Open Access
14 December 2023 Single-distance and dual-slope frequency-domain near-infrared spectroscopy to assess skeletal muscle hemodynamics
Author Affiliations +
Abstract

Significance

Non-invasive optical measurements of deep tissue (e.g., muscle) need to take into account confounding contributions from baseline and dynamic optical properties of superficial tissue (adipose tissue).

Aim

Discriminate superficial and deep tissue hemodynamics using data collected with frequency-domain (FD) near-infrared spectroscopy (NIRS) in a dual-slope (DS) configuration.

Approach

Experimental data were collected in vivo on the forearm of three human subjects during a 3-min arterial occlusion or 1-min venous occlusion. Theoretical data were generated using diffusion theory for two-layered media with varying values of the reduced scattering coefficient (μs) (range: 0.5 to 1.1 mm − 1) and absorption coefficient (μa) (range: 0.005 − 0.015 mm − 1) of the two layers, and top layer thickness (range: 2 to 8 mm). Data were analyzed using diffusion theory for a homogeneous semi-infinite medium.

Results

Experimental data in vivo were consistent with simulated data for a two-layered medium with a larger μs in the top layer, comparable absorption changes in the top and bottom layers during venous occlusion, and smaller absorption changes in the top vs. bottom layers during arterial occlusion.

Conclusions

The dataset generated by DS FD-NIRS may allow for discrimination of superficial and deep absorption changes in two-layered media, thus lending itself to individual measurements of hemodynamics in adipose and muscle tissue.

1.

Introduction

Near-infrared spectroscopy (NIRS) is a non-invasive optical technique that has been applied to measurements of the perfusion and oxygenation of various biological tissues, including skeletal muscle1,2 and brain.3 NIRS studies have employed continuous-wave (CW),47 time-domain (TD),8 and frequency-domain (FD)9,10 systems, which are based on constant, pulsed, and intensity-modulated illumination, respectively. To achieve high accuracy and reliability, non-invasive NIRS measurements of skeletal muscle and brain must take into account potentially confounding contributions to the optical signals from superficial tissue, namely skin and adipose tissue in the case of muscle measurements2 or scalp and skull tissue in the case of brain measurements.11

Focusing on muscle studies, it was observed that NIRS data need to be corrected for the adipose tissue thickness (ATT) to yield measurements that are representative of muscle tissue.6,12,13 To this aim, TD-NIRS data were used either to generate correction curves derived from the ratio of partial optical pathlength in muscle to the total optical pathlength,8 or to yield separate measurements for the top layer (adipose tissue) and bottom layer (muscle) by modeling the tissue as a two-layered medium.14 CW-NIRS for skeletal muscle measurements has been commonly implemented by placing a single illumination point (source) and a single optical collection point (detector) on the tissue surface. This arrangement is commonly referred to as single-distance (SD) since it relies on a single source–detector separation. In this approach, the measured quantity is the optical intensity (I) as well as its changes associated with muscle hemodynamics and deoxygenation (ΔI).6,12 These SD intensity (SDI) measurements are known to have a greater sensitivity to superficial tissue compared to single distance phase (SDϕ) measurements obtained with FD-NIRS.15,16 In the case of FD-NIRS, I and ϕ represent the amplitude and the phase, respectively, of the photon-density waves that result from modulated illumination.

One alternative method that has been applied to the brain,1517 and skeletal muscle18 is the dual-slope (DS) approach. The DS approach aims to measure the changes in the slope of either the linearized intensity (ln(ρ2I)) or phase (ϕ) as a function of source–detector distance (ρ), without any need of calibration for source emission or detector sensitivity factors. This is done by using a special arrangement of two sources and two detectors on the tissue surface, and by averaging the two paired single-slopes that are measured using each single source in combination with the two detectors.19 We have shown that in comparison with SD methods, the DS approach offers preferential sensitivity to deeper localized absorption changes in otherwise homogeneous media,19 and that hemodynamics measured on the human head are more closely representative of cerebral tissue than superficial scalp/skull tissue.16,20 In this work, we explore the ability of DS FD-NIRS measurements in limiting or assessing the contribution of adipose tissue to the NIRS measurements of skeletal muscle.

Two commonly used NIRS protocols to measure blood flow and oxygen consumption in skeletal muscle involve the use of a venous occlusion and an arterial occlusion, respectively. These protocols cause hemodynamic and oxygenation changes in limb tissues (skeletal muscle, skin, adipose tissue, etc.) that are distal to the pneumatic cuff used to perform the vascular occlusion. A venous occlusion is achieved by inflating a pneumatic cuff that is wrapped around the limb to a pressure of 40 to 60 mmHg, which is above venous pressure, effectively causing the accumulation of blood in distal tissues. Such protocols have been used to measure muscle blood flow from the initial rate of accumulation of blood following the onset of the occlusion.2,7,21 Previous work has shown that blood flow in adipose tissue is greater than in skeletal muscle at rest,22 as confirmed by NIRS measurements on a population of 78 subjects with ATT ranging from 1.5 to 9 mm.6 An arterial occlusion is achieved similarly but with a greater cuff inflation pressure of about 200  mmHg or more, which is above the systolic pressure, effectively blocking both the inflow and outflow of blood to the measured tissue.23 During an arterial occlusion, the oxygen consumption is measured through the rate of deoxygenation of heme compounds [hemoglobin (Hb) and myoglobin (Mb)] in muscle. Because the near-infrared absorption spectra of hemoglobin and myoglobin have extremely similar shapes, being both dominated by their heme group(s), NIRS methods can only measure the combined concentrations of hemoglobin and myoglobin in their oxygenated state (oxy[Hb + Mb]) or deoxygenated state (deoxy[Hb + Mb]). In this article, we use the compact notation of O and D for oxy[Hb+Mb] and deoxy[Hb+Mb], respectively, and T for the total heme concentration (T=O+D). In the absence of blood redistribution within the muscle, the oxygen desaturation of heme groups results in equal rates of decrease of O and increase of D. Adipose tissue has been shown to have a lower oxygen consumption than skeletal muscle at rest.6,22,24 The different hemodynamic and deoxygenation behavior of adipose tissue and muscle tissue during a venous or arterial occlusion render these protocols useful tools to investigate the relative sensitivity of NIRS measurements to adipose and muscle tissues.

Here we use the DS method with both I and ϕ measurements on the human forearm during either an arterial or venous occlusion. For each occlusion protocol, we report results on three subjects. We note that DS measurements comprise two sets of SD data collected at two different source–detector separations (25 and 35 mm in this work). The main objective of this work is to investigate the information content of different data types (namely SDI, SDϕ, DSI, and DSϕ) in relation to the in-homogeneous hemodynamics in superficial tissue (skin and adipose layer) and deeper tissue (skeletal muscle) in response to vascular occlusions. To guide our interpretation of the experiments in vivo, we also performed simulations in two-layered media to simulate the effects of different baseline tissue properties, different absorption changes in the two layers, and different top layer thicknesses on the measured optical data generated by FD-NIRS in both SD and DS configurations.

2.

Methods

2.1.

Experimental Protocol and Instrumentation

A total of 4 subjects [3 female (F), 1 male (M); age range 22 to 29 years] participated in a study approved by the Tufts University Institutional Review Board (IRB) for FD-NIRS measurements conducted on the human forearm. Two experimental protocols, one involving a venous occlusion and one involving an arterial occlusion, are schematically illustrated in Fig. 1(a). The venous occlusion protocol consisted of a 1 min baseline, followed by a 1 min venous occlusion achieved by inflating a pneumatic cuff wrapped around the upper arm to a pressure of 60 mmHg and a final 1 min recovery. The arterial occlusion protocol consisted of a 1 min baseline followed by a 3 min arterial occlusion achieved by inflating a pneumatic cuff wrapped around the upper arm to a pressure of 220 mmHg and a final 3 min recovery. Subjects 2 and 4 participated in both protocols, whereas subject 1 elected to only participate in the arterial occlusion protocol and subject 3 only in the venous occlusion protocol, as listed in Table 1. Therefore, this work reports results on three subjects for each occlusion period.

Fig. 1

Experimental setup for frequency-domain (FD) near-infrared spectroscopy (NIRS) measurements on the human forearm during an arterial occlusion or venous occlusion. (a) The timing of the occlusions achieved by inflation of a pneumatic arm cuff using an automatic cuff inflation system [E-20 Rapid Cuff Inflation System, D.E. Hokanson, Inc. (Bellevue, Washington, United States)]. (b) A pneumatic arm cuff [SC12D, D.E. Hokanson, Inc. (Bellevue, Washington, United States)] was placed on the subject’s upper arm. The cuff pressure was measured continuously using a Series 626 Pressure Transmitter Dwyer Instruments Inc. (Michigan City, Indiana, United States) and recorded synchronously with the data collected by the FD-NIRS system [ISS Imagent V2 (Champaign, Illinois, United States)]. An optical probe connected to the FD-NIRS system was placed on the forearm above the brachioradialis muscle. (c) Schematic of the FD-NIRS optical probe with numbers used to label optical sources and letters to label detectors. This arrangement allowed for two single-distance (SD) measurements at a source–detector separation of 25 mm (1A and 2B), two SD measurements at a source–detector separation of 35 mm (1B and 2A), and one dual-slope measurement based on data collected by all combinations of sources and detectors (1AB2).

JBO_28_12_125004_f001.png

Table 1

Summary of the baseline measurements for each subject (sub) before the arterial occlusion and venous occlusion protocols (prot). Adipose tissue thickness (ATT), mean ± standard deviation of the absorption coefficient (μa) and reduced scattering coefficient (μs′) during the baseline period at the two wavelengths of 690 and 830 nm, and generalized mean optical pathlength for intensity (I) (⟨L⟩I) and phase (ϕ) (⟨L⟩ϕ). ⟨L⟩I and ⟨L⟩ϕ are reported for both wavelengths and source–detector distances (ρ): ρ1=25  mm and ρ2=35  mm.

Protμa (mm−1)μs′ (mm−1)⟨L⟩I (mm)⟨L⟩ϕ (mm)
SubATT (mm)690 nm830 nm690 nm830 nm690 nm830 nm690 nm830 nm690 nm830 nm690 nm830 nm
ρ1ρ2ρ1ρ2ρ1ρ2ρ1ρ2
Arterial occlusion12.0 ± 0.30.0172 ± 0.00070.0160 ± 0.00040.58 ± 0.010.56 ± 0.0110615310715410.315.611.116.8
23.4 ± 0.20.0161 ± 0.00020.0190 ± 0.00010.551 ± 0.0030.477 ± 0.00310615292.713310.816.48.1512.2
46.0 ± 0.50.0198 ± 0.00010.01756 ± 0.000080.581 ± 0.0030.596 ± 0.0031001441061538.6413.010.215.4
Venous occlusion23.4 ± 0.20.0150 ± 0.00010.0177 ± 0.00010.560 ± 0.0030.5140.00310915898.614212.018.29.2613.9
34.4 ± 0.20.0156 ± 0.00030.0185 ± 0.00030.547 ± 0.0060.494 ± ±0.00610715495.113611.317.18.5712.9
46.0 ± 0.50.0185 ± 0.00040.01839 ± 0.000090.571 ± 0.0020.598 ± 0.0031021471051519.3014.09.6114.5

All subjects were placed in a seated position with the right forearm on a flat surface. Before starting the experiment, ATT was measured with a skinfold caliper at the measurement location and the ATT values are reported in Table 1 for each subject. First, a pneumatic arm cuff [SC12D, D.E. Hokanson, Inc. (Bellevue, Washington, United States)] was placed on the upper right arm [Fig. 1(b)]. The cuff pressure was controlled using an automatic cuff inflation system [E-20 Rapid Cuff Inflation System, D.E. Hokanson, Inc. (Bellevue, Washington, United States)] to inflate the cuff to the desired pressure. The cuff pressure was measured continuously using a digital manometer [Series 626 Pressure Transmitter Dwyer Instruments Inc. (Michigan City, Indiana, United States)], and the pressure data were recorded synchronously with the FD-NIRS data.

An optical probe for FD-NIRS measurements was placed on the forearm over the brachioradialis muscle and secured via black athletic bandaging to guarantee good contact with the skin. The optical probe contained two pairs of illumination optical fibers and two collection optical fiber bundles that were connected to the light sources and optical detectors, respectively, of a commercial ISS Imagent V2 (Champaign, Illinois, United States) (modulation frequency: 140.625 MHz). Each pair of illumination fibers was connected to two laser diodes that emitted light at 690 and 830 nm. The light sources were time-multiplexed, so only one source was on at any given time. The sampling rate for data collection ranged from 2 to 10 Hz, with lower sampling rates used to increase integration time for each source–detector pair to decrease phase noise. The two optical collection points in the optical probe are labeled A and B, whereas the two source locations (each featuring illumination at two wavelengths) are labeled 1 and 2 [Fig. 1(c)]. This arrangement allows for two SD measurements at a source–detector separation of 25 mm (1A and 2B), two SD measurements at a source–detector separation of 35 mm (1B and 2A), and one DS measurement (1AB2) that combines data collected by all source–detector combinations.

2.2.

Theoretical Simulations

An analytical approach based on diffusion theory for a two-layer medium25,26 with an infinitely thick bottom layer was used to guide the interpretation of the experimental results from both occlusion protocols. The choice of an infinitely thick bottom layer is based on the fact that, for typical optical properties of tissue, all data types considered in this work have a sensitivity of at least 90% to the top 20 mm of tissue.27 Between the adipose tissue layer and the radius bone, one finds the brachioradialis muscle and the extensor carpi radialis, which have a combined thickness of about 18 mm in adults younger than 50 years old.28 Considering a superficial adipose tissue thickness >2  mm, the bone depth is at about 20 mm or more (depending on individual anatomy). We verified that for typical optical properties of tissue, all data types considered in this work have a sensitivity of at least 90% to the top 20 mm of tissue thus justifying our choice to perform simulations for two-layered media in which the bottom layer extends indefinitely. For each simulation, we defined specific values of the absorption coefficient (μa) in the top and bottom layers (μa,top and μa,bot, respectively), the reduced scattering coefficient (μs) in the top and bottom layers (μs,top and μs,bot, respectively), and the top layer thickness (Ltop). This was considered as a simulation of the baseline conditions. We then computed the complex reflectance (R˜), whose amplitude (I) and phase (ϕ) represent the FD-NIRS signal, for each source–detector pair of the probe geometry in Fig. 1(c). This complex reflectance was used to obtain the effective optical properties (i.e., the optical properties of an effective semi-infinite homogeneous medium) at baseline. Absorption perturbations in the top and bottom layers (Δμa,top and Δμa,bot, respectively) were then introduced, and the associated change in R˜ (ΔR˜) was calculated. Finally, the effective absorption coefficient change (Δμa) associated with ΔR˜ were obtained as described in Sec. 2.3. Processing of the experimental and simulated data was done in the same way to allow for a fair comparison.

To study the effect of different parameters of the two-layered medium (Ltop and optical properties of the two layers) on the recovered (effective) Δμa, the top and bottom layer μa, μs, and Ltop were all varied. In all simulations, three cases of Δμa,top and Δμa,bot were considered: (1) Δμa,top<Δμa,bot, (2) Δμa,top=Δμa,bot, and (3) Δμa,top>Δμa,bot. These cases were chosen to represent the different hemodynamic responses in the two layers during the two occlusions considered: lower oxygen consumption in superficial adipose tissue versus muscle tissue at rest (i.e., |Δμa,top|<|Δμa,bot| during an arterial occlusion) and greater blood flow in superficial adipose tissue vs. muscle tissue at rest (i.e., Δμa,top>Δμa,bot during a venous occlusion). The optical properties of the two layers at baseline were chosen so that the effective optical properties (defined in Sec. 2.3) were within the range of those measured experimentally (Table 1).

2.3.

Determining Effective Optical Properties and Effective Absorption Changes

Methods described here were applied to both experimental and simulated reflectance (R˜) data. Effective optical properties, μa and μs (for each wavelength in the case of experimental data), were calculated using a multi-distance iterative recovery method29 based on a semi-infinite homogeneous medium with extrapolated boundary conditions. The optical probe was used in the self-calibrating (SC) configuration,30 which allows one to collect DS intensity (DSI) and DS phase (DSϕ) to yield effective μa and μs. For the experimental data described in Sec. 2.1, effective μa and μs at baseline were taken as the average over time from the start of the experiment to 1 s before the start of cuff inflation. Table 1 reports the effective μa and μs for each subject, experimental protocol, and wavelength.

Changes in I or ϕ recorded over a time interval Δt at a single ρ were translated into associated effective absorption changes (Δμa) as follows:

Eq. (1)

Δμa,M(ρ,Δt)=ΔM(ρ,Δt)LM(ρ),
where M is either ln(ρ2I) or ϕ for SD data and LM is the generalized mean optical pathlength associated with data type M (LM indeed has units of length if M is dimensionless, as it is the case here for both ln(ρ2I) and ϕ (the latter being expressed in radians) [we note that a dimensionless constant factor in the argument of the logarithm for M=ln(ρ2I) cancels out in ΔM]). The generalized mean optical pathlength is essentially the opposite of the partial derivative of the data type M [ln(ρ2I) or ϕ] with respect to μa and was calculated using the effective μa and μs for each optical wavelength (λ) and ρ as reported in Table 1. Note that LM/ρ is the conventional differential pathlength factor (DPF) in the case of I data. We averaged the Δμa calculated using SD data from the two different source–detector pairs that feature the same source–detector separation leading to a single value for LI(ρ1), LI(ρ2), Lϕ(ρ1), and Lϕ(ρ2), where ρ1=25  mm and ρ2=35  mm. From these values of generalized pathlengths, we computed a differential slope factor for data type M (DSFM) as follows:

Eq. (2)

DSFM(ρ1,ρ2)=LM(ρ2)LM(ρ1)ρ2ρ1.

Changes in DS data for data type M (ΔDSM) (i.e., changes in the slope of either ln(ρ2I) or ϕ versus ρ) over time Δt were then converted into associated absorption changes as follows:

Eq. (3)

Δμa,DSM(ρ1,ρ2,Δt)=ΔDSM(ρ1,ρ2,Δt)DSFM(ρ1,ρ2).

Considering that LM(ρ) and DSFM(ρ1,ρ2) are always positive, the minus sign in Eqs. (1) and (3), reflects the fact that an increase in SDM or DSM (i.e., positive values of ΔM(ρ,Δt) or ΔDSM) always corresponds to a decrease in absorption. Finally, we note that both LM and DSFM were calculated using analytical formulas obtained in the semi-infinite homogeneous medium with extrapolated boundary conditions. For this reason, even these parameters can be considered as “effective” since they differ from the parameters obtained in a two-layered medium. However, for conciseness we will not refer to them as “effective.”

3.

Results

3.1.

Measurements In Vivo

3.1.1.

Effective optical properties of tissue

Effective optical properties of the forearm tissue were measured at baseline (described in Sec. 2.3) before performing either vascular occlusion. These effective optical properties were used to calculate the generalized optical pathlength for intensity (LI) and generalized optical pathlength for phase (Lϕ) at each wavelength (690 and 830 nm) and at each ρ (ρ1=25  mm and ρ2=35  mm) by considering a semi-infinite homogeneous medium (see Table 1). LI and Lϕ then yield a DSFI and DSFϕ according to Eq. (2). A comparison of the baseline effective optical properties measured on subjects 2 and 4 before the arterial and venous occlusions (subjects 2 and 4 participated in both protocols) shows differences of less than 6%. These differences can be attributed to different hemodynamic conditions and measurement accuracy errors as they exceed the standard deviations over the baseline period reported in Table 1, which represent the measurement precision.

3.1.2.

Absorption changes associated with vascular occlusions

Figure 2 shows representative time traces of the experimental Δμa (Δμa,Exp) obtained with SD and DS I and ϕ data during an arterial occlusion in subject 4 [Fig. 2(a): 690 nm; Fig. 2(b): 830 nm] and during a venous occlusion in subject 2 [Fig. 2(c): 690 nm; Fig. 2(d): 830 nm]. We remind that Δμa,Exp is the effective absorption change, calculated with the methods described in Sec. 2.3. All temporal traces of Δμa,Exp have been low-pass filtered to 0.2 Hz to help visualization by removing fast fluctuations and oscillations at the heart rate. Note that greater noise is typically associated with phase versus intensity data. The measured Δμa,Exp at the two wavelengths reflect the changes in tissue concentrations of the oxygenated (ΔO) and deoxygenated (ΔD) heme groups during the occlusions.

Fig. 2

Representative time traces of experimentally measured absorption coefficient change (Δμa,Exp) during an arterial occlusion [panels (a) and (b): subject 4] and venous occlusion [panels (c) and (d): subject 2]. For each occlusion type, data are shown at wavelengths of 690 nm [panels (a) and (c)] and 830 nm [panels (b) and (d)]. Each subplot contains time traces of Δμa,Exp calculated using data from single-distance intensity (SDI) at 25 mm (dashed red) and 35 mm (dashed green), single-distance phase (SDϕ) at 25 mm (solid red) and 35 mm (solid green), dual-slope intensity (DSI) (dashed blue), and dual-slope phase (DSϕ) (solid blue). Co-registered time traces of the measured arm cuff pressure are shown in pink and scaled according to the right axes. All measurements have been low-pass filtered to 0.2 Hz.

JBO_28_12_125004_f002.png

During an arterial occlusion, the blood flow in and out of the measured limb is blocked, causing an increase in D (ΔD>0) and a decrease of the same magnitude in O (ΔO<0) due to tissue oxygen consumption. At 690 nm, D has a greater extinction coefficient (ϵ) than O and the opposite is true at 830 nm, which translated into an increase in μa at 690 nm and a decrease in μa at 830 nm, as seen in Figs. 2(a) and 2(b). At the onset of the arterial occlusion, a motion artifact that lasts approximately 1 to 2 s is noticeable at both wavelengths for all data types. This artifact is also visible in DS data types, even though they should be largely insensitive to changes in the optical coupling between sources and detectors.16,31 A possible reason for this artifact to appear in DS data is that the optical coupling change may have occurred faster than the data sampling cycle through all sources and detectors. No motion artifact is visible at the onset of the venous occlusion [Fig. 2(c) and 2(d)], possibly as a result of the lower cuff inflation pressure in this case. During a venous occlusion, only the outflow of blood is blocked (at least initially) from the measured limb so that there is blood pooling and an increase in the total-heme concentration (T) in the tissue. Consequently, μa increases at both wavelengths, as shown in Figs. 2(c) and 2(d).

Differences in the maximum absorption changes obtained with the six data types (namely SDI and SDϕ at 25 mm, SDI and SDϕ at 35 mm, and DSI and DSϕ) during and after the two vascular occlusion protocols can be seen in Fig. 2. These differences indicate a different spatial sensitivity of the various data types to non-homogeneous hemodynamics in the probed tissue. While Fig. 2 reports data collected on two representative subjects, Fig. 3 summarizes the results for all subjects, protocols, and wavelengths. Figure 3 reports the maximum Δμa,Exp resulting from the vascular occlusion, as measured by the difference between the average Δμa,Exp over the last 10 s of the occlusion and the average Δμa at 1 to 2 s after the onset of occlusion (to account for any absorption shifts from baseline due to an artifact occurring at the start of cuff inflation). Error bars in Fig. 3 represent the propagation of standard deviations of measurements over time (1 to 2 s after the onset of occlusion and last 10 s of occlusion).

Fig. 3

Bar plots depicting the maximum measured absorption coefficient change (Δμa,Exp) across each subject for an arterial occlusion (a) and (b) and venous occlusion (c) and (d) protocols. Across all panels, measurements using phase (ϕ) have a hatched bar, measurements with intensity (I) are a solid color, measurements with single-distance (SD) at 25 mm are colored in red, measurements with SD at 35 mm are in green, and measurements with dual-slope (DS) are in blue.

JBO_28_12_125004_f003.png

While the experimental data have been analyzed under the simplifying assumption that the investigated tissue is spatially homogeneous (with baseline effective optical properties reported in Table 1 and effective absorption changes), the layered anatomical structure of the skin, adipose tissue, and skeletal muscle accounts for in-homogeneous baseline optical properties as well as in-homogeneous hemodynamic responses to vascular occlusions. One may think that, regardless of the baseline optical properties, collecting SD optical data at longer source–detector distances would always increase the sensitivity to hemodynamic changes in deeper tissue layers. In the case of homogeneous semi-infinite media with layered absorption changes, this is usually the case, and we have also shown that the sensitivity to absorption changes in the bottom layer is greater for DSϕ than DSI data.16

For all three subjects in the arterial occlusion protocol [Figs. 3(a) and 3(b)] and for subject 2 in the venous occlusion protocol [Figs. 3(c) and 3(d)], the maximum absorption changes measured with I data seem to indicate a greater absorption change (in absolute value) deeper in the tissue. As described in the introduction, this behavior is consistent with a higher increase of oxygen consumption in muscles than in adipose tissue. In fact, the measured absorption change progressively increases in absolute value when obtained from SDI at 25 mm, SDI at 35 mm, and DSI. However, the opposite is true for ϕ data, in which case the measured absorption change progressively decreases in absolute value when obtained from SDϕ at 25 mm, SDϕ at 35 mm, and DSϕ. This paradoxical result is assigned to non-homogeneous baseline optical properties in combination with in-homogeneous hemodynamic changes in the investigated tissue. We also note the positive change (increase) in the absorption coefficient at 830 nm in subject 1 measured with DSϕ during an arterial occlusion [Fig. 3(b)], which is assigned to a spatially heterogeneous increase in heme concentration (resulting from blood redistribution) to which DSϕ data is most sensitive.

No paradoxical behavior is observed in subjects 3 and 4 in the venous occlusion protocol [Figs. 3(c) and 3(d)]. The measured values of Δμa,Exp with the different data types are consistent with a smaller increase in blood volume in deeper tissue (muscle) versus superficial tissue (adipose tissue). This is consistent with reported higher blood flow in adipose tissue than in muscle. In fact, the measured absorption change progressively decreases when it is obtained from SDI at 25 mm, SDI at 35 mm, DSI, SDϕ at 25 mm, SDϕ at 35 mm, and DSϕ. To investigate whether in-homogeneous optical properties at baseline and in-homogeneous hemodynamics during vascular occlusion in the superficial adipose tissue and in the deeper skeletal muscle tissue can account for the observed results, we have modeled the tissue as a two-layer medium as presented in Sec. 3.2.

3.1.3.

Oxygen consumption and blood flow obtained from initial absorption change rates during an arterial or venous occlusion, respectively

Absorption changes at 690 and 830 nm were translated into tissue ΔO and ΔD using the molar extinction coefficient of oxy- and deoxyhemoglobin (each containing four heme groups). Tissue oxygen consumption was obtained from the initial temporal slope of (ΔDΔO)/2 after the start of arterial occlusion. For all subjects, the calculated oxygen consumption was in the range of 0.95 to 2.2  μmol/(100  mL)/min. Tissue blood flow as obtained from the initial temporal slope of ΔT after the start of the venous occlusion. The range for calculated blood flow was 0.48 to 4.7  mL/(100  mL)/min.

3.2.

Two-layer Simulation

The simulations based on diffusion theory for a two-layered medium intend to investigate the effect of baseline optical properties and absorption changes in the two layers on the effective absorption changes obtained with different data types.

3.2.1.

Effect of baseline scattering in the two layers

Simulated Δμa (Δμa,Sim) from data generated in a two-layered medium for different values of the top and bottom layer μs (μs,top and μs,bot, respectively) are shown in Fig. 4. Specifically, we considered cases of greater μs in the top layer, the same μs in the two layers, and smaller μs in the top layer, according to the x-axis of Fig. 4. A homogeneous μa was considered so that the top (μa,top) and bottom (μa,bot) absorption coefficients were both set to 0.015  mm1, and the top layer thickness (Ltop) was set to 5 mm. Furthermore, we examined three cases: Δμa smaller in the top layer than in the bottom layer [Fig. 4(a)], the same Δμa in the two layers [Fig. 4(b)], and Δμa greater in the top layer than in the bottom layer [Fig. 4(c)].

Fig. 4

Simulated absorption coefficient change (Δμa,Sim) obtained with different data types collected on two-layered medium with a different baseline reduced scattering coefficient in the top layer (μs,top) and bottom layer (μs,bot). For all simulations, the top layer thickness (Ltop) was 5 mm, and the absorption coefficient for the top layer (μa,top) and bottom layer (μa,bot) were both set to 0.015  mm1. In all panels, measurements using phase (ϕ) data are represented by a hatched bar, measurements using intensity (I) data by a solid bar, measurements using single-distance (SD) at 25 mm are colored in red, measurements using SD at 35 mm are in green, and measurements using dual-slope (DS) are in blue. The absorption coefficient change for the top (Δμa,top) and bottom (Δμa,bot) layers were simulated for the cases of (a) Δμa,top<Δμa,bot(Δμa,top=0.0008  mm1,Δμa,bot=0.008  mm1); (b) Δμa,top=Δμa,bot(Δμa,top=Δμa,bot=0.008  mm1); and (c) Δμa,top>Δμa,bot(Δμa,top=0.008  mm1,Δμa,bot=0.0008  mm1). These values of baseline optical properties and absorption changes were chosen to be consistent with the results of in vivo measurements. The bottom of the horizontal axis label reports the effective homogeneous absorption coefficient (μa,Eff) and reduced scattering coefficient (μs,Eff) obtained from the combination of DSI and DSϕ data.

JBO_28_12_125004_f004.png

When Δμa,top<Δμa,bot [Fig. 4(a)], when μs,top/μs,bot decreases from values > 1 to values less than 1 (i.e., when the top layer transitions from being more scattering to being less scattering than the bottom layer), the effective Δμa,Sim obtained with ϕ data increases, whereas the effective Δμa,Sim obtained with I data decreases. This shows that as the scattering of the top layer decreases, the phase data becomes more representative of the absorption changes in the bottom layer, and intensity data become more representative of absorption changes in the top layer. A physical interpretation of this result is that as the scattering of the bottom-layer increases in relation to the scattering of the top layer, the spatial sensitivity of intensity data shifts toward the top layer, whereas the sensitivity for phase data shifts toward the bottom layer. This is an example of the complementary information provided by intensity data in relation to the two layers of a layered medium, which is the basis for this work about leveraging FD-NIRS data for discrimination of adipose tissue and skeletal muscle hemodynamics.

When Δμa,top=Δμa,bot [Fig. 4(b)], if μs,top>μs,bot SDI data overestimate the true Δμa while all other data types underestimate the true Δμa. As the scattering mismatch between the two layers decreases, Δμa,Sim obtained with SDI approaches the true Δμa and, in the homogeneous case all Δμa,Sim are less than the true Δμa (as a result of a relatively large Δμa and the nonlinear dependence of the various data types changes on Δμa). As μs,top becomes smaller than μs,bot, Δμa,Sim obtained with ϕ data becomes greater than Δμa,Sim obtained with I data. At the smallest value of μs,top/μs,bot, it is Δμa,Sim obtained with phase data that overestimates the true Δμa. Lastly, when Δμa,top>Δμa,bot [Fig. 4(c)], varying the scattering mismatch between the two layers has no major impact, with the most notable effects being a consistently good agreement between Δμa,Sim obtained with DSϕ data and Δμa,bot, and a worsening agreement between Δμa,bot and Δμa,Sim obtained with DSI data as the ratio μs,top/μs,bot decreases.

Of the cases considered in Fig. 4, there is only one that reproduces the paradoxical behavior observed in vivo during arterial occlusion (in all three subjects) and venous occlusion (in subject 2): it is the case of μs,top>μs,bot and Δμa,top<Δμa,bot [Fig. 4(a)]. For the venous occlusion protocol in subjects 3 and 4, two possible scenarios from the two-layer simulations reproduced the in vivo results. They are (1) μs,top>μs,bot and Δμa,topΔμa,bot and (2) μs,top<μs,bot and Δμa,top>Δμa,bot.

As a final comment on the results of Fig. 4, we note that in the case of a homogeneous medium and a homogeneous absorption change [the set of bars in the center of Fig. 4(b)], the underestimation of the true Δμa by Δμa,Sim is due to the non-linear dependence on Δμa of Δln(ρ2I), Δϕ, and associated slope changes. To mimic the experimental results, in Fig. 4(b) we have used a relatively large Δμa (about 50% of the background μa), which is beyond the linear approximation intrinsic in the DPF and differential slope factor (DSF) approach. We have verified that using a much smaller Δμa (<0.1% of the background μa), Δμa,Sim is about equal to the true Δμa for all data types.

3.2.2.

Effect of baseline absorption in the two layers

Simulated absorption changes (Δμa,Sim) from data generated in a two-layered medium for different values of the top and bottom layer μa are shown in Fig. 5. In this simulation, μs,top=1.1  mm1, μs,bot=0.5  mm1, and Ltop=5  mm. These values of μs in the two layers were chosen as they were most representative of the in vivo results observed in both occlusions protocols, as discussed in relation to Fig. 4. From Fig. 5, it is seen that the values of μa,top and μa,bot do not significantly impact the qualitative relationship between the absorption changes obtained with the six data types. In the case Δμa,top>Δμa,bot [Fig. 5(c)], it is worth noting that the absorption change obtained with all the data types tends to approach the true Δμa,bot as μa,top/μa,bot increases.

Fig. 5

Simulated absorption coefficient change (Δμa,Sim) obtained with different data types collected on two-layered medium with different baseline absorption coefficient in the top layer (μa,top) and bottom layer (μa,bot). For all simulations, the top layer thickness (Ltop) was 5 mm, and the reduced scattering coefficient (μs) was set to 1.1  mm1 for the top layer (μs,top) and to 0.5  mm1 for the bottom layer (μs,bot). In all panels, measurements using phase (ϕ) data are represented by a hatched bar, measurements using intensity (I) data by a solid bar, measurements using single-distance (SD) at 25 mm are colored in red, measurements using SD at 35 mm are in green, and measurements using dual-slope (DS) are in blue. The absorption coefficient change for the top (Δμa,top) and bottom (Δμa,bot) layers were simulated for the cases of (a) Δμa,top<Δμa,bot(Δμa,top=0.0008  mm1,Δμa,bot=0.008  mm1); (b) Δμa,top=Δμa,bot(Δμa,top=Δμa,bot=0.008  mm1); and (c) Δμa,top>Δμa,bot(Δμa,top=0.008  mm1,Δμa,bot=0.0008  mm1). These values of baseline optical properties and absorption changes were chosen to be consistent with the results of in vivo measurements. The bottom of the horizontal axis label reports the effective homogeneous absorption coefficient (μa,Eff) and reduced scattering coefficient (μs,Eff) obtained from the combination of DSI and DSϕ data.

JBO_28_12_125004_f005.png

3.2.3.

Effect of top layer thickness

Simulated absorption changes (Δμa,Sim) from data generated in a two-layered medium for different values of the top layer thickness (in the range 2 to 8 mm) are shown in Fig. 6. In the cases where Δμa,top<Δμa,bot [Fig. 6(a)] or Δμa,top>Δμa,bot [Fig. 6(c)], as the top layer thickness increases Δμa,Sim tends toward Δμa,top for every data type. This is an intuitive result since a thicker superficial layer is expected to result in a greater sensitivity to it. The case where Δμa,top=Δμa,bot [Fig. 6(b)] yields less intuitive results in which Δμa,Sim obtained with I data show a closer agreement to the true homogeneous Δμa than Δμa,Sim obtained with ϕ data. This latter result was also observed in the case of μs,top/μs,bot1, regardless of the baseline absorption coefficients of the two layers [see Figs. 4(b) and 5(b)].

Fig. 6

Simulated absorption coefficient change (Δμa,Sim) obtained with different data types collected on a two-layered medium with different values of the top-layer thickness (Ltop). For all simulations, the absorption coefficient of the top layer (μa,top) was set to 0.005  mm1 and the one of the bottom layer (μa,bot) was set to 0.015  mm1, whereas the reduced scattering coefficient was set to 1.1  mm1 for the top layer (μs,top) and to 0.5  mm1 for the bottom layer (μs,bot). In all panels, measurements using phase (ϕ) data are represented by a hatched bar, measurements using intensity (I) data by a solid bar, measurements using single-distance (SD) at 25 mm are colored in red, measurements using SD at 35 mm are in green, and measurements using dual-slope (DS) are in blue. The absorption coefficient change for the top (Δμa,top) and bottom (Δμa,bot) layers were simulated for the cases of (a) Δμa,top<Δμa,bot(Δμa,top=0.0008  mm1,Δμa,bot=0.008  mm1); (b) Δμa,top=Δμa,bot(Δμa,top=Δμa,bot=0.008  mm1); and (c) Δμa,top<Δμa,bot(Δμa,top=0.008  mm1,Δμa,bot=0.0008  mm1). These values of baseline optical properties and absorption changes were chosen to be consistent with the results of in vivo measurements. The bottom of the horizontal axis label reports the effective homogeneous absorption coefficient (μa,Eff) and reduced scattering coefficient (μs,Eff) obtained from the combination of DSI and DSϕ data.

JBO_28_12_125004_f006.png

4.

Discussion

4.1.

Baseline In Vivo Optical Properties

The effective homogeneous baseline optical properties, measured on the human forearm by a combination of DSI and DSϕ data analyzed with diffusion theory for a semi-infinite homogeneous medium, are reported in Table 1 for all subjects and wavelengths (690 and 830 nm). The μa and μs values are in the range of 0.015 to 0.019  mm1 and 0.48 to 0.60  mm1, respectively, across the various subjects. These values agree with optical properties measured non-invasively on human skeletal muscle (specifically, the vastus lateralis) using FD-NIRS.9 With an assumed 75% water content, these baseline absorption coefficients at the two wavelengths result in the following range of values for the heme-related quantities: O: 40 to 56  μm; D: 24 to 35  μm; T: 69 to 82  μm; and tissue saturation (StO2=O/T): 55% to 70%. These values are also comparable to previous FD-NIRS measurements10,13,32 even though StO2 values are at the lower end of the range of values typically reported, possibly due to correction for water absorption.33 In this work, the relevance of baseline effective optical properties is for the determination of LI and Lϕ used to calculate the Δμa from SD and DS data according to Eqs. (1) and (3).

4.2.

Absorption Changes and Associated Hemodynamics In Vivo

The measured absorption changes of the forearm tissue in response to vascular occlusions vary for different data types (Fig. 3). Because these absorption changes reflect blood accumulation (during a venous occlusion) and blood deoxygenation (during an arterial occlusion), they are the basis for measurements of muscle blood flow (using a venous occlusion protocol) and muscle oxygen consumption (using an arterial occlusion protocol). Even though the experimental Δμa values reported in Fig. 3 are maximum changes during vascular occlusions, whereas blood flow and oxygen consumption are obtained from initial rates of change of ΔO and ΔD during vascular occlusion, we translated the initial rates of change in our results to oxygen consumption and blood flow. The range of values observed for rest blood flow was 0.48 to 4.7  mL/(100  mL)/min and for rest oxygen consumption was 0.95 to 2.2  μmol/(100  mL)/min or 0.02 to 0.05  mL/(100  mL)/min. These values are comparable to those previously reported using NIRS measurements during venous or arterial occlusion, respectively.6,32,34,35 The emphasis of this work is on taking advantage of the different values of Δμa obtained with the different data types considered here (SDI, SDϕ, DSI, and DSϕ) to be able to discriminate the superficial (skin + adipose tissue) and deeper (skeletal muscle) hemodynamics. The goal of this research is to perform hemodynamic measurements that are representative of skeletal muscle considering the fact that adipose tissue and muscle tissue have different baseline optical properties and feature different hemodynamics during external perturbations or physiological processes.

4.3.

Theoretical Simulations in a Two-Layered Medium

We used diffusion theory to simulate data in a two-layered medium of known geometrical and optical properties to reproduce key features of experimental data collected on the forearm muscle during vascular occlusion protocols. To make a fair comparison, the same methods for calculating Δμa was used for simulated and experimental data. A fundamental approach in these methods was that the values of L, which are used to find the DPF and DSF for I and ϕ data, were obtained using effective homogeneous optical properties. Previous work has studied the dependence of the SDI DPF (obtained from the mean photon time-of-flight measured with TD-NIRS) on the optical properties and top layer thickness of a two-layered medium36 but did not study the effect of assuming a homogeneous medium in the calculation of the DPF. To see the impact on our results of obtaining DPF and DSF under the assumption that the tissue is homogeneous, we re-ran the simulations reported in Fig. 4 using the correct DPF and DSF (i.e., using LM values obtained from the actual optical properties of a two-layered medium). Our most surprising and counter-intuitive result is the one reported in Fig. 4(a), where the Δμa (obtained with intensity and phase data) exhibit an opposite dependence on the source–detector distance when μs,top>μs,bot. Our goal was to explore whether this result may be due to the incorrect assumption of a homogeneous medium in the determination of DPF and DSF.

For the specific case of Fig. 4(a) with μs,top=1.1  mm1 and μs,bot=0.5  mm1, Fig. 7(a) shows simulations using DPF and DSF calculated from the effective homogeneous optical properties [as in Fig. 4(a)], whereas Fig. 7(b) shows simulations using DPF and DSF calculated using the correct two-layer medium optical properties. Using the correct two-layer absorption changes [Fig. 7(b)], the values of Δμa obtained with the different data types show the consistent behavior of increasing source–detector distance (and being maximal in the DS configuration) for both I and ϕ data, whereas using the homogeneous pathlength [Fig. 7(a)] results in an opposite behavior of intensity and phase data. This result is now fully consistent with a greater absorption change in the bottom layer and shows that the paradoxical result of Fig. 4(a) for μs,top=1.1  mm1 and μs,bot=0.5  mm1 may indeed follow from the use of effective homogeneous optical properties instead of the correct two-layer properties in the calculation of DPF and DSF. To further test this result on experimental data, we calculated Δμa,Exp for one subject (subject 1) and one wavelength (690 nm) during an arterial occlusion using the DPF and DSF calculated for the two-layer optical properties that best matched the experimental data [as also used in Figs. 7(a) and 7(b)]. The L values derived for I (LI) and ϕ (Lϕ) assuming homogeneous and two-layered optical properties are reported in Table 2. Using the DPF and DSF for a homogeneous medium results in values of Δμa,Exp [reported in Fig. 7(c)] that replicate those of Fig. 4(a). Using the DPF and DSF for a two-layered medium modifies the values of Δμa,Exp [as reported in Fig. 7(d)] similarly to the simulation in Fig. 7(b). This is one particular case in which changing the method of calculating L resulted in a significant difference in our results, but this was not the case for all other subjects.

Fig. 7

Comparison of the absorption coefficient change (Δμa) obtained from a generalized mean optical pathlength (L) assuming either a homogeneous medium [panels (a) and (c)] or a two-layered medium [panels (b) and (d)] in simulations (Δμa,Sim) [panels (a) and (b)] and in experiments in vivo (Δμa,Exp) [panels (c) and (d)] on subject 1 at 690 nm during an arterial occlusion. For all simulations, the top layer thickness (Ltop) was 2 mm, the absorption coefficient (μa) in the top layer (μa,top) was 0.005  mm1 and in the bottom layer (μa,bot) was 0.015  mm1, and the reduced scattering coefficient (μs) for the top layer (μs,top) was 1.1  mm1 and in the bottom layer (μs,bot) was 0.5  mm1.

JBO_28_12_125004_f007.png

Table 2

Generalized mean optical pathlength for intensity (⟨L⟩I) and phase (⟨L⟩ϕ) at 690 nm using either a homogeneous medium having optical properties calculated from experimental data, or a two-layered medium. For the two-layer medium, the top layer thickness (Ltop) was 2 mm, the absorption coefficient in the top layer (μa,top) was 0.005  mm−1 and in the bottom layer (μa,bot) was 0.015  mm−1, the reduced scattering coefficient for the top layer (μs,top′) was 1.1  mm−1, and in the bottom layer (μs,bot′) was 0.5  mm−1. Here, the source–detector distances (ρ) are 25 mm (ρ1) and 35 mm (ρ2).

Sub 1 at 690 nm⟨L⟩I (mm)⟨L⟩ϕ (mm)
ρ1ρ2ρ1ρ2
Homogeneous medium10615310.315.6
Two-layer medium12116812.919.8

In the absence of a knowledge of the optical properties of the top and bottom layer, the assumption of a homogeneous medium to calculate the DPF and DSF is required. Several studies, particularly in CW-NIRS, have assumed a wavelength-independent DPF6,7,12 taken from the literature, which will introduce errors in the measured values of Δμa and in their wavelength dependence.

4.4.

Limitations of Modeling Based on a Two-Layered Medium

Modeling tissue as a two-layered medium assumes lateral homogeneity. Lateral in-homogeneities in tissue optical properties and in tissue hemodynamics under the optical probe can arise for several reasons, including physiological/anatomical differences or uneven contact pressure between the probe and tissue. Spatial mapping of skeletal muscle has shown variability across muscle regions.9,34,37 The datasets for this study resulted in similar Δμa,Exp for the two SDs at the same distance but on opposite sides on the probe [Fig. 1(c)], which supports the assumption of lateral homogeneity. However, in the presence of lateral heterogeneity in the tissue optical properties, tissue hemodynamics, and/or probe-tissue contact pressure, the assumption of lateral homogeneity intrinsic in the two-layered model may lead to incorrect results. Special care must be taken in applying the optical probe with even pressure on the muscle tissue, and the lateral homogeneity of the measured hemodynamics should be verified before applying the methods presented here.

4.5.

Implications of this Study for Optical Measurements of Muscle Hemodynamics

Comparison between the experimental results in vivo for the various data types (Fig. 3) and the simulations on a two-layered medium (Figs. 46) lead to the following conclusions for the baseline optical properties and the absorption changes in the top and bottom tissue layers.

  • 1. In the case of an arterial occlusion, μs,top>μs,bot and |Δμa,top|<|Δμa,bot| (see Fig. 4), and μa,topμa,bot [see Fig. 5(a)]. These results for baseline optical properties are consistent with the higher scattering and lower absorption previously reported for adipose tissue versus muscle tissue.8,38,39 The lower absorption change in the top (adipose) layer versus the bottom (muscle) layer during arterial occlusion implies a lower oxygen consumption in adipose versus muscle tissue at rest, which has previously been reported.6,8,24,40

  • 2. In the case of venous occlusion, the in vivo results in subject 2 are in line with the arterial occlusion results above, whereas in subjects 3 and 4 they are consistent with either of these two scenarios: (1) μs,top>μs,bot and Δμa,topΔμa,bot [see Figs. 4(b) and 4(c)] or (2) μs,top<μs,bot and Δμa,top>Δμa,bot [Fig. 4(c)]. For consistency with the baseline optical properties observed in the arterial occlusion case, scenario (1) is the selected one. The greater or comparable absorption change in the top (adipose) layer versus the bottom (muscle) layer during venous occlusion implies a greater or comparable blood flow in adipose versus muscle tissue at rest. This result is in line with a previous study that reported a constant (or slightly increasing) measured blood flow with NIRS as a function of adipose layer thickness.6

  • 3. The top layer thickness in the range 2 to 8 mm (Fig. 6) significantly impacted the magnitude of Δμa obtained using different data types, but only marginally impacted their relative values.

The results presented here demonstrate the different impact of baseline optical properties and absorption changes of a two-layered medium on the various data types measured with FD-NIRS in a DS configuration. Therefore, these results suggest the feasibility of discriminating baseline optical properties and hemoglobin concentration changes in adipose and muscle tissue by fitting measured and simulated Δμa obtained with the various data types. The results presented here do not identify a single data type (SDI, SDϕ, DSI, and DSϕ) that is most effective at sensing the deeper absorption changes (in muscle tissue) with minimal sensitivity to superficial absorption changes (in skin and adipose tissue). While previous work identified DSϕ as a data type that should indeed feature an optimal preferential sensitivity to deeper tissue,19 such previous work assumed optical properties that are homogeneous at baseline. In this work, we have shown that layered optical properties, and especially a top layer that is more scattering than the bottom layer, may significantly impact the spatial regions of sensitivity of the various data types considered here depending on the top layer thickness and the specific distribution of optical properties at baseline. Consequently, rather than identifying a single most effective data type for non-invasive muscle measurements, we propose to take advantage of the collective information content of all data types collected by DS FD-NIRS, specifically SDI(ρ1), SDI(ρ2), SDϕ(ρ1), SDϕ(ρ2), DSI, and DSϕ, to discriminate the superficial and deeper tissue hemodynamics. This may be done by fitting the data with a two-layered model of the investigated tissue to obtain baseline and dynamic optical properties of the two layers (with the top layer thickness measured independently or also treated as a fitting parameter). While this method is rigorous, it may require data at more than two source–detector distances for an accurate determination of all fitting parameters. Alternatively, we could use the two-layer model in a different way. First, the DS FD-NIRS data are used to measure effective baseline optical properties (μa,Eff and μs,Eff from which DPF and DSF values are obtained), the top layer thickness is measured independently (for example by a skinfold caliper or ultrasound imaging), and the effective Δμa values obtained with the different data types are then measured during protocols that elicit tissue hemodynamic and/or oxygenation changes. The set of effective optical properties and effective Δμa values are then fit with those obtained/assumed from two-layered media, using the four fitting parameters of μs,top, μs,bot, Δμa,top, and Δμa,bot (scattering changes may be neglected, i.e., Δμs,top=Δμs,bot=0, as hemodynamics should mostly result in absorption changes, and the baseline absorption may be considered homogeneous, μa,top=μa,bot=μa,Eff, as the specific values of baseline absorption of the two layers do not significantly impact the absorption changes obtained with the six data types, as shown in Fig. 5).

5.

Conclusion

The objective of this work is to explore the feasibility of discriminating superficial and deep tissue hemodynamics by performing non-invasive optical measurements with rich temporal and spatial information content. FD-NIRS is used to collect the amplitude [or intensity (I)] and the phase (ϕ) of photon-density waves launched into the tissue by intensity-modulated illumination. Two source–detector separations are used to collect data (I and ϕ) that feature different depth sensitivities in tissue. A special, symmetrical arrangement of two sources and two detectors is used to perform self-calibrating measurements of baseline optical properties, and robust measurements of absorption changes that are largely insensitive to instrumental drifts and motion artifacts. This work shows that the data types (I and ϕ) and geometrical arrangements of data collection (SD, DS) result in a set of measurements with a strong dependence on the baseline effective optical properties and absorption changes in the two-layered medium. This can be the basis for discriminating absorption dynamics in superficial and deep tissue by collecting data that are analyzed under the assumption of tissue homogeneity and by fitting them using a two-layered tissue model. In the case of muscle measurements, the thickness of the top layer (adipose tissue) can be measured with a skinfold caliper or ultrasound imaging. It is also possible to further expand the data space beyond what was done in this work, by introducing multiple DS sets of sources and detectors that feature multiple pairs of source–detector distances.

To the best of our knowledge, this is the first time that DS FD-NIRS is applied to muscle studies and that the information content of SD and DS FD-NIRS data is investigated in relation to the optical properties of two-layered media.

The approach proposed here, where measurement of effective optical properties and effective absorption changes are analyzed using diffusion theory for two-layered media, is an alternative to methods of data analysis with two-layered tissue models, as already proposed for studies in muscle8,14,26 and brain.4143 These latter methods, however, may be less robust and require a greater data space (in the temporal and spatial domains) than the one proposed here. Future research will explore practical implementations of the method proposed in this work in comparison with a full-fledged two-layer tissue model for the hemodynamic characterization of layered tissue. This is a significant objective for non-invasive optical studies of biological tissue, with direct implications for skeletal muscle studies as well as human brain studies.

Disclosures

The authors disclose no conflicts of interest.

Code and Data Availability

Code and data utilized in this work are available from the authors upon reasonable request.

Acknowledgments

This work is supported by the National Institutes of Health (NIH) (Grant No. R01-EB029414). G.B. is also supported by grant K12-GM133314.

References

1. 

M. Ferrari, M. Muthalib and V. Quaresima, “The use of near-infrared spectroscopy in understanding skeletal muscle physiology: recent developments,” Philos. Trans. R. Soc. A Math. Phys. Eng. Sci., 369 4577 –4590 https://doi.org/10.1098/rsta.2011.0230 (2011). Google Scholar

2. 

T. J. Barstow, “Understanding near infrared spectroscopy and its application to skeletal muscle research,” J. Appl. Physiol., 126 1360 –1376 https://doi.org/10.1152/japplphysiol.00166.2018 (2019). Google Scholar

3. 

M. Wolf, M. Ferrari and V. Quaresima, “Progress of near-infrared spectroscopy and topography for brain and muscle clinical applications,” J. Biomed. Opt., 12 062104 https://doi.org/10.1117/1.2804899 JBOPFO 1083-3668 (2007). Google Scholar

4. 

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

5. 

N. Olamaei et al., “Characterization and reproducibility of forearm arterial flow during reactive hyperemia,” Physiol. Meas., 31 763 https://doi.org/10.1088/0967-3334/31/6/003 PMEAE3 0967-3334 (2010). Google Scholar

6. 

M. C. van Beekvelt et al., “Adipose tissue thickness affects in vivo quantitative near-IR spectroscopy in human skeletal muscle,” Clin. Sci., 101 21 –28 https://doi.org/10.1042/cs1010021 (2001). Google Scholar

7. 

T. J. Cross and S. Sabapathy, “The impact of venous occlusion per se on forearm muscle blood flow: implications for the near-infrared spectroscopy venous occlusion technique,” Clin. Physiol. Funct. Imaging, 37 (3), 293 –298 https://doi.org/10.1111/cpf.12301 (2017). Google Scholar

8. 

E. Ohmae et al., “Sensitivity correction for the influence of the fat layer on muscle oxygenation and estimation of fat thickness by time-resolved spectroscopy,” J. Biomed. Opt., 19 067005 https://doi.org/10.1117/1.JBO.19.6.067005 JBOPFO 1083-3668 (2014). Google Scholar

9. 

V. Quaresima et 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., 9 413 –420 https://doi.org/10.1117/1.1646417 JBOPFO 1083-3668 (2004). Google Scholar

10. 

L. F. Ferreira, D. M. Hueber and T. J. Barstow, “Effects of assuming constant optical scattering on measurements of muscle oxygenation by near-infrared spectroscopy during exercise,” J. Appl. Physiol., 102 358 –367 https://doi.org/10.1152/japplphysiol.00920.2005 (2007). Google Scholar

11. 

N. M. Gregg et al., “Brain specificity of diffuse optical imaging: improvements from superficial signal regression and tomography,” Front. Neuroenerg., 2 14 https://doi.org/10.3389/fnene.2010.00014 (2010). Google Scholar

12. 

V. M. Niemeijer et al., “The influence of adipose tissue on spatially resolved near-infrared spectroscopy derived skeletal muscle oxygenation: the extent of the problem,” Physiol. Meas., 38 539 https://doi.org/10.1088/1361-6579/aa5dd5 PMEAE3 0967-3334 (2017). Google Scholar

13. 

J. C. Craig et al., “Effect of adipose tissue thickness, muscle site, and sex on near-infrared spectroscopy derived total-hemoglobin + myoglobin,” J. Appl. Physiol., 123 1571 –1578 https://doi.org/10.1152/japplphysiol.00207.2017 (2017). Google Scholar

14. 

R. Re et al., “Effect of a thin superficial layer on the estimate of hemodynamic changes in a two-layer medium by time domain NIRS,” Biomed. Opt. Express, 7 264 –278 https://doi.org/10.1364/BOE.7.000264 BOEICL 2156-7085 (2016). Google Scholar

15. 

T. Pham et al., “Sensitivity of frequency-domain optical measurements to brain hemodynamics: Simulations and human study of cerebral blood flow during hypercapnia,” Biomed. Opt. Express, 12 766 –789 https://doi.org/10.1364/BOE.412766 BOEICL 2156-7085 (2021). Google Scholar

16. 

G. Blaney et al., “Phase dual-slopes in frequency-domain near-infrared spectroscopy for enhanced sensitivity to brain tissue: first applications to human subjects,” J. Biophotonics, 13 (1), e201960018 https://doi.org/10.1002/jbio.201960018 (2020). Google Scholar

17. 

T. Pham et al., “Noninvasive optical measurements of dynamic cerebral autoregulation by inducing oscillatory cerebral hemodynamics,” Front. Neurol., 12 745987 https://doi.org/10.3389/fneur.2021.745987 (2021). Google Scholar

18. 

G. Blaney et al., “Towards functional dual-slope imaging: methods to design, evaluate, and construct a source-detector array,” Proc. SPIE, 11639 116390T https://doi.org/10.1117/12.2577715 PSISDG 0277-786X (2021). Google Scholar

19. 

A. Sassaroli, G. Blaney and S. Fantini, “Dual-slope method for enhanced depth sensitivity in diffuse optical spectroscopy,” J. Opt. Soc. Am. A Opt. Image Sci. Vis., 36 1743 –1761 https://doi.org/10.1364/JOSAA.36.001743 (2019). Google Scholar

20. 

G. Blaney et al., “Dual-slope imaging of cerebral hemodynamics with frequency-domain near-infrared spectroscopy,” Neurophotonics, 10 013508 https://doi.org/10.1117/1.NPh.10.1.013508 (2023). Google Scholar

21. 

R. E. Wood and I. B. Stewart, “Can venous occlusion plethysmography be used to measure high rates of arterial inflow?,” Eur. J. Appl. Physiol., 108 239 –245 https://doi.org/10.1007/s00421-009-1208-1 EJAPFN 1439-6319 (2010). Google Scholar

22. 

S. W. Coppack et al., “Postprandial substrate deposition in human forearm and adipose tissues in vivo,” Clin. Sci., 79 339 –348 https://doi.org/10.1042/cs0790339 (1990). Google Scholar

23. 

S. Jones et al., “Recent developments in near-infrared spectroscopy (NIRS) for the assessment of local skeletal muscle microvascular function and capacity to utilise oxygen,” Artery Res., 16 25 –33 https://doi.org/10.1016/j.artres.2016.09.001 (2016). Google Scholar

24. 

L. M. K. Chin et al., “The relationship between muscle deoxygenation and activation in different muscles of the quadriceps during cycle ramp exercise,” J. Appl. Physiol., 111 1259 –1265 https://doi.org/10.1152/japplphysiol.01216.2010 (2011). Google Scholar

25. 

A. Liemert and A. Kienle, “Light diffusion in a turbid cylinder. II. Layered case,” Opt. Express, 18 9266 –9279 https://doi.org/10.1364/OE.18.009266 OPEXFF 1094-4087 (2010). Google Scholar

26. 

G. Blaney et al., “Broadband diffuse optical spectroscopy of two-layered scattering media containing oxyhemoglobin, deoxyhemoglobin, water, and lipids,” J. Innov. Opt. Health Sci., 15 2250020 https://doi.org/10.1142/S1793545822500201 (2022). Google Scholar

27. 

S. Fantini, G. Blaney and A. Sassaroli, “Transformational change in the field of diffuse optics: from going bananas to going nuts,” J. Innov. Opt. Health Sci., 13 1930013 https://doi.org/10.1142/S1793545819300131 (2020). Google Scholar

28. 

D. Meza-Valderrama et al., “The feasibility and reliability of measuring forearm muscle thickness by ultrasound in a geriatric inpatient setting: a cross-sectional pilot study,” BMC Geriatr., 22 137 https://doi.org/10.1186/s12877-022-02811-3 (2022). Google Scholar

29. 

G. Blaney et al., “Dual-slope diffuse reflectance instrument for calibration-free broadband spectroscopy,” Appl. Sci., 11 1757 https://doi.org/10.3390/app11041757 (2021). Google Scholar

30. 

D. Hueber et al., “New optical probe designs for absolute (self-calibrating) NIR tissue hemoglobin measurements,” Proc. SPIE, 3597 618 –631 https://doi.org/10.1117/12.356784 PSISDG 0277-786X (1999). Google Scholar

31. 

V. Perekatova et al., “VIS-NIR diffuse reflectance spectroscopy system with self-calibrating fiber-optic probe: study of perturbation resistance,” Diagnostics, 13 457 https://doi.org/10.3390/diagnostics13030457 (2023). Google Scholar

32. 

K. Gurley, Y. Shang and G. Yu, “Noninvasive optical quantification of absolute blood flow, blood oxygenation, and oxygen consumption rate in exercising skeletal muscle,” J. Biomed. Opt., 17 075010 https://doi.org/10.1117/1.JBO.17.7.075010 JBOPFO 1083-3668 (2012). Google Scholar

33. 

M.-A. Franceschini et al., “Quantitative spectroscopic determination of hemoglobin concentration and saturation in a turbid medium: analysis of the effect of water absorption,” J. Biomed. Opt., 2 147 –153 https://doi.org/10.1117/12.262145 JBOPFO 1083-3668 (1997). Google Scholar

34. 

C. Casavola et 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., 5 269 –276 https://doi.org/10.1117/1.429995 JBOPFO 1083-3668 (2000). Google Scholar

35. 

L. A. Paunescu et al., “Calf muscle blood flow and oxygen consumption measured with near-infrared spectroscopy during venous occlusion,” Proc. SPIE, 3597 317 –323 https://doi.org/10.1117/12.356790 PSISDG 0277-786X (1999). Google Scholar

36. 

I. Pirovano et al., “Effect of adipose tissue thickness and tissue optical properties on the differential pathlength factor estimation for NIRS studies on human skeletal muscle,” Biomed. Opt. Express, 12 571 –587 https://doi.org/10.1364/BOE.412447 BOEICL 2156-7085 (2021). Google Scholar

37. 

U. Wolf et al., “Regional differences of hemodynamics and oxygenation in the human calf muscle detected with near-infrared spectrophotometry,” J. Vasc. Intervent. Radiol., 18 1094 –1101 https://doi.org/10.1016/j.jvir.2007.06.004 JVIRE3 1051-0443 (2007). Google Scholar

38. 

S. L. Jacques, “Optical properties of biological tissues: a review,” Phys. Med. Biol., 58 R37 https://doi.org/10.1088/0031-9155/58/11/R37 PHMBA7 0031-9155 (2013). Google Scholar

39. 

A. N. Bashkatov, E. A. Genina and V. V. Tuchin, “Optical properties of skin, subcutaneous, and muscle tissues: a review,” J. Innov. Opt. Health Sci., 04 9 –38 https://doi.org/10.1142/S1793545811001319 (2011). Google Scholar

40. 

M. Niwayama et 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 4571 –4575 https://doi.org/10.1063/1.1322578 RSINAK 0034-6748 (2000). Google Scholar

41. 

B. Hallacoglu, A. Sassaroli and S. Fantini, “Optical characterization of two-layered turbid media for non-invasive, absolute oximetry in cerebral and extracerebral tissue,” PLoS ONE, 8 e64095 https://doi.org/10.1371/journal.pone.0064095 POLNCL 1932-6203 (2013). Google Scholar

42. 

R. Saager and A. Berger, “Measurement of layer-like hemodynamic trends in scalp and cortex: implications for physiological baseline suppression in functional near-infrared spectroscopy,” J. Biomed. Opt., 13 (3), 034017 https://doi.org/10.1117/1.2940587 JBOPFO 1083-3668 (2008). Google Scholar

43. 

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

Biography

Cristianne Fernandez received her PhD from Tufts University (Medford, Massachusetts, United States) in 2023 in the DOIT lab under the advisement of Prof. Sergio Fantini. Her work mainly focused on using frequency-domain near-infrared spectroscopy with coherent hemodynamics spectroscopy to measure blood flow oscillations in the brain and measurements of hemodynamics more representative of skeletal muscle. Before her time at Tufts University, she received an undergraduate degree in biomedical engineering from Florida International University (Miami, Florida, United States).

Giles Blaney is a National Institutes of Health Institutional Research and Academic Career Development Award Postdoctoral Scholar in the DOIT lab at Tufts University. He received his PhD from Tufts University (Medford, Massachusetts, United States) in 2022 after working in the same lab with Prof. Sergio Fantini as his advisor. Before that he received an undergraduate degree in mechanical engineering and physics from Northeastern University (Boston, Massachusetts, United States). His current research interests include diffuse optics and its possible applications within and outside of medical imaging.

Jodee Frias is a second-year biomedical engineering PhD student in the DOIT Lab at Tufts University under Prof. Sergio Fantini. Prior to this, she received her undergraduate degree in biomedical engineering from Boston University. Her current research focuses on the use of near-infrared spectroscopy to investigate skeletal muscle and brain hemodynamics.

Fatemeh Tavakoli is a second-year PhD student in the DOIT lab under the advisement of Prof. Sergio Fantini at Tufts University (Medford, Massachusetts, United States). Prior to that, she received an MSc degree in electrical engineering from Islamic Azad University, Science and Research Branch in 2020 (Tehran, Iran). During her master’s degree, she found Fractal geometry to be a unique way of tuning optical properties, and she successfully published several papers in the area of novel nanoscale structures for biosensing applications. Her current research interests include near-infrared spectroscopy to investigate non-invasive techniques for hemodynamic monitoring of skeletal muscle and brain function.

Angelo Sassaroli received his PhD in physics in 2002 from the University of Electro-Communications (Tokyo, Japan). From July 2002 to August 2007, he was a research associate in the research group of Prof. Sergio Fantini at Tufts University. In September 2007, he was appointed by Tufts University as a research assistant professor. His field of research is near-infrared spectroscopy and diffuse optical tomography.

Sergio Fantini is professor of biomedical engineering and principal investigator of the DOIT at Tufts University. His research activities on the application of diffuse optics to biological tissues resulted in about 130 peer-reviewed scientific publications and 13 patents. He co-authored with Prof. Irving Bigio (Boston University, Boston, Massachusetts, United States) a textbook on “Quantitative Biomedical Optics” published by Cambridge University Press in 2016. He is a fellow of SPIE, Optica, and the American Institute for Medical and Biological Engineering.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 International License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Cristianne Fernandez, Giles Blaney, Jodee Frias, Fatemeh Tavakoli, Angelo Sassaroli, and Sergio Fantini "Single-distance and dual-slope frequency-domain near-infrared spectroscopy to assess skeletal muscle hemodynamics," Journal of Biomedical Optics 28(12), 125004 (14 December 2023). https://doi.org/10.1117/1.JBO.28.12.125004
Received: 11 July 2023; Accepted: 20 November 2023; Published: 14 December 2023
Advertisement
Advertisement
KEYWORDS
Absorption

Adipose tissue

Muscles

Optical properties

Hemodynamics

Tissues

Near infrared spectroscopy

Back to Top