Open Access
7 May 2012 Investigation of a diffuse optical measurements-assisted quantitative photoacoustic tomographic method in reflection geometry
Author Affiliations +
Abstract
Photoacoustic tomography provides the distribution of absorbed optical energy density, which is the product of optical absorption coefficient and optical fluence distribution. We report the experimental investigation of a novel fitting procedure that quantitatively determines the optical absorption coefficient of chromophores. The experimental setup consisted of a hybrid system of a 64-channel photoacoustic imaging system with a frequency-domain diffused optical measurement system. The fitting procedure included a complete photoacoustic forward model and an analytical solution of a target chromophore using the diffusion approximation. The fitting procedure combines the information from the photoacoustic image and the background information from the diffuse optical measurements to minimize the photoacoustic measurements and forward model data and recover the target absorption coefficient quantitatively. 1-cm-cube phantom absorbers of high and low contrasts were imaged at depths of up to 3.0 cm. The fitted absorption coefficient results were at least 80% of their true values. The sensitivities of this fitting procedure to target location, target radius, and background optical properties were also investigated. We found that this fitting procedure was most sensitive to the accurate determination of the target radius and depth. Blood sample in a thin tube of radius 0.58 mm, simulating a blood vessel, was also studied. The photoacoustic images and fitted absorption coefficients are presented. These results demonstrate the clinical potential of this fitting procedure to quantitatively characterize small lesions in breast imaging.

1.

Introduction

Photoacoustic tomography (PAT) takes advantage of the high optical absorption of blood-rich tumors and relatively low optical absorption of normal tissues in the near-infrared (NIR) optical spectrum to offer high-resolution and high-contrast images for up to a few centimeters in depth.110 PAT, therefore, has great potential to detect early-stage breast cancers.2,6,10 Oraevsky’s group was one of the first to build a PAT system solely for breast imaging.6,11,12 However, conventional PAT can only provide the distribution of absorbed optical energy density, which is the product of optical absorption coefficient and optical fluence distribution. The absorbed energy density is clearly not the intrinsic functional parameter for breast cancer diagnosis, as it depends on the distribution of the external light illumination, which is a spatially varying function. In the past, our group developed a dual-modality system that uses the location information provided by PAT to guide diffuse optical tomography (DOT) for quantifying the absorption coefficient distribution of targets.13,14 This approach offers low-resolution DOT images and cannot take full advantage of the high resolution provided by the PAT images. Several groups have investigated the reconstruction of the absolute absorption coefficient distribution from PAT images.1522 Jiang et al.1517 have used the finite element method to first recover the absorbed optical energy density, followed by the use of the photon diffusion equation to extract the absorption coefficient. Laufer et al.18,19 developed the spectrum technique to iteratively fit the chromophore concentration from measurements of photoacoustic signals acquired at different wavelengths. Other groups have introduced different methods to reconstruct optical absorption quantitatively, such as using sparse signal representation,20 multiple optical sources,21 or acoustic spectra.22 In all of these methods, however, the importance of the background optical property is not emphasized, although it significantly affects the reconstructed quantitative PAT absorption coefficients. In this paper, we introduce a fitting procedure that uses diffuse optical measurements (DOM) to assist PAT in quantifying tissue-absorption coefficient. The PAT setup used is the reflection geometry because it offers more comfort and convenience to the patient during breast imaging. In Sec. 2 of this paper, the forward model, the fitting procedure, and the experimental system are introduced. In Sec. 3, the simulation and experimental results are presented. It will be shown that the reconstruction accuracy obtained is about 84% for a high contrast target and 88% for a low contrast target embedded in a highly scattering medium. Section 3 also discusses the sensitivities of this fitting procedure to the target parameters (radius and location) and background optical properties.

2.

Methods

2.1.

Forward Model

The complete forward model combines the model of light propagation through the scattering medium and the model of photoacoustic wave propagation. The photon diffusion equation is used to describe the fluence distribution inside the medium. The simple analytical expressions for fluence inside the target and background medium23 are implemented in the model. The fluence inside the target, Φin, and background medium, Φout, are only related to background optical properties, target optical properties, and target position as:

Eq. (1a)

Φout=vS04π|rrs|exp(ikout|rrs|)+l,m[Al,mjl(koutr)+Bl,mnl(koutr)]Yl,m(θ,ϕ),

Eq. (1b)

Φin=l,m[Cl,mjl(kinr)+Dl,mnl(kinr)]Yl,m(θ,ϕ).

In Eqs. (1a) and (1b), Al,m, Bl,m, Cl,m, and Dl,m are coefficients. jl(x) and nl(x) are the spherical Bessel and Neumann functions, respectively, and Yl,m(θ,ϕ) is the spherical harmonics. kout and kin are the complex wave numbers of the background medium and target, respectively, and are related to their optical properties. As a result, the absorbed optical energy distribution (Q) can be determined and related to the background optical properties, target optical properties, and target position as well. The initial photoacoustic pressure wave induced by a short-pulsed laser is proportional to the absorbed optical energy density distribution and Gruneisen coefficient (Γ). The initial pressure, p0(r,μa), and the subsequent acoustic wave, p(r,t), can be expressed by the following equations:24

Eq. (2a)

p0(r,μa)=ΓQ(r,μa),

Eq. (2b)

p(r,t)=14πct|r|=ctp0(rr)ctds.
From Eqs. (1) and (2), it can be seen that the simulated pressure signal is determined by the background optical properties, target optical properties, and the target position.

2.2.

Fitting Procedure

The complete fitting procedure, described by a three-step process, is illustrated in Fig. 1. The first step is to obtain the background optical properties using the diffuse optical measurement system.25 Since a probe with a flat surface was used for housing the ultrasound transducer, optical sources, and detectors, a semi-infinite geometry was a reasonable approximation for the optical measurements. The following set of linear equations were used to relate the calibrated amplitude Iαβ and phase ϕαβ measured at the source-detector pair αβ, to the source-detector separation ραβ as,26

Eq. (3a)

log(ραβ2Iαβ)=kiραβ,

Eq. (3b)

ϕαβ=krραβ,
where kr and ki are the real and imaginary parts of the complex wave number, respectively. We leave the wave number as a variable and use the fitted kr and ki values to calculate the background scattering and absorption coefficients. The importance of background optical properties will be demonstrated in the next section. In the second step, the map of the absorbed optical energy density is reconstructed using the delay-and-sum algorithm.27 The accurate location and radius of the target are read and used in the iteration in the next step. The third and last step is the inversion scheme to recover the absorption coefficient distribution from the photoacoustic measurements. Since the background optical properties and target position are determined in step 1 and step 2 above, the only unknown variables for the simulated pressure signal are the target optical properties. The object function (E) is defined as the difference between the experimental signal and the simulated signal, and is given as:

Eq. (4)

E=[PData(μa,t,μs,t)Psim(μa,t,μs,t)]2.

Fig. 1

Fitting procedure to quantitatively recover the target optical properties.

JBO_17_6_061213_f001.png

The optical properties of the target can be fitted using an iterative method to minimize the object function. The nonlinear regression function fminsearch, used in the minimization procedure, is available in MATLAB. If the optical properties of the background are unknown, then a four-parameter fitting procedure (μa and μs of the background medium, and μa and μs of the target) is needed. In this case, the two-parameter fitting procedure can be modified to a four-parameter one as follows: (a) the first step in the three-step-procedure is skipped; (b) the target location information is determined in the second step as before; and (c) the object function in the third step is modified to incorporate μa and μs of the background medium. This modified procedure hence allows the parameters, μa and μs of the background medium, and μa and μs of the target, to be fitted.

2.3.

Experimental System and Phantom

Figure 2(a) shows the experimental configuration of the diffuse optical measurements-assisted (DOM) PAT system. A Ti:Sapphire (Symphotics TII, LS-2134) laser optically pumped by a Q-switched Nd:YAG laser (Symphotics-TII, LS-2122), delivered 10-ns pulses at 15 Hz and 750 nm. The laser output was coupled into two 1000-micron step-index input fibers (Thorlabs, BFL48-1000) using a convex lens and circular beamsplitter arrangement as shown in the figure. The convex lens, which has 99% transmittance at 750-nm wavelength, has a focal length of 20 cm and focuses the light into the pair of fibers that are placed at its focal point. The beam splitter was used to split the incoming beam into two beams, one for each fiber. It was measured to have a 60% transmitting and 36% reflecting split ratio for horizontally polarized light at the 750 nm that was used for the experiment. The overall coupling efficiency of the setup (including the losses in the lens and beam splitter) was about 85%, and the total output energy was about 18 mJ per pulse. The two output ends of the fibers were then mounted on opposite (longitudinal) sides of a low-frequency linear ultrasound transducer described below and used for photoacoustic imaging.

Fig. 2

(a) Experimental configuration of diffuse-optical-measurement-assisted PAT system; (b) probe geometry used for simulation and phantom experiments.

JBO_17_6_061213_f002.png

A low-frequency ultrasound transducer was employed to maximize the sensitivity for photoacoustic signal detection. The ultrasound transducer manufactured by Vermon (France) had 64 elements with 0.85-mm pitch. The center frequency of the transducer was 1.3 MHz with a 6-dB response from 500 to 1800 kHz. An integrated acoustic lens with 2.5-cm focal length increased the sensitivity at imaging depths for which the optical fluences were low. The photoacoustic signals from the 64-array elements were individually amplified by 60 dB using receiver electronics designed and constructed by our group and multiplexed into eight data acquisition channels with a sampling frequency of 40 MHz and 12-bit precision.13 Due to the multiplexing, eight laser firings were required to generate a single 64-channel capture. Because the laser had 15-Hz repetition rate, the acquisition rate was about two frames per second.

The diffuse optical measurement system was a frequency-domain system and consisted of light source and detection subsystems, as well as a hand-held probe.25 The source had four laser diodes of wavelengths 740, 780, 808, and 830 nm. Each laser diode was modulated at 140 MHz and sequentially coupled into six optical source fibers on the probe through optical switches. On the receiving end, the reflected light from the turbid medium associated with each of the illumination sources was detected simultaneously with 12 optical fiber bundles of 3 mm diameter. The detection fibers then coupled the light into 12 parallel photomultiplier tubes (PMTs) and electronics channels. An aluminum foil-covered plastic plate with dimensions of 9×10cm was used for the hand-held probe as shown in Fig. 2(b). The middle slot was used for the ultrasound transducer, and the light source and detector fibers were deployed around the probe. The reflective aluminum foil prevented light from being absorbed by the probe and generating image artifacts.28 Semi-infinite partial reflection boundary condition was used for the estimation of the background optical properties, and the aluminum cover provided an effective refraction coefficient of 0.6. In this study, the background optical properties obtained from the 780-nm laser diode only were used.

In the phantom experiments, 0.6% Intralipid solution was used to emulate the background breast tissue. The calibrated optical absorption coefficient (μa) and reduced scattering coefficient (μs) of the Intralipid solution were in the range of 0.01 to 0.02 and 4.0 to 8.0cm1, respectively, for the experiments. A 1cm3-cubical soft absorber of higher (μa=0.28cm1), and another of lower (μa=0.08cm1) optical contrast was made from polyvinyl chloride plastisol and used to mimic a malignant and benign lesion respectively. The calibration was performed on a larger piece phantom of dimensions of 10×10×5cm3 using the DOM system. In general, a malignant lesion has higher contrast with optical absorption coefficient ranging from 0.18 to 0.4cm1, and a benign lesion has the lower contrast with absorption coefficient from 0.007 to 0.15cm1.2931

3.

Results

3.1.

Simulation Results with/without Background Optical Properties

Photoacoustic signals with 2% zero-mean Gaussian noise of a target of 1 cm diameter was simulated. In order to mimic breast legions of different contrasts, the simulation was performed for different μa of the target ranging from 0.30 to 0.05cm1 in 0.05cm1 decrements and μs=7.0cm1. The target was submerged in a scattering medium, having optical properties of μa=0.02cm1 and μs=7.0cm1, and was located at the position (x=0, y=0, z=2.0) cm relative to the center of the probe. If one assumes that the target location and size are already known from information obtained from the photoacoustic image, then there remain four unknowns: μa and μs of the target, and μa and μs of the background medium. Figure 3(a) and 3(b) are the results from the four-parameter fitting. The solid lines in Fig. 3(a) represent the true μa value of the target (blue line) and background (black line), and the red stars and green squares are the fitted μa values for the target and background, respectively. The x-axis represents the expected target μa and background μa in cm1, and y-axis is the fitted μa in cm1. When the target has low contrast, such as 0.05cm1, the error is largest: 102% for the fitted target μa, and 399% for the fitted background μa. On the other hand, as the target contrast gets higher, the errors gradually decrease to a minimum and then increase again as the contrast increases. The errors in the fitted target μa were 102%, 34%, 3.1%, 16.5%, 28.4%, and 36% corresponding to the target μa of 0.05, 0.10, 0.15, 0.20, 0.25, and 0.30cm1, respectively. The fitted background μa decreased from 0.099 to 0.022cm1 when the target contrast increased from 0.05 to 0.30cm1 with the error ranging from 400% to 7%. Clearly, four-parameter fitting is not accurate enough to determine the target μa. This is because many unknowns can make the nonlinear fitting algorithm settle at the local minimum of the object function. Figure 3(b) shows the fitted results for the target μs and background μs with the error ranging from 12% to 27% for target μs, and from 20% to 70% for background μs. The fitted results for the case where the background μa and μs are already known from DOM are shown in Fig. 3(c) and 3(d). It is seen from Fig. 3(c) that the errors for the fitted target μa have decreased to 4.6%, 1.2%, 0.7%, 0.55%, 0.5%, and 0.3% corresponding to the true target μa of 0.05, 0.10, 0.15, 0.20, 0.25, and 0.30cm1, respectively. The errors for the fitted target μs also ranged from 27% to 36%. Overall, the fitting accuracy for both the high and low contrast targets can be greatly improved if the total number of unknown parameters is decreased. Besides, the run-time of the fitting algorithm is also greatly reduced. This improvement was realized by using background optical properties obtained from the diffused optical measurements to reduce the four-parameter fitting to a two-parameter one. Figure 3(b) and 3(d) shows that the fitted target μs is not very sensitive to the change in the background optical properties. As a result, the target μa is the only fitted parameter in the phantom experiments.

Fig. 3

Fitting results from the simulation data with 2% noise: (a) fitted background μa and target μa with four unknown parameters; (b) fitted background μs and target μs with four unknown parameters; (c) fitted target μa with two unknown parameters; (d) fitted target μs with two unknown parameters.

JBO_17_6_061213_f003.png

3.2.

Phantom Experiments

The optical properties of the background medium can be calculated using Eqs. (3a) and (3b). Using the frequency domain diffuse optical measurement system described in Sec. 2.3, a set of amplitudes (Iαβ) and phases (ϕαβ) is measured at different distances (ραβ), where α represents the source number and β represents the detector number. Because the parameters for each PMT and individual laser diode varied considerably from one to another, we had to calibrate the gain and phase shift for each channel. Considering the gain difference and phase shift, Eq. (3) can be modified as

Eq. (5a)

log(ραβ2Iαβ)=log[Is(α)]+log[Id(β)]kiραβ

Eq. (5b)

ϕαβ=ϕs(α)+ϕd(β)+krραβ,
where Is(α) and ϕs(α) are the relative gain and phase delay associated with source channel α, and Id(β) and ϕd(β) are similar quantities associated with detector channel β. kr and ki are the real and imaginary parts of the complex wave number. In Eq. (5), the total number of unknowns are 2(α+β)+2, which is far smaller than the total number of measurements α×β. Consequently, we can solve all the Is(α), ϕs(α), Id(β), and ϕd(β) terms.25,26 Figure 4(a) is an example of the calibrated amplitude [log(ραβ2Iαβ)] versus source-detector distance in logarithmic scale, and Fig. 4(b) is an example of the calibrated phase versus s-d distances for 0.6% Intralipid solution, both at 780 nm. As one can see, the calibrated amplitude and phase from various pairs change linearly with distance. Therefore, the slopes of the two figures can be fitted as kr and ki, the real and imaginary parts of the complex wave number respectively. According to the diffusion theory, the complex wave number (k=kr+iki) is related to the medium optical properties as23

Eq. (6)

k2=3(vμaμs+iωμs)v,
where v is the speed of light in the medium and ω is the modulation frequency. From this equation, μa and μs can be solved accordingly. The use of 0.4% to 0.8% Intralipid solutions always yielded scattering and absorption coefficients with rather good accuracy. The results in Fig. 4 were determined by using the source-detector distances between 2.5 and 5.5 cm, due to the sensitivity of the PMT. In this case, the background optical properties for 0.6% Intralipid solution were μa=0.019cm1 and μs=5.24cm1 at 780 nm.

Fig. 4

Measured amplitude and phase profiles of diffuse waves: (a) log(ραβ2Iαβ) versus source-detector distances; (b) phases versus source-detector distances.

JBO_17_6_061213_f004.png

In the phantom experiments, the following calibration process was introduced to scale the simulation data to the measurements from the U.S. transducer array: we started with a phantom object with known optical properties, which was a 1-cm3 cubical soft absorber made of polyvinyl chloride plastisol with the calibrated optical properties, μa=0.30cm1 and μs=6.2cm1. An empirical relation, shown in Eq. (7), was established from PAT simulation trials of the phantom. The equation considered the scan angle of each transducer element in the simulation, and weighted the simulated pressure according to the distance of each pixel from the transducer elements as

Eq. (7)

p(r,t)=Δrp0[r(θ)Δr(θ)]·er(θ)53.8,
where θ is the angle between the line joining a pixel to the transducer element and the vertical line through the element. Figure 5(a) compares the signal obtained from experiment with that from the simulation in one channel of the transducer. In this figure, the x-axis is the depth in units of mm, the y-axis is the amplitude of the signal in units of volts. The phantom center was located at 28 mm depth, and the front face at 23 mm depth. The red dash line in the figure is the PAT signal from experiment, and the blue solid line is that from simulation. It is seen that both the experimental and simulated signals closely match each other in terms of the target location and generated acoustic amplitude. At the shallower depth, however, the mismatch between the simulated and experimental data was high. This is because in the experimental data, photoacoustic signals generated from the transducer’s front face upon light illumination were present.28,32 On the other hand, this signal was absent in the simulated data. However, this does not pose any problem as the fitting procedure only compares the difference between the experimental and simulated signals at the region around the target. In the case of Fig. 5(a), the signals from depths of 18 to 40 mm were summed up for each channel, as shown in Fig. 5(b). The x-axis represents the 64 elements of the transducer; the y-axis is the summation of the amplitude from each channel. The red dash line represents the experimental signal, whereas the blue solid line represents the simulation. The shape and amplitude of these two curves compare favorably well with each other. The calibrations were performed at different depths and used in the experiments.

Fig. 5

Comparison of the experimental and simulated signals: (a) experimental signal and simulated signal obtained from one U.S. transducer element; (b) the total intensities obtained from experimental signals and simulation signals from 64 channels.

JBO_17_6_061213_f005.png

A 1cm3-cubical absorber of higher (μa=0.28cm1) and another of lower (μa=0.08cm1) optical contrast were used as soft tissue targets to mimic a high-contrast malignant lesion and low-contrast benign lesion, respectively. The absorber was placed inside Intralipid with calibrated optical properties of μa=0.019cm1 and μs=5.24cm1, and imaged at depths of 1.5, 2.0, 2.5, and 3.0 cm measured from the transducer face to the center of the absorber. Figure 6(a) shows the PAT images of the higher-contrast target with its center located at 2.0 cm from the transducer. The horizontal axis of the PAT images is the lateral dimension (in mm) in the x-direction, and the vertical axis is the depth (in mm) in the z-direction. The high-resolution PAT images clearly delineate the front and back face of the target, as shown in the circled regions in the figure. The location of the target can be read as (0.2, 2.0) cm in x and z directions. Because the photoacoustic images provided only a two-dimensional (2-D) image, the center of the target in the y dimension was set to y=0, which is the center of the transducer [as shown in Fig. 2(b)]. The fitted target μa converged to 0.255cm1 after 18 iterations, which was 91% of the true value. Table 1 lists the fitted results and errors for the higher-contrast absorber at different depths. The errors in the fitted data range from 7.1% to 16.8%; these are 18% to 37% better than the reconstructed DOT results with guidance from PAT image in Ref. 14. In the referenced paper, the maximum reconstructed absorption coefficients for the higher-contrast target have errors about 25% to 54%.

Fig. 6

(a) PAT images of the high contrast target with center located at 2.0 cm depth; (b) PAT images of the low contrast target with center located at 2.0 cm depth. The x-axis of the PAT images is lateral dimension and vertical z-axis is the depth. The units are mm. The circled region shows the cubic target.

JBO_17_6_061213_f006.png

Table 1

Phantom experiment for single high-contrast target (calibrated μa=0.28  cm−1) and low-contrast target (calibrated μa=0.08  cm−1).

Reconstructed μa (cm1)High-contrast targetLow-contrast target
Fitted valueErrorFitter valueError
Target center depth 1.5 cm0.3007.1%0.0833.8%
Target center depth 2.0 cm0.2558.9%0.0867.5%
Target center depth 2.5 cm0.23316.8%0.07111.2%
Target center depth 3.0 cm0.25510.7%0.0756.2%

Similar experiments were performed for the lower-contrast target. Figure 6(b) shows the PAT images of this target whose center location may be read as (0.5,2.0)cm. The contrast of the PAT images was not as good as that of the higher-contrast one. However, the front and back faces of the target were fairly resolved. After 14 iterations, the fitted target μa had a final converged value of 0.086cm1, which is 108% of the true value. The fitting results at other depths are listed in Table 1, and the errors at different depths range from 3.8% to 11.2%. These values are either comparable or 20% better than the DOT results for low contrast target in Ref. 14.

3.3.

Sensitivity Test

The sensitivity of the two-parameter fitting procedure to the target parameters and background optical properties was evaluated. We found that accurate target information, namely the location (x, y, z) and size, are critical for the fitting procedure. As shown in Fig. 7(a), where the x-axis is the measured radius of target and y-axis is the fitted μa in cm1, if the target’s radius is not estimated accurately, varying from 0.4 to 0.6 cm, the fitted μa is greatly influenced by this error. This is easy to understand because the total signal is related to the volume of the target. When the estimated target size is smaller than that of true size, the fitted absorption coefficient increases to match the measured signal; in this case, the errors increase from 15% to 74%. An example of this is shown in Fig. 7(a) for a target centered at 2.5 and 3.0 cm, and in general is true for all depths. The fitted μa is also strongly related to the target depth as shown in Fig. 7(b). In this figure, the x-axis is the assumed target depth, and we still used the data in which the target was centered at 2.5 and 3.0 cm depths. When the target depth input is not accurate, the fitted error also increases from 15% to 100%. On the other hand, the x-y location of target has small influence on the fitting accuracy as shown in Fig. 7(c). The x-axis represents the target location in the x direction, and y-axis represents the fitted target μa values. When the location of the target was moving in the x-direction away from the true location up to 0.2 cm, the fitted μa varied less than 7%.

Fig. 7

The influence of structural information: (a) radius; (b) depth; (c) x-y plane.

JBO_17_6_061213_f007.png

The background optical properties are also critical to the two-parameter fitting procedure. Figure 8(a) shows that the target fitted μa changes as the input background μa increases. In the figure, the x-axis represents the background μa in cm1, and y-axis is the fitted μa. The error ranges from 16% to 45%. The background μs also has similar influence on the fitting procedure as shown in Fig. 8(b). In this figure, x-axis represents the background μs, and y-axis is the fitted target μa. When the background μs is not accurate, the fitting error also increases from 16% to 60%. On the other hand, the target μs has a small influence on the fitting accuracy as shown in Fig. 8(c). The x-axis represents the target μs and ranges from 1 to 11 in 2cm1 increments, and y-axis represents the fitted target μa. When the target μs was changed, the fitted μa varied less than 20%.

Fig. 8

The influence of optical properties: (a) background μa change; (b) background μs change; (c) target μs change.

JBO_17_6_061213_f008.png

3.4.

Blood Tube Experiments

To test the robustness of the fitting procedure, we also performed experiments with blood collected from a sacrificed mouse. The blood was diluted and injected into a thin tube with 0.58-mm inner diameter and 0.9-mm outer diameter. To estimate the absorption coefficient of the blood, the formula from Ref. 33 was used.

Eq. (8)

μa(λ)=2.303e(λ)·C/(64500gHb/mole),
where e(λ) is molar extinction coefficient and C is the diluted hemoglobin concentration of the blood. From the hemoglobin concentration, the calculated μa is 0.39cm1. Figure 9 shows the PAT image of the blood tube that is submerged inside Intralipid with optical properties of μa=0.02cm1 and μs=4.9cm1 at a depth of 1.5 cm. The converged value for μa was 0.342cm1; this is 88% of estimated true value.

Fig. 9

PAT image of the blood tube.

JBO_17_6_061213_f009.png

4.

Summary and Discussion

Currently, ultrasound (US) is used extensively as an adjunct to x-ray mammography to differentiate cysts from solid lesions and is increasingly used for screening younger patients. However, ultrasound probes mechanical contrast of the lesion. On the other hand, PAT provides functional information of the lesion related to light absorption of hemoglobin content. Due to the different contrast mechanisms, some lesions may not be detectable by a nonoptical modality but yet have high optical contrast. A single contrast usually cannot provide accurate diagnosis because of biological variability. As a result, the combination of PAT and DOM or DOT may likely improve the detection and diagnosis of US occult lesions.

Diffuse optical measurements have shown to provide reliable background optical properties of breast tissue.34,35 For patients with breast lesions, the background optical properties can be measured from the contralateral breasts. Photoacoustic techniques have shown promising results to map breast tissue optical absorption contrast with ultrasound resolution, which is an order of magnitude higher than that of diffuse optical tomography. However, quantitative optical contrast is not readily available by photoacoustic tomography. This study demonstrates that diffuse optical measurements can assist photoacoustic tomography in quantifying target optical contrast by iteratively fitting the forward model with the experimental data. The forward model using the analytical expression for fluence in the medium and photoacoustic wave propagation is faster in terms of computation, when compared with the finite element method.

There are still technical challenges that need to be solved before in vivo studies become feasible. First, the accurate quantification of the target depends on the accurate estimation of target diameter and depth, which in turn depends on the image quality of the photoacoustic images. Since we do not have clinical examples of photoacoustic imaging yet, we used existing ultrasound images obtained from breast lesions of patients to estimate the potential error in future PAT clinical experiments. This is acceptable because photoacoustic and ultrasound images have similar resolutions, albeit different contrasts. Figure 10(a) is an ultrasound image of a benign lesion of a 53-year-old patient obtained prior to her ultrasound-guided core biopsy [case 1 in Fig. 10(b)]. An ultrasound expert marked the boundary of the lesion and estimated the depth of the lesion. In general, the depth of the lesion is relatively accurate, but the radius is more subjective. However, for smaller lesions in the neighborhood of 1 cm, the estimated radii are typically within 80% to 120% of the true size. This is based on our experience obtained from correlating sizes measured from ultrasound images and tumor stages reported from surgical pathology. Accordingly, in the simulation, the measured radius, together with 80% and 120% of it, were used in the fitting procedure, which represented 64% to 144% of the area of interest. The fitting results for five cases are plotted in Fig. 10(b). The first two cases were benign fibroadenoma, and the last three cases were malignant carcinoma. Clearly, the radius changes have influence on the fitting value, but the benign and malignant lesions can be separated. To further improve the photoacoustic imaging capability on delineating tumor margins, contrast agents may be needed to reliably identify the target parameters.

Second, the ultrasound transducer face generates photoacoustic waves upon the absorption of light from source fibers. These waves propagate to the target where they are reflected back to the transducer and picked up as signals, producing artifacts in the photoacoustic images; these artifacts are especially prominent in the low-contrast target images. We have been evaluating different type of materials that can be used at the face of the ultrasound transducer to effectively prevent light absorption and at the same time minimize the acoustic signal attenuation.28 Third, the current system is not portable for clinical experiment. In the current setup, the light is delivered through two 1000-micron step-index fibers, and the output energy density is about 80mJ/cm2 which exceed the ANSI safety standard (20mJ/cm2). We are looking for a compact laser and have cooperation with OFS Inc., which is developing a new fiber assembly using a proprietary power splitting technology to split light energy from one 940-micron core input fiber to nineteen 200-micron core output fibers. The fiber assembly will decrease the energy density output of the source fibers by a factor of 10. Fourth, the current fitting procedure is suitable for a single target. In principle, this method can be extended to multiple-target scenarios by taking care of cross-talk signals between targets; we are currently investigating several fitting procedures to address this issue.

Fig. 10

Simulation of the potential target-size effects on optical contrast quantification using target-size measurements obtained from clinical ultrasound images: (a) ultrasound image with marked boundary obtained from a 53-year-old woman; (b) the box and whisker plot of the maximum absorption of the lesions.

JBO_17_6_061213_f010.png

In summary, we have presented a new fitting method, which can quantitatively recover the absorption coefficient using a diffuse optical measurements-assisted photoacoustic tomography in reflection geometry. The background optical properties provided by the diffuse measurements can reduce the total unknowns in the fitting procedure and significantly improve the target characterization. This hybrid technique may overcome challenges and limitations of each technology alone and has potential in cancer detection and diagnosis.

Acknowledgments

The authors thank the National Institute of Health (R01EB002136), and Army Medical Research and Materiel Command Postdoctal Fellowship award (W81XWH-09-1-0511), for funding and supporting this work.

References

1. 

X. Wanget al., “Noninvasive laser-induced photoacoustic tomography for structural and functional in vivo imaging of the brain,” Nat. Biotechnol., 21 (7), 803 –806 (2003). http://dx.doi.org/10.1038/nbt839 NABIF9 1087-0156 Google Scholar

2. 

L. V. Wang, “Prospects of photoacoustic tomography,” Med. Phys., 35 (12), 5758 –5767 (2008). http://dx.doi.org/10.1118/1.3013698 MPHYA6 0094-2405 Google Scholar

3. 

G. Kuet al., “Thermoacoustic and photoacoustic tomography of thick biological tissues toward breast imaging,” Technol. Cancer Res. Treat., 4 (5), 559 –565 (2005). TCRTBS 1533-0346 Google Scholar

4. 

D. RazanskyC. VinegoniV. Ntziachristos, “Imaging of mesoscopic targets using selective-plane photoacoustic tomography,” Phys. Med. Biol., 54 (9), 2769 –2777 (2009). http://dx.doi.org/10.1088/0031-9155/54/9/012 PHMBA7 0031-9155 Google Scholar

5. 

K. H. Songet al., “Noninvasive photoacoustic identification of sentinel lymph nodes containing methylene blue in vivo in a rat model,” J. Biomed. Opt., 13 (5), 054033 (2008). http://dx.doi.org/10.1117/1.2976427 JBOPFO 1083-3668 Google Scholar

6. 

S. Ermilovet al., “Laser opto-acoustic imaging system for detection of breast cancer,” J. Biomed. Opt., 14 (2), 024007 (2009). http://dx.doi.org/10.1117/1.3086616 JBOPFO 1083-3668 Google Scholar

7. 

H. Brechtet al., “Whole-body three-dimensional optoacoustic tomography system for small animals,” J. Biomed. Opt., 14 (6), 064007 (2009). http://dx.doi.org/10.1117/1.3259361 JBOPFO 1083-3668 Google Scholar

8. 

E. ZhangJ. LauferP. Beard, “Backward-mode multiwavelength photoacoustic scanner using a planar Fabry-Perot polymer film ultrasound sensor for high-resolution three-dimensional imaging of biological tissues,” Appl. Opt., 47 (4), 561 –577 (2008). http://dx.doi.org/10.1364/AO.47.000561 APOPAI 0003-6935 Google Scholar

9. 

J. Lauferet al., “Three-dimensional noninvasive imaging of the vasculature in the mouse brain using a high resolution photoacoustic scanner,” Appl. Opt., 48 (10), 299 –306 (2009). http://dx.doi.org/10.1364/AO.48.00D299 APOPAI 0003-6935 Google Scholar

10. 

D. Piraset al., “Photoacoustic imaging of the breast using the twente photoacoustic mammoscope: present status and future perspectives,” IEEE J. Sel. Top. Quantum Electron, 16 (4), 730 –739 (2010). 0018-9197 Google Scholar

11. 

V. G. Andreevet al., “Opto-acoustic tomography of breast cancer with arc-array transducer,” Proc. SPIE, 3916 36 –47 (2000). http://dx.doi.org/10.1117/12.386339 PSISDG 0277-786X Google Scholar

12. 

A. Oraevskyet al., “Laser opto-acoustic imaging of breast cancer in vivo,” Proc. SPIE, 4256 6 –15 (2001). http://dx.doi.org/10.1117/12.429300 PSISDG 0277-786X Google Scholar

13. 

J. Gamelinet al., “Photoacoustic guidance of diffusive optical tomography with a hybrid reflection geometry probe,” Proc. SPIE, 7177 0701 –0711 (2009). http://dx.doi.org/10.1117/12.813547 PSISDG 0277-786X Google Scholar

14. 

P. D. Kumavoret al., “Target detection and quantification using a hybrid hand-held diffusive optical tomography and photoacoustic tomography system,” J. Biomed. Opt., 16 (4), 046010 (2011). http://dx.doi.org/10.1117/1.3563534 JBOPFO 1083-3668 Google Scholar

15. 

Z. YuanH. Jiang, “Quantitative photoacoustic tomography: recovery of optical absorption coefficient of heterogeneous media,” Appl. Phys. Lett., 88 (23), 213301 (2006). APPLAB 0003-6951 Google Scholar

16. 

Z. YuanQ. WangH. Jiang, “Reconstruction of optical absorption coefficient maps of heterogeneous media by photoacoustic tomography coupled with diffusion equation based regularized Newton method,” Opt. Express, 15 (26), 18076 –18081 (2007). http://dx.doi.org/10.1364/OE.15.018076 OPEXFF 1094-4087 Google Scholar

17. 

L. Yinet al., “Tomographic imaging of absolute optical absorption coefficient in turbid media using combined photoacoustic and diffusing light measurements,” Opt. Lett., 32 (17), 2556 –2558 (2007). http://dx.doi.org/10.1364/OL.32.002556 OPLEDP 0146-9592 Google Scholar

18. 

J. Lauferet al., “Quantitative spatially resolved measurement of tissue chromophore concentrations using photoacoustic spectroscopy: application to the measurement of blood oxygenation and haemoglobin concentration,” Phys. Med. Biol., 52 (1), 141 –168 (2007). http://dx.doi.org/10.1088/0031-9155/52/1/010 PHMBA7 0031-9155 Google Scholar

19. 

J. LauferE. ZhangP. Beard, “Evaluation of absorbing chromophores used in tissue phantoms for quantitative photoacoustic spectroscopy and imaging,” IEEE J. Sel. Top. Quantum. Electron., 16 (3), 600 –608 (2010). http://dx.doi.org/10.1109/JSTQE.2009.2032513 IJSQEN 1077-260X Google Scholar

20. 

A. RosenthalD. RazanskyV. Ntziachristos, “Quantitative opto-acoustic signal extraction using sparse signal representation,” IEEE Trans. Med. Imag., 28 (12), 1997 –2006 (2009). http://dx.doi.org/10.1109/TMI.2009.2027116 ITMID4 0278-0062 Google Scholar

21. 

R. Zemp, “Quantitative photoacoustic tomography with multiple optical sources,” Appl. Opt., 49 (18), 3566 –3572 (2010). http://dx.doi.org/10.1364/AO.49.003566 APOPAI 0003-6935 Google Scholar

22. 

Z. GuoS. HuL. Wang, “Calibration-free absolute quantification of optical absorption coefficients using acoustic spectra in 3-D photoacoustic microscopy of biological tissue,” Opt. Lett., 35 (12), 2067 –2069 (2010). http://dx.doi.org/10.1364/OL.35.002067 OPLEDP 0146-9592 Google Scholar

23. 

D. Boaset al., “Scattering of diffuse photon density waves by spherical inhomogeneities within turbid media—analytic solution and applications,” Proc. Natl. Acad. Sci. USA, 91 (11), 4887 –4891 (1994). http://dx.doi.org/10.1073/pnas.91.11.4887 1091-6490 Google Scholar

24. 

K. Kostliet al., “Temporal backward projection of opto-acoustic pressure transients using Fourier transform methods,” Phys. Med. Biol., 46 (7), 1863 –1872 (2001). http://dx.doi.org/10.1088/0031-9155/46/7/309 PHMBA7 0031-9155 Google Scholar

25. 

Q. Zhuet al., “Optimal probing of optical contrast of breast lesions of different size located at different depths by US localization,” Technol. Cancer Res. Treat., 5 (4), 365 –380 (2006). Google Scholar

26. 

N. G. Chenet al., “Simultaneous near infrared diffusive light and ultrasound imaging,” Appl. Opt., 40 (34), 6367 –6280 (2001). http://dx.doi.org/10.1364/AO.40.006367 APOPAI 0003-6935 Google Scholar

27. 

C. HoelenF. de Mul, “Image reconstruction for photoacoustic scanning of tissue structures,” Appl. Opt., 39 (31), 5872 –5883 (2000). http://dx.doi.org/10.1364/AO.39.005872 APOPAI 0003-6935 Google Scholar

28. 

B. Tavakoliet al., “Effect of ultrasound transducer face reflectivity on the light fluence inside a turbid medium in photoacoustic imaging,” J. Biomed. Opt., 15 (4), 046003 (2010). http://dx.doi.org/10.1117/1.3462930 JBOPFO 1083-3668 Google Scholar

29. 

Q. Zhuet al., “The potential role of optical tomography with ultrasound localization in assisting ultrasound diagnosis of early-stage invasive breast cancers,” Radiology, 256 (2), 367 –378 (2010). http://dx.doi.org/10.1148/radiol.10091237 RADLAX 0033-8419 Google Scholar

30. 

Q. Zhuet al., “Ultrasound-guided optical tomographic imaging of malignant and benign breast lesions: initial clinical results of 19 cases,” Neoplasia, 5 (5), 379 –388 (2003). 1522-8002 Google Scholar

31. 

D. Grosenicket al., “Time-domain scanning optical mammography: Optical II. properties and tissue parameters of 87 carcinomas,” Phys. Med. Biol., 50 (11), 2451 –2468 (2005). http://dx.doi.org/10.1088/0031-9155/50/11/002 PHMBA7 0031-9155 Google Scholar

32. 

P. D. KumavorA. AguirreQ. Zhu, “Reduction of secondary echoes generated from ultrasound transducer face in photoacoustic imaging implemented in reflection geometry,” Biomed. Opt., (2010). Google Scholar

34. 

N. Shahet al., “Noninvasive functional optical spectroscopy of human breast tissue,” Proc. Natl. Acad. Sci, 98 (8), 4420 –4425 (2001). http://dx.doi.org/10.1073/pnas.071511098 0369-3236 Google Scholar

35. 

T. Durduranet al., “Bulk optical properties of healthy female breast tissue,” Phys. Med. Biol., 47 (16), 2847 –2861 (2002). http://dx.doi.org/10.1088/0031-9155/47/16/302 PHMBA7 0031-9155 Google Scholar
© 2012 Society of Photo-Optical Instrumentation Engineers (SPIE) 0091-3286/2012/$25.00 © 2012 SPIE
Chen Xu, Patrick D. Kumavor, Andres Aguirre, and Quing Zhu "Investigation of a diffuse optical measurements-assisted quantitative photoacoustic tomographic method in reflection geometry," Journal of Biomedical Optics 17(6), 061213 (7 May 2012). https://doi.org/10.1117/1.JBO.17.6.061213
Published: 7 May 2012
Lens.org Logo
CITATIONS
Cited by 14 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Optical properties

Absorption

Acquisition tracking and pointing

Optical testing

Photoacoustic spectroscopy

Transducers

Ultrasonography

Back to Top