as the white matter and skeletal muscle show anisotropy due to asymmetric cellular structures. Even though there have been a few studies for anisotropic conductivity image reconstructions in MREIT (Seo et al 2004, Pyo et al 2005, Degirmenci and Eyuboglu 2007, 2012, 2013, Nam and Kwon 2010b), there is no practically reliable method available yet. We, therefore, need a reliable experimental method to obtain additional information on the three-dimensional anisotropic conductivity tensor.

In this paper, we adopt the linear relation between the conductivity tensor C and the water diffusion tensor D (Tuch et al 2001) and propose a novel algorithm to reconstruct the anisotropic conductivity tensor by combining the DT-MRI and MREIT techniques without any referred extra-cellular conductivity and diffusivity information. This idea was first suggested by Ma et al (2013) as diffusion tensor current density impedance imaging (DT-CD-II), where DT-MRI is combined with current density impedance imaging (CDII) for anisotropic conductivity tensor imaging. Since they use CDII, measurements of the vector quantity B are required. To obtain all the three components of B = (Bx, By, Bz), the imaging object must be rotated twice inside the MRI scanner. In this paper, we perform the anisotropic conductivity image reconstruction using only Bz data without rotating the object.

To use the linear relation between the two tensors, we recover the projected current density JP from the measured Bz data as the best approximation of J. Using the reconstructed JP and the water diffusion tensor map, we determine the extra-cellular conductivity and diffusivity ratio (ECDR) denoted as ηext. This enables us to use a new direct algorithm to quantitatively visualize the anisotropic conductivity tensor map.

In the following sections, we will first investigate the relation between C and D. Presenting a novel direct conductivity tensor reconstruction algorithm combining JP and D, we will evaluate its performance through numerical simulations and imaging experiments using a phantom including both isotropic and anisotropic objects.

2. Methods

2.1. Relation between conductivity and water diffusion tensors

The pulsed gradient spin echo (PGSE) in figure 1(a) excites the spin system by applying time-constant magnetic field gradients before and after the π-pulse and acquires the k-space data. Due to the dephasing of magnetization because of diffusion in the period Δ, the acquired signal intensity reflects the amount of diffusion that has occurred during the diffusion gradient pulses.

Figure 1.

Figure 1. (a) DT-MRI pulse sequence and (b) MREIT pulse sequence with low-frequency injection currents.

Standard image High-resolution image

The effective water diffusion tensor D can be written as a positive definite symmetric matrix:

Equation (1)

where the column vectors of SD are the orthogonal eigenvectors of D, the superscript T denotes the transpose and di for i = 1, 2, 3 are the corresponding eigenvalues. The signal intensity ρD of diffusion MRI is given by

Equation (2)

where ρ0 is the signal obtained without diffusion-sensitizing gradient, $\bf g$ is the normalized diffusion-sensitizing gradient vector and b denotes the diffusion-weighting factor depending on the gradient pulse used in the DT-MRI sequence. From figure 1(a),

Equation (3)

where γ = 26.75 × 107 rad Ts−1 is the gyromagnetic ratio of hydrogen and δ and G are the duration and amplitude, respectively, of the diffusion-sensitizing gradient pulse along a given direction. Diffusion in the x-, y- and z-directions can be measured by applying the gradient in the x-, y- and z-directions, respectively.

The eigenvalue ci of the conductivity tensor C can be represented as

Equation (4)

where σext is the extra-cellular conductivity, dint and dext are the intra- and extra-cellular diffusion coefficients, respectively, and $\mathcal {O}(d_{{\rm int}}^2)$ is bounded as $d_{{\rm int}}^2$ tends to infinity (Tuck et al 2001, Haueisen et al 2002). For small intra-cellular diffusion dint ≈ 0, the eigenvalue ci satisfies

Equation (5)

The eigenvalues ci of C and the eigenvectors of D in SD can determine the conductivity tensor as

Equation (6)

2.2. Measurement of current-induced magnetic flux density

In a conventional spin-echo MREIT pulse sequence, both positive and negative currents of the same amplitude and duration are injected. These injection currents with the pulse width of Tc accumulate extra phases. The corresponding k-space MR signals can be described as

Equation (7)

where ρ is the T2 weighted spin density, ϕ is any systematic phase artifact and Ω is a field-of-view (FOV). Here, the superscript of S± denotes a brief notation for S+ and S. For the standard coverage of the k-space, we set

Equation (8)

where Gx is the frequency encoding gradient strength, TE is the echo time, ΔGy is the phase encoding step and Tpe is the phase encoding time. The induced magnetic flux densities generated by the positive and negative injection currents I± are denoted as ±Bz, respectively.

The standard two-dimensional inverse Fourier transform provides the complex MR images ${\mathcal {M}}^\pm$ as

Equation (9)

From (9), the magnetic flux density Bz can be recovered as

Equation (10)

where α and β are the imaginary and real part of ${\mathcal {M}}^+/{\mathcal {M}}^-$, respectively. From the analysis by Scott et al (1992) and Sadleir et al (2005), the noise standard deviation $sd_{B_z}$ of the measured Bz is given by

Equation (11)

where Tc is the current injection duration and ϒM is the signal-to-noise ratio (SNR) of the MR magnitude image.

Using the property that the noise standard deviation is inversely proportional to Tc and |ρ|, a multi-echo imaging sequence which acquires a series of echoes has been developed and optimized to reduce the noise level in the measured Bz data (Nam and Kwon 2010a). Figure 1(b) shows a schematic diagram for the injection current nonlinear encoding (ICNE) gradient-multi-echo pulse sequence.

2.3. Computation of projected current density

Let Ω be a cylindrical domain with its boundary ∂Ω. We may express Ω as a union of the slices perpendicular to the z-axis:

Equation (12)

Here, Ω0 denotes the center slice. In this paper, we use the following two-dimensional estimations:

Equation (13)

Since Bz is the only measurable quantity without rotating the object inside the MRI scanner, we inject current in the orthogonal direction to the main magnetic field through a pair of attached electrodes to maximize Bz. According to the Helmholtz decomposition, Park et al (2007) decomposed the internal current density J into curl-free and divergence-free components as

Equation (14)

Here, J0 := −∇α is the background current density determined by

Equation (15)

where n is the outward unit normal vector on ∂Ω and g denotes the Neumann boundary data subject to the injection current. The vector potential $\boldsymbol{\Psi }$ satisfies

Equation (16)

We define the vector field JP as the projected current density, which can be directly calculated from any vector field J, with

Equation (17)

where the potential ψ solves

Equation (18)

Note that JP may not be identical to J of which the z-component is not negligible. The projected current density JP is the quantity observable from the measured Bz data, which satisfies ∇ · JP = 0 and the following stability condition (Park et al 2007):

Equation (19)

Here, the constant C depends only on Ωt and not on J. The estimated stability between J and JP implies that JP is close to the true current density J depending on the z-component of JJP.

2.4. Reconstruction of conductivity tensor

As described by Ma et al (2013), from (5) and (6), we can get

Equation (20)

The internal current densities Ji for i = 1, 2 subject to two externally injected currents can be represented as

Equation (21)

where ui is the voltage potential corresponding to the injection current gi for i = 1, 2 and

Equation (22)

is a scale factor called the ECDR.

Using the estimated diffusion tensor D, the current Ji satisfies the following relation:

Equation (23)

and it implies

Equation (24)

The term D−1Ji can be defined as the pseudo-current as suggested by Ma et al (2013). With a current density in place of D−1Ji, the relation (24) is also the basis to recover the isotropic conductivity from two internal current densities in CDII (Hasanov et al 2008).

Assuming that the reconstructed projected current density JP, iJi = −ηextDui by injecting transversal currents, the unknown ηext can be represented as

Equation (25)

where D is known. Setting

Equation (26)

the identity (25) can be written as

Equation (27)

where Ei, x and Ei, y are the x- and y-components of Ei, respectively. Using (27), we have the following matrix equation:

Equation (28)

where

Equation (29)

and the unknown variable ${\mathbf {x}}= \big( \frac{\partial \log \eta _{{\rm ext}}}{\partial x}, \frac{\partial \log \eta _{{\rm ext}}}{\partial y} \big)^T$. By solving (28), we get the transversal gradient vector $\tilde{\nabla }\log \eta _{{\rm ext}}$ in the imaging slice Ωt.

The log-scale log ηext is recovered by

Equation (30)

where $\Phi _2(\mathbf {r}-\mathbf {r}^{\prime })=\frac{1}{2\pi }\log |\mathbf {r}-\mathbf {r}^{\prime }|$ is the two-dimensional fundamental solution of the Laplace equation. From a reference background value of ηext, the log-scale log ηext can be iteratively updated based on the equation (30). A similar formula to determine the isotropic conductivity from the measured Bz data was introduced by Oh et al (2003). To guarantee the uniqueness of the solution log ηext, we set log ηext := ζ + φ where

Equation (31)

The following two-dimensional harmonic equation in Ωt is uniquely determined up to a constant:

Equation (32)

Therefore, the scale factor log ηext is uniquely determined up to a constant in Ωt. The determination of the constant depends on a practical situation including a known background conductivity value or a reference region with a known conductivity value such as the hydrogel under the thin carbon electrode.

2.5. Numerical simulation

We built a three-dimensional numerical model using COMSOL (COMSOL Inc., USA). Figure 2(a) shows the cylindrical model with 100 mm diameter and 140 mm height. We attached two pairs of electrodes on its surface to inject currents. Figure 2(b) shows its finite element mesh with 264 020 tetrahedral and 19 130 triangular elements. As shown in figure 2(c), the model includes two types of anomalies: one is the isotropic agar gel object and the other is the anisotropic muscle tissue with different directional diffusion rates. The orientations of three muscle tissues were $0,\frac{\pi }{2}$ and 0.6 radian as marked in figure 2(c). The background of the model was assumed to be saline.

Figure 2.

Figure 2. (a) Cylindrical model and (b) its mesh configuration. (c) Simulated MR image without diffusion-sensitizing gradient. The numbers indicate the orientations of the muscle tissues in radians. (d) Simulated MR images with diffusion-sensitizing gradient. (e) Color-coded FA map.

Standard image High-resolution image

To generate simulated data, we set S0 as the signal obtained without the diffusion-sensitizing gradient:

Equation (33)

With the diffusion-sensitizing gradient, we calculated the echo signal intensity as

Equation (34)

where g represents the normalized diffusion-sensitizing gradient vector and

Equation (35)

where α is the orientation specified in table 1. Figure 2(d) shows the echo signal intensities Sb for the six direction gradients of

Equation (36)

and the used parameters in table 1.

Table 1. Simulation parameters to generate the diffusion tensor map in figure 2.

Parameter Saline Muscle Agar
Proton density 1.0 0.7 0.5
T1 (ms) 3620 1420 883
T2 (ms) 767 31.7 70
d1 (× 10−3 mm2 s−1) 1.70 1.17 2.10
d2 (× 10−3 mm2 s−1) 1.70 0.84 2.10
d3 (× 10−3 mm2 s−1) 1.70 0.70 2.10
Orientation, α (rad) 0 0, π/2, 0.6 0

The most commonly used measure to quantify the anisotropy is the fractional anisotropy (FA). We calculated the FA value as the normalized variance of the three eigenvalues of the water diffusion tensor:

Equation (37)

where $\hat{d}:=\frac{d_1+d_2+d_3}{3}$. Figure 2(e) shows the color-coded FA map of the numerical model to indicate diffusion along the x-axis (red), y-axis (green) and z-axis (blue). We computed the FA map by taking the projection of the FA value on to the principal eigenvector (Jiang et al 2006). For the case of our numerical simulation, the principal eigenvector was [ − cos α, −sin α, 0]T and, therefore, the color-coded FA map was computed as

For the muscle anomaly oriented at 0.6 radian, the FA value was found to be 0.2610 using equation (37). The corresponding color was [R, G, B] = [0.2154, 0.1474, 0], which is between the red and green. Since the FA value is zero for the isotropic object, the agar object is not visible in the FA map.

For the MREIT simulation, the isotropic conductivity values of the background saline and circular agar anomaly regions were 0.2 Sm−1 and 0.5 Sm−1, respectively. The orientations of the eigenvectors of the anisotropic muscle region were assumed to be the same as those of the diffusion eigenvectors and the eigenvalues were set to be (c1, c2, c3) = (1.00, 0.710, 0.675) Sm−1.

2.6. Imaging experiment

Figures 3(a) and (b) illustrate the phantom design. Around the cylindrical phantom with 200 mm diameter and 140 mm height, we attached four carbon-hydrogel electrodes (HUREV Co. Ltd, Korea). We filled it with 1 S m−1 saline and placed one cylindrical isotropic gel object of 2 S m−1 conductivity. We also put three pieces of anisotropic biological tissue (chicken breast) with the dimension of 35 × 35 × 35 mm3. We placed the phantom inside the bore of the 3 T MRI scanner (Achieva, Philips) equipped with the 32-channel receiver RF coil (SENSE-Head-32ch, Philips).

Figure 3.

Figure 3. (a) Phantom design and (b) its top view. (c) T2 weighted MR magnitude image with a b-value of zero. (d) Series of diffusion-weighted MR images for the diffusion weighting factor b = 1000 s mm−2. The gradients were applied along the x-, y- and z-directions, respectively. (e) Color-coded FA map.

Standard image High-resolution image

We acquired the diffusion-weighted data using the single-shot spin-echo echo planar imaging sequence. We applied the diffusion-weighting gradients in 15 directions with b-value of 1000 s mm−2. The imaging parameters were as follows: repetition time TR = 3000 ms, echo time TE = 73 ms, slice thickness = 5 mm, number of excitations (NEX) = 2, FOV = 180 × 180 mm2 and acquisition matrix size = 64 × 64. The NEX means the number of repeated acquisitions to improve the SNR by averaging. Figure 3(c) shows the T2 weighted MR magnitude image. Three images in figure 3(d) are the b-value diffusion-weighted MR magnitude images with the applied gradients in the x-, y- and z-directions. Figure 3(e) shows the color-coded FA map. The left chicken breast object shows diffusion along the x-direction (red), the middle-top object shows diffusion along intermediate directions (red and green) and the right one shows diffusion along the y-direction (green).

We used a custom-designed MREIT current source to sequentially inject 10 mA currents for the horizontal and vertical directions (Kim et al 2011). The current source was synchronized with the ICNE multi-gradient-echo pulse sequence to optimize the multiple Bz, l data for l = 1, ..., NE by minimizing the noise effect (Nam and Kwon 2010a). The imaging parameters were as follows: repetition time TR = 35 ms, first echo time $T_{E_1}$ = 1.88 ms, echo space Es = 2.2 ms, slice thickness = 5 mm, number of echoes NE = 15, NEX = 50, flip angle = 6.65°, FOV = 260 × 260 mm2, imaging matrix = 128 × 128. The imaging time (TR × 128 × NEX) was 3.7 min. For each current injection direction, we acquired the k-space data twice, subject to the positive and negative injection currents to cancel out the systematic phase artifact. Therefore, to obtain the full data set for the horizontal and vertical directions, the total imaging time was 14.8 min. The multiple echoes corresponding to the different echo times $T_{E_l}=1.88+2.2\times (l-1)$ ms for l = 1, ..., 15 were acquired by the alternating read-out gradients.

3. Results

3.1. Simulation results

Figure 4(a) shows the simulated data of Bz, 1 and Bz, 2 subject to the horizontal and vertical current injections, respectively. Figure 4(b) shows the x- and y-components of the recovered projected current densities JP, 1 and JP, 2 for the horizontal and vertical injection currents, respectively.

Figure 4.

Figure 4. (a) Simulated Bz data and (b) computed projected current density JP. Superscripts 1 and 2 denote the horizontal and vertical current injections, respectively. (c) Diffusion tensor map and (d) computed ECDR ηext. (e) Reconstructed images of the anisotropic conductivity tensor C.

Standard image High-resolution image

Figure 4(c) shows the water diffusion tensor D. Since it is symmetric with six variables, we estimated them from the six diffusion-sensitizing gradient vectors and one reference S0 by taking the logarithm on both sides of (34). From the estimated water diffusion tensor, we determined its three eigenvectors by diagonalizing it. Using JP in figure 4(b) and D in (c), we solved the matrix equation (28) to compute the ECDR factor ηext using a regularization factor ξ as

Equation (38)

where I is the 2 × 2 identity matrix. The value of ξ was set to be inversely proportional to the determinant of A. We, therefore, added a relatively large value of ξ where |A| was small and vice versa. Figures 4(d) and (e) show the computed ECDR ηext and the reconstructed conductivity tensor C, respectively.

We estimated the L2-error of the reconstructed C using the following formula:

Equation (39)

where Ωt is the imaging region, C* denotes the true conductivity tensor and Nxy is the number of pixels in Ωt. The L2-error values were 0.0900, 0.0885, 0.0895, 0.0066, 0.000 and 0.000 for σ11, σ22, σ33, σ12, σ13 and σ23, respectively.

To evaluate the performance of the proposed method at different noise levels, we added the white Gaussian random noise to both the diffusion weighted signal and Bz data assuming that the DT-MRI and MREIT experiments were conducted independently. For the DT-MRI experiment, we assumed that the SNR of S0 in the background saline varied from 70 to 80 dB, in steps of 5 dB. Since the SNRs of the different regions depend on their relaxation properties, we used the following relation to compute the SNRs of the gel and tissue regions:

Equation (40)

where T1, 0, T2, 0 and ϒM, 0 were the T1 and T2 relaxation values and the SNR of the background saline region, respectively. The subscript k denotes either the gel or tissue regions and the T1, k and T2, k values are given in table 1. We chose TR and TE as 4000 and 40 ms, respectively, which were similar to those used in the typical experimental study. After computing the SNR in each region, we calculated the noise standard deviation sdk of each region using the following relation:

Equation (41)

where 〈Sb, k〉 is the average signal amplitude in each region and the factor 0.66 accounts for Rayleigh statistics. We generated and added the complex white Gaussian noise with zero mean and the standard deviation given by (41).

For the MREIT experiment, we assumed that the SNR of the background saline region ϒM, 0 varied from 70 to 100 dB, in steps of 15 dB. With TR = 1200 and TE = 15 ms, we computed the expected noise levels for the gel and tissue regions using (40). For the current pulse width, Tc = 30 ms, the expected noise standard deviation, $sd_{B_z,k}$ of Bz, for each region, was

Equation (42)

We tried four different noise levels of $\big(\Upsilon _{M,0}^D,\Upsilon _{M,0}^{B_z}\big)=(70,\infty ),(70,100),(70,85)$ and (70, 70) dB where $\Upsilon _{M,0}^D$ and $\Upsilon _{M,0}^{B_z}$ are the background region SNRs of the DT-MRI and MREIT experiments, respectively. Figure 5 shows images of the ECDR ηext and diagonal elements σ11, σ22, σ33 of C for the chosen noise levels. Table 2 summarizes the estimated L2-errors showing that the proposed method stably reconstructed the conductivity tensor C for the chosen noise levels.

Figure 5.

Figure 5. Computed ECDR ηext and reconstructed images of the diagonal elements of C for different noise levels of $(\Upsilon _{M,0}^D,\Upsilon _{M,0}^{B_z})$. (a): (70, ), (b): (70, 100), (c): (70, 85) and (d): (70, 70) dB.

Standard image High-resolution image

Table 2. L2-errors of the reconstructed conductivity tensor for different noise levels.

$\Upsilon _{M,0}^{B_z}$ (dB) σ11 σ22 σ33 σ12 σ13 σ23
0.0916 0.0909 0.0907 0.0221 0.0213 0.0220
100 0.1097 0.1078 0.1052 0.0225 0.0216 0.0226
 85 0.1132 0.1109 0.1056 0.0235 0.0232 0.0243
 70 0.1240 0.1211 0.1186 0.0247 0.0242 0.0256

3.2. Experimental results

Figure 6(a) shows the measured Bz, 1 and Bz, 2 data from the phantom in figure 3 corresponding to the horizontal and vertical current injections, respectively. Figure 6(b) plots computed JP, 1 and JP, 2 for the horizontal and vertical injection currents, respectively. Figure 6(c) shows the diffusion tensor map D. In the three chicken breast regions with muscle fibers oriented in the x-, y- and xy-directions as shown in figure 3(e), the diffusion tensor map shows different diffusion anisotropy effects in each region. The cylindrical gel region shows the isotropic diffusion. Figure 6(d) and (e) show images of the ECDR ηext and the conductivity tensor C, respectively, reconstructed from the measured D and Bz data without using any referred or assumed extra-cellular and diffusivity values.

Figure 6.

Figure 6. (a) Measured Bz data from the phantom in figure 3 and (b) computed projected current density JP. Superscripts 1 and 2 denote the horizontal and vertical current injections, respectively. (c) Diffusion tensor map and (d) computed ECDR ηext. (d) Reconstructed images of the anisotropic conductivity tensor C.

Standard image High-resolution image

Figure 7 shows the color-coded FA and conductivity values in the three chicken breast tissue regions. The reconstructed conductivity values of the tissues exhibit their anisotropic characteristics depending on the muscle fiber orientation. We can see that σ11 is relatively higher in the left tissue and σ12 is higher in the top middle tissue.

Figure 7.

Figure 7. (a) Color-coded FA map of three chicken breast tissue regions and (b) reconstructed images of the anisotropic conductivity tensor C in the same regions.

Standard image High-resolution image

Table 3 shows the conductivity values, eigenvalues and FA values of the reconstructed conductivity tensor C at five different regions. The maximum eigenvalues of the background saline and the agar gel object were 1.05 and 2.22 S m−1, respectively, which were close to the true values of 1 and 2 S m−1. For the anisotropic muscle regions, the ratio between the maximum and minimum eigenvalues was 1.38.

Table 3. Conductivity values (σij), eigenvalues (ci) and FA values computed from the reconstructed conductivity tensor images of the phantom in figure 3.

  Saline Agar Left tissue Right tissue Top-middle tissue
σ11 1.0271 ± 0.0246 2.1771 ± 0.1444 0.2220 ± 0.0403 0.1587 ± 0.0227 0.1576 ± 0.0504
σ22 1.0335 ± 0.0209 2.1950 ± 0.1538 0.1745 ± 0.0295 0.1969 ± 0.0249 0.1503 ± 0.0449
σ33 1.0187 ± 0.0230 2.1978 ± 0.1637 0.1679 ± 0.0303 0.1541 ± 0.0221 0.1355 ± 0.0410
σ12 −0.0177 ± 0.0062 −0.0069 ± 0.0117 −0.0089 ± 0.0023 0.0113 ± 0.0034 0.0192 ± 0.0057
σ13 −0.0095 ± 0.0053 0.0011 ± 0.0193 −0.0017 ± 0.0016 0.0010 ± 0.0031 0.0039 ± 0.0026
σ23 −0.0039 ± 0.0052 −0.0195 ± 0.0161 −0.0007 ± 0.0040 0.0029 ± 0.0010 0.0023 ± 0.0020
c1 1.0525 ± 0.0240 2.2294 ± 0.1581 0.2241 ± 0.0391 0.1994 ± 0.0532 0.1745 ± 0.0532
c2 1.0202 ± 0.0239 2.1864 ± 0.1499 0.1745 ± 0.0308 0.1579 ± 0.0212 0.1364 ± 0.0422
c3 1.0065 ± 0.0209 2.1542 ± 0.1538 0.1658 ± 0.0302 0.1515 ± 0.0222 0.1326 ± 0.0408
FA 0.0233 ± 0.0050 0.0178 ± 0.0049 0.1664 ± 0.0098 0.1535 ± 0.0161 0.1561 ± 0.0105

4. Discussion

To compare the reconstructed conductivity tensor with the one without using the measured Bz data, we estimated the extra-cellular diffusivity dext in the background saline as

Equation (43)

With the known background conductivity value of σext = 1 S m−1, the extra-cellular conductivity to the diffusivity ratio was given as

Equation (44)

Without using the measured Bz data, we computed the eigenvalues of the conductivity tensor C using the following relation:

Equation (45)

Then, we reconstructed the conductivity tensor image shown in figure 8 by using the relation ${\bf C}={\bf S}_D\tilde{{\bf C}}{\bf S}_D^T$. We note that the agar gel region is not distinguished from the background saline. Since the water diffusion tensor can provide only the direction information of the conductivity tensor, quantitative reconstructions of the conductivity tensor are not possible without using the referred or assumed extra-cellular conductivity values at all pixels. Comparing the reconstructed images of C in figures 6(e) and 8, we can clearly see that the proposed method overcomes the limitations of the conventional DT-MRI and MREIT techniques.

Figure 8.

Figure 8. Reconstructed images of the anisotropic conductivity tensor C without using MREIT data. The images lack quantitative information regarding the absolute conductivity values.

Standard image High-resolution image

Unlike the method by Ma et al (2013) where all the three components of B = (Bx, By, Bz) are used, we are restricted to the Bz data in this paper to avoid rotating the imaging object inside the MRI scanner. Instead of using $\mathbf {J}=\frac{1}{\mu _0}\nabla \times \mathbf {B}$, therefore, we computed the projected current density JP from the measured Bz data subject to two injection currents. Though JP is the best approximation of J, it may have a relatively large error if JP, z is significantly different from J0, z which is the z-component of the current density J0 for the isotropic homogeneous conductivity distribution. This undesirable situation may occur if the isotropic or anisotropic conductivity distribution changes a lot along the z-direction. To reduce the error, it would be advantageous to use long electrodes in the z-direction and choose the imaging slices around the middle of them.

Without measuring Bx and By, we can reconstruct isotropic conductivity images by using two sets of the Bz data subject to two linearly independent injection currents (Woo and Seo 2008, Seo and Woo 2011). In the proposed DT-MREIT method, we take advantage of the diffusion tensor acquired by using DT-MRI and transform the anisotropic conductivity tensor image reconstruction problem into the image reconstruction of the ECDR factor ηext, which is a scalar quantity. Though we speculate that the error in ηext originated from JP instead of J would be small for most cases, further study is needed to quantitatively investigate the effects of using JP on the errors in reconstructed anisotropic conductivity images.

Since the proposed method should utilize the measured Bz data in addition to the water diffusion tensor map, we have to carefully deal with the measurement noise in Bz. In the procedure to estimate JP, we solve the harmonic equation (18) with ∇2Bz as a source term. For those local regions of MR signal void or very short T2 relaxation time, the amount of noise in measured Bz data cannot be reduced. Future studies of in vivo animal and human experiments should adopt denoising techniques to effectively deal with these problematic regions and avoid the noise propagation to other regions (Jeon et al 2010, Lee et al 2011).

To probe the passive material property of the conductivity, we inject current into the imaging object in MREIT. The amount of the injection current must be lower than a certain level so that the generated internal current density does not stimulate the nerve and muscle. For example, the internal current density of 1 to 23 A m−2 was estimated as the threshold to stimulate a nerve with 20 μm diameter (Reilly 1989). Injection of 5 mA through a uniform-current-density electrode with 25 cm2 contact area will produce 2 A m−2 current density underneath the electrode (Song et al 2011). We speculate that current injections of 1 to 5 mA through surface electrodes with a practical size will produce negligible side effects such as nerve or muscle stimulations since the injected current spreads to result in a local current density below a stimulation threshold. However, there must be further studies of carefully designed in vivo animal and human experiments to find the maximum allowed imaging current.

Electrical property tomography (EPT) is another MR-based method to produce conductivity and permittivity images at the Larmor frequency using the B1 mapping technique (Katscher et al 2009, van Lier et al 2013, Zhang et al 2013). Local Maxwell tomography (LMT) also produces conductivity images at the Larmor frequency (Sodickson et al 2013). Noting that the anisotropy is mostly observable at low frequencies and also, that most physiological events occur at relatively low frequencies, MREIT, especially when combined with DT-MRI, seems to be more advantageous for functional imaging applications since it provides conductivity images at low frequencies below 1 KHz. EPT and LMT, which do not require current injection, may find important clinical applications to determine different tissue pathologies and values of local specific absorption rate.

In this study, current density imaging was done as part of the anisotropic conductivity image reconstruction. The projected current density JP may find useful clinical applications in deep brain stimulation (DBS) and transcranial direct current stimulation (tDCS). We suggest further experimental studies where the anisotropic conductivity tensor imaging together with the projected current density is utilized to better understand the underlying mechanisms of DBS and tDCS. Once we obtain the anisotropic conductivity tensor image, we may also perform various numerical analyses for any electrode configuration and stimulation pattern.

5. Conclusion

We proposed the novel method called DT-MREIT to reconstruct the anisotropic conductivity tensor image. By combining the DT-MRI and MREIT techniques, the new method quantitatively recovers the direction and magnitude of the anisotropic conductivity tensor as well as the isotropic conductivity value. Results of the numerical simulations and phantom experiments clearly show the feasibility of the new method.

The DT-MREIT technique suggested in this paper can provide quantitative anisotropic and also isotropic conductivity images together with diffusion tensor images. There could be numerous potential applications for the brain and muscle including ischemia, inflammation, bleeding and tumor (Woo and Seo 2008, Seo and Woo 2011). There is also an attempt to detect local conductivity changes accompanied by neural activities using MREIT (Sadleir et al 2010). EEG and MEG source imaging is to provide non-invasive fast imaging of the brain activity by solving the associated forward and inverse problems (Baillet et al 2001). We expect that quantitative anisotropic conductivity information will be valuable in these source imaging methods to improve their performance (Haueisen et al 2002, Gao et al 2006).

Acknowledgments

This paper resulted from the Konkuk University research support program and was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MEST) (2013R1A2A2A04016066, 2012R1A1A2008477).

Please wait… references are loading.