Open Access
1 November 2007 Functional near-infrared spectroscopy: current status and future prospects
Yoko Hoshi M.D.
Author Affiliations +
Abstract
Near-infrared spectroscopy (NIRS), which was originally designed for clinical monitoring of tissue oxygenation, has been developing into a useful tool for neuroimaging studies (functional near-infrared spectroscopy). This technique, which is completely noninvasive, does not require strict motion restriction and can be used in a daily life environment. It is expected that NIRS will provide a new direction for cognitive neuroscience research, more so than other neuroimaging techniques, although several problems with NIRS remain to be explored. This review demonstrates the strengths and the advantages of NIRS, clarifies the problems, and identifies the limitations of NIRS measurements. Finally, its future prospects are described.

1.

Introduction

In 1977, Jöbsis first described the in vivo application of near-infrared spectroscopy (NIRS);1 this technique was originally designed for clinical monitoring of tissue oxygenation.2, 3, 4 Since the early 1990s, it has also been developing as a useful tool for neuroimaging studies [functional near-infrared spectroscopy (fNIRS)]. 5, 6, 7, 8 Over the past 30 years, the technology has advanced and a wide range of NIRS instruments have been developed. Among them, the instruments for continuous wave (CW) measurements based on the modified Beer-Lambert law (CW-type instruments), which include the earliest NIRS instruments, are the most readily available commercially. Instruments of this type allow us to observe dynamic changes in regional cerebral blood flow (rCBF) in real time by measuring concentration changes in cerebral hemoglobin (Hb). Using fNIRS, various types of brain activities, such as motor and cognitive activities have been assessed. 9, 10, 11, 12, 13 The recent advent of multichannel CW-type instruments has greatly increased the use of NIRS in a variety of fields.

At the same time, however, the accuracy and reliability of NIRS have not yet been widely accepted. This is mainly attributable to incomplete knowledge of which region in the brain is sampled by near-infrared (NIR) light, difficulty in selective detection of NIRS signals arising from the cerebral tissue, and the problem of quantification. Despite a number of theoretical and experimental investigations, NIR light propagation in the human head remains to be fully understood. Because the detected light on the scalp carries information about not only the cerebral tissue but also the extracerebral tissue, and changes in extracerebral blood flow influence the determination of cerebral Hb concentration changes, it is necessary to separate signals originating in the cerebral tissue from those coming from the extracerebral tissue. For this purpose, a multidetector system consisting of CW-type instruments has been developed.14, 15 However, separation of NIR signals was incomplete, and other methods are being explored.

The major problem with NIRS is that concentration changes in Hb cannot be quantified with CW-type instruments, which has hindered NIRS from being widely employed in clinical medicine and research. Over the past 30 years, work in the field of NIRS has concentrated on solving this problem. Many different approaches to quantification have been tried, and the quantitative accuracy of time-resolved spectroscopy (TRS) and of phase-resolved spectroscopy (PRS) has been established. However, the difficulty of quantification has not yet been completely overcome (see Sec. 4.1).

In this review, I first outline the basic theory of NIRS, which will aid in the understanding of the potential and limitations of the technique. Then, focusing mainly on CW measurements, I give specific examples of the strengths and advantages of NIRS measurements over other neuroimaging modalities, and I also clarify the problems and identify the limitations of NIRS measurements. Finally, I describe its future prospects. Optical imaging is omitted because it is described in detail by other authors.

2.

Basic Theory of NIRS

2.1.

NIR Light Propagation in the Head

Knowledge about which regions in the brain are sampled by NIRS light is still incomplete, although a number of theoretical and experimental investigations on NIR light propagation in the human head have been performed. 16, 17, 18, 19 Light propagation is generally approximated by diffusion equation, which is not valid on non- or low-scattering regions, and can be predicted by Monte Carlo simulation. Theoretical analyses of the head models consisting of three- or four-layered slabs, the latter incorporating a clear cerebrospinal fluid (CSF) layer, have demonstrated that light propagation in the adult head is highly affected by the presence of a low-scattering CSF layer.20 This indicated that the light penetration in the adult brain might be limited to the outer cortical gray matter, which has been confirmed by their later studies, in which light propagation in head models generated from a magnetic resonance imaging (MRI) scan was predicted by Monte Carlo simulation.17 A more recent study has reported that a large source-detector spacing only broadens the sampling region on the brain surface and affects the penetration depth in the adult head to a lesser degree, whereas the intensely sensitive region in the neonatal head is confined in the gray matter; however, the deeper region of the white matter is sampled with a large source-detector spacing.19 In theoretical analysis, optical properties in each layer of the head are critical for prediction of the light propagation, though absorption (μa) and reduced scattering coefficients (μs) in each layer used for analysis are different from study to study. This is attributable to the fact that in situ measurements of the optical properties are not feasible. Thus, further investigation is required to confirm the validity of the theoretically predicted light propagation.

2.2.

NIRS Signals in Activated Areas

Regional brain activation is accompanied by increases in rCBF and the regional cerebral oxygen metabolic rate (rCMRO2) . It is widely accepted that the degree of the increase in rCBF exceeds that of the increase in rCMRO2 ,21 which results in a decrease in deoxyhemoglobin (deoxy-Hb) in venous blood. Thus, increases in total hemoglobin (t-Hb) and oxyhemoglobin (oxy-Hb) with a decrease in deoxy-Hb are expected to be observed in activated areas in NIRS measurements. However, deoxy-Hb and t-Hb do not necessarily show these changes: studies have observed both no change in t-Hb with an increase in oxy-Hb and a reciprocal decrease in deoxy-Hb, and an increase or no change in deoxy-Hb accompanying increases in t-Hb and oxy-Hb.6, 22, 23 Using a newly developed perfused rat brain model, we examined the direct effects of each change in CBF and CMRO2 on cerebral hemoglobin oxygenation to interpret NIRS signals.24 We confirmed that the directions of changes in oxy-Hb are always the same as those of rCBF, whereas the direction of changes in deoxy-Hb is determined by changes in venous blood oxygenation and volume. It has also been confirmed that small changes in CBF are not accompanied by those in t-Hb. Thus, oxy-Hb is the most sensitive indicator of changes in rCBF in NIRS measurements.

2.3.

Fluctuations in NIRS Signals at Rest

CW measurements have revealed that even under resting conditions, the Hb oxygenation state fluctuates,25, 26, 27 These fluctuations are divided into two types. One is the fluctuation whose general patterns are systemic and that are related to physiological activities such as the systemic arterial pulse oscillations (1Hz) and respiration (0.2 to 0.3Hz ). The other is the slower Hb wave fluctuation (frequency <0.05Hz ), of which the temporal pattern varies with each brain region (Fig. 1 ). It is well known that the cerebral blood flow velocity (CBFV) measured by transcranial Doppler ultrasound shows slow oscillation, which is thought to be attributed to small pial artery oscillations.28, 29 The frequency and characteristics of fluctuations in the Hb oxygenation state are similar to those of the CBFV oscillations. It is thus likely that the fluctuations in the Hb oxygenation state also originate from small artery oscillations. Our simultaneous measurements with NIRS and electroencephalography (EEG) have suggested the possibility that these oscillations are a result of vasomotor responses to spontaneous neuronal activity.30 Both the faster and slower fluctuations of the Hb oxygenation state sometimes have amplitudes comparable with those of the signals evoked by functional activity. Thus, taking into account these fluctuations is critical for interpretation of NIRS signals.

Fig. 1

Fluctuations in concentrations of oxygenated (dotted line), deoxygenated (broken line), and total hemoglobin (solid line) in the frontal (A) and occipital (B) regions during the resting state. The two regions were measured simultaneously. (Reprinted from Ref. 96 with permission.)

062106_1_008706jbo1.jpg

3.

Functional Near-Infrared Spectroscopy (fNIRS)

3.1.

Neuroimaging Studies with CW-Type Instruments

The strengths and advantages of CW-type instruments are as follows: (1) temporal resolution is high (less than 1sec ) and completely noninvasive, which allow long-time continuous measurements in real time and repeating measurements within short intervals, and (2) measurements can be performed with less motion restriction and in natural environments. Such strengths and advantages of NIRS enable neuroimaging studies on subjects who have not been fully examined until now, such as children, the elderly, and patients with psychoneurological problems, as they are difficult to measure by other neuroimaging techniques, such as positron emission tomography (PET) and functional magnetic resonance imaging (fMRI).

In neonates and infants, NIRS has been mostly applied to investigate evoked responses to stimuli such as visual,31, 32 olfactory,33 and auditory stimulation,34, 35 and passive knee movement.36 These studies have shown that unlike measurements on adult subjects, deoxy-Hb often increases accompanying increases in oxy-Hb and t-Hb, which corresponds to an inverse blood oxygen level–dependent (BOLD) signal.37 This stimulus-related increase in deoxy-Hb was explained by a lower increase in rCBF compared with an increase in the rCMRO2 .31 However, our study has demonstrated that the direction of activation-related changes in deoxy-Hb in neonates varies with each measurement even in the same subject during photic stimulation,32 which has given another possible explanation for the increase in deoxy-Hb. That is, it might be attributed to venous dilation caused by activation-related increases in rCBF. Even more recently, higher order functions, such as response to language, have been investigated in neonates.38, 39, 40 Although only a few NIRS studies on cognitive and socioemotional development have been so far reported,41, 42, 43 the importance of NIRS will soon increase in developmental psychology.

Okada 44 first applied NIRS to evaluating the frontal function in chronic schizophrenics. In the few years following this study, only a few psychiatric applications were reported.45, 46 Lately, however, NIRS has become an increasingly popular method in psychiatry. 47, 48, 49, 50 Several research groups have examined task-related hemodynamic changes in psychiatric patients and found task-dependent abnormalities in frontal hemodynamics in schizophrenia45, 48 and depression.47, 49 Such task-dependent abnormalities were also found in patients with Alzheimer’s disease.51, 52 These results underline the usefulness of NIRS in investigating frontal lobe dysfunction and evaluating psychopathologic conditions in psychiatric patients.

Measurements with less motion restriction in the daily life environment open new dimensions in neuroimaging studies. Using a 30-channel CW-type instrument, Miyai 53 succeed in visualizing cortical activation patterns associated with human gait. This indicated that NIRS was useful for evaluating cerebral activation patterns during pathological movements and rehabilitation intervention. Furthermore, a portable single-channel NIRS instrument combined with a wireless telemetry system (the wearable NIRS system) allows subjects to move during measurements as with portable electrocardiogram (ECG) and EEG instruments. The details of the portable NIRS instrument (HEO 200, Omron Ltd. Inc., Kyoto, Japan) have been reported elsewhere.54 This instrument is connected to the transmitter of a wireless system, and these are packed in a small bag that a subject carries (Fig. 2 ). NIRS signals are then sent by the wireless system to the receiver, which is connected to a laptop computer, on which data are displayed in real time. NIRS signals can be transmitted to a place at a maximum distance of 30m in an open field, but about 10m inside a building. This NIRS system makes it possible to monitor brain activity of freely moving subjects outside of laboratories.55

Fig. 2

Measurement being performed using the wearable NIRS system.

062106_1_008706jbo2.jpg

Using a multichannel CW-type instrument, we can examine spatiotemporal characteristics in hemodynamic changes associated with brain activity. Furthermore, multichannel NIRS instruments have the potential for imaging the sequence of brain activation.56, 57 In Fig. 3 , for an example, three brain regions (the left dorsolateral prefrontal cortex, left BA 8, and right ventrolateral prefrontal cortex) were independently activated during performance of the n -back task, in which the time course of changes in oxy-Hb was different and it appeared that these regions had worked in a complementary manner.57 This also implies that the results obtained in PET and fMRI studies can vary by measurement points. Examining the time course of hemodynamic changes is crucial for understanding the brain function.

Fig. 3

Dynamic oxygenated Hb changes during the 3-0 back task in different brain regions. The start and end of the horizontal axis denotes the start and end point of the task, respectively. D, dolsorateral prefrontal cortex; BA 8, Brodmann area 8; V, ventrolateral prefrontal cortex; LT, left; RT, right. (Reprinted from Ref. 57 with permission.)

062106_1_008706jbo3.jpg

3.2.

Multimodal Measurements

There are various functional neuroimaging modalities, such as fMRI, PET, and magnetoencepahlography (MEG). Each modality detects a different aspect of brain function and has different merits and demerits. Thus, combined measurements with multiple modalities are expected to be complementary and synergistic. Because NIRS can easily be combined with any of the other modalities, a number of combined studies with NIRS and other modalities have already been reported. Simultaneous measurements with PET52, 58 and fMRI23, 59 have also contributed to confirming the validity of NIRS, but the correspondence between the NIRS and the BOLD-fMRI recorded hemodynamic responses is still controversial.60, 61, 62

Neurovascular coupling is the physiological basis of such neuroimaging techniques to measure signals attributed to hemodynamic changes (e.g., PET, fMRI, and NIRS), whereas its mechanisms remain to be elucidated. To investigate the neurovascular coupling in the human brain, NIRS is appropriate because its temporal resolution is high, and it can be combined with electrophysiological methods, such as EEG and MEG. 63, 64, 65, 66 Simultaneous measurements with NIRS and electrophysiological methods have also been employed for studies of higher brain functions.67

Transcranial magnetic stimulation (TMS) is becoming popular as a therapeutic tool for neuropsychological diseases, such as depression and cerebellospinal degeneration, as well as for neurological examination. However, because therapeutic effects of TMS vary with its stimulation conditions, it is required to examine the relationship between stimulation conditions and changes in brain activity, CBF, and cerebral metabolisms. NIRS can be combined much more easily with TMS than PET or fMRI and has recently been employed for this aim.68, 69

4.

Problems of NIRS

4.1.

Selective and Quantitative Detection of NIRS Signals

Quantification of NIRS data has been a central issue in the NIRS field. When Hb concentration changes are global within the tissue, quantification is possible with TRS and PRS, which can determine optical path length. In the case where Hb concentration changes are localized, such as functional brain activation, however, those changes cannot be quantified accurately. The optical path length determined by TRS and PRS is the mean total path length (t-PL) but not the mean partial path length (p-PL) in the cerebral tissue. Because the t-PL is much longer than the p-PL,17, 19 Hb concentration changes are underestimated when the t-PL is substituted for the modified Beer-Lambert law (partial volume reduction). However, measurement of the p-PL is not feasible. To quantify NIRS data obtained from CW-type instruments without measuring the t-PL, the assumption that the ratio of the source-detector separation to the t-PL is constant has often been made.70 Arranging the source-detector separation for each pair at equal distance, in which the t-PL can be considered a constant if the assumption is correct, multichannel CW-type instruments generate topographical images of relative concentration changes in Hb. However, this assumption is not correct. As shown in Fig. 4 , the ratio of the source-detector separation to the t-PL [differential path length factor (DPF)]70 varies with each position. Furthermore, the p-PL is negatively related to the t-PL at a fixed source-detector spacing (Fig. 5 ), and the ratio of the p-PL to the t-PL varies with each wavelength and each measurement position.71 This means that substitution of the t-PL for the Beer-Lambert law provides not only underestimated but also inaccurate results. Thus, for quantification, we have to come up with methods to measure the p-PL or other approaches to quantitative and selective detection of signals arising from the brain that are not based on the modified Beer-Lambert law. In addition to the issue of quantification, amplitudes of NIRS signals vary with the source-detector position.72, 73 Thus, comparing amplitudes across subjects and/or regions within a subject is not valid.

Fig. 4

Different path length factors measured on the left scalp adjacent to the somatosensorimotor cortex in a 23-year-old healthy adult. The wavelength is 759nm .

062106_1_008706jbo4.jpg

Fig. 5

Relationship between the PL (t-PL) and the p-PL (closed circle) and the ratio of the p-PL to the PL (closed triangle) in the three head models. (Reprinted from Ref. 71 with permission.)

062106_1_008706jbo5.jpg

Diffuse optical tomography (DOT), which reconstructs images of Hb concentration changes using multiple light sources and detectors, is a potential technique for quantitative detection of focal changes in cerebral hemodynamics.74 DOT is not based on the modified Beer-Lambert law and can be performed with TRS,75 PRS,76 and CW-type instruments.77 TRS is also a potential tool for this purpose. It provides the temporal point spread function, which carries information about depth-dependent attenuation even for a measurement with a single source-detector distance. Even though several time domain approaches 78, 79, 80, 81 and a multidistance frequency domain approach82 have been proposed, further investigation must be continued to apply these methods to human head measurements.

4.2.

NIRS Data Analysis

Unlike PET and fMRI, there are no standard methods of NIRS data analysis, and various analyses have been performed so far. This is not in itself a problem; however, the validity and reliability of each method should be confirmed. Until recently, to examine whether task-related changes in NIRS signals in an individual are significant or not, comparison of NIRS signals between the resting and activation states has commonly been performed by using a paired t -test. However, because NIRS data are time series data, t -statistics cannot be used for this aim, although they can be used for comparing means of NIRS signal changes between two states within subjects. Autoregressive models are commonly used to analyze time series data, though it is very difficult to derive an autoregressive model for NIRS data. Model-based, event-related, and both combined analyses have recently been tried in some research groups. 60, 83, 84, 85 Although the model-based analysis is widely used for data analysis in fMRI and PET studies, it is unclear whether this analytical method can be applied to NIRS data, because the pattern of hemodynamic changes varies with each measurement and it is difficult to derive proper hemodynamic response functions, which might be also true of PET and fMRI. In the case that the same task can repeatedly be performed without habituation, however, an event-related analysis is available.

As mentioned in Sec. 2.3, NIRS signals are not constant during the resting state, which possibly reflects physiological phenomena,26, 30 and it is often observed that these signals do not return to the original levels immediately after the activation state. In such cases, baseline correction has been performed in some studies. However, it should be noted that the baseline correction could distort actual cerebral hemodynamic changes.

4.3.

Cross Talk

Recently, cross talk between the estimated oxy-Hb and deoxy-Hb concentrations has been intensively studied.72, 86 The estimated concentration change (Δ[]estim) is expressed by Eq. 1

Eq. 1

Δ[A]estim=PΔ[A]real+C[B]real,
where A represents either oxy-Hb or deoxy-Hb, and B represents the other chromophores deoxy-Hb or oxy-Hb, respectively. Δ[]real indicates the real concentration change. P denotes the partial volume reduction in the estimated chromophore concentration, and C denotes the cross talk from the other chromophore.

When two wavelengths ( λ1 , λ2 ) are used, the cross talk is expressed by Eq. 2

2.

C=EL,
E=εR(λ1)εR(λ2)εB(λ1)εA(λ2)εA(λ1)εB(λ2),
L=p-PL(λ2)t-PL(λ2)p-PL(λ1)t-PL(λ1),
where εA and εB are the extinction coefficients of the two chromophores, A and B . To eliminate the cross talk, a wavelength pair that minimizes E is required. Some research groups have recommended to use the 830- to 690-nm pair and reported that the 780- to 830-nm pair, which has traditionally been used, increases E .77, 86, 87 However, L has not been considered. Even though E is minimized in the 830- to 690-nm pair, L might be larger at this wavelength pair than at the 780- to 830-nm pair. In addition, it is unclear whether light propagation between the source and detector at 690nm is the same as that at 830nm and whether scattering changes at 690nm can be ignored. Further investigations are required before the new wavelength pair is employed.

4.4.

Practical Issues

Except for neonates, it is not feasible for NIRS to noninvasively measure deep brain structures such as the diencephalons. It is also difficult to identify the exact brain areas that are beneath the NIRS probes without three-dimensional MRI measurements. However, it has been reported that there is an appropriate relationship between the international 10-20 system of electrode positioning in electroencephalogram and the cortical anatomy.88, 89 Thus, the measured brain area can be roughly deduced by using the location of the 10-20 system as a landmark.

Advances in NIRS technology have enabled simultaneous measurements at multiple brain regions with high temporal resolution, which is desirable not only to investigate regional differences in brain activation but also to detect localized brain activation. To improve spatial resolution, various arrangements of source-detector positions are being designed. However, fundamental problems remain to be solved: it takes time and skill to place many source-detector pairs on the hairy scalp and on the head of neonates and infants. Improvement of NIRS probes is vital for dissemination of NIRS in various fields of medicine and research. In addition, although less motion restriction during measurements is the strength of NIRS, head motion easily causes artifacts. Such artifacts cannot be eliminated automatically with a computer because of difficulty in distinguishing between signals and artifacts and require inspection by the naked eye.

5.

Future Prospects

As is mentioned in Sec. 4, there remain a number of technical issues to be explored and practical difficulties to be solved. Nevertheless, NIRS is a tool distinct from other neuroimaging techniques for the study of brain functions and for the diagnosis, assessment, and treatment of psychoneurological diseases. Thus, a variety of novel applications of NIRS, such as NIRS-based brain-computer interface (BCI), are being tried. BCI provides users with an alternative output channel rather than the normal output path of the brain (i.e., the efferent nervous system and muscles).90 BCI has been given much attention recently as an alternate mode of communication and control for the disabled, such as patients suffering from amyotrophic lateral sclerosis and “locked-in” patients. Most of the current BCI systems rely on the brain’s electrical activity producing scalp EEG signals. Because the scalp EEG signals, however, are inherently noisy and nonlinear, a more accessible interface that uses a more direct measurement of brain function to control an output device is being explored. NIRS is considered as a possible alternative to electrical signals.91

Another optical approach to detect brain activation is also being tried. Conventional NIRS instruments detect signals corresponding to relatively slow hemodynamic responses. In contrast, a much faster signal occurring over a period of tens of milliseconds has been detected by both a frequency-domain system92, 93 and a CW system.94 These fast signals, which are thought to be attributable to scattering changes in neurons, are much weaker than those of hemodynamic origin, and high temporal resolution is required for their detection. The instrumentation and data analysis of the techniques have remarkably improved over the last few years, making it feasible to detect neuronal activity. This new approach is becoming a powerful clinical tool.

In the last few years the development of the multichannel NIRS system, which has abandoned handiness, one of the most characteristic features of NIRS, has been focused on in the NIRS field, the miniaturization of the NIRS system has also been tried. We have been extending the wearable system mentioned in Sec. 3.1 to the multichannel system. Wolf’s group has recently succeeded in miniaturizing near-infrared optical imaging and creating a wireless sensor.95 Such miniaturized NIRS systems will contribute not only to neuroscience research but also to monitoring tissue oxygenation, which was the original aim of NIRS development.

The development of NIRS has taken a long time. Over the last 30 years, however, NIRS has been making steady progress and its strengths and advantages are expected to provide a new direction for functional mapping studies that other neuroimaging techniques have not been able to achieve. Thus, NIRS shows great promise for providing further insight into brain function and as a clinical tool. Lastly, I must say that I am greatly indebted to the profound pioneering research of Dr. Jöbsis, and I would like to pay him all due respects.

References

1. 

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

2. 

J. E. Brazy, D. V. Lewis, M. H. Mitnick, and F. F. Jöbsis, “Noninvasive monitoring of cerebral oxygenation in preterm infants: Preliminary observation,” Pediatrics, 75 217 –225 (1985). 0031-4005 Google Scholar

3. 

P. A. Rea, J. Crowe, Y. Wickramasinghe, and P. Rolfe, “Non-invasive optical methods for the study of cerebral metabolism in the human newborn: A technique for the future?,” J. Med. Eng. Technol., 9 160 –165 (1985). 0309-1902 Google Scholar

4. 

J. S. Wyatt, M. Cope, D. T. Delpy, S. Wray, and E. O. R. Raynolds, “Quantification of cerebral oxygenation and haemodynamics in sick newborn infants by near infrared spectrophotometry,” Lancet, 2 1063 –1066 (1986). https://doi.org/10.1016/S0140-6736(86)90467-8 0140-6736 Google Scholar

5. 

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

6. 

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., 13 516 –520 (1993). 0271-678X Google Scholar

7. 

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

8. 

B. Chance, Z. Zhuang, C. Unah, C. Alter, and L. Lipton, “Cognition-activated low frequency modulation of light absorption in human brain,” Proc. Natl. Acad. Sci. U.S.A., 90 3770 –3774 (1993). https://doi.org/10.1073/pnas.90.8.3770 0027-8424 Google Scholar

9. 

W. N. Colier, V. Quaresima, B. Oeseburug, and M. Ferrari, “Human motor-cortex oxygenation changes induced by cyclic coupled movements of hand and foot,” Exp. Brain Res., 129 457 –461 (1999). https://doi.org/10.1007/s002210050913 0014-4819 Google Scholar

10. 

H. R. Heekeren, H. Obrig, R. Wenzel, K. Eberle, J. Ruben, K. Villringer, R. Kurth, and A. Villringer, “Cerebral haemoglobin oxygenation during sustained visual stimulation—A near-infrared spectroscopy study,” Philos. Trans. R. Soc. London, Ser. B, 352 743 –750 (1997). https://doi.org/10.1098/rstb.1997.0057 0962-8436 Google Scholar

11. 

H. Sato, T. Takeuchi, and K. L. Sakai, “Temporal cortex activation during speech recognition: An optical topography study,” Cognition, 73 B55 –B66 (1999). https://doi.org/10.1016/S0010-0277(99)00060-8 0010-0277 Google Scholar

12. 

S. Shimada, K. Hiraki, G. Matsuda, and I. Oda, “Decrease in prefrontal hemoglobin oxygenation during reaching tasks with delayed visual feedback: A near-infrared spectroscopy study,” Brain Res. Cognit. Brain Res., 20 480 –490 (2004). https://doi.org/10.1016/j.cogbrainres.2004.04.004 0926-6410 Google Scholar

13. 

M. Tanosaki, Y. Hoshi, Y. Iguchi, Y. Oikawa, I. Oda, and M. Oda, “Variation of temporal characteristics in human cerebral hemodynamic responses to electric median nerve stimulation: A near-infra-red spectroscopic study,” Neurosci. Lett., 316 75 –78 (2001). https://doi.org/10.1016/S0304-3940(01)02372-2 0304-3940 Google Scholar

14. 

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

15. 

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

16. 

D. A. Boas, J. P. Culver, J. J. Stott, and A. K. Dunn, “Three dimensional Monte Carlo code for photon migration through complex heterogeneous media including the adult human head,” Opt. Express, 10 159 –170 (2002). 1094-4087 Google Scholar

17. 

M. Firbank, E. Okada, and D. T. Delpy, “A theoretical study of the signal contribution of regions of the adult head to near-infrared spectroscopy studies of visual evoked responses,” Neuroimage, 8 69 –78 (1998). 1053-8119 Google Scholar

18. 

G. Gratton, J. S. Maier, M. Fabiani, W. W. Mantulinm, and E. Gratton, “Feasibility of intracranial near-infrared optical scanning,” Psychophysiology, 31 211 –215 (1994). https://doi.org/10.1111/j.1469-8986.1994.tb01043.x 0048-5772 Google Scholar

19. 

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

20. 

E. Okada, M. Firbank, M. Schweiger, S. R. Arridge, M. Cope, and D. T. Delpy, “Theoretical and experimental investigation of near-infrared light propagation in a model of the adult head,” Appl. Opt., 36 21 –31 (1997). https://doi.org/10.1007/s002459900053 0003-6935 Google Scholar

21. 

P. Fox and M. E. Raichle, “Focal physiological uncoupling of cerebral blood flow and oxidative metabolism during somatosensory stimulation in human subjects,” Proc. Natl. Acad. Sci. U.S.A., 83 1140 –1144 (1986). https://doi.org/10.1073/pnas.83.4.1140 0027-8424 Google Scholar

22. 

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

23. 

A. Kleinschmidt, H. Obrig, M. Requardt, K.-D. Merboldt, U. Dirnagl, A. Villringer, and J. Frahm, “Simultaneous recording of cerebral blood oxygenation changes during human brain activation by magnetic resonance imaging and near-infrared spectroscopy,” J. Cereb. Blood Flow Metab., 16 817 –826 (1996). https://doi.org/10.1097/00004647-199609000-00006 0271-678X Google Scholar

24. 

Y. Hoshi, N. Kobayashi, and M. Tamura, “Interpretation of near-infrared spectroscopy signals: A study with a newly developed perfused rat brain model,” J. Appl. Physiol., 90 1657 –1662 (2001). 8750-7587 Google Scholar

25. 

Y. Hoshi and M. Tamura, “Fluctuations in the cerebral oxygenation state during the resting period in functional mapping studies of the human brain,” Med. Biol. Eng. Comput., 35 328 –330 (1997). https://doi.org/10.1007/BF02534085 0140-0118 Google Scholar

26. 

V. Toronov, M. A. Francecshini, 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., 27 801 –815 (2000). https://doi.org/10.1118/1.598943 0094-2405 Google Scholar

27. 

M. L. Schroeter, O. Schmiedel, and D. Y. von Cramon, “Spontaneous low-frequency oscillation decline in the aging brain,” J. Cereb. Blood Flow Metab., 24 1183 –1191 (2004). https://doi.org/10.1097/01.WCB.0000135231.90164.40 0271-678X Google Scholar

28. 

L. M. Auer and I. Sayama, “Intracranial pressure oscillations (B-waves) caused by oscillations in cerebrovascular volume,” Acta Neurochir., 68 93 –100 (1983). 0001-6268 Google Scholar

29. 

K. F. Lindegaard, T. Lundar, J. Wiberg, D. Sjøberg, R. Aaslid, and H. Normes, “Variations in middle cerebral artery blood flow investigated with noninvasive transcranial blood velocity measurements,” Stroke, 18 1025 –1030 (1987). 0039-2499 Google Scholar

30. 

Y. Hoshi, S. Kosaka, Y. Xie, S. Kohri, and M. Tamura, “Relationship between fluctuations in the cerebral hemoglobin oxygenation state and neuronal activity under resting conditions in man,” Neurosci. Lett., 245 147 –150 (1998). 0304-3940 Google Scholar

31. 

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

32. 

Y. Hoshi, S. Kohri, Y. Matsumoto, K. Cho, T. Matsuda, S. Okajima, and S. Fujimoto, “Hemodynamic responses to photic stimulation in neonates,” Pediatr. Neurol., 23 323 –327 (2000). 0887-8994 Google Scholar

33. 

M. Bartocci, J. Winberg, C. Ruggiero, L. L. Bergqvist, G. Serra, and H. Lagercrantz, “Activation of olfactory cortex in newborn infants after odor stimulation: A functional near-infrared spectroscopy,” Pediatr. Res., 48 18 –23 (2000). https://doi.org/10.1203/00006450-200007000-00006 0031-3998 Google Scholar

34. 

K. Sakatani, S. Chen, W. Lichty, H. Zuo, and Y. Wang, “Cerebral blood oxygenation changes induced by auditory stimulation in newborn infants measured by near infrared spectroscopy,” Early Hum. Dev., 55 229 –236 (1999). https://doi.org/10.1016/S0378-3782(99)00019-5 0378-3782 Google Scholar

35. 

P. Zaramella, F. Freato, A. Amigoni, S. Salvadori, P. Marangoni, A. Suppjei, B. Schiavo, and L. Chiandetti, “Brain auditory activation measured by near-infrared spectroscopy (NIRS) in neonates,” Pediatr. Res., 49 213 –219 (2001). https://doi.org/10.1203/00006450-200102000-00014 0031-3998 Google Scholar

36. 

K. Isobe, T. K. Usaka, K. Nagano, K. Okubo, S. Yasuda, M. Kondo, S. Itoh, and S. Onishi, “Functional imaging of the brain in sedated newborn infants using near infrared topography during passive knee movement,” Neurosci. Lett., 299 221 –224 (2001). https://doi.org/10.1016/S0304-3940(01)01518-X 0304-3940 Google Scholar

37. 

P. Born, H. Leth, M. J. Miranda, E. Rostrup, A. Stensgaard, B. Peitersen, H. B. W. Larsson, and H. C. Lou, “Visual activation in infants and young children studies by functional magnetic resonance imaging,” Pediatr. Res., 44 578 –583 (1998). https://doi.org/10.1203/00006450-199810000-00018 0031-3998 Google Scholar

38. 

M. Peňa, A. Maki, D. Kovačić, G. Dehaene-Lambertz, H. Koizumi, F. Bouquet, and J. Mehler, “Sounds and silence: An optical topography study of language recognition at birth,” Proc. Natl. Acad. Sci. U.S.A., 100 11702 –11705 (2003). https://doi.org/10.1073/pnas.1934290100 0027-8424 Google Scholar

39. 

R. N. Aslin and J. Mehler, “Near-infrared spectroscopy for functional studies of brain activity in human infants: Promise, prospects, and challenges,” J. Biomed. Opt., 10 11009 (2005). 1083-3668 Google Scholar

40. 

T. Wilcox, H. Bortfeld, R. Woods, E. Wruck, and D. A. Boas, “Using near-infrared spectroscopy to assess neural activation during object processing in infants,” J. Biomed. Opt., 10 11010 (2005). 1083-3668 Google Scholar

41. 

A. Baird, J. Kagan, T. Gaudette, K. A. Walz, N. Hershlag, and D. A. Boas, “Frontal lobe activation during object performance: Data from near-infrared spectroscopy,” Neuroimage, 16 1120 –1126 (2002). 1053-8119 Google Scholar

42. 

Y. Hoshi and S.-J. Chen, “Cerebral blood flow changes associated with emotion in children,” Pediatr. Neurol., 27 275 –281 (2002). 0887-8994 Google Scholar

43. 

P. Weber, J. Lütschg, and H. Fahnenstich, “Cerebral hemodynamic changes in response to an executive function task in children with attention-deficit hyperactivity disorder measured by near-infrared spectroscopy,” J. Dev. Behav. Pediatr., 26 105 –111 (2005). 0196-206X Google Scholar

44. 

F. Okada, Y. Tokumitsu, Y. Hoshi, and M. Tamura, “Impaired interhemispheric integration in brain oxygenation and hemodynamics in schizophrenia,” Eur. Arch. Psychiatry Clin. Neurosci., 244 17 –25 (1994). 0940-1334 Google Scholar

45. 

A. J. Fallgatter and W. K. Strik, “Reduced frontal functional asymmetry in schizophrenia during a cued continuous performance test assessed with near-infrared spectroscopy,” Schizophr Bull., 26 913 –919 (2000). 0586-7614 Google Scholar

46. 

S. Saito, D. Yoshikawa, F. Nishihara, T. Morita, Y. Kitani, T. Amaya, and T. Fujita, “The cerebral hemodynamic response to electrically induced seizures in man,” Brain Res., 673 93 –100 (1995). 0006-8993 Google Scholar

47. 

K. Matsuo, N. Kato, and T. Kato, “Decreased cerebral haemodynamic response to cognitive and physiological tasks in mood disorders as shown by near-infrared spectroscopy,” Psychol. Med., 32 1029 –1037 (2002). 0033-2917 Google Scholar

48. 

T. Shinba, M. Nagano, N. Karia, N. Ozawa, T. Shinohara, S. Shimosato, and Y. Hoshi, “Near-infrared spectroscopy analysis of frontal lobe dysfunction in schizophrenia,” Biol. Psychiatry, 55 154 –164 (2004). 0006-3223 Google Scholar

49. 

T. Suto, M. Fukuda, M. Ito, T. Uehara, and M. Mikuni, “Multichannel near-infrared spectroscopy in depression and schizophrenia: Cognitive brain activation study,” Biol. Psychiatry, 55 501 –511 (2004). 0006-3223 Google Scholar

50. 

H. Kuwabara, K. Kasai, R. Takizawa, Y. Kawakubo, H. Yamasue, M. A. Roger, M. Ishijima, K. Watanabe, and N. Kato, “Decreased prefrontal activation during letter fluency task in adults with pervasive developmental disorders: A near-infrared spectroscopy study,” Behav. Brain Res., 172 272 –277 (2006). 0166-4328 Google Scholar

51. 

A. J. Fallgatter, M. Roesler, L. Sitzmann, A. Heidrich, T. J. Mueller, and W. K. Strik, “Loss of functional hemispheric asymmetry in Alzheimer’s dementia assessed with near-infrared spectroscopy,” Brain Res. Cognit. Brain Res., 6 67 –72 (1997). https://doi.org/10.1016/S0926-6410(97)00016-5 0926-6410 Google Scholar

52. 

C. Hock, K. Villringer, F. Müller-Spahn, R. Wenzel, H. Heekeren, S. Schuh-Hofer, M. Hofman, S. Minoshima, M. Schwaiger, U. Dirnagl, and A. Villringer, “Decrease in parietal cerebral hemoglobin oxygenation during performance of a verbal fluency task in patients with Alzheimer’s disease monitored by means of near-infrared spectroscopy (NIRS)—Correlation with simultaneous rCBF-PET measurements,” Brain Res., 755 293 –303 (1997). https://doi.org/10.1016/S0006-8993(97)00122-4 0006-8993 Google Scholar

53. 

I. Miyai, H. C. Tanabe, I. Sase, H. Eda, I. Oda, I. Konishi, Y. Tsunazawa, T. Suzuki, T. Yanagida, and K. Kubota, “Cortical mapping of gait in humans: A near-infrared spectroscopic topography study,” Neuroimage, 14 1186 –1192 (2001). 1053-8119 Google Scholar

54. 

T. Shiga, K. Yamamoto, K. Tanabe, Y. Nakase, and B. Chance, “Study of an algorithm based on model experiments and diffusion theory for a portable tissue oximeter,” J. Biomed. Opt., 2 154 –161 (1997). https://doi.org/10.1117/12.268963 1083-3668 Google Scholar

55. 

Y. Hoshi and S.-J. Chen, “New dimension of cognitive neuroscience research with near-infrared spectroscopy: Free-motion neuroimaging studies,” Progress in Brain Mapping Research, 205 –229 Nova Science, New York (2006). Google Scholar

56. 

Y. Hoshi and M. Tamura, “Near-infrared optical detection of sequential brain activation in the prefrontal cortex during mental tasks,” Neuroimage, 5 292 –297 (1997). 1053-8119 Google Scholar

57. 

Y. Hoshi, B. H. Tsou, V. A. Billock, M. Tanosaki, Y. Iguchi, M. Shimada, T. Shinba, Y. Yamada, and I. Oda, “Spatiotemporal characteristics of hemodynamic changes in the human lateral prefrontal cortex during working memory tasks,” Neuroimage, 20 1493 –1504 (2003). 1053-8119 Google Scholar

58. 

Y. Hoshi, H. Onoe, Y. Watanabe, J. Andersson, M. Bergström, A. Lilija, B. Långström, and M. Tamura, “Non-synchronous behavior of neuronal activity, oxidative metabolism and blood supply during mental tasks in man,” Neurosci. Lett., 172 129 –133 (1994). https://doi.org/10.1016/0304-3940(94)90679-3 0304-3940 Google Scholar

59. 

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., 28 521 –527 (2001). https://doi.org/10.1118/1.1354627 0094-2405 Google Scholar

60. 

T. J. Huppert, R. D. Hoge, S. G. Diamond, M. A. Franceshini, and D. A. Boas, “A temporal comparison of BOLD, ASL, and NIRS hemodynamic responses to motor stimuli in adult humans,” Neuroimage, 29 368 –382 (2006). 1053-8119 Google Scholar

61. 

A. Seiyama, J. Seki, H. C. Tanabe, I. Sase, A. Takatuki, S. Miyauchi, H. Eda, S. Hayashi, T. Imaruoka, T. Iwakura, and T. Yanagida, “Circulatory basis of fMRI signals: Relationship between changes in the hemodynamic parameters and BOLD signal intensity,” Neuroimage, 21 1204 –1214 (2004). 1053-8119 Google Scholar

62. 

N. Fujikawa, K. Sakatani, Y. Katayama, Y. Murata, T. Hoshino, C. Fukaya, and T. Yamamoto, “Evoked-cerebral blood oxygenation changes in false-negative activations in BOLD contrast functional MRI of patients with brain tumors,” Neuroimage, 21 1464 –1471 (2004). 1053-8119 Google Scholar

63. 

H. Obrig, H. Israel, M. Kohl-Bareis, K. Uludag, R. Wenzel, B. Müller, G. Arnold, and A. Villringer, “Habituation of the visually evoked potential and its vascular response: Implications for neurovascular coupling in the healthy adult,” Neuroimage, 17 1 –18 (2002). 1053-8119 Google Scholar

64. 

S. P. Koch, J. Steinbring, A. Villringer, and H. Obrig, “Synchronization between background activity and visually evoked potential is not mirrored by focal hyperoxygenation: Implications for the interpretation of vascular brain imaging,” J. Neurosci., 26 4940 –4948 (2006). https://doi.org/10.1523/JNEUROSCI.3989-05.2006 0270-6474 Google Scholar

65. 

B.-M. Mackert, G. Wubbeler, S. Leistner, K. Uldag, H. Obrig, A. Villringer, L. Trahms, and G. Curio, “Neurovascular coupling analyzed non-invasively in the human brain,” NeuroReport, 19 63 –66 (2004). https://doi.org/10.1097/00001756-200401190-00013 0959-4965 Google Scholar

66. 

J. Fuster, M. Guiou, A. Ardestani, A. Cannestra, S. Sheth, Y.-D. Zhou, A. Toga, and M. Bonder, “Near-infrared spectroscopy (NIRS) in cognitive neuroscience of the primate brain,” Neuroimage, 26 215 –220 (2005). 1053-8119 Google Scholar

67. 

S. G. Horovitz and J. C. Gore, “Simultaneous event-related potential and near-infrared spectroscopic studies of semantic processing,” Hum. Brain Mapp, 22 110 –115 (2004). 1065-9471 Google Scholar

68. 

Y. Noguchi, E. Watanabe, and K. L. Sakai, “An event-related optical topography study of cortical activation induced by single-pulse transcranial magnetic stimulation,” Neuroimage, 19 156 –162 (2003). 1053-8119 Google Scholar

69. 

H. Mochizuki, Y. Ugawa, Y. Terao, and K. L. Sakai, “Cortical hemoglobin-concentration changes under the coil induced by single-pulse TMS in humans: A simultaneous recording with near-infrared spectroscopy,” Exp. Brain Res., 169 302 –310 (2006). https://doi.org/10.1007/s00221-005-0149-0 0014-4819 Google Scholar

70. 

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 1433 –1442 (1988). https://doi.org/10.1088/0031-9155/33/12/008 0031-9155 Google Scholar

71. 

Y. Hoshi, M. Shimada, C. Sato, and Y. Iguchi, “Reevaluation of near-infrared light propagation in the adult human head: Implications for functional near-infrared spectroscopy,” J. Biomed. Opt., 10 064032 (2005). https://doi.org/10.1117/1.2142325 1083-3668 Google Scholar

72. 

G. Strangman, M. A. Franceschini, and D. A. Boas, “Factors affecting the accuracy of near-infrared spectroscopy concentration calculation for focal changes in oxygenation parameters,” Neuroimage, 18 865 –879 (2003). 1053-8119 Google Scholar

73. 

Y. Fukui, T. Yamamoto, T. Kato, and E. Okada, “Analysis of light propagation in a three-dimensional realistic head model for topographic imaging by finite difference method,” Opt. Rev., 10 470 –473 (2003). https://doi.org/10.1007/s10043-003-0470-4 1340-6000 Google Scholar

74. 

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

75. 

H. Gao, H. Zhao, and Y. Yamada, “Improvement of image quality in diffuse optical tomography using full time-resolved data,” Appl. Opt., 41 778 –791 (2002). https://doi.org/10.1364/AO.41.000778 0003-6935 Google Scholar

76. 

B. W. Pogue, M. S. Patterson, H. Jiang, and K. D. Paulsen, “Initial assessment of a simple system for frequency domain diffuse optical tomography,” Phys. Med. Biol., 40 1709 –1729 (1995). https://doi.org/10.1088/0031-9155/40/10/011 0031-9155 Google Scholar

77. 

D. A. Boas, T. Gaudette, G. Strangman, X. Cheng, J. J. A. Marota, and J. B. Mandeville, “The accuracy of near infrared spectroscopy and imaging during focal changes in cerebral hemodynamics,” Neuroimage, 13 76 –90 (2001). 1053-8119 Google Scholar

78. 

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

79. 

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

80. 

M. Shimada, Y. Hoshi, and Y. Yamada, “A simple algorithm for measurement of the absorption coefficients of a two-layered medium by spatially and time-resolved reflectance,” Appl. Opt., 44 7554 –7563 (2005). https://doi.org/10.1364/AO.44.007554 0003-6935 Google Scholar

81. 

C. Sato, M. Shimada, Y. Hoshi, and Y. Yamada, “Extraction of depth-dependent signals from time-resolved reflectance in layered turbid media,” J. Biomed. Opt., 10 064008 (2005). https://doi.org/10.1117/1.2136312 1083-3668 Google Scholar

82. 

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

83. 

M. L. Schroeter, S. Zysset, T. Kupka, F. Kruggel, and D. Yves von Cramon, “Near-infrared spectroscopy can detect brain activity during a color-word matching Stroop task in an event-related design,” Hum. Brain Mapp, 17 61 –71 (2002). 1065-9471 Google Scholar

84. 

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

85. 

M. M. Plichta, S. Heinzel, A.-C. Ehlis, P. Pauli, and A. J. Fallgatter, “Mode-based analysis of rapid event-related functional near-infrared spectroscopy (NIRS) data: A parametric validation study,” Neuroimage, 35 625 –634 (2007). 1053-8119 Google Scholar

86. 

K. Uludağ, J. Steinbrink, A. Villringer, and H. Obrig, “Separability and cross talk: Optimizing dual wavelength combinations for near-infrared spectroscopy of the adult head,” Neuroimage, 22 583 –589 (2004). 1053-8119 Google Scholar

87. 

N. Okui and E. Okada, “Wavelength dependent of crosstalk in dual-wavelength measurement of oxy- and deoxy-hemoglobin,” J. Biomed. Opt., 10 011015 (2005). https://doi.org/10.1117/1.1846076 1083-3668 Google Scholar

88. 

V. L. Towle, J. Bolanos, D. Suarez, K. Tan, R. Grzeszczuk, D. N. Levin, R. Cakmur, S. A. Frank, and J. Spire, “The spatial location of EEG electrodes: Locating the best-fitting sphere relative to cortical anatomy,” Electroencephalogr. Clin. Neurophysiol., 86 1 –6 (1993). https://doi.org/10.1016/0013-4694(93)90061-Y 0013-4649 Google Scholar

89. 

M. Okamoto, H. Dan, K. Sakamoto, K. Takeo, K. Shimizu, A. Kohno, I. Oda, S. Isobe, T. Suzuki, K. Kohyama, and I. Dan, “Three-dimensional probabilistic anatomical cranio-cerebral correlation via the international 10–20 system oriented for transcranial functional brain mapping,” Neuroimage, 21 99 –111 (2004). 1053-8119 Google Scholar

90. 

N. Birbaumer and L. G. Cohen, “Brain-computer-interface (BCI): Communication and restoration of movement in paralysis,” J. Physiol. (London), 579 621 –636 (2007). https://doi.org/10.1113/jphysiol.2006.125633 0022-3751 Google Scholar

91. 

R. Sitaram, H. Zhang, C. Guan, M. Thulasidas, Y. Hoshi, A. Ishikawa, K. Shimizu, and N. Birbaumer, “Temporal classification of multi-channel near infrared spectroscopy signals of motor imagery for developing a brain-computer interface,” Neuroimage, 34 1416 –1427 (2007). 1053-8119 Google Scholar

92. 

G. Gratton, M. Fabiani, P. M. Corballis, D. C. Hood, M. R. Goodman-Wood, J. Hirsch, K. Kim, D. Friedman, and E. Gratton, “Fast and localized event-related optical signals (EROS) in the human occipital cortex: Comparisons with the visual evoked potential and fMRI,” Neuroimage, 6 168 –180 (1997). 1053-8119 Google Scholar

93. 

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 1868 –1875 (2002). 1053-8119 Google Scholar

94. 

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

95. 

T. Mühlemann, D. Haensse, and M. Wolf, “Ein drahtloser Sensor für die bildgebende in vivo Nahinfrarotspektroskopie,” (2006) Google Scholar

96. 

Y. Hoshi, “Functional near-infrared spectroscopy: Potential and limitations in neuroimaging studies,” Neuroimaging Part A: International Review of Neurobiology, 238 –266 Elsevier Academic Press, New York (2005). Google Scholar
©(2007) Society of Photo-Optical Instrumentation Engineers (SPIE)
Yoko Hoshi M.D. "Functional near-infrared spectroscopy: current status and future prospects," Journal of Biomedical Optics 12(6), 062106 (1 November 2007). https://doi.org/10.1117/1.2804911
Published: 1 November 2007
Lens.org Logo
CITATIONS
Cited by 360 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Near infrared spectroscopy

Brain

Neuroimaging

Signal detection

Head

Tissue optics

Hemodynamics

Back to Top