Open Access
27 February 2018 Theoretical characterization of annular array as a volumetric optoacoustic ultrasound handheld probe
Author Affiliations +
Abstract
Optoacoustic ultrasound (OPUS) is a promising hybridized technique for simultaneous acquisition of functional and morphological data. The optical specificity of optoacoustic leverages the diagnostic aptitude of ultrasonography beyond anatomy. However, this integration has been rarely practiced for volumetric imaging. The challenge lies in the effective imaging probes that preserve the functionality of both modalities. The potentials of a sparse annular array for volumetric OPUS imaging are theoretically investigated. In order to evaluate and optimize the performance characteristics of the probe, series of analysis in the framework of system model matrix was carried out. The two criteria of voxel crosstalk and eigenanalysis have been employed to unveil information about the spatial sensitivity, aliasing, and number of definable spatial frequency components. Based on these benchmarks, the optimal parameters for volumetric handheld probe are determined. In particular, the number, size, and the arrangement of the elements and overall aperture dimension were investigated. The result of the numerical simulation suggests that the segmented-annular array of 128 negatively focused elements with 1λ  ×  20λ size, operating at 5-MHz central frequency showcases a good agreement with the physical requirement of both imaging systems. We hypothesize that these features enable a high-throughput volumetric passive/active ultrasonic imaging system with great potential for clinical applications.

1.

Introduction

Arguably, clinical diagnoses are becoming more reliant on the competence of imaging systems in visualization of the anatomy and physiology. The distinguishing characteristic of an imaging modality is greatly defined by its mechanism of contrast.1 In this respect, medical ultrasound (US) is subjected to the mechanical impedance mismatches of the investigating media. Although it is sufficient to display the anatomy, pathological situations require far more dynamic contrast. Optoacoustic (OA) imaging benefits from the optical selective absorption associated with the chromophoric molecules and affords manifold contrasts. Through the OA effect, characteristics of absorbing structures are encoded in a broadband acoustic waves. OA can be considered as a passive acoustic imaging in a sense that the acoustic waves are induced inside the tissue. Hence, by employing the array of ultrasonic sensors, combination of OA and US is inherent and worthwhile. The unrivalled advantage of optoacoustic/ultrasound (OPUS) stems from the precise coregistration and correspondingly, the correlation between the detailed anatomical data of ultrasound and physiological information of OA images.

Effectively, ultrasound (US) probes paved the way for various clinical and preclinical applications of OPUS.24 Despite the advantages, employed probes are essentially optimized for ultrasound imaging. For example, an aperture as large as 60λ would suffice for 1 deg spatial resolution,5 whereas in the backward epi-OA imaging systems, the detection view angle of less than 2π results in inaccurate estimation.6,7 Using speckle illumination technique, in Ref. 8, the angular view is increased by confining the boundary of optically excited acoustic sources, albeit at the cost of imaging depth. Analogously, ultrasonic thermal encoding9 has been practiced by extending the idea of nonuniform signal generation for deeper regions. However, the acquisition time for these methods is far from the real-time imaging. In an alternative approach, the effective detecting aperture has been increased virtually via coupling the acoustic reflectors.10,11 Although the success of this approach is practically demonstrated, the required geometry restricts the clinical applications.

Aside from the partial view angle, volumetric imaging is faced with multifold challenges. Given the λ/2 interelement spacing constraint (to avoid grating lobes), 60λ aperture requires high transducer (element) counts and expensive hardware.12 To cope with, variety of undersampling strategies have been pursued,1316 among them, aperiodic sparse arrays are offering 60λ aperture size with maximum 256 elements.5,17,18 In order to optimize the ultrasonic pulse echo beam quality and signal-to-noise ratio (SNR), the density of elements is tapered toward the center of aperture.5,19 Consequently, the efficient integration with light delivery system for OA imaging is not trivial. Reference 20 discusses the optimum trade-off among the illumination, aperture angle of view, SNR, number, and size of the elements for a universal handheld probe. The imaging system seems impeccable for volumetric OA imaging, but the focused arrangement of the integrated staring transducers may not proffer the field of view (FOV), SNR, or the contrast expected from classical ultrasound systems.21

The potentials of planar annular array in terms of achievable resolution and uniform sensitivity for three-dimensional (3-D) OPUS imaging have been theoretically demonstrated.2224 Yet for the otherwise point such as transducers, the directional response, and averaging over the surface of the elements deteriorate the imaging performance. While the averaging behavior increases the SNR, the ill-suited directional response can be treated by negative acoustic lens.25 Withal, in such configurations there is not much degree of freedom to simultaneously evaluate the effect of number and size of the elements. In a comparative study, similar performance between sunflower Fermat-spiral array and circular-annular array has been found (see Ref. 18). The lesser periodicity in Fermat spiral configuration encourages the sparser distribution and subsequently more freedom in the choice of parameters. The sparer array also means a larger aperture, which leads to a larger view angle, more angular frequency, and simply more spatial degree of freedom in defining the true object.26 However, a pervasive analysis framework is required to provide an insight to the imaging performance of the designed system and proceed with further optimization. A common objective assessment technique for imaging systems is based on the point spread function (PSF) and spatial sensitivity.2628 Often considered as the system response to a point source, PSF is highly biased to the merits of the reconstruction algorithm. Alternatively, PSF can be calculated analytically via a linear model based on the spatial impulse response29 that incorporates the properties of transceivers. For assessment of a shift variant system such as OPUS, reliable source localization in terms of standalone contribution and resolvability from the neighboring is crucial.

In this work, we are investigating the effect of the size, number, and arrangement of negatively focused transducers in annular array for a multipurpose 3-D OPUS handheld probe. The proposed geometry has the advantage of mitigating the transducer/channel counts required for 3-D imaging while preserving the spatial resolution and uniform insonification for the FoV.22,23 The inherent central cavity of the probe furnishes the in-line illumination, which is expedient for maximizing the fluence.30 In the next section, we explain the theoretical development of a linear model that comprises the behavior of an imaging system. Two evaluation methods, namely voxel crosstalk matrix28 and eigenanalysis, have been briefly introduced and their ability in interrogating the model and correspondingly assessing the characteristics of the imaging system is discussed. The former provides information w.r.t. the system sensitivity and spatial aliasing while the latter addresses the quantity and quality of spatial frequency components. The assessment is followed by the system optimization with regards to the number, size, and geometry of the detecting aperture, and the results are shown in Sec. 3. Finally, in Sec. 4 the optimal choice of parameters, concluding remarks, and future directions are discussed.

2.

Methods

2.1.

Construction of the Model

Acoustic imaging is concerned with interpreting the detected induced or scattered waves from the surface of medium. In linear regime, the governing equation for a homogeneous medium satisfies31

Eq. (1)

2p(r,t)t2ν22p(r,t)=s(r,t),
where p is the diverging acoustic wave from the source s with the speed of sound ν related to the medium. The source represents the reflectivity R for the insonified scatterers32 (Eq. 2) or the capability of the illuminated absorbers to transform photons energy to acoustic wave. The latter is characterized by the product of absorption coefficient μa of the chromophores, temperature-dependent Grüneisen parameter Γ representing the thermoelastic efficiency of the chromophores and optical fluence ϕ(r) (Eq. 3).

Eq. (2)

sUS(x,y,z)=[uexc(t)*htrIR(t)]R(x,y,z),

Eq. (3)

sOA(x,y,z)=μa(x,y,z)Γ(x,y,z)ϕ(x,y,z),
where uexc(t) is the excitation pulse, * represents the convolution in time, and htrIR the transmit impulse response of the transducer. The projection of the source on the transducer surface is recorded in the form of RF signal urec. This spatiotemporal sensing behavior of the transducers is given by the transfer function (Fig. 1) and is composed of two parts: I. The spatial filtering dictated by the transducer geometry, which is attributed to the averaging effect (integral part in Eq. 4) and is formulated by the spatial impulse response [hSIR(r,t)]. II. The temporal band-pass filtering, which is imposed by the acoustoelectrical impulse response [hAIR(t)] of the transducer. Hence, the recorded pressure field by the active area of the transducer Ad can be expressed via Huygens’ principle33

Eq. (4)

urec(r,t)=hreAIR(t)*Wt{Adδ[t(|rsrre|)/ν]|rsrre|dAd},
where rs and rre represent the source and receiving element distance from the control point, and W denotes an adaptive weighting factor, which conforms to the nature of the source. For US, it can be considered as the tissue acoustic attenuating properties, whereas in OA, the deposited optical energy in the volume of interest. Therefore, the recorded projection urec of any discrete point within the aperture FoV can be formulated as

Eq. (5)

urec(x,y,z,t)=hreAIR(t)*hreSIR(xxre,yyre,zzre,t)*s.

Fig. 1

The projection of acoustic wave emanated from a source within the FoV on the surface of a transducer. For every subelement, a delay is applied corresponding to the β6  dB diverging angle, to simulate the effect of negative focusing.

JBO_23_2_025004_f001.png

Expanding the formula for every voxel within the aperture FOV yields a time-discrete matrix representation of the above formula.34,35

Eq. (6)

Urec=M·S,
where U is a vector containing concatenated time sampled recorded signals, ensemble of all of the recorded projections, MJ×K is the forward model (i.e., transfer matrix) describing the propagation of the K voxels (induced sources) to N detecting elements. For every voxel-element pair, there is an attributing impulse response of size L. The columns of M are containing the response of all elements for a given voxel, being vertically concatenated such that J=L×N. The result is a convolution matrix that correlates a set of sources to a set of elements in the aperture.

Eq. (7)

M=[h(1,1)IRh(1,2)IRh(1,K)IRh(2,1)IRh(2,2)IRh(2,K)IRh(N,1)IRh(N,2)IRh(N,K)IR]J×K,
where h(n,k)IR=h(n,k)AIR*h(n,k)SIR. As for every voxel in the FoV, there is an attributed function, M can be treated as the matrix containing voxel expansion function. The integration over the spanned basis of each voxel implies the weighting value of the corresponding voxel and is proportional to the sensitivity of transducers to that voxel. Accordingly, for the generic assessment of the imagining system, M is exploited by two evaluation methods, the voxel crosstalk matrix28,36,37 and eigenanalysis.26

2.2.

Voxel Crosstalk Matrix

The concept of crosstalk has been established in the assessment of system design, mainly to determine the reliability of the estimable coefficients.38 The reliability in this context is the measure of strength and standalone contribution to individual sources. For a shift variant system such as OPUS, the coefficients are voxels energy. Hence, the effect of aperture in spatial blurring, aliasing, and spatial sensitivity is amenable to an analytic description. Voxel crosstalk matrix can be obtained as follows:

Eq. (8)

H=MTM.

The crosstalk matrix H is essentially a square matrix with ith diagonal elements (Hii) describing the stand alone contribution of corresponding voxel {spatial sensitivity [Fig. 2(c)]} and off-diagonal elements, Hij;ij, representing the parasitic contribution of neighboring voxels, known as spatial aliasing. The value of Hij2/HijHii indicates the resolvability and consequently spatial aliasing between any two voxels. Therefore, a better imaging system is the one with weaker off-diagonal and stronger diagonal elements. The quantitative evaluation is possible through a set of metrics for systematic evaluation of the system voxel crosstalk matrix.37 The utilized metrics in this context are the root mean square error (RMSE) and mean absolute error (MAE). These metrics are complementary to each other, providing a robust systematic approach in quantitative analysis of imaging performance. Together they render information about the variance and degree of error in a model.39,40

Fig. 2

(a) The transfer function hIR correlates the pair of source-element for every 200 control points and 128 elements of annular array. (b) The corresponding voxel crosstalk matrix and (c) the spatial sensitivity of system obtained from the diagonal elements. (d) Spatial Fourier transform of eigenvectors of model matrix and (e) the eigenvalue distribution. For every eigenvector bound to a nonzero eigenvalue, there is a spatial frequency. The dashed line segregated the Mprime from Mnull.

JBO_23_2_025004_f002.png

2.3.

Eigenanalysis

As the name implies, eigendecomposition of the square matrix H features prominently the principal characteristics of the matrix by delivering the basis of eigenvectors q scaled by scalar eigenvalues λ.

Eq. (9)

Hq=λq,

Eq. (10)

H=QΛQ1.

H is a Hermitian matrix with real and positive eigenvalues and linearly independent orthogonal eigenvectors q, Q (square matrix) is the projection of H in the eigenspace with eigenvectors along its columns, and Λ is a diagonal matrix containing eigenvalues λi along its diagonal such that λ1>λ2>>λn. The eigendecomposition can be considered as an expansion of the matrix with respect to its basis (i.e., eigenvectors), each with certain weighting coefficient (i.e., eigenvalues). Alternatively, eigenvectors of H can be calculated by singular value decomposition (SVD) of matrix M, given by M=UΣV. From the mathematical point of view, V=Q and Σ is a diagonal matrix containing singular values σi along its diagonal such that σi=λi. Regardless, the matrix H can be represented simply as weighted and ordered combination of basis. The eigenvectors of the matrix H depict the principal structures of the detected wave components, reserving the original properties of the original matrix.32

Notably, the matrix M is a spatiotemporal matrix, therefore SVD decomposes M into the temporal matrix U and spatial matrix V. In other words, V is composing of K columns equal to the number of voxels while U is containing J columns equal to the number of recorded samples. Therefore, the frequency distribution of V has physical interpretation w.r.t. the sensitivity of system to certain spatial frequencies for the given voxels. This distribution would enable to extract the primary space Mprime (major components) from the null space Mnull (e.g., noise).26,41 In other words, distinguishing the strengths and deficiencies of these systems. The value of σi can be linked to the energy level or stability parameters of these components. In principle, the eigenvalues (or similarly singular values) are providing a measure of fidelity of these components such that the ith eigenvalue exemplifies the ith eigenvector’s strength. In general, the most salient features of the matrix are stored in the first few eigenvectors (Mprime) where σi>0. This boundary is referred to as the rank of matrix κ. Meanwhile, the sum of the eigenvalues or the nuclear norm of the matrix represents the energy of the matrix. Thus, the spectrum of eigenvalues, which is a generalized form of the Fourier power spectrum, allows discrimination between the primary and null space [Figs. 2(d) and 2(f)].

Eq. (11)

M=Mprim+Mnull,

Eq. (12)

Mprim=i=1κσiuivi*,

Eq. (13)

(MMprim)22=σκ+12.

The system with larger κ is imputed to the superior focusing competence26 and more accurate set of acquired data. Eigenvalues with the associated eigenvectors are pointing out the maximum sensitivity of the system toward the direction of incoming wave, hence specifying the object features that are more accurately estimated. For a spatially variant system, this can be estimated by the Fourier of eigenvectors as well. Moreover, these vectors are containing the information about the missing projections, which implies the partial view angle [Fig. 2(d)] and resolvability of the spatial frequency component, hence redundancy of projections.

Subsequently, eigenanalysis estimates whether the formed image misses these angular frequencies, even if OPUS is an ideal reflection mode imaging system (i.e., infinite detectors). The quality and quantity of angular frequencies that an aperture can generate or similarly receive is contained in the model matrix [Fig. 2(d)]. Therefore, eigenanalysis provides a simple but credible mechanism for the comparative appraisal between two or more imaging systems.

2.4.

Array Design

By taking into account the light delivery housing for perpendicular illumination, we are exploring the optimum geometrical properties for the OPUS handheld probe. The design array supposedly provides enough flexibility in elements’ geometry and arrangement such that the periodicity is minimized and the agreement among directivity, sensitivity, and spatial aliasing can be settled. To begin with, segmented annular array of point-like transducers is considered. The elements are distributed equidistantly in a ring fashion around the optical probes, enabling real-time volumetric acquisition. The 11-mm lumen is considered for the optical probe, which is in agreement with the limited 128 elements of 1λ width for the central frequency of 5 MHz with 80% fractional bandwidth.23 Indeed, the transducer frequency response is of crucial importance and has a determinant role in the so called relative spatial resolution; defined as the ratio of the penetration limit to the depth resolution. The aforementioned frequency response would allow the detection of OA sources as deep as 38 mm ex vivo and 20 mm in vivo,42,43 ideally down to the size of 0.3 mm.44

First, we explore the effect of element number, which has a reciprocal relationship to the interelement spacing. The optimization is pursued for the element size and geometrical distribution. The directive angular view (β) for the element edge a (height or width) is compensated by the negative focusing such that β6dB=2arcsin(0.6λc/a),45 for λc is the wavelength corresponding to the central frequency. We simulated negative focusing for every element in the convex structure as if the focus is behind the transducer (Fig. 1). For the numerical calculation and in silico studies we found the well-known FieldII simulations toolbox46 convenient.

3.

Results

3.1.

Voxel Crosstalk Matrix

Herein, the result of the voxel crosstalk analysis that estimates the spatial sensitivity and spatial aliasing for the given model matrix is presented. Figure 3 shows the effect of element number for four annular arrays composed of 32, 64, 128, and 256 point-like elements (0.5λ size) and an array of 128λ-width elements. Three axial slices within the FoV of 18×18×20  mm3 has been visualized, each composed of 100×100 voxels. The illustrated x-y planes are parallel to the surface of the aperture, at the depth of, respectively, 1, 2, and 3 cm away from the aperture [Fig. 3(I)]. The spatial aliasing for the most and least sensitive voxels in the FoV is provided in Fig. 3(II). Presented in Fig. 3(III), the two metrics of RMSE and MAE are negatively oriented scores and their parallel variation in the same order suggests the existence of many small errors. The variance of the errors is corresponding to the absolute difference between the values of two metrics.39 The qualitative and quantitative analyses indicate the improvement in sensitivity and aliasing with the increase in transducer counts. However, for the case where the overall active area is the same, the aperture with larger elements showcases a better performance.

Fig. 3

Crosstalk analysis on the effect of element number for segmented annular array. I: 60 dB logarithmic scale (w.r.t. the pressure at the surface of transducers) spatial sensitivity. II: Parasitic contribution to the central voxels of the plane at z=1  cm (upper row) and marginal voxels (lower row) of the plane at z=3  cm. III: Quantitative metrics.

JBO_23_2_025004_f003.png

Figure 4 presents the results of voxel crosstalk analysis w.r.t. the height of the transducers while the interelement spacing is limited to λ. The larger elements give rise to higher sensitivity and lesser spatial aliasing. More voxels are contributing to the parasitic signals but in lesser values, which also means lesser sidelobe value. The quantitative analysis confirms the qualitative interpretation [Fig. 4(III)]. As the size of element increases, the values of both metrics decrease but the difference between them increases. Due to its point-like transducers, the aperture with 0.5λ element’s height22 is exempt and here is set as the reference for quantitative analysis.

Fig. 4

Crosstalk analysis on the effect of element height for segmented annular array. I: 60 dB logarithmic scale spatial sensitivity. II: Parasitic contribution to the central voxel of the plane at z=1  cm (upper row) and marginal voxel (lower row) of the plane at z=3  cm. III: Quantitative metrics.

JBO_23_2_025004_f004.png

In pursuit of an aperture with lesser periodicity5,18 (i.e., aliasing), Fig. 5 compares the result of voxel crosstalk analysis for three virtual apertures with rather same element size but different arrangements. The performance of segmented annular array of 1λ×20λ element size is shown alongside the annular circular array and annular spiral array of 5λ×5λ element size [Fig. 5(IV)]. For the last two, the uniform distribution of energy over the FoV is expected due to the rather sparser geometry. Yet, the lack of uniform overlap among the elements’ view angle costs the energy in the central area, shown in Fig. 5(I). Figure 5(III) outlines the quantitative analysis for the three apertures. The segmented annular array showcases lesser errors and lesser difference between RMSE and MAE, an indication for lesser variance in the error.39

Fig. 5

Qualitative and quantitative interpretation of crosstalk analysis on the effect of elements arrangement. I: 60 dB logarithmic scale spatial sensitivity. II: Parasitic contribution to the central voxel of the plane at z=1  cm (upper row) and marginal voxel (lower row) of the plane at z=3  cm. III: Quantitative metrics. IV: Array geometries.

JBO_23_2_025004_f005.png

3.2.

Eigenanalysis

In this section, we summarize the effect of aperture partial view on the information loss, using eigenanalysis. The system comparison depends both on the number of effective components and the structure of the singular vectors. As the arrays are circularly symmetrical, we speculate the variations over four 18 mm lines of 200 control points at four different depths. The control points lie on a plane parallel to the aperture. Figure 6 shows the spatial Fourier transform of the right-hand side singular vectors, for four axial planes away from the surface of aperture, respectively, at 1, 10, 20, and 30 mm.

Fig. 6

The spatial Fourier transform of eigenvectors of the model matrix for 200 control points along the central line at four different axial planes. The white-dashed line indicates the borderline between the Mprime and Mnull.

JBO_23_2_025004_f006.png

In general, there are less available spectral components with the depth (columns), recalling that OPUS is a spatially variant imaging system. For the case where the object plane is near the aperture (first row of Fig. 6), the lower spatial frequencies are associated with discrepancies due to the limited directivity of large elements (the left side of the first white-dashed line). On the other side, the higher frequency components in the right side of the second dashed line are not containing information but noise. The rather meaningful information is limited to the area between the two dashed lines, thus less sensitivity to generated OA clutters.47 For the other planes, where the control points are farther from the aperture, the lower angular frequency components are stronger and contain resolvable structures. The separating border between the noise and meaningful data is highlighted by the white-dashed line. The adjacent number is pinpointing the turning point of which segregates the resolvable and unresolvable spectral information. One can notice that for the segmented annular array, the larger element, hence the larger aperture is accompanied by more and finer definition of the frequency components where each meaningful singular vector is associated with narrower spectrum.

By analogy to the segmented annular array, both annular circular and annular spiral arrays are containing more robust DC components for the axial planes in the far-field (1  cm). However, the number of valuable components (marked next to the dashed line) remains comparative. In addition, the rippled arms of annular circular and annular spiral arrays are indicating less resolvability for the frequency components and more aliasing. This is not observed in the segmented annular array where the thinner arms are implying a better resolvability for frequency components. Therefore, the imaging performance of segmented annular array can outperform the other layouts, for both OA and ultrasound imaging.

4.

Discussion and Conclusions

This study has detailed the theory of a framework in which the efficiency of a generic assessment tool in evaluating the performance of array imaging was demonstrated. The main focus of this work was design and characterization of a high compelling volumetric handheld probe for OPUS imaging. On that note, several substantial steps have been taken.

An underpinning step of this work was deriving a discrete linear model that explains the imaging performance of the designed system. The calculation is based on the system impulse response that correlates the sources to the recorded signals via the model matrix of the aperture. The purpose of developing the model was mainly to characterize and further optimize the imaging aptitude of the array as a spatially variant limited view imaging system. It is worth noting that the same model can be employed in conjunction with image reconstruction as an accurate imaging operator.

A fruitful comparison assessment tool that fully investigates the dynamics of the imaging device and allows for optimization w.r.t. the imaging performance has been followed. In particular, the effect of number, size, and geometrical location of the elements and their optimum configurations was considered. By employing the spatial crosstalk, potentially a principled argument is provided for comparative studies among the performances of different imaging systems. On the other hand, the eigenanalysis is providing an argument w.r.t. partial view angle. Together they enable to draw a valid conclusion about imaging performance of the imaging system.

Finally, the avails of the annular probes in the OPUS imaging by elaborating the appraisal analysis were demonstrated. Throughout the analysis, it was found that the annular geometry provides a unique layout for moderately large aperture containing limited number of elements. In addition, the central cavity preserves the housing for in-line arrangement between the optical source and transducers. Owing to its circular symmetry and inherent lumen, uniform excitation/insonification within the FoV is achievable [Fig. 4(I)], which is suitable for volumetric imaging. These studies confirmed that the superior performance is accompanied by increasing the number and size of the elements as long as the directivity is compensated. The latter can take place by the help of negative focusing via either of convex transducer or negative lens. It was shown that for these elements, the least aliasing does not necessarily correspond to the aperture with periodicity. Overall, the performance of segmented annular array in imaging outweighs the rest of geometries, considering the resolution, detectability, and uniform response for both modalities. For the sake of simplicity, we only considered flash-echo mode for US imaging. However, the approach can be extended to more sophisticated beamforming algorithms where the projections and therefore contrast and resolution are increased. One future research line is the objective assessment of the evaluated arrays based on the final image quality. However, the risk of bias evaluation with respect to the merits of reconstruction algorithm remains.

In conclusion, this work presents the development of a multipurpose handheld probe for volumetric OA ultrasound imaging. The theory behind this development has been established, and the dynamics of a new geometry for volumetric OPUS imaging has been assessed. Last but not least, we restricted our analysis to the geometrical properties of the array. Further optimization w.r.t. the transducing properties is deferred to future research.

Disclosures

The authors have no conflicts of interest to declare.

Acknowledgments

This work was supported by the People Programme (Marie Curie Actions) of the European Union’s Seventh Framework Programme FP7/2007-2013/ under REA grant agreement no. 317526 within the framework of OILTEBIA project. D. V. acknowledges the support of LABEX CELYA (ANR-10-LABX-0060) of Universite de Lyon, within the program Investissements d’Avenir (ANR-11-IDEX-0007) operated by the French National Research Agency (ANR).

References

1. 

P. N. T. Wells, Scientific Basis of Medical Imaging, Churchill Livingstone, Edinburgh (1982). https://doi.org/10.1063/1.2915500 Google Scholar

2. 

P. J. van den Berg, K. Daoudi and W. Steenbergen, “Pulsed photoacoustic flow imaging with a handheld system,” J. Biomed. Opt., 21 (2), 026004 (2016). https://doi.org/10.1117/1.JBO.21.2.026004 JBOPFO 1083-3668 Google Scholar

3. 

Y. Wang et al., “In vivo three-dimensional photoacoustic imaging based on a clinical matrix array ultrasound probe,” J. Biomed. Opt., 17 (6), 061208 (2012). https://doi.org/10.1117/1.JBO.17.6.061208 JBOPFO 1083-3668 Google Scholar

4. 

M. Gerling et al., “Real-time assessment of tissue hypoxia in vivo with combined photoacoustics and high-frequency ultrasound,” Theranostics, 4 (6), 604 –613 (2014). https://doi.org/10.7150/thno.7996 Google Scholar

5. 

A. Ramalli et al., “Density-tapered spiral arrays for ultrasound 3-D imaging,” IEEE Trans. Ultrason. Ferroelectr. Freq. Control, 62 (8), 1580 –1588 (2015). https://doi.org/10.1109/TUFFC.2015.007035 ITUCER 0885-3010 Google Scholar

6. 

G. Paltauf, R. Nuster and P. Burgholzer, “Weight factors for limited angle photoacoustic tomography,” Phys. Med. Biol., 54 (11), 3303 –3314 (2009). https://doi.org/10.1088/0031-9155/54/11/002 PHMBA7 0031-9155 Google Scholar

7. 

E. Mercep et al., “Hybrid optoacoustic tomography and pulse-echo ultrasonography using concave arrays,” IEEE Trans. Ultrason. Ferroelectr. Freq. Control, 62 (9), 1651 –1661 (2015). https://doi.org/10.1109/TUFFC.2015.007058 ITUCER 0885-3010 Google Scholar

8. 

J. Gateau et al., “Improving visibility in photoacoustic imaging using dynamic speckle illumination,” Opt. Lett., 38 (23), 5188 –5191 (2013). https://doi.org/10.1364/OL.38.005188 OPLEDP 0146-9592 Google Scholar

9. 

L. Wang et al., “Ultrasonic-heating-encoded photoacoustic tomography with virtually augmented detection view,” Optica, 2 (4), 307 –312 (2015). https://doi.org/10.1364/OPTICA.2.000307 Google Scholar

10. 

R. Ellwood et al., “Photoacoustic imaging using acoustic reflectors to enhance planar arrays,” J. Biomed. Opt., 19 (12), 126012 (2014). https://doi.org/10.1117/1.JBO.19.12.126012 JBOPFO 1083-3668 Google Scholar

11. 

B. Huang et al., “Improving limited-view photoacoustic tomography with an acoustic reflector,” J. Biomed. Opt., 18 (11), 110505 (2013). https://doi.org/10.1117/1.JBO.18.11.110505 JBOPFO 1083-3668 Google Scholar

12. 

J. A. Jensen et al., “Sarus: a synthetic aperture real-time ultrasound system,” IEEE Trans. Ultrason. Ferroelectr. Freq. Control, 60 (9), 1838 –1852 (2013). https://doi.org/10.1109/TUFFC.2013.2770 ITUCER 0885-3010 Google Scholar

13. 

G. R. Lockwood et al., “Optimizing the radiation pattern of sparse periodic linear arrays,” IEEE Trans. Ultrason. Ferroelectr. Freq. Control, 43 (1), 7 –14 (1996). https://doi.org/10.1109/58.484457 ITUCER 0885-3010 Google Scholar

14. 

A. Austeng and S. Holm, “Sparse 2-D arrays for 3-D phased array imaging-design methods,” IEEE Trans. Ultrason. Ferroelectr. Freq. Control, 49 (8), 1073 –1086 (2002). https://doi.org/10.1109/TUFFC.2002.1026019 ITUCER 0885-3010 Google Scholar

15. 

J. T. Yen, J. P. Steinberg and S. W. Smith, “Sparse 2-D array design for real time rectilinear volumetric imaging,” IEEE Trans. Ultrason. Ferroelectr. Freq. Control, 47 (1), 93 –110 (2000). https://doi.org/10.1109/58.818752 ITUCER 0885-3010 Google Scholar

16. 

M. Karaman et al., “Minimally redundant 2-D array designs for 3-D medical ultrasound imaging,” IEEE Trans. Med. Imaging, 28 (7), 1051 –1061 (2009). https://doi.org/10.1109/TMI.2008.2010936 ITMID4 0278-0062 Google Scholar

17. 

J. W. Choe et al., “Real-time volumetric imaging system for CMUT arrays,” in IEEE Int. Ultrasonics Symp. (IUS), 1064 –1067 (2011). https://doi.org/10.1109/ULTSYM.2011.0261 Google Scholar

18. 

O. Martínez-Graullera et al., “2D array design based on fermat spiral for ultrasound imaging,” Ultrasonics, 50 (2), 280 –289 (2010). https://doi.org/10.1016/j.ultras.2009.09.010 ULTRA3 0041-624X Google Scholar

19. 

E. Roux et al., “2-D ultrasound sparse arrays multidepth radiation optimization using simulated annealing and spiral-array inspired energy functions,” IEEE Trans. Ultrason. Ferroelectr. Freq. Control, 63 (12), 2138 –2149 (2016). https://doi.org/10.1109/TUFFC.2016.2602242 ITUCER 0885-3010 Google Scholar

20. 

X. L. Deán-Ben and D. Razansky, “Portable spherical array probe for volumetric real-time optoacoustic imaging at centimeter-scale depths,” Opt. Express, 21 (23), 28062 –28071 (2013). https://doi.org/10.1364/OE.21.028062 OPEXFF 1094-4087 Google Scholar

21. 

T. F. Fehm, X. L. Deán-Ben and D. Razansky, “Four dimensional hybrid ultrasound and optoacoustic imaging via passive element optical excitation in a hand-held probe,” Appl. Phys. Lett., 105 (17), 173505 (2014). https://doi.org/10.1063/1.4900520 APPLAB 0003-6951 Google Scholar

22. 

S. J. Norton, “Synthetic aperture imaging with arrays of arbitrary shape. II. The annular array,” IEEE Trans. Ultrason. Ferroelectr. Freq. Control, 49 (4), 404 –408 (2002). https://doi.org/10.1109/58.996557 ITUCER 0885-3010 Google Scholar

23. 

M. A. Kalkhoran et al., “Volumetric pulse echo and optoacoustic imaging by elaborating a weighted synthetic aperture technique,” in IEEE Int. Ultrasonics Symp. (IUS), 1 –4 (2015). https://doi.org/10.1109/ULTSYM.2015.0238 Google Scholar

24. 

M. A. Kalkhoran, F. Varray and D. Vray, “Dual frequency band annular probe for volumetric pulse-echo optoacoustic imaging,” Phys. Proc., 70 1104 –1108 (2015). https://doi.org/10.1016/j.phpro.2015.08.236 PPHRCK 1875-3892 Google Scholar

25. 

M. Pramanik, G. Ku and L. V. Wang, “Tangential resolution improvement in thermoacoustic and photoacoustic tomography using a negative acoustic lens,” J. Biomed. Opt., 14 (2), 024028 (2009). https://doi.org/10.1117/1.3103778 JBOPFO 1083-3668 Google Scholar

26. 

M. Tanter et al., “Optimal focusing by spatio-temporal inverse filter. I. Basic principles,” J. Acoust. Soc. Am., 110 (1), 37 –47 (2001). https://doi.org/10.1121/1.1377051 JASMAN 0001-4966 Google Scholar

27. 

R. J. Zemp, C. K. Abbey and M. F. Insana, “Linear system models for ultrasonic imaging: application to signal statistics,” IEEE Trans. Ultrason. Ferroelectr. Freq. Control, 50 (6), 642 –654 (2003). https://doi.org/10.1109/TUFFC.2003.1209551 ITUCER 0885-3010 Google Scholar

28. 

M. Roumeliotis et al., “Analysis of a photoacoustic imaging system by the crosstalk matrix and singular value decomposition,” Opt. Express, 18 (11), 11406 –11417 (2010). https://doi.org/10.1364/OE.18.011406 OPEXFF 1094-4087 Google Scholar

29. 

P. R. Stepanishen, “Transient radiation from pistons in an infinite planar baffle,” J. Acoust. Soc. Am., 49 (5B), 1629 –1638 (1971). https://doi.org/10.1121/1.1912541 JASMAN 0001-4966 Google Scholar

30. 

Z. Wang, S. Ha and K. Kim, “A new design of light illumination scheme for deep tissue photoacoustic imaging,” Opt. Express, 20 (20), 22649 –22659 (2012). https://doi.org/10.1364/OE.20.022649 OPEXFF 1094-4087 Google Scholar

31. 

P. M. Morse and K. U. Ingard, Theoretical Acoustics, Princeton university press, Princeton, New Jersey (1968). https://doi.org/10.1119/1.1976432 Google Scholar

32. 

C. Prada et al., “Decomposition of the time reversal operator: detection and selective focusing on two scatterers,” J. Acoust. Soc. Am., 99 (4), 2067 –2076 (1996). https://doi.org/10.1121/1.415393 JASMAN 0001-4966 Google Scholar

33. 

A. Lhémery, “Impulse-response method to predict echo-responses from targets of complex geometry. Part I: theory,” J. Acoust. Soc. Am., 90 (5), 2799 –2807 (1991). https://doi.org/10.1121/1.401877 JASMAN 0001-4966 Google Scholar

34. 

F. Lingvall and T. Olofsson, “On time-domain model-based ultrasonic array imaging,” IEEE Trans. Ultrason. Ferroelectr. Freq. Control, 54 (8), 1623 –1633 (2007). https://doi.org/10.1109/TUFFC.2007.433 ITUCER 0885-3010 Google Scholar

35. 

M.-L. Li, Y.-C. Tseng and C.-C. Cheng, “Model-based correction of finite aperture effect in photoacoustic tomography,” Opt. Express, 18 (25), 26285 –26292 (2010). https://doi.org/10.1364/OE.18.026285 OPEXFF 1094-4087 Google Scholar

36. 

J. Qi and R. H. Huesman, “Wavelet crosstalk matrix and its application to assessment of shift-variant imaging systems,” IEEE Trans. Nucl. Sci., 51 (1), 123 –129 (2004). https://doi.org/10.1109/TNS.2003.822997 IETNAE 0018-9499 Google Scholar

37. 

P. Wong et al., “Objective assessment and design improvement of a staring, sparse transducer array by the spatial crosstalk matrix for 3d photoacoustic tomography,” PLoS One, 10 (4), e0124759 (2015). https://doi.org/10.1371/journal.pone.0124759 POLNCL 1932-6203 Google Scholar

38. 

H. H. Barrett et al., “Objective assessment of image quality. II. Fisher information, Fourier crosstalk, and figures of merit for task performance,” J. Opt. Soc. Am. A, 12 (5), 834 –852 (1995). https://doi.org/10.1364/JOSAA.12.000834 JOAOD6 0740-3232 Google Scholar

39. 

C. J. Willmott and K. Matsuura, “Advantages of the mean absolute error (MAE) over the root mean square error (RMSE) in assessing average model performance,” Clim. Res., 30 (1), 79 –82 (2005). https://doi.org/10.3354/cr030079 CLREEW 0936-577X Google Scholar

40. 

T. Chai and R. R. Draxler, “Root mean square error (RMSE) or mean absolute error (MAE)?–arguments against avoiding RMSE in the literature,” Geosci. Model Dev., 7 (3), 1247 –1250 (2014). https://doi.org/10.5194/gmd-7-1247-2014 Google Scholar

41. 

T. M. Rossi and I. M. Warner, “Rank estimation of emission excitation matrixes using frequency analysis of eigenvectors,” Anal. Chem., 58 (4), 810 –815 (1986). https://doi.org/10.1021/ac00295a035 Google Scholar

42. 

K. Song and L. V. Wang, “Deep reflection-mode photoacoustic imaging of biological tissue,” J. Biomed. Opt., 12 (6), 060503 (2007). https://doi.org/10.1117/1.2818045 JBOPFO 1083-3668 Google Scholar

43. 

L. V. Wang, “Multiscale photoacoustic microscopy and computed tomography,” Nat. Photonics, 3 (9), 503 –509 (2009). https://doi.org/10.1038/nphoton.2009.157 NPAHBY 1749-4885 Google Scholar

44. 

V. G. Andreev, A. A. Karabutov and A. A. Oraevsky, “Detection of ultrawide-band ultrasound pulses in optoacoustic tomography,” IEEE Trans. Ultrason. Ferroelectr. Freq. Control, 50 (10), 1383 –1390 (2003). https://doi.org/10.1109/TUFFC.2003.1244756 ITUCER 0885-3010 Google Scholar

45. 

B. Woodward, S. Hole and W. Forsythe, “Transducer design for a correlation log,” Ultrasonics, 31 (1), 21 –34 (1993). https://doi.org/10.1016/0041-624X(93)90029-Y ULTRA3 0041-624X Google Scholar

46. 

J. A. Jensen, “Field: a program for simulating ultrasound systems,” in 10th Nordicbaltic Conf. on Biomedical Imaging, 351 –353 (1996). Google Scholar

47. 

M. Jaeger, J. C. Bamber and M. Frenz, “Clutter elimination for deep clinical optoacoustic imaging using localised vibration tagging (LOVIT),” Photoacoustics, 1 (2), 19 –29 (2013). https://doi.org/10.1016/j.pacs.2013.07.002 Google Scholar

Biography

Mohammad Azizian Kalkhoran graduated from Heidelberg University, Germany, in 2013. Currently, he is a PhD student at Creatis Laboratory at INSA Lyon. His current research interests include hybridized optoacoustic and ultrasound, optical microscopy, and tomography and fluorescent imaging.

Didier Vray is currently a professor of Signal Processing and Computer Sciences at INSA-Lyon, France. Since he joined the research laboratory CREATIS, his main research interest has focused on ultrasound medical imaging. His research includes vascular imaging, flow imaging, tissue motion estimation, and bimodality US/optics imaging. He is the author of more than 100 scientific publications and 2 issued patents in this field.

© 2018 Society of Photo-Optical Instrumentation Engineers (SPIE) 1083-3668/2018/$25.00 © 2018 SPIE
Mohammad Azizian Kalkhoran and Didier Vray "Theoretical characterization of annular array as a volumetric optoacoustic ultrasound handheld probe," Journal of Biomedical Optics 23(2), 025004 (27 February 2018). https://doi.org/10.1117/1.JBO.23.2.025004
Received: 9 October 2017; Accepted: 25 January 2018; Published: 27 February 2018
Lens.org Logo
CITATIONS
Cited by 4 scholarly publications and 1 patent.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Imaging systems

Ultrasonography

Transducers

Optoacoustics

Acoustics

Spatial frequencies

Imaging arrays

Back to Top