1. Introduction
Photoacoustic (PA) imaging is a hybrid imaging technology that provides multiscale spatial resolution and deep tissue penetration by utilizing intrinsic optical absorption with ultrasound (US) detection [
1,
2,
3]. Currently, the development of real-time PA imaging systems for biomedical applications (e.g., breast tumor and melanoma detection, vascular structure imaging and evaluation of muscle oxygenation) is greatly needed to assess anomalous tissue perfusion resulting from conditions such as ischemia or subcutaneous tumors [
4,
5,
6,
7,
8,
9,
10]. In addition, lots of novel handheld PA imaging systems have been proposed to provide diagnostic information for preclinical implementation [
11,
12,
13]. Most studies have provided their strategies for developing the handheld PA systems, such as using a customized linear transducer array or hemispherical transducer array to acquire real-time imaging [
13]. However, the details of simulation for designing a handheld PA probe have not been explicitly discussed in most articles. Specifically, enough photons need to be delivered and absorbed by the target within the American National Standards Institute (ANSI) safety limit to acquire PA imaging with a high signal-to-noise ratio (SNR) at a deep penetration depth in scattering medium. Therefore, simulations for designing a PA probe are needed.
The random directions of radiation caused by multiscattering and absorption dominate the light propagation in biological tissues, compromising light delivery to selected targets for optical and PA imaging. Therefore, previous research studies have attempted to optimize the designs of PA systems for biomedical applications to increase detection sensitivity and observe targets at a deeper depth [
14,
15,
16]. The radiative transfer equation (RTE) has been commonly employed to simulate light propagation in turbid materials. The Monte Carlo method, based on the RTE, provides a solution for light propagation simulation [
17]. Recently, the Monte Carlo method has been used to evaluate the designs of both dark-field and bright-field PA tomography [
14]. The researchers analyzed how the beam width and length, incident angle, and interval between illumination sources affect the light distribution in scattering medium [
14]. However, the boundary condition of the incident surface and interfaces between different layers of the Monte Carlo model is simplified (i.e., only initial weight is assigned based on the properties of the medium to determine the percentages of reflection and transmission) [
18]. In other words, the Monte Carlo simulation may encounter certain events such as the transition exaggeration between different layers [
19]. Alternatively, the finite element method (FEM) is another approach for solving the RTE and estimating light propagation in biological tissues. FEM solves problems in a finite geometric domain with the application of appropriate boundary conditions, and a diffusion approximation of RTE is normally applied with FEM when dealing with the forward scattering problem of photons. Previous studies have derived the progression of PA wave generation from the diffusion equation to pressure propagation [
20], investigation of PA image reconstruction [
21] and simulation of PA effects at the microscopic level [
22]. Thus, a complete model based on FEM (e.g., COMSOL) should be included for evaluating both light propagation estimation (i.e., the diffusion equation) and PA signal generation (i.e., heat generation, stress change, and the acoustic wave equation) to investigate the significant factors affecting the generation of PA signals. For instance, Wang et al. designed a new light illumination scheme (light catcher) based on a FEM-based simulation of PA imaging [
15]. Thus, FEM simulations were included in this study to provide an optimal way to deliver the light for acquiring PA imaging.
In this study, we will provide a guide to the development of a linear array-based, handheld, real-time PA imaging system designed for neurological applications, along with simulations and experimental results. In our design, the employed laser source with a fast-wavelength-tuning function can generate 100 high-energy pulses per second within the near-infrared (NIR) window (i.e., various wavelengths could be selected for different applications), ensuring the quality of multispectral imaging if multiple chromophores need to be evaluated at different wavelengths. Integration of this laser with a fast data acquisition US platform allows the required information to be analyzed in real-time with a high framerate. First, the results of simulation and phantom to evaluate the performance of the developed system are presented. Next, the system was also utilized to investigate the hemodynamic functions in rodent models of photothrombotic ischemia (PTI) and subcutaneous tumors. This handheld PA imaging system demonstrates real-time observations of symptoms/impairments in vascular functions.
2. Materials and Methods
To ensure that enough photons could be delivered and absorbed by the target for generating PA signals with a high SNR, we first need to conduct simulations using multiple models (e.g., heat transfer) to optimize the parameters. Then, after determination of the parameters for light delivery, we could customize a fiber bundle to integrate with the commercial linear-array transducer as the handheld PA imaging probe. Based on the characteristics of the laser source and US platform, we could further perform the multispectral PA imaging using a handheld probe in real-time.
2.1. Light Propagation Simulation in Scattering Medium
For the simulation in scattering medium, the target medium was 1% Lipofundin (B. Braun Singapore Pte. Ltd., Singapore), which has optical properties similar to those of organic soft tissue (i.e., µ
a = 0.026 cm
−1;
= 6.31 cm
−1; at 785 nm) [
23]. The dimensions of the target medium were 160 mm × 80 mm × 50 mm in this model. The position of the light sources for calculation was defined as the location where large scattering occurs:
where
and
are the absorption and scattering coefficient, respectively, while g is an anisotropic coefficient, defined as the average cosine of scattering angles [
18].
is the reduced scattering coefficient. Z0 is the transport mean free path; within the distance of Z0, photons propagate in its original directions with negligible scattering events. To fit the imaging dimensions of the rectangular-shaped US transducer array detector surface, we defined the dimensions of the laser beam as 16.5 mm × 0.8 mm. For highly scattering materials (i.e.,
), such as biological tissues and 1% Lipofundin, the light distribution follows the diffusion equations [
15,
16,
24]:
where n is the refractive index, c is the speed of light in vacuum (3e
8 m/s),
is the fluence in
, and
is the light source. The simulated laser source was a short pulsed, high-energy laser, with Gaussian energy distributed temporally.
, the energy of each pulse, was 20 mJ in our case. Pulse width (
) was 5 ns.
is the time point where the Gaussian pulse reaches its peak value.
The boundary condition assumption was selected based on the materials (e.g., 1% Lipofundin in this study) for our PA imaging system. For the incident surface, a Robin boundary condition is more appropriate [
15,
16,
25]:
where n is the refractive index of the incident materials (e.g., n = 1.35 for 1% Lipofundin), and R
eff is the internal reflection of uniformly diffuse radiation calculated using a curve fit by Egan and Hilgeman (i.e., Equation (7)). To derive the light fluence (φ) within biological specimens, we used the partial differential equation (PDE) modules of COMSOL Multiphysics (COMSOL Inc., Burlington, MA, USA).
2.2. Simulation of PA Wave Propagation Using COMSOL
In this study, PA simulation was conducted on a fairly small scale to reduce the considerable computation time. To mimic the biological tissue environment, we used a graphite sphere with a radius of 6 µm as the imaging target, while the surrounding medium was 1% Lipofundin with a radius of 300 µm. To evaluate the PA response, we considered the following physical phenomena:
- (i)
Light diffusion/propagation in the scattering medium (calculated in the previous section). We used the acquired fluence intensities (based on different illumination angles and intervals) at 10 mm depth as the fluence input for calculating the PA wave generation.
- (ii)
Heat generation and transfer inside the target and medium. For nanosecond laser pulse radiation, the temperature T was calculated using the following equation [
26,
27,
28]:
where C and k are the heat capacity and thermal conductivity, respectively. ρ is the density of the selected material. Q is the heat source, and it can be represented by the absorbed fluence times a yield Y (i.e., Q = ϕμ
aY).
- (iii)
Target stress-strain distribution. We also considered how the strain and stress of the target would change due to the transient temperature rise. The solid stress-strain model was used to evaluate this thermal expansion characteristic [
15,
27]:
where ɛ is the changed strain, and β represents the thermal expansion coefficient. T
ref is the reference temperature. The mechanical properties of the material were also included as inputs for this calculation, such as Young’s modulus and Poisson’s ratio.
- (iv)
Thermal expansion and PA wave generation/propagation. Thermal expansion was modeled using the boundary acceleration between the target and surrounding medium. The dominant equation is shown below:
where p is the response PA wave. The propagation of the PA wave in a medium follows the equation [
15,
22,
27]:
where c
s is the sound velocity [
16,
27]. Based on the above equations, we could simulate a complete procedure of PA wave generation for evaluating the proposed strategies of the PA imaging system design. Note that a schematic of simulation procedures using the FEM method is shown in
Figure S1 (Supplementary Materials) to illustrate the simulation strategy in details.
2.3. Handheld, Real-Time Photoacoustic Imaging System
In this study, we simulated, designed and developed a handheld, real-time PA imaging system with a fast multispectral function, as shown in
Figure 1. An optical parametric oscillator (OPO) (SpitLight EVO 200 OPO, InnoLas Laser GmbH, Krailling, Germany; 60 × 30 × 9 cm
3, 30 kg) was employed for laser illumination, while the input laser light was generated by a diode-pumped neodymium-doped yttrium aluminum garnet (Nd:YAG) laser at 532 nm to achieve a high signal output energy (27 mJ to 35 mJ based on the wavelength selection). The tuning wavelength range of the signal output is from 680 to 980 nm in the NIR region for deep penetration depth in biological tissues, while the wavelength range from 1180 to 2400 nm can also be provided by the Idler output port. This OPO can tune the wavelength per pulse at a 100-Hz repetition rate (i.e., the wavelength shift time is less than 10 ms, full span) to achieve the goal of fast multispectral functionality. In this study, we used only 800 nm for cerebral blood volume (CBV) evaluation, and 750 nm/850 nm for acquiring HbT and SO
2 information in real-time. The pulse width of the laser light from the OPO is 4–7 ns (suitable to generate the PA signal), while the line width is 10–450 cm
−1. Light from the OPO output was coupled with a customized fiber bundle (CeramOptec GmbH, Bonn, Germany) comprising 597 fibers. The input connector of the fiber bundle was specifically designed based on the dimensions of the output socket of the OPO; thus, the fiber bundle can be directly inserted into the OPO to decrease the laser energy loss. The fiber bundle has the following outputs: one single fiber was designed with a subminiature version A (SMA) connector for monitoring fluctuations in the laser energy; the remaining fibers were evenly distributed into two other arms with a rectangular output size of 16.5 mm × 0.8 mm. Note that a schematic of the customized fiber bundle design is shown in
Figure S2 (Supplementary Materials).
The generated PA signal was detected by a high-frequency US transducer array (L22-14v, Verasonics Inc., Kirkland, WA, USA) designed for imaging fine structures. This array is a 128-element linear transducer array with −6 dB bandwidth of 67% and center frequency at 18.5 MHz, while the pitch of each element is 0.10 mm. A 128-channel research US platform (Vantage 128, Verasonics Inc.; 50 × 30 × 50 cm3, 35 kg) was used for recording the PA signal. For performing high-quality, real-time imaging, 14-bit analog-to-digital (A/D) converters were included in the platform with a sample rate up to 62.5 MHz. For phantom and in vivo studies (e.g., small animal experiments), a customized XYZ 3-axis scanning stage (LS-110 with Hydra and Pollux controllers, Physik Instrumente (PI) GmbH & Co, Karlsruhe, Germany) with a frame holder was used to precisely image the region of interest (ROI). The travel range of each axis is 102 mm, and the step size can be as low as 1 µm, which can cover the entire region of a small animal’s cortex or hind leg with adequate accuracy. The total cost of the developed PA imaging system (including the laser source, US platform, scanning stage and the fiber bundle) is approximately $SGD 380 K.
When testing the system setup with phantoms (e.g., pencil lead and thin hair in a water tank filled with 1% Lipofundin), the arms of the fiber bundle were mounted on two manual linear stages (PMT2, PHOTONIK, Singapore) and rotation stages (MSRP01/M, Thorlabs, Newton, NJ, USA), which were fixed on the Z-axis of the scanning stage. The manual linear stages can adjust the interval between the two arms and ensure that these two arms are at the same height, while the rotation stage can adjust the light incident angle to verify the simulated results. In addition, the Z-axis of the scanning stage can adjust the illumination depth of the excitation beams. Next, the linear-array US transducer was fixed between the two arms of the fiber bundle for recording the generated PA signal, as shown in
Figure 1. The experimental setup for PA imaging of small animals is similar to the setup of the phantom experiments. During the imaging process, both the US transducer array and the output arms of the fiber bundle were immersed in an acrylic water tank with a square hole at the bottom, which was sealed with 15-µm-thick polyethylene film. A thin layer of US gel was applied between the thin film and the small animal’s organ, such as the skin (for tumor imaging) or cortex (for stroke imaging), to ensure proper acoustic coupling for in vivo experiments. The scanning stage was used to scan the ROI of the target in either the X or Y directions. The PA signal recorded by the US transducer was further processed by the US platform to perform real-time PA imaging.
One optimized wavelength (i.e., 800 nm) was employed to monitor the changes in CBV. The CBV was assumed to be proportional to the specific cortical region at 800 nm (very close to the isosbestic point where the absorption of oxygenated and deoxygenated hemoglobin is identical) [
29]. In addition, the total hemoglobin concentration (HbT) and oxygen saturation (SO
2) were estimated by using two wavelengths (i.e., 750 nm and 850 nm), based on the optical absorption coefficient of hemoglobin at different wavelengths [
29]:
where µ
a is the optical absorption coefficient; ε
Hb and ε
HbO2 are the molar extinction coefficients of Hb and HbO
2, respectively; and [Hb] and [HbO
2] are the molar concentrations of Hb and HbO
2, respectively [
29,
30]. For the maximum intensity projection (MIP) calculation, multiple B-scan images were first collected to form a 3D data set. Afterwards, a specific viewpoint (e.g., X-Y plane) was selected, and the image was reconstructed based on the pixels with maximum intensities of each column (e.g., the A-scan line along the Z-direction) covering the ROI.
2.4. Animal Experiments
All animal experimental protocols were approved by the Institutional Animal Care and Use Committee (IACUC) of the National University of Singapore (NUS). We also confirmed that all methods were performed in accordance with the guidelines and regulations of IACUC and the Office of Safety, Health & Environment (OSHE) of NUS. Animals were housed at a constant temperature and humidity with free access to food and water. For the stroke study, three adult male Sprague-Dawley (S-D) rats were used for the experiments. For PTI induction [
2], the photosensitizer Rose Bengal (Sigma, St. Louis, MO, USA) was diluted to 10 mg/mL in 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid (HEPES)-buffered saline and injected into the tail vein at 0.2 mL/100 g rat body weight. Then, a selected cortical blood vessel within a drilled cranial window (8 × 6 mm
2) was illuminated using a 5 mW, 532 nm continuous-wave (CW) laser light (AMGA-005, ONSET, Taipei, Taiwan) to generate an occlusion. PA imaging was conducted pre- and post-PTI to evaluate CBV changes. For the subcutaneous tumor study, 4T1 mammary carcinoma cells were cultured
in vitro before injection into immunocompromised mice. Tumor cells were inoculated subcutaneously at 1 × 10
5 viable cells in the right hind leg, and the tumor region was imaged using our PA imaging system 7 days after inoculation. Total three female NCr-nude mice were imaged for this study. Further details regarding animal preparation have been previously reported [
2,
31,
32]. Note that all data generated or analyzed during this study are included in this article (and its
Supplementary Materials file).
4. Discussion and Conclusions
In this study, we have demonstrated how to design/build a handheld, real-time PA imaging system from simulations (i.e., fluence and pressure simulations using COMSOL software and ultrasonic simulation using Vantage software) for applications in neurological diseases such as stroke and tumors.
FEM simulation is especially powerful for solving the PDE such as the diffusion equation [
15,
36]. This simulation can handle complex geometries, set proper boundary conditions and has advantages in solving inhomogeneity structures. Our simulation with the diffusion equation can simulate light propagation and distribution in a highly scattering medium, such as human tissues, while geometric optics can help us trace the light path and visualize the effect of the incident angle in water. The Monte Carlo simulation is a common tool and was used for the same purpose in a previous study [
14]. The light distribution with the variations of angles and intervals has been investigated using the Monte Carlo simulation. Their results showed that the light fluence dropped less than 7% at a depth of 10 mm when the incident angle changed from 15 degrees to 35 degrees. In addition, at the same depth (i.e., 10 mm), the light fluence decreased more than 25% when the interval between the two fiber arms increased from 10 mm to 14 mm. Their simulation results are consistent with ours (
Figure 3B,C), which support our simulation models/works [
14]. However, compared with the FEM simulation, the Monte Carlo simulation needs more computational capability to derive the precise solution for the RTE. In addition, a previous study cross-validated the Monte Carlo simulation, FEM and analytical solution methods in solving light transmission in skin. Comparing the Monte Carlo and FEM models, there were transition gaps in the photon density profile when light passes through different layers of skin by using the Monte Carlo simulation, whereas the results of FEM showed smooth transitions between different layers [
19]. Another drawback of the Monte Carlo model for simulating the PA effect is that it can be used to assess the optical part only, instead of the entire procedure of PA signal generation. That is, the calculated data must be processed using other tools, such as the k-wave toolbox in MATLAB, to generate the US responses. In contrast, COMSOL can process data from the light distribution to PA signal generation [
15,
16]. Our simulation model of PA signal generation involves the following modules: PDE, heat transfer, solid mechanics and acoustic generation. Specifically, our model considers not only the fluence diffusion but also the effects of material properties, such as thermal conductivity and Young’s modulus. In addition, for SO
2 measurement, to compensate for the differences of scattering and absorption coefficients of different wavelengths in soft tissues, we also employed this simulation model to examine the light propagation of 750 nm and 850 nm wavelengths for adjusting the incident light intensity (e.g., the incident laser intensity at 750 nm should be 12.2% larger than at 850 nm to ensure a consistent light intensity at a depth of 10 mm in scattering medium). Overall, we built the PA simulation model using COMSOL in this study to optimize the light path design in the selected medium and evaluate the performance on PA responses [
24]. According to the simulation results, we customized the PA probe with an optimized incident angle and interval between the two arms of the fiber bundle, allowing the pump beam to penetrate deeper in both water and the scattering medium. In future studies, we will further cross validate our results (using COMSOL) with other simulation methods (such as Monte Carlo + k-wave) to examine the performance of our simulation models for different applications. Note that the white circular area in the center of each subfigure of
Figure 4B is the imaging target. In our study, we did not demonstrate PA generation inside the target due to the module restriction. That is, the normal displacement of the target surface was used as the acoustic pressure source for observation in the medium, and we were unable to assign a boundary condition to the target surface for simultaneously evaluating the PA pressure propagations inside the target and in the medium (i.e., the simulation module disabled the boundary of the target surface when we tried to evaluate both the PA pressure propagations). Thus, we observed only the wave propagation as shown in
Figure 4B.
For in vivo applications, this developed PA imaging system is useful for monitoring disease conditions in multiple disease models. We successfully demonstrated that the symptoms of neurological diseases, such as ischemia during stroke and angiogenesis/hypoxia in a tumor, could be identified by our PA imaging system in real-time (e.g., B-scan image). Additionally, reconstructed MIP, multi-slice C-scan and 3D interactive images could be acquired/displayed after each scanning within a minute, showing the profile of the imaging target from different viewpoints (
Figure 8 and
Figure S4). As an example, in the near future, the middle cerebral artery occlusion (MCAo) disease model (i.e., the global stroke model) would benefit from this current system because of the deep imaging depth of the PA imaging system with an adequate lateral resolution, allowing for the monitoring of stroke progression. In addition, our high-energy/repetition rate pulsed laser can be employed for biomedical applications based not only on intrinsic contrast agents but on specific extrinsic probes such as nanoparticles. For example, PA imaging using dual-mode rare-earth nanoparticles with different excitation wavelengths could be utilized to almost simultaneously visualize multiple targets [
37,
38]. Furthermore, our PA imaging system and customized triple-modality nanoparticles with tumor biomarkers could be used for clinical applications in delineating brain tumor margins in the future.
Moreover, a high-performance US platform (i.e., the Vantage 128 system) was used to achieve real-time PA imaging analysis and display based on its radio frequency data transfer rate via 8 PCI express lanes (up to 6.6 GB/s). Because the data acquisition rate into local memory could be up to 100,000 frames/s, this US platform can also be utilized for ultrahigh-frequency US imaging. That is, we can develop a multimodal system including both PA and ultrafast US imaging to acquire multifaceted information. For example, in brain imaging, the US imaging function of the system could provide information on blood vessel structure and blood flow, while the PA imaging function can provide the status of functional hemodynamics, such as SO2, for preclinical examinations.
In summary, we provide a guide for designing a handheld, real-time PA imaging probe and system for preclinical neurological disease applications. FEM simulations were employed to evaluate the influence of incident angle and interval between the two arms of the fiber bundle for fluence propagation in scattering medium. In addition, a PA wave simulation model was built to assess the design strategy of the PA probe. Ultrasonic simulations were also conducted to determine the actual imaging resolutions of the current US imaging system, and phantom experimental validations were performed to examine the simulation results. The experimental results were consistent with the simulations, indicating that the interval between the two arms of the fiber bundle plays a major role in fluence propagation. Using the designed PA probe based on the given optimal parameters, we successfully reconstructed a color-coded PA image of phantom targets with depth information. Additionally, in vivo PTI and subcutaneous tumor experiments were conducted to verify the performance of the handheld PA imaging system in real-time. In the near future, we will continue to optimize the designs of the fiber bundle with attachable focusing components to achieve a higher energy density to increase the detection sensitivity. For instance, a lens tube will be customized for housing different kinds of optical components, with a specific connector at the end for connecting the fiber bundle. An optical diffuser could be used for the dark-field illumination at a deeper target region, while a cylindrical lens (i.e., to generate the laser line) will be used to design a bright-field handheld probe for delivering the incident light directly under the surface of the US linear-array transducer, increasing the SNR. In addition, we believe that the developed PA model can be employed to design PA nanoparticles for diverse applications, and provide directions for building optical/PA imaging systems to probe neurovascular functions in different disorders using small animal models.