Elena Beretta, Monika Muszkieta, Wolf Naetar, and Otmar Scherzer

6A variational method for quantitative photoacoustic tomography with piecewise constant coefficients

Elena Beretta, Dipartimento di Matematica, Politecnico di Milano, Via Edoardo Bonardi, 20133 Milan, Italy ([email protected])

Wolf Naetar, Computational Science Center, University of Vienna, Oskar Morgenstern-Platz 1, A-1090Vienna, Austria ([email protected])

Monika Muszkieta, Faculty of Pure and Applied Mathematics, Wroclaw University of Technology,Wyb. Wyspianskiego 27, PL-50-370Wroclaw, Poland [email protected]

Otmar Scherzer, Computational Science Center, University of Vienna, Oskar Morgenstern-Platz 1,A-1090 Vienna, Austria and Radon Institute of Computational and Applied Mathematics, AustrianAcademy of Sciences, Altenbergerstr. 69, A-4040 Linz, Austria ([email protected])

Abstract: We consider the inverse problem of determining spatially heterogeneous absorption and diffusion coefficients μ(x), D(x), from a single measurement of the absorbed energy ε(x) = μ(x)u(x), where u satisfies the elliptic partial differential equation

(D(x)u(x)) + μ(x)u(x) =0 in Ω N .

This problem, which is central in quantitative photoacoustic tomography, is in general ill-posed since it admits an infinite number of solution pairs. Using similar ideas as in [31], we show that when the coefficients μ, D are known to be piecewise constant functions, a unique solution can be obtained. For the numerical determination of μ, D, we suggest a variational method based on an Ambrosio Tortorelli approximation of a MumfordShah-like functional, which we implement numerically and test on simulated two-dimensional data.

Keywords: Quantitative photoacoustic tomography, mathematical imaging, inverse problems, MumfordShah functional

AMS: 35R25, 35R30, 65J22, 65K10, 92C55

6.1Quantitative photoacoustic tomography

6.1.1Introduction

Photoacoustic tomography (PAT) is a medical imaging technique that combines electromagnetic excitation (in the visible spectrum) with ultrasound measurements. In a photoacoustic experiment, a translucent sample is illuminated by a laser pulse with wavelength λ (which we assume to be fixed in this article). The absorbed optical energy leads to thermal expansion, generating an ultrasound pressure wave p(x, t) that can be measured outside the sample.

Since the pulses used in PAT are very short, the complete energy is deposited almost instantaneously compared to travel times of acoustic waves, and the pressure wave p can be assumed to have been generated by an initial pressure (x), that is, it satisfies the wave equation

and p|M, where denotes a measurement surface, can be obtained from ultrasound measurements [13, 14, 28, 45].

By solving an inverse problem for the wave equation (see, e.g, [28] for a review of inversion techniques), these ultrasound measurements can be used to estimate the initial pressure (x). Since

that is, (x) is proportional to the absorbed energy ε(x), which is in turn proportional to the optical absorption coefficient μ(x) at the applied wavelength λ and the local fluence u(x) (the time-integrated laser power received at x), the initial pressure visualizes contrast in μ. The constant of proportionality Γ(x) is called the Grüneisen coefficient (or PA efficiency since it describes the efficiency of conversion from absorbed energy to acoustic signal) [13, 14].

In quantitative photoacoustic tomography (qPAT), the goal is to apply PAT to determine (inhomogeneous) optical material properties of the sample (which are of diagnostic interest). To do so, an additional nonlinear inverse problem for light transport has to be solved, since the fluence u(x) is inhomogeneous and itself dependent on the optical properties of the sample [13, 14].

While it is also possible to attack the acoustic and optical inverse problems simultaneously (see, e.g., [9, 24]), here, we assume that the acoustic part of the problem has been solved successfully, that is that (possibly noisy) data δ (with δ signifying the noise level) are available.

In this article, we utilize the diffusion approximation (which is valid in highly scattering media [13, 45]) of the radiative transfer equation to model the fluence distribution. It is, however, also possible to use a radiative transfer model for qPAT (see [17, 33, 38, 44, 46]), albeit at the cost of increased computational and analytical complexity.

The diffusion model (in a Lipschitz domain Ω N) is given by

The parameter D is called diffusion coefficient. The Dirichlet boundary data g (which we assume to be continuous and known in this article) describes the illumination pattern. Note that this is a time-independent model, again due to the fact that energy is deposited almost instantaneously compared to time scales of the acoustic system.

The difficulty of the inverse problem varies depending on which parameters of the model are assumed to be known. We present three inverse problems often considered in the literature. The hardest one is:

Bal and Ren showed in [6] that for arbitrary coefficients μ, D, Γ W1,(Ω), this problem is unsolvable, even if multiple measurements of (with different known boundary illumination patterns g) are available.

In [31], the authors showed that with a restriction to piecewise constant parameters, unique reconstruction of all three unknown parameters μ, D, Γ from multiple measurements is possible (under a condition on the directions of uk, where uk is the fluence of the k-th illumination pattern). Furthermore, an analytical reconstruction procedure was suggested and implemented numerically, which, unfortunately, is relatively sensitive to noise.

Alberti and Ammari [1] also established a unique reconstruction result, based on morphological component analysis (a sparsity approach), in a slightly more general setting (which assumes different degrees of smoothness of the coefficients and the fluence). They also provide numerical reconstructions, which were, however, not tested for noise sensitivity in the case of (P3).

To simplify the problem, it is often assumed that the Grüneisen coefficient Γ isknown or constant, which implies that the absorbed energy can be estimated with (with δ again denoting the noise level). It remains to solve

If only a single measurement of εδ is given, this inverse problem is also ill-posed (since it has infinitely many solutions pairs, see [12, 31, 39]).

However, in [1], the authors were able to recover μ (independently of the light transfer model used) from a single measurement of ε (again using a sparsity method, assuming different degrees of smoothness of the coefficients and the fluence).

In [6, 7] it was shown that this problem is uniquely solvable if two measurements of ε (corresponding to well-chosen boundary illuminations g1, g2) are available. Numerically, this multi-illumination case was treated in [6, 22, 36, 39, 44, 48], using a multitude of different techniques, see also the review paper by Cox et al. [13].

The simplest case works under the assumption that the diffusion coefficient D is also known:

This inverse problem has a unique solution even for a single measurement, which can be seen by substituting μu = εδ in (1.2), providing the possibility to solve for u [8]. For other (numerical) approaches, cf. [13]. To the authors knowledge, this simplified problem is the only case for which practical viability (with experimental data) has been established both for phantoms [47] and biological samples [40, 41].

6.1.2Contributions of this article

In this article, we consider the problem (P2) using a single measurement for our reconstructions, a problem that is in general ill-posed. Similar to [31] (which treats (P3) with multiple measurements), a restriction to piecewise constant μ, D also proves to be useful for this problem. In fact, as shown in Section 6.2, when the parameters μ, D are piecewise constant functions (and noise-free data are given), the inverse problem (P2) can be solved uniquely, without any further assumptions.

In Section 6.3, we present a variational model for the reconstruction of piecewise constant μ, D from noisy data εδ based on the Ambrosio Tortorelli approximation of a MumfordShah-like functional. Compared to the two-step reconstruction process presented in [31] (which was introduced for (P3), but is applicable for (P2)), which first detects the regions where the parameters are constant, then reconstructs the parameter values from jumps of the data and its derivatives, this variational approach is much more robust with respect to noise. This is mainly due to the fact that the numerical approach presented in [31] requires almost perfect jump detection in the second derivatives of εδ to get reasonable estimates of μ, D (since the jumps have to form a full partition of the domain Ω), which is highly challenging in the presence of significant amounts of noise. This is not the case for the variational method presented here.

Finally, a description of our implementation and numerical results can be found in Section 6.4.

6.2Recovery of piecewise constant coefficients

In this section we show that piecewise constant parameters μ, D can be recovered uniquely from a single measurement of the absorbed energy ε(μ, D).

In the following, let be a partition of Ω N into open sets and μ, D piecewise constant on that is, for Ωm open and μm, Dm +

Furthermore, for k , denote by

Jk(f) = Ω {B Ω | B is open and f Ck(B)}

the discontinuities of a function f L(Ω) and its derivatives up to k-th order.

First, we show that coefficient discontinuities can be recovered from the data ε.

Proposition 2.1. Let μ, D be of the form(2.1) and ε = ε(μ, D) satisfy (1.1),(1.2) weakly. Then we have

Proof. First, take an arbitrary open ball B with B (J0(μ) J0(D)) = 0 and notice that, by interior elliptic regularity, u C(B) and hence also ε C(B). It follows that

J0(μ) J0(D) J2(ε) .

By the De GiorgiNashMoser theorem [23, Theorem 8.22], we have u C0(Ω), so

Finally, as u C(Ωm), m = 1, . . . , M, (1.2) holds strongly in all Ωm, and we get

since μm, Dm are scalars. Hence, which shows that Δε cannot be continuous where jumps (since hence

The proposition follows from

Remark 2.2. The proof of Proposition 2.1 also shows that jumps in μ and D have different effects on the data ε. By (2.3), the jumps of μ can be recovered from those in ε. Jumps of D (not coinciding with jumps of μ) on the other hand, are smoothed in the data and have to be obtained from the derivatives of ε. In particular, the discontinuities of Δε suffice to find the jumps in D.

Furthermore, note that from discontinuities of ε (or, equivalently, |ε|2), one can, in general, only recover the part of the jumps of D where u(x) ν(x) ≠ 0 (where ν denotes a normal vector to the hyper-surface of discontinuity in D), as u may be continuous across parts of the jump set of D where u(x) is tangential.

To see this, consider for instance the case of a partition consisting of a homogeneous region Ω1 with (nontouching) smooth inclusions Ω2, . . . , ΩM. In this case, we have for all m = 1, . . . , M (due a result of Li and Nirenberg, see [29]) and the interface conditions

hold point-wise on Ωm for m = 2, . . . , M, where ν, τ denote vectors normal and tangential to Ωm (see, e.g., [31] for a derivation). Now, let

We have J0(D) Λ J1(u) since, by the interface conditions (2.5), u cannot be continuous across J0(D) Λ (it changes length). Conversely, if we take an open ball B Ωk Ω1 (for some 2 k M) with B J0(D) Λ = 0, we have, again by (2.5),

since either Dk = D1 or u|ΩkB ν = u|Ω1B ν = 0. As we have (2.6) implies u C1(B) and therefore hence

This shows that, in general, discontinuities in the second derivatives of the data ε have to be identified in order to get the whole jump set of D.

Once the partition is known, the coefficients μ, D can be recovered from thejumps in ε, and the boundary values of u.

Proposition 2.3. Let μ,D be of the form (2.1) with known. Furthermore, let ε satisfy (1.1),(1.2) for a known boundary illumination g. Then, μ and D can be determined uniquely from ε.

Proof. Let x Ωm Ωn for some m, n {1, . . . , M}. Since

Furthermore, take x Ω Ωk (for some k {1, . . . , M}). We have

Starting in Ωk and using (2.7) on all interfaces, we recover all μm, m = 1, . . . , M.

Finally, to obtain Dm, m = 1, . . . , M, we use (2.4), that is,

Remark 2.4. If g is not known, Proposition 2.3 can be used to determine the parameters μ, D up to a constant.

6.3A MumfordShah-like functional for qPAT

In Section 6.2, we showed that piecewise constant absorption and diffusion coefficients μ, D, can be recovered from noise-free data ε by an analytical procedure whichfirst determines the coefficient jumps, that is the partition and then thenumerical values In [31], such a two-step approach was implementednumerically (for the problem (P3)).

In the presence of significant amounts of noise in the data, however, such an approach is infeasible, since it requires the detection of jumps of derivatives up to second order of the data εδ. In particular, jumps may remain partially undetected (e.g., if the edge detection has to be restricted to first derivatives due to noise, see Remark 2.2 and the numerical examples in Section 6.4), leading to an incomplete estimated partition and therefore highly erroneous parameter estimates.

To overcome this problem, we propose a variational approach favoring piecewise constant solutions that estimates the numerical values of piecewise constant μ, D and their jumps at the same time.

There are multiple different methods for piecewise constant regularization of inverse problems. For instance, a popular class of methods is based on the level set method [32] or variations of it, for example [10, 11, 16, 42, 43]). This approach has been suggested for qPAT (using the radiative transfer model) in [17], however, it was only tested using multiple measurements. In this article, we use an AmbrosioTortorelli approximation of a MumfordShah-like functional (which was first suggested for electrical impedance tomography in [37]). The main advantage of this approximation is that we can utilize incomplete jump information (obtained, for instance, from jumps of the data or other means) to initialize the minimization procedure (see Section 6.3.3). Numerically, this leads to faster convergence (and in some cases improved minimizers). Additionally, the number of segments does not have to be known in advance (in contrast to multiple level set methods) and the minimization with respect to the jump indicator functions is a simple elliptic problem.

We want to minimize the MumfordShah-like functional

where HN1 is the (N 1)-dimensional Hausdorff-measure and ε(μ, D) the operator that maps the (unknown) parameters μ, D to the measurements ε satisfying (1.1),(1.2). The minimum is taken over all μ, D in suitable (that is, point-wise bounded from below and above) subsets of W1,2(Ω Kμ), W1,2(Ω KD) and all closed sets Kμ, KD Ω (the jump sets of the coefficients μ, D).

Functionals of this type, first introduced by Mumford and Shah for image denoising and segmentation in [30], have been applied for a wide range of inverse problems, see, for example [20, 26, 27, 34, 35, 37]). The basic idea behind the functional is as follows. The discrepancy term

forces minimizing coefficients μ̂ , D ̂ to be close (in L2-sense) to a solution of the inverse problem ε(μ, D) = εδ (of which there are infinitely many), while the regularization terms

force both the variation of μ̂ , D ̂ (outside their jump sets K̂ μ , K̂ D) as well as the hypersurface area of K̂ μ, K̂ D to be small. The regularization parameters αμ, αD control the amount of continuous variation the parameters may have, whereas βμ, βD control the complexity of the coefficients jump sets Kμ , K D.

Note that while in general, minimizers of this functional can have some continuous variation, they have to be close to piecewise constant for large values of αμ, αD. In fact, the AmbrosioTortorelli approximation of the functional (3.1), which will be introduced in Section 6.3.2, can also be used for piecewise constant regularization (in the limit αμ, αD , see Remark 3.5).

6.3.1Existence of minimizers

Existence of minimizers of functionals like (3.1) (which are, in general, not unique) was first established in [18] for image segmentation and in [37] for regularization of nonlinear operator equations.

To ensure that ε(μ, D) is well defined and that has a minimizer, point-wise bounds have to be enforced (see [25]), so we have to restrict the coefficients to {f L(Ω) : a f b a.e. in Ω} for some 0 < a, b < a priori. We also require the following Lemma, the proof follows the ideas in [19].

Lemma 3.1. The nonlinear measurement operator

where u(μ, D) solves (1.2), is continuous.

Proof. Let Furthermore, let un := u(μn, Dn) and u := u(μ, D) be solutions of (1.2) corresponding to the given coefficients.

By the usual energy estimates for (1.2) (cf. [21, Chapter 6, Theorem 2]) we have

unW1,2(Ω) C(a, b, Ω)gL2(Ω) .

Hence, there exists a (re-labeled) subsequence (uk )k with u k u weakly in W1,2(Ω) for some u W1,2(Ω). From the weak form of (1.2) we get for all υ W1,(Ω) and k

Taking k and using the fact that the left side acts as a bounded linear functional on uk W1,2(Ω), we obtain for all υ W1,(Ω)

By density of W1,(Ω) W1,2(Ω), this implies Since the same argument also holds for every subsequence of the original sequence (un)n, we get un u weakly in W1,2(Ω) and thus un u strongly in L2(Ω).

Finally, since u L(Ω) by the maximum principle, continuity of ε follows from

Remark 3.2. Since for all the operator L1(Ω)2 L2(Ω) is also continuous.

Following [3, 37], we show the existence of minimizers of a weak form of defined for coefficients μ, D in the space ??(Ω), the special functions of bounded variation, which may have jump discontinuities. For a quick introduction, see Appendix A.

We obtain the functional

where S(f) denotes the approximate discontinuity set of a Lebesgue-measurable function f and μ, D the density of the absolutely continuous part (with respect to the Lebesgue measure) of the respective distributional gradients (see Appendix A).

In this setting, the existence of minimizers can be established using the direct method and the ?? compactness theorem.

Proposition 3.3. The weak MumfordShah-like functional has at least one minimizer (μ̂ , D̂ ) in

Proof. Let be a minimizing sequence of the functional (3.2). Clearly, we have for all n

By the ??-compactness theorem (see Section A), and using Lemma 3.1, first applied to μn, then to the obtained subsequence of Dn, there exist (μ̂ , D̂ ) ??(Ω)2 and a subsequence (μk , D k)k (re-labeled) with

We also have since is closed in Furthermore, it is a minimizer of since the L2(Ω)-norm is weakly lower semi-continuous and is continuous.

6.3.2Approximation

Using the approach of Ambrosio and Tortorelli (cf. [4]), one can obtain functionals

that approximate the functional from (3.2) and are easier to minimize:

It is well known that, for all α, β > 0 and ζ = o(ϵ),

in the sense of Γ-convergence in L1(Ω) (formally, the latter has to be extended to an additional variable, see [4, 37]). Since the data term is continuous in L1(Ω) and Γ-convergence is stable under continuous perturbations, we thus get Γ-limϵ0 ϵ = and therefore L1(Ω)-convergence of (a subsequence of) ϵ-minimizers (which can beshown to lie in a compact set [4]) to minimizers.

The existence of a minimizer of (3.3) can be established using the direct method.

Proposition 3.4. The functional ϵ has at least one minimizer

Proof. F ϵ is coercive with respect to the semi-norm

|μ|W1,2(Ω) + |D|W1,2(Ω) + υμW1,2(Ω) + υDW1,2(Ω) ,

which, combined with the restriction shows that a minimizing sequence is bounded in W1,2(Ω)4, so it contains a subsequence that converges weakly to (μ̂ , D̂ , υ̂μ, υ̂D) in W1,2(Ω)4 (and thus strongly in L2(Ω)4). Moreover, note that the spaces and are closed in L2(Ω). The proposition follows since the discrepancy term is continuous in L2(Ω)2 and the regularization terms are weakly lower semi-continuous in W1,2(Ω)4 (cf. [15, Theorem 1.13] for the α-terms).

Remark 3.5. If we vary αμ, αD with ϵ and take

the minimizers of ϵ converge to piecewise constant (in ??-sense, that is, with gradient vanishing almost everywhere) functions (μ̂ , D̂ ) that minimize

among all piecewise constant μ, D. A proof (for the single variable case) can be found in [4]. This explains, in light of Section 6.2, why this type of regularization is useful for qPAT.

6.3.3Minimization

In this subsection, we proceed formally, that is without proving convergence, existence of minimizers of sub-problems and that the necessary derivatives and adjoints exist. For the minimization of (3.3), we suggest the following alternating directions approach:

Algorithm 3.6.

  1. Choose parameters a, b, ϵ, αμ, αD, βμ, βD > 0.
  2. Find an initial estimate K ̂ of the edge set J0(μ) J0(D) using edge detection.
  3. Initialize with μ0 = D0 1 and
  4. Iterate until convergence:

(i)

(ii)

We now explain the steps in more detail.

Initialization using edge detection

In our numerical simulations, it became clear that using Proposition 2.1 to obtain a reasonable estimate K ̂ of J0(μ) J0(D) leads to faster convergence and better minimizers than simply taking K ̂ = 0.

Given the noisy measurements εδ, we have to estimate the discontinuities (or edges) of εδ, |εδ|2, and |Δεδ|. Since the jumps in these functions are of multiplicative nature (proportional to the local value of u or u ν), it is advantageous to apply a logarithmic transformation prior to edge detection (to obtain constant contrast), see [31].

Because of noise, the data has to be smoothed prior to taking derivatives (turning discontinuities into areas with large gradients). We smooth by convolution with a Gaussian, since this approach has the advantage that differentiation and smoothing can be done in one step (by differentiating the low-pass filter instead of the function). We have

Similar to [31], we proceed as follows:

  1. Detect edges K̂0 of
  2. Detect edges K̂1 of
  3. Detect edges K̂2 of
  4. Take K ̂ = K̂ 0 K̂ 1 K2

Here, 1B acts as a cut-off function for the filters (e.g., using ρ the p-norm ball of radius ρ). The operation A B = {z Ω | (B + z) A} denotes set erosion, which is performed to avoid multiple detection of edges (by ensuring that the cut-off filter does not intersect with already detected edges or the outside of the domain). Note, however, that for large B thismay lead to parts of edges (close to the domain boundary or already detected edges) not being detected. The scalar γ is a minimal value enforced for |εδ|2 (to avoid creating singularities at zeros).

The edge detection itself is performed by applying thresholds ξ0 , ξ1, ξ2 to the functions |f0|, |f1|, |f2| (taking the super-level sets as edge sets). Note that in contrast to [31], no complete segmentation is necessary and cruder edge estimates suffice, that is, the detected edges dont have to be reduced to thin curves.

Iteration

In step (i), to find, for fixed

we use GaussNewton-minimization. That is, in every iteration of an inner loop, we linearly approximate ε(μk + sμ , Dk + sD) ε(μk , Dk) + ε(μk , Dk)(sμ , sD) and take update steps

which we then project into so we update with

(μk+1, Dk+1) = (μk , Dk) + (s a) b .

A straightforward calculation shows that a minimizer s = (sμ , s D) of (3.4) satisfies the weak form of

with homogeneous Neumann boundary conditions for sμ, s D and

Lσ : W1,2(Ω) W1,2(Ω), x (σx) .

The linear operator ε(μ, D) and its (formal) adjoint ε(μ, D) are given by

where u = u(μ, D) and

By introducing auxiliary variables y1, y2, y3 (and taking μ = μk , D = Dk), the equations (3.5),(3.6) can be re-written as the system of (weak) PDE

which is amenable to discretization as a sparse matrix using, for example the finite element method (FEM).

In step (ii) of the outer loop we have to find, for fixed

The minimizer (υμ, υD) satisfies the linear equations

which can be solved separately and implemented numerically using FEM.

6.4Implementation and numerical results

In this section, the proposed algorithm is tested on two sets of simulated two-dimensional data (with different parameter range and level of detail). We also vary the noise on the data since the reconstruction quality strongly depends on the noise level.

The data (see Figures 1 and 2) was generated in the diffusion model (1.2) using self-written (linear-basis) finite element code in MATLAB. For both examples, we took Ω = [0, 5]2 and used a uniform boundary condition g 1. The simulated data were generated on a (400×400)-grid and then down-sampled (by averaging) to (200×200) to avoid inverse crime. After that, Gaussian noise withdifferent intensities (standard deviations of 0.5% and 10% of the average signal value dx) was added to the data.

The edge detector described in Section 6.3.3 was implemented by finite difference approximations of |f0|, |f1|, |f2| using central differences inside and one-sided differences near the boundary of the domain. The functions f1, f2, f3 were calculated by convolution filtering with the MATLAB function imfilter. For all examples, we used a square cutoff function, that is with ρ = 2σk (where k = 0, 1, 2).

The edge detector is used to detect jumps in the derivatives of the data εδ up to second order (to obtain an initial estimate of the parameter jump set J0(μ) J0(D)). Since this process is highly sensitive with respect to noise, we varied the edge detection procedure subject to the amount of noise in the data. In the noise-free examples, we estimated the jumps of all three functions f0, f1, f3, that is, jumps of derivatives of εδ up to second order. We restricted the jump estimation to f0, f1 for the low-noise examples (i.e., jumps of derivatives up to first order) and f0 in the high-noise examples (only jumps in the data εδ itself).

To obtain the parameters μ, D given an initial estimate of their combined edge set J0(μ) J0(D), we used the AmbrosioTortorelli approximation (3.3) of the MumfordShah functional introduced in Section 6.3. To minimize the functional, we used Algorithm 3.6, iterating, for all examples, the outer alternating-directions loop until the functional changes by less than 0.01% and the inner GaussNewton loop until the functional value changes less than 1%. We directly implemented the systems (3.7) and (3.8) using (self-written) linear-basis finite element code. The implementation is not fully conforming, that is, where necessary we used conversions between piecewise linear and piecewise constant functions for simplicity. The discrete systems were solved using the standard MATLAB sparse equation solver mldivide.

Fig. 1: Example A. Parameters μ, D and FEM-generated data E(μ, D).

For all examples, we chose a = 0.01, b = 3, and ϵ = 0.01. The other reconstruction parameters used (which were selected by hand) are listed in Tables 1 and 2.

Reconstruction results and error profiles at different noise levels can be seen in Figures 3 and 4. In both examples, the noise-free reconstructions are very accurate and contain mostly smoothing errors. In the low-noise reconstructions, because ofthe fact that more regularization is necessary, some of the parameter variation is underestimated. In the high-noise examples, most detail in D is lost since a lot of regularization is required to get reasonable results. The fine detail in μ can, however, still be recovered very accurately in both examples.

Fig. 2: Example B. Parameters μ, D and FEM-generated data E(μ, D).

Remark 4.1. In some of the examples, minimization of (3.1) using log-parameters, that is the mapping (log μ, log D) ε(μ, D) instead of ε, gave slightly better results. Note that this approach leads to a different linearization and therefore also a different system (3.7) to be solved in every step. However, to keep the presentation simple, we decided to use the functional as presented in Section 6.3 for our numerical experiments.

Fig. 3: Reconstructions A. Columns: 0%, 0.1%, and 10% noise. Rows: Estimated edge set, reconstructed parameters μest, Dest and error profiles for μest, Dest. The error profiles show values of the true and reconstructed parameters along the image diagonals. The color axis in the images of the reconstructed parameters was fixed to the same range as the true parameters shown in Figure 1.
Fig. 4: Reconstructions B. Columns: 0%, 0.1%, and 10% noise. Rows: Estimated edge set, reconstructed parameters μest, Dest, and error profiles for μest, Dest. The error profiles show values of the true and reconstructed parameters along the image diagonals. The color axis in the images of the reconstructed parameters was fixed to the same range as the true parameters shown as in Figure 2.

Table 1: Parameters used for Figures 3 and 4 (for the functional (3.3)).

Table 2: Parameters used for Figures 3 and 4 (for edge detection).

ASpecial functions of bounded variation and the SBV-compactness theorem

This section briefly introduces the notion of ??-functions and their compactness theorem. For a more comprehensive presentation with proofs, see, for example [5].

For a function f L1(Ω) with distributional gradient Df, we define its total variation by

The space ?(Ω), consisting of all L1(Ω)-functions with finite total variation (i.e., of bounded variation), is a Banach space with the norm

f BV(Ω) := f L1(Ω) + |Df|Ω .

Note that by the RieszMarkov representation theorem, functions f L1(Ω) are of bounded variation if and only if Df is a finite vector Radon measure. The measure Df can be decomposed into three parts

Df = Da f + Dj f + Dc f .

Daf is the part of Df that is absolutely continuous with respect to the Lebesgue measure N, that is

Da f = f N

for some integrable density function f. The jump part Dj f is concentrated on the jump (or approximate discontinuity) set S(f) defined by

S(f) := {x Ω | f (x) < f +(x)} ,

where, denoting with Bρ(x) the ball centered at x with radius ρ,

Furthermore, for HN1-almost all x S(f), there exists a unit normal vector ν(x) and we have

Dj f = (f + f )ν HN1|S(f) .

The remaining Cantor part Dc f is concentrated on a subset of ΩS(f)with intermediateHausdorff dimension between N 1 and N.

A function f ?(Ω) is a special function of bounded variation (or f ??(Ω)) if Dc f = 0, that is,

Df = f N + (f + f )ν HN1|S(f) .

Furthermore, we have the following compactness theorem due to Ambrosio [2]:

Theorem A.1. Let fn be a sequence in ??(Ω) with

for all n and some p > 1, M > 1. Then there exist a subsequence (fnk )k and f ??(Ω) with

Acknowledgment: This work was supported by the Austrian Science Fund (FWF)within the project FSP P26687 Interdisciplinary Coupled Physic Imaging and byIK I059-N funded by the University of Vienna.

Bibliography

[1]G. S. Alberti and H. Ammari. Disjoint sparsity for signal separation and applications to hybrid inverse problems in medical imaging. preprint, ENS, 2015. http://arxiv.org/abs/1502.04540.

[2]L. Ambrosio. A compactness theorem for a new class of functions of bounded variation. Boll. Un. Mat. Ital. B, 3:857881, 1989.

[3]L. Ambrosio. Existence theory for a new class of variational problems. Arch. Ration. Mech. Anal., 111:291322, 1990.

[4]L. Ambrosio and V. M. Tortorelli. On the approximation of free discontinuity problems. Boll. Un. Mat. Ital. B, 6:105123, 1992.

[5]H. Attouch, G. Buttazzo, and G. Michaille. Variational Analysis in Sobolev and BV Spaces: Applications to PDEs and Optimization. SIAM, Society for Industrial and Applied Mathematics,2006.

[6]G. Bal and K. Ren. Multi-source quantitative photoacoustic tomography in a diffusive regime. Inverse Probl., 27(7):075003, 2011.

[7]G. Bal and G. Uhlmann. Inverse diffusion theory of photoacoustics. Inverse Probl., 26:085010, 2010.

[8]B. Banerjee, S. Bagchi, R.M. Vasu, and D. Roy. Quantitative photoacoustic tomography from boundary pressure measurements: noniterative recovery of optical absorption coefficient from the reconstructed absorbed energy map. J. Opt. Soc. Amer. A, 25(9):23472356, 2008.

[9]M. Bergounioux, X. Bonnefond, T. Haberkorn, and Y. Privat. An optimal control problem in photoacoustic tomography. Math. Models Methods Appl. Sci., 24(12):2525, 2014.

[10]M. Burger and S. Osher. A survey on level set methods for inverse problems and optimal design. European J. Appl. Math., 16(02):263301, 2005.

[11]T. Chan and L. Vese. Active contours without edges. IEEE Trans. Image Process., 10(2):266277, 2001.

[12]B. T. Cox, S. R. Arridge, and P. C. Beard. Estimating chromophore distributions from multiwavelength photoacoustic images. J. Opt. Soc. Amer. A, 26(2):443455, 2009.

[13]B. T. Cox, J. G. Laufer, S. R. Arridge, and P. C. Beard. Quantitative spectroscopic photoacoustic imaging: a review. J. Biomed. Opt., 17(6):061202, 2012.

[14]B. T. Cox, J. G. Laufer, and P. C. Beard. The challenges for quantitative photoacoustic imaging. Proc. SPIE, 7177:717713, 2009.

[15]B. Dacorogna. Direct methods in the calculus of variations, volume 78 of Applied Mathematical Sciences. Springer, New York, 2 edition, 2008.

[16]A. De Cezaro, A. Leitao, and X. Tai. On multiple level-set regularization methods for inverse problems. Inverse Probl., 25(3):035004, 2009.

[17]A. De Cezaro, F. Travessini De Cezaro, and J. Sejje Suarez. Regularization approaches for quantitative photoacoustic tomography using the radiative transfer equation. J. Math. Anal. Appl.,429(1):415438, 2015.

[18]E. De Giorgi, M. Carriero, and A. Leaci. Existence theorem for a minimum problem with free discontinuity set. Arch. Ration. Mech. Anal., 108:195218, 1989.

[19]H. Egger and M. Schlottbom. Analysis and regularization of problems in diffuse optical tomography. SIAM J. Math. Anal., 42(5):19341948, 2010.

[20]S. Esedoglu and J. Shen. Digital inpainting based on the MumfordShahEuler image model. European J. Appl. Math., 13:353370, 2002.

[21]L. C. Evans. Partial Differential Equations, volume 19 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 1998.

[22]H. Gao, S. Osher, and H. Zhao. Quantitative photoacoustic tomography. In Mathematical Modeling in Biomedical Imaging II, pages 131158. Springer Berlin, Heidelberg, 2012.

[23]D. Gilbarg and N. Trudinger. Elliptic Partial Differential Equations of Second Order. Classics in Mathematics. Springer Verlag, Berlin, 2001. Reprint of the 1998 edition.

[24]M. Haltmeier, L. Neumann, and S. Rabanser. Single-stage reconstruction algorithm for quantitative photoacoustic tomography. Inverse Probl., 31(6):065005, 2015.

[25]M. Jiang, P. Maass, and T. Page. Regularization properties of the MumfordShah functional for imaging applications. Inverse Probl., 30(3):035007, 2014.

[26]E. Klann. A MumfordShah-like method for limited data tomography with an application to electron tomography. SIAM J. Imaging Sciences, 4(4):10291048, 2011.

[27]T. Kreutzmann and A. Rieder. Geometric reconstruction in bioluminescence tomography. Inverse Probl. Imaging, 8(1):173197, 2014.

[28]P. Kuchment and L. Kunyansky. Mathematics of thermoacoustic tomography. European J. Appl. Math., 19:191224, 2008.

[29]Y. Y. Li and L. Nirenberg. Estimates for elliptic systems from composite material. Comm. Pure Appl. Math., 56(7):892925, 2003.

[30]D. Mumford and J. Shah. Optimal approximations by piecewise smooth functions and associated variational problems. Comm. Pure Appl. Math., 42(5):577685, 1989.

[31]W. Naetar and O. Scherzer. Quantitative photoacoustic tomography with piecewise constantmaterial parameters. SIAM J. Imaging Sciences, 7(3):17551774, 2014. Funded by the Austrian Science Fund (FWF) within the FSP S105 Photoacoustic Imaging. Funded by the University of Vienna within the Vienna Graduate School in Computational Science IK I059-N.

[32]S. Osher and J. A. Sethian. Fronts propagating with curvature-dependent speed: Algorithms based on Hamilton-Jacobi formulations. J. Comput. Phys., 79(1):1249, 1988.

[33]A. Pulkkinen, B. Cox, S. Arridge, J. P. Kaipio, and T. Tarvainen. Quantitative photoacoustic tomography using illuminations from a single direction. J. Biomed. Opt., 20(3):036015, 2015.

[34]R. Ramlau and W. Ring. A MumfordShah level-set approach for the inversion and segmentation of X-ray tomography data. J. Comput. Phys., 221:539557, 2007.

[35]R. Ramlau and W. Ring. Regularization of ill-posed MumfordShah models with perimeter penalization. Inverse Probl., 26:115001, 2010.

[36]K. Ren, Gao H., and H. Zhao. A hybrid reconstruction method for quantitative PAT. SIAM J. Imaging Sciences, 6(1):3255, 2013.

[37]L. Rondi and F. Santosa. Enhanced electrical impedance tomography via the MumfordShah functional. ESAIM Control Optim. Calc. Var., 6:517538, 2001.

[38]T. Saratoon, T. Tarvainen, B. T. Cox, and S. R. Arridge. A gradient-based method for quantitative photoacoustic tomography using the radiative transfer equation. Inverse Probl., 29(7):075006, 2013.

[39]P. Shao, B. Cox, and R. J. Zemp. Estimating optical absorption, scattering, and Grueneisen distributions with multiple-illumination photoacoustic tomography. App. Opt., 50(19):31453154, 2011.

[40]Y. Sun, E. Sobel, and H. Jiang. Quantitative three-dimensional photoacoustic tomography of the finger joints: an in vivo study. J. Biomed. Opt., 14(6):064002, 2009.

[41]Y. Sun, E. Sobel, and H. Jiang. Noninvasive imaging of hemoglobin concentration and oxygen saturation for detection of osteoarthritis in the finger joints using multispectral three-dimensional quantitative photoacoustic tomography. J. Opt., 15(5):055302, 2013.

[42]X. Tai and X. F. Chan. A survey on multiple level set methods with applications for identifying piecewise constant functions. International Journal of Numerical Analysis and Modeling, 1(1):2547, 2004.

[43]X. Tai and H. Li. A piecewise constant level set method for elliptic inverse problems. Appl. Num. Math., 57:686696, 2006.

[44]T. Tarvainen, B. T. Cox, J. P. Kaipio, and S. R. Arridge. Reconstructing absorption and scatteringdistributions in quantitative photoacoustic tomography. Inverse Probl., 28(8):084009, 2012.

[45]L. V. Wang and H. Wu, editors. Biomedical Optics: Principles and Imaging. Wiley-Interscience,New York, 2007.

[46]L. Yao, Y. Sun, and H. Jiang. Quantitative photoacoustic tomography based on the radiative transfer equation. Opt. Letters, 34(12):17651767, 2009.

[47]Z. Yuan and H. Jiang. Quantitative photoacoustic tomography: Recovery of optical absorption coefficient maps of heterogeneous media. Appl. Phys. Lett., 88(23):231101, 2006.

[48]R. J. Zemp. Quantitative photoacoustic tomography with multiple optical sources. App. Opt., 49(18):35663572, 2010.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
18.217.181.166