Open Access
2 May 2013 Continuous correction of differential path length factor in near-infrared spectroscopy
Tanveer Talukdar, Solomon G. Diamond, Jason H. Moore
Author Affiliations +
Abstract
In continuous-wave near-infrared spectroscopy (CW-NIRS), changes in the concentration of oxyhemoglobin and deoxyhemoglobin can be calculated by solving a set of linear equations from the modified Beer-Lambert Law. Cross-talk error in the calculated hemodynamics can arise from inaccurate knowledge of the wavelength-dependent differential path length factor (DPF). We apply the extended Kalman filter (EKF) with a dynamical systems model to calculate relative concentration changes in oxy- and deoxyhemoglobin while simultaneously estimating relative changes in DPF. Results from simulated and experimental CW-NIRS data are compared with results from a weighted least squares (WLSQ) method. The EKF method was found to effectively correct for artificially introduced errors in DPF and to reduce the cross-talk error in simulation. With experimental CW-NIRS data, the hemodynamic estimates from EKF differ significantly from the WLSQ (p<0.001 ). The cross-correlations among residuals at different wavelengths were found to be significantly reduced by the EKF method compared to WLSQ in three physiologically relevant spectral bands 0.04 to 0.15 Hz, 0.15 to 0.4 Hz and 0.4 to 2.0 Hz (p<0.001 ). This observed reduction in residual cross-correlation is consistent with reduced cross-talk error in the hemodynamic estimates from the proposed EKF method.

1.

Introduction

In functional near-infrared spectroscopy (fNIRS) studies, cerebral blood volume and blood oxygenation changes are studied as indicators of brain activity. Hemodynamics are quantified based on localized changes in the concentrations of oxyhemoglobin (HbO) and deoxyhemoglobin (HbR). Three different classes of NIRS instrumentation are used to measure cerebral hemodynamics, and each have certain advantages and disadvantages. Time-domain NIRS (TD) and frequency domain NIRS (FD) are capable of providing absolute quantification of hemodynamics.13 TD provides superior depth resolution by tracking the time of flight of photons.4 However, the complexity of TD and FD systems make them more costly for scale up to high-density measurement probes5 and whole head neuroimaging.6 The instrument of choice for high-density neuroimaging applications with NIRS is continuous-wave (CW).

With CW-NIRS systems, the attenuation of light is measured at multiple wavelengths between source and detector optodes that are separated by a few centimeters on the scalp. The change in light absorption or optical density ΔODλ for a given wavelength, λ, is the negative logarithm of the ratio of the detected light at a given time t2 relative to the detected light at a reference time t1

Eq. (1)

ΔODλ=log10[I(t2)I(t1)].

Changes in HbO and HbR can be calculated by solving a set of linear equations based on the modified Beer-Lambert Law (MBLL):7

Eq. (2)

ΔODλ=εHbO,λ·DPFλ·d·ΔHbO+εHbR,λ·DPFλ·d·ΔHbR,
where ΔHbO and ΔHbR represent changes in concentration of HbO and HbR, respectively, and εHbO,λ and εHbR,λ are the molar extinction coefficients that specify how strongly HbO and HbR absorb light at a given wavelength. The product DPFλ·d is the mean optical path length, where d is the source-detector separation distance and DPFλ is the wavelength-dependent differential path length factor. DPFλ effectively scales the linear distance between the source and detector optodes to account for the increased distance that light travels due to scattering and absorption effects within the tissue.

Unlike TD or FD, CW-NIRS cannot be used to determine DPFλ. Typically, the assumed DPFλ values used in CW-NIRS analysis are obtained from prior studies of healthy subjects and range from three to six.2,8 However, DPFλ can vary between subjects due to differences in the anatomical structure and tissue composition of the brain. Morphological changes in the brain induced by aging or neurological disorders can affect the optical properties of brain tissue, altering DPFλ values.9 Furthermore, DPFλ used in the MBLL is assumed constant for a specific wavelength and cannot account for partial volume effects that arise from assuming global changes in the concentrations of HbO and HbR in a brain volume consisting of multiple layers of tissue.10 Studies have shown that skull thickness and the cerebrospinal fluid (CSF) in the head can affect the mean optical path length and hence DPFλ values.11,12 The reason this is a problem is that systematic errors in DPFλ lead to systematic errors in the quantification of cerebral oxygenation levels. In this work, we propose a method for calculating changes in the concentration of HbO and HbR while continuously correcting for systematic errors in DPFλ.

2.

Theoretical Formalism

CW-NIRS is an increasingly important tool for clinical monitoring and for diagnosis of neurological disorders. However, DPFλ values cannot be known in advance in such patients. Efforts have been made to estimate mean optical path lengths in infants13 as well as in patients undergoing cardiopulmonary bypass.14 Having knowledge of DPFλ values significantly improves the accuracy of calculated cerebral hemodynamics, which are important in the clinical evaluation of patients with stroke or brain injury.15,16

When inaccurate DPFλ values are used in the MBLL, the calculated concentration changes in ΔHbO and ΔHbR become inaccurate in two ways. The first type of inaccuracy is a loss of the absolute scale on the results; the second type of inaccuracy is called cross talk, which is when ΔHbO become contaminated by ΔHbR and vice versa. From a mathematical perspective, the errors C in ΔHbO and ΔHbR are given by:

Eq. (3)

CΔHbO=ΔHbOΔHbO*

Eq. (4)

CΔHbR=ΔHbRΔHbR*,
where ΔHbO and ΔHbR represent solutions obtained from using inaccurate DPFλ values, and the exact solutions that would have been obtained from using the correct DPFλ values are ΔHbR* and ΔHbR*. Scale errors arise when the DPFλ values are systematically too high or too low due to a multiplicative scale factor S and cross-talk errors arise when the DPFλ values are shifted relative to one another due to an offset term ΔDPFλ:

Eq. (5)

DPFλ*=S(DPFλΔDPFλ).

Pure scale errors result when S1 and pure cross talk errors occur when S=1 and ΔDPFλ0 for all but one of the wavelengths. In NIRS spectroscopic calculations involving multiple wavelengths, the errors C in ΔHbO and ΔHbR can be computed as:

Eq. (6)

C=[CΔHbOCΔHbR]=[AA*]Y,
where Y is a vector containing ΔODλ measurements and A and A* are the Moore–Penrose pseudoinverses. The matrix A containing inaccurate path lengths is defined as:

Eq. (7)

A=[Lλ1000Lλ2000LλN][εΔHbO,λ1εΔHbR,λ1εΔHbO,λ2εΔHbR,λ2εΔHbO,λNεΔHbR,λ2]=LE,
which is composed of a diagonal path length matrix L and a matrix of extinction coefficients E. The accurate model matrix A* is composed of scalar factor S, a diagonal matrix (L+ΔL) and the matrix E, where ΔL contains the offset terms for all but the N’th mean optical path length. The scalar factor and offset terms fully account for the inaccurately assumed DPFλ values:

Eq. (8)

A*=S(L+ΔL)E,

Eq. (9)

ΔL=[ΔLλ1000ΔLλ2000ΔLλN],
where ΔLλN=0 so that the scale error and cross-talk error effects are separable. The objective of the present work is to reduce the cross talk errors in ΔHbO and ΔHbR using an algorithm that can continuously correct for offsets in DPFλ during CW-NIRS spectroscopic calculations. Scale factor errors cannot be corrected by the proposed method.

Several methods based on numerical and analytical approaches are already available for calculating DPFλ in simulation. These methods typically rely on modeling light transport in the head using the radiative transport equation (RTE). The diffusion equation, which is derived from the P1 approximation to the RTE17,18 can accurately describe photon transport in a highly scattering medium. However, in a low scattering media like CSF, the diffusion approximation does not hold. Analytical solutions to the diffusion equation, which can be obtained for simple geometries19 are difficult to solve in complex geometries like the head. Instead, numerical approaches like the finite element method (FEM) are preferred. Another numerical approach is the FEM based Monte Carlo (MC) technique.2022 Although computationally expensive,23 FEM based MC has the advantage that it can handle complex geometries like the head and can be used to solve the full RTE. The major disadvantages of simulation approaches to calculating DPFλ values are inaccuracies in FEM representations of head anatomy and a lack of knowledge of the true optical properties in the mesh elements. Therefore, DPFλ values obtained from simulation methods are likely to be different from the actual values, which can lead to cross-talk error in the calculated ΔHbO and ΔHbR.10,22,24 This motivates the need for experimental methods of calculating relative DPFλ values from CW-NIRS data.

In the present work, we focus on calculating relative offsets in differential path length factors (DPFλ) at wavelengths of 690, 785, and 830 nm with respect to a DPFλ value at 808 nm that is held fixed (ΔDPF808=0). We apply the extended Kalman filter (EKF) to calculate the ΔDPFλ values and the changes in concentration of HbO and HbR. Since ΔDPFλ is a relative correction, the calculated ΔHbO and ΔHbR are corrected relative to one another but not corrected to absolute scales. The absolute accuracies of the calculated ΔHbO and ΔHbR values are determined by the a priori value of DPFλ at 808 nm.

3.

Dynamic State Estimation Using the EKF Algorithm

The EKF algorithm operates on a state-space dynamical system that consists of an observation model and a process model. In our case, the observation model is based on the MBLL and is derived by relating ΔODλ to the state variables ΔDPFλ, ΔHbO, and ΔHbR. The subscripts j and k are introduced in the equations below to specify the source-detector pair and the discrete time step, respectively. The MBLL-based observation model in matrix form is:

Eq. (10)

Yj,k=Hj,kXj,k+vj,k.

In Eq. (10), Xj,k is a vector containing state variables that we want to calculate using the EKF algorithm:

Eq. (11)

Xj,k=[ΔHbOj,kΔHbRj,kΔDPFλ1,j,kΔDPFλ2,j,kΔDPFλ3,j,k].

The measurement transition matrix Hj,k determines the dynamic characteristics of the model and is formed from the coefficients of ΔHbO and ΔHbR in the MBLL equation [Eq. (2)] with the addition of ΔDPFλi,j,k terms that correspond to the relative changes in the differential path length factor at wavelength λi for source-detector pair j and time step k. ΔDPFλi,j,k is updated at each time step for λi (i=690, 785, 830) and serves as a correction term for the assumed DPFλi,j. Three additional rows are also included in Hj,k so that we can specify statistical priors on ΔDPFλi,j,k:

Eq. (12)

Hj,k=[(DPFλ690,j+ΔDPFλ690,j,k)djελ690,HbO(DPFλ690,j+ΔDPFλ690,j,k)djελ690,HbR000(DPFλ785j+ΔDPFλ785,j,k)djελ785,HbO(DPFλ785,j+ΔDPFλ785,j,k)djελ785,HbR000(DPFλ830,j+ΔDPFλ830,j,k)djελ830,HbO(DPFλ830,j+ΔDPFλ830,j,k)djελ830,HbR000DPFλ808jdjελ808,HbODPFλ808,jdjελ808,HbR000001000001000001].

We assume that the expected values of ΔDPFλi,j,k are zero. These priors are incorporated in the measurement vector Yj,k along with ΔODλi,j,k obtained for i=690, 785, 808, 830:

Eq. (13)

Yj,k=[ΔODλ690,j,kΔODλ785,j,kΔODλ830,j,kΔODλ808,j,k000].

The observation model [Eq. (10)] contains an observation noise term vj,k, which is assumed to be zero mean white Gaussian noise (WGN) with covariance Rj:

Eq. (14)

Rj=[σΔOD,λ6902000σΔOD,λ7852σΔOD,λ8302σΔOD,λ8082σΔDPF,R2σΔDPF,R2000σΔDPF,R2].

The process model describes the dynamics of the state variables and is assumed to be stochastic:

Eq. (15)

Xj,k+1=FjXj,k+wj,k

Equation (15) contains an innovation term, also known as process noise, defined in vector wj,k, that is assumed to be drawn from a zero mean multivariate normal distribution with covariance Qj:

Eq. (16)

Qj=[σΔHbO2000σΔHbR2σΔDPF,Q2σΔDPF,Q2000σΔDPF,Q2].

The state transition matrix Fj in Eq. (15) is used to describe the relationship between the state variables X at time step k and k+1. In our process model, Fj is a 5 by 5 identity matrix.

The input and output state transition matrices Fj and Hj, along with the output vector Yj,k are used in the EKF algorithm to estimate the state variables in Xj,k. The EKF performs state estimation through an iterative prediction-correction scheme. In the prediction step, the most recent estimated state variables X^j,k1 and error covariance matrix Pj,k1 are projected forward in time to compute a predicted or a prior estimate X^j,k at the current time step k:

Eq. (17)

X^j,k=FjX^j,k1.

The predicted error covariance matrix is given by:

Eq. (18)

Pj,k=FjPj,k1FjT+Qj,
where Fj is the Jacobian matrix of partial derivatives of Fj with respect to each of the state variables, and Qj is the process noise covariance matrix.

In the correction stage, a posterior estimate of the states X^j,k is calculated based on a linear combination of the prior estimate X^j,k and a weighted difference between the predicted measurement HjX^j,k and the actual measurement Yj,k:

Eq. (19)

X^j,k=X^j,k+Kj,k(Yj,kHjX^j,k).

The optimal Kalman gain weighting matrix, Kj,k, is given by:

Eq. (20)

Kj,k=Pj,kHjT(HjPj,k1HjT+Rj)1,
where Hj is the Jacobian matrix obtained from Hj. A posterior error covariance matrix, Pj,k is calculated using the predicted error covariance matrix Pj,k:

Eq. (21)

Pj,k=(IKk,jHj)Pj,k.

The computed X^j,k and Pj,k are then used in the next predictor step.

4.

Methods

4.1.

NIRS Spectroscopic Calculation with EKF Using Simulated ΔODλ

NIRS ΔODλ data was generated at 690, 785, 808, and 830 nm with a 25 Hz sampling rate using the forward equation of MBLL from assumed known sequences of ΔHbO and ΔHbR. Large errors were introduced in the DPFλ values with respect to the assumed known value of 6. We also introduced WGN such that the signal-to-noise ratio (SNR) was 40 dB. Source-detector separation distance for each of the four wavelengths was assumed to be 35 mm.

In order to use the EKF, we need to have knowledge of the variance of the measurement noise and the process noise entries in matrices Rj and Qj, respectively. Measurement noise variance for the simulated data was already known from the statistics of the WGN. In the case of the measurement noise variance prior σΔDPF,R2 for ΔDPFλ, we specified an upper bound of 1.31 for its value. This prior was established heuristically based on the reported standard deviation of DPF values obtained experimentally from over the somatosensory motor region.25 We then estimated the process noise variances of the state variables. The process noise variances σΔHbO2 and σΔHbR2 were calculated from the difference between the original ΔHbO and ΔHbR sequences and their moving averages. The moving average filter had a span of three points. The process noise variance of ΔDPFλ, σΔDPF,Q2, was obtained by calculating the variance of the errors introduced in DPFλ. We initialized the error covariance matrix Pj,0 and the state variables Xj,0 to zero. We then applied the EKF on the simulated ΔODλ data and compared the estimated time series of ΔHbO and ΔHbR with the known sequences. A weighted least squares approach (WLSQ)26 was also used to estimate ΔHbO and ΔHbR for comparison with the solutions from EKF. In the WLSQ method, the measurement data was weighted by the inverse of the measurement noise variances:

Eq. (22)

W=[σΔOD,λ6902000σΔOD,λ7852σΔOD,λ8302000σΔOD,λ8082]1,
where the diagonal entries of the weighting matrix W correspond to the measurement noise variances for the simulated data at 690 nm, 785 nm, 830 nm and 808 nm.

4.2.

Tuning the EKF Parameters Using Simulated ΔODλ

Inaccurate process noise variance specification can affect the convergence properties and accuracy of the EKF state estimates. In our EKF model, the parameters that required tuning were the process noise variances σΔHbO2, σΔHbR2, and σΔDPF,Q2. We estimated σΔHbO2 or σΔHbR2 by calculating the variance of the residual obtained by subtracting low pass filtered ΔHbO and ΔHbR sequences from the original sequences as described in the previous section. Estimating σΔDPF,Q2 is more challenging. In order to determine σΔDPF,Q2, we picked nine test values on a logarithmic scale for σΔDPF,Q2 ranging from 1012 to 100 for use in the EKF with simulated NIRS data. The simulated data had errors in DPFλ of 1.5 at 690 nm, 1.5 at 785 nm, 2.9 at 830 nm, and zero error at 808 nm so that the inaccuracies introduced are only due to cross talk. For each test value of σΔDPF,Q2, we calculated the mean residual sum squared (RSS) error in the estimates of ΔHbO and ΔHbR as percentages of the sum-squared total (SST). We then determined the tuned value for σΔDPF,Q2 as the one that produced the minimum combined percent error.

4.3.

Application of the EKF Algorithm to NIRS Measurements

With the tuned EKF parameters, we ran the algorithm on actual NIRS measurements collected from three human subjects. The data were obtained from an Institutional Review Board (IRB) approved study in which each subject was fitted with a quad-source optode emitting light at wavelengths of 690, 785, 808, and 830 nm. The source optode was positioned on the scalp in a region over the left superior frontal lobe near position F3 on the electroencephalogram (EEG) 10 to 20 system. Data were analyzed from two detectors separated by 33 mm from the source optode. Subjects were asked to breathe normally for the first minute of the experiment to establish a baseline for processing the NIRS data in spectroscopic calculations. Subjects were then asked to breathe normally for 2 min followed by 1 min of deep breathing. This cycle of events were repeated for 8 min. The data sampling rate was 25 Hz and noisy channels were eliminated from the study if their SNR levels were less than 0 dB. Low frequency drifts in the data were removed by fitting a 5th order polynomial to data segments with a time span of 300 s. The SNR was calculated as the ratio of the variances of low to high pass filtered ΔODλ. Low pass filtering was carried out by using a 4th order Butterworth filter with cutoff frequency at 2 Hz. The data was filtered both in the forward and reverse direction to compensate for phase lags introduced by the filter. The high pass filtered signal containing the noise components was obtained by subtracting data smoothed with a moving average filter having a time span of three points from the original detrended data.

NIRS data can also be contaminated by motion artifacts and various methods exist to remove these artifacts.27 In our case, we applied Chauvenet’s criterion28 to reject data segments that contained motion related signal variations. A deviation ratio (DR) at each time point was calculated by dividing the ΔODλ signal deviations from their moving averages by the standard deviation of signal deviations. Signal deviations were calculated as the difference between the raw data and its smoothed version obtained by applying a moving average filter with a span of 30 points (1.2 s duration). Data with a DR greater than the standard Chauvenet’s criterion threshold were eliminated and the discontinuous data segments then spliced together. The resulting motion corrected ΔODλ signals were used as input data in the EKF algorithm. We then compared ΔHbO and ΔHbR calculated using the EKF with values obtained using the WLSQ method. Results were forward and reverse filtered between 0 and 2.5 Hz with a 4th order Butterworth low-pass filter to remove high frequency noise components.

4.4.

Residual Error Analysis in EKF and WLSQ

We directly quantified the percent cross-talk error in ΔHbO and ΔHbR calculated from the simulated data using EKF and WLSQ in three frequency bands that are known to contain physiologically relevant spectral peaks.29 The frequency bands we designated were low frequency oscillations (LF) between 0.04 and 0.15 Hz, high frequency oscillations (HF) between 0.15 and 0.4 Hz, and cardiac oscillations (CARDIAC) between 0.4 and 2.0 Hz. First the calculated hemodynamics and the known simulation sequences were band pass filtered in forward and reverse directions in the LF, HF and CARDIAC frequency bands with 4th order Butterworth filters. These filtered signals were then subtracted from one another to give ΔHbO and ΔHbR residuals. The percentages cross-talk error in ΔHbO and ΔHbR for EKF and WLSQ were then calculated as the ratio of the standard deviations of the residuals and the known sequences in each frequency band.

In case of experimental data, we analyzed the autocorrelation structure of the residuals obtained from the EKF and WLSQ results by calculating the fraction of points outside the 95% confidence interval of the autocorrelation function. For a sample of uncorrelated residuals, one would expect that, on average, 5% of the sample autocorrelation function will lie outside this confidence interval. The first 10 s of lags were not included in the calculation to remove the contribution of the main lobe of the autocorrelation function.

We also performed cross-correlations between the residuals in the LF, HF, and CARDIAC bands. For both the EKF and WSLQ method, we band pass filtered the corresponding residuals at each wavelength in the three frequency bands. We then cross-correlated the filtered residuals from all combinations of wavelength pairs resulting in a total of six combinations of wavelength pairs at each frequency band. These coefficients were then squared and averaged over all wavelength pairs for each source-detector pair resulting in mean squared cross-correlation coefficients. These coefficients were compared to direct measures of cross-talk error in simulation and then used as a proxy for cross talk error in the experimental NIRS data.

In order to statistically evaluate the performance of the proposed EKF algorithm, we conducted ANOVA on the EKF and WLSQ mean squared cross-correlation coefficients. First, we applied the Fisher Z-transformation30 to the EKF and WLSQ mean squared correlation coefficients to obtain normally distributed random variables. These transformed variables were then used as input for the ANOVA with the factors of wavelength pair, subject, and algorithm (i.e., EKF or WLSQ).

It is not possible to obtain direct measurements of cross-talk error from experimental CW-NIRS data. Instead we computed the mean percent signal deviations between the EKF and WLSQ solutions for the ΔHbO and ΔHbR sequences to quantify the magnitude of the differences in the EKF solutions. ΔHbO and ΔHbR sequences from EKF and WLSQ were first filtered in the LF, HF, and CARDIAC bands. Each filtered sequence was then split into short segments having a span of 30 points (1.2 s duration). A residual sequence was then obtained by subtracting each EKF segment from the corresponding WLSQ segment. The percent signal deviation in ΔHbO or ΔHbR was then calculated from the ratio of standard deviations of the residual sequences and the WLSQ segments. Finally, the mean percent signal deviations in ΔHbO and ΔHbR were computed by averaging all the percent signal deviations calculated from the short segments.

5.

Results

The EKF and WLSQ solutions for the simulated NIRS data are shown in Fig. 1. The solutions are displayed after low-pass filtering with a 2.5 Hz cutoff frequency using a Butterworth filter of order 4. The filter was applied both in the forward and reverse direction to compensate for phase lags introduced by the filter. Figure 1(a) shows the EKF and WLSQ solutions for ΔHbO, which agree closely with the known ΔHbO sequence as evidenced by R2 statistics of 0.99 for EKF and 0.96 for WLSQ. The residual errors between the known ΔHbO sequence and the solution obtained using EKF and WLSQ are shown in Fig. 1(b). Similarly, Fig. 1(c) and 1(d) shows ΔHbR computed from EKF and WLSQ and their corresponding residual errors in ΔHbR. It is evident in Fig. 1(c) that ΔHbR calculated using EKF agrees almost perfectly with the known sequence (R2=0.99). In case of WLSQ, there are large discrepancies between the calculated ΔHbR and the known sequence (R2=0.66). We can observe from the residual error plots for EKF that the magnitude of the error decreases rapidly during the first 2 s. This is approximately the amount of time it takes for the ΔDPFλ values to converge to the known values in the simulation [Fig. 1(e)]. In case of WLSQ, the residual errors do not decrease over time [Fig. 1(b) and 1(d)]; they tend to show some degree of serial correlation and contain oscillatory components.

Fig. 1

EKF and WLSQ solutions for simulated NIRS data: (a) ΔHbO calculated using EKF and WLSQ and the true known sequence; (b) residual errors in ΔHbO for EKF and WLSQ; (c) ΔHbR calculated using EKF and WLSQ; (d) residual errors for ΔHbR; (e) ΔDPFλ terms calculated using EKF. The true ΔDPFλ errors are shown in dotted lines.

JBO_18_5_056001_f001.png

In the simulated data, we had explicit prior knowledge of the measurement noise variance σΔOD,λi2 based on the simulated WGN that was introduced in ΔODλ. The process noise variances σΔHbO2 and σΔHbR2 were found to be approximately 1×1012(μM)2. Results from tuning the process noise variance in ΔDPFλ are shown in Fig. 2. A minimum value in the percent mean RSS error occurs at σΔDPF,Q2=1×106. This value of σΔDPF,Q2 was used in the EKF applied to both the simulated and experimental NIRS data.

Fig. 2

Percent mean RSS error between known ΔHbO and ΔHbR and the solutions computed using EKF for nine evaluated levels of σΔDPF,Q2. The minimum RSS error corresponds to a process noise variance of 1×106.

JBO_18_5_056001_f002.png

Figure 3(a) shows the percent cross-talk error in the solutions obtained from EKF and WLSQ. The percent cross-talk error in ΔHbO is significantly smaller in EKF relative to that found for WLSQ across all frequency bands (paired t-test, p<0.05). Although the cross-talk error in ΔHbR is slightly higher for EKF in the HF and CARDIAC band, the corresponding values in WLSQ are much higher at 45.9% and 35.1%, respectively.

Fig. 3

Simulated NIRS data residual analysis: (a) percent cross-talk error in ΔHbO and ΔHbR solution obtained using EKF ad WLSQ. Error bars show standard error; (b) mean sum squared cross-correlation coefficients for LF, HF, and CARDIAC filtered residuals for EKF and WLSQ.

JBO_18_5_056001_f003.png

Figure 3(b) shows the mean of the sum squared cross correlation coefficients obtained for the LF, HF, and CARDIAC frequency bands for EKF and WLSQ applied to simulated NIRS data. We can observe that the EKF has reduced mean sum squared cross-correlation coefficient values in the LF band relative to WLSQ. In case of the HF and CARDIAC bands, the mean squared cross-correlation coefficient values for EKF and WLSQ are comparable.

The results from EKF and WLSQ applied to experimental NIRS data segments are shown in Fig. 4. ΔHbO and ΔHbR computed using EKF agree closely with the WLSQ results [Fig. 4(a) and 4(b)]. Figure 4(c) shows the ΔDPFλ values calculated for wavelengths of 690, 785, and 830 nm. Large deviations in the ΔDPFλ values occur between 125 and 140 s. During this time interval, the difference between the EKF and WLSQ solution for ΔHbO and ΔHbR is also large [Fig. 4(d)]. The R2 values for the plotted time interval are 0.97 for ΔHbO and 0.98 for ΔHbR.

Fig. 4

EKF and WLSQ results obtained for experimental NIRS data collected from subject 3 (a) ΔHbO estimated using WLSQ and EKF (b) corresponding ΔHbR estimated using WLSQ and EKF; (c) EKF estimates of ΔDPFλ at 690, 785, and 830 nm; (d) difference between WLSQ and EKF hemodynamic signals.

JBO_18_5_056001_f004.png

Figure 5 shows ΔHbO and ΔHbR results in the LF, HF and CARDIAC bands for EKF and WLSQ. The differences between the EKF and WLSQ solutions are also plotted.

Fig. 5

Band pass filtered ΔHbO (upper row) and ΔHbR (lower row) sequences from EKF and WLSQ for subject 3: (a) and (d) LF (0.04 to 0.15 Hz); (b) and (e) HF (0.15 to 0.4 Hz); (c) and (f) CARDIAC (0.4 to 2.0 Hz).

JBO_18_5_056001_f005.png

Figure 6 shows the residuals obtained for WLSQ and EKF and their autocorrelation sequences for subject 3. The WLSQ and EKF residuals shown in Fig. 6(a) and 6(b) appears to contain physiological signal components such as cardiac oscillations in the EKF residuals. There also appear to be low frequency components in both the WLSQ and EKF residuals. The autocorrelation structures shown in Fig. 6(c) and 6(d) shows reduced serial correlation at larger lags in EKF compared to WLSQ. The autocorrelation values for the 690 nm residual appear to be substantially smaller in EKF compared to WLSQ.

Fig. 6

(a) WLSQ residuals for wavelengths 690, 758, 808, and 830 nm; (b) EKF residuals for the four wavelengths; (c) autocorrelation coefficients for residuals from WLSQ with the 95% confidence interval shown in solid gray lines (d) autocorrelation coefficient of residuals from EKF with the 95% confidence interval shown in solid gray lines.

JBO_18_5_056001_f006.png

Table 1 summarizes the mean percentages of points in the autocorrelation functions for WLSQ and EKF that fall outside of the 95% confidence intervals. The percentages of points outside the confidence bounds for EKF are consistently smaller than for WLSQ.

Table 1

Mean percentages of points that fall outside of the 95% confidence intervals for the EKF and WLSQ autocorrelation functions.

Wavelength (nm)WLSQ (%)EKF (%)
69039.64.1
78541.934.4
80841.129.2
83038.421.7

Figure 7 compares the mean squared cross-correlation coefficient values obtained in the analysis of EKF and WLSQ residuals. All three subjects have smaller mean squared correlation coefficient values for EKF relative to those found for WLSQ (see Table 2 for ANOVA). The lowest mean squared cross-correlation coefficients were found for the HF band in EKF.

Fig. 7

Mean squared cross-correlation coefficients of residuals from the analysis of experimental data from three subjects. Results are shown for LF, HF and CARDIAC frequency bands with standard error bars.

JBO_18_5_056001_f007.png

Table 2

NOVA of mean squared cross-correlation values shown in Fig. 7.

FactorSum Sq.d.f.Mean Sq.FP-value
LF (0.04–0.15 Hz)
Wavelength6.06051.2127.39<0.001
Subject0.02720.0130.080.9217
Algorithm2.40012.40014.63<0.001
Error10.34630.164
Total18.8271
LF (0.04–0.15 Hz)
Wavelength6.49951.29910.90<0.001
Subject0.26920.1341.130.3289
Algorithm1.41511.41511.87<0.001
Error7.512630.119
Total15.69771
CARDIAC (0.4–2.0 Hz)
Wavelength5.32951.06512.83<0.001
Subject0.00220.0010.010.9863
Algorithm2.69912.69932.49<0.001
Error5.234630.083
Total13.2771

ANOVA results are summarized in the Table 2. Algorithm (EKF versus WLSQ) contributes significantly (p<0.001) in all three frequency bands. Therefore, we can reject the null hypothesis that the mean squared cross-correlation coefficient values found for EKF and WLSQ are the same at those frequency bands. It is interesting to note that the wavelength factor also contributes significantly to the variability in the mean squared cross-correlation coefficient values indicating that the EKF corrections vary significantly with wavelength.

Figure 8 shows the mean percent deviations between EKF and WLSQ solutions in the LF, HF and CARDIAC frequency bands. For each frequency band, the mean percent deviations in ΔHbO and ΔHbR are significantly different from zero (two-tailed t-test: p<0.001). The mean deviation over all frequency bands was found to be 4.04% for ΔHbO and 4.92% for ΔHbR. In the LF band, the mean percent deviations are comparable for both ΔHbO and ΔHbR. The mean percent signal deviation is smaller in ΔHbO compared to ΔHbR in the HF and CARDIAC bands (p<0.05).

Fig. 8

Mean percent deviations between EKF and WLSQ solutions filtered in the LF, HF, and CARDIAC frequency bands shown with 95% confidence intervals from a paired t-test.

JBO_18_5_056001_f008.png

6.

Discussion and Conclusion

The proposed EKF algorithm is able to calculate ΔHbO and ΔHbR from simulated NIRS data by continuously updating ΔDPFλ values. We observed in simulation that the residual errors in ΔHbO and ΔHbR decrease rapidly within the first 2 s as the EKF tracked the errors introduced in DPFλ. In the analysis of experimental data, we observed reduced autocorrelation of residuals for EKF compared to WLSQ. We believe that the superior performance of EKF is achieved due to weighting of the state variable innovations by the optimal Kalman gain at each time step. The application of the Kalman gain in this way helps to minimize the posterior error covariance over time. In WLSQ, the measurement data is weighted by constant factors over all time steps; so even though the solution is optimal in the least squares sense, WLSQ lacks the flexibility of EKF to account for time variation in DPFλ. As a result, EKF outperforms WLSQ in the estimation of ΔHbO and ΔHbR.

We observed reduced cross-correlations in the residuals in wavelength pairs for EKF in the LF, HF and CARDIAC bands compared to WLSQ, which is consistent with the reduction of cross-talk error. Convergence of the EKF state estimates to the true states is dependent on prior knowledge of the process noise and measurement noise variance statistics. In the proposed method, we demonstrate how to obtain prior estimates of those statistics based on minimizing the residual error in EKF.

We also quantified the differences in the hemodynamic signals calculated using EKF and WLSQ in physiologically relevant frequency bands. We found significant differences in the ΔHbR sequences in the LF and HF bands. The average of differences between EKF and WLSQ solutions across all frequency bands was 3.75% for ΔHbO and 4.57% for ΔHbR. We believe that this difference is due to a reduction of cross-talk error in ΔHbO and ΔHbR obtained using EKF. This is supported by the observation that the EKF solution obtained from simulated NIRS data had reduced cross-talk error relative to the WLSQ solutions. In the simulation, the errors in ΔHbO and ΔHbR from the known sequences were significantly lower in the EKF results compared to the WLSQ results.

The proposed EKF method offers several advantages over other methods that may be used to calculate DPFλ. Unlike other studies where DPFλ is estimated based on the assumption that it is constant over time,2,25,31 the proposed EKF algorithm continuously updates DPFλ and it can account for variation in DPFλ due to differences in anatomical structure and tissue composition of the brain. In cases where relative calibration of ΔHbO and ΔHbR is desired, the proposed method eliminates the need to measure DPFλ using TD or FD instruments or applying simulation techniques discussed in Sec. 2.

The proposed EKF algorithm has potential applications in the clinical investigation of brain injury. A number of recent studies have used NIRS to monitor patients during stroke rehabilitation and recovery.3234 A limitation in these studies is the inability to accurately quantify cerebral hemodynamics in a region of interest. In the above referenced stroke studies, DPFλ is assumed to be constant over time and the studies did not account for possible variation due to the stroke. However, DPFλ could deviate substantially in stroke because of the high absorption coefficient of hemoglobin. As a result, the calculated ΔHbO and ΔHbR may not have accurately reflected the hemodynamics in those patients. The proposed algorithm could also be applied to other clinical diagnostic applications of NIRS such as breast cancer detection.35 The optical properties of cancerous tissue are different from those of healthy tissue in the breast. Accurate assessment of hemodynamic changes in breast tissue also relies on using accurate estimates of DPF in the NIRS spectroscopic calculations. Further investigation is required to determine if the proposed EKF algorithm is able to correct path length errors in NIRS applications in stroke, cancer, and other clinical conditions.

References

1. 

D. T. DelpyM. Cope, “Quantification in tissue near-infrared spectroscopy,” Phil. Trans. R. Soc. Lond. B, 352 (1354), 649 –659 (1997). http://dx.doi.org/10.1098/rstb.1997.0046 PTRBAE 0962-8436 Google Scholar

2. 

A. Duncanet al., “Optical pathlength measurements on adult head, calf and forearm and the head of the newborn infant using phase resolved optical spectroscopy,” Phys. Med. Biol., 40 (2), 295 –304 (1995). http://dx.doi.org/10.1088/0031-9155/40/2/007 PHMBA7 0031-9155 Google Scholar

3. 

D. T. Delpyet al., “Estimation of optical pathlength through tissue from direct time of flight measurement,” Phys. Med. Biol., 33 (12), 1433 –1442 (1988). http://dx.doi.org/10.1088/0031-9155/33/12/008 PHMBA7 0031-9155 Google Scholar

4. 

J. SelbD. K. JosephD. A. Boas, “Time-gated optical system for depth-resolved functional brain imaging,” J. Biomed. Opt., 11 (4), 044008 (2006). http://dx.doi.org/10.1117/1.2337320 JBOPFO 1083-3668 Google Scholar

5. 

B. R. WhiteJ. P. Culver, “Quantitative evaluation of high-density diffuse optical tomography: in vivo resolution and mapping performance,” J. Biomed. Opt., 15 (2), 026006 (2010). http://dx.doi.org/10.1117/1.3368999 JBOPFO 1083-3668 Google Scholar

6. 

M. A. Franceschiniet al., “Diffuse optical imaging of the whole head,” J. Biomed. Opt., 11 (5), 054007 (2006). http://dx.doi.org/10.1117/1.2363365 JBOPFO 1083-3668 Google Scholar

7. 

M. Copeet al., “Methods of quantitating cerebral near infrared spectroscopy data,” Adv. Exp. Med. Biol., 222 183 –189 (1988). Google Scholar

8. 

J. S. UltmanC. A. Piantadosi, “Differential pathlength factor for diffuse photon scattering through tissue by a pulse-response method,” Math. Biosci., 107 (1), 73 –82 (1991). http://dx.doi.org/10.1016/0025-5564(91)90072-Q MABIAR 0025-5564 Google Scholar

9. 

C. Bonnéryet al., “Changes in diffusion path length with old age in diffuse optical tomography,” J. Biomed. Opt., 17 (5), 056002 (2012). http://dx.doi.org/10.1117/1.JBO.17.5.056002 JBOPFO 1083-3668 Google Scholar

10. 

G. StrangmanM. A. FranceschiniD. A. Boas, “Factors affecting the accuracy of near-infrared spectroscopy concentration calculations for focal changes in oxygenation parameters,” NeuroImage, 18 (4), 865 –879 (2003). http://dx.doi.org/10.1016/S1053-8119(03)00021-1 NEIMEF 1053-8119 Google Scholar

11. 

K. Yoshitaniet al., “Effects of hemoglobin concentration, skull thickness, and the area of the cerebrospinal fluid layer on near-infrared spectroscopy measurements,” Anesthesiology, 106 (3), 458 –462 (2007). http://dx.doi.org/10.1097/00000542-200703000-00009 ANESAV 0003-3022 Google Scholar

12. 

E. OkadaD. T. Delpy, “Near-infrared light propagation in an adult head model. I. Modeling of low-level scattering in the cerebrospinal fluid layer,” Appl. Opt., 42 (16), 2906 –2914 (2003). http://dx.doi.org/10.1364/AO.42.002906 APOPAI 0003-6935 Google Scholar

13. 

D. A. BenaronC. D. KurthJ. M. Steven, “Transcranial optical path length in infants by near-infrared phase-shift spectroscopy,” J. Clin. Monit., 11 (2), 109 –117 (1995). http://dx.doi.org/10.1007/BF01617732 JCMOEH 0748-1977 Google Scholar

14. 

K. Yoshitaniet al., “Measurements of optical pathlength using phase-resolved spectroscopy in patients undergoing cardiopulmonary bypass,” Anesth. Analg., 104 (2), 341 –346 (2007). http://dx.doi.org/10.1213/01.ane.0000253508.97551.2e AACRAT 0003-2999 Google Scholar

15. 

S. Muehlschlegelet al., “Feasibility of NIRS in the neurointensive care unit: a pilot study in stroke using physiological oscillations,” Neurocrit. Care, 11 (2), 288 –295 (2009). http://dx.doi.org/10.1007/s12028-009-9254-4 NCEACB 1541-6933 Google Scholar

16. 

M. N. Kimet al., “Noninvasive measurement of cerebral blood flow and blood oxygenation using near-infrared and diffuse correlation spectroscopies in critically brain-injured adults,” Neurocrit. Care, 12 (2), 173 –180 (2010). http://dx.doi.org/10.1007/s12028-009-9305-x NCEACB 1541-6933 Google Scholar

17. 

Y. OkiH. KawaguchiE. Okada, “Validation of practical diffusion approximation for virtual near infrared spectroscopy using a digital head phantom,” Opt. Rev., 16 (2), 153 –159 (2009). http://dx.doi.org/10.1007/s10043-009-0026-3 1340-6000 Google Scholar

18. 

F. GaoH. NiuH. Zhao, “The forward and inverse models in time-resolved optical tomography imaging and their finite-element method solutions,” Image Vis. Comput., 16 (9), 703 –712 (1998). http://dx.doi.org/10.1016/S0262-8856(98)00078-X IVCODK 0262-8856 Google Scholar

19. 

S. R. ArridgeM. CopeD. T. Delpy, “The theoretical basis for the determination of optical pathlengths in tissue: temporal and frequency analysis,” Phys. Med. Biol., 37 (7), 1531 –1560 (1992). http://dx.doi.org/10.1088/0031-9155/37/7/005 PHMBA7 0031-9155 Google Scholar

20. 

Y. FukuiY. AjichiE. Okada, “Monte Carlo prediction of near-infrared light propagation in realistic adult and neonatal head models,” Appl. Opt., 42 (16), 2881 –2887 (2003). http://dx.doi.org/10.1364/AO.42.002881 APOPAI 0003-6935 Google Scholar

21. 

E. Okadaet al., “Theoretical and experimental investigation of near-infrared light propagation in a model of the adult head,” Appl. Opt., 36 (1), 21 –31 (1997). http://dx.doi.org/10.1364/AO.36.000021 APOPAI 0003-6935 Google Scholar

22. 

K. Uludaget al., “Cross talk in the Lambert-Beer calculation for near-infrared wavelengths estimated by Monte Carlo simulations,” J. Biomed. Opt., 7 (1), 51 –59 (2002). http://dx.doi.org/10.1117/1.1427048 JBOPFO 1083-3668 Google Scholar

23. 

M. Hiraokaet al., “A Monte Carlo investigation of optical pathlength in inhomogeneous tissue and its application to near-infrared spectroscopy,” Phys. Med. Biol., 38 (12), 1859 –1876 (1993). http://dx.doi.org/10.1088/0031-9155/38/12/011 PHMBA7 0031-9155 Google Scholar

24. 

N. OkuiE. Okada, “Wavelength dependence of crosstalk in dual-wavelength measurement of oxy- and deoxy-hemoglobin,” J. Biomed. Opt., 10 (1), 011015 (2005). http://dx.doi.org/10.1117/1.1846076 JBOPFO 1083-3668 Google Scholar

25. 

H. Zhaoet al., “Maps of optical differential pathlength factor of human adult forehead, somatosensory motor and occipital regions at multi-wavelengths in NIR,” Phys. Med. Biol., 47 (12), 2075 –2093 (2002). http://dx.doi.org/10.1088/0031-9155/47/12/306 PHMBA7 0031-9155 Google Scholar

26. 

T. J. Huppertet al., “HomER: a review of time-series analysis methods for near-infrared spectroscopy of the brain,” Appl. Opt., 48 (10), D280 –D298 (2009). http://dx.doi.org/10.1364/AO.48.00D280 APOPAI 0003-6935 Google Scholar

27. 

F. C. RobertsonT. S. DouglasE. M. Meintjes, “Motion artifact removal for functional near infrared spectroscopy: a comparison of methods,” IEEE Trans. Biomed. Eng., 57 (6), 1377 –1387 (2010). http://dx.doi.org/10.1109/TBME.2009.2038667 IEBEAX 0018-9294 Google Scholar

28. 

J. P. Holman, Experimental Methods for Engineers, 4th ed.McGraw-Hill Inc., New York (1984). Google Scholar

29. 

A. Cammet al., “Heart rate variability-Standards of measurement, physiological interpretation, and clinical use,” Circulation, 93 (5), 1043 –1065 (1996). http://dx.doi.org/10.1161/01.CIR.93.5.1043 CIRCAZ 0009-7322 Google Scholar

30. 

R. A. Fisher, “On the ‘probable error’ of a coefficient of correlation deduced from a small sample,” Metron, 1 (4), 3 –32 (1921). 0026-1424 Google Scholar

31. 

S. Umeyama, “New method of estimating wavelength-dependent optical path length ratios for oxy-and deoxyhemoglobin measurement using near-infrared spectroscopy,” J. Biomed. Opt., 14 (5), 054038 (2009). http://dx.doi.org/10.1117/1.3253350 JBOPFO 1083-3668 Google Scholar

32. 

G. Strangmanet al., “Near-infrared spectroscopy and imaging for investigating stroke rehabilitation: test-retest reliability and review of the literature,” Arch. Phys. Med. Rehabil., 87 (12), 12 –19 (2006). http://dx.doi.org/10.1016/j.apmr.2006.07.269 APMHAI 0003-9993 Google Scholar

33. 

J. C. Eliassenet al., “Brain-mapping techniques for evaluating poststroke recovery and rehabilitation: a review,” Top. Stroke Rehabil., 15 (5), 427 –450 (2008). http://dx.doi.org/10.1310/tsr1505-427 1074-9357 Google Scholar

34. 

P. Y. Linet al., “Review: applications of near infrared spectroscopy and imaging for motor rehabilitation in stroke patients,” J. Med. Biol. Eng., 29 (5), 210 –221 (2009). IYSEAK 0021-3292 Google Scholar

35. 

S. FantiniA. Sassaroli, “Near-infrared optical mammography for breast cancer detection with intrinsic contrast,” Ann. Biomed. Eng., 40 (2), 398 –407 (2012). http://dx.doi.org/10.1007/s10439-011-0404-4 ABMECF 0090-6964 Google Scholar
© 2013 Society of Photo-Optical Instrumentation Engineers (SPIE) 0091-3286/2013/$25.00 © 2013 SPIE
Tanveer Talukdar, Solomon G. Diamond, and Jason H. Moore "Continuous correction of differential path length factor in near-infrared spectroscopy," Journal of Biomedical Optics 18(5), 056001 (2 May 2013). https://doi.org/10.1117/1.JBO.18.5.056001
Published: 2 May 2013
Lens.org Logo
CITATIONS
Cited by 27 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Near infrared spectroscopy

Error analysis

Hemodynamics

Linear filtering

Diamond

Tissue optics

Process modeling

Back to Top