11institutetext: School of Biomedical Engineering, University of British Columbia 22institutetext: Department of Integrative Oncology, BC Cancer Research Institute 33institutetext: Department of Radiology, University of British Columbia 44institutetext: Department of Physics, University of British Columbia55institutetext: Department of Department of Earth, Ocean and Atmospheric Sciences
55email: [email protected], 55email: [email protected], 55email: [email protected]

Beyond Conventional Parametric Modeling: Data-Driven Framework for Estimation and Prediction of Time Activity Curves in Dynamic PET Imaging

Niloufar Zakariaei 1122    Arman Rahmim 11223344    Eldad Haber 55
Abstract

Dynamic Positron Emission Tomography (dPET) imaging and Time-Activity Curve (TAC) analyses are essential for understanding and quantifying the biodistribution of radiopharmaceuticals over time and space. Traditional compartmental modeling, while foundational, commonly struggles to fully capture the complexities of biological systems, including non-linear dynamics and variability. This study introduces an innovative data-driven neural network-based framework, inspired by Reaction Diffusion systems, designed to address these limitations. Our approach, which adaptively fits TACs from dPET, enables the direct calibration of diffusion coefficients and reaction terms from observed data, offering significant improvements in predictive accuracy and robustness over traditional methods, especially in complex biological scenarios. By more accurately modeling the spatio-temporal dynamics of radiopharmaceuticals, our method advances modeling of pharmacokinetic and pharmacodynamic processes, enabling new possibilities in quantitative nuclear medicine.

Keywords:
Dynamic PET Reaction-Diffusion Neural Network Predictive Modeling

1 Introduction

Positron Emission Tomography (PET) is a medical imaging technique that uses radioactive pharmaceuticals to visualize and measure body physiological processes, widely applied in oncology, neurology, and cardiology [32, 12, 34]. Dynamic PET, through sequential imaging, allows for the detailed visualization and quantification of radiopharmaceutical distribution within the body over time, providing valuable insights into physiological processes such as blood flow, glucose metabolism, and receptor binding [23].

The processing of dynamic PET data involves image reconstruction followed by generation and analysis of Time-Activity Curves (TACs). TACs are generated by plotting the radioactivity concentration over time in specific regions of interest (ROI) [28]. Subsequently, pharmacokinetic modeling is used to estimate physiological parameters related to the radiopharmaceutical’s behavior. Commonly, parametric models are used. Such models are based on compartmental modeling, incorporating different physiological compartments and their interconnections. These models are dubbed as parametric models because they contain a small number of parameters (typically less than 10) that govern the shape of the TAC [9, 7].

In principle, a reasonable parametric pharmacokinetic model should fit a TAC and be flexible enough to accommodate different TACs from different regions and patients. In particular, Two-Tissue Compartment Model (2TCM) and Three-Tissue Compartment Model (3TCM) are common in practice [9, 34]. Since such models contain only a few parameters, they are straightforward to use in order to explain dynamics. However, these models rely on strong assumptions that may not hold under complex biological conditions, potentially limiting their applicability in realistic settings [14, 8]. For instance, these models assume it is possible to homogenize each compartment to a constant rate of exchange between compartments, neglecting the variability and non-linear dynamics characteristic of biological systems. Furthermore, these models do not commonly consider the impact of physiological changes in the course of imaging or the influence of underlying pathologies that may alter radiopharmaceutical kinetics and distribution, which can lead to inaccuracies in PET signal interpretation [1, 31, 19]. There are also issues of patient and/or organ movements over time that may not be fully compensated even with motion correction methods, impacting kinetic models [17]. Finally, research in fields such as emergence [6] suggests that combining even simple properties (like the decay at each cell) can yield a much more complex behavior of the homogenized system.

These limitations highlight the need to extend or refine compartmental models to better capture the complexity of biological systems and enhance PET analysis reliability. A motivating example for this is demonstrated in Figure 2 where 1 slice of a liver are given at different times.

Refer to caption Refer to caption Refer to caption Refer to caption
t1subscript𝑡1t_{1}italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT t2subscript𝑡2t_{2}italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT t3subscript𝑡3t_{3}italic_t start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT t4subscript𝑡4t_{4}italic_t start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT
Refer to caption Refer to caption Refer to caption Refer to caption
t5subscript𝑡5t_{5}italic_t start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT t6subscript𝑡6t_{6}italic_t start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT t7subscript𝑡7t_{7}italic_t start_POSTSUBSCRIPT 7 end_POSTSUBSCRIPT t8subscript𝑡8t_{8}italic_t start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT
Refer to caption
Figure 1: Dynamic liver images captured at eight distinct time points.
Refer to caption
Figure 2: The average of the TAC curve over the liver, along with its standard deviation. Data is fitted using a three-tissue compartment model (3TCM) based on the early 6 time frames.

While some decay is evident initially, it is challenging to justify a specific parametric form. Moreover, averaging across the entire organ yields a curve (see Figure 2) that does not exhibit typical compartment behavior of single- or multi-exponential biological decay. Observation of the data reveals spatial patterns that are that are not resolved by averaging activity over the organ. These patterns suggest variability in liver function, which could be insightful for therapeutic treatments.

The goal of this paper is to propose a data-driven methodology for the estimation of TACs, that vary both temporally and spatially. Instead of using a particular parametric form for a TAC, we utilize a carefully designed neural network architecture that aligns with the physical behavior of the phenomena we observe, calibrating its parameters based on the data at hand. Neural network models, far more complex than simple compartmental models, are justified only by their enhanced predictive power. In this work, we show that indeed, these models outperform simple compartmental models, at least for the datasets we have tested.

Neural network architectures that mimic physical systems are now common in many fields of science and engineering [11, 13, 24]. Such architectures, based on Partial Differential Equations in high dimensions, can be tailored to different physical phenomena, contrasting with compartmental models based on fitting very few parameters in an Ordinary Differential Equations. As highlighted in the literature [11], this approach allows for more tailored and nuanced simulations of physical systems, offering potential advantages over compartmental models in terms of both precision and applicability. Here, we choose an architecture based on a Reaction Diffusion system. The data is used to learn the diffusion coefficients and the reaction term. We show that such a network can fit spatio-temporal patterns observed in the data, obtaining stronger predictive power than classical parametric models. These models can be effectively trained, uncovering new patterns in the data that simple parametric models cannot resolve. Moreover, by using a physically motivated neural network architecture, our network can be seen as an extension of a compartment model into high dimensions.

2 Deriving a reaction diffusion neural network for the modeling of TACs

Reaction Diffusion systems have been used extensively in biology [20, 27, 16], from modeling the propagation of electromagnetic waves in the heart [22] to patterns generated on butterfly wings [16]. The equations describe the interaction between a number of species (that can be chemicals or different populations) and their spatial dynamics.

Hinted by its name, the equation comprises two parts: reaction and diffusion. The reaction term is local, meaning it is pointwise. Compartment models can, in fact, be considered as local reaction terms. The diffusion represents the spatial dynamics. Typically, different species have different diffusion coefficients. The interaction between the diffusion and reaction terms is the cause of pattern formation (see the classical work by Turing [33] and references within). The reaction-diffusion equation can be written as

𝐮t=𝜿Δ𝐮+R(𝐮,t;𝜽)𝐮𝑡𝜿Δ𝐮𝑅𝐮𝑡𝜽\displaystyle{\frac{\partial{\bf u}}{\partial t}}={\bm{\kappa}}\Delta{\bf u}+R% ({\bf u},t;{\bm{\theta}})divide start_ARG ∂ bold_u end_ARG start_ARG ∂ italic_t end_ARG = bold_italic_κ roman_Δ bold_u + italic_R ( bold_u , italic_t ; bold_italic_θ ) (1)

with appropriate initial and boundary conditions. Here, 𝐮=[𝐮1,,𝐮c]𝐮subscript𝐮1subscript𝐮𝑐{\bf u}=[{\bf u}_{1},\ldots,{\bf u}_{c}]bold_u = [ bold_u start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , bold_u start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ] is a vector representing c𝑐citalic_c different species (or in the context of deep networks, channels), and the coefficients 𝜿=[𝜿1,,𝜿s]𝜿subscript𝜿1subscript𝜿𝑠{\bm{\kappa}}=[{\bm{\kappa}}_{1},\ldots,{\bm{\kappa}}_{s}]bold_italic_κ = [ bold_italic_κ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , bold_italic_κ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ] are the diffusion coefficients for each species. The reaction term R(𝐮,t;𝜽)𝑅𝐮𝑡𝜽R({\bf u},t;{\bm{\theta}})italic_R ( bold_u , italic_t ; bold_italic_θ ) couples the different species through nonlinear interaction. Finally, the parameters 𝜽𝜽{\bm{\theta}}bold_italic_θ are trainable parameters in the reaction term.

While reaction diffusion systems have been applied to physical and biological systems for a long time, recent developments in neural network technology have demonstrated that it is possible to derive a neural network interpreted as a reaction diffusion system in high dimensions [24]. In the context of dynamic PET, the image under consideration can be thought of as a weighted sum of different species. In the 2 or 3 compartment model, only 2 or 3 species are used; however, with a deep network, one can employ an arbitrarily large number of species and learn the reaction term, that is, the interaction between them.

Let I(t,𝐱)𝐼𝑡𝐱I(t,\mathbf{x})italic_I ( italic_t , bold_x ) represent the PET image we aim to model, where both 𝐱𝐱\mathbf{x}bold_x and t𝑡titalic_t are discretizations of space and time. Initially, the network embeds I(𝐱,t)𝐼𝐱𝑡I(\mathbf{x},t)italic_I ( bold_x , italic_t ) into a higher-dimensional state using a so-called Multilayer Preceptron (MLP) [18]. Let 𝐮(𝐱,t)=[𝐮1(𝐱,t),,𝐮c(𝐱,t)]𝐮𝐱𝑡subscript𝐮1𝐱𝑡subscript𝐮𝑐𝐱𝑡\mathbf{u}(\mathbf{x},t)=[\mathbf{u}_{1}(\mathbf{x},t),\ldots,\mathbf{u}_{c}(% \mathbf{x},t)]bold_u ( bold_x , italic_t ) = [ bold_u start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( bold_x , italic_t ) , … , bold_u start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( bold_x , italic_t ) ] be the embedded state. We assume 𝐮𝐮\mathbf{u}bold_u adheres to the reaction diffusion equation 1 and discretize it in space-time. A common approach for the discretization of such a system is the Implicit-Explicit method (IMEX) (see [2, 25, 26, 3] and references within). Let 𝐀𝐀\mathbf{A}bold_A be a discretization of the negative Laplacian. Then the discretization of the system reads

𝐮~k+1𝐮k=hκ𝐀𝐮k+1and𝐮k+1𝐮~k+1=hR(𝐮~k+1,t;𝜽)formulae-sequencesubscript~𝐮𝑘1subscript𝐮𝑘𝜅subscript𝐀𝐮𝑘1andsubscript𝐮𝑘1subscript~𝐮𝑘1𝑅subscript~𝐮𝑘1𝑡𝜽\displaystyle\tilde{\mathbf{u}}_{k+1}-\mathbf{u}_{k}=-h\kappa\mathbf{A}\mathbf% {u}_{k+1}\quad{\rm and}\quad\mathbf{u}_{k+1}-\tilde{\mathbf{u}}_{k+1}=hR(% \tilde{\mathbf{u}}_{k+1},t;\bm{\theta})over~ start_ARG bold_u end_ARG start_POSTSUBSCRIPT italic_k + 1 end_POSTSUBSCRIPT - bold_u start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = - italic_h italic_κ bold_Au start_POSTSUBSCRIPT italic_k + 1 end_POSTSUBSCRIPT roman_and bold_u start_POSTSUBSCRIPT italic_k + 1 end_POSTSUBSCRIPT - over~ start_ARG bold_u end_ARG start_POSTSUBSCRIPT italic_k + 1 end_POSTSUBSCRIPT = italic_h italic_R ( over~ start_ARG bold_u end_ARG start_POSTSUBSCRIPT italic_k + 1 end_POSTSUBSCRIPT , italic_t ; bold_italic_θ ) (2)

where hhitalic_h is the time step. The first equation is implicit and requires the solution of the system

(𝐈+hκ𝐀)𝐮~k+1=𝐮k𝐈𝜅𝐀subscript~𝐮𝑘1subscript𝐮𝑘\displaystyle(\mathbf{I}+h\kappa\mathbf{A})\tilde{\mathbf{u}}_{k+1}=\mathbf{u}% _{k}( bold_I + italic_h italic_κ bold_A ) over~ start_ARG bold_u end_ARG start_POSTSUBSCRIPT italic_k + 1 end_POSTSUBSCRIPT = bold_u start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT (3)

The solution of a linear system is, in general, slow. However, since the image is defined on a regular grid, the inversion of the method can be efficiently achieved in order nlog(n)𝑛𝑛n\log(n)italic_n roman_log ( italic_n ) (where n𝑛nitalic_n is the number of pixels) by using a cosine transform [21]. The reaction term equation 2 is modeled by a two-layer MLP with the form

R(𝐮,t)=𝐊2σ(𝐊1𝐮+𝐭e).𝑅𝐮𝑡subscript𝐊2𝜎subscript𝐊1𝐮subscript𝐭𝑒\displaystyle R(\mathbf{u},t)=\mathbf{K}_{2}\sigma(\mathbf{K}_{1}\mathbf{u}+% \mathbf{t}_{e}).italic_R ( bold_u , italic_t ) = bold_K start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_σ ( bold_K start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT bold_u + bold_t start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ) . (4)

Here, 𝐊1subscript𝐊1\mathbf{K}_{1}bold_K start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and 𝐊2subscript𝐊2\mathbf{K}_{2}bold_K start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are so-called 1×1111\times 11 × 1 convolutions; i.e., they only mix the channels of 𝐮𝐮\mathbf{u}bold_u. The function σ𝜎\sigmaitalic_σ is an activation function (here we have used the silu activation) and 𝐭esubscript𝐭𝑒\mathbf{t}_{e}bold_t start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT is an embedding of the time. Here, similar to diffusion-based methods [30], we have used an MLP

𝐭e=𝐖σ(t𝐛)subscript𝐭𝑒𝐖𝜎𝑡𝐛\displaystyle\mathbf{t}_{e}=\mathbf{W}\sigma(t\mathbf{b})bold_t start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = bold_W italic_σ ( italic_t bold_b ) (5)

where t𝑡titalic_t is the scalar input time and 𝐛𝐛\mathbf{b}bold_b and 𝐖𝐖\mathbf{W}bold_W are learnable parameters.

The output of the network is an image: therefore, the network has one final so-called closing layer that compresses the output of the reaction diffusion discretization into a single channel. A sketch of our network is plotted in Figure 3.

Refer to caption
Figure 3: The reaction diffusion network architecture: An opening layer takes the input image into a higher embedded dimension, where a reaction diffusion network with learned diffusivity and reaction operates. In the example above, 3 channels are utilized to open the image.

Given the time-dependent data, our network learns the MLP that embeds the image, the diffusion coefficients 𝜿𝜿{\bm{\kappa}}bold_italic_κ, the convolution matrices 𝐊1subscript𝐊1{\bf K}_{1}bold_K start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, 𝐊2subscript𝐊2{\bf K}_{2}bold_K start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, and the time embeddings 𝐛𝐛{\bf b}bold_b and 𝐖𝐖{\bf W}bold_W. Finally, the network learns the closing MLP layer.

Note that similar to Long short-term memory (LSTM’s) our network shares parameters between different time steps [29] and also unlike LSTM our network uses time embedding, which makes the function explicitly depend on time. We observed that this allows us to better predict the time behaviour of the system.

3 Training the network and Dataset

As stated in the introduction, a neural network model can be justified to use if it yields better predictions. Our training process is geared to demonstrate that. Let I(tj,𝐱),j=1,,Nformulae-sequence𝐼subscript𝑡𝑗𝐱𝑗1𝑁I(t_{j},{\bf x}),j=1,\ldots,Nitalic_I ( italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT , bold_x ) , italic_j = 1 , … , italic_N be a set of images obtained from a single patient. For shorthand we define 𝐈j=I(tj,𝐱)subscript𝐈𝑗𝐼subscript𝑡𝑗𝐱{\bf I}_{j}=I(t_{j},{\bf x})bold_I start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_I ( italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT , bold_x ). To this end, we have divided the data set into two groups: a training group 𝐈1,,𝐈ssubscript𝐈1subscript𝐈𝑠{\bf I}_{1},\ldots,{\bf I}_{s}bold_I start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , bold_I start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and a validation set 𝐈s+1,,𝐈Nsubscript𝐈𝑠1subscript𝐈𝑁{\bf I}_{s+1},\ldots,{\bf I}_{N}bold_I start_POSTSUBSCRIPT italic_s + 1 end_POSTSUBSCRIPT , … , bold_I start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT. Our goal is to train on the training images, in times t1,,tssubscript𝑡1subscript𝑡𝑠t_{1},\ldots,t_{s}italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and for the network to predict the later images that are in the validation set. In the training, we assume to have an image 𝐈jsubscript𝐈𝑗{\bf I}_{j}bold_I start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT and we attempt to predict 𝐈j+1subscript𝐈𝑗1{\bf I}_{j+1}bold_I start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT, that is,

𝐈j+1pred=f(𝐈j,𝜽)subscriptsuperscript𝐈pred𝑗1𝑓subscript𝐈𝑗𝜽\displaystyle{\bf I}^{\rm pred}_{j+1}=f({\bf I}_{j},{\bm{\theta}})bold_I start_POSTSUPERSCRIPT roman_pred end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT = italic_f ( bold_I start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT , bold_italic_θ ) (6)

where f(,)𝑓f(\cdot,\cdot)italic_f ( ⋅ , ⋅ ) is the neural network described above and 𝜽𝜽{\bm{\theta}}bold_italic_θ are the neural network parameters. To calibrate the parameters we minimize the standard Mean Squared Error (MSE) loss 12j𝐈j+1pred(𝜽)𝐈j+1212subscript𝑗superscriptnormsubscriptsuperscript𝐈pred𝑗1𝜽subscript𝐈𝑗12{\frac{1}{2}}\sum_{j}\|{\bf I}^{\rm pred}_{j+1}({\bm{\theta}})-{\bf I}_{j+1}\|% ^{2}divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ∥ bold_I start_POSTSUPERSCRIPT roman_pred end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT ( bold_italic_θ ) - bold_I start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT ∥ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. The training method is commonly used for LSTM networks [4] and has also been recently proposed for training reaction diffusion networks in the context of graph neural networks [10].

For the minimization process, we utilize the Adam optimizer [15], employing gradient clip** to ensuring the stability. To test the predictability of our model, our network is trained only on the first s𝑠sitalic_s time steps, and then it is used to predict the subsequent activity. As mentioned in the introduction, the justification for using a complex model such as a neural network is its ability to predict the TAC beyond its training point, which is crucial for validation. As demonstrated in the following section, our model successfully achieves this objective.

Our research employs a dataset from 7 male patients undergoing [18F]DCFPyL imaging for prostate cancer. The collected data reveal an average age of 70.29±1.89plus-or-minus70.291.8970.29\pm 1.8970.29 ± 1.89 years, a mean weight of 93.71±18.20plus-or-minus93.7118.2093.71\pm 18.2093.71 ± 18.20 kg and an average administered dose of 7.08±1.31plus-or-minus7.081.317.08\pm 1.317.08 ± 1.31 mCi.

4 Numerical Results and Evaluation

As outlined in Section 3, we used the early time frames (here the first 11 frames) of our dataset to train our network to predict the latest frames (last 4 frames). Similarly, we applied this method to the conventional 3TCM using non-linear least squares analysis. Finally, we compared the reaction diffusion network’s performance against the conventional 3TCM, benchmarking both against ground truth. Figure 4 clearly illustrates that the reaction diffusion network aligns more closely with the actual data than the 3TCM in the latter four time frames.

Ground Truth

Refer to caption

t=68min𝑡68mint=68\,\text{min}italic_t = 68 min

Refer to caption

t=73min𝑡73mint=73\,\text{min}italic_t = 73 min

Refer to caption

t=78min𝑡78mint=78\,\text{min}italic_t = 78 min

Refer to caption

t=83min𝑡83mint=83\,\text{min}italic_t = 83 min

3TCM

Refer to caption

t=68min𝑡68mint=68\,\text{min}italic_t = 68 min

Refer to caption

t=73min𝑡73mint=73\,\text{min}italic_t = 73 min

Refer to caption

t=78min𝑡78mint=78\,\text{min}italic_t = 78 min

Refer to caption

t=83min𝑡83mint=83\,\text{min}italic_t = 83 min

Proposed Model

Refer to caption

t=68min𝑡68mint=68\,\text{min}italic_t = 68 min

Refer to caption

t=73min𝑡73mint=73\,\text{min}italic_t = 73 min

Refer to caption

t=78min𝑡78mint=78\,\text{min}italic_t = 78 min

Refer to caption

t=83min𝑡83mint=83\,\text{min}italic_t = 83 min

Refer to caption
Figure 4: A selected slice of one of the patients, showcasing the latest time points predicted by our proposed model versus the 3TCM and the original images.

The superiority of our model is further evidenced in Figure  5, which provides a detailed examination of small, defined regions within each organ (15-30 pixels), as detailed in Table  2 and Table  2. Here, we show the result for 3 patients as an example. The neural network’s predicted TACs are shown to more accurately reflect the ground truth compared to those produced by 3TCM, which is less accurate when the data doesn’t follow a decaying pattern.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 5: Predicted vs. actual [18F]DCFPyL concentrations over time in 3 patients. Training data are shown with solid markers, test data with hollow markers. Predictions from the reaction diffusion model and traditional 3TCM are compared with original test. Trend lines connect data points, showing our model’s enhanced accuracy in capturing distribution dynamics.

The efficacy of our model is quantitatively substantiated in Table  2 and Table  2 , which uses the MSE metric, calculated over the test time frames, highlighting our model’s predictive strength for future time points. For a robust analysis, five characteristic slices of the targeted organ from each patient were analyzed, aligning with the test time frames.

Table 1: Proposed Model

Patient

Parotid

Blood

Liver

Spleen

Kidney

1 0.685 0.195 0.134 0.207 0.343
2 1.974 0.441 0.115 0.062 0.212
3 0.553 0.665 0.105 0.054 0.121
4 2.752 0.424 0.505 0.151 1.037
5 0.657 1.045 0.687 0.351 1.099
6 1.404 1.085 0.072 0.051 0.942
7 0.750 4.167 0.250 0.189 0.236
Table 2: 3TCM

Patient

Parotid

Blood

Liver

Spleen

Kidney

1 18.08 1.927 1.671 2.611 8.767
2 55.27 1.199 4.701 13.59 17.71
3 27.13 1.822 2.103 0.007 19.05
4 44.63 1.651 3.856 0.715 18.84
5 28.28 0.282 1.661 0.059 2.969
6 27.53 1.076 4.073 2.694 19.60
7 23.36 4.167 3.498 1.758 11.27
Comparison across 5 organs in 7 patients shows the proposed model’s predictive accuracy vs. 3TCM over all test time frames, using MSE metric.

5 Discussion and Conclusions

In this paper, we propose the use of a reaction diffusion neural network to model TACs in Dynamic PET imaging. These models replace the commonly used two- and three-compartment models. Although more complex, they offer higher flexibility in fitting data, allowing for better prediction of non-trivial TACs. It is important to note that reaction diffusion neural networks differ from "generic" neural networks and tend to behave similarly to their continuous analogs as extensively used in chemistry and biology [33]. As such, these models can be considered "biological", similar to other reaction diffusion models in biology. The results obtained by our networks suggest they are far more predictive compared to compartmental models. In fact, our predictions of TACs are superior to any parametric model known to us, suggesting that our model should be considered when a model is needed to predict TAC’s. This is particularly meaningful, given the development of complex radiopharmaceuticals (e.g., [18F][^{18}\text{F}][ start_POSTSUPERSCRIPT 18 end_POSTSUPERSCRIPT F ]DCFPyL and next-gen theranostics) with mechanisms less known than those of older radiopharmaceuticals [5].

One important question remains unanswered: why are reaction diffusion models superior to simpler models that are pointwise (i.e., models that do not include a spatial term)? We hypothesize that the diffusion predicted by the model does not represent only diffusion but also the blurring of the image due to unaccounted-for recovery, motion errors and partial volume effects that leads to blurring which is equivalent to diffusion. The reaction term used in our network is simple, involving only two layers with a single non-linearity. Approximation theory suggests that such an architecture is sufficient to model any continuous function [35]. Thus, while we have presented a data-driven approach to parameter estimation for TACs in dynamic PET, our model’s superior predictive power can be both motivated and explainable physically.

References

  • [1] Anderson, R.M., May, R.M.: Infectious Diseases of Humans: Dynamics and Control. Oxford University Press (1991)
  • [2] Ascher, U.: Stabilization of invariants of discretized differential systems. Numerical Algorithms 14, 1–23 (1997)
  • [3] Ascher, U.: Numerical methods for Evolutionary Differential Equations. SIAM, Philadelphia (2010)
  • [4] Chen, J., Xu, X., Wu, Y., Zheng, H.: GC-LSTM: Graph Convolution Embedded LSTM for Dynamic Link Prediction. arXiv preprint arXiv:1812.04206 (2018)
  • [5] Cri\textcommabelowsan, G., Moldovean-Cioroianu, N.S., Timaru, D.G., Andrie\textcommabelows, G., Căinap, C., Chi\textcommabelows, V.: Radiopharmaceuticals for pet and spect imaging: A literature review over the last decade. International Journal of Molecular Sciences 23(9),  5023 (2022)
  • [6] Cucker, F., Smale, S.: On the mathematics of emergence. Japanese Journal of Mathematics 2, 197–227 (2007)
  • [7] De Benetti, F., Simson, W., Paschali, M., Sari, H., Rominger, A., Shi, K., Navab, N., Wendler, T.: Self-supervised learning for physiologically-based pharmacokinetic modeling in dynamic pet. In: Medical Image Computing and Computer Assisted Intervention – MICCAI 2023. pp. 290–299. Springer Nature Switzerland, Cham (2023)
  • [8] Diekmann, O., Heesterbeek, H., Britton, T.: Mathematical Tools for Understanding Infectious Disease Dynamics. Princeton University Press (2012)
  • [9] Dimitrakopoulou-Strauss, A., Pan, L., Sachpekidis, C.: Kinetic modeling and parametric imaging with dynamic pet for oncological applications: general considerations, current clinical applications, and future perspectives. European journal of nuclear medicine and molecular imaging 48, 21–39 (2021)
  • [10] Eliasof, M., Haber, E., Treister, E.: Adr-gnn: Advection-diffusion-reaction graph neural networks. arXiv preprint arXiv:2307.16092 (2023)
  • [11] Haber, E., Ruthotto, L.: Stable architectures for deep neural networks. arxiv preprint 1705.03341 abs/1705.03341, 1–21 (2017), http://arxiv.longhoe.net/abs/1705.03341
  • [12] Jadvar, H., Parker, J.: Clinical PET and PET/CT. Springer-Verlag London, 1 edn. (2005). https://doi.org/10.1007/b138777, https://doi.org/10.1007/b138777
  • [13] **, K.H., McCann, M.T., Froustey, E., Unser, M.: Deep convolutional neural network for inverse problems in imaging. IEEE Transactions on Image Processing 26(9), 4509–4522 (2017)
  • [14] Keeling, M.J., Rohani, P.: Modeling Infectious Diseases in Humans and Animals. Princeton University Press (2008)
  • [15] Kingma, D.P., Ba, J.: Adam: A method for stochastic optimization. arXiv preprint arXiv:1412.6980 (2014)
  • [16] Kondo, S., Miura, T.: Reaction-diffusion model as a framework for understanding biological pattern formation. science 329(5999), 1616–1620 (2010)
  • [17] Kotasidis, F.A., Tsoumpas, C., Rahmim, A.: Advanced kinetic modelling strategies: towards adoption in clinical pet imaging. Clinical and Translational Imaging 2, 219–237 (2014)
  • [18] Krizhevsky, A., Sutskever, I., Hinton, G.: Imagenet classification with deep convolutional neural networks. Adv Neural Inf Process Syst 61, 1097–1105 (2012)
  • [19] Morris, E.D., Endres, C.J., Schmidt, K.C., Christian, B.T., Muzic, R.F., Fisher, R.E.: Kinetic modeling in positron emission tomography. Emission tomography: The fundamentals of PET and SPECT 46(1), 499–540 (2004)
  • [20] Murray, J.: Parameter space for Turing instability in reaction diffusion mechanisms: a comparison of models. J. Theoretical Biology 98, 143–162 (1982)
  • [21] Nagy, J., Hansen, P.: Deblurring Images. SIAM, Philadelphia (2006)
  • [22] Panfilov, A., Dierckx, H., Volpert, V.: Reaction–diffusion waves in cardiovascular diseases. Physica D: Nonlinear Phenomena 399, 1–34 (2019)
  • [23] Rahmim, A., Lodge, M., Karakatsanis, N., et al.: Dynamic whole-body pet imaging: principles, potentials and applications. European Journal of Nuclear Medicine and Molecular Imaging 46, 501–518 (2019). https://doi.org/10.1007/s00259-018-4153-6, https://doi.org/10.1007/s00259-018-4153-6
  • [24] Ruthotto, L., Haber, E.: Deep neural networks motivated by partial differential equations. arXiv preprint arXiv:1804.04272 (2018)
  • [25] Ruuth, S.: Implicit-explicit methods for reaction–diffusion problems. In preparation (1993)
  • [26] Ruuth, S.: Efficient Algorithms for Diffusion-Generated Motion by Mean Curvature. Ph.D. thesis, Institute of Applied Mathematics, University of British Columbia (1996)
  • [27] Schnackenberg, J.: Simple chemical reaction systems with limit cycle behaviour. J. Theoretical Biology 81, 389–400 (1979)
  • [28] Scipioni, M., Pedemonte, S., Santarelli, M.F., Landini, L.: Probabilistic graphical models for dynamic pet: A novel approach to direct parametric map estimation and image reconstruction. IEEE Transactions on Medical Imaging 39(1), 152–160 (2020). https://doi.org/10.1109/TMI.2019.2922448
  • [29] Seo, Y., Defferrard, M., Vandergheynst, P., Bresson, X.: Structured Sequence Modeling with Graph Convolutional Recurrent Networks. In: International Conference on Neural Information Processing. pp. 362–373. Springer (2018)
  • [30] Song, Y., Ermon, S.: Generative modeling by estimating gradients of the data distribution. Advances in Neural Information Processing Systems 32 (2019)
  • [31] Tolles, J., Luong, T.: Modeling epidemics with compartmental models. JAMA 323(24), 2515–2516 (2020). https://doi.org/10.1001/jama.2020.8420
  • [32] Trotter, J., Pantel, A.R., Teo, B.K.K., Escorcia, F.E., Li, T., Pryma, D.A., Taunk, N.K.: Positron emission tomography (pet)/computed tomography (ct) imaging in radiation therapy treatment planning: A review of pet imaging tracers and methods to incorporate pet/ct. Advances in Radiation Oncology (2023). https://doi.org/10.1016/j.adro.2023.101212, https://doi.org/10.1016/j.adro.2023.101212, open Access
  • [33] Turing, A.M.: The chemical basis of morphogenesis. Bulletin of mathematical biology 52(1), 153–197 (1990)
  • [34] Wang, G., Rahmim, A., Gunn, R.N.: Pet parametric imaging: Past, present, and future. IEEE Transactions on Radiation and Plasma Medical Sciences 4(6), 663–675 (2020). https://doi.org/10.1109/TRPMS.2020.3025086
  • [35] Wu, Z., Pan, S., Chen, F., Long, G., Zhang, C., Philip, S.Y.: A comprehensive survey on graph neural networks. IEEE Transactions on Neural Networks and Learning Systems (2020)