Brought to you by:
Paper

Group sparse-based Taylor expansion method for liver pharmacokinetic parameters imaging of dynamic fluorescence molecular tomography

, , , , , , , , , and

Published 14 May 2024 © 2024 Institute of Physics and Engineering in Medicine
, , Citation Yansong Wu et al 2024 Phys. Med. Biol. 69 115006 DOI 10.1088/1361-6560/ad4084

0031-9155/69/11/115006

Abstract

Objective. Pharmacokinetic parametric images obtained through dynamic fluorescence molecular tomography (DFMT) has ability of capturing dynamic changes in fluorescence concentration, thereby providing three-dimensional metabolic information for applications in biological research and drug development. However, data processing of DFMT is time-consuming, involves a vast amount of data, and the problem itself is ill-posed, which significantly limits the application of pharmacokinetic parametric images reconstruction. In this study, group sparse-based Taylor expansion method is proposed to address these problems. Approach. Firstly, Taylor expansion framework is introduced to reduce time and computational cost. Secondly, group sparsity based on structural prior is introduced to improve reconstruction accuracy. Thirdly, alternating iterative solution based on accelerated gradient descent algorithm is introduced to solve the problem. Main results. Numerical simulation and in vivo experimental results demonstrate that, in comparison to existing methods, the proposed approach significantly enhances reconstruction speed without a degradation of quality, particularly when confronted with background fluorescence interference from other organs. Significance. Our research greatly reduces time and computational cost, providing strong support for real-time monitoring of liver metabolism.

Export citation and abstract BibTeX RIS

1. Introduction

Monitoring liver metabolism poses a significant challenge due to the liver's compensatory abilities (Stanger et al 2007). Dynamic fluorescence molecular tomography (DFMT) provides a efficient approach for monitoring liver metabolism owing to non-radiation, non-invasive, high sensitivity, high specificity, and cost-effectiveness. This technique involves monitoring the absorption, distribution, and metabolism of specific targeted probes. Due to the pharmacokinetic rates of fluorescence within tissues, it can provide additional specificity and quantitative physiological and pathological information for assessing organ function (Zhang et al 2018). DFMT boundary measurements can be transformed into images of pharmacokinetic parameters through the application of compartmental models (Milstein et al 2005, Shinohara et al 1996, Gurfinkel et al 2000), which are known as parametric images and can provide valuable physiological information (Alacam and Yazici 2009). Ultimately, DFMT allows for highly accurate characterization of pharmacokinetics in small animals (James and Gambhir 2012, Zhang et al 2022, Wang et al 2023).

Over the past few decades, DFMT has been applied as an extended application of FMT, and some optical transmission models and reconstruction algorithms on FMT have been applied to it. Due to the high scattering characteristics of light in biological tissues and the limitation of surface fluorescence data, the reconstruction of DFMT has always been a challenging problem (Wang et al 2004). At present, the reconstruction methods of DFMT mainly include data-driven methods (Mu and Zeng 2019) and model-driven methods (Qin et al 2009, He et al 2010, 2018). Model-driven methods is mainly include the norm regularization methods, which focused on the typical prior (smooth for L2, sparse for L1) of reconstruction target (Qin et al 2009, He et al 2010, 2018). Moreover, structural prior has also become one of the hot topics in current research for morphological reconstruction of fluorescence distribution. And data-driven methods aim to learn a mapping relationship from input data to label data, in order to accurately predict the corresponding label when given new input data (Mu and Zeng 2019). Compared to model-driven approaches, data-driven methods have advantages in terms of speed and robustness. However, they still face challenges related to model prior information, model interpretability, data domain adaptation, as well as training computational and time cost. Therefore, the reconstruction methods of DFMT mainly rely on model-driven methods.

At present, the model-driven reconstruction methods of DFMT problems mainly divide into indirect method and direct method. The indirect method holds the assumption that the concentration of the fluorophores remains constant during the data acquisition process (Zhang et al 2013). The indirect method is simple and easy to implement, but the parametric image obtained by this method is often low spatial resolution (Alacam and Yazici 2009, Wang and Qi 2009). Therefore, the direct method is proposed (Zhang et al 2014), which allows the fluorophores concentration in each projection to change, thus fully utilizing the temporal correlation of boundary measurements to further improve the reconstruction quality of parameter images (Zhang et al 2015b). Although the direct method improves the reconstruction performance, the reconstruction time increases. Considering that monitoring liver metabolism is online procedure, it is crucial for accelerating the speed of reconstruction. To deal with issues, GPU-accelerated is the efficient solution (Chen et al 2016), which would increase the cost of equipment. Thus, many researchers proposed dimensionality reduction techniques, such as principal components analysis (PCA) (Cao et al 2013, Zhang et al 2015a), the piecewise-constant level-set (Liu et al 2010b) and the shape-based tomographic reconstruction method (Gottam et al 2019). However, these dimensionality reduction methods will lead to loss of reconstruction accuracy. Therefore, it is still a challenge to improve the reconstruction speed with better reconstruction performance.

In the field of DFMT, the distribution characteristics of fluorescence change over time exhibit a certain degree of spatiotemporal continuity and diversity. Additionally, it holds a wealth of structural prior information and internal connections, such as structural sparse regularization prior (Jiang et al 2016, Liu et al 2017, Zhao et al 2021, Guo et al 2022). Hence, the primary challenge in this study is to devise an efficient, suitable and speedy reconstruction method based on these prior knowledge. Therefore, we propose a group sparse-based Taylor expansion (GSTE) method. Firstly, Taylor expansion framework of temporal sequences signal is introduced to reduce time and computational cost. Secondly, group sparsity based on structural prior is introduced to improve reconstruction accuracy. On this basis, alternating iterative solution based on accelerated gradient descent (AGD) algorithm (Qiao et al 2019) is introduced to solve this problem. In order to verify the performance of the method, we conducted numerical simulations and in vivo experiments, which prove that GSTE method reduces time cost without degradation of quality.

The paper is organized as follows, section 2 describes the proposed acceleration method for DFMT and the evaluation metrics. Section 3 shows the setups of numerical simulations and mouse experiments. Section 4 shows the results of numerical simulations and mouse experiments. Finally, section 5 discusses the results of the study and draws conclusions.

2. Methods

2.1. DFMT inverse reconstruction model

The forward problem of DFMT can be described by a set of coupled diffusion equations (Joshi et al 2006, Wang and Wu 2012), as shown in equation (1)

Equation (1)

where Ω and δ(r) represent imaging region and the Dirac function, respectively, and Θ is a calibration factor, which accounts for the unknown gain of the DFMT system. The subscripts x and m indicate the excited and emitted progressing, respectively. Dx and Dm represent optical diffusion coefficient of excited and emitted progressing, respectively. μax and μam represent absorption coefficient of excited and emitted progressing, respectively. η μaf (r, t) represents the unknown fluorescence yield distribution to be reconstructed.

Solving this equation by the finite element method yields the following linear relationship between the photon flow rate distribution Φ(t) in the biological tissue during the emission of the fluorescent light source and the distribution vector X(t) of the fluorophores concentration in the organism (Schweiger et al 1995), t is the time variation of fluorophores concentration

Equation (2)

where M is the system matrix expressed as m*n, m is the number of measurements on the imaging mouse surface, and n is the total number of nodes in the mouse; X is an n*t dimensional matrix; and Φ is an m*t dimensional matrix.

Our proposed method is mainly aimed at monitoring metabolism in the liver. The distribution of fluorophores X(t) will change due to metabolism in organs and tissues. This change can be described by the biexponential model based on compartmental modeling (Cuccia et al 2003, Milstein et al 2005, Alacam and Yazici 2009)

Equation (3)

where $A={\left[{A}_{1},\cdots ,{A}_{n}\right]}^{T}$, $B={\left[{B}_{1},\cdots ,{B}_{n}\right]}^{T}$, $\alpha ={\left[{\alpha }_{1},\cdots ,{\alpha }_{n}\right]}^{T}$, $\beta ={\left[{\beta }_{1},\cdots ,{\beta }_{n}\right]}^{T}$. And i (i = 1, ..., n) is employed to denote the spatial locations of nodes when a domain is discretized into n nodes. The parameters A (a.u.), B (a.u.), α (min−1) and β (min−1) are the pharmacokinetic parameters describing the metabolic process of indocyanine green (ICG) (Gurfinkel et al 2000), when the gain of the DFMT system is unknown, the pharmacokinetic parameters A and B (a.u.) have arbitrary unit (a.u.) (Zhang et al 2013). α (min−1) and β (min−1) are the uptake and excretion rates of fluorophores, which have physiological significance for quantitative evaluation of organ function (Shinohara et al 1996, Gurfinkel et al 2000).

2.2. Taylor expansion method based on group sparsity

Expanding the double-exponential model in equation (3) with Taylor formula, we can get

Equation (4)

where $\varphi =\left[{A}_{i}-{B}_{i},{A}_{i}{\alpha }_{i}-{B}_{i}{\beta }_{i},{A}_{i}{\alpha }_{i}^{2}-{B}_{i}{\beta }_{i}^{2},\cdots ,{A}_{i}{\alpha }_{i}^{k}-{B}_{i}{\beta }_{i}^{k}\right]$ $=f\left(\left[{A}_{i},{B}_{i},{\alpha }_{i},{\beta }_{i}\right]\right)$, $T={\left[-1,t,-\tfrac{1}{2}{t}^{2},\cdots ,\tfrac{-1}{k!}{\left(-t\right)}^{k}\right]}^{T}$.

In the following method introduction, k is Taylor expansion order (k = 3, 4...). We will verify which order is the optimal choice to balance time, computational cost and reconstruction accuracy through experiments. At this time, combined with equations (2) , (3) and (4), we convert the problem into the following form

Equation (5)

Based on equation (5), combined with the typical DFMT problem, we can construct the following objective function

Equation (6)

L is the weight matrix obtained from anatomical structure (Schulz et al 2009).

Considering the metabolism of ICG by other organs, we define fluorophores concentration as a linear combination of liver region fluorophores concentration and other region fluorophores concentration

Equation (7)

among them, X1 represents the fluorescence yield in the liver, and X2 represents the total fluorescence yield in all organs outside the liver and muscles of mouse, X1 and X2 are all n*t dimensional matrix

Equation (8)

among them, φ1 represents the distribution of pharmacokinetic parameters within the liver, and φ2 represents the total distribution of pharmacokinetic parameters in all organs and tissues outside the liver of mouse. φ, φ1 and φ2 are all the n*(k+1) matrix, and k is Taylor expansion order.

Considering the prior information of histomorphology and the group sparsity of fluorescence yield distribution in liver region, we combine equations (6) and (8) to construct the following objective function

Equation (9)

2.3. Alternating iterative solution based On AGD

To solve equation (9), we use alternating optimization based on AGD (Qiao et al 2019). The objective function can be split into the following two sub problems

Equation (10)

Equation (11)

and λ1 and λ2 are regularization parameter of L2,1 norm regularization term and Frobenius norm regularization term, respectively.

The two sub-problems can be generalized to the following form

Equation (12)

where φ* represents φ1 or φ2, $\tilde{{\rm{\Phi }}}{\left(t\right)}_{\mathrm{new}}$ represents $\tilde{{\rm{\Phi }}}(t)-M{\varphi }_{2}$ in equation (10), or $\tilde{{\rm{\Phi }}}(t)-M{\varphi }_{1}$ in equation (11). P represents L2,1-norm or F-norm.

AGD algorithm is used to solve this problem (Qiao et al 2019)

Equation (13)

where ${\gamma }^{j}=\max \left(2{\gamma }^{j},{\parallel M\left({\varphi }^{j}-{s}^{j}\right)\parallel }_{F}/{\parallel {\varphi }^{j}-{s}^{j}\parallel }_{F}\right)$, $\phi =\tfrac{1}{2}{\parallel M\varphi -\tilde{{\rm{\Phi }}}{\left(t\right)}_{\mathrm{new}}\parallel }_{F}^{2}$. j is iterations of alternating optimization.

When φ satisfying the condition is calculated, we can use φ to solve pharmacokinetic parameters A, B, α, β, that is, to solve the following equations

Equation (14)

Finally, the distribution of pharmacokinetic parameters A, B, α, β in mouse can be obtained. Flow chart of the GSTE method is shown in figure 1. And more details about alternating iterative solution based on AGD are described in section S2 of supplementary data.

Figure 1.

Figure 1. Flow chart of GSTE method.

Standard image High-resolution image

3. Experimental settings

3.1. Numerical simulations

In the numerical simulation, the digimouse Atlas (Dogdas et al 2007) is employed to construct a 3D simulation model. As shown in figure 2(a), The 3D digital mouse is divided into muscle, liver, kidneys, and lungs with specific optical parameters (Ale et al 2012). Their relevant optical properties are shown in table S1 of the supplementary data, where the excitation wavelength is 650 nm and the emission wavelength is 670 nm. The excitation points, represented by the black points, are located one transport mean free path beneath the surface of mouse. As shown in figure 2(c), the fluorescence signals are acquired from the other side of the laser illumination, within a 120° field of view (FOV). As shown in figure S1 of supplementary data, the imaging small animal needs to be continuously rotated for multiple circles (Liu et al 2010a), in order to monitor the metabolic process of fluorescent probe in the body. In each circle, eight projections are acquired during the rotational process of the imaging small animal, and each projection can produce lots of measurement points. We will concatenate the eight measurements matrices from the 8 views along the direction of matrix column as the fluorescence signals measurements at one time point. Figure 2(b) shows the fluorescent probe concentration curve, simulating metabolic processes of fluorescent probe in the mouse within 180 minutes. These fluorescent probe concentration curves are generated based on equation (3) and the corresponding pharmacokinetic parameters in table 1. In figure 2(d), we present the ground truth of the parametric images A (a.u.), B (a.u.), α (min−1) and β (min−1) of the digital mouse obtained according to table 1 (Hillman and Moore 2007, Chen et al 2011) at z = 16.5 mm.

Figure 2.

Figure 2. Numerical simulation settings. (a) The 3D digital mouse model used in the simulation. The mouse torso from the neck to the bottom of the kidneys is selected as the investigated region, totally 34 mm in length. Four parts are included in the model: liver, lungs, kidneys and muscle. We set up 8 excitation points in digimouse at z = 16.5 mm shown as black points. (b) Fluorescent probe concentration curves simulating the metabolic processes of fluorescent probe in different metabolic regions, the corresponding pharmacokinetic parameters of the curves are shown in table 1. (c) The fluorescence signals are acquired from the other side of the laser illumination, within a 120° field of view (FOV). (d) The true parametric images A (a.u.), B (a.u.), α (min−1) and β (min−1) of the digimouse at z = 16.5 mm are obtained according to table 1. The black points represent the excitation point source locations.

Standard image High-resolution image

Table 1. Pharmacokinetic parameters used in the numerical simulation.

Regions A (a.u.) B (a.u.) α (min−1) β (min−1)
Liver1.01.00.4350.011
Lungs0.80.80.2960.02
Kidneys1.31.30.2540.016
Other tissues0.50.50.3480.009

The digital mouse model is discretized into a uniform tetrahedral mesh consisting of 6096 nodes and 30 123 tetrahedral elements for reconstruction in Amira (Amira, Visage Imaging, Australia). In order to verify the performance of this method, the following experiments are designed as follows: Firstly, different order of Taylor expansion are adopted for comparison to determine the order of Taylor expansion that achieves the best performance with the method. Secondly, different time intervals within a metabolic time of 180 min (10 800 s) are used to find the optimal fluorescence data collection interval for the best performance. Finally, GSTE method proposed in this article compared with direct method using graphics processing units (direct method (GPU)) (Chen et al 2016), indirect method (Zhang et al 2013), and TE method, respectively. Direct method (GPU) is a very reliably accurate method for reconstructing parametric images in DFMT state of art. Direct method (GPU) reduces the reconstruction time by ∼90% without a degradation of quality compared with the method proposed by predecessors. The speed of indirect method in reconstructing parametric images is fast, but its accuracy is far inferior to the direct method; TE method only use Taylor expansion framework, AGD algorithm and the L2-norm as the regularization term of the objective function, without taking advantage of the group sparsity derived from the structural prior.

In order to verify the reconstruction accuracy and speed of our proposed GSTE method, time cost is recorded to evaluate reconstruction speed among the methods, and relative difference(RD) is calculated to evaluate the reconstruction accuracy among the methods. Each study is conducted five times to ensure stability of method. And we randomly add 13dB noise to each group of experiments

Equation (15)

3.2. In vivo experiments

In vivo experiments are conducted on 7 week old BALB/c nude mice (Keaoke Biotechnology Co., Ltd, Xi'an, China). All operations followed the Laboratory Animal Management and Welfare Ethics, Northwest University, China (Permit Number: NWU-AWC-20210901 M). To reduce the pain of mice during the experiment and reduce the data collection error, mice are anesthetized with 2% isoflurane–air mixture gas (RWD Life Science Co., Ltd, Shenzhen, China) throughout the experiment. As shown in figure 3(a), our experiment is mainly conducted in a multimodal optical imaging system. The mice are placed on a rotating turntable, and then, fluorescence images are taken. The excitation source is obtained by a continuous-wave semiconductor laser at 780 nm with a power of 450 mW, and fluorescence imaging is achieved using an 840 nm emission filter. Their relevant optical properties are shown in table S1 of the supplementary data, where the excitation wavelength is 780 nm and the emission wavelength is 840 nm. Then, we inject a prepared ICG solution (0.02 ml of 1.5 mg ml−1) into the tail vein of mouse. All fluorescence images are composed of a 512 × 512 pixel, Recorded by a CCD camera cooled at −80°C (Andor, Belfast, Northern Ireland, UK). The exposure time of CCD is set to 0.1 s, and the total imaging time of DFMT is about 180 min. Each FMT fluorescence image is collected every second, we collected a total of 10 800 FMT images, taking 180 min in total.

Figure 3.

Figure 3. Multimodal optical imaging system and in vivo experimental settings and data. (a) The multi-modal optical imaging system. (b) Mean fluorescence intensity on surface collected within 180 min. (c) The white light image of mouse. (d) Representative coronal x-ray CT slice. (e)–(g) Segmentation results for transversal x-ray CT slices indicated by the green, cyan and purple dashed lines in (d), all transversal x-ray CT slices getting from the mouse torso in (d) are manually segmented into the liver, lungs, kidneys and other tissues. (h1)–(h5) Fluorescent images of five representative frames. (i1)–(i5) The registration results of fluorescence signals from (h1)–(h5) with the CT data from (d).

Standard image High-resolution image

After collecting the fluoresence data, two doses of iohexol are injected into the mouse body, with an intraperitoneal injection of 0.3 ml each time, with an interval of 30 min between the two doses, to enhance the anatomical structure of the liver in micro-CT imaging. 30 min after the second injection, the mouse is immobilized and scanned using an x-ray source with a micro-focused cone beam (L9181-02, Hamamatsu Photonics, Hamamatsu Japan). In the x-ray scanning, the system is operated at a source voltage of 90 kV and power of 27W. An x-ray flat-panel detector (C7942CA-22, Hamamatsu Photonics, Hamamatsu Japan), employed in high-resolution CT imaging, is used to obtain a total of 600 x-ray projections with an interval of 0.6° and integrating time of 0.5 s. The 3D anatomical structures are segmented from the CT data (Dogdas et al 2007), Amira is used to segment the lungs, liver, kidneys and muscle, then, we use the segmentation results constructing a Laplacian matrix (Schulz et al 2009).

Figure 3(b) shows mean fluorescence intensity on surface collected within 180 min, and the fitting curves are represented by green lines, reflecting the metabolic pattern of ICG in mouse. Five representative frames are selected at 5, 33, 63, 120, and 150 min. Figure 3(h1)–(h5) show the corresponding fluorescence images captured by the CCD camera. In addition, figure 3(i1)–(i5) show measured data mapped on the three-dimensional surface of the mouse torso.

Similarly to the numerical simulation, in vivo mouse model is discretized into a uniform tetrahedral mesh consisting of 6115 nodes and 28 917 tetrahedral elements for reconstruction. In order to verify the performance of GSTE method in vivo experiments, we use the optimal parameters obtained from numerical simulations (i.e. the optimal order of Taylor expansion and the optimal fluorescence data collection interval for 180 min data collection). Then, We compare the performance of GSTE method with direct method (GPU), indirect method, and TE method in vivo experiments. Each study is conducted five times to ensure stability of method.

4. Experimental results

4.1. Numerical simulations

4.1.1. Parameter selection experiments

To assess the impact of the Taylor expansion order and fluorescence data collection interval for 180 min data collection on the GSTE method, numerical simulations with different Taylor expansion orders and fluorescence data collection intervals are conducted to select the optimal Taylor expansion order and fluorescence data collection interval for 180 min data collection from reconstructed results. Figure 4(a) shows RDs and time cost of the pharmacokinetic parameters reconstructed by the GSTE method when the k value is set to 3, 4, 5, 6 and 7, respectively. It can be seen that with the increase of the order, the RDs of the reconstruction gradually decreases and the reconstruction time gradually increases. A compromise method is to choose a suitable k to balance reconstruction accuracy and time. We found that when k is bigger than 4, the reduction of RDs becomes less significant. However, the increase of reconstruction time is still obvious when k is bigger than 4. Therefore, an appropriate trade-off between reconstruction accuracy and time can be achieved when k = 4.

Figure 4.

Figure 4. Time(s) consumptions and RDs of the different Taylor expansion orders and fluorescence data collection intervals for simulations.

Standard image High-resolution image

Figure 4(b) shows the RDs and time cost of the pharmacokinetic parameters reconstructed by the GSTE method when the fluorescence data collection interval for 180 min data collection is set to 540, 360, 270, 216 and 180, respectively (i.e. the number of columns corresponding Φ(t) is 20, 30, 40, 50, 60). It should be noted that the RDs here is the mean of the RDs of the four pharmacokinetic parameters. It can be seen that when the fluorescence data collection interval is smaller, that is, the surface fluorescence measurement data is richer, the RDs of the pharmacokinetic parameters reconstruction is lower, and the time cost of reconstruction is higher. When the fluorescence data collection interval for 180 min data collection is less than 270, the RDs has been reduced to 0.06 and gradually stabilized. When the fluorescence data collection time interval is greater than 270, the increase of reconstruction time is still obvious. Therefore, considering the reconstruction time and reconstruction accuracy, the fluorescence data collection interval is set to 270 for 180 min data collection (i.e. the number of columns corresponding Φ(t) is set to 40).

4.1.2. Comparison of pharmacokinetic parameters reconstruction using different methods

In order to test the performance of GSTE method, after obtaining the optimal Taylor expansion order and the optimal fluorescence data collection interval for 180 min data collection, the GSTE method is compared with other methods on the basis of using these optimal parameters.

Figure 5(a) shows the time cost of the four methods after taking the logarithm. We can get such information from the figure: the GSTE method and TE method which use Taylor expansion framework is much faster than the other two methods, they can reduce the time consumption by ∼90%, which also shows the superiority of the strategy using Taylor expansion framework of temporal sequences signal. However, in figure 5(b), we can see that the pharmacokinetic parameters reconstruction accuracy of GSTE method and direct method (GPU) is much better than the other two methods. The pharmacokinetic parameters reconstruction accuracy of indirect method is worst. GSTE method reduces the RDs by ∼45% compared to TE in average and almost keep the same RDs with direct method (GPU). The detailed experimental data of time consumptions and RDs in numerical simulations are shown in table 2. In order to further compare the pharmacokinetic parameters reconstruction accuracy of GSTE method and direct method (GPU) in each organ, we listed the 3D parametric images of the RDs between the reconstructed results of GSTE method and the ground truth, as shown figure 5(c1)–(c4), and the cross-sectional images at Z = 16.5 mm of figure 5(c1)–(c4), as shown figure 5(d1)–(d4), the 3D parametric images of the RDs between the reconstructed results of direct method (GPU) and the ground truth, as shown figure 5(e1)–(e4), and the cross-sectional images at Z = 16.5 mm of figure 5(e1)–(e4), as shown figure 5(f1)–(f4). The results demonstrate that, the pharmacokinetic parameters reconstruction accuracy of GSTE method is relatively higher in organs, especially in liver, and the reconstruction accuracy of direct method (GPU) is relatively higher in muscle. Moreover, as shown in figure 5(b), pharmacokinetic parameters reconstruction accuracy in all organs and tissues of GSTE method and direct method (GPU) are almost identical. Therefore, considering both reconstruction accuracy and time cost, the GSTE method performs better than the other three methods.

Figure 5.

Figure 5. (a) Log of time consumptions of the different method. (b) RDs of the different methods for A (a.u.), B (a.u.), α (min−1) and β (min−1). (c1)–(c4) 3D images of RDs with GSTE method and ground truth. (d1)–(d4) Cross-sectional images of the (c1)–(c4). (e1)–(e4) 3D images of RDs with direct method (GPU) and ground truth. (f1)–(f4) Cross-sectional images of the (e1)–(e4).

Standard image High-resolution image

Table 2. Time consumptions and RDs of the different methods in numerical simulations.

  RD (A) RD (B) RD (α) RD (β)Time (s)
GSTE0.032 ± 0.0030.032 ± 0.0020.070 ± 0.0020.102 ± 0.00510.6 ± 0.55
Direct method (GPU)0.029 ± 0.0010.033 ± 0.0050.072 ± 0.0050.096 ± 0.004424.0 ± 3.67
Indirect method0.124 ± 0.0050.124 ± 0.0040.168 ± 0.0030.215 ± 0.004526.8 ± 3.03
TE0.050 ± 0.0100.043 ± 0.0090.151 ± 0.0030.160 ± 0.01210.6 ± 0.55

4.2. In vivo experiments

Similarly to the numerical simulation, in order to verify the performance of GSTE method in vivo experiments, we use the optimal parameters obtained from numerical simulations (i.e. the optimal order of the Taylor expansion and the optimal fluorescence data collection interval for 180 min data collection). Parameter selection experiments in vivo experiments are shown in figure S2 of the supplementary data.

Figure 6(a1)–(a4) shows 3D images of reconstructed results of Full Direct method, because Full Direct method is the most accurate method for reconstructing pharmacokinetic parameters state of art. Figure 6(b1)–(b4) shows Cross-sectional images of figure 6(a1)–(a4), (g) shows the time cost of the four methods after taking the logarithm. We can get such information from the figure: the method using Taylor expansion framework is much faster than the other two methods,they can reduce the time consumption by ∼90%, which also shows the superiority of the strategy using Taylor expansion framework of temporal sequences signal. However, in figure 6(h), we can see that the pharmacokinetic parameters reconstruction accuracy of GSTE method and direct method (GPU) is much better than the other two methods. The pharmacokinetic parameters reconstruction accuracy of indirect method is worst. GSTE method reduces the RDs by ∼42% compared to TE method in average and almost keep the same RDs with direct method (GPU). The detailed experimental data of time consumptions and RDs in vivo experiments are shown in table S3 of the supplementary data. In order to further compare the pharmacokinetic parameters reconstruction accuracy of GSTE method and direct method (GPU) in each organ, we listed the 3D parametric images of the RDs between the reconstructed results of GSTE method and the reconstructed results of Full Direct method, as shown figure 6(c1)–(c4), and the cross-sectional images at Z = 68 mm of figure 6(c1)–(c4), as shown figure 6(d1)–(d4), the 3D parametric images of the RDs between the reconstructed results of direct method (GPU) and the the reconstructed results of Full Direct method, as shown figure 6(e1)–(e4), and the cross-sectional images at Z = 68 mm of figure 6(e1)–(e4), as shown figure 6(f1)–(f4). The results demonstrate that, the pharmacokinetic parameters reconstruction accuracy of GSTE method is relatively higher in organs, especially in liver, and the reconstruction accuracy of direct method (GPU) is relatively higher in muscle. Therefore, considering both reconstruction accuracy and time cost, the GSTE method performs better than the other methods.

Figure 6.

Figure 6. (a1)–(a4) 3D images of the reconstruction results for A (a.u.), B (a.u.), α (min−1) and β (min−1) with Full Direct Method. (b1)–(b4) Cross-sectional images of the (a1)–(a4). (c1)–(c4) 3D images of the RDs of reconstruction results for A (a.u.), B (a.u.), α (min−1) and β (min−1) with GSTE and Full Direct Method. (d1)–(d4) Cross-sectional images of the (c1)–(c4). (e1)–(e4) 3D images of the RDs of reconstruction results for A (a.u.), B (a.u.), α (min−1) and β (min−1) with direct method (GPU) and Full Direct Method. (f1)–(f4) Cross-sectional images of the (e1)–(e4). (g) Log of time consumptions of the different methods. (h) RDs of the different methods for A (a.u.), B (a.u.), α (min−1) and β (min−1).

Standard image High-resolution image

5. Discussion and conclusion

DFMT can be used for tumor detection, drug development, and metabolic research (Milstein et al 2005, Alacam and Yazici 2009, Zhang et al 2013). However, due to the inherent serious ill-posed nature of inverse problem, liver metabolism monitoring using DFMT still has many challenges in biological applications. The previously proposed reconstruction methods can improve reconstruction quality. However, due to the high time and computational cost, the implementation of these reconstruction method is limited.

In this study, the Taylor expansion method based on structural prior proposed in this paper is used to reconstruct pharmacokinetic parameters. Firstly, we use Taylor expansion framework of temporal sequences signal to reduce the computational and time cost. Then, we design a group sparsity model based on the structural prior to improve the reconstruction accuracy obviously. Thirdly, the AGD algorithm is used to solve the problem. For the GSTE method, appropriate parameter should be carefully selected to balance the calculation speed and the reconstruction accuracy (Cao et al 2013). We have selected the optimal Taylor expansion order and fluorescence data collection interval for 180 min data collection through experiments(i.e. Taylor expansion order is set to 4 and fluorescence data collection interval is set to 270 for 180 min data collection). Furthermore, the GSTE method is compared with other methods, the performance of GSTE method is confirmed by numerical simulations and in vivo experiments. The results demonstrate that GSTE method has greatly reduced the time cost on the basis of previous research, with nearly no quality degradation. However, there are still many deficiencies that need to be addressed in our research, such as how to improve the accuracy of pharmacokinetic parameters reconstruction in muscle while ensuring the accuracy of pharmacokinetic parameters in liver and other organs.

In future work, we have the following ideas: firstly, we can embed other algorithm such as fast iterative shrinkage-thresholding algorithm (FISTA) (Beck and Teboulle 2009), incomplete variables truncated conjugate gradient method (IVTCG) (He et al 2010), stochastic gradient descent algorithm (SGD) (Newton et al 2018) into Taylor expansion framework to better balance the reconstruction speed and accuracy. Secondly, we can also use deep learning methods such as long short term memory (LSTM) (Yu et al 2019) or transformer (Vaswani et al 2017) with GSTE method into DFMT to improve reconstruction method performance (Zhang et al 2021, Li et al 2023). Moreover, early monitoring of liver disease requires the ability to distinguish and analyze normal and injured liver tissues. If the reconstruction accuracy of our method reaches a certain level, i hope that our research can devote to the liver injury monitoring, which needs the ability to distinguish and analyze normal and injured liver tissues (Zhao et al 2023). All of these require us to continue conducting experiments to verify their feasibility.

Acknowledgments

National Natural Science Foundation of China, Grant/Award Nos. (11871321, 12271434, 61901374, 61906154, 61971350, 62201459, 62271394, 82 071914); Natural Science Basic Research Plan in Shaanxi Province of China, Grant/Award Numbers 2023-JC-JQ-57.

Data availability statement

Data underlying the in vivo results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request. The data that support the findings of this study are available upon reasonable request from the authors.

Disclosures

The authors declare no potential conflict of interests.

Please wait… references are loading.

Supplementary data (0.3 MB PDF)

10.1088/1361-6560/ad4084