1.IntroductionCerebral blood flow (CBF) is essential for monitoring metabolic oxygenation,1,2 neurovascular coupling,3,4 and metabolic response to functional stimuli.5,6 For example, CBF abnormalities are caused by ischemic strokes,7 head trauma,8 or brain injury.9,10 There are several blood flow measurement techniques, including computed tomography,11 magnetic resonance imaging,12 and positron emission tomography.13 However, although they are well-established, they cannot provide continuous, long-term measurements at the bedside. Laser Doppler flowmetry can measure microvascular blood flow but only probe shallow tissues.14 Doppler ultrasound techniques can only measure blood flow in larger vasculatures and are unsuitable for longitudinal monitoring for unstable probe orientations.15 Near-infrared diffuse optical methods are becoming popular in blood flow measurements as they are noninvasive, nonionized, portable, and faster. Among them is diffuse correlation spectroscopy (DCS),16,17 using a laser with a long coherence length () to illuminate tissue surfaces and collect remitted scattered light at a distance, typically 1 to 3 cm, away from the incident position. The scattered light from flowing red blood cells causes a speckle pattern fluctuating at a rate proportional to the flow rate. This blood-flow-dependent information can be quantified based on the normalized temporal intensity autocorrelation function (ACF) , where is the measured scattered light and is the correlation lag time.17,18 DCS can measure blood flow in vivo in small animals,19,20 human brains,16 and muscles.21 Traditionally, to derive blood flow index (BFi), the measured is fitted with a homogenous semi-infinite one-layer analytical model22 or the Monte Carlo model.23 This fitting process typically utilizes nonlinear least-square methods (NLSMs) with Levenberg–Marquardt optimization or trust-region-reflective methods.24–26 However, treating biological tissues with a homogenous semi-infinite model is not quite realistic, as significant signal contamination from superficial tissue layers (e.g., scalp/skull) occurs when measuring deep flow in the brain. Research has been conducted to minimize the discrepancy, with the diffusion equation for layered geometries developed for fitting methods, including two-27,28 and three-layer analytical models.29,30 Unfortunately, multilayer models highly rely on a priori knowledge of each layer’s optical properties (namely the absorption coefficient and reduced scattering coefficient ) and thickness to estimate blood flow within each layer. Commonly, layer optical properties and thicknesses are assumed from literature, and the errors in these assumed values can lead to significant errors in brain blood flow estimations. Additionally, the multilayer model is susceptible to measurement noise, especially for the three-layer model, although its accuracy in BFi estimations has been validated.24,31 Moreover, these methods are iterative and time-consuming. To overcome these limitations, the ’th-order linear (NL) algorithm,32,33 least-absolute minimization (L1 norm), and the support vector regression (SVR)34 were proposed. However, under the NL framework, BFi extraction is significantly influenced by the linear regression approach adopted.34 Although L1 norm and SVR are new approaches to processing DCS data, they are sensitive to signal deviations.35,36 Additionally, the BFi computing time is 28.07 and 52.93 s (using L1 norm and SVR, respectively), still slow for practical applications, particularly for real-time monitoring.34 Deep learning, an increasingly popular method, has been widely applied to biomedical time sequence data, including electroencephalogram (EEG) and electrocardiogram (ECG),37,38 but has yet to be broadly used in DCS. Very recently, Zhang et al.39 proposed the first recurrent neural network (RNN) regression model to DCS, followed by 2D convolution neural networks (2D CNNs),40 long short-term memory (LSTM),41 and ConvGRU.42 LSTM, as a typical RNN structure, has proven stable and robust for quantifying relative blood flow in previous studies in phantom and in vivo experiments.41 2D CNN, on the other hand, tends to require large training datasets for complex structures, demanding massive memory resources. ConvGRU, the newest deep learning method introduced to DCS, has also exhibited excellent performances in BFi extraction. Nevertheless, all existing algorithms are designed for a single source–detector distance (), corresponding to a specific depth in biological tissues. To accommodate a wider range of , retraining the model becomes necessary. Inspired by a recently published one-dimensional convolutional neural network (1D CNN)43 for fluorescence lifetime imaging (FLIM), we proposed the DCS neural network (DCS-NET) based on 1D CNN for quantifying the coherent factor and BFi. The primary objective of this work is to present and evaluate an artificial intelligence (AI) framework, called DCS-NET, in β and BFi estimations. We established the Monte Carlo simulation model based on the open-source tool Monte Carlo eXtreme (MCX) developed by Fang and Boas44 to generate emulating experiment data. The DCS-NET training, validation, and testing datasets are from the semi-infinite geometry model.22 We investigated DCS-NET’s performance on absolute BFi and relative BFI (rBFi)’s estimations and compared them with semi-infinite and three-layer model fitting methods. To best link our work with actual outcomes expected in practice, we modeled DCS measurement noise based on realistic experimental conditions, considering various noise levels controlled by the integration time (). We define a metric that accounts for the intrinsic sensitivity of the brain blood flow and evaluate it between DCS-NET and traditional fitting methods. We also show BFi estimation errors induced by the inaccurate assumptions about layer optical properties and thicknesses when using fitting methods based on the semi-infinite and three-layer solutions of the correlation diffusion equation. Figure 1 summarizes the main concept of our work. All essential parameters are defined in Table 6 in the Appendix to facilitate our discussion. 2.Methods2.1.DCS TheoryThe transport of the unnormalized electric field auto-correlation function, , is well described by the correlation diffusion equation:17,45 where is the wavenumber of light, and are the refractive index and wavelength in the scattering medium, respectively. is the fraction of dynamic photon scattering events in the medium. is the mean squared displacement of scatterers in the turbid medium during a time interval . is the point source located at ; is the source–detector distance. and are the tissue’s absorption and reduced scattering coefficients, respectively. For a semi-infinite medium, the solution of Eq. (1) using the extrapolated boundary condition for continuous-wave DCS is where , , , , and to be consistent with Ref. 46. Previous studies have shown that the scatters’ Brownian diffusion motion model18,47 aligns well with in vivo DCS experiments, and therefore, the mean-squared displacement can be derived as , where represents the effective diffusion coefficient. BFi in DCS is typically defined as .21,48 is linked to the normalized electric field auto-correlation function as where is a constant accounting for the collection setup, such as the number of detected speckles and the numerical aperture of the detection fiber.However, realistic biological tissues49 show multiple layers with different physiological and optical properties. Using DCS to conduct in vivo CBF measurements, light must propagate through different layers, including the scalp and skull.50,51 Thus, layered analytical models have been proposed for BFi extraction. These include the two-27,28 and three-layer analytical models.24,30,31,52 This study considers the three-layer analytical model, where a turbid medium consisting of slabs is considered, as shown in Fig. 2(c). Each slab has its thickness, , , 2, 3, where are the coordinates along the -axis and , and are absorption and scattering coefficients. To solve Eq. (1) in the layered medium (along direction), we can use the Fourier transform for the transverse coordinate as where is the radial spatial frequency. Equation (1) can then be rewritten as where , , and .We divided the top layer into two sublayers: layer 0 () identified by , and layer 1 (). Then, the solution of Eq. (5) inside the ’th layer () can be written as where and are coefficients for each layer determined by the boundary conditions where and are the extrapolation lengths accounting for internal reflections at the tissue surface () and the back surface (), respectively. is the photon diffusion coefficient in layer , and is the speed of light.Substituting Eq. (6) into Eq. (7), and can be determined (, 2, 3), and we obtain the solution of Eq. (5) at as where Num and Denom (when and ) areTherefore, by performing the inverse Fourier transform of Eq. (8) with respect to , the field ACF at can be written as where denotes the zero-order Bessel function of the first kind. The integral bound for in Eq. (11) should theoretically be from 0 to . However, in practice, the numerical integration is performed with a limited range as advised in Ref. 29.2.2.Noise ModelsThis study evaluates the impact from noise on BFi and . We employed a broadly accepted noise model proposed by Zhou et al.53 The standard deviation (, noise) of is given as where is the bin width of the correlator, is the bin index corresponding to . is the average number of photons detected within the bin time, where is the detected photon count rate, and is the integration time (e.g., measurement duration). is the decay rate of , which is obtained from fitting the measured to the theoretical . Gaussian noise54,55 was added to based on a statistical noise model to determine the noise (). Considering realistic photon budgets, the photon count rate at 785 nm was assumed to be 8.05 kcps.55 Three different noise levels were defined according to (= 1, 10, or 30 s).2.3.Intrinsic Sensitivity EstimationTo evaluate the sensitivity to changes in blood flow in the deeper layer, we fixed the effective diffusion coefficient in layer 1 and increased in layer 3 as , is an integer and . The physiological and optical parameters listed in Table 1 are taken as baseline conditions. Similar to Ref. 54, the intrinsic sensitivity () is defined as where and represent the estimated BFi (, , or , meaning DCS-NET, the semi-infinite, and three-layer fitting methods) for the perturbed and baseline conditions, respectively, and and are in layer 3 for the perturbed and baseline conditions, respectively.Table 1Physiological and optical parameters56 at 785 nm in the human head model.
2.4.Monte Carlo SimulationsWe utilized a simplified model comprising three layers to emulate the scalp (5 mm), skull (7 mm), and brain (50 mm, large enough so that we can treat the medium as semi-infinite), respectively.57 All layers were assumed homogeneous, as demonstrated in Fig. 2(a), and their corresponding optical properties are summarized in Table 1. MCX utilized an anisotropic factor () of 0.89 and a refractive index () of 1.3744 for all layers. We launched photons from a source with a diameter of 1 mm and set the detector radii to 0.13, 0.28, 0.45, 0.7, 1, and 1.5 mm for , 10, 15, 20, 25, and 30 mm, respectively, recording data from multiple distances simultaneously. An example of the source and the detector was arranged as shown in Fig. 2(a). MCX records the path lengths and momentum transfer from the detected photons for obtaining the electric field ACF :22 where is the number of detected photons, is the number of tissue types (3 for our simulations), and and stand for the total momentum transfer and the total path length of photon in layer , respectively. is the absorption coefficient, and is the mean square displacement of the scattered particles in layer . Here, , where is the effective diffusion coefficient of layer . The simulated is normalized to , and then we can obtain using the Siegert relationship with . In this simulation, the delay time (127 data points) was used for .2.5.Deep Learning Architecture DesignThe structure of DCS-NET is shown in Fig. 3(a). DCS-NET takes to estimate and BFi independently. DCS-NET consists of (1) a shared branch for temporal feature extraction and (2) two subsequent independent branches for estimating and BFi, with a similar structure to the shared branch. The two CNN layers in the shared branch have a wider sliding window with a larger kernel size of 13 and a giant stride of 5. They are expected to capture more general features of the auto-correlation decay curves. The batch normalization (BN) layer58 is employed after each convolutional layer. It reduces the shift of internal covariance and accelerates network training when processing normalized data. To implement feature pooling and effectively reconstruct and BFi, we use a pointwise convolution layer with a kernel size of 1 after the convolutional neural network, followed by the activation function, the Sigmoid function. The model input is measured (here, we used data from MCX) , of which the size is . Both the estimated and BFi have a size of . Note that the simulated was normalized to (0, 1] before being fed into the model. 2.6.Training Dataset PreparationThe training datasets can be easily obtained using synthetic data based on the homogenous semi-infinite analytical model, as shown in Fig. 2(b). Thus, according to Eqs. (2) and (3), 200,000 training datasets () were generated and split into the training (80%) and the validation (20%) groups. Each dataset consists of the input, , and its corresponding labels are BFi and , which are the output. The training batch size is 128, with 800 training epochs. We used an early stopping callback with 20 patient epochs to prevent overfitting. To match the realistic experiments, in the dataset, we set , , , , and , where stands for a uniform distribution. training datasets contain noisy and noiseless (the noise model has been described in Ref. 53) ACFs, as shown in Fig. 3(c). The green, yellow, and red lines represent noisy , and the blue line represents noiseless . We used the optimizer Adam59 for the training process, with the learning rate fixed at in the standard back-propagation. We used the mean square error loss function for updating the network by controlling the following problem: where is the network output (estimated BFi or ), and is the corresponding label (true BFi or ) in the ’th training pairs. is the mapping function, is the trainable weights of our networks, and is the number of training pairs. Figure 3(b) shows that the training and validation losses decrease rapidly and reach the plateau after 85 epochs. The training process’s best score reaches a small value of 0.000725, indicating that the network is well trained as the estimated and BFi are close to the ground truth. The model was conducted in Python using Pytorch with Intel (R) Core (TM) i9-10900KF CPU @3.70 GHz.3.Results3.1.Absolute BFi Recovery Versus Detection DepthsTo investigate how the absolute BFi and behave in terms of among DCS-NET, semi-infinite, and three-layer fitting approaches, we generated via MCX Monte Carlo simulations for , 10, 15, 20, 25, and 30 mm, as described in Sec. 2.4. Table 1 shows all the relevant parameters used in MCX simulations. The absolute BFi in this study corresponds to the Brownian diffusion coefficient (assumed ). When using DCS-NET, was fed into the pre-trained model. For the semi-infinite fitting procedure, was fitted to Eqs. (2) and (3), and we assumed , , for the brain layer (layer 3), as provided in Table 1. We also fitted the simulated with the three-layer model, Eqs. (11) and (12), and , , , , , , , , , , and . Meanwhile, we set and as the initial guesses. For the fitting, we used NLSM ( in MATLAB with the Levenberg–Marquardt optimization) to minimize the unweighted least squares objective function, where is the number of sampled , and is from Eq. (3) or Eq. (12). Fitting was performed on from 1 to .Table 2 presents the true and BFi and estimated and BFi using DCS-NET, semi-infinite, and three-layer fitting methods. All input parameters for fitting are assumed as described above, and . We define , , and (also , , and ) for DCS-NET, the semi-infinite, and three-layer fitting methods, respectively. We define , where is the BFi error with DCS-NET. Similarly, and are the BFi estimated errors with the semi-infinite and three-layer fitting methods. Table 2BFi in the brain estimated using DCS-NET, homogeneous semi-infinite and three- layer fitting models.
Table 2 shows when the semi-infinite model is used, the estimated BFi is closer to layer 1 (), even for , suggesting that a homogenous fitting procedure is more sensitive to the superficial layers’ dynamic properties. This finding is consistent with the results reported by Gagnon et al.27 Using the three-layer fitting model, we obtained , close to when . This is because the mean light penetration depth is to .19 When is small, most detected photons predominantly travel through layer 1. As increases (), the estimated BFi decreases, reaching at , with of 6.17%. This is because as increases, the detected photons penetrate inside the skull layer (), resulting in an increased contribution of layer 2. This phenomenon is expected, because the three-layer modeling can remove the contribution from superficial layers52 to obtain accurate BFi. Interestingly, when using DCS-NET, the estimated BFi increases as increases, reaching with of 4.83% at . These results suggest that the AI model is capable of recognizing the depth. Regarding estimation, there is no significant difference among the three methods. 3.2.Absolute BFi Recovery with NoiseFigure 3(c) displays the semi-infinite analytical example curves with noise using the model proposed by Zhou et al.53 The curves were obtained with at different noise levels (), , and with an assumed . To assess DCS-NET’s performance in practical scenarios, we modified the Monte Carlo code to generate curves including noise according to Zhou et al.’s noise model.53 We generated 100 sets for each noise level (including noiseless). Still, we minimized Eq. (17) using the Levenberg–Marquardt optimization routine. We performed the residual analysis to assess the efficiency of the semi-infinite and three-layer models. We define the residual and resnorm (the squared 2-norm of the residual) as where is the lag time index, and is the length of the time trace. is the fitted obtained from fitting methods based on analytical models at the lag time , and the corresponding true value is from MCX. The fitting results using the semi-infinite and three-layer analytical models are presented in Fig. 4, in which noisy curves from MCX (blue star-shaped) and fitted curves (red lines) at different noise levels are shown. Figures 4(a)(i)–4(a)(iv) show the MCX-generated and fitted using the semi-infinite model, and they exhibit an increasing trend in , ranging from to , indicating that the semi-infinite method becomes inaccurate when the noise level increases. Additionally, reaches 3.02 when . Similar behaviors are observed in the three-layer fitting, as shown in Figs. 4(b)(i)–4(b)(iv).We also calculated the mean BFi and over 100 trials. As for , we arrive at the same conclusion as Sec. 3.1 that all three methods exhibit similar behaviors at the same noise level. A high noise level () leads to a significant standard deviation, as shown in Fig. 5(a). Figure 5(b) shows the estimated BFi. The estimated BFi for the semi-infinite model deviates significantly from the ground truth. When using the three-layer fitting method, is 82.30% at the lower noise level (). As the noise level increases, also increases, with reaching 390.10% at the high noise level (). Furthermore, a high noise level leads to a more significant standard deviation, indicating that BFi estimation is highly sensitive to noise when the three-layer fitting method is applied, in accordance with previous findings.52 In contrast, (using DCS-NET) at a high noise level () is 12.87%, whereas at a low noise level (), it is only 1.93%, indicating that DCS-NET is not susceptible to noise. Figure 5(b) also shows that when the three-layer fitting method is used, the BFi precision can be enhanced through increasing . 3.3.Relative Blood FlowIn practice, we do not aim to obtain absolute BFi measurements. Instead, the relative variation in blood flow (e.g., ) is oftener used.19 To evaluate DCS-NET for extracting rBFi in the brain, we assigned , in layer 3 (brain) and fixed in other layers. Figure 6 presents rBFi calculated on noiseless data at . In Fig. 6, rBFi calculated by DCS-NET, the semi-infinite, and three-layer fitting methods on noiseless data for for ranges from to () with a step of . , we define the estimated BFi as at the start point. To compare the accuracy of the three different methods in quantifying rBFi, we defined the error in rBFi as (, , or ), meaning the rBFi estimation error using DCS-NET, the semi-infinite, and three-layer fitting methods, respectively. We can observe that (red star) is close to the true (blue solid line) with ranging from 0.15% to 8.35%. By contrast, the semi-infinite and three-layer methods result in more significant errors of and , respectively. As expected, the semi-infinite homogenous solution resulted in significant errors in rBFi, in agreement with Ref. 33. 3.4.Intrinsic SensitivityAs described in Sec. 2.3, the input in layer 3, denoted as , serves as the base point, and its corresponding recovered BFi is denoted as . Similarly, we assigned ( is an integer; ), and it is referred to as the perturbed blood flow . We also define a perturbation level . We calculated the corresponding BFi for , and then used Eq. (14) to obtain , and . We considered physiological noise by utilizing the noise model described in Sec. 2.2. Figure 7(a) shows the noiseless intrinsic sensitivity, demonstrating that DCS-NET exhibits . The intrinsic sensitivity reaches 2.5 × when , then decreases with increasing. In comparison, the three-layer fitting method achieved , whereas the semi-infinite fitting method yielded of only 14.12% on noiseless data. Figures 7(b)–7(d) illustrate sensitivity curves at various noise levels. Especially noteworthy are the instances where at and . Conversely, with the semi-infinite and three-layer fitting models, predominantly assumes negative values, underscoring the considerable impact of measurement noise on sensitivity. Furthermore, the impact of measurement noise on the sensitivity overgrows, particularly for the three-layer fitting method, as apparent in Fig. 7(d). 3.5.BFi Extraction with Varied Optical Properties and Scalp/Skull ThicknessesIn practical applications, a patient’s head parameters can vary significantly, and the ideal scenario is to measure them before conducting DCS measurements. However, it is not always straightforward, and we usually assume average values. However, we must evaluate the impact of assumed errors on BFi estimation. Since and are typically unknown and have to be measured separately or taken from literature. We examined how and of layer 3 (brain) impact BFi extraction. Changing the scalp/skull thickness also varies BFi, which can be observed using the multi-layered model fitting method. Here, we use the three-layer fitting method, and all BFi were obtained at . Additional details are presented in Table 3. Table 3Varying optical properties and scalp (Δ1) and skull (Δ2) thicknesses.
3.5.1.variationTo study how impacts BFi, we set , 0.015, 0.019, 0.023, and 0.027 and in MCX. The baseline is at , with and variation. In this case, two BFi groups were calculated. The first group was calculated assuming a constant (0%), defined as , and the calculated BFi is defined as . The second group was calculated using the known set in MCX, which we considered as true , and the corresponding calculated BFi is considered as . 3.5.2.variationSimilarly, we conducted simulations with , 0.888, 1.110, 1.332, and and a fixed to investigate how impacts BFi estimation. We define the estimated BFi as when (at 0%, defined as ). Additionally, was calculated using the known set in MCX, considered as true . The mean and standard deviation of the estimated BFi (versus ) over 100 trials are shown in Fig. 8(a). We also compare and . The blue () and green () dashed lines are for the semi-infinite model, whereas the red () and purple () dashed line are for the three-layer model. The red solid () and black dashed lines are for DCS-NET. Similarly, the BFi’s mean and standard deviation (versus ) over 100 trials are shown in Fig. 8(b). Figure 9 shows the BFi variation (in %) versus the and variations (in %). The percentage error for is defined as . Similarly, we define the percentage error for as . The BFi error (in %) caused by assumed error in or is defined as . Figures 8 and 9 show that is positively related to and negatively related to for semi-infinite and three-layer fitting models, in good agreement with previous findings.26,31 On the other hand, curves obtained from DCS-NET are close and are not sensitive to and . This result is expected, as from Eq. (2), should yield a more pronounced impact compared to , primarily due to the second-order contribution from and observed in biological tissues. Extreme examples are shown in Fig. 9, namely, a more extensive results in and results in . When reaches +62%, reaches and gives . The results from the three-layer fitting model show similar behaviors. Namely, is positively related to and negatively related to in layer 3, this result aligns well with the conclusions from Zhao et al.’ conclusion.31 In contrast, DCS-NET only shows in caused by and (blue solid and brown solid lines for and , respectively in Fig. 9), indicating that the variations in and have negligible impact on BFi estimation. 3.5.3.Scalp thickness variationTo investigate ’s impact on BFi, we varied (= 3, 4, 5, 6, and 7 mm) and fixed in MCX. We define the estimated BFi as when (0%, defined as ). Additionally, was calculated using the known set in MCX, considered as true . 3.5.4.Skull thickness variationSimilarly, to investigate ’s impact on BFi, we varied (= 4.2, 5.6, 7.0, 8.4, and 9.8 mm) and fixed in MCX. We define the estimated BFi as calculated when (0%, defined as ). Additionally, was calculated using the known set in MCX, considered as true . Figure 10(a) presents BFi’s mean value (represented by bar plots) and standard deviation (depicted by error bars) over 100 trials versus . The rightmost bar group represents the results obtained with . Figure 10(b) shows BFi’s mean value and standard deviation versus , the rightmost bar group represents the results obtained with . Still, we can see that the semi-infinite model cannot provide accurate BFi at a deeper layer. When changed, falls into [the bar group 1 in Fig. 10(a)] when using DCS-NET, whereas falls into [the bar group 3 in Fig. 10(a)] using the three-layer fitting model, slightly larger than that using DCS-NET. However, increases to when estimation error occurs using the three-layer fitting method [shown in the rightmost bar group in Fig. 10(a)]. Whereas for the variation in , falls into when DCS-NET is used [the bar group 1 in Fig. 10(b)], whereas falls into when the three-layer fitting method is used [the bar group 3 in Fig. 10(b)]. Both present similar accuracy. However, when is not accurate, becomes more pronounced and reaches [the rightmost bar group in Fig. 10(b)]. Figure 11 shows the BFi variation (in %) versus the and variations (in %). The percentage error for is defined as . Similarly, we define the percentage error for as . The BFi error (in %) caused by assumed error in and is defined as . As it is commonly known, and cause a significant . Figures 10(a) and 10(b) demonstrate a positive correlation between and (and ). Furthermore, as observed in Fig. 11, resulting from ranges from to +43.68%. In contrast, caused by ranges from to +53.47%. This error range is significantly narrower than that caused by the skull thickness, agreeing with the findings in Ref. 31. For DCS-NET, caused by both and falls within the limited range of to . 3.6.BFi Estimation TimeIn addition, the BFi estimation time is also an important parameter, especially in real-time measurements, and Table 4 compares the three extraction methods. We record it for single decays and batch decays (e.g., 100 trials) at different noise levels. It is clear that DCS-NET is promising for real-time applications. All data reported in Table 4 are standard deviations and means for repeating three times after discarding the first few runs that usually take longer. The analysis were performed using the workstation (CPU: Intel(R) Core(TM) i9-10900X @3.70 GHz; Memory: 128 GB; graphics processing unit (GPU): NVIDIA Quadro RTX 5000). Table 4The BFi estimation time (with Matlab parfor for semi-infinite and three-layer fitting models).
4.DiscussionOur study shows that DCS-NET can robustly quantify DCS-based blood flow measurements. We used DCS-NET to analyze the ACFs generated from MCX. The proposed network is based on 1DCNN,43 which is straightforward, quicker to train, and faster than high-dimension CNNs for time sequence analysis, such as FLIM data.43,60 To evaluate DCS-NET, we compared it with the semi-infinite, three-layer fitting methods by changing tissue optical properties ( and ), depths (related to ), and scalp/skull thicknesses ( and ). BFi estimated by DCS-NET shows a small error range induced by and (see Fig. 9) and a slightly wider error range induced by and (see Fig. 11). For rBFi, the error from DCS-NET (8.35%) is much less than that of the semi-infinite and three-layer fitting methods (43.76% and 19.66%, respectively). Moreover, DCS-NET yields more than 71.34% sensitivity to brain blood flow, whereas the semi-infinite and three-layer fitting methods yield 14.12% and 61.96%, respectively [Fig. 7(a)]. We considered measurement noise using a stochastic noise model53 to reflect experimental realities. With DCS-NET, is 12.87% at a high noise level (), whereas it increases to 390.10% when using the three-layer fitting method. At a low noise level (), the three-layer fitting model yields of 82.30%, much worse than 1.93% obtained by DCS-NET, suggesting that DCS-NET is less sensitive to noise [see Fig. 5(b)]. Figures 10(a) and 10(b) show that the three-layer analytical method (modeling the head, i.e., scalp, skull, and brain) can minimize the influence of extracerebral layers on measured DCS signals. However, this model requires a priori knowledge of layer optical properties and thicknesses. Therefore, accurately estimating scalp and skull thicknesses is required for reliable CBF estimation when using a three-layer analytical model. Besides accuracy and robustness, the computational cost is a critical factor that impacts practical applications, especially for real-time monitoring. Table 4 reveals that it took 0.004 s for DCS-NET to quantify 100 curves with 127 data points. In contrast, it took 0.160 and 181.697 s, respectively, for the semi-infinite fitting and three-layer fitting procedures. For quantifying a single autocorrelation decay curve, it only took 0.001 s for DCS-NET. In contrast, it took 0.032 and 17.496 s, respectively, for the semi-infinite fitting and three-layer fitting procedures. DCS-NET is the fastest among the three, around 17,000-fold faster than the three-layer model and 32-fold faster than the semi-infinite model. Table 5 lists existing deep learning methods applied to DCS techniques. It shows that DCS-NET’s training is much faster than 2DCNNs,40 approximately 140-fold faster. Although the remaining models, RNN,39 LSTM,41 and ConvGRU,42 have fewer total layers, they are limited to a specific . Table 5Comparison of existing AI methods for BFi estimation.
Notes: the training parameters of RNN and CNN(2D) are not given in the literature; we calculate them according to the structure shown in the literature. Although DCS-NET is more robust than the semi-infinite and three-layer fitting methods, our study has several limitations. First, DCS-NET’s training datasets were generated using the semi-infinite diffusion model as advised in Ref. 40. Nevertheless, this model does not consider scalp and skull thicknesses, which could potentially explain why the error range caused by and is much broader than that caused by and (Figs. 9 and 11). The complexity of including training datasets generated from a layered model is beyond the scope of this study, given this report’s already long length. In future, we will train new networks using datasets generated from a layered model, and alternatively, obtaining training datasets from in vivo measurements, as demonstrated in Refs. 41 and 42 will also be considered. Second, current rBFi calculations do not consider variations in optical properties between the baseline and activation states. Indeed, and in the brain can vary according to interventions (e.g., functional activation), which are recognized to impact perfusion. Failing to account for these changes could introduce additional uncertainties in rBFi measurements. Third, we did not include a comparison with the two-layered analytical model in this report; it may be worth further investigation. Fourth, as we all know, analytical fitting methods suffer from partial volume effects and recover only a fraction of the actual change; still, the relationship between the recovered change and the actual change remains linear. However, from Fig. 7(a), we can see the BFi values from DCS-NET reflect various degrees of the relative ground truth change according to the relative change; thus, they have a non-linear relationship with actual brain blood flow. This suggests processing data with our DCS-NET could result in non-physiological distortions. We will further investigate this and improve our network models in future studies. Finally, our study was solely conducted using simulation data. In the future, we will perform phantom and in vivo experiments to validate our findings. 5.ConclusionWe compared the proposed DCS-NET against the semi-infinite and the three-layer models for estimating , BFi and rBFi. We used Monte Carlo simulations to validate their performances. This study evaluated the cerebral sensitivity using a deep learning method and the influence of scalp/skull thickness and variations on BFi extraction. Additionally, we examined the impact of noise. Our findings revealed that the homogenous model is sensitive to superficial layers. In contrast, the three-layer model performs better in estimating BFi in deeper layers but is more susceptible to measurement noise. Furthermore, DCS-NET outperforms the semi-infinite and three-layer fitting models in rBFi recovery. Using DCS-NET, variations in and have less impact on BFi, unlike variations in scalp and skull thicknesses, which show a more significant error in BFi. Moreover, iterative fitting methods are much slower and unsuitable for real-time “online” processing. In contrast, our DCS-NET is 32-fold faster than the semi-infinite model and 17,000-fold faster than the three-layer model, showing great potential for continuous real-time clinical applications. 6.AppendixTable 6 shows all essential parameters used in the throughout article, ensuring accessibility to comprehensive details for interested readers. Table 6Essential parameters list.
Code and Data AvailabilityThe data and code supporting the findings of this study are available from the corresponding author upon reasonable request. FundingThis work has been funded by the Engineering and Physical Sciences Research Council (Grant No. EP/T00097X/1): the Quantum Technology Hub in Quantum Imaging (QuantiC) and the University of Strathclyde. Author ContributionsQ.W. conceived the presented idea, performed the analysis, and derived the theoretical models. Q.W. and Z.Z. developed the neural network models. M.L. contributed to data analysis. D.L. devised and supervised the project and the findings of this work. All authors contributed to the writing of this paper. AcknowledgmentsWe thank Professor Stefan A. Carp, Massachusetts General Hospital, Harvard Medical School, for his valuable advice on Monte Carlo simulations by MCX. We also acknowledge Saeed Samaei, Department of Medical Physics, University of Western Ontario, Canada, for fruitful discussions. ReferencesK. Uludağ et al.,
“Coupling of cerebral blood flow and oxygen consumption during physiological activation and deactivation measured with fMRI,”
NeuroImage, 23
(1), 148
–155 https://doi.org/10.1016/j.neuroimage.2004.05.013 NEIMEF 1053-8119
(2004).
Google Scholar
T. Durduran and A. G. Yodh,
“Diffuse correlation spectroscopy for non-invasive, micro-vascular cerebral blood flow measurement,”
NeuroImage, 85 51
–63 https://doi.org/10.1016/j.neuroimage.2013.06.017 NEIMEF 1053-8119
(2014).
Google Scholar
X. Liang et al.,
“Coupling of functional connectivity and regional cerebral blood flow reveals a physiological basis for network hubs of the human brain,”
Proc. Natl. Acad. Sci. U. S. A., 110
(5), 1929
–1934 https://doi.org/10.1073/pnas.1214900110
(2013).
Google Scholar
A. Devor et al.,
“Frontiers in optical imaging of cerebral blood flow and metabolism,”
J. Cereb. Blood Flow Metab., 32
(7), 1259
–1276 https://doi.org/10.1038/jcbfm.2011.195
(2012).
Google Scholar
C. Cheung et al.,
“In vivo cerebrovascular measurement combining diffuse near-infrared absorption and correlation spectroscopies,”
Phys. Med. Biol., 46
(8), 2053
–2065 https://doi.org/10.1088/0031-9155/46/8/302 PHMBA7 0031-9155
(2001).
Google Scholar
V. Quaresima, S. Bisconti and M. Ferrari,
“A brief review on the use of functional near-infrared spectroscopy (fNIRS) for language imaging studies in human newborns and adults,”
Brain Lang., 121
(2), 79
–89 https://doi.org/10.1016/j.bandl.2011.03.009
(2012).
Google Scholar
R. Delgado-Mederos et al.,
“Transcranial diffuse optical assessment of the microvascular reperfusion after thrombolysis for acute ischemic stroke,”
Biomed. Opt. Express, 9
(3), 1262 https://doi.org/10.1364/BOE.9.001262 BOEICL 2156-7085
(2018).
Google Scholar
E. O’Keeffe et al.,
“Dynamic blood–brain barrier regulation in mild traumatic brain injury,”
J. Neurotrauma, 37
(2), 347
–356 https://doi.org/10.1089/neu.2019.6483 JNEUE4 0897-7151
(2020).
Google Scholar
J. Sutin et al.,
“Time-domain diffuse correlation spectroscopy,”
Optica, 3
(9), 1006 https://doi.org/10.1364/OPTICA.3.001006
(2016).
Google Scholar
P. Joris et al.,
“Cerebral blood flow measurements in adults: a review on the effects of dietary factors and exercise,”
Nutrients, 10
(5), 530 https://doi.org/10.3390/nu10050530
(2018).
Google Scholar
H. Yonas, R. R. Pindzola and D. W. Johnson,
“Xenon/computed tomography cerebral blood flow and its use in clinical management,”
Neurosurg. Clin. North Am., 7
(4), 605
–616 https://doi.org/10.1016/S1042-3680(18)30349-8
(1996).
Google Scholar
E. R. Muir et al.,
“Quantitative cerebral blood flow measurements using MRI,”
Cerebral Angiogenesis, 1135 205
–211 Springer New York, New York, NY
(2014). Google Scholar
J. J. Vaquero and P. Kinahan,
“Positron emission tomography: current challenges and opportunities for technological advances in clinical and preclinical imaging systems,”
Annu. Rev. Biomed. Eng., 17
(1), 385
–414 https://doi.org/10.1146/annurev-bioeng-071114-040723 ARBEF7 1523-9829
(2015).
Google Scholar
G. Holló, T. J. T. P. van den Berg and E. L. Greve,
“Scanning laser Doppler flowmetry in glaucoma,”
Int. Ophthalmol., 20
(1–3), 63
–70 https://doi.org/10.1007/BF00212948
(1996).
Google Scholar
A. Oglat et al.,
“A review of medical Doppler ultrasonography of blood flow in general and especially in common carotid artery,”
J. Med. Ultrasound, 26
(1), 3 https://doi.org/10.4103/JMU.JMU_11_17 JMEUEV
(2018).
Google Scholar
T. Durduran et al.,
“Diffuse optical measurement of blood flow, blood oxygenation, and metabolism in a human brain during sensorimotor cortex activation,”
Opt. Lett., 29
(15), 1766 https://doi.org/10.1364/OL.29.001766 OPLEDP 0146-9592
(2004).
Google Scholar
D. A. Boas and A. G. Yodh,
“Spatially varying dynamical properties of turbid media probed with diffusing temporal light correlation,”
J. Opt. Soc. Am. A, 14
(1), 192 https://doi.org/10.1364/JOSAA.14.000192 JOAOD6 0740-3232
(1997).
Google Scholar
T. Durduran et al.,
“Diffuse optics for tissue monitoring and tomography,”
Rep. Prog. Phys., 73
(7), 076701 https://doi.org/10.1088/0034-4885/73/7/076701 RPPHAG 0034-4885
(2010).
Google Scholar
E. M. Buckley et al.,
“Diffuse correlation spectroscopy for measurement of cerebral blood flow: future prospects,”
Neurophotonics, 1
(1), 011009 https://doi.org/10.1117/1.NPh.1.1.011009
(2014).
Google Scholar
Y. Shang et al.,
“Diffuse optical monitoring of repeated cerebral ischemia in mice,”
Opt. Express, 19
(21), 20301 https://doi.org/10.1364/OE.19.020301 OPEXFF 1094-4087
(2011).
Google Scholar
G. Yu et al.,
“Time-dependent blood flow and oxygenation in human skeletal muscles measured with noninvasive near-infrared diffuse optical spectroscopies,”
J. Biomed. Opt., 10
(2), 024027 https://doi.org/10.1117/1.1884603 JBOPFO 1083-3668
(2005).
Google Scholar
D. A. Boas et al.,
“Establishing the diffuse correlation spectroscopy signal relationship with blood flow,”
Neurophotonics, 3
(3), 031412 https://doi.org/10.1117/1.NPh.3.3.031412
(2016).
Google Scholar
M. M. Wu et al.,
“Improved accuracy of cerebral blood flow quantification in the presence of systemic physiology cross-talk using multi-layer Monte Carlo modeling,”
Neurophotonics, 8
(1), 015001 https://doi.org/10.1117/1.NPh.8.1.015001
(2021).
Google Scholar
J. Li et al.,
“Noninvasive detection of functional brain activity with near-infrared diffusing-wave spectroscopy,”
J. Biomed. Opt., 10
(4), 044002 https://doi.org/10.1117/1.2007987 JBOPFO 1083-3668
(2005).
Google Scholar
L. Dong et al.,
“Simultaneously extracting multiple parameters via fitting one single autocorrelation function curve in diffuse correlation spectroscopy,”
IEEE Trans. Biomed. Eng., 60
(2), 361
–368 https://doi.org/10.1109/TBME.2012.2226885 IEBEAX 0018-9294
(2013).
Google Scholar
D. Irwin et al.,
“Influences of tissue absorption and scattering on diffuse correlation spectroscopy blood flow measurements,”
Biomed. Opt. Express, 2
(7), 1969 https://doi.org/10.1364/BOE.2.001969 BOEICL 2156-7085
(2011).
Google Scholar
L. Gagnon et al.,
“Investigation of diffuse correlation spectroscopy in multi-layered media including the human head,”
Opt. Express, 16
(20), 15514 https://doi.org/10.1364/OE.16.015514 OPEXFF 1094-4087
(2008).
Google Scholar
J. Wu et al.,
“Two-layer analytical model for estimation of layer thickness and flow using diffuse correlation spectroscopy,”
PLoS ONE, 17
(9), e0274258 https://doi.org/10.1371/journal.pone.0274258 POLNCL 1932-6203
(2022).
Google Scholar
J. Li et al.,
“Analytical models for time-domain diffuse correlation spectroscopy for multi-layer and heterogeneous turbid media,”
Biomed. Opt. Express, 8
(12), 5518 https://doi.org/10.1364/BOE.8.005518 BOEICL 2156-7085
(2017).
Google Scholar
K. Verdecchia et al.,
“Assessment of a multi-layered diffuse correlation spectroscopy method for monitoring cerebral blood flow in adults,”
Biomed. Opt. Express, 7
(9), 3659 https://doi.org/10.1364/BOE.7.003659 BOEICL 2156-7085
(2016).
Google Scholar
H. Zhao, E. Sathialingam and E. M. Buckley,
“Accuracy of diffuse correlation spectroscopy measurements of cerebral blood flow when using a three-layer analytical model,”
Biomed. Opt. Express, 12
(11), 7149 https://doi.org/10.1364/BOE.438303 BOEICL 2156-7085
(2021).
Google Scholar
Y. Shang and G. Yu,
“A Nth-order linear algorithm for extracting diffuse correlation spectroscopy blood flow indices in heterogeneous tissues,”
Appl. Phys. Lett., 105
(13), 133702 https://doi.org/10.1063/1.4896992 APPLAB 0003-6951
(2014).
Google Scholar
Y. Shang et al.,
“Extraction of diffuse correlation spectroscopy flow index by integration of Nth-order linear model with Monte Carlo simulation,”
Appl. Phys. Lett., 104
(19), 193703 https://doi.org/10.1063/1.4876216 APPLAB 0003-6951
(2014).
Google Scholar
P. Zhang et al.,
“Approaches to denoise the diffuse optical signals for tissue blood flow measurement,”
Biomed. Opt. Express, 9
(12), 6170 https://doi.org/10.1364/BOE.9.006170 BOEICL 2156-7085
(2018).
Google Scholar
V. N. Vapnik,
“An overview of statistical learning theory,”
IEEE Trans. Neural Netw., 10
(5), 988
–999 https://doi.org/10.1109/72.788640 ITNNEP 1045-9227
(1999).
Google Scholar
V. N. Vapnik, The Nature of Statistical Learning Theory, 35
–40 Springer-Verlag, New York
(1999). Google Scholar
Y. Zhang et al.,
“An investigation of deep learning models for EEG-based emotion recognition,”
Front. Neurosci., 14 622759 https://doi.org/10.3389/fnins.2020.622759 1662-453X
(2020).
Google Scholar
X. Liu et al.,
“Deep learning in ECG diagnosis: a review,”
Knowl. Based Syst., 227 107187 https://doi.org/10.1016/j.knosys.2021.107187
(2021).
Google Scholar
P. Zhang et al.,
“Signal processing for diffuse correlation spectroscopy with recurrent neural network of deep learning,”
in IEEE Fifth Int. Conf. on Big Data Comput. Service and Appl. (BigDataService),
328
–332
(2019). https://doi.org/10.1109/BigDataService.2019.00058 Google Scholar
C.-S. Poon, F. Long and U. Sunar,
“Deep learning model for ultrafast quantification of blood flow in diffuse correlation spectroscopy,”
Biomed. Opt. Express, 11
(10), 5557 https://doi.org/10.1364/BOE.402508 BOEICL 2156-7085
(2020).
Google Scholar
Z. Li et al.,
“Quantification of blood flow index in diffuse correlation spectroscopy using long short-term memory architecture,”
Biomed. Opt. Express, 12
(7), 4131 https://doi.org/10.1364/BOE.423777 BOEICL 2156-7085
(2021).
Google Scholar
J. Feng et al.,
“Cerebral blood flow monitoring using a ConvGRU model based on diffuse correlation spectroscopy,”
Infrared Phys. Technol., 129 104541 https://doi.org/10.1016/j.infrared.2022.104541 IPTEEY 1350-4495
(2023).
Google Scholar
D. Xiao, Y. Chen and D. D.-U. Li,
“One-dimensional deep learning architecture for fast fluorescence lifetime imaging,”
IEEE J. Sel. Top. Quantum Electron., 27
(4), 1
–10 https://doi.org/10.1109/JSTQE.2021.3049349 IJSQEN 1077-260X
(2021).
Google Scholar
Q. Fang and D. A. Boas,
“Monte Carlo simulation of photon migration in 3D turbid media accelerated by graphics processing units,”
Opt. Express, 17
(22), 20178 https://doi.org/10.1364/OE.17.020178 OPEXFF 1094-4087
(2009).
Google Scholar
D. A. Boas, L. E. Campbell and A. G. Yodh,
“Scattering and imaging with diffusing temporal field correlations,”
Phys. Rev. Lett., 75
(9), 1855
–1858 https://doi.org/10.1103/PhysRevLett.75.1855 PRLTAO 0031-9007
(1995).
Google Scholar
X. Cheng et al.,
“Development of a Monte Carlo-wave model to simulate time domain diffuse correlation spectroscopy measurements from first principles,”
J. Biomed. Opt., 27
(8), 083009 https://doi.org/10.1117/1.JBO.27.8.083009 JBOPFO 1083-3668
(2022).
Google Scholar
J. Dong et al.,
“Diffuse correlation spectroscopy with a fast Fourier transform-based software autocorrelator,”
J. Biomed. Opt., 17
(9), 097004 https://doi.org/10.1117/1.JBO.17.9.097004 JBOPFO 1083-3668
(2012).
Google Scholar
G. Yu et al.,
“Noninvasive monitoring of murine tumor blood flow during and after photodynamic therapy provides early assessment of therapeutic efficacy,”
Clin. Cancer Res., 11
(9), 3543
–3552 https://doi.org/10.1158/1078-0432.CCR-04-2582
(2005).
Google Scholar
A. Kienle et al.,
“Noninvasive determination of the optical properties of two-layered turbid media,”
Appl. Opt., 37
(4), 779 https://doi.org/10.1364/AO.37.000779 APOPAI 0003-6935
(1998).
Google Scholar
L. Gagnon et al.,
“Double-layer estimation of intra- and extracerebral hemoglobin concentration with a time-resolved system,”
J. Biomed. Opt., 13
(5), 054019 https://doi.org/10.1117/1.2982524 JBOPFO 1083-3668
(2008).
Google Scholar
F. Lesage, L. Gagnon and M. Dehaes,
“Diffuse optical-MRI fusion and applications,”
Proc. SPIE, 6850 68500C https://doi.org/10.1117/12.766745 PSISDG 0277-786X
(2008).
Google Scholar
H. Zhao and E. M. Buckley,
“Influence of source–detector separation on diffuse correlation spectroscopy measurements of cerebral blood flow with a multilayered analytical model,”
Neurophotonics, 9
(3), 035002 https://doi.org/10.1117/1.NPh.9.3.035002
(2022).
Google Scholar
C. Zhou et al.,
“Diffuse optical correlation tomography of cerebral blood flow during cortical spreading depression in rat brain,”
Opt. Express, 14
(3), 1125 https://doi.org/10.1364/OE.14.001125 OPEXFF 1094-4087
(2006).
Google Scholar
D. Mazumder et al.,
“Optimization of time domain diffuse correlation spectroscopy parameters for measuring brain blood flow,”
Neurophotonics, 8
(3), 035005 https://doi.org/10.1117/1.NPh.8.3.035005
(2021).
Google Scholar
S. A. Carp et al.,
“Diffuse correlation spectroscopy measurements of blood flow using 1064 nm light,”
J. Biomed. Opt., 25
(9), 097003 https://doi.org/10.1117/1.JBO.25.9.097003 JBOPFO 1083-3668
(2020).
Google Scholar
J. Selb et al.,
“Sensitivity of near-infrared spectroscopy and diffuse correlation spectroscopy to brain hemodynamics: simulations and experimental findings during hypercapnia,”
Neurophotonics, 1
(1), 015005 https://doi.org/10.1117/1.NPh.1.1.015005
(2014).
Google Scholar
J. Selb et al.,
“Improved sensitivity to cerebral hemodynamics during brain activation with a time-gated optical system: analytical model and experimental validation,”
J. Biomed. Opt., 10
(1), 011013 https://doi.org/10.1117/1.1852553 JBOPFO 1083-3668
(2005).
Google Scholar
S. Ioffe and C. Szegedy,
“Batch normalization: accelerating deep network training by reducing internal covariate shift,”
in Int. Conf. on Mach. Learn.,
448
–456
(2015). Google Scholar
D. P. Kingma and J. Ba,
“Adam: a method for stochastic optimization,”
(2017). Google Scholar
Q. Wang et al.,
“Simple and robust deep learning approach for fast fluorescence lifetime imaging,”
Sensors, 22
(19), 7293 https://doi.org/10.3390/s22197293 SNSRES 0746-9462
(2022).
Google Scholar
BiographyQuan Wang received his master’s degree in optics from Xi’an Technological University, Shaanxi, China, in 2018. From 2018 to 2020, he worked as a production technician at Electro Scientific Industries (MKS) Pte Ltd and as an optical engineer at KLA-Tencor Pte Ltd in Singapore. He is pursuing a PhD in the Department of Biomedical Engineering at the University of Strathclyde, Glasgow, United Kingdom. His current research focuses on fluorescence lifetime imaging systems, flow cytometry, and diffuse correlation spectroscopy. Mingliang Pan holds a bachelor’s degree in telecommunications engineering from Anhui University, Hefei, China. He further pursued and obtained his master’s degree in optical engineering from the University of Shanghai for Science and Technology, Shanghai, China. Currently, he is a PhD candidate in the Department of Biomedical Engineering at the University of Strathclyde, Glasgow, United Kingdom. His research interests include diffuse correlation spectroscopy, Raman spectroscopy, and microfluidics. Zhenya Zang is a PhD student at the University of Strathclyde, Glasgow, United Kingdom. His research interests include computational imaging, machine learning, and high-performance reconfigurable hardware design. David Day-Uei Li received his PhD in electrical engineering from National Taiwan University, Taipei, Taiwan, in 2001. He then joined the Industrial Technology Research Institute, working on complementary metal-oxide-semiconductor (CMOS) optical and wireless communication chipsets. From 2007 to 2011, he worked at the University of Edinburgh, Edinburgh, on two European projects focusing on CMOS single-photon avalanche diode sensors and systems. He then took the lectureship in biomedical engineering at the University of Sussex, Brighton, in mid-2011, and in 2014, he joined the University of Strathclyde, Glasgow, as a senior lecturer. He has published more than 100 research articles and patents. His research interests include time-resolved imaging and spectroscopy systems, mixed-signal circuits, CMOS sensors and systems, embedded systems, optical communications, and field programmable gate array/GPU computing. His research exploits advanced sensor technologies to reveal low-light but fast biological phenomena. |
Blood circulation
Analytic models
Education and training
Data modeling
Spectroscopy
Brain
Error analysis