Open Access
8 March 2023 Unrolled-DOT: an interpretable deep network for diffuse optical tomography
Yongyi Zhao, Ankit Raghuram, Fay Wang, Stephen Hyunkeol Kim, Andreas H. Hielscher, Jacob T. Robinson, Ashok Veeraraghavan
Author Affiliations +
Abstract

Significance

Imaging through scattering media is critical in many biomedical imaging applications, such as breast tumor detection and functional neuroimaging. Time-of-flight diffuse optical tomography (ToF-DOT) is one of the most promising methods for high-resolution imaging through scattering media. ToF-DOT and many traditional DOT methods require an image reconstruction algorithm. Unfortunately, this algorithm often requires long computational runtimes and may produce lower quality reconstructions in the presence of model mismatch or improper hyperparameter tuning.

Aim

We used a data-driven unrolled network as our ToF-DOT inverse solver. The unrolled network is faster than traditional inverse solvers and achieves higher reconstruction quality by accounting for model mismatch.

Approach

Our model “Unrolled-DOT” uses the learned iterative shrinkage thresholding algorithm. In addition, we incorporate a refinement U-Net and Visual Geometry Group (VGG) perceptual loss to further increase the reconstruction quality. We trained and tested our model on simulated and real-world data and benchmarked against physics-based and learning-based inverse solvers.

Results

In experiments on real-world data, Unrolled-DOT outperformed learning-based algorithms and achieved over 10× reduction in runtime and mean-squared error, compared to traditional physics-based solvers.

Conclusion

We demonstrated a learning-based ToF-DOT inverse solver that achieves state-of-the-art performance in speed and reconstruction quality, which can aid in future applications for noninvasive biomedical imaging.

1.

Introduction

Light scattering is one of the primary limitations for optical imaging in biological tissue. It leads to lower spatial resolution, contrast and signal-to-noise ratio (SNR). Many optical imaging methods attempt to mitigate the effects of scattering through ballistic imaging, filtering unscattered photons from heavily scattered photons. A number of ballistic imaging techniques have been developed, including spatial filtering (i.e., confocal microscopy), polarimetry, and time-of-flight (ToF) imaging.14 Unfortunately, the number of ballistic photons decreases exponentially with optical thickness. For example, optical imaging through the human skull requires imaging through 6.5  mm of tissue with scattering coefficient 9  mm1. This corresponds to 60 mean free paths, which indicates on average a photon will undergo 60 scattering events while propagating through a scatterer of this thickness. Therefore, only one ballistic photon would remain for every e60=1.14*1026 photons from the light source. As a result, most ballistic imaging methods fail to image beyond a few millimeters of tissue. The limited depth of ballistic imaging has led to great interest in diffuse optical tomography, a technique that uses ballistic and diffuse photons to image through densely scattering media.

In diffuse optical tomography (DOT), an array of light sources and detectors, which are placed on the subject, capture a set of optical measurements. An associated reconstruction algorithm converts these measurements into a high resolution, human-interpretable images. To establish this reconstruction algorithm, DOT uses the predictable relationship between spatial variations in the optical properties and the spatial-angular distribution of light, the radiance.5,6 This relationship can be represented as a function known as the forward model, which maps the spatial distribution of optical properties to the radiance. To perform image reconstruction, this forward model must be inverted, establishing a mapping from the radiance to the spatial distribution of optical properties. This is generally referred to as the inverse problem. It is widely accepted that an accurate forward model and inverse solver are essential to high accuracy DOT image reconstruction.7 Accurate forward models and inverse solvers were essential in several recent papers that achieved high resolution reconstructions through densely scattering media.5,811

Despite the recent progress in optimizing the DOT inverse solver, it still possesses several limitations. First, inverse solvers can be slow. Non-linear algorithms may require several hours to complete.12 Even linear inverse algorithms, which are regarded as the fastest solvers, may require minutes to reconstruct a single image.9 This precludes the possibility of real-time imaging of biological events, which may require subsecond timing resolution. Second, many inverse solvers require several hyperparameters, which are hand-tuned. Several prior works use optimization-based inverse solvers that include regularization terms, such as Tikhonov or L1 norm regularization. Scaling the weights of these regularization terms can impact the resulting image reconstruction. For example, increasing L1 norm regularization increases image sparsity. This can improve the image reconstruction quality. However, since the parameters are hand-tuned, there is no guarantee that the final result achieves the highest fidelity to the original image. Third, the image reconstruction quality is highly sensitive to the system calibration. Reconstruction algorithms require knowledge of the system parameters, i.e., the source and detector positions, the target geometry, and the camera transform. While these parameters can sometimes be estimated through a calibration procedure, this can be difficult for dynamic or living targets. This can lead to model mismatch in estimating the background measurement13 or the mesh geometry14 and can lead to inaccurate reconstructions of the target’s optical properties or position.

To address these challenges, we propose to use algorithm unrolling as the DOT inverse solver, a technique we have dubbed “Unrolled-DOT.” In algorithm unrolling, the iterations of a standard linear inverse solver are converted into fully connected (FC) layers of a neural network.15,16 This has been shown to improve the image reconstruction quality and runtime of the solver (Fig. 1). As shown by Lecun et al.,16 and as we will demonstrate in Sec. 6, the unrolled algorithm can achieve a significantly faster algorithm runtime than traditional physics-based solvers. In addition to improving the runtime, Unrolled-DOT also improves on traditional physics-based solvers since it directly learns the hyperparameters and calibration from data. In addition to outperforming physics-based solvers, there are also benefits to using the unrolled algorithm versus other learning-based models: the unrolled algorithm is interpretable, and can be trained with fewer parameters than traditional networks.15 While many traditional deep learning models are considered “black-box models,” the learned parameters of the unrolled-algorithm correspond to physically meaningful values. For example, the learned activation parameter in the unrolled network corresponds to sparsity regularization. In addition, traditional methods may possess tens of millions of parameters and require millions of examples for training.17,18 We will demonstrate that Unrolled-DOT can be trained with just a few thousand examples.

Fig. 1

Comparison of unrolled algorithms versus standard iterative linear solvers: (a) classical inverse solvers, such as ISTA perform reconstruction by iteratively applying a gradient step followed by the proximal operator, i.e., the shrinkage function with predetermined parameter λ. (b) This iterative procedure can be unrolled. The gradient steps become FC layers of a neural network. The non-differentiable operation becomes an activation function, such as a shrinkage function with learnable parameter θ. Because the forward model is learned, unrolled algorithms can reduce the number of layers and perform faster reconstruction compared to linear solvers. The learned model can also reduce model mismatch and increase the image reconstruction quality.

JBO_28_3_036002_f001.png

Thus, we report the following contributions.

  • 1. Our algorithm achieves higher image reconstruction quality compared to both physics-based and learning-based solvers for DOT/ToF-DOT image reconstructions. Our method reduces the mean-squared error (MSE) by over an order of magnitude on real-world data, compared with physics-based linear solvers.

  • 2. Our method is faster than traditional linear DOT solvers, achieving over an order of magnitude reduction in the runtime compared to all other physics-based methods in our comparison. High-quality image reconstruction can be achieved with just one layer, approximately corresponding to a single iteration of a standard linear solver. In contrast, other linear inverse solvers may require tens of iterations.

  • 3. We collected the first real-world DOT/ToF-DOT dataset with thousands of measurement and ground truth pairs (5000 total), which we made available to the public.19 In generating DOT tissue phantoms, it is difficult to precisely place the absorbing targets and even harder to achieve on a large scale. Our method uses an Eink display to programmatically control the target shape and placement, which provides a more accurate ground truth. This is important for supervised learning algorithms that require measurement and ground-truth pairs. In addition, these supervised models typically require thousands of training examples for DOT reconstruction,2022 which could only be achieved with synthetic datasets in prior works. In the future, our real-world dataset could be used to train and benchmark models on data that inherently possesses the features (noise, contrast, etc.) of real-world measurements. This can lead to more accurate benchmarks and trained algorithms that better account for model mismatch with real-world measurements.

While our approach makes several improvements over existing methods, we also acknowledge that there are certain limitations that can be addressed in future work. First, our prototype uses a single laser source and single-pixel detector, which are raster-scanned by galvo mirrors. In the future, an imaging system that possesses an on-chip array of sources and detectors would improve the measurement acquisition speed. In addition, the measurements are currently captured for only two-dimensional imaging targets and can be extended to three-dimensional targets. In future work, it may be possible to extend to other imaging modalities through transfer learning or by learning only the parameters associated with the physical apparatus. These learned parameters can be treated as a calibration. To perform image reconstruction, the physics-based models can still be applied on top of this calibration. Finally, applying these algorithms to in-vivo targets is reserved as future work.

2.

Related Works

2.1.

DOT/ToF-DOT

There is a long history of imaging through scattering media using techniques such as near-infrared spectroscopy (NIRS) and diffuse optical tomography (DOT). Early works in NIR spectroscopy include monitoring of blood oxygenation levels in the brain, which began in the 1970s.23 It was later discovered that neural activity was correlated to blood oxygenation levels, and could be measured using NIRS, which led to the development of functional near infrared spectroscopy in the 1990’s.24 In the present day, DOT has been applied to many applications in biomedical imaging, including functional neuroimaging, breast tumor detection, and small animal imaging.2527 DOT techniques can be generally categorized into three modalities: continuous-wave (CW), frequency domain (FD), and time domain (TD), which is also referred to as ToF.

In CW-DOT, each recorded measurement between a source and detector is a scalar intensity value. Due to the simplicity of the associated hardware, this technology is the most cost-efficient and can be easily integrated into wearable devices.28 CW-DOT has been shown to image up to centimeters deep into tissue, but generally achieves poorer spatial resolution on the order of centimeters.6 Recent progress with high-density DOT, which uses a dense array of sources and detectors, has shown improvement in spatial resolution.25 Another DOT imaging modality is FD diffuse optical tomography (FD-DOT). FD-DOT uses a frequency-modulated optical signal for the light source. The optical properties of the target change the amplitude and phase of the resulting measurements, which can improve the image reconstruction quality using greater measurement diversity.29

Finally, the third DOT imaging modality is TD or ToF diffuse optical tomography (ToF-DOT), which enhances the imaging quality using TD measurements on the order of picoseconds. While the theory for ToF-DOT has existed since the 1990s, development of the technology was limited by hardware performance and cost.30 In recent years, ToF-DOT imaging systems have risen in popularity as the technology for TD detectors has matured. Lyons et al.9 demonstrated that a transmission-geometry ToF-DOT system could achieve approximately millimeter-scale spatial resolution through a 2.5-cm-thick scatterer. Unfortunately, ToF-DOT systems were still very slow due to the large volume of spatial-temporal measurements that had to be processed. Thus, confocal scanning geometries (CToF-DOT) have been applied to ToF-DOT, reducing the number of measurements that are required while maintaining a high spatial resolution.5,10

2.2.

Physics-based DOT Inverse Solvers

While many imaging modalities exist, all DOT techniques require a forward model of light propagation as well as an inverse solver to recover a spatial distribution of optical properties from optical measurements. Both the forward model and inverse solver have an enormous impact on the image reconstruction quality and the overall runtime. Most algorithms exhibit a tradeoff between the reconstruction quality and runtime. Therefore, the forward and inverse algorithm selection is critical. The most accurate method of light propagation is generally a full solution to the radiative transfer equation using the Monte Carlo algorithm.6,3133 While this approach is accurate, it can also be quite slow. Solutions can also often be estimated using mesh-based methods, such as finite-element method (FEM) or boundary element method,3436 and finite volume methods (FVM).12 Faster but more inaccurate forward models using the diffusion approximation37,38 or truncated spherical harmonics approximations3941 have also been proposed. While physics-based methods have allowed DOT to achieve high resolution imaging through densely scattering media, we believe they still possess certain limitations, including a tradeoff between speed and reconstruction quality, a need for hyperparameter tuning,5 and a need for system calibration.13,14 We believe that a learning-based method may help to mitigate these challenges.

2.3.

Data-driven Models

While a lot of efforts have been placed on physics-based approaches, there has been a rise in the popularity of machine-learning methods. Table 1 summarizes these recent works. Yedder et al. was one of the first to use a deep neural network as the DOT inverse solver.20 Subsequent work by Sabir et al.21 and Yoo et al.22 also applied neural networks, more specifically convolutional neural network architectures, for solving the DOT inverse problem. Nizam et al. and Deng et al. have demonstrated progress in improving CW diffuse optical tomography reconstruction quality using a variation of the automated transform by manifold approximation (AUTOMAP) architecture.4345 Li et al.46 has applied deep learning and DOT to clinical applications, improving the accuracy of breast tumor imaging. Deep learning has also been applied for generating data in DOT applications.47 Recently, Zou et al.42 used an autoencoder to learn the forward model and inverse problem of FD-DOT. While their training data included experimental data, it was limited to only 30 unique configurations. Our proposed approach aims to improve on prior work by constructing a neural network that is more interpretable and is trained on a real-world dataset with thousands of examples to account for model mismatch.

Table 1

Comparison of DOT dataset size in recent works: Above, we summarize the important attributes of datasets from recent works on machine learning and DOT. For the dataset size, we break down whether the dataset was obtained through simulated or real-world data. In addition, we note the size of the data for training and testing. Most data-driven DOT algorithms are trained on large simulated datasets or a limited number of real-world measurements. Our method is the first to use a dataset of thousands of real-world measurements.

DatasetDOT modality (CW, FD, ToF)Real-world training data? (Y/N)Simulated dataset size (number of Measurements)Real dataset size (number of measurements)
Yedder et al.20CWN450032 (testing only)
Sabir et al.21FDN65002 (testing only)
Yoo et al.22FDN15003 (testing only)
Zou et al.42FDYNA (trained on real-world data)114 (24 inhomogeneous/ 90 homogeneous phantoms)
Unrolled (ours)CW/ToFYNA (trained on real-world data)5000(trainingand testing)

The techniques described above all applied machine learning to improve the inverse solver for recovering spatial distributions of optical properties. There are also many other works that apply machine learning to other modalities of optical imaging,48,49 for example fluorescence molecular tomography,50,51 fluorescence lifetime imaging,5254 and direct optical imaging without an inverse solver.55 In addition, machine learning has also been used in hybrid imaging modalities. For example, deep learning has improved the reconstruction quality of magnetic resonance imaging (MRI)-guided DOT/NIRS in the applications of breast tumor detection56 and cerebral subdural hematoma monitoring;57 however, because these methods also rely on MRI, they may lack the cost-efficiency and portability of an exclusively optical imaging approach.

2.4.

Algorithm Unrolling

Neural networks are commonly treated as black boxes, hindering interpretability. This lack of explainability has limited their adoption in several domains, especially for medical imaging, in which the interpretability of results is essential. To increase the explainability of these networks, a new technique known as algorithm unrolling has risen in popularity.15 Gregor and Lecun were the first to propose this type of neural network architecture, an algorithm that they coined “learned iterative shrinkage thresholding algorithm” (LISTA).16 The key idea of this approach is to represent the iterations of the ISTA as layers of a neural network. This model is more interpretable than a standard neural network because the learnable parameters represent the forward matrix of the underlying physical model. This work has led to the “unrolling” of other algorithms, such as the alternating direction method of multipliers (ADMMs)-CSNET algorithm, and has been used in several imaging applications, such as lensless imaging and MRI.58,59 Recent works have also tested unrolled networks that do not train the network weights. Liu et al.60 introduced a network with analytically determined weights and learnable stepsizes and thresholds. In the application of diffuse optical tomography, Hua et al.61 used an unrolled network for fluorescence diffuse optical tomography. However, their method learns only the regularizer term rather than the entire forward model, and was trained only on simulated data, which may lead to model mismatch.

3.

Physics-Based DOT Forward Model

The key to image reconstruction in DOT is an accurate model of light propagation, generally referred to as the forward model. The forward model establishes the relationship between the spatial distribution of optical parameters, i.e., the target image that we ultimately want to solve for, and the set of recorded measurements. Because of this relationship, the accuracy of the forward model is essential to obtaining high-quality, accurate image reconstructions. For example, if the forward model assumes an inaccurate position for the detector, this can lead to misplacement of the targets or other artifacts in the reconstructed image.

To establish an accurate forward model, we employ the radiative transfer equation (RTE). The RTE, and consequently most DOT forward models, uses the particle model of light because the wave nature of light has negligible impact on heavily scattered light. While the RTE is generally considered the most accurate model of light propagation,39 it can be computationally expensive or even intractable to solve for an arbitrary target geometry and imaging setup.

We model the propagation of light using a linear forward model. This model assumes a linear relationship between small changes in the optical parameters (perturbations) and the optical measurements that we collect. This is valid under the Born approximation for small optical changes, relative to the background optical properties.31 The linear forward model can be expressed as

Eq. (1)

Δm=JΔμ.

Here, Δm, Δμ, and J represent the differential measurements, the target, and the Jacobian (or sensitivity) matrix, respectively. The target Δμ is modeled as an optical perturbation. For example, neural activity can lead to optical changes in the associated regions of the brain due to the hemodynamic response.25 The differential measurements Δm=mbkgmtarget correspond to the difference between the detected intensity in the unperturbed (mbkg) and perturbed (mtarget) state. The Jacobian matrix relates these two quantities, i.e., a change in the optical parameters (Δμ) corresponds to a change in the detected intensity (Δm). Thus, each entry of the Jacobian matrix is a partial derivative: Jrs=mrμs.

Under this linear setup, the Jacobian matrix is essential to both generating synthetic test data and image reconstruction with a traditional, physics-based forward model. Several methods exist for simulating the Jacobian matrix, including analytical expressions, FEM, and the Monte Carlo (MC) algorithm.31,37,62,63 We use MC in this work since it is regarded as the most accurate and flexible of these methods. The details of the implementation are described in Sec. 5; here, we will provide a brief description of the algorithm itself. MC uses randomly sampled photon trajectories to simulate the propagation of light.64 These photon trajectories are sampled from a distribution that incorporates the optical parameters of the target medium. By averaging the information over many samples, we obtain an accurate estimate of the distribution of light. MC is an unbiased estimator and can be easily adapted to arbitrary geometries; however, it may require many random samples and consequently a long run time to converge to a sufficiently low-variance estimate.32

While the RTE and MC can be used to model the underlying physics of the imaging system, they require knowledge of the imaging setup and target parameters. These parameters are provided as input to the algorithm. Some examples of input parameters include the source–detector positions which are generally a vector of coordinates and the target geometry which can be represented as a mesh. Certain parameters may be simple to estimate for a particular setup, such as the source–detector separation.65 Other parameters may be challenging to estimate through calibration, which can lead to model mismatch. For example, Reisman et al. showed that assuming an incorrect target geometry led to an incorrect depth estimation.14 To avoid potential reconstruction errors from inaccurate forward model parameters, we use the unrolled network to learn the system parameters. Thus, our training procedure can be interpreted as a calibration step. By training on real-world measurements and their associated ground truth images, our model reduces the potential for model mismatch and improves the image reconstruction quality.

4.

Unrolled-DOT

4.1.

Unrolled Network Design

Our ultimate goal is to solve the inverse problem: determine the target parameters from a given set of measurements. Using the linear forward model described in Eq. (1), the inverse can be formulated as a convex optimization problem

Eq. (2)

Δμ*=minΔμJΔμΔm2+Λ(Δμ).

Here, as in Sec. 3, the terms Δμ, J, Δm represent the target, Jacobian matrix, and observed differential measurements, respectively. (·)* denotes the optimum. This equation states that the desired reconstruction (Δμ*) minimizes the MSE between the predicted and the observed measurements (JΔμΔm2) and includes a penalty term (Λ(Δμ)) that enforces prior information regarding the target. For example, Λ(Δμ)=λ1|Δμ|1 would enforce an image sparsity prior. A suite of techniques exist for solving linear inverse problems that follow the form of Eq. (2). This includes the ISTA, fast ISTA (FISTA) an accelerated version of ISTA, and the ADMM.66,67 We will focus on ISTA because this algorithm forms the foundation of our proposed method. However, our results will include comparisons with these other standard solvers.

ISTA can be considered as a variation of gradient descent. As stated before, the problem setup generally follows a similar structure to Eq. (2), where Δμ in our notation is the standard unknown variable, and J is the standard forward matrix (often referred to as A). Using ISTA, the solution is obtained by iteratively performing the gradient step and applying a proximal operator for the non-differentiable portion of the objective. Mathematically, a single step of this procedure can be expressed as66

Eq. (3)

Δμk+1=h(Δμkγ(JTJΔμkJTΔm);λ),
where, γ(JTJΔμkJTΔm) is the gradient step with step-size γ, and h(·,λ) is the proximal operator. Generally, the gradient step solves the linear inverse problem, while the proximal operator is used to enforce regularization constraints. For example, an L1 norm penalty can be used to enforce image sparsity in the reconstruction. In this case, the proximal operator is the shrinkage function with tuning parameter λ. An overview of this procedure is also shown in Fig. 1(a). While this procedure is simple and sufficiently general that it can account for even non-differentiable terms in the objective function, there are also several challenges to using this approach, particularly for DOT image reconstruction.

First, the ISTA algorithm possesses relatively slow convergence. Even for algorithms such as FISTA, which accelerates the convergence of ISTA, tens of iterations may be needed to converge to a sufficiently accurate reconstruction. For this reason, unrolled algorithms have been proposed with the potential to increase the reconstruction speed. This gain was demonstrated by Gregor et al.16 and is shown empirically in our results in Sec. 6. In addition to improvements in efficiency, there are also potential benefits to the image reconstruction quality. Any misalignment between the expected forward model and the underlying system that generates the measurements can reduce the efficacy of standard linear solvers. To address this, we propose a deep network to learn the parameters of the standard inverse solver. More specifically, we apply the learned-ISTA algorithm.16

The learned algorithm follows a very similar structure to ISTA. Each “layer” of the algorithm corresponds to an iteration of ISTA. The i’th layer can be represented mathematically as the function16

Eq. (4)

Δμi+1=fi(Δμi)=h(SΔμi+WΔm;λ).

Here, W, S, and λ correspond to learnable parameters of the network. h(·;λ) is the non-linearity applied after each layer of the network and corresponds to the shrinkage function used in ISTA. Using this function, we can express the full network with L-layers as

Eq. (5)

FL=fLfL1,,f0.

Here f is the function: fi(y)=h(Sfi1(y)+Wy;λ), and f0(y)=h(Wy) is the “0-th” layer. An overview of this full pipeline is shown in Fig. 2(b). Next, we can show the relationship between the ISTA and the unrolled network. If we set W=γJT and S=IγJTJ, we return to the original ISTA expression. In this case, the layers of the unrolled-network would be equivalent to the iterations of ISTA, as represented in Eq. (3) and stated in Gregor et al.16 One of the benefits of this type of “unrolling” algorithm is interpretability: the learned parameters provide insight into the physics of the underlying system because the layers correspond to iterations of a standard linear solver. Additional details about this are provided in Sec. 6.

Fig. 2

Unrolled-DOT network architecture: (a) the full algorithm pipeline is shown in the grey box. The set of measurements is first passed through an unrolled network to perform image reconstruction. Additional details are shown in panel (b); a key of relevant terms is shown in panel (c). The unrolled network uses a shared FC layer and K independent FC layers. Additionally, we use the shrinkage function as the non-linear activation function. The shrinkage functions are applied after each layer. After the unrolled network, the output is passed through a U-Net, which further improves the image reconstruction quality. The loss is a weighted sum of a standard loss (MSE or MAE) and the Visual Geometry Group (VGG) perceptual loss with weight parameter β. Backpropagation is then applied to train the unrolled network and U-Net.

JBO_28_3_036002_f002.png

Training is performed using the standard backpropagation approach, with either synthetic or experimentally collected data

Eq. (6)

W*,S*,λ*=minW,S,λF(Δm;W,S,λ)Δμ2.

As before, W, S, and λ are parameters of the learned network, F. Here, we use MSE as the loss function to determine optimality, though other metrics can also be chosen. The measurements Δm are generated either synthetically by applying the physics-based forward model or captured experimentally using a prototype system. For training, ground truth data is accessible both in simulation and in experiments. In simulation, the ground truth and forward model directly compute the measurements. In experiments, we display the ground truth on an electronic ink (E-ink) display, which can programmatically control the spatial distribution of optical absorbers and scatterers. We collect measurements corresponding to the ground truth target with the experimental prototype. Additional details about our full training and testing procedure are described in Sec. 5.

4.2.

Additional Perceptual Enhancement

While the unrolled network offers a powerful learning-based model for solving linear inverse problems, the quality of our image reconstructions can still be further improved. To this end, we also incorporated a refinement U-Net and VGG perceptual loss in our network design and training procedure. In adding these steps, we were inspired by Khan et al.,68 who demonstrated that the U-Net and VGG-loss could enhance image reconstruction quality in the application of lensless imaging.

The U-Net architecture is a convolutional neural network that uses a contracting and expanding path. This network architecture was originally designed by Ronneberger et al. for the application of image segmentation.69 It was subsequently shown that the U-Net architecture could also be applied to image reconstruction.70 In addition to the refinement U-Net, we also apply a VGG perceptual loss to further enhance the image reconstruction quality.71,72 The resulting loss term can be expressed as

Eq. (7)

W*,S*,λ*,θ*=minW,S,λ,θUθ(F(Δm;W,S,λ))Δμ2+fVGG(Uθ(F(Δm;W,S,λ)))fVGG(Δμ)2.

For the VGG-loss, the ground truth image and the reconstructed image, i.e., the output of the refinement U-Net, are passed through a pretrained VGG network. The VGG loss is then calculated from the standard (MSE or MAE) loss between intermediate outputs of the VGG network for the predicted and true images. The overall loss then becomes the weighted sum of the standard loss and the VGG loss, as shown in Eq. (7). Here, Uθ(·) refers to the U-Net with trainable parameters θ, and fVGG(·) is a function that maps the input to the intermediary outputs of the VGG network. Because the VGG network is pretrained, its parameters are not updated during the Unrolled-DOT training. Finally, this overall loss is backpropagated to train the unrolled network and U-Net.

5.

Materials and Methods

5.1.

Computational Processing

The computational processing in these experiments includes the generation of synthetic data, machine learning for training and testing our models, and computational post-processing, which includes testing other standard linear solvers. The model training and testing were implemented in Python using the Pytorch library. In addition, the structural similarity (SSIM) loss in Sec. 6.3 used the Kornia library.73 MATLAB was generally used for post-processing, such as plotting and visualization, as well as implementing the FISTA, and ADMM solvers for the algorithm comparison in Sec. 6.4.

The synthetic data was generated using an in-house Monte Carlo simulator, which implemented the Monte Carlo multi-layered (MCML) algorithm.64 The standard MNIST and Fashion MNIST images were scaled to 41×41  pixels, where each pixel corresponds to 1 mm in world coordinates. The pixel values of the images were converted to absorption coefficients, with the maximum pixel intensity scaling to μa=1.0  mm1, and negligible background absorption. In addition, the background scattering coefficient was set to μs=9.0  mm1. For the simulated experiments testing the generalizability of the model and the comparison of reconstruction with and without the U-Net and VGG loss a 10×10 array of sources and detectors is used. For each source–detector pair, a TD measurement consisting of four timebins is simulated. Measurements are normalized to 1.0. For the circular phantoms experiments in the supplementary material, a circular grid of 32 source–detector pairs are distributed uniformly over the circle. Again, each time-domain measurement consists of 4 timebins. We used a precomputed Jacobian matrix to apply the forward model on the ground truth and generate the full synthetic dataset. Jacobian matrix construction is described in reference.31 In the experimental reconstruction and simulated generalizability experiments, a non-negativity threshold and normalization between 0 and 1 was applied to the reconstructed images. The synthetic dataset consisted of 3000 training images and 100 testing images from the MNIST and fashion-MNIST datasets.

In our experiments, we tested the unrolled network with and without the refinement U-Net and VGG-loss. When testing only the unrolled network, the batch size was equal to the training set size; however, when the U-Net and VGG-loss are applied, a smaller batch size is used to meet the greater memory demands and the GPU memory limitations. In addition, to meet the input size requirement of the U-Net and VGG-Net, the images were zero padded to increase the size from 41×41 to 64×64. The U-Net consisted of four downsampling levels (i.e. images were downsampled by a factor of 24=16) followed by four upsampling levels to bring images back to size 64×64. Finally, the weight factor for the VGG perceptual loss was set to 0.01 for all experiments.

Computational processing used a combination of both CPU and GPU hardware. Highly parallelizable processes, including the model training, applying the learned forward model, and generating synthetic data with Monte Carlo, were performed on Nvidia 1080/2080 Ti’s. Other post-processing was performed on a standard desktop, running on an Intel Xeon CPU.

5.2.

Real-World Data Capture Procedure

Figure 3 shows a rendering of the prototype that was used to capture experimental data. Our setup consisted of a light source, detector, and target. For the light source, we used a pulsed laser (the NKT supercontinuum laser) set to 670 to 680 nm wavelength. Our detector is the MPD FastGatedSPAD, which is a single-pixel single-photon avalanche diode (SPAD). The detector also uses an associated MPD picosecond delayer and a Picoquant Hydraharp to perform the time-correlated single-photon counting.

Fig. 3

Unrolled-DOT experimental prototype setup: a rendering of our prototype system is shown above. Images displayed on an E-ink monitor served as the targets. A tissue phantom was placed in front of the target to mimic the effects of biological scattering. Our imaging setup consisted of a laser and a single pixel single photon avalanche diode (SPAD). The laser and virtual detector are raster-scanned over the surface of the tissue phantom by a pair of galvo mirrors.

JBO_28_3_036002_f003.png

Our target consisted of an E-ink display and an optically scattering tissue phantom. The E-ink display allows us to control the spatial distribution of light absorbers and scatterers programmatically.74 We used a Form 3 stereolithography three-dimensional (3D) printer to synthesize the optical tissue phantom. The optical properties can be controlled by mixing the Formlabs white and clear resins. By measuring the temporal point spread function of the transmitted light through the phantom, we determined the scattering coefficient of the tissue phantom to be 9  mm1, which is within the acceptable range of biological tissue.75 More details regarding the phantom synthesis and characterization procedure can be found in our supplementary materials and the following references.5,10,76,77

We captured a total of 5000 measurements, which we split into a training and testing set of 4500 and 500 measurements, respectively. Ground truth targets consisted of examples from MATLAB’s digits dataset, which we then programmatically projected onto the E-ink display. The measurements consisted of 5×5 source and 5×5 detector scan points. The minimum source–detector separation was 4 mm, and the inter-source separation was 8 mm. In the raw data, each transient was initially binned into 4000 timebins, which were each 1-ps in duration. In post-processing, the data were further binned into 400 timebins, which were each 10 ps in duration. For all TD measurements, we applied temporal filtering, specifically the Laplace-temporal-filter78 with s=104. This helped reduce the amount of data that needed to be processed and improved the image reconstruction quality. Because the temporal-filtering reduces the transient to a scalar value, the input is a B×N matrix, where B is the batch size and N is the number of source–detector pairs. For generating CW-DOT measurements, these time-domain transients can be integrated into a scalar. Using this system, we captured, to the best of our knowledge, the first dataset of DOT/ToF-DOT measurements with thousands of examples for training supervised machine learning models.

6.

Results and Discussion

In this section, we report the results of our experiments, which include tests on both synthetic and real-world data. We began by testing the image reconstruction algorithm on a simulated dataset to show the benefits of applying a refinement U-Net and VGG-loss, and to show the algorithm’s generalizability. We also applied Unrolled-DOT to experimentally collected (real-world) data. Our first experiment on the real-world data is to verify the validity of our learned model itself. To do this, we visualize both the output feature maps of the pretrained VGG-loss as well as the learned model. In addition, we also demonstrate that our model is able to account for model mismatch by testing how the network performs when initialized with a misaligned source–detector array. After verifying our results, we conduct a performance analysis. The first test in this performance analysis is to compare the results of training the unrolled network and U-Net end-to-end versus training the U-Net after training the unrolled-network (two-step training). Next we conduct an ablation study, in which we test how the model performs when various training parameters are varied including the loss function, the number of training iterations, and number of layers. Finally, we compare our approach to traditional physics-based linear solvers as well as learning-based inverse solvers. We conducted this test on real-world data and compared both a standard ToF-DOT arrangement (measurements from all source–detector pairs) as well as a CToF-DOT.

6.1.

Image Reconstruction on Synthetic Data

6.1.1.

Comparing performance with and without U-Net+VGG on simulated data

We began by testing the performance of our model on synthetic data. We first demonstrated the benefits of adding a refinement U-Net and VGG perceptual loss to the unrolled network. Here, we trained solely on images from the fashion MNIST dataset. In both cases, a three-layer unrolled network was trained with a learning rate of LR=104 and MSE-loss. When training just the unrolled network a batch size equal to the training set (3000 measurements) was trained for 2000 iterations. With the U-Net and VGG-loss, the model was trained for 200 iterations with a batch size of 500 measurements. In Fig. 4, we see that adding the U-Net and VGG-loss to the network architecture dramatically improves the visual quality of the reconstructed image. The benefits of adding these components are further confirmed by quantitative metrics of the image reconstruction quality: the MSE and correlation factor. By incorporating the U-Net and VGG-loss, the MSE decreased from 6.40*102 to 1.50*102 while the correlation coefficient increased from 0.897 to 0.961. We have also shown that using Unrolled-DOT, it is possible to reconstruct more complex images that have varying contrast levels rather than sparse, binary images that are typically used as targets.

Fig. 4

Improvement in image reconstruction quality using U-Net and VGG perceptual loss: We trained and tested our Unrolled-DOT algorithm on the fashion MNIST dataset. We compared the reconstruction quality with and without the refinement U-Net and VGG perceptual loss. We can see that incorporating the U-Net and VGG loss dramatically improves the reconstruction quality, based on both visual inspection and the average MSE across all test images. We have also shown that with these two enhancements, the reconstruction algorithm can accurately reconstruct images with varying levels of contrast.

JBO_28_3_036002_f004.png

6.1.2.

Generalizability test

Next, we also demonstrated the generalizability of Unrolled-DOT. We trained two separate models on structurally different images: the MNIST and fashion-MNIST datasets, respectively. We then performed image reconstruction on measurements generated from MNIST dataset images. In this case, we tested only the unrolled-network without the U-Net and VGG-loss. We see that both models achieve good reconstruction of the test images, though model A does achieve slightly higher reconstruction quality than model B (an increase in the correlation coefficient from 0.94 to 0.97) because it is trained and tested on images of the same class (Fig. 5).

Fig. 5

Generalizability of the Unrolled-DOT model. We trained two models: one on the MNIST dataset (model A) and another on the fashion MNIST dataset (model B). Both models performed similarity in reconstructing test images on the MNIST dataset, demonstrating the generalizability of our learning approach.

JBO_28_3_036002_f005.png

6.2.

Interpretability of the Trained Model

6.2.1.

Visualization of the VGG-loss features

In addition to a standard training loss, the Unrolled-DOT algorithm also uses a VGG-loss. To calculate the VGG-loss, the ground truth training images and reconstructed images are passed through a pretrained VGG network. The VGG-loss is then determined by the MSE/MAE error between the VGG network outputs for the ground truth and reconstructed images. Examples of feature maps that are generated by the VGG network are visualized in Fig. 6. The outputs used in the VGG loss are the ReLU 2-2 and the ReLU 4-3 features, which are the network outputs between the second convolution and second max pool and the fourth convolution and third maxpool, respectively.68,72 Johnson et al.72 observed that the ReLU 2-2 features corresponded to the low-level features of the image, while the ReLU 4-3 features appeared to correspond to the higher-level structures of the image. Therefore, using both features for the VGG-loss aids in preserving the fidelity of both low-level features and overall structure. We see in Fig. 6 that ReLU 2-2 seems to correspond to features of the input image, such as edges. While the ReLU 4-3 feature maps are very heavily downsampled, we see that for both sets of VGG outputs, there is good agreement between the ground truth and reconstructed images.

Fig. 6

Visualization of VGG-network feature maps: The VGG loss uses the ReLU 2-2 (output of the ReLU between second convolution and second max pool) and ReLU 4-3 (output of the ReLU between fourth convolution and third max pool).68 Johnson et al.72 have noted that ReLU 2-2 corresponds to low-level features while ReLU 4-3 corresponds to higher-level structures. Thus, these two features were chosen for the VGG loss to encourage a better fit for both low-level features and overall structure.

JBO_28_3_036002_f006.png

6.2.2.

Visualization of the learned Jacobian

The layers of our learned network correspond to the iterations of a standard linear solver. As a result, the learnable parameters can provide insight into the underlying physical parameters of the system. Recall from Sec. 4 that the learnable parameter W corresponds to the adjoint of the Jacobian matrix, i.e., W=γJT in the unlearned algorithm. Therefore, by visualizing the learned W parameter, we can analyze the learned sensitivity matrix.

In Fig. 7, we show a visualization of rows of the physics-based Jacobian matrix, and the learned sensitivity profile W, which was trained on real-world measurements. Because the Jacobian matrix describes the sensitivity of the measurements to small perturbations in the optical parameters, it can also be interpreted as the likelihood that a photon will pass through a particular point in space. In this visualization, we see the sensitivity profile as a function of (x,y) spatial location, for a fixed depth z. The regions of highest sensitivity, i.e., the bright yellow regions, correspond to the position of the light source/detector. This follows from the intuition that the sensitivity profile describes the likelihood of photons passing through a point in space: regions closer to the source/detector will have a higher likelihood of a photon passing through and therefore a greater sensitivity.

Fig. 7

System parameter estimation using learned model. (a) In traditional DOT solvers, the forward model is assumed and must be passed in as an algorithm input. (b) With Unrolled-DOT, the forward model is estimated from data. Visualizations for rows of the physics-based and learned-forward models are shown on the bottom row.

JBO_28_3_036002_f007.png

The interpretability of the learned model allows for greater insight into the underlying system. In future work, the learned Jacobian could be used to estimate parameters of the underlying physical model. This could be useful for calibrating the source–detector positions, or estimating the optical parameters.

6.2.3.

Mitigating model mismatch with Unrolled-DOT

We also wish to demonstrate that Unrolled-DOT is able to mitigate certain cases of model mismatch through training. To demonstrate this, we use an incorrect forward matrix to initialize the Unrolled-DOT model, as shown in Fig. 8. In Fig. 8(a), we visualize the layers of the unrolled network. More specifically, we visualize rows of the matrix, which correspond to the network layer. Recall that the layers of the Unrolled-DOT model correspond to the Jacobian, or sensitivity matrix. The rows of this matrix should correspond to the sensitivity for a specific source–detector pair. Therefore, we should expect to see highest sensitivity (bright yellow regions) at locations corresponding to the source–detector position. The true positions of the sources and detectors are overlaid on the visualization of the network layers as a red circle and blue square, respectively. Initially, we see that the regions of high sensitivity do not correlate well with the source–detector positions. However, after 2000 iterations of training, we see that the regions of highest sensitivity shift to the correct locations, corresponding to the source–detector positions. As expected, in Fig. 8(b), we see that mitigating this model mismatch also improves the image reconstruction quality. The average MSE is reduced from 5.6 to 0.020 and the average correlation coefficient is increased from 0.29 to 0.88.

Fig. 8

Overcoming model mismatch using Unrolled-DOT: the model was initialized with a weight matrix, which was generated with incorrect source–detector positions. Panel (a) shows a visualization of the rows of the forward matrix at different stages of training. The regions of highest sensitivity (bright yellow region) correspond approximately to the position of the source–detector pair. We see that at 0 iterations, the regions of highest sensitivity do not correspond well to the true source–detector positions. After 2000 iterations of training, we see there is good agreement between the source–detector positions and regions of high sensitivity. In addition, in panel (b), we see that training the network and reducing model mismatch also improves the image reconstruction quality.

JBO_28_3_036002_f008.png

6.3.

Performance Analysis

6.3.1.

End-to-end training versus two-step training

The Unrolled-DOT model consists of two networks: an unrolled network and a U-Net. These two networks can be trained end-to-end or in a two-step procedure, training the unrolled network first and subsequently the U-Net. The first test in our performance analysis is to compare the end-to-end training procedure with the two-step training procedure. The results of this experiment are shown in Fig. 9. The end-to-end training procedure achieves a small improvement compared to the two-step training procedure; however, the improvement is relatively small since the MSE is only reduced from 0.013 to 0.11. From the reconstructed images shown in Fig. 9, we also see that the two methods achieve comparable image reconstruction quality.

Fig. 9

Comparison of end-to-end training with two-step training procedure. The Unrolled-DOT algorithm uses two networks: an unrolled network and a U-Net. The two networks can be trained end-to-end or in two steps: training the unrolled network first, then the U-Net. Above, average reconstruction MSE and correlation coefficients are reported for the full test set. We see there is only a marginal reduction in the MSE from 0.013 with the two-step training to 0.011 with the end-to-end training and a 0.01 increase in the correlation coefficient. Visual inspection of the example reconstruction images (shown above) appears to confirm that both methods achieve similar reconstruction quality.

JBO_28_3_036002_f009.png

6.3.2.

Ablation study

In addition to performing image reconstruction, we also characterized the performance of our algorithm as a function of different training parameters. Our ablation study is conducted on real-world data. For simplicity, we tested only the unrolled network, without the U-Net or VGG-loss. In general, we trained a three-layer unrolled network for 2000 iterations with MAE loss, and a learning rate of LR=2*104. The parameter is varied if it is the dependent variable for a particular experiment. For example, in testing the impact of the number of network layers, the model is still trained for 2000 iterations with MAE loss, and a learning rate of LR=2*104; however, the number of layers is varied between 1 and 10.

In this analysis, we specifically studied the performance as we varied the loss function, number of training iterations, and the number of layers in our network (Fig. 10). First, we compared three different loss functions. The loss functions we chose were based on the standard metrics used to judge image reconstruction quality: the MSE, the mean absolute error (MAE), and the SSIM loss.79,80 In comparing the loss functions, it appears the MAE produces the highest quality reconstructions, with less blurring and artifacts as compared to MSE and SSIM. It is interesting that the MAE (L1) objective achieves a lower overall MSE on the test set than the MSE (L2) objective; however, this phenomenon was also observed by Zhao et al.81 They postulated this is because MAE is able to avoid local minima better than MSE.

Fig. 10

Ablation study on objective function, number of training iterations, and number of network layers: we tested the image reconstruction quality as a function of several training parameters. (a) The MAE loss provides marginal improvements in the image reconstruction quality compared to SSIM and MSE. (b) In addition, with a reasonable learning rate, the reconstruction quality increases with more iterations in training. (c) Finally, we found the image reconstruction quality increases with the number of layers, though there seem to be diminishing returns beyond seven layers.

JBO_28_3_036002_f010.png

Second, we tested the number of iterations used to train the network. From this experiment, we found a fairly straightforward result: given an appropriate learning rate, the image reconstruction quality increased as the number of iterations increased. In Fig. 10, this is particularly noticeable in the reconstruction of the number “9,” in which we see that the blurring effect is significantly decreased as the number of iterations is increased.

Third, we tested how the number of layers in our learned model affects the reconstruction quality. We found that by increasing the number of layers, there is an increase in the image reconstruction quality, which is in agreement with the standard intuition and prior work.16 However, this improvement does seem to produce diminishing returns as the number of layers is increased beyond seven layers.

6.4.

Image Reconstruction on Experimentally Collected Data

We trained and tested our model on real-world data that we collected from our prototype system. The model was trained on a dataset of handwritten digits. Our model was tested on measurements from both a standard ToF-DOT setup (i.e., scanning all pairs of sources and detectors), as well as a CToF-DOT. Our Unrolled-DOT model was trained with a learning rate LR=2*104. Measurements are normalized to 5.0. With the U-Net and VGG-loss, the model was trained for 400 iterations with a batch size of 900 measurements. For training just the unrolled network, the model was trained for 2000 iterations with a batch size equal to the training set size (4500 measurements).

To show the benefits of our approach, we compared our method with two physics-based inverse solvers and two machine-learning-based inverse solvers. The physics-based solvers that we compare against are FISTA with L1 sparsity for CToF-DOT reconstruction, which was proposed by Zhao et al.,5 and ADMM with total-variation (TV) regularization for DOT reconstruction, which was proposed by Lu et al.82 We used in-house implementations for each of these inverse solvers and applied the Monte Carlo forward model. For FISTA, both the Jacobian and measurements are scaled to 1.0; while for ADMM the Jacobian and measurements are normalized to 2.1*104 and 6*105, respectively. Since the solvers are applied to a linear inverse problem, the normalization factors should not impact the final reconstruction as long as the non-linear hyperparameter terms are scaled appropriately. Additional hyperparameters for the physics-based solvers are listed in the supplementary material.

In addition, we also compare our method with two machine-learning-based methods: a standard FC multilayer perceptron, which will be referred to as the FC-net, and the AUTOMAP architecture.45 FC-net possesses an input layer, output layer, and a single hidden layer of width equal to the output size and uses the tanh activation function. FC-net is trained with MSE-loss. We also compare our method with AUTOMAP. Similar to FC-net, the first portion of the AUTOMAP network consists of FC layers: an input layer, and two hidden layers with width equal to the output size, and tanh activation functions. After the FC layers, the intermediate vector is reshaped into a two-dimensional image with dimensions equal to the output. The network then applies two convolutional layers with a kernel size of 5 and ReLU activation functions, followed by a deconvolution layer with a kernel size of 7. AUTOMAP is also trained with MSE-loss. More details about the AUTOMAP architecture are provided by Zhu et al.45

6.4.1.

Unrolled-DOT for confocal scanning geometry

We first tested Unrolled-DOT on a sparse set of measurements, collected from a 5×5 array of sources and detectors with a CToF-DOT. Therefore, measurements from only 25 source–detector pairs are used for image reconstructions. For this experiment, we used a three-layer unrolled network without the U-Net and VGG-loss and trained with MSE-loss. The results are shown in Fig. 11. From the reconstructed images in Fig. 11(a), we see that Unrolled-DOT achieves a higher image reconstruction quality compared with all other methods. This is particularly true compared to the physics-based solvers, which display severe reconstruction artifacts. In Fig. 11(b), we show a plot of the reconstruction quality metrics and reconstruction runtime. Each algorithm corresponds to a different symbol, shown on the key in the bottom right. The left y-axis (blue) corresponds to SSIM, while the right y-axis (red) corresponds to MSE. A higher SSIM and lower MSE correspond to improved image reconstruction quality, while lower runtime corresponds to a faster algorithm. Therefore, the “ideal” algorithm would be in the top left and bottom left for the blue and red plots, respectively. Based on the increase in the SSIM and the reduction in the MSE, shown in Fig. 11(b), we see that Unrolled-DOT outperforms all other algorithms in this comparison (exact SSIM, MSE, and runtime values are included in our supplementary material). Unrolled-DOT is also faster than all other algorithms except the FC-net; however, the increased speed of FC-net comes at the cost of reduced image reconstruction quality compared to Unrolled-DOT.

Fig. 11

Unrolled-DOT image reconstruction on experimental data with sparse 5×5 confocal scanning geometry: the reported metrics are averaged over the entire test set. In panel (b), we see there is a dramatic reduction in the MSE from 0.102 using FISTA and 0.218 with ADMM, compared to 0.042 with our approach. Similarly the SSIM increases from 0.20 and 0.36 for ADMM and FISTA, respectively to 0.72 with our approach. In addition, Unrolled-DOT achieves a faster runtime than AUTOMAP and an increased SSIM and reduced MSE compared to both AUTOMAP and the FC network. Visually, in panel (a), we also see from the reconstructed images that the Unrolled-DOT algorithm shows good agreement with the ground truth and removes artifacts from the sparse source–detector array.

JBO_28_3_036002_f011.png

6.4.2.

Unrolled-DOT for reconstruction of full ToF-DOT measurements

Next, we tested the Unrolled-DOT algorithm on a full set of ToF-DOT measurements, i.e., all source–detector pairs. The measurements were captured from a 5×5 source array and 5×5 detector array. Therefore, measurements were collected from 25 sources and 25 detectors, a total of 625 source–detector pairs. Here, we tested a 1-layer unrolled network with MAE loss. We tested the algorithm with and without the U-Net and VGG-loss. From the reconstructed images in Fig. 12, we see that our Unrolled-DOT method achieves a noticeably improved image reconstruction. In addition to visual inspection, we also see the performance of our method using quantitative metrics: the SSIM and MSE [shown in Fig. 12(b) and Table S2 in the Supplementary Material]. We see that, even with just a one-layer unrolled network, the SSIM increases from 0.18 to 0.60 with ADMM and FISTA, respectively, to 0.88 with Unrolled-DOT. In addition, when the U-Net and VGG-loss are added to our training pipeline, the image reconstruction quality further increases.

Fig. 12

Unrolled-DOT image reconstruction on experimental data from 625 source–detector pairs. We demonstrate the improvements both visually (a) and in the reconstruction metrics and runtime (b). Unrolled-DOT outperforms traditional physics-based linear solvers since it avoids imaging artifacts and poor reconstruction quality that results from model mismatch and sub-optimal hyperparameters. In addition, we see that the unrolled network, even without the U-Net and VGG loss, achieves better reconstruction quality than the fully connected network and comparable image reconstruction quality to AUTOMAP, but with significantly reduced runtime. With the U-Net and VGG loss, Unrolled-Dot achieves a significantly higher SSIM and lower MSE.

JBO_28_3_036002_f012.png

There are two primary reasons for our improved reconstruction quality. First, our trained method is able to learn the underlying physical model and can avoid potential model mismatch. Various parameters of the imaging system and target are automatically calibrated. This includes the source–detector positions and optical parameters of the target. Secondly, our trained model directly optimizes the hyperparameters for image reconstruction. In contrast, traditional linear solvers require several hand-tuned parameters as inputs. A full table of all relevant hyperparameters and the procedure for tuning these parameters is included in the supplementary materials, Table S1 in the Supplementary Material. At a minimum, this includes tuning the number of iterations and the weight of the regularization terms. For the FISTA algorithm, we incorporated L1 regularization to tune the image sparsity; for the ADMM algorithm, we incorporated TV-regularization. Since these hyperparameter values are generally hand-tuned, the performance of the corresponding inverse solver will depend on the tuning heuristics, which may not be optimal. As a result, the performance is improved using a learning-based algorithm that directly optimizes the hyperparameters.

In Fig. 12(b). we also report the runtimes to reconstruct the full test set, which consists of 500 images. This plot follows a similar format to Fig. 11(b). We see that the unrolled network is over an order of magnitude faster than other physics-based linear solvers. This speed-up was achieved with all algorithms in the comparison running on the same hardware: the Intel Xeon CPU. We will note that there were also implementation-specific details across the different methods. First, the learning-based methods were implemented with Python and Pytorch, while the other methods used MATLAB. In addition, the TV-regularization for the ADMM reconstruction used a computationally expensive matrix multiplication to calculate the image gradient. However, the speed-up from using our approach is not only demonstrated by the timing experiments. Unrolled-DOT requires just one layer to obtain the desired image reconstruction. FISTA and ADMM may require tens of iterations. Because Unrolled-DOT can achieve high-quality reconstruction with just one layer, and each layer performs a comparable number of operations to a single iteration of a standard solver, our method should be faster regardless of implementation-specific variations.

We also see that the refinement U-Net and VGG-perceptual loss further improve the image reconstruction quality. With these enhancements, the model seems to be better at rejecting spurious artifacts and increasing the sharpness and contrast. Quantitatively, we see in Table S2 in the Supplementary Material that this led to an increase in the SSIM from 0.88 to 0.92 and a decrease in the MSE from 1.67*102 to 1.14*102. Though the U-Net does slow down the algorithm, it is useful in applications that prioritize the reconstruction quality over the run time. Furthermore, the benefits of the U-Net and VGG-loss should be more distinct for image reconstruction on more complex natural images.

Finally, we see that our model performs well compared to other trained models. Without the U-Net and VGG-loss, Unrolled-DOT is faster than all other algorithms in our comparison, except the FC-Net. However, it is able to achieve an increased SSIM and reduced MSE compared to the FC-Net and comparable image reconstruction quality to AUTOMAP. With the addition of the U-Net and VGG-loss, the reconstruction metrics are further improved compared to AUTOMAP and FC-Net.

7.

Conclusion

We have proposed a learning-based method for solving the DOT inverse problem. Our method is able to obtain high-quality image reconstructions through densely scattering media, perform reconstructions efficiently, and can be trained end-to-end on learned data. Additionally, our network is interpretable, which may aid in system calibration. We demonstrated our results on both synthetic and real-world data, and prepared a dataset of 5000 ToF-DOT measurements. Ultimately, our learned model is a step towards fast high-resolution imaging through scattering media.

Disclosures

The authors have no relevant financial interests in the manuscript and no other potential conflicts of interest to disclose.

Acknowledgments

The authors would like to thank Yicheng Wu, Vivek Boominathan, and Salman Khan for their advice and insightful discussions in this project. This research was developed with funding from the Defense Advanced Research Projects Agency (DARPA) (Grant No. N66001-19-C-4020). The views, opinions and/or findings expressed are those of the authors and should not be interpreted as representing the official views or policies of the Department of Defense or the U.S. Government. In addition, this project was also funded in part by the NSF Expeditions in Computing (Grant No. 1730574). Author Yongyi Zhao was supported by a training fellowship from the NLM Training Program (Grant No. T15LM007093); author Fay Wang was supported by a National Science Foundation Graduate Research Fellowship (Grant No. DGE-2036197). In addition, this project was also funded in part by the NSF CAREER award (Grant No. 1652633).

Code, Data, and Materials Availability

Code is available at: https://github.com/yyiz/Unrolled-DOT. Our real-world DOT/ToF-DOT dataset is available at the following link: https://doi.org/10.5281/zenodo.7654959 (Ref. 19).

References

1. 

J. M. Schmitt, A. H. Gandjbakhche and R. F. Bonner, “Use of polarized light to discriminate short-path photons in a multiply scattering medium,” Appl. Opt., 31 6535 –6546 https://doi.org/10.1364/AO.31.006535 APOPAI 0003-6935 (1992). Google Scholar

2. 

M. G. Tanner et al., “Ballistic and snake photon imaging for locating optical endomicroscopy fibres,” Biomed. Opt. Express, 8 4077 –4095 https://doi.org/10.1364/BOE.8.004077 BOEICL 2156-7085 (2017). Google Scholar

3. 

H. Horinaka et al., “Extraction of quasi-straightforward-propagating photons from diffused light transmitting through a scattering medium by polarization modulation,” Opt. Lett., 20 1501 –1503 https://doi.org/10.1364/OL.20.001501 OPLEDP 0146-9592 (1995). Google Scholar

4. 

M. Minsky, “Microscopy apparatus,” U.S. Patent 3 013 467A (filed Nov. 1957, granted Dec. 1961).

5. 

Y. Zhao et al., “High resolution, deep imaging using confocal time-of-flight diffuse optical tomography,” IEEE Trans. Pattern Anal. Mach. Intell., 43 (7), 2206 –2219 https://doi.org/10.1109/TPAMI.2021.3075366 ITPIDJ 0162-8828 (2021). Google Scholar

6. 

D. A. Boas et al., “Imaging the body with diffuse optical tomography,” IEEE Signal Process. Mag., 18 (6), 57 –75 https://doi.org/10.1109/79.962278 ISPRE6 1053-5888 (2001). Google Scholar

7. 

M. A. Ansari et al., “Diffuse optical tomography: image reconstruction and verification,” J. Lasers Med. Sci., 5 (1), 13 –18 (2014). Google Scholar

8. 

H. K. Kim et al., “Ultrafast and ultrahigh-resolution diffuse optical tomography for brain imaging with sensitivity equation based noniterative sparse optical reconstruction (sensor),” J. Quant. Spectrosc. Radiat. Transf., 276 107939 https://doi.org/10.1016/j.jqsrt.2021.107939 JQSRAE 0022-4073 (2021). Google Scholar

9. 

A. Lyons et al., “Computational time-of-flight diffuse optical tomography,” Nat. Photonics, 13 (8), 575 –579 https://doi.org/10.1038/s41566-019-0439-x NPAHBY 1749-4885 (2019). Google Scholar

10. 

D. B. Lindell and G. Wetzstein, “Three-dimensional imaging through scattering media based on confocal diffuse tomography,” Nat. Commun., 11 4517 https://doi.org/10.1038/s41467-020-18346-3 NCAOBW 2041-1723 (2020). Google Scholar

11. 

G. Satat et al., “All photons imaging through volumetric scattering,” Sci. Rep., 6 33946 https://doi.org/10.1038/srep33946 SRCEC3 2045-2322 (2016). Google Scholar

12. 

H. K. Kim and A. H. Hielscher, “A PDE-constrained SQP algorithm for optical tomography based on the frequency-domain equation of radiative transfer,” Inverse Probl., 25 (1), 15010 https://doi.org/10.1088/0266-5611/25/1/015010 INPEEY 0266-5611 (2008). Google Scholar

13. 

Y. Ardeshirpour et al., “Artifact reduction method in ultrasound-guided diffuse optical tomography using exogenous contrast agents,” J. Biomed. Opt., 16 (4), 046015 https://doi.org/10.1117/1.3569088 JBOPFO 1083-3668 (2011). Google Scholar

14. 

M. D. Reisman et al., “Structured illumination diffuse optical tomography for noninvasive functional neuroimaging in mice,” Neurophotonics, 4 (2), 021102 https://doi.org/10.1117/1.NPh.4.2.021102 (2017). Google Scholar

15. 

V. Monga, Y. Li and Y. C. Eldar, “Algorithm unrolling: interpretable, efficient deep learning for signal and image processing,” IEEE Signal Process. Mag., 38 (2), 18 –44 https://doi.org/10.1109/MSP.2020.3016905 ISPRE6 1053-5888 (2021). Google Scholar

16. 

K. Gregor and Y. LeCun, “Learning fast approximations of sparse coding,” in Proc. 27th Int. Conf. Int. Conf. Mach. Learn., ICML’10, 399 –406 (2010). Google Scholar

17. 

A. Krizhevsky, I. Sutskever and G. E. Hinton, “Imagenet classification with deep convolutional neural networks,” Commun. ACM, 60 84 –90 https://doi.org/10.1145/3065386 CACMA2 0001-0782 (2017). Google Scholar

18. 

K. He et al., “Deep residual learning for image recognition,” in IEEE Conf. Comput. Vis. and Pattern Recognit. (CVPR), 770 –778 (2016). https://doi.org/10.1109/CVPR.2016.90 Google Scholar

19. 

Y. Zhao et al., “Dataset for Unrolled-DOT: an interpretable deep network for diffuse optical tomography,” https://doi.org/10.5281/zenodo.7654959 (2023). Google Scholar

20. 

H. Ben Yedder et al., “Deep learning based image reconstruction for diffuse optical tomography,” Lect. Notes Comput. Sci., 11074 112 –119 https://doi.org/10.1007/978-3-030-00129-2_13 LNCSD9 0302-9743 (2018). Google Scholar

21. 

S. Sabir et al., “Convolutional neural network-based approach to estimate bulk optical properties in diffuse optical tomography,” Appl. Opt., 59 1461 –1470 https://doi.org/10.1364/AO.377810 APOPAI 0003-6935 (2020). Google Scholar

22. 

J. Yoo et al., “Deep learning diffuse optical tomography,” IEEE Trans. Med. Imaging, 39 (4), 877 –887 https://doi.org/10.1109/TMI.2019.2936522 ITMID4 0278-0062 (2020). Google Scholar

23. 

F. F. Jöbsis, “Noninvasive, infrared monitoring of cerebral and myocardial oxygen sufficiency and circulatory parameters,” Science, 198 1264 –1267 https://doi.org/10.1126/science.929199 (1977). Google Scholar

24. 

M. Ferrari and V. Quaresima, “A brief review on the history of human functional near-infrared spectroscopy (fNIRS) development and fields of application,” NeuroImage, 63 (2), 921 –935 https://doi.org/10.1016/j.neuroimage.2012.03.049 NEIMEF 1053-8119 (2012). Google Scholar

25. 

M. D. Wheelock, J. P. Culver and A. T. Eggebrecht, “High-density diffuse optical tomography for imaging human brain function,” Rev. Sci. Instrum., 90 (5), 051101 https://doi.org/10.1063/1.5086809 RSINAK 0034-6748 (2019). Google Scholar

26. 

K. Lee, “Optical mammography: diffuse optical imaging of breast cancer,” World J. Clin. Oncol., 2 64 –72 https://doi.org/10.5306/wjco.v2.i1.64 (2011). Google Scholar

27. 

V. Venugopal, J. Chen and X. Intes, “Development of an optical imaging platform for functional imaging of small animals using wide-field excitation,” Biomed. Opt. Express, 1 143 –156 https://doi.org/10.1364/BOE.1.000143 BOEICL 2156-7085 (2010). Google Scholar

28. 

H. Zhao and R. J. Cooper, “Review of recent progress toward a fiberless, whole-scalp diffuse optical tomography system,” Neurophotonics, 5 (1), 011012 https://doi.org/10.1117/1.NPh.5.1.011012 (2017). Google Scholar

29. 

B. W. Pogue et al., “Instrumentation and design of a frequency-domain diffuse optical tomography imager for breast cancer detection,” Opt. Express, 1 391 –403 https://doi.org/10.1364/OE.1.000391 OPEXFF 1094-4087 (1997). Google Scholar

30. 

A. Pifferi et al., “New frontiers in time-domain diffuse optics, a review,” J. Biomed. Opt., 21 (9), 091310 https://doi.org/10.1117/1.JBO.21.9.091310 JBOPFO 1083-3668 (2016). Google Scholar

31. 

R. Yao, X. Intes and Q. Fang, “Direct approach to compute Jacobians for diffuse optical tomography using perturbation Monte Carlo-based photon ‘replay’,” Biomed. Opt. Express, 9 (10), 4588 –4603 https://doi.org/10.1364/BOE.9.004588 BOEICL 2156-7085 (2018). Google Scholar

32. 

J. Chen and X. Intes, “Time-gated perturbation Monte Carlo for whole body functional imaging in small animals,” Opt. Express, 17 19566 –19579 https://doi.org/10.1364/OE.17.019566 OPEXFF 1094-4087 (2009). Google Scholar

33. 

C. Zhang et al., “A differential theory of radiative transfer,” ACM Trans. Graph., 38 (6), 1 –16 https://doi.org/10.1145/3355089.3356522 ATGRDF 0730-0301 (2019). Google Scholar

34. 

W. Xie et al., “Boundary element method for diffuse optical tomography,” in Seventh Int. Conf. Image and Graph., 5 –8 (2013). https://doi.org/10.1109/ICIG.2013.7 Google Scholar

35. 

A. Joshi, W. Bangerth and E. M. Sevick-Muraca, “Adaptive finite element based tomography for fluorescence optical imaging in tissue,” Opt. Express, 12 5402 –5417 https://doi.org/10.1364/OPEX.12.005402 OPEXFF 1094-4087 (2004). Google Scholar

36. 

M. Schweiger, “GPU-accelerated finite element method for modelling light transport in diffuse optical tomography,” Int. J. Biomed. Imaging, 2011 403892 https://doi.org/10.1155/2011/403892 (2011). Google Scholar

37. 

C. Liu et al., “High resolution diffuse optical tomography using short range indirect subsurface imaging,” in IEEE Int. Conf. Comput. Photogr., (2020). https://doi.org/10.1109/ICCP48838.2020.9105173 Google Scholar

38. 

B. Chen, K. Stamnes and J. J. Stamnes, “Validity of the diffusion approximation in bio-optical imaging,” Appl. Opt., 40 6356 –6366 https://doi.org/10.1364/AO.40.006356 APOPAI 0003-6935 (2001). Google Scholar

39. 

H. K. Kim et al., “Frequency-domain optical tomographic image reconstruction algorithm with the simplified spherical harmonics (SP(3)) light propagation model,” Int. J. Therm. Sci., 116 265 –277 https://doi.org/10.1016/j.ijthermalsci.2017.03.004 (2017). Google Scholar

40. 

N. Naik et al., “Fully nonlinear SP3 approximation based fluorescence optical tomography,” IEEE Trans. Med. Imaging, 36 (11), 2308 –2318 https://doi.org/10.1109/TMI.2017.2718028 ITMID4 0278-0062 (2017). Google Scholar

41. 

D. Yang et al., “Performance investigation of SP3 and diffusion approximation for three-dimensional whole-body optical imaging of small animals,” Med. Biol. Eng. Comput., 53 805 –814 https://doi.org/10.1007/s11517-015-1293-8 MBECDY 0140-0118 (2015). Google Scholar

42. 

Y. Zou et al., “Machine learning model with physical constraints for diffuse optical tomography,” Biomed. Opt. Express, 12 5720 –5735 https://doi.org/10.1364/BOE.432786 BOEICL 2156-7085 (2021). Google Scholar

43. 

B. Deng, H. Gu and S. A. Carp, “Deep learning enabled high-speed image reconstruction for breast diffuse optical tomography,” Proc. SPIE, 11639 116390B https://doi.org/10.1117/12.2577736 PSISDG 0277-786X (2021). Google Scholar

44. 

N. I. Nizam et al., “Wide-field diffuse optical tomography using deep learning,” in Biophotonics Congr.: Biomed. Opt. 2022 (Transl., Microsc., OCT, OTS, BRAIN), OW4D.7 (2022). https://doi.org/10.1364/OTS.2022.OW4D.7 Google Scholar

45. 

B. Zhu et al., “Image reconstruction by domain-transform manifold learning,” Nature, 555 487 –492 https://doi.org/10.1038/nature25988 (2018). Google Scholar

46. 

S. Li et al., “Difference imaging from single measurements in diffuse optical tomography: a deep learning approach,” J. Biomed. Opt., 27 (8), 086003 https://doi.org/10.1117/1.JBO.27.8.086003 JBOPFO 1083-3668 (2022). Google Scholar

47. 

N. I. Nizam et al., “Monte Carlo-based data generation for efficient deep learning reconstruction of macroscopic diffuse optical tomography and topography applications,” J. Biomed. Opt., 27 (8), 083016 https://doi.org/10.1117/1.JBO.27.8.083016 JBOPFO 1083-3668 (2022). Google Scholar

48. 

G. M. Balasubramaniam et al., “Tutorial on the use of deep learning in diffuse optical tomography,” Electronics, 11 (3), 305 https://doi.org/10.3390/electronics11030305 ELECAD 0013-5070 (2022). Google Scholar

49. 

J. T. Smith et al., “Deep learning in macroscopic diffuse optical imaging,” J. Biomed. Opt., 27 (2), 020901 https://doi.org/10.1117/1.JBO.27.2.020901 JBOPFO 1083-3668 (2022). Google Scholar

50. 

F. Long, “Deep learning-based mesoscopic fluorescence molecular tomography: an in silico study,” J. Med. Imaging, 5 (3), 036001 https://doi.org/10.1117/1.JMI.5.3.036001 (2018). Google Scholar

51. 

H. Wang et al., “Fluorescence molecular tomography reconstruction of small targets using stacked auto-encoder neural networks,” IEEE Access, 8 37657 –37663 https://doi.org/10.1109/ACCESS.2020.2975807 (2020). Google Scholar

52. 

M. A. K. Sagar et al., “Machine learning methods for fluorescence lifetime imaging (flim) based label-free detection of microglia,” Front. Neurosci., 14 931 https://doi.org/10.3389/fnins.2020.00931 1662-453X (2020). Google Scholar

53. 

J. T. Smith et al., “Fast fit-free analysis of fluorescence lifetime imaging via deep learning,” Proc. Natl. Acad. Sci., 116 (48), 24019 –24030 https://doi.org/10.1073/pnas.1912707116 (2019). Google Scholar

54. 

V. Mannam et al., “Machine learning for faster and smarter fluorescence lifetime imaging microscopy,” J. Phys.: Photonics, 2 042005 https://doi.org/10.1088/2515-7647/abac1a (2020). Google Scholar

55. 

M. T. Chen et al., “GANPOP: generative adversarial network prediction of optical properties from single snapshot wide-field images,” IEEE Trans. Med. Imaging, 39 1988 –1999 https://doi.org/10.1109/TMI.2019.2962786 ITMID4 0278-0062 (2020). Google Scholar

56. 

J. Feng et al., “Deep-learning based image reconstruction for mri-guided near-infrared spectral tomography,” Optica, 9 264 –267 https://doi.org/10.1364/OPTICA.446576 (2022). Google Scholar

57. 

H. Wang et al., “Continuous monitoring method of cerebral subdural hematoma based on mri guided dot,” Biomed. Opt. Express, 11 2964 –2975 https://doi.org/10.1364/BOE.388059 BOEICL 2156-7085 (2020). Google Scholar

58. 

Y. Yang et al., “ADMM-CSNet: a deep learning approach for image compressive sensing,” IEEE Trans. Pattern Anal. Mach. Intell., 42 (3), 521 –538 https://doi.org/10.1109/TPAMI.2018.2883941 ITPIDJ 0162-8828 (2020). Google Scholar

59. 

K. Monakhova et al., “Learned reconstructions for practical mask-based lensless imaging,” Opt. Express, 27 28075 –28090 https://doi.org/10.1364/OE.27.028075 OPEXFF 1094-4087 (2019). Google Scholar

60. 

J. Liu et al., “ALISTA: analytic weights are as good as learned weights in LISTA,” in Int. Conf. Learn. Represent., (2019). Google Scholar

61. 

Y. Hua et al., “Interpretable model-driven projected gradient descent network for high-quality fdot reconstruction,” Opt. Lett., 47 2538 –2541 https://doi.org/10.1364/OL.459456 OPLEDP 0146-9592 (2022). Google Scholar

62. 

D. Hyde, “Improving forward matrix generation and utilization for time domain diffuse optical tomography,” Worcester Polytechnic Institute, (2002). Google Scholar

63. 

M. Schweiger and S. R. Arridge, “The Toast++ software suite for forward and inverse modeling in optical tomography,” J. Biomed. Opt., 19 (4), 040801 https://doi.org/10.1117/1.JBO.19.4.040801 JBOPFO 1083-3668 (2014). Google Scholar

64. 

L. Wang, S. L. Jacques and L. Zheng, “MCML - Monte Carlo modeling of light transport in multi-layered tissues,” Comput. Methods Programs Biomed., 47 (2), 131 –146 https://doi.org/10.1016/0169-2607(95)01640-F CMPBEK 0169-2607 (1995). Google Scholar

65. 

C. Li and H. Jiang, “A calibration method in diffuse optical tomography,” J. Opt. A: Pure Appl. Opt., 6 844 https://doi.org/10.1088/1464-4258/6/9/005 (2004). Google Scholar

66. 

A. Beck and M. Teboulle, “A fast iterative shrinkage-thresholding algorithm,” Soc. Ind. Appl. Math. J. Imaging Sci., 2 (1), 183 –202 https://doi.org/10.1109/ICASSP.2009.4959678 (2009). Google Scholar

67. 

S. Boyd et al., “Distributed optimization and statistical learning via the alternating direction method of multipliers,” Found. Trends Mach. Learn., 3 1 –122 https://doi.org/10.1561/2200000016 (2011). Google Scholar

68. 

S. Khan et al., “FlatNet: towards photorealistic scene reconstruction from lensless measurements,” IEEE Trans. Pattern Anal. Mach. Intell., 44 1934 –1948 https://doi.org/10.1109/TPAMI.2020.3033882 ITPIDJ 0162-8828 (2020). Google Scholar

69. 

O. Ronneberger, P. Fischer and T. Brox, “U-net: convolutional networks for biomedical image segmentation,” Lect. Notes Comput. Sci., 9351 234 –241 https://doi.org/10.1007/978-3-319-24574-4_28 LNCSD9 0302-9743 (2015). Google Scholar

70. 

L. Jin et al., “Deep learning extended depth-of-field microscope for fast and slide-free histology,” Proc. Natl. Acad. Sci., 117 (52), 33051 –33060 https://doi.org/10.1073/pnas.2013571117 (2020). Google Scholar

71. 

K. Simonyan and A. Zisserman, “Very deep convolutional networks for large-scale image recognition,” in Int. Conf. Learn. Represent., (2015). Google Scholar

72. 

J. Johnson, A. Alahi and L. Fei-Fei, “Perceptual losses for real-time style transfer and super-resolution,” Lect. Notes Comput. Sci., 9906 694 –711 https://doi.org/10.1007/978-3-319-46475-6_43 LNCSD9 0302-9743 (2016). Google Scholar

73. 

E. Riba et al., “Kornia: an open source differentiable computer vision library for pytorch,” in Winter Conf. Appl. Comput. Vis., (2020). https://doi.org/10.1109/WACV45572.2020.9093363 Google Scholar

74. 

B. Comiskey et al., “An electrophoretic ink for all-printed reflective electronic displays,” Nature, 394 (6690), 253 –255 https://doi.org/10.1038/28349 (1998). Google Scholar

75. 

F. Bevilacqua et al., “In vivo local determination of tissue optical properties: applications to human brain,” Appl. Opt., 38 (22), 4939 –4950 https://doi.org/10.1364/AO.38.004939 APOPAI 0003-6935 (1999). Google Scholar

76. 

J.-P. Bouchard et al., “Reference optical phantoms for diffuse optical spectroscopy. Part 1 – error analysis of a time resolved transmittance characterization method,” Opt. Express, 18 11495 –11507 https://doi.org/10.1364/OE.18.011495 OPEXFF 1094-4087 (2010). Google Scholar

77. 

L. A. Dempsey et al., “Geometrically complex 3d-printed phantoms for diffuse optical imaging,” Biomed. Opt. Express, 8 1754 –1762 https://doi.org/10.1364/BOE.8.001754 BOEICL 2156-7085 (2017). Google Scholar

78. 

M. Schweiger and S. R. Arridge, “Application of temporal filters to time resolved data in optical tomography,” Phys. Med. Biol., 44 1699 –1717 https://doi.org/10.1088/0031-9155/44/7/310 PHMBA7 0031-9155 (1999). Google Scholar

79. 

S. Hao and S. Li, “A weighted mean absolute error metric for image quality assessment,” in IEEE Int. Conf. Visual Commun. and Image Process. (VCIP), 330 –333 (2020). https://doi.org/10.1109/VCIP49819.2020.9301889 Google Scholar

80. 

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

81. 

H. Zhao et al., “Loss functions for image restoration with neural networks,” IEEE Trans. Comput. Imaging, 3 (1), 47 –57 https://doi.org/10.1109/TCI.2016.2644865 (2017). Google Scholar

82. 

W. Lu et al., “Graph- and finite element-based total variation models for the inverse problem in diffuse optical tomography,” Biomed. Opt. Express, 10 2684 –2707 https://doi.org/10.1364/BOE.10.002684 BOEICL 2156-7085 (2019). Google Scholar

Biographies of the authors are not available.

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.
Yongyi Zhao, Ankit Raghuram, Fay Wang, Stephen Hyunkeol Kim, Andreas H. Hielscher, Jacob T. Robinson, and Ashok Veeraraghavan "Unrolled-DOT: an interpretable deep network for diffuse optical tomography," Journal of Biomedical Optics 28(3), 036002 (8 March 2023). https://doi.org/10.1117/1.JBO.28.3.036002
Received: 19 September 2022; Accepted: 9 February 2023; Published: 8 March 2023
Advertisement
Advertisement
KEYWORDS
Image restoration

Education and training

Reconstruction algorithms

Machine learning

Data modeling

Image quality

Diffuse optical tomography

Back to Top