Open Access
9 June 2014 Sensitivity correction for the influence of the fat layer on muscle oxygenation and estimation of fat thickness by time-resolved spectroscopy
Etsuko Ohmae, Shinichiro Nishio, Motoki Oda, Hiroaki Suzuki, Toshihiko Suzuki, Kyoichi Ohashi, Shunsaku Koga, Yutaka Yamashita, Hiroshi Watanabe
Author Affiliations +
Abstract
Near-infrared spectroscopy (NIRS) has been used for noninvasive assessment of oxygenation in living tissue. For muscle measurements by NIRS, the measurement sensitivity to muscle (SM ) is strongly influenced by fat thickness (FT). In this study, we investigated the influence of FT and developed a correction curve for SM with an optode distance (3 cm) sufficiently large to probe the muscle. First, we measured the hemoglobin concentration in the forearm (n=36 ) and thigh (n=6 ) during arterial occlusion using a time-resolved spectroscopy (TRS) system, and then FT was measured by ultrasound. The correction curve was derived from the ratio of partial mean optical path length of the muscle layer ⟨LM ⟩ to observed mean optical path length ⟨L⟩ . There was good correlation between FT and ⟨L) at rest, and ⟨L⟩ could be used to estimate FT. The estimated FT was used to validate the correction curve by measuring the forearm blood flow (FBF) by strain-gauge plethysmography (SGP_FBF) and TRS (TRS_FBF) simultaneously during a reactive hyperemia test with 16 volunteers. The corrected TRS_FBF results were similar to the SGP_FBF results. This is a simple method for sensitivity correction that does not require use of ultrasound.

1.

Introduction

Near-infrared spectroscopy (NIRS) in the 700 to 900 nm region is useful for continuous and noninvasive assessment of oxygenation in living tissue. This method is based on the relatively high transparency of tissue to NIR light and the dependence of light absorption in this region on the hemoglobin (Hb) concentration in the tissue. The conventional continuous wave (CW) method is used to determine relative changes in Hb concentrations,14 while time-resolved spectroscopy (TRS),58 phase modulation spectroscopy,912 and broadband CW13 methods have been developed for quantitation of Hb concentrations in scattering media. However, NIRS signals also arise from other tissues, such as scalp, bone, and cerebrospinal fluid (CSF) for head measurements, and skin and fat for muscle measurements. The effects of multilayer tissue structures have been studied in simulations and phantom experiments.1423

For brain measurement, the effects of CSF and skin blood flow have been discussed.14,22,24 Many studies have shown the feasibility of assessing cerebral hemodynamics with NIRS.2527 Increasing the optode spacing in experimental measurements can reportedly allow for brain signal detection by NIRS.2830 The contribution of intracerebral tissue at 4 cm was 70% in multidistance TRS measurement.31,32 Furthermore, the TRS method with diffusion theory6 is reportedly affected less by skin blood flow than the CW method.3335

For muscle measurements, the measurement sensitivity of muscle (SM) is strongly influenced by the fat thickness (FT).3638 There are considerable differences between individuals for FT, but fat tissue structure is not as complicated as brain tissue. A thicker fat layer decreases the partial mean optical path length of the muscle layer LM and results in underestimation of muscle oxygenation. Therefore, quantitative comparison of NIRS data between individuals is difficult because of FT differences. Several approaches have been proposed to solve this problem, including proportional assessment using transient arterial occlusion,39 time constant (τ), and the recovery time for parameters obtained by NIRS. Additionally, correction curves for SM against FT have been estimated from LM of Monte Carlo simulations.17,18,23

In the TRS system, the distribution of mean optical path lengths L can be measured directly. Its temporal profile is analyzed using diffusion theory,6 which enables determination of the Hb concentration. In recent years, TRS has been applied in many areas, including monitoring of cerebral oxygen metabolism during surgery4042 and for neonates,43,44 bed side monitoring of subarachnoid hemorrhage,45 the study of breast cancer,46 and hemodynamic study of muscle.47,48 However, the fundamental issue described above has remained unresolved.

In this study, we used LM/L from TRS to develop a correction curve for SM for changes in the muscle. The correction curve was validated by measuring forearm blood flow (FBF) simultaneously by TRS and strain-gauge plethysmography (SGP) during reactive hyperemia. We also studied the relationship between L and FT.

2.

Materials and Methods

2.1.

TRS

In this study, L, the reduced scattering coefficient (μs), the absorption coefficient (μa), the Hb concentration, and oxygen saturation (SO2) were measured by TRS (TRS-10 for one channel,49,50 TRS-20 for two channels,47,51 Hamamatsu Photonics K.K., Hamamatsu, Japan).

Light pulses were emitted at 760, 795, and 830 nm, with full-width at half maximum of 100 ps, pulse rate of 5 MHz, and an average power of 200 μW. The time-correlated single-photon counting (TCSPC) method was used to measure the temporal profile of detected photons.

2.2.

TRS Data Analysis

L is the mean distance that photons travel through the tissue. This parameter is obtained directly from the mean time delay of the temporal profile by TRS measurement, assuming that the photons travel at a constant speed within the scattering medium.52 Usually, a deconvolution method is used to remove the influence of the measuring system, but this is a complicated process. Zhang et al. proposed a simple subtraction method for rapid and accurate calculation of L without deconvolution.53 In this study, we used Zhang’s method, which subtracts the center of gravity of the instrument response function (IRF) from that of the observed temporal profile. The validity of this method has been well proven numerically and experimentally.

The optical parameters (μs and μa) were derived from the temporal profiles as described below. The behavior of a photon within scattering and absorption media, such as the human body, is expressed by the photon diffusion equation [Eq. (1)].6

Eq. (1)

1νtϕ(r,t)D2ϕ(r,t)+μaϕ(r,t)=S(r,t),
where ϕ(r,t) is the diffuse photon fluence rate at position r and time t, D is the photon diffusion coefficient as expressed by D=1/3μs, ν is the velocity of light within the media, and S(r,t) is the light source.

Solutions using this equation are found under different boundary conditions. We used the solution of a semi-infinite homogeneous model with a zero boundary condition in reflectance mode6 for the TRS data analysis. In this solution, R(d,t) is expressed by Eq. (2) as a function of the optode spacing, μs, and μa.

Eq. (2)

R(d,t)=(4πDν)32z0t52exp(μaνt)exp(d2+z024Dνt),
where d is the optode spacing and z0=1/μs.

Using the weighted nonlinear least-squares method based on the Levenberg-Marquardt method, we fitted Eq. (2) for an instantaneous point source to the observed temporal profiles obtained from TRS and determined μs and μa at each wavelength, while taking into account the effect of IRF.54,55 The weights for the least-squares calculations were obtained from Poisson distribution because the TRS data measured by TCSPC method have an error obeying the Poisson distribution.

We first assumed that absorption in the 700 to 900 nm range arose from oxygenated hemoglobin (oxyHb), deoxygenated hemoglobin (deoxyHb), and water. It is difficult to distinguish between myoglobin (Mb) and Hb because they have similar absorption spectra. In this study, most of the signals were assumed to arise from Hb, and the influence of Mb was ignored.56,57 The μaλ at the measured wavelengths λ (760, 795, and 830 nm) is expressed as shown in Eq. (3).

Eq. (3)

μa760nm=εoxyHb760nmCoxyHb+εdeoxyHb760nmCdeoxyHb+μaH2O760nm,μa795nm=εoxyHb795nmCoxyHb+εdeoxyHb795nmCdeoxyHb+μaH2O795nm,μa830nm=εoxyHb830nmCoxyHb+εdeoxyHb830nmCdeoxyHb+μaH2O830nm,
where εmλ is the molar extinction coefficient of substance m at wavelength λ and Cm is the concentration of substance m. The water absorption (μaH2Oλ) was measured using a conventional spectral photometer (U-3500, Hitachi High-Technologies Corporation, Tokyo, Japan).

After subtracting the water absorption from μa at each wavelength, assuming the volume fraction of the water content was 60%,58 we determined the concentrations of oxyHb and deoxyHb using the least-squares fitting method. The total concentration of hemoglobin (totalHb) and SO2 were calculated from Eqs. (4) and (5).

Eq. (4)

totalHb=oxyHb+deoxyHb,

Eq. (5)

SO2=oxyHbtotalHb×100.

2.3.

Estimation of SM

We estimated SM based on the method of Kohri et al.31 Figure 1 shows a model of fat and muscle layers during arterial occlusion. The optode spacing, d, was sufficient for photons to reach the muscle layer. The regions where light passed through were divided into a fat layer and a muscle layer. Each of these layers was homogeneous within the layer, and the two layers had different μa and μs. FT was defined as the distance from skin surface to muscle, and the influence of the skin was ignored.3335

Fig. 1

Model of fat and muscle layers under arterial occlusion. The photon travels from the source to the detector in curved path. FT, fat thickness (skin + fat layer); L, observed mean optical path length; LF, partial mean optical path length of the fat layer; LM, partial mean optical path length of the muscle layer; Δμa, change in observed absorption coefficient; ΔμaF, change in absorption coefficient of the fat layer; ΔμaM, change in absorption coefficient of the muscle layer; d, optode spacing.

JBO_19_6_067005_f001.png

The observed L was the sum of the partial mean optical path length of the fat layer LF and LM, as follows:

Eq. (6)

L=LF+LM.

Propagation of photons in the fat layer (LF) was regarded as the sum of the optical distances in two directions, from the irradiation point to the muscle layer and from the muscle layer to the detection point. If LF depends on FT, LF is expressed using the differential path length factor of FT with an optode distance d (DPFfat: partial mean path length for one direction per unit FT), as follows:

Eq. (7)

LF=2×FT×DPFfat.

The modified Lambert-Beer’s law for the two-layered structure is expressed as the summation of the optical density of each layer17,59 as follows:

Eq. (8)

OD=L×μa+X=LF×μaF+LM×μaM+X,
where μa is the observed absorption coefficient; μaF and μaM are the absorption coefficients of the fat layer and the muscle layer, respectively; and X is the optical attenuation attributable to scattering.

When the absorption changes only in the muscle layer in arterial occlusion, the change in the observed absorption coefficient (Δμa) is expressed as Eq. (9), where X is cancelled out.

Eq. (9)

Δμa=ΔODL=LML×ΔμaM=(LLF)L×ΔμaM=SM×ΔμaM,
where ΔOD is the change in the optical density, ΔμaM is the change in the absorption coefficient of the muscle layer, and SM is the measurement sensitivity to muscle with optode spacing d, which is expressed by dividing LM by L. The change in absorption coefficient in the fat layer was ignored because fat has very low blood volume and oxygen consumption.6065

In general, deoxyHb has been used as an index of microvascular O2 extraction in muscle studies by NIRS.48,6669 The change in observed deoxyHb (ΔHb) showed an increase in arterial occlusion, and was expressed as a function of SM and the change in deoxyHb of muscle (ΔHbM) instead of Δμa and ΔμaM [Eq. (10)], because ΔHb is directly proportional to Δμa.

We assumed that the differences among individuals for muscle oxygen consumption (mVO2) at rest were small and that the influence of FT on NIRS was larger than any differences among individuals.37,70 Then we determined DPFfat and ΔHbM so that the value of E in Eq. (11) was minimized using the least-squares method for all individuals.

Eq. (10)

ΔHb=SM×ΔHbM=L2×FT×DPFfatL×ΔHbM,

Eq. (11)

E=i=1n{ΔHb(i)[L(i)2×FT(i)×DPFfatL(i)×ΔHbM]}2(n=number of volunteers).

If FT was known, SM was estimated by dividing LM by L using the determined DPFfat, as shown in Eq. (12). Moreover, ΔHb divided by SM provides an estimate of ΔHbM.

Eq. (12)

SM=LML=LLFL=L2×FT×DPFfatL.

2.4.

Protocol

2.4.1.

Arterial cuff occlusion test on the forearm and thigh

Informed consent was obtained from all subjects before the experiment. Arterial cuff occlusion on the forearm was performed on 21 male and 15 female healthy volunteers (37.6±10.3 years of age), and on the thigh on six male healthy volunteers (23.9±5.0 years of age). The optical probe for the forearm was attached to one point on the skin over the left brachioradialis muscle using the TRS-10 system. The optical probes for the thigh were attached to four places on the skin over the right vastus lateralis (VL) and the rectus femoris (RF) muscles at the distal and proximal sites using two TRS-20 systems. All optode distances were 3 cm. Sampling times for the forearm and thigh were 1 and 5 s, respectively. Arterial cuff occlusion tests were performed at rest with the subject in the supine position. TRS data recorded before occlusion were defined as the resting state. The occlusion cuff was placed on the upper forearm or the top of the thigh and connected to a pneumatically powered rapid inflator (E20, D. E. Hokanson Inc., Bellevue, Washington). The cuff on the forearm was inflated for 3 min at 250 mmHg and then deflated. The cuff on the thigh was inflated until the changes in deoxyHb and oxyHb plateaued at 250 mmHg, and then it was deflated. The FT at each optical measurement point was also measured by B-mode ultrasound (USD N-500, NIHON KOHDEN CORP., Tokyo, Japan or Logiq 400, GE-Yokogawa Medical Systems, Japan).

2.4.2.

Simultaneous FBF measurement by SGP and TRS during reactive hyperemia

Informed consent was obtained from all subjects before the experiment. FBF at rest and after vascular endothelial stimulus by arterial occlusion was measured on the left forearm of 12 male and 4 female healthy volunteers (37.9±10.5 years of age). These measurements were recorded by SGP and TRS-10 simultaneously. The venous occlusion method was used to measure FBF.57,71 FBF was observed as an increase in blood volume by interrupted venous flow and influx of arterial blood. By contrast, vascular endothelial stimulus by arterial occlusion leads to vasodilatation and an increase in FBF (reactive hyperemia). The volunteers were in the supine position at rest. The cuff for venous occlusion was placed proximally on the left arm and connected to an automatic pneumatic inflator (E20, D. E. Hokanson Inc.) set to 40 mmHg.7274 A second cuff was placed above the first cuff and connected to another pneumatic cuff inflator to provide arterial occlusion. To record measurements from the same region, wire strain-gauge for SGP (EC-5R, D. E. Hokanson Inc.) and an optical probe were placed in close proximity to each other on the skin over the brachioradialis muscle. Optode spacing was 3 cm, and the sampling time was 1 s. FT was estimated from the relationship between L and FT without measurement by ultrasound.

FBF measurements were performed with continuous cycles of 10 s of occlusion and 10 s of deflation of the venous cuff. First, three resting basal flow measurements were recorded. Next, rapid arterial occlusion of the left arm was executed by inflation of the arterial cuff for 3 min at 250 mmHg. This protocol stimulated reactive hyperemia. After the arterial cuff was suddenly deflated, FBF measurements by venous occlusion were repeated until the FBF value stabilized.

SGP is a popular method for measuring FBF using forearm volume changes during venous occlusion. Results from FBF obtained with SGP (SGP_FBF) are expressed in mL/min/100mL. By contrast, FBF obtained with TRS (TRS_FBF) is defined as the maximum increase in totalHb per unit time (calculated as the slope of a linear fit of totalHb data over 5 s) during venous occlusion. Concentration changes in totalHb (ΔtHb) are expressed in μM/s and are converted to mL/min/100mL by taking into account the Hb concentration (14g/dl) of blood and the molecular weight of Hb (64,500g/mol), as shown in Eq. (13). For adjusting for the influence of fat, the obtained TRS_FBF results were corrected (cTRS_FBF) using Eq. (14).

Eq. (13)

TRS\_FBF=ΔtHb×64,500×6014×105,

Eq. (14)

cTRS\_FBF=TRS\_FBFSM.

2.5.

Statistical Analysis

Quantitative variances are expressed as mean±standard deviation. Comparisons of SGP_FBF and TRS_FBF before and after correction were analyzed using the squared Pearson’s correlation coefficient (r2) and the Bland-Altman test. In the Bland-Altman test, the precision of the bias was defined as two standard deviations of the mean difference, and the presence of fixed and proportional biases was investigated. The influence of FT on the TRS parameter was compared between forearm and thigh results by multiple regression analysis using a dummy variable. Statistical significance was defined with P<0.05.

3.

Results

Figure 2 shows a typical case (case 4, M, 28 years of age, VL at distal site) of TRS parameters (L, μs, μa at three wavelengths), the concentrations of oxyHb, deoxyHb, totalHb, and SO2, and the change in concentration of Hb during arterial occlusion of the thigh. During occlusion, although μs at the three wavelengths was almost constant, L and μa at 760 nm were different from that at the other wavelengths. A decrease in oxyHb and a simultaneous increase in deoxyHb were observed with depletion of local available O2, whereas totalHb remained almost constant. Similar trends were observed in the forearm data.

Fig. 2

Time course of time-resolved spectroscopy (TRS) parameters during arterial occlusion in a typical case [case 4, M, 28 years of age, vastus lateralis (VL) at distal site]: (a) L, (b) μs, (c) μa, (d) Hb concentration, (e) SO2, and (f) change in Hb concentration.

JBO_19_6_067005_f002.png

Figure 3 shows variations of L, μs, and μa at 795 nm, totalHb, and SO2 in forearm and thigh under resting conditions plotted against FT obtained by ultrasound. The FT of the forearm was 0.49±0.08 (range 0.34 to 0.64 cm), while FT of the thigh was 0.44±0.11 (range 0.25 to 0.65 cm). As FT increased, L and μs increased, and μa and totalHb decreased. Moreover, the relationships between FT and each of the TRS parameters, except for L, were significantly different for the forearm and thigh results. By contrast, SO2 was almost constant regardless of FT.

Fig. 3

The changes in (a) L, (b) μs, (c) μa, (d) totalHb, and (e) SO2 of forearm and thigh at rest plotted against FT.

JBO_19_6_067005_f003.png

Figure 4 shows the relationship between FT and ΔHb and the change in SO2 (ΔSO2) during arterial occlusion for the forearm and four thigh positions (VL and RF at distal and proximal sites). Changes in these parameters were estimated in the first 2 min (selected part of totalHb is constant) during arterial occlusion to avoid the effect of Mb.56,57,7577 When FT increased, ΔHb decreased and ΔSO2 increased. Changes in these parameters in the thigh tended to be lower than in the forearm.

Fig. 4

Relationship between FT and the change in (a) deoxyHb and (b) SO2 during arterial occlusion for the forearm and four positions [VL and rectus femoris (RF) at distal and proximal sites] in the thigh.

JBO_19_6_067005_f004.png

DPFfat and ΔHbM of the forearm and thigh were determined by Eq. (11) using FT, ΔHb, and L at 795 nm during rest (Table 1). These parameters were estimated in each position because changes in mVO2 of the thigh depend on the position of the thigh during positron emission tomography (PET).78 The estimated DPFfat did not vary much (average 9.57±0.66), but ΔHbM was substantially different between the forearm and thigh. Figure 5 shows the relationship between FT and SM. SM was obtained from Eq. (12) using DPFfat, L at 795 nm, and FT. SM decayed exponentially with FT.

Table 1

DPFfat and ΔHbM determined using FT and ΔHb and 〈L〉 at 795 nm, μs′ and μa at 795 nm, totalHb and SO2 of each position at rest.

nDPFfatΔHbM (μM)FT (cm)ΔHb (μM)L795nm (cm)μs795nm (cm1)μa795nm (cm1)TotalHb (μM)SO2 (%)
Forearm368.9423.770.47±0.086.16±2.4613.93±1.578.69±1.020.2623±0.037125.9±19.073.9±2.8
VLProximal69.3411.900.41±0.074.85±1.8913.02±1.618.31±1.060.2532±0.0353115.9±17.171.1±0.6
Distal69.5815.400.32±0.077.11±2.0211.25±0.796.43±0.770.2505±0.0270116.0±14.475.7±3.7
ThighRFProximal69.3212.140.51±0.084.82±1.6615.66±1.2710.15±0.980.2175±0.0211103.9±11.374.0±3.1
Distal610.6713.810.51±0.064.67±1.0516.43±1.0510.62±0.730.2070±0.019298.0±8.873.5±1.7
VL, vastus lateralis; RF, rectus femoris.

Fig. 5

Relationship between FT and SM with 3 cm optode spacing estimated from experimental data in the forearm and four positions (VL and RF at distal and proximal sites) in the thigh.

JBO_19_6_067005_f005.png

Figure 6(a) shows typical (case 11, M, 30 years of age) changes in the concentrations of oxyHb, deoxyHb, and totalHb during the forearm reactive hyperemia test. We determined TRS_FBF from the change in totalHb by venous occlusion. The FBF at the time of the first venous occlusion, which occurred immediately after release of arterial occlusion (250 mmHg), was excluded because the measurement conditions at this time differed from those at other times in the test. Figure 6(b) shows the results for SGP_FBF and TRS_FBF. Both these methods showed a large increase in FBF immediately after vascular endothelial stimulus compared with the resting state. The FBF then reduced gradually until rest. Additionally, we corrected TRS_FBF using the correction curve (Fig. 5) and obtained cTRS_FBF. FT was estimated from the relationship [Fig. 3(a)] between FT and L at 795 nm. FT was not measured by ultrasound in this protocol. The result for cTRS_FBF [Fig. 6(b)] was similar to that obtained by SGP_FBF. Table 2 shows L at 795 nm, FT estimated from L, and SM for each volunteer (n=16). The estimated FT was 0.43±0.08 (range 0.28 to 0.57 cm). SM was 0.43±0.05 (range 0.31 to 0.56).

Fig. 6

Time course of changes in (a) concentration of oxyHb, deoxyHb, and totalHb and (b) SGP_FBF, TRS_FBF, and cTRS_FBF during forearm reactive hyperemia tests in a typical case (case 11, M, 30 years of age).

JBO_19_6_067005_f006.png

Table 2

〈L〉 at 795 nm, estimated FT from 〈L〉 and SM in each volunteer (n=16).

Case no.L795nm (cm)Estimated FT (cm)SM
112.770.3950.439
214.860.5250.335
315.610.5680.306
412.590.3820.451
512.810.3970.437
613.480.4410.399
713.250.4260.411
813.550.4460.395
911.860.3310.502
1012.790.3960.438
1114.740.5180.340
1213.360.4340.412
1313.380.4350.405
1414.850.5250.335
1513.370.4340.405
1611.150.2780.560
Average13.400.4330.410
Standard deviation1.160.0750.065

Figure 7 shows the linear regression plot and the Bland-Altman plot of SGP_FBF, TRS_FBF, and cTRS_FBF. The slope improved from 0.28 [Fig. 7(a)] to 0.78 [Fig. 7(b)]. The correlation coefficient (r2) improved slightly from 0.45 [Fig. 7(a)] to 0.53 [Fig. 7(b)] and showed good linearity. The bias and the precision (±2 standard deviations) were 4.64±6.14mL/min/100mL between SGP_FBF and TRS_FBF [Fig. 7(c)], indicating the presence of fixed and proportional biases. By contrast, the bias and the precision between SGP_FBF and cTRS_FBF [Fig. 7(f)] were 0.48±6.25mL/min/100mL, indicating the absence of fixed or proportional biases.

Fig. 7

Correlation between SGP_FBF and (a) TRS_FBF and (b) cTRS_FBF. Difference plots using Bland-Altman for SGP_FBF and (c) TRS_FBF and (d) cTRS_FBF.

JBO_19_6_067005_f007.png

4.

Discussion

4.1.

Influence of FT on TRS Measurement Parameters

The influence of the fat layer on the NIRS signal has been reported for the CW method.3638 Although the TRS method enables estimation of the absorption of the inner layer in media with two layers,7984 we adapted a homogeneous model to facilitate practical application of TRS data analysis. The observed optical parameters (μs and μa), L, and the Hb concentration in the resting state were all influenced by FT (Fig. 3). As FT increased, μs increased and μa decreased. Compared with muscle, fat causes more scattering and has a lower blood volume (absorption).64,65 With increasing FT, the forearm and thigh results showed different trends for μs, μa, and totalHb. Similar results were reported in an earlier in vivo study at 800 nm,85 where the forearm μs was 6.9cm1 and μa was 0.23cm1 and the calf μs was 8.9cm1 and μa was 0.17cm1. In contrast to the other parameters, SO2 was constant and not dependent on FT. This result shows that SO2 in the fat layer and muscle layer are similar at rest. Even if the oxygen metabolism is different in the fat and muscle layers,37,56 a balance between oxygen supply and demand is maintained.

4.2.

Estimation of DPFfat and ΔHbM from Arterial Occlusion

The change in SO2 with arterial occlusion was affected by FT [Fig. 4(b)]. The change in deoxyHb showed the opposite trend to SO2 with the change in FT [Fig. 4(a)]. The changes in SO2 and deoxyHb were caused by an imbalance in O2 consumption between the fat and muscle layers because oxygen metabolism is much higher in muscle than in fat.37,56

We estimated DPFfat and ΔHbM from ΔHb and FT according to Eq. (11), assuming that ΔHb depends on oxygen metabolism in the muscle layer. ΔHbM values differed greatly between the forearm and thigh (Table 1). We compared ΔHbM with those obtained using conventional methods and basal metabolism, by converting ΔHbM to mVO2 values.18 These values were 0.20 and 0.11±0.01mL/100g/min for the forearm and thigh, respectively. By comparison, a resting forearm mVO2 of 0.15mL/100g/min was obtained by the conventional invasive Fick method,86 and 0.16mL/100g/min by phosphorus magnetic resonance spectroscopy.87,88 These values are similar to that obtained in the present study. By contrast, earlier resting leg mVO2 values of 0.18mL/100g/min from a muscle biopsy of human VL89 and 0.19mL/100g/min from PET78 were higher than our result. In the PET study, images suggested that oxygen consumption and blood volume near the surface were lower than in deeper tissue.78 Because NIRS measurements arise from relatively shallow tissue regions, the thigh mVO2 obtained in the present study may have originated from a different tissue depth than the values obtained by the other methods. Johnson et al. reported that deeper tissue has more type 1 fibers than tissue closer to the surface in VL muscle.90 By contrast, the cross-sectional area of the forearm is smaller than that of the thigh. Therefore, the forearm ΔHbM may reflect the signal origin from the deeper tissue of the brachioradialis muscle.

We estimated the thigh ΔHbM in four positions, specifically for the right VL and RF muscles at distal and proximal sites. These measurements were used to investigate the presence of spatially heterogeneous muscle oxygen metabolism. The thigh ΔHbM at the proximal sites were smaller than those at the distal sites. Mizuno et al.78 used PET to study thigh muscle after exercise and reported that mVO2 changes became progressively larger as sampling gradually moved from proximal to distal sites in the muscle. By contrast, in the same study, mVO2 at rest decreased as sampling moved from proximal to distal sites. Our result agreed with the trend observed by Mizuno et al.78 after exercise. Arterial occlusion was performed at rest, but we recognize rapid change occurred in transient hypoxia in our experimental situation.

4.3.

Correction Curve for Sensitivity to Muscle Oxygenation

Our correction curve showed lower sensitivity than the results of Niwayama et al.18 Their correction curve was obtained from a Monte Carlo simulation and in vivo measurement (ergometer exercise and ischemia tests) by the CW method. Specifically, the curve was normalized using the ratio of LM to an FT of 0 mm, with LM calculated from simulations using optical properties from the literature and in vivo data using DPF=4.0. By comparison, our correction curve was derived from a ratio of LM to L as described in Fig. 5. With TRS, L can be measured directly without simulation or DPF. Because L is largely dependent on FT [Fig. 3(a)], the correction curve obtained using the observed L is expected to be more accurate than that of Niwayama, et al.18

4.4.

Prediction of FT by L

Interestingly, the relationship between L and FT was not significantly different between the forearm and thigh [Fig. 3(a)]. In a preliminary examination of this relationship, L at 10 positions on the skin over the VL and RF muscles of each volunteer (n=3) obtained using TRS and FT values obtained by ultrasound (average 0.64±0.08cm, range 0.49 to 0.82 cm) were used. The relationship was consistent with the results of this study (data not shown). These results suggest that L could be used to estimate FT without requiring other methods, such as ultrasound or body fat calipers.

4.5.

Validation of the Correction Curve

The correction curve was validated by comparing SGP_FBF and TRS_FBF results obtained during a reactive hyperemia test. The increase in blood volume from occlusion was ascribed to blood flow. In earlier studies without correction, NIRS_FBF results showed good agreement with SGP_FBF results,9194 although the NIRS values were generally two-to-three times lower than those from SGP.56,86,92 In the present study, the slope of the curve for the comparison of the SGP_FBF and TRS_FBF results before correction was similar to these earlier reports. The reason NIRS_FBF is underestimated is because SGP reflects the total forearm blood flow of adipose, bone, skin, and skeletal muscle tissues, whereas NIRS reflects only the local flow in the region of interest and the signal primarily arises from small vessels.

In addition to the reason discussed above, we also demonstrated that FT had a large influence on NIRS data. cTRS_FBF results were similar to SGP_FBF results, and fixed and proportional biases disappeared (Fig. 7). The slope between cTRS_FBF and SGP_FBF improved from 0.28 to 0.78, and this result was in agreement with reports that hematocrit in capillaries is underestimated by 25% compared with whole body results.56,95

The TRS approach provides better reproducibility than SPG. Watanabe et al. reported that intrasubject variation from day-to-day with TRS was much lower than that with SGP, and SGP was susceptible to small body movements.94 Therefore, the motion error in SGP should also be considered when comparing these techniques.

5.

Conclusion

We determined SM with a 3 cm optode spacing using arterial occlusion data obtained with a TRS system. A curve to correct for sensitivity to changes in muscle oxygenation was obtained using DPFfat, FT, and L. cTRS_FBF results corrected using this curve were similar to the SGP_FBF values in 16 volunteers. In addition, there was good correlation between L and FT in the experimental data. This result indicates FT could be estimated from L without ultrasound. TRS is a simple method to correct for sensitivity to muscle oxygenation without ultrasound. The correction curve is suitable for FT from 0.2 to 0.6 cm. Further work is required to extend the FT range and measurement position to allow for broader application of this method.

Acknowledgments

The authors would like to thank Mr. A. Hiruma for his constant support and encouragement.

References

1. 

M. CopeD. T. Delpy, “System for long-term measurement of cerebral blood and tissue oxygenation on newborn infants by near infra-red transillumination,” Med. Biol. Eng. Comput., 26 (3), 289 –294 (1988). http://dx.doi.org/10.1007/BF02447083 MBECDY 0140-0118 Google Scholar

2. 

P. F. Masonet al., “The assessment of cerebral oxygenation during carotid endarterectomy utilising near infrared spectroscopy,” Eur. J. Vasc. Surg., 8 (5), 590 –594 (1994). http://dx.doi.org/10.1016/S0950-821X(05)80596-1 0950-821X Google Scholar

3. 

T. J. FarrellM. S. PattersonB. Wilson, “A diffusion theory model of spatially resolved, steady-state diffuse reflectance for the noninvasive determination of tissue optical properties in vivo,” Med. Phys., 19 (4), 879 –888 (1992). http://dx.doi.org/10.1118/1.596777 MPHYA6 0094-2405 Google Scholar

4. 

S. Suzukiet al., “Tissue oxygenation monitor using NIR spatially resolved spectroscopy,” Proc. SPIE, 3597 582 –592 (1999). http://dx.doi.org/10.1117/12.356862 PSISDG 0277-786X Google Scholar

5. 

B. Chanceet al., “Time-resolved spectroscopy of hemoglobin and myoglobin in resting and ischemic muscle,” Anal. Biochem., 174 (2), 698 –707 (1988). http://dx.doi.org/10.1016/0003-2697(88)90076-0 ANBCA2 0003-2697 Google Scholar

6. 

M. S. PattersonB. ChanceB. C. Wilson, “Time resolved reflectance and transmittance for the non-invasive measurement of tissue optical properties,” Appl. Opt., 28 (12), 2331 –2336 (1989). http://dx.doi.org/10.1364/AO.28.002331 APOPAI 0003-6935 Google Scholar

7. 

M. Odaet al., “A simple and novel algorithm for time-resolved multiwavelength oximetry,” Phys. Med. Biol., 41 (3), 551 –562 (1996). http://dx.doi.org/10.1088/0031-9155/41/3/015 PHMBA7 0031-9155 Google Scholar

8. 

Y. Yamashitaet al., “Continuous measurement of oxy- and deoxyhemoglobin of piglet brain by time-resolved spectroscopy,” OSA Trends in Optics and Photonics, 22 205 –207 (1998). OTOPFZ 1094-5695 Google Scholar

9. 

M. S. Pattersonet al., “Frequency-domain reflectance for the determination of the scattering and absorption properties of tissue,” Appl. Opt., 30 (31), 4474 –4476 (1991). http://dx.doi.org/10.1364/AO.30.004474 APOPAI 0003-6935 Google Scholar

10. 

J. B. FishkinE. Gratton, “Propagation of photon-density waves in strongly scattering media containing an absorbing semi-infinite plane bounded by a straight edge,” J. Opt. Soc. Am. A, 10 (1), 127 –140 (1993). http://dx.doi.org/10.1364/JOSAA.10.000127 JOAOD6 0740-3232 Google Scholar

11. 

S. FantiniM. A. FranceschiniE. Gratton, “Semi-infinite-geometry boundary problem for light migration in highly scattering media: a frequency-domain study in the diffusion approximation,” J. Opt. Soc. Am. B, 11 (10), 2128 –2138 (1994). http://dx.doi.org/10.1364/JOSAB.11.002128 JOBPDE 0740-3224 Google Scholar

12. 

Y. TsuchiyaT. Urakami, “Frequency domain analysis of photon migration based on the microscopic Beer-Lambert law,” Jpn. J. Appl. Phys., 35 (Part 1, No. 9A), 4848 –4851 (1996). http://dx.doi.org/10.1143/JJAP.35.4848 JJAPA5 0021-4922 Google Scholar

13. 

H. Z. Yeganehet al., “Broadband continuous-wave technique to measure baseline values and changes in the tissue chromophore concentrations,” Biomed. Opt. Express, 3 (11), 2761 –2770 (2012). http://dx.doi.org/10.1364/BOE.3.002761 BOEICL 2156-7085 Google Scholar

14. 

M. FirbankM. SchweigerD. T. Delpy, “Investigation of light piping through clear regions of scattering objects,” Proc. SPIE, 2389 167 –173 (1995). http://dx.doi.org/10.1117/12.209964 PSISDG 0277-786X Google Scholar

15. 

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

16. 

E. OkadaD. T. Delpy, “Effect of scattering of arachnoid trabeculae on light propagation in the adult brain,” in Biomedical Optical Spectroscopy and Diagnostics, (2000). Google Scholar

17. 

L. Linet al., “Influence of a fat on muscle oxygenation measurement using near-IR spectroscopy: quantitative analysis based on two-layered phantom experiments and Monte Carlo simulation,” Front. Med. Biol. Eng., 10 (1), 43 –58 (2000). http://dx.doi.org/10.1163/15685570052061531 0921-3775 Google Scholar

18. 

M. Niwayamaet al., “Quantitative measurement of muscle hemoglobin oxygenation using near-infrared spectroscopy with correction for the influence of a subcutaneous fat layer,” Rev. Sci. Instrum., 71 (12), 4571 –4575 (2000). http://dx.doi.org/10.1063/1.1322578 RSINAK 0034-6748 Google Scholar

19. 

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

20. 

E. OkadaD. T. Delpy, “Near-infrared light propagation in an adult head model. II. Effect of superficial tissue thickness on the sensitivity of the near-infrared spectroscopy signal,” Appl. Opt., 42 (16), 2915 –2922 (2003). http://dx.doi.org/10.1364/AO.42.002915 APOPAI 0003-6935 Google Scholar

21. 

S. MisonooE. Okada, “Adult head modeling based on time-resolved measurement for NIR instrument,” Proc. SPIE, 4250 522 –529 (2001). http://dx.doi.org/10.1117/12.434527 PSISDG 0277-786X Google Scholar

22. 

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

23. 

O. SayliE. B. AkselA. Akin, “Crosstalk and error analysis of fat layer on continuous wave near-infrared spectroscopy measurements,” J. Biomed. Opt., 13 (6), 064019 (2008). http://dx.doi.org/10.1117/1.3028008 JBOPFO 1083-3668 Google Scholar

24. 

T. Takahashiet al., “Influence of skin blood flow on near-infrared spectroscopy signals measured on the forehead during a verbal fluency task,” Neuroimage, 57 (3), 991 –1002 (2011). http://dx.doi.org/10.1016/j.neuroimage.2011.05.012 NEIMEF 1053-8119 Google Scholar

25. 

E. Kelleret al., “Changes in cerebral blood flow and oxygen metabolism during moderate hypothermia in patients with severe middle cerebral artery infarction,” Neurosurg. Focus, 8 (5), e4 (2000). http://dx.doi.org/10.3171/foc.2000.8.5.4 1092-0684 Google Scholar

26. 

J. T. Elliottet al., “Quantitative measurement of cerebral blood flow in a juvenile porcine model by depth-resolved near-infrared spectroscopy,” J. Biomed. Opt., 15 (3), 037014 (2010). http://dx.doi.org/10.1117/1.3449579 JBOPFO 1083-3668 Google Scholar

27. 

R. Aroraet al., “Preservation of the metabolic rate of oxygen in preterm infants during indomethacin therapy for closure of the ductus arteriosus,” Pediatr. Res., 73 (6), 713 –718 (2013). http://dx.doi.org/10.1038/pr.2013.53 PEREBL 0031-3998 Google Scholar

28. 

P. W. McCormicket al., “Intracerebral penetration of infrared light. Technical note,” J. Neurosurg., 76 (2), 315 –318 (1992). http://dx.doi.org/10.3171/jns.1992.76.2.0315 JONSAC 0022-3085 Google Scholar

29. 

D. N. Harriset al., “NIRS in adults—effects of increasing optode separation,” in Oxygen Transport to Tissue XV, 837 –840 (1994). http://dx.doi.org/10.1007/978-1-4615-2468-7 Google Scholar

30. 

T. J. Germonet al., “Optode separation determines sensitivity of near infrared spectroscopy to intra- and extracranial oxygenation changes,” J. Cereb. Blood Flow Metab., 15 617 (1996). JCBMDN 0271-678X Google Scholar

31. 

S. Kohriet al., “Quantitative evaluation of the relative contribution ratio of cerebral tissue to near-infrared signals in the adult human head: a preliminary study,” Physiol. Meas., 23 (2), 301 –312 (2002). http://dx.doi.org/10.1088/0967-3334/23/2/306 PMEAE3 0967-3334 Google Scholar

32. 

E. Ohmaeet al., “Cerebral hemodynamics evaluation by near-infrared time-resolved spectroscopy: correlation with simultaneous positron emission tomography measurements,” Neuroimage, 29 (3), 697 –705 (2006). http://dx.doi.org/10.1016/j.neuroimage.2005.08.008 NEIMEF 1053-8119 Google Scholar

33. 

A. H. Hielscheret al., “Time-resolved photon emission from layered turbid media,” Appl. Opt., 35 (4), 719 –728 (1996). http://dx.doi.org/10.1364/AO.35.000719 APOPAI 0003-6935 Google Scholar

34. 

C. Satoet al., “Intraoperative monitoring of depth-dependent hemoglobin concentration changes during carotid endarterectomy by time-resolved spectroscopy,” Appl. Opt., 46 (14), 2785 –2792 (2007). http://dx.doi.org/10.1364/AO.46.002785 APOPAI 0003-6935 Google Scholar

35. 

Y. Hoshi, “Towards the next generation of near-infrared spectroscopy,” Philos. Trans. A Math. Phys. Eng. Sci., 369 (1955), 4425 –4439 (2011). http://dx.doi.org/10.1098/rsta.2011.0262 1364-503X Google Scholar

36. 

S. HommaT. FukunagaA. Kagaya, “Influence of adipose tissue thickness on near infrared spectroscopic signal in the measurement of human muscle,” J. Biomed. Opt., 1 (4), 418 –424 (1996). http://dx.doi.org/10.1117/12.252417 JBOPFO 1083-3668 Google Scholar

37. 

M. C. van Beekveltet al., “Adipose tissue thickness affects in vivo quantitative near-IR spectroscopy in human skeletal muscle,” Clin. Sci. (Lond), 101 (1), 21 –28 (2001). http://dx.doi.org/10.1042/CS20000247 CSCIAE 0143-5221 Google Scholar

38. 

D. GeraskinH. BoethM. Kohl-Bareis, “Optical measurement of adipose tissue thickness and comparison with ultrasound, magnetic resonance imging, and callipers,” J. Biomed. Opt., 14 (4), 044017 (2009). http://dx.doi.org/10.1117/1.3184425 JBOPFO 1083-3668 Google Scholar

39. 

T. Hamaokaet al., “Near-infrared spectroscopy/imaging for monitoring muscle oxygenation and oxidative metabolism in healthy and diseased humans,” J. Biomed. Opt., 12 (6), 062105 (2007). http://dx.doi.org/10.1117/1.2805437 JBOPFO 1083-3668 Google Scholar

40. 

E. Ohmaeet al., “Clinical evaluation of time-resolved spectroscopy by measuring cerebral hemodynamics during cardiopulmonary bypass surgery,” J. Biomed. Opt., 12 (6), 062112 (2007). http://dx.doi.org/10.1117/1.2804931 JBOPFO 1083-3668 Google Scholar

41. 

K. Yoshitaniet al., “Clinical validity of cerebral oxygen saturation measured by time-resolved spectroscopy during carotid endarterectomy,” J. Neurosurg. Anesthesiol., 25 (3), 248 –253 (2013). http://dx.doi.org/10.1097/ANA.0b013e31827ee0cf JNANEV 0898-4921 Google Scholar

42. 

K. Yamazakiet al., “Cerebral oxygen saturation evaluated by near-infrared time-resolved spectroscopy (TRS) in pregnant women during caesarean section—a promising new method of maternal monitoring,” Clin. Physiol. Funct. Imaging, 33 (2), 109 –116 (2013). http://dx.doi.org/10.1111/cpf.2013.33.issue-2 CPFICA 1475-0961 Google Scholar

43. 

S. Ijichiet al., “Developmental changes of optical properties in neonates determined by near-infrared time-resolved spectroscopy,” Pediatr. Res., 58 (3), 568 –573 (2005). http://dx.doi.org/10.1203/01.PDR.0000175638.98041.0E PEREBL 0031-3998 Google Scholar

44. 

K. Koyanoet al., “The effect of blood transfusion on cerebral hemodynamics in preterm infants,” Transfusion, 53 (7), 1459 –1467 (2013). http://dx.doi.org/10.1111/trf.2013.53.issue-7 TRANAT 0041-1132 Google Scholar

45. 

N. Yokoseet al., “Bedside assessment of cerebral vasospasms after subarachnoid hemorrhage by near infrared time-resolved spectroscopy,” in Oxygen Transport to Tissue XXXI, 505 –511 (2010). http://dx.doi.org/10.1007/978-1-4419-1241-1 Google Scholar

46. 

Y. Uedaet al., “Time-resolved optical mammography and its preliminary clinical results,” Technol. Cancer Res. Treat., 10 (5), 393 –401 (2011). TCRTBS 1533-0346 Google Scholar

47. 

S. Kogaet al., “Methodological validation of the dynamic heterogeneity of muscle deoxygenation within the quadriceps during cycle exercise,” Am. J. Physiol. Regul. Integr. Comp. Physiol., 301 (2), R534 –541 (2011). http://dx.doi.org/10.1152/ajpregu.00101.2011 0363-6119 Google Scholar

48. 

L. M. Chinet al., “The relationship between muscle deoxygenation and activation in different muscles of the quadriceps during cycle ramp exercise,” J. Appl. Physiol., 111 (5), 1259 –1265 (2011). http://dx.doi.org/10.1152/japplphysiol.01216.2010 JAPYAA 0021-8987 Google Scholar

49. 

M. Odaet al., “Near-infrared time-resolved spectroscopy system for tissue oxygenation monitor,” Proc. SPIE, 3597 611 –617 (1999). http://dx.doi.org/10.1117/12.356809 PSISDG 0277-786X Google Scholar

50. 

Y. Hoshiet al., “Resting hypofrontality in schizophrenia: a study using near-infrared time-resolved spectroscopy,” Schizophr. Res., 84 (2–3), 411 –420 (2006). http://dx.doi.org/10.1016/j.schres.2006.03.010 SCRSEH 0920-9964 Google Scholar

51. 

C. Satoet al., “Estimating the absorption coefficient of the bottom layer in four-layered turbid mediums based on the time-domain depth sensitivity of near-infrared light reflectance,” J. Biomed. Opt., 18 (9), 097005 (2013). http://dx.doi.org/10.1117/1.JBO.18.9.097005 JBOPFO 1083-3668 Google Scholar

52. 

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

53. 

H. Zhanget al., “Simple subtraction method for determining the mean path length traveled by photons in turbid media,” Jpn. J. Appl. Phys., 37 (Part 1, No. 2), 700 –704 (1998). http://dx.doi.org/10.1143/JJAP.37.700 JJAPA5 0021-4922 Google Scholar

54. 

K. Suzukiet al., “Quantitative measurement of optical parameters in the breast using time-resolved spectroscopy. Phantom and preliminary in vivo results,” Invest. Radiol., 29 (4), 410 –414 (1994). http://dx.doi.org/10.1097/00004424-199404000-00003 INVRAV 0020-9996 Google Scholar

55. 

A. Liebertet al., “Fiber dispersion in time domain measurements compromising the accuracy of determination of optical properties of strongly scattering media,” J. Biomed. Opt., 8 (3), 512 –516 (2003). http://dx.doi.org/10.1117/1.1578088 JBOPFO 1083-3668 Google Scholar

56. 

G. Yuet al., “Time-dependent blood flow and oxygenation in human skeletal muscles measured with noninvasive near-infrared diffuse optical spectroscopies,” J. Biomed. Opt., 10 (2), 024027 (2005). http://dx.doi.org/10.1117/1.1884603 JBOPFO 1083-3668 Google Scholar

57. 

V. Quaresimaet al., “Spatial distribution of vastus lateralis blood flow and oxyhemoglobin saturation measured at the end of isometric quadriceps contraction by multichannel near-infrared spectroscopy,” J. Biomed. Opt., 9 (2), 413 –420 (2004). http://dx.doi.org/10.1117/1.1646417 JBOPFO 1083-3668 Google Scholar

58. 

C. Fuschet al., “Body water, lean body and fat mass of healthy children as measured by deuterium oxide dilution,” Isotopenpraxis Isotopes Environ. Health Stud., 29 (1–2), 125 –131 (1993). http://dx.doi.org/10.1080/10256019308046145 1025-6016 Google Scholar

59. 

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

60. 

S. W. Coppacket al., “Postprandial substrate deposition in human forearm and adipose tissues in vivo,” Clin. Sci. (Lond), 79 (4), 339 –348 (1990). Google Scholar

61. 

P. ArnerJ. Bulow, “Assessment of adipose tissue metabolism in man: comparison of Fick and microdialysis techniques,” Clin. Sci. (Lond), 85 (3), 247 –256 (1993). Google Scholar

62. 

K. N. FraynS. W. CoppackS. M. Humphreys, “Subcutaneous adipose tissue metabolism studied by local catheterization,” Int. J. Obes. Relat. Metab. Disord., 17 (Suppl 3), S18 –21 (1993). Google Scholar

63. 

L. SimonsenJ. BulowJ. Madsen, “Adipose tissue metabolism in humans determined by vein catheterization and microdialysis techniques,” Am. J. Physiol., 266 (3 Pt 1), E357 –365 (1994). AJPHAP 0002-9513 Google Scholar

64. 

K. MatsushitaS. HommaE. Okada, “Influence of adipose tissue on muscle oxygenation measurement with an NIRS instrument,” Proc. SPIE, 3194 159 –165 (1998). http://dx.doi.org/10.1117/12.301048 PSISDG 0277-786X Google Scholar

65. 

C. R. Simpsonet al., “Near-infrared optical properties of ex vivo human skin and subcutaneous tissues measured using the Monte Carlo inversion technique,” Phys. Med. Biol., 43 (9), 2465 –2478 (1998). http://dx.doi.org/10.1088/0031-9155/43/9/003 PHMBA7 0031-9155 Google Scholar

66. 

D. S. DeLoreyJ. M. KowalchukD. H. Paterson, “Relationship between pulmonary O2 uptake kinetics and muscle deoxygenation during moderate-intensity exercise,” J. Appl. Physiol., 95 (1), 113 –120 (2003). JAPYAA 0021-8987 Google Scholar

67. 

B. Grassiet al., “Muscle oxygenation and pulmonary gas exchange kinetics during cycling exercise on-transitions in humans,” J. Appl. Physiol., 95 (1), 149 –158 (2003). JAPYAA 0021-8987 Google Scholar

68. 

L. F. Ferreiraet al., “Muscle capillary blood flow kinetics estimated from pulmonary O2 uptake and near-infrared spectroscopy,” J. Appl. Physiol., 98 (5), 1820 –1828 (2005). http://dx.doi.org/10.1152/japplphysiol.00907.2004 JAPYAA 0021-8987 Google Scholar

69. 

D. R. Joneset al., “Reply to Quaresima and Ferrari,” J. Appl. Physiol., 107 372 –373 (2009). http://dx.doi.org/10.1152/japplphysiol.00314.2009 JAPYAA 0021-8987 Google Scholar

70. 

M. C. van Beekveltet al., “In vivo quantitative near-infrared spectroscopy in skeletal muscle during incremental isometric handgrip exercise,” Clin. Physiol. Funct. Imaging, 22 (3), 210 –217 (2002). http://dx.doi.org/10.1046/j.1475-097X.2002.00420.x CPFICA 1475-0961 Google Scholar

71. 

F. Harelet al., “Arterial flow measurements during reactive hyperemia using NIRS,” Physiol. Meas., 29 (9), 1033 –1040 (2008). http://dx.doi.org/10.1088/0967-3334/29/9/003 PMEAE3 0967-3334 Google Scholar

72. 

C. Casavolaet al., “Blood flow and oxygen consumption with near-infrared spectroscopy and venous occlusion: spatial maps and the effect of time and pressure of inflation,” J. Biomed. Opt., 5 (3), 269 –276 (2000). http://dx.doi.org/10.1117/1.429995 JBOPFO 1083-3668 Google Scholar

73. 

M. Annuket al., “Impaired endothelium-dependent vasodilatation in renal failure in humans,” Nephrol. Dial. Transplant., 16 (2), 302 –306 (2001). http://dx.doi.org/10.1093/ndt/16.2.302 NDTREA 0931-0509 Google Scholar

74. 

J. R. de Berrazuetaet al., “Endothelial dysfunction, measured by reactive hyperaemia using strain-gauge plethysmography, is an independent predictor of adverse outcome in heart failure,” Eur. J. Heart Fail., 12 (5), 477 –483 (2010). http://dx.doi.org/10.1093/eurjhf/hfq036 EJHFFS 1388-9842 Google Scholar

75. 

J. R. Wilsonet al., “Noninvasive detection of skeletal muscle underperfusion with near-infrared spectroscopy in patients with heart failure,” Circulation, 80 (6), 1668 –1674 (1989). http://dx.doi.org/10.1161/01.CIR.80.6.1668 CIRCAZ 0009-7322 Google Scholar

76. 

P. A. Moleet al., “Myoglobin desaturation with exercise intensity in human gastrocnemius muscle,” Am. J. Physiol., 277 (1 Pt 2), R173 –180 (1999). AJPHAP 0002-9513 Google Scholar

77. 

R. S. Richardsonet al., “Effect of mild carboxy-hemoglobin on exercising skeletal muscle: intravascular and intracellular evidence,” Am. J. Physiol. Regul. Integr. Comp. Physiol., 283 (5), R1131 –1139 (2002). 0363-6119 Google Scholar

78. 

M. Mizunoet al., “Regional differences in blood flow and oxygen consumption in resting muscle and their relationship during recovery from exhaustive exercise,” J. Appl. Physiol., 95 (6), 2204 –2210 (2003). JAPYAA 0021-8987 Google Scholar

79. 

A. Kienleet al., “Noninvasive determination of the optical properties of two-layered turbid media,” Appl. Opt., 37 (4), 779 –791 (1998). http://dx.doi.org/10.1364/AO.37.000779 APOPAI 0003-6935 Google Scholar

80. 

J. Ripollet al., “Recovery of optical parameters in multiple-layered diffusive media: theory and experiments,” J. Opt. Soc. Am. A Opt. Image Sci. Vis., 18 (4), 821 –830 (2001). http://dx.doi.org/10.1364/JOSAA.18.000821 JOAOD6 0740-3232 Google Scholar

81. 

A. Sassaroliet al., “Performance of fitting procedures in curved geometry for retrieval of the optical properties of tissue from time-resolved measurements,” Appl. Opt., 40 (1), 185 –197 (2001). http://dx.doi.org/10.1364/AO.40.000185 APOPAI 0003-6935 Google Scholar

82. 

F. MartelliS. Del BiancoG. Zaccanti, “Procedure for retrieving the optical properties of a two-layered medium from time-resolved reflectance measurements,” Opt. Lett., 28 (14), 1236 –1238 (2003). http://dx.doi.org/10.1364/OL.28.001236 OPLEDP 0146-9592 Google Scholar

83. 

M. ShimadaY. HoshiY. Yamada, “Simple algorithm for the measurement of absorption coefficients of a two-layered medium by spatially resolved and time-resolved reflectance,” Appl. Opt., 44 (35), 7554 –7563 (2005). http://dx.doi.org/10.1364/AO.44.007554 APOPAI 0003-6935 Google Scholar

84. 

L. Gagnonet al., “Double-layer estimation of intra- and extracerebral hemoglobin concentration with a time-resolved system,” J. Biomed. Opt., 13 (5), 054019 (2008). http://dx.doi.org/10.1117/1.2982524 JBOPFO 1083-3668 Google Scholar

85. 

S. J. MatcherM. CopeD. T. Delpy, “In vivo measurements of the wavelength dependence of tissue-scattering coefficients between 760 and 900 nm measured with time-resolved spectroscopy,” Appl. Opt., 36 (1), 386 –396 (1997). http://dx.doi.org/10.1364/AO.36.000386 APOPAI 0003-6935 Google Scholar

86. 

M. C. Van Beekveltet al., “Performance of near-infrared spectroscopy in measuring local O(2) consumption and blood flow in skeletal muscle,” J. Appl. Physiol., 90 (2), 511 –519 (2001). JAPYAA 0021-8987 Google Scholar

87. 

T. Hamaokaet al., “Noninvasive measures of oxidative metabolism on working human muscles by near-infrared spectroscopy,” J. Appl. Physiol., 81 (3), 1410 –1417 (1996). JAPYAA 0021-8987 Google Scholar

88. 

T. Sakoet al., “Validity of NIR spectroscopy for quantitatively measuring muscle oxidative metabolic rate in exercise,” J. Appl. Physiol., 90 (1), 338 –344 (2001). JAPYAA 0021-8987 Google Scholar

89. 

R. C. Harriset al., “The effect of circulatory occlusion on isometric exercise capacity and energy metabolism of the quadriceps muscle in man,” Scand. J. Clin. Lab Invest., 35 (1), 87 –95 (1975). http://dx.doi.org/10.3109/00365517509068010 SJCLAY 0036-5513 Google Scholar

90. 

M. A. Johnsonet al., “Data on the distribution of fibre types in thirty-six human muscles. An autopsy study,” J. Neurol. Sci., 18 (1), 111 –129 (1973). http://dx.doi.org/10.1016/0022-510X(73)90023-3 JNSCAG 0022-510X Google Scholar

91. 

A. D. Edwardset al., “Measurement of hemoglobin flow and blood flow by near-infrared spectroscopy,” J. Appl. Physiol., 75 (4), 1884 –1889 (1993). JAPYAA 0021-8987 Google Scholar

92. 

R. A. De Blasiet al., “Noninvasive measurement of forearm blood flow and oxygen consumption by near-infrared spectroscopy,” J. Appl. Physiol., 76 (3), 1388 –1393 (1994). JAPYAA 0021-8987 Google Scholar

93. 

S. Hommaet al., “Near-infrared estimation of O2 supply and consumption in forearm muscles working at varying intensity,” J. Appl. Physiol., 80 (4), 1279 –1284 (1996). JAPYAA 0021-8987 Google Scholar

94. 

H. Watanabeet al., “Assessment of forearm reactive hyperemia using near-infrared time resolved spectroscopy: a new tool to evaluate endothelial function,” in Proc. of Asian Symp. on BOPM, 164 –165 (2002). Google Scholar

95. 

P. VaupelF. KallinowskiP. Okunieff, “Blood flow, oxygen consumption and tissue oxygenation of human tumors,” Adv. Exp. Med. Biol., 277 895 –905 (1990). http://dx.doi.org/10.1007/978-1-4684-8181-5 AEMBAP 0065-2598 Google Scholar

Biographies of the authors are not available.

© 2014 Society of Photo-Optical Instrumentation Engineers (SPIE) 0091-3286/2014/$25.00 © 2014 SPIE
Etsuko Ohmae, Shinichiro Nishio, Motoki Oda, Hiroaki Suzuki, Toshihiko Suzuki, Kyoichi Ohashi, Shunsaku Koga, Yutaka Yamashita, and Hiroshi Watanabe "Sensitivity correction for the influence of the fat layer on muscle oxygenation and estimation of fat thickness by time-resolved spectroscopy," Journal of Biomedical Optics 19(6), 067005 (9 June 2014). https://doi.org/10.1117/1.JBO.19.6.067005
Published: 9 June 2014
Lens.org Logo
CITATIONS
Cited by 23 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Fourier transforms

Near infrared spectroscopy

Absorption

Tissue optics

Skin

Tissues

Ultrasonography

Back to Top