Open Access
5 June 2024 Distribution-informed and wavelength-flexible data-driven photoacoustic oximetry
Author Affiliations +
Abstract

Significance

Photoacoustic imaging (PAI) promises to measure spatially resolved blood oxygen saturation but suffers from a lack of accurate and robust spectral unmixing methods to deliver on this promise. Accurate blood oxygenation estimation could have important clinical applications from cancer detection to quantifying inflammation.

Aim

We address the inflexibility of existing data-driven methods for estimating blood oxygenation in PAI by introducing a recurrent neural network architecture.

Approach

We created 25 simulated training dataset variations to assess neural network performance. We used a long short-term memory network to implement a wavelength-flexible network architecture and proposed the Jensen–Shannon divergence to predict the most suitable training dataset.

Results

The network architecture can flexibly handle the input wavelengths and outperforms linear unmixing and the previously proposed learned spectral decoloring method. Small changes in the training data significantly affect the accuracy of our method, but we find that the Jensen–Shannon divergence correlates with the estimation error and is thus suitable for predicting the most appropriate training datasets for any given application.

Conclusions

A flexible data-driven network architecture combined with the Jensen–Shannon divergence to predict the best training data set provides a promising direction that might enable robust data-driven photoacoustic oximetry for clinical use cases.

1.

Introduction

Blood oxygen saturation (sO2) is an important indicator of individual health status used routinely in patient management.1 Photoacoustic (PA) imaging (PAI) is a promising medical imaging modality for real-time non-invasive spatially resolved measurement of sO2,2 with early clinical applications3 shown in, for example, inflammatory bowel disease4 and cardiovascular diseases.5 In cancer, alterations in localized sO2 levels have been linked to angiogenesis and hypoxia,6 the key hallmarks of cancer that are known to affect treatment outcomes7 and are measurable through PAI.

Unfortunately, it remains difficult to apply PAI to derive quantitative values for sO2 from multispectral PA measurements.8,9 Linear unmixing (LU) remains the de facto standard for sO2 estimation from PAI measurements10 because of its simplicity and flexibility but has well-understood limitations in its applicability and accuracy.11 The limitations of LU are significant in the context of artifacts arising from: the optical processes (e.g., non-linear light fluence distribution12 leading to spectral coloring), acoustic processes (e.g., reflection artifacts13), or reconstruction algorithms (e.g., model mismatch in sound speed14). Using LU is particularly challenging in the presence of highly absorbing tissues, such as the epidermis,15 which can introduce reflection artifacts and a spectral bias that leads to an overestimation of sO2, which increases with darker skin tone.16

Data-driven unmixing schemes have shown promise to alleviate some of the shortcomings of LU1719 but suffer from three major drawbacks: (1) inflexibility to receiving different input data after training;20 (2) performance determined by the composition of the training dataset;21 and (3) limited testing on diverse and representative use cases.22 In comparison to LU, data-driven methods are often inflexible regarding the input data after training, impacting generalizability, making them difficult to use, and requiring laborious tailoring to a specific application and imaging system. Thus, data-driven sO2 estimation methods can struggle to translate promising findings from in silico to in vivo data.23,24 The lack of high-quality annotated training data and reliable validation data has made it difficult to implement data-driven methods robustly. One way of tackling this challenge lies in bridging the gap between simulated and actual PA data, exploiting realistic phantoms,25 an approach that has recently started to be explored.2628 Furthermore, many data-driven approaches use single-pixel input spectra for their inversion, as with LU, even though 3D inversions would be preferred for realistic use cases.29 Solving the full 3D problem is typically computationally intensive, limiting its success in vivo.30 To make the inverse problem more tractable without full 3D information, some approaches use priors in the inversion scheme,31 differential image analysis,9 or multiple measurements with differences in illumination.32

In this work, we set out to address the three aforementioned limitations. We improve the flexibility of data-driven sO2 estimation using a long short-term memory (LSTM) network that enables input wavelength flexibility. We propose a method to inform the choice of training data, which can either be used to choose the best pre-trained model for a target application or to inform the choice of simulation parameters when creating a training data set to underpin a new model. We test these methods on diverse data sources across simulations, phantoms, small animals, and humans.

2.

Materials and Methods

We begin by investigating sensitivity to changes in training dataset simulation parameters, by defining a baseline dataset (BASE) with typical assumptions on the tissue geometry and functional parameter ranges, then adapt it into 24 variations [Fig. 1(a)]. We use the Jensen–Shannon divergence to determine the ideal training dataset for a given use case. We propose a deep learning network architecture based on an LSTM network that is flexible regarding the input wavelengths [Fig. 1(b)] and use a testing strategy that comprises computational studies in silico, phantoms in gello,33 and in vivo data [Fig. 1(c)].

Fig. 1

Overview of the methods used. (a) Photon transport was simulated with a Monte Carlo light model for each of the 24 distinct datasets adapted from a baseline tissue assumption (BASE). Wavelength-dependent initial pressure spectra (right) were extracted from the vessels simulated in the leftmost panel. (b) A deep learning network based on an LSTM network was introduced to enable greater flexibility regarding input wavelengths for analysis. The hidden state of the LSTM was passed to fully connected layers, which output the estimated blood oxygenation sO2. (d) The performance of the LSTM-based method when trained on datasets with different tissue simulation parameters was tested across different datasets, ranging from in silico simulations and in gello phantom measurements, to in vivo measurements. This figure was created with Inkscape using BioRender assets.

JBO_29_S3_S33303_f001.png

2.1.

In Silico Datasets

Twenty-five in silico datasets were simulated in Python using the SIMPA toolkit34 [Fig. 1(a)]. A Monte Carlo model35 was used for the optical forward model with a 50 mJ Gaussian beam using 107 photons and a 20 mm radius, simulating at 41 wavelengths from 700 to 900 nm in 5 nm steps, and assuming an anisotropy of 0.9. For the datasets that include acoustic modeling, a 2D k-space pseudo-spectral time domain method implemented in the k-Wave toolbox36 was used. We assumed a uniform speed of sound of 1500  ms1, a density of 1000  gcm3, and disregarded acoustic absorption. For most datasets, a generic linear ultrasound detector array was placed in the center of the simulated volume. The generic array consists of 100 detection elements (modeled as rectangular elements) with a pitch of 0.18 mm, a length of 0.5 mm, a center frequency of 4 MHz, a bandwidth of 55%, and a sampling rate of 40 MHz. For the datasets mimicking the setup of commercial instruments, we used the built-in device definitions provided in SIMPA.

2.1.1.

Baseline dataset simulation parameters

The parameters of the BASE dataset were chosen to reflect the typical parameter choices found in the literature.20,29,3740 A 19.2×19.2×19.2  mm cube was simulated with a 0.3 mm voxel size, resulting in 643 voxels per volume. The background started from the second voxel from the top and was modeled as muscle tissue with a blood volume fraction of 1%20 with 70% oxygenation.41 We added 0 to 9 cylinders with a diameter randomly chosen from U[0.3 mm, 2.0 mm], where U denotes a uniform distribution. The tissue was divided into 3×3 equal-sized compartments in the imaging plane, and a vessel was included with a 50% probability. Each vessel purely contained blood with a randomly drawn sO2 value from U[0%, 100%]. To prevent vessel overlap, they were located within their respective compartments’ boundaries.

We defined 24 variations of BASE (Table 1). The variations encompassed a range of tissue backgrounds, vessel sizes, illumination geometries, and resolutions, as well as the addition of a skin layer, performing acoustic simulation, and using digital twins of two commercial instruments (MSOT Acuity Echo and MSOT InVision 256-TF; iThera Medical GmbH, Munich, Germany). The MSOT Acuity Echo is a handheld clinical PAI device with 256 detection elements and an angular coverage of 125 deg, whereas the MSOT InVision 256-TF is a tomographic pre-clinical PAI device with 256 detection elements and a 270 deg view angle. The full details for the digital twin parameters of these devices can be found in prior work26,34 and are available within the SIMPA toolkit. Background heterogeneities were introduced by applying a 3D Gaussian filter of size 1.2 mm to uniform noise.

Table 1

Summary of datasets used in the study.

Dataset identifierChanges from BASE
BG_0-100Background sO2 randomly chosen from U(0%, 100%)
BG_60-80Background sO2 randomly chosen from U(60%, 80%)
BG_H2OBackground modeled as water only
HET_0-100Background sO2 heterogeneously varied between 0% and 100%
HET_60-80Background sO2 heterogeneously varied between 60% and 80%
RES_0.15Simulation grid spacing: 0.15 mm
RES_0.15_SMALLSimulation grid spacing: 0.15 mm; radii of vessels halved
RES_0.6Simulation grid spacing: 0.6 mm
RES_1.2Simulation grid spacing: 1.2 mm
SKIN0.3 mm melanin layer; melanosome fraction: U(0.1%, 5%)
ILLUM_5mmRadius of incident beam: 5 mm
ILLUM_POINTRadius of incident beam: 0 mm
SMALLRadii of the vessels halved
ACOUSAcoustic modeling with linear transducer array
WATER_2cm2 cm H2O layer added between illumination and tissue
WATER_4cm4 cm H2O layer added between illumination and tissue
MSOTVolume extended: 75 × 19.2 × 19.2 mm; MSOT Acuity twin
MSOT_SKINMSOT + SKIN
MSOT_ACOUSMSOT + ACOUS
MSOT_ACOUS_SKINMSOT + SKIN + ACOUS
INVISVolume extended: 90 × 25 × 90 mm; MSOT InVision 256-tf twin. Single vessel in a 10 mm radius tubular background
INVIS_SKININVIS + SKIN
INVIS_ACOUSINVIS + ACOUS
INVIS_SKIN_ACOUSINVIS + SKIN + ACOUS
Each row identifies the changes in dataset creation parameters performed for each of the training datasets relative to the BASE dataset (specified in Sec. 2.1). The left column shows an identifier assigned to each dataset that is used throughout the paper and the right column summarizes the deviation from BASE. U denotes a uniform distribution.

2.1.2.

Data preprocessing

We extracted pixel-wise PA spectra from simulations of initial pressure or reconstructed signal amplitude if acoustic simulations were performed. For reconstruction, the backprojection algorithm implemented in the PATATO toolbox was used.42 Simulations were processed to extract the spectra in two steps: first, all spectra from blood vessels were selected; second, spectra where the signal intensity at 800 nm was less than 10% of the maximum were discarded. If less than 10% of voxels were chosen this way, we selected the 10% of voxels with the highest signal amplitude from the dataset. We enforced this selection criterion to effectively exclude voxels with low signal-to-noise ratio caused by optical attenuation, as previous works have shown that idealized simulations can contain spectra that display stronger spectral coloring than those actually seen in experimental data from tissues.20,32

The number of extracted spectra from the different dataset variations ranged from 25 thousand to 60 million, compared with BASE with 7 million spectra; the mean over all datasets was just above 6 million. The large difference in extractable spectra is primarily caused by two factors: (1) typical simulations yield 3D p0 distributions, but adding acoustic forward modeling leads to a 2D reconstructed image; and (2) some datasets only include a single vessel in the center of the phantom tube, mimicking a blood flow phantom43 (described below). To mitigate performance differences caused by this discrepancy, we stratified the dataset sizes by randomly sampling 300,000 spectra with replacement per dataset, which represents a balanced compromise between undersampling larger datasets and oversampling smaller ones.44

We performed z-score normalization on each spectrum, setting the mean (μ) to 0 and variance (σ2) to 1, which discards signal intensity information and eliminates the need for quantitative simulation calibration. Since we performed the same spectra-wise z-score normalization on experimental data, this normalization allows us to apply sO2 estimation algorithms trained in silico to experimental in gello and in vivo data.

2.2.

Deep Learning Algorithm

To address the limited flexibility of data-driven oximetry methods,20,40 a custom unmixing network architecture was composed that contains an LSTM network. Due to the recurrent nature of an LSTM, it can process sparse spectra containing zeros at arbitrary positions [see Fig. 1(b)].

The network input size was fixed at 41, representing the maximum number of wavelengths (between 700 and 900 nm in 5 nm steps) we consider during inference. The number is a trade-off between maximizing the spectral resolution of the input features and simulation time efficiency. The network started with an LSTM layer with a hidden size of 100. A masking layer was used to identify missing values and instruct the LSTM to ignore them. The LSTM output was then flattened into a fixed-length encoding. Following the LSTM, a three-layer fully connected neural network was used with input size 4200, hidden size 1000, and output size 1. A leaky rectified linear unit was used after each layer, and the activation function after the final layer was a sigmoid function to constrain sO2 predictions between 0 and 1 [Fig. 1(b)].

The deep learning networks were trained for 100 epochs, where each epoch included the entire training set. The parameters were optimized with the Adam optimizer from the Keras framework45 using an initial learning rate of 103 and a mean absolute error loss. The learning rate was halved upon a 5-epoch plateau of the validation loss.

2.3.

In Gello Blood Flow Phantom Imaging

Two variable oxygenation blood flow phantoms were imaged using a previously described protocol.43 Briefly, agar-based cylindrical phantoms with a radius of 10 mm were created, and a polyvinyl chloride tube (inner diameter 0.8 mm, outer diameter 0.9 mm) was embedded in the center before the agar was allowed to set at room temperature. The base mixture comprised 1.5% (w/v) agar (Sigma-Aldrich 9012-36-6, St. Louis, Missouri, United States) in Milli-Q water and was heated until dissolved. After cooling to 40°C, 2.08% (v/v) of pre-warmed intralipid (20% emulsion, Merck, 68890-65-3) and 0.74% (v/v) Nigrosin solution (0.5  mg/mL in Milli-Q water) were added, mixed, and poured into the mold. Imaging was performed using the MSOT InVision 256TF (iThera Medical GmbH, Munich, Germany) according to a previously established standard operating procedure.46 PA images of the phantom were acquired in the range between 700 and 900 nm with 20 nm increments. The first phantom was imaged using deuterium oxide (D2O, heavy water) as the coupling medium within the system, while the second was immersed in normal water (H2O) during imaging.

2.4.

In Vivo Human Forearm Imaging

Human forearm imaging was performed as part of the PAI Skin Tone study, which started in June 2023 following approval by the East of England—Cambridge South Research Committee (Ref: 23/EE/0019). The study was conducted in accordance with the Declaration of Helsinki and written informed consent was obtained from all study participants. Participants were excluded if they could not give consent, were under the age of 20 or over 80, or had a body mass index outside the range between 18.5 and 30. Imaging was performed using the MSOT Acuity Echo (iThera Medical GmbH, Munich, Germany) using laser light between 660 and 1300 nm, averaging over 10 scans each, and analysis was performed at five wavelengths (700, 730, 760, 800, and 850 nm). One forearm scan from N=7 randomly chosen subjects with Fitzpatrick type 1 or 2 was selected for the purposes of testing the method proposed in this work. The authors manually segmented the radial artery in each scan using the medical imaging interaction toolkit (MITK).47

2.5.

In Vivo Mouse Imaging

All animal procedures were conducted under project and personal licenses (PPL no PE12C2B96, PIL no I53057080), issued under the United Kingdom Animals (Scientific Procedures) Act, 1986, and compliance was approved locally by the CRUK Cambridge Institute Biological Resources Unit. Nine healthy 9-week-old female C57BL/6 albino mice were imaged using the MSOT InVision 256TF (iThera Medical GmbH, Munich, Germany) according to a previously established standard operating procedure.46 In addition, six 28-week-old healthy female BALB/c nude mice were imaged while inhaling 100% CO2 as their terminal procedure. In both cases, imaging was performed at 10 wavelengths equally spaced between 700 and 900 nm averaging over 10 scans each. The mouse body, kidneys, spleen, spine, and aorta were manually segmented by the authors using MITK.

2.6.

Performance Evaluation

The performance of the LSTM-based method was evaluated using the median absolute error (ϵsO2) between the estimate (sO^2) and the ground truth/reference sO2

Eq. (1)

ϵsO2=median(|sO2sO^2|).

Ground truth values are available for the in silico and in gello datasets. For the in vivo measurements, reference values were based on the literature. We assumed sO2 of mixed murine blood to be 60% to 70%41 and of arterial murine blood sO2 under anesthesia to be 94% to 98%.48 For the CO2 terminal procedure, we assumed that CO2 binds to hemoglobin, forming carbaminohemoglobin, which leads to oxygen unloading49 and has an absorption spectrum similar to deoxyhemoglobin,50,51 thus continuously decreasing the actual52 and measured global blood sO2. In humans, arterial blood sO2 of 95% to 100% was assumed.53

2.7.

Simulation Gap Measure

To predict the best-fitting training data set for a target application, one could use the sO2 estimates of a trained algorithm to calculate error metrics, such as the absolute estimation error, but this is only possible when ground truth or reference sO2 values for a representative dataset are available. For in vivo applications, this is typically not the case, and unsupervised methods for performance prediction in the context of PA oximetry remain largely unexplored.

Our alternative solution to this problem uses the Jensen–Shannon divergence54 (DJS). DJS measures the distance between distributions and finds application in, e.g., the training of generative adversarial networks.55 We compute DJS between spectra drawn from a reference and a target distribution (i.e., the training and the test data set) and calculate the Pearson correlation coefficient between the resulting DJS and ϵsO2 of the LSTM estimates. DJS measures the distance between unpaired samples drawn from two probability distributions P and Q. DJS is a symmetric version of the Kullback–Leibler divergence56 (DKL) and is defined as

Eq. (2)

DJS(P||Q)=12DKL(P||M)+12DKL(Q||M),
where M=12(P+Q) and the discrete DKL is defined as the relative entropy between two probability distributions

Eq. (3)

DKL(P||Q)=xXP(x)log(P(x)Q(x)).

To apply these measures, it is important to consider (1) handling the multidimensional probability distributions arising from multi-wavelength measurements and (2) the transformation of two sample distributions into the same sample space. We calculate an aggregate DJS by calculating the mean over the distance for each wavelength in the spectrum

Eq. (4)

DJS¯=1NλλΛDJS(Pλ||Qλ),
where Nλ is the number of all available wavelengths Λ. To standardize the sample space, a z-score normalization is performed for each spectrum, and a histogram with 100 bins ranging from 3σ to 3σ is created. A Python implementation of the Jensen–Shannon distance, available in the Scipy (v1.10.1) package,57 was used. Using this definition of DJS, only the intersection of two different sets of wavelengths can be compared.

3.

Results

3.1.

LSTM-Based Method Achieves Accurate Results across a Range of Available Input Wavelengths

The accuracy of the LSTM-based network architecture was first tested on the BASE data set, when varying the numbers of available wavelengths (Nλ) for training or inference. ϵsO2 was extracted when training and testing the network on a certain fixed Nλ, ranging from 3 to 41 wavelengths [Fig. 2(a)]. We found that ϵsO2 decreased as more wavelengths were used for training. Nevertheless, ϵsO2 for Nλ=10 was only slightly higher than for Nλ=41 wavelengths. For Nλ<10, ϵsO2 starts to rapidly increase, which aligns with prior literature31 and is potentially exacerbated due to the “vanishing gradient problem”58 in LSTMs, which arises when a substantial portion of the input parameter space consists of zeroes.

Fig. 2

LSTM-based method shows wavelength flexibility. (a) LSTMs were trained with varying numbers of wavelengths (Nλ) to show that with an increasing number of wavelengths, the accuracy of the predictions increases. (b) LSTM trained at a given Nλ can be applied to data with different Nλ but yields the best results when Nλ of the test spectra matches that of the training spectra (indicated by the green vertical line).

JBO_29_S3_S33303_f002.png

Next, a network trained at a fixed Nλ (in this case Nλ=20) was tested on data with a different Nλ [Fig. 2(b)]. Accuracy was found to decrease rapidly if fewer wavelengths were used for testing, but the error remains low if slightly more wavelengths are used. Nevertheless, the results show that the LSTM-based network performs best if the Nλ used during inference matches the Nλ during training.

3.2.

In Silico Cross-Validation Reveals the Effect of Changing Simulation Parameters

For each of the 24 simulation parameter variations of BASE, 500 data points with random spatial vessel distributions and a fixed random number generator seed for reproducibility were generated. To investigate the sensitivity of data-driven sO2 estimates to changes in training dataset parameters, we first trained LSTM-based networks using all 41 available wavelengths on each simulated training dataset and one on a mixed dataset (ALL). We then performed cross-validation on all datasets by applying every trained network to each of the datasets and calculating the median estimation error ϵsO2 [Fig. 3(a)]. The ϵsO2 values range from 0.5% to over 35%. As expected, the best performance occurs when testing on the training set; however, it is important to note that ϵsO2 is not zero (instead ranging from 0.5% to 5%), suggesting that the network has not overfitted the training data.

Fig. 3

Dimensionality reduction and cross-validation reveal systematic differences among training datasets. (a) An LSTM-based network trained on each dataset is then applied to every other dataset, and all ϵsO2 (median absolute error in percentage points) can be visualized as a performance matrix. Dataset names are shortened for visibility but are detailed in Table 1. (b) UMAP projections of the four representative example datasets onto an embedding of all training data. (c) Mapping the ground truth sO2 onto the same projection reveals a correlation along the first UMAP axis.

JBO_29_S3_S33303_f003.png

We calculated a Uniform Manifold Approximation and Projection (UMAP)59 embedding of 200,000 randomly chosen spectra from all datasets. With UMAP, we visualized the location of the spectra of representative datasets on this embedding [Fig. 3(b)]. Examining this visualization indicates that some changes in the dataset parameters result in highly different spectra (e.g., adding a layer of water on top of the tissue), while others lead to only minor variations (e.g., changing the background oxygenation). Labeling the UMAP embedding with the corresponding ground truth sO2 values reveals a correlation from low to high oxygenation along the horizontal UMAP axis [Fig. 3(c)].

From the in silico cross-validation heatmap [Fig. 3(a)], we can derive several key observations concerning the design of simulated data:

  • 1. Variation in background sO2 has a minimal effect with the used 1% blood volume fraction; however, this could become more significant at higher blood volume fractions.

  • 2. Resolution matters: Performance improves with higher spatial resolution simulations in the training data, suggesting that fine details in the spectral data are important for accurate sO2 estimation.

  • 3. Illumination matters: When changing from a Gaussian to a point source illumination, the error increased.

  • 4. Chromophore inclusion: When the test dataset contained melanin, but not the training data, the estimation error increased by an average of 5.8 percentage points. When designing a training dataset, all chromophores relevant to the target application should thus be included.

  • 5. Acoustic modeling causes systematic changes: Using acoustic modeling and image reconstruction introduced systematic spectral changes that increase εsO2 can have detrimental effects on the estimation accuracy and should be considered during training data simulation.

  • 6. Training on a combined dataset is better: Including random samples from all training datasets yielded more accurate estimates for all test datasets in silico. It should already be noted that this finding was not reproducible on the experimental datasets, suggesting that the LSTM-based method was not able to generalize better by training on a combined dataset.

3.3.

Jensen–Shannon Divergence Correlates with the Estimation Error and Can Therefore Be Used to Identify the Best Training Data Set

Given the variance in performance introduced by the choice of training data, it is desirable to automatically determine the best training dataset for a given algorithm and target application. The Jensen–Shannon divergence (DJS) allows one to quantify the distance between the data distribution of each dataset and the target data. We calculated the correlation between DJS and the median absolute sO2 estimation error (ϵsO2). When applying all networks, each trained on a distinct training dataset, to the BASE dataset, we found that DJS correlates strongly with ϵsO2 [Pearson correlation coefficient R=0.76, Fig. 4(a)]. We found the same results when correlating DJS with the mean squared error (R=0.77). We randomly sampled 100.000 spectra from the entire BASE dataset 10 times and computed the DJS score for each training dataset. The RES_0.15 dataset, which is the dataset simulated at the highest resolution, and not the BASE dataset, achieved the best DJS score. A possible reason for this could be that the BASE dataset is a subset of the RES_0.15 dataset and that we drew independent random samples from the entire training distribution.

Fig. 4

Jensen–Shannon divergence (DJS) can predict estimation performance. (a) DJS correlates with the median absolute sO2 estimation error ϵsO2 when applying all networks, each trained on a distinct training dataset, to the BASE dataset. (b) DJS for the D2O flow phantom data, which shows a similar correlation with ϵsO2. (c) After removing two outliers, DJS shows the same degree of correlation with ϵsO2 for the H2O flow phantom data.

JBO_29_S3_S33303_f004.png

Extending the analysis to experimentally acquired in gello D2O flow phantom data showed a similar correlation (R=0.74). The network trained on SMALL achieved the best score with DJS=0.35 and ϵsO2=4.2% [Fig. 4(b)]. The application of DJS to the H2O flow phantom experiment initially revealed no correlation (R=0.1); however, networks trained on ILLUM_POINT and MSOT_ACOUS_SKIN were outliers and after removing these, and the correlation was comparable to other datasets [R=0.72, Fig. 4(c)]. The network trained on WATER_4cm achieved the best score with DJS=0.47 and ϵsO2=12.6%. The presence of outliers emphasizes the importance of expert oversight when applying summary metrics such as DJS.

DJS correlates with ϵsO2 across multiple simulated and experimental data sets, providing evidence that the Jensen–Shannon divergence can predict algorithm performance. This is particularly relevant for previously unseen datasets where the true sO2 is unknown. For each training dataset, a DJS value can be computed by drawing random samples from both the training and unseen dataset, as outlined in Sec. 2.7. An LSTM-based network pre-trained on the dataset with the lowest corresponding DJS would then be chosen for data analysis since lower DJS correlates with a lower ϵsO2. The same strategy could also be used to guide an optimization process to tailor the simulation parameters to create a new training dataset that matches the target application.

3.4.

In Gello Testing Shows That the LSTM Method Outperforms Learned Spectral Decoloring (LSD)

Algorithm performance on the oxygenation flow phantom was compared with LU as the de facto state of the art and with a previously proposed LSD method.20 We show three example PA intensity images at 700 nm of the D2O flow phantom at three time points t = [0 min, 44 min, 70 min] [Fig. 5(a)], annotated with reference oxygenation (sO2ref) calculated from pO2 reference measurements using the Severinghaus equation.60

Fig. 5

Estimation of flow phantom data highlights performance dependence on the training dataset. Three example images of the D2O flow phantom are shown at different time points (0, 22, 44 min) displaying (a) the photoacoustic signal intensity at 700 nm with a red contour marking the blood-carrying tube and (b) the sO2 estimations from different methods. We visually compare the performance of LU (c), LSD (d), and the LSTM-based method (e) by plotting the sO2 estimations over the same image section and time points shown in panel (a).

JBO_29_S3_S33303_f005.png

Comparing the estimates of all methods trained on the SMALL training dataset by plotting the estimated sO2 over time [Fig. 5(b)] reveals that the LSTM-based method is, on average, more than twice as accurate as the LSD method and four times as accurate compared with LU. We show the LU estimates for the three example images [Fig. 5(c)], demonstrating the restricted dynamic range of LU estimates from t=0  min to t=44  min, ranging from 80% to 32%, compared with a ground truth of 99% to 1%. Example images for LSD [Fig. 5(d)] and the LSTM-based method [Fig. 5(e)] are shown as well, where the latter can recover the widest dynamic range, extending from 97% to 5%. For both methods, we chose SMALL as the training data set, as it was assigned the lowest DJS score. We compute the mean over the tube area, outlined in red, to exclude artifacts introduced by the limited transducer bandwidth and the reconstruction algorithm.

3.5.

Static In Vivo Testing Shows the Applicability of the Proposed Method to Different Use Cases

Having examined performance in the phantom system with known ground truth, we next apply the data-driven methods to static photoacoustic measurements of seven human forearms [Figs. 6(a)6(f)] and seven mouse abdomens [Figs. 6(g)6(l)]. For each, we show an example PA image, calculate DJS on the data distribution, and compare the estimated blood oxygenation in a region of interest (human forearm: radial artery; mouse: aorta and spine) with literature references.

Fig. 6

Jensen–Shannon divergence (DJS) proves valuable for in vivo data. LU and LSTM applied to measurements of the human forearm (a)–(f) and mice abdomens (g)–(l). Panels (a) and (g) show the photoacoustic signal at 800 nm, and panels (b) and (h) show the spread of DJS estimates for the training datasets. Panels (c) and (i) show boxplots of the highlighted regions of interest over all N=7 subjects. The horizontal lines show expected sO2 values for arterial blood (red) and mixed blood (blue). sO2 images are shown for models trained on a good fit [(d), (j)], a bad fit [(e), (k)], and the BASE dataset [(f), (l)] as predicted by DJS. On the bottom right of these images, the value distribution is shown as a grey histogram with the mean values of the regions of interest highlighted in their respective color.

JBO_29_S3_S33303_f006.png

For the forearm data [Fig. 6(a)], the MSOT_ACOUS_SKIN dataset is objectively the best fit and was assigned the second highest DJS score, whereas the ILLUM_POINT dataset was the worst fit [Fig. 6(b)]. Some estimated sO2 values were close to the expected radial artery sO2 value of 95% to 100%. Notably, the network trained on the MSOT_ACOUS_SKIN dataset results in sO290%, while the network trained on the ILLUM_POINT dataset produces sO295% [Fig. 6(c)].

The sO2 estimates of the network trained on the MSOT_ACOUS_SKIN dataset [Fig. 6(d)] seem to have three primary modes: high values >80% from the vessel structures, values in the 60% to 80% range in the surrounding tissue, and low values from 10% to 50% in the skin and deep tissue. The sO2 estimates of the network trained on the ILLUM_POINT dataset [Fig. 6(e)], on the other hand, are concentrated on high sO2 values in all superficial tissue and only seem to be below 85% in the skin and in deep tissue. The network trained on the BASE dataset [Fig. 6(f)] estimates low sO2 values throughout the entire tissue and does not exceed 80%. The ILLUM_POINT dataset, while seemingly successful if only considering values from the radial artery, was assigned the highest DJS value. The estimates and marginal histograms show that many estimates are mapped to <20% and >90%, explaining the good score in the radial artery. This finding demonstrates a common limitation of data-driven oximetry methods, where the estimated value distributions do not agree with expectations based on human physiology. The combination of all datasets (ALL) results in an extremely low sO2 estimate in the radial artery (median sO2<50%), which contradicts the in silico cross-validation results and indicates overfitting of the method to the training datasets.

For mouse images, the aorta and the area around the spinal cord are examined [Fig. 6(g)], assuming from literature a physiological arterial sO2 of 92% to 98% and for the spinal cord, a mixed arterial and venous blood with sO2 of 60% to 70%. WATER_4cm is objectively the best matching dataset and was assigned the highest score according to DJS [Fig. 6(h)]. All data-driven methods significantly increase the sO2 estimate in the aorta and lie within the desired bounds in the spinal cord [Fig. 6(i)]. The BASE [Fig. 6(l)] and WATER_4cm [Fig. 6(j)] datasets estimate a broad distribution and yield a higher sO2 estimate in the aorta and a larger spread between sO2 in the aorta and spinal cord compared with LU. The limitations of the ILLUM_POINT [Fig. 6(k)] dataset are even more evident in the mouse data, where even more pixels are either assigned 0% or 100%.

3.6.

Dynamic In Vivo Testing Demonstrates That the LSTM Method Can Reveal Physiological Processes

To provide a quantifiable decrease in the in vivo sO2 levels in mice that could test the capability of the LSTM-based method, we imaged N=6 mice when experiencing asphyxiation breathing 100% CO2.61 sO2 estimates were extracted from the major visible organs in the scan (spleen, kidneys, spinal cord, and aorta) at 3 min before and 10 min after CO2 asphyxiation.

CO2 asphyxiation [before, Figs. 7(a)7(d); after, Figs. 7(e)7(f)] increases the PA signal amplitude at 800 nm [Figs. 7(a) and 7(e)] in the superficial organs up to a depth of 3  mm, while the center of the mouse shows a decrease in signal. Pixels with negative signal intensities at 800 nm were excluded from the analysis (shown in black). sO2 values in the examined organs before CO2 asphyxiation are generally consistent between LU and data-driven unmixing methods, with the network trained on WATER_4cm estimating slightly lower sO2 values (5 to 8 percentage points lower) (Table 2). Notably, the direction of predicted effects on sO2 levels aligns well between LU and data-driven unmixing methods. Still, the effect sizes are up to three times greater when utilizing data-driven approaches, demonstrating a wider dynamic range. Intriguingly, in the case of the aorta, LU predicts an increase in sO2 levels despite the expected global decrease in sO2 levels due to CO2 exposure. This may be caused by the aforementioned increase in absorption coefficient in the periphery leading to an increase in spectral coloring in depth.

Fig. 7

Data-driven methods estimate an increased sO2 dynamic range during CO2 delivery compared to LU. A single representative mouse is shown here. Panels (a)–(d) show the photoacoustic image (a) and sO2 estimation results (b)–(d) 3 min before asphyxiation, and panels (e)–(h) show the photoacoustic image (e) and sO2 estimation results (f)–(h) 10 min after asphyxiation. We show sO2 estimates for LU [(b), (f)], the BASE [(c), (g)], and the best dataset as predicted by the Jensen–Shannon divergence [WATER_4cm (d), (h)]. Panels (a) and (e) show the outlines of the full-organ segmentations, and all other panels (b)–(d) and (f)–(h) show outlines of the segmented regions used in Table 2.

JBO_29_S3_S33303_f007.png

Table 2

sO2 decreases during CO2 asphyxiation.

Dataset →LUBASEWATER_4cm
Organ ↓Before (%)ΔsO2 (%)Before (%)ΔsO2 (%)Before (%)ΔsO2 (%)
Body46±14−2 (n.s.)44±20−7 (**)37±17−8 (**)
Spleen45±10−10 (**)40±19−24 (**)35±15−26 (**)
Kidney52±6−14 (**)54±17−29 (**)47±11−32 (**)
Spine55±6−14 (**)60±11−35 (**)52±11−34 (**)
Aorta67±6+5 (n.s.)76±5−7 (*)68±17−18 (*)
Reported are the mean ± standard deviation of sO2 measurement before CO2 asphyxiation (Before) and the change in the mean (ΔsO2) after the procedure. Values are reported at a maximum depth of 3 mm into the mouse body for: LU, an LSTM-based network trained on the baseline training dataset (BASE), and on the best fitting dataset according to the Jensen–Shannon divergence (WATER_4cm). p-values for ΔsO2 are calculated using a non-parametric Mann–Whitney U test and indicated by (n.s. = not significant; *p<0.05; **p<0.01). The segmentation masks are adjusted in Figs. 7(b)–7(d), 7(f)–7(h) to show the region of interest considered when calculating the results.

4.

Discussion

We present an LSTM-based method for estimating sO2 from multispectral PA images. We demonstrate that it can yield superior inference results compared with the previously proposed LSD method while at the same time being usable in a flexible manner, making it a promising candidate to replace LU as the de facto state of the art. We also show that the performance of the trained networks is highly dependent on the training data and that changes in simulation parameters can lead to drastically different data distributions. We thus propose to use the Jensen–Shannon divergence (DJS) to complement the LSTM-based method. DJS correlates with the median absolute sO2 estimation error and can thus be used to select the best-fitting training dataset or to optimize the training data distribution to fit the target application.

We highlight how the interplay of the LSTM-based method with DJS can be used on a diversity of in vivo human and mouse data acquired with different scanners and demonstrate that the LSTM-based method can reveal significant dependencies in sO2 changes that conventional LU would fail to identify. The LSTM-based method further consistently outperforms LU, with estimated sO2 values aligning better with ground truth measurements in gello and literature references in vivo. LU also shows significant outliers in regions where imaging artifacts are present, which are either not present or less pronounced when using the LSTM-based method.

Our in silico cross-validation reveals that acoustic modeling and image reconstruction introduce systematic spectral changes not explained by the initial pressure spectrum alone. Thus, an accurate digital model of the clinically used device is crucial during data simulation to ensure the best algorithm performance. While the combined dataset showed promising results in silico, these were not replicated in the experimental datasets, which may suggest that the network is overfitting and able to differentiate between the different simulated datasets.

DJS appears to be a valuable measure for determining the optimal training dataset for the LSTM-based method, as it correlates with the median absolute sO2 estimation error ϵsO2 in silico (R=0.76) and in gello (R=0.74/0.72). DJS predicts a plausible training dataset for all three in vivo applications tested in this study, where the predicted value range was 0.4 to 0.7 on the mouse data, 0.5 to 0.8 on the forearm data, and 0.3 to 0.7 on the CO2 data. With the development of fast and auto-differentiable simulation pipelines,62 it should be possible to optimize the simulation parameters for accurate sO2 estimates by iteratively minimizing DJS. When using differentiable implementations of distribution distance measures, it might even be possible to integrate this optimization into an unsupervised training routine.

The in gello experiments with H2O as the coupling medium had high ϵsO2 errors for all sO2 estimation methods and DJS was consistently high. The wavelength-dependent absorption of light by the water couplant likely adds further spectral coloring, which is not present in most of the simulated data sets. The predicted best training dataset was WATER_4cm and the worst ILLUM_POINT, which is consistent with the in vivo mouse experiments also having H2O as the coupling medium. Contrary to DJS prediction, ILLUM_POINT has the lowest ϵsO2, resulting in no correlation (R=0.11); after removing outliers, the correlation was on par with the other experiments (R=0.72). In both cases, the estimation error was lower than predicted by DJS; while the distributions were different, the trained networks still managed to estimate accurate sO2 values from the training data. Specifically, the ILLUM_POINT dataset, judging from the quasi-bimodal distributions of the network’s estimates on in vivo data, appears to have achieved a good agreement with high sO2 values purely by chance. More generally for the InVision experiments, the data sets that mimicked the InVision system were not the best-performing according to DJS, indicating greater systematic spectral differences compared with the more generic datasets. Outliers at one extreme, and more subtle impacts of different simulation parameters at another, are obscured by summary measures such as DJS, thus expert oversight for the use and interpretation of summary measures is needed.

The in vivo experiments with CO2 asphyxiation showed an increase in signal amplitude at 800 nm in the periphery with a decrease in signal in the center of the mouse. These phenomena suggest an overall increase in the absorption coefficient, which could be caused by a range of factors, including blood coagulation,63 erythrocyte aggregation,64 the presence of bicarbonate ions (HCO3) in the blood formed by the dissociation of carbonic acid into bicarbonate and hydrogen ions,49 or an increase in blood volume due to blood stasis. Vasoconstriction of the capillaries leading to pallor mortis could also play a role in the better visibility of the superficial organs.65

Having applied the LSTM-based method combined with DJS to a diverse range of data, we believe they provide a promising route to replacing LU for pixel-based PA oximetry. Combining the flexibility in the application of LU with the increase in accuracy of deep learning-based unmixing methods is attractive. The CO2 experiment suggests that LU can underestimate the effect size of sO2 changes due to its compressed dynamic range and susceptibility to artifacts. We thus recommend that LU should be complemented by a deep learning-based estimation method. The codes and data of this study are available open-source, facilitating its widespread testing and future application.

Nonetheless, there remain limitations to this study that should be the subject of further research. In this work, we investigated DJS predictions on a predefined set of datasets. Based on the results, we believe that DJS might be suitable to be used within a minimization scheme to either manually or automatically determine the best choice of simulation parameters for a given data set, but this remains to be investigated.

Low-resolution 2D acoustic modeling was used to limit the computational overhead, yet we found that the acoustic forward model and image reconstruction algorithm introduce systematic changes to the spectra. In the future, high-resolution 3D acoustic simulations that realistically follow the target hardware and model-based image reconstruction algorithms should be conisdered66 to limit the influence of artifacts introduced by the simulation or reconstruction algorithms. To account for spectral coloring artifacts more robustly, the full 2D—or better 3D—tissue context should be taken into account within the neural network, as quantitative PAI is only possible with the full 3D context.12,21 In addition, we have shown that good training data are key for deep learning-based methods for PA oximetry, as such, simulating data as realistically as possible is important. One direction toward this is to make use of domain adaptation methods27,28,67 that adapt simulated training data to appear more realistic.

5.

Conclusion

The presented LSTM-based approach for sO2 estimation from multispectral PA images surpasses the performance of LU and a previously reported data-driven oximetry method, making it a promising candidate to replace LU as the state of the art. We address the impact of training data variations by introducing the Jensen–Shannon divergence (DJS) as a valuable complement, enabling the selection of optimal datasets and fine-tuning for specific applications. Our LSTM-based method consistently outperforms LU, aligning well with ground truth measurements and literature references, while mitigating outliers in regions prone to imaging artifacts. The combination of the flexibility of the novel LSTM-based method with DJS for training data optimization is a promising direction to make data-driven oximetry methods robustly applicable for clinical use cases.

Disclosures

The authors declare no conflict of interest regarding this work.

Code and Data Availability

The data and code to reproduce the findings of this study are openly available. The data are available under the CC-BY 4.0 license at: https://doi.org/10.17863/CAM.105987. The code is available under the MIT license at: https://github.com/BohndiekLab/LearnedSpectralUnmixing.

Acknowledgments

This work was funded by Deutsche Forschungsgemeinschaft [DFG, German Research Foundation] (JMG; GR 5824/1), Cancer Research UK (SB, TRE; C9545/A29580), Cancer Research UK RadNet Cambridge (EVB; C17918/A28870), Against Breast Cancer (LH), and the Engineering and Physical Sciences Research Council (SB, EP/R003599/1). The work was supported by the NVIDIA Academic Hardware Grant Program and utilized two Quadro RTX 8000. The authors would like to thank Dr Mariam-Eleni Oraiopoulou for the helpful discussions. This work was supported by the International Alliance for Cancer Early Detection, a partnership among Cancer Research UK (C14478/A27855), the Canary Center at Stanford University, the University of Cambridge, OHSU Knight Cancer Institute, University College London, and the University of Manchester. This research was supported by the NIHR Cambridge Biomedical Research Centre (Grant No. NIHR203312). The views expressed are those of the authors and not necessarily those of the NIHR or the Department of Health and Social Care.

References

1. 

A. Fawzy et al., “Racial and ethnic discrepancy in pulse oximetry and delayed identification of treatment eligibility among patients with COVID-19,” JAMA Internal Med., 182 (7), 730 –738 https://doi.org/10.1001/jamainternmed.2022.1906 (2022). Google Scholar

2. 

P. Beard, “Biomedical photoacoustic imaging,” Interface Focus, 1 (4), 602 –631 https://doi.org/10.1098/rsfs.2011.0028 (2011). Google Scholar

3. 

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

4. 

F. Knieling et al., “Multispectral optoacoustic tomography for assessment of Crohn’s disease activity,” New Engl. J. Med., 376 (13), 1292 –1294 https://doi.org/10.1056/NEJMc1612455 NEJMBH (2017). Google Scholar

5. 

A. Karlas et al., “Cardiovascular optoacoustics: from mice to men—a review,” Photoacoustics, 14 19 –30 https://doi.org/10.1016/j.pacs.2019.03.001 (2019). Google Scholar

6. 

M. W. Dewhirst, Y. Cao and B. Moeller, “Cycling hypoxia and free radicals regulate angiogenesis and radiotherapy response,” Nat. Rev. Cancer, 8 (6), 425 –437 https://doi.org/10.1038/nrc2397 NRCAC4 1474-175X (2008). Google Scholar

7. 

D. Hanahan, “Hallmarks of cancer: new dimensions,” Cancer Discov., 12 (1), 31 –46 https://doi.org/10.1158/2159-8290.CD-21-1059 (2022). Google Scholar

8. 

J. Laufer et 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 https://doi.org/10.1088/0031-9155/52/1/010 PHMBA7 0031-9155 (2006). Google Scholar

9. 

C. Bench and B. Cox, “Quantitative photoacoustic estimates of intervascular blood oxygenation differences using linear unmixing,” in J. Phys.: Conf. Ser., 012001 (2021). https://doi.org/10.1088/1742-6596/1761/1/012001 Google Scholar

10. 

S. Tzoumas and V. Ntziachristos, “Spectral unmixing techniques for optoacoustic imaging of tissue pathophysiology,” Philos. Trans. R. Soc. A Math. Phys. Eng. Sci., 375 (2107), 20170262 https://doi.org/10.1098/rsta.2017.0262 (2017). Google Scholar

11. 

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

12. 

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

13. 

M. K. A. Singh et al., “In vivo demonstration of reflection artifact reduction in photoacoustic imaging using synthetic aperture photoacoustic-guided focused ultrasound (PAFUSion),” Biomed. Opt. Express, 7 (8), 2955 –2972 https://doi.org/10.1364/BOE.7.002955 BOEICL 2156-7085 (2016). Google Scholar

14. 

J. Jose et al., “Speed-of-sound compensated photoacoustic tomography for accurate imaging,” Med. Phys., 39 (12), 7262 –7271 https://doi.org/10.1118/1.4764911 MPHYA6 0094-2405 (2012). Google Scholar

15. 

Y. Mantri and J. V. Jokerst, “Impact of skin tone on photoacoustic oximetry and tools to minimize bias,” Biomed. Opt. Express, 13 (2), 875 –887 https://doi.org/10.1364/BOE.450224 BOEICL 2156-7085 (2022). Google Scholar

16. 

T. R. Else et al., “The effects of skin tone on photoacoustic imaging and oximetry,” (2023). Google Scholar

17. 

G. P. Luke et al., “O-net: a convolutional neural network for quantitative photoacoustic image segmentation and oximetry,” (2019). Google Scholar

18. 

S. Agrawal et al., “Learning optical scattering through symmetrical orthogonality enforced independent components for unmixing deep tissue photoacoustic signals,” IEEE Sens. Lett., 5 (5), 7001704 https://doi.org/10.1109/LSENS.2021.3073927 (2021). Google Scholar

19. 

V. Grasso, R. Willumeit-Römer and J. Jose, “Superpixel spectral unmixing framework for the volumetric assessment of tissue chromophores: a photoacoustic data-driven approach,” Photoacoustics, 26 100367 https://doi.org/10.1016/j.pacs.2022.100367 (2022). Google Scholar

20. 

J. Gröhl et al., “Learned spectral decoloring enables photoacoustic oximetry,” Sci. Rep., 11 (1), 6565 https://doi.org/10.1038/s41598-021-83405-8 SRCEC3 2045-2322 (2021). Google Scholar

21. 

J. Gröhl et al., “Deep learning for biomedical photoacoustic imaging: a review,” Photoacoustics, 22 100241 https://doi.org/10.1016/j.pacs.2021.100241 (2021). Google Scholar

22. 

H. Assi et al., “A review of a strategic roadmapping exercise to advance clinical translation of photoacoustic imaging: from current barriers to future adoption,” Photoacoustics, 32 100539 https://doi.org/10.1016/j.pacs.2023.100539 (2023). Google Scholar

23. 

V. Sandfort et al., “Data augmentation using generative adversarial networks (cyclegan) to improve generalizability in CT segmentation tasks,” Sci. Rep., 9 (1), 16884 https://doi.org/10.1038/s41598-019-52737-x SRCEC3 2045-2322 (2019). Google Scholar

24. 

L. Maier-Hein et al., “Surgical data science—from concepts toward clinical translation,” Med. Image Anal., 76 102306 https://doi.org/10.1016/j.media.2021.102306 (2022). Google Scholar

25. 

L. Hacker et al., “Criteria for the design of tissue-mimicking phantoms for the standardization of biophotonic instrumentation,” Nat. Biomed. Eng., 6 (5), 541 –558 https://doi.org/10.1038/s41551-022-00890-6 (2022). Google Scholar

26. 

J. Gröhl et al., “Moving beyond simulation: data-driven quantitative photoacoustic imaging using tissue-mimicking phantoms,” (2023). Google Scholar

27. 

A. K. Susmelj et al., “Signal domain adaptation network for limited-view optoacoustic tomography,” Med. Image Anal., 91 103012 https://doi.org/10.1016/j.media.2023.103012 (2023). Google Scholar

28. 

K. K. Dreher et al., “Unsupervised domain transfer with conditional invertible neural networks,” (2023). Google Scholar

29. 

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 https://doi.org/10.1117/1.JBO.25.8.085003 JBOPFO 1083-3668 (2020). Google Scholar

30. 

B. T. Cox, S. R. Arridge and P. C. Beard, “Estimating chromophore distributions from multiwavelength photoacoustic images,” J. Opt. Soc. Am. A, 26 (2), 443 –455 https://doi.org/10.1364/JOSAA.26.000443 JOAOD6 0740-3232 (2009). Google Scholar

31. 

S. Tzoumas et al., “Eigenspectra optoacoustic tomography achieves quantitative blood oxygenation imaging deep in tissues,” Nat. Commun., 7 (1), 12121 https://doi.org/10.1038/ncomms12121 NCAOBW 2041-1723 (2016). Google Scholar

32. 

T. Kirchner and M. Frenz, “Multiple illumination learned spectral decoloring for quantitative optoacoustic oximetry imaging,” J. Biomed. Opt., 26 (8), 085001 https://doi.org/10.1117/1.JBO.26.8.085001 JBOPFO 1083-3668 (2021). Google Scholar

33. 

S. Manohar, “‘in gello’ imaging,” (2023). Google Scholar

34. 

J. Gröhl et al., “SIMPA: an open-source toolkit for simulation and image processing for photonics and acoustics,” J. Biomed. Opt., 27 (8), 083010 https://doi.org/10.1117/1.JBO.27.8.083010 JBOPFO 1083-3668 (2022). Google Scholar

35. 

Q. Fang and D. A. Boas, “Monte Carlo simulation of photon migration in 3D turbid media accelerated by graphics processing units,” Opt. Express, 17 (22), 20178 –20190 https://doi.org/10.1364/OE.17.020178 OPEXFF 1094-4087 (2009). Google Scholar

36. 

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 https://doi.org/10.1117/1.3360308 JBOPFO 1083-3668 (2010). Google Scholar

37. 

C. Cai et al., “End-to-end deep neural network for optical inversion in quantitative photoacoustic imaging,” Opt. Lett., 43 (12), 2752 –2755 https://doi.org/10.1364/OL.43.002752 OPLEDP 0146-9592 (2018). Google Scholar

38. 

T. Chen et al., “A deep learning method based on u-net for quantitative photoacoustic imaging,” Proc. SPIE, 11240 112403V https://doi.org/10.1117/12.2543173 PSISDG 0277-786X (2020). Google Scholar

39. 

K. Hoffer-Hawlik and G. P. Luke, “absO2luteU-net: tissue oxygenation calculation using photoacoustic imaging and convolutional neural networks,” (2019). Google Scholar

40. 

I. Olefir et al., “Deep learning-based spectral unmixing for optoacoustic imaging of tissue oxygen saturation,” IEEE Trans. Med. Imaging, 39 (11), 3643 –3654 https://doi.org/10.1109/TMI.2020.3001750 ITMID4 0278-0062 (2020). Google Scholar

41. 

D. G. Lyons et al., “Mapping oxygen concentration in the awake mouse brain,” Elife, 5 e12024 https://doi.org/10.7554/eLife.12024 (2016). Google Scholar

42. 

T. R. Else et al., “PATATO: a Python photoacoustic tomography analysis toolkit,” J. Open Source Softw., 9 (93), 5686 https://doi.org/10.21105/joss.05686 (2024). Google Scholar

43. 

M. Gehrung, S. E. Bohndiek and J. Brunker, “Development of a blood oxygenation phantom for photoacoustic tomography combined with online pO2 detection and flow spectrometry,” J. Biomed. Opt., 24 (12), 121908 https://doi.org/10.1117/1.JBO.24.12.121908 JBOPFO 1083-3668 (2019). Google Scholar

44. 

A. Halevy, P. Norvig and F. Pereira, “The unreasonable effectiveness of data,” IEEE Intell. Syst., 24 (2), 8 –12 https://doi.org/10.1109/MIS.2009.36 (2009). Google Scholar

45. 

F. Chollet et al., “Keras,” https://keras.io (2015). Google Scholar

46. 

J. Joseph et al., “Evaluation of precision in optoacoustic tomography for preclinical imaging in living subjects,” J. Nucl. Med., 58 (5), 807 –814 https://doi.org/10.2967/jnumed.116.182311 JNMEAQ 0161-5505 (2017). Google Scholar

47. 

I. Wolf et al., “The medical imaging interaction toolkit,” Med. Image Anal., 9 (6), 594 –604 https://doi.org/10.1016/j.media.2005.04.005 (2005). Google Scholar

48. 

A. M. Loeven et al., “Arterial blood sampling in male CD-1 AND C57BL/6J mice with 1% isoflurane is similar to awake mice,” J. Appl. Physiol., 125 (6), 1749 –1759 https://doi.org/10.1152/japplphysiol.00640.2018 (2018). Google Scholar

49. 

H. Abramczyk et al., “Hemoglobin and cytochrome c. reinterpreting the origins of oxygenation and oxidation in erythrocytes and in vivo cancer lung cells,” Sci. Rep., 13 (1), 14731 https://doi.org/10.1038/s41598-023-41858-z SRCEC3 2045-2322 (2023). Google Scholar

50. 

E. Dervieux et al., “Measuring hemoglobin spectra: searching for carbamino-hemoglobin,” J. Biomed. Opt., 25 (10), 105001 https://doi.org/10.1117/1.JBO.25.10.105001 JBOPFO 1083-3668 (2020). Google Scholar

51. 

M. Taylor-Williams et al., “Noninvasive hemoglobin sensing and imaging: optical tools for disease diagnosis,” J. Biomed. Opt., 27 (8), 080901 https://doi.org/10.1117/1.JBO.27.8.080901 JBOPFO 1083-3668 (2022). Google Scholar

52. 

P. R. Huber et al., “CO2 angiography,” Catheter. Cardiovasc. Interv., 55 (3), 398 –403 https://doi.org/10.1002/ccd.10123 (2002). Google Scholar

53. 

A. J. Williams, “Assessing and interpreting arterial blood gases and acid-base balance,” BMJ, 317 (7167), 1213 –1216 https://doi.org/10.1136/bmj.317.7167.1213 (1998). Google Scholar

54. 

J. Lin, “Divergence measures based on the Shannon entropy,” IEEE Trans. Inf. Theory, 37 (1), 145 –151 https://doi.org/10.1109/18.61115 IETTAW 0018-9448 (1991). Google Scholar

55. 

I. Goodfellow et al., “Generative adversarial nets,” in Adv. Neural Inf. Process. Syst., (2014). Google Scholar

56. 

S. Kullback and R. A. Leibler, “On information and sufficiency,” Ann. Math. Stat., 22 (1), 79 –86 https://doi.org/10.1214/aoms/1177729694 AASTAD 0003-4851 (1951). Google Scholar

57. 

P. Virtanen et al., “SciPy 1.0: fundamental algorithms for scientific computing in Python,” Nat. Methods, 17 261 –272 https://doi.org/10.1038/s41592-019-0686-2 1548-7091 (2020). Google Scholar

58. 

R. M. Schmidt, “Recurrent neural networks (RNNs): a gentle introduction and overview,” (2019). Google Scholar

59. 

L. McInnes, J. Healy and J. Melville, “UMAP: Uniform Manifold Approximation and Projection for dimension reduction,” (2018). Google Scholar

60. 

J. W. Severinghaus, “Simple, accurate equations for human blood O2 dissociation computations,” J. Appl. Physiol., 46 (3), 599 –602 https://doi.org/10.1152/jappl.1979.46.3.599 (1979). Google Scholar

61. 

M. N. Fadhel et al., “Fluence-matching technique using photoacoustic radiofrequency spectra for improving estimates of oxygen saturation,” Photoacoustics, 19 100182 https://doi.org/10.1016/j.pacs.2020.100182 (2020). Google Scholar

62. 

T. Rix et al., “Efficient photoacoustic image synthesis with deep learning,” Sensors, 23 (16), 7085 https://doi.org/10.3390/s23167085 SNSRES 0746-9462 (2023). Google Scholar

63. 

R. Su et al., “Optoacoustic 3D visualization of changes in physiological properties of mouse tissues from live to postmortem,” Proc. SPIE, 8223 82230K https://doi.org/10.1117/12.910975 PSISDG 0277-786X (2012). Google Scholar

64. 

R. K. Saha and M. C. Kolios, “A simulation study on photoacoustic signals from red blood cells,” J. Acoust. Soc. Am., 129 (5), 2935 –2943 https://doi.org/10.1121/1.3570946 JASMAN 0001-4966 (2011). Google Scholar

65. 

A. T. Schäfer, “Colour measurements of pallor mortis,” Int. J. Legal Med., 113 81 –83 https://doi.org/10.1007/PL00007713 IJLMEA 1427-1596 (2000). Google Scholar

66. 

C. Dehner et al., “A deep neural network for real-time optoacoustic image reconstruction with adjustable speed of sound,” Nat. Mach. Intell., 5 (10), 1130 –1141 https://doi.org/10.1038/s42256-023-00724-3 (2023). Google Scholar

67. 

J. Li et al., “Deep learning-based quantitative optoacoustic tomography of deep tissues in the absence of labeled experimental data,” Optica, 9 (1), 32 –41 https://doi.org/10.1364/OPTICA.438502 (2022). Google Scholar

Biography

Janek Gröhl completed his MSc degree in medical computer science in 2016 and his PhD in April 2021 under Prof. Lena Maier-Hein at the German Cancer Research Center. He is a postdoctoral fellow with Prof. Sarah Bohndiek funded by the Walter Benjamin Programme of the German Research Foundation focusing on using machine learning methods and physical modeling to tackle the problem of quantitative photoacoustic imaging.

Kylie Yeung completed her MRes in connected electronic and photonic systems in 2022, jointly between University College London and the University of Cambridge, during which she investigated the effect of different simulated training datasets for quantitative photoacoustic imaging. She is currently pursuing a PhD at the University of Oxford.

Kevin Gu completed his MSci degree in experimental and theoretical physics in 2022 at the University of Cambridge. His dissertation was focused on a wavelength-flexible approach to data-driven photoacoustic oximetry. He now works as a quantitative researcher in the Netherlands.

Thomas R. Else completed his MSci degree in experimental and theoretical physics in 2019 and is currently pursuing a PhD in medical sciences, both at the University of Cambridge. His research focuses on the clinical translation of photoacoustic imaging, with a focus on the equitable application of the technology through open-access software and evaluation of skin tone biases.

Monika Golinska completed her PhD in oncology at the University of Cambridge in 2012. During her postdoctoral research at Prof. Sarah Bohndiek’s lab, she studied the relationship between sex hormones and tumor microenvironment in breast cancer. She is currently an MCSA fellow at the Medical University of Lodz and researches endometriosis and its link with ovarian cancer.

Ellie V. Bunce completed her MPharmacol degree (Hons) at the University of Bath in 2021. She is currently pursuing a PhD in medical sciences at the University of Cambridge, investigating the potential of vascular-targeted therapies to enhance cancer radiotherapy response using novel imaging approaches.

Lina Hacker is a junior research fellow at the Department of Oncology at the University of Oxford, United Kingdom. Her research is focused on the medical and technical validation of novel approaches for cancer imaging, specifically relating to tumor hypoxia. She received her PhD in medical sciences at the University of Cambridge, United Kingdom, and holds a master’s and bachelor’s degree in biomedical engineering and molecular medicine, respectively.

Sarah E. Bohndiek completed her PhD in radiation physics at the University College London in 2008 and then was a postdoctoral fellow in molecular imaging in both the United Kingdom (at Cambridge) and the United States (at Stanford). Since 2013, she has been a group leader at the University of Cambridge, where she is jointly appointed in the Department of Physics and the Cancer Research UK Cambridge Institute. She was appointed as a full professor of biomedical physics in 2020. Sarah was recently awarded the CRUK Future Leaders in Cancer Research Prize and the SPIE Early Career Achievement Award in recognition of her innovation in biomedical optics.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 International License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Janek Gröhl, Kylie Yeung, Kevin Gu, Thomas R. Else, Monika Golinska, Ellie V. Bunce, Lina Hacker, and Sarah E. Bohndiek "Distribution-informed and wavelength-flexible data-driven photoacoustic oximetry," Journal of Biomedical Optics 29(S3), S33303 (5 June 2024). https://doi.org/10.1117/1.JBO.29.S3.S33303
Received: 19 March 2024; Accepted: 17 May 2024; Published: 5 June 2024
Advertisement
Advertisement
KEYWORDS
Education and training

Computer simulations

Blood

In vivo imaging

Tissues

Error analysis

Oximetry

Back to Top