Elsevier

NeuroImage

Volume 85, Part 1, 15 January 2014, Pages 6-27
NeuroImage

Review
A review on continuous wave functional near-infrared spectroscopy and imaging instrumentation and methodology

https://doi.org/10.1016/j.neuroimage.2013.05.004Get rights and content

Highlights

  • Comprehensive review on continuous wave functional near infrared imaging

  • Overview of currently available commercial near infrared imaging instrumentation

  • Review of technical aspects such as light sources, detectors and sensor arrangements

  • Review of methodological aspects, algorithms, and data analysis and its tool boxes

Abstract

This year marks the 20th anniversary of functional near-infrared spectroscopy and imaging (fNIRS/fNIRI). As the vast majority of commercial instruments developed until now are based on continuous wave technology, the aim of this publication is to review the current state of instrumentation and methodology of continuous wave fNIRI. For this purpose we provide an overview of the commercially available instruments and address instrumental aspects such as light sources, detectors and sensor arrangements. Methodological aspects, algorithms to calculate the concentrations of oxy- and deoxyhemoglobin and approaches for data analysis are also reviewed.

From the single-location measurements of the early years, instrumentation has progressed to imaging initially in two dimensions (topography) and then three (tomography). The methods of analysis have also changed tremendously, from the simple modified Beer-Lambert law to sophisticated image reconstruction and data analysis methods used today. Due to these advances, fNIRI has become a modality that is widely used in neuroscience research and several manufacturers provide commercial instrumentation. It seems likely that fNIRI will become a clinical tool in the foreseeable future, which will enable diagnosis in single subjects.

Introduction

Continuous light has been used to non-invasively investigate human tissue such as the breast, head and testes by transmitting the light through the body as early as at least in the 19th century (Bright, 1831, Curling, 1856, Cutler, 1929). More specifically, already in 1862 Hoppe-Seyler from Germany, described the spectrum of oxy-hemoglobin (O2Hb) (Perutz, 1995) and in 1864 Stokes from the United Kingdom added the spectrum of deoxy-hemoglobin (HHb) and consequently discovered the importance of hemoglobin for the oxygen transport (Perutz, 1995). In 1876 von Vierordt, also from Germany, analyzed tissue by measuring the spectral changes of light penetrating tissue when the blood circulation was occluded (Severinghaus, 2007, von Vierordt, 1876) and in 1894 Hüfner from Germany spectroscopically determined absolute and relative amounts of O2Hb and HHb in vitro (Hüfner, 1894). After decades of no relevant research in this field, in the 1930s the work on spectroscopic determination of tissue oxygenation was continued by several researchers. For example Nicolai, Germany, repeated the study of von Vierordt (Nicolai, 1932a, Nicolai, 1932b), and Matthes and Gross, Germany, demonstrated for the first time the spectroscopic determination of O2Hb and HHb in human tissue using two wavelengths, one in the red and near-infrared region (Matthes and Gross, 1938a, Matthes and Gross, 1938b, Matthes and Gross, 1938c).

In terms of quantification, an important first step was the discovery of the Beer–Lambert law first by the French mathematician Bouguer in 1729 (Bouguer, 1729). It is often attributed to the Swiss Lambert, although he cited Bouguers work in 1760 himself (Lambert, 1760). The law was extended by the German Beer to quantify concentrations in 1852 (Beer, 1852). Since the Beer–Lambert law is only valid in non-scattering media, it cannot be applied to biological tissue. Relatively recently therefore the modified Beer–Lambert law (MBLL) was developed by the British Delpy (Delpy et al., 1988), to take into account the light scattering. The MBLL is often used by many instruments described in this review. Further important steps were also analytical solutions of the diffusion equation (e.g. Arridge et al., 1992, Patterson et al., 1989) to quantitatively describe light transport in tissue.

Based on the insight of the relative transparency of the tissue including the skull in the near infrared range in 1977 Jobsis from the USA first demonstrated the feasibility to continuously and non-invasively monitor the concentration of O2Hb and HHb ([O2Hb] and [HHb]) in the brain (Jobsis, 1977). Therefore he is considered to be the initiator of near-infrared spectroscopy (NIRS).

His discovery led to designing and building of several NIRS instruments (Ferrari and Quaresima, 2012). All these instruments were continuous wave (CW) instruments. The term “continuous wave” means that the instrumentation is solely based on a light intensity measurement, i.e. near-infrared light is sent into the tissue and the intensity of the re-emerging (i.e. diffusely reflected) light is measured. This is in contrast to time resolved techniques such as time and frequency domain techniques, which, additionally to the intensity measurements also measure the time of flight, i.e. the time that the light needs to travel through the tissue. For a visualization of the three different techniques please refer to Fig. 1.

The disadvantage of CW systems is that they cannot fully determine the optical properties of tissue (i.e. light scattering (μs′) and absorption (μa) coefficients) and therefore the [O2Hb] and [HHb] cannot be determined absolutely. However, with a few reasonable assumptions it is possible to quantify changes in [O2Hb] and [HHb]. Therefore, during the first years, NIRS instruments were mostly trend monitors, employed to study various physiological conditions and clinical interventions. Much research was aimed at obtaining absolute values either by physiological maneuvers (e.g. Edwards et al., 1988, Wyatt et al., 1990) or enhancing the instrumentation (e.g. Matcher et al., 1995a, Matcher et al., 1995b, Wolf et al., 1997). Later time resolved techniques were developed and became available and enabled to determine absolute values. This will not be discussed further, because it is not within the scope of this review.

1993 was a crucial year in the development of functional NIRS (fNIRS) of the brain. In the same year four research groups published results and demonstrated that it is possible to non-invasively investigate brain activity using fNIRS (Chance et al., 1993, Hoshi and Tamura, 1993, Kato et al., 1993, Villringer et al., 1993). Brain activity leads to an increase in oxygen consumption, which is accompanied by an increase in cerebral blood flow due to neurovascular coupling. This leads to a change in the local [O2Hb] and [HHb] (Wolf et al., 2002), which can be detected non-invasively by fNIRS. These first measurements were carried out with simple instruments, which measured at one or a few locations. Since brain activity in response to a stimulation occurs only at specific locations in the brain, when measuring just at one location it is often difficult to find the correct position on the head for the measurement. In addition, there is a scientific interest in measuring a spatial pattern, how the brain activity affects an area of the brain.

A next major step in the development was to design imaging instruments that covered a larger area of the head and enabled mapping of brain activity, i.e. to deliver topographic information (Ferrari and Quaresima, 2012, Maki et al., 1995). This had several advantages: it enabled to localize brain activity and the precise localization of the sensor was less important. This technology is called functional near infrared imaging (fNIRI). On the one hand, it was quite clear that it is highly important to expand the interrogated area by using imaging systems. On the other hand, quantification is not that important in neuroscience, i.e. it is more important to statistically significantly detect a change in brain activity than to quantify it in absolute terms. For these reasons up to today most imaging systems are based on CW technology. In addition, time resolved systems have a lower time resolution, are more expensive and the time of flight is generally a more noisy parameter than the intensity and therefore not useful for detecting small functional activations. In contrast, CW systems are relatively low cost, can be miniaturized and wireless systems and can be applied unobtrusively in everyday life situations or even freely moving animals (Muehlemann et al., 2008).

As a next step, sensor arrangements where several source detector distances are measured simultaneously, so-called overlapping measurements, enabled to apply tomographic approaches, i.e. image reconstructions in three dimensions (Joseph et al., 2006).

Today fNIRI has entered neuroscience as a research tool. It has been shown that fNIRI is reliable and trustworthy for research based on investigating groups of subjects, although reliability in single subjects is not sufficient yet (Kono et al., 2007, Plichta et al., 2006, Plichta et al., 2007a, Plichta et al., 2007b, Schecklmann et al., 2008). Consequently, the number of publications on fNIRI in neuroscience has increased exponentially within the last years.

One of the next aims is to apply fNIRI clinically. For this purpose it will be compulsory to ensure a high reproducibility in single subjects. However, a sufficient reliability on the single subject level has not yet been achieved (Biallas et al., 2012a, Biallas et al., 2012b, Kono et al., 2007, Plichta et al., 2006, Plichta et al., 2007a, Plichta et al., 2007b, Schecklmann et al., 2008). Hence, some research is currently focused on improving the reliability. Possible reasons for the lack in reliability are the superficial tissue (i.e. light has to penetrate several tissue layers such as e.g. skin and skull before it reaches the brain) or systemic physiological changes, which contaminate the signal of the brain and possibly instrumental shortcomings such as an insufficiency in spatial resolution and/or signal to noise ratio (SNR). Generally there are several possibilities to improve reliability: On the instrumental level, it is important to select appropriate wavelengths, light sources, detectors, and geometrical arrangements to avoid crosstalk and ensure a high SNR; on a methodological level the aim is to reduce the influence of superficial tissue or systemic components.

The aim of this paper is to review the current state of instrumentation and methodology of CW fNIRI. For this purpose we will give an overview of the commercially available instruments and address instrumental aspects such as light sources, detectors and sensor arrangements. Methodological aspects such as algorithms to determine [O2Hb] and [HHb] and data analysis will also be reviewed.

There is a variety of terms used for NIRS. In general the term NIRS is often used as an overarching term for the whole technology, but in principle it only refers to NIRS systems measuring at single locations with up to four sensors, but without imaging capacity. Imaging systems, which we will call NIRI here, have more than 4 channels and produce two or three dimensional images. In the literature, 2D imaging systems are also called near infrared topography or mapping, or diffuse optical imaging. 3D imaging systems are also called near infrared tomography or mapping, or diffuse optical tomography or imaging. The term “diffuse” refers to the fact that due to the high μs′ in tissue the propagation of photons through tissue can be modeled as a diffusion process. Unfortunately, the term “diffuse” is often misinterpreted as blurred or fuzzy, i.e. referring to a negative connotation of the technique, which clearly is not meant or adequate. As for magnetic resonance imaging (MRI), which was formerly referred to as nuclear magnetic resonance, the term “nuclear” falsely evoked the association of a radioactive method, we propose to abandon the term “diffuse” in the future to avoid this negative connotation. The spatially resolved measurement of brain activity in two and three dimensions we call functional NIRI (fNIRI). All systems (except for one) and methods reviewed here are generally about CW fNIRI and therefore the term “CW” will be omitted below.

Section snippets

Overview of commercially available imaging instrumentation

There is already a wide variety of commercially available fNIRI devices in the market. Therefore, researchers are likely to find devices suited for their respective needs. System complexity ranges from few sources and detectors suitable to image certain brain areas to systems covering the whole head. Some devices use sensor patches (D2, D3, D4, D7, D14, see Table 1) with integrated components whereas all others use optical fibers and flexible head-caps allowing for adjustments to the individual

Technological design aspects of fNIRI

From an engineer's point of view, CW fNIRI merely requires to switch on a NIR light source, couple the emitted light into the scalp, and measure the diffuse reflectance that re-emerges from the tissue a few centimeters distant from that source (Ferrari and Quaresima, 2012, Ferrari et al., 2004, Giacometti and Diamond, 2013, Hoshi, 2003, Strangman et al., 2002, Wolf et al., 2007). Hence, CW fNIRI instruments come in comparably simple setups, allowing specialization towards wearable, miniaturized

Approaches to determine the [O2Hb] and [HHb]

There are different algorithms to convert raw light intensity data into [O2Hb], [HHb] and the total hemoglobin concentration ([tHb], i.e. the sum of [O2Hb] and [HHb]) or tissue oxygen saturation (StO2). The most common ones in fNIRI are the MBLL (Delpy et al., 1988) and multi-distance (MD) approaches, making use of several source–detector distances.

In the following we will discuss properties and assumptions of the different approaches. Independent of the approach when using CW devices it is not

Future directions of fNIRI

As one can see in the exponential growth of the number of publications on fNIRI, the development of fNIRI is continuing and much progress has already been made during the last 35 years. It is foreseeable that this progress will be sustained, because the potential of fNIRI is still far from being fully exploited.

For continuous wave fNIRI instrumentation, there will be an incentive to build instruments that incorporate an increased number of light sources and detectors in order to take more

Conclusions

fNIRI celebrates its 35th birthday. From the single-location measurements at the beginning the instrumentation has developed into first two dimension (topography) and then three dimensions (tomography). Also the methods of analysis have changed tremendously, from the simple modified Beer-Lambert law to sophisticated image reconstruction and data analysis methods. Due to these advances, fNIRI has become a modality that is widely used in neuroscience research and a number of manufacturers provide

Acknowledgments

We acknowledge funding from the Swiss National Science Foundation, Nano-Tera, CHIRP1 ETH Research Grant CH1-02 09-3, and the Clinical Research Priority Programs Tumor Oxygenation and Molecular Imaging. FS would like to thank Dr. Ilias Tachtsidis (University College London, UK) for fruitful discussion about fNIRI signal processing matters.

Conflict of interest statement

We declare that we have no conflict of interest.

References (314)

  • H. Ellis

    Anatomy of head injury

    Surgery

    (2012)
  • M. Ferrari et al.

    A brief review on the history of human functional near-infrared spectroscopy (fNIRS) development and fields of application

    Neuroimage

    (2012)
  • L. Gagnon et al.

    Improved recovery of the hemodynamic response in diffuse optical imaging using short optode separations and state-space modeling

    Neuroimage

    (2011)
  • L. Gagnon et al.

    Short separation channel location impacts the performance of short channel regression in NIRS

    Neuroimage

    (2012)
  • L. Gagnon et al.

    Further improvement in reducing superficial contamination in NIRS using double short separation measurements

    Neuroimage

    (2014)
  • P.J. Goadsby et al.

    Stimulation of an intracranial trigeminally-innervated structure selectively increases cerebral blood flow

    Brain Res.

    (1997)
  • L. Gygax et al.

    Prefrontal cortex activity, sympatho-vagal reaction and behaviour distinguish between situations of feed reward and frustration in dwarf goats

    Behav. Brain Res.

    (2013)
  • C. Habermehl et al.

    Somatosensory activation of two fingers can be discriminated with ultrahigh-density diffuse optical tomography

    Neuroimage

    (2012)
  • L. Holper et al.

    Between-brain connectivity during imitation measured by fNIRS

    Neuroimage

    (2012)
  • L. Holper et al.

    The teaching and the learning brain: a cortical hemodynamic marker of teacher–student interactions in the Socratic dialog

    Int. J. Educ. Res.

    (2013)
  • Y. Hoshi et al.

    Detection of dynamic changes in cerebral oxygenation coupled to neuronal function during mental work in man

    Neurosci. Lett.

    (1993)
  • C.B. Akgül et al.

    Extraction of cognitive activity-related waveforms from functional near-infrared spectroscopy signals

    Med. Biol. Eng. Comput.

    (2006)
  • P.G. Al-Rawi et al.

    Evaluation of a near-infrared spectrometer (NIRO 300) for the detection of intracranial oxygenation changes in the adult head

    Stroke

    (2001)
  • S.R. Arridge et al.

    The theoretical basis for the determination of optical pathlengths in tissue: temporal and frequency analysis

    Phys. Med. Biol.

    (1992)
  • S.R. Arridge et al.

    A finite element approach for modeling photon transport in tissue

    Med. Phys.

    (1993)
  • Artinis, The Netherlands
  • H. Atsumori et al.

    Development of wearable optical topography system for mapping the prefrontal cortex activation

    Rev. Sci. Instrum.

    (2009)
  • A. Beer

    Bestimmung der Absorption des rothen Lichts in farbigen Flüssigkeiten

    Ann. Phys. Chem.

    (1852)
  • P.B. Benni et al.

    Validation of the CAS neonatal NIRS system by monitoring vv-ECMO patients: preliminary results

    Adv. Exp. Med. Biol.

    (2005)
  • M. Biallas et al.

    Reproducibility and sensitivity of detecting brain activity by simultaneous electroencephalography and near-infrared spectroscopy

    Exp. Brain Res.

    (2012)
  • M. Biallas et al.

    Multimodal recording of brain activity in term newborns during photic stimulation by near-infrared spectroscopy and electroencephalography

    J. Biomed. Opt.

    (2012)
  • D.A. Boas et al.

    Can the cerebral metabolic rate of oxygen be estimated with near-infrared spectroscopy?

    Phys. Med. Biol.

    (2003)
  • D.A. Boas et al.

    Improving the diffuse optical imaging spatial resolution of the cerebral hemodynamic response to brain activation in humans

    Opt. Lett.

    (2004)
  • P. Bouguer

    Essai d'optique, sur la gradation de la lumiere

    (1729)
  • A. Bozkurt et al.

    Safety assessment of near infrared light emitting diodes for diffuse optical measurements

    Biomed. Eng. Online

    (2004)
  • A. Bozkurt et al.

    A portable near infrared spectroscopy system for bedside monitoring of newborn brain

    Biomed. Eng. Online

    (2005)
  • R. Bright

    Diseases of the brain and nervous system

  • E. Bullmore et al.

    Statistical methods of estimation and inference for functional MR image analysis

    Magn. Reson. Med.

    (1996)
  • D. Canova et al.

    Inconsistent detection of changes in cerebral blood volume by near infrared spectroscopy in standard clinical tests

    J. Appl. Physiol.

    (2011)
  • B. Chance et al.

    Cognition-activated low-frequency modulation of light absorption in human brain

    Proc. Natl. Acad. Sci. USA

    (1993)
  • J. Choi et al.

    Noninvasive determination of the optical properties of adult brain: near-infrared spectroscopy approach

    J. Biomed. Opt.

    (2004)
  • K. Çiftçi et al.

    Constraining the general linear model for sensible hemodynamic response function waveforms

    Med. Biol. Eng. Comput.

    (2008)
  • M. Cope et al.

    A CCD spectrophotometer to quantitate the concentration of chromophores in living tissue utilising the absorption peak of water at 975 nm

    Adv. Exp. Med. Biol.

    (1989)
  • A. Corlu et al.

    Uniqueness and wavelength optimization in continuous-wave multispectral diffuse optical tomography

    Opt. Lett.

    (2003)
  • A. Corlu et al.

    Diffuse optical tomography with spectral constraints and wavelength optimization

    Appl. Opt.

    (2005)
  • T. Correia et al.

    Identification of the optimal wavelengths for optical topography: a photon measurement density function analysis

    J. Biomed. Opt.

    (2010)
  • J.P. Culver et al.

    Optimization of optode arrangements for diffuse optical tomography: a singular-value analysis

    Opt. Lett.

    (2001)
  • T.B. Curling

    A Practical Treatise on the Diseases of the Testis, and of the Spermatic Cord and Scrotum

    (1856)
  • M. Cutler

    Transillumination as an aid in the diagnosis of breast lesions

    Surg. Gynecol. Obstet.

    (1929)
  • S.N. Davie et al.

    Impact of extracranial contamination on regional cerebral oxygen saturation: a comparison of three cerebral oximetry technologies

    Anesthesiology

    (2012)
  • Cited by (1339)

    View all citing articles on Scopus
    View full text