Open Access
1 November 2007 Progress of near-infrared spectroscopy and topography for brain and muscle clinical applications
Author Affiliations +
Abstract
This review celebrates the 30th anniversary of the first in vivo near-infrared (NIR) spectroscopy (NIRS) publication, which was authored by Professor Frans Jöbsis. At first, NIRS was utilized to experimentally and clinically investigate cerebral oxygenation. Later it was applied to study muscle oxidative metabolism. Since 1993, the discovery that the functional activation of the human cerebral cortex can be explored by NIRS has added a new dimension to the research. To obtain simultaneous multiple and localized information, a further major step forward was achieved by introducing NIR imaging (NIRI) and tomography. This review reports on the progress of the NIRS and NIRI instrumentation for brain and muscle clinical applications 30 years after the discovery of in vivo NIRS. The review summarizes the measurable parameters in relation to the different techniques, the main characteristics of the prototypes under development, and the present commercially available NIRS and NIRI instrumentation. Moreover, it discusses strengths and limitations and gives an outlook into the "bright" future.

1.

Introduction

This review celebrates the 30th anniversary of the first in vivo near-infrared (NIR) spectroscopy (NIRS) publication,1 which was authored by Frans Jöbsis, who described his discoveries in two papers published in the Journal of Biomedical Optics 22 years after his original publication.2, 3

Starting with the pioneering work of Jöbsis, noninvasive NIRS was first utilized to investigate cerebral oxygenation experimentally and clinically and, later on, muscle oxidative metabolism. In addition, since 1993, multichannel NIRS instruments have been largely applied to investigate the functional activation of the human cerebral cortex in adults 4, 5, 6, 7 and later in newborns.8 A number of recent detailed reviews describing the principles, the limitations, and the applications of NIRS have appeared in the literature. 9, 10, 11, 12, 13, 14, 15, 16, 17, 18 The same is true for reviews describing the applications of NIRS on cerebral oxygenation monitoring in newborns and adults. 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29

The most recently available NIRS technology for monitoring cerebral oxygenation can contribute to the identification of deficits in cerebral oxygenation. Monitoring such deficits supports certain forms of therapy in reversing cerebral oxygenation issues and thereby preventing long-term neurological sequelae. Recently, it has been demonstrated that quantitative thresholds for cerebral oxygenation led to the identification of cerebral ischemia in the adult brain and thus increased the scope of clinical use of NIRS.29 A number of recent detailed reviews describe the use of NIRS and NIRS imaging (NIRI) for human brain mapping 15, 30, 31, 32, 33, 34, 35, 36 and muscle exercise pathophysiology. 37, 38, 39, 40, 41, 42, 43

This review reports on the progress of the NIRS and NIRI instrumentation for brain and muscle clinical applications, 30 years after the discovery of in vivo NIRS. The review summarizes the measurable parameters in relation to the different NIRS techniques, the main characteristics of the prototypes under development, and the present commercially available NIRS and NIRI instrumentation. Moreover, a discussion on the strengths and limitations of NIRS and/or NIRI and an outlook into the “bright” future are reported.

2.

Methods

Papers were retrieved by the authors through different strategies. First, a search on the two databases MEDLINE and INSPEC was performed using the keywords: “near infrared,” “near infrared oximetry,” “cerebral oximetry,” “muscle oxygenation,” “optical imaging,” and/or “instrument.” The references were screened and the full texts of relevant publications were retrieved. Next, the references of reviews were hand searched. The research was restricted to literature on the NIRS and/or NIRI instrumentation suitable for human muscle and brain measurements published or made available up to February 2007. Breast imaging instrumentation was not included, because its progress has recently been reviewed.44, 45, 46 In addition, three-dimensional tomography was excluded, because it is covered by another paper in honor of Professor F. F. Jöbsis. The very recent proceedings of conferences organized by the following societies: Optical Society of America, The International Society of Optical Engineering (SPIE), Organization of Human Brain Mapping, American College of Sports Medicine, and the Polish Academy of Sciences were also consulted. Research groups known to be active in the field were contacted for gathering further information. The Web sites of the commercial systems were searched and visited for exploring the specifications of the instruments. After collecting all the documentation, a consensus was made by all authors to properly select material eligible for inclusion in this review. The material was sorted according to the type of NIRS and NIRI instrumentation and the parameters measured. Tables were generated to report the origin and properties of each instrument and all the measurable parameters.

3.

Results

NIR from the 650- to 950-nm wavelength penetrates tissue relatively deeply. In this region of wavelength, chromophores such as oxyhemoglobin ( O2Hb in micromolar concentration), deoxyhemoglobin (HHb in micromolar concentration), cytochrome oxidase, water, lipids, and indocyanine green absorb light. Thus their concentration can in principle be measured by NIRS and NIRI. However, besides the light absorption, the strong light scattering of tissue in the NIR has to be taken into consideration. To quantify the measurements, theoretical models describing light transportation in tissue have been developed.47 Because a general mathematical approach is not feasible, all the mathematical models rely on assumptions and approximations to simplify matters.47 It is important to ensure that these assumptions are fulfilled, when applying NIRS and NIRI.

The most widely used approximations are the differential pathlength factor (DPF) method48, 49, 50 and the diffusion approximation. 47, 51, 52, 53 The DPF method is a relatively simple model that enables us to quantify changes in chromophore concentration. Absolute values cannot be obtained directly by the DPF method. Only changes in light attenuation are measured, and it is assumed that these changes reflect changes in the chromophore concentration. If geometrical or structural changes occur, they will be misinterpreted as changes in the chromophore concentration, which, for instance, might occur during motion artifacts. In addition, the DPF method assumes that the tissue and the change in chromophore concentration are homogeneous. To obtain quantitative values, the DPF, which accounts for the increased pathlength due to light scattering, has to be measured or taken from the literature. 54, 55, 56, 57, 58, 59

The diffusion approximation of the Boltzmann transport equation is another widely used mathematical model. The diffusion approximation has analytical solutions under the following assumptions: (1) tissue is homogeneous, (2) scattering is much larger than absorption, and (3) the tissue has a specific geometry—infinite, semi-infinite, slab, or two-layered.47 To obtain correct values, it is again vital to observe these boundary conditions. The DPF method is in agreement with the diffusion approximation. The diffusion approximation can be used to measure absolute values of the absorption and scattering coefficients of tissue and from the absorption coefficient, absolute values of the chromophore concentration can be calculated. Generally, this requires measuring light intensity and the time of flight (i.e., the time the light takes to pass through the tissue).

Several techniques to physically carry out the measurements have been described and applied. Table 1 summarizes the different types of instruments and indicates key features, advantages, and disadvantages. The parameters that can be measured are outlined in Table 2 .

Table 1

Near-infrared spectroscopy and imaging instrumentation: Characteristics and main parameters directly measured.

Single-Distance CW Photometers1 or 2 Channels OximetersImagers
Parameters measured and instrument characteristicsDiscrete wavelengthsBroadband second derivativeDWSSRS CWPMS MDPMS MFTRSCWPMSTRS
[O2Hb] , [HHb] , [tHb] yes, changes1yes, absolute valuenoyes, changes1yes, absolute valueyes, absolute valueyes, absolute valueyes, changes1yes, absolute valueyes, absolute value
Blood flow measurementnonoyes, relativenonononononono
Scattering and absorption coefficient and pathlength measurementnoyes, pathlengthnonoyesyesyesnoyesyes
Tissue O2Hb saturation measurement ( SO2 ,%)noyesnoyesyesyesyesnoyesyes
Penetration depth with a 4-cm source-detector separationlowlowlowlow, but deep for SO2 deeplowlowlowdeeplow
Sampling rate (Hz) 100 1 5 6 100 1 6 100 50 1
Spatial resolution (cm)n.a.n.a.feasiblen.a.n.a.n.a.n.a. 1 1 1
Instrument sizevery smallmediummediumsmallsmallmediummediumsome bulky, some smallbulkybulky
Instrument stabilizationn.r.n.r.required.n.r.n.r.n.r.requiredn.r.n.r.required
Transportabilityeasyeasyfeasibleeasyeasyeasyeasysome easy, some feasiblefeasiblefeasible
Instrument costlowmoderatehighmoderatemoderatehighhighsome low, some highvery highvery high
Caution for eye exposure to coherent sourcesn.r.n.r.requiredn.r.n.r.n.r.requireddepends on instrumentrequiredrequired
Stable optical contactcriticalcriticalcriticalnot criticalnot criticalcriticalnot criticalcriticalnot criticalnot critical
Precise anatomical localizationnononononononoscarcescarcescarce
Telemetryavailablen.a.n.a.availabledifficultdifficultdifficultavailabledifficultnot easy
Discrimination between cerebral and extracerebral tissue (scalp, skull, CSF)n.a.n.a.feasiblen.a.feasiblen.a.feasiblen.a.feasiblefeasible
Possibility to measure deep brain structuresfeasible on newbornsfeasible on newbornsfeasible on newbornsfeasible on newbornsfeasible on newbornsfeasible on newbornsfeasible on newbornsfeasible on newbornsfeasible on newbornsfeasible on newborns

1When the differential pathlength factor (DPF) is included to calculate the tissue pathlength [ =DPF× (source-detector separation)].

CSF=cerebrospinal fluid, CW=continuous wave, DWS=diffusing-wave spectroscopy or diffuse correlation spectroscopy, HHb=deoxyhemoglobin, MD=multidistance geometry, MF=multifrequency measurement, n.a.=not available, n.r.=not required, O2Hb =oxyhemoglobin, PMS=phase modulation spectroscopy, SRS=spatially resolved spectroscopy, tHb=O2Hb+HHb , TRS=time resolved spectroscopy.

Table 2

Parameters measured directly and indirectly by near-infrared spectroscopy and imaging instrumentation.

ParameterUnitsModalityApplicability (during muscle exercise)Author [reference]
ΔO2Hb , ΔHHb , ΔtHb ,Delpy 1997129
ΔoxCCO ,a.u., μM×cm , μM DYesTisdall 200764
OIGrassi 1999130
D (by SRS)YesMatcher 1995,75 De Blasi 1993,1994,104, 105 Quaresima 2002,131 Cuccia 200580
Tissue O2 saturation%D (by PMS)YesFantini 199576
D (by TRS)YesOda 1996132
D (by calibration)YesBenni 2005133
Second differentialNoMatcher 1994, Cooper 199658, 134
Muscle SvO2 %I (by VOM)NoYoxall 1997135
DNoFranceschini 2002136
Muscle tHb μM D (by PMS)YesFranceschini 1997126
a.u.D (by DWS)NoDurduran 200395
Muscle BF mL100mLmin I (by VOM)NoDe Blasi 1994105
I (by ICG)YesBoushel 2000137
Muscle Hb flow μMmin I (by VOM)NoWolf 2003106
Muscle VO2 mL100gmin I (by VOM)NoDe Blasi 1993, 1994104, 105
I (by AOM)
Muscle recovery timesDNoChance 1992138
Muscle compliancemL/L/mmHgINoBinzoni 2000139
Cerebral SvO2 %I (by VOM)NoYoxall 1995140
DNoWolf 199772
Cerebral tHb μM D (by PMS)YesChoi 200478
I (by O2 swing)NoWolf 2002141
I (by O2 swing)NoWyatt 1990,101 Wolf 2002141
Cerebral BV mL100mL SRS and second differentialNoLeung 2006142
I (by ICG)NoHopton 1999143
a.u.D (by DWS)NoDurduran 2004,96 Li 200594
Cerebral BF mL100mLmin I (by O2 swing)NoEdwards 1988100
I (by ICG)Roberts 1993,144 Keller 2000145
Cerebral VO2 mL100gmin Combination cerebral SvO2 and BFNoElwell 2005146
Δ =Relative changes from arbitrary baseline, AOM=arterial occlusion Method, a.u.=arbitrary units, BF=blood flow, BV=blood volume, DWS=diffusing-wave spectroscopy, D=directly, I=indirectly, ICG=indocyanine green, OI=oxygenation index (ΔO2Hb-ΔHHb) , oxCCO=cytochrome c oxidase redox state, PMS=phase modulation spectroscopy, SRS=spatially resolved spectroscopy, SvO2 =venous O2 saturation, tHb=O2Hb+HHb , TRS=time resolved spectroscopy, VO2 =oxygen consumption, VOM= venous occlusion method.

Most of the parameters are based on the measurement of O2Hb and HHb . In addition, NIRS’s measurement of the changes in the redox state of oxidized cytochrome c oxidase (ΔoxCCO) , as first proposed by Jöbsis,1 has the potential to provide a unique method for monitoring changes of intracellular O2 delivery.9, 60 Although much work has been done on the refinement of NIRS hardware and algorithms (utilized to deconvolute the light absorption signal), recent years have seen a vivid discussion in the literature on the possibility of measuring ΔoxCCO by NIRS. To improve the accuracy of the measurement of this NIRS parameter, most of the recent animal61, 62 and human63, 64 NIRS studies have been performed using a broadband approach with a continuous white light spectrum.

Continuous wave (CW) means that only changes in the light intensity are measured. Usually at least two different wavelengths are multiplexed to obtain spectral information. The ambient light level is also measured and subtracted by the NIRS instrument. CW can easily be used for imaging by using many source-detector pairs, which are distributed on the tissue of interest. 15, 65, 66, 67, 68, 69, 70 This method only allows the continuous quantification of relative values (except for absolute values of venous oxygen saturation71, 72) and usually relies on the DPF method. Another disadvantage is represented by the fact that it is relatively sensitive to motion artifacts. The advantages are that CW is inexpensive and can be miniaturized to the extent of a wireless instrument,73 even for imaging (Fig. 1 ). In addition, in many situations (e.g., studies of functional activity of the brain or intervention studies for testing reactions on drugs or changes in treatment15, 32, 74) relative values are sufficient (Fig. 2 ).

Fig. 1

Wireless imaging instrument attached to a newborn infant’s head. The squares (blue) represent the detector locations, while the circles (red) depict source locations, each equipped with light emitting diodes at two wavelengths (730 and 830nm ). The electronics to the right includes a Bluetooth device for wireless transmission, drivers for the light emitting diodes, filters, analog-to-digital converters, a microprocessor, and a power supply based on a battery. The instrument weighs as little as 40g , has a sample rate of 100Hz , and the battery lasts for approximately 3h . The wireless technology is comfortable to wear, easy to apply, and enables measurements in moving subjects and everyday situations. (Color online only.)

062104_1_007706jbo1.jpg

Fig. 2

A sample of a functional NIRs measurement with a 100-Hz sampling rate in a healthy neonate. The upper trace (red) depicts O2Hb , and the lower trace (blue) HHb and the straight line (black) depict the duration of the visual stimulation. A number of physiological phenomena can be observed: (1) The arterial pulsations are visible in the O2Hb tracing. The pulsations can be used to calculate the heart rate and arterial oxygen saturation. (2) Approximately every 10s , there are fluctuations in the blood circulation (the so-called slow vasomotion). These changes are particularly evident in the O2Hb tracing. (3) The O2Hb increases and the HHb decreases during the stimulation. This corresponds to a typical functional cortical activation. Although the slow vasomotion partially masks the activation, the measurement can be repeated several times and thus the functional activation can be revealed statistically. (Color online only.)

062104_1_007706jbo2.jpg

Table 4

Main recently developed near-infrared prototypes.

Name of the instrument or town of the universityTechniqueNumber of channelsUniversity or firmAuthor [reference]
OximetersIrvineBroadband PMS1Irvine Univ., USAPham 2000,147 Lee 2006148
KeelePMS1Keele Univ., UKAlford 2000149
KoblenzBroadband SRS1Koblenz Univ., GermanyGeraskin 2005150
NeoBrainCW8Helsinki Univ., FinlandNissila 2002151
PhiladelphiaMultidistance SRS1NIM, Inc., USANelson 2006152
IRIS-3CW1INFM, ItalyGiardini153
TSNIR-3Multidistance SRS1Tsinghua Univ., ChinaTeng 2006154
ZurichPMS1Univ. Hospital Zurich, SwitzerlandBrown 2004155
ImagersArlingtonCW64Univ. of Texas, Arlington, USAKashyap 2007156
BerlinCW22Charité, GermanyBoden 2007157
LondonCW20Univ. College London, UKEverdell 2005158
NIROXCOPE 201CW16BoğaziÇi Univ., TurkeyAkin, 2006159
NanjingCW16Southeast Univ., ChinaLi 2005160
New YorkCWvar.Columbia Univ., USASchmitz 2002161
PhiladelphiaCW16Drexel Univ., USALeon-Carrion 2006162
St. LouisCW300Washington Univ., USACulver 2006163
Zurich1CW16Univ. Hospital Zurich, SwitzerlandMühlemann 200673
BerlinTRS16Physikalisch Technische Bundesaustalt, GermanyLiebert 200682
BostonTRS32Harvard Univ., USASelb 2006164
HamamatsuTRS16Hamamatsu, JapanUeda 2005165
MilanTRS16Politecnico of Milan, ItalyContini 2006166
MonstirTRS32Univ. College London, UKSchmidt 2000167
StrasbourgTRS8Strasbourg Univ., FranceMontcel 2004168
WarsawTRS16Academy of Sciences, PolandLiebert 2005169
HelsinkiPMS16Helsinki Univ., FinlandNissila 2005170
SeoulPMS16Yonsei Univ., South KoreaHo 2007171
HokkaidoSRS64Hokkaido Univ., JapanKek 2006172
IrvineSRSCCDIrvine Univ., USACuccia 200580

1Wearable instrument.

CCD=charge coupled device, the instrument uses a noncontact camera; CW=continuous wave; PMS=phase modulation spectroscopy; SRS=spatially resolved spectroscopy; TRS=time resolved spectroscopy; Univ.=university; var.=variable.

Spatially resolved spectroscopy (SRS) is also called multidistance spectroscopy and is based on light intensity being measured at several different source-detector distances.75, 76 One problem of NIRS and/or NIRI is that the light coupling between the optodes and the tissue is unknown, difficult to measure, and sensitive to changes on the tissue surface over time. SRS techniques assume that the coupling is the same for the different source-detector distances and, by measuring the intensity as a function of the distance, determine a parameter that is independent of the coupling.76 This allows the determination of ratios of O2Hb to total hemoglobin (O2Hb+HHb) and thus tissue oxygen saturation. The application of cerebral NIRS in adults has been hampered by concerns over contamination from extracerebral tissues. Using SRS,77 the brain was identified as the anatomic source of the signal on adult patients undergoing carotid endarterectomy. A change in brain oxygen saturation was predominantly associated with internal carotid artery clamping. The reason is that using a SRS approach, the superficial layers of tissue affect all the light bundles similarly and therefore their influence cancels out. Only deeper tissue layers have an effect on the values.78, 79 Using a single source-detector distance, however, the influence of the superficial tissues on the signals is relatively large. It depends on the source-detector separation. It can be minimized using large separations and a correction for an extracranial sample volume or both.9

The enhanced type of SRS, called spatial frequency domain measurements,80 projects several bar patterns of different distances between bright bars and dark bars on the tissue. This type of imaging is able to determine absolute values.

Time resolved spectroscopy (TRS), also known as time domain spectroscopy, 49, 81, 82, 83, 84, 85 is a technique that measures the time of flight in addition to the light intensity. It does so by emitting a short (100ps) pulse of light into the tissue and measuring the time point spread function of the light after it passes through the tissue. Due to the scattering process, the pulse will broaden and, due to absorption, the intensity will be reduced. The result of such a measurement is a histogram of the number of photons on the y axis and their arrival times on the x axis. The histogram also contains information about the depth of the photonic path, because photons that arrive later have a higher probability to have traveled deeper. The absorption and the reduced scattering coefficients are calculated from the histogram and the absorption coefficients are utilized to calculate the absolute values of the chromophores concentration. This technique is also used for three-dimensional imaging and tomography.14, 85, 86 Thus, from the physicist’s point of view, TRS is an excellent method because it yields a lot of information relatively rapidly and with a high dynamic range. However, it requires sophisticated instrumentation that is so far commercially unavailable. Because the instrumentation usually operates in photon counting mode, it is highly sensitive and can penetrate relatively large tissues (e.g., the head of a neonate). However, due to the low number of photons, TRS measurements are also characterized by a relatively high level of noise. From a clinical point of view, the disadvantages are represented by the physical size of the instrumentation, the use of glass fibers, and the photomultiplier tubes (i.e., the danger of destroying these detectors by excess ambient light). In the near future, technological advances in this field, in particular the miniaturization and reduction in cost of the instrumentation, will promote this technology.

Phase modulation spectroscopy (PMS) is also called intensity modulated or frequency domain spectroscopy. This technique is in principle equivalent to TRS except that it operates in the Fourier domain. This means that the light sources are intensity modulated at radio frequencies ( 50MHz to 1GHz ). After passing through the tissue, the mean intensity (DC), amplitude (AC), and phase of the emerging wave are measured. The phase contains information about the time of flight. To obtain the same information as TRS, PMS requires scanning through all frequencies from 50MHz to 1GHz . 76, 87, 88, 89, 90, 91, 92 The result is a Fourier transform of the time point spread function of TRS. Only a few instruments are operated in scanning mode (also called multifrequency mode) 89, 90, 91, 92 because the time resolution is relatively low. Most of the instruments are single frequency instruments and use a multidistance or SRS geometry.76, 87, 88 It has been shown that the latter type of instrument is technically much simpler than TRS and provides measurements with a good signal-to-noise ratio and a high time resolution. In addition, unlike TRS instruments, SRS instruments can deal with a higher number of photons at the detector and thus with a higher signal-to-noise ratio. From a clinical point of view, the advantages are represented by the easier transportability and the commercial availability. However, compared to TRS, if only one frequency is used, PMS provides less information about the tissue. In addition, from a clinical point of view, the disadvantages are represented by the use of the glass fibers and the sensitivity of the photomultiplier tubes to excess light. In the near future, this technology might profit from technological advances and developments in the mobile communications industry, which lead to the miniaturization, optimization, and dramatic reduction in cost of crucial components such as synthesizers or demodulators.

Broadband imaging, or second differential spectroscopy, 89, 90, 91, 92, 93 means that white light is used instead of discrete wavelengths and, at the detection site, a spectrometer measures the whole range of wavelengths. The advantage is that a whole spectrum is available, which allows the discrimination of chromophores within the tissue with higher accuracy and less crosstalk. Using the second differential, even absolute values can be obtained if a certain water concentration can be assumed.58 One disadvantage of second differential spectroscopy is that taking the derivative magnifies the noise level and thus measurements have a lower signal-to-noise ratio. Some groups also use a combination of broadband and PMS to be absolutely quantitative.92 The disadvantage is that to utilize all wavelengths, the power of the light source needs to be higher and tissue warming may be a dangerous consequence.

Diffusing-wave spectroscopy (DWS), also called diffuse correlation spectroscopy, allows using lasers with a long coherence length and the speckle pattern that is created in the tissue. 94, 95, 96, 97, 98 Speckles, a pattern of bright and dark spots, are a result of the interference of light. This interference occurs when light with large coherence length (laser light) is going through the tissue by different paths, which may lead to constructive or destructive interference. Because in a tissue there is also movement, mainly of the blood, this interference pattern changes in time. The autocorrelation of the speckle pattern contains information about the blood flow. This technique is related to laser Doppler flowmetry, which measures superficial blood flow and is not included in this review. DWS is the fruit of a relatively recent development and the technology is relatively expensive. In the future, efforts for understanding the factors that affect the autocorrelation must be made to completely quantify blood flow.

NIRI, also called diffuse optical imaging (DOI) or topography, reconstructs two-dimensional images of the chromophore concentrations in tissue. The term “diffuse” in DOI refers to the fact that the theory is based on the diffusion approximation. This type of instrumentation operates usually in reflection mode. The resolution of the images achieved today is on the order of 1cm .

Table 3 includes the main commercially available instruments, and Table 4 provides an overview of the most important recent noncommercial prototypes.

Table 3

Main commercial near-infrared clinical instrumentation.

InstrumentTechniqueNumber of channelsCompanyWeb site
PhotometersBOM-L1 TRSingle-distance CW1Omegawave, Japanwww.omegawave.co.jp
HEO-2001, 2Single-distance CW1OMRON, Japann.a.
Micro-RunMan1Single-distance CW1NIM, Inc., USAn.a.
OXYMON MkIIISingle-distance CW1 to 96Artinis, The Netherlandswww.artinis.com
OximetersFORE-SIGHT3Multidistance1Casmed, USAwww.casmed.com
INVOS 5100C3Multidistance2 or 4Somanetics, USAwww.somanetics.com
InSpectra 3253Multidistance1Hutchinson, USAwww.htbiomeasurement.com
NIMOMultidistance1NIROX, Italywww.nirox.it
NIRO-100Multidistance2Hamamatsu, Japanwww.hamamatsu.com
NIRO-200Multidistance2Hamamatsu, Japanwww.hamamatsu.com
O2CBroadband2LEA, Germanywww.lea.de
ODISsey4Multidistance2Vioptix, Inc., USAwww.vioptix.com
OM-220Multidistance2Shimadzu, Japanwww.med.shimadzu.co.jp
OxiplexTSMultidistance PMS1 or 2ISS, USAwww.iss.com
TRS-20Multidistance TRS2Hamamatsu, Japanwww.hamamatsu.com
ImagersDynotCWup to 32NIRx, USAwww.nirx.net
ETG-40003CW44Hitachi, Japanwww.hitachimed.com
ETG-70003CW72Hitachi, Japanwww.hitachimed.com
ImagentPMSup to 128ISS, USAwww.iss.com
LED IMAGERCW16NIM, Inc., USAn.a.
nScan D1200CW16 to 32Arquatis, Switzerlandwww.arquatis.com
nScan W1200Wireless CW16Arquatis, Switzerlandwww.arquatis.com
NIRO-200CW8Hamamatsu, Japanwww.hamamatsu.com
NIRS 4/58CW4 or 58TechEn, Inc, USAwww.nirsoptix.com
OMM-2001CW42Shimadzu, Japanwww.med.shimadzu.co.jp
OMM-3000CW64Shimadzu, Japanwww.med.shimadzu.co.jp

1Wearable instrument.

2No longer commercially available.

3USA Food and Drug Administration’s approval.

430-min battery backup.

CW=continuous wave, n.a.=not available, PMS=phase modulation spectroscopy, SRS=spatially resolved spectroscopy, TRS=time resolved spectroscopy.

4.

Discussion

Table 3 shows that quite a number of oximeters and imagers are commercially available. The presence of three big Japanese companies developing such devices underlines the consistent efforts made by this country in the field of NIRS and NIRI development. Unfortunately, so far, very few instruments have the approval of the American Food and Drug Administration. Therefore, their distribution has been limited to Japan and/or the European Community. Considering the high cost and the restricted clinical applications of the imagers, more oximeters than imagers have been sold particularly for monitoring adult brain oxygenation during heart surgery. It is not possible to report the exact number of the oximeters sold because the companies do not release such figures. However, it is possible to estimate that more than 2000 clinical oximeters are presently operating for different clinical applications.

The development of instrumentation and methodology has been proceeding in steps. At first, only CW instruments with one channel were available. These instruments allowed measurement of relative values only (i.e., changes in chromophore concentration). They provided useful information in many instances, particularly in intervention studies in which, for instance, the safety of drugs was tested (e.g., Ref. 99) or functional brain activity was investigated. In brain studies, absolute values of hemoglobin concentration or blood flow can be obtained using changes in oxygenation. 100, 101, 102, 103 In muscle studies, the combination of relative concentration changes with a venous or arterial occlusion provides absolute quantitation of the oxygenation and blood flow.104, 105, 106 In a second step, instrumentation based on spatially resolved or time resolved (TRS or PMS) methods led to the measurement of absolute values of concentration.76 This considerable evolution enhanced the value of the NIRS measurements, because it allowed the comparison of concentration and oxygen saturation values among patients without any interventions. This paves the way for monitoring patients during treatment (e.g., in intensive care). In a third step, the use of multichannel instruments enhanced the scope of the measurements from single locations to two or three dimensions. This was another big step, because the single location measurements usually assumed that the values at a given location were representative for the whole area or organ. Imaging studies however showed that (1) this assumption is not true and (2) there may be considerable local variability in volume and/or flow and oxygenation. 106, 107, 108, 109

This leaves us with new problems that have to be solved to enhance NIRI advancement, for example, the placement of multiple channels, the handling of large amounts of data, and the algorithms for reconstructing images. However, NIRS and/or NIRI instrument development can be considered constant as witnessed, for instance, by the fact that every 2 to 3 years new models have been replacing the previous ones, particularly as far as oximeters are concerned. Usually, the new models are characterized by lower dimensions, weight, and cost, as well as improved data presentation, software, and precision. In addition, new techniques have been proposed and are under evaluation for improving the quantitation of oximeters (Table 4).

A large effort refers to the development of imaging instrumentation and image reconstruction algorithms. The main problem of imaging relies on the existence of the strong scattering of light in the NIR range and the very low number of light bundles. Most of the commercial imagers are based on CW light sources and are still very bulky and expensive (Table 3). The fact that several prototypes have been developed by industries and academic institutions using TRS and PMS approaches could suggest that these techniques would be utilized by the next generation of commercial clinical imagers (Table 4). But why are there so many different instruments? One reason is that, unlike the other well-established imaging modalities such as magnetic resonance imaging (MRI) or computerized tomography (CT), the setup of NIRS and/or NIRI is highly dependent on the application performed and the tissue measured. Thus, each of the instruments optimizes a certain aspect. For example in neonatology, it is less important to utilize high sensitivity detectors, because neonatal tissue is relatively transparent, and neonatal measurements require soft and flexible probes to prevent lesions of the sensitive skin. An instrument, which optimally incorporates all the physical aspects of the technique (such as highly sensitive detectors) and therefore is capable of providing all the measurable parameters, might be impractical for any kind of clinical application because, for instance, the detector is too sensitive to excess light and could therefore be easily destroyed.

The possibility to map the whole cerebral cortex convinced many cognitive neuroscience research groups to utilize NIRI instrumentation for human brain mapping studies. In this framework, sophisticated data processing methods have recently been investigated and applied to the analysis of NIRI data. Principal component analysis has been utilized for analyzing the spatial and spectral features of diffuse reflectance data from brain tissue110 and for suppressing systemic physiological contributions to the evoked hemoglobin-related signals.111 Independent component analysis112 and the continuous wavelet transform113 have been proposed to detect activated cortical areas, whereas lagged covariance methods have been proposed to explore functional brain connectivity from event-related optical signals.114 In the attempt to characterize the contributions of systemic parameters, such as the heart rate and the mean arterial blood pressure to the low-frequency oscillations in cerebral oxygenation,115 researchers applied information transfer analysis. The recent and quickly growing emphasis placed on data processing procedures for NIRI data shows the importance that the NIRI field is attributing to the development of powerful and reliable data analysis tools. However, no standardized approach for NIRI data analysis has been established yet, laying further emphasis on the development of standard data processing schemes to elevate NIRI into a well-established human cortical imaging modality.116

One measured but not widely explored variable is the light scattering, which is related to tissue structure, cell membranes, and mitochondria. Unfortunately scattering and scattering changes are often disregarded when the focus is on the absorption. One example showing the potential value of scattering changes is their association with the neuronal activity.117 The latter leads to small changes in light scattering at the neuronal level. Because the changes are small, they are difficult to detect. Although several groups report the detection of such changes, 118, 119, 120, 121, 122 there are some controversies.123 New algorithms able to better separate the other physiological signals from the scattering changes might help to resolve this issue. Light scattering has also been investigated in NIR mammography for breast cancer detection.124

The progress of NIRS and/or NIRI is not as rapid as expected and hoped for.22, 125 There are several reasons for this. In fact, NIRS and NIRI have many pitfalls and limitations. Some typical examples can be summarized as follows. (1) For correct measurements, it is necessary to precisely know the assumptions in the physical models and to make sure that they are fulfilled (e.g., the boundary conditions assumed in the algorithms have to correspond to the geometry of the tissue under investigation). (2) An incorrect attachment of the sensor might lead to light piping and consequently large errors. (3) Heterogeneous tissue cannot be measured if the physical model assumes a homogenous tissue. (4) The different NIRS and NIRI approaches show a different degree of susceptibility to movement artifacts, single distance measurements are highly sensitive while multidistance geometries are relatively inert.126 Often these pitfalls lead to errors that in turn are wrongly used to disqualify NIRS and/or NIRI results. A strong interdisciplinary collaboration between clinicians and scientists could facilitate the correct use of NIRS and NIRI. Another explanation for the slow progress is that there is not a unique ideal NIRS and NIRI instrument. Instead, different instruments could be optimal for a given clinical application. Another problem is that the clinical studies for understanding the meaning of a new parameter (such as tissue oxygen saturation), for establishing its normal values, and for determining limits requiring therapeutic or corrective actions (e.g., the administration of oxygen) call for time-consuming, extensive, and very expensive clinical studies.

There is considerable technical progress that, leading to a higher precision of the measurements and resolution of the images, could partly overcome the limitations of the technique. Also from the clinical perspectives there is considerable progress in view of the first clinical applications entering routine. 21, 22, 25, 26, 27, 28, 127 It can be predicted that the evolution of this progress will consist of an increasing variety of clinical applications in which NIRS and/or NIRI will become established techniques in hospitals.

5.

Conclusion

Thirty years after NIRS’s discovery, NIRS and NIRI are currently at a stage of transition from basic clinical research to an adjuvant in clinical applications. On average, two to three papers per day about the clinical applications of NIRS and NIRI are reported on MEDLINE and “Current Contents Connect” (Thomson Scientific, USA). In addition, several technical papers are published in journals not included in MEDLINE. In the next 5 years, additional efforts are expected in technology developments, commercialization, and clinical validation of oximetry and imager instrumentation. In particular, oximeters are expected to become capable of measuring absolute values, and this will give a consistent contribution for the expansion of their clinical applications. Multimodality imaging systems will be developed to integrate NIRI with various other well-established brain imaging techniques such as MRI and positron emission tomography.128 Structural information of brain tissue that is obtained from conventional imaging tools, such as CT, MRI, and ultrasound, will provide highly useful coregistration and guidance that will ultimately improve the accuracy of NIRI image reconstruction. Because NIRS and/or NIRI have an inherently high contrast, technological and computational advances will enable image reconstruction with higher spatial resolution and sensitivity. NIRI techniques show a tremendous potential for noninvasive brain imaging by providing functional and metabolic maps of the activated brain cortex. The complementary information provided by changes in O2Hb and HHb ; the coregistration with electroencephalography and systemic parameters such as the heart rate, blood pressure, and respiratory rate; and the development of dedicated data processing algorithms are critically important for the analysis and interpretation of NIRI data.

In summary, although NIRS and NIRI have been growing slowly but constantly, NIRS and NIRI are on the verge of entering clinical everyday applications and have already brought many valuable insights in clinical research. There are good prospects that NIRS and/or NIRI will light up in the future, shed light on many physiological issues, and brighten the perspectives of many illnesses.

Acknowledgments

The authors, in particular Marco Ferrari, wish to thank Professor Frans Jöbsis for inspiring their scientific research careers. This research was supported in part by PRIN 2005 (VQ, MF) and the county of Zurich, Switzerland (MW). The authors thank Mark Adams for revising the English language. We thank the parents for the consent to publish the picture of their infant, who was not harmed in this process.

References

1. 

F. F. Jöbsis, “Noninvasive, infrared monitoring of cerebral and myocardial oxygen sufficiency and circulatory parameters,” Science, 198 (4323), 1264 –1267 (1977). https://doi.org/10.1126/science.929199 0036-8075 Google Scholar

2. 

F. F. Jöbsis-VanderVliet, “Discovery of the near-infrared window into the body and the early development of near-infrared spectroscopy,” J. Biomed. Opt., 4 (3), 392 –396 (1999). https://doi.org/10.1117/1.429952 1083-3668 Google Scholar

3. 

F. F. Jöbsis-VanderVliet and P. D. Jöbsis, “Biochemical and physiological basis of medical near-infrared spectroscopy,” J. Biomed. Opt., 4 (3), 397 –402 (1999). https://doi.org/10.1117/1.429953 1083-3668 Google Scholar

4. 

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

5. 

Y. Hoshi and M. Tamura, “Dynamic multichannel near-infrared optical imaging of human brain activity,” J. Appl. Physiol., 75 (4), 1842 –1846 (1993). 8750-7587 Google Scholar

6. 

Y. Hoshi and M. Tamura, “Detection of dynamic changes in cerebral oxygenation coupled to neuronal function during mental work in man,” Neurosci. Lett., 150 (1), 5 –8 (1993). https://doi.org/10.1016/0304-3940(93)90094-2 0304-3940 Google Scholar

7. 

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

8. 

J. H. Meek, M. Firbank, C. E. Elwell, J. Atkinson, O. Braddick, and J. S. Wyatt, “Regional hemodynamic responses to visual stimulation in awake infants,” Pediatr. Res., 43 (6), 840 –843 (1998). https://doi.org/10.1203/00006450-199806000-00019 0031-3998 Google Scholar

9. 

M. Ferrari, L. Mottola, and V. Quaresima, “Principles, techniques, and limitations of near infrared spectroscopy,” Can. J. Appl. Physiol., 29 (4), 463 –487 (2004). 1066-7814 Google Scholar

10. 

J. C. Hebden, S. R. Arridge, and D. T. Delpy, “Optical imaging in medicine: I. Experimental techniques,” Phys. Med. Biol., 42 (5), 825 –840 (1997). https://doi.org/10.1088/0031-9155/42/5/007 0031-9155 Google Scholar

11. 

A. P. Gibson, J. C. Hebden, and S. R. Arridge, “Recent advances in diffuse optical imaging,” Phys. Med. Biol., 50 (4), R1 –R43 (2005). https://doi.org/10.1088/0031-9155/50/4/R01 0031-9155 Google Scholar

12. 

S. R. Arridge and J. C. Hebden, “Optical imaging in medicine. II. Modelling and reconstruction,” Phys. Med. Biol., 42 (5), 841 –853 (1997). https://doi.org/10.1088/0031-9155/42/5/008 0031-9155 Google Scholar

13. 

S. R. Arridge, “Optical tomography in medical imaging,” Inverse Probl., 15 (2), R41 –R93 (1999). https://doi.org/10.1088/0266-5611/15/2/022 0266-5611 Google Scholar

14. 

J. C. Hebden, “Advances in optical imaging of the newborn infant brain,” Psychophysiology, 40 (4), 501 –510 (2003). https://doi.org/10.1111/1469-8986.00052 0048-5772 Google Scholar

15. 

Y. Hoshi, “Functional near-infrared optical imaging: utility and limitations in human brain mapping,” Psychophysiology, 40 (4), 511 –520 (2003). https://doi.org/10.1111/1469-8986.00053 0048-5772 Google Scholar

16. 

B. Chance, M. Cope, E. Gratton, N. Ramanujam, and B. Tromberg, “Phase measurement of light absorption and scatter in human tissue,” Rev. Sci. Instrum., 69 (10), 3457 –3481 (1998). https://doi.org/10.1063/1.1149123 0034-6748 Google Scholar

17. 

H. Owen-Reece, M. Smith, C. E. Elwell, and J. C. Goldstone, “Near infrared spectroscopy,” Br. J. Anaesth., 82 (3), 418 –426 (1999). 0007-0912 Google Scholar

18. 

P. Rolfe, “In vivo near-infrared spectroscopy,” Annu. Rev. Biomed. Eng., 2 715 –754 (2000). https://doi.org/10.1146/annurev.bioeng.2.1.715 1523-9829 Google Scholar

19. 

P. L. Madsen and N. H. Secher, “Near-infrared oximetry of the brain,” Prog. Neurobiol., 58 (6), 541 –560 (1999). 0301-0082 Google Scholar

20. 

H. L. Edmonds Jr., B. L. Ganzel, E. H. Austin 3rd, “Cerebral oximetry for cardiac and vascular surgery,” Semin. Cardiothorac. Vasc. Anesth., 8 (2), 147 –166 (2004). Google Scholar

21. 

J. D. Tobias, “Cerebral oxygenation monitoring: near-infrared spectroscopy,” Expert Rev. Med. Devices, 3 (2), 235 –243 (2006). Google Scholar

22. 

G. Greisen, “Is near-infrared spectroscopy living up to its promises?,” Semin. Fetal Neonatal Med., 11 (6), 498 –502 (2006). Google Scholar

23. 

K. R. Ward, R. R. Ivatury, R. W. Barbee, J. Terner, R. Pittman, I. P. Filho, and B. Spiess, “Near infrared spectroscopy for evaluation of the trauma patient: a technology review,” Resuscitation, 68 (1), 27 –44 (2006). 0300-9572 Google Scholar

24. 

P. G. Al-Rawi, “Near infrared spectroscopy in brain injury: Today’s perspective,” Acta Neurochir. Suppl. (Wien), 95 453 –457 (2005). 0065-1419 Google Scholar

25. 

A. J. Wolfberg and A. J. du Plessis, “Near-infrared spectroscopy in the fetus and neonate,” Clin. Perinatol., 33 (3), viii707 –728 (2006). 0095-5108 Google Scholar

26. 

G. M. Hoffman, “Pro: Near-infrared spectroscopy should be used for all cardiopulmonary bypass,” J. Cardiothorac Vasc. Anesth., 20 (4), 606 –612 (2006). 1053-0770 Google Scholar

27. 

A. Casati, E. Spreafico, M. Putzu, and G. Fanelli, “New technology for noninvasive brain monitoring: continuous cerebral oximetry,” Minerva Anestesiol., 72 (7–8), 605 –625 (2006). Google Scholar

28. 

M. C. Taillefer and A. Y. Denault, “Cerebral near-infrared spectroscopy in adult heart surgery: Systematic review of its clinical efficacy,” Can. J. Anaesth., 52 (1), 79 –87 (2005). 0832-610X Google Scholar

29. 

P. G. Al-Rawi and P. J. Kirkpatrick, “Tissue oxygen index: Thresholds for cerebral ischemia using near-infrared spectroscopy,” Stud. Cercet Endocrinol., 37 (11), 2720 –2725 (2006). 0039-3924 Google Scholar

30. 

D. A. Boas, A. M. Dale, and M. A. Franceschini, “Diffuse optical imaging of brain activation: Approaches to optimizing image sensitivity, resolution, and accuracy,” Neuroimage, 23 (Suppl. 1), S275 –S288 (2004). 1053-8119 Google Scholar

31. 

E. Gratton, V. Toronov, U. Wolf, M. Wolf, and A. Webb, “Measurement of brain activity by near-infrared light,” J. Biomed. Opt., 10 (1), 11008 (2005). 1083-3668 Google Scholar

32. 

H. Obrig and A. Villringer, “Beyond the visible—Imaging the human brain with light,” J. Cereb. Blood Flow Metab., 23 (1), 1 –18 (2003). https://doi.org/10.1097/00004647-200301000-00001 0271-678X Google Scholar

33. 

J. Steinbrink, A. Villringer, F. Kempf, D. Haux, S. Boden, and H. Obrig, “Illuminating the BOLD signal: Combined fMRI-fNIRS studies,” Magn. Reson. Imaging, 24 (4), 495 –505 (2006). https://doi.org/10.1016/j.mri.2005.12.034 0730-725X Google Scholar

34. 

Y. Hoshi, “Functional near-infrared spectroscopy: Potential and limitations in neuroimaging studies,” Int. Rev. Neurobiol., 66 237 –266 (2005). 0074-7742 Google Scholar

35. 

G. Strangman, D. A. Boas, and J. P. Sutton, “Non-invasive neuroimaging using near-infrared light,” Biol. Psychiatry, 52 (7), 679 –693 (2002). 0006-3223 Google Scholar

36. 

S. C. Bunce, M. Izzetoglu, K. Izzetoglu, B. Onaral, and K. Pourrezaei, “Functional near-infrared spectroscopy,” IEEE Eng. Med. Biol. Mag., 25 (4), 54 –62 (2006). 0739-5175 Google Scholar

37. 

R. Boushel, H. Langberg, J. Olesen, J. Gonzales-Alonzo, J. Bulow, and M. Kjaer, “Monitoring tissue oxygen availability with near infrared spectroscopy (NIRS) in health and disease,” Scand. J. Med. Sci. Sports, 11 (4), 213 –222 (2001). 0905-7188 Google Scholar

38. 

R. Boushel and C. A. Piantadosi, “Near-infrared spectroscopy for monitoring muscle oxygenation,” Acta Physiol. Scand., 168 (4), 615 –622 (2000). https://doi.org/10.1046/j.1365-201x.2000.00713.x 0001-6772 Google Scholar

39. 

V. Quaresima, R. Lepanto, and M. Ferrari, “The use of near infrared spectroscopy in sports medicine,” J. Sports Med. Phys. Fitness, 43 (1), 1 –13 (2003). 0022-4707 Google Scholar

40. 

M. Ferrari, T. Binzoni, and V. Quaresima, “Oxidative metabolism in muscle,” Philos. Trans. R. Soc. London, Ser. B, 352 (1354), 677 –683 (1997). https://doi.org/10.1098/rstb.1997.0049 0962-8436 Google Scholar

41. 

K. K. McCully and T. Hamaoka, “Near-infrared spectroscopy: What can it tell us about oxygen saturation in skeletal muscle?,” Exerc Sport Sci. Rev., 28 (3), 123 –127 (2000). 0091-6331 Google Scholar

42. 

Y. N. Bhambhani, “Muscle oxygenation trends during dynamic exercise measured by near infrared spectroscopy,” Can. J. Appl. Physiol., 29 (4), 504 –523 (2004). 1066-7814 Google Scholar

43. 

J. P. Neary, “Application of near infrared spectroscopy to exercise sports science,” Can. J. Appl. Physiol., 29 (4), 488 –503 (2004). 1066-7814 Google Scholar

44. 

S. Fantini and P. Taroni, “Optical mammography,” Cancer Imaging: Lung and Breast Carcinomas, 449 –458 Elsevier, New York (2007). Google Scholar

45. 

R. X. Xu and S. P. Povoski, “Diffuse optical imaging and spectroscopy for cancer,” Expert Rev. Med. Devices, 4 (1), 83 –95 (2007). Google Scholar

46. 

D. R. Leff, O. J. Warren, L. C. Enfield, A. Gibson, T. Athanasiou, D. K. Patten, J. Hebden, G. Z. Yang, and A. Darzi, “Diffuse optical imaging of the healthy and diseased breast: A systematic review,” Breast Cancer Res. Treat., 0167-6806 Google Scholar

47. 

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

48. 

D. T. Delpy, S. R. Arridge, M. Cope, D. Edwards, E. O. Reynolds, C. E. Richardson, S. Wray, J. Wyatt, and P. van der Zee, “Quantitation of pathlength in optical spectroscopy,” Adv. Exp. Med. Biol., 248 41 –46 (1989). 0065-2598 Google Scholar

49. 

D. T. Delpy, M. Cope, P. van der Zee, S. Arridge, S. Wray, and J. Wyatt, “Estimation of optical pathlength through tissue from direct time of flight measurement,” Phys. Med. Biol., 33 (12), 1433 –1442 (1988). https://doi.org/10.1088/0031-9155/33/12/008 0031-9155 Google Scholar

50. 

S. Wray, M. Cope, D. T. Delpy, J. S. Wyatt, and E. O. Reynolds, “Characterization of the near infrared absorption spectra of cytochrome aa3 and haemoglobin for the non-invasive monitoring of cerebral oxygenation,” Biochim. Biophys. Acta, 933 (1), 184 –192 (1988). https://doi.org/10.1016/0005-2728(88)90069-2 0006-3002 Google Scholar

51. 

M. A. O’Leary, “Imaging with diffuse photon density waves,” University of Pennsylvania, (1996). Google Scholar

52. 

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

53. 

S. Fantini, M. A. Franceschini, and E. 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). 0740-3224 Google Scholar

54. 

A. Duncan, J. H. Meek, M. Clemence, C. E. Elwell, L. Tyszczuk, M. Cope, and D. T. Delpy, “Optical pathlength measurements on adult head, calf and forearm and the head of the newborn infant using phase resolved optical spectroscopy,” Phys. Med. Biol., 40 (2), 295 –304 (1995). https://doi.org/10.1088/0031-9155/40/2/007 0031-9155 Google Scholar

55. 

A. Duncan, J. H. Meek, M. Clemence, C. E. Elwell, P. Fallon, L. Tyszczuk, M. Cope, and D. T. Delpy, “Measurement of cranial optical path length as a function of age using phase resolved near infrared spectroscopy,” Pediatr. Res., 39 (5), 889 –894 (1996). https://doi.org/10.1203/00006450-199605000-00025 0031-3998 Google Scholar

56. 

M. Essenpreis, M. Cope, C. E. Elwell, S. R. Arridge, P. van der Zee, and D. T. Delpy, “Wavelength dependence of the differential pathlength factor and the log slope in time-resolved tissue spectroscopy,” Adv. Exp. Med. Biol., 333 9 –20 (1993). 0065-2598 Google Scholar

57. 

S. Fantini, D. Hueber, M. A. Franceschini, E. Gratton, W. Rosenfeld, P. G. Stubblefield, D. Maulik, and M. R. Stankovic, “Non-invasive optical monitoring of the newborn piglet brain using continuous-wave and frequency-domain spectroscopy,” Phys. Med. Biol., 44 (6), 1543 –1563 (1999). https://doi.org/10.1088/0031-9155/44/6/308 0031-9155 Google Scholar

58. 

S. J. Matcher, M. Cope, and D. T. Delpy, “Use of the water absorption spectrum to quantify tissue chromophore concentration changes in near-infrared spectroscopy,” Phys. Med. Biol., 39 (1), 177 –196 (1994). https://doi.org/10.1088/0031-9155/39/1/011 0031-9155 Google Scholar

59. 

J. S. Wyatt, M. Cope, D. T. Delpy, P. van der Zee, S. Arridge, A. D. Edwards, and E. O. Reynolds, “Measurement of optical path length for cerebral near-infrared spectroscopy in newborn infants,” Drug Metab. Dispos., 12 (2), 140 –144 (1990). 0090-9556 Google Scholar

60. 

C. E. Cooper, M. Cope, V. Quaresima, M. Ferrari, E. Nemoto, R. Springett, S. Matcher, P. Amess, J. Penrice, L. Tyszczuk, J. Wyatt, and D. T. Delpy, “Measurement of cytochrome oxidase redox state by near infrared spectroscopy,” Adv. Exp. Med. Biol., 413 63 –73 (1997). 0065-2598 Google Scholar

61. 

V. Quaresima, R. Springett, M. Cope, J. T. Wyatt, D. T. Delpy, M. Ferrari, and C. E. Cooper, “Oxidation and reduction of cytochrome oxidase in the neonatal brain observed by in vivo near-infrared spectroscopy,” Biochim. Biophys. Acta, 1366 (3), 291 –300 (1998). https://doi.org/10.1016/S0005-2728(98)00129-7 0006-3002 Google Scholar

62. 

R. Springett, J. Newman, M. Cope, and D. T. Delpy, “Oxygen dependency and precision of cytochrome oxidase signal from full spectral NIRS of the piglet brain,” Am. J. Physiol. Heart Circ. Physiol., 279 (5), H2202 –H2209 (2000). 0363-6135 Google Scholar

63. 

K. Uludag, J. Steinbrink, M. Kohl-Bareis, R. Wenzel, A. Villringer, and H. Obrig, “Cytochrome-c-oxidase redox changes during visual stimulation measured by near-infrared spectroscopy cannot be explained by a mere cross talk artefact,” Neuroimage, 22 (1), 109 –119 (2004). 1053-8119 Google Scholar

64. 

M. M. Tisdall, I. Tachtsidis, T. S. Leung, C. E. Elwell, and M. Smith, “Near-infrared spectroscopic quantification of changes in the concentration of oxidized cytochrome c oxidase in the healthy human brain during hypoxemia,” J. Biomed. Opt., 12 (2), 024002 (2007). https://doi.org/10.1117/1.2718541 1083-3668 Google Scholar

65. 

M. A. Franceschini, S. Fantini, J. H. Thompson, J. P. Culver, and D. A. Boas, “Hemodynamic evoked response of the sensorimotor cortex measured noninvasively with near-infrared optical imaging,” Psychophysiology, 40 (4), 548 –560 (2003). https://doi.org/10.1111/1469-8986.00057 0048-5772 Google Scholar

66. 

M. Wolf, U. Wolf, V. Toronov, A. Michalos, L. A. Paunescu, J. H. Choi, and E. Gratton, “Different time evolution of oxyhemoglobin and deoxyhemoglobin concentration changes in the visual and motor cortices during functional stimulation: A near-infrared spectroscopy study,” Neuroimage, 16 (3 Pt. 1), 704 –712 (2002). 1053-8119 Google Scholar

67. 

V. Toronov, M. A. Franceschini, M. Filiaci, S. Fantini, M. Wolf, A. Michalos, and E. Gratton, “Near-infrared study of fluctuations in cerebral hemodynamics during rest and motor stimulation: temporal analysis and spatial mapping,” Med. Phys., 27 (4), 801 –815 (2000). https://doi.org/10.1118/1.598943 0094-2405 Google Scholar

68. 

V. Toronov, A. Webb, J. H. Choi, M. Wolf, A. Michalos, E. Gratton, and D. Hueber, “Investigation of human brain hemodynamics by simultaneous near-infrared spectroscopy and functional magnetic resonance imaging,” Med. Phys., 28 (4), 521 –527 (2001). https://doi.org/10.1118/1.1354627 0094-2405 Google Scholar

69. 

H. R. Heekeren, R. Wenzel, H. Obrig, J. Ruben, J. P. Ndayisaba, Q. Luo, A. Dale, S. Nioka, M. Kohl, U. Dirnagl, A. Villringer, and B. Chance, “Towards noninvasive optical human brain mapping improvements of the spectral, temporal and spatial resolution of near-infrared spectroscopy,” Proc. SPIE, 2979 847 –857 (1997). https://doi.org/10.1117/12.280233 0277-786X Google Scholar

70. 

M. Tamura, Y. Hoshi, and F. Okada, “Localized near-infrared spectroscopy and functional optical imaging of brain activity,” Philos. Trans. R. Soc. London, Ser. B, 352 (1354), 737 –742 (1997). https://doi.org/10.1098/rstb.1997.0056 0962-8436 Google Scholar

71. 

M. A. Franceschini, A. Zourabian, J. B. Moore, A. Arora, S. Fantini, and D. A. Boas, “Local measurement of venous saturation in tissue with non-invasive, near-infrared respiratory-oximetry,” Proc. SPIE, 4250 164 –170 (2001). https://doi.org/10.1117/12.434535 0277-786X Google Scholar

72. 

M. Wolf, G. Duc, M. Keel, P. Niederer, K. von Siebenthal, and H. U. Bucher, “Continuous noninvasive measurement of cerebral arterial and venous oxygen saturation at the bedside in mechanically ventilated neonates,” Crit. Care Med., 25 (9), 1579 –1582 (1997). 0090-3493 Google Scholar

73. 

T. Mühlemann, D. Haensse, and M. Wolf, “Ein drahtloser sensor für die bildgebende in-vivo nahinfrarotspektroskopie,” P60 (2006). Google Scholar

74. 

A. Villringer and B. Chance, “Non-invasive optical spectroscopy and imaging of human brain function,” Trends Neurosci., 20 (10), 435 –442 (1997). https://doi.org/10.1016/S0166-2236(97)01132-6 0166-2236 Google Scholar

75. 

S. Matcher, P. Kirkpatrick, K. Nahid, M. Cope, and D. T. Delpy, “Absolute quantification methods in tissue near infrared spectroscopy,” Proc. SPIE, 2389 486 –495 (1995). https://doi.org/10.1117/12.209997 0277-786X Google Scholar

76. 

S. Fantini, M. A. Franceschini, J. S. Maier, S. A. Walker, B. Barbieri, and E. Gratton, “Frequency-domain multichannel optical detector for noninvasive tissue spectroscopy and oximetry,” Opt. Eng., 34 32 –42 (1995). https://doi.org/10.1117/12.183988 0091-3286 Google Scholar

77. 

P. G. Al-Rawi, P. Smielewski, and P. J. Kirkpatrick, “Evaluation of a near-infrared spectrometer (NIRO 300) for the detection of intracranial oxygenation changes in the adult head,” Stroke, 32 (11), 2492 –2500 (2001). https://doi.org/10.1161/hs1101.098356 0039-2499 Google Scholar

78. 

J. Choi, M. Wolf, V. Toronov, U. Wolf, C. Polzonetti, D. Hueber, L. P. Safonova, R. Gupta, A. Michalos, W. Mantulin, and E. Gratton, “Noninvasive determination of the optical properties of adult brain: Near-infrared spectroscopy approach,” J. Biomed. Opt., 9 (1), 221 –229 (2004). https://doi.org/10.1117/1.1628242 1083-3668 Google Scholar

79. 

M. A. Franceschini, S. Fantini, L. A. Paunescu, J. S. Maier, and E. Gratton, “Influence of a superficial layer in the quantitative spectroscopic study of strongly scattering media,” Appl. Opt., 37 (31), 7447 –7458 (1998). 0003-6935 Google Scholar

80. 

D. J. Cuccia, F. Bevilacqua, A. J. Durkin, and B. J. Tromberg, “Modulated imaging: Quantitative analysis and tomography of turbid media in the spatial-frequency domain,” Opt. Lett., 30 (11), 1354 –1356 (2005). 0146-9592 Google Scholar

81. 

A. Liebert, H. Wabnitz, D. Grosenick, M. Moller, R. Macdonald, and H. Rinneberg, “Evaluation of optical properties of highly scattering media by moments of distributions of times of flight of photons,” Appl. Opt., 42 (28), 5785 –5792 (2003). https://doi.org/10.1364/AO.42.005785 0003-6935 Google Scholar

82. 

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

83. 

R. Cubeddu, A. Pifferi, P. Taroni, A. Torricelli, and G. Valentini, “Compact tissue oximeter based on dual-wavelength multichannel time-resolved reflectance,” Appl. Opt., 38 (16), 3670 –3680 (1999). 0003-6935 Google Scholar

84. 

A. Torricelli, A. Pifferi, L. Spinelli, P. Taroni, V. Quaresima, M. Ferrari, and R. Cubeddu, “Multi-channel time-resolved tissue oximeter for functional imaging of the brain,” 1980 –1983 (2004) Google Scholar

85. 

A. Torricelli, A. Pifferi, P. Taroni, C. D’Andrea, and R. Cubeddu, “In vivo multidistance multiwavelength time-resolved reflectance spectroscopy of layered tissues,” Proc. SPIE, 4250 290 –295 (2001). https://doi.org/10.1117/12.434500 0277-786X Google Scholar

86. 

J. C. Hebden, “Optical tomography: Development of a new medical imaging modality,” AIP Conf. Proc., 440 79 –90 (1998). 0094-243X Google Scholar

87. 

J. B. Fishkin, P., T. C. So, A. E. Cerussi, S. Fantini, M. A. Franceschini, and E. Gratton, “Frequency-domain method for measuring spectral properties in multiple-scattering media: Methemoglobin absorption spectrum in a tissuelike phantom,” Appl. Opt., 34 (7), 1143 –1155 (1995). 0003-6935 Google Scholar

88. 

E. Gratton, S. Fantini, M. A. Franceschini, G. Gratton, and M. Fabiani, “Measurements of scattering and absorption changes in muscle and brain,” Philos. Trans. R. Soc. London, Ser. B, 352 (1354), 727 –735 (1997). https://doi.org/10.1098/rstb.1997.0055 0962-8436 Google Scholar

89. 

F. Bevilacqua, A. J. Berger, A. E. Cerussi, D. Jakubowski, and B. J. Tromberg, “Broadband absorption spectroscopy in turbid media by combined frequency-domain and steady-state methods,” Appl. Opt., 39 (34), 6498 –6507 (2000). 0003-6935 Google Scholar

90. 

T. H. Pham, F. Bevilacqua, T. Spott, J. S. Dam, B. J. Tromberg, and S. Andersson-Engels, “Quantifying the absorption and reduced scattering coefficients of tissuelike turbid media over a broad spectral range with noncontact Fourier-transform hyperspectral imaging,” Appl. Opt., 39 (34), 6487 –6497 (2000). 0003-6935 Google Scholar

91. 

A. Cerussi, R. Van Woerkom, F. Waffarn, and B. Tromberg, “Noninvasive monitoring of red blood cell transfusion in very low birthweight infants using diffuse optical spectroscopy,” J. Biomed. Opt., 10 (5), 51401 (2005). 1083-3668 Google Scholar

92. 

T. H. Pham, O. Coquoz, J. B. Fishkin, E. Anderson, and B. J. Tromberg, “Broad bandwidth frequency domain instrument for quantitative tissue optical spectroscopy,” Rev. Sci. Instrum., 71 (6), 2500 –2513 (2000). https://doi.org/10.1063/1.1150665 0034-6748 Google Scholar

93. 

K. Tanner, E. D’Amico, A. Kaczmarowski, S. Kukreti, J. Malpeli, W. W. Mantulin, and E. Gratton, “Spectrally resolved neurophotonics: A case report of hemodynamics and vascular components in the mammalian brain,” J. Biomed. Opt., 10 (6), 64009 (2005). 1083-3668 Google Scholar

94. 

J. Li, G. Dietsche, D. Iftime, S. E. Skipetrov, G. Maret, T. Elbert, B. Rockstroh, and T. Gisler, “Noninvasive detection of functional brain activity with near-infrared diffusing-wave spectroscopy,” J. Biomed. Opt., 10 (4), 44002 (2005). 1083-3668 Google Scholar

95. 

T. Durduran, Y. Guoqiang, Z. Chao, G. Lech, B. Chance, and A. G. Yodh, “Quantification of muscle oxygenation and flow of healthy volunteers during cuff occlusion of arm and leg flexor muscles and plantar flexion exercise,” Proc. SPIE, 4955 (1), 447 –453 (2003). https://doi.org/10.1117/12.476880 0277-786X Google Scholar

96. 

T. Durduran, G. Yu, M. G. Burnett, J. A. Detre, J. H. Greenberg, J. Wang, C. Zhou, and A. G. Yodh, “Diffuse optical measurement of blood flow, blood oxygenation, and metabolism in a human brain during sensorimotor cortex activation,” Opt. Lett., 29 (15), 1766 –1768 (2004). https://doi.org/10.1364/OL.29.001766 0146-9592 Google Scholar

97. 

T. Durduran, M. G. Burnett, G. Yu, C. Zhou, D. Furuya, A. G. Yodh, J. A. Detre, and J. H. Greenberg, “Spatiotemporal quantification of cerebral blood flow during functional activation in rat somatosensory cortex using laser-speckle flowmetry,” J. Cereb. Blood Flow Metab., 24 (5), 518 –525 (2004). https://doi.org/10.1097/00004647-200405000-00005 0271-678X Google Scholar

98. 

Y. Guoqiang, T. Durduran, G. Lech, Z. Chao, B. Chance, E. R. Mohler, and A. G. Yodh, “Time-dependent blood flow and oxygenation in human skeletal muscles measured with noninvasive near-infrared diffuse optical spectroscopies,” J. Biomed. Opt., 10 (2), 24027 (2005). 1083-3668 Google Scholar

99. 

A. D. Edwards, J. S. Wyatt, C. Richardson, A. Potter, M. Cope, D. T. Delpy, and E. O. Reynolds, “Effects of indomethacin on cerebral haemodynamics in very preterm infants,” Lancet, 335 (8704), 1491 –1495 (1990). https://doi.org/10.1016/0140-6736(90)93030-S 0140-6736 Google Scholar

100. 

A. D. Edwards, J. S. Wyatt, C. Richardson, D. T. Delpy, M. Cope, and E. O. Reynolds, “Cotside measurement of cerebral blood flow in ill newborn infants by near infrared spectroscopy,” Lancet, 2 (8614), 770 –771 (1988). https://doi.org/10.1016/S0140-6736(88)92418-X 0140-6736 Google Scholar

101. 

J. S. Wyatt, M. Cope, D. T. Delpy, C. E. Richardson, A. D. Edwards, S. Wray, and E. O. Reynolds, “Quantitation of cerebral blood volume in human infants by near-infrared spectroscopy,” J. Appl. Physiol., 68 (3), 1086 –1091 (1990). 8750-7587 Google Scholar

102. 

M. Wolf, N. Brun, G. Greisen, M. Keel, K. von Siebenthal, and H. Bucher, “Optimising the methodology of calculating the cerebral blood flow of newborn infants from near infra-red spectrophotometry data,” Med. Biol. Eng. Comput., 34 (3), 221 –226 (1996). https://doi.org/10.1007/BF02520077 0140-0118 Google Scholar

103. 

M. Wolf, H. U. Bucher, V. Dietz, M. Keel, K. von Siebenthal, and G. Duc, “How to evaluate slow oxygenation changes to estimate absolute cerebral haemoglobin concentration by near infrared spectrophotometry in neonates,” Adv. Exp. Med. Biol., 411 495 –501 (1997). 0065-2598 Google Scholar

104. 

R. A. De Blasi, M. Cope, C. Elwell, F. Safoue, and M. Ferrari, “Noninvasive measurement of human forearm oxygen consumption by near infrared spectroscopy,” Eur. J. Appl. Physiol., 67 (1), 20 –25 (1993). https://doi.org/10.1007/BF00377698 0301-5548 Google Scholar

105. 

R. A. De Blasi, M. Ferrari, A. Natali, G. Conti, A. Mega, and A. Gasparetto, “Noninvasive measurement of forearm blood flow and oxygen consumption by near-infrared spectroscopy,” J. Appl. Physiol., 76 (3), 1388 –1393 (1994). 8750-7587 Google Scholar

106. 

U. Wolf, M. Wolf, J. H. Choi, M. Levi, D. Choudhury, S. Hull, D. Coussirat, L. A. Paunescu, L. P. Safonova, A. Michalos, W. W. Mantulin, and E. Gratton, “Localized irregularities in hemoglobin flow and oxygenation in calf muscle in patients with peripheral vascular disease detected with near-infrared spectrophotometry,” J. Vasc. Surg., 37 (5), 1017 –1026 (2003). 0741-5214 Google Scholar

107. 

U. Wolf, M. Wolf, J. H. Choi, L. A. Paunescu, L. P. Safonova, A. Michalos, and E. Gratton, “Mapping of hemodynamics on the human calf with near infrared spectroscopy and the influence of the adipose tissue thickness,” Adv. Exp. Med. Biol., 510 225 –230 (2003). 0065-2598 Google Scholar

108. 

V. Quaresima, W. N. Colier, M. van der Sluijs, and M. Ferrari, “Nonuniform quadriceps O2 consumption revealed by near infrared multipoint measurements,” Biochem. Biophys. Res. Commun., 285 (4), 1034 –1039 (2001). https://doi.org/10.1006/bbrc.2001.5292 0006-291X Google Scholar

109. 

V. Quaresima, M. Ferrari, M. A. Franceschini, M. L. Hoimes, and S. Fantini, “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). https://doi.org/10.1117/1.1646417 1083-3668 Google Scholar

110. 

K. Yokoyama, M. Watanabe, Y. Watanbe, and E. Okada, “Interpretation of principal components of the reflectance spectra obtained from multispectral images of exposed pig brain,” J. Biomed. Opt., 10 (1), 11005 (2005). 1083-3668 Google Scholar

111. 

Y. Zhang, D. H. Brooks, M. A. Franceschini, and D. A. Boas, “Eigenvector-based spatial filtering for reduction of physiological interference in diffuse optical imaging,” J. Biomed. Opt., 10 (1), 11014 (2005). 1083-3668 Google Scholar

112. 

G. Morren, U. Wolf, P. Lemmerling, M. Wolf, J. H. Choi, E. Gratton, L. De Lathauwer, and S. Van Huffel, “Detection of fast neuronal signals in the motor cortex from functional near infrared spectroscopy measurements using independent component analysis,” Med. Biol. Eng. Comput., 42 (1), 92 –99 (2004). https://doi.org/10.1007/BF02351016 0140-0118 Google Scholar

113. 

Y. D. Liu, G. H. Zang, F. Y. Liu, L. R. Yan, M. Li, Z. T. Zhou, and D. W. Hu, “Spatial and temporal analysis for optical imaging data using CWT and tICA,” Lect. Notes Comput. Sci., 3765 508 –516 (2005). 0302-9743 Google Scholar

114. 

E. Rykhlevskaia, M. Fabiani, and G. Gratton, “Lagged covariance structure models for studying functional connectivity in the brain,” Neuroimage, 30 (4), 1203 –1218 (2006). 1053-8119 Google Scholar

115. 

T. Katura, N. Tanaka, A. Obata, H. Sato, and A. Maki, “Quantitative evaluation of interrelations between spontaneous low-frequency oscillations in cerebral hemodynamics and systemic cardiovascular dynamics,” Neuroimage, 31 (4), 1592 –1600 (2006). 1053-8119 Google Scholar

116. 

M. L. Schroeter, M. M. Bucheler, K. Muller, K. Uludag, H. Obrig, G. Lohmann, M. Tittgemeyer, A. Villringer, and D. Y. von Cramon, “Towards a standard analysis for functional near-infrared imaging,” Neuroimage, 21 (1), 283 –290 (2004). 1053-8119 Google Scholar

117. 

R. A. Stepnoski, A. LaPorta, F. Raccuia-Behling, G. E. Blonder, R. E. Slusher, and D. Kleinfeld, “Noninvasive detection of changes in membrane potential in cultured neurons by light scattering,” Proc. Natl. Acad. Sci. U.S.A., 88 (21), 9382 –9386 (1991). https://doi.org/10.1073/pnas.88.21.9382 0027-8424 Google Scholar

118. 

G. Gratton, C. R. Brumback, B. A. Gordon, M. A. Pearson, K. A. Low, and M. Fabiani, “Effects of measurement method, wavelength, and source-detector distance on the fast optical signal,” Neuroimage, 32 (4), 1576 –1590 (2006). 1053-8119 Google Scholar

119. 

M. A. Franceschini and D. A. Boas, “Noninvasive measurement of neuronal activity with near-infrared optical imaging,” Neuroimage, 21 (1), 372 –386 (2004). 1053-8119 Google Scholar

120. 

M. Wolf, U. Wolf, J. H. Choi, R. Gupta, L. P. Safonova, L. A. Paunescu, A. Michalos, and E. Gratton, “Functional frequency-domain near-infrared spectroscopy detects fast neuronal signal in the motor cortex,” Neuroimage, 17 (4), 1868 –1875 (2002). 1053-8119 Google Scholar

121. 

M. Wolf, U. Wolf, J. H. Choi, V. Toronov, L. A. Paunescu, A. Michalos, and E. Gratton, “Fast cerebral functional signal in the 100-ms range detected in the visual cortex by frequency-domain near-infrared spectrophotometry,” Psychophysiology, 40 (4), 521 –528 (2003). https://doi.org/10.1111/1469-8986.00054 0048-5772 Google Scholar

122. 

J. Steinbrink, M. Kohl, H. Obrig, G. Curio, F. Syre, F. Thomas, H. Wabnitz, H. Rinneberg, and A. Villringer, “Somatosensory evoked fast optical intensity changes detected non-invasively in the adult human head,” Neurosci. Lett., 291 (2), 105 –108 (2000). https://doi.org/10.1016/S0304-3940(00)01395-1 0304-3940 Google Scholar

123. 

J. Steinbrink, F. C. Kempf, A. Villringer, and H. Obrig, “The fast optical signal—robust or elusive when non-invasively measured in the human adult?,” Neuroimage, 26 (4), 996 –1008 (2005). 1053-8119 Google Scholar

124. 

N. Shah, A. E. Cerussi, D. Jakubowski, D. Hsiang, J. Butler, and B. J. Tromberg, “Spatial variations in optical and physiological properties of healthy breast tissue,” J. Biomed. Opt., 9 (3), 534 –540 (2004). https://doi.org/10.1117/1.1695560 1083-3668 Google Scholar

125. 

S. E. Nicklin, I. A. Hassan, Y. A. Wickramasinghe, and S. A. Spencer, “The light still shines, but not that brightly? The current status of perinatal near infrared spectroscopy,” Arch. Dis. Child Fetal Neonatal Ed., 88 (4), F263 –F268 (2003). 1359-2998 Google Scholar

126. 

M. A. Franceschini, D. Wallace, B. Barbieri, S. Fantini, W. W. Mantulin, S. Pratesi, G. P. Donzelli, and E. Gratton, “Optical study of the skeletal muscle during exercise with a second generation frequency-domain tissue oximeter,” Proc. SPIE, 2979 807 –814 (1997). https://doi.org/10.1117/12.280229 0277-786X Google Scholar

127. 

E. Keller, A. Nadler, H. Alkadhi, S. S. Kollias, Y. Yonekawa, and P. Niederer, “Noninvasive measurement of regional cerebral blood flow and regional cerebral blood volume by near-infrared spectroscopy and indocyanine green dye dilution,” Neuroimage, 20 (2), 828 –839 (2003). 1053-8119 Google Scholar

128. 

J. C. Gore, S. G. Horovitz, C. J. Cannistraci, and P. Skudlarski, “Integration of fMRI, NIROT and ERP for studies of human brain function,” Magn. Reson. Imaging, 24 (4), 507 –513 (2006). https://doi.org/10.1016/j.mri.2005.12.039 0730-725X Google Scholar

129. 

D. T. Delpy and M. Cope, “Quantification in tissue near-infrared spectroscopy,” Philos. Trans. R. Soc. London, Ser. B, 352 649 –659 (1997). https://doi.org/10.1098/rstb.1997.0046 0962-8436 Google Scholar

130. 

B. Grassi, V. Quaresima, C. Marconi, M. Ferrari, and P. Cerretelli, “Blood lactate accumulation and muscle deoxygenation during incremental exercise,” J. Appl. Physiol., 87 (1), 348 –355 (1999). 8750-7587 Google Scholar

131. 

V. Quaresima, T. Komiyama, and M. Ferrari, “Differences in oxygen re-saturation of thigh and calf muscles after two treadmill stress tests,” Zentralbl Bakteriol Mikrobiol. Hyg., Abt. 1, Orig. B, 132 (1), 67 –73 (2002). 0174-3015 Google Scholar

132. 

M. Oda, Y. Yamashita, G. Nishimura, and M. Tamura, “A simple and novel algorithm for time-resolved multiwavelength oximetry,” Phys. Med. Biol., 41 (3), 551 –562 (1996). https://doi.org/10.1088/0031-9155/41/3/015 0031-9155 Google Scholar

133. 

P. B. Benni, B. Chen, F. D. Dykes, S. F. Wagoner, M. Heard, A. J. Tanner, T. L. Young, K. Rais-Bahrami, O. Rivera, and B. L. Short, “Validation of the CAS neonatal NIRS system by monitoring vv-ECMO patients: Preliminary results,” Adv. Exp. Med. Biol., 566 195 –201 (2005). 0065-2598 Google Scholar

134. 

C. E. Cooper, C. E. Elwell, J. H. Meek, S. J. Matcher, J. S. Wyatt, M. Cope, and D. T. Delpy, “The noninvasive measurement of absolute cerebral deoxyhemoglobin concentration and mean optical path length in the neonatal brain by second derivative near infrared spectroscopy,” Pediatr. Res., 39 (1), 32 –38 (1996). https://doi.org/10.1203/00006450-199601000-00005 0031-3998 Google Scholar

135. 

C. W. Yoxall and A. M. Weindling, “Measurement of venous oxyhaemoglobin saturation in the adult human forearm by near infrared spectroscopy with venous occlusion,” Med. Biol. Eng. Comput., 35 (4), 331 –336 (1997). https://doi.org/10.1007/BF02534086 0140-0118 Google Scholar

136. 

M. A. Franceschini, D. A. Boas, A. Zourabian, S. G. Diamond, S. Nadgir, D. W. Lin, J. B. Moore, and S. Fantini, “Near-infrared spiroximetry: Noninvasive measurements of venous saturation in piglets and human subjects,” J. Appl. Physiol., 92 (1), 372 –384 (2002). 8750-7587 Google Scholar

137. 

R. Boushel, H. Langberg, J. Olesen, M. Nowak, L. Simonsen, J. Bulow, and M. Kjaer, “Regional blood flow during exercise in humans measured by near-infrared spectroscopy and indocyanine green,” J. Appl. Physiol., 89 (5), 1868 –1878 (2000). 8750-7587 Google Scholar

138. 

B. Chance, M. T. Dait, C. Zhang, T. Hamaoka, and F. Hagerman, “Recovery from exercise-induced desaturation in the quadriceps muscles of elite competitive rowers,” Am. J. Physiol., 262 (3 Pt. 1), C766 –C775 (1992). 0002-9513 Google Scholar

139. 

T. Binzoni, V. Quaresima, M. Ferrari, E. Hiltbrand, and P. Cerretelli, “Human calf microvascular compliance measured by near-infrared spectroscopy,” J. Appl. Physiol., 88 (2), 369 –372 (2000). 8750-7587 Google Scholar

140. 

C. W. Yoxall, A. M. Weindling, N. H. Dawani, and I. Peart, “Measurement of cerebral venous oxyhemoglobin saturation in children by near-infrared spectroscopy and partial jugular venous occlusion,” Pediatr. Res., 38 (3), 319 –323 (1995). 0031-3998 Google Scholar

141. 

M. Wolf, K. von Siebenthal, M. Keel, V. Dietz, O. Baenziger, and H. U. Bucher, “Comparison of three methods to measure absolute cerebral hemoglobin concentration in neonates by near-infrared spectrophotometry,” J. Biomed. Opt., 7 (2), 221 –227 (2002). https://doi.org/10.1117/1.1463044 1083-3668 Google Scholar

142. 

T. S. Leung, I. Tachtsidis, M. Smith, D. T. Delpy, and C. E. Elwell, “Measurement of the absolute optical properties and cerebral blood volume of the adult human head with hybrid differential and spatially resolved spectroscopy,” Phys. Med. Biol., 51 (3), 703 –717 (2006). https://doi.org/10.1088/0031-9155/51/3/015 0031-9155 Google Scholar

143. 

P. Hopton, T. S. Walsh, and A. Lee, “Measurement of cerebral blood volume using near-infrared spectroscopy and indocyanine green elimination,” J. Appl. Physiol., 87 (5), 1981 –1987 (1999). 8750-7587 Google Scholar

144. 

I. Roberts, P. Fallon, F. J. Kirkham, A. Lloyd-Thomas, C. Cooper, R. Maynard, M. Elliot, and A. D. Edwards, “Estimation of cerebral blood flow with near infrared spectroscopy and indocyanine green,” Lancet, 342 (8884), 1425 (1993). https://doi.org/10.1016/0140-6736(93)92786-S 0140-6736 Google Scholar

145. 

E. Keller, G. Wietasch, P. Ringleb, M. Scholz, S. Schwarz, R. Stingele, S. Schwab, D. Hanley, and W. Hacke, “Bedside monitoring of cerebral blood flow in patients with acute hemispheric stroke,” Crit. Care Med., 28 (2), 511 –516 (2000). 0090-3493 Google Scholar

146. 

C. E. Elwell, J. R. Henty, T. S. Leung, T. Austin, J. H. Meek, D. T. Delpy, and J. S. Wyatt, “Measurement of CMRO2 in neonates undergoing intensive care using near infrared spectroscopy,” Adv. Exp. Med. Biol., 566 263 –268 (2005). 0065-2598 Google Scholar

147. 

T. H. Pham, O. Coquoz, J. B. Fishkin, E. Andersen, D. V. Gelfand, J. Milliken, T. Waddington, and B. J. Tromberg, “Broad bandwidth frequency domain instrument for quantitative tissue optical spectroscopy,” Rev. Sci. Instrum., 71 2500 –2513 (2000). https://doi.org/10.1063/1.1150665 0034-6748 Google Scholar

148. 

J. Lee, D. J. Saltzman, A. E. Cerussi, D. V. Gelfand, J. Milliken, T. Waddington, B. J. Tromberg, and M. Brenner, “Broadband diffuse optical spectroscopy measurement of hemoglobin concentration during hypovolemia in rabbits,” Physiol. Meas, 27 (8), 757 –767 (2006). 0967-3334 Google Scholar

149. 

K. Alford and Y. Wickramasinghe, “Intensity modulated near infrared spectroscopy: Instrument design issues,” Proc. SPIE, 3911 330 –337 (2000). https://doi.org/10.1117/12.384920 0277-786X Google Scholar

150. 

D. Geraskin, B. Platen, J. Franke, and M. Kohl-Bareis, “Algorithms for muscle oxygenation monitoring corrected for adipose tissue thickness,” 33 –39 (2005) Google Scholar

151. 

I. Nissila, K. Kotilahti, K. Fallström, and T. Katila, “Instrumentation for the accurate measurement of phase and amplitude in optical tomography,” Rev. Sci. Instrum., 73 3306 –3331 (2002). https://doi.org/10.1063/1.1497496 0034-6748 Google Scholar

152. 

L. A. Nelson, J. C. McCann, A. W. Loepke, J. Wu, B. B. Dor, and C. D. Kurth, “Development and validation of a multiwavelength spatial domain near-infrared oximeter to detect cerebral hypoxia-ischemia,” J. Biomed. Opt., 11 (6), 064022 (2006). https://doi.org/10.1117/1.2393251 1083-3668 Google Scholar

153. 

M. E. Giardini and S. Trevisan, “Portable high-end instrument for in-vivo infrared spectroscopy using spread spectrum modulation,” 860 –863 (2004) Google Scholar

154. 

Y. Teng, H. Ding, Q. Gong, Z. Jia, and L. Huang, “Monitoring cerebral oxygen saturation during cardiopulmonary bypass using near-infrared spectroscopy: The relationships with body temperature and perfusion rate,” J. Biomed. Opt., 11 (2), 024016 (2006). https://doi.org/10.1117/1.2187422 1083-3668 Google Scholar

155. 

D. Brown, R. Hornung, D. Haensse, M. Jacoma, M. Meerstetter, G. Morren, M. Stahel, D. Fink, H. U. Bucher, and M. Wolf, “Frequency-domain near-infrared spectroscopy measures tissue concentration of hemoglobin, lipids and water,” (2004) Google Scholar

156. 

D. R. Kashyap, N. Chu, A. Apte, B. P. Wang, and H. Liu, “Development of broadband multichannel NIRS (near-infrared spectroscopy) imaging system for quantification of spatial distribution of hemoglobin derivatives,” Proc. SPIE, 6434 64341X (2007). https://doi.org/10.1117/12.700894 0277-786X Google Scholar

157. 

S. Boden, H. Obrig, C. Köhncke, H. Benav, P. Koch, and J. Steinbrink, “The oxygenation response to functional stimulation: Is there a physiological meaning to the lag between parameters?,” Neuroimage, 36 (1), 100 –107 (2007). 1053-8119 Google Scholar

158. 

N. L. Everdell, A. P. Gibson, I. D. C. Tullis, T. Vaithianathan, J. C. Hebden, and D. T. Delpy, “A frequency multiplexed near-infrared topography system for imaging functional activation in the brain,” Rev. Sci. Instrum., 76 093705 (2005). https://doi.org/10.1063/1.2038567 0034-6748 Google Scholar

159. 

A. Akin and D. Bilensoy, “Cerebrovascular reactivity to hypercapnia in migraine patients measured with near-infrared spectroscopy,” Brain Res., 1107 (1), 206 –214 (2006). 0006-8993 Google Scholar

160. 

C. J. Li, H. Gong, Z. Gan, S. Q. Zeng, and Q. M. Luo, “Verbal working memory load affects prefrontal cortices activation: Evidence from a functional NIRS study in humans,” Proc. SPIE, 5696 33 –40 (2005). https://doi.org/10.1117/12.590222 0277-786X Google Scholar

161. 

C. H. Schmitz, M. Locker, J. M. Lasker, A. H. Hielscher, and R. L. Barbour, “Instrumentation for fast functional optical tomography,” Rev. Sci. Instrum., 73 (2), 429 –439 (2002). https://doi.org/10.1063/1.1427768 0034-6748 Google Scholar

162. 

J. Leon-Carrion, J. Damas, K. Izzetoglu, K. Pourrezai, J. F. Martin-Rodriguez, J. M. Barroso y Martin, and M. R. Dominguez-Morales, “Differential time course and intensity of PFC activation for men and women in response to emotional stimuli: A functional near-infrared spectroscopy (fNIRS) study,” Neurosci. Lett., 403 (1–2), 90 –95 (2006). 0304-3940 Google Scholar

163. 

J. P. Culver, B. L. Schlaggar, H. Dehghani, and B. W. Zeff, “Diffuse optical tomography for mapping human brain function,” (2006) Google Scholar

164. 

J. Selb, D. K. Joseph, and D. A. Boas, “Time-gated optical system for depth-resolved functional brain imaging,” J. Biomed. Opt., 11 (4), 044008 (2006). https://doi.org/10.1117/1.2337320 1083-3668 Google Scholar

165. 

Y. Ueda, T. Yamanaka, D. Yamashita, T. Suzuki, E. Ohmae, M. Oda, and Y. Yamashita, “Reflectance diffuse optical tomography: Its application to human brain mapping,” Jpn. J. Appl. Phys., Part 1, 44 (38), L1203 –L1206 (2005). https://doi.org/10.1143/JJAP.44.L1203 0021-4922 Google Scholar

166. 

D. Contini, A. Torricelli, A. Pifferi, L. Spinelli, P. Taroni, V. Quaresima, M. Ferrari, and R. Cubeddu, “Multichannel time-resolved tissue oximeter for functional imaging of the brain,” IEEE Trans. Instrum. Meas., 55 85 –90 (2006). https://doi.org/10.1109/TIM.2005.861502 0018-9456 Google Scholar

167. 

F. E. Schmidt, M. E. Fry, E. M. Hillman, J. C. Hebden, and D. T. Delpy, “A 32-channel time-resolved instrument for medical optical tomography,” Rev. Sci. Instrum., 71 256 –265 (2000). https://doi.org/10.1063/1.1150191 0034-6748 Google Scholar

168. 

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

169. 

A. Liebert, M. Kacprzak, and R. Maniewski, “Time-resolved reflectometry and spectroscopy for assessment of brain perfusion and oxygenation,” 113 –121 (2005) Google Scholar

170. 

I. Nissila, T. Noponen, K. Kotilahti, T. Katila, L. Lipiäinen, T. Tarvainen, M. Schweiger, and S. Arridge, “Instrumentation and calibration methods for the multichannel measurement of phase and amplitude in optical tomography,” Rev. Sci. Instrum., 76 044302 (2005). https://doi.org/10.1063/1.1884193 0034-6748 Google Scholar

171. 

K. Kwon, D. Ho, G. Eom, S. Lee, and B. Kim, “Trust region method for DOT image reconstruction,” Proc. SPIE, 6434 643428 (2007). https://doi.org/10.1117/12.701729 0277-786X Google Scholar

172. 

K. J. Kek, M. Samizo, T. Miyakawa, N. Kudo, and K. Yamamoto, “Imaging of regional differences of muscle oxygenation during exercise using spatially resolved NIRS,” 2622 –2625 (2005). Google Scholar
©(2007) Society of Photo-Optical Instrumentation Engineers (SPIE)
Martin Wolf, Marco Ferrari, and Valentina Quaresima "Progress of near-infrared spectroscopy and topography for brain and muscle clinical applications," Journal of Biomedical Optics 12(6), 062104 (1 November 2007). https://doi.org/10.1117/1.2804899
Published: 1 November 2007
Lens.org Logo
CITATIONS
Cited by 520 scholarly publications and 2 patents.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Near infrared

Continuous wave operation

Near infrared spectroscopy

Brain

Tissue optics

Tissues

Phase modulation

Back to Top