Open Access
16 March 2022 Normalization of optical fluence distribution for three-dimensional functional optoacoustic tomography of the breast
Author Affiliations +
Abstract

Significance: In three-dimensional (3D) functional optoacoustic tomography (OAT), wavelength-dependent optical attenuation and nonuniform incident optical fluence limit imaging depth and field of view and can hinder accurate estimation of functional quantities, such as the vascular blood oxygenation. These limitations hinder OAT of large objects, such as a human female breast.

Aim: We aim to develop a measurement-data-driven method for normalization of the optical fluence distribution and to investigate blood vasculature detectability and accuracy for estimating vascular blood oxygenation.

Approach: The proposed method is based on reasonable assumptions regarding breast anatomy and optical properties. The nonuniform incident optical fluence is estimated based on the illumination geometry in the OAT system, and the depth-dependent optical attenuation is approximated using Beer–Lambert law.

Results: Numerical studies demonstrated that the proposed method significantly enhanced blood vessel detectability and improved estimation accuracy of the vascular blood oxygenation from multiwavelength OAT measurements, compared with direct application of spectral linear unmixing without optical fluence compensation. Experimental results showed that the proposed method revealed previously invisible structures in regions deeper than 15 mm and/or near the chest wall.

Conclusions: The proposed method provides a straightforward and computationally inexpensive approximation of wavelength-dependent effective optical attenuation and, thus, enables mitigation of the spectral coloring effect in functional 3D OAT imaging.

1.

Introduction

Optoacoustic tomography (OAT), also known as photoacoustic computed tomography (PACT), is an emerging imaging modality that shows promise in sensitivity for breast cancer detection, especially in dense breasts.14 OAT images exhibit greater detection sensitivity for highly vascularized, i.e., aggressive, breast tumors,5 and greater diagnostic specificity for all tumors over other modalities, such as x-ray mammography and ultrasound.1,5,6 Another advantage of OAT over x-ray mammography is that OAT does not involve ionizing radiation.1,2 Due to advances in OAT systems design and image reconstruction, a three-dimensional (3D) volumetric scan of the entire breast is now possible.3,5,7 At a single near-infrared illumination wavelength, natural chromophores in the breast tissue, such as hemoglobin, act as endogenous OAT contrast agents. From external ultrasound measurements of the pressure induced by the laser pulses, the spatial distribution of the chromophores can be estimated; this provides a qualitative, anatomical measure of the blood vasculature.3,5,710 Measurements from multiple illumination wavelengths matching the local maximum, minimum, and isosbestic point of deoxy- and oxyhemoglobin can be reconstructed into quantitative estimates of the blood oxygen saturation.5,1116 This technology referred to as quantitative OAT (qOAT), also known as quantitative PACT, when combined with ultrasound, provides both anatomical and functional information of the breast that can facilitate detecting tumor angiogenesis and hypoxia.1

In 3D OAT breast imaging, it is not feasible to deliver the optical fluence uniformly throughout the whole breast volume, due to optical attenuation in tissue and design constraints of the imaging systems.57,9,13,17 Whereas distributions of the optical absorption coefficient are determined by tissue types and physiological status, initial pressure distributions decay with depth in the tissue because of light attenuation. Furthermore, the attenuation is wavelength-dependent. Therefore, direct application of spectral linear unmixing methods to reconstructed OAT images, which correspond to the estimated initial pressure distributions, results in inaccurate estimates of blood oxygen saturation.14,15,18,19

To improve visualization of the reconstructed volumetric images, the optoacoustic imaging community has utilized a commercial tool, AMIRA (Thermo Fisher Scientific),20 and open source, interactive tools such as ImageJ (Wayne Rasband),21 ParaView (Kitware),22 and 3D Slicer (Kitware).23 Also, a 3D PHOVIS (POSTECH, Korea)24 has been recently released that is developed specifically for optoacoustic imaging. However, these tools do not provide physics-informed image processing for contrast enhancement at depth in reconstructed OAT volumes. In addition, these visualization tools are semi-automatic and require substantial manual intervention by the user. Pattyn et al.25 proposed a model-based method to compensate for the optical fluence distribution within a heterogeneous physical phantom that mimics a breast. Monte Carlo (MC) simulation was employed, and known optical properties of the phantom were assumed. However, in practice, the distributions of optical properties within the breast are generally unknown.

A straightforward physics-informed image processing method is proposed to compensate for both the nonuniform incident optical fluence at the breast surface and the wavelength- and depth-dependent optical attenuation within the breast, and the impact of the proposed method on accuracy of the linear unmixing of deoxy- and oxyhemoglobin is investigated. The contributions of this paper are twofold. First, this study establishes an implementation of the linear unmixing method for use with a large object such as a female breast. Second, the proposed method improves sensitivity of the 3D OAT breast imaging by improving contrast at depth.

The paper is organized as follows. Background materials, including existing illumination systems in 3D OAT breast imaging and spectral linear unmixing, are provided in Sec. 2. The proposed method is explained in Sec. 3, and the study description and evaluation metrics are provided in Sec. 4. Results from computer-simulation and experimental studies are presented in Sec. 5, and a discussion is given in Sec. 6. The conclusions of the study are provided in Sec. 7.

2.

Background

In OAT imaging, a short laser pulse is employed to irradiate an object at time t=0 and conversion of the absorbed optical energy into the thermal energy results in the generation of an initial pressure distribution p0(r,λ), where r=(x,y,z)R3 and λ is a wavelength. The pressure distribution subsequently propagates and is measured by multiple ultrasonic transducers located on a measurement aperture Ω0R3 that partially or completely surrounds the object. The propagated pressure wavefields, i.e., the optoacoustic signals, at time t>0 are denoted as p(r,t).

By solving the associated acoustic inverse problem,26 an estimate of the absorbed energy density distribution within the object can be obtained. Functional quantities such as the vascular blood oxygenation can be reconstructed via qOAT from multiwavelength measurements.5,1115,27

In most implementations of 3D OAT breast imaging, the patient lies prone on a bed with their breast located inside a water-filled bowl just below the surface plane of the bed. The breast is illuminated at a near-infrared wavelength with a short laser pulse.3,5,7,9,17,28 The induced pressure waves p(r,t) propagate out of the breast and are measured with the ultrawide-band ultrasonic transducers.16

2.1.

Illumination in 3D OAT Breast Imaging

Several different 3D OAT breast imaging systems have been proposed and established, but their data-acquisition principles are similar.3,5,7,9,17,28 Existing systems for 3D OAT breast imaging are equipped with two subsystems: an illumination system and an optoacoustic signal detection system. The focus here is on the illumination system. A common feature in the existing illumination systems is that the light is delivered in a radially symmetric pattern to the breast surface from either single3,7,9,17,28 or multiple5 light sources. The laser fluences involved are well below the maximum permissible exposure for skin defined by the American National Standards Institute (ANSI).29

Table 1 presents salient design features of the light-delivery systems employed by the five OAT breast imaging systems shown in Fig. 1. These systems deliver the light to the entire breast surface. However, less optical energy per unit area (optical fluence) is delivered to regions near the chest wall compared with the center of the breast, and this imbalanced distribution of the incident optical fluence causes lower voxel brightness near the chest wall in the reconstructed OAT images.

Table 1

Light-delivery systems employed by the five OAT breast imaging systems in Fig. 1.

Ref.λ (nm)Salient design features
7,9755, 795• A conical laser beam is emitted from below and then reflected through a planoconvex lens and a holographic diffuser [Fig. 1(a)].
• The breast is contained by a 0.5-mm-thick polyethylene terephthalate glycol-modified (PETG) cup which is optically and acoustically transparent.
31064• A donut-shaped laser beam is emitted from below and then reflected through an axicon lens and an engineered diffuser [Fig. 1(b)].
• An agar pillow is used to slightly compress the breast.
28755, 1064• A laser beam is split into a bottom beam and side beams, and then those are diverged via concave lenses [Fig. 1(c)].
• The side beams are emitted slightly upward from nine optical fiber bundles that rotate around the breast in discrete steps.
17532• A ring-shaped light beam is formed via a cone-shaped reflector, stationary conical reflector, and mobile conical reflector [Fig. 1(d)].
• The mobile conical reflector vertically moves together with the ring-shaped detector array during the scan.
5755 to 1064• Laser beams are emitted from five fiber-optic segments that are constrained to the surface of a light paddle that rotates around the breast in discrete steps [Fig. 1(e)].
• The PETG breast cup is 0.1 mm thick.

Fig. 1

Illumination in 3D OAT imaging systems for the breast: (a) Kruger et al.9 and Toi et al.,7 (b) Lin et al.,3 (c) Schoustra et al.,28 (d) Alshahrani et al.,17 and (e) Oraevsky et al.5

JBO_27_3_036001_f001.png

2.2.

Spectral Linear Unmixing

Several qOAT methods have been proposed to estimate the optical absorption coefficient μa(r,λ) and/or oxygen saturation distribution.1315,3037 Among them, a two-step spectral linear unmixing approach has been widely used.5,13,18,3036 The first step of the method is OAT image reconstruction (i.e., acoustic inversion) to estimate the initial pressure distribution p0(r,λ) that is induced via the optical absorption and subsequent nonradiative relaxation of electronic energy into heat. The second step is to approximate the oxygen saturation distribution from the estimates of p0(r,λi)=Γμa(r,λi)ϕ(r,λi)11,38 acquired at multiple wavelengths (i=1,,n,nN). Here, Γ is the dimensionless Grüneisen parameter that can be considered constant for soft tissues.11,38,39

In unmixing methods, reconstructed estimates of p0(r,λi) are considered as surrogates of μa(r,λi). Unmixing procedures are predicated upon the relationship μa(r,λi)=ϵHb(λi)CHb(r)+ϵHbO2(λi)CHbO2(r), where ϵHb and ϵHbO2 are known wavelength-dependent molar extinction coefficients and CHb and CHbO2 denote molar concentrations of deoxy- and oxyhemoglobin, respectively.1113,18 The molar concentrations of CHb(r) and CHbO2(r) are computed as

Eq. (1)

[CHb(r)CHbO2(r)]=ϵ+[μa(r,λ1)μa(r,λn)],ϵ[ϵHb(λ1)ϵHbO2(λ1)ϵHb(λn)ϵHbO2(λn)],and  n2,
where ϵ+ is a pseudoinverse of the molar extinction coefficient matrix ϵ. Given CHb(r) and CHbO2(r), the oxygen saturation distribution is computed as sO2(r)=CHbO2(r)CHb(r)+CHbO2(r)×100%. The distribution of the total hemoglobin concentration is calculated as CtHb(r)=CHb(r)+CHbO2(r), and subsequently the blood vasculature can be detected.

In practice, the distribution of the optical fluence ϕ(r,λ) in breast tissues is not constant because of nonuniform illumination and optical attenuation. Once the light reaches the breast, the optical fluence decreases approximately exponentially with depth; this is described by the well-known Beer–Lambert law: ϕ(d,λ)=ϕ0exp(μeff(λ)d). Here, ϕ(d,λ) denotes optical fluence at a depth d and a wavelength λ, ϕ0 is the incident optical fluence to the breast surface (d=0), and μeff(λ) is an effective attenuation coefficient at the wavelength of λ that reflects both the scattering and absorption of light in tissues.6,12,40 In addition, it is challenging to uniformly deliver the light to the breast surface in 3D OAT imaging. Hence, the reconstructed p0(r,λ) exhibits undesirable variations in the voxel brightness according to design of the illumination system.5,6 This limits the visible depth and field of view in the reconstructed p0(r,λ).

Most significantly, the effective attenuation coefficient μeff(λ) is wavelength-dependent, which is known as the “spectral coloring effect.”41 Thus, in the linear unmixing employing multiple wavelength estimates of p0(r,λ) as surrogates of μa(r,λ), the oxygen saturation distribution cannot be accurately estimated without compensation for μeff(λ).14,15,19,42 Besides, p0(r,λ) is exponentially attenuated with depth, so the compensation for μeff(λ) is more important for large organs such as a female breast, that usually has the maximum depth larger than 20 mm, compared to relatively small organs such as skin with the maximum illumination depth of just a few millimeters.

3.

Normalization of Optical Fluence Distribution

The proposed method seeks to estimate and compensate for the nonuniform optical fluence distribution. This is referred as normalization of the optical fluence distribution hereafter. The method was designed based on a common feature of the existing illumination systems; specifically, that radially symmetric light delivery to the breast surface is employed. In this section, the details of the method are described based on the reference imaging system shown in Fig. 2, where a patient lies prone on the examination bed and the patient’s breast is located inside the breast cup. A spherical coordinate system is assumed, with the origin corresponding to the center of the breast cup [see Figs. 2(a) and 3(d)]. Here, θ is defined to be a polar angle from the positive z axis with 0θπ.

Fig. 2

3D OAT scan using LOUISA-3D: (a) breast scan schematic and (b) photograph of phantom scan. (b) Nonuniform illumination is observed on the surface of the tissue-mimicking physical phantom.

JBO_27_3_036001_f002.png

Fig. 3

(a)–(c) 3D OAT breast image (α^) of a healthy volunteer at a wavelength of 755 nm, scanned by TomoWave Laboratories using LOUISA-3D5 at MD Anderson Cancer Center, and (d) breast shapes in a given breast cup. Maximum voxel brightness projection (MVBP) along (a) x axis and cross-sections on (b) y-z plane at x=0 mm and (c) x-y plane at z=30  mm. The slice is indicated with a white dotted line in (a). The image (α^) was reconstructed using filtered backprojection (FBP)43 method. The brightness range of the images was adjusted for better visibility. In panels (b) and (c), a white solid circle indicates the location of the brightest voxel in the cross-section, and a cyan dashed line represents the approximated breast boundary.

JBO_27_3_036001_f003.png

In the proposed method, the nonuniform distribution of the optical fluence within the breast is estimated from the voxel values in the reconstructed 3D OAT image α^ that is an estimate of p0(r,λ) discretized employing a uniform Cartesian lattice. The following reasonable assumptions are made:

  • A1. The distribution of the incident optical fluence varies along the polar direction (θ) and is radially symmetric on x-y planes [see Fig. 3(a)];

  • A2. Blood vessels absorb more optical energy than other breast tissues do because deoxy- and oxyhemoglobin of red blood cells are the only optically absorbing chromophores at near-infrared wavelengths that are typically used in OAT breast imaging. Moreover, for wavelengths near 800 nm, artery and vein are nearly indistinguishable from each other in the reconstructed image α^12,44 [see Figs. 3(a)3(c)];

  • A3. Anatomically, at least one voxel corresponding to a blood vessel near the skin layer exists at any polar angle in the reconstructed image α^ [see Figs. 3(a)3(c)];

  • A4. The shape of the breast located inside a hemispherical stabilizer cup is a partial spheroid and static [see Fig. 3(d)];

  • A5. The Beer–Lambert law6,12,40 can be used to approximate the optical fluence distribution within the breast.

The normalization of the optical fluence distribution is conducted in the following order: (1) compensation for nonuniform incident optical fluence, (2) estimation of breast surface and depth of each voxel relative to the breast surface, and (3) compensation for the effective optical attenuation. The location of the breast surface must be known for optical attenuation compensation, but breast surface detection is highly challenging because the top skin layer (epidermis) can appear dimmer than the noise surrounding the breast due to the nonuniform incident optical fluence [see Figs. 3(a)3(c)]. Therefore, the incident optical fluence needs to be normalized in advance of the breast surface detection.

In the proposed method, a hemispherical breast stabilizer cup is assumed, as it is employed in several 3D OAT breast imaging systems.5,7,9 The cup is selected for each breast size, so it maintains the breast shape radially symmetric. Therefore, the only possible breast shape is a partial spheroid, as shown in Fig. 3(d). Whereas the nipple and areola absorb more light than the other breast tissues due to their high concentration of pigment, their impact on the optical fluence distribution within the breast is insignificant because of their relatively small volume.45 Thus, in the proposed method, the region 160  deg<θ180  deg, which was determined by an average diameter ratio of the breast and areola,45 is excluded from consideration.

The flowchart of the proposed method is provided in Fig. 4, and the details of each step are given in the following sections.

Fig. 4

Flowchart of normalization of optical fluence distribution in 3D OAT breast images.

JBO_27_3_036001_f004.png

3.1.

Compensation for Nonuniform Incident Optical Fluence

Under assumption A1, the distribution of the incident optical fluence is radially symmetric, which means it can be interpreted as a function of polar angle [Fig. 3(d)]. If strong optical absorbers that have similar μa values, such as blood vessels, are densely located near the object surface (d=0), the spatial distribution of p0 on the surface is proportional to the distribution of the incident optical fluence (p0ϕ). As mentioned earlier, the blood vessels in subdermal regions appear as the brightest voxels at any polar angle in the reconstructed image α^ [assumption A3; see Figs. 3(b) and 3(c)]. Thus, the distribution of the incident optical fluence can be approximated by the voxel brightness of the blood vessels near the breast skin layer (subdermal), i.e., maximum voxel brightness projection (MVBP) at discretized polar angles [θ]n(θiΔθ2,θi+Δθ2) with an increment Δθ of 1 deg. The polar angle [θ]n is calculated as cos1(zn/rn), where rn=xn2+yn2+zn2 is the distance of the n’th voxel from the origin, and xn, yn, and zn are the x, y, and z-coordinates of the n’th voxel in the uniform Cartesian grid, respectively.

An L-degree polynomial curve qL(θ) is fitted to the extracted maximum voxel brightness according to the discretized polar angles within the range of interest (90 deg to 160 deg), where the region containing the nipple and areola is excluded, as shown in Fig. 5(a). The estimated distribution of the incident optical fluence ϕ^0 is described as

Eq. (2)

[ϕ^0]n=qL([θ]n),n=1,N,
where N is the total number of voxels. The L is set depending on the illumination pattern. It was found that L=1 and L=2 were sufficient for accurately fitting the data in the experimental studies (Sec. 4.2) and computer-simulation studies (Sec. 4.1), respectively.

Fig. 5

Compensation for nonuniform incident optical fluence: (a) estimated incident optical fluence as a function of polar angle and (b) MVBP of 3D OAT breast image after the compensation (α^N0) along x axis. The results were obtained from Fig. 3(a). In panel (a), a black solid line indicates the maximum voxel brightness according to polar angles, and a red solid line represents a first-degree polynomial curve q1([θ]i) fitted to the maximum voxel brightness. The polar angles to the right of the blue dashed line correspond with nipple and areola in (a).

JBO_27_3_036001_f005.png

Elementwise division of ϕ^0 is applied to the reconstructed image α^, to compensate for the nonuniform incident optical fluence as shown in Fig. 5(b):

Eq. (3)

[α^N0]n=[α^]n[ϕ^0]n,n=1,,N.

A comparison of the images before and after the compensation is shown in Fig. 6

Fig. 6

Comparison of images before and after compensation for non-uniform incident optical fluence: MVBP of a 5-mm-thick y slice at y=0 mm (a) before (α^) and (b) after (α^N0) the compensation along y axis and (c) MVBP of their difference (α^N0α^) along y axis. The results were obtained from Fig. 3(a). The voxel brightness near the chest wall (θ=90  deg) in (a) is lower than in the region near the areola (θ160  deg) and, accordingly, the compensation procedure leads to a higher gain near the chest wall as shown in (b) and (c).

JBO_27_3_036001_f006.png

3.2.

Estimation of Breast Surface and Depth

Under assumption A3, the blood vessels located in subdermal regions can be employed to infer the breast surface in the reconstructed 3D OAT image α^. The average range of skin thickness of healthy female human breasts is between 0.5 and 2.4 mm.46 This is a relatively thin layer that attenuates light negligibly in comparison to attenuation in the bulk. To extract the voxels that belong to blood vessels in the close proximity of the breast surface, first a 3D median filter is applied to reduce the noise. Subsequently, the contrast of the resulting image is increased by elementwise square operation:

Eq. (4)

[α^N0]n=([med{α^N0}]n)2,n=1,,N,
where med{·} is a 3D median filter function with a 3-by-3-by-3 filter.

The voxels corresponding to the blood vessels near the breast surface are extracted using Otsu thresholding47 applied to α^N0. The set of the extracted voxels is defined as V={nN:[α^N0]nT} where T is an Otsu’s threshold. For each polar angle, the voxels in V that are the farthest from the z axis are selected to estimate the breast boundary. The estimated radius of the breast in the cross-section of slice (j’th z-slice), is given as

Eq. (5)

[ρz]j=max{[ρ]n}nSjV,
where [ρ]n is the distance of the n’th voxel from the z-axis calculated as xn2+yn2, and Sj denotes the set of all voxels, on the j’th z-slice, i.e., the x-y plane.

In Fig. 7(a), blue circle markers correspond to an estimate ρz for each z-slice. An elliptical curve is fit to ρz to obtain a smooth representation of the breast boundary according to assumption A4. The ellipse equation is (zzC)2a2+ρ2b2=1. Here, a and b are lengths along semimajor and semiminor axes, respectively, zC is a z-coordinate of the ellipse center. These parameters are determined by the elliptical curve fitting. The surface formed by rotation of the estimated elliptical curve [a red solid line in Fig. 7(a)] around the z axis is then used to approximate the breast surface as shown in Fig. 7(b). From this, a breast mask MBreast is built by assigning the value “1” to voxels inside the surface and “0” outside.

Fig. 7

Breast surface estimation: (a) estimated radii on x-y planes ρ^z; (b) estimated breast surface; and (c) estimated breast boundary on y-z plane at x=0 overlaid on the MVBP of α^N0 along the x axis.

JBO_27_3_036001_f007.png

Finally, the depth d at each voxel that is used in the Beer–Lambert law is approximated as the minimum distance from the breast surface, i.e., the estimated spheroid surface, using Newton’s method.48

3.3.

Compensation for the Effective Optical Attenuation

With consideration of assumptions A2 and A5, optical attenuation can be approximated as a function of depth from the decay of the voxel brightness inside the blood vessels in the reconstructed 3D OAT image. Specifically, such an approximation uses the maximum voxel brightness value in a certain depth range of (dmΔd2,dm+Δd2), where Δd is an increment of 1 voxel:

Eq. (6)

[α^BV]m=maxnCm{[α^]n}.
Here, Cm denotes a set of voxels in the m’th depth range. Figure 8 shows α^BV according to depth.

Fig. 8

Estimation of optical attenuation at a wavelength of 755 nm. A black solid line indicates maximum voxel brightness α^BV at a certain depth range of (dmΔd2,dm+Δd2), and a blue curve is the estimated optical attenuation ϕ^a. The μeff estimated from the 3D OAT breast image was 1.0041  cm1.

JBO_27_3_036001_f008.png

An exponential curve based on the Beer–Lambert law (assumption A5) is fit to α^BV as shown in Fig. 8. The estimated optical attenuation is expressed as

Eq. (7)

[f^a]n=c*exp(μeff*[d]n),n=1,,N,
where c* and μeff* are the prefactor and the effective optical attenuation coefficient estimated from the curve fitting, respectively.

Elementwise division of ϕ^a is applied to the image after normalization of nonuniform incident optical fluence α^N0 as follows:

Eq. (8)

[α^N]n={[α^N0]n[ϕ^a]n,if nMBreast0,otherwise.

4.

Study Description and Evaluation Metrics

4.1.

Computer-Simulation Studies

The proposed method was investigated and evaluated by use of physical measures of image quality and its impact on spectral linear unmixing. Results from the proposed method were compared with those from a general-purpose image contrast enhancement method. To investigate the impact of the proposed method on spectral linear unmixing, ground truth data are required. Accordingly, a realistic numerical breast phantom was generated, and a computer-simulation study was conducted.

4.1.1.

Realistic numerical breast phantom

The numerical breast phantom was created using a computational framework for virtual 3D OAT breast imaging trials developed by the authors of Ref. 49. This framework employs an open source tool from the U.S. Food and Drug Administration50 with modifications for use in OAT imaging.49,51 The produced numerical breast phantom consists of fat, skin, glandular, nipple, arterial, and venous tissues. The breast shape is a hemisphere with a radius of 60 mm, which is compatible with use of a breast stabilizer cup. The entire phantom was discretized using a uniform 3D Cartesian lattice with a voxel size of 0.25 mm.

4.1.2.

Functional, optical, and acoustic properties

Wavelength-dependent optical properties were assigned to each breast tissue by prescribing the composition of each tissue type in terms of total hemoglobin concentration, oxygen saturation, and volume fractions of blood, water, fat, and melanin.44 Illumination wavelengths of 757, 800 (the isosbestic point of deoxy- and oxyhemoglobin), and 850 nm were selected from near-infrared-I range (650 to 950 nm) that is commonly used in OAT breast imaging.5,7,9 While at least two wavelengths are required for the linear unmixing of deoxy- and oxyhemoglobin, additional wavelengths lead to more stable estimates. As data acquisition time increases proportionally to the number of illumination wavelengths used, OAT images at only a few wavelengths can be collected in clinically relevant settings. For this reason, two- and three-wavelength linear unmixing methods were utilized in the numerical studies. Regarding the acoustic properties of the numerical breast phantom, homogeneous speed of sound and no acoustic attenuation were assumed.52

4.1.3.

Simulation of optoacoustic signals

The optoacoustic signals were simulated in three dimensions using the GPU-accelerated MCXLAB software.53,54 The illumination geometry was configured to mimic LOUISA-3D5 (Sec. 2) where laser beams are cylindrically emitted from five linear fiber-optic segments on the surface of an arc-shaped light paddle that rotates around the breast in 20 discrete steps. In the MC simulation, to mimic the beam from each fiber-optic segment, five cone beams with a half-angle of 12.5 deg were employed. Consequently, a total of 500 cone beams were simulated for 20 illumination views. The light source positions were evenly distributed along the linear fiber-optic segments [Fig. 2(a)]. The incident beam direction was specified as perpendicular to the linear segments toward the origin of coordinates [Fig. 3(d)]. The number of photons simulated was 108 per cone beam, and the simulation duration was 5 ns. The size of a simulation domain was 340×340×170  voxels with a voxel size of 0.5 mm. Subsequently, the true initial pressure p0(r,λ) was computed via elementwise multiplication of the simulated optical fluence and optical absorption coefficient. A Grüneisen parameter Γ=1 was assumed, as is commonly done as constant for soft tissues.2,55

4.1.4.

Acoustic wave propagation and data acquisition

Acoustic wave propagation and data acquisition were simulated using the GPU-accelerated k-Wave toolbox.56 The measurement geometry was defined as in the LOUISA-3D system:5 an arc-shaped probe with 96 transducers collecting pressure data at 320 tomographic views (Fig. 2). A total of 1536 time samples were acquired at virtual transducers with a sampling frequency of 20 MHz. Idealized point-like transducers were assumed. Thermal acoustic noise was modeled as Gaussian noise with zero mean and standard deviation equal to 1% of the maximum signal strength, as was determined based on the in vivo breast data.

4.1.5.

Image reconstruction and processing

Noisy synthetic data were reconstructed using a GPU-accelerated FBP,43 implemented in C++ and CUDA.38,57 The size of the reconstructed volume was 480×480×240  voxels (120×120×60  mm3). Elapsed time for the image reconstruction was 40  s using four NVIDIA GeForce GTX 1080 GPUs. After the image reconstruction, k-means singular value decomposition dictionary learning58 was applied to reduce the noise.

For physical image quality evaluation, numerical results from the proposed method were compared with those from contrast limited adaptive histogram equalization (CLAHE),59 a method to enhance local contrast that is commonly used in medical images, such as ultrasound images,60 mammograms,61 and optical microangiographies.62 In OAT imaging, CLAHE is employed in a multispectral OAT system (iThera Medical, Germany).63 While several implementations of CLAHE are available for one- and two-dimensional images, an extension to 3D images was implemented for use in this study.

For detection and visualization of the blood vasculature, multiscale vessel enhancement filtering63,64 and Otsu thresholding47 were applied to the reconstructed initial pressure with no optical fluence normalization, CLAHE, and the proposed method. The vessel enhancement filter, also known as Frangi filter, detects tubular structures based on an eigenvalue analysis of the Hessian matrix of the image at multiple scales.64 The thicknesses of the detected structures are controlled through a set of scale parameters. In this work, the parameters with widths ranging from 1 to 5 voxels (0.25 to 1.25 mm) were chosen, as they are representatives of vessel diameters in the breast.65

To investigate spectral linear unmixing, molar concentrations of deoxy- and oxyhemoglobin were computed via Eq. (1), from which total hemoglobin concentration CtHb(r) and oxygen saturation sO2(r) were subsequently calculated. Results from the two- and three-wavelength linear unmixing methods, with no optical fluence normalization, CLAHE, and the proposed method, were compared.

4.2.

Studies with Clinical Data

Two clinical data sets corresponding to the right and left breast of a healthy volunteer were acquired by TomoWave Laboratories (Houston, Texas) using LOUISA-3D5 at MD Anderson Cancer Center and processed by the authors with institutional review board approval. The breasts were illuminated at a wavelength of 755 nm. The details of the illumination geometry of LOUISA-3D5 were given in Sec. 2. Acoustic measurements were collected with ultrawide-band (50 kHz to 6 MHz) ultrasonic transducers of size of 1.1×1.1  mm2. Additional details of the measurement geometry and image reconstruction were provided in Sec. 4.1. The image reconstruction and processing were conducted identically to those in computer-simulation studies. Experimental results from the proposed method were compared with those from CLAHE.59

4.3.

Evaluation Metrics

4.3.1.

Physical measures of image quality

The peak signal-to-noise ratio (PSNR)66 and structure similarity (SSIM) index67 were calculated. They are defined as

Eq. (9)

PSNR=10logMAXα2MSE,
and

Eq. (10)

SSIM(x,y)=(2μxμy+C1)(2σxy+C2)(μx2+μy2+C1)(σx2+σy2+C2).

In Eq. (9), MAXα is the maximum possible value of voxel brightness (e.g., 255 in 8-bit voxel values), and MSE is the mean squared error with respect to the ground truth, i.e., true μa distribution of the blood vessels. In Eq. (10), μx, μy, σx, σy, and σxy are the local means, standard deviations, and covariance of images x and y. Here, x and y correspond to the μa estimate, i.e., the reconstructed initial pressure with no normalization, CLAHE, and the proposed method, and the true μa distribution of the blood vessels, respectively. C1=(K1MAXα)2 and C2=(K2MAXα)2 in Eq. (10) denote stabilization constants, where K1=0.01 and K2=0.03 are the default values in the Image Processing Toolbox of MATLAB.

4.3.2.

Task-based measures of image quality

Blood vessel detectabily and artery/vein classification accuracy were used to evaluate the impact of the proposed method on spectral linear unmixing.

Based on the estimate of total hemoglobin concentration, the blood vessel voxels were detected via multiscale vessel enhancement filtering and Otsu thresholding. The detection performance was assessed using the detectability index DET defined as

Eq. (11)

DET=N^AVNA+NV×100%,
where NA and NV are the numbers of all voxels corresponding to arteries and veins in the numerical phanton, and N^AV is the number of voxels correctly detected as vasculature, respectively.

The accuracy of artery/vein classification was assessed under two different scenarios. In the calculation of the classification accuracy index ACC, the vascular structures of the numerical phantom were assumed as known, while in the detection-classification accuracy index (DACC), the vascular structures were estimated from the reconstructed OAT images as explained above. In both scenarios, an oxygenation level of 83.5% was used as the threshold for the classification of arteries and veins. This threshold corresponds to the arithmetic mean of the oxygenation level assigned to arteries (97%) and veins (70%) in the numerical phantom.

Specifically, in the scenario of known vascular structures, true artery rate (TAR), true vein rate (TVR), and classification accuracy index (ACC) were defined as

Eq. (12)

TAR=NTANA×100%,TVR=NTVNV×100%,andACC=NTA+NTVNA+NV×100%,
where NTA and NTV are the numbers of the voxels that are correctly classified as arteries and veins, respectively.

Similarly, in the scenario in which the vascular structures are also estimated from the spectral unmixing of the OAT images, detected true artery rate (DTAR), detected true vein rate (DTVR), and DACC were defined as

Eq. (13)

DTAR=N^TANA×100%,DTVR=N^TVNV×100%,andDACC=N^TA+N^TVNA+NV×100%,
where N^TA and N^TV are the numbers of the voxels that are detected and correctly classified as arteries and veins, respectively.

5.

Results

5.1.

Computer-Simulation Studies Results

5.1.1.

Physical measures of image quality

Figure 9 shows results obtained by applying CLAHE and the proposed optical fluence normalization method to the reconstructed estimate of p0. In the results from CLAHE [Fig. 9(c)] and the proposed method [Fig. 9(d)], more structures at depths deeper than 5 mm (green to red color) were revealed compared to the reconstructed initial pressure distribution in Fig. 9(b). The vasculature in the image produced using the proposed optical fluence normalization in Fig. 9(d) is visually more similar to that in the ground truth [Fig. 9(a)] than that produced by CLAHE in Fig. 9(c).

Fig. 9

Comparison between distributions of (a) the optical absorption coefficient μa, (b) initial pressure estimate reconstructed from the noisy measurements, simulated at a wavelength of 800 nm, using FBP with no normalization, and (c) images processed via CLAHE and (d) optical fluence normalization method. The images are presented as MVBP along y axis and color-encoded by depth. A depth range of 0 to 30 mm was visualized. A Jet color map in MATLAB was used to illustrate the breast tissues at different depths.

JBO_27_3_036001_f009.png

Figure 10 shows PSNR [Eq. (9)] and SSIM [Eq. (10)] comparisons between CLAHE [Fig. 9(c)] and the proposed method [Fig. 9(d)]. As shown in Fig. 10, the results of the proposed method showed higher PSNR and SSIM than those of no normalization and CLAHE for all three wavelengths.

Fig. 10

Comparison on PSNR and SSIM between no normalization (black), CLAHE (cyan), and the proposed method (red).

JBO_27_3_036001_f010.png

5.1.2.

Task-based measures of image quality

Figure 11 shows the detected blood vasculature and the estimated blood oxygenation using two- and three-wavelength linear unmixing methods with no optical fluence normalization [Fig. 11(a)], CLAHE [Fig. 11(b)], and the proposed method [Fig. 11(c)].

Fig. 11

Estimates of vascular blood oxygenation obtained using (a) no optical fluence normalization, (b) CLAHE, and (c) the proposed method. The used wavelength pairs are 757 and 800 nm (first column), 757 and 850 nm (second column), 800 and 850 nm (third column), and 757, 800, and 850 nm (fourth column). The vascular blood oxygenation of the numerical phanton (ground truth) is shown on the top right. Paraview22 was used for volume rendering.

JBO_27_3_036001_f011.png

With respect to the blood vasculature detection, the majority of the voxels corresponding to the blood vessels were not detected without normalization of the optical fluence [Fig. 11(a)]. Although many of the blood vessel voxels near the breast surface were detected via CLAHE, the voxels in the region deeper than 15 mm [orange to red color in Fig. 9(a)] were not detected [Fig. 11(b)]. The proposed method significantly improved the blood vasculature detectability [Fig. 11(c)]. With respect to estimation of the vascular blood oxygenation, the proposed method [Fig. 11(c)] enhanced the estimation accuracy in regions deeper than 18 mm [orange to red color in Fig. 9(a)] for all choices of wavelengths [Fig. 11(c)]. The results of vascular oxygenation estimation from CLAHE [Fig. 11(b)] were relatively inaccurate regardless of the wavelength pairs and voxel location, compared to the proposed method [Fig. 11(c)].

Table 2 provides detectability (DET), detectability-classification accuracy, and classification accuracy with respect to arteries (TAR and DTAR), veins (TVR and DTVR), and both (ACC and DACC).

Table 2

Artery/vein detectability and classification accuracy (%).

λs (nm)NormalizationDETTARTVRACCDTARDTVRDACC
757, 800None12.0089.8175.9781.583.6217.6911.99
CLAHE33.5372.1579.1676.3216.7138.8329.86
Proposed method78.8977.8897.2189.3866.8576.6972.71
757, 850None11.3974.7888.7183.062.7817.0811.32
CLAHE31.0462.5490.8279.3616.2137.1228.65
Proposed method76.1971.6296.5286.4358.9874.3868.14
800, 850None11.8730.4185.6163.242.8815.7110.51
CLAHE34.7525.8186.0761.6516.3930.3724.70
Proposed method79.5148.6677.4765.7937.3565.3353.99
757, 800, 850None11.5878.3588.0684.123.0717.2811.52
CLAHE32.1364.9190.0379.8516.9638.0529.51
Proposed method77.1173.1296.8087.2161.0175.2469.47
Note: For each metric and each choice of wavelengths, the entry corresponding to the best performing method is in bold.

As presented in Table 2, the DET of the proposed method was, on average, 6.66 and 2.37 times greater than no optical fluence normalization and CLAHE, respectively. The proposed method showed slightly better ACC compared with the others in Table 2. The proposed method increased the DACC by 5.81 and 2.34 times on average compared to no optical fluence normalization and CLAHE, respectively. The distribution of artery/vein voxels was not uniform with respect to depth. There were more voxels that correspond to the blood vessels (veins in particular) near the surface. Thus, further analysis of the classification accuracy according to depth will be presented henceforward [Fig. 12(b)].

Fig. 12

(a) Artery/vein detectability and (b) classification accuracy of no optical fluence normalization (black color), CLAHE (cyan color), and the proposed method (red color), according to 10 mm-depth ranges. The used wavelength pairs are 757 and 800 nm (dashed lines); 757 and 850 nm (dash-dotted lines); 800 and 850 nm (dotted lines); and 757, 800, and 850 nm (solid lines).

JBO_27_3_036001_f012.png

Figure 12 shows (a) vasculature detectability and (b) artery/vein classification accuracy of no optical normalization, CLAHE, and the proposed method as a function of depth. In this analysis, depth was quantized using 5 bins with a width of 10 mm. At all depth ranges, the proposed method [red color in Fig. 12(a)] outperformed the other two [black and cyan colors in Fig. 12(a)] in blood vessel detectability. The artery/vein voxels located deeper than 20 and 30 mm were not detected when no optical fluence normalization [black color in Fig. 12(a)] and CLAHE [cyan color in Fig. 12(a)] were applied, respectively. As shown in Fig. 12(b), in the depth ranges of 10 to 20 mm and 20 to 30 mm, the ACC of the proposed method (red color) was higher than the others (black and cyan colors), and the ACC largely dropped in the results of all three methods at a depth deeper than 30 mm. It is speculated that this is because the strength of the attenuated optoacoustic signals at depths deeper than 30 mm is similar to or lower than that of the noise. The ACC of CLAHE [cyan color in Fig. 12(b)] was either lower or slightly higher, up to 2.53%, than that of no optical fluence normalization [black color in Fig. 12(b)].

5.1.3.

Results from Experimental Studies

Figure 13 shows reconstructed 3D OAT images with no optical fluence normalization [Figs. 13(a) and 13(b)], CLAHE [Figs. 13(c) and 13(d)], and the proposed method [Figs. 13(e) and 13(f)]. A depth range of 0 to 30 mm was visualized. In Fig. 13, the region deeper than 15 mm (orange to red color) is nearly invisible to the human eye in the CLAHE results [Figs. 13(c) and 13(d)] while it is clearly visible in the results of the proposed method [Figs. 13(e) and 13(f)]. Additional visualization of the comparison in Figs. 13(a), 13(c), and 13(e) is provided in Video 1, showing a z-slice (x-y plane) of the breast image at each descretized z location with an increment of 0.25 mm (from 46 to 2.25  mm). In Video 1, the visibility of the blood vessels seated deeper than 15 mm (orange to red color) is consistent with the results in Fig. 13. The effective optical attenuation coefficient estimated from the left breast [Figs. 13(a), 13(c), and 13(e)] was 1  cm1 and that from the right breast [Figs. 13(b), 13(d), and 13(f)] was 0.98  cm1.

Fig. 13

Comparison between reconstructed images with [(a), (b)] no optical fluence normalization, [(c), (d)] CLAHE, and [(e), (f)] the proposed method. The used wavelength was 755 nm. Images in the left column [(a), (c), and (e)] are from the left breast and those in the right column [(b), (d), and (f)] are from the right breast. The images (a) to (f) are presented in MVBP of the entire breast volume along y axis. The still images in panel (g) of Video 1 (Video 1, MP4, 944 kB [URL: https://doi.org/10.1117/1.JBO.27.3.036001.1]) illustrate a z slice (x-y plane) at z=10  mm. The images were color-encoded by depth using the Jet color map in MATLAB.

JBO_27_3_036001_f013.png

6.

Discussion

In spite of the method’s simplicity, the numerical results demonstrated that the proposed method significantly improved vasculature detectability by compensating for optical attenuation and increased estimation accuracy of the vascular blood oxygenation by mitigating the spectral coloring effect (Figs. 11 and 12, and Table 2). Voxel brightness in the reconstructed estimate of p0(r,λi) decreased with depth due to optical attenuation. This resulted in severely underestimating total hemoglobin concentration at depths deeper than 10 mm when applying spectral linear unmixing directly to p0(r,λi), rather than μa(r,λi). On the other hand, in the estimation of oxygen saturation (CHbO2(r)/CtHb(r)), the effect of the optical attenuation could not be canceled out because of its dependence on wavelengths, i.e., the spectral coloring effect. Thus, the classification accuracy constantly decreased with depth without optical fluence normalization in Fig. 12(b). The proposed method ameliorated such reduction [Fig. 12(b)].

Furthermore, the value of the effective optical attenuation coefficient (Fig. 8), which was estimated from the in vivo 3D OAT breast images using the proposed method, correlates well with experimental measurements (1  cm1) that were reported in previous studies.44,6870 The proposed method is completely measurement-data-driven, therefore, a prior knowledge of the optical properties of the breast tissues, anatomy of the vascular network, and precise characterization of the illumination pattern and incident fluence is not required.

The proposed method was specifically implemented for the 3D OAT breast imaging system presented in Fig. 2. However, the general framework for the normalization of the optical fluence distribution is not limited to breast imaging and to this specific system. For example, the curve fitting for incident optical fluence estimation can be opportunely modified to account for different optical illumination patterns.

Although these studies demonstrated qualitative and quantitative enhancement achieved via use of the proposed method, there remain limitations. First and foremost, the proposed method assumes a constant effective optical attenuation coefficient when estimating the fluence map within the breast. Errors in the estimation of the fluence map due to neglecting spatial variations of effective optical attenuation coefficient may introduce bias in the optical energy absorption estimates.

Besides, to obtain 3D quantitative images of the vascular blood oxygenation from in vivo data, further investigations should address acoustic heterogeneity of breast tissue and noise removal (thermal acoustic noise from the medium and thermal noise from ultrasound transducer and electromagnetic interference). Because the proposed method compensates for the depth-dependent optical attenuation by amplifying the image brightness at each voxel in the reconstructed 3D OAT images as a function of depth, existing noise and artifacts are also amplified depending on the depth. Application of the proposed method to images reconstructed using advanced regularization techniques can reduce such noise and artifacts, thus extending imaging depth.

Future directions include investigation of the proposed method to imaging of breasts with benign and malignant lesions and other 3D OAT imaging applications, such as transcranial imaging and small animal imaging (whole or partial body). It is expected that the performance of the proposed method largely depends on the distributions of the optical properties within the target. For example, in whole mouse imaging, hemoglobin-concentrated organs, such as liver, kidneys, and colon, locally occupy a certain extent as bulk. This causes a locally varying imbalance in the optical fluence distribution. In such case, the assumptions of the proposed method are invalid, thus, further investigation is required, including the use of more sophisticated numerical models to estimate the fluence distribution, such as MC photon transport simulation or simplified spherical harmonics approximation of radiative transfer equations.12

7.

Conclusion

In this work, a straightforward physics-based method to normalize optical fluence distributions in 3D OAT breast images was proposed. The method is based on generally accepted assumptions on breast anatomy and optical properties as well as common features of light delivery in existing 3D OAT breast imagers. In the proposed method, both distributions of incident optical fluence and optical attenuation within the breast tissues are estimated solely from the voxel brightness in the reconstructed images, thus, a prior knowledge of the breast and specific geometry of the light-delivery system is not required.

Numerical studies demonstrated that the proposed method—in conjunction with spectral linear unmixing—significantly enhanced blood vasculature detectability and improved estimation accuracy of vascular blood oxygenation down to a depth of 30 mm, when compared with no optical fluence normalization and a general-purpose image contrast enhancement technique called CLAHE. In addition, the proposed method outperformed CLAHE, in terms of PSNR and SSIM. It was also demonstrated that the proposed method can be applied to in vivo data. In particular, the effective optical attenuation coefficients estimated from the reconstructed 3D OAT breast images via the proposed method were found to be consistent with those experimentally measured in in vivo studies. With further investigations on acoustic heterogeneity, noise removal, and vascular segmentation, the use of the proposed method can potentially achieve 3D in vivo functional OAT images of the whole breast.

Disclosures

The authors A. A. Oraevsky and R. Su disclose ownership interest in TomoWave Laboratories, Inc. Other authors have no relevant financial interests and no potential conflicts of interest to disclose.

Acknowledgments

This work was supported in part by NIH Awards NS102213, EB028652, and R01NS099429.

Code, Data, and Materials Availability

Data and code are available upon request. Please contact Dr. Park at sp33@illinois.edu.

References

1. 

A. A. Oraevsky et al., “Clinical optoacoustic imaging combined with ultrasound for coregistered functional and anatomical mapping of breast tumors,” Photoacoustics, 12 30 –45 (2018). https://doi.org/10.1016/j.pacs.2018.08.003 Google Scholar

2. 

Y. Lou et al., “Generation of anatomically realistic numerical phantoms for photoacoustic and ultrasonic breast imaging,” J. Biomed. Opt., 22 (4), 041015 (2017). https://doi.org/10.1117/1.JBO.22.4.041015 JBOPFO 1083-3668 Google Scholar

3. 

L. Lin et al., “Single-breath-hold photoacoustic computed tomography of the breast,” Nat. Commun., 9 2352 (2018). https://doi.org/10.1038/s41467-018-04576-z Google Scholar

4. 

A. A. Oraevsky et al., “Laser optoacoustic imaging of breast cancer in vivo,” Proc. SPIE, 4256 6 –15 (2001). https://doi.org/10.1117/12.429300 PSISDG 0277-786X Google Scholar

5. 

A. Oraevsky et al., “Full-view 3D imaging system for functional and anatomical screening of the breast,” Proc. SPIE, 10494 104942Y (2018). https://doi.org/10.1117/12.2318802 PSISDG 0277-786X Google Scholar

6. 

S. Park, “Compensation for non-uniform illumination and optical fluence attenuation in three-dimensional optoacoustic tomography of the breast,” Proc. SPIE, 10878 108784X (2019). https://doi.org/10.1117/12.2514750 PSISDG 0277-786X Google Scholar

7. 

M. Toi et al., “Visualization of tumor-related blood vessels in human breast by photoacoustic imaging system with a hemispherical detector array,” Sci. Rep., 7 41970 (2017). https://doi.org/10.1038/srep41970 SRCEC3 2045-2322 Google Scholar

8. 

X. L. Deán-Ben et al., “Volumetric hand-held optoacoustic angiography as a tool for real-time screening of dense breast,” J. Biophotonics, 9 (3), 253 –259 (2016). https://doi.org/10.1002/jbio.201500008 Google Scholar

9. 

R. A. Kruger et al., “Photoacoustic angiography of the breast,” Med. Phys., 37 (11), 6096 –6100 (2010). https://doi.org/10.1118/1.3497677 Google Scholar

10. 

N. Nyayapathi et al., “Dual scan mammoscope (DSM)—a new portable photoacoustic breast imaging system with scanning in craniocaudal plane,” IEEE Trans. Biomed. Eng., 67 1321 –1327 (2020). https://doi.org/10.1109/TBME.2019.2936088 IEBEAX 0018-9294 Google Scholar

11. 

K. Wang, M. A. Anastasio, “Photoacoustic and thermoacoustic tomography: image formation principles,” Handbook of Mathematical Methods in Imaging, 781 –815 Springer, New York (2011). Google Scholar

12. 

L. V. Wang and H.-I. Wu, Biomedical Optics: Principles and Imaging, John Wiley & Sons Inc., Hoboken, New Jersey (2012). Google Scholar

13. 

M. Li, Y. Tang and J. Yao, “Photoacoustic tomography of blood oxygenation: a mini review,” Photoacoustics, 10 65 –73 (2018). https://doi.org/10.1016/j.pacs.2018.05.001 Google Scholar

14. 

B. T. Cox et al., “Quantitative spectroscopic photoacoustic imaging: a review,” J. Biomed. Opt., 17 (6), 061202 (2012). https://doi.org/10.1117/1.JBO.17.6.061202 JBOPFO 1083-3668 Google Scholar

15. 

B. T. Cox, J. G. Laufer and P. C. Beard, “The challenges for quantitative photoacoustic imaging,” Proc. SPIE, 7177 717713 (2009). https://doi.org/10.1117/12.806788 PSISDG 0277-786X Google Scholar

16. 

A. A. Oraevsky, “Optoacoustic tomography: from fundamentals to diagnostic imaging of breast cancer,” Biomedical Photonics Handbook, PM222 715 –757 2nd ed.CRC Press, Florida (2014). Google Scholar

17. 

S. S. Alshahrani et al., “All-reflective ring illumination system for photoacoustic tomography,” J. Biomed. Opt., 24 (4), 046004 (2019). https://doi.org/10.1117/1.JBO.24.4.046004 JBOPFO 1083-3668 Google Scholar

18. 

M. Sivaramakrishnan et al., “Limitations of quantitative photoacoustic measurements of blood oxygenation in small vessels,” Phys. Med. Biol., 52 1349 –1361 (2007). https://doi.org/10.1088/0031-9155/52/5/010 PHMBA7 0031-9155 Google Scholar

19. 

R. Hochuli et al., “Estimating blood oxygenation from photoacoustic images: can a simple linear spectroscopic inversion ever work?,” J. Biomed. Opt., 24 (12), 121914 (2019). https://doi.org/10.1117/1.JBO.24.12.121914 JBOPFO 1083-3668 Google Scholar

20. 

D. Stalling, M. Westerhoff, H.-C. Hege, “Amira: a highly interactive system for visual data analysis,” The Visualization Handbook, 749 –770 Elsevier Science, Netherlands (2005). Google Scholar

21. 

C. A. Schneider, W. S. Rasband and K. W. Eliceiri, “NIH image to ImageJ: 25 years of image analysis,” Nat. Methods, 9 (7), 671 –675 (2012). https://doi.org/10.1038/nmeth.2089 1548-7091 Google Scholar

22. 

J. Ahrens, B. Geveci, C. Law, “Paraview: an end-user tool for large data visualization,” Visualization Handbook, 717 –732 Elsevier Science, Netherlands (2005). Google Scholar

23. 

A. Fedorov et al., “3D slicer as an image computing platform for the quantitative imaging network,” Magn. Reson. Imaging, 30 (9), 1323 –1341 (2012). https://doi.org/10.1016/j.mri.2012.05.001 MRIMDQ 0730-725X Google Scholar

24. 

S. Cho et al., “3D PHOVIS: 3D photoacoustic visualization studio,” Photoacoustics, 18 100168 (2020). https://doi.org/10.1016/j.pacs.2020.100168 Google Scholar

25. 

A. Pattyn et al., “Model-based optical and acoustical compensation for photoacoustic tomography of heterogeneous mediums,” Photoacoustics, 23 100275 (2021). https://doi.org/10.1016/j.pacs.2021.100275 Google Scholar

26. 

J. Poudel, Y. Lou and M. A. Anastasio, “A survey of computational frameworks for solving the acoustic inverse problem in three-dimensional photoacoustic computed tomography,” Phys. Med. Biol., 64 14TR01 (2019). https://doi.org/10.1088/1361-6560/ab2017 PHMBA7 0031-9155 Google Scholar

27. 

X. L. Deán-Ben and D. Razansky, “Adding fifth dimension to optoacoustic imaging: volumetric time-resolved spectrally enriched tomography,” Light Sci. Appl., 3 (1), e137 –e137 (2014). https://doi.org/10.1038/lsa.2014.18 Google Scholar

28. 

S. M. Schoustra et al., “Twente photoacoustic mammoscope 2: system overview and three-dimensional vascular network images in healthy breasts,” J. Biomed. Opt., 24 (12), 121909 (2019). https://doi.org/10.1117/1.JBO.24.12.121909 JBOPFO 1083-3668 Google Scholar

29. 

“American national standard for safety use of lasers,” Orlando, Florida (2014). Google Scholar

30. 

H. Xu and B. W. Rice, “In-vivo fluorescence imaging with a multivariate curve resolution spectral unmixing technique,” J. Biomed. Opt., 14 (6), 064011 (2009). https://doi.org/10.1117/1.3258838 JBOPFO 1083-3668 Google Scholar

31. 

S. T. M. Gammon et al., “Spectral unmixing of multicolored bioluminescence emitted from heterogeneous biological sources,” Anal. Chem., 78 1520 –1527 (2006). https://doi.org/10.1021/ac051999h ANCHAM 0003-2700 Google Scholar

32. 

R. Lansford, G. Bearman and S. E. Fraser, “Resolution of multiple green fluorescent protein color variants and dyes using two-photon microscopy and imaging spectroscopy,” J. Biomed. Opt., 6 (3), 311 –318 (2001). https://doi.org/10.1117/1.1383780 JBOPFO 1083-3668 Google Scholar

33. 

E. Fux and C. Mazel, “Unmixing coral fluorescence emission spectra and predicting new spectra under different excitation conditions,” Appl. Opt., 38 486 –494 (1999). https://doi.org/10.1364/AO.38.000486 APOPAI 0003-6935 Google Scholar

34. 

D. Chorvat et al., “Spectral unmixing of flavin autofluorescence components in cardiac myocytes,” Biophys. J., 89 L55 –L57 (2005). https://doi.org/10.1529/biophysj.105.073866 BIOJAU 0006-3495 Google Scholar

35. 

H. Tsurui et al., “Seven-color fluorescence imaging of tissue samples based on Fourier spectroscopy and singular value decomposition,” J. Histochem. Cytochem., 48 653 –662 (2000). https://doi.org/10.1177/002215540004800509 JHCYAS 0022-1554 Google Scholar

36. 

F. Nadrigny et al., “Detecting fluorescent protein expression and co-localisation on single secretory vesicles with linear spectral unmixing,” Eur. Biophys. J. Biophys. Lett., 35 533 –547 (2006). https://doi.org/10.1007/s00249-005-0040-8 Google Scholar

37. 

J. Glatz et al., “Blind source unmixing in multi-spectral optoacoustic tomography,” Opt. Express, 19 3175 –3184 (2011). https://doi.org/10.1364/OE.19.003175 OPEXFF 1094-4087 Google Scholar

38. 

K. Wang et al., “Accelerating image reconstruction in three-dimensional optoacoustic tomography on graphics processing units,” Med. Phys., 40 (2), 023301 (2013). https://doi.org/10.1118/1.4774361 MPHYA6 0094-2405 Google Scholar

39. 

E. V. Petrova, A. A. Oraevsky and S. A. Ermilov, “Red blood cell as a universal optoacoustic sensor for non-invasive temperature monitoring,” Appl. Phys. Lett., 105 (9), 094103 (2014). https://doi.org/10.1063/1.4894635 APPLAB 0003-6951 Google Scholar

40. 

D. F. Swinehart, “The Beer-Lambert law,” J. Chem. Educ., 39 (7), 333 –335 (1962). https://doi.org/10.1021/ed039p333 JCEDA8 0021-9584 Google Scholar

41. 

L. Ding et al., “Constrained inversion and spectral unmixing in multispectral optoacoustic tomography,” IEEE Trans. Med. Imaging, 36 (8), 1676 –1685 (2017). https://doi.org/10.1109/TMI.2017.2686006 ITMID4 0278-0062 Google Scholar

42. 

A. N. Bashkatov et al., “Measurement of tissue optical properties in the context of tissue optical clearing,” J. Biomed. Opt., 23 (9), 091416 (2018). https://doi.org/10.1117/1.JBO.23.9.091416 JBOPFO 1083-3668 Google Scholar

43. 

M. Xu and L. V. Wang, “Universal back-projection algorithm for photoacoustic computed tomography,” Phys. Rev. E, 71 016706 (2005). https://doi.org/10.1103/PhysRevE.71.016706 Google Scholar

44. 

S. L. Jacques, “Optical properties of biological tissues: a review,” Phys. Med. Biol., 58 (11), R37 –R61 (2013). https://doi.org/10.1088/0031-9155/58/11/R37 PHMBA7 0031-9155 Google Scholar

45. 

D. J. Hauben et al., “Breast–areola–nipple proportion,” Ann. Plastic Surg., 50 (5), 510 –513 (2003). https://doi.org/10.1097/01.SAP.0000044145.34573.F0 APCSD4 Google Scholar

46. 

Jr. T. L. Pope et al., “Breast skin thickness: normal range and causes of thickening shown on film-screen mammography,” J. Can. Assoc. Radiol., 35 (4), 365 –368 (1984). JCARAU 0008-2902 Google Scholar

47. 

N. Otsu, “A threshold selection method from gray-level histograms,” IEEE Trans. Syst. Man Cybern., 9 62 –66 (1979). https://doi.org/10.1109/TSMC.1979.4310076 Google Scholar

48. 

P. Heckbert, Graphics Gems IV (IBM Version), 1st edMorgan Kaufmann, United States (1994). Google Scholar

49. 

S. Park et al., “Realistic three-dimensional optoacoustic tomography imaging trials using the VICTRE breast phantom of FDA (Conference Presentation),” Proc. SPIE, 11240 112401H (2020). https://doi.org/10.1117/12.2552380 PSISDG 0277-786X Google Scholar

50. 

A. Badano et al., “Evaluation of digital breast tomosynthesis as replacement of full-field digital mammography using an in silico imaging trial,” JAMA Network Open, 1 e185474 –e185474 (2018). https://doi.org/10.1001/jamanetworkopen.2018.5474 Google Scholar

51. 

J. Harris et al., Diseases of the Breast, 5th edWolters Kluwer Health, United States (2014). Google Scholar

52. 

P. Hasgall et al., “IT’IS database for thermal and electromagnetic parameters of biological tissues (Version 4.0),” (2018). Google Scholar

53. 

Q. Fang and D. A. Boas, “Monte Carlo simulation of photon migration in 3D turbid media accelerated by graphics processing units,” Opt. Express, 17 20178 –20190 (2009). https://doi.org/10.1364/OE.17.020178 OPEXFF 1094-4087 Google Scholar

54. 

L. Yu et al., “Scalable and massively parallel Monte Carlo photon transport simulations for heterogeneous computing platforms,” J. Biomed. Opt., 23 (1), 010504 (2018). https://doi.org/10.1117/1.JBO.23.1.010504 JBOPFO 1083-3668 Google Scholar

55. 

C. Bench, A. Hauptmann and B. Cox, “Toward accurate quantitative photoacoustic imaging: learning vascular blood oxygen saturation in three dimensions,” J. Biomed. Opt., 25 (8), 085003 (2020). https://doi.org/10.1117/1.JBO.25.8.085003 JBOPFO 1083-3668 Google Scholar

56. 

B. E. Treeby and B. T. Cox, “k-wave: MATLAB toolbox for the simulation and reconstruction of photoacoustic wave fields,” J. Biomed. Opt., 15 (2), 021314 (2010). https://doi.org/10.1117/1.3360308 JBOPFO 1083-3668 Google Scholar

57. 

J. Nickolls et al., “Scalable parallel programming with cuda,” Queue - GPU Comput., 6 (2), 40 –53 (2008). https://doi.org/10.1145/1365490.1365500 Google Scholar

58. 

M. Aharon, M. Elad and A. Bruckstein, “K-SVD: an algorithm for designing of overcomplete dictionaries for sparse representation,” IEEE Trans. Signal Process., 54 (11), 4311 –4322 (2006). https://doi.org/10.1109/TSP.2006.881199 ITPRED 1053-587X Google Scholar

59. 

K. Zuiderveld, “Contrast limited adaptive histogram equalization,” 474 –485 (1994). Google Scholar

60. 

P. Singh, R. Mukundan and R. D. Ryke, “Feature enhancement in medical ultrasound videos using contrast-limited adaptive histogram equalization,” J. Digit. Imaging, 33 273 –285 (2020). https://doi.org/10.1007/s10278-019-00211-5 JDIMEW Google Scholar

61. 

M. Sundaram et al., “Histogram modified local contrast enhancement for mammogram images,” Appl. Soft Comput., 11 (8), 5809 –5816 (2011). https://doi.org/10.1016/j.asoc.2011.05.003 Google Scholar

62. 

S. Yousefi et al., “Uniform enhancement of optical micro-angiography images using Rayleigh contrast-limited adaptive histogram equalization,” Quant. Imaging Med. Surg., 3 (1), 5 –17 (2013). https://doi.org/10.3978/j.issn.2223-4292.2013.01.01 Google Scholar

63. 

E. Merčep et al., “Transmission-reflection optoacoustic ultrasound (TROPUS) computed tomography of small animals,” Light Sci. Appl., 8 18 (2019). https://doi.org/10.1038/s41377-019-0130-5 Google Scholar

64. 

A. F. Frangi et al., “Multiscale vessel enhancement filtering,” Lect. Notes Comput. Sci., 1496 130 –137 (1998). https://doi.org/10.1007/BFb0056195 LNCSD9 0302-9743 Google Scholar

65. 

G. Xi et al., “Label-free imaging of blood vessels in human normal breast and breast tumor tissue using multiphoton microscopy,” Scanning, 2019 5192875 (2019). https://doi.org/10.1155/2019/5192875 SCNNDF 0161-0457 Google Scholar

66. 

A. Horé and D. Ziou, “Image quality metrics: PSNR vs. SSIM,” in 20th Int. Conf. Pattern Recognit., 2366 –2369 (2010). https://doi.org/10.1109/ICPR.2010.579 Google Scholar

67. 

Z. Wang et al., “Image quality assessment: from error visibility to structural similarity,” IEEE Trans. Image Process., 13 (4), 600 –612 (2004). https://doi.org/10.1109/TIP.2003.819861 IIPRE4 1057-7149 Google Scholar

68. 

J. Sandell and T. Zhu, “A review of in-vivo optical properties of human tissues and its impact on PDT,” J. Biophotonics, 4 (11-12), 773 –787 (2011). https://doi.org/10.1002/jbio.201100062 Google Scholar

69. 

H. Heusmann, J. G. Koelzer and G. Mitic, “Characterization of female breasts in vivo by time-resolved and spectroscopic measurements in the near infrared spectroscopy,” J. Biomed. Opt., 1 (4), 425 –434 (1996). https://doi.org/10.1117/12.250669 JBOPFO 1083-3668 Google Scholar

70. 

K. Suzuki M.D. et al., “Quantitative measurement of optical parameters in normal breasts using time-resolved spectroscopy: in vivo results of 30 Japanese women,” J. Biomed. Opt., 1 (3), 330 –334 (1996). https://doi.org/10.1117/12.239902 JBOPFO 1083-3668 Google Scholar

Biography

Seonyeong Park is a postdoctoral research associate at the Department of Bioengineering, University of Illinois Urbana–Champaign, Urbana, Illinois. She received her BS degree in electronics, computer, and telecommunication engineering and her MS degree in information and communications engineering from Pukyong National University, Busan, Korea, in 2011 and 2013, respectively, and her PhD in electrical and computer engineering from Virginia Commonwealth University, Richmond, Virginia, USA, in 2017. Her research interests include virtual imaging trials, photoacoustic computed tomography, functional imaging, image reconstruction, and inverse problem.

Frank J. Brooks received his PhD in physics from Washington University (WU) in Saint Louis. His dissertation specialization was in biophysics, rate equations, and statistical physics. He applied these skills to various problems in statistical image analysis, image acquisition, and radio-tracer kinetics at the Department of Radiation Oncology at the Washington University School of Medicine (WUSM) and again at the Department of Radiology at the Mallinckrodt Institute of Radiology, also at WUSM. In 2017, he accepted a position at the WU Department of Biomedical Engineering where he began research at the intersection of machine learning and imaging science. In 2019, he began his current research professorship at the Grainger College of Engineering Department of Bioengineering at the University of Illinois Urbana–Champaign where he is a senior member of the Computational Imaging Science Laboratory. His current research interests are in the creation of stochastic object models and the evaluation of generative adversarial networks.

Umberto Villa is a research assistant professor of electrical and systems engineering, WU in St. Louis, St. Louis, Missouri, USA. He received his BS and MS degrees in mathematical engineering from Politecnico di Milano, Milan, Italy, in 2005 and 2007, respectively, and his PhD in mathematics from Emory University, Atlanta, Georgia, USA, in 2012. His research interests lie in the computational and mathematical aspects of large-scale inverse problems, imaging science, and uncertainty quantification. A strong component of his work includes developing scalable efficient algorithms for integrating data (images, experimental measurements, or observations) and mathematical models with applications to biomedical and engineering-relevant problems.

Richard Su is a senior scientist at TomoWave Laboratories. He received his PhD in biomedical engineering from the University of Houston, Texas, USA, in 2021. He has been involved in research projects in the field of optoacoustic tomography for the past ten years and published 22 papers. His research is focused on visualization and detection of breast and prostate tumors, molecular and functional imaging in small animal models, as well as monitoring temperature changes in the course of thermal therapy. He has also significant experience in development and implementation of signal and image processing for laser ultrasound and optoacoustic applications, which is essential for obtaining high image quality and quantitative accuracy of functional information.

Mark A. Anastasio is a Donald Biggar Willett Professor in engineering and the head of the Department of Bioengineering, University of Illinois at Urbana–Champaign, Urbana, Illinois, USA. He received his PhD from the University of Chicago, Chicago, Illinois, USA, in 2001. His research broadly addresses computational image science, inverse problems in imaging, and machine learning for imaging applications. He has contributed broadly to emerging biomedical imaging technologies that include photoacoustic computed tomography, ultrasound computed tomography, and x-ray phase-contrast imaging. His work has been supported by numerous NIH grants, and he served for 2 years as the chair of the NIH EITA Study Section. He is a fellow of the Society of Photo-Optical Instrumentation Engineers (SPIE), the American Institute for Medical and Biological Engineering (AIMBE), and the International Academy of Medical and Biological Engineering (IAMBE).

Alexander A. Oraevsky has extensive experience in leading research and development laboratories in academia and small businesses. He is a pioneer in the field of biomedical optoacoustics. He is a fellow of SPIE, founder and chair of the largest SPIE/BIOS annual conference: “Photons plus Ultrasound: Imaging and Sensing.” Currently, he is CEO and chief technology officer at TomoWave Laboratories and holds an adjunct professor position at the Biomedical Engineering Department of the University of Houston. He is the primary inventor of 23 patents, has published 11 book chapters, and over 200 scientific papers dealing with novel laser technologies applicable in biology and medicine. As a principal investigator, he successfully accomplished aims of 30 projects funded by NIH and other research foundations. He is the recipient of multiple research awards advancing biomedical applications of the optoacoustic imaging sensing and monitoring, including Berthold Leibinger Innovations Prize.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 Unported License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Seonyeong Park, Frank J. Brooks, Umberto Villa, Richard Su, Mark A. Anastasio, and Alexander A. Oraevsky "Normalization of optical fluence distribution for three-dimensional functional optoacoustic tomography of the breast," Journal of Biomedical Optics 27(3), 036001 (16 March 2022). https://doi.org/10.1117/1.JBO.27.3.036001
Received: 17 November 2021; Accepted: 22 February 2022; Published: 16 March 2022
Lens.org Logo
CITATIONS
Cited by 10 scholarly publications.
Advertisement
Advertisement
KEYWORDS
Breast

3D image processing

Tissue optics

Signal attenuation

Optoacoustics

Blood

3D image reconstruction

Back to Top