Open Access
1 November 2005 Extraction of depth-dependent signals from time-resolved reflectance in layered turbid media
Author Affiliations +
Abstract
We try a new approach with near-IR time-resolved spectroscopy, to separate optical signals originated in the upper layer from those in the lower layer and to selectively determine the absorption coefficient (µa) of each layer in a two-layered turbid medium. The difference curve in the temporal profiles of light attenuation between a target and a reference medium is divided into segments along the time axis, and a slope of each segment is calculated to determine the depth-dependent µa. The depth-dependent µa values are estimated under various conditions in which µa and the reduced scattering coefficient (µ′s) of each layer are changed with a Monte Carlo simulation and in phantom experiments. Temporal variation of them represents the difference in µa between two layers when µ′s of a reference is the same as that of the upper layer of the target. The discrepancies between calculated µa and the real µa depend on the ratio of the real µa of the upper layer to that of the lower layer, and our approach enables us to estimate the ratio of µa between the two layers. These results suggest the potential that µa of the lower layer can be determined by our procedure.

1.

Introduction

Near-infrared spectroscopy (NIRS) has often been used for noninvasive evaluation of tissue oxygenation. Recently, this technique has been developed as a tool for human brain mapping by measuring the hemodynamic changes of cerebral cortex associated with neuronal activation.1, 2, 3 NIRS has some advantages over other methods for cerebral hemodynamics evaluation such as positron emission tomography (PET) and functional magnetic resonance imaging (fMRI) in terms of, for example, temporal resolution, easier handling, and less motion restriction during measurement. Some problems, however, remain to be solved before the application of NIRS to measurements of the human head. The head is a layered structure consisting of cerebral and extracerebral tissues (scalp, skull, and cerebrospinal fluid), and each layer has different optical properties. When light is applied to the scalp and diffusely reflected light is collected at a position on the scalp a few centimeters away from the incident point, the detected light carries information concerning not only the cerebral but also extracerebral tissues. Because the changes in extracerebral blood flow influence the determination of cerebral hemoglobin (Hb) concentration changes, some reports4, 5, 6 have questioned the validity of cerebral NIRS. It is thus necessary to separate signals originating in the cerebral tissue from those in the extracerebral tissue. As one of the ways for this purpose, a multidetector system consisting of continuous-wave-type (cw-type) instruments has been developed.7, 8 In this approach, however, separation of NIR signals attributed to cerebral and extracerebral tissues was incomplete because the measurement by different distances of light guide pairs might cause a discrepancy in the position where signals originate. In addition, with instruments of this type, it was difficult to quantify the Hb concentration.9, 10 The absolute value of Hb concentration is indispensable because neuronal activation and hemodynamic changes seem to significantly depend on the baseline conditions,11, 12 and also to compare the magnitude of the changes between different cortical areas and/or individuals.

In contrast, time-resolved reflectance measurement with short-pulsed light [time-resolved spectroscopy (TRS)], which gives a temporal profile of detected light intensity, has high potential to overcome these issues. It is believed to carry information about depth-dependent attenuation based on the correlation of detection time to penetration depth of photons.13, 14, 15, 16, 17 Furthermore, it makes possible the determination of the optical properties in turbid media such as the absorption coefficient18, 19 (μa) of which acquisition at multiple wavelengths can give quantitative Hb concentrations. In most cases of conventional TRS measurements for the human head, μa was determined by fitting an analytical solution of the photon diffusion equation for a homogenous medium to the time-resolved reflectance profile on the assumption that the target medium is a semiinfinite homogeneous. The analysis based on this assumption not only causes an error in estimation of the optical properties of the human head20, 21, 22, 23, 24 but also cannot bring out the depth-dependent information. Therefore, new analytical methods that are suitable for layered media are required. To overcome this problem, some methods have been provided, such as by multilayered (time-dependent) diffusion equation9, 25, 26, 27 and by moments of distribution of time of flight.16, 28 Most of these analyses are employed by combination with the measurements at two or more distances, whereas a single-distance measurement is more favorable for practical measurements on the head due to the heterogeneity of the superficial layer and its curvature. On the other hand, the analysis with time-dependent mean partial path length by Steinbrink 16 could be applied to a single-distance measurement, but it did not enable us to estimate μa in a given condition but only the change in μa from one condition to another.

In this paper, we attempt to develop a simple analytical method to separate optical signals originated in shallower layers from those in deeper layers and to selectively determine the μa values of them by TRS measurements with a single fixed source-detector spacing. As the first step of an analysis of a multilayer model such as the human head, a two-layered medium was considered. Our approach to this purpose is “time-segment analysis,” which divides the temporal profile of detected light intensity into time segments (e.g., every 500ps ), leading to an estimation of depth dependent μa . In this paper, we first demonstrate μa estimated by the slope of each time segment (time-segmented μa ) and its discrepancy from the real μa of the lower layer under various conditions in which μa and reduced scattering coefficient (μs) of each layer were changed with a Monte Carlo simulation and in phantom experiments. We also show the relationship between such discrepancy and the difference in μa between the upper and lower layers. Then we discuss a correction of time-segmented μa that was estimated smaller than the real μa with the ratio of time-segmented μa in an earlier time segment to that in a later one. In addition, we consider a way to calculate an appropriate μa in the case that μs of medium was unknown. Finally, we suggest the applicability of our approach in practical measurement.

2.

Method for Determining μa

2.1.

Theoretical Consideration

When a light impulse is incident on the surface of a semiinfinite homogeneous medium (object), the time-resolved intensity of reflected light at time t can be expressed as29

Eq. 1

I(t)=S(t)exp(μact),
where μa is the absorption coefficient of the medium, c is the light velocity in the medium, and S(t) is a scattering function that is dependent on their reduced scattering coefficient (μs) in the photon diffusion regime. We consider a reference medium besides the object. The reflected light intensity for the reference at t is written as IR(t)=SR(t)exp(μa,RcRt) , where the subscripts R to I(t) , μa , c , and S(t) indicate a reference. Attenuation (A) is defined as logarithm of the inverse of reflectance and according to the time-resolved Beer-Lambert law,30 difference between the attenuation of an object [A(t)] and that of a reference [AR(t)] , Adiff(t) , can be expressed as

Eq. 2

Adiff(t)=A(t)AR(t)=ln[IR(t)I(t)]=(μacμa,RcR)t+Sdiff(t),
where Sdiff(t)=ln[SR(t)S(t)] . When μs and the refractive index are the same for the reference and the object, Sdiff(t) vanishes and Adiff(t) becomes linear to t and its slope is denoted as μadiffc , where μadiff=μaμa,R , and c=cR .

However, in the case where the object is an inhomogeneous medium (and the reference is a homogeneous one), Adiff(t) is no longer linear to t even if Sdiff(t) can be ignored. When μadiff varies due to the layer structure, dAdiff(t)dt reflects the depth dependence in μadiff . Here we introduce time-dependent apparent absorption coefficient, μa(t) , given by iμaili(t)L(t) , where the subscript i refers to the layer number in the object medium, li(t) is the mean partial path length in the i ’th layer for the photons detected at time t , and L(t) is total path length defined as L(t)=ct . Then we define time dependent difference in apparent absorption coefficient μadiff(t) , which is the difference between μa(t) and μa,R at each time. In the case of two-layered media, the i ’th layer simply corresponds to either upper or lower layer and dAdiff(t)dt is expressed as follows:

Eq. 3

dAdiff(t)dt=(ddt)[μadiff(t)ct]+dSdiff(t)dt=μa,upper[dlupper(t)dt]+μa,lower[dllower(t)dt]μa,Rc+dSdiff(t)dt,
where μa,upper and μa,lower are the μa values of the upper and lower layer for an object, and lupper(t) and llower(t) are the mean partial path length in the upper and lower layers at time t , respectively. Figure 1 shows time dependence of the number of scattering events for the photons detected by reflectance mode in the upper part of the depth from 2to36mm for a homogeneous semiinfinite medium ( μa=0mm1 , μs=1.0mm1 , source-detector distance ρ=30mm , n=1.37 ), predicted by a Monte Carlo simulation. As for photons detected later in time in the measurement of reflectance mode, the number of scattering events in shallow layers is almost constant. This is typically seen after around 3ns in layers that are shallower than 10mm . Because the number of scattering events is converted to the photon path length by mean free pass, the mean partial path lengths of the shallow layers can be considered constant after 3ns . At a later time, therefore, the mean partial path length in deeper layers is dominant to the total path length.

Fig. 1

Number of scattering events in the upper part of the depth from 2to36mm for a homogeneous semiinfinite medium ( μa=0mm1 , μs=1.0mm1 , ρ=30mm , n=1.37 ), predicted by Monte Carlo simulation. The vertical axis represents the number of scattering events per photon detected at time t .

064008_1_010506jbo1.jpg

2.2.

Time-Segmented μa and Method for Obtaining μa of the Lower Layer

Next, we introduced mean absorption differences obtained by ΔAdiff(t)Δt instead of dAdiff(t)dt in consideration of errors by getting the regression curve and to simplify an analysis process. For this aim, we divided the temporal profile of detected light intensity into segments to extract time-dependent μa (time-segment analysis). When we assume that μs and refractive index are the same for reference and object,

Eq. 4

ΔAdiff(t)Δt=(1Δt)(μa,upperΔlupper(t) +μa,lowerΔllower(t)μa,RcΔt) ={μa,upper[Δlupper(t)ΔL(t)] +μa,lower[Δllower(t)ΔL(t)]}cμa,Rc,
where ΔL(t) is total path length change, which is Δlupper(t)+Δllower(t)=cΔt . As shown in Fig. 1, if Δlupper(t) is negligible after 3ns , the slopes of Adiff(t) against ct at later time segments converge to (μa,lowerμa,R) . If we know μa,R , μa,lower can thus be determined. We refer to {μa,upper[Δlupper(t)ΔL(t)]+μa,lower[Δllower(t)ΔL(t)]} as the time-segmented μa,μaseg , which represents a depth-dependent absorption coefficient at each time segment. It depends on the contribution of the change in the mean partial path length to that in the total path length between two different times. Accordingly, Eq. 4 could be rewritten again as ΔAdiff(t)Δt=(μasegμa,R)c .

In the following, a calculation process is explained for the case of the two-layered object and homogeneous reference. First, from the two sets of data for the object and reference, we calculate Adiff(t) profile and divide it into segments. Then Adiff(t) at each time segment is approximated linear to time and its slope (time-segmented slope) is estimated by linear regression. From the time-segmented slope, μaseg can be obtained from μa,R and c based on Eq. 4. In this study, time range of 500ps was selected for the segment size considering the reliability of regression analysis of the time-segmented slope. Before calculations of Adiff(t) , temporal profiles of light intensity measured by TRS are deconvoluted by the pulse response of the system that can be determined in every experiment. Deconvolution can be performed by the Fourier transform, and higher frequency components over 34 of the maximum are eliminated as they are judged to be noise.

3.

Simulation and Experiment

3.1.

Monte Carlo Simulation

To analyze the light propagation and calculate the reflectance from a two-layered semiinfinite medium, we used the Monte Carlo code developed by Wang and Jacques.31 The code was modified to correspond to our measurement system. The number of scattering events in layers ranging from z to z+2mm in depth, and reflectance for the medium were calculated with the time step of 10ps . Source photons were perpendicularly irradiated on the surface of the semiinfinite media, and photons that were emitted from a detector position on the surface were all detected. The source-detector distance was 30mm . The calculation was repeated until the number of the detected photons reached 1,000,000. For each layer separately, μs were given as 1.0 and 1.5mm1 and μa were varied at the range of 0.005to0.02mm1 . These values were chosen close to the optical properties of the tissues in human heads, μs and μa of the head32 were 0.7 to 1.0 and 0.012to0.016mm1 at 825nm , and μs and μa of gray matter33 were 0.4 to 0.7 and 0.018mm1 at 849nm , respectively. To compare with the results of phantom experiments described in the following, a case for smaller μa value (0.005mm1) was also examined. We compared some reflectance profiles simulated in an anisotropic setting (scattering coefficient μs=10mm1 , anisotropy parameter g=0.9 ) with those in an isotropic one (μs=1.0mm1) and confirmed that the difference between the two was small enough to be ignored. Therefore, isotropic scattering was assumed to reduce the calculation time. A refractive index of 1.37 was used for both the upper and lower layers.

3.2.

Phantom Preparation

Homogeneous and two-layered gelatin phantoms, the sizes of which were 230×230×55to70mm , were prepared. The thicknesses of the upper layers in two-layered phantoms were 10 and 15mm . The base material of phantoms was an 8wt% gelatin (Wako Pure Chemical Industries, Ltd., Osaka, Japan) solution. The gelatin solution itself was transparent and its μa is 0.003mm1 at 760nm . We adjusted μs and μa to 1.0 to 1.5 and 0.003to0.02mm1 at 760nm , respectively, by adding Intralipid (Fresenius Kabi AB, Upsala, Sweden) as scatterers and ink (greenish brown ink; Chugai Kasei Co., Musashino, Japan; μa at 760, 800, and 830nm were 0.247, 0.180, and 0.125mm1 , respectively, in 1vol% aqueous solution) as an absorber. A homogeneous phantom without ink (S0; μs=1.0mm1 , μa=0.003mm1 ) was also prepared as a reference for analysis. We determined μs and μa of each phantom by curve fitting the solution of photon diffusion equation to time-resolved data (the details are described in Sec. 3(D)). To confirm the reproducibility of the results, we prepared eight homogeneous phantoms on different days and/or at different times and compared their optical properties. The thickness of the upper layer and optical properties of phantoms are listed in Table 1 .

Table 1

Sets of optical properties at 760nm of gelatin phantoms used in experiment, where μa and μs′ are theoretical values based on the concentration of ink and Intralipid, respectively.

Phantom NameOptical Coefficient at 760nm (Prepared Value)/ mm−1
Homogeneous μa μs′
S00.0031.0
S10.0081.0
Two-LayeredUpper LayerLower Layer
Thickness/ mm−1 μa μs′ μa μs′
L0100.0031.00.0031.0
L110,150.0131.00.0081.0
L210,150.0031.00.0081.0
L3150.0031.00.0041.0
L4100.0031.00.0151.0
L5100.0031.00.0171.0
L610,150.01051.00.0081.0
L710,150.0191.00.0151.0
L8100.0031.00.0081.2
L9100.0181.00.0201.2
L10100.0031.20.0041.0
L11100.0031.20.0081.0
L12100.0131.50.0081.0
L13100.0031.20.0081.2

3.3.

Instrumentation

A single-channel TRS instrument34 (TRS-10, Hamamatsu Photonics KK, Hamamatsu, Japan) was employed in this study. In the TRS-10, three laser diodes with different wavelengths (760, 800, and 830nm ) generate optical light pulses having duration of around 100ps (full width at half maximum; FWHM) at the repetition rate of 5MHz . After adjusting the intensity by an optical attenuator, light pulses are delivered to the sample through an optical light guide [GI type, 200μm core diameter, numerical aperture (NA)=0.25 ]. The strongest power irradiated to the sample is around 30μW at each wavelength. The light pulses passing through the sample are collected by a fiber bundle ( 3mm diameter, NA=0.21 ), and transmitted to a high-speed photomultiplier tube with S-25 photocathode (H6279-MOD, Hamamatsu Photonics KK, Japan) for single-photon detection in the NIR light region. A circuit for time-resolved measurement based on time-correlated single-photon counting (TCSPC) method measures the temporal profile. Minimal data acquisition time is 100ms .

3.4.

Time-Resolved Measurement and Determination of Optical Properties of Phantoms

In phantom experiments, as shown in Fig. 2 , incident and detecting light guide with a separation of 30mm were placed on the upper surface of a phantom. Measurements with an accumulation time of 10s were performed at more than three different positions within an area 40mm apart from the edge of the phantom to avoid the distortion of photon diffusion due to the edge. The count rate was adjusted to 100to150kcps at each wavelength by optical attenuator to prevent the pile-up distortion. We avoided specular reflection and light leakage by employing a black light-guide holder. In every experiment, the instrumental responses of the TRS-10 were measured, facing the input and receiving fibers each other through a neutral density filter in a black tube. The instrumental response was around 150ps FWHM at each wavelength.

Fig. 2

Schematic of the measurement for samples. It performed by reflectance mode, in which both light guides for radiation and detection are arranged in parallel. A mat black plate that has holes for light guides is on the samples as a light guide holder to decrease the error due to specular reflection and light leakage.

064008_1_010506jbo2.jpg

Optical properties in homogeneous phantoms and in each layer of the two-layered phantoms were determined by fitting an analytical solution of the photon diffusion equation to the measured temporal profile with nonlinear least squares regression. The data in a time range of 0to5400ps were selected for fitting. In obtaining the analytical solution, the extrapolated boundary condition of reflectance mode35, 36 was employed. The theoretical profile was convoluted with the measured incident pulse shape in the fitting process. Light velocity in gelatin was assumed to be 0.225mmps , corresponding to its refractive index of 1.33. The same values of these were also used for the calculation of μaseg .

4.

Results

4.1.

Monte Carlo Simulation

4.1.1.

Temporal variations of μaseg

First we examined the temporal variations of μaseg under several conditions in which the μa of each layer were changed. As a matter of convenience, the segments were numbered as I, II, III, etc. in order of time starting from t=500ps (e.g., the segment by time range from 500to1000ps was denoted as segment I and that from 2500to3000ps was as segment V, and so on). Figure 3 shows the variations of μaseg(K) ( K=segment number) of media, the upper layer thickness of which was 10mm and μs of both layers were 1mm1 . The μaseg(K) are expressed as the ratio to the real μa,lower . A homogeneous medium was used as a reference profile ( μs=1mm1 , μa=0.005mm1 ). Roughly, three types of the μaseg(K) curves were observed. The μaseg(K) in the case of the identical μa between the two layers were almost constant in all the time segments. In the case where μa,upper was larger than μa,lower , the ratio of μaseg(K) to the real μa,lower was above 1 at 0.5to1ns (segment I) and gradually decreased with time, converging on 1 in 2.5to3.5ns (segments V and VI). On the other hand, in the case where μa,upper was smaller than μa,lower , μaseg(I) was below 1 at 0.5to2ns and increased with time. Even in a later time period, μaseg(K) did not reach the μa,lower .

Fig. 3

Temporal variations of μaseg(K) of the two-layered media, where the upper layer thickness was 10mm and μs of both layers were 1mm1 , predicted by Monte Carlo simulation. The vertical axis represents the ratio of μaseg(K) to the real μa,lower . A homogeneous medium was used as a reference profile ( μs=1mm1 , μa=0.005mm1 ).

064008_1_010506jbo3.jpg

4.1.2.

Effects of the difference in scattering on μaseg

Next, using two additional two-layered models in which upper μs was different from the lower μs , we examined the effects of the difference in μs between the two layers on the μaseg(K) . Figure 4(a) shows the temporal variations of μaseg(K) of the media with different μs between the two layers as well as those of media with the same μs . As the reference profile, a homogeneous medium in which μs was the same as that of the upper layer ( μs=1 or 1.5mm1 , μa=0mm1 ) was used. The vertical axis represents the ratio of μaseg(K) to the real μa,lower . Even though upper and lower μs values were different from each other in the range from 1.0to1.5mm1 , the temporal variation of μaseg(K) was very similar to that in the case in which both layers had the same μs .

Fig. 4

(a) Temporal variations of μaseg(K) of two-layered media with different μs between the two layers together with those of media with the same μs predicted by Monte Carlo simulation. The upper layer thickness was 10mm . As a reference profile, a homogeneous medium in which μs was the same as that of the upper layer ( μs=1 or 1.5mm1 , μa=0.005mm1 ) was used. The vertical axis represents the ratio of μaseg(K) to the real μa,lower . (b) Temporal variations of μaseg(K) of two-layered media (upper layer: μs=1mm1 , μa=0.015mm1 , lower layer: μs=1mm1 , μa=0.01mm1 ; upper layer thickness=10mm ) predicted by Monte Carlo simulation and analyzed using three different homogeneous references [ μs=0.8 (▵), 1.0 (∎), 1.2mm1 (엯); μa=0.005mm1 ].

064008_1_010506jbo4.jpg

Figure 4(b) shows the effects of the μs difference between an object and a reference on the temporal variations of μaseg(K) . Each μaseg(K) was calculated by a reference in which μs was the same (1.0mm1) as, larger (1.2mm1) than, and smaller (0.8mm1) than that of upper layer of an object. When we used a reference with larger μs , μaseg(K) in the earlier time segments were larger than μa estimated by a reference with the same μs . In contrast, when a reference with smaller μs was used, μaseg(I) was smaller than that in the case of the same μs . The μaseg(VI) calculated by a 1.2-mm1 reference and the μaseg(V) and (VI) by the 0.8-mm1 reference were almost equal to that calculated by the reference with the same μs(1.0mm1) .

4.1.3.

Effect of the μa difference between the upper and lower layer on the μaseg in later segments

Figure 5 shows the μaseg(IV) (V), and (VI) of two-layered media (upper layer thickness=10mm ) predicted by Monte Carlo simulation as a function of the ratio of μa,upper to μa,lower (μaratioupperlower) . The data obtained in the cases in which μs was equal (the models in Sec. 4.1.1) and different (the models in Sec. 4.1.2) between the two layers were plotted together. As the reference profile, a homogeneous medium in which μs was the same as that of the upper layer was used. When the μaratioupperlower was equal to or more than 1, the μaseg(IV) , (V), and (VI) nearly represented the real μa . The error was less than 5.0% in segment VI. On the other hand, the μaseg(K) even in these later time periods were smaller than the real μa in the range of the μaratioupperlower less than 1. The deviation grew larger as the μaratioupperlower decreased.

Fig. 5

The μaseg (IV) (◻), (V) (◆), and (VI) (▵) of two-layered media predicted by Monte Carlo simulation as a function of the ratio of μa,upper to μa,lower(μaratioupperlower) . Upper layer thickness was 10mm . The data in both cases that μs was identical or different between the two layers were plotted together. A homogeneous medium in which μs was the same as that of the upper layer was used as the reference profile.

064008_1_010506jbo5.jpg

4.2.

Phantom Experiment

4.2.1.

Optical properties of phantoms

Maximum deviation of the optical properties of phantoms in intra- and inter-samples from the mean value were 3 and 8%, respectively. Temporal profiles of detected light intensity in homogeneous (S0) and two-layered phantom (L0), which consists of two layers with the same optical properties as those of S0, were almost in agreement (data not shown). It was thus concluded that the effect of the interface between two layers on the pattern of photon propagation was negligible in present study conditions.

4.2.2.

Temporal variation of μaseg

Whichever wavelength (760, 800, and 830nm ) we used, the results obtained from the following analyses of phantoms were almost the same. Thus, we present the data at 760nm as representative results. In the phantoms with values of μa larger than 0.015mm1 , the profiles of Adiff(t) were reliable only in the time range of 0.5to2ns . Therefore, we mainly show the results of phantoms with μa of 0.003to0.013mm1 , for which our instrument shows adequate performance, although they were smaller than that of human head generally estimated. Figure 6 shows the typical time-resolved reflectance [Fig. 6(a)] and Adifft curve [Fig. 6(b)] obtained by phantom measurements. In this condition, the data were reliable in the time period from 0.35to3ns . In the early time period before 0.25ns and in the later time period after 3ns , data were noisy because of the interaction of incident light and/or the decrease in detected photon numbers. Therefore, analysis for slope of Adiff(t) (calculation of μaseg ) was performed in the range of 0.5to3ns . The segments were numbered as I, II, III, etc. in order of time in the same manner as the case in Sec. 4(A).

Fig. 6

Typical time-resolved reflectance and Adifft curve obtained by phantom measurements: (a) temporal profiles of incident light (IN), intensity of the detected light through a homogeneous [IR(t)] and a two-layered medium [I(t)] and (b) Adifft curve obtained from the data in (a) in the time period of 0to4ns . The data of detected light intensity through the phantoms are deconvoluted by the temporal profile of the incident light and Adiff at each time was calculated. The Roman numerals show numbers of segments of the curve parted every 0.5ns to get the μaseg .

064008_1_010506jbo6.jpg

Figure 7 shows temporal variations of μaseg(K) in two-layered phantoms (measured optical properties of L1: μs=1.14±0.02mm1 , μa,upper=0.0143±0.0001mm1 , μa,lower=0.0088±0.0002mm1 ; L2: μs=1.12±0.01mm1 , μa,upper=0.0030±0.0001mm1 , μa,lower=0.0081±0.0001mm1 ) with the upper layer thickness of 10mm as well as that of the homogeneous phantom (S1: μs=1.06±0.02mm1 , μa=0.0084±0.0002mm1 ). The μaseg(K) are expressed as the ratio to the real μa,lower . The μs of the reference was the same as that of the upper and lower layers of these phantoms. The μaseg(K) of S1 were almost constant in all the time segments, and the ratio to the real μa,lower was 0.96±0.06 . As for L1 (μaratioupperlower=1.66) , the ratio of μaseg(I) to the real μa,lower was 1.37±0.05 . It gradually decreased with time to 0.98±0.07 in segment IV and 0.91±0.01 in segment V. On the other hand, in L2 (μaratioupperlower=0.40) , μaseg(I) was 0.56±0.04 and increased with time to 0.80±0.05 in 3ns (segment V).

Fig. 7

Temporal variations of μaseg(K) of two-layered phantoms (L1 and L2), of which upper layer thickness was 10mm , and homogeneous phantom of S1 (L1: μs=1.14±0.02mm1 , μa,upper=0.0143±0.0001mm1 , μa,lower=0.0088±0.0002mm1 ; L2: μs=1.12±0.01mm1 , μa,upper=0.0030±0.0001mm1 , μa,lower=0.0081±0.0001mm1 ; S1: μs=1.06±0.05mm1 , μa=0.0084±0.0002mm1 ). A homogeneous phantom S0 ( μs=1.01±0.04mm1 , μa=0.0030±0.0001mm1 ) was used as a reference. The plots represent mean ±SD (standard deviation) ratio of μaseg(K) to the real μa,lower (L1, L2: n=3 positions ×2 samples; S1: n=3 positions ×8 samples). Top dotted line and bottom broken line indicate the real μa,upper shown as the ratio to the real μa,lower in L1 and L2, respectively.

064008_1_010506jbo7.jpg

Figure 8 shows the temporal variations of μaseg(K) of phantoms with an upper layer thickness of 15mm . The optical properties of each layer in L1 and L2 were the same as those of L1 and L2, respectively. For comparison, μaseg(K) of S1 is also shown. Like in the phantoms with an upper layer thickness of 10mm shown in Fig. 7, the pattern of μaseg(K) was dependent on μa,upper . When μa,upper was larger than μa,lower(L1) , the μaseg(V) indicated μa,lower .

Fig. 8

Temporal variations of μaseg(K) of two-layered phantoms ( L1 and L2 ), of which the upper layer thickness was 15mm , and homogeneous phantom of S1 at 760nm using S0 ( μs=1.0mm1 , μa=0.003mm1 ) as a reference. The optical properties of each layer in L1 and L2 are the same as L1 and L2. The notation of the vertical axis is the same as that of Fig. 7. Values are means ±SD ( L1 , L2 : n=3 positions ×2 samples, S1: n=3 positions ×8 samples). Top dotted line and bottom broken line represent the real μa,upper shown as the ratio to the real μa,lower in L1 and L2 , respectively.

064008_1_010506jbo8.jpg

The effects of the scattering difference on the temporal variations of μaseg(K) were found in a similar manner to the results by Monte Carlo simulation (data not shown).

4.2.3.

Effect of the μa difference between the upper and lower layer on the μaseg in later segments

Figure 9 shows the μaseg in segment IV [Fig. 9(a)] and segment V [Fig. 9(b)] as a function of μaratioupperlower by measurements of phantoms in Table 1. As in the Monte Carlo simulation, the μaseg(K) in later time periods depended on the difference in μa values between the upper and the lower layers: it was smaller than the real μa,lower in the range of the μaratioupperlower less than 1, while it almost represented the real μa,lower when the μaratioupperlower was equal to or more than 1. In the case in which upper layer thickness was 15mm , the μaseg was much smaller in the range of the μaratioupperlower below 1, while it was larger than the real values in the range of the μaratioupperlower above 1 in segment IV, but not in segment V. When the μs value of the lower layer was different from that of the reference, there were no significant changes in the relationship between the μaratioupperlower and the μaseg(IV) or (V).

Fig. 9

Plots of μaseg in (a) segment IV and (b) segment V in reference to the ratio of μa,upper to μa,lower (μaratioupperlower) of three kinds of phantoms (Table 1); upper layer thickness is 10mm and μs of both layers are the same as the reference (∎), upper layer thickness is 10mm and μs of the lower layer is different from the reference (◻), and upper layer thickness is 15mm and μs of both layers are the same as the reference (●). The vertical axis represents the ratio of μaseg(K) to the real μa,lower . Each plot was average value by several measurements for one sample.

064008_1_010506jbo9.jpg

5.

Discussion

5.1.

Relationship Between μaseg and Real μa of Each Layer

Time segment analysis, dividing the temporal profile of detected light intensity into segments, leads to some findings on path distribution in the direction of depth of photons detected at each time for layered turbid media with absorbers.

The results obtained from the analyses for Monte Carlo simulation data were generally confirmed by those of phantom experiments data. As one of the mutual results, it was found that the μaseg values in later segments depended on the difference in real μa between two layers. This represents that in later time, the partial path length in each layer depends on both μa values of the layers because the probability that the photons mainly passed in the larger μa layer are detected declines. When μa,upper is smaller than μa,lower , lupperllower at a later time is relatively high and Δlupper cannot be ignored even in later time segments. Thus, μaseg was underestimated by the effect of μa,upper . In contrast, when μa,upper is equal to or larger than μa,lower , llower of photons detected later in time is sufficiently longer than lupper and steadily increase with t , that is, (ΔlupperΔL) can be considered nearly zero in later time periods, as expected. The presented results, for which μaseg were underestimated if μa,upper was smaller than μa,lower , were in agreement with a study by Hielscher, 22 who employed a curve-fitting method with homogeneous diffusion equation to later part of the time-resolved reflectance.

Similarly, μaseg(I) did not give the real μa,upper values correctly, which were also affected by the difference in μa between two layers (data not shown). As shown in Fig. 1, the photons detected at 500to1000ps have a low probability of passing through layers lower than the depth of 10mm . Although Δlupper is dominant to ΔL , ΔllowerΔL cannot be taken as zero even in this period, especially when the thickness of the upper layer is less than 10mm . To estimate μa,upper more properly, it might be effective to use an earlier time segment of Adifft curve (e.g., at 350to700ps ). But due to the interference of incident light, it seems difficult to select the optimal time period in practical measurements.

5.2.

Estimation of μaratioupperlower by the Ratio of μaseg at an Early Time to that at a Later Time

In spite of the discrepancy of μaseg from the real μa , our method gives us enough information to know the difference in μa between the upper and lower layers. Moreover, based on the relationship between the extent of underestimation of μa and the ratio of the real μa of two layers (Fig. 5), we would derive μa,lower by correction of μaseg under the condition where μs of the upper layer and that of a reference are the same and upper layer thickness is almost known. For judgment of the μaratioupperlower , we attempted to use the μaseg at an early time (segment I) and at a later time (segment V) (μasegratioIV) . First, under the condition where μs of the reference was the same as that of the upper layer of the medium, we examined the μasegratioIV of various media with different μaratioupperlower predicted by the Monte Carlo simulation. Figure 10 shows the plots of μaratioupperlower to μasegratioIV . There was a strong linear correlation between these values (r2=0.99) . Then we checked the relationship between the ratio of the μaseg (V) to the real μa,lower and the μasegratioIV for both data sets by Monte Carlo simulation [Fig. 11(a) ] and by phantom experiment [Fig. 11(b)]. In the range of this ratio below 1, the μaseg(V)μa,lower linearly related to the μasegratioIV , i.e., we could express them as μaseg(V)μa,lower=a×μasegratioIV+b , where a and b are constants. By regression analysis, the relationships and the correlation coefficients were a=0.85 , b=0.16 , and r=0.80 in the plot with Monte Carlo data, and a=0.84 , b=0.18 , and r=0.96 in that by the phantom experiment, which were almost in agreement. Segment IV and segment VI could also be used for the estimation of μa,lower . But under the condition of this study, segment V was the most adequate to analyze μa,lower , because segment IV included more information of the upper layer [cf. the case of the upper thickness of 15mm in Fig. 9(a)] and segment VI had higher noise than segment V.

Fig. 11

Plots of μaseg (V) with reference to the ratio of the μaseg (I) to (V) (μasegratioIV) in (a) Monte Carlo simulation data and (b) two-layered phantoms with an upper layer thickness of 10mm (◻) and 15mm (●), analyzed by a reference of which μs was the same as that of the upper layer. The vertical axis represents the ratio of μaseg(K) to the real μa,lower . Broken lines are regression lines within the range of μasegratioIV below 1.

064008_1_010506jbo11.jpg

Fig. 10

Plots of the ratio of μa,upper to μa,lower ( μaratioupperlower ) in reference to the ratio of μaseg (I) to (V) ( μasegratioIV ) for the Monte Carlo simulation data.

064008_1_010506jbo10.jpg

Consequently, we could evaluate μaratioupperlower from μasegratioIV and we could estimate the μa,lower by dividing μaseg(K) in later time by the value of (a×μasegratioIV+b) . In Fig. 11, the data in both cases in which μs was equal and different between the two layers within the range of 1.0to1.5mm1 were plotted together. Therefore, the equation for the relation between μaseg(V)μa,lower and μasegratioIV holds its validity at least under the condition that μs is 1.0to1.5mm1 and μa is 0.003to0.02mm1 of each layer in the two-layered media.

5.3.

Effect of the Scattering Coefficient on Estimation of μa of the Lower Layer

As shown in Fig. 4(a), although the μs of the lower layer was different from that of the reference, the μaseg values were almost identical with those for the case where the μs of the lower layer was the same as that of the reference. In contrast, as shown in Fig. 4(b), the difference in μs between the upper layer and the reference significantly influenced the μaseg in earlier time segments. These results indicate that Sdiff(t) in Eq. 2 is almost equal to the difference in scattering function only between the shallower part of the media and the reference. In addition, our results showed that the dependence of μaseg after 2.5ns on the difference in μs between the upper layer and the reference was weak. These results are explained by the analytical solution derived by Patterson [Eqs. (7) and (8), Ref. 18]. In the case of homogeneous media, time differential of Sdiff(t) in Eq. 2 [or in the first line of Eq. 3] in this paper can be represented as follows: dSdiff(t)dt=[ρ2(4DRcR)ρ2(4DOc)]t2 , where DR and DO are the diffusion coefficients for the reference and the object, respectively. As [ρ2(4DRcR)ρ2(4DOc)] is constant, dSdiff(t)dt is in proportion to the reciprocal of the square of time. Thus, the dSdiff(t)dt is negligible at the later time. Similarly, in the case of the two-layered media, it can be considered that dSdiff(t)dt is very small at the later time, i.e., the effect of the μs difference on the μaseg at the later time segments can be negligible. In contrast, the scattering term cannot be ignored in the segment I.

5.4.

Determination of Proper Reference with Time Segment Analysis

We have showed so far the results mainly in the case where μs of the reference was the same as that of the upper layer of the medium. However, for proper analysis with the correction in Sec. 5(B), we must select an adequate reference with μs very close to that of the upper layer of the target medium that was actually unknown. We can show one of the methods to determine the adequate reference.

As for a homogeneous medium, if μs of an object and a reference are the same, which implies Sdiff(t) is zero in Eq. 2, the Adifft relation becomes linear. That is, if we select a reference that makes Adiff(t) a straight line, the μs of the reference is the same as that of the object.37 In layered media, as mentioned in Sec. 5(C), Sdiff(t) mainly depended on the difference in μs between the upper layer and the reference. Thus, we applied the precedings concept that Sdiff(t) is estimated based on the degree of linear relation for Adifft curve to two-layered media to select an adequate reference. Figure 12(a) shows Adifft curves of two-layered medium (upper layer, μs=1mm1 , μa=0.01mm1 , and lower layer, μs=1.5mm1 , μa=0.015mm1 , upper layer thickness=10mm ) in 400to1000ps , predicted by Monte Carlo simulation and analyzed using five references that had different μs values within the range of 0.6to1.4mm1 (μa=0mm1) . We examined the coefficient of determination r2 of each curve by linear regression analysis. It could be found that when a reference was used whose μs was the same as that of the upper layer of an object, r2 was extremely close to 1. It indicates that when thickness of the upper layer is larger than 10mm , the change in μadiff(t) with t is so small as compared with that of Sdiff(t) in time segment I and in time earlier than segment I. Thus, using this method of investigating linear relation for Adifft at an earlier time, we can select an appropriate reference. Figure 12(b) shows estimated μa values of the lower layer and the errors for this model analyzed by various references already described. The correction of μaseg (V) with equation in Sec. 5(B) was used to estimate these μa . It was found that when a reference was used whose μs value was the same as that of the upper layer of an object, the estimated μa was almost the same as the real μa . We also found that when the difference in μs between a reference and the upper layer of an object was smaller than 0.5mm1 , the error of estimated μa was less than 10%. In other words, we should prepare a series of reference in which μs were changed by every 0.5mm1 within the range of the human extracerebral tissue to estimate reliable μa by this method.

Fig. 12

(a) The Adifft curves of a two-layered medium (upper layer: μs=1mm1 , μa=0.01mm1 ; and lower layer: μs=1.5mm1 , μa=0.015mm1 ; upper layer thickness=10mm ) in 400to1000ps , predicted by Monte Carlo simulation and analyzed using five references that had various μs values within the range of 0.6to1.4mm1 (μa=0mm1) . (b) Estimated value of μa,lower and the error caused by analyzing with each reference which μs was within the range of 0.8to1.6mm1 . The correction of μaseg (V) with the equation in Sec. 5(B) was used to estimate these μa values.

064008_1_010506jbo12.jpg

5.5.

Possibility of Applying the Present Method to Human Head Measurements

The time segment analysis of time-resolved reflectance in two-layered media enabled us to simply acquire information on the μa of each layer and was implemented for a reflectance by measurement with a single source-detector distance. Moreover, we found that the estimated value of μa,lower depended on the value of μaratioupperlower , i.e., μaseg were underestimated if μa,upper was smaller than μa,lower , which was almost the same as the results in the case of fitting analysis for the later part of time-resolved reflectance by a solution of diffusion equations for homogeneous media.22 In our method, however, the difference in μa(μaratioupperlower) can be evaluated using μaseg both at earlier time and at later time (μasegratioIV) .

For the adult human forehead, the total thickness of scalp and skull is28, 38 about 10mm . Thus, the μa of the brain may be estimated by μaseg (IV) to (VI) if the human head could be approximated to two-layered model. On the other hand, the human head is a multilayered structure and the scattering phenomenon must be more complex. Therefore we require further experiments with a more sophisticated model of the human head to confirm the utility and limitations of this method. For example, relation between the discrepancies of estimated μa from the real μa and the ratio of μaseg at an earlier time to that at a later time in multilayered models, and dependence of the layer thickness on the relationship should be examined. As the segment size, a time range of 500ps was chosen in this study. For a multilayered model, however, we should evaluate the segment size to obtain better results. Moreover reliable estimation requires determining μs values of the scalp and skull, which vary with individuals and each position on the head.

6.

Conclusion

We confirmed that the “time segment analysis” of time resolved reflectance for a two-layered media could selectively determine the values of μa,lower . Under the condition where μs of a reference profile is the same as that of the upper layer of an object medium, this analysis enables us to estimate the deference in μa between the upper and the lower layers. It was also possible to determine μa,lower by correcting the μaseg with the ratio of μaseg in an earlier time segment and to that in a later one if we have rough knowledge of the upper layer thickness. In conclusion, the presented approach has a potential to selectively determine the value of μa and to quantitatively evaluate concentrations of the Hb in human cerebral tissue.

Acknowledgments

The authors very much appreciate the helpful discussion we had with Mr. Y. Yamashita and Dr. M. Oda at Hamamatsu Photonics KK.

References

1. 

Y. Hoshi and M. Tamura, “Detection of dynamic changes in cerebral oxygenation coupled to neuronal function during mental work in man,” Neurosci. Lett., 150 5 –8 (1993). https://doi.org/10.1016/0304-3940(93)90094-2 0304-3940 Google Scholar

2. 

T. Kato, A. Kamei, S. Takashima, and T. Ozaki, “Human visual cortical function during photic stimulation monitoring by means of near-infrared spectroscopy,” J. Cereb. Blood Flow Metab., 13 (3), 516 –520 (1993). 0271-678X Google Scholar

3. 

A. Villringer, J. Planck, C. Hock, L. Schleinkofer, and U. Dirnagl, “Near infrared spectroscopy (NIRS): a new tool to study hemodynamic changes during activation of brain function in human adults,” Neurosci. Lett., 154 (1–2), 101 –104 (1993). https://doi.org/10.1016/0304-3940(93)90181-J 0304-3940 Google Scholar

4. 

T. J. Germon, A. E. Young, A. R. Manara, and R. J. Nelson, “Extracerebral absorption of near infrared light influences the detection of increased cerebral oxygenation monitored by near infrared spectroscopy,” J. Neurol., Neurosurg. Psychiatry, 58 (4), 477 –479 (1995). 0022-3050 Google Scholar

5. 

D. N. Harris and S. M. Bailey, “Near infrared spectroscopy in adults. Does the Invos 3100 really measure intracerebral oxygenation?,” Anaesthesia, 48 (8), 694 –696 (1993). 0003-2409 Google Scholar

6. 

H. Owen-Reece, C. E. Elwell, W. Harkness, J. Goldstone, D. T. Delpy, J. S. Wyatt, and M. Smith, “Use of near infrared spectroscopy to estimate cerebral blood flow in conscious and anaesthetized adult subjects,” Br. J. Anaesth., 76 (1), 43 –48 (1996). 0007-0912 Google Scholar

7. 

P. W. McCormick, M. Stewart, M. G. Goetting, M. Dujovny, G. Lewis, and J. I. Ausman, “Noninvasive cerebral optical spectroscopy for monitoring cerebral oxygen delivery and hemodynamics,” Crit. Care Med., 19 (1), 89 –97 (1991). 0090-3493 Google Scholar

8. 

T. J. Germon, P. D. Evans, N. J. Barnett, P. Wall, A. R. Manara, and R. J. Nelson, “Cerebral near infrared spectroscopy: emitter-detector separation must be increased,” Br. J. Anaesth., 82 (6), 831 –837 (1999). 0007-0912 Google Scholar

9. 

A. Kienle, T. Glanzmann, G. Wagnière, and H. van den Bergh, “Investigation of two-layered turbid media with time-resolved reflectance,” Appl. Opt., 37 (28), 6852 –6862 (1998). 0003-6935 Google Scholar

10. 

M. Larsson, H. Nilsson, and T. Strömberg, “In vivo determination of local skin optical properties and photon path length by use of spatially resolved diffuse reflectance with application in laser Doppler flowmetry,” Appl. Opt., 42 (1), 124 –134 (2003). 0003-6935 Google Scholar

11. 

H. Laufs, A. Kleinschmidt, A. Beyerle, E. Eger, A. Salek-Haddadi, C. Preibisch, and K. Krakow, “EEG-correlated fMRI of human alpha activity,” Neuroimage, 19 (4), 1463 –1476 (2003). 1053-8119 Google Scholar

12. 

M. Moosmann, P. Ritter, I. Krastel, A. Brink, S. Thees, F. Blankenburg, B. Taskin, H. Obrig, and A. Villringer, “Correlates of alpha rhythm in functional magnetic resonance imaging and near infrared spectroscopy,” Neuroimage, 20 (1), 145 –158 (2003). 1053-8119 Google Scholar

13. 

S. D. Bianco, F. Martelli, and G. Zaccanti, “Penetration depth of light re-emitted by a diffusive medium: Theoretical and experimental investigation,” Phys. Med. Biol., 47 4131 –4144 (2002). https://doi.org/10.1088/0031-9155/47/23/301 0031-9155 Google Scholar

14. 

B. Montcel, R. Chabrier, and P. Poulet, “Detection of cortical activation with time-resolved diffuse optical method,” Appl. Opt., 44 (10), 1942 –1947 (2005). https://doi.org/10.1364/AO.44.001942 0003-6935 Google Scholar

15. 

J. Selb, J. J. Stott, M. A. Franceschini, A. G. Sorensen, and D. A. Boas, “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 (2005). https://doi.org/10.1117/1.1852553 1083-3668 Google Scholar

16. 

J. Steinbrink, H. Wabnitz, H. Obrig, A. Villringer, and H. Rinneberg, “Determining changes in NIR absorption using a layered model of the human head,” Phys. Med. Biol., 46 879 –896 (2001). https://doi.org/10.1088/0031-9155/46/3/320 0031-9155 Google Scholar

17. 

G. Zaccanti, D. Contini, M. Guriori, A. Ismaelli, H. Liszka, and A. Sassaroli, “Detectability of inhomogeneities within highly diffusing media,” Proc. SPIE, 2389 755 –762 (1995). 0277-786X Google Scholar

18. 

M. S. Patterson, B. Chance, and B. C. Wilson, “Time resolved reflectance and transmittance for the non-invasive measurement of tissue optical properties,” Appl. Opt., 28 (12), 2331 –2336 (1989). 0003-6935 Google Scholar

19. 

E. M. Sevick, B. Chance, J. Leigh, S. Nioka, and M. Maris, “Quantification of time- and frequency-resolved optical spectra for the determination of tissue oxygenation,” Anal. Biochem., 195 330 –351 (1991). https://doi.org/10.1016/0003-2697(91)90339-U 0003-2697 Google Scholar

20. 

F. Fabbri, M. E. Henry and S. Fantini, “Optical measurements of absorption changes in two-layered diffusive media,” Phys. Med. Biol., 49 1183 –1201 (2004). https://doi.org/10.1088/0031-9155/49/7/007 0031-9155 Google Scholar

21. 

T. J. Farrell, M. S. Patterson, and M. Essenpreis, “Influence of layered tissue architecture on estimates of tissue optical properties obtained from spatially resolved diffuse reflectometry,” Appl. Opt., 37 (10), 1958 –1972 (1998). 0003-6935 Google Scholar

22. 

A. H. Hielscher, H. Liu, B. Chance, F. K. Tittel, and S. L. Jacques, “Time-resolved photon emission from layered turbid media,” Appl. Opt., 35 719 –728 (1996). 0003-6935 Google Scholar

23. 

R. J. Hunter, M. S. Patterson, T. J. Farrell, and J. E. Hayward, “Haemoglobin oxygenation of a two-layer tissue-simulating phantom from time-resolved reflectance: effect of top layer thickness,” Phys. Med. Biol., 47 (2), 193 –208 (2002). https://doi.org/10.1088/0031-9155/47/2/302 0031-9155 Google Scholar

24. 

A. Kienle, M. S. Patterson, N. Dögnitz, R. Bays, G. Wagnières, and H. van den Bergh, “Noninvasive determination of the optical properties of two-layered turbid media,” Appl. Opt., 37 779 –791 (1998). 0003-6935 Google Scholar

25. 

A. Kienle and T. Glanzmann, “In vivo determination of the optical properties of muscle with time-resolved reflectance using a layered model,” Phys. Med. Biol., 44 2689 –2702 (1999). https://doi.org/10.1088/0031-9155/44/11/301 0031-9155 Google Scholar

26. 

F. Martelli, A. Sassaroli, Y. Yamada, and G. Zaccanti, “Analytical approximate solutions of the time-domain diffusion equation in layered slabs,” J. Opt. Soc. Am. A, 19 (1), 71 –80 (2002). 0740-3232 Google Scholar

27. 

F. Martelli, S. Del Bianco, G. Zaccanti, A. Pifferi, A. Torricelli, A. Bassi, P. Taroni, and R. Cubeddu, “Phantom validation and in vivo application of an inversion procedure for retrieving the optical properties of diffusive layered media from time-resolved reflectance measurements,” Opt. Lett., 29 (17), 2037 –2039 (2004). https://doi.org/10.1364/OL.29.002037 0146-9592 Google Scholar

28. 

H. Wabnitz, H. Obrig, M. Moller, R. Macdonald, A. Villringer, and H. Rinneberg, “Time-resolved multidistance near-infrared spectroscopy of the adult head: intracerebral and extracerebral absorption changes from moments of distribution of times of flight of photons,” Appl. Opt., 43 (15), 3037 –3047 (2004). https://doi.org/10.1364/AO.43.003037 0003-6935 Google Scholar

29. 

K. Furutsu and Y. Yamada, “Diffusion approximation for a dissipative random medium and the applications,” Phys. Rev. E, 50 3634 –3640 (1994). https://doi.org/10.1103/PhysRevE.50.3634 1063-651X Google Scholar

30. 

Y. Nomura, O. Hazeki, and M. Tamura, “Relationship between time-resolved and non-time-resolved Beer-Lambert law in turbid media,” Phys. Med. Biol., 42 1009 –1022 (1997). https://doi.org/10.1088/0031-9155/42/6/002 0031-9155 Google Scholar

31. 

L. Wang and S. L. Jacques, “Monte Carlo modeling of light transport in multi-layered tissues in standard C,” (1992) Google Scholar

32. 

G. Zaccanti, A. Taddeucci, M. Barilli, P. Bruscaglioni, and F. Martelli, “Optical properties of biological tissues,” Proc. SPIE, 2389 513 –521 (1995). 0277-786X Google Scholar

33. 

J. H. Choi, M. Wolf, V. Toronov, U. Wolf, C. Polzonetti, D. Hueber, L. P. Safonova, R. Gupta, A. Michalos, W. Mantulin, and E. Gratton, “Noninvasive determination of the optical properties of adult brain: near-infrared spectroscopy approach,” J. Biomed. Opt., 9 (1), 221 –219 (2004). https://doi.org/10.1117/1.1628242 1083-3668 Google Scholar

34. 

M. Oda, Y. Yamashita, T. Nakano, A. Suzuki, K. Shimizu, I. Hirano, F. Shimomura, E. Ohmae, T. Suzuki, and Y. Tsuchiya, “Nearinfrared time-resolved spectroscopy system for tissue oxygenation monitor,” Proc. SPIE, 3579 611 –617 (1999). 0277-786X Google Scholar

35. 

M. S. Patterson, S. J. Madsen, J. D. Moulton, and B. C. Wilson, “Diffusion equation representation of photon migration in tissue,” IEEE 1991 Microwave Symp. Digest, 905 –908 (1991) Google Scholar

36. 

A. H. Hielscher, S. L. Jacques, L. Wang, F. K. Tittel, “The influence of boundary conditions on the accuracy of diffusion theory in time-resolved reflectance spectroscopy of biological tissues,” Phys. Med. Biol., 40 (11), 1957 –1975 (1995). https://doi.org/10.1088/0031-9155/40/11/013 0031-9155 Google Scholar

37. 

K. Tanaka, Y. Tanikawa, R. Araki, Y. Yamada, and E. Okada, “Measurement of optical properties of biological tissues by using time-resolved transmittance spectroscopy,” Kougaku, 31 (12), 886 –892 (2002). Google Scholar

38. 

H. Zhao, Y. Tanikawa, F. Gao, Y. Onodera, A. Sassaroli, K. Tanaka, and Y. Yamada, “Maps of optical differential pathlength factor of human adult forehead, somatosensory motor, occipital regions at multi-wavelengths in NIR,” Phys. Med. Biol., 47 1 –18 (2002). https://doi.org/10.1088/0031-9155/47/1/301 0031-9155 Google Scholar
©(2005) Society of Photo-Optical Instrumentation Engineers (SPIE)
Chie Sato, Miho Shimada, Yoko Hoshi M.D., and Yukio Yamada "Extraction of depth-dependent signals from time-resolved reflectance in layered turbid media," Journal of Biomedical Optics 10(6), 064008 (1 November 2005). https://doi.org/10.1117/1.2136312
Published: 1 November 2005
Lens.org Logo
CITATIONS
Cited by 17 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Reflectivity

Monte Carlo methods

Photons

Scattering

Optical properties

Head

Picosecond phenomena

Back to Top