Open Access
27 January 2024 Quantification of blood flow index in diffuse correlation spectroscopy using a robust deep learning method
Quan Wang, Mingliang Pan, Zhenya Zang, David Day-Uei Li
Author Affiliations +
Abstract

Significance

Diffuse correlation spectroscopy (DCS) is a powerful, noninvasive optical technique for measuring blood flow. Traditionally the blood flow index (BFi) is derived through nonlinear least-square fitting the measured intensity autocorrelation function (ACF). However, the fitting process is computationally intensive, susceptible to measurement noise, and easily influenced by optical properties (absorption coefficient μa and reduced scattering coefficient μs) and scalp and skull thicknesses.

Aim

We aim to develop a data-driven method that enables rapid and robust analysis of multiple-scattered light’s temporal ACFs. Moreover, the proposed method can be applied to a range of source–detector distances instead of being limited to a specific source–detector distance.

Approach

We present a deep learning architecture with one-dimensional convolution neural networks, called DCS neural network (DCS-NET), for BFi and coherent factor (β) estimation. This DCS-NET was performed using simulated DCS data based on a three-layer brain model. We quantified the impact from physiologically relevant optical property variations, layer thicknesses, realistic noise levels, and multiple source–detector distances (5, 10, 15, 20, 25, and 30 mm) on BFi and β estimations among DCS-NET, semi-infinite, and three-layer fitting models.

Results

DCS-NET shows a much faster analysis speed, around 17,000-fold and 32-fold faster than the traditional three-layer and semi-infinite models, respectively. It offers higher intrinsic sensitivity to deep tissues compared with fitting methods. DCS-NET shows excellent anti-noise features and is less sensitive to variations of μa and μs at a source–detector separation of 30 mm. Also, we have demonstrated that relative BFi (rBFi) can be extracted by DCS-NET with a much lower error of 8.35%. By contrast, the semi-infinite and three-layer fitting models result in significant errors in rBFi of 43.76% and 19.66%, respectively.

Conclusions

DCS-NET can robustly quantify blood flow measurements at considerable source–detector distances, corresponding to much deeper biological tissues. It has excellent potential for hardware implementation, promising continuous real-time blood flow measurements.

1.

Introduction

Cerebral 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 (>5  m) 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) g2(τ)I(t)I(t+τ)I(t)2, where I(t) is the measured scattered light and τ 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 g2(τ) 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.2426 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 μa and reduced scattering coefficient μs) 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 N’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 (ρ), corresponding to a specific depth in biological tissues. To accommodate a wider range of ρ, 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 β 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 g2(τ) 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 (Tint). 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.

Fig. 1

Flowchart of the proposed analysis. Step 1 generates the ACF g2(τ) from MCX at different source–detector distances (5, 10, 15, 20, 25, and 30 mm), optical properties (μa1,2,3, μs1,2,3), scalp/skull thicknesses (Δ1, Δ2), and different noise levels using the three-layer slab. Step 2 obtains training datasets containing noise. The datasets are generated using a semi-infinite diffusion model with μa(0.01,1]  mm1, μs(0.5,  1.6]  mm1, β(0,1], and BFi[108,105]  mm2/s. Then, the simulated g2(τ) from step 1 is analyzed by the pretrained model to predict β and BFi. Step 3 fits the simulated data from step 1 with semi-infinite and three-layer models with known/assumed optical properties/thicknesses to extract β and BFi. Step 4 assesses BFi and β estimations and concludes the intrinsic sensitivity and errors in terms of the variations in μa, μs, Δ1, and Δ2.

JBO_29_1_015004_f001.png

2.

Methods

2.1.

DCS Theory

The transport of the unnormalized electric field auto-correlation function, G1(ρ,τ)E*(ρ,t)·E(ρ,t), is well described by the correlation diffusion equation:17,45

Eq. (1)

(13μs2+μa+13αk02μsΔr2(τ))G1(ρ,τ)=S(ρ),
where k0=2πn/λ is the wavenumber of light, n and λ are the refractive index and wavelength in the scattering medium, respectively. α is the fraction of dynamic photon scattering events in the medium. Δr2(τ) is the mean squared displacement of scatterers in the turbid medium during a time interval τ. S(ρ) is the point source located at ρ; ρ is the source–detector distance. μa and μs 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

Eq. (2)

G1(ρ,τ)=3μs4π(exp(Kr1)r1exp(Kr2)r2),
where K=3μaμs+αμs2k02Δr2(τ), r1=ρ2+z02, r2=ρ2+(z0+2zb)2, z0=(μa+μs)1, and zb=5/(3μs) 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 Δr2(τ)=6Dbτ, where Db represents the effective diffusion coefficient. BFi in DCS is typically defined as αDb.21,48 g2(τ) is linked to the normalized electric field auto-correlation function as

Eq. (3)

g2(ρ,τ)=1+βg1(ρ,τ)2;g1(ρ,τ)=|G1(ρ,τ)G1(ρ,τ=0)|,
where β 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 N slabs is considered, as shown in Fig. 2(c). Each slab has its thickness, Δp=LpLp1, p=1, 2, 3, where L0,1,2,3 are the coordinates along the z-axis and μa1,2,3, and μs1,2,3 are absorption and scattering coefficients. To solve Eq. (1) in the layered medium (along z direction), we can use the Fourier transform G(r,τ) for the transverse coordinate ρ as

Eq. (4)

G^(q,z,τ)=d2ρG(r,τ)exp(iq·ρ),
where q is the radial spatial frequency. Equation (1) can then be rewritten as

Eq. (5)

[2z2Θ(p)2(q,τ)]G^(q,z,τ)=3μs(p)δ(zz),
where Θ(p)2(q,τ)=3μa(p)μs(p)+6k02μs(p)2Db(p)τ+q2, z=1/μs1, and p=1,2,3.

Fig. 2

Simulation layered model and analytical models. (a) A large slab representing a human brain consisting of three layers of the scalp (5 mm), skull (7 mm), and brain (50 mm). (b) The homogenous semi-infinite analytical model used for fitting methods and generating deep-learning training datasets. (c) Three-layer geometric scheme including the position of the source and detector, each layer has its own thickness Δ1,2,3 and characterized by the absorption coefficient μa1,2,3 and reduced scattering coefficient μs1,2,3.

JBO_29_1_015004_f002.png

We divided the top layer into two sublayers: layer 0 (0<z<z) identified by p=0, and layer 1 (z<z<Δ1). Then, the solution of Eq. (5) inside the p’th layer (p=1,2,3) can be written as

Eq. (6)

G^(q,z,τ)=A(p)exp(Θ(p)z)+B(p)exp(Θ(p)z),
where A(p) and B(p) are coefficients for each layer determined by the boundary conditions

Eq. (7)

G^0(q,z,τ)z0zG^0(q,z,τ)=0,z=0,G^0(q,z,τ)=G^1(q,z,τ),z=z,zG^0(q,z,τ)=zG^1(q,z,τ)+3μs1,z=z,G^p(q,z,τ)=G^p+1(q,z,τ),z=Lp,  p=1,2,DpzG^p(q,z,τ)=Dp+1zG^p+1(q,z,τ)z=Lp,  p=1,2,G^3(q,z,τ)+z3zG^3(q,z,τ)=0,z=L3,
where z01/μs1 and z31/μs3 are the extrapolation lengths accounting for internal reflections at the tissue surface (z=0) and the back surface (z=L3), respectively. Dp=c/3μs(p) is the photon diffusion coefficient in layer p, and c is the speed of light.

Substituting Eq. (6) into Eq. (7), A(p) and B(p) can be determined (p=1, 2, 3), and we obtain the solution of Eq. (5) at z=0 as

Eq. (8)

G^(q,z=0,τ)=NumDenom,
where Num and Denom (when p=3 and Δ3) are

Eq. (9)

Num=3μs1(z0Θ1D1cosh(Θ1(Δ1z))(Θ2D2cosh(Θ2Δ2)+Θ3D3sinh(Θ2Δ2))+Θ2D2(Θ3D3cosh(Θ2Δ2)+Θ2D2sinh(Θ2Δ2))sinh(Θ1(Δ1z))),

Eq. (10)

Denom=Θ2D2cosh(Θ2D2)(Θ1(D1+Θ3D3z0)cosh(Θ1D1)+(Θ3D3+Θ12D1z0)sinh(Θ1D1))+(Θ1(Θ3D1D3+Θ22D22z0)cosh(Θ1D1)+(Θ22D22+Θ12Θ3D1D3z0)sinh(Θ1D1))sinh(Θ2Δ2).

Therefore, by performing the inverse Fourier transform of Eq. (8) with respect to q, the field ACF at z=0 can be written as

Eq. (11)

G(ρ,z=0,τ)=1(2π)2d2qG^(q,z=0,τ)exp(iq·ρ)=12πdqG^(q,z=0,τ)qJ0(q·ρ),

Eq. (12)

g2(ρ,z=0,τ=0)=G(ρ,z=0,τ)G(ρ,z=0,τ=0),
where J0 denotes the zero-order Bessel function of the first kind. The integral bound for q in Eq. (11) should theoretically be from 0 to +. However, in practice, the numerical integration is performed with a limited range as [0  mm1,30  mm1] advised in Ref. 29.

2.2.

Noise Models

This study evaluates the impact from noise on BFi and β. We employed a broadly accepted noise model proposed by Zhou et al.53 The standard deviation (σ(τ), noise) of g2(τ) is given as

Eq. (13)

σ(τ)=TbTint[β2(1+e2ΓTb)(1+e2Γτ)+2m(1e2ΓTb)e2Γτ1e2ΓTb  +2n1β(1+e2Γτ)+n2(1+eΓτ)]1/2,
where Tb is the bin width of the correlator, m is the bin index corresponding to τ. nITb is the average number of photons detected within the bin time, where I is the detected photon count rate, and Tint is the integration time (e.g., measurement duration). Γ is the decay rate of g2(τ), which is obtained from fitting the measured g2(τ) to the theoretical g2(τ)1+βexp(Γτ). Gaussian noise54,55 was added to g2(τ) based on a statistical noise model to determine the noise (σ(τ)). 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 Tint (= 1, 10, or 30 s).

2.3.

Intrinsic Sensitivity Estimation

To evaluate the sensitivity to changes in blood flow in the deeper layer, we fixed the effective diffusion coefficient Db=1×106  mm2/s in layer 1 and increased Db in layer 3 as αDb=[1+0.1×(w1)]×6×106  mm2/s, w is an integer and w=1,2,,11. The physiological and optical parameters listed in Table 1 are taken as baseline conditions. Similar to Ref. 54, the intrinsic sensitivity (ηH) is defined as

Eq. (14)

ηH=(BFiHBFi0)/BFi0(CBFperturbCBF0)/CBF0×100%,
where BFiH and BFi0 represent the estimated BFi (H=D, S, or T, meaning DCS-NET, the semi-infinite, and three-layer fitting methods) for the perturbed and baseline conditions, respectively, and CBFperturb and CBF0 are Db in layer 3 for the perturbed and baseline conditions, respectively.

Table 1

Physiological and optical parameters56 at 785 nm in the human head model.

LayerThickness (mm)μa (mm−1)μs′ (mm−1)Blood flow index (mm2/s)
Scalp (Δ1)50.0190.6601×106
Skull (Δ2)70.0140.8600
Brain500.0191.1106×106

2.4.

Monte Carlo Simulations

We 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 (g) of 0.89 and a refractive index (n) of 1.3744 for all layers. We launched 2×109 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 ρ=5, 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 G1(τ):22

Eq. (15)

G1(τ)=1Nps=1Npexp(13k02i=1NtYs,iΔr2(τ)i)exp(i=1Ntμa,iLs,i),
where Np is the number of detected photons, Nt is the number of tissue types (3 for our simulations), and Ys,i and Ls,i stand for the total momentum transfer and the total path length of photon s in layer i, respectively. μa,i is the absorption coefficient, and Δr2(τ)i is the mean square displacement of the scattered particles in layer i. Here, Δr2(τ)i=6Diτ, where Di is the effective diffusion coefficient of layer i. The simulated G1(τ) is normalized to G1(0), and then we can obtain g2(τ) using the Siegert relationship with β=0.5. In this simulation, the delay time 1  μsτ<10,000  μs (127 data points) was used for g2(τ).

2.5.

Deep Learning Architecture Design

The structure of DCS-NET is shown in Fig. 3(a). DCS-NET takes g2(τ) to estimate β and BFi independently. DCS-NET consists of (1) a shared branch for temporal feature extraction and (2) two subsequent independent branches for estimating β 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 β 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) g2(τ), of which the size is 1×127. Both the estimated β and BFi have a size of 1×1. Note that the simulated g2(τ) was normalized to (0, 1] before being fed into the model.

Fig. 3

Design and evaluation of the convolution neural network (CNN). (a) The proposed DCS-NET includes a CNN, BN, and sigmoid activation layers. The convolution layer parameters are the filter number × the kernel size × the stride. (b) Training and validation losses of DCS-NET. (c) g2(τ) with noise-free (blue), and with realistic noise added, assuming an 8.05 kcps at 785 nm at different noise levels with Tint=1, 10, and 30 s.

JBO_29_1_015004_f003.png

2.6.

Training Dataset Preparation

The 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 (200,000×127) were generated and split into the training (80%) and the validation (20%) groups. Each dataset consists of the input, g2(τ), and its corresponding labels are BFi and β, 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 μaU(0.01,1]  mm1, μsU(0.5,1.6]  mm1, βU(0,1], BFi[108,105]  mm2/s, and ρU[5,30]  mm, where U stands for a uniform distribution. g2(τ) 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 g2(τ), and the blue line represents noiseless g2(τ). We used the optimizer Adam59 for the training process, with the learning rate fixed at 1×105 in the standard back-propagation. We used the mean square error loss function for updating the network by controlling the following problem:

Eq. (16)

L(𝒫)=1MiMF(Xi,𝒫)Yi22,
where X is the network output (estimated BFi or β), and Y is the corresponding label (true BFi or β) in the i’th training pairs. F is the mapping function, 𝒫 is the trainable weights of our networks, and M 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 β 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.

Results

3.1.

Absolute BFi Recovery Versus Detection Depths

To investigate how the absolute BFi and β behave in terms of ρ among DCS-NET, semi-infinite, and three-layer fitting approaches, we generated g2(τ) via MCX Monte Carlo simulations for ρ=5, 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 Db (assumed α=1). When using DCS-NET, g2(τ) was fed into the pre-trained model. For the semi-infinite fitting procedure, g2(τ) was fitted to Eqs. (2) and (3), and we assumed μa=0.019  mm1, μs=1.099  mm1, for the brain layer (layer 3), as provided in Table 1.

We also fitted the simulated g2(τ) with the three-layer model, Eqs. (11) and (12), and Db1=1×106  mm2/s, Db2=0  mm2/s, Db3=6×106  mm2/s, μa1=0.019  mm1, μs1=0.635  mm1, μa2=0.014  mm1, μs2=0.851  mm/s, μa3=0.019  mm1, μs3=1.099  mm1, Δ1=5  mm, and Δ2=7  mm. Meanwhile, we set β=0.3 and Db3=2×107  mm2/s as the initial guesses. For the fitting, we used NLSM (lsqcurvefit(·) in MATLAB with the Levenberg–Marquardt optimization) to minimize the unweighted least squares objective function,

Eq. (17)

argminj=1j=Nτ[g2(τ)MCXg2(τ)H]2,H=(S,T),
where Nτ is the number of sampled g2(τ), and g2(τ)H is from Eq. (3) or Eq. (12). Fitting was performed on τ from 1 to 10,000  μs.

Table 2 presents the true β and BFi and estimated β and BFi using DCS-NET, semi-infinite, and three-layer fitting methods. All input parameters for fitting are assumed as described above, and βGT=0.5. We define BFiD, BFiS, and BFiT (also βD, βS, and βT) for DCS-NET, the semi-infinite, and three-layer fitting methods, respectively. We define εBFi,D(%)=|BFiDBFiGT|/BFiGT×100%, where εBFi,D is the BFi error with DCS-NET. Similarly, εBFi,S and εBFi,T are the BFi estimated errors with the semi-infinite and three-layer fitting methods.

Table 2

BFi in the brain estimated using DCS-NET, homogeneous semi-infinite and three- layer fitting models.

ρ (mm)LayerBFiGT (mm2/s)BFiD (mm2/s)BFi estimated by fitting methods (mm2/s)
BFiSBFiT
511×106βD=0.521βS=0.501βT=0.493
20BFiD=8.45×107BFiS=7.15×107BFiT=7.15×107
36×106
1011×106βD=0.509βS=0.499βT=0.493
20BFiD=7.36×107BFiS=5.47×107BFiT=2.17×105
36×106
1511×106βD=0.501βS=0.498βT=0.504
20BFiD=1.03×106BFiS=4.79×107BFiT=1.43×105
36×106
2011×106βD=0.499βS=0.495βT=0.506
20BFiD=2.07×106BFiS=4.57×107BFiT=8.17×106
36×106
2511×106βD=0.499βS=0.493βT=0.505
20BFiD=4.82×106BFiS=4.63×107BFiT=5.63×106
36×106
3011×106βD=0.499βS=0.491βT=0.505
20BFiD=5.71×106BFiS=4.88×107BFiT=4.97×106
36×106

Table 2 shows when the semi-infinite model is used, the estimated BFi is closer to layer 1 (αDb=1×106  mm2/s), even for ρ=30  mm, 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 BFiT=7.15×107  mm2/s, close to 1×106  mm2/s when ρ=5  mm. This is because the mean light penetration depth is ρ/3 to ρ/2.19 When ρ is small, most detected photons predominantly travel through layer 1. As ρ increases (ρ10  mm), the estimated BFi decreases, reaching 5.63×106  mm2/s at ρ=25  mm, with εBFi,T of 6.17%. This is because as ρ increases, the detected photons penetrate inside the skull layer (αDb=0  mm2/s), 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 ρ increases, reaching 5.71×106  mm2/s with εBFi,D of 4.83% at ρ=30  mm. These results suggest that the AI model is capable of recognizing the depth. Regarding β estimation, there is no significant difference among the three methods.

3.2.

Absolute BFi Recovery with Noise

Figure 3(c) displays the semi-infinite analytical example g2(τ) curves with noise using the model proposed by Zhou et al.53 The curves were obtained with ρ=30  mm at different noise levels (Tint=1,10,30  s), μa=0.019  mm1, and μs=1.099  mm1 with an assumed BFi=2×107  mm2/s. To assess DCS-NET’s performance in practical scenarios, we modified the Monte Carlo code to generate g2 curves including noise according to Zhou et al.’s noise model.53 We generated 100 g2 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 δ and resnorm (the squared 2-norm of the residual) ϵ as

Eq. (18)

δ=f(β,BFi,τq)g2(τq),ϵ=q=1q=Qδ2,
where q is the lag time index, and Q is the length of the time trace. f(β,BFi,τq) is the fitted g2(τ) obtained from fitting methods based on analytical models at the lag time τq, and the corresponding true value is g2(τq) from MCX. The fitting results using the semi-infinite and three-layer analytical models are presented in Fig. 4, in which noisy g2(τ) curves from MCX (blue star-shaped) and fitted g2(τ) curves (red lines) at different noise levels are shown. Figures 4(a)(i)4(a)(iv) show the MCX-generated and fitted g2 using the semi-infinite model, and they exhibit an increasing trend in δ, ranging from (0.0025,0.0025) to (0.5,0.5), indicating that the semi-infinite method becomes inaccurate when the noise level increases. Additionally, ϵ reaches 3.02 when Tint=1  s. Similar behaviors are observed in the three-layer fitting, as shown in Figs. 4(b)(i)4(b)(iv).

Fig. 4

MCX-generated (scattered stars) and fitted (red solid lines) g2 curves using semi-infinite and three-layer fitting methods. [(a) (i)–(iv), respectively] noisy MCX simulated data (scattered star-shaped) at different noise levels fitted with the semi-infinite homogeneous model; [(b) (i)–(iv)] noisy MCX-generated data fitted with the for the three-layer fitting procedure. The corresponding residual δ and resnorm ϵ curves are also included.

JBO_29_1_015004_f004.png

We also calculated the mean BFi and β over 100 trials. As for β, 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 (Tint=1  s) 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, εBFi,T is 82.30% at the lower noise level (Tint=30  s). As the noise level increases, εBFi,T also increases, with εBFi,T reaching 390.10% at the high noise level (Tint=1  s). 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, εBFi,D (using DCS-NET) at a high noise level (Tint=1  s) is 12.87%, whereas at a low noise level (Tint=30  s), 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 Tint.

Fig. 5

Estimated β by DCS-NET, semi-infinite, and three-layer fitting methods at different noise levels (Tint=1, 10, and 30 s). The bar height means the average value for estimated BFi or β, the error bar means the standard deviations σ. (b) The estimated BFi by the three methods at different noise levels. The red dot line stands for the ground truth. (All the average values were obtained over 100 trials.)

JBO_29_1_015004_f005.png

3.3.

Relative Blood Flow

In practice, we do not aim to obtain absolute BFi measurements. Instead, the relative variation in blood flow (e.g., rBFi=BFi/BFi0) is oftener used.19 To evaluate DCS-NET for extracting rBFi in the brain, we assigned αDb(w)=[1+0.05×(w1)]×6×106  mm2/s, w=1,2,,21 in layer 3 (brain) and fixed αDb in other layers. Figure 6 presents rBFi calculated on noiseless data at ρ=30  mm.

Fig. 6

rBFi calculated by DCS-NET, the semi-infinite, and three-layer fitting methods on noiseless data for ρ=30  mm for αDb ranges from 6×106  mm2/s to 1.2×105  mm2/s (w=1,,21) with a step of 0.05×106. rBFi=BFi/BFi0, we define the estimated BFi as BFi0 at the start point.

JBO_29_1_015004_f006.png

In Fig. 6, rBFi calculated by DCS-NET, the semi-infinite, and three-layer fitting methods on noiseless data for ρ=30  mm for αDb ranges from 6×106 to 1.2×105  mm2/s (w=1,,21) with a step of 0.05×106. rBFi=BFi/BFi0, we define the estimated BFi as BFi0 at the start point.

To compare the accuracy of the three different methods in quantifying rBFi, we defined the error in rBFi as εrBFi,H=|rBFiHrBFiGT|/rBFiGT×100% (H=D, S, or T), meaning the rBFi estimation error using DCS-NET, the semi-infinite, and three-layer fitting methods, respectively. We can observe that rBFiD (red star) is close to the true rBFi (blue solid line) with εrBFi,D ranging from 0.15% to 8.35%. By contrast, the semi-infinite and three-layer methods result in more significant errors of 3.41%εrBFi,S43.76% and 0.36%εrBFi,T19.66%, respectively. As expected, the semi-infinite homogenous solution resulted in significant errors in rBFi, in agreement with Ref. 33.

3.4.

Intrinsic Sensitivity

As described in Sec. 2.3, the input Db in layer 3, denoted as CBF0=6×106  mm2/s, serves as the base point, and its corresponding recovered BFi is denoted as BFi0. Similarly, we assigned αDb=[1+0.05×(w1)]×6×106  mm2/s (w is an integer; w=1,2,,21), and it is referred to as the perturbed blood flow CBFperturb. We also define a perturbation level ζ=(CBFperturbCBF0)/CBF0×100%. We calculated the corresponding BFi for αDb, and then used Eq. (14) to obtain ηD, ηS and ηT. 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 ηD>71.34%. The intrinsic sensitivity reaches 2.5 × when ζ=20%, then decreases with ζ increasing. In comparison, the three-layer fitting method achieved ηT=61.96%, whereas the semi-infinite fitting method yielded ηS 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 ηD  >0 at Tint=10  s and Tint=30  s. Conversely, with the semi-infinite and three-layer fitting models, η 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).

Fig. 7

(a) Intrinsic sensitivity on noiseless data. (b)–(d) The sensitivities for noise with Tint=30  s, Tint=10  s, Tint=1  s, respectively. η is the intrinsic sensitivity that defined in Eq. (14), and ζ is the perturbation level in layer 3 (brain). Red, blue, and dark lines present ηD, ηT, and ηS, respectively. The perturbation levels in the graphs start at ζ=20%.

JBO_29_1_015004_f007.png

3.5.

BFi Extraction with Varied Optical Properties and Scalp/Skull Thicknesses

In 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 μa and μs are typically unknown and have to be measured separately or taken from literature. We examined how μa and μs 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 ρ=30  mm. Additional details are presented in Table 3.

Table 3

Varying optical properties and scalp (Δ1) and skull (Δ2) thicknesses.

−40%−20%0%+20%+40%
μa (mm1)0.0110.0150.0190.0230.027
μs (mm1)0.6590.8791.0991.3191.539
Δ1 (mm)3.0004.0005.0006.0007.000
Δ2 (mm)4.2005.6007.0008.4009.800

3.5.1.

μa variation

To study how μa impacts BFi, we set μa=0.011, 0.015, 0.019, 0.023, and 0.027 and μs=1.099  mm1 in MCX. The baseline is at μa=0.019  mm1, with ±20% and ±40% variation. In this case, two BFi groups were calculated. The first group was calculated assuming a constant μa=0.019  mm1 (0%), defined as μa,m, and the calculated BFi is defined as BFim. The second group was calculated using the known μa set in MCX, which we considered as true μa, and the corresponding calculated BFi is considered as BFiGT.

3.5.2.

μs variation

Similarly, we conducted simulations with μs=0.666, 0.888, 1.110, 1.332, and 1.554  mm1 and a fixed μa=0.019  mm1 to investigate how μs impacts BFi estimation. We define the estimated BFi as BFim when μs=1.099  mm1 (at 0%, defined as μs,m). Additionally, BFiGT was calculated using the known μs set in MCX, considered as true μs.

The mean and standard deviation of the estimated BFi (versus μa) over 100 trials are shown in Fig. 8(a). We also compare BFim and BFiGT. The blue (BFiGT) and green (BFim) dashed lines are for the semi-infinite model, whereas the red (BFiGT) and purple (BFim) dashed line are for the three-layer model. The red solid (BFiGT) and black dashed lines are for DCS-NET. Similarly, the BFi’s mean and standard deviation (versus μs) over 100 trials are shown in Fig. 8(b).

Fig. 8

(a) Estimated BFi versus μa, the green and purple dashed lines are for BFim assuming μa=0.019  mm1, the red solid and black dashed lines are for BFiGT and BFiD, respectively, and the red and blue dashed lines are for BFiGT using the three-layer and semi-infinite fitting methods. (b) Estimated BFi versus μs, the green and purple dashed lines are for BFim assuming μs=1.10  mm1, the red solid and black dashed lines are for BFiGT and BFiD, respectively, and the red and blue dashed lines are for BFiGT using the three-layer and semi-infinite fitting methods.

JBO_29_1_015004_f008.png

Figure 9 shows the BFi variation (in %) versus the μa and μs variations (in %). The percentage error for μa is defined as Eμa=[μa,mμaμa]×100%. Similarly, we define the percentage error for μs as Eμs=[μs,mμsμs]×100%. The BFi error (in %) caused by assumed error in Eμa or Eμs is defined as EBFi=[BFimBFiGTBFiGT]×100%.

Fig. 9

BFi error (in %) versus errors in the μa and μs variation (in %) among DCS-NET, semi-infinite, and three-layer fitting methods.

JBO_29_1_015004_f009.png

Figures 8 and 9 show that EBFi is positively related to Eμa and negatively related to Eμs for semi-infinite and three-layer fitting models, in good agreement with previous findings.26,31 On the other hand, EBFi curves obtained from DCS-NET are close and are not sensitive to Eμa and Eμs. This result is expected, as from Eq. (2), μs should yield a more pronounced impact compared to μa, primarily due to the second-order contribution from μs and μsμa observed in biological tissues. Extreme EBFi examples are shown in Fig. 9, namely, a more extensive Eμa+62% results in EBFi+25% and Eμa30% results in EBFi10%. When Eμs reaches +62%, EBFi reaches 50% and Eμs30% gives EBFi+70%.

The results from the three-layer fitting model show similar behaviors. Namely, EBFi is positively related to Eμa   and negatively related to Eμs in layer 3, this result aligns well with the conclusions from Zhao et al.’ conclusion.31 In contrast, DCS-NET only shows 1%+5% in EBFi caused by Eμa and Eμs (blue solid and brown solid lines for μa and μs, respectively in Fig. 9), indicating that the variations in μa and μs have negligible impact on BFi estimation.

3.5.3.

Scalp thickness variation

To investigate Δ1’s impact on BFi, we varied Δ1 (= 3, 4, 5, 6, and 7 mm) and fixed Δ2=7  mm in MCX. We define the estimated BFi as BFim when Δ1=5  mm (0%, defined as Δ1,m). Additionally, BFiGT was calculated using the known Δ1 set in MCX, considered as true Δ1.

3.5.4.

Skull thickness variation

Similarly, to investigate Δ2’s impact on BFi, we varied Δ2 (= 4.2, 5.6, 7.0, 8.4, and 9.8 mm) and fixed Δ1=5  mm in MCX. We define the estimated BFi as BFim calculated when Δ2=7.0  mm (0%, defined as Δ2,m). Additionally, BFiGT was calculated using the known Δ2 set in MCX, considered as true Δ2.

Figure 10(a) presents BFi’s mean value (represented by bar plots) and standard deviation (depicted by error bars) over 100 trials versus Δ1. The rightmost bar group represents the results obtained with Δ1=5  mm. Figure 10(b) shows BFi’s mean value and standard deviation versus Δ2, the rightmost bar group represents the results obtained with Δ2=7  mm. Still, we can see that the semi-infinite model cannot provide accurate BFi at a deeper layer. When Δ1 changed, εBFi,D falls into 1.17%8.33% [the bar group 1 in Fig. 10(a)] when using DCS-NET, whereas εBFi,T falls into 4.30%14.66% [the bar group 3 in Fig. 10(a)] using the three-layer fitting model, slightly larger than that using DCS-NET. However, εBFi,T increases to 11.67%16.05% when Δ1 estimation error occurs using the three-layer fitting method [shown in the rightmost bar group in Fig. 10(a)]. Whereas for the variation in Δ2, εBFi,D falls into 0.33%10.33% when DCS-NET is used [the bar group 1 in Fig. 10(b)], whereas εBFi,T falls into 1.50%13.33% when the three-layer fitting method is used [the bar group 3 in Fig. 10(b)]. Both present similar accuracy. However, when Δ2 is not accurate, εBFi,T becomes more pronounced and reaches 41.09%193.40% [the rightmost bar group in Fig. 10(b)].

Fig. 10

(a) BFi’s mean value and standard deviation versus Δ1, and the rightmost bar group represents the results obtained with Δ1=5  mm. (b) BFi’s mean value and standard deviation versus Δ2, and the rightmost bar group represents the results obtained with Δ2=7  mm. Each bar in the plot represents the average BFi over 100 trials calculated using three different methods, whereas the error bar stands for the standard deviation of BFi over 100 trials.

JBO_29_1_015004_f010.png

Figure 11 shows the BFi variation (in %) versus the Δ1 and Δ2 variations (in %). The percentage error for Δ1 is defined as EΔ1=[Δ1,mΔ1Δ1]×100%. Similarly, we define the percentage error for Δ2 as EΔ2=[Δ2,mΔ2Δ2]×100%. The BFi error (in %) caused by assumed error in EΔ1 and EΔ2 is defined as EBFi=[BFimBFiGTBFiGT]×100%.

Fig. 11

BFi error (in %) versus errors in Δ1 and Δ2 (in %) between DCS-NET and three-layer fitting methods.

JBO_29_1_015004_f011.png

As it is commonly known, EΔ1 and EΔ2 cause a significant EBFi. Figures 10(a) and 10(b) demonstrate a positive correlation between EBFi and EΔ1 (and EΔ2). Furthermore, as observed in Fig. 11, EBFi resulting from EΔ2 ranges from 176.41% to +43.68%. In contrast, EBFi caused by EΔ1 ranges from 44.29% 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, EBFi caused by both Δ1 and Δ2 falls within the limited range of 6% to +8%.

3.6.

BFi Estimation Time

In 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 4

The BFi estimation time (with Matlab parfor for semi-infinite and three-layer fitting models).

Noise level1 trial100 trials
DCS-NET (s)Semi-infinite (s)Three-layer (s)DCS-NET (s)Semi-infinite (s)Three-layer (s)
Tint=30  s0.001±4.567×1040.034±0.01217.141±2.0270.004±8.728×1040.159±0.030180.176±3.029
Tint=10  s0.001±2.846×1040.031±0.02116.946±3.1570.004±7.367×1040.157±0.029181.023±2.025
Tint=1  s0.001±2.544×1040.030±0.03917.946±4.5870.004±2.765×1040.160±0.032187.118±5.398
Noiseless0.001±3.007×1040.032±0.00117.002±1.2480.004±9.415×1040.156±0.012180.169±3.017

4.

Discussion

Our 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 (μa and μs), depths (related to ρ), and scalp/skull thicknesses (Δ1 and Δ2). BFi estimated by DCS-NET shows a small error range 1%+5% induced by μa and μs (see Fig. 9) and a slightly wider error range 6%+8% induced by Δ1 and Δ2 (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, εBFi,D is 12.87% at a high noise level (Tint=1  s), whereas it increases to 390.10% when using the three-layer fitting method. At a low noise level (Tint=30  s), the three-layer fitting model yields εBFi,T 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 g2 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 ρ.

Table 5

Comparison of existing AI methods for BFi estimation.

ModelTraining parametersTraining timeTotal layerρ (mm)Year
DCS-NET25,50613  min185 to 302023
RNN39174,080N/A20252019
CNN(2D)4075,55230.5  h16127.52020
LSTM411161N/A2152021
ConvGRU4211,557N/A10202022
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 (6%+8%) caused by Δ1 and Δ2 is much broader than that (1%+5%) caused by μa and μs (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, μa and μs 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.

Conclusion

We compared the proposed DCS-NET against the semi-infinite and the three-layer models for estimating β, 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 μa/μs 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 μa and μs 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.

Appendix

Table 6 shows all essential parameters used in the throughout article, ensuring accessibility to comprehensive details for interested readers.

Table 6

Essential parameters list.

VariableVariable full names
αThe fraction of dynamic photon scattering events in the medium
βCoherent factor
μaAbsorption coefficient
ηDIntrinsic sensitivity using DCS-NETDefined in Sec. 2.3
ηSIntrinsic sensitivity using the semi-infinite fitting methodDefined in Sec. 2.3
ηTIntrinsic sensitivity using the three-layer fitting methodDefined in Sec. 2.3
EμaThe percentage error of the assumed μaDefined in Sec. 3.5
μsReduced scattering coefficient
EμsThe percentage error of the assumed μsDefined in Sec. 3.5
ρSource–detector distance
DbBrownian diffusion coefficient
DpPhoton diffusion coefficient
qRadial spatial frequency
BFiBlood flow index estimated in DCS (i.e., αDb)
BFi0Baseline BFi
BFimEstimated BFi when assumed μa, μs, Δ1, Δ2 are constant at 0%Defined in Sec. 3.5
BFiGTGround-truth blood flowDefined in Sec. 3.5
BFiDBlood flow index estimated by DCS-NETDefined in Sec. 3.1
BFiSBlood flow index estimated by the semi-infinite fitting methodDefined in Sec. 3.1
BFiTBlood flow index estimated by the three-layer fitting methodDefined in Sec. 3.1
EBFi (%)The BFi error (in %) between BFim and BFiGTDefined in Sec. 3.5
εBFi,DError percentage of BFi using DCS-NETDefined in Sec. 3.1
εBFi,SError percentage of BFi using the semi-infinite fitting methodDefined in Sec. 3.1
εBFi,TError percentage of BFi using the three-layer fitting methodDefined in Sec. 3.1
rBFiRelative blood flow indexDefined in Sec. 3.3
rBFiGTRelative blood flow index for ground truthDefined in Sec. 3.3
rBFiDrBFi estimated by DCS-NETDefined in Sec. 3.3
rBFiSrBFi estimated by the semi-infinite fitting methodDefined in Sec. 3.3
rBFiTrBFi estimated by the three-layer fitting methodDefined in Sec. 3.3
εrBFi,DError percentage of rBFi using DCS-NETDefined in Sec. 3.3
εrBFi,SError percentage of rBFi using the semi-infinite fitting methodDefined in Sec. 3.3
εrBFi,TError percentage of rBFi using the three-layer fitting methodDefined in Sec. 3.3
CBFCerebral blood flowDefined in Sec. 3.4
CBF0Baseline cerebral blood flowDefined in Sec. 3.4
CBFperturbαDb during perturbed conditionsDefined in Sec. 3.4
ζPerturbation levelDefined in Sec. 3.4
Δ1Scalp thickness
Δ2Skull thickness
TintIntegration time
TbThe bin width of the correlator
mBin index
τLag time
g1Normalized electric auto-correlation function
g2Normalized intensity auto-correlation function
ΓDecay rate of g2
wInteger, w=1,2,3
δResidual from fitting proceduresDefined in Eq. (18)
ϵResnorm (the square 2-norm of the residual)Defined in Eq. (18)

Disclosures

The authors declare no conflict of interest.

Code and Data Availability

The data and code supporting the findings of this study are available from the corresponding author upon reasonable request.

Funding

This 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 Contributions

Q.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.

Acknowledgments

We 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.

References

1. 

K. 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

2. 

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

3. 

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

4. 

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

5. 

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

6. 

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

7. 

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

8. 

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

9. 

J. Sutin et al., “Time-domain diffuse correlation spectroscopy,” Optica, 3 (9), 1006 https://doi.org/10.1364/OPTICA.3.001006 (2016). Google Scholar

10. 

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

11. 

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

12. 

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

13. 

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

14. 

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

15. 

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

16. 

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

17. 

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

18. 

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

19. 

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

20. 

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

21. 

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

22. 

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

23. 

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

24. 

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

25. 

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

26. 

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

27. 

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

28. 

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

29. 

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

30. 

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

31. 

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

32. 

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

33. 

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

34. 

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

35. 

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

36. 

V. N. Vapnik, The Nature of Statistical Learning Theory, 35 –40 Springer-Verlag, New York (1999). Google Scholar

37. 

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

38. 

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

39. 

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

40. 

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

41. 

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

42. 

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

43. 

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

44. 

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

45. 

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

46. 

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

47. 

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

48. 

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

49. 

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

50. 

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

51. 

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

52. 

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

53. 

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

54. 

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

55. 

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

56. 

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

57. 

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

58. 

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

59. 

D. P. Kingma and J. Ba, “Adam: a method for stochastic optimization,” (2017). Google Scholar

60. 

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

Biography

Quan 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.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 International License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Quan Wang, Mingliang Pan, Zhenya Zang, and David Day-Uei Li "Quantification of blood flow index in diffuse correlation spectroscopy using a robust deep learning method," Journal of Biomedical Optics 29(1), 015004 (27 January 2024). https://doi.org/10.1117/1.JBO.29.1.015004
Received: 12 June 2023; Accepted: 2 January 2024; Published: 27 January 2024
Advertisement
Advertisement
KEYWORDS
Blood circulation

Analytic models

Education and training

Data modeling

Spectroscopy

Brain

Error analysis

Back to Top